University of South Florida Scholar Commons

Graduate Theses and Dissertations Graduate School

6-16-2008 Found in Raw Sewage and Their otP ential to Indicate Fecal Pollution in Coastal Environments Erin M. Symonds University of South Florida

Follow this and additional works at: https://scholarcommons.usf.edu/etd Part of the American Studies Commons

Scholar Commons Citation Symonds, Erin M., "Viruses Found in Raw Sewage and Their otP ential to Indicate Fecal Pollution in Coastal Environments" (2008). Graduate Theses and Dissertations. https://scholarcommons.usf.edu/etd/522

This Thesis is brought to you for free and open access by the Graduate School at Scholar Commons. It has been accepted for inclusion in Graduate Theses and Dissertations by an authorized administrator of Scholar Commons. For more information, please contact [email protected].

Viruses Found in Raw Sewage and Their Potential

to Indicate Fecal Pollution in Coastal Environments

by

Erin M. Symonds

A thesis submitted in partial fulfillment of the requirements for the degree of Master of Science College of Marine Science University of South Florida

Major Professor: Mya Breitbart, Ph.D. John H. Paul, Ph.D. Dale W. Griffin, Ph.D.

Date of Approval: June 16, 2008

Keywords: Water Quality, Fecal-Associated Pathogenic Viruses, Viral Diversity, Microbial Indicators,

© Copyright 2008, Erin M. Symonds

ACKNOWLEDGEMENTS

I am extremely grateful for the guidance and support that I have received from Dr.

Mya Breitbart, my major professor, throughout the course of my degree. She is an outstanding advisor whose creativity, patience, good humor, and dedication facilitated the formation and execution of this study. Funding for this work was provided by the U.S.

Environmental Protection Agency (Grant # X7-96465507-0). I would also like to thank the members of my committee, Dale Griffin and John Paul, for their advice, expertise, and support throughout this research as well as in the revision of my thesis.

Special thanks to members of the Breitbart Lab (Camille Daniels, Dawn

Goldsmith, Terry Ng, Kim Pause, and Karyna Rosario) for their knowledge, advice, and support throughout this project. Their willingness to help ensured successful collection of samples. Without their support in the lab, the completion of this project would have been much more difficult. Thanks to Mike Gray (USGS) for his expertise in phylogenetics.

I would like to recognize several parties for their assistance in collecting samples

for this project. Thanks to Kim Ritchie and Eric Bartels of Mote Marine Laboratory as

well as the crew of the R/V Bellows for their support in collecting samples. Thanks to

Dale Griffin, Valerie Harwood and her lab, the Rosario Cora family, Gabe Vargo, and the

Suncoast Seabird Sanctuary for supplying samples. Finally, I would like to express

gratitude to all of the wastewater treatment facilities for supplying wastewater samples.

The completion of my thesis would not have happened without the support of my

family and friends. Thanks to all of you who listened over the last couple of year.

TABLE OF CONTENTS

LIST OF TABLES iii

LIST OF FIGURES iv

ABSTRACT v

INTRODUCTION 1

PROJECT OBJECTIVES 8

MATERIALS AND METHODS 9

Concentration of Viruses in Wastewater Samples 9 Wastewater Viral Concentration Efficiency 10 Identification of Targeted Viruses 11 Characterizing the Diversity of Picobirnaviruses in Raw Sewage 18 Determination of Assay Sensitivity 20 Isolation of Viruses from Large-scale Seawater Samples 20 Seawater Viral Isolation Methods Comparison 25 Determination of Viral Stability in Seawater 27

RESULTS 29

Viral Detection in Raw Sewage 29 Viral Detection in Final Effluent 30 PCR Assay Sensitivity 30 Wastewater Viral Concentration Efficiency 31 Diversity of Picobirnaviruses in Raw Sewage 33 Viral Isolation and Detection in Large-Scale Seawater Samples 37 Seawater Viral Isolation Methods Comparison 39 Stability of Pepper Mild Mottle and in Seawater 41

DISCUSSION 43

Viruses Detected in Raw Sewage 43

i Viruses Detected in Final Effluent 44 Wastewater Viral Concentration Efficiency 46 Diversity of Picobirnaviruses in Raw Sewage 46 Viral Isolation and Detection in Large-Scale Seawater Samples 48 Seawater Viral Isolation Methods Comparison 48 Stability of Pepper Mild Mottle Virus and Picobirnavirus in Seawater 49 Limitations 50 Potential of PMMoV and Picobirnaviruses as Water Quality Indicators 51

CONCLUSIONS 54

REFERENCES 55

ii

LIST OF TABLES

Table 1 Viruses under investigation in raw sewage, final effluent, and the marine environment. 5

Table 2 Date and location of raw sewage and final effluent sampling from municipal wastewater treatment facilities. 10

Table 3 Primer nucleotide sequences used to PCR amplify viruses from 10 viral groups. 12

Table 4 Characteristics of viral families found in raw sewage (35). 24

Table 5 Primer nucleotide sequences used for PCR amplification of poliovirus (38). 27

Table 6 Summary table of the targeted viruses detected in 12 raw sewage samples collected from throughout the coastal United States (two samples came from Florida). 29

Table 7 Summary table of the targeted viruses detected in 12 final effluent samples collected from 11 different coastal states. 30

Table 8 The sensitivities of each PCR assay employed in this study, which is identical to previous studies. 31

Table 9 The top BLASTN hits of cloned picobirnavirus PCR products from U.S. raw sewage samples. 34

Table 10 The concentration of viruses (viral-like particles/milliliter) at each step of the concentration and purification process for each seawater sample. 38

Table 11 Direct counts of virus-like particles (VLP) per milliliter of seawater for each step in the TFF Method’s concentration and purification protocol with samples from the three spiking experiments. 39

Table 12 Summary of wastewater treatment processes at each of the wastewater treatment facilities identified by location. 45

iii

LIST OF FIGURES

Figure 1 Sites of large-scale seawater sampling in Florida. 22

Figure 2 Agarose gel electrophoresis (2%) of adenovirus PCR products from the raw sewage spiking experiment. 32

Figure 3 Agarose gel electrophoresis (2%) of adenovirus PCR products from the final effluent spiking experiment. 32

Figure 4 A condensed, neighbor joining (Jukes-Cantor model) phylogenetic tree of a ~200 bp segment of the picobirnavirus RNA-dependent RNA polymerase gene from 207 U.S. raw sewage sequences. 35

Figure 5 Rarefaction analyses of the picobirnavirus sequences from U.S. raw sewage. 37

Figure 6 Agarose gel electrophoresis (2%) showing the effectiveness of the TFF Method (lanes 2-8) and the Public Health Method (lanes 9-15) from the first spiking experiment using polioviruses. 40

Figure 7 Agarose gel electrophoresis (2%) showing the effectiveness of the TFF Method (lanes 2-8) and the Public Health Method (lanes 9-15) from the second spiking experiment using adenoviruses. 40

Figure 8 Agarose gel electrophoresis (2%) showing the effectiveness of the TFF Method (lanes 2-8) and the Public Health Method (lanes 9-15) from the third spiking experiment with an overnight incubation using adenoviruses. 41

Figure 9 Agarose gel electrophoresis (2%) depicting picobirnavirus PCR products (~200 bp) from two experiments assessing the stability of picobirnaviruses from raw sewage in seawater. 42

Figure 10 Agarose gel electrophoresis (2%) of PMMoV PCR products from two experiments assessing the stability of PMMoV from raw sewage in seawater. 42

iv

VIRUSES FOUND IN RAW SEWAGE AND THEIR POTENTIAL TO INDICATE FECAL POLLUTION IN COASTAL ENVIRONMENTS

Erin M. Symonds

ABSTRACT

The presence of pathogenic viruses in coastal environments is an important tool in evaluating water quality and health risks. Millions of viruses are excreted in fecal matter and bacterial indicators do not correlate with the presence of pathogenic viruses.

Enteroviruses have been used to identify fecal pollution in the environment; however, other viruses shed in fecal matter could be used to indicate fecal pollution. The purpose of this research is to develop a baseline understanding of the diversity of viruses found in raw sewage and to assess their presence in the marine environment. PCR was used to detect adenoviruses, herpesviruses, hepatitis B viruses, , , papillomaviruses, pepper mild mottle viruses, picobirnaviruses, reoviruses, , and sapporoviruses in raw sewage collected from throughout the United States and from five marine environments ranging in their proximity to dense human populations.

Adenoviruses, noroviruses, pepper mild mottle viruses, and picobirnaviruses were detected in raw sewage but absent in the marine environment, making these viruses potential indicators of fecal pollution in marine environments. These viruses were also found in many of the final effluent samples. Pepper mild mottle viruses may be useful for source tracking fecal contamination since it was consistently found in human sewage and is not expected in the feces of other animals due to its dietary origin. Furthermore, this

v research uncovered previously unknown sequence diversity in human picobirnaviruses.

This baseline understanding of viruses in raw sewage and the marine environment will enable educated decisions to be made regarding the use of viruses in water quality assessments.

vi

INTRODUCTION

Nearly 40% of Americans live in coastal environments and they produce

approximately 1.0x10 10 gallons of wastewater per day. The majority of treated wastewater, as well as untreated sewage, drains into the marine environment (2).

Millions of viruses and bacteria are excreted in human fecal matter (6, 17, 98) and current methods of sewage treatment do not always effectively remove these organisms (90-92,

94). In addition to treated wastewater discharge, other point sources of fecal pollution include the following: combined sewer overflows, municipal storm sewer systems, concentrated animal feeding operations, and meat-, fish-, and shellfish-processing facilities (33).

While point sources of fecal pollution can have significant impacts to the health of coastal environments, non-point sources of fecal pollution can pose an even bigger threat due to difficulties in identifying and mitigating the source of pollution. In 2002, nearly half of the pollution sources responsible for beach closures in the United States were from unknown and non-point sources (33). Run-off, farm animals, wildlife, septic systems, swimmers, and faulty sanitary sewer lines are all examples of non-point sources of fecal contamination (31, 33). Non-point source fecal pollution is a major problem in areas such as the Florida Keys, where the community relies on the use of ~25,000 septic tanks (some of which are outdated), ~5-10,000 illegal cesspits, and ~1,207 aerobic treatment facilities (34, 70). Additionally, the permeable nature of Florida’s soils and limestone bedrock combined with the use of inappropriate wastewater disposal systems

1 increases the risk of surface and ground water contamination with fecal-associated microorganisms (65, 78).

As a means to understand the fate of wastewater in the Florida Keys, several viral tracer studies were conducted in the late 1990s. Viral tracers from a domestic septic tank were shown to move into surrounding canals and coastal waters in as little as 3.5 hours and 23 hours respectively (76, 77). In the case of injection wells, viral tracers were found in groundwater and marine waters within 8 hours and 10 – 53 hours respectively after seeding (78). An additional study investigating the movement of viral tracers from injection wells and septic tanks revealed that it was possible for the viral tracers to move from the on-site disposal systems into neighboring canals and out to coastal marine surface waters at a rate of 1.7 to 57.5 meters per hour for septic systems and 66.8 to 141 meters per hour for injection wells (76). These studies convincingly demonstrated that septic systems and injection wells can be non-point sources of wastewater pollution in the

Florida Keys.

Due to the high pathogen loads and nutrient concentrations present in raw sewage, fecal pollution poses a threat to environmental and human health. Marine water containing excess nutrients from wastewater pollution, can cause hypoxic events, eutrophication, a decline in seagrass and coral reef biomass, as well as a reduction in finfish and shellfish populations (52). Many epidemiological investigations have shown an increased risk of disease in people who have swam in polluted waters identified by high concentrations of fecal bacteria (2, 26, 37, 51, 74). Furthermore, ingestion of seafood exposed to fecal pollution has also been shown to increase one’s risk of disease

(29, 87). The coastal environment is an important economic resource with tourism,

2 recreational activities, and commercial fishing generating significant revenues. As a result of the known health risks associated with swimming in polluted waters and consuming seafood contaminated by wastewater, the US government passed legislation to address this problem.

The Beaches Environmental Assessment and Coastal Health Act, enacted in 2000,

holds each state responsible for monitoring and reporting the presence of fecal pollution

in all recreational waters via indicator organisms (Public Law 106-284). An ideal

indicator organism of fecal pollution, as defined by the Environmental Protection Agency

(EPA), should posses several qualities (33). First, detection of the organism should be

simple and affordable. Second, the organism should not be present in the environment

naturally in the absence of pollution. Third, the organism should be present in polluted

waters in concentrations correlated with the amount of fecal pollution. Finally, the

indicator organism should have a decay rate comparable to that of the pathogen of

concern.

Currently, the US EPA mandates the use of bacterial indicators, such as fecal

coliforms and enterococci, to assess water quality. Although monitoring of these bacteria

is simple and inexpensive, it has been shown that fecal-associated bacteria are not ideal

indicators of fecal pollution for several reasons. One unsuitable characteristic of bacterial

indicators is their ability to live in sediments (19, 36, 62). If bacterial indicators are able

to survive in sediments, then their re-suspension into the water column can mask

correlations between their concentrations and the extent of current fecal pollution (i.e.,

false positives). Furthermore, it is also likely that the die-off rate of bacterial indicators

does not match that of fecal associated pathogens, like viruses.

3 Another unfavorable characteristic of current bacterial indicators is their inability to predict or correlate with the presence of pathogenic viruses associated with wastewater

(45, 46, 96). Human pathogenic viruses associated with feces are generally more robust than enteric bacteria and are not as easily eliminated by current methods of wastewater treatment (50, 96). A California study showed that adenoviruses are more resilient to tertiary wastewater treatment and ultraviolet-disinfection than are bacterial indicators of fecal pollution (90). Additionally, adenoviruses and noroviruses have been detected in treated wastewater discharge in Europe (91, 94). Since concentrations of fecal coliforms and enterococci inadequately detect fecal pollution and therefore inaccurately depict the risks to human health, many have proposed the use of an alternative viral indicator of wastewater contamination (39, 46, 69).

Over the last ten years researchers have investigated the distribution and abundance of enteric viruses in Florida as a means of understanding fecal pollution and public health risks. Enteric viruses describe a group of eukaryotic viruses (including viruses belonging to the , , Picornaviridae , and families) transmitted via the fecal-oral route. These viruses have been found throughout the marine environment (i.e. in seawater, sponge, and coral mucus samples) adjacent to the Florida

Keys (28, 39, 44, 45, 64, 96). They have also been detected in seawater and freshwater samples from Sarasota, Florida (65). Despite the identification of pathogenic viruses within the coastal environment of Florida, the presence of other, potentially hazardous viruses transmitted via the fecal-oral route still remains unknown because they have yet to be studied. The occurrence of emerging diseases in humans and marine organisms appears to be rising and some of these diseases may be caused by human-associated fecal

4 flora (48, 71, 73, 85). While it is impractical to monitor the presence of all viral pathogens related to wastewater pollution, the development of an accurate viral indicator of sewage contamination is needed for enhanced water quality monitoring.

To develop a comprehensive water quality indicator protocol, it is first necessary to establish a broad, baseline understanding of the many pathogenic, eukaryotic viruses in raw sewage. Currently no broad baseline data exists on the presence of eukaryotic viruses in raw sewage in the United States. Several studies have identified adenoviruses, noroviruses, reoviruses, rotaviruses, and other (e.g. polioviruses, coxsackie viruses, echo-viruses) in untreated raw sewage in Australia, Europe and South Africa (30,

54, 66, 91, 92, 94). Since no comprehensive study has been executed to identify the prevalence of many different eukaryotic viruses in raw sewage from the United States, the first objective of this research was to identify the presence of ten viral groups known to be transmitted via the fecal-oral route in raw sewage samples collected throughout the country.

TABLE 1. Viruses under investigation in raw sewage, final effluent, and the marine environment .

Viral Family Virus Type Adenoviridae Human Adenoviruses, A - F dsDNA Caliciviridae Noroviruses & Sapporoviruses ssRNA (+ strand) Hepatitis B Viruses reverse transcribing Human Herpesviruses dsDNA Human Papillomaviruses dsDNA Morbilliviruses ssRNA (- strand) Mammal Reoviruses Reoviridae Rotaviruses, Group A dsRNA Unassigned Pepper Mild Mottle Viruses ssRNA Unclassified Human Picobirnaviruses dsRNA

The presence of the viral families listed in Table 1 was analyzed in raw sewage

samples collected from around the United States. All of the viral families under

5 investigation are known to infect humans except pepper mild mottle viruses, which infects plants . The majority of these viral families, excluding studies of Adenoviridae,

Caliciviridae, and Reoviridae , have not been studied in sewage despite their fecal-oral transmission. Picobirnaviruses and pepper mild mottle viruses (PMMoV) have been detected in individual fecal samples (13, 83, 95, 98); however, their presence has never been analyzed in collective waste nor have they been proposed as markers of fecal pollution.

Although PMMoV are plant pathogens, these viruses were most abundant in a metagenomic survey of RNA viruses in human feces (98). PMMoV is a rod-shaped single-stranded RNA virus that infects ornamental peppers. Through human consumption of processed pepper products (ie. hot sauce), the PMMoV is excreted in human feces on the order of nearly one billion viruses per gram dry weight of feces. Interestingly, these viruses are stable enough to survive passage through the gut and remain infectious to plants (98). The prevalence of PMMoV in human feces, regardless of its importance for human health, makes it a potential indicator of fecal pollution.

After determining which pathogenic viruses were found in raw sewage, their presence in final effluent and the marine environment was determined. Since accurate indicators would not be found in the marine environment in the absence of pollution, the natural presence of these potential viral markers in seawater samples ranging in their exposure to human populations was determined. The results of this research will have important applications to water quality monitoring programs mandated by the US EPA throughout the country. Characterizing the presence and diversity of a broad range of viral families suspected to be found in feces in raw sewage from around the United States

6 will narrow the list of possible viral indicators of fecal pollution. The conclusions of this research will identify viruses that can potentially be used to indicate wastewater pollution in coastal environments. This discovery phase is a critical first step before rapid and sensitive assays can be developed for these indicator viruses.

7

PROJECT OBJECTIVES

Current bacterial indicators of fecal pollution inadequately assess the extent of fecal

pollution in recreational waters; accordingly, many have proposed the use of viruses to

indicate fecal pollution in the environment. Before a viral indicator of fecal pollution can

be identified, many factors must be investigated to include the prevalence of different

viral types in sewage, their correlation to the presence of other pathogens, and their

behavior in coastal environments. This study aims to begin to assess the use of viruses as

indicators of fecal contamination. The objectives of this study are:

 To determine the presence of ten types of viruses that transmit via the fecal-oral route in raw sewage collected from 11 states in the U.S.

 To characterize the diversity of picobirnaviruses in raw sewage collected from 11 states in the U.S.

 To determine the presence of the viruses that were detected in raw sewage in the following sample types:

o Final effluent collected from 11 states. o Seawater samples collected from locations ranging in their proximity to dense human populations.

 To investigate the stability of picobirnaviruses and PMMoV in seawater.

8

MATERIALS AND METHODS

Concentration of Viruses in Wastewater Samples

A total of 12 raw sewage and 12 final effluent samples were collected from

wastewater treatment facilities in the United States as listed in Table 2. For each sample,

10 ml was filtered through a 0.45 µm polyether sulfone membrane filter cartridge

(Sterivex; Millipore, Billerica, MA, USA) to separate large bacteria and other particles from viruses. The filtrate was then concentrated using a centrifugal concentration device

(Centriplus YM-50; Millipore, Billerica, MA, USA) to a volume less than one milliliter.

Samples were further concentrated to less than 200 µl using another centrifugal concentration device (Microcon Ultracel YM-30; Millipore, Billerica, MA, USA).

Nucleic acid was extracted from 200 µl of filtered, concentrated sample using the

MinElute Virus Spin Kit (QIAGEN ©, Valencia, CA, USA). Immediately following nucleic acid extraction, cDNA was synthesized from the extracted RNA using First

Strand Synthesis Superscript III Reverse Transcription Kit (Invitrogen © Carlsbad, CA,

USA) with random hexamers. Whole amplication using GenomiPhi V2 (GE

Healthcare, Piscataway, NJ, USA) was utilized for every raw sewage and final effluent sample as a means of increasing the concentration of total DNA and cDNA prior to virus identification.

9 TABLE 2. Date and location of raw sewage and final effluent sampling from municipal wastewater treatment facilities.

Date Location Abbreviation 11/5/2007 Alabama AL 10/18/2007 California CA 11/14/2007 Connecticut CT 11/30/2007 Florida FL 12/8/2006 Florida Keys FL K 11/12/2007 Louisiana LA 11/5/2007 Maine ME 11/24/2007 Maryland MD 11/13/2007 New Jersey NJ 11/13/2007 North Carolina NC 11/13/2007 Oregon OR 11/8/2007 Washington WA

Wastewater Viral Concentration Efficiency

Known quantities of adenoviruses were added to a raw sewage and a final effluent

sample in order to determine the efficiency of the methods employed to concentrate

viruses from raw sewage and final effluent. Twenty milliliters of raw sewage and final

effluent were autoclaved for 15 minutes at 121 °C as a means of eliminating the presence

of adenovirus particles while minimizing the alteration of particulates in the sample. Ten

milliliters of raw sewage and final effluent were spiked to a final concentration of

8.48x10 5 Adenovirus-20 particles per milliliter. The remaining ten milliliters of

autoclaved raw sewage and final effluent were not spiked; rather, they served as a

negative control throughout the viral concentration, nucleic acid isolation, and viral

detection processes. Viruses from all four samples were concentrated using the methods

previously described for raw sewage and final effluent. Adenoviruses were assayed for

in 1:10 serial dilutions of extracted DNA from spiked and control raw sewage and final

effluent using nested PCR using previously published primers in Table 3 and conditions

(3).

10 Identification of Targeted Viruses

To identify the presence and diversity of the targeted viruses in raw sewage, PCR

or RT-PCR for each of the targeted viral groups was executed using previously published

primers and conditions described in the following subsections for each viral family. All

PCR products were visualized using agarose gel (2%) electrophoresis stained with

ethidium bromide. If no positive PCR products were amplified, the presence of inhibitors

was analyzed by assaying serially diluted (to 10 -4) sample DNA spiked with positive control. Positive PCR products with one distinct band were purified using the UltraClean

PCR Clean-up Kit (MO BIO Laboratories, Inc., Carlsbad, CA, USA). If more than one band was present, the UltraClean Gelspin Kit (MO BIO Laboratories, Inc., Carlsbad, CA,

USA) was utilized to gel-purify the positive PCR product of the correct size. All purified positive PCR products were sequenced with their respective primers. Sequences were trimmed using Sequencher (Gene Codes Corporation, Ann Arbor, MI, USA) and the identity of the positive PCR products were confirmed by comparing the sequence against the GenBank non-redundant database using BLASTN (4, 5). If BLASTN searches did not result in sequence identities to the targeted virus, then TBLASTX searches were executed in an attempt to detect weaker sequence similarities to known sequences in GenBank.

11 TABLE 3. Primer nucleotide sequences used to PCR amplify viruses from 10 viral groups. Note that mixed bases represent the following nucleotides: H (A, C, or T), M (A or C), R (G or A), S (G or C), V (G, C, or A), W (A or T), and Y (T or C). Note that I has base-paring bias to C, A, G, and then T.

Target Viruses Primer Name Primer Sequence (5'-3') Reference AV-A1 (+) GCC GCA GTG GTC TTA CAT GCA CAT C Human Adenoviruses, AV-A2 (-) CAG CAC GCC GCG GAT GTC AAA GT 3 A - F AV-B1 (+) GCC ACC GAG ACG TAC TTC AGC CTG AV-B2 (-) TTG TAC GAG TAC GCG GTA TCC TCG CGG TC Noroviruses & P290 (+) GAT TAC TCC AAG TGG GAC TCC AC 57 Sapporoviruses P289 (-) TGA CAA TGT AAT CAT CAC CAT A HBS-1 (+) ATC AGG ATT CCT AGG ACC C HBS-R1 (-) AGG ACA AAC GGG CAA CAA C Hepatitis B Viruses 60 HBS-11 (+) GCG GGG TTT TTC TTG TTG AC HBS-R11 (-) GAA CCA ACA AGA AGA TGA GGC FP1 (+) GAY TTY GCI AGY YTI TAY CC FP2 (+) TCC TGG ACA AGC AGC ARI YSG CIM TIA A Herpesviruses FR1 (-) GTC TTG CTC ACC AGI TCI ACI CCY TT 12 FP3 (+) TGT AAC TCG GTG TAY GGI TTY ACI GGI GT RP4 (-) CAC AGA GTC CGT RTC ICC RTA IAT upstream ATG TTT ATG ATC ACA GCG GT Morbilliviruses 11 downstream ATT GGG TTG CAC CAC TTG TC FAP59 TAA CWG TIG GIC AYC CWT ATT Papillomaviruses 40 FAP64 CCW ATA TCW VHC ATI TCI CCA TC Human PicoB25 (+) TGG TGT GGA TGT TTC 83 Picobirnaviruses PicoB43 (-) ART GYT GGT CGA ACT T PMMV-F AAC CTT TCC AGC ACT GCG PMMoV 98 PMMV-R GCG CCT ATG TCG TCA AGA CT L1.rv5 (+) GCA TCC ATT GTA AAT GAC GAG TCT G L1.rv6 (-) CTT GAG ATT AGC TCT AGC ATC TTC TG Mammal Reoviruses 61 L1.rv7 (+) GCT AGG CCG ATA TCG GGA ATG CAG L1.rv8 (-) GTC TCA CTA TTC ACC TTA CCA GCA G RV1 (-) GTC ACA TCA TAC AAT TCT AAT CTA AG RV2 (+) CTT TAA AAG AGA GAA TTT CCG TCT G Rotaviruses, Group A 43 RV3 (+) TGT ATG GTA TTG AAT ATA CCA C RV4 (-) ACT GAT CCT GTT GGC CAW CC

Adenoviruses were identified using a nested primer set (Table 3) that can detect

47 adenovirus serotypes infecting humans (3) and has been previously used to detect adenoviruses in the environment (39). These primers target the adenovirus hexon gene and have a sensitivity of 100 targets. Adenovirus-20 DNA, extracted from viral particles

(number VR-1097 ATCC®, Manassas, VA, USA), was used as a positive control. For the first round of PCR, a total 50 µl reaction mixture contained 2 µl of sample DNA, 1X

REDTaq PCR Reaction Buffer (10.0 mM Tris-HCl pH 8.3, 50.0 mM KCl, 1.1 mM

12 MgCl 2, 0.01% gelatin; Sigma-Aldrich © St. Louis, MO, USA), an additional 0.4mM

MgCl 2, 0.25 mM each dNTP, 1 µM AV-A1 primer, 1 µM AV-A2 primer, and 1 U

REDTaq DNA Polymerase (Sigma-Aldrich © St. Louis, MO, USA). The second PCR reaction was identical to the first, except the template was 5 µl of the first round product and primers were replaced with 1 µM AV-B1 and 1 µM AV-B2. The amplicon of the

first reaction was 300 bp and 143 bp for the second reaction. Both reactions were

incubated for 4 min at 94oC, followed by 40 cycles of [92 o C for 30 sec, 60 o C for 30 sec,

and 72 o C for 1 min], followed by incubation at 72 o C for 5 min.

One primer pair capable of amplifying the RNA polymerase of noroviruses and sapporoviruses was used in this study (Table 3) (57). Human genotype II cDNA, extracted and reverse transcribed from Freon extracted norovirus genotype II from stool, was used as a positive control. This primer pair produced an amplicon of 319 bp for noroviruses and 331 bp for sapporoviruses. The 50 µl PCR reaction mixture was composed of 2 µl of target cDNA, 1X REDTaq PCR Reaction Buffer (10.0 mM Tris-HCl

pH 8.3, 50.0 mM KCl, 1.1 mM MgCl 2, 0.01% gelatin; Sigma-Aldrich © St. Louis, MO,

USA), an additional 0.4 mM MgCl 2, 100 µg/ml BSA, 0.25 mM each dNTP, 1 µM of each

primer, and 1 U REDTaq DNA Polymerase (Sigma-Aldrich © St. Louis, MO, USA). The

reaction mixture was incubated for 3 min at 94 oC, followed by 40 cycles of [94 o C for 30

s, 49 o C for 80 s, and 72 o C for 1 min], and a final extension of 72 o C for 10 min.

Nested PCR targeting the S gene was used to analyze the presence of hepatitis B virus (Table 3) (60). The initial amplicon had a length of 310 bp and the final amplicon was 241 bp. Human hepatitis B DNA (number 45020 ATCC®, Manassas, VA) was used as a positive control. Both rounds of PCR had 50 µl reaction volumes. The initial round

13 of PCR contained 2 µl of target DNA, 1X REDTaq PCR Reaction Buffer (10.0 mM Tris-

HCl pH 8.3, 50.0 mM KCl, 1.1 mM MgCl 2, 0.01% gelatin; Sigma-Aldrich © St. Louis,

MO, USA), an additional 0.4 mM MgCl 2, 0.25 mM each dNTP, 1 µM HBS-1 primer, 1

µM HBS-R1 primer, and 1 U REDTaq DNA Polymerase (Sigma-Aldrich © St. Louis,

MO, USA). The final round of PCR was the same as the initial round, except the template

was 2 µl of initial PCR product and the initial primers were replaced with 1 µM HBS-11

and 1 µM HBS-R11. The PCR conditions, used for both rounds of PCR, were as follows:

5 min at 95 oC, followed by 30 cycles of [95 o C for 30 s, 55 o C for 40 s, and 72 o C for 40 s].

A degenerate, nested primer set was used to detect a range of herpesviruses, including eight human strains (Table 3) (12). This nested reaction targeted the DNA polymerase and yielded an initial amplicon of 800 bp and a final amplicon between 215 bp and 235 bp. Human herpesvirus simplex 1 DNA was used as a positive control. The final volume of both PCR reactions was 50 µl. The first PCR reaction contained 1 µl of

extracted DNA, 1X REDTaq PCR Reaction Buffer (10.0 mM Tris-HCl pH 8.3, 50.0 mM

KCl, 1.1 mM MgCl 2, 0.01% gelatin; Sigma-Aldrich © St. Louis, MO, USA), 0.25 mM each dNTP, 1 µM FP1 primer, 1 µM FP2 primer, 1 µM FR1 primer, and 1 U REDTaq

DNA Polymerase (Sigma-Aldrich © St. Louis, MO, USA). The second PCR reaction was identical to the first, except the template was 5 µl of the first round product and primers were replaced with 1 µM FP3 and 1 µM RP2. Both reactions were incubated for 2 min at

94 oC, followed by 55 cycles of [94 o C for 20 s, 46 o C for 30 s, and 72 o C for 30 s],

followed by incubation at 72 o C for 10 min.

14 An assay able to detect 87% of human papillomaviruses using a pair of degenerate primers was utilized to screen samples (Table 3) (40). These primers targeted the L1 gene and amplified a 478 bp product with a sensitivity ranging from 1-10 copies.

Human papillomavirus 11 DNA (number 45151 ATCC®, Manassas, VA) was used as a positive control. The total reaction mixture, containing 2 µl of target DNA, was 50 µl and had final concentrations of 1X REDTaq PCR Reaction Buffer (10.0 mM Tris-HCl pH

8.3, 50.0 mM KCl, 1.1 mM MgCl 2, 0.01% gelatin; Sigma-Aldrich © St. Louis, MO,

USA), an additional 1.4 mM MgCl 2, 0.25 mM each dNTP, 1 µM of each primer, and 1 U

REDTaq DNA Polymerase (Sigma-Aldrich © St. Louis, MO, USA). The PCR reaction was incubated for 10 min at 94 oC, followed by 45 cycles of [94 o C for 1 min, 50 o C for 1

min, and 72 o C for 1 min], followed by incubation at 72 o C for 5 min.

One primer set, amplifying a region of the phosphoprotein gene, was used to detect the presence of morbilliviruses and produced a 429 bp amplicon (Table 3) (11).

Measles virus cDNA, extracted and reverse transcribed from Measles virus Edmonston strain particles (number VR-24 ATCC®, Manassas, VA), was used as a positive control.

Two µl of cDNA was added to a 48 µl PCR reaction mixture of 1X REDTaq PCR

Reaction Buffer (10.0 mM Tris-HCl pH 8.3, 50.0 mM KCl, 1.1 mM MgCl 2, 0.01% gelatin; Sigma-Aldrich © St. Louis, MO, USA), an additional 0.4 mM MgCl 2, 1 µM of each primer, 0.25 mM each dNTP, and 1 U REDTaq DNA Polymerase (Sigma-Aldrich ©

St. Louis, MO, USA). The PCR reaction underwent 35 cycles of [94 o C for 1.5 min, 25 o C for 2 min, and 72 o C for 2 min], followed by incubation at 72 o C for 7 min.

The presence of reoviruses and rotaviruses, both members of the Reoviridae family, were analyzed in this study. The details of the primer sets used in PCR

15 amplification of each of these groups is found in Table 3. The reovirus PCR assay, consisting of a nested reaction targeting the L1 gene, yielded a 344 bp amplicon (61).

This assay was able to detect one copy of 44 reovirus isolates from human, bovine, and murine strains. Human reovirus cDNA, extracted and reverse transcribed from human reovirus type 1 strain Lang (number VR-230 ATCC®, Manassas, VA), was used as a positive control.

For the primary PCR reaction, 2 µl of target cDNA was added to the 50 µl reaction mixture with final concentrations of the following components: 1X REDTaq

PCR Reaction Buffer (10.0 mM Tris-HCl pH 8.3, 50.0 mM KCl, 1.1 mM MgCl 2, 0.01% gelatin; Sigma-Aldrich © St. Louis, MO, USA), an additional 1.4 mM MgCl 2, 0.25 mM each dNTP, 1 µM L1.rv5 primer, 1 µM L1.rv6 primer, and 1 U REDTaq DNA

Polymerase (Sigma-Aldrich © St. Louis, MO, USA). The secondary PCR reaction was carried out by adding 1 µl of product from the primary PCR reaction to a 25 µl reaction mixture containing: 1X REDTaq PCR Reaction Buffer (10.0 mM Tris-HCl pH 8.3, 50.0 mM KCl, 1.1 mM MgCl 2, 0.01% gelatin; Sigma-Aldrich © St. Louis, MO, USA), an additional 0.9 mM MgCl 2, 0.25 mM each dNTP, 1 µM L1.rv7 primer, 1 µM L1.rv8

primer, and 1 U REDTaq DNA Polymerase (Sigma-Aldrich © St. Louis, MO, USA).

Both primary and secondary reactions were initially incubated for 1 min at 94 oC, followed by 35 cycles of [94 o C for 20 s, 50 o C for 30 s, and 72 o C for 30 s], followed by

incubation at 72 o C for 10 min.

Rotaviruses were detected using the primer set described in Table 3. This nested

PCR assay targeted the VP7 gene of group A rotaviruses and produced an initial 1059 bp amplicon and a final amplicon of 346 bp (43). It has been previously used to detect

16 rotaviruses in river water. Simian SA-11 cDNA, extracted and reverse transcribed from viral particles extracted from tissue, was used as a positive control. The initial PCR reaction had 2 µl of target cDNA and a total volume of 50 µl with the

following concentrations: 1X REDTaq PCR Reaction Buffer (10.0 mM Tris-HCl pH 8.3,

50.0 mM KCl, 1.1 mM MgCl 2, 0.01% gelatin; Sigma-Aldrich © St. Louis, MO, USA), an additional 0.4 mM MgCl 2, 1 µM R1 primer, 1 µM R2 primer, 0.25 mM each dNTP, 2

µg/ml BSA, and 1 U REDTaq DNA Polymerase (Sigma-Aldrich © St. Louis, MO,

USA). The initial PCR reaction was incubated for 1 min at 94 oC, followed by 25 cycles

of [94 o C for 30 s, 55 o C for 30 s, and 72 o C for 1 min], followed by incubation at 72 o C

for 3 min. Then 2 µl of product from the initial reaction was added to the final PCR

reaction. The final 50 µl PCR reaction was composed of the following: 1X REDTaq PCR

Reaction Buffer (10.0 mM Tris-HCl pH 8.3, 50.0 mM KCl, 1.1 mM MgCl 2, 0.01% gelatin; Sigma-Aldrich © St. Louis, MO, USA), an additional 2.4 mM MgCl 2, 1 µM R3 primer, 1 µM R4 primer, 0.25 mM each dNTP, 2 µg/ml BSA, and 1 U REDTaq DNA

Polymerase (Sigma-Aldrich © St. Louis, MO, USA). The final PCR reaction was incubated for 1 min at 94 oC, followed by 40 cycles of [94 o C for 30 s, 55 o C for 30 s, and

72 o C for 30 s], followed by incubation at 72 o C for 3 min.

PMMoV were detected using primers that amplify a 201 bp region of the helicase

gene (98). Two microliters of cDNA was added to a 50 µl final volume PCR reaction

mixture of 1X REDTaq PCR Reaction Buffer (10.0 mM Tris-HCl pH 8.3, 50.0 mM KCl,

1.1 mM MgCl 2, 0.01% gelatin; Sigma-Aldrich © St. Louis, MO, USA), 1 µM of each primer, 0.25 mM each dNTP, and 1 U REDTaq DNA Polymerase (Sigma-Aldrich © St.

Louis, MO, USA). The PCR reaction underwent the following: incubation at 95 o C for 5

17 min, followed by 40 cycles of [94 o C for 1 min, 50 o C for 45 sec, and 72 o C for 1 min], and an incubation at 72 o C for 5 min. Raw sewage was used as a positive control.

A portion of the genomic segment 2 of genotype I human picobirnaviruses

ranging from 123 to 246 bp was amplified using the primer set described in Table 3 (83).

Genotype I human picobirnaviruses were targeted for this study since genotype II

picobirnaviruses are rarely found in human populations (13). These primers have been

previously used to detect genotype I picobirnaviruses in feces (13). Raw sewage was used

as a positive control. The PCR reaction had a final volume of 50 µl and had 2 µl of target

cDNA, 1X REDTaq PCR Reaction Buffer (10.0 mM Tris-HCl pH 8.3, 50.0 mM KCl, 1.1

mM MgCl 2, 0.01% gelatin; Sigma-Aldrich © St. Louis, MO, USA), an additional 0.9 mM MgCl 2, 1 µM of each primer, 0.25 mM each dNTP, and 1 U REDTaq DNA

Polymerase (Sigma-Aldrich © St. Louis, MO, USA). The PCR reaction was incubated for 2 min at 94 oC, followed by 40 cycles of [94 o C for 1 min, 49 o C for 2 min, and 72 o C

for 3 min], followed by incubation at 72 o C for 7 min.

Characterizing the Diversity of Picobirnaviruses in Raw Sewage

In order to gain a deeper understanding of the diversity of picobirnaviruses in raw

sewage, all positive PCR products were cloned into pCR4-TOPO (Invitrogen © Carlsbad,

CA, USA) and transformants were screened for inserts by PCR with the M13F and M13R

primers. UltraClean PCR Clean-up Kit (MO BIO Laboratories, Inc., Carlsbad, CA, USA)

was used to purify PCR products of the proper size. Positive PCR products were

sequenced using M13F. All sequences were trimmed using Sequencher (Gene Codes ©

Ann Arbor, MI, USA) and the identity of the positive PCR products were confirmed by

18 comparing the sequence against the GenBank non-redundant database using BLASTN and TBLASTX (4, 5).

All sequences with insignificant identities (E-values > 0.001 and/or identities less

than 50 bp in length) to known picobirnaviruses were removed. All of the sequences from

raw sewage were de-replicated to 99% sequence identity with gaps using the online

program FastGroup II (97). The de-replication process and the representative sequences

for each group were visually confirmed. In addition to de-replicating sequences,

FastGroup II also executed Chao1 (24, 25) and rarefaction analyses (49, 53),

measurements of diversity (97). The phylogenetic relationship of the picobirnaviruses

detected in raw sewage to known picobirnaviruses was identified through aligning de-

replicated raw sewage sequences with known sequences from GenBank. These known

sequences from GenBank had the best BLASTN identity to at least one of the sequences

from raw sewage. All sequences were trimmed to ~200 bp. Since reference sequence S

and T were less than 200 bp, they were not included in the alignment. Alignments were

executed using ClustalX v.1.8 (89). The resulting alignment was verified visually for

accuracy and then imported into MEGA v. 4 for phylogenetic analyses (88).

The phylogenetic analysis of cloned picobirnavirus PCR products from raw

sewage was executed similar to previous picobirnavirus studies (13, 83). Prior to creating

a phylogenetic tree, the average pairwise Jukes-Cantor distance was calculated in order to

determine the appropriateness of creating a neighbor joining tree. Since the average pair

wise Jukes-Cantor distance was less than 1.0 (0.356), a neighbor joining phylogenetic

tree was calculated using a Jukes-Cantor model along with bootstrap analyses for 2000

19 replicates. Branches with insignificant bootstrap values ( ≤ 50) were condensed to create

the final phylogenetic tree.

Determination of Assay Sensitivity

The amount of nucleic acid in the cleaned up PCR positive control after completion

of PCR amplification was quantified using a NanoDrop ND-1000 (NanoDrop

Technologies ©, Wilmington, DE, USA). If the PCR reaction was nested, then the

positive control from the first round of PCR amplification was quantified. The number of

targets was back-calculated using the measured concentration of DNA and knowledge of

the amplicon size using the following equation, where Y is the concentration of DNA

measured by the Nanodrop and Z is the length of the amplicon:

Y ng 1 g 1 mol bp 02.6 x10 23 bp 1 target # targets * * * * = 1 µl 10 9 ng 660 g 1 mol bp Z bp µl

The cleaned up PCR positive control was then serially diluted to 1 target/ µl. The entire

dilution series underwent the appropriate PCR or nested PCR previously described for

that viral group, with the exception that only 1 µl of target DNA/cDNA was added to the

reaction mixture. PCR products were visualized on a 2% agarose gel stained with

ethidium bromide and the sensitivity of the assay was determined.

Isolation of Viruses from Large-scale Seawater Samples

A total of four 200 liter samples of seawater were collected from four different sites

ranging in their proximity to dense human populations. These environmental samples

were ultimately analyzed for any viruses detected in raw sewage. Seawater was collected

from North Shore Beach (N27 °46’57.42”, W82 °37’26.63”), a frequently polluted Saint

20 Petersburg City, Florida beach, on the morning of 25 September 2007. Saint Petersburg, located along Tampa Bay, is home to nearly 250,000 people (U.S. Census Bureau, 2006

Population Estimates) and serves as the beach closest to dense human populations in this study. Based upon water quality analyses conducted by the Pinellas County Health

Department using current bacterial indicators of fecal pollution, North Shore beach had

‘good’ concentrations of fecal coliform (<200 organisms/100ml seawater) and enterococcus (<36 organisms/100 ml seawater) on the day of sampling. However, North

Shore beach had been closed for nearly two weeks prior to sampling due to high concentrations of bacterial indicators.

Seawater was also collected from three other locations. One sample was gathered

from a canal with a nearby injection well in Lower Matacumbe Key, Florida

(N24 °51’29.26”, W80 °43’40.27”) on 10 Dec 2006. This site in Lower Matacumbe is

located adjacent to an on-site disposal system and surrounded by a moderate human

population. In past studies, large amounts of enteric viruses have been found this Lower

Matacumbe Key sample site (45, 96). Another seawater sample was collected from a near

shore reef at Looe Key buoy #10 (about N24 °33’03.16”, W81 °24’27.51”) on 11 Dec

2006. Due to Looe Key’s proximity to shore, it will serve as the second furthest location

from dense human population. Aboard the R/V Bellows, the final seawater sample was

collected 31 May 2007 offshore of the Dry Tortugas (N24 °36.548’, W82 °50.464’). This site is furthest from shore, closest to the Gulf Stream, and furthest from dense human populations.

21 Saint Petersburg, FL

Lower Matacumbe, FL

Looe Key, FL

Dry Tortugas Decreasing proximity to dense human populations populations human dense to Decreasing proximity Influence Human

FIGURE 1. Sites of large-scale seawater sampling in Florida. The colored dots represent the four areas sampled: North Shore Beach (blue), Lower Matacumbe canal (red), Looe Key (yellow), and offshore the Dry Tortugas (green).

Large-scale seawater samples collected from North Shore Beach, the Florida Keys,

and off-shore the Dry Tortugas were concentrated for viral isolation using a combination

of differential filtration with tangential flow filters (TFF), density-dependent

centrifugation in cesium chloride (CsCl), and polyethylene glycol (PEG) precipitation

(18, 22). All of the size-based filtration and density-dependent isolation steps were

chosen based upon the sizes and densities of the viral families under investigation (Table

4). The particles larger than the TFF pore size are concentrated into the retentate;

conversely, the filtrate will contain particles smaller than the TFF pore size. First, a 0.2

µm TFF (GE Healthcare, Piscataway, NJ) was used to remove bacteria, eukaryotes, and other large particles from 200 liters of seawater. The 0.2 µm TFF retentate, containing

22 everything > 0.2 µm, was filtered through a 0.22 µm polyether sulfone membrane filter cartridge (Sterivex; Millipore, Billerica, MA) and the filtrate was added to the 0.2 µm

TFF filtrate. The bacteria and eukaryotes concentrated into the 0.2 µm TFF retentate were

collected on the 0.22 µm polyether sulfone membrane filter cartridge.

Then the resulting 0.2 µm TFF filtrate, containing the viral particles, was

concentrated through a 100 kDa TFF (GE Healthcare, Piscataway, NJ) to generate a final

volume of <100 ml. It is possible that herpesviruses and morbilliviruses may have been

excluded from the viral concentrate by the initial 0.2 µm filtering. If either of these viruses were detected in raw sewage, then total DNA or RNA was extracted directly from the 0.2 µm retentate collected on the 0.22 µm polyether sulfone membrane filter cartridge

(Sterivex; Millipore, Billerica, MA) using the PowerSoil DNA or RNA Isolation Kit

(MO BIO Laboratories, Inc., Carlsbad, CA, USA). Purification of the viral concentrate

and removal of potential PCR inhibitors was achieved through loading the sample on to a

cesium chloride (CsCl) density gradient. The gradient was ultracentrifuged at 61,120 X g

(22,000 rpm in a SW40Ti rotor) at 4 °C for 2 hours and the 1.2-1.5 g/ml fraction (i.e.

CsCl Viral fraction) was collected to isolate the target viruses. The remaining CsCl

density gradient (i.e. CsCl Other fraction) was also collected to measure the effectiveness

of the CsCl density gradient via epifluorescent microscopy.

PEG precipitation was used to further concentrate each sample before nucleic acid

extraction (86). A final concentration of 10% PEG 8000 and 1 M NaCl was added to the

viral concentrate. After the mixture was incubated overnight at 4 °C, it was spun for 30

min at a temperature of 4 °C and a speed of 13,000 X g. The supernatant was poured off

and the pellet was re-suspended. Two samples deviated slightly from the above

23 concentration steps. For seawater samples collected from North Shore Beach and the Dry

Tortugas, viruses were further concentrated using PEG precipitation before density- dependent centrifugation in cesium-chloride. Hereafter this method will be referred to as the “TFF Method”. An additional concentration step using centrifugal concentration filters (Centriplus YM-50; Millipore, Billerica, MA, USA) was used for seawater collected from the Dry Tortugas.

TABLE 4. Characteristics of viral families found in raw sewage (35).

Viral Family Virus Size Range (nm) Density CsCl (g/cm 3) Adenoviridae Human Adenoviruses, A - F 70 - 90 1.30 - 1.37 Caliciviridae Noroviruses & Sapporoviruses 27 - 40 1.33 - 1.41 Hepadnaviridae Hepatitis B Viruses 42 - 50 1.25 Herpesviridae Human Herpesviruses 125 - 200 1.22 - 1.28 Papillomaviridae Human Papillomaviruses 55 1.34 - 1.35 Paramyxoviridae Morbilliviruses > 150 1.23 Mammal Reoviruses Reoviridae 60 - 80 1.36 - 1.39 Rotaviruses, Group A Unassigned Pepper Mild Mottle Viruses 35 - 41 1.38 - 1.40 Unclassified Human Picobirnaviruses 30 - 40 1.4

To ensure purification and concentration of the viral sample, a portion of the

concentrate after each step was stained with SYBR Gold nucleic acid stain (Invitrogen ©

Carlsbad, CA, USA) and direct counts were made using epifluorescent microscopy (75).

QIAamp MinElute Virus Spin Kits (QIAGEN ©, Valencia, CA, USA) were used to

isolate DNA and RNA from 200 µl of each of the purified, final viral concentrates. In

order to detect RNA viruses, cDNA was immediately synthesized from viral RNA using

random hexamers and the First Strand Synthesis Superscript III Reverse Transcription

Kit (Invitrogen © Carlsbad, CA, USA). Whole genome amplication using GenomiPhi

V2 (GE Healthcare, Piscataway, NJ, USA) was utilized for each seawater sample from

24 North Shore Beach, the Florida Keys, and off-shore the Dry Tortugas to increase the concentration of total DNA and cDNA prior to virus identification.

Seawater Viral Isolation Methods Comparison

In order to confirm and compare the effectiveness of the large-scale seawater viral

concentration, purification, and nucleic acid extraction techniques employed in this study,

four spiking experiments were completed to compare the TFF Method previously

described for North Shore Beach seawater to widely accepted Public Health Method (58).

Over the last 10 years, the TFF Method has been developed to concentrate viruses from

seawater. These methods include the use of TFF filters and subsequent purification and

concentration steps (e.g. density-dependent centrifugation and PEG precipitation). The

TFF Method was modified for the comparison experiments to a starting volume of

seawater of 15 liters. The Public Health Method used in this experiment were a modified

version of the Katayama et al. method and consisted of the following steps (39).

Acetic acid (10%) was added to 1 liter of seawater until the sample was acidified to

a pH between 3.5 and 4.0. Following acidification, the sample was filtered through a type

HA, negatively charged, 0.47 mm diameter, 0.45 µm pore size membrane (Millipore,

Billerica, MA). One hundred milliliters of H 2SO 4 (0.5mM) was filtered through the

membrane and viruses were eluted using 10 ml of NaOH (1mM). The viral eluate was

collected in a container with 100 µl of H 2SO 4 (50mM) and 100 µl of 100X Tris-EDTA

(TE) buffer to neutralize the viral concentrate. Viral concentrates were stored at -20 oC until further concentrated using a centrifugal concentration device (Centriplus YM-50;

Millipore, Billerica, MA, USA) to a volume less than 2 ml. For both methods, nucleic

25 acid was extracted from 200 µl of viral concentrate using the MinElute Virus Spin Kit

(QIAGEN ©, Valencia, CA, USA). To detect RNA viruses, cDNA was synthesized immediately after nucleic acid extraction using random hexamers in the First Strand

Synthesis Superscript III Reverse Transcription Kit (Invitrogen © Carlsbad, CA, USA).

A total of three spiking experiments were executed to assess the concentration efficiency of the TFF and Public Health Methods. Three experiments were executed to determine the methods’ efficiency at concentrating different types of viruses (DNA and

RNA) as well as the methods’ ability to separate viruses potentially attached to particles.

For the first experiment, the effectiveness of each method was tested by initially spiking water samples with a final concentration of 6.14Ex10 2 poliovirus particles per milliliter of seawater. Nested PCR, with a reported sensitivity of about 4 viral particles, was executed using the primers in Table 5 on cDNA synthesized from each method (39). The sensitivity of this assay was confirmed by calculating the number of Poliovirus particles cDNA was synthesized from, serially diluting the poliovirus cDNA to 10 -5 (~ 4 poliovirus

particles), and completing the nested PCR reaction on each dilution.

To roughly determine the efficiency of each method, the cDNA from each method

was serially diluted to 10 -6 and nested PCR was also carried out on each dilution. For the first round of PCR, the total reaction volume was 25 µl and contained 3 µl of cDNA, 2X

Promega PCR Master Mix (3 mM MgCl 2, 50 U/ml TaqDNA Polymerase pH 8.5, 0.40

mM each dNTP; Promega ©, Madison, WI, USA), 1 µM JP UP primer, and 1 µM PAN

ENT DWN primer. The first round PCR reaction underwent 40 cycles of [95 oC for 30 s,

57.7 o C for 30 s, and 72 o C for 45 s] and an extension of 72 oC for 5 min. The second round of PCR, 25 µl, contained 1 µl of the first round PCR product, 1X REDTaq PCR

26 Reaction Buffer (10.0 mM Tris-HCl pH 8.3, 50.0 mM KCl, 1.1 mM MgCl 2, 0.01% gelatin; Sigma-Aldrich © St. Louis, MO, USA), an additional 1 mM MgCl 2, 1 µM PAN

ENT UP primer, 1 µM JP DWN primer, 0.25 mM each dNTP, and 1 U REDTaq DNA

Polymerase (Sigma-Aldrich © St. Louis, MO, USA). The second round PCR reaction

underwent 40 cycles of [95 oC for 30 s, 56.5 o C for 30 s, and 72 o C for 30 s] and an extension of 72 oC for 5 min.

TABLE 5. Primer nucleotide sequences used for PCR amplification of poliovirus (38).

Primer Name Primer Sequence 5' - 3' Nested Target Gene Amplicon JP UP TTA AAA CAG CTC TGG GGT TG PAN ENT DWN CTA ACC GGT AGG CCA Outside untranslated 600 bp PAN ENT UP CCT CCG GCC CCT GAA TG region JP DWN CCG ACG AAT ACC ACT GTT A Inside 154 bp

The second method efficiency comparison was completed following the same

protocol as previously described except seawater samples were spiked with 4.24x10 4

Adenovirus-20 particles per milliliter of seawater. Nested PCR was used to identify the presence of these viruses in the extracted DNA and the 1:10 serial dilutions of this DNA for each method using previously published conditions and primers in Table 3 (3). The day before the third trial, 500 ml of seawater was spiked with Adenovirus-20. Following incubation overnight at room temperature while horizontally shaking, the third trial was executed as described for the second trial. This last trial was executed to understand if the two methods’ concentrating efficiency changed after allowing the viruses to attach to particles overnight.

Determination of Viral Stability in Seawater

The stability of PMMoV and picobirnaviruses in seawater was approximated by spiking 950 ml of seawater with 50 ml of raw sewage in a 1 liter clear plastic bottle. Two

27 sets of experiments were executed, beginning on 6 February 2008 and 19 March 2008.

For both experiments, ten milliliter samples were collected prior to spiking as a control

(T c). Before this bottle was placed off the seawall in Bayboro Harbor (Saint Petersburg,

Florida), the sewage spiked seawater was mixed for 10 minutes and an initial 10 ml sample (T o) was collected. In the first experiment, samples were subsequently taken on days 1 (T 1), 2 (T 2), 3 (T 3), 4 (T 4), 5 (T 5), 8 (T 8), and 39 (T 39 ). During this time period, the seawater surface temperature fluctuated between 20 °C and 24 °C (NOAA CO-OPS station 8726520). The second experiment measured the stability of the viruses over a similar period of time but samples were collected on days 4 (T 4), 7 (T 7), 14 (T 14 ), 21

(T 21 ), and 28 (T 28 ). The seawater surface temperature varied between 22 °C and 27 °C

during this time (NOAA CO-OPS station 8726520).

Nucleic acid from the viruses in each 10 ml sample was isolated the same day they

were collected using the methods previously outlined for wastewater. cDNA was

synthesized from the extracted RNA using First Strand Synthesis Superscript III Reverse

Transcription Kit (Invitrogen © Carlsbad, CA, USA) with random hexamers. The

presence of PMMoV and Picobirnaviruses was determined through PCR using the

methods previously described for identification in raw sewage. In the presence of PCR

products of many sizes, the identity of key positive PCR products (control and last day

seen) were confirmed through purifying PCR products and sequencing as previously

described.

28

RESULTS

Viral Detection in Raw Sewage

Raw sewage samples were collected from 12 wastewater treatment facilities located

in 11 different states. The nucleic acid of viruses from 10 ml of raw sewage was isolated

and analyzed for the presence of ten types of viruses (Table 6). Three types of viruses

were found in 100% of the raw sewage samples: adenoviruses, picobirnaviruses, and

PMMoV. Noroviruses were detected in 7 out of 12 samples. After whole genome

amplification, papillomaviruses were identified in 2 out of 12 samples. Herpesviruses,

reoviruses, rotaviruses, morbilliviruses, and hepadnaviruses were not detected in any of

the raw sewage samples; however, it is possible that they were present in concentrations

below the detection limit of their assays.

TABLE 6. Summary table of the targeted viruses detected in 12 raw sewage samples collected from throughout the coastal United States (two samples came from Florida). Positive identities to known targeted viruses in GenBank were those with a BLASTN hit with an E value ≤ 0.001. All positive identifications are from original nucleic acid samples unless otherwise marked. Two asterisks illustrate that the positive identifications were from GenomiPhi amplified samples.

Viral Group Raw Sewage Adenoviruses 100% (12/12) Noroviruses & Sapporoviruses 58.3% (7/12) Hepadnaviruses 0% Herpesviruses 0% Morbilliviruses 0% Papillomaviruses ** 16.7% (2/12) Picobirnaviruses 100% (12/12) PMMoV 100% (12/12) Reoviruses 0% Rotaviruses 0%

29 Viral Detection in Final Effluent

Final effluent samples were collected on the same day of raw sewage collection

from 11 states throughout the U.S. Viruses in 10 ml of final effluent were concentrated

and their nucleic acid was extracted to identify the ten viruses under study. The results of

the PCR analyses are listed in Table 7. PMMoV was detected in all but one final effluent

sample. In addition to detecting PMMoV, four other viruses were found. Picobirnaviruses

was identified in a third of the final effluent samples, including one sample that

underwent whole genome amplification prior to PCR. Three of the 12 final effluent

samples were positive for adenoviruses. Finally, noroviruses and reoviruses (after whole

genome amplification) were identified in only one of the final effluent samples.

TABLE 7. Summary table of the targeted viruses detected in 12 final effluent samples collected from 11 different coastal states. Positive identities to known targeted viruses in GenBank were those with a BLASTN hit with an E value ≤ 0.001. All positive identifications are from original nucleic acid samples unless otherwise marked. An asterisk denotes a positive identification from a GenomiPhi amplified sample.

Viral Group Final Effluent Adenoviruses 25% (3/12) Noroviruses & Sapporoviruses 8.3% (1/12) Hepadnaviruses 0% Herpesviruses 0% Morbilliviruses 0% Papillomaviruses 0% Picobirnaviruses *33.3% (4/12) PMMoV 91.7% (11/12) Reoviruses *8.3% (1/12) Rotaviruses 0% PCR Assay Sensitivity

The sensitivity of the assays used to identify the viruses under investigation ranged from 1 target to 100,000 targets and was identical to those previously reported

(Table 8). According to the manufacturer, whole genome amplification increases the total amount of DNA or cDNA prior to PCR amplification up to 600-fold (GenomiPhi V2, GE

30 Healthcare, Piscataway, NJ, USA; reviewed by (14)); thus, the sensitivity of each assay was approximately increased 100- fold.

TABLE 8. The sensitivities of each PCR assay employed in this study, which is identical to previous studies.

Detection Limit GenomiPhi Adjusted Primer Reference Target Virus (targets) Detection Limit (targets) Human Adenoviruses, 3 100 1 A - F Noroviruses & 57 10000 100 Sapporoviruses 60 Hepatitis B Viruses 10000 100

12 Herpesviruses 10 1

40 Papillomaviruses 100 1

11 Morbilliviruses 10000 100

61 Mammal Reoviruses 1 1 Rotaviruses, 43 100000 1000 Group A 98 PMMoV 100 1 Human 83 1000 10 Picobirnaviruses

Wastewater Viral Concentration Efficiency

The efficiency of the centrifugal concentration devices used to concentrate raw

sewage and final effluent were evaluated by spiking autoclaved samples with a final

concentration of 8.48x10 5 adenovirus particles/ml. Both the autoclaved un-spiked raw

sewage and final effluent samples were run in parallel to the spiked samples and the lack

of PCR product ensured the elimination of pre-existing adenovirus particles (far right

well in Figure 2 & Figure 3). This methods experiment produced identical results for raw

sewage and final effluent. Assuming near 100% recovery of adenovirus DNA from the

nucleic acid extraction kit, a 10 -3 dilution (equivalent to ~100 viruses) of the original

31 DNA should have resulted in an amplicon. For both samples, a dilution of 10 -3 was the most diluted DNA sample to produce a PCR product (Figure 2 & Figure 3), which suggests that the methods used to concentrate wastewater samples were successful in concentrating viruses.

FIGURE 2. Agarose gel electrophoresis (2%) of adenovirus PCR products from the raw sewage spiking experiment. The numbers in each lane indicate the extent of DNA dilution prior to PCR amplification. ‘C’ represents the spiking experiment control. This control was autoclaved raw sewage.

32 Diversity of Picobirnaviruses in Raw Sewage

Since picobirnaviruses were found in all U.S. raw sewage samples and limited

knowledge exists on this virus, the diversity of picobirnaviruses in raw sewage was

analyzed. A total of 288 (~ 22 per location) cloned picobirnavirus PCR products of the

RNA-dependent RNA polymerase gene were sequenced by Agencourt® (Beverly, MA,

USA) and 72% of these sequences had significant identities (E-value ≤ 0.001 and

identities over 50 bp) to known picobirnaviruses in GenBank. Those sequences without

significant BLASTN hits also did not show significant TBLASTX similarities to

GenBank, and were therefore not used in subsequent analyses. The significant BLASTN

sequence identities ranged from 83% to 100% over regions ranging from 56 bp to 194 bp

to known human and porcine picobirnaviruses. . The sequences from GenBank with the

greatest identity to the sequences from raw sewage were extracted for alignment purposes

(Table 9).

33 TABLE 9. The top BLASTN hits of cloned picobirnavirus PCR products from U.S. raw sewage samples. These sequences were included in the alignment of raw sewage sequences as a means of understanding the identity and relatedness of picobirnaviruses present in raw sewage. Those sequences highlighted in gray were too short to include in the alignment.

Letter GenBank Accession Number Description A dbj|AB186898.1| Human B dbj|AB193349.1| Human C emb|AJ504794.1|HPI504794 Human strain 1-HUN-01 D emb|AM419115.1| Human isolate castellon-3880 E emb|AM706365.1| Porcine strain c10 clone 6 F emb|AM706366.1| Porcine strain c10 clone 14 G emb|AM706367.1| Porcine strain D4 clone 1 H emb|AM706368.1| Porcine strain D4 clone 2 I emb|AM706374.2| Porcine strain D6 clone 1 J emb|AM706379.1| Porcine strain D6 clone 10 K emb|AM706380.1| Porcine strain D6 clone 11 L emb|AM706397.1| Porcine strain E4 clone 14 M emb|AM706399.1| Porcine strain E4 clone 30 N gb|AF245701.1|AF245701 Human strain 2-GA-91 O gb|AF246612.1|AF246612 Human strain 1-GA-91 P gb|AF246935.1|AF246935 Human strain 202-FL-97 Q gb|AF246936.1|AF246936 Human strain 203-FL-97 R gb|AY805390.1| Human strain 745-ARG-99 S gb|EU104359.1| Porcine strain 1a T gb|EU104360.1| Porcine strain 2 U gb|EU104362.1| Porcine strain 4

Prior to alignment, a total of 70 groups were created from the 207 raw sewage picobirnavirus sequences by FastGroup II (97) set to group sequences with 99% identity with gaps. While the majority of groups were composed of one or two sequences, seven groups contained 3 sequences or more. The de-replicated sequences along with the top

BLASTN hits (Table 9) were aligned in ClustalX v.1.8 (89) and phylogenetic analyses were carried out in MEGA v. 4 (88). Given an average pair wise Jukes-Cantor distance of

0.356, a neighbor joining tree was created using the Jukes-Cantor model with bootstrap replication of 2000. The final phylogenetic tree with insignificant (bootstrap value < 50) branches condensed is displayed in Figure 4.

34

FIGURE 4. A condensed, neighbor joining (Jukes-Cantor model) phylogenetic tree of a ~200 bp segment of the picobirnavirus RNA-dependent RNA polymerase gene from 207 U.S. raw sewage sequences. Each branch represents a sequence or a group of sequences (99% identical with gaps) depending upon the block color. Every colored block is numbered and each number corresponds to the adjacent table. This table explains the locations represented by each block. Each letter represents a reference sequence from GenBank (Table 9). The pink pig next to a letter represents porcine sequences. Letters without a pig indicate that the sequence is from a human.

35 Three main points can be interpreted from the condensed phylogenetic tree and table in Figure 4. First, the high number of single sequences (blue blocks) suggest that the

RNA-dependent RNA polymerase of picobirnaviruses from raw sewage has a great deal of sequence diversity. Furthermore, it suggests that more cloned PCR products would need to be sequenced to fully understand the diversity of picobirnaviruses. The need for further sequencing is further supported by the calculated Chao1 value and the rarefaction curve created in FastGroup II (97). Chao1 is a minimum estimator of the number of unique sequences (i.e. richness) (24, 25). Chao1 predicted a minimum of 200 unique picobirnaviruses in U.S. raw sewage but only 70 were sampled in this study. This suggests that more clones need to be sequenced in order to describe adequately the diversity of picobirnaviruses in U.S. raw sewage.

The rarefaction analyses executed in FastGroupII further support the need to sequence more clones to gain a complete view of picobirnavirus sequence diversity in

U.S. raw sewage (97). The rarefaction analyses plot the number of unique sequences versus the number of clones sequenced (Figure 5) (49, 53). Sufficient clones have been sequenced when the curve reaches an asymptote. Since it is unclear where the curve would asymptote, more clones will need to be sequenced in order to ensure a complete analysis of picobirnavirus diversity in U.S. raw sewage.

36

FIGURE 5. Rarefaction analyses of the picobirnavirus sequences from U.S. raw sewage.

Some of the picobirnavirus sequences from raw sewage had the greatest identities to porcine picobirnaviruses in GenBank (Table 9) and this grouping is illustrated in the condensed phylogenetic tree. It is also interesting to note that the picobirnavirus sequences from raw sewage do not seem to have any sort of geographic distribution based upon visual analysis. The blocks on the phylogenetic tree not colored blue show that identical (or >99% identical) sequences were recovered from multiple states. For example, 8 out of 10 sequences from North Carolina were unique (or <99% identical).

Viral Isolation and Detection in Large-Scale Seawater Samples

Viral concentrates from all large-scale seawater samples were collected and purified using a combination of TFF filtration, density-dependent centrifugation in CsCl, and PEG precipitation. The combination of these protocols effectively concentrated viruses 10,000 fold from seawater samples (Table 10). Despite successful viral

37 concentration, some viruses were lost in the PEG supernatant and CsCl Other fractions

(Table 10). The amount of viruses lost is approximately equivalent to concentrating

100% of viruses in 2.5 L of seawater. Although viruses were lost during the concentration and purification processes, the results of the methods comparison experiments illustrate the effectiveness of the TFF Method at concentrating viruses.

TABLE 10. The concentration of viruses (viral-like particles/milliliter) at each step of the concentration and purification process for each seawater sample. The sample from which nucleic acid was extracted from is displayed in bold text.

Environmental Sample, Concentration VLP/ml TFF Method Step Lower Matacumbe Looe Key Dry Tortugas North Shore Beach

Original 4.84E+07 3.06E+06 3.06E+06 9.72E+07 TFF Concentrate 1.70E+10 5.75E+09 9.56E+08 2.24E+10 CsCl Concentrate 1.93E+10 1.07E+10 5.62E+09 1.72E+11 CsCl Other 5.20E+09 2.96E+08 1.37E+08 7.09E+09 PEG Concentrate 2.21E+11 1.18E+10 3.48E+09 2.47E+11 PEG Supernatant 2.11E+08 5.20E+06 1.53E+08 4.81E+08 Centriplus Concentrate 5.19E+10

All environmental samples were tested for the presence of PCR inhibitors and no

significant inhibition was observed. None of the viruses under investigation were

detected by PCR in any of the large-scale seawater samples, regardless of their proximity

to dense human populations. Therefore, the viruses found in raw sewage (adenoviruses,

noroviruses, picobirnaviruses, and PMMoV) were either absent in the marine

environment or present at concentrations below the detection limit of the assays at the

time of sampling. The presence of these viruses in raw sewage coupled with their absence

in the marine environment suggests that they would be good candidates for fecal

pollution indicators.

38 Seawater Viral Isolation Methods Comparison

Two methods, TFF and Public Health, for viral concentration were evaluated by

spiking seawater samples with known concentrations of viral particles. Although three

spiking experiments were executed using two different types of viruses, the results of all

three experiments concluded that the TFF Method was at least 10 times more effective at

concentrating viruses than the Public Health Method. Despite losing viruses in each step

of the viral concentration and purification method (Table 11), the TFF Method was able

to detect polioviruses better than the Public Health Method (Figure 6). Similar results

were noted in the adenovirus spiking experiments (Figure 7), including the third trial that

involved the incubation of adenovirus particles overnight (Figure 8). These experiments

prove that the TFF Method effectively concentrates DNA and RNA viruses. Furthermore,

the results of the third study indicate that the TFF Method can concentrate eukaryotic

viruses even after they are allowed to attach to particles. The sensitivities of the nested

PCR assays for poliovirus and adenovirus were equivalent to previous studies at ~ 4 virus

particles and ~100 virus particles respectively (3, 38, 39).

TABLE 11. Direct counts of virus-like particles (VLP) per milliliter of seawater for each step in the TFF Method’s concentration and purification protocol with samples from the three spiking experiments. Viruses were concentrated from seawater using TFF and PEG precipitation. The PEG Pellet was further purified by extracting the ‘viral’ fraction of the cesium chloride gradient after centrifugation. Viruses in the PEG supernatant and the CsCl Other fraction were lost. Nucleic acid was extracted from the CsCl Viral step highlighted in bold text.

Spiking Experiment Trials Spiked with Spiked with Spiked with Adenovirus 20 TFF Method Steps Poliovirus Adenovirus 20 (incubated overnight) VLP per ml VLP per ml VLP per ml Original Seawater 7.00E+06 1.22E+07 9.76E+06 TFF Viral Concentrate 5.63E+09 4.49E+09 6.36E+09 PEG Pellet 3.50E+10 2.39E+10 3.10E+10 PEG Supernatant 6.78E+08 8.90E+08 2.68E+08 CsCl Viral 1.34E+11 3.50E+10 8.04E+10 CsCl Other 4.31E+09 4.47E+08 2.30E+09

39

40

FIGURE 8. Agarose gel electrophoresis (2%) showing the effectiveness of the TFF Method (lanes 2-8) and the Public Health Method (lanes 9-15) from the third spiking experiment with an overnight incubation using adenoviruses. The numbers in each lane indicate the extent of DNA dilution prior to PCR amplification.

Stability of Pepper Mild Mottle Virus and Picobirnavirus in Seawater

The stability of two viruses in raw sewage, PMMoV and picobirnaviruses, was

estimated via a bottle experiment in which seawater was spiked with a known quantity of

raw sewage. This bottle was placed into the water off the seawall and sampled

intermittingly. Two trials of this basic experiment were executed and varied in their

duration. The viruses were detected using PCR and gel electrophoresis. The results of

these two experiments are displayed separately for picobirnaviruses (Figure 9) and

PMMoV (Figure 10). The results of the picobirnavirus stability experiments are difficult

to interpret due to the non-specific amplification of nucleic acid from seawater samples

(Figure 9). Based upon sequence confirmation, the faint band observed in the un-spiked

seawater (control) of the picobirnavirus experiments is actually to the chloroplast of

Ostrecoccus tauri . Despite the non-specific binding of the primers and lack of sequence confirmation, picobirnaviruses from raw sewage clearly remained stable in seawater no longer than 4 to 7 days.

41

FIGURE 9. Agarose gel electrophoresis (2%) depicting picobirnavirus PCR products (~200 bp) from two experiments assessing the stability of picobirnaviruses from raw sewage in seawater.

It is easier to interpret the stability of PMMoV than the stability of picobirnavirus due to the specificity of the primers (Figure 10). Although a light band appears in the control of experiments 1 and 2, the spiking of raw sewage is apparent in both experiments. PMMoV in raw sewage appears to be stable in seawater for approximately a week before returning to background levels. Since only a single PCR product was produced, these results were not confirmed via sequencing.

FIGURE 10. Agarose gel electrophoresis (2%) of PMMoV PCR products from two experiments assessing the stability of PMMoV from raw sewage in seawater.

42

DISCUSSION

Viruses Detected in Raw Sewage

The primary goal of this study was to identify possible viral indicators of fecal

pollution for the marine environment. In order to achieve this goal, the presence of ten

viral groups, known to be shed in human feces, were analyzed via PCR in raw sewage

collected from throughout the country. In summary, five different types of viruses were

detected in one or more of the raw sewage samples. Noroviruses, previously found in

European raw sewage (66, 91), was detected in less than 60% of the raw sewage samples;

thus, their use as a marker of fecal pollution could potentially underestimate the extent of

fecal contamination. Papillomaviruses were detected in 16.7% of raw sewage samples

after whole genome amplification. Although their presence is interesting,

papillomaviruses were found in too few samples to be considered a potential indicator of

fecal pollution.

Three of these viruses, adenoviruses, picobirnaviruses, and PMMoV, were found in

100% of the raw sewage samples. The results of this study support prior findings

regarding the prevalence of adenoviruses in raw sewage (15, 42, 90, 92, 94).

Furthermore, the use of adenoviruses to indicate fecal contamination has been

demonstrated by numerous studies in a variety of environments (1, 27, 32, 38, 55, 56, 64,

82). This study is the first of its kind to demonstrate widespread prevalence of

picobirnaviruses and PMMoV in raw sewage, suggesting that these viruses are good

43 indicators of sewage. In addition, the detection of these viruses in the marine environment will suggest a recent contamination event since neither were found in the marine environment in the absence of pollution.

Interestingly, rotaviruses and reoviruses were not among those viruses detected in raw sewage samples even though their presence in raw sewage has been documented in other countries (30) and they have been used in prior fecal pollution studies (7, 8, 16, 21,

59, 72, 81). It is possible that these viruses were present at concentrations below the detection limit of their assays and/or that they were not prevalent on the day of sampling.

Alternatively, it is plausible that these viruses are not abundant in the U.S. Further studies need to be completed in order to verify the absence of Reoviridae in U.S. raw sewage.

Viruses Detected in Final Effluent

It is important to note that the raw sewage and final effluent samples were collected at approximately the same time; therefore, the final effluents in this study do not necessarily represent the final product of the raw sewage analyzed. The wastewater treatment process for each of the wastewater treatment facilities is summarized in Table

12 to assist in analyzing the results of viral detection in final effluent. Five viral groups

(adenoviruses, noroviruses, reoviruses, pepper mild mottle viruses, and picobirnaviruses) were found in 10 ml samples of final effluent. In general, no correlation appears to exist between the viruses found in raw sewage and those found in the final effluent, regardless of wastewater treatment.

Noroviruses were detected in final effluent from Louisiana while reoviruses were found in final effluent from New Jersey after whole genome amplification. Since noroviruses and reoviruses were detected in few or no raw sewage samples, it is not

44 surprising that they were detected in a few final effluent samples. Adenoviruses were detected in the final effluent samples from Maine, Florida, and Oregon. The detection of adenoviruses in final effluent in this study supports previous studies that have demonstrated the prevalence of adenoviruses in treated sewage (15, 27, 90, 92, 94).

Picobirnaviruses were detected in a third of the final effluent samples (Connecticut,

Maine, Oregon, and Washington). It is important to note that the presence of chlorine has been shown to interfere with the detection of enteroviruses ( pers. com. Dale Griffin) and that residual chlorination in final effluent samples could have interfered with the detection of viruses in this study.

TABLE 12. Summary of wastewater treatment processes at each of the wastewater treatment facilities identified by location. FL K describes the Florida Keys .

U.S. State Primary Treatment Secondary Treatment Tertiary Treatment AL grit removal via screen aeration basin, 2 o clarifier chlorination CA sedimentation activated sludge system gravity filters, chlorination CT sedimentation activated sludge system chlorination FL none aeration basin, 2 o clarifier sand filters, chlorination FL K grit removal via screen activated sludge system fabric filtration, UV, chlorination LA none activated sludge system final clarifiers, chlorination MD sedimentation activated sludge system sand filters, chlorination ME sedimentation rotating biological contactors chlorination NC grit removal via screen activated sludge system deep bed filter, chlorination NJ sedimentation activated sludge system multimedia filter, chlorination OR sedimentation activated sludge system chlorination WA sedimentation activated sludge system chlorination

Of the three groups of viruses detected in 100% of the raw sewage samples,

PMMoV was the only virus to be found in all final effluent samples except the Florida

Keys. These viruses are known to be extremely stable and have been shown to survive

food processing and passage through the gut (98). It is reasonable to believe that PMMoV

was not identified in the Florida Keys final effluent due to the extensive treatment

process that includes fabric filtration, UV radiation, and chlorination. Since PMMoV was

45 detected in raw sewage and final effluent samples, its presence may not correlate with or predict the infectivity of the pathogens of concern in raw sewage. However, these viruses may be useful ultraconservative indicators of fecal contamination.

Wastewater Viral Concentration Efficiency

Control experiments demonstrated that the centrifugal concentration devices

effectively recovered adenoviruses that were spiked into raw sewage and final effluent.

No viruses were observed in the flow-through via epifluorescent microscopy and the

expected dilution of final DNA resulted in amplicon via PCR. The use of a centrifugal

concentration device is included in methods used by Public Health protocols (58). These

experiments suggest that centrifugal concentration devices can be used to effectively

concentrate enteric viruses from raw sewage and effluent samples.

Diversity of Picobirnaviruses in Raw Sewage

Picobirnaviruses are currently an unclassified group of viruses even though they

have been recently proposed to belong to the family Picobirnaviridae (10). As their name suggests, these viruses have a bisegmented double stranded RNA genome (23, 83).

Picobirnavirus particles are fairly small (35 nm), non-enveloped, and spherical. They have been found in the feces of a wide range of mammals, including: humans (13, 83,

95), pigs (10), rats (80), chickens (63), calves (93), giant anteaters (47), rabbits (41, 67), guinea pigs (79), water foals (20), Choiques, Chinese geese, American ostriches, pelicans, donkeys, orangutans, armadillos, and gloomy pheasants (68). They are currently unculturable and their pathogenesis is unknown (23). These viruses have been implicated

46 as possible enteric pathogens and have occasionally been associated with

(13, 23).

Several important conclusions can be drawn from the phylogenetic analysis of the

picobirnavirus sequences from raw sewage. First, it appears that a complete

understanding of the sequence diversity of picobirnaviruses remains unknown given the

large number of single sequences. This is further supported by the Chao1 and rarefaction

analyses. The great diversity of genotype I human picobirnaviruses, based on the RNA-

dependent RNA polymerase gene, have been previously reported (9, 13). Similarly,

extensive diversity has been observed in genotype I porcine picobirnaviruses and their

existence as quasispecies has been postulated (10). In order to fully grasp the sequence

diversity of the RNA-dependent RNA polymerase of genotype I human picobirnaviruses

in raw sewage, it would be necessary to sequence more clones.

It is also interesting to observe that the picobirnaviruses found in raw sewage

group with both human and porcine picobirnaviruses. The strong sequence identity

among human and porcine picobirnaviruses has been previously identified (10); however,

Banyai et al. still observed separate groupings of human and porcine picobirnaviruses.

While the majority of picobirnavirus sequences from raw sewage grouped with known

human sequences, many raw sewage sequences grouped closely with porcine

picobirnaviruses. The relatedness observed in this study among porcine and human

genotype I picobirnaviruses suggests the inability of picobirnaviruses to source track

fecal pollution based upon the sequence alone. Although picobirnaviruses may not be

suitable for identifying sources of fecal pollution, they have the potential to be effective

indicators of fecal pollution throughout the wastewater treatment process as well as in

47 coastal environments because of their widespread abundance in raw sewage and their absence from unpolluted marine environments.

Viral Isolation and Detection in Large-Scale Seawater Samples

None of the viruses that were detected in raw sewage (human adenoviruses,

noroviruses, human picobirnaviruses, and PMMoV) were found in any of the large-scale

seawater samples, regardless of their proximity to dense human populations. Although it

is possible that these viruses were present in concentrations below the detection limit of

the assays at the time of sampling, it is reasonable to believe their absence in the natural

marine environment is due to a lack of hosts. This concept is supported by the lack of

adenoviruses and other human DNA virus sequence similarities to marine metagenomic

sequences in the CAMERA (Community Cyberinfrastructure for Advanced Marine

Microbial Ecology Research and Analysis) database (84). Since none of the viruses found

in raw sewage were detected in the marine environment, the presence of any of these

viruses (adenovirus, norovirus, picobirnavirus, and PMMoV) should identify a

wastewater input.

Seawater Viral Isolation Methods Comparison

The TFF Method was developed for the concentration of marine ;

therefore, it is possible that they could have selected against the collection of the targeted

eukaryotic viruses. Enteric viruses are relatively small and should have been included in

the filtrate of the 0.2 µm TFF filter. However, they are known to become associated with

solids and could have been excluded from collection in the 0.2 µm TFF retentate (39).

Additionally, it is possible that enteric viruses could have been lost through further

48 concentration using PEG precipitation and CsCl density-dependent centrifugation. While all of these are possible mechanisms of enteric viral loss, the results of the methods comparison experiments suggest that the TFF Method was 10 to 1000-fold more effective than the commonly used Public Health Method at concentrating polioviruses and adenoviruses. These results contest the status quo among environmental public health scientists who believe that TFF concentration is less effective than other commonly used methods (44). Despite demonstrating the usefulness of the TFF Method, it is expensive in terms of cost and time and is not recommended for routine water quality monitoring.

Stability of Pepper Mild Mottle Virus and Picobirnavirus in Seawater

The stability of PMMoV and picobirnavirus from raw sewage in seawater was ~1

week and less than a week, respectively. Unfortunately, the results of this study only

approximate the stability of picobirnaviruses in seawater due to the non-specific binding

of primers, the presence of light bands in the control samples, the inability to sequence

key PCR products, and the intermittent disappearance of positive PCR products. Before

picobirnaviruses can be considered as a marker of fecal pollution, it is essential that the

specificity of the picobirnavirus primers be increased to reduce the amount of non-

specific binding observed in environmental samples. Similarly, intermittent PCR

amplification made estimating the stability of PMMoV difficult. The

appearance/disappearance of amplicon is likely the result of the patchiness of the seeded

viruses.

The stability of enteric adenoviruses has been previously compared to type 1

polioviruses and the hepatitis A viruses (32). Enriquez et al. found that enteric

adenoviruses were considerably more stable than poliovirus type 1 and hepatitis A

49 viruses. Specifically, the T 99 (the time it takes for 99% inactivation of the original titer),

for adenovirus-40 and -41 is 77 days and 85 days in seawater, respectively (32). This is

over four times longer than the 18 day T99 of poliovirus type 1 in seawater (32). Since the

methods employed by Enriquez et al. are much more robust, it is difficult to compare the

stability of enteric adenoviruses in seawater to the estimated stability of picobirnaviruses

and PMMoV as outlined in this research. However, it appears that picobirnaviruses and

PMMoV exhibit stability more similar to poliovirus type 1 in seawater.

Limitations

It is important to recognize a few shortcomings of this research. First, it is possible

that any undetected or infrequently detected viruses were present in concentrations at or

below the detection limit of their assays. It is also possible that these under-identified

viruses were not prevalent on the day of sampling. Furthermore, it is important to

emphasize that the results are based upon a relatively small volume of raw sewage, final

effluent, or seawater. Additionally, the detection of any viral group in raw sewage or final

effluent does not imply infectivity.

This study is the first of its kind to illustrate diversity of picobirnaviruses in raw

sewage. It is important to recognize that the phylogenetic analyses in this study were

dependent upon a short region of one gene. In order to obtain a clear picture of

picobirnavirus phylogenetics, it would be necessary to analyze both the RNA-dependent

RNA polymerase as well as the protein. It is also important to remember that the

primers utilized only amplify genotype I picobirnaviruses (13, 83). Furthermore,

additional sequences need to be collected from human and animal picobirnaviruses to

better understand the diversity of this viral group.

50 Potential of PMMoV and Picobirnaviruses as Water Quality Indicators

In order to protect public health, it is necessary to identify a practical method for assessing fecal pollution in recreational waters. This is often accomplished through the use of indicators, which are used to approximate the presence of pathogens. The current bacterial indicators of fecal contamination are not good indicators of wastewater pollution or human health risk (19, 36, 45, 46, 62, 96). The results of this study suggest that picobirnaviruses and PMMoV may be utilized as indicators of fecal pollution since they were detected in all raw sewage samples collected in the United States. It is also clear that these viruses do not replicate in the marine environment since they were not detected in any of the seawater samples. Before picobirnaviruses and PMMoV can be used as indicators to monitor water quality, it is important to determine several correlations to prove their utility as indicators of fecal pollution.

The characteristics of an ideal indicator are defined by the US EPA and include the following correlations (33). First, it would be necessary to understand if the decay rate of the proposed indicator correlates with pathogens of concern throughout the wastewater treatment process and in the marine environment. Pathogens of concern in raw sewage include viruses, bacteria, and protists. Future studies will need to determine if the presence of picobirnaviruses and PMMoV correlates with non-viral pathogens of interest throughout the wastewater treatment process as well as in coastal environments.

The results of this study indicate that picobirnaviruses may correlate with pathogens of concern since they were detected in a third of final effluent samples and appear to remain stable in the marine environment for less than a week. Given the presence of PMMoV in 91.7% of final effluent samples and its estimated stability of one

51 week in seawater, it is likely that this virus may overestimate the presence of sewage associated pathogens. It is possible that the utility of PMMoV is as an ultraconservative indicator of fecal pollution since it was found in most of the final effluent samples. While an ultraconservative indicator of fecal pollution could misjudge health risks, an ultraconservative indicator could indicate the possible presence of risk associated contaminants (i.e. pharmaceuticals, nutrients, etc) and thus has the potential to be a useful tool for environmental investigations.

Another important correlation to determine is if the proposed viral indicator concentrations correlate with the amount of fecal pollution in the marine environment.

Although it is difficult to measure the true extent of fecal pollution in the environment given the lack of correlation among the current indicators and pathogens of interest, thorough analysis of a variety of contaminated environmental sites will facilitate addressing this issue. This study has demonstrated the absence of PMMoV and picobirnaviruses from five marine sites that ranged in their proximity to human populations.

It is interesting to note that PMMoV, but not picobirnaviruses, was detected using methods described for wastewater in one 20 ml Hillsborough River sample collected in knee deep water adjacent to a sewage outfall pipe. This site was characterized to have poor water quality based upon current bacterial indicators (sample & bacterial indicator data courtesy of the Harwood Laboratory, University of South Florida). Furthermore,

PMMoV was detected in four out of seven 50 ml ground water samples that contained enteroviruses (samples & data courtesy of Dale Griffin, United States

Geological Survey). These ground water samples were collected in rural areas, which rely

52 on on-site disposal systems, outside of Tallahassee, Florida. The detection of PMMoV in small volumes of water exposed to wastewater sources supports its potential use as an indicator.

The most useful indicator of fecal pollution would also be able to distinguish human from animal sewage contamination. Since the majority of human fecal pollution comes from non-point sources, it is often difficult to identify the source of pollution. If an indicator is able to differentiate between human and animal fecal contamination, then the ability to identify the source of pollution increases. As a result, it is important to assess the potential of PMMoV and picobirnaviruses at source tracking tools.

The strong sequence identity between picobirnaviruses from human sewage and

GenBank reference sequences from pigs in this research suggests the inability of picobirnaviruses to source track fecal pollution using sequencing alone. The lack of source tracking capability of picobirnaviruses using only the sequence is further supported by their prevalence in a wide range of mammals. Alternatively due to its dietary origin, PMMoV is not expected in animal feces and to date has yet to be detected in animal feces ( pers. com. Karyna Rosario). The following collective fecal samples have

tested negative for PMMoV: laughing gulls (n = 8) from the Suncoast Seabird Sanctuary

in Pinellas County, Florida, bovine feces from a dairy farm in Puerto Rico, equine feces

from a horse racetrack in Puerto Rico (samples courtesy of the Rosario family), a rooster

(n = 1), Grey Horned Owl (n = 1), Barred Owl (n = 1), Bald Eagle (n = 1), and Turkey

Vultures (n = 2) from the Boyd Hill Nature Park (samples courtesy of Gabe Vargo) in St.

Petersburg, Florida.

53

CONCLUSIONS

Adenoviruses, picobirnaviruses, and PMMoV were the only viruses detected in

100% of raw sewage samples collected from 11 U.S. coastal states. Some of these viruses were also detected in a range of final effluent samples from the same wastewater treatment facilities but none of these viruses were detected in the marine environment in the absence of identifiable pollution. While adenoviruses are known human pathogens and have been proposed as markers of fecal pollution in previous studies, the results of this research demonstrate the potential use of picobirnaviruses and PMMoV as indicators of fecal pollution. Further research will be needed to determine if these candidate viruses have the necessary characteristics of a microbial water quality indicator.

54

REFERENCES

1. Albert, M. J. 1986. Enteric adenoviruses - brief review. Archives of Virology 88: 1-17.

2. Alexander, L. M., A. Heaven, A. Tennant, and R. Morris. 1992. Symptomatology of children in contact with sea-water contaminated with sewage. Journal of Epidemiology and Community Health 46: 340-344.

3. Allard, A., B. Albinsson, and G. Wadell. 1992. Detection of adenoviruses in stools from healthy-persons and patients with diarrhea by 2-step polymerase chain-reaction. Journal of Medical Virology 37: 149-157.

4. Altschul, S. F., W. Gish, W. Miller, E. W. Myers, and D. J. Lipman. 1990. Basic local alignment search tool. Journal of Molecular Biology 215: 403-410.

5. Altschul, S. F., T. L. Madden, A. A. Schaffer, J. H. Zhang, Z. Zhang, W. Miller, and D. J. Lipman. 1997. Gapped BLAST and PSI-BLAST: a new generation of protein database search programs. Nucleic Acids Research 25: 3389- 3402.

6. Apajalahti, J. H. A., A. Kettunen, P. H. Nurminen, H. Jatila, and W. E. Holben. 2003. Selective plating underestimates abundance and shows differential recovery of bifidobacterial species from human feces. Applied and Environmental Microbiology 69: 5731-5735.

7. Aulicino, F. A., L. Mauro, M. Marranzano, M. Biondi, A. Ursino, and M. Carere. 2000. Microbiological quality of the Catania coastal sea water. Annali d'igiene 12: 533-41.

8. Aulicino, F. A., P. Orsini, M. Carere, and A. Mastrantonio. 2001. Bacteriological and virological quality of seawater bathing areas along the Tyrrhenian coast. International Journal of Environmental Health Research 11: 5- 11.

9. Banyai, K., F. Jakab, G. Reuter, J. Bene, M. Uj, B. Melegh, and G. Szucs. 2003. Sequence heterogeneity among human picobirnaviruses detected in a gastroenteritis outbreak. Archives of Virology 148: 2281-91.

55 10. Banyai, K., V. Martella, A. Bogdan, P. Forgach, F. Jakab, E. Meleg, H. Biro, B. Melegh, and G. Szucs. 2008. Genogroup I picobirnaviruses in pigs: evidence for genetic diversity and relatedness to human strains. Journal of General Virology 89: 534-539.

11. Barrett, T., I. K. G. Visser, L. Mamaev, L. Goatley, M. F. Vanbressem, and A. Osterhaus. 1993. Dolphin and porpoise morbilliviruses are genetically distinct from phocine distemper Virus. Virology 193: 1010-1012.

12. Benson, K. A. S., C. A. Manire, R. Y. Ewing, J. T. Saliki, F. I. Townsend, B. Ehlers, and C. H. Romero. 2006. Identification of novel alpha- and gammaherpesviruses from cutaneous and mucosal lesions of dolphins and whales. Journal of Virological Methods 136: 261-266.

13. Bhattacharya, R., G. C. Sahoo, M. K. Nayak, K. Rajendran, P. Dutta, U. Mitra, M. K. Bhattacharya, T. N. Naik, S. K. Bhattacharya, and T. Krishnan. 2007. Detection of genogroup I and II human picobirnaviruses showing small genomic RNA profile causing acute watery diarrhoea among children in Kolkata, India. Infection Genetics and Evolution 7: 229-238.

14. Binga, E. K., R. S. Lasken, and J. D. Neufeld. 2008. Something from (almost) nothing: the impact of multiple displacement amplification on microbial ecology. Isme Journal 2: 233-241.

15. Bofill-Mas, S., N. Albinana-Gimenez, P. Clemente-Casares, A. Hundesa, J. Rodriguez-Manzano, A. Allard, M. Calvo, and R. Girones. 2006. Quantification and stability of human adenoviruses and polyomavirus JCPyV in wastewater matrices. Applied and Environmental Microbiology 72: 7894-6.

16. Brassard, J., K. Seyer, A. Houde, C. Simard, and Y. L. Trottier. 2005. Concentration and detection of hepatitis A virus and rotavirus in spring water samples by reverse transcription-PCR. Journal of Virological Methods 123: 163-9.

17. Breitbart, M., I. Hewson, B. Felts, J. M. Mahaffy, J. Nulton, P. Salamon, and F. Rohwer. 2003. Metagenomic analyses of an uncultured viral community from human feces. Journal of Bacteriology 185: 6220-6223.

18. Breitbart, M., P. Salamon, B. Andresen, J. M. Mahaffy, A. M. Segall, D. Mead, F. Azam, and F. Rohwer. 2002. Genomic analysis of uncultured marine viral communities. Proceedings of the National Academy of Sciences of the United States of America 99: 14250-14255.

19. Brownell, M. J., V. J. Harwood, R. C. Kurz, S. M. McQuaig, J. Lukasik, and T. M. Scott. 2007. Confirmation of putative stormwater impact on water quality at a Florida beach by microbial source tracking methods and structure of indicator organism populations. Water Research 41: 3747-3757.

56

20. Browning, G. F., R. M. Chalmers, D. R. Snodgrass, R. M. Batt, C. A. Hart, S. E. Ormarod, D. Leadon, S. J. Stoneham, and P. D. Rossdale. 1991. The prevalence of enteric pathogens in diarrhoeic thoroughbred foals in Britain and Ireland. Equine Veterinary Journal 23: 405-9.

21. Caballero, S., F. X. Abad, F. Loisy, F. S. Le Guyader, J. Cohen, R. M. Pinto, and A. Bosch. 2004. Rotavirus virus-like particles as surrogates in environmental persistence and inactivation studies. Applied and Environmental Microbiology 70: 3904-9.

22. Casas, V., and F. Rohwer. 2007. Phage metagenomics. Methods in Enzymology 421: 259-68.

23. Chandra, R. 1997. Picobirnavirus, a novel group of undescribed viruses of mammals and birds: a minireview. Acta Virologica 41: 59-62.

24. Chao, A. 1987. Estimating the population size for capture-recapture data and unequal catchability. Biometrics 43: 783 - 791.

25. Chao, A. 1984. Non-parametric estimation of the number of classes in a population. Scand J Stat 11: 783 - 791.

26. Cheung, W. H. S., K. C. K. Chang, R. P. S. Hung, and J. W. L. Kleevens. 1990. Health-effects of beach water-pollution in Hong-Kong. Epidemiology and Infection 105: 139-162.

27. Crabtree, K. D., C. P. Gerba, J. B. Rose, and C. N. Haas. 1997. Waterborne adenovirus: A risk assessment. Water Science and Technology 35: 1-6.

28. Donaldson, K. A., D. W. Griffin, and J. H. Paul. 2002. Detection, quantitation and identification of enteroviruses from surface waters and sponge tissue from the Florida Keys using real-time RT-PCR. Water Research 36: 2505-2514.

29. Dowell, S. F., C. Groves, K. B. Kirkland, H. G. Cicirello, T. Ando, Q. Jin, J. R. Gentsch, S. S. Monroe, C. D. Humphrey, C. Slemp, D. M. Dwyer, R. A. Meriwether, and R. I. Glass. 1995. A multistate outbreak of oyster-associated gastroenteritis - implications for interstate tracing of contaminated shellfish. Journal of Infectious Diseases 171: 1497-1503.

30. Dubois, E., F. LeGuyader, L. Haugarreau, H. Kopecka, M. Cormier, and M. Pommepuy. 1997. Molecular epidemiological survey of rotaviruses in sewage by reverse transcriptase seminested PCR and restriction fragment length polymorphism assay. Applied and Environmental Microbiology 63: 1794-1800.

57 31. Elmir, S. M., M. E. Wright, A. Abdelzaher, H. M. Solo-Gabriele, L. E. Fleming, G. Miller, M. Rybolowik, M. T. P. Shih, S. P. Pillai, J. A. Cooper, and E. A. Quaye. 2007. Quantitative evaluation of bacteria released by bathers in a marine water. Water Research 41: 3-10.

32. Enriquez, C. E., C. J. Hurst, and C. P. Gerba. 1995. Survival of the enteric adenovirus-40 and adenovirus-41 in tap, sea, and waste-water. Water Research 29: 2548-2553.

33. Environmental Protection Agency. 2006. Implementing the BEACH Act of 2000 Report to Congress . EPA-823-R-06-001.

34. Environmental Protection Agency. 1999. Water quality concerns in the Florida Keys: sources, effects, and solutions. EPA 904-R-99-05.

35. Fauquet, C. M., M. A. Mayo, M. A. Maniloff, J. Maniloff, J. Desselberger, and L. A. Ball (ed.). 2005. Virus Taxonomy 8th ed. Elsevier Academic Press, Amsterdam.

36. Ferguson, D. M., D. F. Moore, M. A. Getrich, and M. H. Zhowandai. 2005. Enumeration and speciation of enterococci found in marine and intertidal sediments and coastal water in southern California. Journal of Applied Microbiology 99: 598-608.

37. Fleisher, J. M., D. Kay, R. L. Salmon, F. Jones, M. D. Wyer, and A. F. Godfree. 1996. Marine waters contaminated with domestic sewage: Nonenteric illnesses associated with bather exposure in the United Kingdom. American Journal of Public Health 86: 1228-1234.

38. Fong, T. T., D. W. Griffin, and E. K. Lipp. 2005. Molecular assays for targeting human and bovine enteric viruses in coastal waters and their application for library-independent source tracking. Applied and Environmental Microbiology 71: 2070-8.

39. Fong, T. T., and E. K. Lipp. 2005. Enteric viruses of humans and animals in aquatic environments: Health risks, detection, and potential water quality assessment tools. Microbiology and Molecular Biology Reviews 69: 357-371.

40. Forslund, O., A. Antonsson, P. Nordin, B. Stenquist, and B. G. Hansson. 1999. A broad range of human papillomavirus types detected with a general PCR method suitable for analysis of cutaneous tumours and normal skin. Journal of General Virology 80: 2437-2443.

41. Gallimore, C., D. Lewis, and D. Brown. 1993. Detection and characterization of a novel bisegmented double-stranded RNA virus (picobirnavirus) from rabbit faeces. Archives of Virology 133: 63-73.

58 42. Gerba, C. P., D. M. Gramos, and N. Nwachuku. 2002. Comparative inactivation of enteroviruses and adenovirus 2 by UV light. Applied and Environmental Microbiology 68: 5167-9.

43. Gilgen, M., D. Germann, J. Luthy, and P. Hubner. 1997. Three-step isolation method for sensitive detection of enterovirus, rotavirus, hepatitis A virus, and small round structured viruses in water samples. International Journal of Food Microbiology 37: 189-199.

44. Griffin, D. W., K. A. Donaldson, J. H. Paul, and J. B. Rose. 2003. Pathogenic human viruses in coastal waters. Clinical Microbiology Reviews 16: 129-143.

45. Griffin, D. W., C. J. Gibson, E. K. Lipp, K. Riley, J. H. Paul, and J. B. Rose. 1999. Detection of viral pathogens by reverse transcriptase PCR and of microbial indicators by standard methods in the canals of the Florida Keys. Applied and Environmental Microbiology 65: 4118-4125.

46. Griffin, D. W., E. K. Lipp, M. R. McLaughlin, and J. B. Rose. 2001. Marine recreation and public health microbiology: Quest for the ideal indicator. Bioscience 51: 817-825.

47. Haga, I. R., S. S. Martins, S. T. Hosomi, F. Vicentini, H. Tanaka, and M. S. Gatti. 1999. Identification of a bisegmented double-stranded RNA virus (Picobirnavirus) in faeces of giant anteaters (Myrmecophaga tridactyla). Veterinary Journal 158: 234-6.

48. Harvell, C. D., K. Kim, J. M. Burkholder, R. R. Colwell, P. R. Epstein, D. J. Grimes, E. E. Hofmann, E. K. Lipp, A. Osterhaus, R. M. Overstreet, J. W. Porter, G. W. Smith, and G. R. Vasta. 1999. Review: Marine ecology - Emerging marine diseases - Climate links and anthropogenic factors. Science 285: 1505-1510.

49. Heck, K. L., G. Van Belle, and D. Simberloff. 1975. Explicit calculation of the rarefaction diversity measurement and the determination of sufficient sample size. Ecology 56: 1459 - 1461.

50. Hejkal, T. W., F. M. Wellings, A. L. Lewis, and P. A. Larock. 1981. Distribution of viruses associated with particles in waste-water. Applied and Environmental Microbiology 41: 628-634.

51. Henrickson, S. E., T. Wong, P. Allen, T. Ford, and P. R. Epstein. 2001. Marine swimming-related illness: Implications for monitoring and environmental policy. Environmental Health Perspectives 109: 645-650.

59 52. Howarth, R. W. 1993. Managing wastewater in coastal urban areas: the role of nutrients in coastal waters. National Academy Press, Washington, D.C.

53. Hurlbert, S. H. 1997. The nonconcept of species diversity: a critique and alternative parameters. Ecology 52: 577 - 586.

54. Irving, L. G., and F. A. Smith. 1981. One-year survey of enteroviruses, adenoviruses, and reoviruses isolated from effluent at an activated-sludge purification plant. Applied and Environmental Microbiology 41: 51-59.

55. Jiang, S., R. Noble, and W. Chu. 2001. Human adenoviruses and coliphages in urban runoff-impacted coastal waters of Southern California. Applied and Environmental Microbiology 67: 179-84.

56. Jiang, S. C. 2006. Human adenoviruses in water: occurrence and health implications: a critical review. Environmental Science & Technology 40: 7132- 7140.

57. Jiang, X., P. W. Huang, W. M. Zhong, T. Farkas, D. W. Cubitt, and D. O. Matson. 1999. Design and evaluation of a primer pair that detects both norwalk- and sapporo-like caliciviruses by RT-PCR. Journal of Virological Methods 83: 145-154.

58. Katayama, H., A. Shimasaki, and S. Ohgaki. 2002. Development of a virus concentration method and its application to detection of enterovirus and norwalk virus from coastal seawater. Applied and Environmental Microbiology 68: 1033- 1039.

59. Kittigul, L., S. Ekchaloemkiet, F. Utrarachkij, K. Siripanichgon, D. Sujirarat, S. Pungchitton, and A. Boonthum. 2005. An efficient virus concentration method and RT-nested PCR for detection of rotaviruses in environmental water samples. Journal of Virological Methods 124: 117-22.

60. Koike, K., M. Kobayashi, M. Gondo, I. Hayashi, T. Osuga, and S. Takada. 1998. Hepatitis B virus DNA is frequently found in liver biopsy samples from hepatitis C virus-infected chronic hepatitis patients. Journal of Medical Virology 54: 249-255.

61. Leary, T. P., J. C. Erker, M. L. Chalmers, A. T. Cruz, J. D. Wetzel, S. M. Desai, I. K. Mushahwar, and T. S. Dermody. 2002. Detection of mammalian reovirus RNA by using reverse transcription-PCR: Sequence diversity within the lambda 3-encoding L1 gene. Journal of Clinical Microbiology 40: 1368-1375.

62. Lee, C. M., T. Y. Lin, C. C. Lin, G. A. Kohbodi, A. Bhattl, R. Lee, and J. A. Jay. 2006. Persistence of fecal indicator bacteria in Santa Monica Bay beach sediments. Water Research 40: 2593-2602.

60 63. Leite, J. P. G., S. P. Monteiro, A. M. Fialho, and H. G. Pereira. 1990. A novel avian virus with tri-segmented double-stranded RNA and further observations of previously described similar viruses with bi-segmented genome. Virus Research 16: 119-126.

64. Lipp, E. K., J. Carrie Futch, and D. W. Griffin. 2007. Analysis of multiple enteric viral targets as sewage markers in coral reefs. Marine Pollution Bulletin 54: 1897-1902.

65. Lipp, E. K., S. A. Farrah, and J. B. Rose. 2001. Assessment and impact of microbial fecal pollution and human enteric pathogens in a coastal community. Marine Pollution Bulletin 42: 286-293.

66. Lodder, W. J., and A. M. D. Husman. 2005. Presence of noroviruses and other enteric viruses in sewage and surface waters in The Netherlands. Applied and Environmental Microbiology 71: 1453-1461.

67. Ludert, J. E., L. Abdul-Latiff, A. Liprandi, and F. Liprandi. 1995. Identification of picobirnavirus, viruses with bisegmented double stranded RNA, in rabbit faeces. Research in Veterinary Science 59: 222-5.

68. Masachessi, G., L. C. Martinez, M. O. Giordano, P. A. Barril, B. M. Isa, L. Ferreyra, D. Villareal, M. Carello, C. Asis, and S. V. Nates. 2007. Picobirnavirus (PBV) natural hosts in captivity and virus excretion pattern in infected animals. Archives of Virology 152: 989-998.

69. Metcalf, T. G., J. L. Melnick, and M. K. Estes. 1995. Environmental virology - from detection of virus in sewage and water by isolation to identification by molecular-biology - a trip of over 50 Years. Annual Review of Microbiology 49: 461-487.

70. Monroe County Health Department. 2000. Aerobic treatment unit – annual report.

71. Munn, C. B. 2006. Viruses as pathogens of marine organisms - from bacteria to whales. Journal of the Marine Biological Association of the United Kingdom 86: 453-467.

72. Muscillo, M., G. La Rosa, C. Marianelli, S. Zaniratti, M. R. Capobianchi, L. Cantiani, and A. Carducci. 2001. A new RT-PCR method for the identification of reoviruses in seawater samples. Water Research 35: 548-56.

73. Nathanson, N. 1998. Emergence of new viral infections: Implications for the blood supply. Biologicals 26: 77-84.

61 74. Noble, R. T., and J. A. Fuhrman. 2000. Rapid virus production and removal as measured with fluorescently labeled viruses as tracers. Applied and Environmental Microbiology 66: 3790-3797.

75. Noble, R. T., and J. A. Fuhrman. 1998. Use of SYBR Green I for rapid epifluorescence counts of marine viruses and bacteria. Aquatic Microbial Ecology 14: 113-118.

76. Paul, J. H., M. R. McLaughlin, D. W. Griffin, E. K. Lipp, R. Stokes, and J. B. Rose. 2000. Rapid movement of wastewater from on-site disposal systems into surface waters in the Lower Florida Keys. Estuaries 23: 662-668.

77. Paul, J. H., J. B. Rose, J. Brown, E. A. Shinn, S. Miller, and S. R. Farrah. 1995. Viral tracer studies indicate contamination of marine waters by sewage disposal practices in Key Largo, Florida. Applied and Environmental Microbiology 61: 2230-2234.

78. Paul, J. H., J. B. Rose, S. C. Jiang, X. T. Zhou, P. Cochran, C. Kellogg, J. B. Kang, D. Griffin, S. Farrah, and J. Lukasik. 1997. Evidence for groundwater and surface marine water contamination by waste disposal wells in the Florida Keys. Water Research 31: 1448-1454.

79. Pereira, H. G., H. P. De Araujo, A. M. Fialho, L. De Castro, and P. Monteiro. 1989. A virus with bi-segmented double-stranded RNA genome in guinea pigs intestines. Memórias do Instituto Oswaldo Cruz 83: 137-140.

80. Pereira, H. G., T. H. Flewett, J. A. Candeias, and O. M. Barth. 1988. A virus with a bisegmented double-stranded RNA genome in rat (Oryzomys nigripes) intestines. Journal of General Virology 69 ( Pt 11): 2749-54.

81. Pianetti, A., W. Baffone, B. Citterio, A. Casaroli, F. Bruscolini, and L. Salvaggio. 2000. Presence of enteroviruses and reoviruses in the waters of the Italian coast of the Adriatic Sea. Epidemiology and Infection 125: 455-62.

82. Rigotto, C., T. C. M. Sincero, C. M. O. Simoes, and C. R. M. Barardi. 2005. Detection of adenoviruses in shellfish by means of conventional-PCR, nested- PCR, and integrated cell culture PCR (ICC/PCR). Water Research 39: 297-304.

83. Rosen, B. I., Z. Y. Fang, R. I. Glass, and S. S. Monroe. 2000. Cloning of human picobirnavirus genomic segments and development of an RT-PCR detection assay. Virology 277:316-329.

62 84. Rusch, D. B., A. L. Halpern, G. Sutton, K. B. Heidelberg, S. Williamson, S. Yooseph, D. Y. Wu, J. A. Eisen, J. M. Hoffman, K. Remington, K. Beeson, B. Tran, H. Smith, H. Baden-Tillson, C. Stewart, J. Thorpe, J. Freeman, C. Andrews-Pfannkoch, J. E. Venter, K. Li, S. Kravitz, J. F. Heidelberg, T. Utterback, Y. H. Rogers, L. I. Falcon, V. Souza, G. Bonilla-Rosso, L. E. Eguiarte, D. M. Karl, S. Sathyendranath, T. Platt, E. Bermingham, V. Gallardo, G. Tamayo-Castillo, M. R. Ferrari, R. L. Strausberg, K. Nealson, R. Friedman, M. Frazier, and J. C. Venter. 2007. The Sorcerer II global ccean sampling expedition: northwest Atlantic through eastern tropical Pacific. PLoS Biology 5: 398-431.

85. Saliki, J. T., E. J. Cooper, and J. P. Gustavson. 2002. Emerging infections of marine mammals - Development of two diagnostic approaches, p. 51-59, Domestic Animal/Wildlife Interface: Issue for Disease Control, Conservation, Sustainable Food Production, and Emerging Diseases, vol. 969.

86. Sambrook, J., and D. W. Russell. 2001. Molecular Cloning A Laboratory Manual, 3rd ed. Cold Spring Harbor Laboratory Press, Cold Spring Harbor, NY.

87. Simmons, G., G. Greening, W. Z. Gao, and D. Campbell. 2001. Raw oyster consumption and outbreaks of viral gastroenteritis in New Zealand: evidence for risk to the public's health. Australian and New Zealand Journal of Public Health 25: 234-240.

88. Tamura, K., J. Dudley, M. Nei, and S. Kumar. 2007. MEGA4: Molecular evolutionary genetics analysis (MEGA) software version 4.0. Molecular Biology and Evolution 24: 1596-9.

89. Thompson, J. D., T. J. Gibson, F. Plewniak, F. Jeanmougin, and D. G. Higgins. 1997. The CLUSTAL_X windows interface: flexible strategies for multiple sequence alignment aided by quality analysis tools. Nucleic Acids Research 25: 4876-4882.

90. Thompson, S. S., J. L. Jackson, M. Suva-Castillo, W. A. Yanko, Z. El Jack, J. Kuo, C. L. Chen, F. P. Williams, and D. P. Schnurr. 2003. Detection of infectious human adenoviruses in tertiary-treated and ultraviolet-disinfected wastewater. Water Environment Research 75: 163-170.

91. van den Berg, H., W. Lodder, W. van der Poel, H. Vennema, and A. M. D. Husman. 2005. Genetic diversity of noroviruses in raw and treated sewage water. Research in Microbiology 156: 532-540.

92. Van Heerden, J., M. M. Ehlers, W. B. Van Zyl, and W. O. K. Grabow. 2003. Incidence of adenoviruses in raw and treated water. Water Research 37: 3704- 3708.

63 93. Vanopdenbosch, E., and G. Wellemans. 1989. Birna-type virus in diarrheic calf faeces. The Veterinary Record 125: 610.

94. Vantarakis, A., and M. Papapetropoulou. 1999. Detection of enteroviruses, adenoviruses and hepatitis A viruses in raw sewage and treated effluents by nested-PCR. Water Air and Soil Pollution 114: 85-93.

95. Wakuda, M., Y. Pongsuwanna, and K. Taniguchi. 2005. Complete nucleotide sequences of two RNA segments of human picobirnavirus. Journal of Virological Methods 126: 165-169.

96. Wetz, J. J., E. K. Lipp, D. W. Griffin, J. Lukasik, D. Wait, M. D. Sobsey, T. M. Scott, and J. B. Rose. 2004. Presence, infectivity, and stability of enteric viruses in seawater: relationship to marine water quality in the Florida Keys. Marine Pollution Bulletin 48: 698-704.

97. Yu, Y., M. Breitbart, P. McNairnie, and F. Rohwer. 2006. FastGroupII: a web- based bioinformatics platform for analyses of large 16S rDNA libraries. BMC Bioinformatics 7: 57.

98. Zhang, T., M. Breitbart, W. H. Lee, J. Q. Run, C. L. Wei, S. W. L. Soh, M. L. Hibberd, E. T. Liu, F. Rohwer, and Y. J. Ruan. 2006. RNA viral community in human feces: Prevalence of plant pathogenic viruses. PLoS Biology 4: 108-118.

64