<<

University of Kentucky UKnowledge

University of Kentucky Doctoral Dissertations Graduate School

2008

MICROBIAL COMMUNITY STRUCTURE DYNAMICS IN OHIO RIVER SEDIMENTS DURING REDUCTIVE DECHLORINATION OF PCBS

Andres Enrique Nunez University of Kentucky

Right click to open a feedback form in a new tab to let us know how this document benefits ou.y

Recommended Citation Nunez, Andres Enrique, "MICROBIAL COMMUNITY STRUCTURE DYNAMICS IN OHIO RIVER SEDIMENTS DURING REDUCTIVE DECHLORINATION OF PCBS" (2008). University of Kentucky Doctoral Dissertations. 679. https://uknowledge.uky.edu/gradschool_diss/679

This Dissertation is brought to you for free and open access by the Graduate School at UKnowledge. It has been accepted for inclusion in University of Kentucky Doctoral Dissertations by an authorized administrator of UKnowledge. For more information, please contact [email protected]. ABSTRACT OF DISSERTATION

Andres Enrique Nunez

The Graduate School

University of Kentucky

2008 MICROBIAL COMMUNITY STRUCTURE DYNAMICS IN OHIO RIVER

SEDIMENTS DURING REDUCTIVE DECHLORINATION OF PCBS

ABSTRACT OF DISSERTATION

A dissertation submitted in partial fulfillment of the requirements for the degree of Doctor of Philosophy in the College of Agriculture at the University of Kentucky

By

Andres Enrique Nunez

Director: Dr. Elisa M. D’Angelo

Lexington, KY

2008

Copyright © Andres Enrique Nunez 2008 ABSTRACT OF DISSERTATION

MICROBIAL COMMUNITY STRUCTURE DYNAMICS IN OHIO RIVER SEDIMENTS DURING REDUCTIVE DECHLORINATION OF PCBS

The entire stretch of the Ohio River is under fish consumption advisories due to contamination with polychlorinated biphenyls (PCBs). In this study, natural attenuation and biostimulation of PCBs and microbial communities responsible for PCB transformations were investigated in Ohio River sediments. Natural attenuation of PCBs was negligible in sediments, which was likely attributed to low temperature conditions during most of the year, as well as low amounts of available nitrogen, phosphorus, and organic carbon. Moreover, surface sediments were relatively oxidized, as indicated by the prevalence of aerobic such as beta- , alpha-Proteobacteria, , and Nitrospira in 16S rRNA sediment clone libraries. On the other hand, several reductive dechlorinators were detected in sediments, including Dehalococcoides, Desulfitobacterium spp. which suggested that reductive dechlorination might be possible in sediments under certain biogeochemical conditions. Considerable amounts of PCBs were transformed by reductive dechlorination (80% in 177 days by pattern N) when sediments were maintained under anaerobic conditions, amended with nutrients and organic carbon, and incubated at 25 ºC in lab microcosms. Analysis of 16S rRNA clone libraries from these treatments revealed that , and were enriched and Proteobacteria were depleted compared to clone libraries from treatment without organic amendments. Reductive dechlorination was decreased in sediments incubated at 10 and 40 ºC, and was not affected by FeSO4 amendments compared to unamended sediments incubated at 25 ºC. Transformations of PCB-153 were investigated in sediments under anaerobic, aerobic and sequential anaerobic and aerobic conditions. Transformations were only observed in treatments with an anaerobic phase, which occurred by reductive dechlorination by pattern N. Neither PCB-153 nor dechlorination products PCB-99 or PCB-47 were transformed under aerobic conditions. Analysis of 16S rRNA clone libraries revealed that Bacteoridetes, Chloroflexi, and Firmicutes were enriched under anaerobic conditions and Proteobacteria were enriched under aerobic conditions. Results from this study revealed that natural attenuation and biostimulation were not effective at removing PCBs from Ohio River sediments. Hence, other remediation methods will need to be employed to decrease PCB levels in this ecosystem.

Keywords: Polychlorinated biphenyls, Reductive Dechlorination, Phylogenetic Analysis, Natural attenuation, Biostimulation

Student’s signature

Date

MICROBIAL COMMUNITY STRUCTURE DYNAMICS IN OHIO RIVER

SEDIMENTS DURING REDUCTIVE DECHLORINATION OF PCBS

By

Andres Enrique Nunez

Director of Dissertation

Director of Graduate Studies

RULES FOR THE USE OF DISSERTATIONS

Unpublished dissertations submitted for the Doctor's degree and deposited in the University of Kentucky Library are as a rule open for inspection, but are to be used only with due regard to the rights of the authors. Bibliographical references may be noted, but quotations or summaries of parts may be published only with the permission of the author, and with the usual scholarly acknowledgments.

Extensive copying or publication of the dissertation in whole or in part also requires the consent of the Dean of the Graduate School of the University of Kentucky.

A library that borrows this dissertation for use by its patrons is expected to secure the signature of each user.

Name Date

DISSERTATION

Andres Enrique Nunez

The Graduate School

University of Kentucky

2008 MICROBIAL COMMUNITY STRUCTURE DYNAMICS IN OHIO RIVER

SEDIMENTS DURING REDUCTIVE DECHLORINATION OF PCBS

DISSERTATION

A dissertation submitted in partial fulfillment of the requirements for the degree of Doctor of Philosophy in the College of Agriculture at the University of Kentucky

By

Andres Enrique Nunez

Director: Dr. Elisa M. D’Angelo

Lexington, KY

2008

Copyright © Andres Enrique Nunez 2008

To Lilia, Nicole and Samuel

ACKNOWLEDGEMENTS

First I would like to thank Dr. Elisa D’Angelo for givingme the opportunity to work under her supervision, and especially for her remarkable patience in guiding me through the writing of this dissertation. Next, I want to thank the members of my committee: Dr. Paul Cornelius, Dr. Mark Coyne, Dr. John Grove, Dr. Chris Matocha, and

Dr. Olw Wendroth for providing the appropriate comments to improve the quality of this dissertation and for enlightening me through the courses I took while being a graduate student at the University of Kentucky. To the outsider examiner, Dr. Richard Warner, for acceding to evaluate this work despite all the delays I had throughout the writing. To

Martin Vandiviere and Georgia Ziegler, the lab technicians I worked with during my dissertation, for helping with many of the analytical procedures I had to perform. To my fellow graduate students, which I will not try to list here so I will not leave anyone out, for providing a friendly and enjoyable atmosphere during all these years. To my family for always being there supporting me despite the distance, and last but not least to Lilia, for having so much patience with me, for always being there when I needed her, and for giving me my other two precious possessions, Nicole and Samuel, that just by taking a look at them make me forget about all the problems around and make feel like a kid again.

iii

Table of Contents

Acknowledgements ...... iii List of tables ...... vii List of figures ...... viii List of files ...... xi CHAPTER 1 INTRODUCTION ...... 1 Overview of the problem ...... 1 Nomenclature, production and uses of PCBs ...... 3 Toxicology of PCBs ...... 7 PCB contamination in the Ohio River ...... 11 Fate processes of PCBs in aquatic environments ...... 14 Aerobic biotransformations ...... 19 Anaerobic biotransformations ...... 21 Objectives and Hypothesis ...... 33 Dissertation Format ...... 35 CHAPTER 2 NATURAL ATTENUATION OF PCBS AND BACTERIAL COMMUNITY COMPOSITION IN OHIO RIVER SEDIMENTS...... 37 Introduction ...... 37 Materials and Methods ...... 41 Site description, sample preparation, and collection ...... 41 PCB extraction ...... 43 DNA extraction and PCR amplification, cloning , and sequencing ...... 43 DGGE analysis...... 46 Analytical methods ...... 46 Phylogenetic analysis of clone library data ...... 48 Statistical analysis of DGGE and sediment chemistry data ...... 49 Results and Discussion ...... 50 Surface water temperature and sediment porewater chemistry ...... 50 PCB losses from the sediment profile ...... 57

iv

DGGE analysis of bacterial communities ...... 60 Clone library analysis of bacterial communities ...... 60 CHAPTER 3 TEMPERATURE AND ELECTRON DONOR AVAILABILITY EFFECTS ON PCB REDUCTIVE DECHLORINATION RATES AND MICROBIAL COMMUNITY COMPOSITION IN OHIO RIVER SEDIMENTS ...... 79 Introduction ...... 79 Materials and Methods ...... 83 Site description and sample collection ...... 83 Microcosm treatments and sampling ...... 83 PCB extraction ...... 85 DNA extraction, PCR, cloning and sequencing...... 85 Phylogenetic analysis ...... 87 Analytical methods ...... 88 Results and Discussion ...... 89 Treatment effects on PCB removal ...... 92 River sediment bacterial community composition ...... 102 Temperature Effects on Anaerobic Bacterial Community Composition ...... 102

FeSO4 Amendment Effects on Anaerobic Bacterial Community Composition ..108 Organic Carbon Effects on Anaerobic Bacterial Community Composition ...... 109 CHAPTER 4 SEQUENTIAL ANAEROBIC-AEROBIC TREATMENT OF PCB-153 IN OHIO RIVER SEDIMENTS ...... 113 Introduction ...... 113 Materials and Methods ...... 115 Site description, sample preparation, and collection ...... 115 PCB extraction ...... 117 DNA extraction and PCR amplification, cloning, and sequencing ...... 117 Reductive dehalogenase gene analysis ...... 119 Phylogenetic analysis ...... 121 Analytical methods ...... 121

v

Results and Discussion ...... 123 Sequential anaerobic aerobic treatment of PCB-153 ...... 123 Detection of reductive dehalogenase genes ...... 126 Changes in microbial community structure under anaerobic and aerobic conditions ...... 130 CHAPTER 5 SUMMARY AND CONCLUSIONS ...... 142 REFERENCES ...... 146 VITA ...... 168

vi

List of Tables

Table 1.1 Selected Properties of Polychlorinated Biphenyls ...... 5 Table 1.2 Uses of PCBs...... 6 Table 1.3 Fish consumption advisories for states bordering the Ohio River ...... 13 Table 1.4 Comparison of dechlorination activities ...... 25 Table 2.1 Tukey-Kramer test for multiple comparisons of porewater chemistry parameters in Ohio River sediments ...... 56 Table 2.2 Distribution of bacteria in major taxonomic groups in libraries from Ohio River sediments during in situ bioremediation of PCBs for four months, as determined by the Classifier tool available at the Ribosomal Database Project (RDP) ...... 65 Table 3.1 Distribution of bacteria in major taxonomic groups in duplicate libraries (ORS FC1 and ORS FC2) from Ohio River sediments before anaerobic treatments, as determined by the Classifier tool available at the Ribosomal Database Project (RDP) ...... 95 Table 3.2 Distribution of bacteria in major taxonomic groups in libraries from Ohio River sediments after exposure to anaerobic treatment for 177 d, as determined by the Classifier tool available at the Ribosomal Database Project (RDP) ...... 103 Table 4.1 Distribution of bacteria in major taxonomic groups in libraries from Ohio River sediments during sequential anaerobic-aerobic treatment of PCB-153 for 250 d, as determined by the Classifier tool available at the Ribosomal Database Project (RDP) ...... 131

vii

List of Figures

Figure 1.1 Structure of Polychlorinated Biphenyls ...... 4 Figure 1.2 Metabolism of PCBs ...... 10 Figure 1.3 Ohio River and its basin...... 12 Figure 1.4 Environmental fate of PCBs ...... 15 Figure 1.5 Bioaccumulation of a chemical along a generic food chain ...... 18 Figure 1.6 Aerobic degradation of PCBs in the “upper pathway” ...... 20 Figure 1.7 3,4-dioxygenation of polychlorinated biphenyls ...... 22 Figure 1.8 Reductive dehalogenation of chloroorganic compounds...... 23 Figure 1.9 Phylogeny of known dechlorinating bacteria ...... 26 Figure 1.10 Mechanistic model for the corrinoid iron-sulfur reductive dehalogenase .....28 Figure 1.11 Standard Gibbs free energy of common redox reaction in the environment with hydrogen as electron donor ...... 30 Figure 1.12 Determination of the Gibbs free energy of reaction with PCB-153 as electron acceptor and H2 as electron donor ...... 31 Figure 2.1 Average temperatures in the Ohio River at Lloyd Greenup Lock during year 2006...... 51 Figure 2.2 Concentrations of dissolved porewater constituents in the Ohio River sediment profile ...... 52 Figure 2.3 Gas chromatograms of PCB mixtures determined at different depth after four months incubation in the Ohio River sediment profile ...... 58 Figure 2.4 Reductive dechlorination of PCB-6 to PCB-23 ...... 59 Figure 2.5 Agarose gel electrophoresis detection of genomic DNA extraction obtained from sediments collected at different depths in the Ohio River sediment profile (A), and corresponding PCR amplification products using primers 341F-GC/907R for DGGE analysis (B)...... 61 Figure 2.6 Denaturating gel gradient electrophoresis analysis of PCR-amplified 16S rDNA obtained from three replicates of four Ohio River sediment depth increments ...... 62

viii

Figure 2.7 Rarefaction curves generated from clone libraries obtained from four Ohio River sediment depth increments, as determined by the furthest neighbor assignment algorithm ...... 63 Figure 2.8 Bootstrapped neighbor-joining phylogenetic tree of 16S rDNA sequences obtained from the in situ reductive dechlorination study and known dehalogenating bacteria ...... 75 Figure 2.9 Bootstrapped neighbor-joining phylogenetic tree of 16S rDNA sequences obtained from the in situ reductive dechlorination study and known aerobic PCB degraders ...... 77 Figure 3.1 Anaerobic losses of 14C-2,4,5,2’,4’,5’-hexachlorobiphenyl (PCB-153) from Ohio River sediments under different treatment conditions, including (a) 10 ºC, 25 ºC, and 40 ºC, and (b) peptone, volatile fatty acid (VFAs) mixture...... 90 Figure 3.2 Polychlorinated biphenyl concentrations during reductive dechlorination of PCB-153 to PCB-99 and PCB-47 using peptone as organic carbon amendment ..93 Figure 3.3 Bootstrapped neighbor-joining phylogenetic tree of sequences in the phyla Firmicutes obtained from the present study and closest relatives in the Ribosomal Database Project release 9.48 ...... 110 Figure 4.1 PCB-153 conversion under anaerobic, aerobic and sequential anaerobic/aerobic conditions ...... 124 Figure 4.2 PCR amplification of reductive dehalogenase DNA from pure cultures ...... 127 Figure 4.3 PCR amplification products of 16S rDNA of Dehalococcoides spp...... 128 Figure 4.4 PCR amplification products of cprA gene of Desulfitobacterium dehalogenans with primers cprA1023F/cprA1202R ...... 129 Figure 4.5 Bootstrapped minimum-evolution phylogenetic tree of 16S rDNA sequences isolated from Ohio River sediments incubated anaerobically for 250 days, as well as known aerobic PCB degraders, and dehalogenating bacteria...... 137 Figure 4.6 Bootstrapped minimum-evolution phylogenetic tree of 16S rDNA sequences isolated from Ohio River sediments incubated aerobically for 250 days, as well as known aerobic PCB degraders, and dehalogenating bacteria ...... 139 Figure 4.7 Bootstrapped minimum-evolution phylogenetic tree of 16S rDNA sequences isolated from Ohio River sediments incubated anaerobically for 150 days and then

ix

aerobically for 100 days, as well as known aerobic PCB degraders, and dehalogenating bacteria ...... 140

x

List of Files

Anunez_2008_Dissertation.pdf (1.34 MB)

xi

CHAPTER 1

INTRODUCTION

Overview of the problem

Polychlorinated biphenyls (PCBs) are synthetic compounds that were synthesized

in the US and other countries between 1929 and 1977 for a wide range of industrial uses.

Since their introduction, several hundred million pounds have been released to the

environment, which has caused adverse effects to humans and biota due to their toxicity

and tendency to bioaccumulate in the food chain. Although most PCBs are chemically

and thermally inert, several aerobic bacterial strains such as , Burkholderia

and Rhodococcus spp. can transform the biphenyl ring of PCBs to relatively innocuous

products, including benzoate, CO2 and H2O. Unfortunately, aerobic biotransformation of

PCBs and other pollutants are limited in environments devoid of oxygen (e.g. anaerobic

sediments from aquatic environments). Aerobic biotransformation of PCBs are further

restricted to PCBs with a low number of chlorines on the biphenyl ring (e.g. <5). In many cases, PCBs have been released as complex mixtures (referred to as Aroclors and many other names) with between 10% (Aroclor 1242) and 100% (Aroclor 1260) of congeners having >5 chlorine atoms per biphenyl ring (Erickson, 1997). A prerequisite

for the removal of these congeners from the environment is the removal of several

chlorine atoms from the biphenyl ring (Abramowicz, 1995; Mohn and Tiedje, 1992).

Experiments with anaerobic sediments have shown that chlorines can be removed

from the biphenyl ring by microbial communities by a process referred to as reductive

dechlorination. Polychlorinated biphenyls tend to be dechlorinated preferentially at the

meta and para positions of the biphenyl ring, leading to the accumulation of ortho-

1

chlorinated congeners in environments with active dechlorination activity. To date, very

few bacteria with the ability to dechlorinate PCBs have been characterized, which include

Dehalococcoides spp. (Fennel et al., 2004), ortho dechlorinator o-17 (Cutter et al., 2001)

and double-flanked dechlorinator DF-1 (Wu et al., 2002). The capacity of these bacteria

to dechlorinate PCBs and other naturally-occurring and synthetic chlorinated

compounds, is attributed to the presence of one or more reductive dehalogenase genes, which couple the oxidation of simple electron donor compounds (e.g. H2, acetate) to the reduction of the PCBs in an energy-yielding process (Smidt and de Vos, 2004).

Reductive dechlorination is a highly syntrophic process in which dechlorination populations depend on electron donor substrates provided by other fermenting, and non- dechlorinating populations such as sulfate reducers. For this reason, it has been difficult to isolate and characterize reductively dechlorinating bacteria from anaerobic environments, determine their relationships to other microbial groups, or elucidate the environmental factors that affect dechlorinating microbial consortia and dechlorination activity. The main motivation of the research in this dissertation was to address questions such as how temperature, and electron donor/acceptor availability affect the microbial community composition in Ohio River sediments during reductive dechlorination of

PCBs.

2

Nomenclature, production, and uses of PCBs

Polychlorinated biphenyls (PCBs) are synthetic compounds that were manufactured in large quantities (1.4 billion lbs) from 1929 to 1977 in the United States for a multitude of industrial uses. They include 209 congeners that differ in the number of chlorines attached to the biphenyl ring (Figure 1.1). Congeners with the same number of chlorine atoms are called homologs, so that the pentachlorobiphenyls, for example, contain 46 congeners (Table 1.1). They have been commercialized as complex mixtures with different degrees of chlorination, under various trade names including Aroclor

(Monsanto, US and UK), Clophen (Bayer, Germany), Kanechlor (Kanegafuchi, Japan),

Phenoclor (Prodelec, France), and Fenclor (Caffaro, Italy). The more common products are Aroclors 1242–1260, Clophens A30–A60 and Kanechlors 300–600. For Aroclors, the first two digits of the numerical designation indicate the number of carbon atoms in the biphenyl ring system and the second two digits indicate the percent chlorine mass in the product. Monsanto manufactured Aroclor mixtures until 1970 in Anniston, AL, and until

1977 in Sauget, IL. Initially, these products were used as coolant/dielectric fluids in transformers and capacitors, as heat transfer fluids, and as flame resistante wood coatings. Due to their inertness, their use was later expanded to paints, inks, pesticides, and other uses (Table 1.2).

The extensive use of PCBs over several decades has resulted in widespread contamination of the environment. It has been estimated that several hundred million pounds have been released to the environment (Bedard, 2003). Unfortunately, many of the same chemical properties that make PCBs excellent choices for so many industrial

3

Cl 2 2' Cl 3 3' ortho=C2, C6 1 1' 4 4' meta=C3, C5 para=C4 5 6 6' 5'

Figure 1.1 Structure of polychlorinated biphenyls

4

Table 1.1 Selected Properties of Polychlorinated Biphenyls (Erickson, 1997)

Vapor Water Isomer Molecular Molecular No. of E at 25 ºC a Pressure solubility Log KOW Log KOC BAF in fish group formula Weight compounds (g m-2 h-1) (Pa) (mg L-1)

3 Mono-CB C12H9Cl 188.7 3 1.1 4.0 4.7 4.0 2.5×10 0.25

3 Di-CB C12H8Cl2 223.1 12 0.24 1.6 5.1 4.5 6.3×10 0.065

4 Tri-CB C12H7Cl3 257.5 24 0.054 0.65 5.5 4.9 1.6×10 0.017

4 -3 Tetra-CB C12H6Cl4 292.0 42 0.012 0.25 5.9 5.3 4.0×10 4.2×10

-3 5 -3

5 Penta-CB C12H5Cl5 326.4 46 2.6×10 0.099 6.3 5.8 1.0×10 1.0×10

-4 5 -4 Hexa-CB C12H4Cl6 360.9 42 5.4×10 0.038 6.7 6.2 2.5×10 2.5×10

-4 5 -5 Hepta-CB C12H3Cl7 395.3 24 1.3×10 0.014 7.1 6.6 6.3×10 6.2×10

-5 -3 6 -5 Octa-CB C12H2Cl8 429.8 12 2.8×10 5.5×10 7.5 7.0 1.6×10 1.5×10

-6 -3 6 -6 Nona-CB C12H1Cl9 464.2 3 6.3×10 2.0×10 7.9 7.5 4.0×10 3.5×10

-6 -4 7 -7 Deca-CB C12Cl10 498.7 1 1.4×10 7.6×10 8.3 7.9 1.0×10 8.5×10

Note: Most of the values presented are averages from congeners in group. BAF, bioaccumulation factor; E, evaporation rate

aLog Koc values estimated according to Girvin and Scott, 1997.

Table 1.2 Uses of PCBs (as percentage)

Before After Category Type of product 1971 1971

Transformer, capacitors, other electrical Closed systems 61 100 insulating/cooling applications

Nominally Hydraulic fluids, heat transfer fluids, lubricants 13 0 closed systems

Plasticizers, surface coatings, ink and dye carriers, Open-end adhesives, pesticide extenders, carbonless copy 26 0 applications paper, dyes.

6

uses also make them some of the worst materials that can be released to the environment.

Even by 1937, their toxic effect was evident in exposed workers, and by 1970,

production was banned and the use of PCBs was restricted to “closed systems” such as

those shown in Table 1.2 (Becton et al., 1979). Currently, PCBs are classified as the

twelfth most persistent organic pollutant in the world, according to the United Nations

Environment Programme Stockholm Convention (UNEP, http://www.pops.int/).

Toxicology of PCBs

Polychlorinated biphenyls are classified as probable carcinogens to humans

(ATSDR, 2001). Due to the extensive contamination of aquatic systems with PCBs, fish consumption represents the largest PCB exposure route to humans. As of 2004, the EPA issued 873 fish consumption advisories due to PCBs in 39 US states. According to the

Code of Federal Regulations, the tolerance for PCBs in fish is about 2 ppm

(21CFR109.30) and the maximum contaminant level for PCBs in drinking water to reduce cancer risk is 0.5 ppb (40CFR141.61). The Occupational Safety and Health

Administration (29CFR1910.1000) has also established permissible inhalation exposure levels of 0.5-1.0 mg/m3 in air for 8-h work periods.

The toxicity of commercial mixtures of PCBs depends on several factors, such as chlorine content and purity of the mixture, animal , age, gender, and route and duration of exposure. Symptoms of PCB exposure include elevated serum lipid levels, serum enzymes, chloracne, dermal lesions, hepatic responses, respiratory problems, eye irritation, decreased birth weight, and variable effects on cancer formation. The acute,

7

subacute, and subchronic toxic responses to commercial PCBs are characterized by lethal

dose (LD50) values that are usually > 1000 mg/kg. In chronic feeding studies, most of the

above toxic responses were observed, with liver being the primary target site (Safe,

1993). The University of California in Berkeley has developed the Carcinogenic Potency

Database (CPDB, http://potency.berkeley.edu/) providing tumor dose (TD50) values for

several chemicals including PCBs. This value indicates the required dose in mg kg-1 d-1

for the appearance of tumors in 50% of the test animals when the control animals do not

exhibit any tumors. Although individual PCB congeners were not tested, the TD50 values

for Aroclor mixtures range from 53.9 in Aroclor 1016 to 2.81 mg kg-1 d-1 in Aroclor

1260, indicating a higher potency of higher chlorinated mixtures as carcinogenic.

Commercial PCB mixtures produce a wide range of biochemical and toxic

responses and most of these are similar to those caused by dioxins (Safe, 1994). In

particular, three PCB congeners, PCB-77 (3,3’,4,4’-tetraCB), PCB-126 (3,3’,4,4’,5- pentaCB), and PCB-169 (3,3’,4,4’,5,5’-hexaCB) have the same spectrum of toxic and biochemical responses observed for dioxins, and therefore are the most toxic among all congeners (Safe, 1994). Maximum dioxin-like activities are associated with presence of two para chlorines, two or more meta chlorines, and no ortho chlorines (Safe, 1993) which allows for coplanarity of the biphenyl rings, thus resembling dioxin structures.

In eukaryotes, the metabolism and toxicity of PCBs is associated with activation of the Aryl Hydrocarbon Receptor (AhR), which is a ligand-dependent transcription factor that regulates transcription of certain key enzymes involved in the metabolism of xenobiotics (Hankinson, 1995; Denison and Nagy, 2003; Marlowe and Puga, 2005).

Although dioxins are the most potent ligands (Hankinson, 1995, Mimura and Fujii-

8

Kuriyama, 2003), PCBs can bind to the AhR, resulting in a complex that is transported to

the nucleus where it binds to specific genomic sequences to induce gene transcription.

Enzymes activated by this mechanism, are also known as Phase I-Phase II enzyme

system (Rose and Hodgson, 2004). As xenobiotics are usually lipophilic and tend to

accumulate in the lipid membranes and be transported with lipoproteins in the blood,

Phase I enzymes introduce polar reactive groups into the molecule to increase water solubility and to make it a more favorable substrate for Phase II enzymes. During Phase

II, conjugation reactions introduce bulky groups such as sulfates, sugars or amino acids that make the xenobiotics much more soluble and easily excreted from the organism (Fig.

1.2). Phase I enzymes usually involve cytochrome P450 (CYPs) and flavin-containing

(FMOs) monooxygenases, and several other dehydrogenases, oxidases, cyclooxygenases, reductases, and hydroxylases, while some Phase II enzymes include glucuronidases, sulfotransferases, methyltransferases, glutathione transferases, and acetyl transferases

(Rose and Hodgson, 2004).

The Phase I-Phase II enzyme system metabolizes a wide range of foreign substances in the organism, however, Phase I enzymes and especially CYPs can produce reactive intermediates such as arene oxides which are powerful electrophiles that can associate with nucleic acids and cause mutations leading to the activation of protooncogenes or inactivation of tumor suppressor genes (Hankinson, 1995).

9

Cly Clx

P450

Arene oxide Proteins Macromolecular Phenols intermediates RNA, DNA adducts

Epoxide hydrolase Glutathione S-transferase Dihydrodiols enzymes Phase Phase Dehydrogenase Glutathione Phase I I Phase enzymes conjugates Catechols II Multiple pathways Phenol Methyl sulfonyl conjugates metabolites

Figure 1.2 Metabolism of PCBs (adapted from Safe, 1994)

10

PCB contamination in the Ohio River

Many aquatic ecosystems near industrial areas are contaminated with PCBs. The most commonly studied systems include the Hudson River in New York, the Housatonic

River in Massachusetts, and the Fox River in Wisconsin. Although less studied, the Ohio

River is also extensively contaminated with PCBs, which impacts the health of wildlife and millions of people that utilize the resource (ORSANCO, 2002).

The Ohio River starts at the confluence of the Allegheny and Monongahela Rivers in Pittsburg, Pennsylvania, and travels 981 miles until it joins the Mississippi River in

Cairo, Illinois. Six states border the Ohio River (Illinois, Indiana, Kentucky, Ohio,

Pennsylvania, and West Virginia) and several other states (New York, North Carolina,

Maryland, Tennessee, and Virginia) are located in the Ohio River basin, which covers about 190,000 square miles and is inhabited by more than 25 million people (Fig. 1.3).

The Ohio River has four designated uses, including warm water aquatic habitat, public water supply, contact recreation, and fish consumption. Unfortunately, the river does not fully support the fish consumption use, due to contamination with PCBs, dioxin, and mercury (ORSANCO, 2006) (Table 1.3).

Several non-point and point sources are believed to contribute PCBs to the Ohio

River. According to an ORSANCO report (2002) the main non-point sources of PCBs to the river are atmospheric deposition, resuspension of contaminated sediments, overland runoff, and inflow of groundwater. On the other hand, the main point sources of PCBs identified so far are wastewater treatment plants, but several other industries are listed as potential PCB sources including several power generation and manufacturing industries.

11

#4830 #2024

#2015

#876

#2094 #104031 #4839 #104032

Figure 1.3 Ohio River and its basin. Fish consumption advisories are indicated according to Table 1.3. Arrows indicate extension of advisory.

12

Table 1.3 Fish Consumption Advisories for States Bordering the Ohio River

State Location From RM 0 to RM 31.7 (#2024) Pennsylvania From RM 31.7 to RM 40.0 (#4830)

Entire length in West Virginia West Virginia (#876)

Ohio Entire Length in Ohio (#2015)

From RM 317.1 to RM 436.2 Kentucky (#2094)

From RM 436.2 to RM 605.0 (#104031)

Kentucky From RM 605.0 to RM 981.0 (#104032)

Indiana Entire length in Indiana

Entire length in Illinois Illinois (#4839)

13

Polychlorinated biphenyls in Ohio River waters range from 0.54 to 4.55 ng/L, but in sediments this range goes up to 8 ppm, with hot spots located near zones with high industrial activity. For example, 8 ppm of PCBs were found in sediments at river miles

71.4, and 122.9, locations associated with steel and aluminum manufacturing industries, and wastewater treatment plants.

Fate processes of PCBs in aquatic environments

Once PCBs are released to aquatic environments, they may undergo numerous fate processes that depend largely on chemical properties of the PCB congeners and the ecosystem. Some of the most important fate processes of PCBs are depicted in Fig. 1.4, and key chemical properties of PCBs that govern the extent of processes are provided in

Table 1.1.

The extent of PCB volatilization into the atmosphere depends largely on the vapor pressure of individual PCBs, which range from 10-6 Pa for highly chlorinated congeners to 1 Pa for monochlorinated congeners (Table 1.1). Thus, the less chlorinated congeners would be expected to volatilize to a greater extent and be more widely distributed in the environment than highly chlorinated congeners. On the other hand, the highly chlorinated congeners would be expected to be relatively enriched in contaminated environments.

Once PCBs enter the atmosphere, they can remain in the vapor phase or associate with particulate matter. Reactions with OH radicals are the main atmospheric degradation pathway (Anderson and Hites, 1996) although photolysis with solar radiation or UV rays is also possible (Manzano et al., 2004). PCBs are eventually removed from the

14

Figure 1.4 Environmental fate of PCBs (from Safe et al., 1985).

15

atmosphere either by dry deposition (in which particulate materials settle down to the

ground or water bodies), or by wet deposition (in which they are removed from the

atmosphere by rain, snow, etc).

The concentration of dissolved PCBs in the water column of lakes and rivers

depends largely on the extent to which PCBs partition in the suspended and deposited

sediments. The extent that chemicals partition between the sediment and water is

described by the equilibrium sorption coefficient (Kd), or by the soil organic carbon

content normalized adsorption coefficient (KOC). The KOC is defined as the sorption

coefficient divided by the organic carbon content in soil or sediment. The KOC values can

be estimated from the KOW values, according to the following expression: Log(KOC) =

1.07×Log(KOW)-0.98 (Girvin and Scott, 1997). Once the KOC is estimated, Kd can be

determined by the following relationship: Kd=KOC×fOC, where fOC is the fraction of

organic carbon in the soil or sediment. As shown in Table 1.1, log KOC values for PCBs

range from 4.0, for less chlorinated PCBs, to 7.9 for highly chlorinated PCBs. Based on

log KOC values, lower chlorinated PCBs would be expected to occur at higher

concentrations in the water column than higher chlorinated congeners. This is important

because the amount of PCBs in the water column largely determines PCB mobility,

bioavailability, and volatilization processes. In addition, PCBs with high KOC values

would tend to accumulate in sediments, which could act as PCB sources to other parts of

the aquatic system via resuspension/desorption events.

Polychlorinated biphenyls may also accumulate in aquatic biota, particularly in

the fatty tissue due to the hydrophobic characteristics of PCBs. For hydrophobic

contaminants, the concentration of a chemical in biota is commonly described by the

16

equilibrium octanol-water coefficient (KOW) of the compound. The KOW is defined as the

concentration of a chemical in octanol divided by the concentration of a chemical in

water at equilibrium, with units of L water/L octanol. Octanol is commonly used because

it has the same carbon:oxygen ratio as lipids, and KOW values have a good correlation

with lipid-water partition coefficients (Shea, 2004). As shown in Table 1.1, the log KOW of PCBs ranges between 4.9 for low chlorinated PCBs to 8.3 for highly chlorinated

PCBs. Thus, the highly chlorinated PCBs would tend to accumulate to a greater extent in aquatic biota than lower chlorinated PCBs in contaminated environments.

Several terms are commonly used to describe the extent and processes by which contaminants accumulate in biota, including bioconcentration, biomagnification, and bioaccumulation (IUPAC, 1993; Leblanc, 2004). Bioconcentration is defined as the ratio of the chemical concentration in an organism and the chemical concentration in surrounding media. Biomagnification refers to the increase in chemical concentration with each progressive link in the food chain. Bioaccumulation is defined as the process by which organisms accumulate chemicals from surrounding media and dietary sources.

An example of how a chemical with a bioaccumulation factor (BAF) of 2 increases in the food chain is depicted in Figure 1.5. According to the example in the figure, the chemical concentration in the water has an arbitrary value of one. Assuming that the first trophic level accumulates the chemical only from the surrounding media, the concentration in the organism is two. The second trophic level can accumulate chemicals from both the surrounding media (two units) and the first trophic level (four units), yielding an overall concentration of chemical in the organism of six. BAF values are

related to the KOW values as follows: BAF=0.048×KOW (Mackay, 1982).

17

Figure 1.5 Bioaccumulation of a chemical along a generic food chain. Assumptions in this model include water concentration of 1, and bioaccumulation factor of 2 either from water or from one trophic level to another. Circled numbers represent the concentration in

the respective compartment. Arrow numbers represent the concentration of chemical

transferred from one compartment to another (from Leblanc, 2004).

18

Aerobic biotransformations

Polychlorinated biphenyls can be transformed by several bacteria under aerobic

conditions, including gram-negative genera Achromobacter, Acinetobacter, Alcaligenes,

Burkholderia, Comamonas, Moraxella, Pseudomonas, Ralstonia, and Sphingomomonas,

as well as gram-positive genera such as , Bacillus, , and

Rhodococcus, (Abraham et al., 2002; Field and Sierra-Alvarez, 2008). In most cases,

these organisms transform PCBs cometabolically, i.e. they do not gain carbon or energy from the reaction. For this reason, biphenyl or monochlorobiphenyls are commonly used to trigger the production of biphenyl dioxygenase enzymes that can also attack PCBs

(Borja et al., 2005). Bacteria capable of aerobic PCB degradation have been isolated from most of contaminated site samples (Field and Sierra-Alvarez, 2008) indicating the ubiquitous presence of aerobic PCB-degrading bacteria. Unfortunately, cometabolic reactions often lead to the production of dead-end and toxic products, such as chlorobenzoates, that can accumulate unless other microorganisms in the environment can mineralize these compounds (Pieper, 2005).

A common pathway by which aerobic bacteria transform PCBs is the 2,3- dioxygenase pathway or “upper pathway” that refers to the conversion of PCBs to chlorobenzoic acids (Borja et al., 2005, Pieper, 2005) (Figure 1.6). In this pathway, the

PCB is hydroxylated at positions 2 and 3 of the aromatic ring, the aromatic ring is

cleaved, and chlorobenzoate intermediates are generated. For this reaction to occur, no

chlorine substituents can occur at positions 2 or 3 of the aromatic rings. Several species

transform PCBs using this pathway, including gram negative strains of Pseudomonas,

19

Figure 1.6 Aerobic degradation of PCBs in the “Upper Pathway”

20

Alcaligenes, Achromobacter, Acinetobacter, and Moraxella and gram positive strains of

Arthrobacter and Rhodococcus (Furukawa et al., 1989).

Some bacterial strains, such as Ralstonia eutrophus H850 (Bedard et al., 1987;

Field and Sierra-Alvarez, 2008) and Pseudomonas sp. LB400 (Haddock et al., 1995) also possess 3,4-dioxygenase enzymes that can transform PCBs through 3,4-dioxygenation

(Fig. 1.7). Positions 2 and 3 are the preferred for attack in the upper pathway. However, if the ortho positions are blocked, the 3,4-dioxygenases provide the flexibility of attacking at positions 3 and 4 if no chlorines are present on those positions (Bedard et al., 1987).

This increases the substrate range for these genera and thus the aerobic degradation potential of these compounds. Both the 2,3- and 3,4-dioxygenation pathways are mostly limited to lower chlorinated congeners with no chlorine atoms in the indicated positions due to steric hindrance and low water solubility of highly chlorinated congeners.

Anaerobic biotransformations

Polychlorinated biphenyls can also be transformed by bacteria under anaerobic conditions by a process called reductive dechlorination. In this process, two-electrons and

a proton are transferred to the PCB, which results in the release of a chloride ion and a

less chlorinated PCB congener (Fetzner and Lingens, 1994) (Fig. 1.8). The redox reaction

is exergonic, and many organisms gain energy from the reaction by a process called

dehalorespiration (Löffler et al., 2003).

Another benefit of reductive dechlorination is that it converts highly chlorinated congeners to less chlorinated congeners, which are often not as toxic and are more amenable to aerobic biotransformations. A number of other halogenated chemicals

21

Figure 1.7 3,4-dioxygenation of polychlorinated biphenyls (Haddock et al., 1995)

22

Electron Electron Donor R-Cl Acceptor Reduced Oxidized e- n AT P

Electron Electron Donor R-H Acceptor Oxidized + Reduced H+ + Cl-

Figure 1.8 Reductive dehalogenation of chloroorganic compounds

23

besides PCBs can be reductively dechlorinated, including hexachlorobenzene, tetrachoroethylene, chlorobenzoates, chlorophenols, chlorobenzenes, chloromethanes, chloroethanes, and chloroethenes (Holliger et al., 2003). Early evidence of the importance of reductive dechlorination of PCBs in the environment was alterations in congener profiles in PCB contaminated aquatic sediments compared to those of the original commercial mixtures (Brown et al., 1987; Abramowicz, 1995; Wu et al., 1998).

Typically, less chlorinated, ortho-chlorinated congeners tend to accumulate and highly chlorinated meta- and para-chlorinated congeners tend to be depleted in these environments (Quensen et al., 1988). More recently, several other PCB reductive dechlorination patterns have been described, which are summarized in Table 1.4 (Bedard,

2003), and it is believed that different bacteria are responsible for the various pathways.

Several bacterial species with reductive dechlorination abilities have been identified (Fig. 1.9). However, only three microorganisms are recognized to reductively dechlorinate PCBs, which include o-17 (Cutter et al., 2001), DF-1 (Wu et al., 2002), and

Dehalococcoides spp. (Cutter et al., 2001). Reductive dechlorination is catalyzed by reductive dehalogenase enzymes (RD), such as trichloroethylene RD and vinyl chloride

RD from Dehalococcoides ethenogenes (Magnuson et al., 2000) chlorophenol RD from

Desulfitobacterium dehalogenans (van de Pas et al., 1999), and tetrachloroethylene RD

(pceA) from Desulfitobacterium hafniense (Maillard et al., 2004). Although no enzymes have been identified to specifically dechlorinate PCBs, reductive dehalogenases are usually associated with the cytoplasmic membrane and they are involved in electron transport-coupled phosphorylation processes for energy production (Futagami et al.,

2008). Structurally, RDs contain two Fe-S clusters in the catalytic center that are

24

Table 1.4 Comparison of dechlorination activities (Bedard, 2003)

Homolog Reactive Primary Dechlorination Targeted substrate chlorophenyl chlorophenyl activity chlorine range groupsa products 34, 234, 245, (23), 25, 235, P Flanked para 4-6 2345, 23456 2356

H Flanked parab 4-7 34, 234, 245, 2345 3, 24, 25, 235

23, 34, 234, 245, H’ Flanked parab,c 3-5 2, 3, 24, 25, 235 2345 234, 236, 245, N Flanked meta 5-9 24, 25, 26, 246 2345, 2346, 23456 Flanked & 3, 23, 25, 34, 234, M 2-4 2, 4, 24, 26 unflanked meta 236 Flanked & 4, 23, 24, 34, 234, Q unflanked 2-4 2, 3, 25, 26 245, 246 parab,c Flanked & LP 3-6 24, 245, 246 2, 25, 26 unflanked parad Doubly flanked T 7-8 2345 245 metae a The targeted chlorine(s) for each chlorophenyl groups is (are) underlined b The doubly flanked meta chlorine of 234-chlorophenyl groups is also targeted c the meta chlorine of 23-chlorophenyl groups is also targeted d The substrate range of this dechlorination process has not been completely characterized e This dechlorination process has been observed only at 50 to 60 ºC.

25

Figure 1.9 Phylogeny of known dechlorinating bacteria (from Löffler et al., 2003).

26

involved in the electron transfer to the chlorinated substrate, and most of these enzymes

also contain a corrinoid cofactor, a Co(I) complex that is similar in structure to vitamin

B12. A mechanistic model for the reductive dechlorination of organochlorine compounds is shown in Figure 1.10 (Krasotkina et al., 2001). Two pathways are presented in this mechanism. The first one (dashed arrows) involves formation of an organocobalt adduct, in which two electrons are transferred to the compound with removal of chlorine while the second pathway (solid arrows) involves transfer of one electron to the compound with formation of a radical anion, which finally loses a chlorine atom after transfer of another electron.

One of the most important factors governing reductive dechlorination is the redox status of the environment. In oxidized environments, bacteria use O2 as electron acceptor

to gain energy from the oxidation of electron donors. Due to the high reduction potential

of O2, bacteria using O2 as an electron acceptor gain high amounts of energy, grow

quickly, and outcompete other organisms that conduct lower energy-yielding reactions. In water-saturated environments, where O2 is consumed faster than it is supplied by diffusion, bacteria must use alternate electron acceptors for energy generation. The order that electron acceptors are reduced in the environment can be largely predicted from the amount of energy associated with the redox reactions, which can be calculated from the relationships ∆Gº = Σ∆Gºf(products) - Σ∆Gºf(substrates), and ∆Gº = -nF∆Eº, where n is the

number of electrons transferred during the reaction, F is the Faraday constant (96.48

kJ/V) and ∆Eº is the difference in potentials of the coupled oxidation-reduction reaction.

Using these relationships and tables of thermodynamic values of common electron

acceptors (Dolfing, 2003; Madigan et al., 2003a), the amount of free energy generated

27

Figure 1.10 Mechanistic model for the corrinoid iron-sulfur reductive dehalogenase.

Dashed arrows pathway (upper) involves formation of an organocobalt aduct, while solid

arrows pathway (lower) involves electron transfer from Co(I) to the aromatic ring

(Krasotkina et al., 2001).

28

from different environmental redox reactions using H2 as electron donor are shown in

Figure 1.11.

Based on these calculations, the predicted order of electron acceptor reduction is

4+ - 3+ 2- O2, Mn , NO3 , Fe , SO4 , and CO2. As mentioned earlier, some bacteria can also use

PCBs as electron acceptors. The amount of energy generated from the reaction can also

be derived from the Gibb’s free energies of formation of the reactants and

products, which were given by Holmes et al. (1993). Based on thermodynamic

calculations, about 116 kJ would be obtained from the reductive dechlorination of PCB-

153 (Fig. 1.12), which is less than the amount of energy derived from denitrification and greater than the amount of energy derived from sulfate reduction. Therefore, it is

predicted that reductive dechlorination of PCB-153 would be inhibited by aerobic and denitrifying activities, but not under sulfate-reducing and methanogenic conditions. In fact, Zwiernik et al. (1998) have shown that adding ferrous sulfate to sediment microcosms stimulated extensive meta and para dechlorination of PCBs. They observed an initial dechlorination inhibition upon sulfate addition, but after sulfate was depleted, amended microcosms showed greater dechlorination extention than unamended controls.

They proposed that the initial inhibition was due to electron acceptor shift to sulfate, and the greater extention of dechlorinaton was caused by an increase in the population of sulfate reducers. Interestingly, Gibbs free energies of reductive dechlorination reactions with highly chlorinated PCBs tend to be more negative (more favorable) than those with less chlorinated PCBs, which may partially explain bacterial preferences for highly chlorinated congeners as electron acceptors. Other environmental factors that affect

29

1 2 O22+ H  → HO 2 ∆Gº = -237.3 kJ/reaction

MnO2+ H 23 CO + H 2  → MnCO3 + 2 H 2 O ∆Gº = -194.9 kJ/reaction

−− NO32+ H  → NO 22 + H O ∆Gº = -162.1 kJ/reaction

2(FeOH )3+ 2 HCO 23 + H 2  →2 FeCO3 + 6 HO 2 ∆Gº = -117.7 kJ/reaction

=+− SO42+ H +44 H  → HS + H 2 O ∆Gº = -36.7 kJ/reaction

CO22+42 H  → CH 42 + H O ∆Gº = -32.8 kJ/reaction

Figure 1.11 Standard Gibbs free energy of common redox reactions in the environment

with hydrogen as electron donor.

30

Cl Cl Cl Cl Cl Cl + - + H2 + H + Cl Cl Cl Cl Cl Cl

     ∆GG = ∆ +∆ G+ +∆ G- - ∆ G +∆ G rd ( f aq PCB -99 f aq H f aq Cl ) ( f aq PCB -153f aq H2 )

 kJ kJ kJ kJ ∆Grd =(167.2reaction + 0 +−( 131.3reaction )) −( 151.8reaction + 0) =− 115.9 reaction

Figure 1.12 Determination of the Gibbs free energy of reaction with PCB 153 as electron

acceptor and H2 as electron donor.

31

reductive dechlorination include temperature, pH, and carbon source availability (Wiegel and Wu, 2000). Therefore, these factors must be considered when evaluating the potential for natural attenuation of PCBs in the environment. Temperature plays an important role in regulating reductive dechlorination by affecting bioavailability and transport of PCBs (Carlson and Hites, 2005; Hermanson et al., 2003), as well as controlling the growth and physiological activity of microbial populations (Madigan et al., 2003b). Reductive dechlorination of PCBs has been observed between 8 and 34 ºC, and 50 to 60 ºC, with optimal dechlorination between 18-30 ºC (Wiegel and Wu, 2000).

Dechlorination pathways may also be affected by temperature, probably by

affecting the composition of dechlorinating populations (Wu et al., 1997). For example,

the removal of flanked meta chlorines, was predominant at temperatures between 8 and

15 ºC, while removal of flanked meta and para chlorines was observed from 18 to 30 ºC

(Wu et al., 1997).

pH may affect the interaction between different dehalogenating and non-

dehalogenating microbial population, and also affects adsorption equilibrium of PCBs

with organic matter and thus influences their bioavailability (Jota and Hassett, 1991).

Maximum dechlorination activity has been observed in the pH range between 5.0 and 8.0

(Wiegel and Wu, 2000), but pH can also affect the dechlorination pathway. For example,

flanked meta dechlorination has been observed at pH 5.0-8.0, unflanked para

dechlorination at pH 6.0-8.0, and ortho dechlorination at pH 6.0-7.5 (Wiegel and Wu,

2000).

The availability of organic C also plays a major role in affecting reductive

dechlorination of PCBs (Nollet et al., 2005). During reductive dechlorination, fermenting

32

bacteria such as Clostridium spp. degrade organic C compounds derived from algae, plants, and animals in the environment to simpler compounds such as acetate and H2 which are utilized as electron donors by dehalogenating bacteria. If the concentrations of acetate and H2 are low due to low amounts of organic matter or consumption by methanogens, sulfate reducers, and acetogens (Smidt and de Vos, 2004; Chen et al.,

2006), then rates of reductive dechlorination will be decreased (Nies and Vogel, 1990;

Wiegel and Wu, 2000).

Objectives and Hypothesis

The overall goals of this project are to evaluate in situ PCB removal and the effects of several environmental factors on microbial community composition and PCB removal from Ohio River sediments, with particular emphasis on reductive dechlorination. Knowledge of the parameters that control the reductive dechlorination would be useful in designing strategies to improve removal of these pollutants from the

Ohio River where PCBs have been found in its entire lenght.

Specific objectives of the research are:

Objective 1. Determine the chemical profiles, bacterial community composition, and PCB loss rates in surface sediments from a specific location in the Ohio River.

Hypothesis: Several factors govern biological transformations of PCBs in the environment, such as temperature, nutrient composition, electron donor and acceptor levels, as well as the presence of microorganisms capable of transforming PCBs. The hypothesis tested in this study is that differences in chemical composition with depth will

33

significantly affect bacterial community composition and PCB removal rates in the

sediment profile. For example, it is expected that more reduced conditions in the deeper

sections of the sediment will be more conducive to anaerobic bacterial groups and

reductive dechlorination, while more oxidized conditions in the upper parts of the

sediment profile will favor aerobic transformations of PCBs. Such information is needed

to determine the potential for natural attenuation of PCBs, as well as determine the main

factors that limit biological PCB removal from this aquatic ecosystem, especially in tide

waters where saturated and unsaturated conditions may change often in the upper position

of the sediment in the river bank.

Objective 2. Determine effects of temperature and electron donor availability on

microbial community and PCB reductive dechlorination rates in Ohio River sediment

microcosms.

Hypothesis: Previous studies have shown that PCB dechlorination rates and pathways

temperature dependent. Different carbon sources and alternative electron acceptors have

also been shown to have an impact on these reactions. The hypothesis tested in this experiment is that these effects are connected with changes in microbial population composition, for example, adding sulfate would increase the population of sulfate reducers in these sediments which in turn may provide hydrogen as electron donor for dechlorinating organisms resulting in an increase in reductive dechlorination rates of

PCBs. Therefore, DNA analysis of clone libraries obtained from these microcosms should reveal an enrichment of these microorganisms. This information would be useful for the selection and establishment of environmental conditions favorable to

34

microorganisms involved in the reductive dechlorination of PCBs, which would lead to the improvement of removal rates of these contaminants from the environment.

Objective 3. Determine feasibility of sequential anaerobic-aerobic treatment for the complete removal of PCB-153 from Ohio River sediment microcosms, and the associated changes in microbial community composition during the process.

Hypothesis: The reductive dechlorination of PCBs is often limited to congeners with five or more chlorine atoms. Inversely, aerobic degradation is limited to congeners with four chlorine atoms or less. This is due to steric hindrance of the chlorine atoms that impede attack of dioxygenase enzymes at position 2,3 or 3,4 in the biphenyl ring. The hypothesis tested in this experiment is that the reductive dechlorination products of PCB-153

(2,2’,4,4’,5,5’-hexachlorobiphenyl) will undergo aerobic oxidation after oxygen exposure of microcosms that have been previously incubated under anaerobic conditions. PCB-153 is one of the most persistent PCBs in the environment and in human tissues. The chlorine substitution pattern in this congener blocks any possible attack of dioxygenase enzymes, however, either meta or para dechlorination will lead to free 2,3 or 3,4 positions where aerobic attack can now be possible.

Dissertation Format

Individual chapters in this dissertation were prepared as individual manuscripts intended for future publication. In this Chapter, a general introduction of PCB nomenclature, production, toxicology, and fate processes was presented. In Chapter 2, in situ PCB removal and bacterial community composition in an Ohio River sediment

35

profile was investigated. In Chapter 3, the effects of temperature, FeSO4, and electron donors such as peptone and volatile fatty acids on the PCB reductive dechlorination rate and the bacterial community composition in Ohio River sediments were described. In

Chapter 4, the effect of sequential anaerobic and aerobic treatments on removal of PCB-

153 in Ohio River sediments was described. Finally, a summary and the conclusions for the present project will be provided in Chapter 5.

36

CHAPTER 2

NATURAL ATTENUATION OF PCBS AND BACTERIAL COMMUNITY

COMPOSITION IN OHIO RIVER SEDIMENTS

Introduction

The Ohio River is one of the most extensively PCB-contaminated rivers in the

U.S. According to ORSANCO (2006) and Rostad et al. (1993), sediment samples collected along the 981 mile stretch of the river contained PCB concentrations ranging between 0.002 and 8 ppm. In addition, many fish species in the river contain > 2 ppm

PCBs. As a result, the EPA has instituted fish consumption advisories to warn the public

about the health hazards associated with consuming PCB-contaminated fish obtained

from the river (ORSANCO, 2006).

PCB contamination of the Ohio River impacts the lives of millions of people that

utilize the resource for aquatic life preservation, recreation, fish consumption, and as a

public water supply. The river starts at the confluence of the Allegheny and Monongahela

Rivers in Pittsburgh, Pennsylvania, and ends in Cairo, Illinois, where it flows into the

Mississippi River. Six states border the Ohio River including Illinois, Indiana, Kentucky,

Ohio, Pennsylvania, and West Virginia, and another five states are part of the river’s

190,000 square-mile basin, including New York, North Carolina, Maryland, Tennessee,

and Virginia. The Ohio River is believed to be the main source of PCBs to the

Mississippi River (Rostad et al., 1995), which extends for another 950 miles from their

confluence until it ends in the Gulf of Mexico.

It has been difficult to identify the main sources of PCBs to the Ohio River.

Effluents from wastewater treatment plants have been identified as important point

37

sources of PCBs (ORSANCO, 2002), but other potential point sources include discharges

from steel and aluminum manufacturing plants, power generation plants, and petroleum

distribution sites. High levels of PCBs have been detected in the air near areas of high

industrial activity, which is believed to contribute significant amounts of PCBs to the

river by atmospheric deposition (ORSANCO, 2002). Additional non-point sources of

PCBs include runoff and leakage from municipal landfills and hazardous wastes sites.

In many other PCB contaminated water bodies such as Fox River, Hudson River,

and Lake Hartwell, SC, it has been shown that PCBs undergo reductive dechlorination in

the sediments (Brown et al., 1987; Imamoglu et al., 2004; Magar et al., 2005). In this

process, chlorine substituents are replaced by hydrogen atoms, a process that is catalyzed

by specialized anaerobic bacteria that inhabit sediments such as o-17, DF-1, and

Dehalococcoides ethenogenes (Cutter et al., 2001; Wu et al., 2002; Fennell et al., 2004).

Although PCBs are not natural compounds, it is believed that reductive dechlorination is a common process due to the widespread distribution of other chlorinated compounds that are produced by a variety of abiotic and biotic processes in the environment (Gribble

1996; 1998; 2003).

Typically, reductive dechlorination leads to an enrichment of ortho-chlorinated

congeners compared to the original PCB congener distribution, which suggests that most

bacteria responsible for the process remove chlorines from the meta and para positions of

the biphenyl ring (Brown et al., 1987; Lake et al., 1992; Sokol et al., 1994; Bedard and

May, 1996). However, other dechlorination patterns have also been observed in some

environments, which suggest that other types of bacteria may be involved in the process.

38

Several factors influence reductive dechlorination in the environment, including temperature, pH, and electron donor/acceptor availability (Wiegel and Wu, 2000).

Therefore, these factors must be considered when evaluating the potential for natural attenuation of PCBs in the environment. Temperature plays an important role in regulating reductive dechlorination by affecting bioavailability and transport of PCBs

(Hermanson et al., 2003; Carlson and Hites, 2005), as well as controlling the growth and physiological activity of microbial populations (Madigan et al., 2003 pp 151-155). In most studies, reductive dechlorination has been studied at room temperature (25ºC), but the reaction has been found to occur at significant rates over a wide temperature range of between 8 and 34 ºC, and 50 to 60 ºC, with maximum rates observed between 18-30 ºC

(Wiegel and Wu, 2000). Temperature was also found to control PCB dechlorination pathways, presumably by affecting the composition of dechlorinating populations in

Woods Pond sediments (Wu et al., 1997). For example, process N (removal of flanked meta chlorines) was predominant at temperatures between 8 and 15 ºC, while process LP

(removal of flanked para chlorines) was observed along with process N, from 18 to 30

ºC.

The pH of the environment has also been found to be an important factor controlling reductive dechlorination of PCBs. pH may affect the interaction between different dehalogenating and non-dehalogenating microbial populations, and also affects adsorption equilibrium of PCBs with organic matter and thus influences their bioavailability (Jota and Hassett, 1991). Maximum dechlorination activity has been observed in the pH range between 5.0 and 8.0 (Wiegel and Wu, 2000). However, the selectivity of chlorine removal has been shown to be affected by pH. For example,

39

flanked meta dechlorination has been observed at pH 5.0-8.0, unflanked para

dechlorination at pH 6.0-8.0, and ortho dechlorination at pH 6.0-7.5 (Wiegel and Wu,

2000). Researchers have speculated that changes in dechlorination patterns are due to

alterations in the bacterial populations responsible for these reactions.

The availability of organic C also plays a major role in affecting reductive

dechlorination of PCBs (Nies and Vogel, 1990). During reductive dechlorination,

organic C compounds derived from algae, plants, and animals in the environment are

degraded to simpler compounds, such as acetate and H2 by fermenting bacteria such as

Clostridium spp., which are in turn utilized as electron donors by bacteria that reductively

dechlorinatie PCBs. If the concentrations of acetate and H2 are low due to low amounts of organic matter or consumption of these electron donors by methanogens, sulfate reducers, and acetogens (Smidt and de Vos, 2004; Chen et al., 2006), then rates of reductive dechlorination will be decreased (Nies and Vogel, 1990; Wiegel and Wu, 2000).

The presence of electron acceptors with reduction potentials greater than 0.5-0.6

V, such as O2 and NO3, are expected to inhibit reductive dechlorination of PCBs because organisms that utilize these electron acceptors outcompete reductive dechlorinators for electron donors and other substances required for microbial growth (Chang et al., 2004;

Newell and Aziz, 2004). Zwiernik et al. (1998) showed that adding ferrous sulfate to

sediment microcosms promoted reductive dechlorination of PCBs, presumably by two

effects. First, by providing sulfate as electron acceptor that stimulates the growth of

sulfate reducers which in turn might be directly involved in reductive dechlorination.

Second, by providing ferrous iron that removes sulfide as insoluble iron sulfide, reducing

bioavailability of toxic anions. Once sulfate is depleted, the increased population of

40

sulfate reducers can turn to PCBs as electron acceptors and thus to reductive

dechlorination.

In this chapter, the specific objective is to determine the local chemical profiles,

the bacterial community composition, and the PCB loss rates at a specific location in the

Ohio River. Factors governing biological transformations of PCBs in the environment

will be determined, in particular, nutrient concentrations, electron donor and acceptor levels, as well as the presence of microorganisms capable of transforming PCBs.

Sediment redox status will be determined based on the porewater chemistry. It is expected that differences in geochemical conditions will significantly affect bacterial community composition and thus, PCB removal rates in the sediment profile. For example, reduced conditions deeper in the sediment will favor anaerobic bacterial groups and reductive dechlorination, while more oxidized conditions in the upper parts of the sediment profile will favor aerobic transformations of PCBs. Such information is needed to determine the potential for natural attenuation of PCBs, as well as determine the main factors that limit biological PCB removal from this aquatic ecosystem.

Materials and Methods

Site description, sample preparation, and collection

Sediments from the littoral zone of river mile 369.5 of the Ohio River

(38°37'14.34"N, 83°9'30.84"W) were collected using a petite ponar grab sampler.

Approximately 10 L of sediment was collected up to a depth of 30 cm, and was

homogenized in a plastic container. 400 mL of sediment were dispensed into three

separate 1-L wide-mouth glass containers with Teflon-coated lids. Sediments in the

41

containers were spiked with 1 mL of a PCB mixture prepared in acetone containing 4000

µg mL-1 of each of the following PCBs: 3,3’-dichlorobiphenyl (PCB-11); 4,4’-

dichlorobiphenyl (PCB-15); 2,2’,5,5’-tetrachlorobiphenyl (PCB-52); 2,2’,3,3’-

tetrachlorobiphenyl (PCB-40); 2,3,4,5-tetrachlorobiphenyl (PCB-61); 2,2’4,4’,6,6’-

hexachlorobiphenyl (PCB-155); 3,3’,4,4’-tetrachlorobiphenyl (PCB-77); 2,2’,4,4’,5,5’-

hexachlorobiphenyl (PCB-153); 2,2’,3,3’,4,4’,5,5’-octachlorobiphenyl (PCB-194); and decachlorobiphenyl (PCB-209, 2000 µg mL-1). These congeners were selected to represent a wide range of chlorine substitution patterns that should provide information about the potential dechlorination activities present in these sediments. Except for PCB-

209, the final concentration of each PCB in the sediment was 10 mg kg-1 on a dry

sediment basis. Sediments with PCBs were poured into 35-8 mL cells of a porewater

equilibrator that was then covered with a 0.45 µm pore size membrane and a plastic mesh

fabric to allow equilibration of porewater constituents in the sediments, but not to allow

exchange of sediment with the surrounding environment. Three equilibrators were driven

into the river sediment to cover a depth of 35 cm from the surface and were equilibrated

for four months. At the end of this period, equilibrators were removed, placed in an ice

chest, and stored in the laboratory at -20 ºC until further analysis. While frozen, the

contents of each cell in the chamber were transferred to separate glass serum vials and

closed with Teflon-coated rubber septa with aluminum crimp. Once the vials were

equilibrated to room temperature, the headspace was analyzed by gas chromatography for

methane and carbon dioxide. After centrifugation at 2500 rpm for 15 min, the supernatant

was removed and saved for analysis of nutrients, metals, and dissolved organic carbon.

The sediment was saved for PCB and DNA extraction and analysis.

42

PCB extraction

Approximately 2 g of wet sediment was mixed with 10 g of anhydrous sodium sulfate (to remove excess moisture) and 25 mL of 5:1 hexane:acetone mixture in a wide-

mouth 60-mL extraction flask for 6 h on a horizontal shaker. The extracted PCBs were

separated from solids by centrifugation at 2500 rpm for 15 min, and 1 mL of the crude

extract was cleaned up with HNO3- and acetone-washed copper powder (1 g) to remove

elemental sulfur that was present at elevated levels and interfered with PCB analysis by

producing false peaks when using gas chromatography with electron capture detector.

DNA extraction and PCR amplification, cloning, and sequencing

Total community DNA was extracted from sediments (0.5 g) with an UltraClean

Soil DNA Isolation Kit (MoBio, Solana Beach, CA) using the maximum yield protocol

according to the manufacturer’s instructions. This extraction protocol includes a

combination of chemical detergent and bead beating steps to lyse cells, and was found to

be effective for evaluating bacterial and fungal diversity in soils, sediments, and other

environmental samples (Gomes et al., 2003; Hackl et al., 2004; Fierer et al., 2005). The

presence of high molecular weight DNA was verified by comparison with molecular

weight standards using gel electrophoresis with 0.8% agarose and 1× Tris-acetate-EDTA

buffer stained with ethidium bromide dye. The DNA was stored at -20°C until

polymerase chain reaction (PCR) amplification.

Separate PCR reactions were conducted for preparation of clone libraries and

DGGE profile analysis of 16S rDNA genes in the sediment samples. For clone libraries,

a 918-bp fragment of the bacterial 16S rDNA gene was amplified by PCR using 27F

43

forward primer 5'-AGA GTT TGA TC(A/C) TGG CTC AG-3' and 907R reverse primer

5'-CCG TCA ATT C(A/C)T TT(A/G) GTT T-3' (Lane, 1991; Muyzer et al., 1993). The

PCR reaction mixture consisted of 6.5 µL distilled water, 12.5 µL 2× IQ Supermix (Bio-

Rad, Hercules, CA), 2.5 µL 2 µM forward and reverse primers, and 1 µL of five times

diluted DNA sample. Each PCR reaction had the following final concetrations: 50 mM

KCl, 20 mM Tris-HCl, pH 8.4, 0.2 mM each dNTP (dATP, dCTP, dGTP, dTTP), 25

U/ml iTaq DNA polymerase, and 3 mM MgCl2. DNA extracts were diluted to reduce

anomalies such as chimeras, heteroduplex product or mutations that can occur during

PCR amplification (Qiu et al., 2001). The PCR was carried out in a MyCycler thermocycler (BioRad, Hercules, CA) with the following temperature program: initial

enzyme activation and denaturation of 5 min at 94ºC, followed by 30 cycles of 1 min at

94ºC, 1 min at 60ºC, and 2 min of extension at 72ºC, and a final extension of 10 min at

72ºC. Thirty cycles was the mininum number required to give faint bands of the expected

size as determined by comparison with molecular weight standards using gel

electrophoresis with 1% agarose and 1× Tris-acetate-EDTA buffer stained with ethidium bromide dye.

For DGGE analysis, a 560-bp fragment of the bacterial 16S rDNA gene was amplified by PCR using 341F-GC forward primer (clamp sequence in bold) 5'-CGC

CCG CCG CGC CCC GCG CCC GTC CCG CCG CCC CCG CCC GCC TAC

GGG AGG CAG CAG-3' and 907R reverse primer 5'-CCG TCA ATT C(A/C)T TT(A/G)

GTT T-3' (Muyzer et al., 1993; Muyzer and Smalla, 1998). The same PCR reaction conditions were used as described above except that a different temperature program was used: initial enzyme activation and denaturation of 3 min at 94ºC, followed by 30 cycles

44

of 30 s at 94ºC, 60 s at 50ºC, and 90 s of extension at 72ºC, with a final extension of 10

min at 72ºC.

Based on the sediment pore water chemistry data, four redox zones were

identified at 1-9 cm, 10-18 cm, 19-26 cm, and 27-35 cm depth in the sediment profile.

PCR products from these zones were combined and purified using Promega Wizard SV

Gel and PCR Clean-Up System (Promega, Madison, WI). The purified products were

ligated into Invitrogen pCRII-TOPO cloning vector and transformed into chemically- competent Escherichia coli TOP10F' cells (Invitrogen, Carlsbad, CA). The resulting clones were isolated on Luria Bertani agar plates with 50 µg mL-1 kanamycin at 37ºC according to the manufacturer’s instructions. Ninety-four randomly selected white colonies per sample were grown overnight in 2-mL 96-well microplates containing 1.5 mL Luria Broth media and 50 µg mL-1 kanamycin at 37ºC on an orbital shaker at 300

rpm.

Plasmids from clones were isolated in 96-well blocks with a Perfectprep Plasmid

96 Vac Direct Bind Kit (Eppendorf, Westbury, NY) and Biomek FX Laboratory

Automated Workstation (Beckman Coulter, Fullerton, CA). Nucleotide sequencing

reactions were performed with the vector M13F primer and ABI PRISM BigDye v3.1

Terminator Cycle Sequencing Kit (Applied Biosystems, Foster City, CA), and resulting

sequences were read with an ABI PRISM 3730 DNA analyzer at the University of

Kentucky Advanced Genetic Technology Center (Lexington, KY).

45

Denaturating Gradient Gel Electrophoresis analysis

The DGGE was carried out on the PCR amplicons according to the methods described by Muyzer et al. (1993) and Muyzer (1999). Briefly, the PCR amplicons (25

μl) were separated on 16.5×16.5 cm, 0.7-mm-thick 6% polyacrylamide (37.5:1 [w/v] acrylamide:bisacrylamide) gel with urea/formamide denaturing gradient. The polyacrylamide gel was made with a denaturing gradient ranging from 40 to 60% (where

100% denaturant contained 7M urea and 40% [v/v] formamide), and the electrophoresis was carried out at 60 V for 17.5 h at 60 ºC in 1×TAE buffer (20 mM Tris, 10 mM acetate,

0.5 mM EDTA, pH 7.4) in a BioRad-DCODE electrophoresis system. The gels were stained for 30 min in GelStar® nucleic acid stain (Lonza, Rockland, ME) and visualized in a GelDoc 2000 imaging system (BioRad, Hercules, CA).

Analytical methods

Sediment pH was measured with calibrated pH electrode and pH meter model

6071 (Jenco Electronics, LTD, Shanghai). Carbon dioxide in the headspace of vials with sediments was determined using a Shimadzu GC-8A gas chromatograph equipped with thermal conductivity detector (30 ºC), with helium as carrier gas and stainless steel

Porapak N column (0.3 cm by 2 m) (Supelco, Bellefonte, PA) maintained isothermally at

25 ºC. Methane in the headspace of vials with sediments was determined using a

Shimadzu GC-14A gas chromatograph equipped with flame ionization detector (FID) at

160 ºC, with nitrogen as carrier gas and a stainless steel, 45/60 mesh-packing, Carboxen

1000 column (0.3 cm by 2 m) (Supelco, Bellefonte, PA) maintained isothermally at 110

ºC.

46

Several porewater constituents in sediments were determined after centrifuging sediments and filtering with a 0.45 µm membrane, including metals (Fe, Mn, Al, Ca, Mg) dissolved organic C, sulfate, nitrate+nitrite, ammonium, and phosphate. One mL of supernatant designated for metal cation analysis was acidified with 200 µL of 65% nitric acid, diluted 10 times with deionized and distilled water, and analyzed using a Thermo

Jarrell Ash Model 61 Inductively Coupled Argon Plasma (Franklin, MA) at the

University of Kentucky Soil Testing Lab. One mL supernatant for dissolved organic C analysis was acidified with 200 µL of 36% hydrochloric acid, diluted five times with deionized and distilled water, and analyzed with a Shimadzu TOC-5000A total organic carbon analyzer (Columbia, MD). Two hundred microliters of supernatant designated for phosphate analyses was mixed with 40 µL of 1.75% (w/v) ammonium heptamolybdate solution in 6.3 N sulfuric acid in a disposable microplate for five min on a microplate mixer and mixed again with 40 µL of 0.035% of malachite green carbinol hydrochloride in 0.35% aqueous polyvinyl alcohol. Color development in the wells was determined by measuring absorbance at 630 nm using a µQuant MQX200 microplate spectrophotometer

(Bio-Tek Instruments, Winooski, VE) (D’Angelo et al., 2001). One hundred and twenty microliters of supernatant designated for ammonium analysis was mixed with 50 µL of

0.04% sodium nitroprusside in 2.26% aqueous phenol and 50 µL of 0.084% sodium hypochlorite in 1 % sodium hydroxide in a disposable microplate for 30 min on a microplate mixer. Color development in the wells was determined by measuring absorbance at 630 nm using a microplate reader. Twenty microliters of supernatant for nitrate+nitrite analysis was mixed with 200 µL of ammonium chloride buffer (pH 8.5) in the presence of an activated cadmium wire brush in a disposable microplate for 45 min

47

on a microplate mixer. After nitrate was reduced to nitrite, samples were mixed with 60

µL of Griess reagent (0.5 % sulfanilamide, and 0.05% N-(1-naphtyl)ethylendiamine dihydrochloride in 1.5 M hydrochloric acid) for 5 min, and the absorbance was determined at 542 nm using a microplate reader. Sulfate in the supernatant was analyzed by ion chromatography with a 1.8 mM carbonate/1.7 mM bicarbonate mobile phase using a Shimadzu HPLC equipped with an IonPac AS4A guard column (4x50 mm) and analytical column (4x250 mm) (Dionex Corp. Sunnyvale, CA), and Model 335 solid phase chemical suppressor module from Alltec (Deerfield, IL).

PCBs in cleaned up extracts were analyzed using a Shimadzu GC-14A gas

chromatograph equipped with 63Ni electron capture detector (320°C), autoinjector

(250°C), and RTX-1 capillary column (30 m by 0.32 mm with 0.25-µm phase thickness)

(Restek Corp., Bellefonte, PA) with the following temperature program: 100°C held for 1

min, ramp to 240°C at 3 ºC min-1, and held at 240°C for 10 min. PCB congeners were

identified by comparing retention times with authentic standards obtained from

Accustandard (New Haven, CT).

Phylogenetic analysis of clone library data

Vector contamination on the sequences obtained from the AGTC was detected

with the VecScreen application on the NCBI website

(http://www.ncbi.nlm.nih.gov/VecScreen/VecScreen.html). Vector contamination

removal, and sequence editing was performed with MEGA4 software (Tamura et al.,

2007). Potential chimeric sequences were detected using the Bellerophon Server (Huber

et al., 2004; http://foo.maths.uq.edu.au/~huber/bellerophon.pl), and were removed from

48

the clone libraries. Operational Taxonomic Units (OTUs) for construction of rarefaction curves were assigned to each sequence using DOTUR (Schloss and Handelsman, 2005; http://www.plantpath.wisc.edu/fac/joh/dotur.html) based on the distance matrix of aligned sequences for each clone library. Taxonomic hierarchy classification, sequence match with nearest neighbors, and clone library comparison for significant differences were performed with the tools available at the Ribosomal Database Project II

(http://rdp.cme.msu.edu/).

Statistical analysis of DGGE and sediment chemistry data

Microbial community data from DGGE analysis was analyzed by non-parametric multivariate analysis of variance (perMANOVA) using PCORD 5, a software package designed for multivariate analysis of ecological data (Anderson, 2001; McCune and

Grace, 2002). This is a distance matrix method that determines a central location for all the observations and central locations for each group of observations to be compared. If different groups have significantly different central locations in multivariate space, then the among-group distances will be relatively large compared to the within group distances, and the resulting test-statistic will be relatively large. The main advantage of perManova is that it does not make assumptions about the distribution of the data that often confound analysis of ecological data by many other multivariate methods (e.g. principal component analysis) (Anderson, 2001).

Four redox zones were identified in the sediment profile, based on porewater chemistry, and one-way analysis of variance for each parameter was obtained using

49

SAS® 9.1 (SAS Institute Inc. Cary, NC). Means for each zone were compared using the

Tukey-Kramer test for multiple comparisons with a 0.05 significance level.

Results and Discussion

Surface water temperature and sediment porewater chemistry

Surface water temperature in the vicinity of the sampling site during the

incubation period ranged from 25 ºC in August 1st to 8 ºC in No/vember (Fig. 2.1)

(USACE, 2006). Wu et al. (1996) determined that dechlorination activity was most rapid

in sediments form the Woods Pond in the Housatonic River between 18-34 ºC. In the

Ohio River, average temperatures in this range only occur between June and September.

Therefore, it is likely that reductive dechlorination would be restricted during most times

of the year due to low temperature conditions in the Ohio River.

The pH in the sediment profile ranged between 6.8 and 7.1, and did not show a strong gradient with sediment depth (Fig. 2.2A). Wiegel and Wu (2000) indicated that reductive dechlorination of PCBs can occur between pH 5.0 and 8.0, with an optimum at

7.0-7.5. Because pH of Ohio River sediments was in this range, it is unlikely that pH would affect this process in these sediments.

Ammonium concentrations ranged between 3.0 and 92 µM, and generally increased with depth particularly in the 25-30 cm interval (Fig. 2.2B). However, even in the lower depths, ammonium levels in sediments were considerably below the 5.5 mmol

L-1 required for optimal growth of bacteria such as Pseudomonas putida (Annuar et al.,

2006). Therefore, it is possible that nitrogen availability might not be high enough to

support PCB transforming populations in the sediments.

50

30

25

20 18 ºC

15

10

5 Temperature (ºC) 0

-5

-10

-15 Jan Feb Mar Apr May Jun Jul Aug Sep Oct Nov Dec Month

Figure 2.1 Average temperatures in the Ohio River at Lloyd Greenup Lock during year

2006 (USACE, 2006)

51

pH A Ammonium B Phosphate C DOC D Nitrate+Nitrite E (µmol/L) (µmol/L) (µmol/L) (µmol/L)

6.6 6.8 7.0 7.2 0 20 40 60 80 100 0.0 1.0 2.0 3.0 4.0 0 2000 4000 6000 0 30 60 90 120 0 0 0 0 0

5 5 5 5 5

10 10 10 10 10

15 15 15 15 15

20 20 20 20 20 Depth (cm) Depth (cm) Depth (cm) Depth (cm) Depth (cm)

25 25 25 25 25

30 30 30 30 30

35 35 35 35 35 Iron F Manganese G Sulfate H Methane I Carbon Dioxide J (µmol/L) (µmol/L) (µmol/L) (µmol/L) (µmol/L) 52 0 30 60 90 120 150 0 20 40 60 80 100 0 1000 2000 0 50 100 150 200 0 2000 4000 6000

0 0 0 0 0

5 5 5 5 5

10 10 10 10 10

15 15 15 15 15

20 20 20 20 20 Depth (cm) Depth (cm) Depth (cm) Depth (cm) Depth (cm)

25 25 25 25 25

30 30 30 30 30

35 35 35 35 35

Figure 2.2 Concentrations of dissolved porewater constituents in the Ohio River sediment profile (Dotted lines indicate separation

between redox zones identified in this study).

Dissolved phosphate concentrations were lower than 4.0 µM, and did not show a

consistent trend with depth except for elevated levels at the 6 cm, 15 cm, and 24 cm

depths (Fig. 2.2C). Dissolved phosphate levels in sediments were comparable with the

Michaelis-Menten constants for phosphate, which range between 0.01 and 0.24 µM for

many bacteria (Vadstein and Olsen, 1989). However, the P:C ratio was lower than 6 µg P

(mg C)-1, thus, it is possible that concentrations of dissolved phosphate limited bacterial growth in the sediments (Vadstein and Olsen, 1989).

Dissolved organic C concentrations ranged between 1500 and 5300 µM, and tended to increase below the 25 cm sediment depth (Fig. 2.2D). High amounts of dissolved organic C and ammonium in the lower sediment depths are likely attributable to anaerobic decomposition of organic matter in these layers. This was supported by an increase of methane from 26.8 to 183 µM in the lower parts of the sediment profile.

Dissolved organic C levels were less than the 0.01% that was reported to improve the reductive dechlorination of PCBs (Klasson et al., 1996; Wiegel and Wu, 2000).

Therefore, organic carbon availability could be a limiting factor in these sediments. It is likely that organic amendments would improve the reductive dechlorination of PCBs in these sediments because additional carbon sources can stimulate microbial population growth, not only of PCB dechlorinators but also other syntropic populations, e.g. hydrogen producing bacteria.

Nitrate+nitrite concentrations ranged between 6.9 and 105 µM and were highest at the 10 cm and 17 cm depths (Fig. 2.2E). These results suggested that nitrification and denitrification were important reactions at these depths. It is unlikely that reductive dechlorination would be important in these zones of the sediment profile because the

53

energy yield associated with these electron acceptors is higher than that corresponding to

PCBs; therefore nitrate and nitrite will be consumed first.

Dissolved Fe levels ranged between 13.8 and 139 µM, and did not show a consistent trend with depth except for slightly higher levels at the 7 cm, 14 cm, and 34 cm depths (Fig. 2.2F). It is likely that the dissolved Fe measured in this study reflected amounts of Fe2+ produced from the reduction of Fe3+ oxyhydroxides that commonly occur in water-saturated environments with circumneutral pH. Low levels of Fe2+ at other depths may be due to precipitation as iron carbonate and sulfide minerals. Dissolved Mn levels ranged between 33.6 and 86.2 µM (Fig. 2.2G), and followed a similar depth trend as dissolved Fe. It is likely that dissolved Mn reflected amounts of Mn2+ produced from the reduction of Mn oxide minerals. The biological reduction of Fe3+ and Mn4+ would tend to inhibit the reductive dechlorination of PCBs at these depths as they have a similar energy yield compared to PCBs, but the limited solubility of PCBs might prevent their utilization in presence of iron and manganese.

Dissolved sulfate concentrations ranged between 200 and 2300 µM (6.4-73.6 mg

L-1), and tended to decrease below the 25 cm sediment depth (Fig. 2.2H). From 2003 to

2005, sulfate levels in surface water were 44-90 mg L-1 upstream and 33-130 mg L-1 downstream of the sampling site (ORSANCO, 2006). Sulfate levels were within these ranges in the upper 25 cm of the profile, but were lower deeper in the profile. These results indicated that sulfate was an important electron acceptor at lower sediment depths.

Sulfate reduction in the sediments was consistent with high levels of elemental sulfur in the crude PCB extracts that were detected by gas chromatography (data not shown).

54

Methane and carbon dioxide also accumulated to high levels in the lower

sediment depths (Fig. 2.2I and J). Methane concentrations were between 26.8 and 183

µM, while carbon dioxide was between 1978 and 5042 µM. Interestingly, methane averaged 38.6 µM in the first 22 cm of the profile while the average below this depth was

106 µM, which suggests that methanogenesis was an important microbial process in Ohio

River sediments.

Based on the porewater chemistry of the sediment profile, four redox zones were

defined as follows: Denitrifying zone, from 1 to 9 cm; Iron-reducing zone, from 10 to 18

cm; Sulfate-reducing zone, from 19 to 26 cm; and Methanogenic zone, from 27 to 35 cm.

Table 2.1 shows the means for each porewater constituent in zones defined above. It is

predicted that reductive dechlorination would be restricted in the upper 25 cm sediment

depths due to the prevalent amounts and reduction of alternate electron acceptors (O2,

nitrate+nitrite, Fe3+, Mn4+, and sulfate) and low amounts of dissolved organic C (Wiegel

and Wu, 2000). Below this depth, however, sediments contained higher amounts of

ammonium, dissolved organic C, and methane, which indicated that nutrient, electron

donor, and redox conditions were more conducive to reductive dechlorination. However,

even at the lower sediment depths, reductive dechlorination may be restricted due to low

temperature conditions in the sediments.

55

Table 2.1 Tukey-Kramer Test for Multiple Comparisons of Porewater Chemistry

Parameters in Ohio River Sediments (All values in µM except for pH)

Parameter 1-9 cm 10-18 cm 19-26 cm 27-35 cm pH 6.94 6.95 7.02 6.96

Ammonium 21.5a 18.6a 29.8a 51.8b

Nitrite+Nitrate 14.2a 46.7b 19.2a,b 27.3a,b

Phosphate 0.67 0.39 0.64 0.23

Sulfate 1453a 1670a 1332a 561b

Iron 28.9 42.7 17.1 33.3

Manganese 50.0a,b 59.7a 39.6b 58.7a

Methane 37.8a 39.3a 54.1a 122.3b

Carbon Dioxide 2885a 2847a 3052a 4535b

Dissolved Organic Carbon 2532a 2391a 2004a 3552b

Note: Values with letters in common are not significantly different at α = 0.05

56

PCB losses from the sediment profile

Small amounts of PCB-61 (2,3,4,5-tetrachlorobiphenyl) (< 10%) were lost at all

depths in the sediments during the 4-month incubation period (Fig. 2.3). The appearance of a new peak on gas chromatograms at a retention time of 9.1 min was identified as

PCB-23 (2,3,5-trichlorobiphenyl). The production of PCB-23 suggested that PCB-61 was dechlorinated at the para position of the biphenyl ring, as shown in Figure 2.4.

Interestingly, other PCBs with chlorines at the para position were not dechlorinated (e.g.

4,4’-dichlorobiphenyl, 2,2’4,4’,6,6’-hexachlorobiphenyl, 3,3’,4,4’-tetrachlorobiphenyl,

2,2’,4,4’,5,5’-hexachlorobiphenyl, and 2,2’,3,3’,4,4’,5,5’-octachlorobiphenyl), which suggested that the chlorination of the adjacent aromatic ring inhibited reductive dechlorination of a PCB congener. Since PCB-61 was reductively dechlorinated at all depths, it appears that conditions in the sediments (e.g. temperature, dissolved organic C, electron acceptors, presence of reductive dechlorinating populations, etc) allowed for low rates of reductive dechlorination of this congener. On the other hand, the concentrations of other PCB congeners did not change at any depth during the incubation period, which suggested that in situ conditions were not favorable for reductive dechlorination of these congeners, as discussed in the previous section. Therefore, natural attenuation does not seem to be a viable alternative for removal of most PCBs from Ohio River sediments.

57

In situ reductive dechlorination of PCBs in Ohio River sediments.

ECD ECD ECD ECD ECD 10PCB 1 ppm ISP01 ISP11 ISP22 ISP34 153 194 61

600 PCB - PCB - 600 PCB - 77 52 40 209 500 155 500 PCB - PCB - PCB - PCB - PCB -

400 400 11 15 23 PCB - PCB - PCB -

300 300 mVolts mVolts

200 200 34 cm 22 cm 100 100 11 cm 1 cm

0 1 ppm std 0

7 8 9 10 11 12 13 14 15 16 17 18 19 20 Minutes

Figure 2.3 Gas chromatograms of PCB mixtures determined at different depths after four months incubation in the Ohio River sediment profile. Note the appearance of a new peak

at 9.1 min retention time, which represents PCB-23.

58

Cl Cl - + - + H2 + 2e + H + Cl Cl Cl Cl Cl Cl

Figure 2.4 Reductive dechlorination of PCB-61 to PCB-23

59

DGGE analysis of bacterial communities

The sizes of the genomic DNA extracted from sediments and PCR amplicons using the

341F-GC and 907R primers were in the expected size ranges of 14-15 kb and 620 bp

respectively, (Fig. 2.5), which indicated that methods used in the study were effective for

extracting and amplifying DNA from the samples.

The number of bands and banding patterns in the different DGGE lanes shown in

Fig. 2.6 represent the dominant types of bacterial populations in various redox layers of the sediment profile. PerMANOVA analysis of relative band intensities in the lanes did not reveal any significant differences in microbial community composition in the sediment profile at a p-value of 0.05. One disadvantage of DGGE analysis is that it is often not sensitive enough to show differences in microbial community composition from various habitats, particularly those with extremely high microbial diversity (Burr et al.,

2006). Therefore, microbial community composition of the sediments was further explored using the clone library approach.

Clone library analysis of bacterial communities

Four clone libraries were prepared for each redox zone by PCR amplification with

primers 27F/907R, with each library containing 53 to 71 chimera-free clones.

Differences in microbial richness in various redox zones were investigated using

rarefaction analysis, in which the number of Operational Taxonomic Units (OTUs) in the

sample is plotted as a function of the number of sequences evaluated (Fig. 2.7).

60

A

B

Figure 2.5 Agarose gel electrophoresis detection of genomic DNA extraction obtained from sediments collected at different depths in the Ohio River sediment profile (A), and

corresponding PCR amplification products using primers 341F-GC/907R for DGGE

analysis (B)

61

Figure 2.6 Denaturating gel gradient electrophoresis analysis of PCR-amplified 16S rDNA obtained from three replicates of four Ohio River sediment depth increments.

62

70 60 50 Species level (3% difference) 40 Denitrifying 30 Iron-reducing 20 Sulfate-reducing 10 Methanogenic Number of OTUs Observed 0 0 10 20 30 40 50 60 70 80 Number of Sequences Sampled 60

50 Family/Class level 40 (10% difference)

30

20

10

Number of OTUs Observed 0 0 10 20 30 40 50 60 70 80 Number of Sequences Sampled 35 30 25 Phylum level (20% difference) 20 15 10 5 Number of OTUs Observed 0 0 10 20 30 40 50 60 70 80 Number of Sequences Sampled

Figure 2.7 Rarefaction curves generated from clone libraries obtained from four Ohio

River sediment depth increments, as determined by the furthest neighbor assignment

algorithm (Schloss and Handelsman, 2005)

63

Rarefaction curves were prepared at the 3%, 10% and 20% difference levels, which corresponded to comparisons of richness at approximately the species, family/class, and phyla levels, respectively (Schloss and Handelsman, 2005). The highest number of OTUs observed at a given number of sequences were similar for all redox zones at the species and family/class levels (60 and 50, respectively). The number of OTUs at the phyla level was higher for the sulfate-reducing zone with a suggested phyla richness of 31 compared to 25-29 phyla in the other redox zones. Unfortunately, these curves do not approach a plateau, which is observed when there are not enough sequences to represent the population diversity. Therefore, rarefaction provides an underestimation of population richness when a small number of sequences is used for the analysis, and alternative richness indicators might be needed.

Differences in microbial community composition in various redox zones were investigated using the Classification and Library Compare tools available in the online

RDP software package (Table 2.2). Bacteria from 20 phyla were represented in clone libraries from all redox zones. However, only 10 or 11 phyla were found in clone libraries from different redox zones, which indicated that there were no significant differences in phyla richness in the different redox zones. These results generally supported the DGGE results. Greater than 90% of clones in the libraries prepared from different redox zones were classified into eight phyla, which generally followed the order

Proteobacteria (36-53%), Chloroflexi (11-17%), Bacteroidetes (5-21%), Firmicutes (6-

13%), (3-11%), (3-6%), Nitrospira (0-4%), and

Planctomycetes (0-3%).

64

Table 2.2 Distribution of bacteria in major taxonomic groups in libraries from Ohio River sediments during in situ bioremediation of

PCBs for four months, as determined by the Classifier tool available at the Ribosomal Database Project (RDP) (Cole et al., 2007).

Depth (cm)†

1-9 10-18 19-26 27-35 Phylum Class Order Representative genera in libraries (n=71) (n=53) (n=68) (n=67) ------%------Gp21 (11), Gp7 (2), Gp16 (3), Acidobacteria Acidobacteria Acidobacteriales 10.0 11.3 2.9 9.0 Gp18(2), Gp8(1), Gp23(1), Gp6(8), Gp17(3) Actinobacteria Actinobacteria 4.2 5.7 2.9 1.5

Coriobacteridae 0.0 0.0 0.0 1.5

65 Rubrobacteridae 2.8 1.9 0.0 1.5 Conexibacter, Patulibacter

Actinobacteridae 1.4 3.8 1.5 0.0

Acidimicrobidae 0.0 0.0 1.5 0.0

Bacteroidetes 9.9a 20.8b 5.9a 4.5a

Bacteroidetes Bacteroidales 5.6 3.8 0.0 1.5 Paludibacter

Flavobacteria Flavobacteriales 1.4 0.0 1.5 1.5 Flavobacterium

Haliscomenobacter (1), Sphingobacteria Sphingobacteriales 2.8a 17.0b 4.4a 1.5a Terrimonas (1), Niastella (1),Cardimium (2) BRC1 - - 0.0 0.0 1.5 1.5

Chlorobi Chlorobia Chlorobiales 0.0 0.0 1.5 1.5

Table 2.2. (continued)

Depth (cm)†

1-9 10-18 19-26 27-35 Phylum Class Order Representative genera in libraries (n=71) (n=53) (n=68) (n=67) ------%------Chloroflexi 14.0 11.3 16.2 13.4

Anaerolineae Anaerolineales 16.9 11.3 10.3 13.4 Anaerolinea (1), Caldilinea (1), Levilinea (7) Chloroflexi Chloroflexales 0.0 0.0 5.9 0.0 Roseiflexus (1) Chrysiogenetes Chrysiogenetes Chrysiogenales 1.4 0.0 0.0 0.0

66 Deferribacteres Deferribacteres Deferribacterales 0.0 1.9 0.0 0.0

Firmicutes 8.5 5.7 10.3 13.4

Bacilli Bacillales 2.8 0.0 0.0 0.0 Bacillus (2) 5.6 5.7 10.3 13.4

Clostridiales 0.0a 3.8a 5.9a 10.4b Fusibacter (1), Gracilibacter (1), Thermoanaero- 5.6 1.9 4.4 3.0 , bacteriales Gemmatimonadetes Gemmatimonadales 0.0 0.0 1.5 0.0

Lentisphaerae Lentisphaerae Victivallales 0.0 1.9 0.0 0.0

Nitrospira Nitrospira Nitrospirales 4.2 1.9 1.5 1.5 Magnetobacterium (5)

Table 2.2. (continued)

Depth (cm)†

1-9 10-18 19-26 27-35 Phylum Class Order Representative genera in libraries (n=71) (n=53) (n=68) (n=67) ------%------OD1 - - 0.0 0.0 0.0 3.0

OP10 - - 0.0 0.0 1.5 0.0

Planctomycetes Planctomycetacia Planctomycetales 1.4 1.9 0.0 0.0 Planctomyces (1) Proteobacteria 35.0 35.8 52.9 42

67 α-proteobacteria 4.2ab 1.9b 11.8a 7.5a

Bradyrhizobium (3), Methylocystis Rhizobiales 2.8 1.9 8.8 6.0 (1), Starkeya (1), Ancalomicrobium (1) 0.0 0.0 1.5 1.5 Rhodobacter (1) Sphingomonadales 0.0 0.0 1.5 0.0 Sphingomonas (1) β-proteobacteria 19.7a 18.9a 29.4a 4.5b

Derxia (2), Hydrogenophaga (2), 14.0a 7.5a,b 17.6a 1.5b Rhodoferax (3), Aquabacterium (1),Eurybacter (1) Methylophilales 0.0 0.0 0.0 1.5

Nitrosomonadales 0.0 5.7 2.9 0.0 Gallionella (3)

Rhodocyclales 2.8 5.7 8.8 0.0 Ferribacterium (1) Table 2.2. (continued)

Depth (cm)†

1-9 10-18 19-26 27-35 Phylum Class Order Representative genera in libraries (n=71) (n=53) (n=68) (n=67) ------%------δ-proteobacteria 10.0a 7.5a 5.9a 18.0b

Desulfobacterales 0.0a 3.8a 0.0a 10.4b Desulfobacterium (2), Desulfocapsa (1) Desulfurellales 1.4 0.0 1.5 0.0

Desulfuromonales 4.2 3.8 0.0 3.0 Geobacter (2)

68 Myxococcales 2.8 0.0 0.0 0.0

Anaeromyxobacter (1) Syntropho-bacterales 2.8 0.0 4.4 9.0 Desulfobacca (3), Smithella (3) ε-proteobacteria Campylobacterales 0.0 0.0 1.5 0.0

γ-proteobacteria 1.4a 7.5b,c 4.4a,b 11.9c

Chromatiales 0.0 1.9 1.5 4.5

Legionellales 0.0 0.0 0.0 1.5

Methylococcales 1.4 1.9 0.0 0.0 Methylomonas (1), Methylobacter (1) Oceanospirillales 0.0 1.9 1.5 3.0

Table 2.2. (continued)

Depth (cm)†

1-9 10-18 19-26 27-35 Phylum Class Order Representative genera in libraries (n=71) (n=53) (n=68) (n=67)

------%------

Pseudomonadales 0.0 0.0 1.5 0.0

Thiotrichales 0.0 1.9 0.0 0.0

Xanthomonadales 0.0 0.0 0.0 1.5 69

Spirochaetes Spirochaetales 1.4 0.0 0.0 0.0

Thermomicrobia Thermomicrobia Thermomicrobiales 2.8 0.0 0.0 0.0

Verrucomicrobia Verrucomicrobiae Verrucomicrobiales 1.4 1.9 0.0 1.5

WS3 - - 0.0 0.0 1.5 3.0

† Values followed by a different letter are significantly different at p= 0.05; n represents number of sequences in clone library.

Many of the groups that were dominant in Ohio River sediments were also

prevalent in other river sediments. Using fluorescent in situ hybridization (FISH), for

example, Fazi et al. (2005) discovered that alpha- and beta-Proteobacteria were dominant in three Italian river sediments. Kloep et al. (2006) observed that alpha-, beta-, and gamma-Proteobacteria, Cytophaga-Flavobacteria, and Planctomycetales were dominant in sediments from the Elbe River in Germany, and Peplies et al. (2006) found that beta-

Proteobacteria, sulfate-reducers, and methanotrophs bacteria were prevalent in four other

German river sediments. Unfortunately, only a small fraction of the bacterial community could be detected by FISH in these studies (<40%), suggesting that more probes would be needed to fully characterize bacterial diversity in these systems (Peplies et al., 2006).

This might explain the differences in bacterial composition observed in the various studies. The present study is one of the first to characterize bacterial community composition using the clone library approach.

There were significant differences in the taxonomic makeup of the bacterial communities in the different redox zones (Table 2.2). Bacteria in the class β-

Proteobacteria made up the largest fraction of the clone libraries from the denitrifying zone (20%) and also considerable fractions of clone libraries in the iron-reducing zone

(19-29%). However, this group made up a smaller fraction of the clone library from the methanogenic zone (<5%). Most of the bacteria in this class were in the order

Burkholderiales, which mostly included bacteria with aerobic heterotrophic type metabolism such as Aquabacterium, Hydrogenophaga, Leptothrix, and Rhodoferax spp.

High levels of bacteria in this group suggested that surface sediments were relatively oxidized compared to sediments in the lower depths. Because reductive dechlorination

70

requires strongly reducing conditions, it is unlikely that this process would occur in the 0-

26 cm depths. On the other hand, presence of these aerobic bacteria suggests that there is

potential for aerobic degradation of PCBs. Although known aerobic PCB degraders were

not detected in the clone libraries, phylogenetic analysis suggests presence of related

species and will be discussed later.

The second most prevalent class of bacteria in the clone library from the 0-9 depth

increment was bacteria in the Anaerolinea within subdivision I of the phylum

Chloroflexi (17% of the clone library) (Table 2.2). There were no differences in the

numbers of Anaerolinea in clone libraries from other sediment depths. Bacteria in this

genus are obligately anaerobic syntrophs that ferment and amino acids

primarily to acetate and H2 (Sekiguchi et al., 2003). Interestingly, Anaerolinea spp. have

been commonly found in many anaerobic dechlorinating environments, and probably

support the reaction in these environments by supplying electron donors, such as

hydrogen to dechlorinating populations (Yamada et al., 2006). High numbers of both

Anaerolinea and aerobic bacteria in the class β-Proteobacteria suggested that anaerobic

and aerobic microsites were prevalent in the 0-26 cm depth. Because PCBs have been

released to the environment as complex mixtures of varying chlorine content, there is not

a single process that, by itself, could totally remove PCBs from the environment. Lower chlorinated congener can be readily degraded by aerobic bacteria, but for higher chlorinated congeners it is necessary to reduce the number of chlorines before they can be aerobically degraded, and it is likely that a sequential anaerobic aerobic treatment can completely remove PCBs as shown previously (Master et al., 2002). Therefore, existence

71

of a mixed aerobic-anaerobic environment in these sediments suggests the possibility of

having both processes for the complete removal of PCBs from Ohio River sediments.

The third most important class of bacteria in the clone library from the

denitrifying zone was δ-Proteobacteria (11% of the clone library composition) (Table

2.2). The number of bacteria in this class was similar in clone libraries from the iron-

reducing and sulfate-reducing zones (6-8%), but it was significantly lower than the clone

library from the methanogenic zone (22%). There were differences in the types of δ-

proteobacteria in the different depths. In the denitrifying zone, the dominant δ-

proteobacteria were in the order Desulfuromonales, and included bacteria in the genera

Geobacteria and Geopsychrobacter that have been associated with electricity harvesting

from sediments using graphite electrodes as their sole electron acceptors (Holmes et al.,

2004), and with oxidation of aromatic hydrocarbons under iron reducing conditions

(Rooney-Varga et al., 1999). More recently, Sung et al. (2006) showed that Geobacter

lovleyi can couple oxidation of organic substrates not only with metal reduction but also

with reduction of tetrachloroethene and triclorethene to dicloroethene. Therefore,

presence of these organisms may suggest favorable conditions for reductive

dechlorination reations.

In the methanogenic zone, the dominant δ-proteobacteria were in the orders

Desulfurobacterales and Syntrophobacterales. The dominant genera in

Desulfurobacterales order included sulfate-reducers such as Desulfobacterium and

Desulfocapsa spp. At this depth, about half of the clones were classified in the order

Desulfobacterales, which include sulfate-reducing organisms Desulfobacterium,

Desulfonema, and Desulfocapsa spp. The dominant genera in the Syntrophobacterales

72

order included strictly anaerobic propionate-degrading syntrophe Smithella (Liu et al.,

1999) and acetate degrading sulfate reducer Desulfobacca (Elferink et al., 1999).

Changes in the relative amounts of sulfate reducers in the sediment profile

supported the porewater data, which indicated that sulfate reduction was an important

process in the lower sediment depths. It is possible that large numbers of sulfate reducers

could inhibit reductive dechlorination of PCBs, by competing with reductive

dechlorinators for electron donors such as hydrogen, acetate or lactate (Madigan et al.,

2003).

The next most important class of bacteria in the clone library from the

denitrifying zone was Acidobacteria (10% of the clone library composition) (Table 2.2).

The number of bacteria in this class was not significantly different between redox zones.

This group of bacteria is of recent discovery and their physiological role is not well

understood (Quaiser et al., 2003). However, Eichorst et al. (2007) showed that members

of this phylum grow under mildly acidic conditions (pH 6-7), and in the presence of oxygen, but are unable to grow under anaerobic conditions using nitrate or iron as electron acceptors. Therefore, their presence suggests aerobic conditions in the sediment profile with low nutrient levels that may have a negative impact on the reductive dechlorination of PCBs in these environments.

In the phylum Bacteroidetes, the order Sphingobacteriales was abundant in the clone libraries. This group is highly diverse and includes anaerobic species such as the

Cytophaga species that are also found in marine environments and may be involved in the breakdown of complex organic matter in sediments (Rosello-Mora et al., 1999). The

presence of these microorganisms supports the establishment of a mixed aerobic-

73

anaerobic environment that does not offer favorable conditions for reductive

dechlorination.

The order Clostridiales was also higher in the bottom layer and this order includes

Desulfitobacterium and Dehalobacter spp. These are obligate anaerobes that have been isolated from environments contaminated with halogenated organic compounds

(Villemur et al., 2005). Desulfitobacterium spp. can use a wide range of electron acceptors for including nitrate, sulfate, metals, humic acids, and halogenated organic compounds either from natural or anthropogenic origin. Although the substrate specificity varies with the species, they can dehalogenate compounds such as trichloro- and tetrachloroethylene, and haloaromatics such as chlorophenols.

Phylogenetic analysis of the sequences in the clone libraries indicated that several bacteria in sediments were closely related to bacteria with reductive dechlorinating abilities (Fig. 2.8). Eleven clones present at all depths were related to Desulfitobacterium spp. Although bacteria in this genus are not known to dechlorinate PCBs, they are known to dechlorinate closely-related chlorinated aromatic compounds, including chlorophenols, bromophenols, hexachlorobenze, and even hydroxylated PCBs (Villemur et al., 2006).

Several clones were related to several other dehalogenating bacteria such as

Desulfuromonas spp, and Trichlorobacter thiogenens, which can dechlorinate solvents like tetrachloro- and trichloroethylene (Smidt and de Vos, 2004) and Desulfomonile tiedjei which can grow with chlorobenzoates as electron acceptor (Mohn and Tiedje,

1992). Additionally, six clones were closely related to Dehalococcoides spp., which are

the only known genera with the capacity to reductively dechlorinate PCBs (Field and

Sierra-

74

89 InSitu2 058 100 InSitu3 040 55 InSitu4 057 Dehalococcoides ethenogenens (AF004928) 96 100 Dehalococcoides sp. BAV1 (AY165308) 95 Dehalococcoides sp. CDBD1 (AF230641) InSitu1 028 53 42 InSitu3 086 100 InSitu3 012 35 InSitu3 018 58 InSitu3 041 InSitu3 033 100 25 InSitu3 066 InSitu3 005 29 InSitu1 071 91 InSitu2 070 InSitu3 035 34 InSitu4 032 84 InSitu4 081 39 InSitu1 056 100 InSitu1 082 InSitu1 093 100 Desulfitobacterium chlororespirans (U68528) 99 Desulfitobacterium hafniense Y51 (AP008230) 99 76 Dehalobacter restrictus (Y10164) InSitu4 012 100 InSitu4 017 64 100 InSitu1 014 61 100 InSitu1 079 InSitu1 086 96 54 InSitu1 008 InSitu3 090 84 InSitu2 013 51 InSitu3 075 Anaeromyxobacter dehalogenans 2CP-1 (AF382396) 57 InSitu1 081 97 Desulfomonile tiedjei (AM086646) 36 Trichlorobacter thiogenens (AF223382) 100 Desulfuromonas chloroethenica (U49748) 100 Desulfuromonas michiganensis BB1 (AF357915) 100 Desulfuromonas michiganensis BRS1 (AF357914) Dehalospirillum multivorans (X82931) 60 dechloracetivorans (AF230530)

0.02

Figure 2.8 Bootstrapped neighbor-joining phylogenetic tree of 16S rDNA sequences obtained from the in situ reductive dechlorination study and known dehalogenating bacteria. Sequences obtained in the present study are indicated as follows:

InSitu1=Denitrifying zone, InSitu2=Iron-reducing zone, InSitu3=Sulfate-reducing zone, and InSitu4=Methanogenic zone. Numbers at branch points are percentages of 500 resampling bootstrap analysis. The scale bar represents 0.02 changes per nucleotide.

75

Alvarez, 2008; Yan et al., 2006). The presence of this organism suggests that reductive

dechlorination may be possible in Ohio River sediments under suitable biogeochemical

conditions.

Several other clones found in the river sediments showed a closer relationship

with aerobic PCB degraders such as Burkholderia xenovorans, Comamonas testosteroni,

and Pseudomonas spp., with differences less than 5% in their sequences (Fig. 2.9) and the

presence of bacterial groups with wide O2 tolerances in the sediments, ranging from

obligately aerobic to strictly anaerobic, suggested that the sediment profile contained a

strong redox gradient. Such a gradient would tend to support sequential reduction and

oxidation of PCBs, although there was no evidence of this during this experiment.

No extensive dechlorination of PCBs was observed in the field study, suggesting

that a different set of conditions, besides redox potential, controlled the extent of the

reaction. For instance, availability of other electron acceptors such as oxygen, nitrate, iron or manganese may limit the reductive dechlorination of PCBs. In a dynamic system like the river, constant perturbation and mixing of the sediments promotes recycling of redox sensitive elements (Meade and D’Angelo, 2005) that may inhibit the reductive dechlorination. Temperature may have a big impact on the rate, as at temperatures below

20 ºC a decrease in dechlorination is observed (Wu et al., 1996; Palekar et al., 2003), and the average temperature in the river during the experiment ranged from 23 ºC in August

to 7 ºC in November when the equilibration chambers were retrieved from the field.

Based on this fact it is more likely that the small rate of dechlorination observed for PCB-

61, occurred only during the first two months of incubation, when the temperatures were

between 19 and 23 ºC.

76

98 InSitu2 050 InSitu2 052 62 InSitu2 046 59 InSitu2 045 93 99 InSitu2 063 Burkholderia sp. JB1 (X92188) 99 Burkholderia xenovorans LB400 (U86373) 99 InSitu1 074 InSitu1 080 85 97 Comamonas testosteroni (AM937260) 99 InSitu1 023 99 InSitu1 066 InSitu4 092 InSitu2 081 91 Pseudomonas pseudoalcaligenes (AB257323) 64 Pseudomonas putida (D87108) 89 Pseudomonas pseudoalcaligenes (DQ777729) 90 InSitu3 027 88 Pseudomonas sp. SA-6 (DQ854841) 99 strain SA-1 (DQ854840) Pseudomonas sp. JN18 A60 A4 (DQ168654) 97 Pseudomonas sp. JN18 A17 R (DQ168645) InSitu4 012 InSitu2 082 74 InSitu2 086 InSitu1 011 99 InSitu4 074 InSitu3 002 95 InSitu3 083 97 Sphingomonas sp. R1 (DQ858957) 91 Sphingomonas aromaticivorans SMCC F199 (U20756) 97 InSitu2 059 99 InSitu2 060 70 InSitu3 019 Bacillus sp. OUCZ63 (AY785743) 52 InSitu2 041 InSitu2 051 InSitu4 056 InSitu3 080 InSitu4 055 InSitu4 048 InSitu1 027 74 InSitu2 033 99 InSitu3 061 InSitu3 048 64 InSitu1 035 98 Rhodococcus sp. OUCZ204 (AY785749) 99 Rhodococcus ruber strain Z57 (DQ858963) 71 Rhodococcus erythropolis strain D14 (DQ858961)

0.02

Figure 2.9 Bootstrapped neighbor-joining phylogenetic tree of 16S rDNA sequences

obtained from the in situ reductive dechlorination study and known aerobic PCB

degraders. Sequences obtained in the present study are indicated as follows:

InSitu1=Denitrifying zone, InSitu2=Iron-reducing zone, InSitu3=Sulfate-reducing zone,

and InSitu4=Methanogenic zone. Numbers at branch points are percentages of 500

resampling bootstrap analysis. The scale bar represents 0.02 changes per nucleotide.

77

In conclusion, porewater chemistry revealed a rather oxidized environment in the

surface sediments at the moment of sampling. Below 20 cm, microbial activities were

dominated by sulfate reducing processes as well as methanogenesis, which are favorable

conditions for reductive dechlorination. Microbial community composition revealed the

presence of microorganisms closely related to known dehalogenating bacteria, such as

Desulfitobacterium spp., and especially Dehalococcoides spp. which have been shown to

dechlorinate PCBs. However, PCB dechlorination was not extensive as only one single product, PCB-23, was detected at the end of the experiment. Other nutrients such as phosphate and organic carbon were limited, and might have had a negative impact on rates and extent of reductive dechlorination. These limitations along with cold temperatures during part of the experiment might explain the observed low conversion of

PCBs into lower chlorinated congeners, and suggest that additional treatments are required to improve natural attenuation of PCBs in this environment. Therefore, it is expected that controls on temperature and nutrient availability might improve the reductive dechlorination of PCBs in Ohio River sediments, which will be discussed in the following chapter.

78

CHAPTER 3

TEMPERATURE AND ELECTRON DONOR AVAILABILITY EFFECTS ON

PCB REDUCTIVE DECHLORINATION RATES AND MICROBIAL

COMMUNITY COMPOSITION IN OHIO RIVER SEDIMENTS.

Introduction

Polychlorinated biphenyls (PCBs) refer to a family of compounds that were manufactured for many decades as coolants in transformers, capacitors, and other electrical equipment. Millions of tons of PCBs have been dispersed globally in the environment, which have been found to accumulate in aquatic sediments and the food chain due to their hydrophobic and stable chemical characteristics. The main PCB exposure route to humans is consumption of sportfish caught from contaminated environments. Overconsumption of PCB-contaminated fish tissue can cause learning and memory losses, reproductive and immunity deficiencies, and skin and liver damage

(Robertson and Hanson, 2001; ATSDR, 2001)

In many aquatic environaments, PCBs can be transformed through the concerted activity of anaerobic and aerobic bacteria that inhabit contaminated sediments. In anaerobic zones, bacteria can replace chlorine atoms of highly substituted PCBs with hydrogen atoms through a process known as reductive dechlorination (Brown et al.,

1987; Quensen et al., 1988; Bedard and Quensen, 1995; Holliger et al., 1999; Wiegel and

Wu, 2000; Smidt and de Vos, 2004). In aerobic zones, less chlorinated PCB congeners can be degraded by many bacteria, such as Alcaligenes, Burkholderia, Pseudomonas,

Bacillus, Corynebacterium, and Rhodococcus (Williams et al., 1997; Maltseva et al.,

79

1999). In aquatic systems with established oxidized and reduced zones, reductive

dechlorination and aerobic catabolism can work in tandem to mineralize PCBs to CO2,

H2O and chloride ions.

Reductive dechlorination of PCBs has been documented in many freshwater,

estuarine, and marine environments; little is known, however, about the identity,

physiology, or ecology of dechlorinating communities. Ye et al. (1992) concluded that

anaerobic spore-forming bacteria dechlorinated PCBs in Hudson River sediments based

on results with pasteurized samples. Based on results with specific microbial inhibitors,

Williams (1997) concluded that anaerobic Gram-positive bacteria dechlorinated PCBs in

Hudson River sediments. Hou and Dutta (2000) concluded that Clostridium species

dechlorinated PCBs in Lake Medinah sediments based on phylogenetic analysis of

anaerobic enrichment cultures. Using a similar approach, Cutter et al. (2001) and Wu et

al. (2002) concluded that Chloroflexi-related strains o-17 (Genbank accession no.

AF294958) and Dehalobium chlorocoercia DF-1 (Genbank accession no. AF393781) dechlorinated PCBs in Baltimore Harbor sediments. To date, Dehalococcoides ethenogenens strain 195 is the only strain available in pure culture with PCB-

dechlorinating activity (Fennel et al., 2004; Seshadri et al., 2005).

Reductive dechlorination in most anaerobic environments is a syntrophic process

between dechlorinating bacteria and many non-dechlorinating bacterial populations.

Dechlorinating bacteria typically use simple electron donors such as H2 and acetate that are provided during oxidation of more complex substrates by fermenting and syntrophic bacteria (Drzyzgal and Gottschal, 2002; Becker et al., 2005). Conversely, it has been known for many years that hydrogenotrophic bacteria (e.g. methanogens and sulfate-

80

reducers) lower H2 to levels that allow fermentation of organic substrates to be

thermodynamically favorable (Fukuzaki et al., 1990). In addition, transcription and

activity of dehalogenase enzymes are believed to be strongly regulated by the presence of

electron acceptors including O2 and sulfur oxyanion species (Townsend and Suflita,

1997; Pop et al. 2004), the concentrations of which are largely controlled by aerobic and anaerobic bacterial groups. Thus, it is important to understand the environmental factors that regulate many bacterial groups involved in these interactions in order to optimize conditions for reductive dechlorination in natural habitats.

Several physical and chemical conditions can have major effects on bacterial groups in anaerobic environments. In methanogenic paddy soils, for example, thermal regime profoundly influenced fermentative and syntrophic production of H2 and low

molecular weight fatty acids, which in turn affected methanogenesis rates in the system

(Chin and Conrad, 1995). It is expected that temperature would also play an important

role in controlling reductive dechlorinators. In addition, Wu et al. (1996; 1997)

hypothesized that temperature influences on PCB dechlorinating populations caused

changes in dechlorination rates and patterns in Woods Pond sediments. This could be

important in the Ohio River as temperatures ranged widely between 5 and 35 ºC during

the year (USACE, 2006).

The types of electron acceptors available in the system can have positive or

negative effects on PCB dechlorination. Ferrous sulfate amendments to Hudson River

sediments, for example, stimulated PCB dechlorination presumably by enriching for

sulfate-reducing, PCB-dechlorinating bacteria and by removing toxic sulfide species as

iron sulfide precipitates (Zwiernik et al., 1998). In another study, nitrate and sulfate

81

inhibited Aroclor 1242 dechlorination in Hudson River sediments, presumably by acting

as preferred electron acceptors for some members of the dechlorinating community (Rhee

et al., 1993). Results from the previous chapter suggest that several electron acceptors,

such as nitrate, Fe+3, Mn+4 and sulfate, could be important controllers of reductive

dechlorination in Ohio River sediments. While environmental conditions are likely to

play a large role in regulating bacterial community structure and processes, few studies

have investigated these effects in anaerobic sediments.

The objectives of this study were to: (i) characterize the bacterial composition of

river sediments using the clone library approach, (ii) determine the effects of anaerobic

treatment conditions (e.g. temperature, electron donors, and FeSO4) on bacterial

community composition in PCB-contaminated sediments, and (iii) evaluate treatment effects on removal of 2,4,5-2’,4’,5’-hexachlorobiphenyl (PCB-153), which is one of the dominant PCB congeners in many sediments including those from the Ohio River. It was hypothesized that treatments would elicit major changes in key bacterial populations that are either directly or indirectly involved in PCB dechlorination. Association between bacterial taxa and treatment conditions identified in this study could be useful for assessing in situ geochemical conditions that regulate anaerobic PCB removal dechlorination in these types of environments.

82

Materials and Methods

Site description and sample collection

The Ohio River, a 1600-km river located in the east-central US, is extensively contaminated with PCBs. Sediments from the river contain between 0.01 and 8.5 ppm

PCBs and many fish species contain >2 ppm PCBs. As a result, the river is under several fish consumption advisories due to PCB contamination (ORSANCO, 2002; USEPA,

2004). PCBs in the river mainly consist of penta- and hexa-chlorinated biphenyls, particularly PCB-153 and 2,3,5-2’,4’,5’, hexachlorobiphenyl (Gundersen et al., 1998;

USEPA, 2006).

Sediments from the littoral zone of the Ohio River (river mile 369.5) were collected using a petite ponar grab in May, 2004. Sediments were transported to the laboratory in an ice chest, and subsamples for phylogenetic analysis and PCB degradation experiments were stored at -80°C and 4°C, respectively. Sediments up to 20 cm had in situ redox potential of 40±37 mV, temperature 15oC, pH 7.9, total carbon 3.2%, dissolved

organic carbon (11 mg kg-1), sulfate (90 mg kg-1), total nitrogen 0.11%, and Mehlich III

extractable nutrients (mg kg-1) P (22), K (79), Ca (2310), Mg (246), Zn (18), Cd (0.3), Cr

(0.5), Ni (4), Pb(0.7), Cu (1.9), Mo (0). At this location, sediments did not have detectable amounts of PCBs, as determined by gas chromatography with an electron capture detector (detection limit<0.05 ppm).

Microcosm treatments and sampling

Thirty-six microcosms were prepared to evaluate bioremediation treatment effects on changes in [14C]-PCB-153 levels and bacterial community composition in sediments

83

over a period of 177 d. Congener PCB-153 was selected because it is one of the most important congeners in the river (USEPA 2006). Sediments (10-mL containing 7 g dry

sediment) were added to 60-mL glass serum bottles, sealed with teflon-coated rubber

stoppers, made anaerobic by purging with N2, and amended with deoxygenated inorganic nutrient and vitamin solution (5-mL) (Owens et al., 1979).

After one week, when methane was detected in the headspace by GC-FID, one set of 18 microcosms was spiked with [14C]-PCB-153 solution (0.05 mL of 30,940 µg mL-1

acetone and 4.2 × 106 disintegrations min-1 (dpm) mL-1 (Sigma, St. Louis, MO).

Polyurethane foam plugs (PUF) (0.5 cm × 0.5 cm) and vials (2 mL) with 1 M NaOH (1

mL) were suspended inside the bottles with stainless steel wire to trap [14C]-volatile

14 organic (VOC) and CO2, respectively, and the microcosms were made anaerobic again by purging with nitrogen gas. The second set of 18 microcosms designated for community DNA analysis was prepared the same way as the first set except that the microcosms were spiked with non-labelled PCB-153 was used

Six treatments prepared in triplicate were imposed on the two sets of microcosms, which included incubations at 10°C, 25°C, 40°C, amendment with 5 mM FeSO4 at 25°C, amendment with 22.5 mM peptone on a carbon basis at 25°C, and amendment with 2.5 mM each of acetate, propionate, and butyrate at 25°C. On days 0, 35, 104, and 177,

14 14 14 14 microcosms with [ C]-PCB-153 were sampled for headspace CO2, CH4, and [ C]-

14 VOC analysis. Headspace CO2 was determined by analyzing radioactivity in the 1 M

NaOH trap. [14C]-VOCs were determined by analyzing radioactivity trapped in the PUF

14 14 plugs. Headspace CH4 was purge with air into a 900 °C furnace, and the CO2 from

combustion was in a 1 M NaOH. After each sampling, new PUF plugs and 1 M NaOH

84

traps were added and the headspace was purged with N2 to re-establish anaerobic conditions. Microcosms with non-labelled PCB-153 were also sampled for PCBs and

genomic DNA (day 177 only).

PCB extraction

PCBs were extracted from sediments (1 mL) in 16 mL serum bottles with 5:1 hexane:acetone (10 mL) on a horizontal shaker for 6 h. After centrifugation at 2500 × g

for 5 min, crude extracts were cleaned up with HNO3- and acetone-washed copper

powder (1 g) and subsequently analyzed for PCB congeners by gas chromatography with

an electron capture detector. Preliminary spike recovery experiments showed that this

procedure yielded >95% recovery of PCBs from sediments.

DNA extraction, PCR, cloning, and sequencing

Total community DNA was extracted from sediments (0.5 g) with an UltraClean

Soil DNA Isolation Kit (MoBio, Solana Beach, CA) using the maximum yield protocol

according to the manufacturer’s instructions. This extraction protocol includes a

combination of chemical detergent and bead beating steps to lyse cells, and was found to

be effective for evaluating bacterial and fungal diversity in soils, sediments, and other

environmental samples (Gomes et al., 2003; Hackl et al., 2004; Fierer et al., 2005). The

presence of high molecular weight DNA was verified by comparison with molecular

weight standards and gel electrophoresis with 0.8% agarose and 1× Tris-acetate-EDTA

buffer stained with ethidium bromide dye. The DNA was stored at -20°C until

polymerase chain reaction (PCR) amplification.

85

A 1.36 kb fragment of the bacterial 16S rDNA gene was amplified by PCR using

27F forward primer 5'-AGAGTTTGATC(A/C)TGGCTCAG-3' and 1392R reverse

primer 5'-ACGGGCGGTGTGT(A/G)C-3' (Lane, 1991). The PCR reaction mixture consisted of 19 µL distilled water, 10 µL 5× TaqMaster (Eppendorf, Westbury, NY), 5

µL 10× PCR buffer, 5 µL 2 mM dNTPs, 5 µL 2 µM forward and reverse primers, 0.25 µL

5 U µL-1 Taq DNA polymerase, and 1 µL of five times diluted DNA sample. DNA

extracts were diluted in order to dilute DNA and contaminants that can cause PCR bias

(Qiu et al., 2001). PCR amplification was carried out in a MyCycler thermocycler

(BioRad, Hercules, CA) with the following conditions: initial enzyme activation and

denaturation of 5 min at 94ºC, followed by 30 cycles of 1 min at 94ºC, 1 min at 60ºC, and

2 min of extension at 72ºC, and a final extension of 10 min at 72ºC. Thirty cycles was the

mininum number required to give faint bands of the expected size as determined by

comparison with molecular weight standards and gel electrophoresis with 1% agarose

and 1× Tris-acetate-EDTA buffer stained with ethidium bromide dye.

PCR products were purified using Promega Wizard SV Gel and PCR Clean-Up

System (Promega, Madison WI) and pooled from the three replicates in order to reduce stochastic PCR bias (Acinas et al., 2005). The pooled, purified products were ligated into

Invitrogen pCRII-TOPO cloning vector and transformed into chemically-competent

Escherichia coli TOP10F' cells (Invitrogen, Carlsbad, CA). The resulting clones were

isolated on Luria Bertani agar plates with 50 µg mL-1 kanamycin at 37ºC according to the manufacturer’s instructions. Ninety-four randomly selected white colonies per sample were grown overnight in 2-mL 96-well microplates containing 1.5 mL Luria Broth media

86

and 50 µg mL-1 kanamycin at 37ºC on an orbital shaker at 300 rpm, and two wells were

reserved for UK-AGTC to performed quality controls.

Plasmids from clones were isolated in 96-well blocks with a Perfectprep Plasmid

96 Vac Direct Bind Kit (Eppendorf, Westbury, NY) and Biomek FX Laboratory

Automated Workstation (Beckman Coulter, Fullerton, CA). Nucleotide sequencing

reactions were performed with the vector M13F primer and ABI PRISM BigDye v3.1

Terminator Cycle Sequencing Kit (Applied Biosystems, Foster City, CA), and resulting

sequences were read with an ABI PRISM 3730 DNA analyzer at the University of

Kentucky Advanced Genetic Technology Center (Lexington, KY).

Phylogenetic analysis

Sequences were aligned against a phylogenetically-diverse set of 16S rRNA sequences in the GreenGenes database using the Nearest Alignment Space Termination

(NAST) algorithm (http://greengenes.lbl.gov/cgi-bin/nph-NAST_align.cgi) and Clustal

W algorithm available in MEGA3 (http://www.megasoftware.net/; Kumar et al., 2004).

Aligned sequences were evaluated for potential anomalies (e.g. chimeras) using

MALLARD with E. coli as the reference sequence

(http://www.cardiff.ac.uk/biosi/research/biosoft/). Anomalous sequences were excluded

from further analysis.

Phylogenetic affiliations of 16S rDNA sequences were determined by the Basic

Local Alignment Search Tool (BLAST) 2.2.14 available at the National Center for

Biotechnology Information (NCBI) webpage (http://www.ncbi.nih.gov/BLAST/Blast.cgi) and the Classifier and Sequence Match tools available at Ribosomal Database Project II

87

(RDP II) Release 9.48 webpage (http://rdp.cme.msu.edu/index.jsp) (Cole et al., 2007).

Differences in community composition between libraries were determined using the

Library Compare tool available at RDP. Bacterial nomenclature followed Bergey's

Manual of Systematic Bacteriology Release 6.0.

Sequences generated in this study were deposited in the GenBank database under the accession numbers EF392901-EF393578.

Analytical methods

Total carbon and nitrogen of oven-dried (60°C) sediment samples were determined with a LECO CN-2000 analyzer (St. Joseph, MI). Soil pH was determined in

a 1:1 soil/water ratio. Concentrations of bioavailable elements (P, K, Ca, Mg, Zn, Cd, Cr,

Ni, Pb, Cu, Mo) were estimated by equilibrating sediment (2.0 g) for 5 min with Mehlich

III solution (20-mL) and filtering the extract with a Whatman #2 filter paper (Mehlich,

1984). Elements in the filtered extracts were analyzed with a Thermo Jarrell Ash Model

61 Inductively Coupled Argon Plasma (Franklin, MA). Sulfate was determined by ion chromatography with a Shimadzu LC-10AD VP liquid chromatograph (Columbia, MD).

Dissolved organic carbon of sediment pore water was determined using a Shimadzu

TOC-5000A Analyzer.

Radioactivity trapped in 1 M NaOH traps was determined by mixing an aliquot with scintillation cocktail and measuring activity using a Packard 2200 CA Tri-Carb liquid scintillation counter (Downers Grove, IL). Radioactivity in PUF was determined by extracting the PUF plug with acetone and mixing with scintillation cocktail as described above.

88

PCB transformation products in hexane:acetone extracts were analyzed using a

Shimadzu Model 14A gas chromatograph equipped with 63Ni electron capture detector

(320°C), autoinjector (250°C), and RTX-1 capillary column (30 m by 0.32 mm with

0.25-µm phase thickness) (Restek Corp., Bellefonte, PA) with the following temperature program: 100°C held for 1 min, ramp to 240°C at 3 min-1, and held at 240°C for 10 min.

PCB congeners were identified by comparing retention times with authentic standards

obtained from Accustandard (New Haven, CT).

Results and Discussion

Treatment Effects on PCB removal

Preliminary experiments were conducted to evaluate natural PCB attenuation

potential in surficial sediments by monitoring in situ changes in Aroclor 1254 congener

concentrations between August and November, when temperatures were in the optimal

range for PCB degradation (USACE, 1995; Wu et al., 1996; 1997). No PCB congeners

were lost from sediments in these investigations (data not shown), which signified that

natural attenuation of PCBs was not an important process at this location or time frame in

the river.

Because PCBs were not degraded under in situ conditions, additional laboratory

studies were performed to evaluate whether several anaerobic treatments that were shown

to be effective at promoting reductive dechlorination in other studies would be effective

in Ohio River sediments. In the lab investigations, up to 67% of PCB-153 was removed

in 177 d, which signified that there was considerable potential for PCBs to be removed

from sediments under certain environmental conditions (Fig. 3.1).

89

Temperature

300 a.

250

200

150 153 (mg/kg) - 100

PCB 10 ºC 25 ºC 50 40 ºC 0 0 50 100 150 200 Time (d) Carbon Source 300 b. 250 VAF Peptone 200 VAF+Br

150 153 (mg/kg) - 100 PCB 50

0 0 50 100 150 200 Time (d)

Figure 3.1 Anaerobic losses of 14C-2,4,5-2’,4’,5’-hexachlorobiphenyl (PCB-153) from

Ohio River sediments under different treatment conditions: (a) 10°C, 25°C, and 40°C, (b) peptone, volatile fatty acid (VFAs) mixture. PCB-153 losses in the FeSO4 treatment were

similar to the 25°C treatment. Each bar represents the standard deviation of the average

of three replicates.

90

As expected, temperature played an important role in controlling PCB-153

removal rates in the lab-incubated samples (Fig. 3.1). PCB-153 removal was greater at

25ºC than at 40ºC or 10ºC, which suggested that losses were microbially-mediated rather than lost by volatilization or other abiotic processes. Similar temperature effects on removal rates were observed in Hudson River and Woods Pond sediments (Mohn and

Tiedje, 1992; Wu et al., 1996; 1997). The fact that PCBs were lost in the lab incubation under anaerobic conditions at 25ºC but not in the in situ experiment was likely attributed to differences in redox and temperature conditions in the two experiments that significantly affected key bacterial populations, as discussed below. Results from the lab incubation suggested that considerable amounts of PCBs could be removed if sediments were maintained at 25ºC under anaerobic conditions for extended periods.

In contrast to temperature, FeSO4 amendments did not significantly affect PCB

removal in lab-incubated sediments (data not shown). This was in contrast to results in

Hudson River sediments, in which FeSO4 amendments improved PCB removal,

presumably by enriching for sulfate-reducing, PCB-transforming bacteria (Zwiernik et

al., 1998). One likely explanation for the lack of FeSO4 response in the current study was

that sediments contained high sulfate levels of 1 mM. Under these conditions, additional

sulfate inputs would not be expected to have a major effect on bacterial groups or PCB

dechlorination rates.

Both the peptone and the fatty acid mixture amendments significantly increased

PCB removal from sediments (Fig. 3.1). Similar organic carbon amendment effects on anaerobic PCB removal were observed in other studies (Nies and Vogel, 1990). Organic carbon likely stimulates PCB removal by serving as a carbon and energy source for PCB

91

transforming bacteria and other beneficial organisms in the dechlorinating community.

Indeed, many bacterial groups associated with reductive dechlorinating activity were

enriched in these treatments, as discussed below. These results, as well as dissolved

organic carbon levels in sediments of only 11 ppm, indicated that PCB transformers in

native sediments were electron donor-limited and that such amendments would likely improve anaerobic PCB removal rates in this system.

In all treatments in which PCB-153 was decreased, there were concomitant increases in 2,4,5-2’,4’-pentachlorobiphenyl (PCB-99) and 2,4-2’,4’-tetrachlorobiphenyl

14 14 14 (PCB-47) (Fig. 3.2). CO2, CH4 and [ C]-VOC were not detected in any treatments, which indicated that PCB-153 mineralization and volatilization were not major processes in the anaerobic sediments. These results demonstrated that PCB-153 and PCB-99 were lost primarily by reductive dechlorination, and that chlorines were preferentially removed from the meta positions of the biphenyl ring. Chlorine removal from the meta position is consistent with Pathway N, although additional pathways may have become apparent if additional congeners were evaluated (Bedard and Quensen, 1995; Williams, 1997). PCB

dechlorination at the meta position is common in other aquatic sediments, including those

from the Hudson River (NY), Woods Pond (MA), and Silver Lake (MA) (reviewed by

Bedard and Quensen, 1995). In the in situ study, the lower chlorinated congeners in

Aroclor 1254 were stable and so if produced would also pose health risks.

River Sediment Bacterial Community Composition

One of the main goals of this study was to evaluate the composition of the

dominant bacterial inhabitants in sediments because of their possible relationship to

92

250 PCB-153

PCB-47

200 PCB-99

150

100

PCB conc. (mg/kg) conc. PCB 50

0 0 50 100 150 200 Time (d)

Figure 3.2 Polychlorinated biphenyl concentrations during reductive dechlorination of

PCB-153 to PCB-99 and PCB-47 using peptone as organic carbon amendment. Each bar

represents the standard deviation of the average of three replicates.

93

important physical and chemical conditions and PCB removal in this system. Duplicate

libraries were prepared from the freshly collected sediments (ORSFC1 and ORSFC2),

which were found to contain similar distributions of bacteria (Table 3.1). Diverse groups

of bacteria were detected, including difficult-to-lyse Gram-positive bacteria. These results suggested that the DNA and PCR methods were reproducible and that many of the artifacts commonly associated with these methods were minimal in this study.

Most sequences in the duplicate libraries (83-88%) were distributed into ten phyla, including Proteobacteria (primarily in classes beta-, alpha-, and delta-

Proteobacteria) > Bacteroidetes (mainly in order Sphingobacterales) = Chloroflexi

(Anaerolineales) > Nitrospira = Actinobacteria = Gemmatimonadetes = Acidobacteria >

Planctomycetes = Spirochaetes = Firmicutes (Table 3.1). The remaining 12-17% of sequences could not be classified with confidence into any known phyla.

None of the sequences in the clone libraries from freshly-collected sediments were related to PCB dechlorinators (e.g. Dehalococcoides, o-17, DF-1 and Clostridium)

or aerobic PCB degraders, which could at least partially explain why Aroclor 1254 was

so stable in sediments in the in situ study. PCB dechlorinators were obviously present in

sediments, however, as evidenced by PCB-153 conversion to PCB-99 and PCB-47 in the lab studies. It is possible that PCB dechlorinators would have been detected if more clones or specific primers were evaluated. Nevertheless, these groups did not appear to be numerically dominant or active in the natural environment, which was probably attributed to unfavorable physical and chemical conditions as suggested by the types of other bacteria that were dominant in the sediments.

94

Table 3.1 Distribution of bacteria in major taxonomic groups in duplicate libraries (ORS FC1 and ORS FC2) from Ohio River

sediments before anaerobic treatments, as determined by the Classifier tool available at the Ribsomal Database Project (RDP) (Cole et

al., 2007). Values in parentheses are number of sequences in the libraries. There were no significant differences between libraries at p-

value=0.05, as determined by the Library Compare tool available in RDP.

Replicate Replicate

ORS FC1 ORS FC2 Representative genera in libraries

Phylum Class Order n=64 n=69 n=133

% of clones

95 Proteobacteria 44 52

Alphaproteobacteria 11 12

Rhizobiales 6 3 Bradyrhizobium (1), Hyphomicrobium (1)

Sphingomonadales 2 6 Sphingopyxis (3), Novosphingobium (2)

Rhodospirillales 2 1

Rhodobacterales 0 0

Caulobacterales 0 1 Brevundimonas (1)

unclassified 2 0

Table 3.1. (continued)

Replicate Replicate

ORS FC1 ORS FC2 Representative genera in libraries

Phylum Class Order n=64 n=69 n=133

% of clones

Betaproteobacteria 22 16

Burkholderiales 11 12 Leptothrix (1), Aquabacterium (3)

Rhodocyclales 5 1 Rhodocyclus (1), Propionivibrio (1)

Nitrosomonadales 2 0 Nitrosomonas (1) 96

Methylophilales 0 0

unclassified 5 3

Gammaproteobacteria 3 3

Xanthomonadales 2 0

Methylococcales 0 1 Methylobacter (1)

Pseudomonadales 0 0

unclassified 2 1

Table 3.1. (continued)

Replicate Replicate

ORS FC1 ORS FC2 Representative genera in libraries

Phylum Class Order n=64 n=69 n=133

% of clones

Deltaproteobacteria 6 19

Desulfuromonales 5 6 Desulfuromonas (1), Malonomonas (1), Geobacter (5) Syntrophobacterales 0 0

Myxococcales 0 3 97

Desulfobacterales 0 4 Desulfonema (1), Desulfocapsa (2)

unclassified 2 6

unclassified 2 0

Chloroflexi Anaerolineae Anaerolinaeles 11 9 Anaerolinea (13)

Bacteroidetes 8 12

Sphingobacteria Sphingobacteriales 5 9 Chitinophaga (6)

Bacteroidetes Bacteroidales 2 0 Bacteroides (1)

Flavobacteria Flavobacteriales 2 3 Algoriphagus (1), Cryomorpha (1), Flavobacterium (1)

Table 3.1. (continued)

Replicate Replicate

ORS FC1 ORS FC2 Representative genera in libraries

Phylum Class Order n=64 n=69 n=133

% of clones

Firmicutes 0 1

Clostridia Clostridiales 0 0

unclassified 0 0

Bacilli Bacillales 0 0 98

unclassified 0 0

Actinobacteria 3 4

Actinobacteria Coriobacteriales 0 0

Actinomycetales 0 4 Asanoa (2), Verrucosispora (1)

Rubrobacterales 2 0

Acidimicrobiales 2 0 Acidomicrobium (1)

unclassified 0 0

Gemmatimonadetes Gemmatimonadetes Gemmatimonadales 6 1 Gemmatimonas (5)

Table 3.1. (continued)

Replicate Replicate

ORS FC1 ORS FC2 Representative genera in libraries

Phylum Class Order n=64 n=69 n=133

% of clones

Planctomycetes Planctomycetacia Planctomycetales 0 3 Isosphaera (1)

Acidobacteria Acidobacteria Acidobacteriales 2 4 Acidobacterium (3)

Nitrospira Nitrospira Nitrospirales 8 1 Magnetobacterium (2), Nitrospira (4)

Verrucomicrobia Verrucomicrobiae Verrucomicrobiales 2 0 99

Spirochaetes Spirochaetales 2 0 Spirochaeta (1)

WS 3 0 0

OP 10 0 0

BRC 1 0 0

unclassified 17 12

The physiological characteristics of bacterial groups that made up a sizeable

fraction of clone libraries (>3% of sequences in libraries) provided several clues as to the chemical conditions and reasons for PCB recalcitrance in sediments. The most

abundantsequences were classified in the genus Anaerolinea within subdivision I of the

phylum Chloroflexi (9% of sequences in both of the duplicate libraries). Anaerolinea

spp. are obligately anaerobic syntrophs that ferment carbohydrates and amino acids

primarily to acetate and H2 (Sekiguchi et al., 2003). Interestingly, Anaerolinea spp. are

commonly found in anaerobic dechlorinating environments, and probably support the

reaction by supplying electron donors to dechlorinating populations. The large presence

of this group suggests that at least part of the sediment profile was anaerobic and would

be able to support reductive dechlorination in sediments under certain conditions.

Also prevalent in the clone libraries were sequences classified in the genus

Gemmatimonas within subdivision 1 of phylum Gemmatimonadetes (3-8% of sequences)

(Table 3.1). Gemmatimonas are aerobic and mesophilic bacteria that utilize many types

of simple sugars, acetate, formate, and benzoate as carbon and energy sources (Zhang et

al., 2003). The large presence of this group suggested that part of the sediment profile

was aerobic. Reductive dechlorination would not be expected in these zones. However, if

PCBs were transformed by aerobic bacteria, then Gemmatimonas could play a role in

PCB removal by degrading benzoate, which is a major and toxic intermediate of aerobic

PCB degradation (Bevinakatti and Ninnekar, 1992; Boyle et al., 1992; Vrana et al.,

1996).

Sequences classified in the genus Chitinophaga within the order

Sphingobacteriales and phylum Bacteroidetes were also abundant in clone libraries from

100

freshly-collected sediments (2-7% of sequences) (Table 3.1). Chitinophaga are aerobic/microaerophilic bacteria that hydrolyze a wide range of biopolymers (e.g. chitin, proteins, and polysaccharides) and degrade many simple carbohydrates to organic acids

(Pankratov et al., 2006). The presence of Chitinophaga spp. also suggested that part of the sediment profile was oxidized.

Sequences affiliated with the genus Geobacter within the order Desulfuromonales under the class delta-Proteobacteria were also prevalent in clone libraries from freshly- collected sediments (3-4% of sequences) (Table 3.1). Geobacter are not known to transform PCBs. However, they may play a role in PCB removal by degrading toxic PCB degradation intermediates or other pollutants under different terminal electron acceptor reducing conditions that are likely to occur in contaminated sediments (Lovley et al 1993;

Wischgoll et al., 2005; Sung et al., 2006).

The presence of bacterial groups with wide O2 tolerances in the sediments, ranging from obligately aerobic to strictly anaerobic, suggested that the sediment profile contained a strong redox gradient. Such a gradient would tend to support sequential reduction and oxidation of PCBs, although there was no evidence of this in the in situ experiment. This suggested that other factor(s), such as electron donor availability, limited these activities in the sediments, as shown in the lab study.

Many of the groups that were dominant in Ohio River sediments were also prevalent in river sediments from other systems, which were determined by fluorescent in situ hybridization (FISH) using various probes for different organisms (Fazi et al., 2005;

Kloep et al., 2006; Peplies et al., 2006). For example, Fazi et al. (2005) discovered that alpha- and beta-Proteobacteria were dominant in three Italian river sediments, Kloep et

101

al. (2006) observed that alpha-, beta-, and gamma-Proteobacteria, Cytophaga-

Flavobacteria, and Planctomycetales were dominant in sediments from the Elbe River in

Germany, and Peplies et al. (2006) found that beta-Proteobacteria, sulfate-reducers, and

methanotrophic bacteria were prevalent in four other German river sediments. In these

studies, only a small fraction of the bacterial community was detected (<40%),

suggesting that many more probes would be needed to fully characterize bacterial

diversity in these systems (Peplies et al., 2006). Results from this study suggest that

additional probes to evaluate bacterial diversity in river sediments might include those for

Anaerolinea, Gemmatimonas, Chitinophaga, Geobacter, and other groups listed in

Tables 3.1 and 3.2.

Temperature Effects on Anaerobic Bacterial Community Composition

It was expected that temperature would significantly affect bacterial community

composition, based on its influence on PCB removal in this and other studies. Indeed,

proteobacterial sequences were significantly higher in the 10°C library compared to the

25°C and 40°C libraries, especially with regard to sequences in the class alpha-

Proteobacteria and order Rhizobiales (Table 3.2). The opposite trend was observed with

Bacteroidetes-related sequences. Most of the Bacteroidetes sequences that were enriched

in the 25°C library were classified as hydrolytic bacteria Chitinophaga spp., which was

surprising considering the O2 requirements of this genus. These results suggest that this group might play a critical role in organic carbon cycling in both anaerobic and aerobic sediments. Most of the Bacteroidetes-related sequences enriched in the 40°C library

were affiliated with the strictly anaerobic fermentative bacteria Anaerophaga (Denger et

102

Table 3.2 Distribution of bacteria in major taxonomic groups in libraries from Ohio River sediments after exposure to ,anaerobic

treatments for 177 d, as determined by the Classifier tool available at the Ribsomal Database Project (RDP) (Cole et al., 2007). Values

in parentheses are number of sequences in the libraries.

Anaerobic treatments

Temperature† Amendment and 25°C‡

Fatty acid Phylum Class Order 10°C 25°C 40°C FeSO Peptone 4 mixture Representative genera in libraries (75) (59) (71) (72) (62) (62) (407) % of clones Proteobacteria

Alphaproteobacteria 52a 27b 30b 33 24 27 103 Mesorhizobium (1), Filomicrobium Rhizobiales 15a 9ab 1b 7 8 5 (1), Pedomicrobium (3), Rhodoplanes (3), Hyphomicrobium (2), Bradyrhizobium (1), Nitrobacter (1) Sphingomonadales 8 3 0 3 5 3 Sphingopyxis (2), Novosphingobium (3) Rhodospirillales 4 2 0 1 0 0 Tistrella (1)

Rhodobacterales 0 2 1 0 2 2 Amaricoccus (1), Rhodobacter (3)

Caulobacterales 3 2 0 3 0 0 Phenylobacterium (1)

unclassified 0 0 0 0 2 0

Betaproteobacteria 11 7 7 7 5 6

Burkholderiales 7 0 3 6 2 2 Acidovorax (1), Cupriavidus (2), Aquabacterium (2)

Table 3.2. (continued)

Anaerobic treatments

Temperature† Amendment and 25°C‡

Fatty acid Phylum Class Order 10°C 25°C 40°C FeSO Peptone 4 mixture Representative genera in libraries (75) (59) (71) (72) (62) (62) n=407 % of clones Rhodocyclales 1 2 0 1 2 0 Dechloromonas (1)

Nitrosomonadales 0 0 0 0 0 0

Methylophilales 0 0 0 0 0 2 Methylophilus (1)

unclassified 3 5 4 0 2 3

104 12 3 3 8 2 3

Xanthomonadales 7 2 3 4 2 2 Thermomonas (1), Lysobacter (1)

Methylococcales 0 0 0 1 0 0 Methylobacter (1)

Pseudomonadales 0 0 0 0 0 2 Pseudomonas (1)

unclassified 5 2 0 3 0 0

Deltaproteobacteria 15 12 17 11 8 13

Desulfuromonales 4 2 0 4 0 3 Geobacter (3), Desulfurmonas (3)

Syntrophobacterales 4a 3a 11b 1 2 3 Desulfomonile (1), Desulfobacca (1), Smithella (12) Myxococcales 3 0 0 4 2 6 Haliangium (1), Cystobacter (3), Mellitangium (1), Chondromyces (1) Desulfobacterales 1 2 0 1 0 0 Desulfonema (1), Desulfobacterium (2)

Table 3.2. (continued)

Anaerobic treatments

Temperature† Amendment and 25°C‡

Fatty acid Phylum Class Order 10°C 25°C 40°C FeSO Peptone 4 mixture Representative genera in libraries (75) (59) (71) (72) (62) (62) n=407 % of clones unclassified 3 5 6 0 5 0

unclassified 0 0 1 0 2 0

Chloroflexi Anaerolineae Anaerolinaeles 9 5 17 11 16 14 Anaerolinea (48)

Bacteroidetes 3a 15b 14b 8 6 5

105 Sphingobacteria Sphingobacteriales 1 5 0 6 2 2 Chitinophaga (7)

Bacteroidetes Bacteroidales 1a 7b 13b 1 3 0 Anaerophaga (4), Rickenella (1)

Flavobacteria Flavobacteriales 0 2 0 1 2 0 Flavobacterium (1), Cryomorpha (1)

unclassified 0 2 1 0 0 3

Firmicutes 8 5 4 6 14* 26*

Aminomonas (1), Cryptanaerobacter Clostridia Clostridiales (1), Dehalobacter (3), 5 3 3 6 14* 23* Desulfitobacterium (11), Clostridium (2), Acetivibrio (1) unclassified 0 0 0 0 0 2

Bacilli Bacillales 2 0 1 0 0 0 Bacillus (2)

unclassified 0 2 0 0 0 2

Table 3.2. (continued)

Anaerobic treatments

Temperature† Amendment and 25°C‡

Fatty acid Phylum Class Order 10°C 25°C 40°C FeSO Peptone 4 mixture Representative genera in libraries (75) (59) (71) (72) (62) (62) n=407 % of clones Actinobacteria 4 7 3 7 8 0

Actinobacteria Coriobacteriales 1 2 0 1 2 0

Actinomycetales 1 2 1 0 3 0 Nocarioides (1), Propionicimonas (1)

Rubrobacterales 1 2 1 3 3 0 Solirubrobacter (1), Conexibacter (4)

106 Acidimicrobiales 0 0 0 1 0 0

unclassified 0 2 0 1 0 0

Acidobacteria Acidobacteria Acidobacteriales 0 5 0 0 0 0 Acidobacterium (1)

Nitrospira Nitrospira Nitrospirales 1 0 0 4 0 0 Magnetobacterium (2), Nitrospira (2)

Verrucomicrobia Verrucomicrobiae Verrucomicrobiales 0 0 0 0 0 2 Prosthecobacter (1)

Spirochaetes Spirochaetales 0 2 1 1 0 3

Gemmatimonadetes Gemmatimonadetes Gemmatimonadales 1 0 1 1 2 0 Gemmatimonas (4)

Planctomycetes Planctomycetacia Planctomycetales 0 0 1 0 0 3

WS3 3 2 0 1 0 0 WS3 (3)

Table 3.2. (continued)

Anaerobic treatments

Temperature† Amendment and 25°C‡

Fatty acid Phylum Class Order 10°C 25°C 40°C FeSO Peptone 4 mixture Representative genera in libraries (75) (59) (71) (72) (62) (62) n=407 % of clones

OP10 0 2 3 0 3 0 OP10 (5)

BRC1 0 0 0 0 2 0 BRC1 (1)

unclassified 19 27 25 26 24 19

† Values followed by a different letter are significantly different at p=0.05. 107 ‡ Values followed by * are significantly different than values in the 25°C library at p=0.05.

al., 2002). Interestingly, at least one Anaerophaga strain produces biosurfactant that increases the solubility of hydrophobic compounds. Syntrophaceae sequences within the class delta-Proteobacteria were also significantly enriched in the 40°C library compared to the 10°C and 25°C libraries. Most of these sequences were classified as Smithella, an anaerobic propionate-degrading syntroph (Liu et al., 1999). The fact that fermentative and syntrophic bacteria were significantly reduced at 10°C was consistent with previous results that showed this group to be particularly sensitive to low temperature in methanogenic paddy soils (Chin and Conrad, 1995).

Reduced levels of fermentative and syntrophic populations in the 10°C library could at least partially explain why PCB dechlorination rates were negligible in this treatment, since these groups are essential for providing electron donors to PCB dechlorinators. The fact that PCB dechlorination was low in the 40°C treatment, despite having high levels of these groups, suggested that PCB dechlorinators were strongly inhibited at this temperature.

FeSO4 Amendment Effects on Anaerobic Bacterial Community Composition

It was expected that FeSO4 amendments to anaerobic sediments would increase levels of sulfate reducing bacteria, particularly strains with PCB dechlorination abilities as proposed by Zwiernik et al. (1998). However, FeSO4 amendments did not have major effects on any bacterial group compared to unamended sediments (Table 3.2). As mentioned previously, a likely explanation is that sediment sulfate concentrations of 1 mM did not limit the growth of sulfate reducing bacteria, which included

Desulfobacterium, Desulfuromonas, Desulfonema, Desulfobacca and Desulfitobacterium.

108

Under high sulfate conditions, it is likely that some other factor limited sulfate reducing

bacteria growth, such as electron donor availability.

Organic Carbon Effects on Anaerobic Bacterial Community Composition

It was expected that amendment of sediments with peptone and a mixture of fatty acids would significantly affect bacterial communities, based on their effects on PCB removal rates as discussed earlier. As expected, both treatments significantly increased levels of Clostridiales sequences compared to unamended sediments (Table 3.2). A

majority of these sequences, which made up 11-13% of clone libraries, were closely related (94% similar) to two uncultured clones from other anaerobic reductive

dechlorinating environments, namely clone SJA-118, a member of an anaerobic

trichlorobenzene-transforming microbial consortium (von Wintzingerode et al., 1999) and clone Peptococcaceae KB-1 1, a member of an anaerobic chloroethene- dechlorinating microcosm (GenBank accession number AY780555) (Fig. 3.3). These sequences were also closely related (90-91%) to several cultured reductive dechlorinating

Desulfitobacterium strains, including tetrachloroethene-dechlorinators

Desulfitobacterium sp. Viet-1 (GenBank accession number AF357919),

Desulfitobacterium hafniense Y51 (Villemur et al., 2002), and chlorophenol- dechlorinator Desulfitobacterium hafniense (previously D. frappieri) PCP-1 (Thibodeau et al., 2004; Villemur et al., 2005).

Although Desulfitobacterium spp. are not known to dechlorinate PCBs, at least two lines of evidence suggest that this population may have conducted the reaction in

Ohio River sediments. First, the Desulfitobacterium population(s) was closely related to

109

100 Clone ORSFAM_c07; EF393260 + 7 additional ORSFAM clones 99 Clone ORSPEP_b12; EF393529 + 6 additional ORSPEP clones 55 Clone SJA-118; TCB dechlorinating consortia; AJ009489 100 Clone KB-1 1; Chloroethene-dechlorinating culture; AY780555 Clone ORSFES_c12; EF393492 Desulfitobacterium sp. Viet-1; PCE-dechlorinating culture; AF357919 44 Peptococcaceae 100 D. hafniense PCP-1; PCP-dechlorinating culture; U40078 100 D. hafniense Y51; AP008230 98 Desulfosporosinus sp 5apy; Raw and drinking water; AF159120 99 Desulfosporosinus sp. LauIII; Anaerobic lake sediments; AJ302078 63 88 Desulfosporosinus sp. 44a-T3a; ZnS-producing biofilm; AY082482 Clone ORSFAM_d07; EF393298 Clone ORS10C_e12; EF392927 Clone ORSFAM_h03; EF393286 71 unclassified Clostridia 100 Clone SHA-86; Dichloropropane-dechlorinating culture; AJ306756 40 79 Clone ORSPEP_b09; EF392527 63 Clone AKIW850; Urban aerosol; DQ129557 58 Clone RL176_aah44c06; Human gut; DQ793847 Clostridiaceae 100 Clone ORSPEP_h05; EF393547 C. bifermentans TYR6, DSM 13560; Oil mill wastewater isolate; AF320283 100 Clone ORS10C_h02; EF392935 60 79 Clone D282120; Oxic-anoxic interphase of flooded paddy soil; AJ617861 67 Clone ORS10C_g03; EF392964 B. luciferensis LMG 18422; Volcanic soil ; AJ419629 Bacillaceae 100 Clone AKAU3813; Uranium contaminated soil; DQ125717 59 100 Clone ORS40C_a03; EF393468 42 Clone AKIW1036; Urban aerosol; DQ129447 Clone ORS25C_b09; EF393012 Clostridiaceae 99 Clone WCHB1-21; Chlorinated-solvent-contaminated aquifer; AF050580 100 Clone ORSFES_f12; EF393468 100 Clone W4-B29; Lake sediments; AY345500 unclassified Clostridia Clone SHA-32; Dichloropropane-dechlorinating consortia; AJ306752 79 Clone ORSFAM_d01; EF393263 100 Clone N21; VFAs-degrading anaerobic sludge; AB195866 Syntrophomonadaceae Clone TSBX14; Polychlorinated-dioxin dechlorinating consortia; AB186800 94 Clone ORS40C_e04;EF393052 85 Clone BE325FW032701CTS_hole1-1; Continental crust; DQ088772 55 Clone ORS10C_c02; EF392946 Clone ORSFAM_d04; EF393266 unclassified Clostridia 100 Clone SHA-16; Dichloropropane-dechlorinating consortia; AJ306751 68 98 Clone ORS25C_e12; EF392996 99 Clone C-131; Anaerobic swine lagoon; DQ018807

0.02

Figure 3.3 Bootstrapped neighbor-joining phylogenetic tree of sequences in the phyla

Firmicutes obtained from the present study and closest relatives in the Ribosomal

Database Project release 9.48. Sequences obtained in the present study are indicated in

bold with the following abbreviations: ORSFC1 and ORSFC2=freshly-collected

sediment (before anaerobic treatment), ORS10C=10°C anaerobic treatment,

ORS25C=25°C anaerobic treatment, ORS40C=40°C anaerobic treatment,

ORSFAM=volatile fatty acid mixture anaerobic treatment, ORSFES=FeSO4 anaerobic

treatment, ORSPEP=peptone anaerobic treatment. Numbers at branch points are percentages of 500 resampling bootstrap analyses. The scale bar represents 0.02 changes

per nucleotide.

110

strains with well-known dechlorination abilities, some of which are able to dehalogenate compounds with structures very similar to PCBs, such as hydroxylated PCBs (Wiegel et al., 1999). Second, the Desulfitobacterium population was largest in samples with the greatest PCB dechlorination, which would be expected if the reaction was coupled to energy generation by dehalorespiration, a capability possessed by some members of this genus (Finneran et al., 2002; Villemur et al., 2006). Results from several other studies suggest that PCB dechlorination was conducted by bacteria with Desulfitobacterium-like

characteristics, including anaerobic, spore-forming bacteria (Ye et al., 1992), anaerobic

Gram-positive bacteria (Williams, 1997), and sulfate-reducing bacteria (Zwiernik et al.,

1998). As pointed out by others (Nonaka et al., 2006; Villemur et al., 2006), additional

work is needed to isolate desulfitobacteria populations and characterize their

dechlorination potential, particularly because of their metabolic versatility and ability to

survive in a wide range of environments.

In conclusion, results from this study showed that Ohio River sediment

populations held large potential for PCB reductive dechlorination, although this potential

may not often be realized because of unfavorable temperature, redox, and substrate

availability conditions. Changes in these conditions were associated with several bacterial

taxa, suggesting that they could be used as indicators to guide and monitor anaerobic

bioremediation activities in situ. Further work is needed to substantiate the accuracy and

sensitivity of these potential indicator taxa for predicting geochemical conditions and

processes in other anaerobic systems. Additional work is also needed to determine

potential mineralization of PCB reductive dechlorination products in a sequential

anaerobic/aerobic treatment, conditions that can be established in the field due to

111

fluctuations in river level. Finally, much debate is still present regarding the microorganisms responsible for PCB reductive dechlorination. PCR amplification of gene targeting reductive dehalogenases may provide additional information on the presence of potential PCB dechlorinators that cannot be identified or detected using the 16S rDNA- based clone library approach. In chapter 4, these objectives will be addressed using PCB-

153 as model compound and PCR primers targeting reductive dehalogenase genes that have been identified.

112

CHAPTER 4

SEQUENTIAL ANAEROBIC-AEROBIC TREATMENT OF PCB-153 IN OHIO

RIVER SEDIMENTS.

Introduction

In the previous chapter it was shown that reductive dechlorination of PCBs in

Ohio River sediments has great potential, and that controlled conditions of nutrient

availability, temperature and redox status, have a major impact on dechlorination rates.

For higher chlorinated PCBs this is a requirement to improve their aerobic degradation

that otherwise would be minimal. Lower chlorinated congeners are readily oxidized by

several aerobic bacteria such as Pseudomonas, Alcaligenes, Achromobacter,

Burkholderia, Comamonas, Sphingomomonas, Ralstonia, Acinetobacter, Rhodococcus,

Corynebacterium, and Bacillus (Abraham et al., 2002, Field and Sierra-Alvarez, 2008).

This degradation occurs via dioxygenation of the aromatic ring that can occur at positions

2,3 or 3,4. The enzymes responsible for this reaction are known as biphenyl dioxygenases

(Borja et al., 2005; Furukawa, 2000; Pieper, 2005). However, the aerobic degradation rate usually decreases with increased chlorine content, and PCBs with ortho chlorines are particularly resistant, therefore, aerobic degradation of highly chlorinated PCBs is severely restricted, and reduction of chlorine content is required for this reaction to occur.

Reductive dechlorination reaction is an alternative for degradation of PCBs. In this reaction, PCBs are used as electron acceptors for energy production, resulting in the substitution of chlorine atoms with hydrogen, and thus producing lower chlorinated congeners. Early evidence of the importance of reductive dechlorination of PCBs in the environment were alterations in congener profiles in PCB contaminated aquatic

113

sediments compared to those of the original commercial mixtures (Brown et al., 1987;

Abramowicz, 1995; Wu et al., 1998). Typically, less chlorinated, ortho-chlorinated

congeners tend to accumulate and highly chlorinated meta- and para-chlorinated congeners tend to be depleted in these environments (Quensen et al., 1988).

Interestingly, toxicity is also reduced during reductive dechlorination because PCB toxicity depends largely on substitutions at those positions because of structural similarities with polychlorinated dioxins (PCCDs).

As both aerobic degradation and reductive dechlorination have been well studied

(Borja et al., 2005; Field and Sierra-Alvarez, 2008), it is expected that exposure to anaerobic conditions followed by aerobic conditions could result in the complete removal of PCBs from the environment. Because reductive dechlorination does not completely eliminate the PCB problem under anaerobic conditions, highly chlorinated congeners will be transformed into lower chlorinated congeners that otherwise could not be transformed under aerobic conditions into open ring intermediates and finally into carbon dioxide and water. Previous studies have demonstrated the feasibility of sequential anaerobic-aerobic treatment of PCBs. For example, Master et al. (2002) showed that weathered Aroclor

1260 can be treated in a sequence of four months of anaerobic incubation followed by 28 days of aerobic incubation. The results were an average substitution of about one chlorine atom per molecule during the anaerobic phase, with no changes in molar concentration, and about 60% total removal of PCBs by the end of the aerobic phase. Evans et al.

(1996) also showed that the sequential anaerobic-aerobic treatment of weathered Aroclor

1248 improved the removal of PCBs from 67% to 70% compared to aerobic treatment alone. Although this might represent a small improvement in PCB removal, the

114

proportion of penta and hexachlorinated congeners in the aerobic treatment was higher,

proving that the overall treatment reduces both the concentration and the chlorine content

in the mixture. In both studies the aerobic phase was inoculated with enriched cultures of

Burkholderia sp. LB400, a known PCB degrader (Rein et al., 2007; Rodrigues et al.,

2006) to improve the degradation rates, but the existence of a native microbial

community capable of aerobic degradation was not verified.

Therefore, one of the objectives in the research reported in this chapter is to

determine the feasibility of sequential anaerobic-aerobic degradation of PCBs in Ohio

River sediments, particularly PCB-153. Additionally, previous results from the in situ

bioremediation in the Ohio River suggests that potential aerobic degraders are present,

and therefore enriched cultures will not be used. Finally, the ultimate goal in this project is to identify the species involved in the reductive dechlorination of PCBs in this environment, and it is expected that using primers designed to detect particular reductive dehalogenase genes will help answer this question.

Materials and Methods

Site description, sample preparation, and collection

Sediments from the littoral zone at river mile 369.5 of the Ohio River

(38°37'14.34"N, 83°9'30.84"W) were collected using a petite ponar grab sampler.

Sediments were transported to the laboratory in an ice chest, and subsamples for DNA

analysis and PCB degradation experiments were stored at -80°C and 4°C, respectively.

At this location, sediments did not have detectable amounts of PCBs, as determined by

gas chromatography with an electron capture detector (detection limit<0.05 ppm).

115

Approximately 10 L of sediment was collected and homogenized in a plastic container,

and dispensed into 2.5 L glass bottles equipped with teflon-lined screw caps. Sediments

were stored at 4 ºC until further use.

To evaluate the potential for a sequential anaerobic-aerobic bioremediation of

PCBs, a total of twenty microcosms were prepared. Briefly, sediments (10-mL containing

7 g dry sediment) were added to 120-mL glass serum bottles, sealed with teflon-coated

rubber stoppers, and amended with deoxygenated inorganic nutrient and vitamin solution

(5-mL) (Owens et al., 1979). These microcosms were designated for PCB and DNA analysis and were spiked with [14C]-PCB-153 (0.05 mL of 31 mg mL-1 acetone and 4.2 ×

106 disintegrations min-1 (dpm) mL-1 (Sigma, St. Louis, MO) to a final concentration of

100 mg kg-1 on a dry sediment basis. Fifteen of these microcosms were purged with nitrogen gas, the rest with air, and then incubated in the dark for 150 days. Air-purged microcosms were repeatedly purged with air every 3 days to ensure aerobic conditions were established, and the purged headspace was collected through polyurethane foam

(PUF) plugs to trap volatile organic compounds, and then through 1 M aqueous NaOH to trap carbon dioxide that may have been produced during the process. At the end of the

150 day period, three serum bottles incubated anaerobically and three incubated aerobically were removed for PCB and DNA analysis. The rest of the bottles were incubated for another 100 days under the same conditions with the exception of six anaerobic bottles that were purged with air and treated aerobically for the rest of the incubation period. These bottles represent the anaerobic-aerobic sequential treatment.

After 50 and 100 days (200 and 250 days total incubation time), three bottles from each

116

treatment, anaerobic, aerobic, and anaerobic-aerobic, were removed for PCB and DNA analysis.

PCB extraction

Approximately 2 g of wet sediment was mixed with 10 g of anhydrous sodium sulfate (to remove excess moisture) and 25 mL of 5:1 hexane:acetone mixture in a wide-

mouth 60-mL extraction flask for 6 h on a horizontal shaker. The extracted PCBs were

separated from solids by centrifugation at 2500 rpm for 15 min, and 1 mL of the crude

extract was cleaned up with HNO3 and acetone-washed copper powder (1 g) to remove elemental sulfur that was present at elevated levels and interfered with PCB analysis by gas chromatography with electron capture detector.

DNA extraction and PCR amplification, cloning, and sequencing

Total community DNA was extracted from sediments (0.5 g) with an UltraClean

Soil DNA Isolation Kit (MoBio, Solana Beach, CA) using the maximum yield protocol

according to the manufacturer’s instructions. This extraction protocol includes a

combination of chemical detergent and bead beating steps to lyse cells, and was found

effective for evaluating bacterial and fungal diversity in soils, sediments, and other

environmental samples (Gomes et al., 2003; Hackl et al., 2004; Fierer et al., 2005). The

presence of high molecular weight DNA was verified by comparison with molecular

weight standards using gel electrophoresis with 0.8% agarose and 1× Tris-acetate-EDTA

buffer stained with ethidium bromide dye. The DNA was stored at -20°C until

polymerase chain reaction (PCR) amplification.

117

Separate PCR reactions were conducted for preparation of clone libraries and

DGGE profile analysis of 16S rDNA genes in the sediment samples. For clone libraries,

a 918-bp fragment of the bacterial 16S rDNA gene was amplified by PCR using 27F

forward primer 5'-AGA GTT TGA TC(A/C) TGG CTC AG-3' and 907R reverse primer

5'-CCG TCA ATT C(A/C)T TT(A/G) GTT T-3' (Lane, 1991; Muyzer et al., 1993). The

PCR reaction mixture consisted of 6.5 µL distilled water, 12.5 µL 2× IQ Supermix (Bio-

Rad, Hercules, CA), 2.5 µL 2 µM forward and reverse primers, and 1 µL of five times

diluted DNA sample. Each PCR reaction had the following final concetrations: 50 mM

KCl, 20 mM Tris-HCl, pH 8.4, 0.2 mM each dNTP (dATP, dCTP, dGTP, dTTP), 25

U/ml iTaq DNA polymerase, and 3 mM MgCl2. DNA extracts were diluted in order to

dilute DNA and contaminants that can cause PCR bias (Qiu et al., 2001). The PCR was

carried out in a MyCycler thermocycler (BioRad, Hercules, CA) with the following temperature program: initial enzyme activation and denaturation of 5 min at 94ºC, followed by 30 cycles of 1 min at 94ºC, 1 min at 60ºC, and 2 min of extension at 72ºC, and a final extension of 10 min at 72ºC. Thirty cycles was the mininum number required to give faint bands of the expected size as determined by comparison with molecular weight standards using gel electrophoresis with 1% agarose and 1× Tris-acetate-EDTA buffer stained with ethidium bromide dye.

Amplified products from three replicates of each treatment were pooled in order to reduce stochastic PCR bias (Acinas et al., 2005) and purified using Promega Wizard

SV Gel and PCR Clean-Up System (Promega, Madison WI). Purified products were ligated into Invitrogen pCRII-TOPO cloning vector and transformed into chemically- competent Escherichia coli TOP10F' cells (Invitrogen, Carlsbad, CA). The resulting

118

clones were isolated on Luria Bertani agar plates with 50 µg mL-1 kanamycin at 37ºC according to the manufacturer’s instructions. Ninety-four randomly selected white colonies per sample were grown overnight in 2-mL 96-well microplates containing 1.5 mL Luria Broth media and 50 µg mL-1 kanamycin at 37ºC on an orbital shaker at 300

rpm.

Plasmids from clones were isolated in 96-well blocks with a Perfectprep Plasmid

96 Vac Direct Bind Kit (Eppendorf, Westbury, NY) and Biomek FX Laboratory

Automated Workstation (Beckman Coulter, Fullerton, CA). Nucleotide sequencing

reactions were performed with the vector M13F primer and ABI PRISM BigDye v3.1

Terminator Cycle Sequencing Kit (Applied Biosystems, Foster City, CA), and resulting

sequences were read with an ABI PRISM 3730 DNA analyzer at the University of

Kentucky Advanced Genetic Technology Center (Lexington, KY).

Reductive dehalogenase gene analysis

For the detection of reductive dehalogenase genes in this study several primer sets

were used. Similar PCR conditions as described before were used, except for the PCR

temperature program which is described after each primer set:

• DHAR1000F [G(A/T)A GCA GGT (C/T)TR GGA (G/C)AA] and DHU1350R

[CC(A/G) TAG CC(A/C/G) AA(G/T) AT(A/T) TCA TC(A/C) AT] developed based

on three known reductive dehalogenase (RDH) genes (pceA from Dehalospirillum

multivorans, cprA from Desulfitobacterium dehalogenans, tceA from

Dehalococcoides ethenogenes (Rhee et al., 2003) The following PCR thermocycling

program was used: 2 min of denaturation at 94°C, followed by 35 cycles of 30-s at

119

94°C, 1-min at 46°C, and 30-s at 67°C with a final elongation step of 7 min at 65ºC,

and finally cooling at 4°C.

• SpDr11F (TTG GAT GAG GCC TTG AAC GC) and SpDR9R (GCG CTG CAT

AAT AGC CAA GC), developed based on pceA gene from Dehalobacter restrictus

and Desulfitobacterium spp. TCE1, PCE-S, and Y51 (Regeard et al., 2004). The

following PCR thermocycling program was used: 3 min of denaturation at 94°C,

followed by 26 cycles of 30-s at 94°C, 30-s at 55°C, and 1 min at 72°C with a final

elongation step of 10 min, and finally cooling at 4°C.

• SpDr1F (CGT TGG ACC TAT TCC ACC TG) and SpDr1R (CAA GAA CGA AGG

CAA TCA CA) also based on pceA gene from Dehalobacter restrictus and

Desulfitobacterium spp. Y51 (AY013365) (Regeard et al., 2004). The following PCR

thermocycling program was used: 3 min of denaturation at 94°C, followed by 26

cycles of 30-s at 94°C, 30-s at 53°C, and 1 min at 72°C with a final elongation step of

10 min, and finally cooling at 4°C.

• SpSm1F (TCG TTG CAG GTA TCG CTA TG) and SpSm1R (TTC AAC AGC AAA

GGC AAC TG) based on pceA gene from Sulfospirillum multivorans. (AF022812)

(Regeard et al., 2004). The following PCR thermocycling program was used: 3 min of

denaturation at 94°C, followed by 26 cycles of 30-s at 94°C, 30-s at 52°C, and 1 min

at 72°C with a final elongation step of 10 min, and finally cooling at 4°C.

• SpDe1F (GCT TTG GCG GTG ATG ATA AG) and SpDe1R (GTT ATA GCC AAG

GCC TGC AA) based on tceA gene from Dehalococcoides ethenogenens

(AF228507) (Regeard et al., 2004). The following PCR thermocycling program was

used: 3 min of denaturation at 94°C, followed by 26 cycles of 30-s at 94°C, 30-s at

120

50°C, and 1 min at 72°C with a final elongation step of 10 min, and finally cooling at

4°C.

• DHC974F (GGG AGT ATC GAC CCT CTC) and DHC1212R (GGA TTA GCT

CCA GTT CAC ACT G) based on 16S rDNA sequences of Dehalococcoides spp.

(Hendrickson et al., 2002). The following PCR thermocycling program was used: 2

min of denaturation at 95°C, followed by either 30 cycles of 1 min at 94°C, 1 min at

55°C, and 1 min at 72°C and finally cooling at 4°C.

• cprA1023F (AAA CGA CGC GAT TAC CTT TG) and cprA1202R (CCT GCT TTA

TGG AAC CAG GA) based on the cprA gene from Desulfitobacterium dehalogenans

(AF115542) (This study). This set of primers was designed using Primer3Plus

(http://www.bioinformatics.nl/cgi-bin/primer3plus/primer3plus.cgi), and the PCR

thermocycling program used was as follows: 3 min of denaturation at 94°C, followed

by 26 cycles of 30-s at 94°C, 30-s at 54°C, and 1 min at 72°C with a final elongation

step of 10 min, and finally cooling at 4°C.

Phylogenetic analysis

Sequences were aligned against a phylogenetically-diverse set of 16S rRNA sequences in the GreenGenes database using the Nearest Alignment Space Termination

(NAST) algorithm (http://greengenes.lbl.gov/cgi-bin/nph-NAST_align.cgi) and Clustal

W algorithm available in MEGA4 (http://www.megasoftware.net/; Kumar et al., 2004).

Aligned sequences were evaluated for potential anomalies (e.g. chimeras) using the

Bellerophon Server (Huber et al., 2004, and were removed from the clone libraries.

Phylogenetic affiliations of 16S rDNA sequences were determined by the Basic Local

121

Alignment Search Tool (BLAST) 2.2.14 available at the National Center for

Biotechnology Information (NCBI) webpage (http://www.ncbi.nih.gov/BLAST/Blast.cgi) and the Classifier and Sequence Match tools available at Ribosomal Database Project II

(RDP II) Release 9.48 webpage (http://rdp.cme.msu.edu/index.jsp) (Cole et al., 2007).

Differences in community composition between libraries were determined using the

Library Compare tool available at RDP. Bacterial nomenclature followed Bergey's

Manual of Systematic Bacteriology Release 6.0.

Analytical methods

PCBs in cleaned up extracts were analyzed using a Shimadzu GC-14A gas

chromatograph equipped with 63Ni electron capture detector (320°C), autoinjector

(250°C), and RTX-1 capillary column (30 m by 0.32 mm with 0.25-µm phase thickness)

(Restek Corp., Bellefonte, PA) with the following temperature program: 100°C held for 1

min, ramp to 240°C at 3 min-1, and held at 240°C for 10 min. PCB congeners were identified by comparing retention times with authentic standards obtained from

Accustandard (New Haven, CT). Preliminary spike recovery experiments showed that this procedure yielded >95% recovery of PCBs from sediments.

Headspace methane was measured using a Shimadzu GC-14A gas chromatograph equipped with flame ionization detector (FID) at 160 ºC, with nitrogen as carrier gas and a stainless steel, 45/60 mesh-packing, Carboxen 1000 column (0.3 cm by 2 m) (Supelco,

Bellefonte, PA) maintained isothermally at 110 ºC. Headspace carbon dioxide was measured using a Shimadzu GC-8A gas chromatograph equipped with thermal conductivity detector (30 ºC), with helium as carrier gas and stainless steel Porapak N

122

column (0.3 cm by 2 m) (Supelco, Bellefonte, PA) maintained isothermally at 25 ºC.

Sulfate was analyzed by ion chromatography using a Shimadzu (Columbia, MD) SLC-

10A system controller with the following modules: LC-10AD solvent delivery module,

FCV-10AL gradient flow controller, CDD-6A conductivity detector, all from Shimadzu, and a Model 335 solid phase chemical suppressor module from Alltec (Deerfield, IL).

Separation was achieved under isocratic conditions with carbonate buffer (1.8 mM carbonate/1.7 mM bicarbonate) in an IonPac AS4A guard colum (4x50 mm)/analytical colum (4x250 mm) system (Dionex Corp. Sunnyvale, CA).

Radioactivity trapped in 1 M NaOH traps was determined by mixing a 500 µL

aliquot with scintillation cocktail and measuring activity using a Packard 2200 CA Tri-

Carb liquid scintillation counter (Downers Grove, IL). Radioactivity in PUF plugs was determined by extracting the PUF plug with acetone and mixing with scintillation cocktail as described above.

Results and discussion

Sequential anaerobic aerobic treatment of PCB-153

Anaerobic treatment of PCB-153 occurred with conversions between 60 and 68%

at the end of the 150 and 250 days, respectively (Fig. 4.1). PCB-99 and PCB-47 were the main detected dechlorination products, with a greater accumulation of PCB-47. There is no evidence that reductive dechlorination reactions removed two chlorine atoms per molecule in a single step. However, when PCB-99 is formed, the molecule might still be in close proximity to the catalytic active site of the reductive dehalogenase

123

270

240 PCB-47 PCB-99 210 PCB-153 ) 1 - 180

150

120

90 PCB concentration (nmol g (nmol concentration PCB 60

30

0 200 d 250 d 150 d 200 d 250 d 150 d 200 d 250 d Anaerobic/aerobic Anaerobic Aerobic Treatment

Figure 4.1 PCB-153 conversion under anaerobic, aerobic, and sequential anaerobic/aerobic conditions. Initial PCB-153 concentration was 270 nmol/g on a dry

basis. Each value represents the mean of three replications ± one standard deviation.

124

(RDH), thus decreasing diffusion problems and energetic barriers associated with

transport of these molecules to the active site. In other words, transport to the RDH active

site is a limiting step for PCB-153, but not for PCB-99, and that might be the reason for a greater accumulation of PCB-47. Under aerobic conditions, there were no significant changes on PCB-153 concentration. This congener does not have free positions 2 and 3, or 3 and 4, for a potential biphenyl dioxygenase attack, and therefore it is particularly resistant to aerobic degradation. However, meta dechlorination provides free 2,3 positions in one ring for PCB-99, and in both rings for PCB-47, making them amenable for aerobic degradation in these sediments. At the end of the 250 days period on the sequential anaerobic-aerobic treatment, PCB-47 concentration appears to be smaller than at 200 days, but not significantly different. However, from PCB-153 concentration it can be concluded that its concentration is smaller in this set of microcosms. A mass balance on total PCB molar concentration indicates that initially there were 250-270 nmol PCBs per gram of dry sediment. But after 250 days of treatment, the anaerobic-aerobic treatment contained only 200 nmol per gram. Assuming that PCB-153 was not lost under aerobic conditions, this suggests that lower chlorinated products were transformed under aerobic conditions. Analysis of the PUF and aqueous NaOH traps, however, did not reveal any detectable radiation, which suggested that any aerobic transformation products remained in the system and might correspond to hydroxylated PCBs or chlorobenzoic acids that are not detected by regular PCB extraction and analysis procedures.

125

Detection of reductive dehalogenase genes

PCR amplification reactions using the degenerate primer set DHAR1000F-

DHU1350R did not show the expected 450 bp fragment in the anaerobic treatment.

Therefore this set was tested with DNA isolated from pure cultures of known dehalogenators such as Desulfitobacterium spp., Dehalobacter restrictus,

Desulfuromonas michinagensis, and Geobacter lovleyi (Fig. 4.2). Only

Desulfitobacterium hafniense and Dehalobacter restrictus yield strong bands for the expected PCR product while Desulfitobacterium dehalogenans yields a faint band.

Therefore it is possible that these primers might not capture all the RDH genes present in the anaerobic microcosms. In addition, PCR products were not detected using more specific primers targeting individual RDH genes such as SpDr11F/SpDr9R,

SpDr1F/SpDr1R, and SpSm1F/SpSm1R, and SpDe1F/SpDe1R, which suggested organisms such as D. restrictus, S. multivorans, and Dehalococcoides spp. were not present in these sediments. These results were surprising considering that

Dehalococcoides spp. were detected using 16S rDNA based primers

DHC974F/DHC1212R that were designed to detect these species (Fig. 4.3). Interestingly, this primer set also provided positive results for microcosms exposed to aerobic conditions suggesting that Dehalococcoides can survive under aerobic conditions for prolonged periods.

PCR amplification using primer pair cprA1023F/cprA1202R was successful at detecting RDH genes corresponding to cprA, which is consistent with the detection of

Desulfitobacterium spp. in the sediment clone libraries (Fig. 4.4). Although cprA has not been shown to be involved in the reductive dechlorination of PCBs, it has been shown to

126

Figure 4.2 PCR amplification of reductive dehalogenase DNA from pure cultures

(Primers DHAR1000F and DHAR1350R)

127

Figure 4.3 PCR amplification products of 16S rDNA of Dehalococcoides spp.

with primers DHC974F/DHC1212R

128

Figure 4.4 PCR amplification products of cprA gene of Desulfitobacterium

dehalogenans with primers cprA1023F/cprA1202R

129

be important in reductive dechlorination of other chlorinated aromatic compounds including hydroxylated PCBs and chlorophenols (Villemur et al., 2006).

Changes in microbial community structure under anaerobic and aerobic conditions

There were significant differences in microbial community structure in sediments under different redox conditions. The greatest differences were between the anaerobic treatment and treatments that contained an aerobic phase. For example, the clone library from the 250 d anaerobic treatment (Table 4.1) was enriched in bacteria in the phyla

Bacteroidetes (20% vs 3-5% in the aerobic treatments), Chloroflexi (18% vs 6-9%), and

Firmicutes (26% vs 9%) but it was relatively depleted in bacteria from the phylum

Proteobacteria (22% vs 52-56%). The enrichment/depletion pattern observed in the anaerobic clone library is similar to clone libraries from anaerobic treatments using organic amendments shown in Chapter 3. Additionally, the observed pattern for the aerobic clone libraries is similar to the freshly collected sediment shown in Chapter 3, suggesting that river sediments were mostly under aerobic conditions at the time of collection.

Several clones closely related to reductive dechlorinators were detected in sediments incubated anaerobically for 250 days (Fig. 4.5). These results further validate the use of the clone library approach and bacterial groups to indicate redox conditions in aquatic sediments. About 19 clones closely related to species such as Desulfitobacterium chlororespirans and Desulfitobacterium hafniense Y51, and several clones related to

130

Table 4.1 Distribution of bacteria in major taxonomic groups in libraries from Ohio River sediments during sequential anaerobic-

aerobic treatment of PCB-153 for 250 days, as determined by the Classifier tool available at the Ribosomal Database Project (RDP)

(Cole et al., 2007).

Anaerobic/ Aerobic Anaerobic Phylum Class Order Aerobic Representative genera (n=206) (n=247) (n=196) %

Acidobacteria Acidobacteria Acidobacteriales 2.9 3.1 1.6

Actinobacteria Actinobacteria 2.9ab 5.1a 1.2b

Acidimicrobiales 1.0 1.5 0.4 Acidimicrobium 131

Clavibacter, Arthrobacter, Actinomycetales 1.5 1.5 0.4 Thermomonospora Coriobacteriales 0.5 1.0 0.0

Rubrobacterales 0.0 1.0 0.4

Bacteroidetes 3.4a 4.6a 19.8b

Bacteroidetes Bacteroidales 1.0a 0.5a 16.2b

Flavobacteria Flavobacteriales 0.5 0.0 1.2 Flavobacterium

Sphingobacteria Sphingobacteriales 1.9 4.1 2.4 Flexibacter, Cytophaga, Lewinella

BRC1 0.5 0.0 0.0

Table 4.1. (continued) Anaerobic/ Aerobic Anaerobic Phylum Class Order Aerobic Representative genera (206) (247) (196) %

Chlamydiae Chlamydiales 0.5 0.0 0.0

Chloroflexi Anaerolineae 6.3a 8.7a 18.2b

Anaerolinaeles 0.0 0.5 1.6 Anaerolinea, Bellilinea, Longilinea

Caldilineales 6.3a 8.2a 16.6b Levilinea, Caldilinea

Crenarchaeota Thermoprotei Thermoproteales 0.0 0.0 0.4

132 Cyanobacteria Chloroplast 0.5 0.0 0.0

Deferribacteres Deferribacteres Deferribacterales 0.0 0.0 2.4

Dehalococcoides 0.0 0.0 1.2 Dehalococcoides

Deinococcus-Thermus Deinococci 1.5 0.5 0.0

Deinococcales 0.0 0.5 0.0

Thermales 1.5 0.0 0.0

Dictyoglomi Dictyoglomi Dictyoglomales 1.0 0.0 0.0

Firmicutes 8.7a 9.2a 25.5b

Table 4.1. (continued) Anaerobic/ Aerobic Anaerobic Phylum Class Order Aerobic Representative genera (206) (247) (196) %

Firmicutes Bacilli Bacillales 0.5 0.5 0.0 Bacillus

Clostridia 8.3a 8.7a 25.1b

Sporacetigenium, Clostridium, Clostridiales 4.9a 4.6a 19.4b Eubacterium, Desulfosporosinus,

Desulfitobacterium 3.4 4.1 5.7 Thermodesulfobium

Erysipelotrichi Erysipelotrichales 0.0 0.0 0.4

133 Gemmatimonadetes Gemmatimonadetes Gemmatimonadales 3.4a 1.0b 1.2b Gemmatimonas

Lentisphaerae Lentisphaerae 1.0 0.0 0.0

Lentisphaerales 0.5 0.0 0.0

Victivallales 0.5 0.0 0.0

Nitrospira Nitrospira Nitrospirales 1.5 1.0 2.0 Magnetobacterium, Thermodesulfovibrio

OD1 1.0 4.1 0.8

OP10 2.9ab 5.6a 0.8b

Planctomycetes Planctomycetacia Planctomycetales 3.9 2.6 2.0 Pirellula, Planctomyces

Table 4.1. (continued) Anaerobic/ Aerobic Anaerobic Phylum Class Order Aerobic Representative genera (206) (247) (196) %

Proteobacteria 55.8a 52.0a 21.5b

Alphaproteobacteria 15.5a 10.2b 2.4c

Parvularculales 0.5 1.5 0.0

Rhodopseudomonas, Bradyrhizobium, Rhizobiales 4.4 3.6 2.0 Ochrobactrum, Hyphomicrobium,

Methylobacterium Pannonibacter, Rhodobacter, Thioclava, Rhodobacterales 1.9 0.5 0.0 134 Paracoccus Phaeospirillum, Azospirillum, Rhodospirillales 6.3a 3.1ab 0.4b Magnetospirillum Rickettsiales 1.0 0.0 0.0

Porphyrobacter, Erythrobacter, Sphingomonadales 1.5 1.5 0.0 Novosphingobium, Sphingomonas Betaproteobacteria 11.2a 18.9b 2.0c

Burkholderiales 6.3 4.6 2.0 Cupriavidus, Burkholderia

Hydrogenophilales 1.5a 8.7b 0.0a Thiobacillus

Nitrosomonadales 1.0 3.6 0.0 Gallionella, Nitrosospira, Nitrosomonas

Rhodocyclales 2.4 2.0 0.0 Azoarcus, Zoogloea, Sterolibacterium

Table 4.1. (continued) Anaerobic/ Aerobic Anaerobic Phylum Class Order Aerobic Representative genera (206) (247) (196) %

Proteobacteria Deltaproteobacteria 19.4a 9.7b 15.4ab

Bdellovibrionales 0.5 0.0 0.0

Desulfobacterales 1.9 1.0 1.6 Desulfatibacillum, Desulfobulbus

Desulfovibrionales 0.5 0.0 1.2

Desulfurellales 0.0 0.0 1.2

135 Desulfuromonales 6.3a 2.6b 0.8b Desulfuromonas, Geobacter, Pelobacter

Hyalangium, Anaeromyxobacter, Myxococcales 8.7a 4.6a 2.8b Myxococcus, Sorangium, Chondromyces Syntrophus, Smithella, Desulfobacca, Syntrophobacterales 1.5a 1.5a 7.7b Desulfomonile, Syntrophobacter Epsilonproteobacteria 0.0 2.0 0.4

Campylobacterales 0.0 2.0 0.0 Sulfuricurvum

Nautiliales 0.0 0.0 0.4

Gammaproteobacteria 9.7a 11.2a 1.2b

Aeromonadales 0.0 0.5 0.0

Table 4.1. (continued) Anaerobic/ Aerobic Anaerobic Phylum Class Order Aerobic Representative genera (206) (247) (196) %

Proteobacteria Gammaproteobacteria Chromatiales 1.9 5.1 0.8

Legionellales 0.0 0.0 0.4 Legionella

Oceanospirillales 1.0 0.5 0.0

Pseudomonadales 3.9a 0.0b 0.0b Pseudomonas

Thiotrichales 1.9ab 4.6a 0.0b Achromatium, Beggiatoa

136 Siderooxidans, Lysobacter, Xanthomonadales 1.0 0.5 0.0 Pseudoxanthomonas

Spirochaetes Spirochaetes Spirochaetales 0.0 0.0 0.8

Thermodesulfobacteria Thermodesulfobacteria Thermodesulfobacteriales 0.0 0.5 0.4

Thermomicrobia Thermomicrobia Thermomicrobiales 0.0 0.5 0.0

Verrucomicrobia Verrucomicrobiae Verrucomicrobiales 1.5 1.5 0.0 Verrucomicrobium

WS3 1.0 0.0 0.0

Values followed by a different letter are significantly different at p= 0.05, Number in parentheses indicate number of sequences in clone library.

99 19 clones 97 70 Desulfitobacterium chlororespirans (U68528) 62 100 Desulfitobacterium hafniense Y51 (AP008230) 97 Dehalobacter restrictus (Y10164) Anaerobic - 275 Bacillus sp. OUCZ63 (AY785743) 99 Anaerobic - 004 61 Anaerobic - 114 Anaerobic - 058 Rhodococcus ruber strain Z57 (DQ858963) 73 100 Rhodococcus erythropolis strain D14 (DQ858961) 91 Rhodococcus sp. OUCZ204 (AY785749) Anaerobic - 143 95 Anaerobic - 099 Anaerobic - 128 99 Dehalobium chlorocoercia DF-1 (AF393781) Anaerobic - 186 99 Dehalococcoides ethenogenens (AF004928) 100 Dehalococcoides sp. CDBD1 (AF230641) 50 62 Dehalococcoides sp. BAV1 (AY165308) Anaerobic - 063 81 Anaerobic - 136 100 Sphingomonas sp. R1 (DQ858957) Sphingomonas aromaticivorans SMCC F19... 100 Desulfuromonas michiganensis BB1 (AF357915) 100 Desulfuromonas michiganensis BRS1 (AF357914) 97 Desulfuromonas chloroethenica (U49748) Trichlorobacter thiogenens (AF223382) 78 Anaerobic - 077 Desulfomonile tiedjei (AM086646) 57 Anaerobic - 181 Anaerobic - 177 100 Anaeromyxobacter dehalogenans 2CP-1 (AF382396) Anaerobic - 270 99 Anaerobic - 133 Anaerobic - 146 51 100 Pseudomonas sp. SA-6 (DQ854841) 75 Pseudomonas aeruginosa strain SA-1 (DQ854840) 100 55 Pseudomonas sp. JN18 A17 R (DQ168645) Pseudomonas pseudoalcaligenes (AB257323) Anaerobic - 168 Anaerobic - 040 99 Burkholderia sp. JB1 (X92188) 100 Burkholderia xenovorans LB400 (U86373) Comamonas testosteroni (AM937260) 4 clones 100 Dehalospirillum multivorans (X82931) Desulfovibrio dechloracetivorans (AF230530)

0.05

Figure 4.5 Bootstrapped minimum-evolution phylogenetic tree of 16S rDNA sequences

isolated from Ohio River sediments incubated anaerobically for 250 days, as well as

known aerobic PCB degraders, and dehalogenating bacteria. Numbers at branch points

are percentages of 500 resampling bootstrap analyses. The scale bar represents 0.05

changes per nucleotide.

137

other dehalogenators such as Desulfomonile tiedjei and Anaeromyxbacter dehalogenans, and Dehalococcoides spp. were detected. As mentioned previously Dehalococcoides spp. are the only recognized PCB dechlorinators, and because the number of clones is small, no significant differences were detected between the treatments. Evidence indicates that both Desulfitobacterium and Dehalococcoides spp. are present and most probably they are forming part of a consortia that reductively dechlorinates PCBs in Ohio

River sediments. This tree also indicates that there are clones related to aerobic PCB degraders such as Pseudomonas, Burkholderia, and Comamonas spp. Because of redox conditions, aerobic degradation of PCB-153 was not an important process in this treatment, however the presence of these organisms suggests that they can tolerate and survive under these conditions and develop into an aerobic PCB degrading consortium, once appropriate conditions are established.

Phylogenetic trees were also constructed for the aerobic and sequential anaerobic- aerobic treatment of PCB-153 (Figs. 4.6 and 4.7 respectively). In the aerobic treatment, there were more clones related to aerobic PCB degraders than related to reductive dechlorinators. There were approximately 28 clones associated with aerobic PCB degraders mentioned previously, while the presence of closely related dechlorinators was scarce. In this case, presence of bacteria related to dechlorinators suggests tolerance to aerobic conditions, presumably by inhabiting anaerobic microsites that persist during the aerobic treatment. A similar result was observed for the sequential anaerobic-aerobic treatment, in which a considerably higher number of clones related to aerobic PCB degraders were present (11 clones), when compared to the anaerobic treatment (5 clones).

138

66 12 clones 100 5 clones 98 Aerobic - 109 Aerobic - 185 55 Aerobic - 224 Comamonas testosteroni (AM937260) Burkholderia sp. JB1 (X92188) 100 Burkholderia xenovorans LB400 (U86373) Aerobic - 072 Pseudomonas pseudoalcaligenes (AB257323) 52 100 Pseudomonas sp. SA-6 (DQ854841) 100 Pseudomonas aeruginosa strain SA-1 (DQ854840) 98 78 Aerobic - 119 Pseudomonas sp. JN18 A17 R (DQ168645) 97 6 clones 83 Aerobic - 236 100 Aerobic - 040 100 Sphingomonas sp. R1 (DQ858957) 62 Sphingomonas aromaticivorans SMCC F19... Aerobic - 247 100 Aerobic - 122 100 Aerobic - 128 74 Anaeromyxopbacter dehalogenans 2CP-1 (AF382396) 75 Aerobic - 261 51 Desulfomonile tiedjei (AM086646) 99 Aerobic - 053 58 Aerobic - 094 100 Trichlorobacter thiogenens (AF223382) Aerobic - 003 99 Aerobic - 146 99 100 Aerobic - 227 Aerobic - 136 99 Desulfuromonas chloroethenica (U49748) 99 Desulfuromonas michiganensis BB1 (AF357915) 100 Desulfuromonas michiganensis BRS1 (AF357914) 96 Dehalococcoides sp. BAV1 (AY165308) 100 Dehalococcoides sp. CDBD1 (AF230641) 79 90 Dehalococcoides ethenogenens (AF004928) 99 Aerobic - 173 Dehalobium chlorocoercia DF-1 (AF393781) Aerobic - 201 97 Rhodococcus erythropolis strain D14 (DQ858961) 100 Rhodococcus sp. OUCZ204 (AY785749) Rhodococcus ruber strain Z57 (DQ858963) 100 Desulfitobacterium chlororespirans (U68528) 94 Desulfitobacterium hafniense Y51 (AP008230) 93 71 Dehalobacter restrictus (Y10164) Bacillus sp. OUCZ63 (AY785743) 99 89 Aerobic - 093 Aerobic - 160 57 Aerobic - 163 Aerobic - 035 100 100 Aerobic - 189 97 Aerobic - 061 100 Aerobic - 279 Desulfovibrio dechloracetivorans (AF230530) Dehalospirillum multivorans (X82931)

0.05

Figure 4.6 Bootstrapped minimum-evolution phylogenetic tree of 16S rDNA sequences isolated from Ohio River sediments incubated aerobically for 250 days, as well as known

aerobic PCB degraders, and dehalogenating bacteria. Numbers at branch points are percentages of 500 resampling bootstrap analyses. The scale bar represents 0.05 changes

per nucleotide.

139

94 8 clones 93 99 Burkholderia xenovorans LB400 (U86373) 94 Burkholderia sp. JB1 (X92188) Anaerobic/Aerobic - 146 85 91 Anaerobic/Aerobic - 149 Comamonas testosteroni (AM937260) 100 87 Anaerobic/Aerobic - 020 Pseudomonas pseudoalcaligenes (AB257323) 79 Pseudomonas sp. JN18 A17 R (DQ168645) 100 Pseudomonas aeruginosa strain SA-1 (D... 92 100 Pseudomonas sp. SA-6 (DQ854841) Anaerobic/Aerobic - 165 Anaerobic/Aerobic - 236 85 94 .Sphingomonas aromaticivorans SMCC F199 (U20756) 50 Sphingomonas sp. R1 (DQ858957) 99 Anaerobic/Aerobic - 266 100 Anaerobic/Aerobic - 200 100 Desulfuromonas michiganensis BRS1 (AF357914) 100 Desulfuromonas michiganensis BB1 (AF357915) 86 Desulfuromonas chloroethenica (U49748) 56 72 Trichlorobacter thiogenens (AF223382) Desulfomonile tiedjei (AM086646) Anaerobic/Aerobic - 257 Anaerobic/Aerobic - 198 74 Anaerobic/Aerobic - 124 51 97 3 clones 94 Anaerobic/Aerobic - 117 54 Anaeromyxobacter dehalogenans 2CP-1 (AF382396) Anaerobic/Aerobic - 150 63 6 clones 58 92 Rhodococcus sp. OUCZ204 (AY785749) 100 Rhodococcus erythropolis strain D14 (DQ858961) 100 Rhodococcus ruber strain Z57 (DQ858963) 88 Anaerobic/Aerobic - 109 Anaerobic/Aerobic - 120 Anaerobic/Aerobic - 016 59 61 4 clones Anaerobic/Aerobic - 071 Bacillus sp. OUCZ63 (AY785743) Dehalobacter restrictus (Y10164) 79 100 Desulfitobacterium hafniense Y51 (AP008230) 88 Desulfitobacterium chlororespirans (U68528) 97 5 clones 95 100 Anaerobic/Aerobic - 118 Anaerobic/Aerobic - 244 Dehalobium chlorocoercia DF-1 (AF393781) 76 Dehalococcoides ethenogenens (AF004928) 100 Dehalococcoides sp. BAV1 (AY165308) 100 88 Dehalococcoides sp. CDBD1 (AF230641) Dehalospirillum multivorans (X82931) Anaerobic/Aerobic - 230 99 Anaerobic/Aerobic - 264 90 Anaerobic/Aerobic - 119 100 100 Anaerobic/Aerobic - 258 Desulfovibrio dechloracetivorans (AF230530)

0.02

Figure 4.7 Bootstrapped minimum-evolution phylogenetic tree of 16S rDNA sequences

isolated from Ohio River sediments incubated anaerobically for 150 days and then aerobically for 100 days, as well as known aerobic PCB degraders, and dehalogenating

bacteria. Sediments Numbers at branch points are percentages of 500 resampling

bootstrap analysis. The scale bar represents 0.02 changes per nucleotide

140

In conclusion, the anaerobic treatment of PCB-153 produced PCB-99 and PCB-47

as dechlorination products, while its aerobic treatment did not cause any changes, despite

the presence of bacteria that were closely related to aerobic PCB degraders. The

dechlorination is also consistent with the presence of microorganism related to known

dehalogenating bacteria such as Desulfitobacterium and Dehalococcoides spp. which was confirmed not only by phylogenetic analysis of 16S rDNA sequences but also by specific

PCR primer design targeting both ribosomal and reductive dehalogenase genes for these species. The sequential anaerobic-aerobic treatment of PCB-153 suggests that aerobic degradation of PCB dechlorination products is possible after changing conditions from anaerobic to aerobic, but it was not conclusive. Therefore additional analysis is required with particular attention to PCB metabolites such as hydroxylated PCBs to determine the true potential of sequential anaerobic-aerobic treatment of PCBs in Ohio River sediments.

141

CHAPTER 5

SUMMARY AND CONCLUSIONS

In situ studies were conducted to evaluate factors affecting the natural attenuation

of PCBs in Ohio River sediments and local microbial community composition (Chapter

2). Laboratory studies were conducted to determine temperature and nutrient availability

effects on PCB reductive dechlorination rates (Chapter 3) and redox conditions effects on

PCB bioremediation (Chapter 4). During the course of this project it was shown that PCB

reductive dechlorination is a feasible process in Ohio River sediment microcosms, under

controlled conditions of nutrient availability, temperature and redox status.

Unfortunately, local conditions in the river are not favorable, mainly because of limiting

factors such as organic carbon, phosphorus or nitrogen, and also because temperatures in

the river are below optimal values during most of the year.

Are local conditions conducive to natural attenuation of PCBs in Ohio River

sediments? PCB losses during the in situ study were negligible. The chemical profile

determined at this location indicated a redox gradient from denitrifying to methanogenic

conditions, but nutrients such as ammonium, phosphate and organic carbon were present

in limited supplies. 16S rRNA clone libraries obtained from these sediments indicated the

prevalence of aerobic bacteria such as Sphingomonas, Pseudomonas and Burkholderia spp. supporting the presence of a rather oxidized environment, but also indicated the presence of potential aerobic PCB degraders that might contribute to natural attenuation of PCBs. Known dehalogenating bacteria such as Dehalococcoides, Desulfitobacterium,

Clostridium and Bacillus spp. were also detected in clone libraries from sediments,

142

indicating that reductive dechlorination can occur. However, cold temperatures during most of the year, limited availability of nutrients in the river, and rather oxidized sediments have created a set of unfavorable conditions that restrict natural attenuation and therefore it would not be an important process for PCBs in the Ohio River.

Did temperature and carbon amendements influence PCB reductive dechlorination rates and microbial community structure in Ohio River sediments?

Cold (10 ºC) and high (40 º) temperatures cause a negative impact on dechlorination rates. Cold temperatures completely inhibited PCB dechlorination while dechlorination at high temperatures was reduced to only 10% compared to the 25 ºC treatment. Microbial community structure also changed under the influence of temperature. Low temperatures caused enrichment in the Alphaprotobacteria class and particularly in the Rhizobiales order but a decrease in the phylum Bacteroidetes. On the other hand, high temperatures caused Bacteroidetes to increase and Alphaproteobacteria to decrease, so Bacteroidetes is proposed as an indicator species for temperature conditions in the river. Carbon amendments caused increased bacteria in the phyla

Firmicutes and Chloroflexi but decreased bacteria in the phylum Bacteroidetes, thus, it is proposed that bacteria in these groups can serve as indicators for organic carbont availability in the river. PCB reductive dechlorination rates were also greater in anaerobic treatments with carbon amendments, and, consequently, it is proposed that the Firmicutes may play an important role in these reactions. However, the role as active PCB dechlorinators is still unclear as Dehalococcoides spp. were also detected in these treatments and they are known to reductively dechlorinate PCBs.

143

Is it possible to achieve complete removal of PCBs through a sequential anaerobic-aerobic treatment?

Control over redox status in Ohio River sediment microcosms improved PCB reductive dechlorination rates. However, this process alone cannot completely remove

PCB from the environment. Laboratory studies showed that PCB-153 reductive dechlorination rates can be improved by adding organic amendments with an average of two chlorine atoms removed per molecule. However, aerobic degradation of its dechlorination products was not detected to a great extent, and additional work may be required to focus on potential metabolites to determine whether the biphenyl ring is undergoing oxygenation and possibly ring-opening reactions. DNA analysis showed that sediments with high PCB dechlorination are associated with enrichment of bacteria from phyla Bacteroidetes, Chloroflexi and Firmicutes, but with a decrease in the population of

Proteobacteria. Particularly Desulfitobacterium spp. were detected in high number in the clone libraries, but also Dehalococcoides spp, were detected, suggesting that these species are actively involved in the reductive dechlorination of PCBs. Dehalococcoides spp. have been shown to dechlorinate PCBs, but the role of Desulfitobacterium spp. might still be in debate because these microorganisms have an aromatic reductive dehalogenase that might also be directly involved in PCB dechlorination. This study also showed the presence of microorganism related to aerobic PCB degraders in these sediments that, although aerobic degradation products were not detected, may accomplish the ultimate goal of complete mineralization of PCBs in the environment.

144

Future Directions

Up to this point, primer design for the detection of RDH genes has been limited for the small number of sequences available in the current databases to describe these genes. However, more studies dedicated to characterization of these enzymes will also increase the number of sequences available, and thus will allow more efficient primer design that might more successful in capturing the diversity of dechlorinating bacteria in aquatic environments, such as the Ohio River. Using such primers at different locations in the river, especially those located near “hot spots” or power generation plants, might reveal not only the diversity of dechlorinating bacteria in this ecosystem but also the potential differences in the spatial distribution of these organisms along the river.

145 REFERENCES

Abraham, W., B. Nogales, P.N. Golyshin, D.H. Pieper, K.N. Timmis, 2002.

Polychlorinated biphenyl-degrading microbial communities in soils and

sediments, Curr. Opin. Microbiol. 5: 246-253.

Abramowicz, D.A., 1995 Aerobic and anaerobic PCB biodegradation in the

Environment. Environ Health Perspect 103(Suppl 5):97-99.

Acinas, S.G., R. Sarma-Rupavtarm, V. Klepac-Ceraj, and M.F. Polz, 2005. PCR-induced

sequence artifacts and bias: insights from comparison of two 16S rRNA clone

libraries constructed from the same sample. App. Environ. Microbiol. 71: 8966-

8969.

Agency for Toxic Substances and Disease Registry (ATSDR), 2001. Toxicological

profile for polychlorinated biphenyls (PCBs). Agency for Toxic Substances &

Disease Registry. (Available at http://www.atsdr.cdc.gov/tfacts17.html, last

retrieved on May 5th, 2008)..

Anderson, M.J., 2001. A new method for non-parametric multivariate analysis of

variance. Aust. Ecol. 26:32-46.

Anderson, P.N., and R.A. Hites, 1996. OH radical reactions: The major removal pathway

for polychlorinated biphenyls from the atmosphere. Environ. Sci. Technol.

30:1756-1763.

Annuar, M.S.M., I.K.P. Tan, S. Ibrahim, and K.B. Ramachandran, 2006. Ammonium

uptake and growth kinetics of Pseudomonas putida PGA1. AsPac J. Mol. Biol.

Biotechnol. 14:1-10.

146 Becker, J.G., G. Berardesco, B.E. Rittmann, and D.A. Stahl, 2005. The role of syntrophic

associations in sustaining anaerobic mineralization of chlorinated organic

compounds. Environ. Health Perspect. 113: 310-316.

Bedard, D. L. 2003. Polychlorinated biphenyls in aquatic sediments: Environmental fate

and outlook for biological treatment p. 443-465. In M. M. Häggblom and I. D.

Bossert (eds), Dehalogenation: Microbial Processes and Environmental

applications,. Kluwer Academic Publishers. New York, NY.

Bedard, D.L., M.L. Haberl, R.J. May, and M.J. Brennan, 1987. Evidence for novel

mechanisms of polychlorinated biphenyl metabolism in Alcaligenes eutrophus

H850. Appl Environ Microbiol. 53:1103-1112.

Bedard D.L., and R. J. May, 1996. Characterization of the polychlorinated biphenyls

(PCBs) in the sediments of Woods Pond: Evidence for microbial dechlorination in

situ. Environ. Sci. Technol., 30:237-245.

Bedard, D.L., and J.F. Quensen III, 1995. Microbial reductive dechlorination of

polychlorinated biphenyls, p. 127-216. In L.Y. Young and C.E. Cerniglia (eds.),

Microbial transformations and degradation of toxic organic chemicals. John

Wiley & Sons, Inc., New York, N.Y.

Bedard, D. L., K. M. Ritalahti, and F. E. Löffler, 2007. The Dehalococcoides population

in sediment-free mixed cultures metabolically dechlorinates the commercial

polychlorinated biphenyl mixture Aroclor 1260. App. Environ. Microbiol.

73:2513-2521.

Benvinakatti, B.G., and H.Z. Ninnekar, 1992. Degradation of biphenyl by a

species. Appl. Microbiol. Biotechnol. 38: 273-275.

147 Borja, J., D.M. Taleon, J. Auresenia, and S. Gallardo, 2005. Polychlorinated biphenyls

and their biodegradation. Process Biochem. 40:1999-2013.

Boyle, A.W., C.J. Silvin, J.P. Hassett, J.P. Nakas,.and S.W. Tanenbaum, 1992. Bacterial

PCB biodegradation. Biodegradation 3: 285-298.

Brown, J. F., D. L. Bedard, M. J. Brennan, J. C. Carnahan, H. Feng, and R. E. Wagner.

1987. Polychlorinated biphenyl dechlorination in aquatic sediments. Science

236:709-712.

Burr, M.D., S.J. Clark, C.R. Spear, and A.K. Camper, 2006. Denaturing gradient gel

electrophoresis can rapidly display the bacterial diversity contained in 16S rDNA

clone libraries. Microb. Ecol. 51:479-486.

Carlson, D.L., and R.A. Hites, 2005. Temperature dependence of atmospheric PCB

concentrations. Environ. Sci. Technol. 39:740-747.

Chang, C., S. Tseng, C. Chang, and C.Ho, 2004. Degradation of 2-chlorophenol via a

hydrogenotrophic biofilm under different reductive conditions, Chemosphere

56:989-997.

Chen, W., S. Chen, S.K. Khanal, S. Sung, 2006. Kinetic study of biological hydrogen

production by anaerobic fermentation. Int. J. Hydrogen Energy 31:2170-2178.

Chin, K.J., and R. Conrad, 1995. Intermediary metabolism in methanogenic paddy soil

and the influence of temperature. FEMS Microbiol. Ecol. 18: 85-102.

Cole, J.R., B. Chai, R.J Farris, Q. Wang, A.S. Kulam-Syed-Mohideen, and D.M.

McGarrell, 2007. The ribosomal database project (RDP-II): Introducing myRDP

space and quality controlled public data. Nucleic Acids Res 35: D169-D172.

148 Cutter, L.A., J.E.M. 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.

D'Angelo, E.M., J. Crutchfield, and M. Vandiviere, 2001. Rapid, sensitive, microscale

determination of phosphate in water and soil. J. Environ. Qual. 30:2206-2209.

Denger, K., R. Warthmann, W. Ludwig and B. Schink, 2002. Anaerophaga

thermohalophila gen. nov., sp. nov., a moderately thermohalophilic, strictly

anaerobic fermentative bacterium. Int. J. Sys. Evol. Microbiol. 52: 173-178.

Denison, M.S., and S.R. Nagy, 2003. Activation of the aryl hydrocarbon receptor by

structurally diverse exogenous and endogenous chemicals. Annu. Rev. Pharmacol.

Toxicol. 43:309-334.

Dolfing, J., 2003. Thermodynamic considerations for dehalogenation. In M.M. Häggblom

and I.D. Bossert eds. Dehalogenation: Microbial processes and environmental

applications pp 89-114. KluwerAcademic Publishers New York NY.

Dolfing. J., and B. K. Harrison, 1992. Gibbs free energy of formation of halogenated

aromatic compounds and their potential role as electron acceptors in anaerobic

environments. Environ. Sci. Technol. 26:2213-2218.

Drzyzga1, O. and J.C. Gottschal, 2002. Tetrachloroethene dehalorespiration and growth

of Desulfitobacterium frappieri TCE1 in strict dependence on the activity of

Desulfovibrio fructosivorans. Appl. Environ. Microbiol. 68: 642-649.

Eichorst, S.A., J.A. Breznak, and T.M. Schmidt, 2007. Isolation and characterization of

soil bacteria that define Terriglobus gen. nov., in the phylum Acidobacteria. Appl.

Environ. Microbiol. 73:2708-2717.

149 Elferink O.S.J., V.W.M. van Akkermans, J.J Bogte, A.J. Stams, 1999. Desulfobacca

acetoxidans gen. nov., sp. nov., a novel acetate-degrading sulfate reducer isolated

from sulfidogenic granular sludge. Int J Syst Bacteriol. 49:345-350.

Erickson, M., 1997. Analytical chemistry of PCBs, 2nd ed. CRC Press LLC, Boca Raton,

FL. USA.

Evans, B.S., C.A. Dudley, and K.T. Klasson, 1996. Sequential anaerobic-aerobic

biodegradation of PCBs in soil slurry microcosms. Appl. Biochem. Biotechnol.

57-58:885-894.

Fazi, S., S. Amalfitano, J. Pernthaler, A. Puddu, 2005. Bacterial communities associated

with benthic organic matter in headwater stream microhabitats. Environ.

Microbiol. 7: 1633-1640.

Fennel, D.E., I. Nijenhuis, S.F. Wilson, S.H. Zinder, and M.M. Haggblom, 2004.

Dehalococcoides ethenogenens strain 195 reductively dechlorinates diverse

chlorinated aromatic pollutants. Environ. Sci. Technol. 38:2075-2081.

Fetzner, S., and F. Lingens, 1994. Bacterial dehalogenases: Biochemistry, genetics, and

biotechnological applications. Microbiol. Rev. 58:641-685.

Field, J.A., and R. Sierra-Alvarez, 2008. Microbial transformation and degradation of

polychlorinated biphenyls. Environ. Pollut. 155:1-12.

Fierer, N., Jackson, J.A., Vilgalys, R., and Jackson, R.B. 2005 Assessment of soil

microbial community structure by use of taxon-specific quantitative PCR assays.

Appl Envir Microbiol 71: 4117-4120.

Finneran, K.T., H.M. Forbush, C.V. Gaw VanPraagh and D.R. Lovley, 2002.

Desulfitobacterium metallireducens sp. nov., an anaerobic bacterium that couples

150 growth to the reduction of metals and humic acids as well as chlorinated

compounds. Int. J. Sys. Evol. Microbiol. 52: 1929-1935

Fukuzaki, S., N. Nishio, M. Shobayashi, and S. Nagai, 1990. Inhibition of the

fermentation of propionate to methane by hydrogen, acetate, and propionate.

Appl. Environ. Microbiol. 56: 719-723.

Furukawa, K., 2000. Biochemical and genetic bases of microbial degradation of

polychlorinated biphenyls (PCBs). J. Gen. Appl. Microbiol. 46:283-296.

Furukawa, K., N. Hayase, K. Taira, and N. Tomizuka, 1989. Molecular relationship of

chromosomal genes encoding biphenyl/polychlorinated biphenyl catabolism:

Some soil bacteria possess a highly conserved bph operon. J. Bacteriol. 17:5467-

5472

Futagami, T., M. Goto,K. Furukawa, 2008. Biochemical and genetic bases of

dehalorespiration. Chem. Rec. 8:1–12.

Girvin, D.C., and A.J. Scott, 1997. Polychlorinated biphenyl sorption by soil:

Measurement of soil-water partition coefficients at equilibrium. Chemosphere

35:2007-2025.

Gomes, N.C.M., O. Fagbola, R. Costa, N.G. Rumjanek, A. Buchner, L. Mendona-Hagler,

and K. Smalla, 2003. Dynamics of fungal communities in bulk and maize

rhizosphere soil in the tropics. Appl. Environ. Microbiol. 69: 3758-3766.

Gribble, G.W., 1996. The diversity of natural organochlorines in living organisms. Pure

Appl. Chem 68:1699-1712.

Gribble G.W., 1998. Naturally occurring organohalogen compounds. Acc. Chem. Res. 31,

141-152.

151 Gribble G.W., 2003. The diversity of naturally produced organohalogens. Chemosphere

52: 289–297.

Gundersen D.T., M.D. Krahling, J.J. Donosky, R.G. Cable, S.D. Mims, 1998.

Polychlorinated biphenyls and chlordane in the gonads of paddlefish, Polyodon

spathula, from the Ohio River. Bull. Environ. Contam. Toxicol. 61: 650-657.

Hackl, E., Zechmeister-Boltenstern, S., Bodrossy, L., and Sessitsch, A. 2004.

Comparison of diversities and compositions of bacterial populations inhabiting

natural forest soils. App Environ Microbiol 70: 5057-5065.

Haddock, J.D., J.R. Horton, and D.T. Gibson, 1995. Dihydroxylation and dechlorination

of chlorinated biphenyls by purified biphenyl 2,3-dioxygenase from Pseudomonas

sp. Strain LB400. J. Bacteriol. 177:20-26.

Hankinson, O., 1995. The aryl hydrocarbon receptor complex. Annu. Rev. Pharmacol.

Toxicol. 35:307-340.

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. 68:485-495.

Hermanson, M.H., C.A. Scholten, and K. Compher, 2003. Variable air temperature

response of gas-phase atmospheric polychlorinated biphenyls near a former

manufacturing facility. Environ. Sci. Technol. 37:4038-4042.

Holliger, C., C. Regeard, and G. Diekert, 2003. Dehalogenation by anaerobic bacteria p.

115-157. In M.M. Häggblom and I.D. Bossert (eds), Dehalogenation: Microbial

152 processes and environmental applications, Kluwer Academic Publishers. New

York, NY.

Holliger, C., F. Wohlfarth, and G. Diekert, 1999. Reductive dechlorination in the energy

metabolism of anaerobic bacteria. FEMS Microbiol Rev 22:383-398.

Holmes, D.A., B.K. Harrison, and J. Dolfing, 1993. Estimation of Gibbs free energies of

formation for polychlorinated biphenyls. Environ. Sci. Technol. 27:725-737.

Holmes, D.E., J.S. Nicoll, D.R. Bond, and D.R. Lovley, 2004. Potential role of a novel

psychrotolerant member of the family Geobacteraceae, Geopsychrobacter

electrodiphilus gen. nov., sp. nov., in electricity production by a marine sediment

fuel cell. Appl. Environ. Microbiol. 70:6023-6030.

Hou, L.H. and S.K. Dutta, 2000. Phylogenetic characterization of several para and meta-

PCB dechlorinating Clostridium species: 16S rDNA sequence analyses. Lett Appl

Microbiol 30: 238-243.

Huber, T., G. Faulkner and P. Hugenholtz, 2004. Bellerophon; a program to detect

chimeric sequences in multiple sequence alignments. Bioinformatics 20:2317-

2319.(http://foo.maths.uq.edu.au/~huber/bellerophon.pl)

Imamoglu, I., K. Li, E.R. Christensen, and J.K. McMullin, 2004. Sources and

dechlorination of polychlorinated biphenyl congeners in the sediments of Fox

River, Wisconsin. Environ. Sci. Technol. 38:2574-2583.

International Union of Pure and Applied Chemistry (IUPAC), 1993. Glossary for

chemists of terms used in toxicology. Pure Appl. Chem. 65:2003-2122

153 Jota, M.A.T. and J.P. Hassett, 1991 Effects of environmental variables on binding of a

PCB congener by dissolved humic substances. Environ. Toxicol. Chem. 10:483-

491.

Klasson, K.T., J.W. Barton, B.S. Evans, and M.E. Reeves, 1996. Reductive microbial

dechlorination of indigenous polychlorinated biphenyls in soil using a sediment-

free inoculum. Biotechnol. Prog. 12:310-315.

Kloep F., Manz, W., Röske, I. 2006. Multivariate analysis of microbial communities in

the River Elbe (Germany) on different phylogenetic and spatial levels of

resolution. FEMS Microbiol Ecol 56: 79-94.

Krasotkina, J., T. Walters, K.A. Maruya, and S.W. Ragsdale, 2001. Characterization of

the B12- and iron-sulfur-containing reductive dehalogenase from

Desulfitobacterium chlororespirans. J. Biol. Chem. 276:40991-40997.

Kumar, S., K. Tamura, and M. Nei, 2004. MEGA3: Integrated software for molecular

evolutionary genetics analysis and sequence alignment. Brief. Bioinform. 5: 150-

163.

Lake, J. L., R. J. Pruell, and F. A. Osterman, 1992. An examination of dechlorination

processes and pathways in New Bedford harbor sediments. Mar. Environ. Res.

33:31-47.

Lane, D. J. 1991. 16S/23S rRNA sequencing, p. 115-175. In E. Stackebrandt and M.

Goodfellow (ed.), Nucleic acid techniques in bacterial systematics. John Wiley &

Sons, Ltd., Chichester, United .

154 Leblanc, G.A., 2004. Basics of environmental toxicology. In E. Hodgson (ed), A

textbook of modern toxicology”. 3rd ed. John Wiley & Sons, Inc. Hoboken, New

Jersey pp 463-478.

Liu, Y, D.L. Balkwill, H.C. Aldrich, G.R. Drake, and D.R. Boone, 1999. Characterization

of the anaerobic propionate-degrading syntrophs Smithella propionica gen. nov.,

sp. nov. and Syntrophobacter wolinii. Int J Syst Bacteriol. 49: 545-556.

Löffler, F.E., J.R. Cole, K.M. Ritalahti, and J.M. Tiedje. 2003. Diversity of

dechlorinating bacteria p. 53-87. In M.M. Häggblom and I.D. Bossert (eds)

Dehalogenation: Microbial processes and environmental applications, Kluwer

Academic Publishers. New York, NY.

Lovley, D.R., S.J. Giovannoni, D.C. White, J.E. Champine, E.J. Phillips, Y.A. Gorby, S.

Goodwin, 1993. Geobacter metallireducens gen. nov. sp. nov., a microorganism

capable of coupling the complete oxidation of organic compounds to the

reduction of iron and other metals. Arch Microbiol 159: 336-44.

Mackay, D., 1982. Correlation of bioconcentration factors. Environ. Sci. Technol.

16:274-278.

Madigan, M.T., J.M. Martinko, J. Parker, 2003a. Brock biology of microorganisms, 10th

ed. Pearson Education Inc., Upper Saddle River, NJ. Appendix I.

Madigan, M.T., J.M. Martinko, J. Parker, 2003b. Brock biology of microorganisms, 10th

ed. Pearson Education Inc., Upper Saddle River, NJ. pp 151-155.

Magar, V.S., G.W. Johnson, R.C. Brenner, J.F. Quensen III, E. Foote, G. Durell, J.A.

Ickes, and C. Peven-McCarthy, 2005. Long-term recovery of PCB-contaminated

155 sediments at the Lake Hartwell superfund site: PCB dechlorination. 1. End-

member characterization. Environ. Sci. Technol. 39:3538-3547.

Magnuson, J.K., M.F.Romine, D.R. Burris, M.T. Kingsley, 2000. Trichloroethene

reductive dehalogenase from Dehalococcoides ethenogenes: Sequence of tceA

and substrate range characterization. Appl. Environ. Microbiol. 66:5141-5147.

Maillard, J., C. Regeard, and C. Holliger, 2004. Isolation and characterization of Tn-

Dha1, a transposon containing the tetrachloroethene reductive dehalogenase of

Desulfitobacterium hafniense strain TCE1. Environ. Microbiol. 7:107–117.

Maltseva, O., T.V. Tsoi, J.F. Quensen III, M. Fukuda, and J.M. Tiedje, 1999.

Degradation of anaerobic reductive dechlorination products of Aroclor 1242 by

four aerobic bacteria. Biodegradation 10: 363-371.

Manzano, M.A., J.A. Perales, D. Sales, and J.M. Quiroga, 2004. Using solar and

ultraviolet light to degrade PCBs in sand and transformer oils, Chemosphere 57:

645-654.

Marlowe, J.L. and A. Puga, 2005. Aryl hydrocarbon receptor, cell cycle regulation,

toxicity, and tumorigenesis. J. Cell. Biochem. 96:1174–1184.

Master, E.R., V.W. Lai, B. Kuipers, W.R. Cullen, and W.W. Mohn, 2002. Sequential

anaerobic-aerobic treatment of soil contaminated with weathered Aroclor 1260.

Environ. Sci. Technol. 36:100-103.

McCune, B., and J. B. Grace, 2002. Analysis of ecological communities. Mjm Software

Design, Gleneden Beach, OR, US.

156 Meade, T., and E.M. D’Angelo, 2005. [14C]Pentachlorophenol mineralization in the rice

rhizosphere with established oxidized and reduced soil layers. Chemosphere

61:48-55.

Mehlich, A. 1984. Mehlich-3 soil test extractant: a modification of Mehlich-2 extractant.

Commun. Soil Sci. Plant Anal. 15: 1409-1416.

Mimura, J., and Y. Fujii-Kuriyama, 2003. Functional role of AhR in the expression of

toxic effects by TCDD. Biochim. Biophys. Acta. 1619:263-268

Mohn, W. W., and J. M. Tiedje, 1992. Microbial reductive dehalogenation. Microbiol.

Rev. 56:482-507.

Muyzer , G., E.C. De Waal, and A.G. Uitterlinden, 1993. Profiling of complex microbial

populations by denaturing gradient gel electrophoresis analysis of polymerase

chain reaction-amplified geens coding for 16S rRNA. Appl Environ. Microbiol.

59:695-700.

Muyzer, G., and K. Smalla, 1998. Application of denaturing gradient gel electrophoresis

(DGGE) and temperature gradient gel electrophoresis (TGGE) in microbial

ecology. Antonie van Leeuwenhoek. 73:127-141.

Becton, A. M., Allen, J. R., Andren, A., Cristol, S., Durfee, R., Freeman, A., Metcalf, R.,

Nisbet, L, Sanders, J., Sodergren, A., and Zitko, V.: 1979, Polychlorinated

Biphenyls, National Academy of Sciences, Washington, DC, 182 pp.

Newell, C.J., and C.E. Aziz, 2004. Long-term sustainability of reductive dechlorination

reactions at chlorinated solvents sites. Biodegradation 15:387–394.

Nies, L, and T.M. Vogel, 1990. Effects of organic substrates on dechlorination of Aroclor

1242 in anaerobic sediments. Appl Environ. Microbiol. 56:2612-2617.

157 Nollet, H., I. Van de Putte, L. Raskin, W. Verstraete, 2005. Carbon/electron source

dependence of polychlorinated biphenyl dechlorination pathways for anaerobic

granules. Chemosphere 58:299-310.

Nonaka, H., G. Keresztes, Y. Shinoda, Y. Ikenaga, M. Abe, K. Naito, et al., 2006.

Complete genome sequence of the dehalorespiring bacterium Desulfitobacterium

hafniense Y51 and comparison with Dehalococcoides ethenogenes 195. J

Bacteriol 188: 2262-2274.

Ohio River Valley Water Sanitation Commission, 2002. Ohio River total maximum daily

load for PCBs: Ohio River Miles 40.0 to 317.1, Final Report. Cincinnati, OH.

(Available at http://www.epa.gov/reg3wapd/tmdl/wv_tmdl/Ohio/OhioRep.pdf,

last retrieved on May 5th, 2008).

Ohio River Valley Water Sanitation Commission, 2006. 2006 biennial assessment of

Ohio River water quality conditions. 2006 Ohio River 305(b) Report. Cincinnati,

OH. (Available at www.orsanco.org/rivinfo/305b.asp, Last retrieved on May 21st,

2008).

Owens, W.F., D.C. Stuckey, J.B. Healy Jr., L.Y. Young, and P.L. McCarty, 1979.

Bioassay for monitoring biochemical methane production and anaerobic toxicity.

Water Res 13: 485-492.

Palekar, L. D., K. A. Maruya, J. E. Kostka, and J. Wiegel, 2003. Dehalogenation of 2,6-

dibromobiphenyl and 2,3,4,5,6-pentachlorobiphenyl in contaminated estuarine

sediment, Chemosphere 53:593-600.

Pankratov, T.A., I.S. Kulichevskaya, W. Lieasck, and S.N. Dedysh, 2006. Isolation of

aerobic, gliding, xylanolytic and laminarinolytic bacteria from acidic Sphagnum

158 peatlands and emended description of Chitinophaga arvensicola. Int. J. Syst.

Evol. Microbiol. 56: 2761-2764.

Peplies, J., C. Lachmund, F.O. Glockner, and W. Manz, 2006. A DNA microarray

platform based on direct detection of rRNA for characterization of freshwater

sediment-related prokaryotic communities. Appl. Environ. Microbiol. 72: 4829-

4838.

Pieper, D.H., 2005. Aerobic degradation of polychlorinated biphenyls. Appl Microbiol

Biotechnol. 67:170–191.

Pop, S.M., R.J. Kolarik, and S.W. Ragsdale, 2004. Regulation of anaerobic

dehalorespiration by the transcriptional activator CprK. J. Biol. Chem. 279:

49910-49918.

Qiu, X., L. Wu, H. Huang, P.E. McDonel, A.V. Palumbo, J.M. Tiedje, et al., 2001.

Evaluation of PCR-generated chimeras, mutations, and heteroduplexes with 16S

rRNA gene-based cloning. App. Environ. Microbiol. 67: 880-887.

Quaiser, A., T. Ochsenreiter, C. Lanz, S.C. Schuster, A.H. Treusch, J. Eck, and C.

Schleper, 2003. Acidobacteria form a coherent but highly diverse group within

the bacterial : evidence from environmental genomics. Mol. Microbiol.

50:563-575.

Quensen, J.F., J.M. Tiedje, and S.A. Boyd, 1988. Reductive dechlorination of

polychlorinated biphenyls by anaerobic microorganisms from sediments. Science

242: 752-754.

159 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 Met. 56:107-118.

Rein, A., M.M. Fernqvist, P. Mayer, S. Trapp, M. Bittens, U.G. Karlson, 2007.

Degradation of PCB congeners by bacterial strains. Appl. Microbiol. Biotechnol.

77:469–481.

Rhee, G.Y., B. Bush, C.M. Bethoney, A. DeNucci, H.M. Oh, and R.C. Sokol, 1993.

Anaerobic dechlorination of Aroclor 1242 as effected by some environmental

conditions. Environ. Toxicol. Chem. 12: 1033-1039.

Rhee, S.K., D. E. Fennell, M. M. Häggblom, and L. J. Kerkhof. 2003. Detection by PCR

of reductive dehalogenase gene motifs in a sulfidogenic 2-bromophenol-

degrading consortium enriched from estuarine sediment. FEMS Microbiol. Ecol.

43:317-324

Robertson, L.W. and L.G. Hanson, 2001. PCBs: Recent advances in environmental

toxicology and health effects. The University of Kentucky Press, Lexington, KY,

USA.

Rodrigues, J.L., C.A. Kachel, M.R. Aiello, J.F. Quensen, O.V. Maltseva, T.V. Tsoi, J.M.

Tiedje, 2006. Degradation of Aroclor 1242 dechlorination products in sediments

by Burkholderia xenovorans LB400 (ohb) and Rhodococcus sp. strain RHA1

(fcb). Appl Environ Microbiol. 72:2476-2482.

Rooney-Varga, J.N., R.T. Anderson, J.L. Fraga, D. Ringelberg, and D.R. Lovley, 1999.

Microbial communities associated with anaerobic benzene degradation in a

petroleum-contaminated aquifer. Appl. Environ. Microbiol. 65:3056-3063.

160 Rose, R.L., and E. Hodgson, 2004. Metabolism of toxicants. In E. Hodgson (ed). “A

textbook of modern toxicology”, 3rd ed. John Wiley & Sons, Inc. Hoboken, New

Jersey.

Rossello-Mora, R., B. Thamdrup, H. Schafer, R. Weller, and R. Amann, 1999. The

response of the microbial community of marine sediments to organic carbon input

under anaerobic conditions. Syst. Appl. Microbiol. 22:237–48.

Rostad, C.E., L.M. Bishop, G.S. Ellis, T.J. Leiker, S.G. Monsterleet, and W.E. Pereira,

1993. Organic contaminants associated with suspended sediment collected during

five cruises of the Mississippi River and its principal tributaries, May 1988 to

June 1990. USGS Open-File Report 93-360. (Available at

http://pubs.usgs.gov/of/1993/ofr93360/, Last retrieved on June 17th, 2008).

Rostad, C.E., L.M. Bishop, G.S. Ellis, T.J. Leiker, S.G. Monsterleet, and W.E. Pereira,

1995. Polychlorinated biphenyls and other synthetic organic contaminants

associated with sediments and fish in the Mississippi River: in Contaminants in

the Mississippi River, 1987-92. Edited by R.H. Meade USGS Circular 1133.

(Available at http://pubs.usgs.gov/circ/circ1133/polychlorinated.html, Last

retrieved on June 17th, 2008).

Safe, S., 1993. Toxicology, structure-function relationship, and human and

environmental health impacts of polychlorinated biphenyls: Progress and

problems. Environ. Health Persp. 100:259-268.

Safe, S., 1994. Polychlorinated biphenyls (PCBs): Environmental impact, biochemical

and toxic responses, and implication for risk assessment. Crit. Rev. Toxicol.

24:87-149.

161 Safe, S., L. Safe, and M. Mullin, 1985. Polychlorinated biphenyls: Congener-specific

analysis of a commercial mixture and a human milk extract J. Agric. Food Chem.

33:24-29

Schloss P. D., and J. Handelsman, 2005. Introducing DOTUR, a computer program for

defining operational taxonomic units and estimating species richness. Appl

Environ Microbiol. 71:1501-1506.

Schneider, A.R.,E.T. Porter, and J.E. Baker, 2007. Polychlorinated biphenyl release from

resuspended Hudson River sediment. Environ Sci Technol.41:1097-1103.

Sekiguchi Y., T. Yamada, S. Hanada, A. Ohashi, H. Harada, and Y. Kamagata, 2003.

Anaerolinea thermophila gen. nov., sp. nov. and Caldilinea aerophila gen. nov.,

sp. nov., novel filamentous thermophiles that represent a previously uncultured

lineage of the domain Bacteria at the subphylum level. Int. J. Syst. Evol.

Microbiol. 53: 1843-1851.

Seshadri, R., L. Adrian, D.E. Fouts, J.A. Eisen, A.M. Phillippy, B.A. Methe, et al. 2005.

Genome sequence of the PCE-dechlorinating bacterium Dehalococcoides

ethenogenes. Science 307: 105-108.

Shea, D., 2004. Transport and fate of toxicants in the environment. In E. Hodgson (ed),

“A textbook of modern toxicology”, 3rd ed. John Wiley & Sons, Inc., Hoboken,

NJ.

Smidt, H., and W. M. de Vos, 2004. Anaerobic microbial dehalogenation. Annu. Rev.

Microbiol. 58:43-73.

162 Sokol, R. C., S. Kwon, C. M. Bethoney, and G. Y. Rhee, 1994. Reductive dechlorination

of polychlorinated biphenyls in St. Lawrence River sediments and variations in

dechlorination characteristics. Environ. Sci. Technol. 28:2054-2064.

Sung, Y., K.E. Fletcher, K.M. Ritalahti, R.P. Apkarian, N. Ramos-Hernandez, R.A.

Sanford, N.M. Mesbah,and F.E. Löffler, 2006. Geobacter lovleyi sp. nov. Strain

SZ, a Novel Metal-Reducing and Tetrachloroethene-Dechlorinating Bacterium.

Appl. Environ. Microbiol. 72:2775-2782.

Tamura, K., J. Dudley, M. Nei, and S. Kumar, 2007. MEGA4: Molecular evolutionary

genetics analysis (MEGA) software version 4.0. Mol. Biol. Evol. 24:1596-1599.

Tamura, K., M. Nei, and S. Kumar, 2004. Prospects for inferring very large phylogenies

by using the neighbor-joining method. Proceedings of the National Academy of

Sciences (USA) 101:11030-11035.

Thibodeau, J., A. Gauthier, M. Duguay, R. Villemur, F. Lepine, P. Juteau, et al. 2004.

Purification, cloning, and sequencing of a 3,5-dichlorophenol reductive

dehalogenase from Desulfitobacterium frappieri PCP-1. Appl. Environ.

Microbiol. 70: 4352-4537.

Townsend, G.T., and J.M. Suflita, 1997. Influence of sulfur oxyanions on reductive

dehalogenation activities in Desulfomonile tiedjei. Appl. Environ. Microbiol. 63:

3594-3599.

U.S. Army Corps of Engineers (USACE), 2006. Ohio River Water Management

Information. 2006 Water Management Division Monthly Reports (Available at

http://www.lrd-wc.usace.army.mil/wmrep.html, Last retrieved on June 27th,

2008).

163 U.S. Army Corps of Engineers (USACE), 1995. Zebra Mussel Research Technical Notes.

Technical Note ZMR-1-32. US Army Corps of Engineers Waterways Experiment

Station. Vicksburg MS. [WWW document] URL

www.sgnis.org/publicat/papers/zmr_1_32.pdf

USEPA, 2004. National listing of fish advisories. Office of Water. Environmental

Protection Agency EPA-828-F-04-016. (Available at

www.epa.gov/waterscience/fish/advisories/factsheet.pdf, last retrieved on May

5th, 2008).

USEPA, 2006. The Great River News Letter. Environmental Protection Agency

EPA/620/N-06/002. (Available at www.epa.gov/emap/greatriver/March06.pdf,

last retrieved on May 5th, 2008).

Vadstein, O., and Y. Olsen, 1989. Chemical composition and phosphate uptake kinetics

of limnetic bacterial communities cultured in chemostats under phosphorus

limitation. Limnol. Oceanogr. 34:939-946. van de Pas, B. A., H. Smidt, W. R. Hagen, J. van der Oost, G. Schraa, A. J. M. 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. 274:20287–20292.

Villemur, R., M. Lanthier, R. Beaudet, F. Lépine, 2006. The Desulfitobacterium genus.

FEMS Microbiol. Rev. 30: 706-733.

Villemur, R., M. Saucier, A. Gauthier, and R. Beaudet, 2002. Occurrence of several

genes encoding putative reductive dehalogenases in Desulfitobacterium

164 hafniense/frappieri and Dehalococcoides ethenogenes. Can. J. Microbiol. 48:

697-706.

Villemur, R., R. Beaudet, M. Lanthier, A. Gauthier, A. Boyer, J. Thibodeau, et al. 2005.

Molecular analysis of Desulfitobacterium frappieri pcp-1 involved in reductive

dehalogenation of pentachlorophenol. Water Sci. Technol. 52: 101-106. von Wintzingerode, F., B. Selent, W. Hegemann, and U.B. Goebel, 1999. Phylogenetic

analysis of an anaerobic TCB-transforming microbial consortium. Appl. Environ.

Microbiol. 65: 283-286.

Vrana, B., K. Dercová, Š. Baláž and A. Ševčíková, 1996. Effect of chlorobenzoates on

the degradation of polychlorinated biphenyls (PCB) by Pseudomonas stutzeri.

World J. Microbiol. Biotechnol. 12: 323-326.

Wiegel, J. and Q. Wu, 2000. Microbial reductive dechlorination of polychlorinated

biphenyls. FEMS Microbiol. Ecol. 32: 1-15.

Wiegel, J., X. Zhang, and Q. Wu, 1999. Anaerobic dehalogenation of hydroxylated

polychlorinated biphenyls by Desulfitobacterium dehalogenans. Appl. Environ.

Microbiol. 65: 2217-2221.

Williams, W.A., J.H. Lobos, and W.E. Cheetham, 1997. A phylogenetic analysis of

aerobic polychlorinated biphenyl-degrading bacteria. Int. J. Sys. Bacteriol. 47:

207-210.

Wischgoll, S., D. Heintz, F. Peters, A. Erxleben, E. Sarnighausen, R. Reski, et al. 2005.

Gene clusters involved in anaerobic benzoate degradation of Geobacter

metallireducens. Mol. Microbiol. 58: 1238-52.

165 Wu, Q., D.L. Bedard, and J. Wiegel, 1996. Influence of incubation temperature on the

microbial reductive dechlorination of 2,3,4,6-tetrachlorobiphenyl in two

freshwater sediments. Appl. Environ. Microbiol. 62: 4174-4179.

Wu, Q., D.L. Bedard, and J. Wiegel, 1997. Temperature determines the pattern of

anaerobic microbial dechlorination of Aroclor 1260 primed by 2,3,4,6-

tetrachlorobiphenyl in Woods Pond sediment. Appl. Environ. Microbiol. 63:4818-

4825.

Wu, Q., K. R. Sowers, and H. D. May. 1998. Microbial reductive dechlorination of

Aroclor 1260 in anaerobic slurries of estuarine sediments. Appl. Environ.

Microbiol. 64:1052-1058.

Wu, Q., J. E. M. 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. App. Environ.

Microbiol. 68:807-812.

Yamada, T., Y. Sekiguchi, S. Hanada, H. Imachi, A. Ohashi, H. Harada and Y.

Kamagata, 2006. Anaerolinea thermolimosa sp. nov., Levilinea saccharolytica

gen. nov., sp. nov. and Leptolinea tardivitalis gen. nov., sp. nov., novel

filamentous anaerobes, and description of the new classes Anaerolineae classis

nov. and Caldilineae classis nov. in the bacterial phylum Chloroflexi. Int. J. Syst.

Evol. Microbiol. 56:1331-1340.

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

166 involvement of phylogenetically similar Dehalococcoides-like bacterial

populations. FEMS Microbiol Ecol. 55:248–261.

Ye, D., J.F. Quensen III, J.M. Tiedje and S.A. Boyd, 1992. Anaerobic dechlorination of

polychlorobiphenyls (Aroclor 1242) by pasteurized and ethanol-treated

microorganisms from sediments. Appl. Environ. Microbiol. 58: 1110-1114.

Zhang, H., Y. Sekiguchi1, S. Hanada, P. Hugenholtz, H. Kim, Y. Kamagata and K.

Nakamura, 2003. Gemmatimonas aurantiaca gen. nov., sp. nov., a Gram-

negative, aerobic, polyphosphate-accumulating micro-organism, the first cultured

representative of the new bacterial phylum Gemmatimonadetes phyl. nov. Int. J.

Syst. Evol. Microbiol. 53: 1155-1163.

Zwiernik, M.J., J.F. Quensen III, and S.A. Boyd, 1998. FeSO4 amendments stimulates

extensive anaerobic PCB dechlorination. Environ. Sci. Technol. 32, 3360-3365.

167 VITA

Name: Andres E. Nunez Date of Birth: June 13, 1973 Place of Birth: San Felix, Venezuela

Education: M.S. Chemistry, Simon Bolivar University (1996-2000) Caracas Venezuela B.S. Chemistry, Simon Bolivar University (1990-1996) Caracas Venezuela

Professional positions held:

Research Assistant, Plant and Soil Sciences, University of Kentucky (2004-2008)

Assistant Professor, Simon Bolivar University, Caracas, Venezuela (2000-2003)

Graduate Research Assistant, Simon Bolivar University, Caracas, Venezuela (1999-2001)

Lab Analyst, Simon Bolivar University, Environmental Management Unit, Caracas Venezuela (1999-2000)

Field Specialist, Simon Bolivar University, Environmental Management Unit, Caracas Venezuela (1996-1997)

Graduate Research Assistant, Simon Bolivar University, Caracas, Venezuela (1996-1997)

Scholastic and professional honors:

Recipient of a Kentucky Opportunity Fellowship, University of Kentucky (2005-2007)

Publications:

Nunez, A., D. Berroteran, and O. Nunez, 2003. Hydrolysis of cyclic phosphoramides. Evidence for syn lone pair catalysis, Org. Biomol. Chem. 1:2283-2289.

Guedez, T., A. Nunez,; E. Tineo, and O. Nunez, 2002. Ring size configuration effect and the transannular intrinsic rates in bislactam macrocycles. J. Chem. Soc., Perkin Trans. 2, 2078-2082.

168 Martinez, M., A. Nunez, R. Lopez, F. Morales, and O. Nunez, 1999. Phenols degradation using TiO2 and UV light. Kinetic measurements in an ample pH range. Acta Cientifica Venezolana, 50(Supl. 1), 81-86.

Nunez, A., and O. Nunez, 1996. Bell-shaped pH-rate profile in a reaction involving a pentacoordinated phosphorus intermediate. J. Org. Chem. 61:8386-8390.

Abstracts

Nunez, A., and E.M. D’Angelo 2006. Phylogenetic characterization of a polychlorinated-biphenyl-dechlorinating microbial community under different anaerobic treatments. 18th World Congress of Soil Science. Philadelphia, PA.

Nunez, A., and E.M. D’Angelo 2005. Microbial population dynamics during anaerobic polychlorinated biphenyls bioremediation in Ohio River sediment. 2005 ASA-CSSA- SSSA International Annual Meetings, Salt Lake City, UT.

Nuñez, O., A. Nuñez, and D. Berroterán 2001. Hydrolysis of five and six membered cyclic phosphoramides. Kinetics and product formation ratio. Sixth Latin-American Conference on Physical Organic Chemistry. Porlamar, Venezuela.

Nunez, A., and O. Nunez 1999. Kinetic study of the acid hydrolysis of 5-membered cyclic phosphoramides. XLIX AsoVAC annual meeting. Maracay, Venezuela.

Martinez, M., A. Nunez, R. Lopez, F. Morales, and O. Nunez, 1999. Phenols degradation using TiO2 and UV light. Kinetic measurements in an ample pH range. XLIX AsoVAC annual meeting. Maracay, Venezuela.

Nunez, A., and O. Nunez 1996. Reaction mechanism of p-nitrophenol and hexakis(imidazolyl)cyclotriphosphazine in aqueous THF. XLVI AsoVAC annual meeting. Barquisimeto, Venezuela.

Nunez, A., and O. Nunez 1995. Kinetic study of the reaction between p-nitrophenol and hexakis(amino)cyclotriphosphazine in aqueous THF. XLV AsoVAC annual meeting. Caracas, Venezuela.

169