new approaches and technologies for quantifying the impact

179
New Approaches and Technologies for Quantifying Fecal Contamination in Tidal Creek and Coastal Receiving Waters Curtis H. Stumpf A dissertation submitted to the faculty of the University of North Carolina at Chapel Hill in partial fulfillment of the requirements for the Doctorate of Philosophy in the Department of Marine Sciences Chapel Hill 2011 Approved by: Rachel T. Noble, Ph.D. Christopher S. Martens, Ph.D. Michael F. Piehler, Ph.D. Taro P. Smith, Ph.D. Jill R. Stewart, Ph.D.

Upload: others

Post on 17-Apr-2022

2 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: New approaches and technologies for quantifying the impact

New Approaches and Technologies for Quantifying Fecal Contamination in Tidal

Creek and Coastal Receiving Waters

Curtis H. Stumpf

A dissertation submitted to the faculty of the University of North Carolina at Chapel Hill in

partial fulfillment of the requirements for the Doctorate of Philosophy in the Department of Marine Sciences

Chapel Hill

2011

Approved by:

Rachel T. Noble, Ph.D.

Christopher S. Martens, Ph.D.

Michael F. Piehler, Ph.D.

Taro P. Smith, Ph.D.

Jill R. Stewart, Ph.D.

Page 2: New approaches and technologies for quantifying the impact

ii

©2011

Curtis H. Stumpf

ALL RIGHTS RESERVED

Page 3: New approaches and technologies for quantifying the impact

iii

ABSTRACT

Curtis H. Stumpf: New Approaches and Technologies for Quantifying Fecal

Contamination in Tidal Creek and Coastal Receiving Waters

(Under the direction of Rachel T. Noble)

The coastal waters of North Carolina (NC) are important recreational areas for

bathing, fishing, and shellfish harvesting. However, much of these coastal waters are

contaminated with fecal indicator bacteria (FIB) (Escherichia coli (EC) and Enterococcus

spp. (ENT)) which are used as proxies for pathogens of fecal origin. This fecal contamination

largely occurs from improperly functioning septic systems, illicit discharges from wastewater

treatment plants, and overland transport to water bodies during rainfall events.

Little is known about loading concentrations and characteristics (transport) of FIB in

tidal creeks of the New River Estuary, NC. FIB frequently exceeded regulatory standards for

fecal contamination in headwater portions of tidal creeks. Total loads of 109 -1012 EC and

ENT cells occurred over the course of storm events, and storm loading was as much as 30 to

37 times greater than baseflow loading for EC and ENT, respectively. Within the first 30% of

creek storm volume for all storms and all creeks combined, a mean cumulative load of only

37% and 44% of the total EC and ENT cells, respectively, were discharged, indicating these

creeks are not demonstrating first flush loading characteristics.

Understanding source of contamination (human vs. wildlife) of these fecal indicators

is useful for understanding risk from exposure to contaminated waters and developing

mitigation strategies to improve water quality. A “toolbox” approach including conventional

Page 4: New approaches and technologies for quantifying the impact

iv

indicators, three alternative fecal indicator Bacteroides assays, and optical brighteners was

utilized to make more robust source estimates of fecal contamination in four tidal creeks.

Indicators of human contamination were found in all creeks despite relatively undeveloped

land surfaces. One creek (Southwest) in particular had suggestive human fecal

contamination, as determined by agreement among the entire suite of indicators. Alternative

indicator testing methods using Bacteroides spp. DNA assays and Quantitative Polymerase

Chain Reaction (QPCR) proved useful at distinguishing human from animal fecal

contamination. However, QPCR for monitoring fecal indicators in complex tidal creek water

matrices were less practical due to confounding factors.

The future of water quality monitoring will likely involve the use of rapid biosensors.

However, current biosensor technologies are often limited by sensitivity and specificity of

analysis. Methods were developed to concentrate, purify, and capture EC, ENT, and

Bacteroides spp. from marine and freshwater samples. Methods were optimized to

concentrate 5 L samples by factors of 221 – 317, with average recovery from concentration

of 58.6 ± 20.0%, 48.3 ± 25.5 %, and 27.5 ± 19.1 % for EC, ENT, and B. thetaiotaomicron,

respectively. An immunomagnetic separation (IMS) technique was also optimized to

selectively recover EC and ENT from samples using antibody labeled magnetic beads. IMS

resulted in moderate recoveries of nearly 40% of EC and ENT from marine water samples,

enabling detection in biosensor platforms.

Overall, this research provides approaches for calculating and characterizing FIB

loading in headwater tidal creeks, using a multiple indicator approach for robustly

determining sources of fecal contamination, and optimized methods for sample processing

for downstream application of emerging biosensor technology. Results indicate the need for

Page 5: New approaches and technologies for quantifying the impact

v

improved assessment of fecal contamination in order to inform mitigation strategies for

reduction of fecal contamination in tidal creeks and marine waters.

Page 6: New approaches and technologies for quantifying the impact

vi

ACKNOWLEDGEMENTS

This research was conducted under the Defense Coastal/Estuarine Research Program

(DCERP), funded by the Strategic Environmental Research and Development Program

(SERDP). Views, opinions, and/or findings contained in this thesis are those of the author

and should not be construed as an official U.S. Department of Defense position or decision

unless so designated by other official documentation. Additional financial support was

provided by Crystal Diagnostics and enumeration materials support from IDEXX. Thanks to

Karen Bates at the Morehead City waste water treatment plant for assistance with influent

collection, Tony and Patricia Rodriguez for assistance with particle size analysis and use of

laboratory equipment, and Kathy Conn and Reagan Converse for technical review. Special

thanks to the Pearl and Piehler laboratories, especially Susan Thompson and Rebecca

Schwartz for assistance in sample collection at Marine Corps Base Camp Lejeune (MCBCL),

and Susan Cohen, MCBCL DCERP Coordinator. In addition, this research would not have

been possible with the help of individuals (past and present) in the Noble Lab at the

University of North Carolina Institute of Marine Sciences, including Denene Blackwood,

Raul Gonzalez, Angie Coulliette, Steve Fries, Jenn Wetz, Mussie Habtestelassie, and Monica

Greene. Special thanks to Crystal Diagnostics for the funding and direction in accomplishing

much of this sample concentration and immunomagnetic separation work, including staff

members Gary Niehaus, Weidong Zhao, Christine Ammons, Jeff Hill, Dan Minardi, and Bob

Bunting. Finally, a great deal of appreciation to my wife Amy Stumpf, and our two children,

Page 7: New approaches and technologies for quantifying the impact

vii

Lila and Max, my parents Rose and Joe, my in- laws, Ralph and Merle Heckroth, and my dog,

Molly B.

Page 8: New approaches and technologies for quantifying the impact

viii

DEDICATION

For my children Max and Lila, my wife Amy, and my parents Rose and Joe, for which

this would never have been possible without your love, support, and guidance.

Page 9: New approaches and technologies for quantifying the impact

ix

TABLE OF CONTENTS

List of Tables ............................................................................................................................xi

List of Figures ......................................................................................................................... xii List of Abbreviations ..............................................................................................................xvi

Chapter 1: Introduction ..............................................................................................................1 Background ................................................................................................................................2

Stormwater Microbial Loading of Tidal Creeks ........................................................................3

Sources of Fecal Contamination ................................................................................................4

New Methods for Biosensor Detection of FIB .........................................................................5

References ..................................................................................................................................8

Chapter 2: Loading of fecal indicator bacteria in North Carolina tidal creek headwaters: hydrographic patterns and terrestrial runoff relationships .......................................................13

Introduction ..............................................................................................................................13

Methods....................................................................................................................................16

Results ......................................................................................................................................20 Discussion ................................................................................................................................24

Conclusion ...............................................................................................................................32

References ................................................................................................................................45

Chapter 3: Quantifying fecal contaminant sources in North Carolina tidal creeks utilizing

three Bacteroides spp. assays, conventional fecal indicators, and optical brighteners ............51

Page 10: New approaches and technologies for quantifying the impact

x

Introduction ..............................................................................................................................51

Methods....................................................................................................................................54

Results ......................................................................................................................................61 Discussion ................................................................................................................................66

Conclusions ..............................................................................................................................74

References ................................................................................................................................87

Chapter 4: Filtration and elution techniques for concentration of intact Escherichia coli,

Enterococcus spp., and Bacteroides thetaiotaomicron cells from seawater............................93 Introduction ..............................................................................................................................93

Methods....................................................................................................................................96

Results and Discussion ..........................................................................................................101

Conclusions ............................................................................................................................109

References ..............................................................................................................................120

Chapter 5: Optimized capture and concentration of Enterococcus spp. and E. coli from water

using Immunomagnetic Separation (IMS)................................................................................125

Introduction ............................................................................................................................125

Methods..................................................................................................................................128

Results ....................................................................................................................................134 Discussion ..............................................................................................................................137

Conclusions ............................................................................................................................142

References ..............................................................................................................................151

Chapter 6: Conclusion............................................................................................................155

Appendices.............................................................................................................................159

Page 11: New approaches and technologies for quantifying the impact

xi

LIST OF TABLES

Table

2.1 Watershed size, predominant soil and land use types for each tidal creek ........................34

2.2 Loading differences between storm and baseflow (times greater), and days required during dry weather to equal total E. coli and Enterococcus spp. loading during an average

storm for each tidal creek.........................................................................................................35

2.3 Spearman rank correlations (R2) for all creeks combined and individual tidal creeks for loads versus rainfall metrics and t-test results for average event mean concentrations

for EC and ENT during both storm and equivalent baseflow periods. ** Significant at p<0.01 * Significant at p<0.5..................................................................................................36

3.1 Reaction conditions and standards for all QPCR assays ...................................................77

3.2 Description of tidal creek predominant land cover, fecal indicator detection, and

percent exceedence of regulatory standards for conventional indicators ................................78

3.3 Bivariate correlations among log transformed data from all creeks using two-tailed Spearman Rho rank correlation, and expressed as correlation coefficient r. ** indicates

p<0.01 .....................................................................................................................................79

4.1 Different methods and treatments tested for bacterial concentration, elution, and particulate removal.................................................................................................................111

4.2 Recovery of E. coli and particulate removal efficiency for seawater spiked with E. coli concentrations of 103 MPN/100 mL ...........................................................................112

5.1 Variables tested for Immunomagnetic Separation to optimize recovery of target cells

and reduce particles in sample ...............................................................................................143

Page 12: New approaches and technologies for quantifying the impact

xii

LIST OF FIGURES

Figure

2.1 Streamflow (m3/s) versus log fecal indicator bacteria concentration (Most Probable Number (MPN)/100 mL) and Total Suspended Solids............................................................38

2.2 Box plots for all creeks across all storms of mean (A) E. coli (B) Enterococcus spp.

(C) flow (D) and total suspended solids ..................................................................................39

2.3 Pollutograph representation of intra-storm variation in log concentrations of E. coli (EC) and Enterococcus spp. (ENT) (MPN/100 mL), and total suspended solids (TSS)

(mg/L) during one representative storm in the four headwater tidal creeks ............................40

2.4 Average mass to volume (M/V) plots, to determine fecal indicator bacteria load in volume of creek discharge, for (A) E. coli and (B) Enterococcus spp. for each tidal creek ...41

2.5 Mean E. coli and Enterococcus spp. concentrations (MPN/g) for instream sediment

during summer baseflow conditions in each tidal creek. Error bars represent standard deviation...................................................................................................................................42

2.6 Relationship between log concentrations (MPN/100 mL) of E. coli and Enterococcus

spp. during wet weather (rainfall > 1.27 cm) stormflow samples, and dry weather (rainfall = 0 cm) baseflow samples ..........................................................................................43

2.7 Study site including tidal creeks and map of Marine Corps Base Camp Lejeune .............44

2.8 (A) E. coli and total suspended solids (log mg/L) during all conditions (B) and Enterococcus spp. and total suspended solids (log mg/L) during all conditions. ** Significant at p<0.01 ...........................................................................................................45

3.1 Marine Corps Base Camp Lejeune tidal creeks of study. Arrows indicate creeks examined for Bacteroides spp. using three discrete assays .....................................................80

3.2 Regulatory exceedences for E. coli (right half) and Enterococcus spp. (top half) for

each creek during all sampling events. E. coli and Enterococcus spp. exceedences are based on regulatory standards of 320 and 104 MPN/100 mL, respectively ............................81

Page 13: New approaches and technologies for quantifying the impact

xiii

3.3 Mean log indicator concentration for each creek. Error bars represent the standard error of the mean ......................................................................................................................82

3.4 Box-plots of log mean concentrations of conventional and Bacteroides indicators for

all creeks combined..................................................................................................................83

3.5 Relationship of BacHum to HF183 results when HF183 assay produced positive results. Spearman’s rho= 0.793 and p<0.05 .............................................................................84

3.6 A representative sample for optical brightener decay curves for each tidal creek and Tide 2x standard. Southwest Creek and the 2x Tide standard are positive for optical brighteners................................................................................................................................85

3.7 Relationships among mean log indicator concentrations from each tidal creek headwater and percent land-development for each watershed.................................................86

4.1 Diagram of the Multi- filter Method (MFM) for concentration and elution of bacteria

for sample volumes less than 1.5 L........................................................................................113

4.2 Diagram of the In- line Method (ILM) for concentration and elution of bacteria for sample volumes less than 5 L ................................................................................................114

4.3 Seawater particulate diameter distribution composition for A. Raw seawater B. Nitex

filtered seawater C. Geotech Dispos-a-filter seawater after filtration D. Seawater centrifuged for 1 min. at 600×g and top 200 mL aspirated. (1) Represents artificial spike

with a known grain size to achieve obscuration of the CILAS 1180 particle size analyzer, and (2) represent actual remaining particles in sample after applied filtration or centrifuging treatment ............................................................................115

4.4 Mean total recovery (due to prefiltration, concentration, and elution step) percentage using optimized method for E. coli (EC), Enterococcus spp. (ENT), and B. thetaiotaomicron (BTIM) for Multi- filter Method (MFM) vs. In-Line Filter Method

(ILM) (E. coli only). Error bars represent standard deviation ..............................................116

Page 14: New approaches and technologies for quantifying the impact

xiv

4.5 Concentration of Bacteroides thetaiotaomicron using Sybr Green bacterial staining enumeration method (A) 5 mL seawater (SW) before concentration (B) 100 µl SW after

concentration ..........................................................................................................................117

4.6 Differences in concentration factor of MFM vs. ILM for E. coli ....................................118

4.7 Percent recovery of E. coli from seawater using Multi- filter Method (MFM) with

212-300 um glass beads vs. no beads, and elution with PBS, PBST-80, and PET solutions. Error bars represent standard deviation .................................................................119

5.1 The Immunomagnetic Separation Method (IMS) showing A) mixing, B) capture,

C) washing, and D) resuspension of cleaned sample with target bacterial cells ...................144

5.2 Recovery of Enterococcus spp. using M280 streptavidin coated DynaBeads and Virostat biotin conjugated polyclonal Enterococcus spp. antibody. Seawater samples

amended with sewage, cultured E. faecalis, and E. faecium, and two un-amended tidal creek samples (Gillets and Courthouse) were tested with for recovery of Enterococcus. Immunomagnetic separation indicates moderate capture of sewage amended and

environmental samples and poor capture of samples amended with cultured cells ..............145

5.3 Recovery percentage with Dynal tosylactivated beads (concentration of 3.55 × 107) and Meridian E. coli and Enterococcus antibody, mixed with particulates and sewage

spiked (8 × 103 per mL E. coli, 3 × 103 per mL Enterococcus). All samples mixed with beads, incubated at 37o C at 25 rpm and enumerated after two washes ................................146

5.4 Bead recovery of E. coli from bead concentrations of 106 to 107 beads per sample.

E. coli concentrations were ~7 × 103 per sample, and samples were incubated at 37o C in a shaker incubator for 30 minutes. Average taken from combined Colilert-18 and

mTEC results. Error bars represent standard deviation .........................................................147

5.5 Representative result from immunomagnetic separation using Dynabead M-280 tosylactivated beads for E. coli capture from a 1 mL sewage sample. E. coli recovery is

28.4%, and each subsequent wash using PBS results in loss of a small amount of bacteria ................................................................................................................................................148

Page 15: New approaches and technologies for quantifying the impact

xv

5.6 Chemical additions to particle amended sewage spiked sample for 5, 15, and 30 minutes at 37° C and high speed mixing. Raw is the initial sample, and centrifuge control

had no chemical added. Control(CNTRL) was 1x PBS, Quercetin(QC) 0.001M, Gallic Acid (GA) 0.001M, Ellagic Acid (EA) 0.0025M, Epigallocatechin Gallate (EA) 0.03

mg/ mL, and Catechin Hydrate (CH) 0.5 mg/ mL. All samples above reference line show greater bacterial availability in sample suspension after treatment .......................................149

5.7 Epifluorescent microscopy of E. coli bound to particulates and beads. Images were

taken after 30 min incubation with magnetic beads and 2 washes. Particulates are still very abundant in sample, and E. coli are highly attached to particles ...................................150

Page 16: New approaches and technologies for quantifying the impact

xvi

ABBREVIATIONS

Ab: Antibody

APHA: American Public Health Association

ATCC: American Type Culture Collection

BMP: best management practices

CDx: Crystal Diagnostics

CE: cell equivalent

CFU: colony forming units

Ct: cycle threshold

DST: defined substrate technology

DNA: deoxyribonucleic acid

EPA: United States Environmental Protection Agency

EC: Escherichia coli

ENT: Enterococcus species

EMC: event mean concentrations

FC: fecal coliforms

FIB: fecal indicator bacteria

ILM: in- line method

IMS: immunomagnetic separation

LCLC: lyotropic chromonic liquid crystal

LOD: limit of detection

MCBCL: Marine Corps Base Camp Lejeune

MF: membrane filtration

Page 17: New approaches and technologies for quantifying the impact

xvii

MFM: multi- filter method

MPN: most probable number

M/V: mass to volume

NC: North Carolina

NCDWQ: North Carolina Division of Water Quality

ND: non-detects

NRE: New River Estuary

OB: optical brighteners

PBS: phosphate buffered saline

PBST: phosphate buffered saline + tween-20

PC: Polycarbonate

QPCR: quantitative polymerase chain reaction

RNA: ribonucleic acid

SE: standard error

SD: standard deviation

SPC: specimen processing control

TMDL: total maximum daily load

TSS: total suspended solids

Page 18: New approaches and technologies for quantifying the impact

CHAPTER 1

Introduction and Background

Marine waters are important aesthetically, for recreation activities, and as economic

stimulants for coastal communities (Lipp et al., 2001; Jeng et al., 2005; Stewart et al., 2007).

However, increasing microbial fecal contamination continues to decrease the availability of

these waters for recreational (swimming) and commercial uses (shellfish harvesting) and

increases risk of disease acquisition from fecal pathogens (Washburn et al., 2003; Fuhrman et

al., 2005). For instance, users of fecally contaminated water bodies for bathing and

consumers of shellfish from these waters have a higher risk of illnesses such as

gastrointestinal illness (diarrhea), skin rashes, respiratory related illness (Cabelli et al., 1979;

Cabelli et al., 1982; Wade et al., 2006; Colford et al., 2007), and even death in cases of

contaminated shellfish consumption (Rippey, 1994). Fecal contamination can originate from

livestock, human waste (wastewater treatment plants as well as septic systems), and wildlife.

Human source contamination in waters can occur from improperly functioning septic

systems, problems with wastewater treatment plants resulting in illicit discharges, and

overland transport to water bodies during rainfall events. Conventional fecal indicator

bacteria (FIB) (E. coli and Enterococcus) are used to determine if a water body is

contaminated with fecal material, and therefore pathogens (i.e. harmful bacteria and viruses).

The coastal waters of North Carolina (NC) are important recreational areas for

bathing, fishing, and shellfish harvesting. However fecal contamination is is one of the

Page 19: New approaches and technologies for quantifying the impact

2

biggest problems in North Carolina (NC) waters, resulting in closure of streams and estuarine

waters (Mallin et al., 2000; Coulliette and Noble, 2008; Line et al., 2008). Within the New

River Estuary (NRE) (the study site for much of this thesis research), 11.3 km2 of estuarine

shellfish waters are impaired, and 62.8 freshwater km are listed as stressed due to fecal

contamination, with 163 freshwater km yet to be assessed (NCDWQ 2007). The New River

Estuary is of particular importance, due to its location within Marine Corps Base Camp

Lejeune (MCBCL). MCBCL is the largest amphibious training base in the United States,

simultaneously hosting more than 47,000 Marines. The need for clean water within this

watershed is paramount, due to its use for amphibious military training as well as recreational

activities such as bathing and shellfishing.

Despite the ubiquity of fecal contamination in many of NC estuaries and costal

waters, little is known about transport (loading) or source of contamination (human vs.

wildlife) of these fecal indicators. Furthermore, current methods for detection are slow,

requiring 18-24 hours to results, increasing risk to water users, and establishing the need for

rapid biosensor technologies for water quality analysis. This thesis has focused on ways to

more rapidly and accurately determine fecal pathogens in NC estuarine waters. Research was

conducted to better understand how these microbes are transported within a coastal syste m,

new human-specific bacterial indicators were tested to differentiate between human and

wildlife sources, and methods were developed to assist with emerging detection technologies

(i.e. biosensors).

Background by Chapter

Stormwater Microbial Loading of Tidal Creeks

Page 20: New approaches and technologies for quantifying the impact

3

In eastern North Carolina (NC), tidal creeks are heavily utilized for shellfish

harvesting, boating, fishing, swimming, and in the NRE, for amphibious military training

(MCBCL, 2006). Loading of FIB to headwater, primarily freshwater, portions of tidal creeks

can have cumulative affects downstream in high priority estuarine waters. Understanding FIB

loading characteristics of these tidal creek headwaters during storm and baseflow periods is

important for overall understanding of impairment and loss of beneficial use of waters of the

NRE.

Current sampling methods estimate fecal pollution of waters based on geometric

mean sampling or single sample thresholds (USEPA 1986). Due to the rapid fluctuations

which can occur in fecal concentrations in stormwater, single sample strategies are often

insufficient and could lead to inaccurate water quality designations (DiDonato et al., 2009).

An intensive, multi-sampling strategy throughout storm hydrograph (Hyer and Moyer, 2004)

and during baseflow is essential in characterization of fecal associated hydrologic patterns,

and accurate fecal concentration and loading estimations. There have been no previous

attempts to determine hydrological loading characteristics of fecal contamination in tidal

creeks of the NRE. Because stormwater runoff can often account for a much greater portion

of overall FIB loading in creeks when compared to non-event baseflow loading (Reeves et

al., 2004; Surbeck et al. 2006; Krometis et al., 2007), characterizing loading patterns is

important in order to accurately manage waters.

Chapter 2 examines how transport of pathogens occurs in the New River Estuary,

using intensive storm and non-storm sampling, and new methods for determining transport

(loading) of these pathogens from creeks to the estuary. Patterns and characteristics of FIB

loading in the tidal creeks of the NRE over a range of meteorological conditions were

Page 21: New approaches and technologies for quantifying the impact

4

quantified during both base and storm flow conditions. FIB load relationships to stream flow

and a first flush characterization are presented. Defining these loading trends will assist water

quality managers in determining magnitude of water quality impairment, time needed for

attenuation of pollutants, and making presumptive closures.

Sources of Fecal Contamination

Conventional indicators (fecal coliforms, Enterococcus spp., and E. coli) are limited

in their ability to determine sources of fecal contamination. Understanding source of fecal

contamination is paramount, since many pathogens of human concern are derived from

human specific sources (Yan and Sadowsky, 2007). Some non-point source contaminated

waters have little or very weak correlation with FIB and the occurrence of viral pathogens of

human concern (Noble and Fuhrman, 2001; Jiang et al., 2004). The result of such findings

could indicate that many waters polluted by non-point source runoff could be overestimating

risks to human health (Colford et al., 2008). Therefore it is important to support conventional

indicator results of fecal pollution with an indicator that is able to connect source of fecal

pollution to human or non-human sources.

Species of Bacteroides have been proposed as alternatives to conventional indicators

and as a better predictor of human contamination (Rolfe et al., 1977; Fiksdal et al., 1985).

Bacteroides have several advantages over current indicators: (1) they are obligate anaerobes,

and presence is representative of relatively recent contamination (1-2 days) (Kreader, 1998)

(2) they are highly abundant in feces of warm blooded animals, and dominant enteric bacteria

in humans, with individual Bacteroides species as high as 1010 per g of human feces (Fiksdal

et al., 1985) (3) they do not proliferate once released into the environment (Dick and Field,

2004) and (4) they show significant correlation with certain pathogens (Savichtcheva et al.,

Page 22: New approaches and technologies for quantifying the impact

5

2007). Previous research has found connections to human fecal contamination and

Bacteroides spp. using discrete assays (Bernhard et al., 2003; Dick and Field, 2004; Carson

et al., 2005; Noble et al., 2006). Throughout the NRE, many creek and estuary waters are

closed for shellfishing due to high FIB concentrations, though the sources of these fecal

contaminants are unclear. Previous work has speculated that in certain tidal creeks (Freeman

Creek), contamination from troop activities could be contributing to contamination during

storm runoff (C2HM Hill Inc., 2000). Determining presence of Bacteroides spp. in tidal

creek headwaters will increase understanding of sources of fecal contamination when

combined with conventional fecal indicator data.

In Chapter 3, three Bacteroides spp. assays were examined in an effort to determine

source (human vs. wildlife) of fecal contamination. Streams determined to have prevalent

conventional FIB contamination were targeted. Optical brightener analysis was also

conducted in order to examine if relationships existed between conventional indicators,

Bacteroides molecular assays, and optical brightener results. This “toolbox approach” will

help elucidate the potential sources of fecal contamination and risk of exposure to military

personnel during training activities, as well as recreational water users.

New Methods for Biosensor Detection of FIB

New rapid methods are necessary for more accurate and simple detection of fecal

contamination of water (Noble and Weisberg, 2005). Current methods for enumeration of

FIB such as multiple tube fermentation, membrane filtration, and chromogenic substrate test

kits require 18-24 hours for quantification of the target FIB (Dick and Field, 2004). Though

rapid molecular techniques (QPCR) have been developed, they require well trained staff with

expertise in multiple preparatory and analytical steps (Ivnitski et al., 1999). Due to the

Page 23: New approaches and technologies for quantifying the impact

6

drawbacks in current technology, numerous rapid and simple biosensors for detection of

bacteria and pathogens in waters are in research and development. Biosensors are

collectively defined as instruments which bind and enumerate targeted analytes, such as

pathogens in water, food, and clinical settings (Helfinstine, 2006).

The use of liquid crystal is one such potentially rapid technology under research and

development for use as a biosensor medium. Specifically lyotropic chromonic liquid crystal

(LCLC) is non-toxic to bacteria (Woolverton et al., 2005), and when used within a readable

cell, could be sensitive and selective, producing results in less than 1 hour (Niehaus, 2004).

Liquid crystals have a mesophase in which molecules form a crystal- like alignment, while

still maintaining a liquid characteristic (Tam-Chang and Huang, 2008). Crystal Diagnostics

(CDx) is currently developing a biosensor utilizing LCLC for detection of fecal indicators

and other potential biohazards in both fresh and marine waters. The current prototype utilizes

microspheres (magnetic beads) with labeled antibodies, which bind to specific target antigens

to form immunocomplexes, resulting in disruption of the LC alignment and changes in

polarized light. Liquid crystal disturbances, or events, caused by aggregate

immunocomplexes can then be enumerated to quantify target cells (Shiyanovskii et al.,

2005).

The CDx biosensor is currently in development to detect fecal indicators, including E.

coli and Enterococcus spp. However, before testing of bacteria and pathogens can continue,

the device must be optimized for dealing with water samples. Current limitations to the

success of this prototype and other biosensors are the interference of particulates and

unwanted debris (Tam-Chang and Huang, 2008). In addition, approximately 5 L volumes of

sample must be concentrated before analysis in order to address sensitivity limitations of the

Page 24: New approaches and technologies for quantifying the impact

7

device. These are common problems with many new microbial detection technologies, as

water quality thresholds have relatively low cell counts (e.g. 35 CFU/100 mL geometric

mean, or 104 CFU/100 mL per single sample Enterococcus spp. thresholds at bathing

beaches), while most devices can handle less than 1 mL volumes, requiring concentration of

samples (Noble and Weisberg, 2005; Stewart et al., 2008). Chapters 4 and 5 examine sample

preparation techniques with seawater samples, including concentration, filtration,

centrifugation, and immunomagnetic separation (IMS) in order to optimize concentration of

cells and removal of particulates.

Page 25: New approaches and technologies for quantifying the impact

8

REFERENCES

Bernhard, A.E., Goyard, T., Simonich, M.T., Field, K.G. (2003). Application of a rapid

method for identifying fecal pollution sources in a multi-use estuary. Water Research, 37, 909–913.

Cabelli, V.J., Dufour, A.P., Levin, M.A., McCabe, L.J., Haberman, P.W. (1979).

Relationship of microbial indicators to health effects at marine bathing beaches. American Journal of Public Health, 69(7), 690-696.

Cabelli, V.J., Dufour, A.P., McCabe, L.J., Levin, M.A. (1982). Swimming-associated

gastroenteritis and water quality. American Journal of Epidemiology, 115(4), 606-616.

Carson, C.A., Christiansen, J.M., Yampara-Iquise, H., Benson, V.W., Baffaut, C., Davis, J.V., Broz, R.R., Kurtz, W.B., Rogers, W.M., Fales, W.H. (2005). Specificity of a

Bacteroides thetaiotaomicron marker for human feces. Applied and Environmental Microbiology, 71(8), 4945-4949.

Colford, J.M., Wade, T.R., Schiff, K.C., Wright, C.C., Griffith, J. F., Sukhminder, S.K.,

Burns, S., Sobsey, M., Lovelace, G., Weisberg, S.B. (2007). Water quality indicators and the risk of illness at beaches with nonpoint sources of fecal contamination. Epidemiology, 18(1), 27-35.

Coulliette. A.D., Noble, R.T. (2008). Impacts of rainfall on the water quality of the Newport River Estuary (Eastern North Carolina, USA). Journal of Water and Health, 6(4), 473-482.

C2HM Hill, Inc., (2000). Freeman Creek evaluation and assessment. Prepared for US Army

Corps of Engineers, Contract No. DACA01-96-0028

Dick, L.K., Field, K.G. (2004). Rapid estimation of numbers of fecal Bacteroidetes by use of a quantitative PCR assay for 16S rRNA genes. Applied and Environmental Microbiology,

70(9), 5695-5697.

DiDonato, G.T., Stewart, J.R., Sanger, D.M., Robinson, B.J., Thompson, B.C., Holland, A.F., Van Dolah, R.F. (2009). Effects of changing land use on the microbial water quality of

tidal creeks. Marine Pollution Bulletin, 58(1), 97-106.

Page 26: New approaches and technologies for quantifying the impact

9

Fiksdal, L., Maki, J.S., LaCroix, S.J., Staley, J.T. (1985). Survival and detection of

Bacteroides spp. prospective indicator bacteria. Applied and Environmental Microbiology, 49(1), 148-150.

Fuhrman, J.A., Liang, X., Noble, R.T. (2005). Rapid detection of enteroviruses in small

volumes of natural waters by real- time quantitative reverse transcriptase PCR. Applied and Environmental Microbiology, 71(8), 4523-4530.

Helfinstine, S.L., Lavrentovich, O.D., Woolverton, C.J. (2006). Lyotropic liquid crystal as a

real time detector of microbial immune complexes. Letters in Applied Microbiology, 43, 27-32.

Hyer, K. E., Moyer, D. L. (2004). Enhancing fecal coliform total maximum daily load

models though bacterial source tracking. Journal of American Water Resources Association, 40(6), 1511-1526.

Ivnitski, D., Abdel-Hamid, I., Atanasov, P., Wilkins, E. (1999). Review: Biosensors for

detection of pathogenic bacteria. Biosensors and Bioelectronics, 14, 599-624.

Jeng, H.A.C., Englande, A.J., Bakeer, R.M., Bradford, H.B. (2005). Impact of urban stormwater runoff on estuarine environmental quality. Estuarine Coastal and Shelf Science,

63, 513-526.

Jiang, S.C., Chu, W. (2004). PCR detection of pathogenic viruses in southern California urban rivers. Journal of Applied Microbiology, 97(1), 17-28.

Kreader, C.A. (1998). Persistence of PCR-detectable Bacteroides distasonis from human feces in river water. Applied and Environmental Microbiology, 64(10), 4103-4105.

Krometis, L.H., Characklis, G.W., Simmons, O.D., Mackenzie, J.D., Likirdopulos, C.A.,

Sobsey, M.D. (2007). Intra-storm variability in microbial partitioning and microbial loading rates. Water Research, 41(2), 506-516.

Line, D.E., White, N.M., Kirby-Smith, W.W., Potts, J.D. (2008). Fecal coliform export from

four coastal North Carolina areas. Journal of American Water Resources Association, 44(3), 606-617.

Page 27: New approaches and technologies for quantifying the impact

10

Lipp, E.K., Farrah, S.A., Rose, J.B. (2001). Assessment and impact of microbial fecal pollution and human enteric pathogens in a coastal community. Marine Pollution Bulletin,

42, 286-293.

Mallin, M.A., Williams, K.E., Esham, E.C., Lowe, R.P. (2000). Effect of human development on bacteriological water quality in coastal watersheds. Ecological Applications,

10(4), 1047-1056.

MCBCL- Marine Corps Base Camp Lejeune (2006). Integrated Natural Resource Management Plan. http://www.lejeune.usmc.mil/emd/INRMP/INRMP.htm.

Niehaus, G.D. (2004). Self contained assay device for rapid detection of biohazardous

agents. United States Patent Publication, Patent # 0185551 A1.

NCDWQ, North Carolina Division of Water Quality (2007). White Oak River Basinwide Water Quality Plan. Planning Section.

Noble, R.T., Fuhrman, J.A. (2001). Enteroviruses detected in the coastal waters of Santa Monica Bay, CA: low correlation to bacterial indicators. Hydrobiologia, 460, 175-184.

Noble, R.T., Weisberg, S.B. (2005). Technologies for rapid detection of bacteria in

recreational waters. Journal of Water and Health, 3(4), 381-392.

Noble, R.T., Griffith, J.F., Blackwood, A.D., Fuhrman, J.A., Gregory, J.B., Hernandez, X., Liang, X., Berra, A.A., Schiff, K., (2006). Multitiered approach using quantitative PCR to

track sources of fecal pollution affecting Santa Monica Bay, California. Applied and Environmental Microbiology, 72(2), 1604-1612.

Reeves, R.L., Grant, S.B., Mrse, R.D., Copil-Oancea, C.M., Sanders, B.F., Boehm, A.B.,

(2008). Scaling and management of fecal indicator bacteria in runoff from a coastal urban watershed in Southern California. Environmental Science and Technology, 38(9), 2637-2648.

Rippey, S.R. (1994). Infectious diseases associated with molluscan shellfish consumption. Clinical Microbiology Reviews, 7(4), 419-425.

Page 28: New approaches and technologies for quantifying the impact

11

Rolfe, R.D., Hentges, D.J., Barrett, J.T., Campbell, B.J. (1977). Oxygen tolerance of human intestinal anaerobes. The American Journal of Clinical Nutrition, 30, 1762-1769.

Savichtcheva, O., Okayama, N., Okabe, S. (2007). Relationships between Bacteroides 16S

rRNA genetic markers and presence of bacterial enteric pathogens and conventional fecal indicators. Water Research, 41, 3615-3628.

Shiyanovskii, S.V., Schneider, T., Smalyukh, I. I., Ishikawa, T., Niehaus, G.D., Doane, K.J.,

Woolverton, C.J., Lavrentovich, O.D., (2005). Real- time microbe detection based on director distortions around growing immune complexes in lyotropic chromonic liquid crystals,

Physical Review E, 71(2), 020702, 1-4.

Stewart, J.R., Santo Domingo, J.W., Wade, T.J. (2007). “Fecal pollution, public health, and microbial source tracking.” Eds. Domingo J.W., Sadowsky, M.J. Microbial Source

Tracking; Santo ASM Press: Washington, DC, Chapter 1.

Stewart, J.R., Gast R.J., Fujioka, R.S., Solo-Gabriele, H.M., Meschke, J.S., Amaral-Zettler, L.A., del Castillo, E., Polz, M.F., Collier, T.K., Strom, M.S., Sinigalliano, C.D., Moeller,

P.D.R., Holland, A.F. (2008). The coastal environment and human health: microbial indicators, pathogens, sentinels and reservoirs. Environmental Health, 7(Suppl 2):S3, 1-14.

Surbeck, C.Q., Jiang, S.C., Ahn, J.H., Grant, S.B. (2006). Flow fingerprinting fecal pollution

and suspended solids in stormwater runoff from an urban coastal watershed. Environmental Science and Technology, 40(14), 4435-4441.

Tam-Chang, S., Huang, L. (2008). Chromonic liquid crystals: properties and applications as

functional materials. Chemical Communications, 1957-1967.

USEPA, United States Environmental Protection Agency, (1986). Ambient water quality criteria for bacteria-1986. Office of Water; Regulations and Standards. Washington D.C.

Wade, T.J., Calderon, R.L., Sams, E., Beach, M., Brenner, K.P., Williams, A.H., Dufour,

A.P. (2006). Rapidly measured indicators of recreational water quality are predictive of swimming-associated gastrointestinal illness. Environmental Health Perspectives, 114(1),

24–28.

Page 29: New approaches and technologies for quantifying the impact

12

Washburn, L., McClure, K.A., Jones, B.H., Bay, S.M. (2003). Spatial scales and evolution of

stormwater plumes in Santa Monica Bay. Marine Environmental Research, 56, 103-125.

Woolverton, C.J., Gustely, E., Li, L., Lavrentovich, O.D. (2005). Liquid crystal effects on bacterial viability. Liquid Crystals, 32(4), 417–423.

Yan, T., Sadowsky, S.J. (2007). Determining sources of fecal bacteria in waterways. Environmental Monitoring Assessment, 129, 97-106.

Page 30: New approaches and technologies for quantifying the impact

CHAPTER 2

Loading of fecal indicator bacteria in North Carolina tidal creek headwaters: hydrographic patterns and terrestrial runoff relationships*

*Published in Water Research (44:16): pg 4704-4715 (Stumpf et al. 2010)

Introduction

Tidal creeks are important conduits of nutrients and contaminants in stormwater

runoff to receiving waters (Mallin and Lewitus, 2004; Didonato et al., 2009), Fecal indicator

bacteria (FIB), such as fecal coliforms (FC), E. coli (EC), and Enterococcus spp. (ENT), are

used as proxies for pathogens and as the primary means of determining fecal pollution in

tidal creeks and estuaries as regulated by the USEPA (1987). Concentrations of FIB in

surface waters impacted by stormwater runoff can increase dramatically during rainfall

events in comparison to baseline conditions (Wilkinson et al., 1995; Lipp et al., 2001;

Shehane et al., 2005), This increase is often due in part to fecal contamination washed from

impervious surfaces and terrestrially-associated fecal contamination that is scoured from land

and transported via stormwater into receiving waters (Geldreich et al., 1968; Weiskel et al.,

1996; Mallin et al., 2000). Stormwater runoff can account for a much greater portion of

overall FIB loading in creeks when compared to non-event baseflow loading (Reeves et al.,

2004; Surbeck et al., 2006; Krometis et al., 2007). Stormwater runoff to receiving waters

such as tidal creeks has the potential to impact public health in myriad ways. Tidal creeks in

eastern North Carolina (NC) are heavily utilized for shellfish harvesting, boating, fishing,

swimming, and, in the New River Estuary (NRE), for amphibious military training. In

Page 31: New approaches and technologies for quantifying the impact

14

addition, loading of FIB to headwater portions of tidal creeks will eventually be transported

to downstream areas and estuarine receiving waters. Understanding FIB loading

characteristics of these tidal creek headwaters is important for overall understanding of

estuarine and coastal water quality dynamics and beneficial use impairments.

Many tidal creeks and estuaries in NC are listed as “impaired” due to elevated fecal

coliform levels. In eastern North Carolina, previous research has shown that greater than

93% of E. coli are fecal coliforms (n=3020, Noble and Kirby-Smith, unpublished data), and

therefore EC and FC are used interchangeably for this study. In compliance with section

303(d) of the Clean Water Act (USEPA, 1987), total maximum daily loads (TMDLs) of

pollutants are established for each water body based on its designated uses. Fecal

contamination, measured through the use of fecal indicator bacteria, is the second most

common pollutant responsible for failure of designated use (ie. impairment) in assessed

waters in the United States (Stewart et al., 2007). Within the 1,436 km2 New River

Watershed in eastern NC, 11.3 km2 of estuarine shellfish waters are “impaired” and 62.8 km2

of freshwater are “stressed” due to fecal contamination, with 163 km2 of freshwater yet to be

assessed for TMDLs (NCDWQ, 2007). Waters designated as “impaired” have FC counts in

exceedence of standards for specific beneficial uses such as shellfish harvestings, while

“stressed” waters support beneficia l use, but are potentially at risk of impairment.

Precipitation patterns in eastern NC consist of episodic events throughout the year

and seasonal hurricanes and nor’easters that can contribute large amounts of precipitation

over short durations. Dry weather discharges and rainfall associated runoff have led to water

quality impairment from FIB in NC creeks and estuaries (Mallin et al., 2001; Fries et al.,

2008; Line et al., 2008), Rainfall thresholds (for example, 2.54 cm of rainfall will result in

Page 32: New approaches and technologies for quantifying the impact

15

closure of certain shellfish areas in NC based on previous research showing fecal

contamination after this threshold rainfall) are used in some areas of NC to protect public

health; however, these thresholds were based on few samples, and Coulliette and Noble

(2008) noted that the impairment threshold for the Newport River Estuary based on the

shellfish fecal coliform standard occurred even before 2.54 cm of rainfall. In addition,

differences among rainfall metrics, including rainfall duration and antecedent rainfall result

in high levels of variability in the loads of FIB transported to receiving waters during storms.

Studies to specifically examine FIB of headwater tidal creeks and the potential

impact on downstream estuarine shellfishing waters in relationship to storm loading are few.

An important departure of this study from other published wet weather FIB monitoring

studies was the use of automated sampling throughout the duration of a storm (as opposed to

the use of single grab samples to represent an entire storm), which allows for a more accurate

estimate of cumulative storm load and the Event Mean Concentration (EMC). Additionally,

this work was also conducted in rarely sampled headwater tidal creeks, and in primarily

undeveloped watersheds, with particular soil characteristics (primarily fine sands and loam),

and highly productive yet impaired shellfishing waters in eastern NC. Previous studies

utilizing automated sampling, and examining similar loading characteristics have primarily

been conducted in the western United States, in urban creeks, and highly developed/highly

impermeable watersheds. Accurate loading characterization of FIB is imperative for

managing and developing appropriate mitigation strategies for shellfishing waters potentially

impaired from fecal contamination throughout the United States.

The goal of this study was to examine the loading of E. coli (EC) and Enterococcus

spp. (ENT) over a range of rainfall and dry weather conditions in headwater portions of tidal

Page 33: New approaches and technologies for quantifying the impact

16

creeks of the NRE in eastern NC. A bi-monthly and multi-sample flow-paced storm sampling

strategy was employed to quantify load during both base and storm flow conditions. The

load of FIB was analyzed in relationship to rainfall amount, stream flow, storm duration, and

antecedent rainfall. A loading analysis was conducted to determine whether the patterns

observed followed a typical “first flush” scenario. Relationships between EC, ENT, and total

suspended solids (TSS) were investigated in an effort to understand potential pathways

(runoff vs. resuspension of sediment reservoir populations) and begin to understand potential

sources (human vs. nonhuman) of fecal contamination. Finally, instream sediments were

examined to determine if sediment reservoir populations, resuspended during storm events,

were a potentially significant source of FIB to the water column.

Materials and Methods

Site Description

The New River watershed, located within Onslow County, NC, encompasses 1,436

km2 and contains the NRE, a broad, shallow estuary with a surface area of 88.1 km2 (Mallin

et al., 2005). Marine Corps Base Camp Lejeune (MCBCL), the Marine Corp’s largest

amphibious training base with greater than 47,000 Marines, occupies a large portion of the

watershed and is located adjacent to Jacksonville, NC. The estuary is used primarily for

boating, bathing, and commercial and recreational fishing and shellfish harvesting, but also

for military operations. Cogdel, French, Freeman, and Gillets are tidal creeks draining to the

NRE (Figure 2.7.) All are first-order creeks with watersheds less than 8 km2 and average

baseflows less than 0.12 m3/s, common for low-gradient creeks less than 3 m wide and 2 m

deep. The four watersheds have similar soils types predominantly comprised of fine sands,

Page 34: New approaches and technologies for quantifying the impact

17

have highly vegetated land cover, and little impermeable surfaces and/or development, with

the exception of Cogdel Creek (Table 2.1).

Field and Laboratory Methods

Storm and Baseflow Sampling

Surface water samples were collected during 10 storms from December 2007 to

December 2008 using ISCO automated water samplers (Model 6712, ISCO, Lincoln, NE)

equipped with flow sensors (Model 750). The ISCO autosamplers were located on the

downstream side of road culverts (i.e. drainage pipes) in the four headwater tidal creeks, and

used velocity sensors (pointed upstream) to measure stream velocity by acoustic doppler and

water level by pressure transducer, and collected 1L water samples in laboratory-washed

sample bottles. The monitored culverts were below the water surface and did not create

backwater conditions, cascades, or other changes in normal stream velocity or flow. Flow,

rainfall, and water quality data (dissolved oxygen, temperature, and conductivity) were

continuously collected and logged at 30-minute intervals using a YSI sonde, Model 600XL.

The ISCO autosamplers were programmed to sample throughout each storm event and were

enabled when flow exceeded predetermined threshold velocities. Baseflow samples (n=102)

were collected bi-monthly from September 2007 to December 2008 in sterilized, triple-rinsed

1L bottles, then placed on ice and processed within 6 h of collection. The ISCO bottles were

removed within 24 hours of automated sample collection, and processed immediately.

Previous work has shown minor degradation of EC and ENT during 24-hour ISCO storage

period in eastern NC (Noble and Fries, 2005) and was therefore considered inconsequential

for this study.

Page 35: New approaches and technologies for quantifying the impact

18

Rainfall

Rainfall was collected via automated rainfall samplers within the Freeman and

Cogdel Creeks watersheds. Freeman Creek rainfall data were utilized for Gillets Creek due to

close proximity (approximately 2.67 km separation) and Cogdel data were used for French

Creek (approximately 3.7 km separation). Total storm rainfall was calculated in the time

between elevation of hydrograph above baseflow until rainfall ceased. Storm rainfall duration

was determined by addition of consecutive 0.5 h increments when at least 0.025 cm of

rainfall was recorded, until storm completion. Seven-day antecedent rainfall was determined

by summation of all rainfall during the 7 days previous to an event.

Hydrograph Sample Selection

For each storm event, ISCO water samples were selected based on location within the

different stages of the hydrograph, determined using ISCO Flowlink Software (Version

4.01). The four storm stages analyzed were: 1) pre-storm, 2) rising limb, 3) peak, and 4)

falling limb, as determined by visual hydrograph analysis. Specific conditions for each stage

are as follows: “pre-storm” was collected before creek flow increased above baseflow and

less than 24 h before storm occurrence, “rising” samples were collected between baseline and

peak flow, “peak” samples were determined as the nearest single sample to the peak of the

storm hydrograph, and “falling” were samples collected after, but within 24 h of peak flow.

Similar multi- time point sample collection approaches, incorporating pollutograph

representations of contaminates, have previously been utilized (Hyer and Moyer, 2004;

Surbeck et al., 2006; Tiefenthaler et al., 2008).

Page 36: New approaches and technologies for quantifying the impact

19

FIB Enumeration and TSS Measurement

Water samples were processed in duplicate using Colilert-18 and Enterolert (IDEXX®

Laboratories, Westbrook, ME) at a 1:10 dilution to calculate most probable number (MPN)

of EC and ENT. Bottle controls were intermittently analyzed to determine if FIB

accumulated in empty ISCO sample bottles in the period between deployment and sample

collection. Results showed no contamination with EC and ENT. ISCO sample tubing was

flushed before individual samples were collected, and assumed uncontaminated, as previous

research under similar sampling conditions found little sample line contamination (less than 4

CFU/100 mL for E. coli) using similar samplers (Solo-Gabriele et al., 2000). TSS

concentrations were determined using mass collected on 0.7 µm filters, after a volume of

sample was filtered, according to standard method 2540D (APHA, 1998).

Sediments

To test the potential of instream sediment sources of FIB, sediment cores were

collected during summer (May, June, July) baseflow conditions when FIB concentrations

were typically highest, and processed to acquire FIB concentration/gram. Duplicate cores

(n=24) were collected directly above ISCO water sampling locations in each creek following

methods of Fries et al. (2008). Sediments were resuspended in phosphate buffered saline, and

enumerated using Colilert-18 and Enterolert as previously described for water sample

bacterial enumeration.

Data and Statistical Analysis

Page 37: New approaches and technologies for quantifying the impact

20

All statistics were conducted either using SPSS (version 11.0) or Microsoft EXCEL

software. An EXCEL interpolation function was used to estimate FIB concentrations

between sample points, similar to load estimation methods used previously (Ensign et al.,

2006). Volumetric flow (m3/s) was calculated by multiplying velocity (m/s) by wetted cross-

sectional area (m2) of each culvert, which was calculated using the cross sectional geometry

and water level (measured using pressure transducers). Total load (cells (MPN)/time) for

baseflow and storm flow conditions were estimated by multiplying FIB concentration

(MPN/100 mL) by flow volume (m3) per 30 min period.

Relationships among FIB loads, precipitation metrics (total rainfall, storm duration,

and antecedent rainfall), streamflow, and TSS were tested using spearman rank correlation

coefficient analysis. FIB and TSS concentrations were log transformed due to their non-

normal distribution, and log values were used to assess relationships among the indicators

and for comparison to streamflow, precipitation metrics, and TSS. EMCs were log

transformed and tested using a t-test to measure significant differences between storm and

equivalent baseflow periods. Significance was accepted at α = 0.05.

Results

FIB Total Load and EMC

FIB loads during storms exceeded baseflow loads for all creeks. EC loads during

storm events were as much as 30 times greater when compared to antecedent baseflow

loading over an equivalent time period, while ENT loads were as much as 37 times greater

than baseflow loads (Table 2.2). The smallest difference between base loading and total

storm loading for ENT and EC was at Gillets and Cogdel Creek, respectively. The greatest

Page 38: New approaches and technologies for quantifying the impact

21

difference between base loading and total storm loading for both EC and ENT was at

Freeman Creek. Calculations of equivalent time (days) of baseflow to achieve loading for

one average storm event produced variable results among creeks. Freeman had the greatest

temporal disparity, requiring as many as 110 and 134 baseflow days to achieve similar total

loading of EC and ENT, respectively, for one average storm (duration = 1.03 days) (Table

2.2). Gillets required the fewest baseflow days, 97 and 48 days for EC and ENT, respectively

(Table 2.2).

EMCs during storm and baseflow conditions were examined between creeks and all

creeks combined. EMCs are defined as total pollutant mass (M) divided by total flow

volume, or M/V, and as used as means to compare and correct for different flow volume

between creeks and storms (Kayhanian and Stenstrom, 2005). The median storm EMCs were

7.07 × 102 and 1.96 ×102 MPN/100 mL for storm EC and ENT, respectively, and 1.48 ×102

and 4.84 × 101 MPN/100 mL for baseflow EC and ENT, respectively, for all creeks

combined. Average EMCs for individual creek EC and ENT during storms and equivalent

baseflow periods ranged from 8.25 × 100 (MPN/100 mL) to 1.58 × 103. Storm and baseflow

FIB EMCs were statistically different when all creeks were combined (p<.01 for EC and

ENT), and for individual creeks (Table 2.3). Freeman EC EMC and Gillets ENT EMC were

the only creeks with no significant differences between storm and baseflow (Table 2.3).

FIB Patterns

Concentrations of FIB were generally strongly correlated with streamflow during

both storm and baseflow conditions. However, Gillets Creek, had no correlation between

streamflow and ENT (Figure 2.1), and unusually high baseflow levels of FIB in relation to

other creeks. When mean creek FIB concentrations from all creeks were examined

Page 39: New approaches and technologies for quantifying the impact

22

throughout the four stages of the hydrograph (pre, rising, peak, and falling), bacterial

concentrations (EC and ENT) increased during rising and peak stages, and declined with

falling hydrograph flow rates (Figure 2.2). Pollutographs for each storm and each creek

indicate intra-storm variability between EC and ENT concentrations at the different stages of

the hydrograph (Figure 2.3), and was common across all storms and all creeks. Pollutant

mass to discharge volume (M/V) ratios were developed for each creek during each storm

event to determine if a “first flush” phenomenon was observed. A first flush is said to exist

when the greatest contaminant concentrations, and therefore loads, occur at the onset of a

storm (Lee et al., 2002) and decline after the initial contaminant pulse is flushed from the

system (Gupta and Saul, 1996). Bertrand-Krajewski et al. (1998) defined a first flush as 80%

of total storm pollutant mass occurring in the first 30% of the storm discharge volume. All

creeks combined averaged 37 and 44% flushing of EC and ENT loads, respectfully, within

the first 30% of storm volume. Cogdel, Gillets, and French Creeks had similar average M/V

loading ratios, while Freeman Creek had the lowest ratio of M/V in the first 30% of

discharge volume (Figure 2.4).

Rainfall Metrics

Variability in total rainfall, rainfall duration, and 7-day antecedent rainfall was high

among different storms and creeks. Total rainfall for each storm ranged from 1.9 to 8.8 cm.

Combining all creeks and all storms (n=36), there were significant correlations between total

rainfall and total storm load of EC and ENT (Table 2.3), and EMCs and EC (R2=0.37, p<.05)

and ENT (R2=0.51, p<.01). Duration of rainfall during storms ranged from 6.5 to 27.5 h.

Across all watersheds, duration of rainfall was correlated with total ENT loading, but had no

Page 40: New approaches and technologies for quantifying the impact

23

correlation with EC loading (Table 2.3) and no significant correlations occurred between EC

or ENT EMCs and rainfall duration. At individual creeks there was no significant correlation

between rainfall duration and total loading and EMC of EC or ENT (Table 2.3). Seven-day

antecedent rainfall ranged from 0 cm to 5.9 cm, with an average of 1 cm. Antecedent rainfall

was not significantly correlated with FIB load for all creeks combined, or for individual

creeks, with the exception of ENT total load at Cogdel (Table 2.3). Antecedent rainfall was

not significantly correlated with FIB EMC for all creeks combined, or for individual creeks.

TSS and Instream Sediments

TSS concentrations were examined to determine if there were relationships between

FIB concentrations and streamflows. TSS showed a moderate correlation with streamflow in

Gillets and Cogdel (Figure 2.1). For all creeks combined, the largest load of mean TSS

occurred during the rising limb of storm flow, indicating more of a first flush (in terms of

M/V ratio) than FIB (Figure 2.2). However, TSS M/V ratios never reached the previously-

defined 80% first flush threshold in any creek (Bertrand-Krajewski et al., 1998). During all

conditions and in all creeks combined, TSS showed weak correlation with EC and ENT

(Figure 2.8). This was also true for individual creeks, with the exception of French

(correlation of EC to TSS, R2 = 0.28, p<.01 and ENT to TSS, R2=0.27, p<.01). In all creeks,

sediment bacterial average concentrations ranged from 19-168 and 28-451 MPN/g, for EC

and ENT, respectively (Figure 2.5). Freeman creek had the highest instream sediment

concentrations of EC and ENT, while Gillets had the lowest.

EC to ENT Rainfall Relationship

Page 41: New approaches and technologies for quantifying the impact

24

The relationship between EC and ENT changed dramatically depending on rainfall

conditions. For all creeks, dry weather resulted in weak EC and ENT correlation, while wet

weather (rainfall >1.27 cm in 24 h) resulted in a much stronger correlation (Figure 2.6). Wet

weather also exhibited a nearly 1:1 ratio of EC to ENT. When streams were examined

individually, these relationships were even more strongly correlated. For example, at Cogdel

Creek, the EC to ENT correlation was weak during dry weather (R2 = 0.23, p<.05) and strong

during wet weather (R2 = 0.80, p<.01).

Discussion

Loading Characterization

Although the watershed area surrounding MCBCL is generally rural with low density

residential and light industrial land use, concentrations and loading rates in these four

headwater creeks demonstrate prevalent FIB contamination, and elevated wet weather levels

in comparison to baseflows. Freeman and Gillets Creek subwatersheds have little to no

development and are primarily used for military training exercises, while French is used

primarily as an artillery impact area. Despite the limited potential for human fecal

contamination, total loading of FIB during rainfall events ranged from 109-1012 FIB cells per

storm in each creek, and storm EMCs ranged from 4.65 × 101 to 5.31 × 103 (MPN/100 mL),

and 1.53 × 101 to 4.76 × 103 (MPN/100 mL), for EC and ENT, respectively. These FIB loads

are similar to those reported by Line et al. in small rural creeks of eastern NC (2008).

However, total loads and EMCs of FIB in our NC tidal creeks were several orders of

magnitude lower than those reported in studies of urban watersheds in California (Surbeck et

al., 2006; Stein et al., 2007). These results support previous findings of greater concentrations

Page 42: New approaches and technologies for quantifying the impact

25

of FIB in developed and more impervious watersheds (Mallin et al., 2000; Schoonover and

Lockaby, 2006), but also emphasize the magnitude of these differences between urban vs.

primarily undeveloped watersheds.

Overall, these tidal creek headwaters are supplying large FIB contaminant loads to

downstream shellfishing areas. With a shellfish water quality standard for fecal coliforms

(FC) of 14 MPN/100 mL in NC, dilution into larger volume shellfishing areas is inadequate

in buffering delivery of upstream contaminant loads unless significant tidal flushing is

present, which is not the case in these creeks. For instance, average EC load in Freeman

Creek during storm events was 1.3 × 1011 MPN/storm. Assuming no decay, tidal exchange,

or inputs from downstream or in-creek/local sources before or at the shellfishing beds, and

using the average volume (C2HM Hill, 2000) of the entire creek to estimate a rough,

maximum, completely-mixed concentration of FC in downstream shellfishing areas, FC

concentrations would exceed the FC standard by nearly 4.5 times. However, these FC water

quality standards are problematic, due to the potential for natural sources (i.e. birds, plants,

soils and sediments, etc.) to contribute significant concentrations of FC and other FIB to

source waters (Whitlock et al., 2002; Steets and Holden, 2003; Anderson et al., 2005).

Future changes to shellfish standards (ie. higher FC standards, a reference system approach,

alternative indicators, etc.) might improve estimations of true impairment and differentiation

between natural sources and anthropogenic fecal contamination.

First Flush and Pollutographs

Characterization of the patterns, concentrations, and loading trends over a range of

storm sizes and patterns is an important first step toward understanding wet weather

Page 43: New approaches and technologies for quantifying the impact

26

stormwater inputs, receiving water conditions, and appropriate remediation strategies. For

instance, if first flush patterns of FIB delivery to high priority shellfishing waters are

observed, with 80% (a mass:volume ratio typically accepted for characterizing stormwater

pollutant first flush patterns) of fecal coliforms delivered during the initial 30% of storm

runoff volume, then specific types and sizes of best management practices (BMPs) such as

retention basins may be more appropriate for stormwater treatment. There has been no

previous attempt to determine FIB loading characteristics in headwater tidal creeks of NC.

Our data do not support a first flush scenario for either EC or ENT concentrations for

combined creek loading data (Figure 2.4). Mass loading to individual creeks averaged as

much as 62% in the first 30% of volume but never achieved 80% loading. Individual storms

that approached first flush were generally short (<12 h) in duration, and intense (rainfall > 3

cm), but this was not always the case. Examination of patterns of loading of FIB in creeks in

central NC and southern California revealed similar findings (Surbeck et al., 2006; Krometis

et al., 2007). Interestingly, this is contrary to first flush loading analyses that have been

documented for other contaminants such as heavy metals, hydrocarbons, and nutrients (Line

et al., 1997; Lee et al., 2002; Stein et al., 2006; Tiefenthaler et al., 2008; Barco et al., 2008),

though these findings are often in more developed and/or highly impervious watersheds

(Sansalone and Cristina, 2004). We conclude that FIB in the examined tidal creeks

headwaters did not follow a first flush due to highly vegetated land cover, a low percentage

of impervious surfaces, and highly permeable soil conditions, which moderate overland flow,

and initial flushing of FIB. Our findings also indicate that FIB remediation efforts must

address loading throughout the duration of a storm in order to achieve successful water

quality improvement.

Page 44: New approaches and technologies for quantifying the impact

27

Pollutographs of EC, ENT, and TSS were developed to determine how contaminant

concentrations occurred in comparison to creek discharge (streamflow) over time. Results

indicate FIB concentration increase and decrease with similar streamflow changes. For

instance, one representative storm plotted for each creek indicates increases in EC, ENT, and

TSS with the rising limb of the hydrograph, higher levels at streamflow peaks, and decline

with the falling limb (Figure 2.3). Average TSS concentrations showed more of an initial

increase with the rising limb (Figure 2.2), possibly due to more of a first flush type of

loading, and/or streambed armoring (i.e., easy to erode sediment is scoured during the early

part of the storm, the streambed is then “armored” and harder to erode despite similar shear

stresses) (Chin et al., 1994). These pollutograph FIB patterns were similar to other studies in

several different watersheds (Surbeck et al., 2006; Krometis et al., 2007; Stein et al., 2007).

Again, these similar results across different watersheds and geographic areas support the

need for entire storm mitigation strategies for reductions of FIB inputs to receiving waters. In

addition, variability between sample points emphasizes the need for intensive multi-point

sampling throughout storms, to acquire accurate FIB contamination estimates. Current

single-grab sample methods for sampling are an insufficient means to accurately characterize

FIB during storm events due to this intra-storm variability.

Base versus Storm Loading

Load of FIB and EMCs varied among the four sampled creeks; however, total load of

FIB and EMCs for all creeks combined during storm events always exceeded baseflow load

during equivalent antecedent baseflow periods. Gillets Creek exhibited the smallest

excursion between total load of ENT during storm events versus baseflow, most likely due to

Page 45: New approaches and technologies for quantifying the impact

28

high baseflow concentrations. It is unclear what is causing these high baseflow levels of FIB.

Overall, mean storm loads were equivalent to approximately 1.5 - 4.5 months of baseflow

loads. A similar comparison was made in central NC where baseflow to storm load

equivalencies were greater than 1 year (Krometis et al., 2007). However, these streams had

much greater discharge (average flow 0.5 m3/s vs. less than 0.12 m3/s in this study) and

occurred in watersheds characterized by higher levels of impervious surface coverage and

residential use. Despite a smaller baseflow loading component, baseflows contribute

substantial fractions of annual FIB loads and contribute to closures of downstream

shellfishing areas, especially in creeks such as Gillets, with high ambient (i.e. non-storm

related) FIB concentrations. This would be especially true during years with few rainfall

events leading to a greater portion of loading during dry periods. However, in these tidal

creeks, and in geographic regions with large scale rainfall events (such as tropical storms and

nor’easters), storms can contribute concentrations of FIB many orders of magnitude larger

than equivalent baseflow periods.

Rainfall Metric Relationships

Storm load of FIB was most often correlated to the total amount of rainfall, but had a

weak relationship to other rainfall metrics. Total rainfall was strongly correlated with both

EC and ENT load, and EC and ENT EMCs. Similar FIB concentration increases due to

rainfall have been shown in other eastern NC creeks and estuaries (Mallin et al., 2001;

Coulliette and Noble, 2008; Fries et al., 2008; Line et al., 2008). However, duration and

antecedent rainfall showed varying degrees of correlation with FIB loads with all creek data

combined (Table 2.3), and little correlation with individual creeks, with the exception of

Page 46: New approaches and technologies for quantifying the impact

29

Cogdel (Table 2.3). These results demonstrate that rainfall could be a useful predictor, but

also highlight the wide range of variability in FIB in relation to rainfall (storm, ambient, and

antecedent) rainfall scenarios.

The lack of correlation between FIB loading and rainfall metrics could be due to

subwatershed land use and soil conditions. Three of the four subwatersheds under

examination are undeveloped and forested, and have soils classified as moderate to rapidly

permeable, comprised mostly of sands (CH2M Hill, 1998; CH2M Hill, 2000; USDA, 2010).

Terrestrial stormwater runoff generally occurs after soil saturation and reduced infiltration of

rainfall on the land surface (Tyrell and Quinton, 2003). Forested and densely vegetated

landscapes in combination with high soil permeability might increase the storm magnitude

(intensity, duration) necessary to cause significant differences in FIB loading. For instance,

Cogdel Creek, the only developed subwatershed (~31% impervious surface), had the only

significant correlation between total ENT loading and antecedent rainfall. One explanation

for these results is that land use and soil type are moderating storm duration and antecedent

rainfall effects on FIB loading, and tidal creeks in more developed areas with substantial

impervious surfaces and/or in areas with low permeable soils may be more prone to FIB

contamination from runoff.

FIB Runoff Relationships

Terrestrial Runoff

In the four headwater creeks examined, there was a close association among FIB load

and EMCs, rainfall, and streamflow, which is indicative of bacterial loading from terrestrial

runoff or instream resuspension/erosion. FIB loads were strongly associated with streamflow

Page 47: New approaches and technologies for quantifying the impact

30

in all creeks, with the exception of ENT in Gillets (Figure 2.1). Others have pointed out that

the correlation between streamflow and FIB is often related to rainfall and runoff transport of

terrestrial fecal pollutants (Nagels et al., 2002). However, TSS was moderately related to

streamflow and weakly related to FIB. Similar results were found in a study conducted in

nearby watershed examining fecal coliforms (Line et al., 2008). This is surprising,

considering instream sediment resuspension typically increases greatly via erosion with

increased flow velocity and associated shear stress (Steets and Holden, 2003). However, this

does not seem to be the case in these watersheds, possibly due to their heavily vegetated and

minimal gradient landscapes, which likely increases rainfall retention and infiltration, and

limits soil scouring from runoff during storm events.

Instream Sediments

Concentrations of average sediment FIB ranged from 19-451 FIB (MPN/g).

These sediment FIB concentrations are substantially less than concentrations founds in

nearby heavy developed (72% residential land use) tidal creek system (Toothman et al.,

2009), indicating that tidal creek sediment FIB concentrations may be directly related to land

use characteristics. Based on this limited data, sediment resuspension is responsible for only

a small contribution (ie. only 9% of EC in Gillets Creek during one representative storm) of

the total FIB load during wet weather (see Appendix A for sediment calculation used to

determine contribution of FIB to water column). After consideration of continually elevated

FIB concentration throughout storms and after TSS decline, and high FIB concentrations

(109-1012 per storm), exceeding what could solely come from instream sources even with

Page 48: New approaches and technologies for quantifying the impact

31

rampant resuspension, we conclude that FIB concentrations are primarily due to terrestrial

source runoff.

E. coli to Enterococcus spp. Ratios

EC and ENT concentrations were strongly coupled during storm events but not during

baseflow. Strong coupling of EC and ENT concentrations are observed in fresh fecal

material (human and animal), but ENT and EC decay at different rates causing decoupling of

the populations over time, especially on terrestrial landscapes (Kibbey et al., 1978; Crane and

Moore, 1986). ENT tends to persist longer than EC in deposited fecal matter in terrestrial

settings (Ostrolenk et al., 1947; Van Donsel et al., 1967). Given the significant correlation

between ENT and EC during storm events, it appears that FIB contributed to the headwater

tidal creeks are from a relatively fresh, potentially runoff associated source, while FIB in the

water during baseflow periods are decaying at different rates due to a range of environmental

conditions. It is also possible, given the concentrations of FIB observed during baseflow

periods in specific creeks that there is growth, or at least long term persistence, of ENT and

EC in these systems. Previous research has examined fecal coliform (E. coli) to ENT ratios

as a means to assess the likely sources of fecal contamination (Hai and Handao, 1982).

Although there are a myriad factors associated with these ratios, the ratio of EC/ENT during

storms was consistently close to 1. This could suggest a predominant source of contamination

from warm-blooded animals (excluding humans). It is clear that there are strong differences

in the stormwater and baseflow ENT and EC dynamics in these headwater tidal creeks. It

would be useful to use molecular analysis, including source specific targets, and community

analysis to further elucidate the meaning of these interesting relationships.

Page 49: New approaches and technologies for quantifying the impact

32

Conclusions

Understanding the percent load FIB per flow period during a storm is important in

assisting water quality managers to develop BMPs and plan for shellfish harvesting closures

(presumptive closures). The results of this study, when compared to more developed and

impervious watersheds, support previous findings of changing concentrations of FIB based

on watershed land use, but these results also emphasize the magnitude of these differences

between urban vs. primarily undeveloped watersheds. This is especially true in geographic

regions with large scale rainfall events (such as tropical storms and nor’easters), where

storms can contribute concentrations of FIB much larger than equivalent baseflow periods.

Due to intra-storm variability between points throughout the storm, a multi-sampling

approach through entire storm duration is necessary for more accurate estimation of fecal

contamination. Current single-grab sample methods (ie. one sample per storm) are

insufficient to accurately characterize FIB due to this intra-storm variability. Since neither

ENT nor EC followed first flush loading patterns, and are primarily contributed to the

estuarine system during storm events, mitigation strategies for FIB reduction will need to

consider full storm duration. In addition, watershed characteristics (land cover and soil types)

could be responsible for modification of loading patterns of FIB and TSS. Finally, FC

shellfish standards may need to be reassessed due to elevated concentrations from natural

sources, which is likely the case in these undeveloped watersheds.

Acknowledgements

This research was supported by the Defense Coastal/Estuarine Research Program (DCERP),

funded by the Strategic Environmental Research and Development Program (SERDP).

Page 50: New approaches and technologies for quantifying the impact

33

Views, opinions, and/or findings contained in this report are those of the authors and should

not be construed as an official U.S. Department of Defense position or decision unless so

designated by other official documentation. Additional materials support was provided by

IDEXX industries.

Page 51: New approaches and technologies for quantifying the impact

34

1 T. Minter Unpublished Data 2 USDA, WebSoil Survey, 2010

Table 2.1. Watershed size, predominant soil and land use types for each tidal creek

Watershed

1Watershed

Size (m2)

2Predominant Sediment

Types

1Predominant Land

Use Types

Land Use

Percentage

Cogdel 7620939 BaB Baymeade fine sand

Predominantly Pine

Forest 33.5

BmB Baymeade-Urban

Business or Commercial

Area 31.3

MaC Marvyn loamy fine

sand

Mixed Pine and

Hardwood Forest 10.3

Mk Muckalee loam

Bottomland Hardwood

Forest 9.3

On Onslow loamy fine

sand

Forest Plantations Under

10 Years 6.1

To Torhunta fine sandy

loam Other 9.7

French 5872377 Ln Leon fine sand Bare Ground 54.3

KuB Kureb fine sand Shrub or Scrub 23.7

Mu Murville fine sand

Predominantly Pine

Forest 19.5

BaB Baymeade fine sand Other 2.4

Freeman 6251037 BaB Baymeade fine sand

Mixed Pine and

Hardwood Forest 42.5

Mk Muckalee loam

Predominantly Pine

Forest 46.4

KuB Kureb fine sand Shrub or Scrub 6.7

AnB Alpin fine sand Bare Ground 3.7

Mu Murville fine sand Other 0.7

Gillets 4199947 WaB Wando fine sand

Predominantly Pine

Forest 39.2

Mk Muckalee loam

Mixed Pine and

Hardwood Forest 32.4

Ln Leon fine sand

Bottomland Hardwood

Forest 11.4

To Torhunta fine sandy

loam

Forest Plantations Under

10 Years 10.5

BaB Baymeade fine sand Other 6.5

Page 52: New approaches and technologies for quantifying the impact

35

Creek

Loading Difference Between

Storm and Baseflow

Equivalent Baseflow Days to

Equal Average Storm Load

EC ENT EC ENT

Cogdel 22 22 121 121

French 25 20 92 73

Gillets 29 14 97 48

Freeman 30 37 110 134

Table 2.2. Loading differences between storm and baseflow (times greater), and days

required during dry weather to equal total E. coli and Enterococcus spp. loading during an

average storm for each tidal creek

Page 53: New approaches and technologies for quantifying the impact

36

Creeks Loading

Total

Rainfall

Rainfall

Duration

Antecedent

Rainfall

Average Storm Event

Mean Concentrations

Average Baseflow Event

Mean Concentrations All Creeks

Combined

Total

Discharge .294 .313 -0.047

Cogdel E

. co

li T

ota

l

Lo

ad

ing

0.62 -.134 0.294

E. co

li

7.51 × 102 *

3.58 × 102 *

French .810* -.393 -0.464 1.58 × 103

** 1.44 × 102 **

Gillets 0.261 -.067 -0.506 1.17 × 10

3 **

1.92 × 10

2 **

Freeman 0.143 .687 -0.317 5.93 × 102 5.72 × 10

1

All Creeks

Combined .534** .193 -0.145

1.11 × 103

**

2.03 × 102

**

Cogdel

En

tero

co

ccu

s sp

p.

To

tal

Lo

ad

ing

.778** .109 .657*

En

tero

co

ccu

s sp

p.

9.08 × 102

*

1.24 × 102

*

French -0.286 .071 0.146 5.02 × 102

** 7.11 × 101

**

Gillets 0.127 .128 -0.15 1.18 × 102

8.35 × 101

Freeman 0.429 .687 -0.049 9.8 × 101 ** 8.25 × 10

0 **

All Creeks

Combined .537** .393* 0.118

3.87 × 102

**

7.91 × 101

**

Table 2.3. Spearman rank correlations (R2) for all creeks combined and individual tidal creeks for loads versus rainfall metrics

and t-test results for average event mean concentrations for EC and ENT during both storm and equivalent baseflow periods.

** Significant at p<0.01 * Significant at p<0.5

36

Page 54: New approaches and technologies for quantifying the impact

37

Figure 2.1. Streamflow (m3/s) versus log fecal indicator bacteria concentration (Most Probable Number (MPN)/100 mL) and Total

Suspended Solids (TSS) for creeks (A) Cogdel (B) French (C) Gillets (D) Freeman, during all flow conditions. ** Significant at

p<0.01

37

Page 55: New approaches and technologies for quantifying the impact

38

Figure 2.2. Box plots for all creeks across all storms of mean (A) E. coli (B) Enterococcus spp. (C) flow (D) and total

suspended solids

38

Page 56: New approaches and technologies for quantifying the impact

39

Figure 2.3. Pollutograph representation of intra-storm variation in log concentrations of E. coli (EC) and Enterococcus spp.

(ENT) (MPN/100 mL), and total suspended solids (TSS) (mg/L) during one representative storm in the four headwater tidal

creeks

39

Page 57: New approaches and technologies for quantifying the impact

40

Figure 2.4. Average mass to volume (M/V) plots, used to determine fecal indicator bacteria load per volume of creek

discharge, for (A) E. coli and (B) Enterococcus spp. for each tidal creek

40

Page 58: New approaches and technologies for quantifying the impact

41

0

100

200

300

400

500

600

700

800

Cogdel French Gillets Freeman

Creek Sediments

Instr

eam

Sedim

ent

FIB

(M

PN

/g)

E. coli

Enterococcus spp.

Figure 2.5. Mean E. coli and Enterococcus spp. concentrations (MPN/g) for instream

sediment during summer baseflow conditions in each tidal creek. Error bars represent

standard deviation

Page 59: New approaches and technologies for quantifying the impact

42

y = 0.42x + 0.85

R2 = 0.12**

y = 0.82x - 0.05

R2 = 0.56**

0.5

1

1.5

2

2.5

3

3.5

4

4.5

0.5 1 1.5 2 2.5 3 3.5 4

Log E. coli (MPN/100mL)

Log E

nte

rococcus

spp.

(MP

N/1

00m

L)

Stormflow

Baseflow

Figure 2.6. Relationship between log concentrations (MPN/100 mL) of E. coli and

Enterococcus spp. during wet weather (rainfall > 1.27 cm) stormflow samples, and dry

weather (rainfall = 0 cm) baseflow samples. ** Significant at p<0.01

Page 60: New approaches and technologies for quantifying the impact

43

Figure 2.7. Study site including tidal creeks and map of Marine Corps Base Camp Lejeune

Page 61: New approaches and technologies for quantifying the impact

44

R2 = 0.18**

-1

0

1

2

3

4

0 1 2 3 4 5

Log E. coli (MPN/100mL)

Log T

SS

(m

g/L

)

R2 = 0.16**

-1

-0.5

0

0.5

1

1.5

2

2.5

3

3.5

0 1 2 3 4 5

Log Enterococcus spp. (MPN/100mL)

Log T

SS

(m

g/L

)

A.

B.

Figure 2.8. (A) E. coli and total suspended solids (log mg/L) during all conditions (B) and

Enterococcus spp. and total suspended solids (log mg/L) during all conditions. ** Significant

at p<0.01

Page 62: New approaches and technologies for quantifying the impact

45

REFERENCES

Anderson, K.L., Whitlock, J.E., Harwood, V.J. (2005). Persistence and differential survival of fecal indicator bacteria in subtropical waters and sediment. Applied and Environmental

Microbiology, 71(6), 3041-3048.

APHA- American Public Health Association (1998). Standard method for the examination of water and wastewater. 20th Edition. Washington, D.C.

Barco, J., Papiri, S., Stentstrom, M.K. (2008). First flush in a combined sewer system.

Chemosphere, 71, 827-833.

Bertrand-Krajewski, J., Chebbo, G., Saget, A. (1998). Distribution of pollutant mass vs. volume in stormwater discharges and the first flush phenomenon. Water Research, 32(8),

2341-2356.

CH2M Hill (2008). Long term watershed management plan for Cogdel’s Creek; C2HM Hill, Inc., Prepared for US Army Corps of Engineers. Contract No. DACA01-96-D-0028.

CH2M Hill (2000). Freeman Creek evaluation and assessment; C2HM Hill, Inc., Prepared for US Army Corps of Engineers, Contract No. DACA01-96-0028.

Chin, C.O., Melville, B.W., Raudkivi, A.J. (1994). Streambed Armoring. Journal of

Hydraulic Engineering, 120(8), 899-918.

Coulliette, A.D., Noble, R.T. (2008). Impacts of rainfall on the water quality of the Newport River Estuary (Eastern North Carolina, USA). Journal of Water and Health, 6(4), 473-482.

Crane, S.R., Moore, J.A. (1986). Modeling enteric bacterial die-off: a review. Water,

Air, and Soil Pollution, 27, 411-439.

DiDonato, G.T., Stewart, J.R., Sanger, D.M. Robinson, B.J., Thompson, B.C., Holland, A.F., Van Dolah, R.F. (2009). Effects of changing land use on the microbial water quality of

tidal creeks. Marine Pollution Bulletin, 58(1), 97-106.

Page 63: New approaches and technologies for quantifying the impact

46

Ensign, S.H., McMillan, S.K., Thompson, S.P., Piehler, M.F. (2006). Nitrogen and phosphorus attenuation within the stream network of a coastal, agricultural watershed.

Journal of Environmental Quality, 35(4), 1237-1247.

Fries, S.J., Characklis, G.W., Noble, R.T. (2006). Attachment of fecal indicator bacterial to particles in the Neuse River Estuary, N.C. Journal of Environmental Engineering, 132(10),

1338-1345.

Fries, S.J., Characklis, G.W., Noble, R.T. (2008). Sediment–water exchange of Vibrio sp. and fecal indicator bacteria: Implications for persistence and transport in the Neuse River

Estuary, North Carolina, USA. Water Research, 42(4-5), 941-950.

Geldreich, E.E., Best, L.C., Kenner, B.A., Van Donsel, D.J. (1968). The bacteriological aspects of stormwater pollution. Journal of Water Pollution Control Federation, 40(11),

1861-1872.

Gupta, K., Saul, A.J. (1996). Specific relationships for the first flush load in combined sewer flows. Water Research, 20(5), 1244-1252.

Hai, L., Handao, C. (1982). Significance of fecal coliform and fecal streptococcus in water pollution monitoring. Acta Academiae Medicinae Wuhan, 2(4), 251-253.

Hyer, K.E., Moyer, D.L. (2004). Enhancing fecal coliform total maximum daily load

models though bacterial source tracking. Journal of American Water Resources Association, 40(6), 1511-1526.

Kayhanian, M., Stenstrom, M.K. (2005). Mass loading of first flush pollutants with treatment

strategy simulations. Transportation Research Record: Journal of the Transportation Research Board, 1904, 133-143.

Kibbey, H.J., Hagedorn, C., McCoy, E.L. (1978). Use of fecal streptococci as indicators of

pollution in soil. Applied and Environmental Microbiology, 35(4), 711-717.

Krometis, L.H., Characklis, G.W., Simmons, O.D., Mackenzie, J.D., Likirdopulos, C.A., Sobsey, M.D. (2007). Intra-storm variability in microbial partitioning and microbial loading

rates. Water Research, 41(2), 506-516.

Page 64: New approaches and technologies for quantifying the impact

47

Lee, J.H., Bang, K.W., Ketchum, L.H., Choe, J.S., Yu, M.J. (2002). First flush analysis of urban storm runoff. Science of the Total Environment, 293(1-3), 163-175.

Line, D.E., Wu, J., Arnold, J.A., Jennings, G.D., Rubin, A.R., (1997). Water quality of first flush runoff from 20 industrial sites. Water Environment Research, 69(3), 305-310.

Line, D.E., White, N.M., Kirby-Smith, W.W., Potts, J.D. (2008). Fecal coliform export from

four coastal North Carolina Areas. Journal of American Water Resources Association, 44(3), 606-617.

Lipp, E.K., Kurz, R., Vincent, R., Rodriguez-Palacios, C., Farrach, S.R., Rose, J.B. (2001).

The effects of seasonal variability and weather on microbial fecal pollution and enteric pathogens in a subtropical estuary. Estuaries, 42(4), 286-293.

Mallin, M.A., Williams, K.E., Esham, E.C., Lowe, R.P. (2000). Effect of human

development on bacteriological water quality in coastal watersheds. Ecological Applications, 10(4), 1047-1056.

Mallin, M.A., Ensign, S.H., McIver, M.R. Shank, G.C., Fowler, P.K. (2001). Demographic,

landscape, and meteorological factors controlling the microbial pollution of coastal waters. Hydrobiologia, 460(1-3), 185-193.

Mallin, M.A., Lewitus, A.J. (2004). The importance of tidal creek ecosystems. Journal of

Experimental Marine Biology and Ecology, 298(2), 145-149.

Mallin, M.A., McIver, M.R., Wells, H.A., Parsons, D.C., Johnson, V.L. (2005). Reversal of eutrophication following sewage treatment upgrades in the New River Estuary, NC.

Estuaries, 28(5), 750-760.

Nagels, J.W., Davies-Colley, R.J., Donnison, A.M., Muirhead, R.W. (2002). Faecal contamination over flood events in a pastoral agricultural stream in New Zealand. Water

Science and Technology, 45(12), 45-52.

Noble, R.T., Fries, J.S. (2005). Estuarine sediment beds as a reservoir for human pathogens: monitoring transport of populations of Enterococci and Vibrio sp. in the Neuse River

Estuary. Report UNC-WRRI-2005-70214. Water Resources Research Institute of the University of North Carolina, Raleigh, NC, 34pp.

Page 65: New approaches and technologies for quantifying the impact

48

NCDWQ, North Carolina Division of Water Quality, (2007). White Oak River Basin

Wide Water Quality Plan, Planning Section.

Ostrolenk, M., Kramer, N., Cleverdon, R.C. (1947). Comparative studies of Enterococci and Escherichia coli as indices of pollution. Journal of Bacteriology, 53(2), 197-203.

Reeves, R.L., Grant, S.B., Mrse, R.D., Copil-Oancea, C.M., Sanders, B.F., Boehm, A.B. (2004). Scaling and management of fecal indicator bacteria in runoff from a coastal urban watershed in Southern California. Environmental Science and Technology, 38(9), 2637-

2648.

Sansalone, J.J., Cristina, C.M. (2004). First flush concepts for suspended and dissolved solids in small impervious watersheds. Journal of Environmental Engineering, 130(11), 1301-1314.

Schoonover, J.E., Lockaby, B.G. (2006). Land cover impacts on stream nutrients and fecal coliform in the lower piedmont of West Georgia. Journal of Hydrology, 331(3-4), 371-382.

Shehane, S.D., Harwood, V.J., Whitlock, J.E., Rose, J.B. (2005). The influence of rainfall on

the incidence of microbial faecal indicators and the dominant sources of faecal pollution in the Florida River. Journal of Applied Microbiology, 98(5), 1127-1136.

Solo-Gabrielle, H.M., Wolfert, M.A., Desmarais, T.R., Palmer C.J. (2000). Sources of

Escherichia coli in a coastal subtropical environment. Applied and Environmental Microbiology, 66(1), 230-237.

Steets, B.M., Holden, P.A. (2003). A mechanistic model of runoff-associated fecal coliform

fate and transport through a coastal lagoon. Water Research, 37, 589-608.

Stein, E.D., Tiefenthaler, L.L., Schiff, K.C. (2006). Watershed-based sources of polycyclic aromatic hydrocarbons in urban stormwater. Environmental Toxicology and Chemistry, 25,

373-385.

Stein, E.D., Tiefenthaler, L.L., Schiff, K.C. (2007). Sources, patterns and mechanisms of storm water pollutant loading from watersheds and land uses of the greater Los Angeles area,

California, USA. Southern California Coastal Water Research Project, Technical Report 510.

Page 66: New approaches and technologies for quantifying the impact

49

Stewart, J.R., Santo Domingo, J.W., Wade, T.J. (2007). “Fecal Pollution, Public Health, and

Microbial Source Tracking.” In Microbial Source Tracking; Santo Domingo, J.W. Sadowsky, M.J. Eds.; ASM Press: Washington, DC, Chapter 1.

Surbeck, C.Q., Jiang, S.C., Ahn, J.H., Grant, S.B. (2006). Flow fingerprinting fecal pollution

and suspended solids in stormwater runoff from an urban coastal watershed. Environmental Science and Technology, 40(14), 4435-4441.

Tiefenthaler, L.L., Stein, E.D., Schiff, K.C. (2008). Watershed and land use-based sources of

trace metals in urban stormwater. Environmental Toxicology and Chemistry, 27, 277-287.

Tyrell, S.F., Quinton, J.N. (2003). Overland flow transport of pathogens from agricultural land receiving faecal wastes. Journal of Applied Microbiology, 94(s1), 87-93.

Toothman, B.R., Cahoon, L.B., Mallin, M.A. (2009). Phosphorus and carbohydrate limitation of fecal coliform and fecal Enterococcus within tidal creek sediments. Hydrobiologia, 636, 401-412.

USEPA, United States Environmental Protection Agency, (1987). Water Quality Act of 1987. Section 303(d).

USDA- United States Department of Agriculture (2010). National Resource Conservation

Service, Web Soil Survey. Accessed 6/15/2010. http://websoilsurvey.nrcs.usda.gov/app/WebSoilSurvey.aspx

Van Donsel, D.J., Geldreich, E.E., Clarke, N.A. (1967). Seasonal variations in survival of

indicator bacteria in soil and their contribution to storm-water pollution. Applied Microbiology, 15(6), 1362-1370.

Weiskel, P.K., Howes, B.L., Heufelder, G.R. (1996). Coliform contamination of a coastal

embayment: sources and transport pathways. Environmental Science and Technology, 30(6), 1872-1881.

Whitlock, J.E., Jones, D.T., Harwood, V.J. (2002). Identification of the sources of fecal

coliforms in an urban watershed using antibiotic resistance analysis. Water Research, 36(17), 4273-4282.

Page 67: New approaches and technologies for quantifying the impact

50

Wilkinson, J.A., Jenkins, A., Wyer, M., Kay, D. (1995). Modeling faecal coliform dynamics in streams and rivers. Water Research, 29(3), 847-855.

Page 68: New approaches and technologies for quantifying the impact

Chapter 3

Quantifying fecal contaminant sources in North Carolina tidal creeks utilizing three Bacteroides spp. assays, conventional fecal indicators, and optical brighteners

Introduction

Fecal contamination of estuarine and coastal waters is a growing concern. Fecal

indicator bacteria (FIB) including fecal coliforms (FC), Escherichia coli (E. coli, EC, the

dominant subset of FC), and Enterococcus spp. (ENT), are conventional indicators used to

quantify fecal contamination in recreational and shellfish harvesting waters. However, these

conventional FIB cannot typically be used to differentiate between human or animal fecal

contamination (Scott et al., 2002; Noble et al., 2003; Meays et al., 2004), and they may

persist and potentially regrow in tropical and sub-tropical water environments (Solo

Gabrielle et al., 2000; Byappanahalli et al., 2003).

In recent years a “toolbox approach” has been lauded for water quality management,

i.e. the combination of conventional FIB measurements, with the use of rapid, molecular

quantification of alternative source-specific fecal DNA markers. Human specific fecal assays

have been widely developed in last decade, and include Bacteroides spp. (Fiksdal et al.,

1985; Carson et al., 2005; Layton et al., 2006; Savichtcheva et al., 2007; Ahmed et al., 2009)

Methanobrevibacter smithii (Johnston et al., 2010), human polyomavirus (McQuaig et al.,

2006), and pepper mild mottle virus (Rosario et al., 2009). Bacteroides spp. are some of the

more promising of the human specific alternative indicators, due to their high concentration

in the human gut (Drasar, 2003), potentially limited persistence and regrowth in the

Page 69: New approaches and technologies for quantifying the impact

52

environment (Kreader, 1998; Dick and Field, 2004), and evidence for correlation between

Bacteroides spp. concentrations and risk to human health (Wade et al., 2006; Savichtcheva et

al., 2007; Wade et al., 2008) Molecular assays have been developed for Bacteroides spp.

detection, each targeting a different portion of the 16S rRNA gene. For instance, a Fecal

Bacteroides spp. (Converse et al., 2009), human specific (BacHum) Bacteroides (Kildare et

al., 2007), and human fecal contamination specific (HF183) Bacteroides (Bernhard and

Field, 2000) have all been developed targeting Bacteroides spp. as an alternative fecal

indicator. These three Bacteroides spp. assays have been used individually to determine

potential human contamination. However, there is little published comparative data between

these assays and a lack of validation in complex tidal creek waters.

Although the toolbox approach has been suggested for use in recreational beach

waters, the approach must be used carefully in tidal creeks systems. There has been limited

testing of Bacteroides spp. assays in tidal creeks, although tidal creeks are vitally important

for shellfish harvesting and recreational use along the eastern and Gulf coasts. Tidal creeks

in North Carolina (NC) are prone to elevated humic concentrations (termed blackwater

creeks) (Ensign and Mallin, 2001), increased total suspended sediment loads during rainfall

induced runoff events (Stumpf et al., 2010), and tidally influenced flow characteristics,

differentiating them from other fresh and coastal waters (Mallin and Lewitus, 2004).

Alternative indicators and sample processing may be affected by these complex tidal creek

water matrices, and it is important that alternative indicators and molecular assays b e

carefully applied as water quality monitoring tools. In addition, many tidal creek systems and

watersheds remain devoid of development given their soil characteristics, often being more

prone to contamination stemming from upstream sources. This separates them from other

Page 70: New approaches and technologies for quantifying the impact

53

coastal systems in that mitigation scenarios will need to include a wide array of stakeholders

in order to be successful.

Chemical indicator methods have also been used to identify household wastewater in

receiving waters. Chemical indicators of anthropogenic inputs into receiving waters include

coprostanol (Roser and Ashbolt, 2007), caffeine (Peeler et al., 2006), and optical brighteners

(OB) (Cao et al., 2009). OB are particularly appealing because their measurement is cost

effective and relatively quick (Hartel et al., 2007; Dickerson et al., 2007; Ahmed et al.,

2008). Recently a method for improved OB analysis was developed to deal with natural

background fluorescence (Cao et al., 2009), a previous problem for analysis of OB in

environmental water samples. However, this method was has not been tested in tidal creek

waters with naturally high background fluorescence due to high concentrations of dissolved

organic matter.

Because a single indicator may not give a robust assessment of fecal contamination in

receiving waters, many researchers have recommended a multi-tiered approach to water

quality testing (Boehm et al., 2003; Noble et al., 2006; Field and Samadpour, 2007). Our

goal was to rigorously quantify the efficacy of indicators using a multi-tiered/toolbox

approach for assessment of water quality in NC tidal creeks affected by different land uses.

Specifically, we examined conventional FIB (EC and ENT), three Bacteroides spp. based

targets, and optical brighteners in four tidal creek headwaters over a range of loading

conditions. QPCR-based Bacteroides spp. concentrations were analyzed in relation to one

another and conventional FIB concentrations in an attempt to understand sources of fecal

contamination in the tidal creeks. This robust multi-tiered approach can provide a more

accurate determination of fecal contaminant source in tidal creek and coastal waters, and

Page 71: New approaches and technologies for quantifying the impact

54

provide insight into the practical application of optical brightener and molecular methods

(measuring Bacteroides spp.) in a tidal creek environment.

Methods

Study Site and Sample Collection

Tidal creeks were located in the New River watershed, Onslow County, NC. A more

in depth description of these creeks can be found in the methods section of Chapter 2 or

Stumpf et al. (2010). Due to the very limited training in many of MCBCL’s watersheds, this

area is well-suited to the examination of Bacteroides spp. and OB in an undeveloped

landscape. Ten total tidal creeks were sampled throughout this study, but for this chapter,

only four (Freeman, Gillets, Courthouse, and Southwest Creek) were analyzed using QPCR

due to the cost and time associated with this analysis.

Tidal creek headwater surface water samples (n=106) were collected from December

2007 to December 2010 using ISCO automated water samplers (Model 6712, ISCO, Lincoln,

NE) located on the downstream side of road culverts (i.e. drainage pipes) over a range of

sampling conditions (no rain to rainfall events). Samples were collected in 1 L, 5 %

hydrochloric acid washed sample bottles. ISCO bottles were removed within 24 h of

automated sample collection and immediately processed. Flow, rainfall, and water quality

data (dissolved oxygen, temperature, and conductivity) were continuously collected and

logged at 30-min intervals using a YSI sonde, Model 600XL. The ISCO auto-samplers were

programmed to sample throughout each storm event and were enabled when flow exceeded

predetermined threshold velocities as described by Stumpf et al. (2010). Baseflow and tidal

creek mouth samples were manually collected in sterilized, 1 L bottles, triple-rinsed with

Page 72: New approaches and technologies for quantifying the impact

55

sample water, and then placed on ice and processed within 6 h of collection. All tidal creek

samples (n=23) were collected within 2 hours of low tide in order to assess tidal creek fecal

inputs to creek mouth and not inputs carried upstream from the estuary. All tidal creek

samples were collected during ambient (no rainfall) conditions.

Analysis for Chromogenic Substrate Tests

Water samples were processed in the laboratory in duplicate using chromogenic

substrate tests Colilert-18 and Enterolert (IDEXX® Laboratories, Westbrook, ME) at a 1:10

dilution to calculate most probable number (MPN) (Hurley and Roscoe, 1985) of EC and

ENT. Empty ISCO sample bottles were intermittently analyzed to determine if FIB

accumulated during deployment, showing no contamination with FIB. ISCO sample tubing

was flushed before individual samples were collected, and was assumed to be

uncontaminated based on previous research under similar sampling conditions that found

little sample line contamination (less than 4 CFU/100 mL for EC) using similar samplers

(Solo-Gabrielle et al., 2000).

Analysis for Optical Brighteners

Optical brighteners were sampled during 15 different events, for each creek. Samples

for optical brighteners were stored on ice, protected from light, and processed within 6 hours.

Optical brighteners were measured in triplicate according to the methods of Cao et al. (2009).

Samples were poured into 15 mL quartz tubes and stored in the dark until reaching room

temperature (~22o C). A Turner Biosystems AU-10 Fluorometer (Sunnyvale, CA) with a

300-400 nm excitation filter and 436 nm emission filter was calibrated using DI water as a

Page 73: New approaches and technologies for quantifying the impact

56

blank, and 50 µl/L of Tide 2X (Proctor and Gamble, Cincinnati, OH) laundry detergent as a

standard. Samples were placed in 15 mL quartz tubes and measured for fluorescence at 0, 5,

and 10 minute intervals after exposure to fluorescent light using a UVP Blak-Ray® UV

Bench Lamp (Upland, CA) at 365 nm in a ventilated biological cabinet.

Analysis for Bacteroides spp.

Filtration and DNA Extraction

Duplicate one-hundred mL samples were filtered through a 47 mm diameter, 0.45 µm

pore sized Millipore polycarbonate (PC) filter (Bedford, MA) using sterilized filtration

funnels (Pall, East Hills, NY). After complete filtration, PC filters were aseptically removed,

and transferred into sterile 1.5 mL microcentrifuge tubes and stored at -80oC. Prior to DNA

extraction, sample filters were equilibrated to room temperature, and transferred to extraction

tubes for processing using aseptic technique.

DNA extraction was initially performed using a GeneRite DNA-EZ extraction kit

(n=10) (North Brunswick, NJ). After all initial samples were determined to be highly

inhibited (based on relation to the specimen processing control, see next section), subsequent

sample (n=128) DNA was extracted using a QIAGEN DNA Stool Extraction Kit

(Germantown, MD) in an attempt to reduce inhibitory compounds. However, due to

extraction control contamination, these extractions were discarded, and new samples (n=128)

were extracted using a MoBio Power Soil (Carlsbad, CA) DNA extraction kit. MoBio

extracted samples were extracted according to manufacturer’s instructions with the following

exceptions: 1) the sample PC filter was placed into the screwcap tube and 100 ng of Salmon

testes DNA (Sigma, St. Louis, MO) was added per sample as a specimen processing control

Page 74: New approaches and technologies for quantifying the impact

57

(SPC); 2) sample was bead beaten for 2 minutes using an 8-place or 45 place bead beater

(BioSpec, Bartlesville, OK) instead of vortexing for 10 min. ; 3) sample was eluted with a

final volume of 60 µl of final elution buffer instead of 100 µl. Extracted DNA was stored at

-20o C until analysis.

Extraction using the MoBio kit was both time intensive, requiring 2.5-3 h per 20

environmental samples, and had poor extraction efficiency. Oligonucleotide standard

extraction was tested for the BacHum and HF183 assays and resulted in <1% recovery.

Specimen Processing Control

To determine the amount of target DNA loss during extraction and due to

environmental matrix inhibition, a known amount of DNA (i.e. specimen processing control

(SPC)) was extracted with each sample. Salmon sperm (Oncorhynchus keta) testes DNA was

added to the sample before extraction at a final concentration of 100 ng per 500 mL

following the techniques of Converse et al. (2009) and based on previous work by Haugland

et al. (2005). All QPCR sample reactions with a Ct value 0.5 log units higher than that of the

calibrator (SPC) were considered to be inhibited. Inhibited samples were diluted 1:10 with

sterile water and reanalyzed. A PC filter blank was also run with each extraction to ensure

that cross contamination did not occur during the extraction step. Sterile buffer was extracted

concomitantly with each set of samples and run as a control.

Quantitative Polymerase Chain Reaction (QPCR)

Controls, calibration, primers and probes

Page 75: New approaches and technologies for quantifying the impact

58

Calibration standards for the Fecal Bacteroides spp. assay were prepared according to

the methods of Converse et al. (2009). The B. thetaiotaomicron (ATCC 29148) standard was

grown under anaerobic conditions and quantified using a SYBR Green I staining method

(Noble and Fuhrman 1998). A starting concentration cell standard of 1 x 105 cells was used

for each set of DNA extractions. Cell equivalent (CE) calibration standards for the BacHum

and HF183 assays were prepared from serially diluted oligonucleotide standard stocks

(BioSearch Technologies, Novata, CA). DNA oligonucleotide standard stocks were

quantified using a PharmaSpec 1700 spectrophotometer (Shimadzu Scientific Instruments,

Columbia, MD) and diluted to 1 × 105 copies per extraction. One standard was processed

with each extraction, and serially diluted to achieve a duplicate 4-point standard curve.

The SPC, Fecal Bacteroides spp., and BacHum assays were run using TaqMan

chemistry primers and probes. Twenty-five µl reaction volumes were prepared using 5 µL of

extracted sample DNA, 1 mM each of forward and reverse primers, 0.1 mM of the TaqMan

probe, and OmniMix beads. Cycle threshold (Ct) value was determined automatically by the

Cepheid Smart-Cycler II software, after the threshold was manually adjusting to achieve the

highest cycling efficiency.

QPCR

The SPC (Haugland et al., 2005), Fecal Bacteroides spp. (Converse et al., 2009), and

BacHum (Kildare et al., 2007) assays were run using TaqMan chemistry primers and probes.

Primer and probe sequences can be found in the original studies for each assay. Twenty-five

µl reaction volumes were prepared using 5 µL of extracted sample DNA, 1 mM each of

forward and reverse primers, 0.1 mM of the TaqMan probe, and mastermix in the form of

Page 76: New approaches and technologies for quantifying the impact

59

Cepheid OmniMix HS-50 (Sunnyvale, CA), a lyophilized premix with 1.5 units of TaKaRa

hot start Taq polymerase, 200 mM of dNTPs, 4 mM of MgCl2, and 25 mM HEPES with a pH

of 8. The HF183 assay was analyzed using Eurogentec 2X SYBR® Green I No Rox (San

Diego, CA), a liquid master mix of HotGoldStar DNA polymerase, dNTPs, reaction buffer,

MgCl2, Uracil-N-Glycosylase, SYBR® Green I, and stabilizers. Five µl of sample was added

for each reaction regardless of assay. All reactions were performed on a Cepheid Smart-

Cycler II (Sunnyvale, CA) with reactions conditions described in Table 3.1. Standard curves

with either cells or oligonucleotide standards (Table 3.1) with four serial dilutions were

completed for each assay. Cycle threshold (Ct) value was determined automatically by the

Cepheid Smart-Cycler II software, after the threshold was manually adjusting to achieve the

highest cycling efficiency. Amplification efficiencies were > 90% with R2> 0.98 for all

assay standards.

A sewage control at 4 serial 10-fold dilutions was also run with each HF183 assay to

confirm the assay’s success at determining sewage presence in the sample. A melt curve was

run at the conclusion of the HF183 assay, and considered confirmatory for target only when a

positive Ct and a melt in the range of 75.0 to 77.0 occurred. Sewage standards which were

positive had melt peaks ranging from 75.4 to 76.3. Oligonucleotide standard melt peaks

ranged from 75.2 to 75.7.

QPCR Performance Assessment

Tidal creek samples were strongly inhibited (the SPC assay run with each set of

samples resulted in quantification that was >0.5 log lower than the known concentration of

the control), especially during rainfall events. Both Qiagen and MoBio extraction kits were

Page 77: New approaches and technologies for quantifying the impact

60

utilized to remove inhibitors and further purify DNA. Extraction required 2.5-3 h per 20

environmental samples. Oligonucleotide standard extraction recovery was tested for the

BacHum and HF183 assays and resulted in <1% recovery. All MoBio extracted samples had

no extraction contamination, while Qiagen extractions had frequent extraction control

contamination. For this reason, only MoBio extracted samples were utilized. Samples still

inhibited after extraction were diluted 1:10 (based on SPC results). No samples required

greater than 1:10 dilution after extraction with MoBio Power Soil Extraction Kit. All QPCR

assays had high amplification efficiency ≥92.7% and linear four-point standard curves (R2 ≥

0.993) were generated for all runs.

Calculations and Statistics

Optical Brightener Determination

Optical brighteners were considered positive if the decrease in fluorescence after 5

min of exposure was > 8% and the ratio of fluorescence after 10 min of UV exposure to

fluorescence after 5 min of UV exposure was less than or equal to 1.5, according to the

method of Cao et al. (2009). Samples showing positive results for all three samples were

considered positive, while those that tested negative for one or more of the replicates were

considered negative (Cao et al., 2009).

Quantification of DNA

Duplicate samples were analyzed and compared to a duplicate 4-point standard curve

performed at 10-fold dilutions. A ΔCt method outlined by Pfaffl (2001) and previously used

by Converse et al. (2009) was used to quantify target DNA in each sample. To determine the

Page 78: New approaches and technologies for quantifying the impact

61

number of cells for each target, a ratio based on sample target DNA to DNA in the

calibrators, the ΔCt value, and the amount of target DNA in the calibrator were used. Non

detection (ND) results from the QPCR were assigned value of 5 CE/100 mL for statistical

analysis.

Statistics

All statistics were conducted either using SPSS (Chicago, IL) statistical analysis

software (version 11.1) or Microsoft EXCEL software. Data were tested using the Shapiro

Wilk test for normality and normal Q-Q plots. Data were not normally distributed and were

subsequently analyzed using nonparametric statistical tests. Mean raw data concentrations

were log-transformed for graphical purposes. The Spearman Rho rank correlation coefficient

test was used to determine statistical correlations, the Kruskall Wallis H test was used to

determine significant differences between indicators, and the Mann-Whitney U test was used

to determine significant differences between microbial indicators and other watershed fecal

indicators in a highly developed watershed (Courthouse Creek). Significance was accepted at

α = 0.05.

Results

Conventional Indicators

Conventional FIB concentrations in all tidal creeks were intermittently above single

sample regulatory thresholds (104 ENT/100 mL and 320 EC/100 mL), with 39.6 % and 49.1

% for EC and ENT, respectively, in exceedence (Figure 3.2). EC and ENT mean

concentration for all creeks were 868 MPN/100 mL with a range of 10.1 to 1.33 × 103, and

Page 79: New approaches and technologies for quantifying the impact

62

432 MPN/100 mL with a range of 5.00 to 5.28 × 103 MPN/100 mL, respectively. Individual

creeks varied by percentage of samples that exceeded regulatory standards (Table 3.2). Mean

EC and ENT levels were highest at Southwest Creek (1.93 × 103 and 1.46 × 103 MPN/100

mL, respectively), while the lowest mean concentrations occurred at Freeman Creek (181 and

84.7 MPN/100 mL, respectively). Individual creek correlation between log EC and ENT

concentrations varied, with Courthouse (ρ= 0.364, p<.001) and Gillets (ρ= 0.149, p<.016)

showing significant correlation between indicators, and Southwest (ρ= 0.176, p=0.175) and

Freeman creek (ρ=0.0172, p=.523) having no significant correlation.

Bacteroides Assays

Fecal Bacteroides spp.

Nearly all tidal creek headwater samples tested (95.2 %) were positive for the Fecal

Bacteroides spp. assay with wide variation in quantity of target dependent on creek (Table

3.2). Concentrations ranged from 5 cell equivalents (CE)/100 mL (limit of detection) to 2.17

× 104 CE/100 mL. Fecal Bacteroides spp. mean concentration for all creeks was 1051

CE/100 mL. Courthouse had the highest mean Fecal Bacteroides spp. concentration while

Gillets had the lowest mean concentration (Figure 3.3, Table 3.2).

BacHum

BacHum was less prevalent than Fecal Bacteroides spp. and was detected in 81.1% of

samples. BacHum concentrations in all creeks ranged from 21.3 to 1.27 × 106 CE/100 mL.

BacHum mean concentration for all creeks combined was 1.89 × 104 CE/100 mL, the highest

mean concentration of any indicator (Figure 3.4). Southwest had the highest mean log

Page 80: New approaches and technologies for quantifying the impact

63

BacHum concentrations (3.15 CE/100 mL), while Freeman Creek had the lowest (2.15

CE/100 mL) (Figure 3.3).

HF183

The human-specific HF183 assay rarely had detectable concentrations in headwater

creeks. HF183 was found in only 6.60 % of samples, with only one sample from Freeman

Creek and 6 samples from Southwest Creek positive. However, only twelve samples from

Southwest Creek were tested using the HF183 assay, indicating a 50% detection rate. HF183

average concentration for Southwest creek was 6.88 × 104 CE/100 mL, with a range of 111 to

7.76 × 105. There were no positive samples from Gillets Creek and Courthouse Creek for

this assay (Figure 3.3).

Relationships Between Indicators

There were statistical relationships between indicators in certain cases. For all creek

data combined there were weak (but statistically significant) positive correlations between

Fecal Bacteroides spp. and ENT concentrations (ρ=0.252, p<0.009), between BacHum and

EC (ρ=0.339, p<.001), and between BacHum and HF183 (ρ=0.312, p<0.001) (Table 3.3).

When individual creeks were examined, certain creeks had similar relationships among

indicators. In Gillets Creek, there was a moderate positive correlation between BacHum and

EC (ρ=0.540, p<.001) and Fecal Bacteroides spp. and ENT (ρ=0.599, p<0.001); in Freemans

Creek there was a weak positive correlation between BacHum and Fecal Bacteroides spp.

(ρ=0.417, p<.034); and in Southwest Creek there was a strong positive correlation between

Page 81: New approaches and technologies for quantifying the impact

64

BacHum and HF183 (ρ=0.793, p<0.05) (Figure 3.5). However, the HF183 assay was only

found positive with regularity (n=6) in Southwest creek.

Using a Kruskall Wallis H test for nonparametric analysis of variance among each

indicator for each creek, statistically significant differences between the different creeks for

Fecal Bacteroides spp. (H(3)=44.2, p<0.001), BacHum (H(3)=8.38, p<0.039), HF183 (H(3)=

41.6, p<0.001), EC (H(3)=19.2, p<0.001), and ENT (H(3)=18.74, p<0.001) were determined.

When indicators within each individual creek were tested, indicators were also significantly

different according to the Kruskall Wallis H test (p<0.001), with the exception of Southwest

Creek, where no indicators were significantly different from one another (H(4)=4.61,

p=0.330).

Optical Brighteners

Optical brightener results for most creeks were consistently negative, with only one

creek repeatedly positive. Optical brightener levels ranged from 24.3 to 297 Raw

Fluorescence Units (RFU). Decay characteristics showed linear decay trends. Only

Southwest Creek and the 2X Tide standard followed a slightly more precipitous decay curve

resulting in greater than 8% decay in the first 5 min. of UV exposure, and ≤ 1.5 ratio between

the first UV exposure (T=5) and second UV exposure (T=10) (Figure 3.6). Southwest Creek

had the highest positive optical brightener results (20% positive). Freeman (n=2) and Gillets

(n=1) also had positive optical brightener results, but positives were very weakly positive

(ratio of 10/5 minute exposure were slightly < 1.5). Optical brighteners were never positive

for any creek during wet weather (rainfall> 2.54 cm) sampling.

Page 82: New approaches and technologies for quantifying the impact

65

Creek Mouth Samples

A small set of samples (n=23) collected from each tidal creek mouth during low tide

were examined for conventional fecal indicators and the three Bacteroides assays. EC and

ENT ranged from 5.00 to 1.03 × 103 and 5.00 to 283 MPN/100 mL, respectively, for all

creek mouth samples. In all creek samples concentrations were <200 MPN/100 mL and

<63.0 MPN/100 mL for EC and ENT, respectively, with the exception of Gillets EC levels

and Southwest ENT levels, which resulted in much larger ranges. For the three Bacteroides

spp. assays, all samples from Courthouse were negative, Freeman had three samples positive

for BacHum, and one sample positive for Fecal Bacteroides, while Gillets and Southwest

both had more than 3 samples positive for both Fecal Bacteroides spp. and BacHum. Gillets

was the only creek mouth positive for HF183, which occurred in only one sample and at low

concentration (210 CE/100 mL).

Land Use

Land use and type was similar for three of the four watersheds, composed primarily

of forested landscapes (Table 3.2). Courthouse was the exception, with nearly 72.0% of the

landscape as developed industrial and commercial use. Courthouse’s mean indicator

concentrations were generally less than or equal to other creeks with the exception of higher

concentrations of Fecal Bacteroides spp. (Figure 3.7). When a Mann-Whitney U test was

used to determine statistically different concentrations of each indicator compared to

Courthouse Creek concentrations, Fecal Bacteroides spp. were significantly different

(higher) (p<0.05) in Courthouse samples as compared to samples from all other creeks.

However, Courthouse concentrations were also significantly different (lower) than HF183

Page 83: New approaches and technologies for quantifying the impact

66

and ENT concentrations in Southwest Creek (U=105, z=-4.42, p<0.001; U=101.5, -2.65,

p<0.01, respectively), and significantly different (higher) than EC and ENT concentrations in

Freeman (U=259, -2.86, p<0.01; U=277, -2.60, p<0.01, respectively).

Discussion

Sources of Fecal Contamination

Levels of conventional fecal indicators of water quality in headwater tidal creeks of

the New River Estuary are often elevated, especially during wet weather conditions (Stumpf

et al., 2010). These levels would, by USEPA regulatory standards, indicate that these waters

are unsafe for swimming and shellfish harvesting during certain periods. Based on results

from conventional indicators, Bacteroides assays, and optical brighteners, Southwest Creek

has convincing evidence of a human source of fecal contamination. Other creeks also have

concentrations of indicator bacteria that are representative of human sources of

contamination, but results were not consistent between indicators. It is still unclear whether

fecal indictors in Courthouse, Gillets, and Freeman are human, animal, or possibly a

combination of both.

Previous research by Stumpf et al. (2010) indicated that the fecal indicators EC and

ENT in these tidal creek headwaters were strongly coupled during storm events and not

during baseflow and ratios of EC/ENT during storms was consistently close to 1. It was

proposed that this could suggest a predominant source of contamination from warm-blooded

animals (excluding humans). However, Fecal Bacteroides spp. and BacHum were present

and in high concentrations in many of these creek samples. Specifically, Bacteroides

measured by the BacHum assay previously proposed as a specific indicator of human

Page 84: New approaches and technologies for quantifying the impact

67

pollution (Kildare et al., 2007), were present in 81% of all creek samples. This latter data

suggests frequent human fecal contamination in each of these creeks.

Other studies have concluded absence of Bacteroides thetaiotaomicron during dry

weather sampling in non-urban streams in California (Tiefenthaler et al., 2008) as evidence

that fecal contamination was not from human sources. Using Bacteroides presence or

absence qualitatively, this would indicate that fecal contamination in these creeks is due to

human pollution. However, when compared to results of the optical brighteners (few positive

detections in any creek) and the HF183 assay (found only in Southwest Creek, with the

exception of one positive sample Freeman Creek), the potential for human contamination in

all creeks based on BacHum results seems less convincing. The result disparity is likely due

to cross-reaction of the BacHum assay with canine (Kildare et al., 2007), swine, and horse

feces (Ahmed et al., 2009). There is a possibility that many of these detections of BacHum in

tidal creeks of MCBCL are the result of fecal contamination from domestic dog populations,

or wildlife such as deer, raccoon, beaver, and opossum, which are abundant in these

watersheds (CH2M Hill, 2000). Further testing with specific wildlife fecal samples could

help to elucidate the potential non-specificity of the BacHum assay. Another possibility is

that frequent detection of the BacHum is from human waste that is present from military

personnel during multi-day training activities in these watersheds. This has previously been

suggested for elevated fecal indicator bacterial concentrations in Freeman Creek (CH2M Hill

2000). Finally, three swine farms are located near headwater tributary creeks of Southwest

Creek, which could influence FIB in the creek mouth sampling data. However, these swine

farms would not affect our headwater tidal creek sampling site, as this sampling point has no

swine operations on the tributary, and is above the tidal signal from the estuary.

Page 85: New approaches and technologies for quantifying the impact

68

Fecal indicator bacteria associated with human sources have could have higher

epidemiological risk than microbial pollution from non-human sources (Scott et al., 2002;

Colford et al., 2007). Presence of human fecal contamination in these creeks is equivocal,

but results seem to indicate that human fecal contamination could be occurring but diffusely,

with the exception of Southwest Creek. Overall, these results solidly support the need for

multiple indicators to accurately determine human fecal contamination.

Bacteroides Assay Comparison

The three discrete Bacteroides assays showed varying results for the targets, with the

BacHum showing the highest average concentrations, followed by Fecal Bacteroides, while

the HF183 was consistently negative for nearly all samples. Previous to this research, it was

assumed that Fecal Bacteroides spp. will be sensitive for more general Bacteroides spp.

(Converse et al., 2009), BacHum for more human specific (Kildare et al., 2007), and HF183

as the confirmatory determinant of human fecal contamination (Bernhard and Field, 2000;

Seurnick et al., 2005; Ahmed et al., 2009). Surprisingly, our results indicate generally higher

mean concentrations for BacHum vs. Fecal Bacteroides. However, Fecal Bacteroides spp.

results were found more often in samples (positive in 95% of samples), while BacHum was

likely more specific (81% positive). Overall, concentrations of Fecal Bacteroides spp. were

not significantly related to BacHum concentrations.

The HF183 assay proved to be very consistent in identifying sewage and

oligonucleotide standards, but showed little indication of fecal contamination from tidal

creek samples. This could be due to either a lack human fecal contamination in the

environmental samples or the low sensitivity of the assay (Seurnick et al., 2005). Previous

Page 86: New approaches and technologies for quantifying the impact

69

research has shown the HF183 to have a limit of detection of 4.7×104 CE per 100 mL

(Seurnick et al., 2005). However, in limited testing of known sewage contaminated beach

water samples from San Cristobal Island, Ecuador, (Stumpf, Noble, and Gonzalez,

unpublished data) we found the assay to be very effective at detecting sewage fecal

contamination (all samples positive). This, in combination with the continuous positives for

all sewage standards run with the assay seems to strengthen previous research that this assay

is a “gold standard” for determination of human source fecal contamination (Bernhard and

Field 2000; Seurnick et al., 2005; Ahmed et al., 2009), though high concentrations of target

DNA are necessary for detection. During the few occurrences that the HF183 assay was

positive, there was a strong relationship with BacHum (Figure 3.5). This could indicate that

the BacHum is a good test for human contamination as proposed by Kildare et al. (2007), and

simply has a better sensitivity, allowing for detection of much lower concentrations. If this is

the case, there is the potential for widespread contamination of these four tidal creek

headwaters, as BacHum was found through these creeks and in many case, in high

concentrations (>1×103 CE/100 mL).

Multi-Tiered Incorporation of Optical Brighteners

Optical brighteners (OB) indicated little to no contamination from human wastewater

sources in the tidal creeks with the exception of Southwest Creek. Due to high levels of

background fluorescence from high organic concentrations in these creeks, the UV decay

method for OB detection developed by Hartel et al. (2007) and modified by Cao et al. (2009)

was utilized. All samples showed decay of optical brightener signal after exposure to UV at 5

and 10 min. timepoints. However, decay curves were not precipitous, indicative of natural

Page 87: New approaches and technologies for quantifying the impact

70

organic decay and not household wastewater (Figure 3.6). Additionally, samples rarely

tested positive for all three replicates, but often tested positive at least one time, indicating

the need for triplicate analysis due to between sample variation, and the risk for false

positives if only one sample was tested. This might also indicate that the method is less

practical for application in tidal creeks, which have naturally high background fluorescence

from organics, or that the criteria for a positive sample might need to change based on

sampling environment. In addition, human pollution from non-wastewater sources (i.e. direct

pollution from troop activities) would not test positive for optical brighteners since not

directly related to household wastewater. Therefore, negative results in these undeveloped

creeks do not necessarily indicate the absence of human fecal pollution, simply the absence

of human wastewater presence.

Previously, the UV decay method was found to be very accurate (99%) at detecting

OB in a blind study of 180 samples (Hartel et al., 2007). The results of this study comparing

the Bacteroides HF183 assay to OB outcomes support the accuracy of this assay. For

instance, the HF183 Bacteroides assay has been purported as a very accurate determinant of

human sewage associated contamination (Bernhard and Field, 2000; Seurinck et al., 2005;

Ahmed et al., 2009). It would be reasonable to expect the presence of OB in samples that

were found positive for the HF183 assay, since OB are found in wastewaters. Southwest

Creek was the only water body consistently positive for the HF183 assay (50% of samples)

and was also positive for OB (20% of samples), which seems to support the connection

between household wastewater contamination, the presence of OB, and the presence of

HF183.

Page 88: New approaches and technologies for quantifying the impact

71

This OB method is a simple and inexpensive way to test for human wastewater.

However, more testing would be helpful to verify the method with known wastewater

contaminated samples, conducted using a sensitivity study. This would involve using water

from each creek spiked with different concentrations of an OB standard (Tide 2x detergent

was used for this study) to determine the sensitivity of the assay for each water body being

tested. This would assume that no OB is present in the creeks, which complicates the testing.

However, such a test would assist in understanding the sensitivity of the technique for the

water body under study. Generally, the OB method should be incorporated as part of a multi-

tiered approach to determining fecal contamination, and not as a stand-alone test of fecal

contamination.

Creek Mouth and Land Relationships

Creek mouth FIB mean concentrations were low or not detected, and generally lower

than headwater tidal creek mean concentrations. However, these samples were taken during

dry weather periods, so it is unclear whether elevated concentrations occur in creek mouths

after storm events. Similar to headwater creek samples, Gillets and Southwest Creek had the

highest concentrations of fecal indicators at creek mouth sampling location. Interestingly, all

three Bacteroides assays showed much lower mean concentrations in the creek mouth areas

(often ND) as compared to the headwater portions of creeks. Fecal contamination is likely

occurring in headwater portions and attenuated during downstream transport, at least during

periods of dry weather. Similar gradients of fecal indicator bacteria have been documented in

an estuarine creek in Australia, and attributed to tidal dilution, degradation, and physico-

chemical properties (stress on indicators) of the downstream estuary (Mill et al. 2006). Due

Page 89: New approaches and technologies for quantifying the impact

72

to Bacteroides spp. generally short duration in the environment, and quick attenuation from

increased temperature, grazing, and UV exposure (Kreader, 1998; Bell et al., 2009), presence

of these indicators may be better predictors of recent fecal contamination than conventional

EC and ENT in estuarine systems. This is important from a management perspective, since

shellfishing activities are focused in these estuarine/creek mouth areas and not in tidal creek

headwaters.

Though developed land use has previously been associated with higher concentrations

of fecal indicators (Mallin et al., 2000; Mallin et al., 2001; Carle et al., 2005; DiDonato et al.,

2009), there is limited data associated with land use and Bacteroides indicator assays. Due to

the very limited development in many of Marine Corps Base Camp Lejeune (MCBCL)

watersheds (T. Mentor unpublished data) in eastern NC, examination of Bacteroides spp.

concentrations within an undeveloped landscape was possible. Despite a common correlation

with developed land surface and increases in conventional FIB, this was not the case in the

four watersheds in this study. The only indicator showing elevated concentrations associated

with a highly developed watershed (72%) at Courthouse Bay Creek, was the Fecal

Bacteroides spp. indicator. Despite low development at Southwest Creek and Gillets,

indicators were often higher than other creeks. These elevated Bacteroides spp.

concentrations could be indicative of human contamination from infrastructure failure (i.e.

leaking sewer lines or overflowing sewage systems during storm events), from military

personnel during extended outdoor training exercises, false positives from wildlife cross

reactivity with the assays, or other unknown factors. However, a much larger study involving

multiple watersheds with varying land use development would need to be conducted to

Page 90: New approaches and technologies for quantifying the impact

73

confirm these results, as these watersheds were primarily undeveloped and only four

watersheds were examined.

Limitations of QPCR in Tidal Creeks

QPCR has been proposed as a rapid method for detecting fecal contamination in

waters (Dick and Field 2004; Noble et al., 2006; Converse et al., 2009) however, this study

found limitations when using QPCR in tidal creek and estuarine environments. Tidal creek

samples were often inhibited (especially when analyzing storm event samples) and required

use of a Power Soil (MoBio) extraction kit that trades off inhibitor reduction for poor

extraction efficiency. Recovery of a known quantity of oligonucleotide standards were tested

for the BacHum and HF183 assays, resulting in extraction recovery < 1%. This low

extraction recovery, when combined with high limits of detection (LOD) of the HF183 assay

in particular, could explain the large number of NDs for samples with lower initial target

concentrations. Overall extraction efficiency was low (Mumy and Findlay, 2004), but low

extraction recovery has previously been cited as a problem for QPCR (Stewart et al., 2008;

Stoeckel et al., 2009). One possibility is that low extraction efficiency was due to destruction

of single stranded oligonucleotide standard during the extraction process (the recovery of the

Fecal Bacteroides spp. extracted cell standard was not quantified). A QPCR DNA standard

needs to show similar extraction and recovery to accurately quantify the target cell DNA

(Stoeckel et al., 2009). Extraction variation and poor recovery has been cited previously as

one of the most difficult problems for the use of QPCR for water quality samples (Stewart et

al., 2008). Further research to confirm that oligonucleotide standards of the BacHum and

HF183 assay are extracting similarly to the cellular target DNA in these water samples is

Page 91: New approaches and technologies for quantifying the impact

74

needed. In addition, the integrity of the extraction (potential of cross contamination) varied

between extraction kits utilized, and the extraction process itself was very time consuming

(requiring >2.5 h per 20 samples). Extraction contamination and time constraints could be

remedied through use of automated DNA extraction process (Knepp et al., 2003; Beuselinck

et al., 2005), but these automated systems are not economical or available for many water

quality testing laboratories.

Though molecular methods might be effective and rapid for less complex sample

types (i.e. beach water samples have previously shown little inhibition when extracted using

a simple and rapid bead beating and centrifugation method (Converse et al., 2009)),

molecular analysis was not rapid or easy for the tidal creek samples from this study. Though

QPCR could be used in the future in tandem with conventional fecal indicators for federal

water quality monitoring, these results indicate that QPCR in highly turbid tidal creek and

other complex water environments may not be the most appropriate tests for quantification of

fecal contamination.

Conclusions

The goal of this research was to determine if the source of fecal contamination in tidal

creeks was human or animal, using conventional and alternative fecal indicators. Based on

the results of this study, the source of contamination in these tidal creeks is not entirely clear.

This research does indicate that there is suggestive human fecal contamination of Southwest

Creek. However, evidence for human fecal contamination in the other creeks requires further

validation. If further research does conclude that fecal contamination in these creeks is not

anthropogenic, then conventional fecal indicators are not a good determinant of human health

Page 92: New approaches and technologies for quantifying the impact

75

risk from recreational activities such as bathing and shellfishing in these tidal creek waters or

the adjacent estuary. This is likely the case in other highly forested, low development

watersheds that still show signs of high levels of fecal indicator bacteria.

The multi-tiered approach of conventional indicators, Bacteroides assays, and optical

brightener indicators did increase confidence in the results of Southwest creek. However,

none of these techniques for determining fecal contamination stands alone as a sole predictor

of source of contamination. Determining fecal source has traditionally been, and continues to

be one of the most important yet difficult aspects of fecal contaminant monitoring in

environmental waters. Though a suite of indicators can be employed, source is difficult to

determine without a visual presence at the actual event (sewage overflow, wastewater

treatment plant failure, etc.).

Molecular techniques (QPCR) have been presented as a rapid and accurate method of

fecal indicator and alternative indicator detection; however, these methods are not always

appropriate in differing water environments. Necessity of multi-step extractions can be very

time consuming, and result in poor recovery of target. The future of rapid testing for water

quality impairment will include molecular detection assays, but molecular detection will

likely be one type of technology in a suite of tools, including the emergence of faster, more

user friendly, and field portable biosensors.

Acknowledgements

This research was supported by the Defense Coastal/Estuarine Research Program (DCERP),

funded by the Strategic Environmental Research and Development Program (SERDP).

Views, opinions, and/or findings contained in this report are those of the authors and should

Page 93: New approaches and technologies for quantifying the impact

76

not be construed as an official U.S. Department of Defense position or decision unless so

designated by other official documentation.

Page 94: New approaches and technologies for quantifying the impact

77

General

Target

Assay

Abbreviation

Study Reaction Conditions 16S rRNA

fragment

Length

Standard

Salmon Sperm SPC Haugland et al., 2005

95°C for 2 min, followed by 45 cycles of 15 s at 94°C and 30 s at 60°C

Unknown Cell Standard

Fecal Bacteroides

Fecal Bacteroides

Converse et al., 2009

95°C for 2 min, followed by 45 cycles of 95°C for

15 s and 60°C for 30 s

110 base pair

Cell Standard

Human Bacteroides

BacHum Kildare et al., 2007

50°C for 2 min, followed by 95°C for 10 min,

followed by 40 cycles at 95°C for 15 sec and 60°C

for 1 min

81 base pair Oligonucleotide Standard

Sewage

Bacteroides HF183

Bernhard and Field, 2000;

Seurnick et al., 2005

50°C for 2 min, followed by 95°C for 10 min,

followed by 40 cycles at 95°C for 30 sec 53°C for

1 min and 60°C for 1 min. Follow by a melt curve

82 Base

pair

Oligonucleotide

Standard

Table 3.1. Reaction conditions and standards for all QPCR assays

77

Page 95: New approaches and technologies for quantifying the impact

78

Table 3.2. Description of tidal creek predominant land cover, fecal indicator detection, and percent exceedence of regulatory

standards for conventional indicators

Tidal

Creek

Headwater

Predominant

Land Cover/

Type

Indicator Sample # Log Mean

MPN or

Cells/100 mL

# Non-

Detects

%

Exceedence

Courthouse

71.7% Business

or Commercial

EC

35

2.43 0 40

ENT 2.20 0 54

Fecal Bacteroides 3.33 1 −

HF183 0.00 35 −

BacHum 2.97 8 −

OB 15 − 15* 0

Southwest

97.0% Forested

EC

12

2.63 0 92

ENT 2.95 0 83

Fecal Bacteroides 2.68 0 −

HF183 1.81 6 −

BacHum 3.37 2 −

OB 15 − 12* 20

Gillets

94.5% Forested

EC

33

2.61 0 61

ENT 2.11 1 50

Fecal Bacteroides 1.91 2 −

HF183 0.00 33 −

BacHum 3.32 5 −

OB 15 − 14* 7

Freeman

96.3% Forested

EC

26

1.97 0 8

ENT 1.75 0 31

Fecal Bacteroides 2.16 0 −

HF183 0.11 25 −

BacHum 2.83 10 −

OB 15 − 13* 13

*Fluorescence was measured in all samples; Non-Detects= non-positive results

- Not applicable because there are no EPA regulations for Bacteroides spp.

78

Page 96: New approaches and technologies for quantifying the impact

79

Fecal

Bacteroides BacHum HF183 EC ENT

Fecal

Bacteroides 1.000 0.950 0.159 0.049 .252**

BacHum -

1.000 .312** .339** 0.021

HF183 - -

1.000 0.110 0.120

EC - - -

1.000 .449**

ENT - - - -

1.000

Table 3.3. Bivariate correlations among log transformed data from all creeks using two-

tailed Spearman Rho rank correlation.** indicates p<0.01

Page 97: New approaches and technologies for quantifying the impact

80

Figure 3.1. Marine Corps Base Camp Lejeune tidal creeks of study. Arrows indicate creeks

examined for Bacteroides spp. using three discrete assays

Page 98: New approaches and technologies for quantifying the impact

81

Log E. coli (MPN/100mL)

4.54.03.53.02.52.01.51.0.5

Log

Ente

roco

ccu

s (M

PN/1

00m

L)

4.0

3.5

3.0

2.5

2.0

1.5

1.0

.5

Southwest

Gillets

Freeman

Courthouse

Figure 3.2. Regulatory exceedences for E. coli (right half) and Enterococcus spp. (top half)

for each creek during all sampling events. E. coli and Enterococcus spp. exceedences are

based on regulatory standards of 320 and 104 MPN/100 mL, respectively.

Page 99: New approaches and technologies for quantifying the impact

82

Figure 3.3. Mean log indicator concentration for each creek. Error bars represent the standard

error of the mean

Page 100: New approaches and technologies for quantifying the impact

83

Fecal Indicator

ENT

EC

HF183

BacHum

Fecal Bacteroides

Lo

g In

dic

ato

r (M

PN

or

CE

/10

0m

L)

7

6

5

4

3

2

1

0

-1

Figure 3.4. Box-plots of log mean concentrations of conventional and Bacteroides indicators

for all creeks combined.

Page 101: New approaches and technologies for quantifying the impact

84

Figure 3.5. Relationship of BacHum to HF183 results when HF183 assay produced positive

results. Spearman’s rho= 0.793 and p<0.05

Page 102: New approaches and technologies for quantifying the impact

85

Figure 3.6. A representative sample for optical brightener decay curves for each tidal creek

and Tide 2x standard. Southwest Creek and the 2x Tide standard are positive for optical

brighteners

Page 103: New approaches and technologies for quantifying the impact

86

Figure 3.7. Relationships among mean log indicator concentrations from each tidal creek

headwater and percent land-development for each watershed

Page 104: New approaches and technologies for quantifying the impact

87

REFERENCES

Ahmed, W., Goonetilleke, A., Powell, D., Gardner, T. (2009). Evaluation of multiple

sewage-associated Bacteroides PCR markers for sewage pollution tracking. Water Research, 43, 4872-4877.

Ahmed, W., Goonetilleke, A., Gardner, T. (2008). Alternative indicators for detection and

quantification of faecal pollution Water. Journal of the Australian Water Association, 39-45.

Bell, A., Layton, A.C., McKay, L., Williams, D., Gentry, R., Sayler, G.S. (2009). Factors influencing the persistence of Fecal Bacteroides in stream water. Journal of Environmental

Quality, 38, 1224-1232.

Bernhard, A.E., Field, K.G. (2000). Identification of nonpoint sources of fecal pollution in coastal waters by using host-specific 16S ribosomal DNA genetic markers from fecal

anaerobes. Applied and Environmental Microbiology, 66(4), 1587-1594.

Beuselinck, K., Van Ranst, M., Van Eldere, J. (2005). Automated Extraction of Viral-Pathogen RNA and DNA for High-Throughput Quantitative Real-Time PCR. Journal of

Clinical Microbiology, 43(11), 5541-5546.

Boehm, A.B., Fuhrman, J.A., Mrse, R.D., Grant, S.B. (2003). Tiered approach for indentification of human fecal pollution source at a recreational beach: Case study at Avalon

Bay, Catalina Island, California. Environmental Science and Technology, 27, 673-680.

Byappanahalli, M., Fowler, M., Shively, D., Whitman, R. (2003). Ubiquity and persistence of Escherichia coli in a midwestern coastal stream. Applied and Environmental Microbiology,

69(8), 4549-4555

Cao, Y., Griffith, J.F., Weisberg, S.B. (2009). Evaluation of optical brightener photodecay characteristics for detection of human fecal contamination. Water Research, 43, 2273-2279.

Carle, M.V., Halpin, P.N., Stow, C.A. (2005). Patterns of watershed urbanization and

impacts on water quality. Journal of American Water Resources Association, 41(3), 693-708.

Page 105: New approaches and technologies for quantifying the impact

88

Carson, C.A., Christiansen, J.M., Yampara-Iquise, H., Benson, V.W., Baffaut, C., Davis, J.V., Broz, R.R., Kurtz, W.B., Rogers, W.M., Fales, W.H. (2005). Specificity of a

Bacteroides thetaiotaomicron marker for human feces. Applied and Environmental Microbiology, 71(8), 4945-4949.

CH2M Hill (2000). Freeman Creek evaluation and assessment; C2HM Hill, Inc.,

Prepared for US Army Corps of Engineers, Contract No. DACA01-96-0028.

Converse, R.R., Blackwood, A.D., Kirs, M., Griffith, J.F., Noble, R.T. (2009). Rapid QPCR-based assay for Fecal Bacteroides spp. as a tool for assessing fecal contamination in

recreational waters. Water Research, 43(19), 4828-4837.

Colford, J.M., Wade, T.R., Schiff, K.C., Wright, C.C., Griffith, J.F., Sukhminder, S.K., Burns, S., Sobsey, M., Lovelace, G., Weisberg, S.B. (2007). Water quality indicators and the

risk of illness at beaches with nonpoint sources of fecal contamination. Epidemiology, 18(1), 27-35.

Dick, L.K., Field, K.G. (2004). Rapid estimation of numbers of fecal Bacteroidetes by use of

a quantitative PCR assay for 16S rRNA genes. Applied and Environmental Microbiology, 70(9), 5695-5697.

Dickerson, J.W., Hagedorn, C., Hassall, A. (2007). Detection and remediation of human-

origin pollution at two public beaches in Virginia using multiple source tracking methods. Water Research, 41, 3758-3770.

DiDonato, G.T., Stewart, J.R., Sanger, D.M., Robinson, B.J., Thompson, B.C., Holland,

A.F., Van Dolah, R.F. (2009). Effects of changing land use on the microbial water quality of tidal creeks. Marine Pollution Bulletin, 58 (1), 97-106.

Drasar, B.S. (2003). Microbial flora of the gut. In: Mara, D., Horan, N. (Eds.), Handbook of

Water and Wastewater Microbiology. Academic Press, San Diego, pp. 99-104.

Ensign S.H., Mallin, M.A. (2001). Stream water quality changes following timber harvest in a coastal plain swamp forest. Water Research, 35(14), 3381-3390.

Field, K.G., Samadpour, M. (2007). Review: Fecal source tracking, the indicator paradigm, and managing water quality. Water Research, 41(16), 3517-3538.

Page 106: New approaches and technologies for quantifying the impact

89

Fiksdal, L., Maki, J.S., LaCroix, S.J., Staley, J.T., (1985). Survival and detection of Bacteroides spp., prospective indicator bacteria. Applied and Environmental Microbiology,

49(1), 148-150.

Hurley, M.A., Roscoe, M.E. (1983). Automated statistical-analysis of microbial enumeration by dilution series. Journal of Applied Bacteriology, 55(1), 159-164.

Hartel, P.G., McDonald, J.L., Gentit, L.C., Hemmings, S.N.J., Rodgers, K., Smith, K.A.,

Belcher, C.N., Kuntz, R.L., Rivera-Torres, Y., Otero, E., Schroder, E.C. (2007). Improving fluorometry as a source tracking method to detect human fecal contamination.

Estuaries and Coasts, 30(3), 551-561.

Haugland, R.A., Siefring, S.C., Wymer, L.J., Brenner, K.P., Dufour, A.P. (2005). Comparison of Enterococcus measurements in freshwater at two recreational beaches by

quantitative polymerase chain reaction and membrane filter culture analysis. Water Research, 39, 559-568.

Johnston, C., Ufnar, J.A., Griffith, J.F., Gooch, J.A., Stewart, J.R. (2010). A real-time qPCR

assay for the detection of the nifH gene of Methanobrevibacter smithii, a potential indicator of sewage pollution. Applied and Environmental Microbiology, 109(6), 1946-1956.

Kildare, B.J., Leutenegger, C.M., McSwain, B.S., Bambic, D.G., Rajal, V.B., Wuertz, S.

(2007). 16S rRNA-based assays for quantitative detection of universal, human-, cow-, and dog-specific fecal Bacteroidales: A Bayesian approach. Water Research, 4, 3701-3715.

Knepp, J.H., Geahr, M.A., Forman, M.S., Valsamakis, A., (2003). Comparison of Automated

and manual nucleic acid extraction methods for detection of enterovirus RNA. Journal of Clinical Microbiology, 41(8), 3532–3236.

Kreader, C.A. (1998). Persistence of PCR-detectable Bacteroides distasonis from human

feces in river water. Applied and Environmental Microbiology, 64(10), 4103-4105.

Layton, A., McKay, L., Williams, D., Garrett, V., Gentry, R., Sayler, G., (2006). Development of Bacteroides 16S rRNA gene TaqMan-based real-time PCR assays for

estimation of total, human, and bovine fecal pollution in water. Applied and Environmental Microbiology, 72(6), 4214-4224.

Page 107: New approaches and technologies for quantifying the impact

90

Mallin, M.A., Williams, K.E., Esham, E.C., Lowe R.P. (2000). Effect of human development on bacteriological water quality in coastal watersheds. Ecological

Applications, 10(4), 1047-1056.

Mallin, M.A., Ensign, S.H., McIver, M.R., Shank, G.C., Fowler, P.K. (2001). Demographic, landscape, and meteorological factors controlling the microbial pollution of coastal waters.

Hydrobiologia, 460(1-3), 185-193.

Mallin, M.A., Lewitus A.J. (2004). The importance of tidal creek ecosystems. Journal of Experimental Marine Biology and Ecology, 298(2), 145-149.

McQuaig, S.H., Scott, T.M., Harwood, V.J., Farrah, S.R., Lukasik, J.O. (2006). Detection of

human-derived fecal pollution in environmental waters by use of a PCR-based Human Polyomavirus assay. Applied and Environmental Microbiology, 72(12), 7567-7574.

Meays, C.L., Broersma, K., Nordin, R., Mazumder, A. (2004). Source tracking fecal bacteria

in water: a critical review of current methods. Journal of Environmental Management, 73, 71-79.

Mill, A., Schlacher, T., Katouli M., (2006). Tidal and longitudinal variation of faecal

indicator bacteria in an estuarine creek in south-east Queensland, Australia. Marine Pollution Bulletin, 52(8), 881-891.

Mumy, K.L., Findlay, R.H. (2004). Convenient determination of DNA extraction efficiency

using an external DNA recovery standard and quantitative-competitive PCR. Journal of Microbiological Methods, 57(2), 259-268.

Noble, R.T., Allen, S.M., Blackwood, A.D., Chu, W., Jiang, S.C., Lovelace, G.L., Sobsey,

M.D., Stewart, J.R., Wait, D.A. (2003). Use of viral pathogens and indicators to differentiate between human and non-human fecal contamination in a microbial source tracking

comparison study. Journal of Water and Health, 1, 195-207.

Noble , R.T., Griffith, J.F., Blackwood, A.D., Fuhrman, J.A., Gregory, J.B., Hernandez, X., Liang, X., Bera, A.A., Schiff, K. (2006). Multitiered approach using quantitative PCR to

track sources of fecal pollution affecting Santa Monica Bay, California. Applied and Environmental Microbiology, 72(2), 1604-1612.

Page 108: New approaches and technologies for quantifying the impact

91

Noble, R.T., Fuhrman, J.D. (1998). Use of SYBR Green I for rapid epifluorescence counts of marine viruses and bacteria. Aquatic Microbiology and Ecology, 14, 113-118.

Peeler, K.A., Opsahl, S.P., Chanton, J.P. (2006). Tracking anthropogenic inputs using

caffeine, indicator bacteria, and nutrients in rural freshwater and urban marine systems. Environmental Science and Technology, 40(24), 7616-7622.

Pfaffl, M.W. (2001). A new mathematical model for relative quantification in real-time

RTPCR. Nucleic Acids Research, 29(9), e45.

Savichtcheva, O., Okayama, N., Okabe, S. (2007). Relationships between Bacteroides 16S rRNA genetic markers and presence of bacterial enteric pathogens and conventional fecal

indicators. Water Research, 41, 3615-3628.

Rosario, K., Symonds, E.M., Sinigalliano, C., Stewart, J., Breitbart, M. (2009). Pepper Mild Mottle Virus as an indicator of fecal pollution. Applied and Environmental Microbiology,

75(22), 7261-7267.

Roser, D.J., Ashbolt, N.J. (2007). Source water quality and management of pathogens in surface catchments and aquifers. Research Report 29, CRC for Water Quality and Treatment,

Bolivar.

Scott, T.M., Rose, J.B., Jenkins, T.M., Farrah, S.R., Lukasik, J. (2002). Microbial source tracking: current methodology and future directions. Applied and Environmental

Microbiology, 68(12), 5796-5803.

Seurinck, S., Defoirdt, T., Verstraete, W., Siciliano, S.D. (2005). Detection and quantification of the human-specific HF183 Bacteroides 16S rRNA genetic marker with real-

time PCR for assessment of human faecal pollution in freshwater. Environmental Microbiology, 7(2), 249–259.

Solo-Gabriele, H.M., Wolfert, M.A., Desmarais, T.R., Palmer, C.J. (2000). Sources of

Escherichia coli in a coastal subtropical environment. Applied and Environmental Microbiology, 66(1), 230-237.

Page 109: New approaches and technologies for quantifying the impact

92

Stewart, J., Gast, R., Fujioka, R., Solo-Gabriele, H., Meschke, J.S., Amaral-Zettler, L., del Castillo, E., Polz, M., Collier, T., Strom, M., Sinigalliano, C., Moeller, P., Holland, A.F.

(2008). The coastal environment and human health: microbial indicators, pathogens, sentinels and reservoirs, Environmental Health, 7 (Suppl. 2): S3, 1-14.

Stoeckel D.M., Stelzer, E.A., Dick, L.K. (2009). Evaluation of two spike-and-recovery

controls for assessment of extraction efficiency in microbial source tracking studies. Water Research, 43(19), 4820-4827.

Stumpf, C.H., Piehler, M.F., Thompson, S., Noble, R.T. (2010). Loading of fecal indicator

bacteria in North Carolina tidal creek headwaters: Hydrographic patterns and terrestrial runoff relationships. Water Research, 44(16), 4704-4715.

Tiefenthaler, L.L., Stein, E.D., Lyon, G.S. (2008). Fecal indicator bacteria (FIB) levels

during dry weather from Southern California reference streams. Environmental Monitoring Assessment, 155(1-4), 477-492.

Wade, T.J., Calderon, R.L., Sams, E., Beach, M., Brenner, K.P., Williams, A.H., Dufour,

A.P. (2006). Rapidly measured indicators of recreational water quality are predictive of swimming-associated gastrointestinal illness. Environmental Health Perspectives, 114(1), 24-

28.

Wade, T.J., Calderon, R.L., Brenner, K.P., Sams, E., Beach, M., Haugland, R., Wymer, L., Dufour, A.P. (2008). High sensitivity of children to swimming-associated

gastrointestinal illness: results using a rapid assay of recreational water quality. Epidemiology, 19(3), 375-383.

Page 110: New approaches and technologies for quantifying the impact

CHAPTER 4

Filtration and elution techniques for concentration of intact Escherichia coli, Enterococcus spp., and Bacteroides thetaiotaomicron cells from seawater

Introduction

Enterococcus spp. and Escherichia coli (E. coli) are fecal indicator bacteria (FIB)

used to determine fecal contamination in recreational waters. As proxies of fecal

contamination, presence of FIB in drinking, bathing, and shellfishing waters can relate to

potential health risks (Cabelli et al., 1982; Wade et al., 2003; Wade et al., 2006; Marion et

al., 2010). Current EPA-approved methods (both culture based and defined substrate

technologies) for detection of E. coli and Enterococcus spp. require 18-24 h. More rapid

molecular techniques such as quantitative polymerase chain reaction (QPCR) have been

developed for enumeration of E. coli and Enterococcus spp. However, these techniques

involve complex multi-step procedures, can result in false negatives due to inhibition, and are

currently limited to more advanced laboratory testing facilities with highly trained personnel.

Recent studies have shown significant relationships between culture based and QPCR

methods (Lavender and Kinzelman, 2009), and QPCR and human health risk (Wade et al.,

2006; Wade et al., 2008). Despite these relationships to human health risk, current methods

still require significant time to acquire results, which may place water users at risk.

Species of Bacteroides have been proposed as alternatives to conventional indicators

and as a better predictor of human contamination (Fiksdal et al., 1985; Carson et al., 2005;

Savichtcheva et al., 2007; Converse et al., 2009). Bacteroides spp. have several advantages

Page 111: New approaches and technologies for quantifying the impact

94

over current indicators: they are obligate anaerobes and therefore presence is representative

of relatively recent contamination (Kreader, 1998); they are the dominant enteric bacteria in

humans, with individual Bacteroides sp. as high as 1011 cells/g (Drasar, 2003); they do not

proliferate once released into the environment (Dick and Field, 2004); they show significant

correlation with certain pathogens (Savichtcheva et al., 2007); and they have been connected

to human fecal contamination using discrete assays (Bernhard et al., 2003; Dick and Field,

2004; Carson et al., 2005; Noble et al., 2006). Epidemiology studies have also found

relationships to fecal Bacteroides spp. and public health risk (Wade et al., 2006).

Additionally, because of their close association with human fecal contamination, Bacteroides

spp. might be better candidates for source tracking contamination in waters (Layton et al.,

2006; Ahmed et al., 2009; Converse et al., 2009). Specifically, B. thetaiotaomicron are found

at greater concentrations than E. coli or Enterococcus spp. per g of feces (Drasar, 2003).

Since concentrations of B. thetaiotaomicron are higher compared to current indicators, their

use may assist developing technologies with poor sensitivities and low detection limits.

Due to complexity and time restraints of currently used assays for quantification of

FIB, simple and rapid methods are under development, including biosensor technologies

(Ivnitski et al., 1999; Noble and Weisberg, 2005). A biosensor is a device that detects an

analyte(s) through a link between a biological component and a complementary

physicochemical component. Some examples of water quality biosensor technologies under

development include suspension microarrays (i.e. Luminex®100™) (Baums et al., 2007), the

immunomagnetic separation/adenosine triphosphate method (Bushon et al., 2009), bacterial

enzyme fluorometry methods (ie. FluoroQuick™, Coliplage®), surface plasmon resonance

methods (Hoa et al., 2007; Garcia-Aljaro et al., 2008); enrichment based technologies

Page 112: New approaches and technologies for quantifying the impact

95

(Xplorer64™ System) and liquid crystal based method (Shiyanovskii et al., 2005). Many of

these systems require high concentrations of intact, metabolically active cells to overcome

problems with sensitivity and for overall analyte detection.

A simple method for FIB concentration in environmental water samples is needed to

overcome the poor sensitivity of many biosensors. Most developing technologies for

detection of microbial indicators can handle less than 1 mL volumes, while current methods

require sensitivity at about 1 cell per mL (Noble and Weisberg, 2005). Water quality standard

thresholds have relatively low cell counts. For example, bathing beach advisories are posted

when Enterococcus spp. concentrations exceed a geometric mean of 35 colony forming units

(CFU)/100 mL or a single sample maximum of 104 CFU/100 mL. Tangential flow (Reid and

Adlam, 1974; Barthel et al., 1989;) and ultrafiltration (Tsai et al., 1993; Morales-Morales et

al., 2003; Hill et al., 2005; Leskinen and Lim, 2008) have been used previously for

concentration of microbes. However, these techniques utilize large sample volumes and

expensive equipment. Also, due to the shear experienced by cells in the water sample, these

methods are not necessarily optimal for concentration and recovery of intact cells.

Biosensor technologies are also often hindered by inhibition or interference by

particulates in the water samples. Particulates including sand, sediments, and plankton can

interfere with bacterial detection. E. coli, Enterococcus spp., and Bacteroides spp. are often

attached to these particulates (Maugeri et al., 2004; Signoretto et al., 2004; Fries et al., 2006;

Zhang et al., 2007) resulting in inconsistency between different enumeration methods or

increased loss of target bacteria if particulates are removed prior to enumeration. Though

there is no known way to selectively remove all particulates without further reducing target

Page 113: New approaches and technologies for quantifying the impact

96

bacteria concentration, an optimized technique for particulate removal is necessary in order

to maximize cell recovery while removing larger particles.

Our goal was to develop a simple technique for concentration of whole cells while

maximizing removal of large particulates (>30 µm) from marine samples. We examined

different methods of bacterial concentration including varying filter sizes, elution solutions,

and cell recovery techniques (inclusion of glass beads, vortexing, and sonication) and the

potential for particulate removal using various prefilters and centrifuging. We have

developed and optimized two methods for varying sample volumes, which can be used to

concentrate E. coli, Enterococcus spp., and Bacteroides thetaiotaomicron (B.

thetaiotaomicron) cells in marine or freshwater. These methods will assist in the

development of rapid technologies for water quality assessment by providing pre-processing

techniques for sample concentration and large particulate removal.

Materials and Methods

Sample Collection and Preparation

Seawater was collected from ocean wave wash (~2 m offshore) at Fort Macon State

Park, Atlantic Beach, North Carolina (NC), using 10 L sterile carboys, and tested for

background fecal indicator bacteria. Seawater was inoculated with cultured bacteria (E. coli

ATCC 25922, Enterococcus faecalis ATCC 29212, or Bacteroides thetaiotaomicron ATCC

29148), at starting concentrations of 102-105 cells per 100 mL. E. coli was cultured in tryptic

soy broth (TSB) (incubated 18 h. at 35ºC), E. faecalis was cultured in brain heart infusion

(BHI) (incubated 18 h. at 37ºC), and B. thetaiotaomicron was cultured in cooked meat

medium (anaerobic incubation for 48 h. at 37ºC). E. coIi and E. faecalis cultures were

Page 114: New approaches and technologies for quantifying the impact

97

transferred after initial incubation and re-grown until exponential growth phase (determined

using a PharmaSpec 1700 spectrophotometer (Shimadzu Scientific Instruments, Columbia,

MD) when the optical density at 520 nm was between 0.1 and 0.5) and immediately

inoculated into seawater. Methods were tested with inoculated seawater unless stated

otherwise (i.e. freshwater pond samples were occasionally tested for comparison). After

optimization, methods were tested on seawater inoculated with sewage influent from the

Morehead City, NC wastewater treatment plant, and the results were compared.

Multi-Filter Method (MFM)

Particulate removal techniques

Three techniques- prefiltration, centrifugation, and sonication- were explored for their

ability to remove particulates and dissociate bacteria from partic les before concentration.

Prefiltration was tested using 47 mm nylon net filters of 100 µm, 60 µm, and 30 µm pore

sizes (Millipore, Billerica, MA), and filtered using a vacuum pump (Gast, Benton Harbor,

MI) at 25.4 cm Hg. An Allegra 25R Centrifuge (Beckman Coulter, Brea, CA) was used for

centrifugation. Samples were placed in multiple 250 mL conical tubes (Nalgene, Rochester,

NY) and centrifuged at different speeds (200, 400, 600, and 800×g) and for different

durations (60 and 90 sec.). After centrifuging, the top 200 mL of supernatant was aspirated

from each 250 mL conical container without disturbing settled particulates. Sample from

both treatments were enumerated to determine recovery of bacterial cells and remaining size

distribution of particulates. Samples were also sonicated using a Bransonic® ultrasonication

bath (Danbury, CT) for 15 sec. prior to prefiltration or centrifugation to determine if

dissociation of cells from particles before particle removal would reduce bacterial loss.

Page 115: New approaches and technologies for quantifying the impact

98

Concentration Techniques

Concentrations of E. coli and Enterococcus spp. were tested with 0.4 µm 47 mm

diameter polycarbonate (PC) filters (Millipore, Billerica, MA). A 1250 mL sample was

equally distributed onto five individual 0.4 µm PC filters, each placed on top o f a GF/F glass

microfiber backing filters (Whatman, Piscataway, NJ) and held in place with a 47 mm

magnetic filter funnels (Pall, Port Washington, NY). A vacuum was applied to each filtration

unit concurrently using a pump (Gast, Benton Harbor, MI) at 25.4 cm Hg. After complete

filtration, the inside of each filter funnel was rinsed thoroughly with 10 mL of 0.01 M

phosphate buffered saline (PBS) (EMD Chemicals, Darmstadt, Germany) to collect

remaining bacteria from the funnel surface. Using sterile forceps, individual PC filters were

folded in half two times (maintaining the sample inside the folded filter), removed, and

placed into a 1.5 mL microcentrifuge tube containing 0.2 g of acid-washed 212-300 µm glass

beads (Sigma Aldrich, St. Louis, MO). Therefore, for each sample, five microcentrifuge

tubes each containing one filter and beads were ready for elution (Figure 4.1).

Elution Techniques

Five hundred µl of eluent was added to each tube containing a filter and beads. Tubes

were vortexed for 1 min. on high speed using a VX100 vortex (Labnet, Woodbridge, NJ).

Supernatant was aspirated and combined with supernatant from the additional four filters to

acquire ~1.7 mL of concentrated sample (Figure 1). Different eluting solutions were

explored to determine best recovery. Solutions were selected based on previous success in

recovery of other microbes (Inoue et al., 2003). Elution solutions included 0.01 M PBS,

Page 116: New approaches and technologies for quantifying the impact

99

PBST-80 (0.01 M PBS and 0.01% Tween-80), PBST-20 (0.01 M PBS and 0.01% Tween-

20), and PET (sodium pyrophosphate 0.02%, Tween-80 0.01% and trisodium EDTA 0.03%).

Samples were eluted with and without glass beads. Varying tube sizes (1.5, 5, and 50 mL)

and eluant volumes (500 µl, 4 mL, and 20 mL) were also examined to determine if larger

volumes during elution would result in improved bacterial recovery (Table 3.1).

In-Line Method (ILM)

A technique for concentration and elution for larger volumes of marine water was

also developed and optimized (using only sewage spiked seawater samples and tested only

for E. coli). Three different prefilters were tested to determine recovery of cells and removal

of particulates. A 3 µm Versapor®-3000 142 mm prefilter (Pall, Port Washington, NY), a 10

µm cartridge Dispos-a-filter™ (Geotech, Denver, CO), and a 30 µm nitex mesh (Genesee

Scientific, San Diego, CA) were tested. Following prefiltration (using one of the previously

mentioned filters), 5 L of sample was concentrated on a 0.45 µm Supor® (Pall, Port

Washington, NY) filter using a polycarbonate in- line 142mm filter holder (Geotech, Denver,

CO), attached to a Masterflex Easy Load II L/S (Cole-Parmer, Vernon Hills, Illinois)

peristaltic pump, and filtered using positive pressure. After complete filtration, the filter was

removed aseptically using forceps, sliced eight times with sterile scissors while maintaining

an intact filter, folded and placed into a 50 mL Falcon™ (Beckton Dickinson, Franklin

Lakes, NJ) tube with 1 g glass beads (Sigma Aldrich, St. Louis, MO). Three mL of eluent

was added and the tube was vortexed for 1 min. on high using a VX100 vortex (Labnet,

Woodbridge, NJ). Supernatant was aspirated and placed in a separate tube for downstream

applications (Figure 2).

Page 117: New approaches and technologies for quantifying the impact

100

Particulate Analysis

Particulate size range in background seawater and filtrate from the MFM and ILM

particulate removal techniques was measured using a Cilas 1180 particle size analyzer

(Orleans, France). The Cilas required a threshold (10% obscuration) of particles present in

the sample to attain an accurate reading. Many of the samples that were prefiltered contained

particulate levels below this threshold, and were spiked with a 1 mm silica/zirconium beads

(Biospec Corp., Bartlesville, OK) to allow for sample analysis.

Enumeration of Bacteria

Enterococcus spp. and E. coli were enumerated using membrane filtration (MF)

employing the mEI (USEPA, method 1600) and mTEC (USEPA, method 1103.1) methods.

Enumeration was also conducted using a defined substrate technology (DST) approach,

which included Colilert-18 for enumeration of E. coli, and Enterolert (IDEXX® Laboratories,

Westbrook, ME) for enumeration of Enterococcus spp. For the DST approach, positive wells

of the Quantitray-2000™ were counted and total cells calculated using the Most Probable

Number (MPN) method (Hurley and Roscoe, 1983). B. thetaiotaomicron were enumerated

using a SYBR Green I method as described by Noble and Fuhrman (1998).

Calculations and Statistics

Initial cell concentrations and recoveries were determined as averages from testing

samples in triplicate, at three dilutions, and averaging both enumeration methods (DST and

MF). For example, cell concentrations were determined by enumeration of initial sample at 3

Page 118: New approaches and technologies for quantifying the impact

101

dilutions (i.e. 10, 30, and 90 mL samples). Each of the three dilutions was processed in

duplicate, and the duplicate results averaged. Averages of the three dilutions were then

combined to determine method average. This process was conducted using both the MF and

DST method. The two method averages were then combined to acquire overall average

concentration in original sample. This same process was used for enumeration of recovered

cells.

Percent recovery, total percent recovery, and concentration factor calculations are

located in Appendix B. Concentration factor was determined for comparison between the

MFM and ILM methods, but only using E. coli total recovery, and only with sewage spiked

samples. All statistics were conducted either using SPSS (version 11.0) or Microsoft EXCEL

software. Comparisons between means were conducted using a Students t-test (unpaired) to

determine significant differences. Significance was accepted at α = 0.05.

Results and Discussion

Optimal Protocols

Based on the results from all tested techniques, an optimized MFM approach for

pretreatment and recovery was developed. Large particulates were removed using multiple

250 mL conical containers centrifuged for 1 min. at 600×g. The upper 200 mL of each

centrifuged sample was aspirated and combined until the required volume was achieved (1.5

L). This sample was then concentrated by filtering equal volumes on five individual 47 mm

0.4 µm PC filters. The inside of each filter funnel was rinsed with ~10 mL 0.01 M PBS,

filters removed with sterile forceps and each filter was placed in a 1.5 mL microcentrifuge

tube containing 0.2 g of glass beads. 500 µl of 0.01 M PBS was added to each tube and

Page 119: New approaches and technologies for quantifying the impact

102

vortexed for 1 min. on high. The tube was spun briefly with a desktop centrifuge to remove

any liquid from the lid. Supernatant (concentrated sample) from each tube was aspirated and

combined for a total recovered volume of ~1.7 mL, ready for downstream applications

(Figure 4.1). Total filtration time was on average less than 15 min. and from method start to

finish was less than 45 min. per sample. Due to the difficulty of enumeration of B.

thetaiotaomicron, the MFM was first optimized for E. coli and Enterococcus spp. before

testing of recovery was conducted with B. thetaiotaomicron.

An alternative approach (ILM) was developed for larger volume (5 L) samples.

Unlike the MFM, the prefiltration for the ILM required use of a 30 µm Nitex mesh prefilter

instead of centrifugation, due to the time and labor- intensity of centrifuging large volumes.

Following prefiltration, 5 L of sample was concentrated on a 0.45 µm Supor® filter as

described in the methods. The filter was removed using forceps, sliced eight times in equal

clock-wise sections with sterile scissors while maintaining an intact filter, folded and placed

into a 50 mL tube with 1 g of glass beads. Three mL of PBST-20 (PBS + 0.01% Tween-20)

was added and tube vortexed for 1 min. on high. Supernatant was aspirated and placed in a

separate tube for downstream applications (Figure 4.2). (Note: For methods allowing greater

volumes of final concentrate, the filter can be rewashed with additional steps of PBST-20,

vortexing, and supernatant aspiration, resulting in improved recovery of ce lls, as was

previously determined by Frischer et al. (2000), when eluting bacteria from sediment). Total

filtration time ranged from 6-15 min. depending on the amount of particulates in the sample,

and from method start to finish was approximately 15-25 min. per sample.

Recovery and Method Comparison

Page 120: New approaches and technologies for quantifying the impact

103

Total recovery rates (± standard deviation) for the MFM averaged 58.6 % ± 20.0%,

48.3 % ± 25.5 %, and 27.5 % ± 19.1 % for E. coli, Enterococcus spp., and B.

thetaiotaomicron, respectively. Mean total recovery of E. coli using the ILM was 19.3 %

±1.8 % (Figure 4.4). Interestingly, we found mean total recovery in freshwater for the ILM to

be much greater (49.6% ± 13.3%, data not shown). Concentration and recovery of B.

thetaiotaomicron was especially encouraging (Figure 4.5), as these alternative microbial

water quality indicators are obligate anaerobes known to degrade quickly in the natural

environment, and we initially expected recovery rates to be very low.

Though mean recovery percentage for the MFM and ILM were quite different (>29%

for E. coli), the concentration factor was greater for the ILM due to the greater volume of

sample filtered but nearly equal final concentrate volumes (~1.75 mL final concentrate for

MFM and ~2 mL concentrate for ILM). Mean MFM concentration factor was 221 ± 5, while

mean ILM concentration factor was 317 ± 25 (Figure 4.6). When small volume

concentrations (<2 L) are required, the MFM method would be preferred due to higher

recovery. However, if larger concentration factors (>2 L) are required due to lower

sensitivity of the biosensor or monitoring assumptions (i.e. non-storm events with fewer

bacteria present in the sample or higher total suspended solids in sample resulting in quick

filter clogging), the ILM could be utilized.

Elution

Elution Solutions

Recovery of E. coli and Enterococcus spp. varied by elution solution (PBS, PBST-80,

PBST-20, and PET). PBS worked equally or better than all other elution solutions explored,

Page 121: New approaches and technologies for quantifying the impact

104

resulting in average recovery of 51.1 % ± 13.4% of E. coli (Figure 4.7) and 55.6% ± 21.5%

Enterococcus spp. Studies testing different eluting solutions have found different levels of

recovery for other microbes, including improved recovery of Cryptosporidium oocysts using

PET (Inoue et al., 2003). Our results indicate that PBS had equal or greater recovery for the

MFM. However, during optimization of the ILM, PBST (Tween-20 0.01%) increased

recovery by 1.7-6.5 %. This improvement in recovery could be due to the larger filter size

(142 mm) used in the ILM, and/or the differences in filter material between the two methods,

resulting in different optimized elution solutions.

Glass Beads, Tubes, and Elution Volumes

The use of glass beads and different tube sizes/elution volumes was tested to

determine if the presence of a physical scrubbing agent (beads) and larger space with greater

eluting solution volumes would improve recovery. Inclusion of 0.2 g of 212-300 µm glass

beads as an eluting material resulted in significantly better recoveries (t-test, p<0.015) of

cells when compared to recoveries from samples without beads. Combined with a 1 min.

vortex, the inclusion of beads increased recovery by as much as 4 times. The use of 1.5 mL

microcentrifuge tubes and 500 µl of elution solution per filter optimized recovery for the

MFM. Comparison of five mL tubes with elution volumes of 4 mL and 50 mL tubes with 20

mL of elution volume did not improve recovery. For the ILM, the use of 3 mL of elution

solution in a 50 mL Falcon tube with 1 g of beads resulted in the highest recovery of E. coli.

A smaller tube size for the ILM was not an option due to size of the filter utilized (142 mm)

and no more than 3 mL of eluent was used to maintain concentration of target cells.

Page 122: New approaches and technologies for quantifying the impact

105

Particulate Removal

Prefiltration

Raw seawater particulate diameter ranged from 0.2 - 140 µm, with an average of 29.7

µm, and a median of 36.6 µm (Figure 4.3A). Various pore sized nylon net filters were

examined to determine bacterial loss and particulate removal. Nylon net filters resulted in

bacterial loss inversely related to pore size, with mean loss of 1.00% ± 0.58%, 9.90% ±

15.5%, and 49.7% ± 14.3% for 100, 60, and 30 µm nylon net filters, respectively, of E. coli

at multiple concentrations between 102- 103 MPN/ 100 mL (Table 4.2). Although 100 µm

and 60 µm pore size nylon net filters retained little target bacteria, particles >30 µm remained

in the sample. When other filter sizes and types were tested for use with the alternative ILM,

smaller pore sizes resulted in greater loss of bacteria. Prefiltration of 5 L seawater samples

through the 3 µm Versapor-3000 and 10 µm Dispos-a-filter resulted in mean E. coli loss of

96.5 % ± 4.43% and 96.2 % ± 1.41%, respectively. These prefilters are not viable for

prefiltration if attempting to retain bacterial cells; however, they were very effective at

particulate removal (Figure 4.3B). Prefiltration through the 30 µm Nitex mesh had a mean E.

coli loss of 41.2 % ± 10.5%, resulting in the least bacterial loss with large particulate removal

(Figure 4.3C).

Samples with higher particulate concentrations (near shore wave wash vs. seawater

collected beyond surf break, marine water vs. freshwater) nearly always had greater loss of

bacteria when prefiltered. For instance, wave wash samples spiked with 103 E. coli per 100

mL and filtered through 30 µm nylon net filters had losses ranging from 37.7-71.3 %, while

non wave wash samples had losses ranging from 23.3% - 61.1%. These results are similar to

previous work which found high loss of E. coli and Enterococcus spp. in marine water when

Page 123: New approaches and technologies for quantifying the impact

106

prefiltered using a 20 µm pore size filter (Bushon et al., 2009). Interestingly, we found that

particulate size in sample did not always match the expected size class after filtration through

a known pore size. For example, though prefiltration through a 30 µm Nitex mesh filter

removed many of the particulates greater than 30 µm in seawater samples, some larger

particulates (30-60 µm) were present in the filtered sample (Figure 4.3C). This was likely

due to the large volumes (5 L) of sample filtered, the potential for particulates to “squeeze”

through pores after mass aggregation on the filter surface, the irregular size of particulates

(allowing rod like shapes to pass through), and/or defects in the filter membrane. Another

possibility was that the apparent occurrence of particles greater than the expected pore size

was an artifact of the measurement device, or a result of the spiked 1 mm silica/zirconium

beads used to achieve obscuration. Though the device was rinsed between samples, these

results might indicate the limitations of the Cilas 1180 for samples with low particulate

concentrations.

Centrifugation

Centrifuging techniques were explored for particulate removal as an alternative to

prefiltration of the sample. A large portion of particulates were removed when centrifuged

for 60 sec. at 600×g. However, a portion of particulates in the 10 - 60 µm size class remained

in the sample (Figure 4.3D). Optimal speeds for recovery of E. coli and Enterococcus spp.

occurred at 600×g for 60 sec. Testing of centrifugation using seawater spiked with cultured

bacteria at multiple concentrations between 102 - 103 MPN/100 mL, resulted in recoveries of

83.6% ± 12.0% and 82.1% ± 10.7% for E. coli and Enterococcus spp., respectively. Though

previous research indicates bacteria remain in suspension at speeds less than 1000×g

Page 124: New approaches and technologies for quantifying the impact

107

(Stevens and Jaykus, 2004), we found increased bacterial loss with increased centrifugation

time (60 sec. to 90 sec.) at 600×g. This is likely due to particulate removal with attached cells

and/or cell aggregate formation and settling. Interestingly, Enterococcus spp. cell recovery

occasionally exceeded 100% after centrifuging the sample, which could be due to

disaggregation of cell chains and biofilms commonly formed by these bacteria (Moellering,

1992). Similar centrifuging techniques (700×g for 2 min.) were recommended for removal of

bacteria from sediment slurries (Frischer et al., 2000). When compared to prefiltration,

centrifuging resulted in equal or greater cell recovery than any of the p refilters, while

removing many of the larger particulates (Figure 4.3D). However, this method for particulate

removal is only recommended for volumes less than 1.5 L, as it was time consuming and

cumbersome with increasing volume.

Sonication

Sonication before centrifugation and prefiltration showed variable results based on

the particulate removal method applied. Sonication for 15 sec. before nylon net filtration

resulted in decreased recovery (~10 % less than non-sonicated samples) of cultured E. coli

cells at multiple concentrations between 102 - 103 MPN/100 mL, and sonicated vs.

unsonicated samples were not significantly different (t-test, p=0.45). Samples sonicated for

15 sec. prior to centrifugation at 600×g showed increased recovery (~11 % greater than non-

sonicated samples) of E. coli cells at concentrations of 102 - 103 MPN/100 mL, but sonicated

vs. unsonicated samples were not significantly different (t-test, p=0.34). Previous work has

shown sonication to be an effective tool for separating microbes from particulates and filter

membranes (Leskinen and Lin, 2008; Anderson et al., 2003; Jones et al., 2000). However, we

Page 125: New approaches and technologies for quantifying the impact

108

found decreased or moderate improvement in recovery between sonicated and unsonicated

samples. This could be due to the use of cultured bacteria, the lack of bacterial attachment in

the time between inoculation and processing, or instrument differences (i.e. sonication bath

vs. sonication probe). Due to the variance and little improvement in recovery, sample

sonication was not incorporated into the optimized methods. However, this technique should

be explored in greater depth to determine if effective.

Bacterial Loss Factors

Ultimately, the loss of bacteria during pretreatment, concentration, and elution will

depend on three main factors: (1) the size and amount of particles in the sample and percent

of bacteria attached to these materials, (2) bacteria remaining on filters and elution equipment

surfaces during concentration, and (3) cell lysis during processing. An average of 40% of E.

coli and Enterococcus spp. cells were determined attached to particulates in estuarine waters

in North Carolina (Fries et al., 2006), while bacterial attachment to zooplankton (Maugeri et

al., 2004), and Bacteroides attachment and loss during particulate removal (Lucena et al.,

1995) have been documented. Recovery success of Cryptosporidium and Giardia has been

shown to be dependent on particulate concentration (Hsu and Huang, 2000). Our results

confirm these findings, as samples collected ~5 m offshore outside the wash zone (lower

particulate) vs. near shore wave wash, and freshwater (lower particulate) vs. marine water

samples, showed decreased loss of bacteria during the pretreatment and concentration

processes. Bacteria left on filters and concentration equipment (filter funnels,

microcentrifuge tubes, tubing, glass beads) can be minimized through washing of equipment

with a neutral pH buffer such as PBS following filtration and using low adherent materials

Page 126: New approaches and technologies for quantifying the impact

109

(filter funnels, graduated cylinders, etc.). Cell lysis should be minor during processing, as

previous research has shown that cells can survive in adverse conditions for short periods of

time (Stevens and Jaykus, 2004) but may become more influential when working with

obligate anaerobes such as Bacteroides spp., or bacteria with fragile membranes.

Conclusion

Both MFM and ILM were successful at concentrating E. coli, Enterococcus spp. and

B. thetaiotaomicron from seawater. These approaches have been successfully combined with

additional sample preparation techniques such as immunomagnetic separation and/or short

term enrichment. They could assist in overcoming sensitivity issues of some developing

biosensor devices. Both the MFM and ILM could also be applied to other microbial targets

and freshwater samples. Future modifications to these methods could include more refined

use of sonication, and the use of desorptive agents to remove bacteria from particulates.

In summary:

Both methods removed many of the larger particulates (>30 m), but smaller

particulates(< 30 m) remained in seawater

Processing time was <45 min. for MFM and <15 min. for the ILM but will depend on

particulate concentration

These methods are portable to water quality laboratories without a need for expensive

equipment or complex steps

ILM had lower recovery but a higher concentration factor (vs. MFM), and was more

appropriate for concentration/elution of larger (2-5 L) volumes and less contaminated

waters

Page 127: New approaches and technologies for quantifying the impact

110

Both methods show potential for use for sample processing when concentrated, intact,

and metabolically active cells are required

Page 128: New approaches and technologies for quantifying the impact

111

MFM Technique Goal

Variable

tested Conditions

Prefiltration with nylon

net filter Particulate removal Pore size 30, 60, 100 µm

Centrifugation Particulate removal

Centrifuge

speed 200, 400, 600, 800 ×g

Centrifuge time 60, 90 sec.

Sonication Dissociation of bacteria

from part icles Duration 15 sec.

Filtration onto 0.4 µm

polycarbonate filter Concentration of bacteria Diameter 47 mm

Elution

Elution of filtered bacteria

Solution 0.01 M PBS, PBST-80, PBST-20, PET

Physical agent With or without 212-300 µm glass beads

Volume 0.5, 4, 20 mL

Tube size 1.5, 5, 50 mL

ILM Technique Goal

Variable

tested Conditions

Filtration onto 0.45 µm

polyethersulfone filter Concentration of Bacteria Diameter 142 mm

Prefiltration with filter Particulate removal Pore size

3 µm (Versapor-3000)

10 µm (Geotech, Dispos-a-filter)

30 µm (Nitex, mesh)

Table 4.1. Different methods and treatments tested for bacterial concentration, elution, and particulate removal

111

Page 129: New approaches and technologies for quantifying the impact

112

Table 4.2. Recovery of E. coli and particulate removal efficiency for seawater spiked with E. coli concentrations of 103

MPN/100 mL

Pretreatment: Filter

Size and Type,

Centrifugation

Conditions

Manufacturer Mean E. coli

Recovery

Range of

Recovery

Particulate

Removal

Efficiency

100 µm Nylon Net Millipore 99.2% 99.0-100% Poor

60 µm Nylon Net Millipore 90.1% 79.9-100% Poor

30 µm Nylon Net Millipore 50.3% 41.8-62.2% Moderate

30 µm Nitex Genesee Scientific 69.2% 62.3-81.3% Moderate

10 µm Dispos-a-filter Geotech 3.79% 2.80-4.79% Good

3 µm Versapor-3000 Pall 2.81% 0.00-9.37% Good

Centrifuging at 600×g N/A 83.6% 67.9-93.6% Moderate/Good

112

Page 130: New approaches and technologies for quantifying the impact

113

Figure 4.1. Diagram of the Multi- filter Method (MFM) for concentration and elution of

bacteria for sample volumes less than 1.5 L

Page 131: New approaches and technologies for quantifying the impact

114

Figure 4.2. Diagram of the In- line Method (ILM) for concentration and elution of bacteria for

sample volumes less than 5 L

Page 132: New approaches and technologies for quantifying the impact

115

Figure 4.3. Seawater particulate diameter distribution composition for A. Raw seawater

B. Nitex filtered seawater C. Geotech Dispos-a-filter seawater after filtration D. Seawater

centrifuged for 1 min. at 600×g and top 200 mL aspirated. (1) Represents artificial spike with

a known grain size to achieve obscuration of the CILAS 1180 particle size analyzer, and (2)

represent actual remaining particles in sample after applied filtration or centrifuging

treatment

Page 133: New approaches and technologies for quantifying the impact

116

Figure 4.4. Mean total recovery (due to prefiltration, concentration, and elution step)

percentage using optimized method for E. coli (EC), Enterococcus spp. (ENT), and B.

thetaiotaomicron (BTIM) for Multi- filter Method (MFM) vs. In-Line Filter Method (ILM)

(E. coli only). Error bars represent standard deviation

Page 134: New approaches and technologies for quantifying the impact

117

Figure 4.5. Concentration of Bacteroides thetaiotaomicron using Sybr Green bacterial

staining enumeration method (A) 5 mL seawater (SW) before concentration (B) 100 µl SW

after concentration

Page 135: New approaches and technologies for quantifying the impact

118

0

100

200

300

400

500

600

700

800

MFM ILM

Co

ncen

trati

on

Facto

r

Colilert-18

mTEC

Average

Figure 4.6. Differences in concentration factor of MFM vs. ILM for E. coli

Page 136: New approaches and technologies for quantifying the impact

119

0

10

20

30

40

50

60

PBS PBST PET

Eluent

E.

coli

Perc

en

t R

eco

very

Beads

No Beads

Figure 4.7. Percent recovery of E. coli from seawater using Multi- filter Method (MFM) with

212-300 um glass beads vs. no beads, and elution with PBS, PBST-80, and PET solutions.

Error bars represent standard deviation

Page 137: New approaches and technologies for quantifying the impact

120

REFERENCES

Ahmed, W., Goonetilleke, A., Powell, D., Chauhan, C., Gardner, T. (2009). Comparison of molecular markers to detect fresh sewage in environmental waters. Water Research, 43(19),

4908-4917.

Anderson, K.L., Whitlock, J.E., Harwood, V.J. (2005). Persistence and differential survival of fecal indicator bacteria in subtropical waters and sediments. Applied and Environmental

Microbiology, 71(6), 3041-3048.

Barthel, K.G., Schneider, G., Gradinger, R., Lenz. J. (1989). Concentration of live pico and nanoplankton by means of tangential flow filtration. Journal of Plankton Research, 11, 1213-

1221.

Baums, I.B., Goodwin, K.D., Kiesling, T., Wanless, D., Diaz, M.R., Fell, J.W. (2007). Luminex detection of fecal indicators in river samples, marine recreational water, and beach

sand. Marine Pollution Bulletin, 54(5), 521–536.

Bernhard, A.E., Goyard, T., Simonich, M.T., Field, K.G. (2003). Application of a rapid method for identifying fecal pollution sources in a multi-use estuary. Water Research, 37,

909–913.

Bushon, R.N., Brady, A.M., Likirdopulos, C.A., Cireddu, J.V. (2009). Rapid detection of Escherichia coli and enterococci in recreational water using an immunomagnetic separation/

adenosine triphosphate technique. Journal of Applied Microbiology, 106(2), 432-441.

Cabelli, V.J. (1982). Microbial indicator systems for assessing water quality. Journal of Microbiology, 48(6), 613-618.

Carson, C.A., Christiansen, J.M., Yampara-Iquise, H., Benson, V.W., Baffaut, C., Davis,

J.V., Broz, R.R., Kurtz, W.B., Rogers, W.M., Fales, W.H. (2005). Specificity of a Bacteroides thetaiotaomicron marker for human feces. Applied and Environmental

Microbiology, 71(8), 4945-4949.

Converse, R.R., Blackwood, A.D., Kirs, M., Griffith, J.F., Noble, R.T. (2009). Rapid QPCR-based assay for fecal Bacteroides spp. as a tool for assessing fecal contamination in

recreational waters. Water Research, 43(19), 4828-4837.

Page 138: New approaches and technologies for quantifying the impact

121

Dick, L.K., Field, K.G. (2004). Rapid estimation of numbers of fecal Bacteroidetes by use of

a quantitative PCR assay for 16S rRNA genes. Applied and Environmental Microbiology, 70(9), 5695-5697.

Drasar, B.S. (2003). Microbial flora of the gut. In: Mara, D., Horan, N. (Eds.), Handbook of

Water and Wastewater Microbiology. Academic Press, San Diego, pp. 99-104.

Fries, S.J., Characklis, G.S., Noble, R.T. (2006). Attachment of fecal indicator bacteria to particles in the Neuse River Estuary, NC. Journal of Environmental Engineering, 132(10),

1338-1345.

Fiksdal, L., Maki, J.S., LaCroix, S.J., Staley, J.T. (1985). Survival and detection of Bacteroides spp., prospective indicator bacteria. Applied and Environmental Microbiology,

49(1), 148-150.

Frischer, M.E., Danforth, J.M., Newton Healy, M.A., Saunders, F.M. (2000). Whole-Cell versus total RNA extraction for analysis of microbial community structure with 16S rRNA-

targeted oligonucleotide probes in salt marsh sediments. Applied and Environmental Microbiology, 66(7), 3037–3043.

Garcıa-Aljaro, C., Munoz-Berbel, X.A., Jenkins, T.A., Blanch, A.R., Munoz, F.X. (2008).

Surface plasmon resonance assay for the real-time monitoring of somatic coliphages in wastewaters. Applied and Environmental Microbiology, 74(13), 4054-4058.

Hill, V.R., Polaczyk, A.L., Hahn, D., Narayanan, J., Cromeans, T.L., Roberts, J.M.,

Amburgey, J.E. (2005). Development of a rapid method for simultaneous recovery of diverse microbes in drinking water by ultrafiltration with sodium polyphosphate and surfactants.

Applied and Environmental Microbiology, 71(11), 6878–6884.

Hoa, X.D., Kirk, A.G., Tabrizian, M. (2007). Towards integrated and sensitive surface plasmon resonance biosensors: a review of recent progress. Biosensors and Bioelectronics,

23, 151–160.

Hsu, B., Huang, C. (2000). Recovery of Giardia and Cryptosporidium from water by various concentration, elution, and purification techniques. Journal of Environmental Quality, 29,

1587-1593.

Page 139: New approaches and technologies for quantifying the impact

122

Hurley, M.A., Roscoe, M.E. (1983). Automated statistical-analysis of microbial enumeration

by dilution series. Journal of Applied Bacteriology, 55(1), 159-164.

Inoue, M., Rai, S.B., Oda, T., Kimura, K., Nakanishi, M., Hotta, H., Uga, S. (2003). A new filter eluting solution that facilitates improved recovery of Cryptosporidium oocysts in water.

Journal of Microbiological Methods, 55, 679-686.

Ivnitski, D., Abdel-Hamid, I., Atanasov P., Wilkins, E. (1999). Review: Biosensors for detection of pathogenic bacteria. Biosensors and Bioelectronics, 14, 599–624.

Jones, T.H., Brassard, J., Johns, W.M., Gagne, M.J. (2009). The effect of pre-treatment and sonication of centrifugal ultrafiltration devices on virus recovery. Journal of Virological Methods, 161(2), 199-204.

Kreader, C.A. (1998). Persistence of PCR-detectable Bacteroides distasonis from human feces in river water. Applied and Environmental Microbiology, 64(10), 4103-4105.

Lavender, J.S., Kinzelman, J.L. (2009). A cross comparison of QPCR to agar-based or

defined substrate test methods for the determination of Escherichia coli and enterococci in municipal water quality monitoring programs. Water Research, 43, 4967-4979.

Layton, A., McKay, L., Williams, D., Garrett, V., Gentry, R., Sayler, G. (2006). Development

of Bacteroides 16S rRNA gene TaqMan-based real-time PCR assays for estimation of total, human, and bovine fecal pollution in water. Applied and Environmental Microbiology, 72(6), 4214-4224.

Leskinen, S.D., Lim, D.V. (2008). Rapid ultrafiltration concentration and biosensor detection of enterococci from large volumes of Florida recreational water. Applied and Environmental Microbiology, 74(15), 4792-4798.

Lucena, F., Muniesa, M., Puig, A., Araujo, R., Jofre, J. (1995). Simple concentration method for bacteriophages of Bacteroides fragilis in drinking water. Journal of Virological Methods, 54, 121-130.

Page 140: New approaches and technologies for quantifying the impact

123

Maugeri, T.L., Carbone, M., Fera, M.T., Irrera, G.P., Gugliandolo, C. (2004). Distribution of potentially pathogenic bacteria as free living and plankton associated in a marine coastal

zone. Journal of Applied and Microbiology, 97, 354–36.

Marion, J.W., Lee, J., Lemeshow, S., Buckley, T.J. (2010). Association of gastrointestinal illness and recreational water exposure at an inland U.S. beach. Water Research, 44(16),

4796-4804.

Moellering, R.C. (1992). Emergence of Enterococcus as a significant pathogen. Clinical Infectious Disease, 14(6), 1173-1176.

Morales-Morales, H.A., Vidal, G., Olszewski, J., Rock, C.M., Dasgupta, D., Oshima, K.H., Smith, G.B. (2003). Optimization of a reusable hollow-fiber ultrafilter for simultaneous concentration of enteric bacteria, protozoa, and viruses from water. Applied Environmental

Microbiology, 69(7), 4098-4102.

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

Noble, R.T., Weisberg, S.B. (2005). A review of technologies for rapid detection of bacteria in recreational waters. Journal of Water and Health, 3, 381-392.

Noble, R.T., Griffith, J.F., Blackwood, A.D., Fuhrman, J.A., Gregory, J.B., Hernandez, X.,

Liang, X., Berra, A.A., Schiff, K. (2006). Multitiered approach using quantitative PCR to track sources of fecal pollution affecting Santa Monica Bay, California. Applied and Environmental Microbiology, 72(2), 1604-1612.

Reid, D.E., Adlam, C. (1974). Large-scale harvesting and concentration of bacteria by tangential flow filtration. Journal of Applied Bacteriology, 41(2), 321-324.

Savichtcheva, O., Okayama, N., Okabe, S. (2007). Relationships between Bacteroides 16S

rRNA genetic markers and presence of bacterial enteric pathogens and conventional fecal indicators. Water Research, 41, 3615-3628.

Shiyanovskii, S.V., Schneider, T., Smalyukh, I.I., Woolverton, C.J., Niehaus, G.D., Doane,

K.J. (2005). Lyotropic chromonic liquid crystals for biological sensing applications. Molecular Crystals and Liquid Crystals, 434, 259-270.

Page 141: New approaches and technologies for quantifying the impact

124

Signoretto, C., Burlacchini, G., del Mar Lleò, M., Pruzzo, C., Zampini, M., Pane, L., Franzini, G., Canepari, P. (2004). Adhesion of Enterococcus faecalis in the nonculturable

state to plankton is the main mechanism responsible for persistence of this bacterium in both lake and seawater. Applied and Environmental Microbiology, 70(11), 6892-6896.

Stevens, K.A., Jaykus, L. (2004). Bacterial separation and concentration from complex

matrices: A review. Critical Reviews in Microbiology, 30(1), 7-24.

Tsai, Y., Sobsey, M.D., Sangermano, L.R., Palmer, C.J. (1993). Simple method of concentrating Enteroviruses and Hepatitis A Virus from sewage and ocean water for rapid

detection by reverse transcriptase-polymerase chain reaction. Applied Environmental Microbiology, 59(10), 3488-3491.

Wade, T.J., Pai, N., Eisenberg, J.N.S., Colford, J.M. (2003). Do U.S. Environmental

Protection Agency water quality guidelines for recreational waters prevent gastrointestinal illness? A systematic review and meta-analysis. Environmental Health Perspectives, 111(8), 1102-1109.

Wade, T.J., Calderon, R.L., Sams, E., Beach, M., Brenner, K.P., Williams, A.H., Dufour, A.P. (2006). Rapidly Measured Indicators of Recreational Water Quality Are Predictive of Swimming-Associated Gastrointestinal Illness. Environmental Health

Perspectives, 114(1), 24-28.

Wade, T.J., Calderon, R.L., Brenner, K.P., Sams, E., Beach, M., Haugland, R., Wymer, L., Dufour, A.P. (2008). High sensitivity of children to swimming-associated

gastrointestinal illness: results using a rapid assay of recreational water quality. Epidemiology, 19(3), 375-383.

Zhang, R., Liu, B., Lau, S.C.K., Ki., J., Qian., P. (2007). Particle-attached and free- living

bacterial communities in a contrasting marine environment: Victoria Harbor, Hong Kong. FEMS Microbiology and Ecology, 61(3), 496-508.

Page 142: New approaches and technologies for quantifying the impact

CHAPTER 5

Optimized capture and concentration of Enterococcus spp. and E. coli from water using

Immunomagnetic Separation (IMS)

Introduction

Current and emerging technologies for bacterial detection in matrices such as food,

drinking water, shellfish, and recreational waters often require sample pre-preparation

(including concentration and purification) to overcome issues prior to enumeration (Porter

and Pickup, 1999; Stevens and Jaykus, 2004; Bhunia, 2008). Several techniques have been

proposed and tested for pre-sample preparation in order to accomplish concentration and

purification including enrichment of bacterial growth (Cudjoe and Krona, 1997), filtration

followed by elution (Leskinen and Lim, 2008), and sonication (Frischer et al., 2000).

However, given each of these approaches, methods to capture target cells away from non-

target biotic and abiotic particles are often still required. This is especially true of the pre-

sample processing for biosensor technologies dependent on aggregate formation, including

the Crystal Diagnostics (CDx) prototype technology, which will include in quantification

particles that exceed a size threshold (~3 um), treating them as bacterial aggregates and

resulting in false positive responses.

Immunomagnetic separation (IMS) has been proposed as a means to specifically

capture bacteria while cleaning unwanted debris from complex sample matrices including

water (Porter and Pickup, 1999; Lee and Deininger, 2004; Lee et al., 2010), food (Safarik et

Page 143: New approaches and technologies for quantifying the impact

126

al., 1995; Nou et al., 2006) and clinical samples (Olsvik et al., 1994). Immunomagnetic

separation (IMS) utilizes paramagnetic beads, conjugated to a specific antibody (polyclonal

or monoclonal) for capture of an intended target (i.e. cell, protein, DNA). Conjugation of

antibody (Ab) can occur using many different methods, but the streptavidin/biotin (Tu et al.,

2009) and tosylactivated covalent (Lee et al., 2010) are two common binding methods. After

magnetic beads are conjugated with Ab specific to a target (e.g. E. coli or Enterococcus

spp.), sample and beads are mixed and Ab binds to a target cell surface antigen. Beads are

magnetically separated from the supernatant, and washed to remove unwanted debris

(environmental particles, proteins, nontarget cells) in the sample (Porter and Pickup, 1999)

(Figure 5.1). In theory, the resulting sample should only consist of the magnetic beads bound

to the intended target antigen, assuming no cross-reactivity with non-target cells (non-

specific binding). Previous research with IMS has shown variable recoveries from <1% to

>85% for Cryptosporidium and Giardia from environmental waters (DiGiorgio et al., 2002;

Feng et al., 2003). However, few published studies exist examining IMS recovery of fecal

indicator bacteria (FIB) from environmental waters.

Though IMS has been optimized for certain uses (i.e. capture of E. coli O157:H7

from ground beef, Cryptosporidium parvum from swimming pool water), there is little

information available on an optimized method for IMS in environmental water samples, in

particular, capture of E. coli and Enterococcus spp. Factors such as Ab conjugation approach,

methods for mixing the bead with sample (incubation times, incubation shaking vs.

centrifugation, bead to bacteria ratios), and washing procedures need to be optimized for

each application (Neurauter et al., 2007) . Though, incubation of sample with beads usually

require an incubation with gentle mixing and a proposed bead to bacteria ratio of at least 4:1

Page 144: New approaches and technologies for quantifying the impact

127

(Neurauter et al., 2007) neither bead concentrations nor centrifugation during incubation

have been optimized. Finally, washing beads post capture of target is recommended to clean

any unwanted debris from the sample (Porter and Pickup, 1999). However, it is unclear how

recovery is affected by these washing procedures, and the number of washes preformed.

Chemical dispersants, sonication and differences in magnetic separation devices

could also influence target cell capture and purification results. Chemical dispersants have

previously been utilized to increase freely available bacteria and disassociate bacteria from

particles (McDaniel and Capone, 1985; Frischer et al., 2000; Stevens and Jaykus, 2004).

Certain chemical dispersants are intended to release bacteria bound to surfaces and reduce

cell aggregation. Some examples of chemical dispersants include Triton X-100 (McDaniel

and Capone, 1985) methanol (Lanau et al., 2005), and trypsin (Garren and Azam, 2010).

Naturally occurring plant flavonoids are also known for their natural ability to disrupt

bacterial biofilms (Duarte et al., 2006; Vikram et al., 2009). Flavonoids or other natural plant

acids could assist in bacterial dissociation from surfaces while remaining non-toxic to target

bacteria. Sonication has previously shown increases in non-particle bound bacterial

availability (Tso et al., 1997; Frischer et al., 2000; Falcioni et al., 2006). Though sonication

had no significant difference in bacterial recovery when used previous to filtration (see

Chapter 4), a new sonication probe and altered method could prove successful in

disassociating bacteria from particles before the IMS, resulting in increased recovery.

Finally, previous research by Nou et al. (2006) indicated that use of a magnetic PickPen

could improve recovery of magnetic particles and reduction of unwanted particles over

traditional magnetic block separation.

Page 145: New approaches and technologies for quantifying the impact

128

The goal of this research was to conduct the preliminary work necessary to fully

optimize a process for capture of E. coli and Enterococcus spp. from recreational waters

using commercially available antibodies. Magnetic beads with both streptavidin/biotin

conjugation as well as covalent tosylactivated bead conjugation were tested. A range of

variables were tested in relation to target cell recovery, including incubation time, bead and

bacteria concentration, centrifugation vs. shaker incubation, and washing (Table 5.1). The

combined effects of certain variables were also assessed. The potential for increased recovery

using chemical additions and sonication previous to IMS was also briefly explored. Finally a

PickPen® (Bio-Nobile) magnetic probe was tested against typical magnetic block separation

to determine if differences in recovery occurred. These testing procedures were developed in

order to determine if the IMS procedure was practical and timely for concentration of target

cells for biosensors geared toward rapid quantification of FIB in water samples.

Methods

General Method

Seawater samples were collected from Fort Macon State Park, Atlantic Beach, NC in

sterile 10L carboys. Marine water samples were prefiltered through a 30 µm pore size Nitex

filter (Genesee Scientific, San Diego, CA) to remove large particulates similar to the methods

of Lee and Deninger (2004). The filtered sample was concentrated using either filtration

methods described in Chapter 4, or using Elutrasep Cryptonite™ hollow fiber filters (HFF)

(Elutrasep, Inc., Canton, GA). Cultured bacteria (E. coli ATCC 25922, Enterococcus faecalis

ATCC 29212, Enterococcus faecium ATCC 8459), were grown using previously determined

optimal conditions (E. coli in tryptic soy broth (TSB) and incubated at 35ºC; Enterococcus

Page 146: New approaches and technologies for quantifying the impact

129

spp. in brain heart infusion (BHI) and incubated at 37ºC). E. coIi and Enterococcus spp.

cultures were transferred after initial incubation (18-24 hr) and re-grown until exponential

growth phase (determined using a PharmaSpec 1700 spectrophotometer (Shimadzu Scientific

Instruments, Columbia, MD)) and cultured bacteria (in growth media) was immediately

inoculated into seawater to produce concentrations from 1.56 × 102 to 7.89 × 102 mL, E.

coli or Enterococcus spp. Raw wastewater influent was collected from the Morehead City,

NC wastewater treatment plant, and stored at 4° C, until sewage samples were inoculated.

Concentrated samples (1 mL) were mixed with Ab conjugated magnetic beads and

mixed for 15-60 min. using a Lab Companion SI-300 shaker incubator (Jeio Tech Inc.,

Woodburn, MA) to allow Ab/cell attachment. Samples were magnetically separated using a

Dynal® MPC-S magnetic particle separator (Invitrogen, Carlsbad, CA) and washed 1-5

times with PBS or PBST (PBS+ 0.01% Tween-20). Beads were resuspended in PBS, and

enumerated.

Magnetic Bead Preparation

Biotin/Streptavidin Beads

DynaBeads® M-280 Streptavidin (Invitrogen, Carlsbad, CA) coated beads were

tested with both biotin conjugated and unconjugated polyclonal Enterococcus spp. Ab

(ViroStat Inc., Portland, ME). Unconjugated Ab were biotinylated using a Pierce EZ-link

Sulfo-NHS-Biotin biotinylation kit (Thermo Fisher Scientific, Rockford, IL). Beads were

prewashed 3 times with PBS to remove sodium azide, and incubated at room temperature for

60 min. Incubation mixing speeds were varied at 25, 60, and 115 rpm to determine best

conjugation conditions. Beads were washed 4 times with 0.1% PBSB (Phosphate Buffered

Page 147: New approaches and technologies for quantifying the impact

130

Saline pH 7.4, and 0.1% w/v Bovine Serum Albumen) to remove excess Ab and occupy

unused binding sites, and stored at 4o C until ready for use. Experiments were run using

seawater spiked with 1) sewage, or 2) E. faecium, and E. faecalis at ~103 cells per mL.

Similar experiments were run with inoculated PBS at concentrations from 8.10 × 101 to 7.89

× 102 to determine if seawater was inhibiting capture of target cells. Finally, two tidal creek

samples from Marine Corps Base Camp Lejeune were tested to determine capture efficiency

of a real world, but highly complex “environmental” sample using the IMS method.

Tosylactivated Beads

DynaBead® M-280 tosylactivated beads were prepared with polyclonal

Enterococcus Ab and polyclonal E. coli Ab against “O” and “K” antigenic serotypes of E.

coli (Meridian Life Sciences Inc., Saco, ME). Invitrogen instructions were followed for

bead/Ab conjugation except that Buffer B was used (0.1M Na-phosphate buffer at 7.4 pH)

instead of Buffer A (Borate buffer). Buffers C, D, and E were used as described. During

covalent binding of Meridian E. coli or Enterococcus Ab, 355 µl of beads were selected for

each tube (7.10 × 108 beads per tube (10.7 mg) and combined with 1 mL of desired Ab (250

µg/ mL) diluted in Buffer B, and 100 µl Buffer C. Bead/Ab solution was mixed for 1 minute

by medium vortexing, and microcentrifuge tubes placed in a Lab Companion 300-SI (Jeio

Tech Inc., Woodburn, MA) incubator shaker at 25 rpm, for 18-24 hr at 37o C. When present

in the lab, tubes were occasionally inverted to increase beads in suspension, and re-

homogenate sample.

Optimized Sample Mixing and Washing

Page 148: New approaches and technologies for quantifying the impact

131

Bead concentration, incubation time and treatment (shaker incubation vs.

centrifugation), and number of washes to recover the greatest E. coli and Enterococcus spp.

from samples were tested, while minimizing unwanted debris. Dynabead M-280

tosylactivated magnetic beads (Invitrogen, Carlsbad, CA) bound to Meridian E. coli and

Enterococcus Ab (Meridian Life Sciences, Saco, ME) were tested in concentrations ranging

from 7.1 ×106 to 3.55 × 107 beads per sample. A bead to target ratio of at least 4:1 is

recommended (Neurauter et al., 2007), and was exceeded for all experiments. Mixing with

sample was explored using different incubation times and techniques. Incubation times were

tested at 10, 20, 30 and 60 min. using shaker incubation at 37° C, and 60 rpm. Centrifugation

of the beads with sample was also tested as a more rapid method for target binding. Initial

centrifugation tests were run at 10 and 20 min incubations at 800 × g using an Eppendorf

microfuge (Eppendorf North America, Hauppauge, NY) at ambient temperature.

Experiments were later conducted with centrifugation for 5 min., at 1100 ×g and 40°C using

a Mikro 120 centrifuge (Hettich, Beverly, MA). Two to three washes of the magnetic beads

are recommend after incubation with sample to remove unwanted debris. Washing was

conducted using PBST (PBS+0.01% Tween-20) or PBS, and 1-3 washes were enumerated to

determine bacterial loss and remaining particulate content.

Chemical Dispersant Additions

Natural plant acids and flavonoids were tested to determine if these chemical

additions could assist in dispersing/disassociating bacteria from particles. Quercetin, catechin

hydrate, epigallocatechin gallate, gallic acid, ellagic acid, and tannic acid (Sigma-Aldrich

Corp., St. Louis, MO) were tested at different concentrations (0.1 Molar (M) to 0.001 M) and

Page 149: New approaches and technologies for quantifying the impact

132

incubation times (5, 10, 15, 20, 30, and 60 min). Particulate amended marine water was

spiked with sewage, to bring raw concentration of E. coli to 102 per mL. One mL subsamples

of the larger homogenized 100 mL sample were distributed to 2 mL tubes, and 1 mL of

chemical dispersant was added. Samples were placed on a multi-purpose Thermo Scientific

Lab Rotator (Ashville, NC) at high speed and incubated @37 C, for varying times. After

incubation, samples were vortexed for 10 sec. on high speed using a Labnet VX100 Vortex

Mixer (Labnet International Inc., Woodbridge, NJ) and tubes centrifuged at 500 × g for 60

sec. using an Eppendorf microfuge (Eppendorf North America, Hauppauge, NY). Bacteria

remaining in supernatant after centrifugation were assumed to be non-particle attached and

compared to a control with no chemical addition to determine if increased bacteria in

suspension occurred.

Sonication

Sonication was tested to determine if treatment increased freely available cells in

sample suspension. Particulate amended freshwater was spiked with sewage and refrigerated

overnight to allow for particulate bacteria interaction. Previous to sonication, samples were

vortexed for 30 second using Labnet VX100 Vortex on high to homogenize sample. A

vortex/centrifuged sample with no sonication was enumerated, as centrifuge control.

Sonication was conducted at varying times (5, 10, 30 sec) and intensities (4, 7, 10 watts)

using a Misonix Microson XL-2000 variable sonicator with a P-1 1/8” probe (Qsonica LLC.,

Newtown, CT). Following sonication, sample was centrifuged for 60 sec. at 700 × g at

ambient temperature using an Eppendorf microfuge, to pellet debris in sample suspension.

Supernatant was enumerated to determine what fraction of bacteria remained in the

Page 150: New approaches and technologies for quantifying the impact

133

suspension (and was assumed non-particle attached). Sonicated samples were tested using the

previously explained IMS procedure to insure that the sonication did not reduce recovery.

PickPen Magnetic Separation

A PickPen® (Bio-Nobile, Parainen Finland) was tested to determine if this method of

magnetic bead extraction was superior to a Dynal MPC-S magnetic block used previously.

The device contained a magnetic pen- like tip, which was sheathed in a disposable rubber

cover, and inserted into the sample containing magnetic beads. Beads were captured by the

magnetic tip, and extracted from the sample. Once extraction of the beads occurred, the

magnetic tip recoiled back into the pen, allowing the beads to be resuspended in a desired

buffer (Nou et al., 2006). Immunomagnetic separation was conducted in the same manner as

previous treatments, with the exception that the magnetic PickPen® was used to remove

magnetic particles from the supernatant and resuspended in 1 mL of PBS. Samples were

mixed by inversion, and magnetic bead transfer was repeated.

Analysis of Bead Capture

Samples processed using the IMS were examined using epifluorescent SYBR Green I

DNA/RNA stain and microscopy described by Noble and Fuhrman (1998) and Patel et al.

(2007). Bound bacteria were stained, and examined using a Nikon Eclipse E800 (Melville,

NY) compound microscope with 100 W mercury-vapor lamp, a 100× immersion lens (plan

fluor infinity objective), and a 10× magnification eyepiece. Images were taken using an

Olympus DP71 camera (Center Valley, PA). Images were examined to determine particulate

Page 151: New approaches and technologies for quantifying the impact

134

content, characteristics of Ab labeled magnetic beads, and qualitative success of target

bacterial capture.

Enumeration and Statistics

Enumeration of bacteria was conducted using either membrane filtration (USEPA

method 1600 or 1603), or chromogenic substrate tests Colilert-18 or Enterolert (IDEXX,

Westbrook, ME) for quantification of E. coli and Enterococcus spp. All recoveries were

based on the formula:

Recovery % = (Concentration in final sample/Concentration in initial sample) x 100

Enumeration of bacteria and bead complexes using epifluorescence microscopy followed the

Sybr Green bacteria staining method of Noble and Fuhrman (1998). All statistics were

conducted either using SPSS (version 11.0) or Microsoft EXCEL (version 2007) software.

Data was tested using Shapiro-Wilk test for normality and Q-Q normality plot analysis.

Comparisons between means for data that were not normally distributed were conducted

using a Mann- Whitney U Test to determine significant differences. Normally distributed

data were examined for significant differences between means using a Student’s T-test

(unpaired). Significance was accepted at α = 0.05.

Results

Recovery

Streptavidin/Biotin Beads and Enterococcus spp.

Page 152: New approaches and technologies for quantifying the impact

135

Magnetic beads labeled with Virostat polyclonal Enterococcus Ab showed variable

recovery ranging from 14.3 to 54.2 % of Enterococcus spp. cells from sewage amended

seawater and tidal creek environmental samples, but almost no recovery of samples spiked

with cultured E. faecium and E. faecalis (Figure 5.2). Enterococcus cells in sewage

inoculated seawater had recoveries ranging from 14.3 to 32.2 %. Seawater spiked with

cultured E. faecium and E. faecalis resulted in recovery of <1 %. When beads with Ab

biotinylated in the lab were used, recovery of cultured cells amended into seawater increased

to 15.5 to 18.8 % for E. faecium and E. faecalis, respectively. Environmental tidal creek

samples from Gillets and Courthouse Bay showed recoveries of Enterococcus spp. (n=2) of

39.5 and 54.2 %, respectively.

Tosylactivated Beads for E. coli and Enterococcus spp.

DynaBead® M-280 tosylactivated beads labeled with Meridian Ab showed moderate

recovery of E. coli and Enterococcus spp. from sewage spiked seawater samples. Recovery

for E. coli and Enterococcus spp. ranged from 8.50 to 75.9 % and 7.28 to 58.1 %,

respectively, dependent on incubation treatment (time of incubation, shaker vs. centrifuge),

bead concentration, and volume of sample. Optimized recovery occurred at the highest bead

concentration (3.55 × 107) and at 30 minute incubation at a mixing speed of 25 rpm (Figure

3). Average recovery for E. coli and Enterococcus spp. using optimal conditions and

bacterial concentrations of 7 × 103 and 2 × 103 per mL were 38.5 % (SD ± 19.9) and 32.4 %

(SD ± 17.5), respectively. Centrifugation at 800 × g for 20 min. instead of shaker incubator

mixing showed no significant difference in recovery of E. coli (U=53.5, z= -0.429, p=0.674).

Centrifugation at 1100 × g for 5 min at 40°C also showed similar recovery results to shaker

Page 153: New approaches and technologies for quantifying the impact

136

incubation. Centrifugation vs. shaker incubation was not tested for Enterococcus spp. Three

washes with PBS or PBST (0.01% Tween-20), removed larger particles and most unbound

bacteria, but smaller particles remained in the sample.

Concentration Effects of Bead and Bacteria

Concentrations of beads and bacteria affected target cell recovery. For both

streptavidin M-280 and tosylactivated M-280 beads, regardless of Ab, recovery improved

with higher concentrations of beads and lower concentrations of bacteria (Figure 5.3). Bead

concentrations were tested from 2.5 × 106 to 3.55 × 107 per 1 mL sample, with greatest

recovery occurring with addition of 3.55 × 107 beads (Figure 5.4). Recovery was inversely

related to cell concentration, showing a decline in recovery when bacterial concentrations

increased. For instance, at E. coli concentrations of 2.29 × 104 E. coli per mL, average

recovery ranged from 16.9 to 24.6 %. However, when E. coli concentrations were decreased

to 3.40 ×103 per mL, the range of recoveries increased to 42.0 to 73.6 %. Regardless of the

type of bead or Ab, E. coli recovery rates decreased with each additional wash, and

consistently showed greater loss in the first wash as compared to later washes (Figure 5.5).

When a PickPen was used instead of a magnetic block, recoveries were 14.9 to 43.3 %

greater than magnetic block capture. In addition, samples seemed to be visually cleaner (less

particles), though no metric for remaining particulates was assessed.

Chemical Additions

Concentrations of non-particle attached E. coli in each sample increased after short

treatments (5-15 min.) at 37°C and rapid shaking at 130 rpm, of low concentrations (0.001

Page 154: New approaches and technologies for quantifying the impact

137

Molar) of certain chemicals. Gallic acid (0.001 M), catechin hydrate (0.5 mg/ mL), and PBS

(0.01 M) showed increased E. coli concentrations, when mixed with sample for 5, 15, and 30

min. and compared to a non-chemical control (Figure 5.6). Overall, catechin hydrate showed

the greatest increase in E. coli with increased concentrations ranging from 20.4 to 41.3%

more than the centrifuge control with15 min. of mixing, and was the only chemical addition

with a significant difference between the non-chemical control (t-test, p<0.05). E. coli

concentrations in the PBS control sample occasionally exceeded concentrations of the

chemical additions.

Sonication

Sonication increased bacteria in suspension after time and intensity of sonication

were optimized, as determined by comparison to a non sonicated centrifuge control and

enumeration. Samples were sonicated and then centrifuged to remove assumed particle

bound bacteria. At optimized conditions, unsonicated centrifuge controls a lways had lower

concentrations of E. coli after centrifugation than sonicated samples. Optimized sonication

results for 5 ml volumes occurred at 4 watts for 5 sec., and 10 mL volumes occurred at 4

watts for 10 sec. Increased recovery of E. coli from seawater after sonication ranged from

7.70 to 27.7%, but results were not statistically significant from the non-sonicated control (t-

test, p=0.434). When brackish water samples from Gillets Creek and the Neuse River were

tested, sonication recoveries ranged from 41.4 to 101% greater than the unsonicated control.

Discussion

Success of the IMS

Page 155: New approaches and technologies for quantifying the impact

138

Antibody labeled magnetic beads (both streptavidin and tosylactivated beads with

Meridian and Virostat Ab) successfully captured E. coli and Enterococcus spp. cells from

complex environmental samples amended with sewage or cultured cells with varying degrees

of success. These results are encouraging, since human/sewage effluent contaminated waters

are a main source of FIB and the main focal point for a range of EPA epidemiological

studies. Average recoveries for E. coli and Enterococcus spp. were slightly less than 40%.

These recoveries were lower than recovery ranges of 36.0 to 83.0% for Crytosporidium

parvum and 0.5 to 53.0% for Giardia from river water (Bukhari et al., 1998; DiGiorgio et al,

2002; Smith and Nichols, 2010) but similar to recoveries in highly turbid reservoir waters

(Krometis et al., 2009), and turbidity amended waters (DiGiorgio et al., 2002). The IMS for

capture of E. coli and Enterococcus spp. were conducted on concentrated and often highly

complex and concentrated water samples. Rigorous washing was required for IMS to remove

as much of the unwanted debris as possible. This rigorous washing has previously been cited

as one reason for lower recoveries of bacteria during the IMS (Porter and Pickup, 1999), and

coupled with highly complex and concentrated samples was likely the reason for overall

lower average recoveries.

Lower recoveries could also be the result of negatively biased enumeration methods.

Bacterial cells, when bound by Ab coated magnetic beads often form super-aggregates of

many cells, beads, and particulates joined together (Figure 5.7). When visualized under

epifluorescence microscopy and bacterial staining procedures, the presence of large

bead/particulate/E. coli super-aggregates was confirmed. In some cases, these complexes

consisted of over 100 cells. This defies the “one bead-one cell” example often shown in

diagrams of magnetic bead manufacturer’s products. Membrane filtration plating techniques

Page 156: New approaches and technologies for quantifying the impact

139

of post-IMS bead bound bacteria using MTEC and mEI agar plates supported these

conclusions, as irregular large colonies were apparent on the growth medium (Appendix C).

The occurrence of these larger colonies could result in a negative skewing of recovery data,

as more than one cell was likely attached per bead, but when cultured, aggregates were

counted as single colonies. Previous studies with IMS and E. coli and Campylobacter jejuni

had similar findings when using colony forming unit (CFU) counts (Perez et al., 1998; Yu et

al., 2001) and this problem is generally known to occur with IMS (Olsvik et al., 1994). Other

researchers have used a mass balance approach to determine recovery from the IMS (Lee et

al., 2009), instead of the direct plating (MF) or DST used for this study, and recovery

calculation described in the methods of this chapter. Using a mass balance approach would

have yielded slightly higher recoveries, but this approach is based on a range of assumptions

that were questioned during our study, and therefore, a mass balance approach was not used.

Bead/Cell Concentrations and Incubation Time

Recommended ratios for beads to cells are 4:1 to 10:1 (Neurauter et al., 2007). We

found that bead concentrations of 3.55 × 107 were optimal for recovery, which agreed with

results from other researchers (1.5 × 107 beads to 103 - 105 bacteria) (Porter and Pickup 1999,

Christine Lee pers. comm.). We did not see increased recovery if incubation time was greater

than 1 hr. and optimal recovery usually occurred within 30 minutes. Different IMS

incubation times with sample types such as food and clinical samples have shown increased

but finite capture efficiency up to 60 min. (Skjerve et al., l990; Gagne et al., 1998).

However, increased incubation times have also been suspected to increase cross reactivity

(Neurauter et al., 2007), and has been shown to occur using polyclonal antibodies (Lee and

Page 157: New approaches and technologies for quantifying the impact

140

Deininger, 2004). The use of centrifugation at 37° C and 800 -1100 ×g for 5- 20 min. as an

alternative IMS (to 30 min. incubation shaker approach) incubation technique resulted in

shorter time for processing, and equivalent recoveries. To our knowledge, this is the first

time that this has been documented in the literature. However, it is unclear whether this

change to processing increases non-specific binding, or changes bead/bacteria aggregate

characteristics (i.e. size, particle inclusion, etc.). However, if artifacts of the process are

minimal, this method alteration could reduce time for processing by at least 10-25 min.

Chemical and Physical Bacterial Dispersion

Bacteria loss during sample concentration and purification is attributed to bacteria

attachment on the surface of particulates which are removed from the sample during

preprocessing. Bacteria are naturally attracted to surfaces (Goulter et al., 2009) and a

moderate fraction in environmental waters is often attached to particles (Fries et al., 2006).

Both chemical additions and sonication showed potential for dispersing bacteria from

particles previous to the IMS, but with varying degrees of success. Previous studies have

utilized surfactants and chemicals such as Tween-80 (Hill et al., 2005), Triton X-100

(McDaniel and Capone, 1985), tetrasodium pyrophosphate (Lepesteur et al., 2003) and

flavonoids and weak acids to break bacterial bonds to surfaces (Duarte et al., 2006; Vikram

et al., 2010), and increase available bacteria for enumeration from complex sample matrices

(Garren and Azam, 2010). The use of catechin hydrate and weak acids including gallic acid

showed minor success at dispersing E. coli from particles after short time periods (5-15 min).

Specifically, catechin hydrate resulted in the highest concentrations of E. coli indicating

Page 158: New approaches and technologies for quantifying the impact

141

dispersal of cell/cell aggregates and/or cell/particle aggregates. This is one of few studies (if

any) that has been tested with flavonoids and FIB.

Use of sonication resulted in greater bacteria in supernatant after centrifugation at 700

× g for 1 min. Unsonicated samples also had greater bacteria loss from supernatant during

centrifugation, indicating that sonication increases free, non-particle attached bacteria in

sample suspension (see Appendix D for process diagram). Sonication has been shown

previously to disrupt cell binding to particulates, and increase freely available cells in sample

suspension (Leskinen and Lim 2008, Falcioni et al. 2006, Tso and Taghon 1997, McDaniel

and Capone 1985). Previous research has shown that sonication can result in an

improvement in suspended bacteria and separation of bacteria from sample particles (Stumpf,

unpublished data). However, this research was conducted using a non-variable Bransonic

sonication bath. A variable strength sonication probe was optimized (by duration and power

of sonication) to increase freely suspended bacteria, while avoiding cell lysis. Five and 10

mL samples showed increased bacteria in suspension when sonicated for 10 sec. at 4 watts.

Duration, intensity, and sample volume were determined to be highly important during

optimization to avoid cell lysis, while maximizing particle disassociation. For instance,

sonication on 1 mL volumes for 60 sec. and 8 watts resulted in almost complete lysis of all E.

coli present in the sample. Sonication could be added as a rapid technique in pre-processing,

and if followed by a brief centrifugation at 500 × g for 30 sec., could increase unattached

cells and decrease particles in suspension.

The use of sonication and chemical dispersants are highly relevant to the success of

sample purification for downstream FIB analysis. Since environmental samples are complex,

a purification of sample must be conducted to remove unwanted debris. However target

Page 159: New approaches and technologies for quantifying the impact

142

bacteria are often lost during purification. Results from this research indicate that both

sonication and chemical additions can assist in bacterial dispersion from particles. Though

bacterial concentrations were often not statistically different from the controls, it is important

to stress that treated samples were always cleaner. Additionally, these methods can be

combined with further processing steps such as low speed centrifugation and IMS to further

reduce particles, and allow higher rates of bacterial capture.

Conclusion

A dearth of research exists examining the efficacy of IMS in complex environmental

water samples and targeting E. coli and Enterococcus spp. Capture and recovery of E. coli

and Enterococcus spp. from marine water samples was moderate (mean < 40%) using

commercially available Ab, though further optimization and sample treatments could

improve these recoveries. The use of IMS was fairly rapid (total process less than 1 hour for

6 samples), reduced debris in sample, and specifically captured target bacteria. However,

because of the nature of the bead/cell aggregate formation of the IMS and utilized

enumeration methods (MF and DST), recoveries may be negatively skewed. Generally, more

research needs to be conducted to determine if improved recoveries of Enterococcus spp. and

E. coli can be obtained from environmental waters. IMS should be explored for applicat ion to

emerging FIB including alternative bacterial indicators that exist in higher concentrations in

sewage effluent such as Bacteroides spp. Overall, IMS is a promising method for selective

capture and recovery of intact, active cells and could be used in various sample types and a

range of applications from biosensor detection to cell purification.

A comparison of procedures for the se

Page 160: New approaches and technologies for quantifying the impact

143

Variable Goal Variation tested Conditions

Magnetic Bead Optimize Capture Magnetic Bead

Dynal m-280 Streptavidin vs. M-280

Tosylactivated beads

Sample Mixing Optimize Capture

Bead Concentration 7.1 × 106 to 3.55 × 10

7 per mL

Sample Incubation 10, 20, 30 and 60 min.

Centrifugation Incubation 10, 20 min. at 800 × g, 5 min. at 1100

× g

Washing Particle reduction Cell recovery and particulates 1-5 washes

Chemical Dispersant

Particle reduction

Chemical tested quercetin, catechin hydrate,

epigallocatechin gallate, gallic acid,

ellag ic acid, tannic acid, pbs

Concentration 0.1 M, 0.001 M

Time of addition 5, 10, 15, 20, 30, 60 min.

Sonication Particle reduction Intensity 4, 7, 10 watts

Time 5, 10, 30 sec.

PickPen Separat ion Optimize Capture Magnetic separation

PickPen vs. Dynal magnetic block

separation

Table 5.1. Variables tested for immunomagnetic separation to optimize recovery of target cells and reduce particles in sample

143

Page 161: New approaches and technologies for quantifying the impact

144

Figure 5.1. The Immunomagnetic Separation Method (IMS) showing A) mixing, B) capture,

C) washing, and D) resuspension of cleaned sample with target bacterial cells (modified from

Porter and Pickup, 1999)

Page 162: New approaches and technologies for quantifying the impact

145

Figure 5.2. Recovery of Enterococcus spp. using DynaBeads M-280 streptavidin labeled with

Virostat biotin conjugated polyclonal Enterococcus spp. antibody. Seawater samples

amended with sewage, cultured E. faecalis, and E. faecium, and two un-amended tidal creek

samples (Gillets Creek and Courthouse Bay) were tested with for recovery of Enterococcus

spp. Immunomagnetic separation indicates moderate capture of sewage amended and

environmental samples and poor capture of samples amended with cultured cells

Page 163: New approaches and technologies for quantifying the impact

146

Figure 5.3. Immunomagnetic Separation mean recovery using Dynal tosylactivated beads

(3.55 × 107 per sample) labeled with Meridian E. coli and Enterococcus spp. antibody.

Samples prepared from mixed particulates and amended with sewage (8 × 103 per mL E. coli,

3 × 103 per mL Enterococcus). All samples mixed with beads, incubated at 37o C at 25 rpm at

varied incubation times, and enumerated after two washes. Error bars represent range of

recovery

Page 164: New approaches and technologies for quantifying the impact

147

Figure 5.4. Bead recovery of E. coli from bead concentrations of 106 to 107 beads per sample.

E. coli concentrations were ~7 × 103 per sample, and samples were incubated at 37o C in a

shaker incubator for 30 minutes. Average taken from combined Colilert-18 and MTEC

results. Error bars represent standard deviation.

Page 165: New approaches and technologies for quantifying the impact

148

Figure 5.5. Mean recovery from immunomagnetic separation using Dynabead M-280

tosylactivated beads for E. coli capture from a 1 mL sewage sample. Mean E. coli recovery is

28.3%, and each subsequent wash using PBS resulted in loss of a fraction of bacteria. Error

bars represent standard error of the mean

Page 166: New approaches and technologies for quantifying the impact

149

Figure 5.6. Chemical additions to particle amended sewage spiked sample for 5, 15, and 30 minutes at 37° C and high

speed mixing. Raw is the initial sample, and centrifuge control had no chemical added. Control(CNTRL) was 1x PBS,

Quercetin(QC) 0.001M, Gallic Acid (GA) 0.001M, Ellagic Acid (EA) 0.0025M, Epigallocatechin Gallate (EA) 0.03

mg/ mL, and Catechin Hydrate (CH) 0.5 mg/ mL. All samples above reference line show greater bacterial availability

in sample suspension after treatment. Error bars represent standard error

149

Page 167: New approaches and technologies for quantifying the impact

150

Figure 5.7. Epifluorescence microscopy of E. coli bound to particulates and beads. Image

was taken after 30 min incubation with magnetic beads and 2 washes. Particulates are still

very abundant in sample, and E. coli are highly attached to particles

Page 168: New approaches and technologies for quantifying the impact

151

REFERENCES

Bhunia, A.K. (2008). Biosensors and bio-based methods for the separation and detection of

foodborne pathogens. Advances in Food and Nutrition Research, 54, 1-44.

Bukhari, Z., McCuin, R.M., Fricker, C.R., Clancy, J.L. (1998). Immunomagnetic separation of Cryptosporidium parvum from source water samples of various turbidities. Applied and

Environmental Microbiology, 64(11), 4495-4499.

Cudjoe, K.S., Krona, R. (1997). Detection of Salmonella from raw food samples using DynaBeads® anti Salmonella and a conventional reference method. International Journal of

Food Microbiology, 37(1), 55-62.

DiGiorgio, C.L. Gonzalez, D.A., Huitt, C.C. (2002). Cryptosporidium and Giardia recoveries in natural waters by using Environmental Protection Agency method 1623. Applied and

Environmental Microbiology, 68(12), 5952-5955.

Duarte, S., Gregoire, S., Singh, A.P., Vorsa, N., Schaich, K., Bowen, W.H., Koo, H. (2006). Inhibitory effects of cranberry polyphenols on formation and acidogenicity of Streptococcus

mutans biofilms. FEMS Microbiology Letters, 257(1), 50-56.

Falcioni, T., Manti, A., Boi, P., Canonico, B., Balsamo, M., Papa, S., (2006). Comparison of disruption procedures for enumeration of activated sludge floc bacteria by flow cytometry.

Clinical Cytometry, 70B, 149-153.

Feng, Y.Y., Ong, S.L., Hu, J.Y., Song, L.F., Tan, X.L., Ng, W.J. (2003). Effect of particles on the recovery of Cryptosporidium oocysts from source water samples of various turbidities.

Applied and Environmental Microbiology, 69(4), 1898-1903.

Fries, S.J., Characklis, G.W., Noble, R.T. (2006). Attachment of fecal indicator bacterial to particles in the Neuse River Estuary, N.C. Journal of Environmental Engineering, 132(10),

1338-1345.

Frischer, M.E., Danforth, J.M., Newton Healy. M.A., Saunders, F.M. (2000). Whole-cell versus total RNA extraction for analysis of microbial community structure with 16S rRNA

targeted oligonucleotide probes in salt marsh sediments. Applied and Environmental Microbiology, 66(7), 3037-3043.

Page 169: New approaches and technologies for quantifying the impact

152

Gagne, A., Lacouture, S., Broes, A., D’Allaire, S., Gottschalk, M. (1998). Development of an immunomagnetic method for selective Isolation of Actinobacillus pleuropneumoniae

serotype 1 from tonsils. Journal of Clinical Microbiology, 36(1), 251-254.

Garren, M., Azam, F. (2010). A new method for counting bacteria associated with coral mucus. Applied and Environmental Microbiology, 76(18), 6128-6133.

Goulter, R.M., Gentle. I.R., Dykes, G.A. (2009). Issues in determining factors influencing

bacterial attachment: a review using the attachment of Escherichia coli to abiotic surfaces as an example. Letters in Applied Microbiology, 49(1), 1-7.

Hill. V.R., Polaczyk, A.L., Hahn, D., Narayanan, J., Cromeans, T.L., Roberts, J.M.,

Amburgey, J.M. (2005). Development of a rapid method for simultaneous recovery of diverse microbes in drinking water by ultrafiltration with sodium polyphosphate and

surfactants. Applied and Environmental Microbiology, 71(11), 6878-6884.

Krometis, L.A.H., Characklis, G.W., Sobsey, M.D. (2009). Identification of particle size classes inhibiting protozoan recovery from surface water samples via USEPA Method 1623.

Applied and Environmental Microbiology, 75, 6619-6621.

Lanau, M., Lemke, A., Walther, K., Martens-Habbena, W., Simon, M. (2005). An improved method for counting bacteria from sediments and turbid environments by epifluorescence

microscopy. Environmental Microbiology, 7(7), 961-968.

Lee, J., Deininger, R.A. (2004). Detection of E. coli in beach water within 1 hour using immunomagnetic separation and ATP bioluminescence. Luminescence, 19, 31-36.

Lee, C.M., Griffith, J.F., Kaiser, W., Jay J. (2010). Covalently linked immunomagnetic

separation adenosinetriphosphate technique (Cov-IMS ATP) enables rapid in-field detection and quantification of Escherichia coli and Enterococcus spp. in freshwater and marine

environments. Journal of Applied Microbiology, 109(1), 324-333.

Lepesteur, M., Blasdall, S., Ashbolt, N.J. (2003). Particle dispersion for further Cryptosporidium and Giardia detection by flow cytometry. Letters in Applied Microbiology,

37(3), 218-229.

Page 170: New approaches and technologies for quantifying the impact

153

Leskinen, S.D., Lim, D.V. (2008). Rapid ultrafiltration concentration and biosensor detection

of Enterococci from large volumes of Florida Recreational water. Applied and Environmental Microbiology, 74(15), 4792-4798.

McDaniel, J.A., Capone, D.A. (1985). A comparison of procedures for the separation of

aquatic bacteria from sediments for subsequent direct enumeration. Journal of Microbiological Methods, 3, 291-302.

Neurauter, A.A., Bonyhadi, M., Lien, E., Nokleby, L., Ruud, E., Camacho, S., Aarvak, T.

(2007). Cell isolation and expansion using Dynabeads®. Advanced Biochemical Engineering and Biotechnology, 106, 41-73.

Noble, R.T., Fuhrman, J.D. (1998). Use of SYBR Green I for rapid epifluorescence counts of

marine viruses and bacteria. Aquatic Microbial Ecology, 14, 113-118.

Nou, X., Arthur, T.M., Bosilevac, J.M., Brichta-Harhay, D.M., Guerini, M.N., Kalchayanand, N., Koohmaraie, M. (2006). Improvement of immunomagnetic separation for

Escherichia coli O157:H7 detection by the Pickpen magnetic particle separation device. Journal of Food Protection, 69(12), 2870-2874.

Olsvik,O., Popovic T., Skjerve, E., Cudjoe, K.S., Hornes, E., Ugelstad, J., Uhlen M. (1994).

Magnetic separation techniques in diagnostic microbiology. Clinical Microbiology Reviews, 7(1), 43-54.

Patel, A.,Noble R.T., Steele, J.A., Schwalbach, M.S., Hewson I., Fuhrman J.A. (2007). Virus

and prokaryote enumeration from planktonic aquatic environments by epifluorescence microscopy with SYBR Green I. Nature Protocols, 2, 269-276.

Perez, F.G., Mascini, M., Tothill, I.E., Turner, A.P.F. (1998). Immunomagnetic separation

with mediated flow injection analysis amperometric detection of viable Escherichia coli O157. Analytical Chemisty, 70, 2380.

Porter, J., Pickup, R. (1999). Magnetic particle-based separation techniques for monitoring

bacteria from natural environments. Clive Edwards, Ed. Environmental Monitoring of Bacteria; Methods in Biotechnology, Chapter 6, Human Press, Inc. Totowa NJ.

Page 171: New approaches and technologies for quantifying the impact

154

Safarik, I., Safarikova, M., Forsythe, S.J. (1995). A Review: The application of magnetic separations in applied microbiology. Journal of Applied Microbiology, 78(6), 575-585.

Skjerve, E., Rorvik, L.M., Olsvik O. (1990). Detection of Listeria monocytogenese in foods by immunomagnetic separation. Applied and Environmental Microbiology, 56, 3478-3481.

Smith, H.V., Nichols R.A.B. (2010). Cryptosporidium : Detection in water and food.

Experimental Parisitology, 124(1), 61-79.

Stevens,K.A., Jaykus, L.A. (2004) Bacterial separation and concentration from complex sample matrices: a review. Critical Reviews in Microbiology, 30(1), 7-24.

Tu, S., Reed, S., Gehring, A., He, Y., Paoli, G. (2009). Capture of Escherichia coli 0157:H7

using immunomagnetic beads of different size and antibody conjugation chemistry. Sensors, 9, 717-730.

Vikram, A., Jayaprakasha, G.K., Jesudhasan, P.R., Pilla, S.D., Patil, B.S. (2006).

Suppression of bacterial cell–cell signaling, biofilm formation and type III secretion system by citrus flavonoids. Journal of Applied Microbiology, 109(2), 515-527.

Yu, L.S.L., Uknalis, J., Tu, S. (2001). Immunomagnetic separation for the isolation of

Campylobacter jejuni from ground poultry meats. Journal of Immunological Methods, 256, 11-18.

Tso, S.F., Taghon, G.L. (1997). Enumeration of protozoa and bacteria in muddy sediment.

Microbial Ecology, 33(2), 144-148.

Page 172: New approaches and technologies for quantifying the impact

CHAPTER 6

Conclusion

Following a decade of strong advances in water quality research to improve transport

characterization, source determination, and detection of fecal contamination, it is still

apparent that the complexity of environmental samples dictates the need for novel

approaches to sample processing for enumeration of fecal markers relevant to public health.

Sampling methods and processing approaches have improved, and alternative indicators have

been deployed in useful schematics such as multi-tiered and toolbox approaches. The work

comprised within this dissertation has advanced the field in the application of the above

techniques to vitally important tidal creek headwaters. Furthermore, the research has

advanced the field through the identification of sample processing approaches relevant to the

most complex of environmental waters, tidal creeks and estuaries.

This dissertation presents one of the first fully robust sets of tidal creek fecal indicator

loading data to date, employing an intensive sampling framework, and unique assessment

methodology for the characterization of loading within tidal creeks of the New River Estuary.

This improved sample scheme and loading characterization of stormwater is valuable for

understanding inter-storm variability, magnitude, and duration of contamination after large

scale storm events (which are common along the coast of North Carolina), and indicates the

need for improved monitoring efforts. As a result, a more accurate understanding of fecal

contaminant loading was developed, including evidence that no first flush occurred for ENT

or EC and that the majority of FIB loading was occurring during storm events.

Page 173: New approaches and technologies for quantifying the impact

156

Understanding these loading dynamics indicates that mitigation strategies for FIB best

management practices will need to consider full storm duration to have any beneficial impact

on fecal contaminant reductions. In addition, this research emphasizes the possibility that

watershed characteristics (land cover and soil types) could be responsible for modification of

loading patterns of FIB and TSS. Finally, elevated FIB concentrations were found even in

very low development watersheds, which is interesting since previous studies have found

strong a correlation between FIB and percent land development. This indicates the need for

remediation of upstream inputs as these watersheds already have very low impervious land

surface.

A toolbox approach using a combination of conventional fecal indicators, alternative

indicator QPCR-based quantification of three different subsets of the Bacteroides spp., and

optical brightener chemical determinants were used to generate a better estimation of the

fecal contamination signature of tidal creeks in economically important shellfish harvesting

areas. The New River Estuary is also the largest amphibious training area for the US military,

making this work directly relevant to public health. Application of this toolbox approach

determined suggestive contamination of one tidal creek with human source fecal

contamination, and the presence of intermittent and diffuse human fecal contamination in

three other headwater tidal creeks. Tidal creeks are notoriously difficult environments to

conduct water quality based research, given their constantly modified environments (based

upon freshwater flow from upstream, tidal impact, wind driven advective mixing,

stratification, etc.). Despite complex water matrix challenges and the limitations of the

individual fecal indicator detection methods, a framework for robust positive identification of

Southwest Creek as contaminated with human fecal contamination and major concern for

Page 174: New approaches and technologies for quantifying the impact

157

water quality in the New River Estuary was determined. This agreement of several different

methods is rare when using multiple indicators, but in this case extremely encouraging. This

approach could be applied to other such water matrices and geographically diverse areas.

This research also highlights the limitations with the use of QPCR and optical

brightener methods in tidal creeks. Given these limitations, work was conducted to improve

the ability to detect target cell/ pathogens in such complex environmental samples using

alternative technologies. Biosensors hold promise for improving detection and testing of

environmental samples, but complex pre-sample processing is often required to overcome

sensitivity issues and successfully detect microbial targets. This research presents several

techniques for sample concentration, purification, and capture of E. coli, Enterococcus spp.,

and Bacteroides spp. A specific IMS based capture technique was developed to rapidly

capture E. coli and Enterococcus spp. from environmental samples. This research is highly

valuable, since there is a dearth of information on the use of IMS for FIB in environmental

waters. These optimized sample processing methods are valuable for successful detection of

fecal contamination in marine and freshwater using biosensors.

Biosensor detection has the potential to revolutionize microbial water testing.

Biosensors can be rapid, accurate, portable, and easy to use. Future water quality testing

might include in situ fecal detection biosensors (similar to handheld sensors currently

available for reporting water quality parameters such as temperature, dissolved oxygen,

conductivity, etc.) that would alert managers once FIB regulatory thresholds are exceeded.

This potential for biosensor produced real-time data would make great strides in improving

protection of public health for swimmers, shellfish harvesting, and military personnel in

maneuvers involving water contact. Though this potential for microbial biosensors is pushing

Page 175: New approaches and technologies for quantifying the impact

158

the limit of our current water monitoring ability, it will likely occur given technological

advancement and the use of alternative targets including whole cells, DNA, and/or

representative chemical fecal indicators such as caffeine or optical brighteners.

Several other possible techniques and determinations of fecal contamination were

also examined or considered during this research. Specifically, the importance and close

association of turbidity with fecal indicators was consistently found. This may indicate that

presumptive closures are appropriate when turbidity levels reach threshold values, and allow

for more rapid estimates of fecal contamination, even when samples become difficult to

analyze due to high inhibition from sediments. Additionally, reservoirs of fecal

contamination in many water bodies are poorly understood. Though reservoir populations

were not addressed in this thesis, the resuspension of FIB is an important consideration of

any study, and needs greater examination. Finally, fate and transport of indicators is also of

utmost importance. Future studies may be able to better examine transport and survival of

fecal indicator bacteria using radioactive isotopes (such as Thorium) tracers, dye studies, and

other seldom utilized techniques. These out of the box approaches will likely assist in greater

understanding of fecal contamination, and need to be considered.

The overall goal of this research was to improve approaches to "real world"

assessment of fecal contamination of water. Water quality is of utmost importance for North

Carolina’s economy, safety of its military personnel, and promotion of water recreation.

Significant advancements in the approaches and technologies for testing indicators of

pathogenic contamination were developed and optimized. This dissertation research project

advances the field of water quality testing, and the methods and approaches developed could

be applied to better monitor and protect waters from fecal contamination.

Page 176: New approaches and technologies for quantifying the impact

159

APPENDIX

A.

In chapter 2 it is stated that sediment resuspension is responsible for only a small

contribution of the total FIB load during wet weather. This was calculated using the

following formula to determine contribution of FIB to water column from sediments:

Average Sediment FIB Concentration (MPN/g) × (TSS concentration (g/L) ×

(0.1L/100 mL) = FIB (MPN/100 mL) in water column concentration from sediment

This was then compared to the storm concentration of FIB. For instance, in Gillets Creek:

Average sediment EC concentration = 19.17 (MPN/g). Average TSS concentration = 13.5

(g/L)

Calculation: 19.17 (MPN/g) × (13.5 (g/L) × (0.1L/100 mL) = 25.88 (MPN/100 mL)

ENT contributed to water column from sediments = 25.9 (MPN/100 mL) Average ENT in water column during storm = 288.4 (MPN/100 mL)

Therefore, contribution of ENT to water column from sediments was only 9% of total ENT

in the storm flow water column.

Page 177: New approaches and technologies for quantifying the impact

160

B.

Calculations:

Percent recovery of bacteria for each treatment was determined by comparing the final

concentration (concentration × volume) to the initial raw sample concentration (concentration

× volume):

Percent Recovery = (Total Bacterial Concentration of Final Sample / Total Bacterial Concentration of Raw Seawater) × 100

Total recovery for the optimized MFM was determined for E. coli, Enterococcus spp., and B.

thetaiotaomicron, while the optimized ILM was determined using E. coli. Total percent

recovery was calculated including loss from all processing steps for each optimized method:

Total Percent Recovery = ((Bacterial Concentration of Final Sample (Including Loss

Due to Particulate Removal, Concentration, and Elution from Optimized Method))/ Total Bacterial Concentration of Raw Seawater) × 100

Concentration factor was determined by comparing total bacteria concentration per equal

volumes of concentrate to initial raw seawater:

Concentration Factor= (Total Bacterial Concentration Final Sample (1 mL volume) / (Bacterial Concentration of Raw Seawater (1 mL volume))

Page 178: New approaches and technologies for quantifying the impact

161

C.

Seawater particulates concentrated from 5L filtration. MTEC and MEI plates from a time series incubation test run at 10, 20, and 30 minutes in the shaker incubator and DynaBead M-

280 tosylactivated with Meridian E. coli and Enterococcus antibody. Colonies are large and irregular shaped indicating possible particulate attachment and super-aggregate colony formation

Page 179: New approaches and technologies for quantifying the impact

162

D.

Diagram of sonication testing

Concentrated sample

Centrifuge @ 700xg for 60 sec

All Samples Vortexed for 30 Sec.

Control Centrif Only Sonication

Treatment 1

Sonication

Treatment 2

Sonication

Treatment 3

Enumerate

Enumerate

Bacterial Loss

Sonicate for 10, 20, 30 Sec @ 4, 7, 10 Watts

Centrifuge @ 700xg for 60 sec

Particles in pellet

Bacteria Remain

in suspension

Enumerate

Bacterial Loss