reductive dehalogenation of

126
REDUCTIVE DEHALOGENATION OF CHLOROETHENES, DICHLOROPHENOL, AND CHLOROBENZENES A Dissertation Presented to the Faculty of the Graduate School of Cornell University in Partial Fulfillment of the Requirements for the Degree of Doctor of Philosophy by Jennifer Mon Yee Fung August 2008

Upload: others

Post on 04-Feb-2022

11 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: REDUCTIVE DEHALOGENATION OF

REDUCTIVE DEHALOGENATION OF

CHLOROETHENES, DICHLOROPHENOL, AND CHLOROBENZENES

A Dissertation

Presented to the Faculty of the Graduate School

of Cornell University

in Partial Fulfillment of the Requirements for the Degree of

Doctor of Philosophy

by

Jennifer Mon Yee Fung

August 2008

Page 2: REDUCTIVE DEHALOGENATION OF
Page 3: REDUCTIVE DEHALOGENATION OF

REDUCTIVE DEHALOGENATION OF CHLOROETHENES,

DICHLOROPHENOL, AND CHLOROBENZENES

Jennifer Mon Yee Fung, Ph.D.

Cornell University 2008

The improper disposal of chlorinated compounds has become an

environmental concern and its remediation is a priority. Dehalococcoides

ethenogenes strain 195 is able to reductively dehalogenate the toxic

compounds, tetrachloroethene (PCE), trichloroethene (TCE) and 2,3-

dichlorophenol (2,3-DCP) to less toxic products, ethene and 3-

monochlorophenol. Its genome contains 19 putative reductive dehalogenase

(RD) genes including pceA (PCE RD), tceA (TCE RD), DET0162 and

DET1559. Differential transcription was examined when strain 195 was

growing on PCE, TCE, or 2,3-DCP. Gene product and function was confirmed

using proteomic analysis and resting cell assays. When grown on PCE or

TCE transcript levels of tceA, pceA and DET0162 were several folds higher

than housekeeping gene, rpoB. DET1559 was only expressed in PCE-grown

cells. In 2,3-DCP-grown cells, pceA and DET0162 were the only RD genes

expressed and tceA transcript level was the 300-fold lower than in PCE-grown

cells. DET0162 contains a translational stop codon and is presumed

nonfunctional and was not detected in proteomic analysis. PceA was the only

RD detected in 2,3-DCP-grown cells. PCE- and 2,3-DCP-grown resting cells

were able to reductively dehalogenate PCE and 2,3-DCP without lag. PceA

has been identified as the 2,3-DCP reductive dehalogenase. In the case of

dichlorobenzene (DCB) and monochlorobenzene (MCB), no microorganism

Page 4: REDUCTIVE DEHALOGENATION OF

has been identified capable of their reductive dehalogenation. Using

historically chlorobenzene-contaminated sediments, microcosms were used to

enrich for DCB and MCB dehalogenators. Sediment microcosms were able to

reductively dehalogenate all three DCB isomers to MCB and benzene over

5,000 µmoles/liter in some microcosms. If only given MCB, reductive

dehalogenation to benzene only occurred in specific sediment samples and

was considered unreliable. Inoculating sediment microcosms with sediment

slurry from DCB- or MCB-dehalogenating microcosm reduced lag time of MCB

reductive dehalogenation to benzene and allowed for reliable growth on MCB

in the absence of DCBs. A 1,2-DCB mixed culture was established by

transferring sediment slurry from DCB-dehalogenating microcosms into

mineral salts medium supplemented with yeast extract. 16S rRNA gene clone

library of the mixed culture contained sequences identical >99% to

Dehalobacter restrictus, a known dehalogenator. Dehalobacter population,

monitored by real time PCR, was linked to reductive dehalogenation of DCBs

and its 16S rRNA gene sequences accounted for 53% of the total bacterial

sequences in the mixed culture. As the toxicity of the chlorinated compounds

can be altered by reductive dehalogenation, characterization of

dehalogenating microorganisms is needed to design effective bioremediation

strategies.

Page 5: REDUCTIVE DEHALOGENATION OF

BIOGRAPHICAL SKETCH

Jennifer Mon Yee Fung was born on July 17, 1979 in San Francisco,

California to John and Helen Fung. Influenced by family and school trips to

the beach, Lake Tahoe, and Muir Woods, Jennifer pursued an education in

the biological sciences. At Skyline Community College, she was introduced to

the diverse world of microbiology and took a special interest in the field of

bioremediation. In 1999 Jennifer transferred to the University of California, San

Diego, Revelle College, to pursue a bachelor degree in microbiology and

began working in the laboratory of Dr. Bradley Tebo at Scripps Institution of

Oceanography where she studied hexavalent chromium reduction by

Shewanella spp. Determined to gain a deeper understanding of microbially-

mediated bioremediation, Jennifer joined the graduate program in the

Department of Microbiology at Cornell University in 2002. At Cornell

University, Jennifer worked under the guidance of Dr. Stephen Zinder on

reductive dehalogenation of halogenated pollutants. Upon completion of her

degree, Jennifer will begin work on bacterial production of biofuels at

LanzaTech in Auckland, New Zealand.

iii

Page 6: REDUCTIVE DEHALOGENATION OF

This work is dedicated to my parents, John and Helen Fung, as well as my

sisters and brother, Christina, Denise and Clifford Fung, for their unconditional

support during my entire educational pursuit.

iv

Page 7: REDUCTIVE DEHALOGENATION OF

ACKNOWLEDGEMENTS

There are many individuals who I would like to express my eternal gratitude

and appreciation for their help in my pursuit of a Ph.D. at Cornell University.

Dr. Stephen Zinder was a wonderful mentor and has made me into a better

scientist. His knowledge of bacterial metabolism was intimidating and truly,

something to strive towards. More importantly, Steve showed me that a

successful scientist could have a healthy work-life balance. I would also like to

specially thank Dr. Lorenz Adrian and Dr. Robert Morris who seemed pleased

with every experiment that I performed. Their positive reinforcement was

highly appreciated to a young graduate student.

My family has kept me grounded and provided me with support throughout this

time. My parents, John and Helen Fung, have provided an emotional safe

haven during the rougher moments. My parents always believed in my

academic ability despite horrible grades in high school and my short tenure in

community college. They made me realize that you’re never too old to receive

care packages and home was only a plane flight away. My siblings, Christina,

Denise and Clifford, all choose different paths in life and their life experiences

have helped keep my own experiences in perspective.

Special thanks goes out to some influential teachers in my life all of whom

renewed my interest in education: Ms. Debbie Molof, Dr. Christine Case, and

Dr. Anna Obrazstova.

v

Page 8: REDUCTIVE DEHALOGENATION OF

I would also like to thank the many friends and colleagues that have provided

scientific and personal insight throughout the years. Wing Hall would not

function without Shirley Cramer, John Enright, Cathy Shappell, Doreen

Dineen, and Patti Butler. All past Zinder lab members with special thanks to

Suz Brauer, Christine Sun, Jocelyn Bosco, Jenny Nelson, and Heather

Fullerton. Heather O’Neil, Sarita Raengpradub and Abbie Wise (the View) are

the original ice cream crew. Andreas Tobas and Ryan Seipke are my secret

keepers. Pete Choi always kept it real whether I wanted it or not. Letal

Salzberg and Jenna Mendell were my emotional life support when I flatlined.

Gerry M.D.K. Lorigan kept me fed my during my finals months in Ithaca.

Finally, Marisa Fontanoz who told me to find a job in a research lab during our

first year at the University of California, San Diego; something that I never

considered.

vi

Page 9: REDUCTIVE DEHALOGENATION OF

TABLE OF CONTENTS

BIOGRAPHICAL SKETCH………………………………………………………. iii

DEDICATION……………………………………………………………………… iv

ACKNOWLEDGEMENTS…………………………...…………………………… v

LIST OF FIGURES…………………………...……………………………..…… vii

LIST OF TABLES…..…………………………...…………………………..….… ix

CHAPTER ONE: Review of Reductive Dehalogenation …………….……….. 1

CHAPTER TWO: Expression of Reductive Dehalogenase Genes in

Dehalococcoides ethenogenes strain 195 Growing on

Tetrachloroethene, Trichloroethene or 2,3-Dichlorophenol ………………… 25

CHAPTER THREE: Reductive Dehalogenation of Dichlorobenzenes and

Monochlorobenzene to Benzene in Microcosms ……………………………. 51

CHAPTER FOUR: Preliminary Characterization of Microbial Enrichment

Cultures that Reductively Dechlorinate Dichlorobenzenes to

Monochlorobenzene ……………………………………………………………. 72

CHAPTER FIVE: Summary and Implications in the Field of Bioremediation 97

Page 10: REDUCTIVE DEHALOGENATION OF

LIST OF FIGURES

Figure 1.1 Two mechanisms of reductive dehalogenation………………………1

Figure 1.2 Phylogenetic tree of reductive dehalogenases ……………………..6

Figure 1.3 Schematic of RD gene structure…………………………………….11

Figure 2.1 Transcript levels of RD genes in Dehalococcoides ethenogenes

strain 195 in cells grown on perchloroethylene, trichloroethene, or 2,3-

dichlorophenol………………………………………………………………………35

Figure 2.2 Upstream regions of tceA, pceA, infA, and DET1407 from

Dehalococcoides ethenogenes strain 195 and vcrA from strain VS……….…37

Figure 2.3 Protein coverage from Dehalococcoides ethenogenes strain 195

grown on perchloroethylene or 2,3-dichlorophenol………………………….….39

Figure 2.4 Reductive dehalogenation activity of resting cells of

Dehalococcoides ethenogenes strain 195 grown on perchloroethylene or 2,3-

dichlorophenol………………………………………………………………………40

Figure 3.1 Reductive dehalogenation of dichlorobenzene isomers to

monochlorobenzene and benzene in sediment microcosms…………………..59

Figure 3.2 Reductive dehalogenation of individual dichlorobenzene isomers to

monochlorobenzene and benzene in sediment microcosms………………..…60

Figure 3.3 Reductive dehalogenation of monochlorobenzene in microcosms

solely given monochlorobenzene…………………………………………………61

Figure 3.4 Reductive dehalogenation of dichlorobenzene to

monochlorobenzene and benzene in inoculated microcosms where inoculum

was dechlorinating dichlorobenzenes……………….……………………………62

Figure 3.5 Reductive dehalogenation of dichlorobenzene to

monochlorobenzene and benzene in inoculated microcosms where inoculum

was dechlorinating monochlorobenzene…………………………………………63

vii

Page 11: REDUCTIVE DEHALOGENATION OF

Figure 4.1 Reductive dehalogenation of dichlorobenzene isomers to

monochlorobenzene in first and sixth transfer cultures………………………...79

Figure 4.2 Benzene production in the seventh transfer culture……………….81

Figure 4.3 Lost of reductive dehalogenation activity in the presence of

vancomycin………………………………………………………………………….82

Figure 4.4 Reductive dehalogenation of individual dichlorobenzene isomers

by sixth transfer cultures…………………………………………………………...83

Figure 4.5 Comparison of 16S rRNA gene copy number of total bacteria,

Dehalococcoides, and Dehalobacter from fourth transfer cultures grown with

or without dichlorobenzenes………………………………………...…………….85

Figure 4.6 Comparison of 16S rRNA gene copy number of total bacteria,

Dehalococcoides, and Dehalobacter from seventh transfer cultures grown with

or without dichlorobenzenes………………………...…………………………….86

Figure 4.7 Hydrogen-fed mixed culture with sodium sulfide as a reducing

agent and 16S rRNA gene copy number after 25 days………………………...87

viii

Page 12: REDUCTIVE DEHALOGENATION OF

LIST OF TABLES

Table 1.1 Overview of dehalogenating microorganisms…………………..……4

Table 2.1 List of gene targets and primer sequences…………………………32

Table 4.1 Growth conditions used to establish a 1,2-dichlorobenzene mixed

culture.……………………………………………………………………………….80

Table 4.2 Phylogenetic groups identified in clone library from fourth transfer

culture grown with or without dichlorobenzenes………………………………...84

Table 5.1 Sequence identity between the reductive dehalogenases in strain

195………………………………………………………………………………….101

ix

Page 13: REDUCTIVE DEHALOGENATION OF

CHAPTER ONE

REVIEW OF REDUCTIVE DEHALOGENATION

Introduction

The degradation of chlorinated compounds has been the focus of numerous

studies due to their extensive use in industrial applications and subsequent

disposal into the environment. The Environmental Protection Agency (Clean

Water Act) has designated a large number of chlorinated compounds as

priority pollutants, including polychlorinated biphenyls (PCBs), chlorinated

benzenes, ethenes, methanes, phenols, and dioxins

(www.epa.gov/watertrain/cwa/). These compounds are either known or

potential causative agents of serious health complications such as liver and

kidney cancers (20). Since many chlorinated compounds are dense non-

aqueous phase liquids (DNAPLs), they have the ability to migrate into the

anoxic region of the sediment and serve as electron acceptors in a process

known as respiratory reductive dehalogenation (Figure 1.1).

Figure 1.1. Two mechanisms of reductive dehalogenation A) hydrogenolysis of cis-dichloroethene to vinyl chloride as observed in Dehalococcoides and B) dichloroelimination of 1,2-dichloroethane to ethene as observed in Desulfitobacterium dichloroeliminans strain DCA1

1

Page 14: REDUCTIVE DEHALOGENATION OF

This process can be found across different phyla such as Chloroflexi,

Firmicutes, and Proteobacteria in a variety of environments (Table 1.1) (55).

Depending on the microorganism, this physiological process can be the only

or one of many respiratory pathways. Understanding the microbial community

involved in reductive dehalogenation is critical, as the fate of chlorinated

compounds in the environment can be determined by the presence of specific

dehalogenators (19).

The calculated redox potential of halogenated substrates is between 200-400

mV, well situated in the range of anaerobic respiration (11). The halogenated

substrates can vary from complex aromatics like 1,2,3,4-tetrachloronapthalene

to simple two carbon compounds like chloroethenes (55). Some bacteria such

as Geobacter lovleyi and Desulfitobacterium spp. are able to carry out

alternative forms of respiration such as sulfate reduction and can use multiple

substrates as electron donors (59, 64). Other bacteria such as

Dehalococcoides spp. and Dehalobacter spp. are known only to respire

halogenated compounds and utilize hydrogen as the electron donor (55).

Insight into the molecular mechanisms of reductive dehalogenation has been

accumulating due to the sequenced genomes of dehalogenators such as

Dehalococcoides spp. and Desulfitobacterium spp. (26, 44, 54). There has

been much progress in the understanding of the molecular biology of reductive

dehalogenases (RD), the proteins responsible for reductive dehalogenation.

In general, RDs are membrane bound and contain a corrinoid cofactor except

for the 3-chlorobenzoate RD of Desulfomonile tiedjei, which contains a heme

group (42). Most RD genes are flanked by putative regulatory genes, and

2

Page 15: REDUCTIVE DEHALOGENATION OF

several show signs of genetic mobility (26, 30). Some bacteria hold large

reservoirs of RD genes, like Dehalococcoides strain CBDB1 with 32, of which

only two have known function (2, 26). While the large number of RD genes

suggests a great capacity for reductive dehalogenation, little is known about

substrate specificity and range.

The goals of this review are to:

1. Give an overview of the diversity of dehalogenating microorganisms

and their suite of chlorinated substrates

2. Survey the molecular biology of reductive dehalogenation

Microbiology of Reductive Dehalogenation

Dehalococcoides and its relatives

Dehalococcoides spp. and its relatives, strain DF-1 and o-17, are members of

the Chloroflexi phylum and specialize in reductive dehalogenation (9, 55, 67).

Dehalococcoides ethenogenes strain 195 was the first bacterium isolated able

to completely reductively dehalogenate perchloroethylene (PCE) to vinyl

chloride (VC) and eventually, ethene with the last step being cometabolic (34).

The genome of strain 195 reveals 19 potential RD genes including the

characterized pceA (DET0318), the gene responsible for PCE to TCE

reductive dehalogenation and tceA, responsible for TCE to ETH (29, 54).

Strain 195 was later shown to dehalogenate chlorinated aromatics as well as

polybrominated diphenyl ethers (1, 13, 18). Strains BAV1 and VS have been

isolated and reductively dehalogenate VC as an energetically favorable

reaction but are unable to dehalogenate PCE or TCE (8, 17). Strain CBDB1

3

Page 16: REDUCTIVE DEHALOGENATION OF

Table 1.1 Overview of dehalogenating microorganisms

Microorganism Halogenated Compound e-donor Genome

Sequenced Total No. of RD Genes

Identified RD genes Phylum Ref.

Dehalococcoides strain 195

CE, CP, CB, PCB, PCDD/F, CN H2 yes 19 2 (pceA, tceA) Chloroflexi (1, 13, 29, 34,

54)

Dehalococcoides strain CBDB1 CB, CP, CD H2 yes 32 1 (cbrA) Chloroflexi (1-3, 6)

Dehalococcoides strain BAV1 CE H2 yes 10 1 (bvcA) Chloroflexi (17, 25)

Strain DF-1 CB, PCB H2 no unk 0 Chloroflexi (66, 67)

Strain TCA1 CA formate,H2 no unk 0 Firmicutes (58)

Dehalobacter restrictus CE H2 no unk 1 (pceA) Firmicutes (21, 31)

Desulfitobacterium hafiense strain Y51 CE, CA formate, lactate, pyruvate yes 2 1 (pceA) Firmicutes (64)

Desulfitobacterium hafiense strain DCB-2 CP, CE, 3Cl4OHPA, TCMP, TCHQ formate, lactate, pyruvate yes 7 1 (cprA) Firmicutes (64)

Sulfurospirillum multivorans CE formate,H2 no unk 1 (pceA) ε-proteobacteria (40, 51)

Geobacter lovleyi sp. nov. strain SZ CE acetate, H2 yes 2 1 (pceA) δ-proteobacteria (59)

CE, chloroethenes; CP, chlorophenols; CB, chlorobenzenes; PCB; polychlorinated biphenyls; PCDD/F, polychlorinated dibenzno-p-dioxins/furans; CN, chlorinated napthalenes;

CD, chlorinated dioxins; CA, chlorinated ethanes; 3Cl4OHPA, 3-chloro-4-hydroxy-phenylacetate; TCMP, 2,3,5,6-tetrachloro-4-methoxyphenol; TCHQ, tetrachlorohydroquinone.

4

Page 17: REDUCTIVE DEHALOGENATION OF

was the first isolated bacterium able to reductively dehalogenate chlorinated

benzenes and dioxins but it is unable to grow on chlorinated ethenes (3, 6).

The genome sequence of Dehalococcoides strain CBDB1 revealed 32

potential RD genes including cbdbA84 (cbrA), responsible for 1,2,3-

trichlorobenzene reductive dehalogenation (2). Twelve RD genes between

strain 195 and CBDB1 share up to 95.4% amino acid sequence similarity,

suggesting a partial overlap of substrates (26). Multiple RD genes have been

identified in Dehalococcoides strains FL-2, KB1 (Dehalococcoides-containing

mixed culture), and BAV1 (14, 14, and 10, respectively) (22). With their large

reservoir of RD genes, Dehalococcoides spp. are likely to have a diverse

substrate range that will be useful for site remediation (Figure 1.2).

Dehalococcoides has also been shown to be important in bioremediation field

studies and dehalogenating mixed cultures. Field studies done in a variety of

contaminated sites throughout North America and Europe indicate that

Dehalococcoides is the key player in the complete reduction of chlorinated

ethenes (19). In sediment-free mixed culture derived from the Housatonic

River, Massachusetts, Dehalococcoides population growth has been linked to

the reductive dehalogenation of Aroclor 1260 (polychlorinated biphenyl

mixture) (4). Similar results were shown with trichlorinated dibenzo-p-dioxins

with mixed cultures originating from the Spittelwasser, Germany (12). With

sediment microcosms derived from three different sites (New York, California,

and Maryland), Yan et al. (68) demonstrated 2,3,4,5-tetrachlorobiphenyl

dehalogenation and presence of Dehalococcoides-like populations.

Commercially available mixed culture, KB-1, where the bacterial population.

5

Page 18: REDUCTIVE DEHALOGENATION OF

Figure 1.2. Phylogenetic tree of reductive dehalogenases of Dehalococcoides strain 195 (DET), strain CBDB1 (cbdb), strain BAV1 (bav), Desulfitobacterium hafniense strain DCB-2 (Dhaf), Geobacter lovelyi strain SZ (Glov) and Photobacterium profundum strain 3TCK (3TCK). The length of the bar represents 10% sequence diversity. Protein distances were calculated with Prodist using Jones-Taylor-Thornton matrix and tree drawn with Neighbor-joining algorithm.

6

Page 19: REDUCTIVE DEHALOGENATION OF

contains Dehalococcoides spp., has been successfully used to bioremediate

sites contaminated with chlorinated ethenes (32)

Strains DF-1 and o-17 are related to Dehalococcoides with 89% 16S rRNA

gene identity to strain 195 and they reductively dehalogenate PCB from the

para position and the ortho position respectively (9, 67). Strain DF-1 is the

only isolated microorganism shown to be capable of growth by reductively

dechlorinating PCBs. It can only use hydrogen as an electron donor and

requires an unknown growth factor from a Desulfovibrio (33). Strain o-17 is

grown as a mixed culture and uses acetate to drive reductive dehalogenation.

Dehalobacter

Dehalobacter spp. are members of the Firmicutes and are only known to

respire halogenated compounds. Dehalobacter restrictus can reductively

dehalogenate PCE and TCE to cis-dichloroethene (cDCE) and only uses

hydrogen as an electron donor (21). Strain TCA is the first bacterium shown

to dehalorespire chloroethanes and can use formate or hydrogen as an

electron donor (58). Grostern et al. (16) used a combination of culture

independent metod to identify Dehalobacter sequences in a chloroethane

dehalogenating mixed consortium . A Dehalobacter sp. has been identified in

a coculture with Sedimentibacter sp. that is able to reductively dehalogenate

β-hexachlorocyclohexane to MCB and benzene (62).

Desulfitobacterium

Desulfitobacterium spp. are also members of Firmicutes but show a marked

difference in physiological characteristics compared to Dehalobacter and

7

Page 20: REDUCTIVE DEHALOGENATION OF

Dehalococcoides. Desulfitobacterium spp. can utilize a diverse range of

electron acceptors such nitrate, sulfite, and fumarate along with halogenated

substrates. As a genus, it is capable of dehalogenating chloroethenes, -

ethanes, -phenols, -butanes and more. When presented with multiple electron

acceptors: nitrate, sulfite, 3-chloro-4-hydroxyphenylacetic acid (3Cl4OHPA),

and fumarate, Desulfitobacterium dehalogenans reductively dehalogenates

3Cl4OHPA and nitrate simultaneously followed by sulfite and fumarate.

Certain strains such as D. dichloroeliminans strain DCA1 reductively

dehalogenate via dichloroelimination instead of hydrogenolysis (10). The

genome of Desulfitobacterium hafniense strain Y51, a bacterium capable of

reductively dehalogenating chloroethanes and PCE, contains two reductive

dehalogenase genes including pceA (44, 60). The genome of

Desulfitobacterium hafniense strain DCB-2, a bacterium capable of reductively

dehalogenating chlorophenols and 3Cl4OHPA, is being annotated and

tentatively contains seven RD genes. Strain DCB-2 has a larger substrate

range than strain Y51 so a larger library of RD genes was expected. A

genome comparison between this strain and Y51 will reveal evolutionary

insight into the development of reductive dehalogenation.

Other Dehalogenating Microorganisms and Sediment Microcosms

Other dehalogenating bacteria include Sulfurospirillum multivorans (ε-

Proteobacteria) and Geobacter lovleyi strain SZ (δ-Proteobacteria) (51, 59).

S. multivorans can reductively dehalogenate PCE using either hydrogen or

formate as an electron donor. It is also capable of using other electron

acceptors such as fumarate and selenate. Its PCE reductive dehalogenase

8

Page 21: REDUCTIVE DEHALOGENATION OF

(PCE RD) is the best-characterized RD (see below). It is constitutively

expressed, and when grown on PCE is membrane-associated and when

grown on fumarate is located in the cytoplasm. G. lovleyi strain SZ is unique

in its ability to simultaneously dehalogenate PCE and reduce radionuclides

like hexavalent uranium. It may play a major role in bioremediation of sites

contaminated with toxic solvents and metals. Photobacterium profundum

strain 3TCK (δ-Proteobacteria) contains a putative reductive dehalogenase

gene but is not known to perform reductive dehalogenation (63).

Reductive dehalogenation has also been described in different environments

though the activity has not been linked to any specific bacteria. In the case of

dichlorobenzene, it has only been demonstrated in sediment microcosms.

Bosma et al (5) showed that flow-through columns from Rhine River,

Netherlands, sediments were able to reductively dehalogenate all three DCB

isomers to MCB, with 1,2-DCB consumed first after 7 days. A microbial

consortium derived from Rhine River sediments was able to reductively

dehalogenate 1,2,4-TCB to MCB with 1,4-DCB being the major intermediate,

and 1,3- and 1,2-DCB produced in trace amounts (35).

Cometabolic Reductive Dehalogenation

A large number of halogenated compounds are reductively dehalogenated as

cometabolic reactions (18, 34, 49). As mentioned above, the final

dechlorination of vinyl chloride to ethene in strain 195 is cometabolic. The

vinyl chloride is unable to induce tceA and depends on the presence of higher-

chlorinated ethenes. The removal of the final chlorine determines the toxicity

of the end product and is a critical step in remediation. The reductive

9

Page 22: REDUCTIVE DEHALOGENATION OF

dehalogenation of complex aromatic structures like polybrominated diphenyl

ethers (PBDE) has been shown to occur in a variety of bacteria such as strain

195 and S. multivorans but only occurs when grown with PCE. It is unknown if

the PCE RD is responsible for the reductive dehalogenation or if PCE

stimulates a RD of unknown function. Sediment microcosms derived from the

Saale River, Germany show the reductive dehalogenation of MCB to benzene

to only occur during dehalogenation of higher chlorinated benzenes (45).

Benzene is easier to degrade than MCB under anaerobic conditions to non-

toxic endproducts such as carbon dioxide or methane (7).

Biochemistry and Genetics of Reductive Dehalogenation

Common Features

The reductive dehalogenation of halogenated compounds is carried about by

reductive dehalogenases (RDs). Though phylogenetic relationship between

dehalogenating bacteria is distant, features of their RDs are remarkably

similar. Common features shared by these proteins include the presence of a

twin arginine transport signal sequence (ssTAT), often used for transport of

folded proteins containing cofactors into the periplasm, iron-sulfur cluster-

binding motifs, corrinoid, and the presence of a second open reading frame

(ORF) usually named the “B” protein i.e. tceB (Figure 1.3) (55). The B ORFs

are predicted to encode hydrophobic polypeptides believed to serve as

membrane anchors. It is hypothesized that the iron-sulfur cluster and corrinoid

are involved in electron transfer. Chemical studies have shown that free

cobalamin catalyzes reductive chlorination of chloroethenes and -benzenes

while iron-sulfur clusters do not (14). The RD enzymatic cofactor vitamin B12

apparently catalyzes reductive dehalogenation of chlorinated ethenes and

10

Page 23: REDUCTIVE DEHALOGENATION OF

benzenes (14). Schmitz et al. (50) found the reductive dehalogenation of

chloropropene by PceA in S. multivorans to proceed via radical formation from

an donated electron from the B12 cofactor. Isolation of the enzymes from

strain 195 involved in reductive dechlorination of PCE to ETH had shown a

single protein of 51 kDa, PCE RD, and a single protein of 61 kDa, TCE RD

(29). The gene encoding TCE RD was subsequently cloned and sequenced

using degenerate primers and found to share similar features to a previously

isolated RD; PCE RD from S. multivorans and ortho-chlorophenol RD (CprA)

from Desulfitobacterium dehalogenans (40, 61).

Figure 1.3. Schematic of RD gene structure and key protein features.

Substrate specificity in RDs shows a restricted range. Studies on TCE RD of

strain 195 show that while it is able to dehalogenate both branched and

brominated compounds, there is a clear preference for linear chlorinated

compounds of two to three carbons (28). The substrate range of PCE RD of

strain 195 has not been determined. The PCE RD of Dehalobacter restrictus

11

Page 24: REDUCTIVE DEHALOGENATION OF

and Desulfitobacterium sp. strain Y51 is able to dehalogenate a wide range of

compounds from TCE to hexachloroethane but not DCE (31, 60). The

underlying explanation for the differences in substrate range is currently

unknown. Because Dehalococcoides spp. do not grow to very high density it

has been difficult to perform extensive biochemical analysis on their RDs, and

sufficient material for enzyme assay can only be obtained from highly mixed-

cultures containing Dehalococcoides (29, 39). Attempts to functionally

express RD genes in E. coli thus far have not been successful (41, 48, 60).

Other methods such as proteomics have been employed to monitor the

expression and presence of RDs under various physiological conditions to

study the molecular basis of Dehalococcoides spp. unique physiological

abilities. Morris et al. (37) identified peptides of RDs from strain 195, strain

CBDB1, the characterized mixed-culture KB1 and an uncharacterized TCE-

dehalogenating mixed-culture from Savannah River National Laboratory site

that could be used as markers of reductive dehalogenation activity.

Horizontal Gene Transfer

Horizontal gene transfer seems to play an important role in reductive

dehalogenation. The pceA of Desulfitobacterium hafniense strain TCE-1

shares 100% nucleotide identity with the pceA of Desulfitobacterium strain

Y51 and PCE-S and Dehalobacter restrictus (15, 36). Maillard et al (30)

discovered the pceA of D. hafniense strain TCE-1 is located on a catabolic

transposon, Tn-Dha1. Sequence comparison with the pceA region of strain

Y51 and PCE-S revealed the presence of a similar transposon but not in D.

restrictus. GC content, nucleotide usage and recombination gene analysis

12

Page 25: REDUCTIVE DEHALOGENATION OF

reveal that 15 of the 19 RD genes, including tceA, in D. ethenogenes strain

195 are located in atypical foreign regions of the genome (47). Genomic

analysis of strain CBDB1 also reveals markers of RD gene recombination

events (26).

Regulation

Most RD genes are flanked by predicted regulatory genes, indicating

regulation on the transcript level. The RD genes of Dehalococcoides spp. are

located in close proximity to global regulator genes who are members of

families such as Crp/Fnr, MarR, and two-component signaling regulators (25,

26, 39). These families regulate a diverse range of functions from antibiotic

production to osmotic sensing (24, 52, 65). Through physiological and

preliminary transcriptional studies, several RD genes of 195 have shown

indications of being under transcriptional regulation. Resting cell activity

assays with cells grown on TCE were able to reductively dehalogenate PCE,

suggesting that pceA was being co-expressed with tceA or that both are

constitutive (28). The tceA of strain 195 does not have adjacent regulatory

genes. RD gene expression in strain 195 appears to be temporally regulated.

During TCE dehalogenation at the transition point from exponential to

stationary phase, four RD genes of unknown function are up regulated

(DET0173, DET0180, DET1535, and DET1545) (23).

The best-studied RD regulon is the ortho-chlorophenol gene cluster

(cprTKZEBACD) of Desulfitobacterium dehalogens (46, 56, 57, 61). cprA

encodes for the chlorophenol reductive dehalogenase and the regulator of

chlorophenol reductive dehalogenation, CprK, belongs to the fumarate nitrate

13

Page 26: REDUCTIVE DEHALOGENATION OF

regulator family (FNR). cprK is constitutively expressed at low levels but it

only binds upstream of cprB in the presence of chlorophenol and therefore,

regulates chlorophenol reductase though metabolite activation. Activity

studies done in the presence of conventional electron acceptors fumarate,

sulfite, sulfate or nitrate, did not repress cprA expression.

Electron Transfer

The mechanism and electron flow of reductive dehalogenation is not very well

understood. In studies of PceA and TceA using whole cells and crude extracts

of strain 195 and reduced methyl viologen, which is considered unable to

cross the cytoplasmic membrane, as an electron donor, reductive

dehalogenation occurs, indicating that the reaction occurs in the periplasm

(43). This is in contrast to the PceA of D. restrictus, which was found to face

the cytoplasm and required a menaquinone shuttle (53). In Desulfomonile

tiedjei strain DCB-1, a novel cytochrome c was co-induced during reductive

dehalogenation and is thought to be part of the electron shuttling process (27).

In strains 195 and CBDB1 no upstream components in the electron transport

chain have been identified though they each contain five hydrogenases (Hyc,

Ech, Hup, Hym, and Vhu) (26, 54). They also contain an annotated formate

dehydrogenase, though neither strain is known to use formate as a growth

substrate. In strain 195, Hup and formate dehydrogenase show the highest

levels of transcripts and were both detected with high peptide coverage in

proteomic studies (38).

14

Page 27: REDUCTIVE DEHALOGENATION OF

Objectives

There has been a large influx of information into the field of reductive

dehalogenation with increasing numbers of dehalogenating microorganisms,

putative RD gene sequences, and halogenated substrates. While

demonstrating that Dehalococcoides 16S rRNA genes are present at a

contaminated site is useful (19), it is now clear that 16S rRNA genes do not

predict function, since strains FL-2, CBDB1, and BAV1 have essentially

identical 16S rRNA gene sequences, yet show markedly different

dehalogenation spectra. Identifying which genes are expressed during

dehalogenation of different substrates may eventually allow clearer diagnoses

of microbial activity at contaminated sites. My first objective is to define RD

gene function in Dehalococcoides ethenogenes strain 195 (chapter two). My

approach is to examine differential transcript levels and peptide sequences

during growth on different chlorinated compounds.

Little is known about the microorganisms that are involved in the reductive

dehalogenation of dichlorobenzenes (DCBs) and monochlorobenzene (MCB).

Reductive dehalogenation of DCBs has only been observed in sediment

microcosms and no microorganism has been identified to be involved in this

process. Though thermodynamically favorable, MCB reductive

dehalogenation to benzene has rarely been observed. My second objective is

to enrich for dichlorobenzene reductive dehalogenation activity from a

historically chlorobenzene-contaminated site using sediment microcosms

(chapter three) and to characterize the microorganism(s) responsible (chapter

four).

15

Page 28: REDUCTIVE DEHALOGENATION OF

REFERENCES

1. Adrian, L., S. K. Hansen, J. M. Fung, H. Gorisch, and S. H. Zinder.

2007. Growth of Dehalococcoides strains with chlorophenols as electron acceptors. Environ Sci Technol 41:2318-23.

2. Adrian, L., J. Rahnenfuhrer, J. Gobom, and T. Holscher. 2007. Identification of a chlorobenzene reductive dehalogenase in Dehalococcoides sp. strain CBDB1. Appl Environ Microbiol 73:7717-24.

3. Adrian, L., U. Szewzyk, J. Wecke, and H. Gorisch. 2000. Bacterial dehalorespiration with chlorinated benzenes. Nature 408:580-3.

4. Bedard, D. L., J. J. Bailey, B. L. Reiss, and G. V. Jerzak. 2006. Development and characterization of stable sediment-free anaerobic bacterial enrichment cultures that dechlorinate aroclor 1260. Appl Environ Microbiol 72:2460-70.

5. Bosma, T. N. P., van der Meer, J.R., Schraa. G., Tros, M. E., Zehnder, A.J.B. 1988. Reductive dechlorination of all trichloro- and dichlorobenzene isomers. FEMS Microbiology Ecology 53:223-229.

6. Bunge, M., L. Adrian, A. Kraus, M. Opel, W. G. Lorenz, J. R. Andreesen, H. Gorisch, and U. Lechner. 2003. Reductive dehalogenation of chlorinated dioxins by an anaerobic bacterium. Nature 421:357-60.

7. Coates, J. D., R. Chakraborty, and M. J. McInerney. 2002. Anaerobic benzene biodegradation--a new era. Res Microbiol 153:621-8.

8. Cupples, A. M., A. M. Spormann, and P. L. McCarty. 2003. Growth of a Dehalococcoides-like microorganism on vinyl chloride and cis-dichloroethene as electron acceptors as determined by competitive PCR. Appl Environ Microbiol 69:953-9.

16

Page 29: REDUCTIVE DEHALOGENATION OF

9. Cutter, L. A., J. E. Watts, K. R. Sowers, and H. D. May. 2001. Identification of a microorganism that links its growth to the reductive dechlorination of 2,3,5,6-chlorobiphenyl. Environ Microbiol 3:699-709.

10. De Wildeman, S., G. Diekert, H. Van Langenhove, and W. Verstraete. 2003. Stereoselective microbial dehalorespiration with vicinal dichlorinated alkanes. Appl Environ Microbiol 69:5643-7.

11. Dolfing, J., and B. K. Harrison. 1993. Redox and reduction potentials as parameters to predict the degradation pathway of chlorinated benzenes in anaerobic environments. FEMS Microbiology Ecology 13:23-29.

12. Ewald, E. M., A. Wagner, I. Nijenhuis, H. H. Richnow, and U. Lechner. 2007. Microbial dehalogenation of trichlorinated dibenzo-p-dioxins by a Dehalococcoides-containing mixed culture is coupled to carbon isotope fractionation. Environ Sci Technol 41:7744-51.

13. Fennell, D. E., I. Nijenhuis, S. F. Wilson, S. H. Zinder, and M. M. Haggblom. 2004. Dehalococcoides ethenogenes strain 195 reductively dechlorinates diverse chlorinated aromatic pollutants. Environ Sci Technol 38:2075-81.

14. Gantzer, C. J., Wackett, L. P. 1991. Reductive dechlorination catalyzed by bacterial transition-metal coenzyme. Environ Sci Technol 25:715-722.

15. Gerritse, J., V. Renard, T. M. Pedro-Gomes, P. A. Lawson, M. D. Collins, and J. C. Gottschal. 1996. Desulfitobacterium sp. strain PCE1, an anaerobic bacterium that can grow by reductive dechlorination of tetrachloroethene or ortho-chlorinated phenols. Arch. Microbiol. 165:132-140.

16. Grostern, A., and E. A. Edwards. 2006. Growth of Dehalobacter and Dehalococcoides spp. during degradation of chlorinated ethanes. Appl Environ Microbiol 72:428-36.

17. He, J., K. M. Ritalahti, K. L. Yang, S. S. Koenigsberg, and F. E. Loffler. 2003. Detoxification of vinyl chloride to ethene coupled to growth of an anaerobic bacterium. Nature 424:62-5.

17

Page 30: REDUCTIVE DEHALOGENATION OF

18. He, J., Robrock, K. R., Alvarez-Cohen, L. 2006. Microbial reductive debromination of polybrominated diphenyl ethers (PBDEs). Environ Sci Technol 40:4429-4434.

19. Hendrickson, E. R., J. A. Payne, R. M. Young, M. G. Starr, M. P. Perry, S. Fahnestock, D. E. Ellis, and R. C. Ebersole. 2002. Molecular analysis of Dehalococcoides 16S ribosomal DNA from chloroethene-contaminated sites throughoutNorth America and Europe. Appl. Environ. Microbiol. 68:485-495.

20. Henschler, D. 1994. Toxicity of chlorinated organic compounds: effects of the introduction of chlorine in organic molecules. Angewandte Chemie International Edition in English 33:1920-1935.

21. Holliger, C., D. Hahn, H. Harmsen, W. Ludwig, W. Schumacher, B. Tindall, F. Vazquez, N. Weiss, and A. J. Zehnder. 1998. Dehalobacter restrictus gen. nov. and sp. nov., a strictly anaerobic bacterium that reductively dechlorinates tetra- and trichloroethene in an anaerobic respiration. Arch Microbiol 169:313-21.

22. Holscher, T., R. Krajmalnik-Brown, K. M. Ritalahti, F. Von Wintzingerode, H. Gorisch, F. E. Loffler, and L. Adrian. 2004. Multiple nonidentical reductive-dehalogenase-homologous genes are common in Dehalococcoides. Appl Environ Microbiol 70:5290-7.

23. Johnson, D. R., E. L. Brodie, A. E. Hubbard, G. L. Andersen, S. H. Zinder, and L. Alvarez-Cohen. 2008. Temporal transcriptomic microarray analysis of "Dehalococcoides ethenogenes" strain 195 during the transition into stationary phase. Appl Environ Microbiol 74:2864-72.

24. Korner, H., Sofia, H.J., and Zumft, W. 2003. Phylogeny of bacterial superfamily of Crp-Fnr transcription regulators: exploiting the metabolic spectrum by controlling alternative gene programs. FEMS Microbiology Reviews 27:559-592.

25. Krajmalnik-Brown, R., T. Holscher, I. N. Thomson, F. M. Saunders, K. M. Ritalahti, and F. E. Loffler. 2004. Genetic identification of a putative vinyl chloride reductase in Dehalococcoides sp. strain BAV1. Appl Environ Microbiol 70:6347-51.

18

Page 31: REDUCTIVE DEHALOGENATION OF

26. Kube, M., A. Beck, S. H. Zinder, H. Kuhl, R. Reinhardt, and L. Adrian. 2005. Genome sequence of the chlorinated compound-respiring bacterium Dehalococcoides species strain CBDB1. Nat Biotechnol 23:1269-73.

27. Louie, T. M., S. Ni, L. Xun, and W. W. Mohn. 1997. Purification, characterization and gene sequence analysis of a novel cytochrome c co-induced with reductive dechlorination activity in Desulfomonile tiedjei DCB-1. Arch Microbiol 168:520-7.

28. Magnuson, J. K., M. F. Romine, D. R. Burris, and M. T. Kingsley. 2000. Trichloroethene reductive dehalogenase from Dehalococcoides ethenogenes: sequence of tceA and substrate range characterization. Appl Environ Microbiol 66:5141-7.

29. Magnuson, J. K., R. V. Stern, J. M. Gossett, S. H. Zinder, and D. R. Burris. 1998. Reductive dechlorination of tetrachloroethene to ethene by a two-component enzyme pathway. Appl Environ Microbiol 64:1270-5.

30. Maillard, J., C. Regeard, and C. Holliger. 2005. Isolation and characterization of Tn-Dha1, a transposon containing the tetrachloroethene reductive dehalogenase of Desulfitobacterium hafniense strain TCE1. Environ Microbiol 7:107-17.

31. Maillard, J., W. Schumacher, F. Vazquez, C. Regeard, W. R. Hagen, and C. Holliger. 2003. Characterization of the corrinoid iron-sulfur protein tetrachloroethene reductive dehalogenase of Dehalobacter restrictus. Appl Environ Microbiol 69:4628-38.

32. Major, D. W., M. L. McMaster, E. E. Cox, E. A. Edwards, S. M. Dworatzek, E. R. Hendrickson, M. G. Starr, J. A. Payne, and L. W. Buonamici. 2002. Field demonstration of successful bioaugmentation to achieve dechlorination of tetrachloroethene to ethene. Environ Sci Technol 36:5106-16.

33. May, H. D., G. S. Miller, B. V. Kjellerup, and K. R. Sowers. 2008. Dehalorespiration with polychlorinated biphenyls by an anaerobic ultramicrobacterium. Appl Environ Microbiol 74:2089-94.

19

Page 32: REDUCTIVE DEHALOGENATION OF

34. Maymo-Gatell, X., Y. Chien, J. M. Gossett, and S. H. Zinder. 1997. Isolation of a bacterium that reductively dechlorinates tetrachloroethene to ethene. Science 276:1568-71.

35. Middeldorp, P., J. De Wolf, A. Zehnder, and G. Schraa. 1997. Enrichment and properties of a 1,2,4-trichlorobenzene-dechlorinating methanogenic microbial consortium. Appl Environ Microbiol 63:1225-1229.

36. Miller, E., G. Wohlfarth, and G. Diekert. 1997. Comparative studies on tetrachloroethene reductive dechlorination mediated by Desulfitobacterium sp. strain PCE-S. Arch. Microbiol. 168:513-519.

37. Morris, R. M., J. M. Fung, B. G. Rahm, S. Zhang, D. L. Freedman, S. H. Zinder, and R. E. Richardson. 2007. Comparative proteomics of Dehalococcoides spp. reveals strain-specific peptides associated with activity. Appl Environ Microbiol 73:320-6.

38. Morris, R. M., S. Sowell, D. Barofsky, S. Zinder, and R. Richardson. 2006. Transcription and mass-spectroscopic proteomic studies of electron transport oxidoreductases in Dehalococcoides ethenogenes. Environ Microbiol 8:1499-509.

39. Muller, J. A., B. M. Rosner, G. Von Abendroth, G. Meshulam-Simon, P. L. McCarty, and A. M. Spormann. 2004. Molecular identification of the catabolic vinyl chloride reductase from Dehalococcoides sp. strain VS and its environmental distribution. Appl Environ Microbiol 70:4880-8.

40. Neumann, A., G. Wohlfarth, and G. Diekert. 1996. Purification and characterization of tetrachloroethene reductive dehalogenase from Dehalospirillum multivorans. J Biol Chem 271:16515-9.

41. Neumann, A., G. Wohlfarth, and G. Diekert. 1998. Tetrachloroethene dehalogenase from Dehalospirillum multivorans: cloning, sequencing of the encoding genes, and expression of the pceA gene in Escherichia coli. J. Bacteriol. 180:4140-4145.

42. Ni, S., J. K. Fredrickson, and L. Xun. 1995. Purification and characterization of a novel 3-chlorobenzoate-reductive dehalogenase

20

Page 33: REDUCTIVE DEHALOGENATION OF

from the cytoplasmic membrane of Desulfomonile tiedjei DCB-1. J Bacteriol 177:5135-9.

43. Nijenhuis, I., and S. H. Zinder. 2005. Characterization of hydrogenase and reductive dehalogenase activities of Dehalococcoides ethenogenes strain 195. Appl Environ Microbiol 71:1664-7.

44. Nonaka, H., G. Keresztes, Y. Shinoda, Y. Ikenaga, M. Abe, K. Naito, K. Inatomi, K. Furukawa, M. Inui, and H. Yukawa. 2006. Complete genome sequence of the dehalorespiring bacterium Desulfitobacterium hafniense Y51 and comparison with Dehalococcoides ethenogenes 195. J Bacteriol 188:2262-74.

45. Nowak, J., Kirsche, N.H., Hegemann, W., Stan, H.J. 1996. Total reductive dechlorination of chlorobenzenes to benzene by a methanogenic mixed culture enriched from Saale river sediment. Appl Environ Microbiol 45:700-709.

46. Pop, S. M., R. J. Kolarik, and S. W. Ragsdale. 2004. Regulation of anaerobic dehalorespiration by the transcriptional activator CprK. J Biol Chem 279:49910-8.

47. Regeard, C., J. Maillard, C. Dufraigne, P. Deschavanne, and C. Holliger. 2005. Indications for acquisition of reductive dehalogenase genes through horizontal gene transfer by Dehalococcoides ethenogenes strain 195. Appl Environ Microbiol 71:2955-61.

48. Regeard, C., J. Maillard, and C. Holliger. 2004. Development of degenerate and specific PCR primers for the detection and isolation of known and putative chloroethene reductive dehalogenase genes. J Microbiol Methods 56:107-18.

49. Robrock, K. R., P. Korytar, and L. Alvarez-Cohen. 2008. Pathways for the anaerobic microbial debromination of polybrominated diphenyl ethers. Environ Sci Technol 42:2845-52.

50. Schmitz, R. P., J. Wolf, A. Habel, A. Neumann, K. Ploss, A. Svatos, W. Boland, and G. Diekert. 2007. Evidence for a radical mechanism of the dechlorination of chlorinated propenes mediated by the

21

Page 34: REDUCTIVE DEHALOGENATION OF

tetrachloroethene reductive dehalogenase of Sulfurospirillum muftivorans. Environ Sci Technol 41:7370-5.

51. Scholz-Muramathsu, H., Neumann, A. , Messmer, M., Moore, E. , Diekert, G. 1995. Isolation and characterization of Dehalospirillum multivorans gen. nov., sp. nov., a tetrachloroethene-utilizing, strictly anaerobic bacterium. Arch Microbiol 163:48-56.

52. Schumacher, M. A., and R. G. Brennan. 2002. Structural mechanisms of multidrug recognition and regulation by bacterial multidrug transcription factors. Mol Microbiol 45:885-93.

53. Schumacher, W., and C. Holliger. 1996. The proton/electron ration of the menaquinone-dependent electron transport from dihydrogen to tetrachloroethene in "Dehalobacter restrictus". J Bacteriol 178:2328-33.

54. Seshadri, R., L. Adrian, D. E. Fouts, J. A. Eisen, A. M. Phillippy, B. A. Methe, N. L. Ward, W. C. Nelson, R. T. Deboy, H. M. Khouri, J. F. Kolonay, R. J. Dodson, S. C. Daugherty, L. M. Brinkac, S. A. Sullivan, R. Madupu, K. E. Nelson, K. H. Kang, M. Impraim, K. Tran, J. M. Robinson, H. A. Forberger, C. M. Fraser, S. H. Zinder, and J. F. Heidelberg. 2005. Genome sequence of the PCE-dechlorinating bacterium Dehalococcoides ethenogenes. Science 307:105-8.

55. Smidt, H., and W. M. de Vos. 2004. Anaerobic microbial dehalogenation. Annu Rev Microbiol 58:43-73.

56. Smidt, H., D. Song, J. van Der Oost, and W. M. de Vos. 1999. Random transposition by Tn916 in Desulfitobacterium dehalogenans allows for isolation and characterization of halorespiration-deficient mutants. J Bacteriol 181:6882-8.

57. Smidt, H., M. van Leest, J. van der Oost, and W. M. de Vos. 2000. Transcriptional regulation of the cpr gene cluster in ortho-chlorophenol-respiring Desulfitobacterium dehalogenans. J Bacteriol 182:5683-91.

58. Sun, B., B. M. Griffin, H. L. Ayala-del-Rio, S. A. Hashsham, and J. M. Tiedje. 2002. Microbial dehalorespiration with 1,1,1-trichloroethane. Science 298:1023-5.

22

Page 35: REDUCTIVE DEHALOGENATION OF

59. Sung, Y., K. E. Fletcher, K. M. Ritalahti, R. P. Apkarian, N. Ramos-Hernandez, R. A. Sanford, N. M. Mesbah, and F. E. Loffler. 2006. Geobacter lovleyi sp. nov. strain SZ, a novel metal-reducing and tetrachloroethene-dechlorinating bacterium. Appl Environ Microbiol 72:2775-82.

60. Suyama, A., M. Yamashita, S. Yoshino, and K. Furukawa. 2002. Molecular characterization of the PceA reductive dehalogenase of Desulfitobacterium sp. strain Y51. J Bacteriol 184:3419-25.

61. van de Pas, B. A., H. Smidt, W. R. Hagen, J. van der Oost, G. Schraa, A. J. Stams, and W. M. de Vos. 1999. Purification and molecular characterization of ortho-chlorophenol reductive dehalogenase, a key enzyme of halorespiration in Desulfitobacterium dehalogenans. J Biol Chem 274:20287-92.

62. van Doesburg, W., van Eekert, M. H., Middeldorp, P., Balk, M., Schraa, G., Starns, A. 2005. Reductive dechlorination of β-hexachlorocyclohexane (β-HCH) by a Dehalobacter species in coculture with a Sedimentibacter sp. FEMS Microbiol Ecology 54:87-95.

63. Vezzi, A., S. Campanaro, M. D'Angelo, F. Simonato, N. Vitulo, F. M. Lauro, A. Cestaro, G. Malacrida, B. Simionati, N. Cannata, C. Romualdi, D. H. Bartlett, and G. Valle. 2005. Life at depth: Photobacterium profundum genome sequence and expression analysis. Science 307:1459-61.

64. Villemur, R., M. Lanthier, R. Beaudet, and F. Lepine. 2006. The Desulfitobacterium genus. FEMS Microbiol Rev 30:706-33.

65. West, A. H. a. S., A. M. 2001. Histidine Kinases and Response Regulator Proteins in Two Component Signaling Systems. TRENDS in Biochemical Sciences 26:369-376.

66. Wu, Q., C. E. Milliken, G. P. Meier, J. E. Watts, K. R. Sowers, and H. D. May. 2002. Dechlorination of chlorobenzenes by a culture containing bacterium DF-1, a PCB dechlorinating microorganism. Environ Sci Technol 36:3290-4.

23

Page 36: REDUCTIVE DEHALOGENATION OF

67. Wu, Q., J. E. Watts, K. R. Sowers, and H. D. May. 2002. Identification of a bacterium that specifically catalyzes the reductive dechlorination of polychlorinated biphenyls with doubly flanked chlorines. Appl Environ Microbiol 68:807-12.

68. Yan, T., T. M. LaPara, and P. J. Novak. 2006. The reductive dechlorination of 2,3,4,5-tetrachlorobiphenyl in three different sediment cultures: evidence for the involvement of phylogenetically similar Dehalococcoides-like bacterial populations. FEMS Microbiol Ecol 55:248-61.

24

Page 37: REDUCTIVE DEHALOGENATION OF

CHAPTER TWO

EXPRESSION OF REDUCTIVE DEHALOGENASE GENES IN

DEHALOCOCCOIDES ETHENOGENES STRAIN 195 GROWING ON

TETRACHLOROETHENE, TRICHLOROETHENE OR 2,3-

DICHLOROPHENOL1

Abstract

Reductive dehalogenase (RD) gene transcript levels in Dehalococcoides

ethenogenes strain 195 were investigated using reverse transcriptase

quantitative PCR during growth and reductive dechlorination of

tetrachloroethene (PCE), trichloroethene (TCE), or 2,3-dichlorophenol (2,3-

DCP). Cells grown with PCE or TCE had high transcript levels (greater than

that for rpoB) for tceA, which encodes the TCE RD, pceA, which encodes the

PCE RD, and DET0162, which contains a predicted stop codon and is

considered nonfunctional. In cells grown with 2,3-DCP, tceA mRNA was less

than 1% of that for rpoB, indicating that its transcription was regulated. pceA

and DET0162 were the only RD genes with high transcript levels in cells

grown with 2,3-DCP. Proteomic analysis of PCE-grown cells detected both

PceA and TceA with high peptide coverage but not DET0162, and analysis of

2,3-DCP-grown cells detected PceA with high coverage but not TceA,

DET0162, or any other potential RD. Cells grown with PCE or 2,3-DCP were

tested for the ability to dechlorinate PCE, TCE or 2,3-DCP with H2 as the

electron donor. 2,3-DCP-grown cells were unable to dechlorinate TCE but

1 Originally published with copyright 2007 from “ Expression of reductive dehalogenase genes from Dehalococcoides ethenogenes strain 195 growing on tetrachloroethene, trichloroethene, or 2,3-dichlorophenol” by Fung, J.M., Morris, R.M., Adrian, L., and Zinder, S.H. Appl Environ Microbiol 73 (14) 4439-4445. Reprinted with permission from American Society of Microbiology.

25

Page 38: REDUCTIVE DEHALOGENATION OF

dechlorinated PCE to TCE without a lag, and PCE-grown cells dechlorinated

2,3-DCP without a lag. These results show that 2,3-DCP-grown cells do not

produce TceA andthat DET0162 is transcribed but its translation product is not

detectable in cells. They also demonstrate that PceA is bifunctional, also

serving as the 2,3-DCP RD. Chlorophenols naturally occur in soils, and are

good candidates for the original substrates for PceA.

Introduction

Dehalococcoides ethenogenes strain 195 reductively dechlorinates

tetrachloroethene (PCE) and trichloroethene (TCE) to vinyl chloride (VC) and

ethene (ETH) (21, 31). In addition to chlorinated ethenes, strain 195 has been

found to reductively dechlorinate chlorobenzenes and other chloroaromatics

(7), and more recently 2,3-dichlorophenol (2,3-DCP) and 2,3,4 trichlorophenol

in the ortho position to 3-monochlorophenol (3-MCP) or 3,4-dichlorophenol

respectively (1). The reduction of halogenated compounds by

Dehalococcoides is carried out by membrane bound respiratory reductive

dehalogenases (RDs) (12, 19, 26, 27), and although more than 90 RD-

homologous genes have been identified in this genus (11, 15, 26, 30), little is

known about their specific functions. PCE-RD (PceA) and TCE-RD (TceA)

were first characterized in mixed dechlorinating enrichment-cultures containing

strain 195, and were found to reductively dehalogenate PCE to TCE and TCE

to VC and ETH, respectively(19). The gene encoding TCE-RD was

subsequently cloned, sequenced and designated tceA (18).

The genome sequence of strain 195 (30) revealed 17 RD-homologous genes

in addition to tceA (designated DET0079) and pceA (designated DET0318, J.

26

Page 39: REDUCTIVE DEHALOGENATION OF

Magnuson, personal communication). Common features of RDs include the

presence of a putative twin arginine transport signal sequence used for

transport into the periplasm of folded proteins that can contain prosthetic

groups, iron-sulfur cluster-binding motifs, and an adjacent “B” RD gene

predicted to encode a small hydrophobic protein proposed to serve as a

membrane anchor. Two of the RD-homologous genes may not be functional.

One (DET0162) contains a verified TGA stop codon that would truncate the

predicted gene product from 488 to 59 amino acids, and a shorter

corresponding "B" gene (DET0163). DET0088 encodes a protein predicted to

be 153 amino acids long corresponding to the C-terminal of other RDs, and

lacks a corresponding "B" gene. Sixteen of the 19 RD genes in D.

ethenogenes have transcriptional regulator genes in close proximity, including

pceA, suggesting they are transcriptionally regulated. The genome sequence

of Dehalococcoides strains CBDB1 revealed 32 potential RD genes (15),

twelve of which share up to 95.4% amino acid sequence identity with RDs

from strain 195, suggesting a partial overlap of substrates. Multiple RD genes

have been identified in Dehalococcoides strains FL-2, KB1 (mixed culture),

and BAV1 (14, 14, and 10, respectively) (10, 34). Although sequenced

Dehalococcoides genomes share high sequence similarity and synteny among

"housekeeping" genes, isolates harbor different suites of RD genes and exhibit

different dehalogenation spectra. Identifying which RD genes are expressed

during dehalogenation of different substrates can provide insights into

reductive dehalogenase function and their potential activity at contaminated

sites.

27

Page 40: REDUCTIVE DEHALOGENATION OF

Here we report D. ethenogenes RD genes expressed during growth and the

reductive dehalogenation of PCE, TCE and 2,3-DCP by comparing RD gene

expression and corresponding dehalogenation activities in strain 195.

Expression and activity results indicated that tceA was under transcriptional

control, and LC/MS/MS proteomic approaches identified TceA protein

fragments only in the presence of PCE and TCE, but not 2,3-DCP. pceA was

the only intact RD gene expressed in cells grown with 2,3-DCP and is most

likely the 2,3-DCP reductive dehalogenase, which suggests that PceA has

broad substrate specificity.

Materials and Methods

Chemicals. Most chemicals were purchased from Sigma-Aldrich (St. Louis,

Missouri) at the highest purity available and gases were purchased from

Airgas East (Elmira, NY).

Growth conditions. D. ethenogenes strain 195 was cultured with PCE, TCE

or 2,3-dichlorophenol as previously described (1, 20). Briefly, culture inoculum

sizes were 2% (vol/vol) in either 27 ml culture tubes, 120 ml serum vials, or

1000 ml incubation containers containing 10, 50, or 500 ml of growth medium,

respectively. Basal salts medium was amended with 2 mM acetate, a vitamin

solution containing 0.05 mg of vitamin B12 per liter, 10% (vol/vol) filter-

sterilized anaerobic digestor sludge supernatant, and 1% (vol/vol) mixed

butyrate-PCE culture extract (20). Doses of PCE, TCE and H2 were added as

previously described (20), and filter-sterilized stock solution of 2,3-DCP was

added with a syringe at increasing doses of 30, 50, 75 µM. Culture tubes were

sealed with Teflon-coated butyl rubber stoppers and incubated at 35 °C.

28

Page 41: REDUCTIVE DEHALOGENATION OF

Reductive dechlorination of ethenes was monitored using a Perkin-Elmer 8500

gas chromatograph with a flame ionization detector (8). Chlorophenols were

analyzed using high-pressure liquid chromatograph (Beckman Coulter,

Fullerton, CA) equipped with an Alltima C8 3µM-bead diameter column (length

53 mM, ID 7mm, Alltech, Deerfield, IL) at ambient temperature. The solvent

for isocratic elution was acetonitrile:water:glacial acetic acid (50:50:0.1) at

1.5ml/min and chlorophenols were detected by their absorbance at 220nm.

To monitor growth, cells were fixed with 25% formaldehyde, filtered onto

0.2µm GTBP Isopore membrane filters (Millipore, Billerica, MA), and stained

with 5µg/ml 2-(4-Amidinophenyl)-6-indolecarbamidine dihydrochloride (DAPI).

All resting cell assays and RNA and protein extractions were performed with

cells harvested near maximum growth and dechlorination rates.

Nucleic acid extraction. Prior to extraction, 40ml of PCE-fed or 200ml of 2,3-

DCP-fed cultures were placed on ice for 30 minutes, centrifuged at 12,000x g

for 10 minutes at 4oC and resuspended in 500 µl of sterile ultrapure RNase-

free water. DNA extractions were performed according to Fennell et al. (6)

except for the elimination of a glass bead homogenization step. In short, cell

pellets for RNA extractions were processed using the RNeasy mini kit

according to the manufacturer’s instructions (Qiagen, Valencia, CA). RNA

was eluted with 50 µl RNase-free water and quantified by measuring

absorbance at 260 and 280 nm on a ND-1000 spectrophotometer (Nanodrop,

Wilmington, DE). RNA samples were treated twice with RNase-free DNase 1

(Fisher Scientific, Rockville, MD) to eliminate contaminating DNA.

29

Page 42: REDUCTIVE DEHALOGENATION OF

Reverse transcription and quantitative PCR. First-strand cDNA synthesis

reactions were performed with random hexamer primers using the iScript

cDNA Synthesis Kit (BioRad, Hercules, CA) according to manufacturer’s

instructions. Reactions were performed in 20 µl solutions containing 10 ng of

RNA incubated at 25oC for 5 min, 42oC for 30 min, and 85oC for 5 min in a

DNA Engine PTC-200 thermocycler (MJ Research, Hercules, CA).

Quantitative PCR amplifications were performed on triplicate samples using an

ABI 7000 Real Time PCR machine (Applied Biosystems, Foster City, CA).

Individual reactions contained iQ SYBR Green Super Mix (BioRad, Hercules,

CA) with 1 ng of cDNA template and 200nM of primer targeting one of the 19

RD genes or the gene encoding the RNA polymerase beta subunit (rpoB)

(Table 2.1). Primers were designed to target RD genes and not to amplify any

other sequences in the D. ethenogenes genome. Specificity of each primer

set was tested by PCR amplification and sequencing of amplified DNA. All

primers were tested to amplify at minimum of 97% reaction efficiency for each

experiment. PCR amplifications were carried out with the following

parameters: 95oC for 10 min., 30 cycles of 95oC for 15s and 60oC for 1 min.

Melting curve analysis and amplicon sequencing were used to screen for

primer dimers, and RNA samples incubated without reverse transcriptase did

not lead to a PCR product showing that no DNA was present. cDNA target

amplifications were compared to DNA standards obtained by serial dilution of

genomic DNA. RD expression levels were calculated from DNA standard

curves generated during each run and with each primer pair and related to

corresponding rpoB expression levels. In preliminary experiments, we found

that using rpoB expression as standard provided the most uniform and

reproducible results compared to other potential standards, including atpA and

30

Page 43: REDUCTIVE DEHALOGENATION OF

the 16S rRNA gene. Rahm et al. (29) found that rpoB was highly expressed in

mixed cultures containing D. ethenogenes that were actively reductively

dechlorinating chloroethenes. Values presented and their standard deviations

were taken from triplicate samples from at least two different cultures.

Mapping transcriptional start sites. The transcription start sites (TSS) of

genes tceA, pceA, infA (DET0497, initiation factor IF-1), and DET1407 were

identified using the 5’-rapid amplification of cDNA ends (5’-RACE) system

according to manufacturer’s instructions (Invitrogen, Carlsbad, CA). A series

of two or three nested primers was designed starting 20 bp downstream from

the translational start codon (Table 2.1). PCR product was sequenced and

aligned with the upstream region of corresponding gene.

Peptide and identification by NanoLC/MS/MS. Membrane-enriched

proteins were extracted from cells as described in Morris et al. (24) and in-gel-

peptide samples were sent to the Cornell Univeristy Life Sciences Core

Laboratory Center for analysis. Samples were reconstituted in 15 µL of 0.1%

formic acid with 2% acetonitrile prior to mass spectrometry (MS) analysis. The

nanoLC was carried out by an LC Packings Ultimate integrated capillary HPLC

system equipped with a Switchos valve switching unit (Dionex, Sunnyvale,

CA). The gel-extracted peptides were injected using a Famous auto sampler

onto a C18 µ-precolumn cartridge for on-line desalting and then separated on

a PepMap C-18 RP nano column, eluted in a 60-minute gradient of 5% to 45%

acetonitrile in 0.1% formic acid at 250 nL/min. The nanoLC was connected in-

line to a hybrid triple quadrupole linear ion trap mass spectrometer, 4000 Q

31

Page 44: REDUCTIVE DEHALOGENATION OF

Table 2.1 Gene targets and corresponding primer designations and

sequences Gene Target Primer Name Sequence (5' - 3')

rpoB rpo-1648f ATTATCGCTCAGGCCAATACCCGT

rpo-1800r TGCTCAAGGAAGGGTATGAGCGAA

tceA tceA-500f TAATATATGCCGCCACGAATGG

tceA-795r AATCGTATACCAAGGCCCGAGG

DET0088 DET088-153f TTATGATGTAGACCCGGAATGGCG

DET088-399r GCTATCCGGATTCTTCAGCCCATA

DET0162 DET0162-284f CCAAAGAAGAACTCCAGGCTGT

DET162-458r AGAGGCACCCAACTGTTTATAGGTCG

DET0173 DET0173-437f CTTGCTGCGGTTCTTTCTCCGTAT

DET0173-653r GAGGCGGCCAGTTATCCAGATATT

DET0180 DET0180-424f TTCCAGATTGCTATGGACAGGTGG

DET0180-604r TCTGCCAACGCTGTAGAAGAAAGG

DET0235 DET0235-430f TTTGGTCAGAGTCGGTCCATGTTC

DET0235-609r CCCAACGTTCAATACCCTCTTAGTC

DET0302 DET0302-398f GCGGCGCATCTCCATTACTTAT

DET0302-632r TTCTCATTCAGCTCGGCAAACC

DET0306 DET0306-460f TTTACCCGTTTCCCTCGCCGTTTT

DET0306-672r TCTATGCAACCCACATCTTCTGCC

DET0311 DET0311-373f GCTGTCCGGCGTAACATGGATTAT

DET0311-552r AAAGCTGCATTCTGTTCGGGTGTG

DET0318 DET0318-484f ATGGTGGATTTAGTAGCAGCGGTC

DET0318-664r ATCATCAAGCTCAAGTGCTCCCAC

DET0876 DET0876-422f TGAAGACTTGGTTAGGGCCTCAGA

DET0876-611r GTGCGTTACCTGTATCATGGGTTG

DET1171 DET1171-392-f GTCTGGCTGGTTTGTCCACTTTCT

DET1171-646r CGGAGAATCAAATACTGCCCGTCT

DET1519 DET1519-357f CTATGCCGGACCGACTATAAAGGA

DET1519-579r TCATCCAGCATAGTGGTGCCAATC

DET1522 DET1522-400f CGTTCCACTGCTTTGCATAATGCC

DET1522-660r CTATCAAGTTCGGCACAGCCTACA

DET1528 DET1528-26f GGCGGTGATACCTGTATGTGGAAT

DET1528-242r TGTGGCCATAGTTAAAGGTCTCGG

DET1535 DET1535-455f CGGGTGAAATAAACATGGGCGGTA

DET1535-656r TTTCCTGAGCACCTACATCACAGC

DET1538 DET1538-346f GAAGCTATTTCGCTGCTGGATACC

DET1538-537r CTCATAGCAGATTTGAGCAGGGCA

DET1545 DET1545-350f ATACTTACCGGTCAAGGGCGTTAG

DET1545-559r ATGGTCACGATGTTCCTGGGTAAG

DET1559 DET1559-356f CAATTAAAGTGGGTGGTTGGGCTG

DET1559-573r ATCTGTGCCCATATCATCTTGCGG

DET0162-stopb DET0162-48f TTTAGGAGCAACCAGCACGGGAGTT

DET0162-231r GCGGCATCAACGCCTGTAAATGAT a Primer name includes the position number of the first nucleotide the primer binds b Primers were designed to target the predicted translational stop codon in DET0162

32

Page 45: REDUCTIVE DEHALOGENATION OF

Trap from ABI/MDS Sciex (Framingham, MA) equipped with Micro Ion Spray

Head ion source.

The data acquisition on the MS was performed using Analyst 1.4.1 software

(Applied Biosystems) in the positive ion mode for information dependent

acquisition (IDA) analysis. In IDA analysis, after each survey scan for m/z 400

to m/z 1550 and an enhanced resolution scan, the three highest intensity ions

with multiple charge states were selected for tandem MS (MS/MS) with rolling

collision energy applied for detected ions based on different charge states and

m/z values.

MS/MS data generated from nanoLC/ESI-based IDA analysis were

interrogated using the ProID 1.4 search engine (Applied Biosystems) for

database searching against the Dehalococcoides strain 195 database. One

trypsin miscleavage, the carboxyamidomethyl modification of cysteine, and a

methionine oxidation were used for all searches. Initial protein identification

was limited to peptide hits with > 95% confidence. Subsequent identification

was limited to at least one peptide with a Pro Group confidence score >95 and

at least one additional peptide with a Pro Group confidence scores >20.

Proteins identified by a single peptide (Figure 2.3) had confidence scores >95.

Resting cell assay of reductive dehalogenase activity. Reductive

dechlorination resting cell assays were performed as described by Magnuson

et al. and Nijenhuis et al. (18, 27) with the following modifications. Cells grown

on PCE and DCP were concentrated 10- and 100-fold relative to their original

volumes, respectively and were prepared with Ti (III) citrate as a reducing

33

Page 46: REDUCTIVE DEHALOGENATION OF

agent. Assays were set up in 8ml crimp-top vials with either 1ml of cells for

PCE and TCE reductive dechlorination assays, or 2ml of cells for the 2,3-DCP

reductive dechlorination assay, along with 8.3mM Ti (III) citrate, 57% hydrogen

in the headspace and 0.5-1mM chlorinated ethene or 20-50µM 2,3-DCP.

Dechlorination was monitored as described above for chlorinated ethenes and

phenols.

Results

RD gene transcript levels. Transcript levels of each of the 19 potential RD

genes normalized to rpoB (29) in D. ethenogenes 195 was examined during

exponential-phase growth of cultures on either PCE, TCE or 2,3-DCP (Figure

2.1). In cells growing on PCE or TCE, tceA and pceA transcript levels were

several-fold higher than those for rpoB, as were those for DET0162, which is

presumably nonfunctional because of a translational stop codon. We

confirmed the presence of this stop codon in cultures we were studying by

sequencing the PCR product from DET0162 using primers listed in Table 2.1.

In PCE-grown cultures transcript levels of DET1559 approached those of rpoB

but not in TCE-grown cells, a finding we confirmed several times. Transcripts

from DET0180 in TCE-grown cells and DET1545 in PCE grown cells were

near 10% of the levels of rpoB transcripts.

In cells grown with 2,3-DCP, the relative transcript levels of pceA were 10-fold

higher than those in PCE or TCE-grown cells, as were those for DET0162. In

contrast, those for tceA transcripts fell three orders of magnitude compared to

cells grown on PCE or TCE, suggesting transcriptional control. The only other

RD gene with transcripts approaching 10% of that for rpoB was DET0180.

34

Page 47: REDUCTIVE DEHALOGENATION OF

PCETCE2,3-DCP

rpoB

tceA

pceA

DE

T008

8D

ET0

162

DE

T017

3

DE

T018

0D

ET0

235

DE

T030

2

DE

T030

6D

ET0

311

DE

T087

6

DE

T117

1

DE

T151

9D

ET1

522

DE

T152

8

DE

T153

5D

ET1

538

DE

T154

5

DE

T155

9

10-3

10-2

10-1

100

101

102

R'

Figure 2.1. Transcript levels for genes in the genome of D. ethenogenes strain 195 annotated as potential reductive dehalogenases, including tceA (DET0079) and pceA (DET0318) in cells grown on PCE, TCE or 2,3-DCP. All transcript levels were normalized to that of rpoB (R’) and error bars represent standard deviations.

Transcription start sites for tceA and pceA. We used 5'-RACE to

determine TSSs for tceA, pceA, and two "housekeeping" genes not involved in

reductive dehalogenation that are likely to be highly expressed. DET0497 is

annotated as encoding translation initiation factor 1(infA) and is highly

expressed in growing cells of Escherichia coli (3), whereas DET1407 was

found to be one of the most abundant membrane-associated proteins in a

proteomic survey of strain 195, and was hypothesized to encode part of the S-

layer cell wall (25). The TSSs of these genes were determined to be 86-156

35

Page 48: REDUCTIVE DEHALOGENATION OF

bp upstream of the predicted translational start codons (Figure 2.2A). Attempts

to determine TSSs for DET0162 and rpoB were not successful.

Potential promoter regions upstream of these TSSs were examined. The

canonical sigma70 promoter site is TTGACa-(16-19 bp)-TAtAaT-(5-9 bp)-TSS

in which capitalized bases are present more than 50% of the time, and these

two conserved hexamers are called –35 and –10 regions respectively (16). All

four upstream regions had acceptable –10 hexamers, all beginning with TA

and ending with T. Moreover, the infA and DET1407 upstream regions had

TGTG motifs one base upstream of the predicted –10 region, known to obviate

the need for a strong –35 region in Bacillus subtilis (33), and tceA had a TG in

the same position which can play a similar role in E. coli (23).

The only gene we examined with a suitable -35 region was infA, which had the

important TTG as the first three bases. DET1407 had essentially no match

near -35, but curiously had a TTGACA located only 10 bp upstream of the –10

region. The potential –35 region for tceA had a 4/6 bp match, but lacked a

beginning T, and only a 2/6 match could be found for the pceA –35 region.

Matches to other potential sigma factor binding sites were not apparent.

We also examined the leader regions between the TSS and the translation

start sites for homology with other potential leader regions. BLAST analyses

revealed that the upstream region of tceA from strain 195 to be nearly 100%

identical with upstream regions of close tceA homologues sequenced from

diverse Dehalococcoides enrichments and strain FL2 (14) (data not

presented). Unfortunately, these sequences began downstream of the

36

Page 49: REDUCTIVE DEHALOGENATION OF

predicted –10 and –35 regions so that these regions could not be compared.

Comparison of the upstream regions of tceA and the vcrA vinyl chloride RD of

Dehalococcoides strain VS (26) revealed 43% sequence identity overall,

including low identity in the predicted promoter region (Figure 2.2A) which

contains a canonical –35 hexamer. Interestingly, there is a conserved stretch

of 15 nucleotides with a single T to C transition mutation located in

approximately the same location within the two leader regions (Figure 2.2B).

The region upstream of the predicted translation start site of pceA is ca. 80%

identical over 200 bases with that of CbdbA1588 from Dehalococcoides strain

CBDB1 with only five nucleotide differences found in the promoter region (data

not shown). The nucleotide and predicted amino acid sequences of the two

genes are 86.1% and 93.7% identical respectively.

A tceA TTTTTCATTCTTTCgTGgCACGTGCGCTTTCCAAGGTGCTATctTCTACTT-122bp-ATG pceA GGGTATTCGTTTATgcGAatGCTTACCTGTCTTCACTTATcATTAAACTGT-142bp-ATG infA CCAAGTCTACCGAGATTTGAttACCTAGCTGAACTGTGGTATaTTTTGCCA- 86bp-TTG DET1407 TAAGTTAAAATGCccGtaAACTATTGACAGGCTGTGTTATTATATAAGATA-156bp-ATG VSvcrA TGCGTATTTTGTGCAATTGACATTACTATATAAAATGCTAgAATACAAATA-104bp-ATG B vcrA TSS-28bp-GC-TTATGGATATTTGGCGTT-56bp-ATG tceA TSS-37bp-ATGTTATGGATACTTGGCTTG-64bp-ATG ********* ***** *

Figure 2.2. (A) Regions upstream of transcription starts determined by 5'-RACE for tceA, pceA, infA and DET1407 from D. ethenogenes strain 195, and the region upstream of vcrA from strain VS (26). Underlined bases represent potential –10 regions, TG motifs upstream of –10, and potential –35 regions. (B) Transcription leader regions upstream of vcrA and tceA showing region of high sequence identity.

Detection of RDs by LC/MS/MS proteomics. Proteomic analyses were

performed on PCE- and 2,3-DCP-grown cells to detect RD polypeptides.

37

Page 50: REDUCTIVE DEHALOGENATION OF

Similar to previous results (24, 25), TceA and PceA RDs were detected with

high peptide coverage in membrane-enriched fractions from PCE-grown cells,

and were among the top five polypeptides in terms of peptide coverage, along

with the products of three genes annotated as co-chaperonin GroEL

(DET1428), BNR/Asp box repeat domain protein (DET1407), and formate

dehydrogenase (DET0187) (Figure 2.3). Twenty-three unique peptides

corresponding to TceA (54% coverage) and 25 unique peptides corresponding

to PceA (55% coverage) were identified in PCE-grown membrane-enriched

protein fractions, and no peptides from other RDs were detected. The same

dominant proteins were identified in membrane-enriched fractions from 2,3-

DCP-grown cells, with the notable exception of TceA. Although fewer proteins

were detected and overall peptide coverage was lower in samples from 2,3-

DCP-grown cells, which contained less protein than those from PCE-grown

cells because the cultures reached lower cell densities, nineteen unique

peptides corresponding to PceA (51% coverage) were identified, whereas no

peptides from other RDs were detected.

38

Page 51: REDUCTIVE DEHALOGENATION OF

Figure 2.3. D. ethenogenes protein coverage obtained from PCE-grown cells, white bars, and 2,3-DCP-grown cells, grey bars. Peptide fragments identified by ESI and MALDI MS/MS of the five most highly expressed proteins identified in cell membrane-enriched fractions.

RD activity. We examined cells harvested from cultures growing on either

PCE or 2,3-DCP for the ability to utilize either substrate independent of

growth. PCE-grown cells dechlorinated PCE to TCE and small amounts of

DCEs and VC (Figure 2.4A), similar to previous results (27), indicating the

presence of PceA and TceA activities. PCE-grown cells also converted 2,3-

DCP to 3-MCP with no lag and at rates ca. 3-fold lower than PCE

dechlorination (Figure 2.4B). 2,3-DCP-grown cells dechlorinated PCE to TCE,

but less chlorinated ethenes were not detected (Figure 2.4C). These cells

reductively dehalogenated 2,3-DCP ca. 5-fold slower than PCE (Figure 2.4D).

39

Page 52: REDUCTIVE DEHALOGENATION OF

Figure 2.4. Reductive dehalogenation activities of resting cells of D. ethenogenes strain 195 grown on PCE or TCE. (A) PCE dechlorination by PCE-grown cells. (B) 2,3-DCP dechlorination by PCE-grown cells. (C) PCE dechlorination by 2,3-DCP grown cells. (D) 2,3-DCP dechlorination by 2,3-DCP grown cells. Graphs show a representative vial from experiments run in triplicate.

40

Page 53: REDUCTIVE DEHALOGENATION OF

Discussion

From previous results (18, 19) it was expected that the genes encoding TceA

and PceA would be highly expressed in PCE-grown cells of D. ethenogenes,

and indeed the transcript levels of these two genes were higher than that for

rpoB and orders of magnitude higher than genes encoding the other RDs, with

the exception of DET0162, (discussed below). Moreover, peptide coverage

for TceA and PceA was high in membrane-enriched fractions of PCE-grown

cells, in agreement with previous results (25). PceA is not needed for growth

on TCE, yet we found high transcript levels in TCE-grown cells; however,

Maymó-Gatell et al. (20) found that TCE-grown cells of strain 195 showed high

PCE dehalogenation activity, indicating that PceA is present in TCE-grown

cells. Transcripts of DET1559 were detected at levels about 10-fold lower than

those for tceA and pceA in PCE-grown cells (Figure 2.1.) but peptides

corresponding to this potential RD were not detected in membrane-enriched

cell preparations in this study. However, DET1559 peptides were detected

with low coverage in a PCE-grown mixed culture containing D. ethenogenes

(24) as were peptides from DET1545 in a pure culture preparation different

from the one used in these studies. Rahm et al. (29) examined the temporal

expression of a select group of RD genes from a mixed culture containing D.

ethenogenes and found, similar to this study, that transcript levels of tceA,

pceA, DET0162 and DET1559 increased during PCE reductive

dehalogenation. In contrast, DET1545 transcripts were also detected though

they did not reach similar maximum transcript levels like that of tceA until after

all PCE was reductively dehalogenated to VC. Thus, there appears to be

some variability in the detection of RDs expressed at lower levels than tceA

and pceA.

41

Page 54: REDUCTIVE DEHALOGENATION OF

Transcript levels for pceA were higher relative to rpoB in 2,3-DCP-grown cells

than they were in PCE-grown cells. Moreover, PceA was the only RD with

detected in proteomic analyses of 2,3-DCP-grown cells. In contrast, tceA

transcript levels were over two orders of magnitude lower than rpoB, peptides

from TceA were not detected in 2,3-DCP-grown cells, and 2,3-DCP-grown

cells did not have detectable TCE RD activity, indicating that tceA is not

expressed in 2,3-DCP-grown cells.

It is not surprising that evidence for regulation of expression of RD-

homologous genes in Dehalococcoides spp. is beginning to accumulate, since

many are located adjacent to genes predicted to encode transcriptional

regulators (15, 30). Johnson et al. (13), in their studies of a TCE-grown

enrichment culture from Alameda Naval Air Station that contained

Dehalococcoides spp., reported increased levels of tceA mRNA in starved

cells given TCE, cis-DCE, trans-DCE, or 1,1-DCE, but not PCE or VC. These

findings suggest that the molecular mechanism of control over tceA is finely

tuned to recognize specific halogenated compounds. This culture did not use

PCE and therefore could not produce TCE potentially needed to induce tceA.

In strain 195, tceA was induced in cells growing on TCE or PCE (which is

metabolized to TCE) but not in cells growing on 2,3-DCP. Curiously, unlike

most other RD genes, there are no genes with strong resemblances to

transcription regulators adjacent to tceA. Of the adjacent genes, DET0080 is

annotated as having unknown function (www.tigr.org/tdb/mdb), but has

similarity to the ArsR family of regulators below the noise cutoff in a hidden

Markov model search. While this gene product is a possible candidate for

42

Page 55: REDUCTIVE DEHALOGENATION OF

regulating tceA expression, it is also possible that regulatory circuits encoded

elsewhere in the chromosome play a role in regulation.

These results also suggest that the PceA RD is responsible for dechlorination

of 2,3-DCP. It has the highest transcript levels in 2,3-DCP-grown cells and it

was detected with high peptide coverage, whereas no other potential RD was

detected. Further bolstering this proposition is the finding that PCE-grown cells

dechlorinated 2,3-DCP without a lag (Figure 4) indicating the appropriate

reductive dehalogenase was already present. In an analogous experiment,

2,3-DCP-grown cells dechlorinated PCE to TCE without a lag but did not

dechlorinate TCE. In Desulfitobacterium strain PCE1, a member of the

Firmicutes, 2-chlorophenol and PCE dechlorination are carried out by distinct

RD enzymes with little cross reactivity towards the other substrate (32).

However, the 2-chlorophenol RD from this organism does not require adjacent

chlorines, as are present in 2,3-DCP and PCE, so that it clearly has different

substrate specificity from the 2,3-DCP-dechlorinating RD in strain 195, and its

gene sequence is phylogenetically distinct from pceA. In light of the large

number of potential RD genes present in the D. ethenogenes genome (30)

with unknown function, it is surprising that PceA rather than one of the other

RDs uses 2,3-DCP, but this finding does suggest an evolutionary route for

Dehalococcoides spp. to take from utilizing chlorophenols, considered to be

naturally occurring substrates, especially in soils (4), to utilizing PCE,

considered a xenobiotic.

Dehalococcoides strain CBDB1 is adept at dehalogenating chlorinated

aromatics (2), including chlorophenols (1). Similar to strain 195, strain CBDB1

43

Page 56: REDUCTIVE DEHALOGENATION OF

reductively dehalogenates 2,3-DCP from the ortho position and reductively

dehalogenates PCE to TCE and trans-dichloroethene (L. Adrian, personal

observation). Strain CBDB1 can also reductively dehalogenate higher

chlorinated chlorophenols such as 2,3,4-trichlorophenol from the meta

position. In strain 195, the translated sequence of pceA has 93.7% amino acid

identity with CbdbA1588 from strain CBDB1 (15). Downstream of cbdbA1588

are predicted histidine kinase (CbdbA1590) and response regulator

(CbdbA1589) genes of a two component regulatory system that have 92.1%

and 91.1% amino acid identity to their homologs associated with pceA in strain

195. Morris et al. (24) recently found that in cultures of strain CBDB1 grown on

2,3-DCP, CbdbA1588 peptides were detected with 31% coverage, while in

contrast to strain 195, CbdbA080, homologous to DET1559, was also detected

with 13% peptide coverage, and CbdbA088, with no homologs in strain 195,

was detected with 10% peptide coverage. This result suggests that the

Cbdb1588 serves as a 2,3-DCP reductive dehalogenase in strain CBDB1, but

with the detection of multiple RDs the situation may be more complicated in

this organism.

Since little is known about transcription initiation and regulation in

Dehalococcoides spp., we determined the TSSs and examined upstream

regions of two expressed RDs, tceA and pceA, and two "housekeeping" genes

expected to be highly expressed, infA, and DET1407. The genome of strain

195 contains a gene predicted to encode a sigma70 homolog (DET0551) (17)

as well as two smaller genes (DET0169 and DET1348) predicted to encode

sigma W and extracellular sigma factor (ECF) homologs respectively, which

usually regulate accessory functions. The predicted amino acid sequence of

44

Page 57: REDUCTIVE DEHALOGENATION OF

DET0551 contains the conserved residues within its region 2.4 and 4.2 that

are involved in nucleotide contact with the –10 and –35 hexamers in other

sigma70 homologues (5, 9, 22), suggesting that its recognition sequences

should resemble the canonical ones. Thus it is reasonable to expect some

housekeeping and other genes to show sigma70 consensus binding sites. In all

four genes, there was a reasonably good match to the consensus –10 sigma70

binding site, but only infA had a close match to the –35 consensus sequence,

as does the vcrA gene from Dehalococcoides strain VS (Figure 2.2A). Both

infA and DET1407 have potential extended –10 regions with a TGTG motif

(33), and it is likely that this allows DET1407 to be transcribed in the absence

of an acceptable –35 region. The region upstream of the tceA TSS has a TG

motif extending its –10 region which may allow transcription despite the

moderate match of its –35 region (23), whereas the pceA has no TG element

and a poor –35 match. In both RD genes, binding by an activator may be

needed for transcription, and in the case of pceA perhaps this function is

provided by the response regulator of the two-component system predicted to

be encoded by DET0315 and DET0316 adjacent to it. Finally, the conserved

15 bp sequence in the leader regions of tceA from strain 195 and vcrA from

strain VS may bind homologous regulatory proteins.

Since DET0162 was not considered to encode a functional RD, it was

surprising that it showed high transcript levels in cells grown in PCE, TCE, or

2,3-DCP; however, peptides from its translation product were not detected in

any cells in this or a previous (25) study, suggesting that it is either not

translated or that the translation product is unstable, and it is therefore a

nonfunctional pseudogene in the process of degradation. It was recently

45

Page 58: REDUCTIVE DEHALOGENATION OF

suggested that pseudogenes are a common feature of microbial genomes

(28). While we were unable to identify a transcription start for DET0162 using

5'-RACE, there is an acceptable –10 region 108 bases upstream of the

predicted translation start, similar to other genes (Figure 2), with TTGACA, a

perfect –35 region match, 16 bases upstream of that, making its transcription

possible. Why this presumably nonfunctional gene showed such high

transcript levels is unclear.

The list of potential RD genes with unknown functions in Dehalococcoides

spp. has increased rapidly over the past several years, while the list of

halogenated substrates has increased at a slower pace. Dehalococcoides spp.

grow to low densities making traditional protein purification techniques and

genetic analyses of RDs difficult, but PCR-based measurements of transcripts

and sensitive proteomic techniques allowed us to identify RD genes expressed

in D. ethenogenes cells using different electron acceptors, and these

techniques were supplemented by measuring RD enzymatic activity in cells.

This approach should continue to be useful in identifying RDs involved in using

other substrates by this microbial group important to bioremediation.

Acknowledgements

This research was supported by NSF Metabolic Biochemistry Grant MCB -

0236044. Helpful comments by J. Helmann and help with proteomics by S.

Zhang with proteomic analyses are gratefully acknowledged.

46

Page 59: REDUCTIVE DEHALOGENATION OF

REFERENCES 1. Adrian, L., S. K. Hansen, J. M. Fung, H. Gorisch, and S. H. Zinder.

2007. Growth of Dehalococcoides strains with chlorophenols as electron acceptors. Environ Sci Technol 41:2318-23.

2. Adrian, L., U. Szewzyk, J. Wecke, and H. Gorisch. 2000. Bacterial dehalorespiration with chlorinated benzenes. Nature 408:580-3.

3. Allen, T. E., M. J. Herrgard, M. Liu, Y. Qiu, J. D. Glasner, F. R. Blattner, and B. O. Palsson. 2003. Genome-scale analysis of the uses of the Escherichia coli genome: model-driven analysis of heterogeneous data sets. J Bacteriol 185:6392-9.

4. de Jong, E., and J. A. Field. 1997. Sulfur tuft and turkey tail: biosynthesis and biodegradation of organohalogens by basidiomycetes. Annu Rev Microbiol 51:375-414.

5. Dove, S. L., Darst, S.A., and Hochschild, A. 2003. Region 4 of Sigma as a Target for Transcription Regulation. Molecular Microbiology 48:863-874.

6. Fennell, D. E., A. B. Carroll, J. M. Gossett, and S. H. Zinder. 2001. Assessment of indigenous reductive dechlorinating potential at a TCE-contaminated site using microcosms, polymerase chain reaction analysis, and site data. Environ. Sci. Technol. 35:1830-1839.

7. Fennell, D. E., I. Nijenhuis, S. F. Wilson, S. H. Zinder, and M. M. Häggblom. 2004. Dehalococcoides ethenogenes Strain 195 reductively dechlorinates diverse chlorinated aromatic pollutants. Envir Sci Technol 38:2075-2081.

8. Freedman, D. L., and J. M. Gossett. 1989. Biological reductive dechlorination of tetrachloroethylene and trichloroethylene to ethylene under methanogenic conditions. Appl. Environ. Microbiol. 55:2144-2151.

47

Page 60: REDUCTIVE DEHALOGENATION OF

9. Helmann, J. D. 1995. Compilation and analysis of Bacillus subtilis sigma A-dependent promoter sequences: evidence for extended contact between RNA polymerase and upstream promoter DNA. Nucleic Acids Res 23:2351-60.

10. Hölscher, T., H. Görisch, and L. Adrian. 2003. Reductive dehalogenation of chlorobenzene congeners in cell extracts of Dehalococcoides sp. strain CBDB1. Appl Environ Microbiol 69:2999-3001.

11. Hölscher, T., R. Krajmalnik-Brown, K. M. Ritalahti, F. Von Wintzingerode, H. Görisch, F. E. Loffler, and L. Adrian. 2004. Multiple nonidentical reductive-dehalogenase-homologous genes are common in Dehalococcoides. Appl Environ Microbiol 70:5290-7.

12. Jayachandran, G., H. Gorisch, and L. Adrian. 2003. Dehalorespiration with hexachlorobenzene and pentachlorobenzene by Dehalococcoides sp. strain CBDB1. Arch Microbiol 180:411-6.

13. Johnson, D. R., P. K. Lee, V. F. Holmes, A. C. Fortin, and L. Alvarez-Cohen. 2005. Transcriptional expression of the tceA gene in a Dehalococcoides-containing microbial enrichment. Appl Environ Microbiol 71:7145-51.

14. Krajmalnik-Brown, R., Y. Sung, K. M. Ritalahti, F. Michael Saunders, and F. E. Loffler. 2007. Environmental distribution of the trichloroethene reductive dehalogenase gene (tceA) suggests lateral gene transfer among Dehalococcoides. FEMS Microbiol Ecol 59:206-14.

15. Kube, M., A. Beck, S. H. Zinder, H. Kuhl, R. Reinhardt, and L. Adrian. 2005. Genome sequence of the chlorinated compound-respiring bacterium Dehalococcoides species strain CBDB1. Nat Biotechnol 23:1269-73.

16. Lewin, B. 1990. Genes IV, 4 ed. Oxford University Press.

17. Lonetto, M., M. Gribskov, and C. A. Gross. 1992. The sigma 70 family: sequence conservation and evolutionary relationships. J Bacteriol 174:3843-9.

48

Page 61: REDUCTIVE DEHALOGENATION OF

18. Magnuson, J. K., M. F. Romine, D. R. Burris, and M. T. Kingsley. 2000. Trichloroethene reductive dehalogenase from Dehalococcoides ethenogenes: sequence of tceA and substrate range characterization. Appl Environ Microbiol 66:5141-7.

19. Magnuson, J. K., R. V. Stern, J. M. Gossett, S. H. Zinder, and D. R. Burriss. 1998. Reductive dechlorination of tetrachloroethene to ethene by a two-component enzyme pathway. Appl. Environ. Microbiol. 64:1270-1275.

20. Maymó-Gatell, X., T. Anguish, and S. H. Zinder. 1999. Reductive dechlorination of chlorinated ethenes and 1,2-dichloroethane by "Dehalococcoides ethenogenes" strain 195. Appl. Environ. Microbiol. 65:3108-3113.

21. Maymó-Gatell, X., Y. T. Chien, J. M. Gossett, and S. H. Zinder. 1997. Isolation of a bacterium that reductively dechlorinates tetrachloroethene to ethene. Science 276:1568-1571.

22. Meima, R., H. M. Rothfuss, L. Gewin, and M. E. Lidstrom. 2001. Promoter cloning in the radioresistant bacterium Deinococcus radiodurans. J Bacteriol 183:3169-75.

23. Mitchell, J. E., D. Zheng, S. J. Busby, and S. D. Minchin. 2003. Identification and analysis of 'extended -10' promoters in Escherichia coli. Nucleic Acids Res 31:4689-95.

24. Morris, R. M., J. M. Fung, B. G. Rahm, S. Zhang, D. L. Freedman, S. H. Zinder, and R. E. Richardson. 2007. Comparative proteomics of Dehalococcoides spp. reveals strain-specific peptides associated with activity. Appl Environ Microbiol 73:320-326.

25. Morris, R. M., S. Sowell, D. Barofsky, S. Zinder, and R. Richardson. 2006. Transcription and mass-spectroscopic proteomic studies of electron transport oxidoreductases in Dehalococcoides ethenogenes. Environ Microbiol 8:1499-509.

26. Muller, J. A., B. M. Rosner, G. Von Abendroth, G. Meshulam-Simon, P. L. McCarty, and A. M. Spormann. 2004. Molecular identification of the catabolic vinyl chloride reductase from

49

Page 62: REDUCTIVE DEHALOGENATION OF

Dehalococcoides sp. strain VS and its environmental distribution. Appl Environ Microbiol 70:4880-8.

27. Nijenhuis, I., and S. H. Zinder. 2005. Characterization of hydrogenase and reductive dehalogenase activities of Dehalococcoides ethenogenes strain 195. Appl Environ Microbiol 71:1664-7.

28. Ochman, H., and L. M. Davalos. 2006. The nature and dynamics of bacterial genomes. Science 311:1730-3.

29. Rahm, B. G., R. M. Morris, and R. E. Richardson. 2006. Temporal expression of respiratory genes in an enrichment culture containing Dehalococcoides ethenogenes. Appl Environ Microbiol 72:5486-91.

30. Seshadri, R., L. Adrian, D. E. Fouts, J. A. Eisen, A. M. Phillippy, B. A. Methe, N. L. Ward, W. C. Nelson, R. T. Deboy, H. M. Khouri, J. F. Kolonay, R. J. Dodson, S. C. Daugherty, L. M. Brinkac, S. A. Sullivan, R. Madupu, K. E. Nelson, K. H. Kang, M. Impraim, K. Tran, J. M. Robinson, H. A. Forberger, C. M. Fraser, S. H. Zinder, and J. F. Heidelberg. 2005. Genome sequence of the PCE-dechlorinating bacterium Dehalococcoides ethenogenes. Science 307:105-8.

31. Smidt, H., and W. M. de Vos. 2004. Anaerobic microbial dehalogenation. Annu Rev Microbiol 58:43-73.

32. van de Pas, B. A., J. Gerritse, W. M. de Vos, G. Schraa, and A. J. Stams. 2001. Two distinct enzyme systems are responsible for tetrachloroethene and chlorophenol reductive dehalogenation in Desulfitobacterium strain PCE1. Arch Microbiol 176:165-9.

33. Voskuil, M. I., and G. H. Chambliss. 2002. The TRTGn motif stabilizes the transcription initiation open complex. J Mol Biol 322:521-32.

34. Waller, A. S., R. Krajmalnik-Brown, F. E. Loffler, and E. A. Edwards. 2005. Multiple reductive-dehalogenase-homologous genes are simultaneously transcribed during dechlorination by Dehalococcoides-containing cultures. Appl Environ Microbiol 71:8257-64.

50

Page 63: REDUCTIVE DEHALOGENATION OF

CHAPTER THREE

REDUCTIVE DEHALOGENATION OF DICHLOROBENZENES AND

MONOCHLOROBENZENE TO BENZENE IN MICROCOSMS

Abstract

Anaerobic microcosms were constructed using sediments from a historically

chlorobenzene-contaminated site and were provided with yeast extract as an

electron donor. In these methanogenic microcosms, all three isomers of

dichlorobenzene (DCB) were reductively dehalogenated to

monochlorobenzene (MCB) when added together or individually, with 1,2-DCB

dehalogenation being the most rapid and 1,4-DCB the slowest. When nearly

all of the DCBs were consumed, benzene was detected and its accumulation

was concomitant with MCB disappearance. Small amounts of toluene were

also detected along with benzene. Subsequent MCB doses were also

converted to benzene, and benzene reached levels in excess of 5,000 µmol/L

in some microcosms. An initial DCB dose stimulated, and in some cases was

necessary for, MCB dehalogenation. Subsequent doses of DCB and MCB

were dehalogenated more rapidly than previous ones, suggestive of a growth-

related process. Addition of a ca. 4% inoculum from microcosms that had

consumed DCBs or MCB stimulated DCB and MCB dehalogenation in fresh

microcosms, also indicative of growth and suggests that the chlorobenzene-

dehalogenating microorganisms in these microcosms are good candidates for

bioaugmentation at anaerobic DCB or MCB contaminated sites.

51

Page 64: REDUCTIVE DEHALOGENATION OF

Introduction

Dichlorobenzenes (1,2-, 1,3-, and 1,4-DCB) and monochlorobenzene (MCB)

have been widely used in industry as chemical stocks , solvents or industrial

surface cleansers, and over time have become common groundwater

contaminants. In aerobic environments, microorganisms degrade DCBs and

MCB by well-characterized pathways involving hydroxylation by oxygenases

ultimately leading to substrates metabolized via the citric acid cycle (16, 30).

However, DCBs and MCB are dense non-aqueous phase liquids (DNAPLs)

and can migrate to anaerobic regions in the subsurface making them

inaccessible to aerobic degradation. In anaerobic environments, halogenated

compounds such as chloroethenes and chlorobenzenes can serve as terminal

electron acceptors and be reductively dehalogenated by bacteria such as

Dehalococcoides spp., Desulfitobacterium spp., Dehalobacter spp., and

Sulfurospirillum spp. (18, 35). The complete reductive dehalogenation of

DCBs and MCB may prove a practical means of bioremediation, since

benzene, although toxic itself, can be transformed into non-toxic products such

as CO2 or CH4 by anaerobic microbial communities (8, 9, 22). However, little

is known about the microorganisms involved in the reductive dehalogenation

of DCBs and MCB.

In the case of cultured microorganisms, only reductive dehalogenation of

chlorobenzenes with three or more chlorines has been observed.

Dehalococcoides strain CBDB1 reductively dehalogenates

hexachlorobenzene (HCB), pentachlorobenzene (PeCB), all

tetrachlorobenzenes (TeCB) and 1,2,3- and 1,2,4-trichlorobenzene (TCB), with

a preference for doubly flanked chlorines, to a mix of 1,3- and 1,4-DCB (1, 19,

52

Page 65: REDUCTIVE DEHALOGENATION OF

20). Dehalococcoides strain 195 (98% 16S rRNA gene identity to strain

CBDB1), has a smaller chlorinated benzene substrate range with HCB, PeCB,

1,2,3,4-TeCB and 1,2,4,5-TeCB and no clear preference for the removal of

doubly or singly flanked chlorines (14). The more distantly related bacterium

DF-1 (89% 16s rRNA gene identity with strain 195) is able to reductively

dehalogenate HCB, PeCB, and 1,2,3,5-TeCB to mainly 1,3,5-TCB (37).

Dehalococcoides and its relatives also dechlorinate other aromatics such as

chlorophenols, polychlorinated biphenyls, and polychlorinated dibenzofurans

and most likely play an important role in the attenuation of chlorinated

aromatics in the environment (2, 5, 7, 14, 38). Other microorganisms such as

Desulfitobacterium spp. and Desulfomonile, members of the Firmicutes and

Proteobacteria respectively, reductively dehalogenate chlorinated aromatics

such as chlorophenols and chlorobenzoates, but chlorinated benzene

dehalogenation has not been demonstrated in these organisms (15, 28).

Dehalogenation of DCBs to MCB and MCB to benzene has only been

demonstrated in microcosms, and no specific organisms have been identified

as involved in these processes. In studies of HCB and PeCB reductive

dehalogenation in sediments from Niagara Falls, New York, Ramanand et al.

(33) demonstrated the reductive dehalogenation of all three DCB isomers to

MCB in microcosms, though DCBs were never given as the sole substrates.

Bosma et al. (6) showed that flow-through columns from Rhine River,

Netherlands, sediments were able to reductively dehalogenate all three DCB

isomers to MCB, with 1,2-DCB consumed first after 7 days. Masunaga et al.

(24) found that estuarine sediments consumed 1,2-DCB at higher rates than

the other two isomers. Microcosms from various contaminated US sites (32)

53

Page 66: REDUCTIVE DEHALOGENATION OF

also favored 1,2-DCB, which was dehalogenated to MCB, and 1,4-DCB was

dechlorinated the most slowly, if at all. A microbial consortium derived from

Rhine River sediments was able to reductively dehalogenate 1,2,4-TCB to

MCB with 1,4-DCB being the major intermediate, and 1,3- and 1,2-DCB

produced in trace amounts (27).

Though thermodynamically favorable (10), reductive dehalogenation of MCB

to benzene has been rarely observed. In sediment microcosms from the

Saale River, Germany, fed DCBs, small amounts of benzene were produced

along with MCB, but only during DCB reductive dehalogenation in a seemingly

cometabolic process (31). Similar results were obtained with samples from

Robins Air Force Base, Georgia (32). More recently, Nijenhuis and colleagues

(21, 29) used natural carbon isotope fractionation, and 13C-MCB-fed biotraps

to obtain evidence for MCB dechlorination to benzene in the chlorobenzene-

contaminated Bitterfeld site in Germany.

In this study, sediment from a historically chlorobenzene-contaminated site

was used to establish microcosms and reductive dehalogenation activity of

DCBs and MCB was investigated. We describe the complete reductive

dehalogenation of DCBs and MCB to benzene under methanogenic

conditions.

Materials and Methods

Chemicals. All chlorobenzenes, benzene and toluene were purchased from

Sigma-Aldrich at the highest purity available. 1,4-dichlorobenzene was

54

Page 67: REDUCTIVE DEHALOGENATION OF

solubilized in 99.5% ethanol (Aldrich) to make a 1M solution. Gases were

purchased from Airgas East.

Sediment samples. Sediment samples were obtained from a water-saturated

drainage ditch at the DuPont Chambers Works site adjacent to the Delaware

River in Salem County, New Jersey, a chemical synthesis and waste

treatment facility that has been historically contaminated with chlorinated

benzenes and anilines as well as other chemicals. The samples were dark

brown, and no chlorobenzenes or organic compounds other than methane

were detected by gas chromatographic analyses (see below) of headspace

samples from microcosms prepared from these sediments without additions.

Sediment samples were stored in plastic 1-liter bottles aerobically at 4oC in the

dark, and samples for microcosm studies were taken from layers below the

surface and were presumably anaerobic as evidenced by methanogenesis.

Sediments stored this way maintained anaerobic reductive dechlorination and

methanogenic activity for over a year. Three batches of sediment samples of

ca. 4 liters each were used, obtained in December 2005, March 2006, and

June 2007.

Microcosms. Unless stated otherwise, all solutions were prepared

anaerobically by flushing with high purity N2 gas. Microcosms were prepared

inside an anaerobic glove box in 125 ml serum vials and sealed with Teflon™-

coated butyl rubber stoppers and aluminum crimps. To each vial, 20 g of

sediment (wet wt) was added to 50 ml of anaerobic deionized water. For

inoculated microcosms, 2 ml of slurry from an actively reductively

dehalogenating microcosm (ca. 0.8 g sediment) was transferred using a

55

Page 68: REDUCTIVE DEHALOGENATION OF

syringe with an 18 ga. needle into a microcosm containing 18 g fresh sediment

in 50 ml of deionized water. Microcosms were amended with 0. 2 g/L yeast

extract to serve as an electron donor and provide nutrients, and 1.0 g/L

sodium bicarbonate as a buffer. The headspace of all microcosms was

flushed using 70% N2/30% CO2 to remove H2 from the headspace

atmosphere, and chlorinated benzenes were added by syringe either neat, or

as a 1M ethanol solution for 1,4-DCB, and this ethanol also could be

fermented and indirectly serve as an electron donor for reductive

dehalogenation and methanogenesis. Subsequent doses of chlorobenzenes

were accompanied by doses of yeast extract. Killed controls were autoclaved

at 121o C for 45 minutes. Microcosms were incubated shaking at 30o C in the

dark. Results presented are for individual microcosms, but all experiments

were done in triplicate with similar results in replicates, and all experimental

results were repeated at least once.

Analytical Methods. Dichlorobenzenes, benzene and methane were

detected by headspace analysis using Perkin-Elmer 8500 gas chromatograph

with flame ionization detector and Rtx-35 (35% diphenyl-65%dimethyl

polysiloxane, 60 meters_0.53mm with 1.5 µm film thickness) capillary column

(Restek). Both detector and injector temperatures were 210oC. The initial

oven temperature was 75oC, and increased to 152oC at 12oC/min, then to

159oC at 3oC/min, and to 210oC at 20oC/min which was held for 2 minutes.

Compounds were identified using retention times of chemical standards and

peak areas were calculated using Peak Simple software.

56

Page 69: REDUCTIVE DEHALOGENATION OF

Benzene and toluene identifications in initial studies were verified using a

Hewlett Packard 6890 series GC system equipped with a HP 5973 Mass

Selective Detector and HP 5 capillary column (Hewlett-Packard). The detector

was kept at 280o C and injector at 255o C. The oven program had an initial

temperature of 120oC held for 3 minutes followed by an increase to 220oC at

5oC/min, 220oC to 290oC at 10oC/min, and a final hold at 290oC for 2 minutes.

Mass spectra were compared to those for authentic standards.

Results

Reductive Dehalogenation of DCBs. Anaerobic microcosms were initially

fed a mixture containing approximately 400 µmol/L (nominal concentration -

µmoles added per liter of liquid volume) of each DCB isomer. There was

difficulty obtaining an accurate mass balance of chlorinated benzenes in

microcosms due to their apparent adsorption, especially DCBs, to sediment

material and syringes. Therefore, appearance of daughter products was

considered more reliable for quantification than disappearance of substrates.

Figure 3.1A shows results for a typical microcosm receiving all three DCB

isomers. MCB was detected after 2 days of incubation and reached high

concentrations at 7 days (note different scale for MCB), when benzene was

also detected. 1,2-DCB was no longer detected after 7 days, 1,3-DCB after 9

days, and 1,4-DCB after 11 days (Figure 3.1A). The lag times for MCB

appearance varied from 2 to 7 days in microcosms depending on sediment

sample, and once activity started, the trends of reductive dehalogenations

were similar, with disappearance of DCB isomers following the same order,

and appearance of MCB followed by benzene. The identity of benzene was

confirmed by gas chromatography/mass spectrometry. CH4 was the only

57

Page 70: REDUCTIVE DEHALOGENATION OF

product detected in headspaces of microcosms incubated with all additions

except DCBs, and no dechlorination products were detected in autoclaved

microcosms incubated with DCBs (data not presented).

Benzene slowly accumulated while DCBs were still present; however, once

DCBs were consumed, its production increased considerably concomitant with

MCB decrease (Figure 3.1B), indicating that MCB was reductively

dehalogenated to benzene. Subsequent doses of DCBs to microcosms were

consumed more rapidly than the initial dose, with transient MCB accumulation

and further buildup of benzene. In microcosms that had begun producing

benzene, a peak co-migrating with toluene was often detected in gas

chromatograms, always in amounts less than 5% those of benzene. The

identity of this peak as toluene was confirmed by gas chromatography/mass

spectrometery (data not presented). The production of toluene by microcosms

was unpredictable. For example, a duplicate microcosm to the one presented

in Figure 3.1A produced comparable amounts of benzene, but less than 25%

of the amount of toluene, and in some microcosms, toluene was barely

detectable (data not presented).

58

Page 71: REDUCTIVE DEHALOGENATION OF

Figure 3.1. Reductive dehalogenation of dichlorobenzene (DCB) isomers to monochlorobenzene (MCB) and benzene (ben) in (A) microcosm given 1,2-DCB, 1,3-DCB, and 1,4-DCB, and (B) the same microcosm given several doses of DCBs as indicated by the arrows showing increase production of benzene.

When added individually to microcosms, each of the three DCB isomers was

converted to MCB followed by benzene (Figure 3.2A, B, C) with 1,2-and 1,3-

DCB being the fastest and 1,4-DCB the slowest, a pattern found in multiple

repetitions of this experiment. Both 1,2- or 1,3-DCB-fed microcosms

reductively dehalogenated MCB to benzene with 1,3-DCB typically showing

shorter lags than 1,2-DCB, whereas benzene production in 1,4-DCB-fed

microcosms was inconsistent across sediment samples. When benzene

production did occur in microcosms fed 1,4-DCB, it was slower compared to

1,2- or 1,3-DCB-fed microcosms, and reached lower levels.

59

Page 72: REDUCTIVE DEHALOGENATION OF

Figure 3.2. Reductive dehalogenation of individual dichlorobenzene (DCB) isomers to monochlorobenzene (MCB) and benzene (ben) in (A) microcosm given 1,2-DCB and (B) microcosm given 1,3-DCB and (C) microcosm given 1,4-DCB.

Reductive Dehalogenation of Monochlorobenzene. Microcosms derived

from the first two sediment samples we studied (see Methods) given solely

MCB did not show any benzene production after monitoring up to 100 days

(Figure 3.3A) in several attempts using different MCB concentrations. The

third, most recently obtained, sediment sample exhibited MCB reductive

dehalogenation activity (Figure 3.3B) and consumed the first feeding of 1,100

µmol/L MCB in 45 days and the second feeding in 10 days, followed by three

more feedings, accumulating over 5,000 µmol/L benzene. Autoclaved control

microcosms did not show MCB reductive dehalogenation activity (data not

shown).

Microcosms from the first two sediment samples that had converted a dose of

DCBs mainly to benzene were able to reductively dehalogenate subsequent

additions of MCB (Figure 3.3C), in contrast to sediments receiving MCB alone

(Figure 3.3A), thus demonstrating a stimulatory effect of prior consumption of

DCBs on MCB dehalogenation.

60

Page 73: REDUCTIVE DEHALOGENATION OF

igure 3.3. Reductive dehalogenation of monochlorobenzene (MCB) in icrocosm given solely MCB (A) with no activity and (B) with benzene (ben)

C)

icrocosm inoculation experiments. We examined the effect of addition

d MCB

,4-

Fmproduction and additional doses of MCB as indicated by the arrows and (microcosm initially given DCBs and maintained on MCB as indicated by the arrows.

M

of sediments that had consumed doses of DCBs and MCB on microcosms

containing "naïve" sediments. Addition of 2 ml sediment slurry (ca.4% v/v)

from a DCB-fed microcosm that had consumed three doses of DCBs

accelerated the reductive dehalogenation of all three DCB isomers an

(Figure 3.4A). In microcosms fed all three DCBs, 400 µmol/L of 1,2-, 1,3- and

1,4-DCB were nearly undetectable in 3, 3, and 7 days, respectively. whereas

in uninoculated controls, disappearance required 7, 9 and 11 days respectively

(data not presented). After DCBs were nearly depleted, the rate of reductive

dehalogenation of MCB to benzene increased. Inoculation of microcosms

using individual DCB isomers also increased their consumption, including 1

DCB (Figure 3.4B), the complete reductive dehalogenation of which was not

otherwise reliable.

61

Page 74: REDUCTIVE DEHALOGENATION OF

To test whether addition of sediments adapted to MCB led to differences in

3.5A), and

riod,

of

er

ted).

1,4-DCB

performance, a culture that had consumed three doses of MCB following an

initial dose of DCB (Figure 3.4C) was used as an inoculum. DCB

disappearance was rapid in microcosms with these inocula (Figure

in contrast to DCB-fed ones, benzene accumulated rapidly from the beginning

with little evidence of inhibition by DCBs. In microcosms fed MCB alone

(Figure 3.5B), benzene accumulation was detected with little or no lag pe

and subsequent MCB doses were consumed and produced benzene at

increasing rates. Uninoculated control microcosms from the same batch

sediments, which was the second one, did not show benzene production aft

60 days (data not presented). These MCB-dehalogenating microcosms were

used to inoculate a subsequent series of MCB-fed microcosms, and this

activity has been transferred several times subsequently (data not presen

Figure 3.4. Reductive dehalogenation of dichlorobenzene (DCB) to monochlorobenzene (MCB) and benzene (ben) in inoculated microcosms where inoculum was given DCBs. (A) given 1,2-DCB, 1,3-DCB, and and (B) microcosms given 1,4-DCB.

62

Page 75: REDUCTIVE DEHALOGENATION OF

Figure 3.5. Reductive dehalogenation of dichlorobenzene (DCB) to monochlorobenzene (MCB) and benzene (ben) in inoculated microcosms where inoculum was actively dechlorinating solely MCB. (A) given 1,2-DCB, 1,3-DCB, and 1,4-DCB and (B) microcosms given MCB and maintained on

hese studies have demonstrated that microcosms constructed using

orks sediments can reductively dehalogenate all three DCB

most

rines

as

nated organic

MCB as indicated by the arrows.

Discussion

T

Chambers W

isomers to MCB. Similar to other results (6, 31, 32), 1,2-DCB was used

readily and 1,4-DCB was used the slowest. It is common for adjacent chlo

in chloroaromatics to be more readily dehalogenated than those ortho or para

to each other (2, 3, 6, 13). Subsequent DCB doses were consumed more

rapidly than initial ones, which is evidence that DCB dehalogenation to MCB

was a growth-related process. Further evidence for growth was the much

higher activity in microcosms inoculated with material from DCB-

dehalogenating microcosms. This behavior is consistent with DCBs serving

electron acceptors for anaerobic respiration similar to other chlori

compounds including more highly chlorinated benzenes(3).

63

Page 76: REDUCTIVE DEHALOGENATION OF

The greatest difference between these microcosm studies and previous ones

,

urrent with

re

ring

e

mall amounts of toluene accompanying benzene were detected in these

be

d

dehalogenation (34).

is the large amount of benzene produced from DCBs and MCB and the timing

of benzene production. Nowak et al. (31) and Quistorff (32) detected only 10-

50 µmol/L benzene, which was only a small percentage to the MCB produced

and in both cases benzene production occurred only during DCB

dechlorination to MCB, suggesting that it was a side reaction conc

DCB dehalogenation. In contrast, benzene in some of our samples

accumulated to over 5000 µmol/L and its accumulation did not requi

concurrent metabolism of DCBs, and indeed seemed to be inhibited du

DCB dehalogenation in some instances. There was no evidence for benzen

consumption in microcosms where it accumulated in large amounts. Electron

acceptors for benzene oxidation other than bicarbonate were unlikely to be

available since the sediments were actively methanogenic, and the highly

reducing conditions in these microcosms that received large amounts of

electron donor are unlikely to be favorable for the syntrophic reactions

involved in methanogenic benzene utilization (36). Moreover, the high

benzene concentrations may have been toxic to organisms capable of

metabolizing it further.

S

microcosms. The significance of toluene production is unclear but it should

mentioned that small amounts of toluene were also found in methanogenic

benzene-degrading microcosms (36), and may be a side-product of activate

forms of benzene such as radicals that may be formed during reductive

64

Page 77: REDUCTIVE DEHALOGENATION OF

MCB dehalogenation was considerably slower and less robust than 1,2- or

,3-DCB dehalogenation, and in some samples did not occur at all unless an

ion is

re

to

ry

can

,

culture

ehalogenate DCBs and MCB relatively rapidly and our ability to transfer this

sms

1

initial dose of DCBs was added. While the increasing rates of benzene

accumulation in MCB-fed microcosms and the ability to transfer MCB-

dehalogenating activity to naïve microcosms suggest MCB dehalogenat

growth-related, the slower rate, stimulation by a previous dose of a mo

chlorinated substrate, and MCB utilization after DCBs are depleted are

reminiscent of the co-metabolic reductive dehalogenation of vinyl chloride

ethene by Dehalococcoides ethenogenes strain 195 (26). The stimulato

effects of DCBs on MCB dehalogenation are most simply explained by

positing that some or all of the organisms involved in DCB dehalogenation

also use MCB, and after their numbers increase while growing on DCBs

switch to the more slowly-utilized MCB upon DCB depletion. However, there

may be more complex explanations, such as stimulation of growth of MCB

dehalogenators by nutrients produced by DCB dehalogenators. As an

example of nutritional crossfeeding, it was recently shown that growth of the

DF-1 organism on PCBs required unknown factors from a Desulfovibrio

(25). Moreover, the inability of some samples to use MCB unless a prior dose

of DCBs was given is difficult to explain by a simple two-stage growth model.

The ability of the microorganisms in these microcosms to reductively

d

activity to naïve microcosms suggests that organisms in these microco

have potential as inoculants for bioaugmentation at anaerobic sites where

these compounds are persistent, a situation analogous to the addition of

65

Page 78: REDUCTIVE DEHALOGENATION OF

Dehalococcoides-containing cultures to chloroethene-contaminated sites

apparently stalled at dichloroethene (12, 17, 23). However, benzene itself

toxic with a lower EPA-mandated drinking water limit than DCBs or MCB

(

is

http://www.epa.gov/safewater/contaminants/index.html). Thus it is imperative

that treatment schemes also include benzene biodegradation. Moreover, o

findings that significant amounts of benzene can be produced from MCB

should be taken into account in attempts at modeling the environmental fates

of DCBs and MCB.

More needs to be lea

ur

rned about the microorganisms dehalogenating DCBs

nd MCB in these microcosms, especially how well they perform at in-situ

se

n

e thank the DuPont Corporation for providing sediment and financial

ful to Clifford Fung for technical assistance in

a

DCB and MCB concentrations that are typically well below those used in the

studies. We are initiating microbiological and molecular ecological studies o

these microcosms to begin to answer these questions.

Acknowledgements

W

support. We are thank

microcosm work. We also would like to thank Christopher DeRito and

Anthony Hay for their assistance with the GC/MS.

66

Page 79: REDUCTIVE DEHALOGENATION OF

REFERENCES

1. Adrian, L., and H. Gorisch. 2002. Microbial transformation of

chlorinated benzenes under anaerobic conditions. Res Microbiol 153:131-7.

2. Adrian, L., S. K. Hansen, J. M. Fung, H. Gorisch, and S. H. Zinder. 2007. Growth of Dehalococcoides strains with chlorophenols as electron acceptors. Environ Sci Technol 41:2318-23.

3. Adrian, L., U. Szewzyk, J. Wecke, and H. Gorisch. 2000. Bacterial dehalorespiration with chlorinated benzenes. Nature 408:580-3.

4. Ball, H., and Reinhard, M. 1996. Monoaromatic hydrocarbon transformation under anaerobic conditions at Seal Beach, California:

laboratory studies. Environ. Tosicol. Chem. 15:114-122.

5. Bedard, D. L., K. M. Ritalahti, and F. E. Loffler. 2007. The Dehalococcoides population in sediment-free mixed cultures metabolically dechlorinates the commercial polychlorinated biphenyl mixture aroclor 1260. Appl Environ Microbiol 73:2513-21.

6. Bosma, T. N. P., van der Meer, J.R., Schraa. G., Tros, M. E., Zehnder, A.J.B. 1988. Reductive dechlorination of all trichloro- and dichlorobenzene isomers. FEMS Microbiology Ecology 53:223-229.

7. Bunge, M., L. Adrian, A. Kraus, M. Opel, W. G. Lorenz, J. R. Andreesen, H. Gorisch, and U. Lechner. 2003. Reductive dehalogenation of chlorinated dioxins by an anaerobic bacterium. Nature 421:357-60.

8. Coates, J. D., R. Chakraborty, J. G. Lack, S. M. O'Connor, K. A. Cole, K. S. Bender, and L. A. Achenbach. 2001. Anaerobic benzene oxidation coupled to nitrate reduction in pure culture by two strains of Dechloromonas. Nature 411:1039-43.

67

Page 80: REDUCTIVE DEHALOGENATION OF

9. Coates, J. D., R. Chakraborty, and M. J. McInerney. 2002. Anaerobic benzene biodegradation--a new era. Res Microbiol 153:621-8.

10. Dolfing, J., and B. K. Harrison. 1993. Redox and reduction potentials as parameters to predict the degradation pathway of chlorinated benzenes in anaerobic environments. FEMS Microbiology Ecology 13:23-29.

11. Edwards, E. A., L. E. Wills, M. Reinhard, and D. Grbic-Galic. 1992. Anaerobic degradation of toluene and xylene by aquifer microorganisms under sulfate-reducing conditions. Appl Environ Microbiol 58:794-800.

12. Ellis, D. E., E. J. Lutz, J. M. Odom, R. J. Buchanan, C. L. Bartlett, M. D. Lee, M. R. Harkness, and K. A. DeWeerd. 2000. Bioaugmentation for accelerated in situ anaerobic bioremediation. Environ. Sci. Technol. 34:2254-2260.

13. Fathepure, B. Z., J. M. Tiedje, and S. A. Boyd. 1988. Reductive dechlorination of hexachlorobenzene to tri- and dichlorobenzenes in anaerobic sewage sludge. Appl Environ Microbiol 54:327-30.

14. Fennell, D. E., I. Nijenhuis, S. F. Wilson, S. H. Zinder, and M. M. Haggblom. 2004. Dehalococcoides ethenogenes strain 195 reductively dechlorinates diverse chlorinated aromatic pollutants. Environ Sci Technol 38:2075-81.

15. Gerritse, J., V. Renard, T. M. Pedro Gomes, P. A. Lawson, M. D. Collins, and J. C. Gottschal. 1996. Desulfitobacterium sp. strain PCE1, an anaerobic bacterium that can grow by reductive dechlorination of tetrachloroethene or ortho-chlorinated phenols. Arch Microbiol 165:132-40.

16. Haigler, B. E., S. F. Nishino, and J. C. Spain. 1988. Degradation of 1,2-dichlorobenzene by a Pseudomonas sp. Appl Environ Microbiol 54:294-301.

17. Hendrickson, E. R., J. A. Payne, R. M. Young, M. G. Starr, M. P. Perry, S. Fahnestock, D. E. Ellis, and R. C. Ebersole. 2002. Molecular analysis of Dehalococcoides 16S ribosomal DNA from

68

Page 81: REDUCTIVE DEHALOGENATION OF

chloroethene-contaminated sites throughout North America and Europe. Appl Environ Microbiol 68:485-95.

18. Holliger, C., S. Gaspard, G. Glod, C. Heijman, W. Schumacher, R. P. Schwarzenbach, and F. Vazquez. 1997. Contaminated environments in the subsurface and bioremediation: organic contaminants. FEMS Microbiol Rev 20:517-23.

19. Holscher, T., H. Gorisch, and L. Adrian. 2003. Reductive dehalogenation of chlorobenzene congeners in cell extracts of Dehalococcoides sp. strain CBDB1. Appl Environ Microbiol 69:2999-3001.

20. Jayachandran, G., H. Gorisch, and L. Adrian. 2003. Dehalorespiration with hexachlorobenzene and pentachlorobenzene by Dehalococcoides sp. strain CBDB1. Arch Microbiol 180:411-6.

21. Kaschl, A., C. Vogt, S. Uhlig, I. Nijenhuis, H. Weiss, M. Kastner, and H. H. Richnow. 2005. Isotopic fractionation indicates anaerobic monochlorobenzene biodegradation. Environ Toxicol Chem 24:1315-24.

22. Lovley, D. R. 2000. Anaerobic benzene degradation. Biodegradation 11:107-16.

23. Major, D. W., M. L. McMaster, E. E. Cox, E. A. Edwards, S. M. Dworatzek, E. R. Hendrickson, M. G. Starr, J. A. Payne, and L. W. Buonamici. 2002. Field demonstration of successful bioaugmentation to achieve dechlorination of tetrachloroethene to ethene. Environ Sci Technol 36:5106-16.

24. Mansunaga, S., Susarla, Sridhar, and Yonezawa, Yoshitaka. 1996. Dechlorination of chlorobenzenes in anaerobic estuarine sediment. Wat. Sci. Tech. 33:173-180.

25. May, H. D., G. S. Miller, B. V. Kjellerup, and K. R. Sowers. 2008. Dehalorespiration with polychlorinated biphenyls by an anaerobic ultramicrobacterium. Appl Environ Microbiol 74:2089-94.

69

Page 82: REDUCTIVE DEHALOGENATION OF

26. Maymo-Gatell, X., I. Nijenhuis, and S. H. Zinder. 2001. Reductive dechlorination of cis-1,2-dichloroethene and vinyl chloride by "Dehalococcoides ethenogenes". Environ Sci Technol 35:516-21.

27. Middeldorp, P., J. De Wolf, A. Zehnder, and G. Schraa. 1997. Enrichment and properties of a 1,2,4-trichlorobenzene-dechlorinating methanogenic microbial consortium. Appl Environ Microbiol 63:1225-1229.

28. Mohn, W. W., and K. J. Kennedy. 1992. Reductive dehalogenation of chlorophenols by Desulfomonile tiedjei DCB-1. Appl Environ Microbiol 58:1367-70.

29. Nijenhuis, I., N. Stelzer, M. Kastner, and H. H. Richnow. 2007. Sensitive detection of anaerobic monochlorobenzene degradation using stable isotope tracers. Environ Sci Technol 41:3836-42.

30. Nishino, S. F., J. C. Spain, L. A. Belcher, and C. D. Litchfield. 1992. Chlorobenzene degradation by bacteria isolated from contaminated groundwater. Appl Environ Microbiol 58:1719-26.

31. Nowak, J., Kirsche, N.H., Hegemann, W., Stan, H.J. 1996. Total reductive dechlorination of chlorobenzenes to benzene by a methanogenic mixed culture enriched from Saale river sediment. Appl Environ Microbiol 45:700-709.

32. Quistorff, A. S. 1999. Microbially mediated reductive dechlorination of dichlorobenzene. Master of Science. Cornell University, Ithaca.

33. Ramanand, K., M. T. Balba, and J. Duffy. 1993. Reductive dehalogenation of chlorinated benzenes and toluenes under methanogenic conditions. Appl Environ Microbiol 59:3266-72.

34. Schmitz, R. P., J. Wolf, A. Habel, A. Neumann, K. Ploss, A. Svatos, W. Boland, and G. Diekert. 2007. Evidence for a radical mechanism of the dechlorination of chlorinated propenes mediated by the tetrachloroethene reductive dehalogenase of Sulfurospirillum muftivorans. Environ Sci Technol 41:7370-5.

70

Page 83: REDUCTIVE DEHALOGENATION OF

35. Smidt, H., and W. M. de Vos. 2004. Anaerobic microbial dehalogenation. Annu Rev Microbiol 58:43-73.

36. Ulrich, A. C., H. R. Beller, and E. A. Edwards. 2005. Metabolites detected during biodegradation of 13C6-benzene in nitrate-reducing and methanogenic enrichment cultures. Environ Sci Technol 39:6681-91.

37. Wu, Q., C. E. Milliken, G. P. Meier, J. E. Watts, K. R. Sowers, and H. D. May. 2002. Dechlorination of chlorobenzenes by a culture containing bacterium DF-1, a PCB dechlorinating microorganism. Environ Sci Technol 36:3290-4.

38. Wu, Q., J. E. Watts, K. R. Sowers, and H. D. May. 2002. Identification of a bacterium that specifically catalyzes the reductive dechlorination of polychlorinated biphenyls with doubly flanked chlorines. Appl Environ Microbiol 68:807-12.

71

Page 84: REDUCTIVE DEHALOGENATION OF

CHAPTER FOUR

PRELIMINARY CHARACTERIZATION OF MICROBIAL ENRICHMENT

CULTURES THAT REDUCTIVELY DECHLORINATE DICHLOROBENZENES

TO MONOCHLOROBENZENE

Abstract

Microbial reductive dechlorination of highly-chlorinated benzenes is well

characterized, whereas little is known about reductive dechlorination of the

important groundwater pollutants dichlorobenzenes (DCBs) and

monochlorobenzene (MCB). I previously described sediment microcosms

that reductively dechlorinated all three DCB isomers to MCB and benzene.

From these microcosms, transfers (ca 4% v/v) were successfully made and

maintained on a mineral salts medium supplemented with yeast extract,

hydrogen, 2-bromoethanesulfonate, sodium sulfide, and vitamins. Cultures

were given all three DCB isomers together or each individually. Transfers fed

all three DCBs completely reductively dechlorinated 1,2-DCB to MCB with

slight dechlorination of 1,3-DCB and no dechlorination of 1,4-DCB. In

cultures fed 1,3-DCB alone, complete reductive dehalogenation to MCB was

observed after 8 days and transfers with 1,4-DCB showed a longer lag of 19

days. In seventh transfer cultures reductively dehalogenating 1,2-DCB, low

amounts of benzene were detected though never more than 0.44% of MCB

produced. In transfers fed MCB, reductive dechlorination to benzene was not

observed. Transfers into medium containing 0.2 mg/ml vancomycin showed

no reductive dechlorination, indicating that Dehalococcoides spp., which lack a

peptidoglycan cell wall and are vancomycin-resistant, were not the causative

agent. These cultures have now been transferred 11 times and maintained

72

Page 85: REDUCTIVE DEHALOGENATION OF

the ability to dechlorinate 1,2-DCB to MCB. A 16S rRNA gene clone library of

the fourth transfer culture contained sequences matching >99% identical to

that of Dehalobacter restrictus, a known dehalogenator. Further analysis

using quantitative real time PCR and specific primers showed an increase in

Dehalobacter population during reductive dehalogenation. Characterization of

these cultures is underway with the goal of identifying and isolating the

organisms responsible for reductive dechlorination.

Introduction

Dichlorobenzenes (DCBs) and monochlorobenzene (MCB) have been used as

industrial solvents and made their way into the environment to become

prevalent contaminants. In aerobic environments, biomineralization of DCBs

and MCB can be carried out by organisms such as Pseudomonas (29). In

anaerobic environments, halogenated compounds such as chloroethenes and

chlorobenzenes can serve as terminal electron acceptors and be reductively

dehalogenated by bacteria such as Dehalococcoides spp., Desulfitobacterium

spp., Dehalobacter spp., and Sulfurospirillum spp. (17, 34). One potential

approach to DCB bioremediation is to stimulate organisms to carry out their

complete reductive dehalogenation to benzene. While benzene itself is toxic,

anaerobic microbial communities can convert it into non-toxic products like

methane or carbon dioxide (8, 20). However, little is known about

microorganisms that can reductively dehalogenate DCBs.

Dehalococcoides strain CBDB1, strain195 and the related strain DF-1 (89%

16s rRNA gene identity to strain 195) are the only isolated microorganisms

known to reductively dehalogenate chlorinated benzenes (1, 3, 12, 13, 39).

73

Page 86: REDUCTIVE DEHALOGENATION OF

These three strains can dehalogenate higher chlorinated benzenes but are

unable to reductively dehalogenate DCBs. Strain CBDB1 has a preference for

the removal of double flanked chlorines and reductively dehalogenates hexa-

to trichlorobenzenes (HCB, TCB) to a mix of 1,3- and 1,4-DCB. Strain 195

(98% 16S rRNA gene identity to strain CBDB1) has no clear chlorine

preference and can reductively dehalogenate HCB, pentachlorobenzene

(PeCB) and tetrachlorobenzenes (TeCB). Strain DF-1 can reductively

dehalogenate HCB, PeCB, and 1,2,3,5-TeCB to mainly 1,3,5-TCB. Other

microorganisms such as Desulfomonile spp., a member of delta-

Proteobacteria, are capable of dehalogenating other chloroaromatics such as

3-chlorobenzoate and chlorophenols (27, 33).

Mixed culture studies have focused on tri- or higher chlorinated benzenes.

Holliger et al. (19) established a microbial mixed culture enriched with 1,2,3-

TCB from sediment columns derived from the Rhine River, Saale, Germany.

The mixed culture was able to reductively dehalogenate HCB, PeCB, TeCB as

well was 1,2,3- and 1,2,4-TCB to 1,3- or 1,4-DCB depending on initial

substrate. A mixed culture described by Middeldorp et al. (26) was able to

reductively dehalogenate 1,2,4-TCB to MCB via 1,4-DCB. Neither mixed

culture was able to reductively dehalogenate 1,2- or 1,3-DCB.

Reductive dehalogenation of DCBs to MCB and MCB to benzene has been

demonstrated only in sediment microcosms. Both Ramanand et al. (32) and

Bosma et al. (5) showed that sediments were able to reductively dehalogenate

all three DCB isomers to MCB, though DCBs were never given as the sole

substrates. It has been shown in estuarine sediment microcosms that 1,2-DCB

74

Page 87: REDUCTIVE DEHALOGENATION OF

was consumed at higher rates than the other two isomers (22). Microcosms

from various contaminated US sites also favored 1,2-DCB, which was

dehalogenated to MCB, and 1,4-DCB was dechlorinated the most slowly, if at

all (31). Though thermodynamically favorable, reductive dehalogenation of

MCB to benzene has been rarely observed. In DCB-fed fluidized bed reactors

inoculated with sediments from the Saale River, Germany, small amounts of

benzene were produced along with MCB, but only during DCB reductive

dehalogenation in a seemingly cometabolic process (30). Similar results were

obtained with samples from Robins Air Force Base, Georgia (31).

Phylogenetic analysis of microbial communities in the Saale River reactors

found a family of clones to be 98.8-99.4% identical to that of 16s rRNA gene of

Dehalobacter restrictus (37).

The goal of this study was to culture microorganisms involved in the reductive

dehalogenation of DCBs. Enrichment cultures in growth medium containing

DCBs were made from sediment microcosms able to reductively dehalogenate

DCBs to MCB and MCB to benzene (14) and transferred multiple times. Here

we report the reductive dehalogenation of 1,2-DCB by a mixed microbial

culture.

Materials and Methods

Chemicals. All chlorobenzenes and chemicals were purchased from Sigma-

Aldrich or Fisher Scientific at the highest purity available. 1,4-dichlorobenzene

75

Page 88: REDUCTIVE DEHALOGENATION OF

was solubilized in either 99.5% ethanol or acetone to make a 1M or 5M

solution respectively. Gases were purchased from Airgas East.

Enrichments of 1,2-DCB Dehalogenators. Unless stated otherwise, all

solutions were prepared anaerobically by flushing with high-purity N2 gas.

Growth medium was dispensed into 125 ml serum vials (50 ml/vial) inside an

anaerobic glove-box (95%N2/5%H2) and sealed with Teflon™-coated butyl

rubber stoppers and aluminum crimps. The headspaces of all the vials were

flushed using 70% N2/30% CO2 to remove H2 from the headspace

atmosphere, and chlorinated benzenes were added by syringe either neat, or

as a 1M or 5M ethanol or acetone solution, respectively, for 1,4-DCB. The

ethanolthrough fermentation could indirectly serve as an electron donor for

reductive dehalogenation and methanogenesis.

A previously described, sediment slurry microcosms that reductively

dechlorinated all three DCB isomers to MCB and benzene was used as the

inoculum (14). From these microcosms, transfers (ca 4% v/v) were

successfully made into a mineral salts medium developed for

Dehalococcoides strain CBDB1 (1) supplemented with 0.2 g/L yeast extract, 1

g/L sodium bicarbonate, 0.8 mM Ti(III) citrate, and a vitamin solution

containing 0.05 mg of vitamin B12 per liter (25). Resazurin (0.001 g/L) was

added to the medium as an indicator of oxygen contamination. During the

course of this study, the medium supplements were changed and may have

included one or more of the following (final concentrations): 5 mM 2-

bromoethanesulfonate (BES), 0.8 mM Ti(III) NTA (replacing citrate), 0.2 g/L

76

Page 89: REDUCTIVE DEHALOGENATION OF

sodium sulfide, 2 mM sodium butyrate, 2 mM sodium acetate, 3% hydrogen,

and 0.2 g/L vancomycin. All cultures were incubated at 30oC in the dark.

Dichlorobenzenes, benzene and methane were detected by headspace

analysis using Perkin-Elmer 8500 gas chromatograph with flame ionization

detector as previously described (14). Results presented are for individual

vials, but all experiments were done in duplicate with similar results in

replicates, and all experimental results were repeated at least once.

DNA Isolation and Quantitative PCR. Genomic DNA was extracted using

FastDNA for Soil kit (MP Biomedical) and resuspended in 55 µl of sterile

deionized water according to the manufacturer’s instructions. Specific primers

were used to target Eubacteria (Eub331F: 5’-TCCTACGGGAGGCAGCAGT-

3’, Eub797R: 5’-GGACTACCAGGGTATCTAATCCTGTT-3’), Dehalococcoides

(Dhc385 5’-GGGTTGTAAACCTCTTTTCAC-3’, Dhc692R: 5’-

TCAGTGACAACCTAGAAAAC-3’), and Dehalobacter (Dhb477F: 5’-

GATTGACGGTACTGCTTTACGG-3’, Dhb647R: 5’-

TACAGTTTCCAATGCTTTACG-3’), populations (15, 16, 28). Real time PCR

was performed as previously described using iQ SYBR Green Supermix and

MyiQ Single Color Real Time PCR Detection System (Bio-Rad) (13). Copy

number was quantified using plasmids containing PCR fragments of each

individual 16S rRNA gene targets derived from pCR2.1 from TA Cloning Kit

according to manufacturer’s instructions (Invitrogen).

Clone library construction. 16S rRNA gene clone libraries were constructed

from DNA extracted from fourth-generation transfer cultures by PCR

77

Page 90: REDUCTIVE DEHALOGENATION OF

amplification using the universal bacterial primers 27F (5′-

AGAGTTTGATCCTGGCTCAG-3′) and 1492R ( 5'-

ACGGYTACCTTGTTACGACTT-3') followed by library construction using

pCR2.1 from TA Cloning Kit according to manufacturer’s instructions

(Invitrogen). Clones were analyzed using restriction enzymes HhaI and HaeIII

(New England Biolabs). Unique clones were sequenced at the Life Sciences

Core Laboratory Center at Cornell University.

Results Development of a 1,2-DCB-Dehalogenating Enrichment Culture. The first

transfers from microcosms were made directly into medium with all three

dichlorobenzene isomers added (1,2-, 1,3- and 1,4-DCB in ethanol), yeast

extract as the electron donor, and Ti(III) citrate as a reducing agent. 1,2-DCB

was the first DCB isomer to be reductively dehalogenated to

monochlorobenzene (MCB), followed by small amounts of 1,3-DCB. No

reductive dehalogenation of 1,4-DCB was observed (Figure 4.1A, data shown

as nominal concentrations). Methanogenesis also occurred in these transfers

and could potentially be a competing reaction with reductive dehalogenation

for H2.

After the first transfer, the methanogen inhibitor, 2-bromoethanesulfonate, was

always added to the medium. To favor reductive dehalogenation, subsequent

growth conditions were designed to be electron-limiting by using butyrate as

the primary electron donor, which is slowly fermentable and poises the H2

concentrations low enough to allow reductive dechlorination to be competitive

with methanogenesis (11). 1.4-DCB was solubilized in acetone, which was

78

Page 91: REDUCTIVE DEHALOGENATION OF

nonfermentable, instead of ethanol and yeast extract concentrations were

lowered(Table 4.1). These changes allowed for sustained reductive

dehalogenation activity in transfers, although they did not completely eliminate

methanogenesis.

After the third transfer, reductive dehalogenation of 1,4-DCB was still not

observed and it was no longer added to further transfers. Since reductive

dehalogenation of 1,2-DCB was the most rapid, after the fifth transfer the

dehalogenating enrichment culture was maintained solely on 1,2-DCB (Figure

4.1B).

igure 4.1. Reductive dehalogenation of dichlorobenzene isomers (DCB) to onochlorobenzene (MCB) in (A) first transfer given 1,2-DCB, 1,3-DCB, and

Fm1,4-DCB with yeast extract as the electron donor and Ti(III)citrate as the reducing agent and (B) the sixth transfer given 1,2-DCB with butyrate as the electron donor and Ti(III)NTA as the reducing agent.

79

Page 92: REDUCTIVE DEHALOGENATION OF

Table 4.1. Growth conditions used to establish 1,2-DCB mixed culture.

Transfer No. Dichlorobenzenes Reducing Agent Electron Donor 1,4-DCB (solvent)

yeast extract (g/L)

1-2 1,2-, 1,3-, and

1,4- Ti(III) citrate yeast extract ethanol 0.2

3 1,2-, 1,3-, and

1,4- Ti(III) NTA butyrate acetone 0.02 4-5 1,2- and 1,3- Ti(III) NTA butyrate --- --- 6 1,2- Ti(III) NTA butyrate --- ---

7 1,2- Na2S butyrate --- 0.02

8-11 1,2- Na2S H2 --- ---

t the seventh transfer sodium sulfide replaced Ti(III) NTA as the reducing

es

ory

s.

enzene Production. Reductive dehalogenation of MCB to benzene was not

tial

ium

ere

A

agent. Sodium sulfide is the reducing agent used in the cultivation of

Dehalococcoides ethenogenes strain 195 but is toxic to Dehalococcoid

strain CBDB1. In addition, studies have shown sodium sulfide to be inhibit

to some methanogens (4, 6, 7). Sodium sulfide did not inhibit 1,2-DCB-

reductive dehalogenation to MCB and severely reduced methanogenesi

B

observed during the first six transfers though the originating sediment

microcosm showed high benzene production from MCB. During the ini

establishment of 1,2-DCB enrichments, transfers were also made into med

containing solely MCB. No reductive dehalogenation of MCB to benzene was

observed after monitoring for 30 days. At the seventh transfer, in which

sodium sulfide was used as reducing agent, trace amounts of benzene w

detected, but never more than 0.44% of the MCB produced (Figure 4.2).

80

Page 93: REDUCTIVE DEHALOGENATION OF

the seventh transfer during1,2-DCB

ductive dehalogenation. Arrows indicate doses of 1,2-DCB.

is the lack of

peptidoglycan cell wall making it vancomycin-resistant. Vancomycin was

Figure 4. 2. Benzene production in re

Vancomycin Sensitivity. One of hallmarks of Dehalococcoides

a

added to the second transfer and reductive dehalogenation activity was not

observed indicating Dehalococcoides was not the causative agent (Figure

4.3).

81

Page 94: REDUCTIVE DEHALOGENATION OF

Figure 4.3. Lost of reductive dehalogenation activity in second transfer in the presence of vancomycin.

Dechlorination Range of the 1,2-DCB Enrichment. To investigate the

reductive dehalogenation potential of the 1,2-DCB enrichment, the sixth

transfer cultures were given either 1,2-, 1,3- or 1,4-DCB as the sole electron

acceptor. In cultures fed 1,3-DCB alone, complete reductive dehalogenation

to MCB was observed after 8 days and transfers with 1,4-DCB showed a

longer lag of 19 days (Figure 4.4A, B). However, upon change to sodium

sulfide as a reducing agent the culture lost 1,3- and 1,4-DCB reductive

dehalogenation activity (data not shown).

82

Page 95: REDUCTIVE DEHALOGENATION OF

Figure 4.4. Reductive dehalogenation of (A) 1,3-DCB and (B) 1,4-DCB by 1,2-DCB-mixed culture at the sixth transfer.

16S rRNA Gene Clone Library and Bacterial Population Quantification of

a DCB Dehalogenating Culture. Genomic DNA was extracted from fourth

transfer cultures grown either with or without DCBs and was used to construct

clone libraries using universal bacterial primers targeting the 16S rRNA gene.

Extractions were done after 36 days and 823 µmoles/L MCB was produced.

From the culture given DCBs, 44% of the characterized clones (31/70) were of

a restriction type, sequences from which were >99% identical to that from

Dehalobacter restrictus, a known dehalogenator of chloroethenes (Table 4.2)

(18). Other sequences found in the DCB-fed culture belonged to other

members of the Firmicutes. Dehalobacter sequences were not found in

culture not given DCBs. In either growth culture, Dehalococcoides sequences

were not found. Clone library construction and analysis was performed by Dr.

Hinsby Cadillo-Quiroz.

To quantify bacterial populations, real time PCR with SYBR green and specific

primers targeting Eubacteria, Dehalobacter, and Dehalococcoides were used.

83

Page 96: REDUCTIVE DEHALOGENATION OF

In the fourth transfer culture given DCB, the number of Dehalobacter 16s

rRNA gene copies was 3.5 + 0.3 x106 copies/ml, which was 100 times greater

than in culture not given DCBs (Figure 4.5). Dehalococcoides 16S rRNA gene

copies were detected both in the presence and absence of DCBs, 1.7+1.0

x104 copy/ml and 2.1+1.4 x104 copy/ml, respectively (Figure 4.5). In culture

given DCB, Dehalococcoides 16S rRNA gene copies were 0.02% of the total

Eubacteria 16S rRNA gene copies. The low copy number of Dehalococcoides

may have been below the detection limit of the clone library. Dehalobacter

16S rRNA gene copy number was 3% of the total Eubacteria 16S rRNA gene

copies in cultures given DCB. Table 4.2. Phylogenetic groups identified in clone library from fourth transfer culture grown in the presence or absence of DCBs. Group assignment was determined using Ribosomal Database Classifier using an 80% confidence threshold (38).

Restriction Type

Library Phylogenetic Assignment* No. Clones

(sequenced)

+DCB Dehalobacter 31 (11) Clostridiales 28 (7) Clostridiaceae 1 4 (3) Unassigned 7(0) -DCB Parabacteroides 18 (6) Clostridiales 12 (4) Clostridium 9 (2) Spirochaetaceae 9 (5) Peptostreptococcaceae incertae sedis 3 (1) Ruminococcaceae 2 (1) Bacteroidales 1 (1) Bacteroides 1 (1) Desulfovibrio 1 (1) Firmicutes 1 (1) Unassigned 7(0)

84

Page 97: REDUCTIVE DEHALOGENATION OF

Figure 4.5. Comparison of 16S rRNA gene copy numbers from fourth transfer cultures grown with or without DCBs using primers that target Dehalococcoides (Dhc), Dehalobacter (Dhb) and Eubacteria (Eub). Error bars represent standard deviation.

To examine changes in the bacterial population upon using sodium sulfide as

a reducing agent and transferring solely on 1,2-DCB, genomic DNA was

extracted from seventh transfer cultures in the presence and absence of 1,2-

DCB after 0 days and 15 days when the culture given 1,2-DCB produced 933

µmoles/L of MCB. Using real time PCR analysis, the culture without 1,2-DCB

showed a 10-fold decrease in Dehalobacter 16S rRNA gene copies while the

Eubacteria population increased by 20-fold (Figure 4.6A). In culture given 1,2-

DCB both Dehalobacter and Eubacteria 16S rRNA gene copies increased by

20- and 10-fold respectively, indicating growth of Dehalobacter in

dehalogenating conditions (Figure 4.6B). Based on 16S rRNA gene copies,

Dehalobacter was 14% of the total Eubacteria. In both cultures,

Dehalococcoides PCR product was not detected. After the seventh transfer

culture conditions were modified to enrich for Dehalobacter.

85

Page 98: REDUCTIVE DEHALOGENATION OF

Figure 4.6. Comparison of 16S rRNA gene copies of Dehalobacter (Dhb) and Eubacteria (Eub) from seventh transfer culture grown (A) without 1,2-DCB and (B) with 1,2-DCB at day 0 and 15. Error bars represent standard deviation.

Targeted Enrichment of Dehalobacter. As Dehalobacter is known only to

utilize H2 as an electron donor (18), the 1,2-DCB culture was serially diluted up

to 10-8 with H2 as the electron donor. Repeated serial dilutions have not

yielded 1,2-DCB reductive dehalogenation past 10-4 dilution (Figure 4.7A).

After 28 days, the number of Dehalobacter 16S rRNA gene copies was

1.6+0.3x106 copies/ml and 53% of the copies detected with eubacterial

primers, while no PCR product was detected for Dehalococcoides (Figure

4.7B). The inability to transfer dechlorination activity to higher dilutions and

the presence of other Eubacteria in the culture suggests a reliance upon

another microorganism(s) to provide an unknown growth factor(s) similar to

strain DF-1 whose growth relies on extracts of Desulfovibrio spp.(23).

86

Page 99: REDUCTIVE DEHALOGENATION OF

Figure 4.7. Hydrogen-fed mixed culture with sodium sulfide as a reducing agent (A) reductive dehalogenation of 1,2-DCB and (B) 16S rRNA gene copies of Dehalobacter (Dhb) and Eubacteria (Eub) after 25 days. Error bars represent standard deviation.

Discussion

This chapter describes the establishment of a 1,2-DCB reductively

dehalogenating mixed culture from sediment microcosms that can be

successively transferred into growth medium. The mixed culture has been

transferred 11 times over the course of ca. 1.4 years: the first three transfers

were made in the presence of all three DCB isomers (ca. 0.75 year), 1,2- and

1,3-DCB for fourth and fifth transfer (ca. 0.25 year), and solely 1,2-DCB for the

remaining transfers (ca. 0.4 year). In accord with other studies (1, 2, 10), the

mixed culture showed a clear preference for the removal of the chlorine in the

ortho position. Over the course of transfers, several changes were made to

the growth conditions (Table 4.1), with the most notable being the addition of

2-bromoethanesulfonate (BES) to inhibit methanogens, sodium sulfide as a

reducing agent, and the switch from yeast extract to butyrate and eventually

hydrogen as the electron donor.

87

Page 100: REDUCTIVE DEHALOGENATION OF

During the initial transfer, methanogenesis, a competitive reaction with

reductive dehalogenation, was observed. BES has been used with limited

success in reductive dehalogenation studies as it can also inhibit reductive

dehalogenation as in Dehalococcoides strain 195 on chloroethenes (9) but not

strain CBDB1 growing on chlorobenzenes (1). In this mixed culture, it did not

inhibit the overall reductive dehalogenation activity, but it could have inhibited

the activity of specific dehalogenating microorganisms. When examining

TeCB and PeCB reductive dehalogenation to MCB, Middeldorp et al (26)

found that adding BES inhibited several dechlorination pathways in their mixed

consortium. Sodium sulfide is an alternative reducing agent from trivalent

titanium but can be lethal to some dehalogenators such as strain CBDB1, but

not others like Dehalobacter restrictus (1, 18, 25). While low levels of sulfide

can be used as a sulfur source, it is also known that high levels of sulfide can

be inhibitory to methanogens (6, 7). At the fifth transfer, sodium sulfide was

used and methanogenesis was decreased even further with no effect on

reductive dehalogenation.

The ability of the current mixed culture to reductively dehalogenate DCB

isomers other than 1,2-DCB is unclear. While reductive dehalogenation of

1,3- and 1,4-DCB was observed at the fifth transfer, it was not detected at the

seventh transfer when sodium sulfide was used as a reducing agent. It is

possible that there were multiple dehalogenating populations with mixed

sensitivity to sodium sulfide or that sodium sulfide itself inhibits the reductive

dehalogenation of 1,3- and 1,4-DCB. It is common for the physiological

properties of mixed cultures to change over the course of transfers. Holliger et

88

Page 101: REDUCTIVE DEHALOGENATION OF

al. (19) also observed a loss of substrate range with other isomers of TCB with

the transfer of their mixed culture to solely 1,2,3-TCB. When Nowak et al. (30)

adapted their FBR to single DCB isomers, they found that 1,2-DCB or 1,4-

DCB-adapted FBRs could only dehalogenate that specific isomer and 1,3-

DCB-adapted to dehalogenated both 1,2- and 1,3-DCB.

Benzene production was reminiscent of a cometabolic process in that it is only

observed during the reductive dehalogenation of higher chlorinated benzenes

and in minor amounts. This activity does not resemble that of microcosms

used as original inoculum, which produced benzene at much higher

concentrations (>3000 µmoles/L) and in the absence of DCB reductive

dehalogenation (14). It could not be determined if benzene production

occurred only after 1,2-DCB reductive dehalogenation ceased like ethene

production from vinyl chloride in strain 195 (24) or only during the reductive

dehalogenation of 1,2-DCB as previously observed in sediment microcosms

(30, 31). It is likely that another microorganism was performing the MCB

reductive dehalogenation observed in sediment microcosms.

Using a combination of culture independent methods, Dehalobacter

sequences were identified and revealed to be a major component in the 1,2-

DCB mixed culture and its growth to be linked to reductive dehalogenation. Its

cultured representative, Dehalobacter restrictus, is a known dehalogenator of

chlorinated ethenes but is not known to reductively dehalogenate

chloroaromatics (18, 21). Dehalobacter strain TCA1 has been shown to

reductively dehalogenate 1,1,1-trichloroethane (1,1,1-TCA) and can use either

H2 or formate as an electron donor (35). Grostern et al. (15) has shown

89

Page 102: REDUCTIVE DEHALOGENATION OF

Dehalobacter to reductive dehalogenate 1,2-dichloroethane via

dichloroelimination forming vinyl chloride. A Dehalobacter sp. has been

identified in a coculture with Sedimentibacter sp. that is able to reductively

dehalogenate β-hexachlorocyclohexane to MCB and benzene (36).

Dehalobacter may have an unexplored diverse range of halogenated

compounds but has been difficult to isolate due to its fastidious growth

requirements. It is only known to use hydrogen as an electron donor ( strain

TCA1 being the exception) and halogenated compounds as electron

acceptors.

Other identified bacteria in the DCB-fed clone library belong to known

fermentors like Clostridia were most likely using the small amount of yeast

extract present and providing hydrogen via fermentation reactions or other

unknown essential nutrients. Since analysis was done when butyrate was

provided as the electron donor a complex community with fermentors was

expected. It is interesting to note the absence of Dehalococcoides and

Chloroflexi sequences as they represent the only known groups with cultured

chlorinated benzene dehalogenators. The addition of BES in culturing

conditions may have selected against Dehalococcoides. Though no

Dehalococcoides sequences were found in the 1,2-DCB mixed consortium,

they still might play a larger role in the reductive dehalogenation within the

original sediment microcosm, which is known to contain a diverse population

of Chloroflexi (Haitjema, C., personal comm.).

This study is the first description of a 1,2-DCB reductive dehalogenating mixed

culture in sediment-free medium. Dehalobacter sp. has been identified as a

90

Page 103: REDUCTIVE DEHALOGENATION OF

member of the culture and may play a major role in reductive dehalogenation,

as it was not present in clone libraries from cultures not fed 1,2-DCB. In

chloroethane dehalogenating communities Dehalobacter and

Dehalococcoides have been shown to play complementary roles in reductive

dehalogenation (15). Further attempts to isolate the microorganism(s)

responsible for 1,2-DCB dehalogenation are underway. Once the

microorganism has been identified physiological parameters such as

temperature and competing electron acceptors can be tested to better

understand its capabilities as a bioremediation agent.

91

Page 104: REDUCTIVE DEHALOGENATION OF

REFERENCES

1. Adrian, L., and H. Gorisch. 2002. Microbial transformation of

chlorinated benzenes under anaerobic conditions. Res Microbiol 153:131-7.

2. Adrian, L., S. K. Hansen, J. M. Fung, H. Gorisch, and S. H. Zinder. 2007. Growth of Dehalococcoides strains with chlorophenols as electron acceptors. Environ Sci Technol 41:2318-23.

3. Adrian, L., U. Szewzyk, J. Wecke, and H. Gorisch. 2000. Bacterial dehalorespiration with chlorinated benzenes. Nature 408:580-3.

4. Bhatnagar, L., Henriquet, M., Zeikus, J.G., Aubert, J.P. 1984. Utilization of mercapto-2-ethanol as medium reductant for determination of the metabolic repsonse of methanogens towards inorganic sulfur compounds. FEMS Microbiol Letters 22:155-158.

5. Bosma, T. N. P., van der Meer, J.R., Schraa. G., Tros, M. E., Zehnder, A.J.B. 1988. Reductive dechlorination of all trichloro- and dichlorobenzene isomers. FEMS Microbiology Ecology 53:223-229.

6. Brauer, S. L., E. Yashiro, N. G. Ueno, J. B. Yavitt, and S. H. Zinder. 2006. Characterization of acid-tolerant H/CO-utilizing methanogenic enrichment cultures from an acidic peat bog in New York State. FEMS Microbiol Ecol 57:206-16.

7. Cadillo-Quiroz, H., E. Yashiro, J. B. Yavitt, and S. H. Zinder. 2008. Characterization of the archaeal community in a minerotrophic fen and terminal restriction fragment length polymorphism-directed isolation of a novel hydrogenotrophic methanogen. Appl Environ Microbiol 74:2059-68.

8. Coates, J. D., R. Chakraborty, and M. J. McInerney. 2002. Anaerobic benzene biodegradation--a new era. Res Microbiol 153:621-8.

9. DiStefano, T. D., J. M. Gossett, and S. H. Zinder. 1992. Hydrogen as an electron donor for dechlorination of tetrachloroethene by an anaerobic mixed culture. Appl Environ Microbiol 58:3622-9.

92

Page 105: REDUCTIVE DEHALOGENATION OF

10. Fathepure, B. Z., J. M. Tiedje, and S. A. Boyd. 1988. Reductive dechlorination of hexachlorobenzene to tri- and dichlorobenzenes in anaerobic sewage sludge. Appl Environ Microbiol 54:327-30.

11. Fennell, D. E., J. M. Gossett, and S. H. Zinder. 1997. Comparison of butyric acid, ethanol, lactic acid, and propionic acid as hydrogen donors for the reductive dechlorination of tetrachloroethene. Environ. Sci. Technol. 31:918-926.

12. Fennell, D. E., I. Nijenhuis, S. F. Wilson, S. H. Zinder, and M. M. Haggblom. 2004. Dehalococcoides ethenogenes strain 195 reductively dechlorinates diverse chlorinated aromatic pollutants. Environ Sci Technol 38:2075-81.

13. Fung, J. M., R. M. Morris, L. Adrian, and S. H. Zinder. 2007. Expression of reductive dehalogenase genes in Dehalococcoides ethenogenes strain 195 growing on tetrachloroethene, trichloroethene, or 2,3-dichlorophenol. Appl Environ Microbiol.

14. Fung, J. M., Weisenstein, B.P., Mack, E.E., Vidumsky, J.E., Ellis, D., and Zinder, S.H. 2008. Reductive dehalogenation of dichlorobenzenes to monochlorobenzene and benzene in microcosms. Environ Sci Technol submitted.

15. Grostern, A., and E. A. Edwards. 2006. Growth of Dehalobacter and Dehalococcoides spp. during degradation of chlorinated ethanes. Appl Environ Microbiol 72:428-36.

16. Hendrickson, E. R., J. A. Payne, R. M. Young, M. G. Starr, M. P. Perry, S. Fahnestock, D. E. Ellis, and R. C. Ebersole. 2002. Molecular analysis of Dehalococcoides 16S ribosomal DNA from chloroethene-contaminated sites throughout North America and Europe. Appl Environ Microbiol 68:485-95.

17. Holliger, C., S. Gaspard, G. Glod, C. Heijman, W. Schumacher, R. P. Schwarzenbach, and F. Vazquez. 1997. Contaminated environments in the subsurface and bioremediation: organic contaminants. FEMS Microbiol Rev 20:517-23.

93

Page 106: REDUCTIVE DEHALOGENATION OF

18. Holliger, C., D. Hahn, H. Harmsen, W. Ludwig, W. Schumacher, B. Tindall, F. Vazquez, N. Weiss, and A. J. Zehnder. 1998. Dehalobacter restrictus gen. nov. and sp. nov., a strictly anaerobic bacterium that reductively dechlorinates tetra- and trichloroethene in an anaerobic respiration. Arch Microbiol 169:313-21.

19. Holliger, C., G. Schraa, A. J. Stams, and A. J. Zehnder. 1992. Enrichment and properties of an anaerobic mixed culture reductively dechlorinating 1,2,3-trichlorobenzene to 1,3-dichlorobenzene. Appl Environ Microbiol 58:1636-44.

20. Lovley, D. R. 2000. Anaerobic benzene degradation. Biodegradation 11:107-16.

21. Maillard, J., W. Schumacher, F. Vazquez, C. Regeard, W. R. Hagen, and C. Holliger. 2003. Characterization of the corrinoid iron-sulfur protein tetrachloroethene reductive dehalogenase of Dehalobacter restrictus. Appl Environ Microbiol 69:4628-38.

22. Mansunaga, S., Susarla, Sridhar, and Yonezawa, Yoshitaka. 1996. Dechlorination of chlorobenzenes in anaerobic estuarine sediment. Wat. Sci. Tech. 33:173-180.

23. May, H. D., G. S. Miller, B. V. Kjellerup, and K. R. Sowers. 2008. Dehalorespiration with polychlorinated biphenyls by an anaerobic ultramicrobacterium. Appl Environ Microbiol 74:2089-94.

24. Maymo-Gatell, X., T. Anguish, and S. H. Zinder. 1999. Reductive dechlorination of chlorinated ethenes and 1, 2-dichloroethane by "Dehalococcoides ethenogenes" 195. Appl Environ Microbiol 65:3108-13.

25. Maymo-Gatell, X., Y. Chien, J. M. Gossett, and S. H. Zinder. 1997. Isolation of a bacterium that reductively dechlorinates tetrachloroethene to ethene. Science 276:1568-71.

26. Middeldorp, P., J. De Wolf, A. Zehnder, and G. Schraa. 1997. Enrichment and properties of a 1,2,4-trichlorobenzene-dechlorinating methanogenic microbial consortium. Appl Environ Microbiol 63:1225-1229.

94

Page 107: REDUCTIVE DEHALOGENATION OF

27. Mohn, W. W., and K. J. Kennedy. 1992. Reductive dehalogenation of chlorophenols by Desulfomonile tiedjei DCB-1. Appl Environ Microbiol 58:1367-70.

28. Nadkarni, M. A., F. E. Martin, N. A. Jacques, and N. Hunter. 2002. Determination of bacterial load by real-time PCR using a broad-range (universal) probe and primers set. Microbiology 148:257-66.

29. Nishino, S. F., J. C. Spain, L. A. Belcher, and C. D. Litchfield. 1992. Chlorobenzene degradation by bacteria isolated from contaminated groundwater. Appl Environ Microbiol 58:1719-26.

30. Nowak, J., Kirsche, N.H., Hegemann, W., Stan, H.J. 1996. Total reductive dechlorination of chlorobenzenes to benzene by a methanogenic mixed culture enriched from Saale river sediment. Appl Environ Microbiol 45:700-709.

31. Quistorff, A. S. 1999. Microbially mediated reductive dechlorination of dichlorobenzene. Master of Science. Cornell University, Ithaca.

32. Ramanand, K., M. T. Balba, and J. Duffy. 1993. Reductive dehalogenation of chlorinated benzenes and toluenes under methanogenic conditions. Appl Environ Microbiol 59:3266-72.

33. Shelton, D. R., and J. M. Tiedje. 1984. Isolation and partial characterization of bacteria in an anaerobic consortium that mineralizes 3-chlorobenzoic acid. Appl Environ Microbiol 48:840-848.

34. Smidt, H., and W. M. de Vos. 2004. Anaerobic microbial dehalogenation. Annu Rev Microbiol 58:43-73.

35. Sun, B., B. M. Griffin, H. L. Ayala-del-Rio, S. A. Hashsham, and J. M. Tiedje. 2002. Microbial dehalorespiration with 1,1,1-trichloroethane. Science 298:1023-5.

36. van Doesburg, W., van Eekert, M. H., Middeldorp, P., Balk, M., Schraa, G., Starns, A. 2005. Reductive dechlorination of β-hexachlorocyclohexane (β-HCH) by a Dehalobacter species in coculture with a Sedimentibacter sp. FEMS Microbiol Ecology 54:87-95.

95

Page 108: REDUCTIVE DEHALOGENATION OF

37. von Wintzingerode, F., B. Selent, W. Hegemann, and U. B. Gobel. 1999. Phylogenetic analysis of an anaerobic, trichlorobenzene-transforming microbial consortium. Appl Environ Microbiol 65:283-6.

38. Wang, Q., G. M. Garrity, J. M. Tiedje, and J. R. Cole. 2007. Naive Bayesian classifier for rapid assignment of rRNA sequences into the new bacterial taxonomy. Appl Environ Microbiol 73:5261-7.

39. Wu, Q., C. E. Milliken, G. P. Meier, J. E. Watts, K. R. Sowers, and H. D. May. 2002. Dechlorination of chlorobenzenes by a culture containing bacterium DF-1, a PCB dechlorinating microorganism. Environ Sci Technol 36:3290-4.

96

Page 109: REDUCTIVE DEHALOGENATION OF

CHAPTER FIVE

SUMMARY AND IMPLICATIONS IN THE FIELD OF BIOREMEDIATION

Summary and Conclusions: Chapter Two

Dehalococcoides ethenogenes strain 195 has 19 potential RD genes including

pceA and tceA, whose gene products are responsible for the reductive

dehalogenation of perchloroethylene (PCE) to trichloroethene (TCE) and TCE

to vinyl chloride (VC) and ethene (ETH), respectively (33). Though strain 195

is able to dehalogenate a large range of other halogenated compounds, RD

genes responsible have yet to be identified (3, 17, 24). RD gene, DET0162,

contains a verified TGA stop codon that would cause the predicted gene

product to be 59 amino acids rather than the typical size of 500 amino acids

(19, 47). This truncation, and a shorter corresponding "B" gene (DET0163),

thought to encode for a membrane anchoring protein, suggest it is no longer

functional. Sixteen of the 19 RD genes in D. ethenogenes, including pceA,

have predicted transcriptional regulator genes in close proximity suggesting

they are under transcriptional regulation (47). The objective of Chapter Two

was to investigate RD gene function and gain insight into its regulation.

Dehalococcoides ethenogenes strain 195 was grown on PCE, TCE or 2,3-

dichlorophenol (2,3-DCP) and RD gene transcript levels was monitored using

reverse transcription quantitative PCR. Resting cell assays and proteomic

analysis was performed on PCE- and 2,3-DCP grown cells to verify RD gene

products and physiological activity. When grown on PCE or TCE, tceA was

expressed at levels several fold higher than that of rpoB, a housekeeping

gene used as a marker of relevant transcript activity and 300-fold lower levels

97

Page 110: REDUCTIVE DEHALOGENATION OF

when grown on 2,3-DCP. This was the first proof that tceA was regulated on

the transcriptional level, which was interesting as it does not have adjacent

regulatory genes. Both pceA and DET0162 were expressed under all growth

conditions. Though it was known that TCE-grown cells reductively

dehalogenated PCE (35), it was unknown at what level the activity could be

ascribed to and this study showed pceA was transcribed. While DET0162

was detected in transcript studies, proteomic analysis did not detect any

associated peptides and most likely its transcript is not processed into a

functional RD.

DET1559 transcript was detected at significant levels when cells were grown

with PCE but its peptides were not detected in proteomic studies. In a PCE-

dehalogenating mixed culture containing strain 195, DET1559 peptides were

detected (37). Perhaps DET1559 was translated but at lower abundance than

PceA or TceA making it difficult to detect. Using microarrays, Johnson et al.

(28) found DET1559 to be transcribed in TCE-fed cultures though under

different growth condition (defined medium) then this study, which contained

two undefined growth supplements (extract PCE-fed butyrate mix culture and

supernatant of anaerobic sludge digestor) (36).

In this study pceA was identified to be the 2,3-DCP reductive dehalogenase.

In transcript and peptide analysis pceA was the only RD gene detected when

grown with 2,3-DCP. Providing additional evidence for PceA bifunctionality,

PCE- and 2,3-DCP resting cell showed reductive dehalogenation activity for

both PCE and 2,3-DCP. Located upstream of pceA are two-component

system (TCS) regulatory genes. Ten of the nineteen RD genes have

98

Page 111: REDUCTIVE DEHALOGENATION OF

associated TCS genes and their histidine kinases (HK) lack the conserved

membrane-spanning regions and are thought to be soluble (47). It would be of

interest to discover how PCE, TCE or 2,3-DCP each acts as a ligand for the

HK to activate transcription. An alternative scenario is that pceA is

constitutively expressed like that of Sulfurospirillum multivorans (39). In strain

195, pceA has been shown to be under temporal control with varying levels of

transcript but never completely turned off (43).

Insight into RD Gene Evolution

The bifunctionality of pceA may reveal information of its genetic evolution.

Though PCE can be produced by natural sources such as algae (1), the

majority produced is anthropogenic (www.epa.gov). How did strain 195 and

other bacteria evolve to utilize this manmade compound? 2,3-DCP represents

a set of compounds commonly found in nature that can be produced by plants

and fungi (21). It is possible that 2,3-DCP is the natural substrate of pceA and

genetic evolution has modified it to utilize PCE. This concept has been shown

in Escherichia coli and Klebsiella aerogenes adaptation to novel five-carbon

sugar, D-arabinose using the L-fucose pathway (7). The enzymes in the L-

fucose pathway can use D-arabinose but is not expressed in its presence.

Regulatory mutations are the first to occur which make the operon inducible by

the new substrate or constitutively expressed. While it can be energetically

costly to constantly produce proteins, selective advantages of constitutively

expressed metabolic genes are a shorter lag time and improved activity by

protein saturation. PceA might also be undergoing selection for improved

catalytic capacity. In the case of E. coli strain K12 acquiring ethylene glycol

utilization, an anthropogenic substrate not metabolized by wild type, mutations

99

Page 112: REDUCTIVE DEHALOGENATION OF

occur in lactaldehyde dehydrogenase, which improved substrate kinetics for

an alternate substrate (5). It is possible that pceA is undergoing a similar

evolutionary process. Alvarez-Cohen et al. (personal communication) have

been transferring strain 195 continuously on TCE and recently discovered the

loss of pceA transcription. In addition to insight into RD genetic evolution,

mutations that cause loss of transcription or RD gene function would provide

insight into its mechanism and regulation.

Given that over 80 homologous RD genes in Dehalococcoides spp. have been

identified with strain CBDB1 harboring 32 and strain 195 with 19 (27, 31, 47),

one can speculate on the evolutionary circumstances that caused such a

proliferation within the genus and individual strains. Gene duplication is an

efficient method for bacteria to develop new functions without losing existing

function. Paralogous genes involved in metabolism are shown to undergo

preferential retention and also, undergo faster mutation rates (20, 29). Strain

195 contains several RD genes with high amino acid sequence identity

(>30%) which is a possible sign of past gene duplication events (Table 5.1)

(20). What do these duplication events mean in terms of halogenated

substrate range? The two vinyl chloride reductases, BvcA and VcrA, from

strain BAV1 and VS, respectively, are 38% identical at the amino acid level

and 39% and 36% with TceA of strain 195, respectively (30, 38). All three

RDs reductively dehalogenate VC. TceA is able to reductively dehalogenate

TCE, dichloroethene but not PCE. In comparison VcrA is able to reductively

dehalogenate TCE at low rates and dichloroethenes but not PCE. The

transcription and proteomic assay developed in Chapter Two can help address

questions in evolution by identifying RD gene function. It would be of interest

100

Page 113: REDUCTIVE DEHALOGENATION OF

to see if the RDs with higher sequence identity (>50%) such as DET1522 and

DET1535 have distinct or overlapping substrate ranges.

Table 5.1. Comparison of reductive dehalogenases of strain 195. Sequence identities of 30% or greater are in bold. Gene/ identity

(%) 88 TceA 162 876 302* 173 1538 1545* 1535* 1522* PceA* 306* 1538* 1519* 1559 235* 311* 180* 1171*

88 12 12 10 10 9 9 9 12 11 11 10 10 10 12 9 10 7 6 tceA 17 20 21 20 20 21 21 22 22 22 20 21 21 18 21 14 13 162 17 16 15 15 14 16 15 15 15 20 21 21 18 21 14 13 876 51 49 43 40 26 25 27 27 27 26 31 29 25 14 16 302 44 39 37 26 29 28 27 28 25 30 29 27 17 17 173 40 38 26 29 28 28 29 25 34 28 27 16 14

1538 34 25 27 27 25 26 24 27 27 26 17 14 1545 24 27 27 27 28 25 29 26 23 16 14 1535 54 45 47 30 31 26 24 26 14 14 1522 42 42 30 30 25 24 27 16 14 PceA 49 32 30 26 26 25 16 14 306 28 27 26 25 25 17 14

1528 40 28 24 25 16 14 1519 25 25 25 15 13 1559 35 29 18 14 235 30 17 15 311 16 15 180 14

1171

*homologs in strain CBDB1 with amino acid sequence identity between 86.4-95.4% (31)

Improved Assessment of Reductive Dehalogenation Activity

By correlating RD gene transcript levels with physiological activity, I have

strengthened the biological understanding of reductive dehalogenation, which

can be used for improved bioremediation applications. Using a combination of

transcription, proteomics and resting cell assays, RD gene function in strain

195 can now be determined for other halogenated substrates such as

101

Page 114: REDUCTIVE DEHALOGENATION OF

pentachlorobenzene. It is critical to use cultured microorganisms such as

strain 195 to identify RD gene function, or RD analysis on uncharacterized

mixed cultures and in situ field studies will not have relevance.

The ability for a contaminated site to undergo stimulated or natural

bioremediation is usually accessed with DNA-based techniques by the

determining the presence of Dehalococcoides spp. using 16s rRNA genes and

specific RD genes such as tceA or bvcA (16, 25). However, neither of these

markers actually denotes physiological activity. RD gene transcripts or

peptides are dynamic markers that can be used to determine substrate

inhibition, growth stage, or nutrient limitations. Lee et al. (32) monitored

Dehalococcoides population and RD gene transcript of vcrA, bvcA, and tceA in

a TCE-contaminated field site and found changes in RD gene transcript as

chlorinated ethenes were dehalogenated. The tceA of strain 195 is under

genetic regulation, though the nature of control is unknown. Its expression is

not affected by the presence of 2,3-DCP. Superfund sites are contaminated

with multiple organic substrates from chlorobenzenes to chloromethane as

well as heavy metals. It would be of interest to see if the tceA expression of

strain 195 is inhibited under these complex conditions.

RDs can be used to directly assess growth stage and reductive dehalogenase

activity. In a study with strain 195 examining whole genome expression during

the transition between exponential and stationary phase, unique RD genes

(DET0173, DET0180, DET1535, and DET1545) were expressed during

stationary phase (28). During stationary phase in strain 195, growth and

reductive dehalogenation are uncoupled and these stationary phase-RDs can

102

Page 115: REDUCTIVE DEHALOGENATION OF

be used to as indicators of growth state. Using a proteomic approach Morris

et al. (37) was able to identify key RD peptides correlating with vcrA from

dechlorinating cultures enriched from a TCE-plume at the heavily

contaminated Department of Energy, Savannah River National Laboratory

site. For site applications, this technique could be adapted to directly detect

RD from sediment samples instead of enrichment cultures.

The transcript and proteomics approach can be expanded to correlate

physiological states with RD activity. In ocean surface waters, a survey of

community gene expression identified high levels of transcripts involved in

carbon fixation which correlated with the carbon-limited environment (18).

Using proteomics to analyze an acid mine drainage microbial community,

peptides involved in oxidative stress and protein refolding were found to be

abundant (44). These proteins reflect the environmental challenges

microorganisms face living in highly acidic environment (pH<1).

Transcriptome of strain 195 can be generated under different conditions such

as nutrient limiting or in a mixed community and RD activity levels monitored.

Specific conditions that effect RD activity can be identified which will provide a

global perspective on the physiology of Dehalococcoides.

Summary and Conclusions: Chapters Three and Four

Reductive dehalogenation of dichlorobenzenes (DCBs) to monochlorobenzene

(MCB) and MCB to benzene has only been demonstrated in microcosms, and

no organisms have been identified to be involved in these processes (2).

Several sediment microcosms and flow columns from a variety of

103

Page 116: REDUCTIVE DEHALOGENATION OF

contaminated sites reductive dehalogenate DCBs to MCB with preferences for

specific isomers (9, 42, 45). Though thermodynamically favorable (14),

reductive dehalogenation of MCB to benzene has been rarely observed and

only known to occur doing the reductive dehalogenation of higher chlorinated

benzenes (42, 45). The objective of these chapters was to characterize the

reductive dehalogenation of DCB and MCB in sediment from a historically

chlorobenzene-contaminated site (Chapter Three) and a mixed culture

developed from the DCB-dehalogenating sediment microcosms (Chapter

Four).

In Chapter Three, I described the complete reductive dehalogenation of DCBs

and MCB to benzene under methanogenic conditions in sediment

microcosms. Sediment microcosms given all three DCB isomers reductively

dehalogenated them all to MCB with a preference for 1,2- followed by 1,3- and

1,4-DCB. During the reductive dehalogenation of DCB, MCB was slowly

dehalogenated to benzene, but once the DCBs were consumed, MCB

turnover was rapid. This amount of reductive dehalogenation of MCB to

benzene was an unprecedented amount of ca. 3500 µmoles/L compared to 50

µmoles/L observed by Nowak et al. (41). The reductive dehalogenation of

MCB in the absence of DCBs was only observed in microcosms constructed

from one of three sediment samples, indicating differences in dehalogenating

activity in the microbial populations. Reductive dehalogenation was enhanced

when sediment microcosms were inoculated with sediment slurry from a

previously chlorinated benzene-dehalogenating microcosm. The addition of

an inoculum eliminated the initial lag in MCB reductive dehalogenation, and

inoculated microcosms immediately dehalogenated MCB concurrently with

104

Page 117: REDUCTIVE DEHALOGENATION OF

DCBs. More importantly, inoculated microcosms could consistently

dehalogenate MCB in the absence of DCB and this activity was maintained

over several transfers into sediment microcosms.

The mechanism of MCB reductive dehalogenation remains elusive. The

inoculation with dehalogenating sediment seemed to activate the MCB-

dehalogenating population in some unknown manner. It was possible the

inoculation provided a nutrient that the MCB-dehalogenating population could

not generate or lacked the activation energy to initially produce. The reductive

dehalogenation of MCB to benzene in the absence of higher chlorinated

benzenes has never been observed before, and sediment microcosms in this

study have potential uses in the field of bioremediation.

In Chapter Four, a 1,2-DCB-dehalogenating mixed culture was developed

from DCB-dehalogenating sediment microcosms. Sediment slurry from a

microcosm was initially transferred into mineral salts medium with yeast

extract, and all three DCB isomers. Over the course of transferring, the mixed

culture apparently lost its ability to reductively dehalogenate 1,3- and 1,4-DCB.

Culturing conditions were modified over several transfers to minimize

methanogenesis, a competing reaction, including the addition of 2-

bromoethanesulfonate (BES). The DCB reductive dehalogenation activity was

vancomycin sensitive indicating that Dehalococcoides was not the causative

agent. The use of BES may have selected against Dehalococcoides in the

mixed culture, as it can inhibit reductive dehalogenation activity in some

strains. The 1,2-DCB-dehalogenating mixed culture was able to produce trace

amounts of benzene but only if grown on DCB. This metabolism was not

105

Page 118: REDUCTIVE DEHALOGENATION OF

similar to the robust benzene production found in the original sediment

microcosm and may be a cometabolic process. To identify possible candidate

dehalogenators, a 16S rRNA gene clone library was constructed from the

fourth transfer cultures and sequences matching >99% similarity to

Dehalobacter restrictus, a known dehalogenator, was identified. At the

seventh transfer growth of Dehalobacter population was linked to reductive

dehalogenation and accounted for 14% of the total bacterial population.

Culturing conditions were modified to enrich for Dehalobacter, and by the

eighth transfer Dehalobacter accounted for 53% of the total bacterial

population.

Though identified over a decade ago (26), not much is known about the genus

Dehalobacter. The pceA of Dehalobacter restrictus has been identified (34)

and using degenerate primers, two RD genes of unknown function were

identified, rdA1-Dr and rdA2-Dr (46). Unlike its dehalogenating counterparts,

Desulfitobacterium and Dehalococcoides, a representative genome for

Dehalobacter has not been sequenced though they are known to reductively

dehalogenate chlorinated ethenes, ethanes, and hexanes (23, 26, 48, 50).

Due to the lack of genomic and information the reductive dehalogenation

capability of Dehalobacter remains unknown and unexplored.

Potential for Bioremediation Application of MCB-Dehalogenating Microcosm

In addition to anthropogenic introduction of DCB and MCB into the

environment, they can accumulate as end products from the reductive

dehalogenation of higher chlorinated benzenes. It has been observed that

106

Page 119: REDUCTIVE DEHALOGENATION OF

higher chlorinated benzenes are more readily dehalogenated in anaerobic

environments and pollutant profiles of sites can change over time (2, 10).

Sediment analysis spanning several decades of Lake Ketelmeer, which

receives water from the Rhine River in Germany, found an accumulation of

DCBs resulting from the reductive dehalogenation of hexachlorobenzene, the

original pollutant (8). While MCB oxidation and reduction has been

demonstrated (40-42), the low concentration of end products does not alter the

accumulative toxicity of the compounds found on site. However in this study,

significant amounts of DCBs and MCB were reductively dehalogenated to

benzene, a more toxic compound (49). Benzene has been shown to be

recalcitrant in a number of environments (6, 15) and in those cases, the

conversion of MCB to benzene is not desirable. The dynamics of DCB and

MCB reductive dehalogenation needs to be taken into account for a better

understanding of their fate in the environment.

The sediment microcosms in Chapter Three have a potential to be developed

for field applications. The simple technique of inoculating sediment

microcosms eliminated lag time of DCB and MCB reductive dehalogenation. It

would be interesting to test if inoculating a site with actively dehalogenating

sediment could stimulate sustained activity for in situ bioremediation.

However, environmental factors that influence chlorinated benzene

degradation are unknown. The Rhine River has been contaminated with MCB

for decades with very little conversion to benzene (8). Environmental

conditions may exist that inhibit its reductive dehalogenation, and experiments

with mixed sediment microcosms would be informative of the types of

107

Page 120: REDUCTIVE DEHALOGENATION OF

environments in which reductive dehalogenation of DCBs and MCB could

function.

The sediment in this study can also be used to develop a broad range

dehalogenating mixed cultures that degrade higher chlorinated benzenes to

carbon dioxide or methane. The anaerobic degradation of benzene can occur

by oxidation to carbon dioxide or reduction to methane (13). The MCB-

dehalogenating sediment microcosms from this study can be mixed with a

benzene-utilizing culture for complete remediation (11, 12). Mixing cultures

can also relieve substrate inhibition if multiple halogenated substrates inhibit

reactions. Grostern and Edwards (22) were able to overcome chloroethane

inhibition of trichloroethene reductive dehalogenation by combining a

trichloroethane dehalogenating mixed culture with one that dehalogenates

chloroethenes. The sediment in this study has not been tested for its ability to

reductively dehalogenate tetra- or higher chlorinated benzenes but has been

shown to reductively dehalogenate trichlorobenzenes (Fullerton, H., personal

communication). There are sediment sources and pure cultures, like strain

CBDB1, that can reductively dehalogenate higher chlorinated benzenes that

could be used to broaden the halogenated substrate range (2, 4). Due to its

robustness and unique ability utilize DCBs and MCB, the reductive

dehalogenating microorganisms found in the 1,2-DCB mixed culture and within

sediment microcosms have the potential to be highly effective in

bioremediation applications.

108

Page 121: REDUCTIVE DEHALOGENATION OF

REFERENCES

1. Abrahamsson, K., Ekdahl, A., Collen, J., and Pedersen, M. 1995.

Marine algae-a source for trichloroethylene and perchloroethylene. Limnol Oceanogr 40:1321-1326.

2. Adrian, L., and H. Gorisch. 2002. Microbial transformation of chlorinated benzenes under anaerobic conditions. Res Microbiol 153:131-7.

3. Adrian, L., S. K. Hansen, J. M. Fung, H. Gorisch, and S. H. Zinder. 2007. Growth of Dehalococcoides strains with chlorophenols as electron acceptors. Environ Sci Technol 41:2318-23.

4. Adrian, L., U. Szewzyk, J. Wecke, and H. Gorisch. 2000. Bacterial dehalorespiration with chlorinated benzenes. Nature 408:580-3.

5. Aguilar, J., Baldoma, L. 1992. The development of a catabolic pathway for ethylene glycol, p. 31-46. In R. P. Mortlock (ed.), The evolution of metabolic function. CRC Press, Boca Raton.

6. Ball, H., and Reinhard, M. 1996. Monoaromatic hydrocarbon transformation under anaerobic conditions at Seal Beach, California:

laboratory studies. Environ. Tosicol. Chem. 15:114-122.

7. Bartkus, J. M. a. M., R.P. 1992. Natural and experimentally evolved pathways for the utilization of the D-arabinose in enteric bacteria, p. 15-30. In R. P. Mortlock (ed.), The evolution of metabolic function. CRC Press, Boca Raton.

8. Beurskens, J. E. M., C. G. C. Dekker, J. Jonkhoff, and L. Pompstra. 1992 - 1993. Microbial dechlorination of hexachlorobenzene in a sedimentation area of the Rhine River. Biogeochemistry 19: 61-81.

9. Bosma, T. N. P., van der Meer, J.R., Schraa. G., Tros, M. E., Zehnder, A.J.B. 1988. Reductive dechlorination of all trichloro- and dichlorobenzene isomers. FEMS Microbiology Ecology 53:223-229.

109

Page 122: REDUCTIVE DEHALOGENATION OF

10. Brahushi, F., U. Dorfler, R. Schroll, and J. C. Munch. 2004. Stimulation of reductive dechlorination of hexachlorobenzene in soil by inducing the native microbial activity. Chemosphere 55:1477-84.

11. Burland, S. M., and E. A. Edwards. 1999. Anaerobic benzene biodegradation linked to nitrate reduction. Appl Environ Microbiol 65:529-33.

12. Coates, J. D., R. Chakraborty, J. G. Lack, S. M. O'Connor, K. A. Cole, K. S. Bender, and L. A. Achenbach. 2001. Anaerobic benzene oxidation coupled to nitrate reduction in pure culture by two strains of Dechloromonas. Nature 411:1039-43.

13. Coates, J. D., R. Chakraborty, and M. J. McInerney. 2002. Anaerobic benzene biodegradation--a new era. Res Microbiol 153:621-8.

14. Dolfing, J., and B. K. Harrison. 1993. Redox and reduction potentials as parameters to predict the degradation pathway of chlorinated benzenes in anaerobic environments. FEMS Microbiology Ecology 13:23-29.

15. Edwards, E. A., L. E. Wills, M. Reinhard, and D. Grbic-Galic. 1992. Anaerobic degradation of toluene and xylene by aquifer microorganisms under sulfate-reducing conditions. Appl Environ Microbiol 58:794-800.

16. Fennell, D. E., A. B. Carroll, J. M. Gossett, and S. H. Zinder. 2001. Assessment of indigenous reductive dechlorinating potential at a TCE-contaminated site using microcosms, polymerase chain reaction analysis, and site data. Environ Sci Technol 35:1830-9.

17. Fennell, D. E., I. Nijenhuis, S. F. Wilson, S. H. Zinder, and M. M. Haggblom. 2004. Dehalococcoides ethenogenes strain 195 reductively dechlorinates diverse chlorinated aromatic pollutants. Environ Sci Technol 38:2075-81.

18. Frias-Lopez, J., Y. Shi, G. W. Tyson, M. L. Coleman, S. C. Schuster, S. W. Chisholm, and E. F. DeLong. 2008. Microbial community gene expression in ocean surface waters. Proceedings of the National Academy of Sciences 105:3805-3810.

110

Page 123: REDUCTIVE DEHALOGENATION OF

19. Fung, J. M., R. M. Morris, L. Adrian, and S. H. Zinder. 2007. Expression of reductive dehalogenase genes in Dehalococcoides ethenogenes strain 195 growing on tetrachloroethene, trichloroethene, or 2,3-dichlorophenol. Appl Environ Microbiol.

20. Gevers, D., K. Vandepoele, C. Simillon, and Y. Van de Peer. 2004. Gene duplication and biased functional retention of paralogs in bacterial genomes. Trends Microbiol 12:148-54.

21. Gribble, G. W. 2003. The diversity of naturally produced organohalogens. Chemosphere 52:289-97.

22. Grostern, A., and E. A. Edwards. 2006. A 1,1,1-trichloroethane-degrading anaerobic mixed microbial culture enhances biotransformation of mixtures of chlorinated ethenes and ethanes. Appl Environ Microbiol 72:7849-56.

23. Grostern, A., and E. A. Edwards. 2006. Growth of Dehalobacter and Dehalococcoides spp. during degradation of chlorinated ethanes. Appl Environ Microbiol 72:428-36.

24. He, J., Robrock, K. R., Alvarez-Cohen, L. 2006. Microbial reductive debromination of polybrominated diphenyl ethers (PBDEs). Environ Sci Technol 40:4429-4434.

25. Hendrickson, E. R., J. A. Payne, R. M. Young, M. G. Starr, M. P. Perry, S. Fahnestock, D. E. Ellis, and R. C. Ebersole. 2002. Molecular analysis of Dehalococcoides 16S ribosomal DNA from chloroethene-contaminated sites throughout North America and Europe. Appl Environ Microbiol 68:485-95.

26. Holliger, C., D. Hahn, H. Harmsen, W. Ludwig, W. Schumacher, B. Tindall, F. Vazquez, N. Weiss, and A. J. Zehnder. 1998. Dehalobacter restrictus gen. nov. and sp. nov., a strictly anaerobic bacterium that reductively dechlorinates tetra- and trichloroethene in an anaerobic respiration. Arch Microbiol 169:313-21.

27. Holscher, T., R. Krajmalnik-Brown, K. M. Ritalahti, F. Von Wintzingerode, H. Gorisch, F. E. Loffler, and L. Adrian. 2004.

111

Page 124: REDUCTIVE DEHALOGENATION OF

Multiple nonidentical reductive-dehalogenase-homologous genes are common in Dehalococcoides. Appl Environ Microbiol 70:5290-7.

28. Johnson, D. R., E. L. Brodie, A. E. Hubbard, G. L. Andersen, S. H. Zinder, and L. Alvarez-Cohen. 2008. Temporal transcriptomic microarray analysis of "Dehalococcoides ethenogenes" strain 195 during the transition into stationary phase. Appl Environ Microbiol 74:2864-72.

29. Kondrashov, F. A., I. B. Rogozin, Y. I. Wolf, and E. V. Koonin. 2002. Selection in the evolution of gene duplications. Genome Biol 3:RESEARCH0008.

30. Krajmalnik-Brown, R., T. Holscher, I. N. Thomson, F. M. Saunders, K. M. Ritalahti, and F. E. Loffler. 2004. Genetic identification of a putative vinyl chloride reductase in Dehalococcoides sp. strain BAV1. Appl Environ Microbiol 70:6347-51.

31. Kube, M., A. Beck, S. H. Zinder, H. Kuhl, R. Reinhardt, and L. Adrian. 2005. Genome sequence of the chlorinated compound-respiring bacterium Dehalococcoides species strain CBDB1. Nat Biotechnol 23:1269-73.

32. Lee, P. K., T. W. Macbeth, K. S. Sorenson, Jr., R. A. Deeb, and L. Alvarez-Cohen. 2008. Quantifying genes and transcripts to assess the in situ physiology of "Dehalococcoides" spp. in a trichloroethene-contaminated groundwater site. Appl Environ Microbiol 74:2728-39.

33. Magnuson, J. K., R. V. Stern, J. M. Gossett, S. H. Zinder, and D. R. Burris. 1998. Reductive dechlorination of tetrachloroethene to ethene by a two-component enzyme pathway. Appl Environ Microbiol 64:1270-5.

34. Maillard, J., W. Schumacher, F. Vazquez, C. Regeard, W. R. Hagen, and C. Holliger. 2003. Characterization of the corrinoid iron-sulfur protein tetrachloroethene reductive dehalogenase of Dehalobacter restrictus. Appl Environ Microbiol 69:4628-38.

35. Maymo-Gatell, X., T. Anguish, and S. H. Zinder. 1999. Reductive dechlorination of chlorinated ethenes and 1, 2-dichloroethane by

112

Page 125: REDUCTIVE DEHALOGENATION OF

"Dehalococcoides ethenogenes" 195. Appl Environ Microbiol 65:3108-13.

36. Maymo-Gatell, X., Y. Chien, J. M. Gossett, and S. H. Zinder. 1997. Isolation of a bacterium that reductively dechlorinates tetrachloroethene to ethene. Science 276:1568-71.

37. Morris, R. M., J. M. Fung, B. G. Rahm, S. Zhang, D. L. Freedman, S. H. Zinder, and R. E. Richardson. 2007. Comparative proteomics of Dehalococcoides spp. reveals strain-specific peptides associated with activity. Appl Environ Microbiol 73:320-6.

38. Muller, J. A., B. M. Rosner, G. Von Abendroth, G. Meshulam-Simon, P. L. McCarty, and A. M. Spormann. 2004. Molecular identification of the catabolic vinyl chloride reductase from Dehalococcoides sp. strain VS and its environmental distribution. Appl Environ Microbiol 70:4880-8.

39. Neumann, A., H. Scholz-Muramatsu, and G. Diekert. 1994. Tetrachloroethene metabolism of Dehalospirillum multivorans. Arch Microbiol 162:295-301.

40. Nijenhuis, I., N. Stelzer, M. Kastner, and H. H. Richnow. 2007. Sensitive detection of anaerobic monochlorobenzene degradation using stable isotope tracers. Environ Sci Technol 41:3836-42.

41. Nowak, J., Kirsche, N.H., Hegemann, W., Stan, H.J. 1996. Total reductive dechlorination of chlorobenzenes to benzene by a methanogenic mixed culture enriched from Saale river sediment. Appl Environ Microbiol 45:700-709.

42. Quistorff, A. S. 1999. Microbially mediated reductive dechlorination of dichlorobenzene. Master of Science. Cornell University, Ithaca.

43. Rahm, B. G., R. M. Morris, and R. E. Richardson. 2006. Temporal expression of respiratory genes in an enrichment culture containing Dehalococcoides ethenogenes. Appl Environ Microbiol 72:5486-91.

113

Page 126: REDUCTIVE DEHALOGENATION OF

44. Ram, R. J., N. C. VerBerkmoes, M. l. P. Thelen, G. W. Tyson, B. J. Baker, R. C. Blake, M. Shah, R. L. Hettich, and J. F. Banfield. 2005. Community proteomics of a natural microbial biofilm. Science 308:1915-1920.

45. Ramanand, K., M. T. Balba, and J. Duffy. 1993. Reductive dehalogenation of chlorinated benzenes and toluenes under methanogenic conditions. Appl Environ Microbiol 59:3266-72.

46. Regeard, C., J. Maillard, and C. Holliger. 2004. Development of degenerate and specific PCR primers for the detection and isolation of known and putative chloroethene reductive dehalogenase genes. J Microbiol Methods 56:107-18.

47. Seshadri, R., L. Adrian, D. E. Fouts, J. A. Eisen, A. M. Phillippy, B. A. Methe, N. L. Ward, W. C. Nelson, R. T. Deboy, H. M. Khouri, J. F. Kolonay, R. J. Dodson, S. C. Daugherty, L. M. Brinkac, S. A. Sullivan, R. Madupu, K. E. Nelson, K. H. Kang, M. Impraim, K. Tran, J. M. Robinson, H. A. Forberger, C. M. Fraser, S. H. Zinder, and J. F. Heidelberg. 2005. Genome sequence of the PCE-dechlorinating bacterium Dehalococcoides ethenogenes. Science 307:105-8.

48. Sun, B., B. M. Griffin, H. L. Ayala-del-Rio, S. A. Hashsham, and J. M. Tiedje. 2002. Microbial dehalorespiration with 1,1,1-trichloroethane. Science 298:1023-5.

49. Synder, R., Witz, G., and Goldstein, B.D. 1993. The toxicology of benzene. Environ Health Perspect. 100:293-306.

50. van Doesburg, W., van Eekert, M. H., Middeldorp, P., Balk, M., Schraa, G., Starns, A. 2005. Reductive dechlorination of β-hexachlorocyclohexane (β-HCH) by a Dehalobacter species in coculture with a Sedimentibacter sp. FEMS Microbiol Ecology 54:87-95.

114