82
Diversity and Dynamics of Algal Viruses in the Bay of Quinte By Robin Marie Rozon A thesis submitted in conformity with the requirements for the degree of Master of Science Ecology and Evolutionary Biology University of Toronto © Copyright by Robin Marie Rozon 2013

Diversity and Dynamics of Algal Viruses in the Bay of Quinte · Diversity and Dynamics of Algal Viruses in the Bay of Quinte Robin Marie Rozon Master of Science ... and Heather Niblock

  • Upload
    others

  • View
    3

  • Download
    0

Embed Size (px)

Citation preview

Page 1: Diversity and Dynamics of Algal Viruses in the Bay of Quinte · Diversity and Dynamics of Algal Viruses in the Bay of Quinte Robin Marie Rozon Master of Science ... and Heather Niblock

Diversity and Dynamics of Algal Viruses

in the Bay of Quinte

By

Robin Marie Rozon

A thesis submitted in conformity with the requirements

for the degree of Master of Science

Ecology and Evolutionary Biology

University of Toronto

© Copyright by Robin Marie Rozon 2013

Page 2: Diversity and Dynamics of Algal Viruses in the Bay of Quinte · Diversity and Dynamics of Algal Viruses in the Bay of Quinte Robin Marie Rozon Master of Science ... and Heather Niblock

ii

Diversity and Dynamics of Algal Viruses in the Bay of Quinte

Robin Marie Rozon

Master of Science

Ecology and Evolutionary Biology

University of Toronto

2013

Abstract

To initiate algal virus research in the Bay of Quinte, three stations were sampled biweekly

throughout 2011. By targeting algal virus DNA polymerase, major capsid protein genes (MCP),

and a Microcystis aeruginosa cyanophage (Ma-LMM01) tail sheath protein gene, PCR

amplification revealed diverse and unique Phycodnaviruses (viruses of eukaryotic algae) and

cyanophage. When analysed statistically, patterns of virus abundance suggested that the

seasonality of any one virus cannot be generalised to predict that of other viruses, even among

closely related viruses. This study also demonstrated a strong relationship between algal virus

abundance and host biomass. It was found that despite the apparent heterogeneity of virus

abundance across the Bay, virus abundance patterns clustered by sampling date and geographic

location. By providing evidence for diverse algal viruses with complex seasonality, this work

highlights significant gaps in the current understanding of Bay of Quinte phytoplankton

ecology.

Page 3: Diversity and Dynamics of Algal Viruses in the Bay of Quinte · Diversity and Dynamics of Algal Viruses in the Bay of Quinte Robin Marie Rozon Master of Science ... and Heather Niblock

iii

Acknowledgements

First and foremost I would like to thank my supervisor Dr. Steven Short for making this

‘drive-by Masters’ the most rewarding and intellectually challenging experience of my life. It is

because of your enthusiasm that I came to UTM, it was with your encouragement that I learned

to learn from my mistakes, and without your guidance I would never had been able to finish this

thesis. Because of your passion for science I understand what it means to work hard and to truly

love what you do; it’s viral! I am forever grateful for your patience and commitment to my

education. Thank you for welcoming me into your lab family and will never forget what we’ve

accomplished here.

Thank you to the members of my supervisory committee, Dr. Peter Kotanen and Dr.

Nick Collins. Your insights, recommendations and comments throughout my graduate degree

have improved my thesis and my skills as a scientist. I hope I can be as beneficial to others in

the future as you have both been to me.

To Cindy Short, Michael Staniewski, and all the past and present members of the Short

lab. While my stay with all of you was ‘short’-er than other programs, you deserve more than a

‘short’ expression of gratitude. In all seriousness, I thank you from the bottom of my heart for

putting up with my endless questions and helping me at every step of my journey. You are the

heart and soul of our work; the lab would have been lonely without you. I will miss our lab

lunches and geeking out with other like-minded souls. I wish you all the best for your future

endeavours and expect many great things from all of you. Finally, thank you for teaching me

that “adversity builds character; if everything went right all the time you’d be as interesting as a

carrot”. People like you are rare in this world and I look forward to many lasting friendships.

A very big thank you goes to Dr. Mohi Munawar and the people at DFO-GLLFAS who

made this collaboration possible. Without the encouragement received from Mark Fitzpatrick

and Heather Niblock I would never have come this far. To the GLLFAS field and lab crew,

Robert Bonnell, Ashley Bedford and Michele Burley, the DFO samples were collected,

processed and analysed on your backs and I am forever grateful. To the smiling faces of Lisa

Elder, Jennifer Lorimer, Ron Dermott, Kelly Bowen, Marten Koops, and the staff at CCIW

Burlington, thank you for your encouragement throughout my undergraduate and graduate

degrees, and for making my time with you so unforgettable.

Page 4: Diversity and Dynamics of Algal Viruses in the Bay of Quinte · Diversity and Dynamics of Algal Viruses in the Bay of Quinte Robin Marie Rozon Master of Science ... and Heather Niblock

iv

Last but not least thank you to my family and friends who supported me through the

good times and the bad. Thank you for being a sympathetic ear when I needed it and a kick in

the pants when I needed that too. Special thanks to Nathaniel Costa for your understanding and

patience; I owe my current sanity to you. Also, thank you for use of your ‘workspace’ and

assorted computer programs. To my parents, thank you for tolerating my rants against physics,

statistics, traffic jams and incidences of personal electrical interference. Most of all, I truly

appreciate your support over the last six years; it’s the best feeling in the world when your

parents can tell other people, “my daughter is working on her Masters degree. Something about

the little dudes in the water… I think.”

Robin

Page 5: Diversity and Dynamics of Algal Viruses in the Bay of Quinte · Diversity and Dynamics of Algal Viruses in the Bay of Quinte Robin Marie Rozon Master of Science ... and Heather Niblock

v

Dedication

To my parents

Je t’aime toujours

Page 6: Diversity and Dynamics of Algal Viruses in the Bay of Quinte · Diversity and Dynamics of Algal Viruses in the Bay of Quinte Robin Marie Rozon Master of Science ... and Heather Niblock

vi

Table of Contents

Abstract ii

Acknowledgements iii

Table of Contents vi

List of Tables viii

List of Figures ix

List of Abbreviations x

Chapter 1: Introduction

The Bay of Quinte

A Historical Perspective 1

Environmental Degradation 1

Project Quinte 3

Phytoplankton in the Bay of Quinte

Research History 3

Historical Trends 5

Aquatic Viruses

Historical Perspective on Aquatic Virus Research 6

Virus Diversity and Dynamics in Aquatic Systems 9

Objectives and Research Questions 11

Chapter 2: Materials and Methods

Study Sites and Sample Collection 12

PCR, Cloning and Sequencing

AVS 13

MCP 14

Sheath 14

Sequence Analysis and Primer Design 16

Quantitative PCR 17

Analysis 18

Page 7: Diversity and Dynamics of Algal Viruses in the Bay of Quinte · Diversity and Dynamics of Algal Viruses in the Bay of Quinte Robin Marie Rozon Master of Science ... and Heather Niblock

vii

Chapter 3: Results

Algal Virus Diversity Study

Algal virus DNA polymerase genes 20

Algal virus major capsid protein genes 20

Microcystis phage sheath protein genes 21

Quantitative Analysis of Viral Dynamics

Virus Dynamics across Time and Space 21

Statistical Analysis of Viral Dynamics 23

Virus Abundance and Host Biomass 24

Clustering by Taxonomic Classification 25

Clustering by Spatial Distribution 26

Chapter 4: Discussion

Algal Virus Diversity Study 27

Quantitative Analysis of Viral Dynamics

Virus Dynamics across Time and Space 31

Statistical Analysis of Viral Dynamics 35

Virus Abundance and Host Biomass 37

Clustering by Taxonomic Classification 39

Clustering by Spatial Distribution 41

Summary 42

Future Directions 44

Literature Cited 45

Page 8: Diversity and Dynamics of Algal Viruses in the Bay of Quinte · Diversity and Dynamics of Algal Viruses in the Bay of Quinte Robin Marie Rozon Master of Science ... and Heather Niblock

viii

List of Tables

Table 1. Limnological Divisions of the Bay of Quinte 52

Table 2. Sample Locations 53

Table 3. Reagents used in PCR reactions 54

Table 4. Thermocycling parameters for PCR reactions 55

Table 5. Quantitative Primers and Probes 56

Table 6. Reagents used in quantitative PCR 58

Table 7. Test of Primer and Probe Specificity 59

Table 8. Friedman analysis of virus abundances between stations 60

Page 9: Diversity and Dynamics of Algal Viruses in the Bay of Quinte · Diversity and Dynamics of Algal Viruses in the Bay of Quinte Robin Marie Rozon Master of Science ... and Heather Niblock

ix

List of Figures

Figure 1. Map of the Bay of Quinte 61

Figure 2. Neighbor joining phylogeny polB 62

Figure 3. Neighbor joining phylogeny MCP 63

Figure 4. Abundances of individual virus genes plotted against time 64

Figure 5. Virus gene abundances at each station plotted against time 65

Figure 6. Virus gene abundance 66

Figure 7. Regression of the sum of virus gene abundances on chlorophyll a. 67

Figure 8. Cluster analysis of virus abundance 68

Figure 9. A comparison of different proximity measures 69

Figure 10. Cluster analysis of biweekly stations 70

Figure 11. Cluster analysis of spatial stations 71

Page 10: Diversity and Dynamics of Algal Viruses in the Bay of Quinte · Diversity and Dynamics of Algal Viruses in the Bay of Quinte Robin Marie Rozon Master of Science ... and Heather Niblock

x

List of Abbreviations

ºC – Degree(s) Celsius

µ - micro (10-6

)

µL – microliter(s)

µm – micrometers

µM – micromolar

AOC – Area of Concern

B – Belleville sampling station

BLAST – Basic Local Alignment Search Tool

bp – Base pairs

BQ – Bay of Quinte

cm – centimeters

Ct – Cycle threshold

DFO – Fisheries and Oceans Canada

DH5α – E. coli strain of competent cells

DNA – deoxyribonucleic acid

dNTP – deoxyribonucleotide triphosphate

dsDNA – Double stranded DNA

EDTA – ethylenediaminetetraacetic acid

FAM – 6-carboxyfluorescein

FQ – Fluorescein quencher

g – gram(s)

GLLFAS – Great Lakes Laboratory for

Fisheries and Aquatic Sciences

GLWQA – Great Lakes Water Quality

Agreement

GPS – Global Positioning System

HAB(s) – Harmful Algal Blooms

HB – Hay Bay sampling station

HCl – Hydrogen chloride

IBM SPSS – International Business

Machines’s Statistical Package

for the Social Sciences

IJC – International Joint Commission

IPTG - Isopropyl β-D-1-

thiogalactopyranoside

JTT – Jones-Taylor-Thornton

L – liter(s)

LB agar - Lysogeny broth agar

LE agarose – Low electroendosmosis agarose

LO – Lake Ontario

M – Molarity

m – meter(s)

Ma-LMM01 – Strain of Microcystis

aeruginosa cyanophage

MCP – Major capsid protein

Page 11: Diversity and Dynamics of Algal Viruses in the Bay of Quinte · Diversity and Dynamics of Algal Viruses in the Bay of Quinte Robin Marie Rozon Master of Science ... and Heather Niblock

xi

MEGA – Molecular Evolutionary Genetics

Analysis

mg – milligram

MgCl2 – Magnesium chloride

mM – milimolar

mm – millimeter(s)

mol – mole(s)

MUSCLE – Multiple Sequence Comparison

by Log-Expectation

n – nano (10-9

)

N – Napanee sampling station

NCLDV – Nucleocytoplasmic large DNA

virus

ng – nanogram(s)

NR – Napanee River

OTU(s) – Operational taxonomic units

PBCV-1 – Paramecium bursaria Chlorella

virus 1

PCR – Polymerase chain reaction

polB – DNA polymerase

PQ – Project Quinte

PVDF – Polyvinylidene fluoride

qPCR – Quantitative polymerase chain

reaction

RAP – Remedial Action Plan

RNA – Ribonucleic acid

ROX – 6-Carboxyl-X-Rhodamine

rpm – revolutions per minute

s – second(s)

SNPs – Single nucleotide ploymorphisms

SS – Spatial survey

ssDNA – Single stranded DNA

TAE buffer – TrisCl base, acetic acid and

EDTA

TEM – Transmission electron microscope

TP – Total phosphorus

TrisCl – Tris(hydroxymethyl)aminomethane

UPGMA – Unweighted pair group method

with arithmetic mean

UV – ultraviolet light

USA – United States of America

V – volt(s)

VC – viral concentrate

X g – acceleration due to gravity

X-Gal – Bromo-chloro-indolyl-

galactopyranoside

YGDTDS – amino acids encoding for

catalytical domain of a viral

DNA polymerase

Page 12: Diversity and Dynamics of Algal Viruses in the Bay of Quinte · Diversity and Dynamics of Algal Viruses in the Bay of Quinte Robin Marie Rozon Master of Science ... and Heather Niblock

1

Chapter 1: Introduction

The Bay of Quinte

A Historical Perspective

In the Late 1700s, empire loyalists fled north and settled along the northern shores of Lake

Ontario. They preferentially settled in a small, ecologically diverse area which could support

their logging and farming needs, now known as the Bay of Quinte. The Bay of Quinte is a 100

km long Z-shaped embayment on the northern shores of Lake Ontario, located 135 km east of

Toronto and 40 km west of Kingston. It has an area of 254 km² and acts as a watershed for

18,182 km², including four major rivers (Trent, Moira, Salmon, and Napanee) that empty into

the Bay. The Bay of Quinte is narrow and relatively shallow, deepening at the interface to Lake

Ontario (Figure 1). The Bay is subdivided into 3 bays; Upper, Middle and Lower Bay, each with

distinct widths and depths (Table 1). The Lower Bay has the only natural passages to Lake

Ontario (Upper and Lower Gap), however the Upper Bay has a navigational channel which

artificially joins the Bay of Quinte to Lake Ontario to allow for ease of access to the local towns

and cities (Johnson and Hurley, 1986).

There are 12 townships, 3 towns (Picton, Napanee, and Deseronto), 2 cities (Trenton and

Belleville) (Committee, 1990), and two historic Department of National Defence bases within

the Bay of Quinte area (Johnson and Hurley, 1986). Over the years, the uses for the Bay of

Quinte have ranged from agricultural to industrial, to commercial and sport fishery, and other

recreational pursuits (Committee, 1993). Belleville and Deseronto take water from the Bay for

domestic and industrial use, but all the aforementioned municipalities have discharged

wastewater into the Bay. Some of these activities were the cause of the excessive nutrient

enrichment and eutrophication of the Bay of Quinte (Johnson and Hurley, 1986).

Environmental Degradation

Since the arrival of settlers in the 1780s, land use in the area has had a great impact on

water quality in the Bay of Quinte. Increased logging and agriculture gave way to full-scale

urbanization in the late 1800s which included mining and deforestation (Committee, 1990).

Nutrient enrichment in the Bay of Quinte became apparent in 1904 when algal slime began

fouling fishing nets and again in 1938 causing taste and odour issues for the municipal water

Page 13: Diversity and Dynamics of Algal Viruses in the Bay of Quinte · Diversity and Dynamics of Algal Viruses in the Bay of Quinte Robin Marie Rozon Master of Science ... and Heather Niblock

2

treatment plants, all likely due to excess phosphorus from the introduction of untreated sewage

and detergents into the Bay (Minns et al., 2011).

The 1940s saw the collapse of the commercial herring fishery and a shift from pike and

bass (visual predators) to walleye (turbid water predators); all indications of more turbid, algal

prone waters. Eurasian milfoil (Myriophyllum spicatum) invaded the Bay in the 1950s,

outcompeting the native macrophytes. By the 1960s, algal blooms were so severe that other

macrophyte species were shaded and disappeared. This massive die-off caused low dissolved

oxygen at depths, loss of species diversity, repugnant odors and diminished water clarity. To

make matters worse, white perch invaded the Bay around this time, decimating the native

populations of whitefish, lake herring, walleye and northern pike (Committee, 1990).

Finally, in 1970, the low dissolved oxygen levels, loss of species diversity and diminished

water quality drove the International Joint Commission (IJC) to take action against the levels of

phosphorus being discharged into Lake Ontario and the Bay of Quinte by striving to achieve “at

least an 80 percent reduction by 1975” (Minns et al., 1986). Two years later, Project Quinte, a

long-term multi-agency ecosystem project lead by Fisheries and Oceans Canada was developed

to control eutrophication under the newly formed Great Lakes Water Quality Agreement

(GLWQA). In 1975, the Bay of Quinte was listed as a “problem area” mainly due to its

excessive nutrient enrichment, nuisance algal blooms, low dissolved oxygen and general

contamination. By the time the GLWQA was renewed in 1978, municipal phosphorus loadings

into the Bay of Quinte had decreased by 50 % and there were improvements in water quality.

However phosphorus levels in the Bay had only decreased by 35 %, trophic levels remained

unbalanced, ecosystems remained dominated by single algal species, and the Bay remained

unsafe for recreational use (Committee, 1990).

In 1985, the Bay of Quinte was listed as an Area of Concern (AOC) according to the

GLWQA. AOCs are locations where environmental quality and beneficial uses of the aquatic

environment are lessened. The Bay of Quinte had 10 of the 14 potential use impairments,

including wildlife consumption restrictions, degradation of benthos, drinking water restrictions,

and particularly, undesirable algae. Part of the process of being listed as an AOC is the

development of a Remedial Action Plan (RAP) which defines the actions and timetables to

restore the beneficial uses of the AOC through surveillance, monitoring and the control of

pollution sources (Minns et al., 1986).

Page 14: Diversity and Dynamics of Algal Viruses in the Bay of Quinte · Diversity and Dynamics of Algal Viruses in the Bay of Quinte Robin Marie Rozon Master of Science ... and Heather Niblock

3

Project Quinte

What began as a straightforward research objective from 1972, “does phosphorus

reduction to specific levels at point sources increase production of economically important fish

stocks and improve water quality in a reasonable amount of time?” became a multi-agency

project that brought together committed scientists and managers with expertise, interest and

concern (Minns et al., 1986). Project Quinte has become a pilot project for the experimental

management approach; with base line data from before phosphorus-loading restrictions came

into effect, to 40 seasons of annual research and monitoring activities. Such multi-level

ecosystem studies are rare and unique in the world (Minns et al., 2011).

With the successful completion of RAP stage one (Committee, 1990), and stage two

(Committee, 1993), local organisations, alongside federal and provincial governments, are

focusing on the implementation of the 80 recommendations set out in the stage two report and

monitoring of the improvements from these remediation efforts to ensure that the goals are

accomplished. Goals presented in 2000 include: fewer beach closures, healthy primary

producers and protection for habitats restored to acceptable levels. Successful completion of

stage three is followed by a delisting process where the RAP participants decide if all the

remediation goals have been met. At this time, the Area of Concern becomes an Area in

Recovery, where continued diligence is necessary to maintain the health of the ecosystem

(Committee, 1990).

The Bay of Quinte is targeted for delisting in 2013; therefore the current focus is on

meeting those targets, reporting the science and preserving the knowledge as successors carry

on. Project Quinte has become a leader in eutrophication research and ecosystem-wide

modeling, with the goal of promoting whole ecosystem wellbeing (Minns et al., 2011). Because

of Project Quinte, localized phytoplankton research is ongoing and far-reaching.

Phytoplankton in the Bay of Quinte

Research History

Prior to Project Quinte, phytoplankton research in the Bay of Quinte was intermittent. The

first investigation of phytoplankton in the Bay of Quinte was accomplished by Tucker (1948).

Page 15: Diversity and Dynamics of Algal Viruses in the Bay of Quinte · Diversity and Dynamics of Algal Viruses in the Bay of Quinte Robin Marie Rozon Master of Science ... and Heather Niblock

4

By sampling in the Lower Bay during the summer of 1945, the research examined the

relationship between wind and mixing depth, and determined the viability of phytoplankton

enumeration as a measure of productivity (Tucker, 1948). This marked the first examination of

phytoplankton trends and abundances in the Bay. Since then, phytoplankton in the Bay of

Quinte have been the focus of numerous studies (McCombie, 1967; Nicholls and Carney, 1979;

1986; Nicholls and Heintsch, 1986; Nicholls et al., 1986; Nicholls et al., 2002; 2004; Nicholls

and Carney, 2011; Munawar et al., 2011).

Project Quinte, as mentioned, was developed as a case study to assess the response of the

Bay to a reduction in phosphorus loadings. Fourteen years after its inception, a special

publication detailing ten years’ worth of research in the Bay of Quinte was released to detail

how the experimental management approach was progressing. Included in this publication was a

paper by Nicholls and Carney (1986) which demonstrated that nitrogen was limiting to

phytoplankton growth in the highly eutrophic Upper Bay while phosphorus was limiting in the

mesotrophic Lower Bay. By studying the impact of reduced phosphorus loading in the Bay of

Quinte, it was found that periods of nitrogen limitation were much less frequent after the

implementation of the phosphorus abatement program, which coincided with a reduction in

biomass of N-fixing phytoplankton, such as blue-green algae like Anabaena and

Aphanizomenon. In the same special publication, Nicholls et al., (1986) described specific

phytoplankton biomass from 1972 to 1981, precisely during the period of reduced phosphorus

loading. It was found that total phytoplankton biomass and bloom durations decreased

significantly, so much so that odor problems and municipal filter clogging issues had lessened.

It was also found that the phytoplankton communities within the natural divisions of the Bay

were no longer as distinct after 1977 (Nicholls et al., 1986). In addition, Nicholls and Heintsch

compared total phytoplankton levels in the Lower Bay from 1945 to 1981 by duplicating the

methods and analysis of Tucker (1948). The results showed that by 1945, the Bay of Quinte was

already becoming eutrophic (although not as eutrophic as the 1950s and 60s) and that there was

no significant difference between the total biomass in the Lower Bay between these studies

(Nicholls and Heintsch, 1986). Nicholls continues to research phytoplankton in the Bay of

Quinte, focusing on statistical assessments of the impact of reduced phosphorus loading and the

Dreissena spp. invasion on phytoplankton and how it might be possible to guide future

management objectives (Nicholls and Carney, 2011).

Page 16: Diversity and Dynamics of Algal Viruses in the Bay of Quinte · Diversity and Dynamics of Algal Viruses in the Bay of Quinte Robin Marie Rozon Master of Science ... and Heather Niblock

5

Historical Trends

Since Tucker (1948) indexed biological productivity in the Bay of Quinte, Nicholls

repeatedly analyzed biomass and community composition trends of phytoplankton species over

the past three decades (Nicholls et al., 2002; Nicholls, 2010; Nicholls and Carney, 2011). Since

1970, the Bay of Quinte has been subjected to two large environmental alterations; point-source

phosphorus loading reduction in the late 1970s and the Dreissena invasion around 1995.

Following reduced phosphorus loading, there was a 51 % decline in total phytoplankton

biomass. There were significant declines in total biomass of Chlorophyceae (–66 %),

Dinophyceae (–58 %), Bacillariophyceae (–56 %), Cryptophyceae (–52 %), and non-nitrogen

fixing species of Cyanophyceae (–26 %). The dominant diatom species Aulacoseira and

Stephanodiscus also showed biomass declines by 58 and 78 % respectively.

The arrival of zebra mussels 15 years later (~1995), influenced some phytoplankton

populations differently than others; total Bacillariophyceae, total Chrysophyceae, and total

Cryptophyceae biomass did not change significantly, but there was a 42, 51 and 55 % decline in

biomass in Chlorophyceae, Cyanophyceae, and Dinophyceae, respectively. Although total

diatom biomass did not change, there was a shift in diatom communities with certain taxa such

as Stephanodiscus, Synedra, and Tabellaria declining by 82, 84, and 98 %, respectively. Of

particular relevance, there was a 13-fold increase in the biomass of bloom forming, toxin-

producing Microcyctis (Nicholls et al., 2002). Overall, the most important primary producers in

the Bay of Quinte have consistently been Bacillariophyceae and Cyanophyceae, which make up

between 75 and 95 % of the total phytoplankton biomass in the Upper Bay, less so in the Middle

and Lower Bay. Within cyanophyta, Anabeana, Aphanizomenon, and Gloeotrichia (all nitrogen-

fixing species) were abundant prior to the establishment of dreissenids. After the arrival of zebra

mussels, non-nitrogen fixing species, such as Microcystis became more abundant, particularly in

the Middle and Lower Bays (Nicholls and Carney, 2011).

Multi-year patterns in phytoplankton biomass have emerged since the origin of Project

Quinte. In the Upper and Middle Bays, the period 1972-1977 showed very high total biomass

(7-16 mm3/L), followed by a year of low biomass associated with point source phosphorus

control. Biomass steadily increased from 1979 to 1984, then stabilized around 5-10 mm3/L until

the arrival of dreissenids in 1995. For the three years following the invasion of zebra mussels

Page 17: Diversity and Dynamics of Algal Viruses in the Bay of Quinte · Diversity and Dynamics of Algal Viruses in the Bay of Quinte Robin Marie Rozon Master of Science ... and Heather Niblock

6

phytoplankton biomass values were the lowest recorded, but then they rapidly increased in 1998

to pre-phosphorus control levels. A sharp decline followed, so much so that 2000 had the lowest

biomasses since 1970. Currently, phytoplankton levels have stabilized around the Upper Bay

RAP objective levels of 4-5 mm3/L (Nicholls and Millard, 2005; Nicholls and Carney, 2011).

Although these changes in phytoplankton growth in the bay over the last few decades have been

somewhat unpredictable, it is clear that the phosphorus abatement program and the invasion of

zebra mussels have had a major impact on algal growth in the Bay of Quinte. While it is

reassuring that algal biomass has been reduced in recent decades, the observed shifts in

community composition with increased abundance of harmful species like Microcystis represent

a new and unpredicted management challenge. Despite the uniqueness and thoroughness of

Project Quinte’s 40 seasons of field data, it is notable that important constituents in Bay of

Quinte food webs, such as algal viruses, have not been studied. Inclusion of these microbes in

Bay of Quinte research will certainly provide a more complete picture of the Bay’s algal

ecology, and may highlight unforeseen complexity helping explain why shifts in Bay of Quinte

algal communities have been unpredictable.

Aquatic Viruses

Historical Perspective on Aquatic Virus Research

Forty years ago, R. M. Brown wrote a seminal review on little known, little understood

algal viruses (Brown, 1972). Brown summarized previous studies describing the isolation of

viruses that infect the blue-green algae Lyngbya, Plectonema and Phormidium (Safferman and

Morris, 1963), as well as other studies that provided the first evidence that viruses could infect

eukaryotic algae such as Chlorella pyrenoidosa (Zavarzina and Protsenko, 1958). With the

discovery of the bacteriophage nearly 50 years prior, Safferman and Morris (1963) approached

their discovery of a blue-green algae virus with some skepticism. They provided evidence for

viruses of cyanobacteria by showing that filtered material could lyse growing cultures, phage-

like particles were present in electron micrographs, and by demonstrating the virus’s host

specificity (Safferman and Morris, 1963; 1964). Thus, at the time of Brown’s review the idea of

viruses being the cause of rapid algal population changes was not new; a number of prokaryotic

viruses had been isolated and virus-like particles had been observed for a number of eukaryotic

hosts. While these observations of viruses infecting algae were profound, at the time the

Page 18: Diversity and Dynamics of Algal Viruses in the Bay of Quinte · Diversity and Dynamics of Algal Viruses in the Bay of Quinte Robin Marie Rozon Master of Science ... and Heather Niblock

7

possibility that algal viruses were highly abundant and widespread in marine and freshwater

environments was not recognized.

Later, the observation of 5 x 106

to 2.5 x 108 virus-like particles per milliliter of water in

transmission electron micrographs of a variety of seawater samples represented a major

paradigm shift in marine microbiology (Bergh et al., 1989). Bergh’s discovery of abundant and

widespread viruses has since been supported by complementary methods such as

epifluorescence microscopy and flow cytometry which also estimate that all seawater samples

contain from 106

to 108 viruses/mL; there are roughly 10

30 viruses in the ocean making them the

most abundant biological entity in the ocean (Suttle, 2005). A study by Proctor and Fuhrman

(1990) also supported the observation that marine viruses were highly abundant, but also

emphasized that they were ecologically relevant. Similar observations of highly abundant

viruses have been noted for freshwater ecosystems, particularly in samples from numerous lakes

in Quebec, Canada that had on average 1.1x108 virus-like particles/mL (Maranger and Bird,

1995), and Lake Ontario that had 2.6x107 virus-like particles/mL (Gouvêa et al., 2006). With the

discovery of the most abundant biological entity in the ocean came the realisation that little was

known about the ecology of algal viruses; counts provided total virus estimates, but gave no

information on virus diversity or potential hosts.

Eukaryotic algal virus research prior to 1979 was strictly observation and speculation. It

wasn’t until after viruses were successfully isolated that viral infection of eukaryotic algae was

demonstrated. The first eukaryotic algal virus isolated was a lytic virus infecting Micromonas

pusilla, a small marine eukaryotic member of Prasinophyceae (Mayer and Taylor, 1979). Mayer

and Taylor observed small polyhedral virus-like particles within M. pusilla under transmission

electron microscope (TEM) and subsequently lysed otherwise healthy M. pusilla cultures using

filtered medium from an infected culture. Through TEM, they successfully observed the

infection and lysing of a host cell by this algal virus. The first freshwater eukaryotic algal isolate

was obtained shortly after when PBCV-1, a large double-stranded DNA virus which infects

Chlorella-like algae, was isolated from Chlorella NC64A, obtained from Lake Muscatine, USA

(Van Etten et al., 1983). A particularly interesting viral isolate was the Acanthamoeba

polyphaga mimivirus (La Scola et al., 2003). This virus contained a record sized genome (1.2

megabase pairs) and was the largest known virus at the time (Raoult et al., 2004). This

mimivirus has been classified into its own group, Mimiviridae, as part of the nucleocytoplasmic

Page 19: Diversity and Dynamics of Algal Viruses in the Bay of Quinte · Diversity and Dynamics of Algal Viruses in the Bay of Quinte Robin Marie Rozon Master of Science ... and Heather Niblock

8

large DNA virus (NCLDV) family as the closest relative of Phycodnaviridae (viruses that infect

eukaryotic algae). From studies of these model virus systems as well as many others, advances

in algal virus ecology related to knowledge of virus burst sizes (range 102 to 10

5), host

specificities (with the exception of viruses of brown algae, there are no known algal viruses that

infect more than one species), persistence in the environment (thought to range from days to

weeks), and genome composition (RNA, ssDNA, dsDNA, etc) have highlighted the biological

complexity of these algal parasites (Short, 2012). While these concepts are important to the

understanding of virus ecology, they don’t demonstrate the ecological significance of algal

viruses in aquatic food webs.

By observing algal viruses in the laboratory and in situ, the impact of viruses on the

phytoplankton community can be inferred. For example, an early study by Suttle at al. (1990) on

primary productivity of phytoplankton infected by viruses suggested that a wide range of

primary producers, including diatoms, cryptophytes, prasinophytes and cyanobacteria could be

susceptible to viral infection. The infection of a phytoplankton community by viruses obtained

from the same seawater sample resulted in a 78 % reduction of primary production indicating

that viruses and their specific hosts occur in close spatial and temporal proximity. The in situ

observations of Emiliania huxleyi, a marine coccolithophorid, and large viruses-like particles by

Bratbak et al. (1993) suggested that host blooms are succeeded by increased virus abundances

and that viruses accounted for 25-100 % of the net mortality of E. huxleyi under non-limiting

nutrient and low nitrate conditions. It is interesting to note that these studies were repeated over

four years in the same Norwegian Fjords, and different virus and phytoplankton abundance

trends were observed each year (Bratbak et al., 1993). The studies by Suttle at al. (1990) and

Bratbak et al. (1993) suggest that viral infection is an important regulator of phytoplankton

community structure.

Phytoplankton community succession is a well-studied area of research; however, the role

that viruses play in host succession is still unresolved. When host species which are differently

susceptible to different strains of virus coexist, it creates the opportunity for host succession.

Viral infection of the most abundant or most productive host species could lead to the rapid

decline of that specific algae species by “killing the winner”, thereby making way for other

hosts to take advantage of the newly available resources (Thingstad, 2000; Winter et al., 2010).

As host densities decline, the probability of a virus encountering its host also declines, leaving

Page 20: Diversity and Dynamics of Algal Viruses in the Bay of Quinte · Diversity and Dynamics of Algal Viruses in the Bay of Quinte Robin Marie Rozon Master of Science ... and Heather Niblock

9

scientists to speculate which host population will be the next ‘winner’. While these theories of

virus ecology are compelling, empirical support is lacking and viral dynamics are still, for the

most part, unresolved; it is still unknown if some virus species are ‘drivers’ in host blooms

dynamics while some species are ‘passengers’ along for the ride (Short, 2012). Therefore it is

crucial to determine what viruses are present in the environment and in what abundance.

Virus Diversity and Dynamics in Aquatic Systems

With advancements in molecular biology, tools to amplify and examine environmental

DNA became available to scientists in 1983 with the invention of the polymerase chain reaction

(Mullis et al., 1986). With the development of the polymerase chain reaction (PCR), a

cultivation-free method of virus identification became possible. The first efforts to uncover the

diversity of eukaryotic algal viruses came from Chen and Suttle (1995) with the development of

primers specific to the DNA polymerase gene (polB) of Phycodnaviruses. Not only did Chen

and Suttle amplify DNA from the aforementioned virus isolates (Chloroviridae and

Prasinoviridae), but they were also able to amplify DNA from natural virus communities,

creating the first technique to rapidly detect eukaryotic viruses in the environment (Chen et al.,

1996). This work paved the way for advancements in quantitative molecular studies of

eukaryotic algal viruses in both marine and freshwater environments. Of particular significance

to this study are the diversity studies by Short and Short (2008) and Short et al. (2011b). In

2008, Short and Short examined the diversity of Phycodnaviruses using the polB gene by

sampling four sites in two distinct geographic locations. They found and sequenced a large

number of unique gene sequences and through phylogenetic analysis determined that cultured

marine viruses were not genetically distinct from the freshwater viruses identified during their

study. Three years later, Short et al. (2011b) used new and extant polB PCR primers to amplify

and identify gene fragments from Lake Ontario that were related to genes from cultivated

prasinoviruses, chloroviruses, and prymnesioviruses; interestingly this first report of freshwater

prymnesioviruses demonstrated the existence of viruses that had never before been observed in

freshwater. While the polB primers allow an examination of algal virus diversity, they may

provide a limited picture of diversity since they are biased towards prasinoviruses and

chloroviruses (Short 2012). Fortunately these primers are not the only molecular approach

available to examine algal virus diversity.

Page 21: Diversity and Dynamics of Algal Viruses in the Bay of Quinte · Diversity and Dynamics of Algal Viruses in the Bay of Quinte Robin Marie Rozon Master of Science ... and Heather Niblock

10

Sequencing genomes of viruses that infect marine prymnesiophytes and haptophytes

revealed conservation among major capsid protein (MCP) genes. By developing universal algal

virus PCR primers targeting MCP gene fragments, it was possible to infer the genetic

relatedness of MCP sequences from cultured viruses and environmental clones (Larsen et al.,

2008). A recent study by Park et al. (2011) demonstrated that the use of both DNA polymerase

and MCP PCR methods to study the diversity of prasinoviruses, chloroviruses, and

prymnesioviruses simultaneously enabled the discovery of wide genetic diversity within similar

virus groups across different geographic locations.

The isolation of cyanophage infecting the toxic bloom forming cyanobacteria Microcystis

aeruginosa from a freshwater lake in Japan (Yoshida et al., 2006) is particularly important to

HAB (Harmful Algal Bloom) research in general, and this study of viruses in the Bay of Quinte

specifically. By developing PCR primers that specifically target genes for Microcystis

aeruginosa phage tail sheath proteins, it was possible to identify the cyanophage in the

environment and assess its potential impact on M. aeruginosa blooms (Takashima et al., 2007).

Since Chen and Suttle (1995), a number of algal virus genes have been studied and numerous

PCR primers have been developed creating tools that can be used to study the environmental

diversity of a broad spectrum of algal virus (Short et al., 2010). Following the implementation of

methods to examine algal virus diversity in marine and freshwater environments, the dynamics

of these viruses can be explored via the development of real-time, quantitative molecular

techniques.

Quantitative studies of algal virus dynamics are still relatively rare in the literature, but

several studies have used quantitative polymerase chain reaction (qPCR) to investigate algal

virus dynamics. Of particular relevance to this Bay of Quinte study is research on the dynamics

of algal viruses in Lake Ontario. Short and Short (2009) used qPCR to track the abundance of

three viruses over 13 months in weekly samples. Using these seasonal abundance patterns, Short

and Short demonstrated that some algal viruses persisted at low abundances during the winter,

while others were more transient, possibly being brought in by riverine flow. Further, continuing

research over longer monitoring periods demonstrated that many viruses are persistent in the

environment suggesting that seed-bank virus populations may be an ecologically important

factor in phytoplankton community succession (Short et al., 2011a). Although Gouvêa (2006)

sampled one station in the Bay of Quinte for total virus abundance (2.09 x 107 particles/mL), no

Page 22: Diversity and Dynamics of Algal Viruses in the Bay of Quinte · Diversity and Dynamics of Algal Viruses in the Bay of Quinte Robin Marie Rozon Master of Science ... and Heather Niblock

11

research has specifically examined algal viruses in the Bay of Quinte. However, if the paradigm

that diverse phytoplankton and viruses are present in all freshwater environments holds true,

then one would expect to find a diverse and dynamic community of viruses in the Bay of

Quinte.

Objectives and Research Questions

The Bay of Quinte and its rich history of seasonally occurring algal blooms present a

unique opportunity to investigate the roles that viruses play in complex phytoplankton

dynamics. Groundbreaking research into algal viruses in the Bay of Quinte could contribute to

understanding freshwater viruses and their impact on phytoplankton communities, as well as the

ecology of the hazardous algal blooms that have plagued the Bay of Quinte for many years. This

study extended previous research on freshwater viruses by increasing the scope of diversity

studies based on algal virus DNA polymerase genes (Short & Short 2008, Short et al. 2011b) by

including other algal virus marker genes such as MCP (Phycodnaviruses) and sheath protein

(Microcystis phage) to examine diversity in the Bay of Quinte. This study then examined the

dynamics of several distinct algal virus taxa across space and time to uncover patterns in algal

virus seasonality. The outcomes of this project were realized through a series of molecular

approaches designed to address the following questions: What virus populations are present in

the Bay of Quinte at a given location and time? Can the use of complementary PCR methods

targeting several different marker genes reveal greater diversity than methods based on any one

marker alone? How do virus abundances in the Bay of Quinte fluctuate temporally over the

sampling season, and spatially between stations? Does the seasonal timing of increased virus

abundance coincide with increased phytoplankton biomass? Do taxonomically related virus

populations show similar seasonality? And finally, are viruses patchily distributed throughout

the Bay such that trends in virus abundances are similar in proximal locations?

Page 23: Diversity and Dynamics of Algal Viruses in the Bay of Quinte · Diversity and Dynamics of Algal Viruses in the Bay of Quinte Robin Marie Rozon Master of Science ... and Heather Niblock

12

Chapter 2: Materials and Methods

Bay of Quinte algal virus communities were characterized using a variety of approaches.

Algal virus diversity was examined through sequencing clone libraries of marker gene

fragments amplified via PCR. Virus population dynamics were examined using qPCR with

primers and probes that were designed to target ten different virus genes observed in this study

as well as previous studies (Short and Short, 2009; Short et al., 2011a). Individual virus

abundances were compared to examine trends within and between taxa, and a variety of cluster

analyses were used to examine temporal and spatial patterns of virus abundance.

Study Sites and Sample collection

In collaboration with the Great Lakes Laboratory of Fisheries and Aquatics Sciences

(GLLFAS) of Fisheries and Oceans Canada (DFO), samples were collected approximately

biweekly during the 2011 sampling season, from early May to late October, from three long-

term study sites (Belleville, Napanee, and Hay Bay) in the Bay of Quinte. An additional 9

samples were also collected from various stations in the Bay of Quinte during an intensive

spatial survey conducted from September 13th

to 15th

(Table 2). Each water sample was an

integrated epilimnetic sample collected from the surface to one meter above the thermocline, or

in the absence of stratification, to twice secchi depth to a maximum of 1 meter from the bottom,

as determined by a Hydrolab Minisonde 5 (Hach Hydromet; Colorado, USA). As per the Quinte

RAP, complementary field data and laboratory analyses conducted by Fisheries and Oceans

Canada will be made available in the Project Quinte Annual Report for 2011. Of particular

importance here are chlorophyll a concentration estimates which were used as a proxy for host

abundance.

Following collection, water samples were transported on ice back to the Canada Center for

Inland Waters in Burlington, Ontario, for processing within 12 hours of field sampling. As

described in Short and Short (2008), each virus sample was filtered using a 0.45 µm pore-size

Durapore PVDF membrane filter (Millipore; Billerica, USA) and the filtrate was stored at 4°C

until it could be transported to the University of Toronto Mississauga for further processing.

From each sample, 72 mL of the filtrate was centrifuged in a SW32-Ti rotor at 25°C for 3.5

hours at 118,000 X g. Following centrifugation, the supernatant was decanted and 300 μL of 10

mM TrisCl (pH 8.0) was added to each centrifuge tube. All tubes were soaked overnight at 4°C,

Page 24: Diversity and Dynamics of Algal Viruses in the Bay of Quinte · Diversity and Dynamics of Algal Viruses in the Bay of Quinte Robin Marie Rozon Master of Science ... and Heather Niblock

13

vortexed for 30 seconds and pelleted material was resuspended using a pipettor. The

resuspended material, henceforth called a virus concentrate, was transferred into a screw-cap

microcentrifuge tube and stored at -20°C.

PCR, cloning and Sequencing

As described in Short and Short (2008), to prepare samples for PCR, 50 µL of each viral

concentrate was subjected to a freeze-thaw treatment to release viral DNA. This procedure

involved heating the concentrate to 95°C for 2 minutes then freezing at -20°C until solid; this

hot-cold cycling was repeated 3 times for each sample. In an attempt to address past concerns

about target biases of individual algal virus PCR primer sets and obtain a broader perspective on

algal virus diversity in the Bay of Quinte, three different sets of primers were used in this study.

Although the AVS and MCP primer sets described below both amplify Phycodnavirus gene

fragments, past work suggests that they are biased towards different subsets of the

Phycodnaviridae and can be used in a complementary manner (Park et al., 2011). Furthermore,

because there is particular interest in the ecology of toxic bloom-forming cyanobacteria, such as

Microcystis aeruginosa, sheath primers targeting M. aeruginosa phage were also used.

AVS:

The AVS primers specifically target a ≈ 700 bp fragment of virus DNA polymerase

genes (polB) in Phycodnaviruses, viruses that infect eukaryotic algae (Chen and Suttle,

1995). DNA polymerase polB genes were amplified from each sample in two rounds of

PCR. First round PCR was made up of reagents in quantities listed in Table 3 (AVS)

with a negative control where nuclease-free water substituted viral template. First round

PCR thermal cycling followed the conditions as in AVS-40 (Table 4). To increase the

yield of amplified DNA from the AVS primers, a second round of PCR was conducted

using the same quantities of reagents, with the exception of viral template (2 μL). For

second round PCR, viral template was obtained by excising, with disposable glass

Pasteur pipettes, agarose plugs from bands visualized in a gel of first-round PCR

products. The excised plugs were placed in tubes with 100 μL of 10 mM TrisCl and

were heated (65°C for 20 minutes) to elute the DNA from the agarose. Two microlitres

Page 25: Diversity and Dynamics of Algal Viruses in the Bay of Quinte · Diversity and Dynamics of Algal Viruses in the Bay of Quinte Robin Marie Rozon Master of Science ... and Heather Niblock

14

of the resulting eluant was used as template for the second round PCR (Short and Suttle,

2003), and thermal cycling followed AVS-25 (Table 4).

MCP:

The MCP primers target ≈ 400 bp to ≈ 500 bp gene fragments in the conserved region of

the major capsid protein in Phycodnaviruses that infect some prymnesiophytes,

haptophytes and Chlorella-like algae (Larsen et al., 2008). Major capsid protein genes

were amplified from select samples in two rounds of PCR. Quantities of reagents used in

first and second round PCR are listed as “MCP” in Table 3, with a negative control

where nuclease-free water substituted viral template. Second round PCR was identical to

first round PCR with the exception that plugs of gel bands from the first round PCR

electrophoresis (as described above for AVS reactions) were the source of template

DNA. Thermal cycling followed MCP50-35 for both first and second round PCR (Table

4).

Sheath:

Sheath primers target a sheath protein gene fragment ≈ 950 bp long in cyanophages

which specifically infect a toxin producing strain of Microcystis aeruginosa (Yoshida et

al., 2006). Sheath protein genes were amplified from select samples in a single round of

PCR using the primers Sheath F2 & R2 (Takashima et al., 2007). Quantities of reagents

used in PCR are listed as “Sheath” in Table 3, with a negative control where nuclease-

free water substituted viral template. Thermal cycling followed the Sheath profile (Table

4).

After thermal cycling, all of the PCR products were mixed with loading dye and loaded

into a 1.5 % LE agarose gel (Promega; Fitchberg, USA) and were electrophoresed at 6.6 V/cm

for 60 minutes in 1x TAE (40mM TrisCl, 20mM Acetic acid, 1 mM EDTA; pH 8.0).

GeneRuler® 100 bp Plus ladder (ThermoFisher Scientific; Waltham, USA) was used as a

molecular weight marker. The gel was stained with 0.5 μg/ml ethidium bromide for 30 minutes

followed by 15 minutes of destaining in water. PolB products from samples collected on June

7th

, 2011 were an exception; following first and second round PCR, PCR products were

Page 26: Diversity and Dynamics of Algal Viruses in the Bay of Quinte · Diversity and Dynamics of Algal Viruses in the Bay of Quinte Robin Marie Rozon Master of Science ... and Heather Niblock

15

electrophoresed for 90 minutes, but first round PCR products were stained for 35 minutes and

destained for 35 minutes while second round followed the aforementioned staining times. The

gels were visualised using a Molecular Imager Gel Doc XR (Bio-Rad Laboratories; Hercules,

USA). In all cases, agarose gels were not exposed to UV light for more than five seconds.

Sheath PCR bands from Hay Bay on June 7th

and July 19th

, 2011 were excised from the

gel and purified using a QIAquick Gel Purification Kit (QIAGEN; Valencia, USA) following

the manufacturer’s recommendations, with the exception of a three minute spin instead of the

recommended one minute following the ethanol wash, and DNA was eluted in 30 μl of 10 mM

TrisCl and let stand for two minutes as opposed to one minute. The purified PCR product was

quantified using a NanoDrop ND-1000 spectrophotometer (NanoDrop Technologies;

Wilmington, USA) before being sent for direct sequencing from the forward primer at the

Center for Applied Genomics at the Hospital for Sick Children (Toronto, Ontario, Canada).

With regard to AVS and MCP, bands from second round PCR of samples collected from

Belleville on June 7th

, August 16th

(MCP only) and October 12th

, 2011 were excised from the gel

and DNA was extracted using a QIAquick Gel Extraction Kit (QIAGEN) following the

manufacturer’s recommendations, with the exception of a three minute spin instead of the

recommended one minute following the ethanol wash, and DNA was eluted in 30 μl of 10 mM

TrisCl and let stand for two minutes as opposed to one minute. Once purified, the PCR products

from the excised bands were cloned using pGEM-T Vector Systems (Promega) according to the

manufacturer’s recommendations for ligation (three μL of purified DNA fragments from the

PCR reactions were used in each ligation). Using Max efficiency DH5α competent cells

(Invitrogen; Carlsbad, USA), the transformation process followed the manufacturer’s

recommendations for heat shock and recovery. Two hundred microlitres of transformed bacteria

were spread on LB agar (10 g/L tryptone, 5 g/L yeast extract, 10 g/L NaCl and 15 g/L agar)

bacteriological plates that were prepared with 100 μg/mL carbenicillin, and were smeared with

100 μL 0.1M IPTG (isopropyl β-D-1-thiogalactopyranoside) and 20 μL 50 mg/mL of X-Gal

(bromo-chloro-indolyl-galactopyranoside) immediately before use. Plates were incubated at

37°C overnight.

Transformants were screened for recombinant plasmids using a single round of PCR as

per the above conditions for the respective primers, with the exception that bacterial cells were

used as templates for the reactions by transferring cells from the edge of a colony by scraping

Page 27: Diversity and Dynamics of Algal Viruses in the Bay of Quinte · Diversity and Dynamics of Algal Viruses in the Bay of Quinte Robin Marie Rozon Master of Science ... and Heather Niblock

16

with a sterile pipette tip and stirring the pipette tip directly into the PCR reaction mixture. From

each transformation, 18 colonies that were confirmed to contain recombinant plasmids were

picked off the plate and incubated overnight at 37°C, 250 RPM for 16 hours in 5 mL of LB

broth with 100 μg/mL carbenicillin. Following the overnight incubation, plasmid DNA was

purified from the colonies using a QIAprep Spin Miniprep Kit (QIAGEN), and quantified using

a NanoDrop ND-1000 spectrophotometer and sent for automated sequencing from the M13f

priming site on the plasmid backbone at the Center for Applied Genomics at the Hospital for

Sick Children (Short and Short, 2008). Only full-length sequences AVS (≈ 700 bp), MCP (≈ 400

bp or ≈ 500 bp), Sheath (≈ 950 bp), were used for further analysis.

Sequence Analysis and Primer Design

Sequence identity matrices were generated in BioEdit 7.0.9.0 (Hall, 1999) and allowed for

nucleic acid sequences greater than 97 % identical to be grouped together and hereafter referred

to as operational taxonomic units (OTUs). Separate alignments for polB, MCP and Sheath were

created using inferred amino acid sequences using MUSCLE (Multiple Sequence Comparison

by Log-Expectation) with the default parameters in MEGA 5.0 (Molecular Evolutionary

Genetics Analysis; Tamura et al., 2011). Aligned polB sequences and MCP sequences were

compared phylogenetically using MEGA 5.0 by constructing independent neighbor-joining trees

based on the Jones-Taylor-Thornton (JTT) model of amino acid substitution (Jones et al., 1992)

and bootstrap from 500 replicates with complete deletion options. It is important to note that due

to the small number of unique Sheath sequences obtained from this study and available in

GenBank, only BLAST analyses were used to characterize the Bay of Quinte sequences from

Sheath PCR. For the polB, MCP and Sheath clone libraries developed in this study, percent

coverage was determined using the calculation C=1-(N/n) where C was the homologous

coverage, N was the number of singleton sequences and OTUs, and n was the total number of

sequences in the sample. Bay of Quinte OTUs in the polB and MCP phylogenies as well as

sequences from previously generated polB clone libraries (Short and Short, 2009; Short et al.,

2011a) were used to choose targets for qPCR primer and probe design. MEGA 5.0 and Adobe

Illustrator CS (Adobe Systems, USA) were used for tree viewing and drawing.

As previously described in Short and Short (2009), using default parameters in Beacon

Designer 7.0 (Premier Biosoft International, USA), TaqMan probes were designed for four

Page 28: Diversity and Dynamics of Algal Viruses in the Bay of Quinte · Diversity and Dynamics of Algal Viruses in the Bay of Quinte Robin Marie Rozon Master of Science ... and Heather Niblock

17

MCP sequences; two of which were from the Mimivirus infecting prasinophytes group (365-

M5.3; 252-M5.13) and the other two from the Mimivirus infecting prymnesiophytes group (399-

M5.4; 356-M5.14rc), and both Sheath OTUs (Sheath 253; Sheath 282). Each qPCR assay was

optimized for highest quality and greatest number of mismatches with non-target sequences.

Closest non-target sequences were determined by creating alignments of nucleic acid sequences

using MUSCLE in MEGA 5.0, and sequence identity matrices generated in BioEdit 7.0.9.0. The

number of base pair mismatches was counted for the most closely related sequences to each

qPCR target. In addition to the six qPCR assays developed for this study, four extant primer and

probe pairs were used to expand the breadth of virus communities available for quantitative

analysis (Table 5). These assays included two primer and probe sets previously described in

Short and Short (2009), LO1b-49 and LO1a-68, as well as two primer and probe pairs

previously described in Short et al. (2011a), LO.20May09.33 and LO Jul.16.20.

Quantitative PCR

The newly designed quantitative assays included probes which were 5’ labelled with FAM

(6-carboxyfluorescein) as a fluorescent reporter and 3’ labelled with Zen Internal Quencher

(Integrated DNA technologies; Coralville, USA) as a quencher. Of the extant primer and probes

pairs, only LO Jul.16.20 was 3’ labelled with Zen Internal Quencher (Integrated DNA

technologies) as a quencher; the other three assays were 3’ labeled with Iowa Black FQ

(Integrated DNA technologies).

As described in Short and Short (2009), quantitative PCR via the 5’ nuclease assay was

conducted on every Bay of Quinte virus concentrate with each primer and probe set using a

MX3000P qPCR system (Stratagene; Cedar Creek, USA). Each sample measurement was

replicated three times and every set of reactions included eight 10-fold serially diluted standards

(on average, ranging from 3.0 x 100 to 3.0 x 10

7 molecules per reaction) run in duplicate, along

with three no-template controls where template was substituted with nuclease-free water. Each

of the ten standards were created from cloned fragments of the target sequence that were

linearized by restriction digest, purified by agarose gel electrophoresis, extracted via QIAquick

gel extraction kit (QIAGEN), and quantified using a NanoDrop ND-1000 spectrophotometer

(NanoDrop Technologies). Experiments to test the efficacy and specificity of the newly

designed qPCR primers and probes were conducted as described in Short and Short (2009)

Page 29: Diversity and Dynamics of Algal Viruses in the Bay of Quinte · Diversity and Dynamics of Algal Viruses in the Bay of Quinte Robin Marie Rozon Master of Science ... and Heather Niblock

18

where Ct (cycle threshold) values of target and closest non-target at ca. 3 x 107 molecules were

compared. The extant primer and probe pairs had previously undergone validity testing in Short

and Short (2009) and Short et al. (2011a) respectively. Reaction conditions for the newly

designed primer and probe pairs (Mimivirus-Pras 252, Mimivirus-Pras 356, Mimivirus-Prym

399, Mimivirus-Prym 356, Sheath 253 and Sheath 282) follow assay group “A” in Table 6,

while the extant assays Chlorovirus and Prasinovirus49 follow group “B”. The other extant

assays, Prasinovirus16.20 and Prasinovirus68, follow groups “C” and “D”, respectively. The

thermal cycling conditions for each primer and probe pair consisted of an initial denaturation

step of five minutes at 95°C followed by 40 cycles of 15 seconds at 95°C and one minute at

60°C.

Analysis

To illustrate the dynamics of the ten individual viruses examined in this study, the virus

gene abundance was plotted in line graphs, smoothed with a modified cubic spline (Liengme,

2008), by virus and by sampling site. Error bars were calculated based on the standard deviation

of qPCR determined gene abundance, and where the error bars are not visible they were smaller

than the plotted line. A box and whiskers plot was created to illustrate and summarize seasonal

abundances for each virus gene at each station. To illustrate the relationship between the sum of

the gene abundance of all ten viruses quantified in this study and chlorophyll a concentrations

(µg/L) as a proxy for phytoplankton host abundance, linear regression analysis was performed

for each biweekly station. The line graphs, box and whiskers plot as well as the linear regression

were created using Microsoft Excel 2007 (Microsoft Corporation; Redmond, USA).

Tests for normality and sphericity were conducted and since gene abundances were not

normally distributed and were not homoscedastic, non-parameteric tests were used for statistical

analyses. Friedman analysis of variance by ranks was used to compare biweekly virus

abundances from each individual virus gene to each other individual virus gene to determine if

there were significant differences between gene abundances within a station (all ten viruses at

each biweekly station) and between stations (each individual virus between all three biweekly

stations). Results with a significance value of p < 0.05 were subjected to post-hoc analysis with

Wilcoxon Signed-Rank Test with a Bonferroni correction applied. The statistical tests were

completed using IBM SPSS Statistics 19 (IBM; Armonk, USA).

Page 30: Diversity and Dynamics of Algal Viruses in the Bay of Quinte · Diversity and Dynamics of Algal Viruses in the Bay of Quinte Robin Marie Rozon Master of Science ... and Heather Niblock

19

To resolve if taxonomically related virus populations cluster together based on abundance

data, distance measures were created using pair-wise Pearson Correlation and Bray-Curtis

dissimilarity matrices based on the biweekly abundance data from each virus gene. The Pearson

Correlation values were converted into distance measures by 1 – r = D, where r was the Pearson

Correlation coefficient and D was the dissimilarity between two data points. The abundances of

each virus gene were compared to the abundances of every other virus gene at all stations in a

pairwise manner, and similar discrete pair-wise comparisons were conducted for the abundances

at each individual station. In an additional discrete pair-wise comparison, the abundances of

each virus gene were compared to the abundance of each other virus gene at each station

sampled during the September spatial survey. The dissimilarity matrices were transposable so

that separate dendrograms of virus genes and sampling events were created from the same data

set. Using four different clustering methods [average (unweighted pair group method with

arithmetic mean; UPGMA), furthest neighbor (complete-linkage), nearest neighbor (single-

linkage) and Ward’s method], dendrograms were created for each cluster analysis. Matrix

calculations were performed using R Statistical Computing ver. 2.15.1 (R Core Team; Vienna,

Austria) and cluster analysis was done using R Statistical Computing and MEGA (UPGMA

only).

Page 31: Diversity and Dynamics of Algal Viruses in the Bay of Quinte · Diversity and Dynamics of Algal Viruses in the Bay of Quinte Robin Marie Rozon Master of Science ... and Heather Niblock

20

Chapter 3: Results

Algal Virus Diversity Study

Algal virus DNA polymerase genes

PCR with algal virus specific polB primers (AVS) was used to generate clone libraries of

amplified fragments from samples collected from Belleville on June 7th

and October 12th

, 2011.

Sequencing generated 29 usable sequences; 25 sequences were discarded because of incomplete

sequencing of the gene fragments, vague base calling or did not display the highly conserved

polB amino acid motif YGDTDS (Short and Suttle, 2002). Based on a sequence identity matrix

nine unique sequence types were identified from the 29 usable sequences. Of these nine unique

sequences types, five were singletons and four were OTUs, or groups of sequences more than 97

% identical. Representative sequences for each unique sequence type were compared to

cultivated and environmental virus polB gene fragments and all polB gene fragments obtained

from the Bay of Quinte were more closely related to genes from cultivated prasinoviruses than

they were to polB genes from other virus genera (Figure 2). Using C=1-(N/n), the percent

coverage of the polB clone libraries was calculated to be 69.0 %.

Algal virus major capsid protein genes

The Phycodnavirus-specific MCP clone libraries were generated from PCR of samples

collected from Belleville on June 7th

, August 16th

and October 12th

, 2011. Sequencing generated

61 usable sequences; nine were discarded for vague base calling and incomplete gene

sequencing. Based on a sequence identity matrix, 24 unique sequences were identified of which

16 were singletons and eight were OTUs. Representative sequences were compared

phylogenetically to cultivated and environmental virus MCP gene fragments (Figure 3). The

majority of Bay of Quinte MCP sequences were ~500 bp sequences that were most closely

related to one of two types of mimivirus-like viruses that are classified based on the types of

hosts they infect; one group are mimiviruses that infect prymnesiophytes, the other are

mimiviruses that infect prasinophytes. The other MCP sequences (~400 bp) were most closely

related to cultivated viruses belonging to the genus Prasinovirus. The percent coverage of the

MCP clone library was calculated to be 60.7 %.

Page 32: Diversity and Dynamics of Algal Viruses in the Bay of Quinte · Diversity and Dynamics of Algal Viruses in the Bay of Quinte Robin Marie Rozon Master of Science ... and Heather Niblock

21

Microcystis phage sheath protein genes

Microcystis sheath protein clone libraries were created from samples collected from Hay

Bay on June 7th

, and July 19th

, 2011 and generated 13 sequences, that formed two distinct OTUs

comprised of ten (Sheath 282) and three (Sheath 253) sequences. BLAST and a sequence

identity matrix revealed that representatives of the two sheath OTUs, Sheath 253 and Sheath

282, were 95 % and 93 % identical to the cultured Microcystis phage Ma-LMM01 (Yoshida et

al., 2006) over 864 and 873 nucleotides, respectively. However, Sheath 253 and Sheath 282

were only 92 % identical to each other. The percent coverage of the Sheath clone library was

calculated to be 84.6 %.

Quantitative Analysis of Viral Dynamics

Pure DNA of the most closely related non-target gene fragments, based on information

from the sequence identity matrices and phylogenetic analyses, were used to test the primer and

probe specificities for each qPCR assay. QPCR detection of 107 non-target molecules ranged

from a Ct value of 22.95 to below the limit of detection, while qPCR detection of 107 target

molecules ranged from a Ct value of 12.07 to 16.91 (Table 7). This demonstrates that the

detection limits for the most closely related genes were at least three orders of magnitude higher

than the target genes (Sheath 253), and for some assays non-targets were not even detected

(Sheath 282). Amplification efficiencies of all qPCR reactions were consistently between 90-

100 %, as per recommendations by Dorak (2006) for accurate quantification via the 5’ nuclease

assay. It is important to note that this technique of viral quantification required that virus

abundances be inferred based on quantified gene abundances (Short and Short, 2008).

Virus Dynamics across Time and Space

Using qPCR, gene abundances were quantified and compared for water samples collected

from three Bay of Quinte stations throughout the 2011 growing season (May to October). Plots

of virus gene abundance over the year revealed complex and unique patterns for each of the

genes studied (Figure 4). Seasonal abundance patterns of some virus genes were relatively

consistent between stations, some had distinct single peaks in abundance, while others seemed

to experience several boom and bust cycles throughout the year. Virus gene abundance also

Page 33: Diversity and Dynamics of Algal Viruses in the Bay of Quinte · Diversity and Dynamics of Algal Viruses in the Bay of Quinte Robin Marie Rozon Master of Science ... and Heather Niblock

22

varied with virus groups, most notably between the two Microcystis phage genes, but also

between Prasinoviruses and mimiviruses infecting prasinophytes and prymnesiophytes.

Quantitative analysis of one of the putative M. aeruginosa phage genes, Sheath 282,

revealed oscillations that varied dramatically across the Bay of Quinte. At Hay Bay, Sheath 282

abundance peaked in July at around 256,000 gene copies/mL, whereas at Napanee it peaked

near the end of September (~22,800 gene copies/mL). In contrast, Sheath 282 abundances at

Belleville never exceeded ten gene copies/mL. Sheath 253 abundances on the other hand were

much more consistent among the three sampling sites, with peak abundance occurring at

Napanee (5,300 gene copies/mL) in the spring. Chlorovirus abundances were also consistent

between stations, with abundances at both Napanee and Hay Bay peaking in early August three

times higher than at Belleville. Analysis of Prasinovirus49 revealed virus abundances below 500

gene copies/mL at all stations until October where gene copies at Belleville peaked at

approximately 2,500 gene copies/mL. A particularly interesting trend was revealed in the gene

abundances of Prasinovirus68 and Prasinovirus16.20; virus abundances at Belleville remained

below detection for the entire 2011 sampling season and gene detection was decreasingly patchy

with stations further out in the Bay, to where both prasinoviruses were regularly detected at Hay

Bay.

Another algal virus that was regularly detected at consistent levels at all three biweekly

Bay of Quinte stations was Mimivirus-Pras 356, peaking in the late summer at Belleville at

approximately 1,600 gene copies/mL. Contrary to Mimivirus-Pras 356 and historical Bay of

Quinte algae blooms, Mimivirus-Pras 252 was most abundant during the spring and early

summer (never more than 300 gene copies/mL), then in July, all Mimivirus-Pras 252 viral

activity dropped below four gene copies/mL for the rest of the season. Gene abundances of

Mimivirus-Prym 399 also revealed consistently high abundances in the spring and early summer

which fell below detection in July. However, unlike Mimivirus-Pras 252, abundances of

Mimivirus-Prym 399 returned to high levels in the early fall, where Hay Bay diverged from the

Upper Bay stations which peaked in late September (Belleville; 2,250 gene copies/mL). Of the

ten gene fragments examined in this study, the abundances of Mimivirus-Prym 356 were the

most consistent between the historical Bay of Quinte stations over the sampling season; virus

abundances at all three stations peaked from July through August, with the highest abundance

(16,000 gene copies/mL) observed at Belleville.

Page 34: Diversity and Dynamics of Algal Viruses in the Bay of Quinte · Diversity and Dynamics of Algal Viruses in the Bay of Quinte Robin Marie Rozon Master of Science ... and Heather Niblock

23

When the abundances of all ten virus genes are plotted in a single graph for each station,

the highly dynamic oscillations of these viruses over a single growing season become apparent

(Figure 5). With all ten viruses plotted in each graph, virus succession is also more evident. For

example, at Belleville, as Mimivirus-Pras 252 drops to below detection, Prasinovirus49 becomes

more abundant. Some viruses are more consistent throughout the sampling season, particularly

Mimivirus-Prym 356 and Chlorovirus, while some are only present at certain times (i.e.,

Mimivirus-Pras 252). Of particular interest are incidences of repeated boom and bust, as seen in

Sheath 282. The uniqueness and complexity of virus abundance oscillations within the Bay of

Quinte is fascinating; the dramatic seasonality differences, particularly within virus groups were

unexpected. However, figures 4 and 5 illustrate the qualitative differences between virus

abundances only; no statistical significance can be inferred. Therefore, further analyses were

conducted to determine if the abundance trends seen in Figure 4 were significantly different.

Statistical Analysis of Viral Dynamics

Seeing that the virus gene abundance data set was obtained from environmental samples

where abundances can vary over several orders of magnitude, it was not surprising that the

values were non-normally distributed. In order to compare the variability of virus abundances

between viruses within a station and virus abundance across stations, a box and whiskers plot

was created (Figure 6). This was complemented by regression analysis (Figure 7) and a

Friedman non-parametric analysis of variance by ranks.

There was a statistically significant difference between each individual virus gene assay at

each station based on virus abundances from all sampling events. That is to say that abundances

of all ten viruses are significantly different from each other at Belleville (χ²(9) = 68.146, p <<

0.01), Hay Bay (χ²(9) = 59.847, p << 0.01) and Napanee (χ²(9) = 58.003, p << 0.01). Post-hoc

analysis with Wilcoxon Signed-Rank Test with a Bonferroni correction applied was conducted,

resulting in a significance level set at p = 0.001, but did not reveal any significant differences

between the abundances of any individual virus and any other virus, at any station. This was

likely due to the high number of multiple comparisons required, as well as being conservative

for possible type I errors. Nonetheless, inspection of the box and whisker plots reveals apparent

patterns in the abundance data (Figure 6). For instance, at all stations, Chlorovirus and

Mimivirus-Pras 356 are more similar with respect to their median abundance throughout the

Page 35: Diversity and Dynamics of Algal Viruses in the Bay of Quinte · Diversity and Dynamics of Algal Viruses in the Bay of Quinte Robin Marie Rozon Master of Science ... and Heather Niblock

24

Bay of Quinte than to any other virus group. In Belleville, Prasinovirus49 and Mimivirus-Prym

399 have similar range and median abundance, as do Mimivirus-Pras 252 and Sheath 253. In

Hay Bay, both Sheath 282 (which has a largest range) and Mimivirus-Pras 356 (smallest values)

have distinctly different abundance medians and ranges than other viruses. Prasinovirus49,

Prasinovirus16.20, Mimivirus-Pras 252, Sheath 253 and Mimivirus-Prym 399 were all detected

at similar ranges of abundance. In Napanee, Sheath 282 and Mimivirus-Prym 399 have similar

abundance means, as do Prasinovirus68, Prasinovirus16.20 and Mimivirus-Pras 356. The other

viruses, Sheath 253, Mimivirus-Pras 252 and Prasinovirus49 also share similar median

abundances.

By comparing the variance of virus abundance of individual viruses between stations, that

is to say, comparing the abundance of a given virus gene at one biweekly station to its

abundance at the other two biweekly stations, it was determined that there were statistically

significant differences for only four of the quantitative assays: Prasinovirus49, Prasinovirus68,

Prasinovirus16.20 and Sheath 282 (Table 8). Follow up with a post-hoc analysis with Wilcoxon

Signed-Rank Test with a Bonferroni correction applied was conducted, resulting in a

significance level set at p = 0.017. The post-hoc analysis examined specific pair-wise

relationships to determine if there were significant differences between any two stations. With

regard to Prasinovirus49, despite being significantly different overall, there were no significant

differences between any of the stations on a pair-wise basis. Conversely, analysis of the other

three virus groups between stations revealed significant differences between Hay Bay and the

other two stations for both Prasinoviruses, and between Hay Bay and Belleville only for Sheath

282 (Table 8).

Virus Abundance and Host Biomass

As previously discussed, the algae virus abundance reported in this study are inferred from

gene abundances obtained through qPCR and therefore gene abundance was a proxy for virus

abundance. The phytoplankton host abundances in this study were also measured via a proxy;

chlorophyll a. As a pigment found in the chloroplasts of algae, chlorophyll a has historically

been used as a measure of phytoplankton biomass in the Bay of Quinte (Minns et al., 2011).

Chlorophyll a concentrations from every sample from each biweekly station during the 2011

sampling season were plotted against the sum of all ten virus abundances at each station (Figure

Page 36: Diversity and Dynamics of Algal Viruses in the Bay of Quinte · Diversity and Dynamics of Algal Viruses in the Bay of Quinte Robin Marie Rozon Master of Science ... and Heather Niblock

25

7). Regression analysis of the sum of virus abundance versus chlorophyll a at Belleville

revealed a strong positive relationship between these proxies for host biomass and virus

abundance (R2

= 0.5316) with a significantly non-zero slope (p = 0.01). Close inspection of virus

and chlorophyll a data revealed that in Belleville, increases in host biomass and virus abundance

were in phase and peaked at the same time. However in the later portion of the sampling season,

the chlorophyll a and virus gene abundances were not as tightly coupled and ended the season

slightly out of phase (data not shown). Regression analysis of the same data from the Hay Bay

location revealed that the relationship of chlorophyll a and virus gene abundance was not

statistically significant (R2 = 0.036, p = 0.576). However, this station included a virus abundance

data point (Sheath 282 from July 19th

) that was at least an order of magnitude higher than all

others that was responsible for this weak relationship. When this exceptionally high Sheath 282

abundance value was removed from the data set, the relationship between virus and chlorophyll

a was strong (R2 = 0.6048) and statistically significant (p < 0.01). The sharp increase in virus

abundance due to Sheath 282 is not reflected in host biomass, nevertheless, fluctuations in host

biomass and virus abundance remain tightly coupled prior to and after the peak in virus

abundance (data not shown). Regression analysis at Napanee revealed a weak positive

correlation between virus abundance and chlorophyll a (R2 = 0.2591), however the slope of the

regression was not significantly non-zero (p = 0.133), which was likely due to a decoupling

between host biomass and virus abundances in the later part of the growing season (data not

shown).

Clustering by Taxonomic Classification

To explore the patterns of virus gene abundance qualitatively, cluster analysis was

conducted in several ways to determine if there were any robust conclusions that could be

reached with regards to fluctuations of virus gene abundances in the Bay of Quinte. After

matrices were generated through pair-wise comparisons of virus gene abundances at all stations,

as well as at each individual station, cluster analysis of each virus type in the Bay demonstrated

that genetically related viruses very rarely clustered together based on virus abundance, but that

the clustering patterns were very sensitive to sample location; i.e., the branching pattern of the

dendrograms created for all stations and each station independently were distinct (Figure 8).

Page 37: Diversity and Dynamics of Algal Viruses in the Bay of Quinte · Diversity and Dynamics of Algal Viruses in the Bay of Quinte Robin Marie Rozon Master of Science ... and Heather Niblock

26

Including the average linkage (UPGMA) method (Figure 8), four different clustering

methods (average linkage, nearest neighbour, furthest neighbour and Ward’s method) were used

to create dendrograms from each proximity (i.e., distance) measure. Using identical clustering

methods, the proximity measures (Pearson Correlation and Bray-Curtis dissimilarity) produced

different branching patterns from one another (Figure 9), but within a proximity measure, each

clustering method produced comparable or virtually identical branching patterns. Despite

different stations and proximity measures, each clustering method supported the observed

pattern that related viruses did not cluster together based on seasonal abundances.

Clustering by Spatial Distribution

Using the same clustering technique as above, virus abundances for all biweekly stations

were used to cluster sampling events as opposed to clustering of the virus groups. Regardless of

proximity measure or clustering method, cluster analysis revealed that the biweekly stations did

cluster together based on location, but rather more by sampling date (Figure 10). Once again

using the same clustering technique, virus abundances from the intensive, three day September

spatial survey were used to cluster sample sites (Figure 11). All proximity measures and

clustering methods revealed that, for the most part, the spatial survey stations cluster together

based on their spatial distribution (Figure 1; Table 2), with most of the Middle Bay stations

(HB4, HB, and P) clustering together and apart from the other Upper Bay stations (NR1, N,

BQ8, BQ9, BQ6, BQ7, B, and B2). It is also interesting to note that many of the samples

collected from nearby stations (i.e., BQ6, BQ7, and B) also clustered together based on virus

abundance.

Page 38: Diversity and Dynamics of Algal Viruses in the Bay of Quinte · Diversity and Dynamics of Algal Viruses in the Bay of Quinte Robin Marie Rozon Master of Science ... and Heather Niblock

27

Chapter 4: Discussion

Through molecular techniques, this seminal examination of algal virus diversity in the Bay

of Quinte has confirmed the presence of diverse Phycodnaviruses and Microcystis cyanophage

in the Bay. In addition, this study has provided the first evidence that PCR methods based on

Phycodnavirus MCP genes could be used to examine the diversity of freshwater algal viruses.

Quantitative molecular techniques were used to determine that the seasonality of certain virus

taxa differed in abundance across both time and space, regardless of their genetic similarity. By

comparing proxies for virus abundance and host biomass, this study supported the current

paradigm that that viruses are dependent on their hosts which, in turn, are subject to

environmental factors, both biotic and abiotic, that drive cycles of phytoplankton bloom and

decay, hereafter referred to as ecological drivers. Finally, via cluster analysis, this study

demonstrated that taxonomic classification or genetic relatedness cannot be used to predict virus

seasonality, and that the abundance of algal virus communities in the Bay of Quinte may be

more similar seasonally than geographically.

Algal Virus Diversity Study

Using four different sample dates, phylogenetic trees were constructed to examine the

diversity of viruses in the Bay of Quinte. Using the three clone libraries created from these

samples, phylogenetic trees were created (Figures 2 and 3) and percent coverage was calculated.

As the first investigation into algal viruses in the Bay of Quinte, the percent coverage values for

this study were lower than those reported in previous Lake Ontario Phycodnavirus diversity

studies (85 %; Short and Short, 2008; Short et al., 2011a) indicating that there remains the

potential for yet undiscoved algal virus diversity in the Bay. However, to provide a first glimpse

into algal virus diversity in this unique environment, higher clone library coverage through deep

sequencing efforts was not necessary.

As expected, the environmental polB sequences obtained from the Bay of Quinte were

biased towards prasinoviruses (Figure 2), a single genus within the Phycodnaviridae. While this

bias is likely the effect of a PCR primer bias, there is at least some evidence that the observed

bias of the AVS PCR primers towards prasinoviruses could actually be the result of their

predominance in natural environments (Clasen and Suttle, 2009). Whatever the case, almost all

of the environmental polB sequences from the Bay of Quinte form a single group with other

Page 39: Diversity and Dynamics of Algal Viruses in the Bay of Quinte · Diversity and Dynamics of Algal Viruses in the Bay of Quinte Robin Marie Rozon Master of Science ... and Heather Niblock

28

environmental sequences from Lake Ontario (Short and Short, 2008; Short et al., 2011a) that is a

sister clade to sequences from cultivated viruses that infect marine phytoplankton. This could

indicate that these freshwater sequences belong to a distinct group of viruses that are common in

Lake Ontario but have not yet been cultivated. The closest cultured relative of these Bay of

Quinte sequences is an Ostreococcus virus (Derelle et al., 2008) however since viruses infecting

the same types of hosts generally cluster together (Chen and Suttle, 1996) and this host genera is

made up of many closely related prasinophytes, it is not possible to speculate on the actual

identity of the host of these Bay of Quinte viruses. It is interesting to remark that all Bay of

Quinte polB samples cluster with prasinoviruses, yet prasinophytes have never been reported in

Lake Ontario (Munawar and Munawar, 1982). This is likely due to the fact that all past Lake

Ontario phycological studies that have been based on light-microscopy, and it is possible that

prasinophytes are too small (< 3 μm) to be readily differentiated from other picoplankton even

at the highest magnification (Short et al., 2011a). Altogether, studies of algal virus polB

diversity suggest that, through molecular surveys, there are great opportunities for

advancements in Lake Ontario and Bay of Quinte phycology and virology.

This study provided the first evidence that PCR methods targeting MCP genes from

viruses of eukaryotic algae can be used to examine freshwater virus communities (Figure 3).

Most of the environmental MCP sequences obtained during this study were most closely related

to two types of mimivirus-like viruses as defined by Monier et al. (2008a) and Monier et al.

(2008b); one group of sequences were closely related mimivirus-like viruses that infect the

prymnesiophytes Phaeocystis pouchetti and Chrysochromulina ericinia, while others were

related to mimivirus-like viruses that infect the prasinophyte Pyramimonas orientalis (Figure 3).

Another set of shorter (i.e., ~400 bp) MCP sequences from the Bay of Quinte (399-M4.18, 399-

M4.1, and 399-M4.2) were most closely related to viruses belonging to the genus Prasinovirus

(Figure 3). While the presence of introns has been shown in the polB gene of some

Phycodnaviruses (Nagasaki et al., 2005), none have been identified for the MCP gene.

According to Larsen et al. (2008), the different amplicon sizes from various MCP genes are

likely due to structural differences in the proteins between different viruses, particularly

mimivirus-like viruses and prasinoviruses.

Although hosts for Bay of Quinte prasinoviruses have not yet been identified, there are

many potential hosts for mimivirus-like viruses of prymnesiophytes. For example, according to

Page 40: Diversity and Dynamics of Algal Viruses in the Bay of Quinte · Diversity and Dynamics of Algal Viruses in the Bay of Quinte Robin Marie Rozon Master of Science ... and Heather Niblock

29

the clustering of MCP sequences, clone 356-M5.14 (qPCR target Mimivirus-Prym 356) is likely

from a virus that infects a Chrysochromulina ericina-like host (Figure 3). This is intriguing

since C. ericina is a marine phytoplankton species. However, since related freshwater species

such as C. parva are known in Lake Ontario and the Bay of Quinte and have been observed

throughout the year (Munawar and Munawar, 1982), it is reasonable to speculate that C. parva

is the host of this virus. On the other hand, it is much more difficult to speculate about the

identity of the host of other types of viruses. The Bay of Quinte sequences 252-M5.5 and 356-

M5.3 (qPCR target Mimivirus-Pras 356) cluster with sequences from Pyramimonas orientalis

virus, a virus that infects a marine prasinophyte. Again, the argument can be made that these

sequences originate from viruses of a freshwater prasinophyte that has not yet been observed in

Lake Ontario or elsewhere. Ultimately, the identity of these and other Bay of Quinte mimivirus-

like viruses will remain uncertain until they are cultivated. Unfortunately this will first require

that the phytoplankton hosts themselves are isolated and grown in the laboratory and this in

itself is a considerable challenge.

By utilising both AVS and MCP primer sets that are specific for a variety of

Phycodnaviruses, these complementary methods have succeeded in providing a broader, more

complete picture of the viral diversity within the Bay of Quinte. Either technique alone would

have led to an incomplete view of algal virus diversity, but used together they mitigate their

apparent biases. Granted that some Phycodnaviridae taxa were not represented in the clone

libraries generated for this study, but given that relatively few samples were amplified, cloned

and sequenced, and only a superficial sequencing effort was conducted for this study, it appears

that the complementary PCR methods used here were an improvement with respect to capturing

a broader range of Phycodnavirus diversity compared to previous studies based on single primer

sets (Short and Short, 2008; Clasen and Suttle, 2009).

To examine an additional dimension of diversity, a clone library for a specific cyanophage

was created. Using primers that target a sheath protein gene sequence of a Microcystis phage

isolated from Japanese freshwater environments (Yoshida et al., 2006), two unique sheath

sequences were found in the Bay of Quinte. While genetically similar algal viruses have been

detected in distantly separated oceans (Short and Suttle, 2002; Bellec et al., 2010), and

freshwater environments (Short and Short, 2008; Roux et al., 2012), the observation of these

phage sequences in the Bay of Quinte was fascinating nonetheless. The Bay of Quinte sequence

Page 41: Diversity and Dynamics of Algal Viruses in the Bay of Quinte · Diversity and Dynamics of Algal Viruses in the Bay of Quinte Robin Marie Rozon Master of Science ... and Heather Niblock

30

Sheath 253 was more closely related to the only known target for these PCR primers, the sheath

protein gene from the Japanese phage Ma-LMM01, than to the other Bay of Quinte sequence,

Sheath 282. Multiple Sheath sequences in the Bay of Quinte are possible if there are multiple

strains of Microcystis phage for a single host stain, or multiple phage strains for different strains

of M. aeruginosa; the Ma-LMM01 sequence and Sheath 253 could be derived from one phage

strain while Sheath 282 is from a different strain, or it is even possible that all three sequences

represent different phage strains. Unlike the Phycodnaviridae which are extremely host specific,

some freshwater cyanophage isolates have been shown to infect a range of bloom-forming hosts,

including Microcystis, and a range of Anabaena and Planktothrix species (Deng and Hayes,

2008). However, Ma-LMM01 is exceptionally host specific with respect to species strains as it

infects only 1 of 11 screened strains of M. aeruginosa, and did not infect any other blue-green

algae species screened (Yoshida et al., 2006). Therefore, it is likely that the Bay of Quinte is

home to an unidentified strain (or strains) of M. aeruginosa that is susceptible to Ma-LMM01,

but until M. aeruginosa in the Bay of Quinte are isolated and sequenced, multiple strains are

impossible to differentiate.

The sequences obtained from each of the clone libraries generated in this study add

credence to the statement that algal viruses are ubiquitous (Breitbart and Rohwer, 2005) and that

the Bay of Quinte is no exception. It has been remarked that “variation within populations is

often greater than variation between populations” (Wilson et al., 2009), which is an important

argument when comparing diversity between locations; it is common to find more genetic

variation within populations than between them. This could be interpreted for differences

between stations within the Bay of Quinte or between sites within Lake Ontario, such as the Bay

of Quinte and Port Credit. A brief comparison of the DNA polymerase gene phylogeny from the

Bay of Quinte with the polB phylogeny from Short and Short (2009) showed a similar clustering

of environmental samples within the genus Prasinovirus, and the Bay of Quinte and Lake

Ontario sequences often cluster together. Deeper sequencing efforts and an expanded study will

be needed to determine if Bay of Quinte viruses are distinguishable from Lake Ontario viruses,

or if either location is home to unique endogenous viruses. The discovery of such algal virus

diversity in the Bay of Quinte has reinforced the effectiveness of complementary PCR methods

in novel environments and has benefited continued diversity studies in freshwater environments.

Page 42: Diversity and Dynamics of Algal Viruses in the Bay of Quinte · Diversity and Dynamics of Algal Viruses in the Bay of Quinte Robin Marie Rozon Master of Science ... and Heather Niblock

31

Quantitative Analysis of Viral Dynamics

All qPCR assays were performed within acceptable amplification efficiencies after having

been subjected to specificity testing. Each of the qPCR assays had percent nucleotide identities

with non-target sequences less than 94 %, most of which ranged between 69 and 83 % (Table

7), and yet all of the non-target fragments used for each qPCR assay verification were the

environmental sequence most closely related to the target. It is interesting to note that as the

only two OTUs available, the sheath protein gene sequences were each other’s closest non-

target sequence, yet the two assays were remarkably different with respect to the differences in

Ct between target and non-target molecules; with Sheath 253 having the smallest and Sheath

282 having the largest of all ten quantitative assays. This might indicate that the nucleotide

differences (8 %) between the sequences are localized to particular areas of the gene, however

that was not observed; crude inspection of the sequence alignment suggests that the single

nucleotide polymorphisms (SNPs) were distributed throughout the entire ~950 bp fragment.

Since the detection limits for non-target sequences were minimally three orders of magnitude

higher than for target sequences (e.g., for Sheath 253), it was unlikely that any non-target genes

contributed to the estimated target abundance, as argued by Short and Short (2009). It is

important to note that the qPCR assays developed and used in this study were used to quantify

gene abundances, virus abundances were merely inferred. As polB is a single-copy gene in

known viral genomes (Delaroque et al., 2001), MCP is a single-copy gene in PBCV-1 (Graves

and Meints, 1992), and the sheath protein is a single-copy gene in Ma-LMM01 (Yoshida et al.,

2006), it is reasonable to infer virus abundance from gene copy abundances. It is also important

to note that quantitative numeration of virus gene copies does not give an indication of

infectious potential nor does it infer the infectious proportion of a virus population (Short,

2012).

Virus Dynamics across Time and Space

By examining the dynamics of several different virus taxa observed in the Bay of Quinte,

the first and most obvious conclusion is that each virus’ seasonality differs from the others.

Even within a genus (i.e., Prasinovirus) the dynamics of different viruses differed in the Bay

reinforcing the results of previous studies of a just a few different Phycodnavirus genes in Lake

Ontario (Short and Short, 2009; Short et al., 2011a). As obligate parasites of phytoplankton,

Page 43: Diversity and Dynamics of Algal Viruses in the Bay of Quinte · Diversity and Dynamics of Algal Viruses in the Bay of Quinte Robin Marie Rozon Master of Science ... and Heather Niblock

32

Phycodnaviruses and cyanophages are both dependent on hosts that photosynthesise and

flourish during seasonal periods to which they are best adapted. In most temperate waters,

phytoplankton biomass is highest during the spring and declines during nutrient limited summer

months. However, historical biomass trends in the Bay of Quinte repeatedly show late-summer

total biomass peaks (Nicholls, 2001) and blue-green algae blooms (Minns et al., 2011). While

most of the viruses monitored were at peak abundance during the late summer and early fall, the

abundance of some, like Mimivirus-Prym 399 and Mimivirus-Pras 252, were higher in the

spring and declined into the summer, succeeded by potentially faster growing, more abundant

virus populations, like Sheath 282 (Figure 4). Phytoplankton succession from spring species to

summer species has been observed in the Bay of Quinte (Munawar and Munawar, 1982;

Munawar et al., 2011; Project Quinte, 2012) and patterns of succession are often attributed to

changing environmental conditions alone (Reynolds, 2006), but the coexisitance of hosts and

viruses with different susceptibilities could lead to even more complex patterns of succession

than could be driven by changing environmental conditions alone (Thyrhaug et al., 2003; Short,

2012). As the hosts go through species succession, it is the associated oscillations in virus

populations that are captured through qPCR, yet it is still unknown if viruses are tracking or

driving host populations (Short, 2012).

The dynamics of Chlorovirus and Mimivirus-Prym 356 were unique among the viruses

examined in this study; they remained consistently detectable at high levels throughout the

sampling season and peaked as per historical Bay of Quinte late-summer bloom abundance

trends. This is likely an effect of consistent host availability where the host species remained

sufficiently abundant/dense to permit continued viral infection throughout the year, peaking

during the late-summer, early-fall (Munawar et al., 2011). Many freshwater phytoplankton

species, including Chlamydomonas globosa, Scenedesmus bijuga, Chrysochromulina parva and

Chroococcus dispersus, have been shown to remain detectable in Lake Ontario (Munawar and

Munawar, 1982) and the Bay of Quinte (Nicholls and Carney, 2011; Project Quinte, 2012)

throughout the spring, summer and fall. With regards to algal virus phenology (Figures 4 and 5),

distinct seasonal patterns over the sampling season would be expected as phytoplankton

community composition in the Bay of Quinte is not consistent over the growing season

(Nicholls and Carney, 2011; Project Quinte, 2012). An intriguing observation of Chlorovirus

and Mimivirus-Prym 356 oscillation patterns is the absence of expected boom and bust

Page 44: Diversity and Dynamics of Algal Viruses in the Bay of Quinte · Diversity and Dynamics of Algal Viruses in the Bay of Quinte Robin Marie Rozon Master of Science ... and Heather Niblock

33

oscillations. A typical boom/bust oscillation implies that the host-parasite interactions reach

high levels then crash out. However, these viruses appeared stable throughout the sampling

season in comparison to the other viruses in this study; they never crashed and were consistent

between stations (Figure 4). This could be due to a stable host-virus interaction created by the

coexistence of multiple strains of hosts and viruses that could not be discriminated, or these

algae hosts could be expressing some form of phenotypic plasticity, altering their susceptibility

to viral infection, thereby coexisting with viruses in the environment (Thyrhaug et al., 2003). On

a broader scale, some viral oscillations are not even consistent between stations, notably the

rapid spring-time boom and bust of Mimivirus-Pras 252 at Belleville, or the slow growth and

eventual crash later in the year of Prasinovirus16.20 at Hay Bay. These observed abundance

patterns are far more complex than what might have been speculated based on typical, yet

simplistic view of virus-host ecology.

The seasonal patterns revealed by qPCR showed a much more dynamic and variable virus

community than what was expected based on previous virus abundance studies in Lake Ontario

where seasonal virus trends were relatively similar in peak timing and growing season duration

(Short and Short, 2009; Short et al., 2011a). It would also be reasonable to expect virus group

abundances to oscillate in a similar fashion, particularly the closely related Microcystis phage

sequences, Sheath 253 and Sheath 282, but that was not observed. Quantitative PCR revealed

that the Microcystis phage sequences had different seasonal patterns and spatial distributions.

Different seasonalities were also observed for the prasinoviruses, particularly Prasinovirus49;

the abundance of this virus varied over the 2011 sampling season as well as between sampling

stations. Both mimiviruses that infect prasinophytes and both mimiviruses that infect

prymnesiophytes were different in with respect to the timing and duration of their peak

abundances. Therefore, the observations of this study suggest that it is not possible to predict the

seasonal trends of a virus based trends of related viruses.

A number of previous studies have examined several distinct aspect of algae virus

ecology, from host succession due to viruses (Thyrhaug et al., 2003), to estimates of host

density dependence (e.g., Cottrell and Suttle, 1995), to determining how viruses persist in the

environment (Cottrell and Suttle, 1995; Short and Short, 2009; Thomas et al., 2011). While

some aspects of algal virus ecology may seem predictable (Short 2012), such as the reoccurance

of a late-summer blue-green algae bloom in the Bay of Quinte, individual virus-host systems

Page 45: Diversity and Dynamics of Algal Viruses in the Bay of Quinte · Diversity and Dynamics of Algal Viruses in the Bay of Quinte Robin Marie Rozon Master of Science ... and Heather Niblock

34

may have to be considered unique. This viral dynamics study was conducted as a primary

examination of ten species/strains of Phycodnaviruses and cyanophage in the Bay of Quinte

over a relatively short period of 24 weeks during the 2011 sampling season; it was not meant to

be all encompassing, but rather ‘snapshots’ of the progression of particular virus communities

over a single field season. By sampling biweekly, smaller virus abundances fluctuations may

have been missed, but even by looking at these ten different viruses with this sampling regime,

the uniqueness of the dynamics observed at the same locations over the same period is

astounding.

By comparing the Bay of Quinte results using the four previously designed quantitative

assays (Table 5) to identical assays previously performed on water samples from Port Credit,

Lake Ontario (Short and Short, 2009; Short et al., 2011a), intriguing differences and similarities

become apparent. Comparing the results from the Prasinovirus49 assay conducted on Port

Credit samples from summer 2008 and the Bay of Quinte samples from summer 2011 revealed a

similar timing of peak abundance (mid-October), but also demonstrated that Prasinovirus49 was

twice as abundant in Port Credit compared to the Bay of Quinte. The Chlorovirus gene

monitored in Port Credit over a two-year period peaked in July 2008 and in June 2009, while the

same gene fragment peaked at approximately 6.5 times higher abundance, but later in the year

(August) in the Bay of Quinte in 2011. Prasinovirus16.20 and Prasinovirus68 peaked in the Bay

of Quinte only a few weeks earlier than they did previously in Port Credit, but for these gene

fragments the abundances were 20 and 50 times larger, respectfully, in Port Credit than the Bay

of Quinte. It is interesting to note that abundances of Prasinovirus16.20 and Prasinovirus68 in

the Bay of Quinte remained low at Hay Bay, nearly below detection at Napanee, and were not

detected at Belleville. The apparent decrease in abundance of these prasinoviruses with

increasing distance into the Bay could be due to host exclusion through competition with algae

species better adapted to the Upper and Middle Bays of the Bay of Quinte. Since these virus

genes were originally detected in Lake Ontario (Short et al., 2011a), their hosts could be better

adapted to open Lake Ontario water. With riverine flushing in the Upper Bay (Nicholls, 1999),

the dominant exchange flows between the Bay of Quinte and Lake Ontario occur in the Lower

and Middle Bays (Freeman and Prinsenberg, 1986), which corresponds to where these

prasinoviruses were most abundant. Past research has found distinct phytoplankton communities

at each of the historical Bay of Quinte monitoring stations (Nicholls et al., 2002; Munawar et al.,

Page 46: Diversity and Dynamics of Algal Viruses in the Bay of Quinte · Diversity and Dynamics of Algal Viruses in the Bay of Quinte Robin Marie Rozon Master of Science ... and Heather Niblock

35

2011), indicating patchy phytoplankton distribution in the Bay. Efforts to document station

specific Bay of Quinte phytoplankton communities are ongoing (Project Quinte, 2012).

The 2011 sampling year in the Bay of Quinte could be considered an atypical season for a

number of reasons. Historically, the Upper Bay has undergone a late-summer single species

algal bloom (Committee, 1990), but in 2011 late-summer algal growth was co-dominated by

several different species (Munawar, unpublished data). In addition, the highest chlorophyll a

and total phosphorus levels historically observed at Belleville (Nicholls, 1999; Nicholls and

Carney, 2011; Project Quinte, 2012) were highest in Hay Bay in 2011. Therefore it is highly

likely that results from future viral dynamics studies in the Bay of Quinte will differ from the

observations of 2011. Nevertheless, it is very unlikely that the observed differences in

seasonality patterns during 2011 (Figures 4 and 5) could be attributed to a unique growing

season. Therefore, future studies would likely arrive at the same conclusion; each virus is unique

and displays different dynamics over time and space. Of particular interest would be the impact

of a historically typical Bay of Quinte season on Prasinovirus68, Prasinovirus16.20 and Sheath

282 trends as all three viruses were the most abundant at Hay Bay over the majority of the

sampling period; their hosts may be particularly sensitive to total phosphorus levels. As the

effects of climate change are becoming more evident on a global scale, it is possible that this

‘atypical’ sampling season could represent a new norm for the Bay of Quinte. Moreover, the

development of additional prasinovirus quantitative assays specific to the Bay of Quinte and

additional freshwater sampling locations within Lake Ontario would allow for the comparison of

genetically related, geographically separated prasinovirus species.

Statistical Analysis of Virus Dynamics

The results of the non-parametric Friedman analysis show that the abundance of different

algal viruses within a given station are statistically very different from each other (p << 0.001)

even though some of the viruses are within the same family (Prasinoviridae) or represent related

strains of a virus (e.g., Microcystis phage). This supports the statement that abundance and trend

generalizations are simply not possible (Short, 2012). Because of the large number of viruses

examined at each station, post-hoc analysis was not informative; the statistical power of non-

parametric tests is already less than that of parametric tests and in order to limit possible type I

errors the Bonferroni correction was applied further increasing the likelihood of a type II error

Page 47: Diversity and Dynamics of Algal Viruses in the Bay of Quinte · Diversity and Dynamics of Algal Viruses in the Bay of Quinte Robin Marie Rozon Master of Science ... and Heather Niblock

36

(failing to observe a difference when one actually exists), an effect of the conservative nature of

the Bonferroni correction. Nevertheless, boxplots of virus abundance at each biweekly station

highlight the variability in the data (Figure 6). Overall, Belleville appears the most variable of

all of the sample sites with respect virus gene abundance, and virus genes such as Microcystis

phage sheath proteins, Prasinovirus49 polB, and all of the mimivirus-like MCP genes were

highly variable across locations with distinct median values and ranges. On the other hand, some

types of viruses at Hay Bay were very similar with respect to their abundance and variability.

Two prasinovirus genes had similar median values and ranges, and while the sheath proteins

genes were similar in median, their ranges were distinct. Of the closely related viruses at

Napanee, only the sheaths were similar with respect to median abundance, but as with Hay Bay,

their range of abundance was dramatically different. Based on the result of the Friedman

analysis, visual inspection of the seasonality patterns of different virus genes at the different

stations (Figure 5), and the obvious variability of virus gene abundances (Figure 6), it is clear

that there is considerable ecological variability within Bay of Quinte algal virus communities,

regardless of sampling location or virus taxonomic classification. This is very likely due to the

variable ecology of the different hosts of these viruses and the different ecological drivers that

influence their growth.

By using the Friedman analysis to compare the abundances of individual viruses between

stations (e.g., comparing the abundance of Chlorovirus at Belleville to Hay Bay and to

Napanee), only four of the ten viruses were statistically significantly different between biweekly

sampling stations (Prasinovirus49, Prasinovirus68, Prasinovirus16.20, and Sheath 282). A

Wilcoxon Signed-Rank post-hoc analysis with a Bonferroni correction applied demonstrated

that of those four, three had significant differences between specific stations (Prasinovirus68,

Prasinovirus16.20 and Sheath 282). By examining these particular viral seasonal trends (Figure

4) and viral variability across stations (Figure 6), the results of the statistical tests are fairly

obvious. For both Prasinovirus68 and Prasinovirus16.20, abundance at Hay Bay was

significantly different than the other two stations, and for Sheath 282, Belleville was

significantly different from Hay Bay, but not Napanee. Since Hay Bay was the station most

often statistically different than other stations based on variances of virus gene abundances, it

appears to be an atypical Bay of Quinte station compared to Belleville and Napanee. As

mentioned above, the 2011 sampling season was not a ‘typical’ Bay of Quinte summer; it would

Page 48: Diversity and Dynamics of Algal Viruses in the Bay of Quinte · Diversity and Dynamics of Algal Viruses in the Bay of Quinte Robin Marie Rozon Master of Science ... and Heather Niblock

37

appear that the ecological drivers unique to Hay Bay had an impact on viral dynamics and

abundance. Therefore is it possible that future algal virus studies in the Bay of Quinte would

produce different results.

The data in Figures 4, 5 and 6 demonstrate the very large range of virus gene abundances

that were observed during this study of different viruses, from below detection limits to over

256,000 gene copies/mL. Previous quantitative studies of freshwater algal viruses have also

found a wide range of gene copies. For example, Short et al. (2011a) observed prasinovirus

abundance from near detection limits to 11,000 gene copies/mL, and studies from Lake Erie in

2009 found up to 3.1 x 106 gene copies/mL of a single cyanomyovirus gene (Matteson et al.,

2011). These past studies support the validity of this work’s abundance values; it reasonable that

a single strain of Microcystis phage was so highly abundant in certain samples (> 105 gene

copies/mL) while other viruses peaked at a much lower abundance (e.g., Prasinovirus68 peaked

at < 20 gene copies/mL). While these statistical tests have demonstrated that algal viruses in the

Bay of Quinte are significantly different from each other across time, and a few are significantly

different between stations, host abundance dynamics were only part of the analysis indirectly. A

possible avenue of future research into Bay of Quinte algal ecology could be the development of

a multivariate analysis based on environmental parameters and phytoplankton population trends

particularly bloom events. Using this framework, future work can determine which

environmental parameters best predict host abundances, and therefore virus abundances. A

particular goal of algal ecology is the prediction and forecasting of harmful algal events, which

could prevent illnesses to livestock and humans alike. Furthermore, predicting an algal bloom

would enable algal virus ecologists to track a lytic event through host growth, bloom and decay.

Virus Abundance and Host Biomass

To determine if phytoplankton biomass predicted algal virus abundance in the Bay of

Quinte, a regression analysis was conducted to compare the sum of the abundance of all ten

virus genes, a proxy for virus abundance, to chlorophyll a concentration, a proxy for host

biomass (Figure 7). The initial clone libraries of this study captured the most abundant and most

common viral sequences in the environment from specific days, at specific locations, and the

quantitative abundance values reflect this (Figure 5). While this ‘sum of viruses’ does not reflect

the abundance of all algal viruses present in the Bay of Quinte, it does reflect the most abundant

Page 49: Diversity and Dynamics of Algal Viruses in the Bay of Quinte · Diversity and Dynamics of Algal Viruses in the Bay of Quinte Robin Marie Rozon Master of Science ... and Heather Niblock

38

viruses from Belleville on June 7th

, August 16th

and October 12th

and Hay Bay on June 7th

and

July 19th

. Of the ten virus sequences used in this study, Chlorovirus and Mimivirus-Prym 356

were consistent, major contributors to the sum of virus abundances at Belleville (together

averaging 75 % of the sum of viruses over the sampling season), but less so at Hay Bay (51 %)

and Napanee (70 %), where Sheath 282 was another major contributor. Sheath 282 accounted

for, on average, 29 % and 16 % of the sum of viruses at Hay Bay and Napanee respectfully, but

only averaged 0.05 % of the sum of viruses at Belleville. Altogether, these three viruses

accounted for over 75 % of the sum of viruses for each station respectively; therefore it was

these three viruses that were driving relationship patterns with host abundances (Figure 7),

which is interesting as chlorophyll a concentration is the measure of all hosts, not of select

species.

Clearly, this study did not estimate total virus abundances. Assuming the total virus counts

(~2.09 x107 virus-like particles) from a single sample from the Bay of Quinte made in 2003 by

Gouvêa (2006) have not changed and represent year round average abundances this study would

have examined at most 1.3 % of the total virus community (Hay Bay, July 19th

, 2011). For any

given sample, the average percentage of the total virus community (by averaging the sum of all

viruses from the biweekly stations) was only 0.09 %. Put into perspective, a quantitative study

in Lake Erie described up to 4.6 % of the total virus community using the target gene g20

present in Cyanomyoviridae (Matteson et al., 2011). Total virus counts from freshwater

(Wilhelm and Smith, 2000; Gouvêa et al., 2006; Wilhelm et al., 2006; Matteson et al., 2011) and

marine ecosystems (Suttle, 2005) report roughly 107 viruses/mL, indicating that a large

proportion of viruses and virus-like particles are not included in quantification studies. Despite

the large proportion of virus-like particles in the Bay of Quinte not accounted for, virus

abundances and chlorophyll a show strong positive statistically significant relationships at

Belleville and Hay Bay, but not at Napanee (Figure 7). This indicates that the algal viruses

studied track strongly with their hosts and that even a small percentage of the total virus

community is enough to see this trend. This supports Clasen (2008) who also found significant

positive relationships between chlorophyll a and virus abundance in freshwater. It is understood

that the sum of viruses used in Figure 7 is not total virus community, but it is possible that the

chlorophyll a values are not a fully accurate representation of actual phytoplankton biomass and

either could be the cause of the disassociation between hosts and viruses seen at Napanee.

Page 50: Diversity and Dynamics of Algal Viruses in the Bay of Quinte · Diversity and Dynamics of Algal Viruses in the Bay of Quinte Robin Marie Rozon Master of Science ... and Heather Niblock

39

By examining the correlation between the sum of virus abundance to chlorophyll a

concentration (Figure 7), the relationship between host and viruses in the Bay of Quinte

supports the current notion that these algal viruses are dependent on their hosts. Almost all

Phycodnaviruses display extreme host specificity; therefore, it is very likely that the ten

different viruses observed in this study each have unique algal/cyanobacteria strains or sub-

stains as their hosts (Short, 2012). However, these hosts could also be susceptible to other

viruses present in the Bay of Quinte. Viral regulated host succession remains largely unresolved

(Thyrhaug et al., 2003), and in situ “killing the winner” studies are limited (Winter et al., 2010).

The few studies that have examined the “killing the winner” viral succession hypothesis

(Thingstad, 2000) in the environment have not provided convincing evidence that viruses

regulate host community composition (Schwalbach et al., 2004) or cause bloom termination

(Schroeder et al., 2003).

This study confirms the relationship between host and virus abundances, and while it

shows that host biomass can predict algal virus abundance it does not provide evidence that

viruses regulated host bloom dynamics. Establishing viruses as causal agents in bloom

termination would require quantification of host and virus abundances over the course of an

algal bloom’s growth and decline. Quantitative studies of both host and viral isolate are

complicated and comprehensive, limited in the literature and often exclusive to a certain

location, so researchers settle for site specific host proxies and total virus counts. Future studies

in the Bay of Quinte could include total virus counts (Noble and Fuhrman, 1998), more

intensive sampling (increased sample number and field season duration) and interpreting the

impact of different year to year environmental drivers on the virus-host system. This could be

done by culturing host algae and isolating viruses, then tracking each through quantitative

techniques to provide a much better understanding of the interaction between viruses, hosts and

environmental conditions.

Clustering by Taxonomic Classification

Distance measures were created through pair-wise Pearson Correlation and Bray-Curtis

dissimilarity comparisons of virus abundance from each virus gene studied to every other virus

gene for all stations, as well as for each station individually to resolve if related virus

populations cluster together based on virus abundances (Figure 8). In an attempt to converge on

Page 51: Diversity and Dynamics of Algal Viruses in the Bay of Quinte · Diversity and Dynamics of Algal Viruses in the Bay of Quinte Robin Marie Rozon Master of Science ... and Heather Niblock

40

a single consensus cluster analysis dendrogram, multiple clustering techniques were used

(Figure 9). While a consensus regarding absolute clustering was not achieved, almost every one

of these analytical techniques produced the same result: Taxonomic groups do not cluster

together based on virus abundance. The only exception is the clustering of Sheath 282 and

Sheath 253 observed in the Belleville dendrograms, which could be due to Prasinovirus16.20

and Prasinovirus68 being excluded as these viruses were consistently below detection limits at

this station, and therefore it is not given much credence. It would appear that each of the viruses

studied have unique seasonal dynamics which greatly limits the potential for overarching

ecological generalizations to be derived solely from virus taxonomic classification. As new

viruses continue to be isolated, more viral taxa are being recognised each year (almost 200

viruses between 2009 and 2011; King et al., 2012). Current taxonomic classification within the

Phycodnaviridae is based partially on the taxonomic affiliation of the hosts (Van Etten et al.,

2002) and the presence of ‘core-genes’ (i.e., polB, MCP) since it can be difficult to infer genetic

relationship due to horizontal gene transfer between viruses. However, based on current

information, the relatedness of many of the established Phycodnavirus genera is relatively

robust. With regards to this study, although the genes targeted for quantitative analysis were

from the same genera, respectfully (or species in the case of the sheath protein gene), each virus

infects a different host. For instance, the cultivated Chrysochromulina ericina virus is closest to

Mimivirus-Prym 356, so it is possible that freshwater species of Chrysochromulina (e.g., C.

parva) is the host for this virus. However it is probably not the host of the other mimivirus-like

virus of prymnesiophytes, Mimivirus-Prym 399, since it clusters separately among similar Bay

of Quinte environmental clones. As with the differences in Phycodnavirus abundance and

seasonal patterns observed by Short et al. (2011a), this study supports the speculation that even

genetically related viruses can infect hosts which have different seasonality themselves. Using

the prasinoviruses from this study as one example, the very unique dynamics between related

virus types will prevent these viruses from being binned together and treated as a single

population in ecosystem models. Based on this conclusion, it is apparent that virus dynamics

cannot be predicted based solely on virus type. That is to say that viral dynamics determined by

quantitative techniques such as these can only be applied to the single virus type containing the

target gene fragment of interest.

Page 52: Diversity and Dynamics of Algal Viruses in the Bay of Quinte · Diversity and Dynamics of Algal Viruses in the Bay of Quinte Robin Marie Rozon Master of Science ... and Heather Niblock

41

As previously mentioned, the 2011 Bay of Quinte sampling season was atypical with

regard to some environmental traits, which would presumably impact virus-host dynamics and

influence the way in which virus abundances cluster together in the dendrograms. Repeated

cluster analysis in a more historically typical sampling season could produce different results.

There is an interesting question raised by this cluster analysis; how could two phage sequences

originating from a single Microcystis phage strain not show more similar temporal abundance

patterns and therefore cluster together more often. It can be hypothesized that the observed

different patterns of Microcystis phage seasonality were due to the presence of two different

host strains of M. aeruginosa in the Bay, but thorough examination of the dynamics of different

Microcystis species, or even strains of M. aeruginosa would be needed to confirm or refute this

hypothesis. Currently, only bulk data on M. aeruginosa is available so considerations of intra-

species population dynamics are highly speculative.

The conclusion of this cluster analysis, that viral dynamic predictions cannot be based on

virus type, represents an important and fundamental hurdle for continued aquatic viral studies.

With regard to viral ecology, it may not be possible for “observed patterns [to] be scaled up

through statistical analysis and modeling to describe the structure and function of marine

ecosystems” (Cullen et al., 2007). To properly understand virus ecology, the dynamics of each

virus of interest may need to be examined individually, and not extrapolated from seasonality

data from related viruses.

Clustering by Spatial Distribution

Using virus abundances from each virus gene for each sample date at each biweekly

stations, as well as virus abundances from each virus gene for each station sampled during the

September spatial survey, distance measures were created through pair-wise Pearson Correlation

and Bray-Curtis dissimilarity comparisons to resolve if sampling events (Figure 10) or spatially

similar stations (Figure 11) cluster together based on abundances. The different cluster analysis

dendrograms revealed the same pattern regardless of clustering technique used; the biweekly

stations clustered more often by sample date than by station, while the spatial stations clustered,

for the most part, by their location in the Bay. The results from the biweekly stations suggest

that for viruses in the Bay of Quinte, patterns of abundance are more similar at certain times of

the year regardless of the sampling location. This makes sense since phytoplankton biomass (as

Page 53: Diversity and Dynamics of Algal Viruses in the Bay of Quinte · Diversity and Dynamics of Algal Viruses in the Bay of Quinte Robin Marie Rozon Master of Science ... and Heather Niblock

42

estimated through chlorophyll a concentration) was low during the early part of the sampling

season and algal virus abundances were also low. Similarly, samples from the later sampling

dates also tended to cluster together. Thus, virus abundances tended to cluster by date rather

than location (Figure 10). Nonetheless, there remains several nodes where sample location,

particularly from Belleville station, cluster. This is likely due to the uniqueness of

Prasinovirus16.20, Prasinovirus68 and Sheath 282 abundances at Belleville; Belleville is the

only location where several viruses were below detection or present at very low abundance.

With HB5 as the only exception, the virus abundance across spatial stations clustered into

groups consistent with the Upper and Middle Bay delineations (Figure 1; Robinson, 1986).

Based on the geographical location of HB5, far into Hay Bay, it may have been subjected to

similar environmental trends as the Upper Bay stations. Particularly with regards to station

depth, HB5 is shallower (4.7 m) than other Middle Bay stations (averaging 8.2 m), and is closer

in depth to the average depth of the Upper Bay stations (4.9 m; Table 1). Also, as mentioned, the

highest levels of total phosphorus in the Bay were found in Hay Bay, however, total phosphorus

levels at HB5 were more comparable to Upper Bay levels than to other Middle Bay stations

(Quinte, unpublished data). Despite increased exceptions (HB5, B2, and N), three divisions of

the sample locations are visible in Figure 11; the Belleville area group (B, BQ7, BQ6, and B2),

the Hay Bay area group (HB, HB4, HB5, and P) and the Napanee area group (N, BQ8, NR1,

and BQ9). While this could be a result of the spatial distribution of the virus communities in the

Bay of Quinte, owing to sampling and laboratory processing constraints, the spatial sampling

itself was done over three days which coincides exactly with the three aforementioned divisions

of the spatial stations. Therefore, the speculation that virus abundances in the Bay of Quinte are

more similar in time than in space may hold true for the spatial survey cluster analysis as well,

albeit over a very short time scale. Repeated sampling with a mix of each area on different days,

and with samples from all three bays (Upper, Middle and Lower) would help resolve if virus

abundance in the Bay of Quinte samples truly cluster with respect to location.

Summary

Broadly speaking and based on the patterns of virus abundance revealed in the

aforementioned analyses, it is clear that virus communities are heterogeneously distributed

across the Bay and over the phytoplankton growing season. Previous marine research has

Page 54: Diversity and Dynamics of Algal Viruses in the Bay of Quinte · Diversity and Dynamics of Algal Viruses in the Bay of Quinte Robin Marie Rozon Master of Science ... and Heather Niblock

43

demonstrated the microscale patchiness of virus abundance at spatial resolutions of 1 and 5 cm

with ‘hotspot’ concentrations varying up to 17-fold (Seymour et al., 2006) as well as significant

changes in virus abundances with 20-minute temporal resolutions (Wommack and Colwell,

2000). When virus abundances can differ between the tips of a multi-channel pipettor (Steven

Wilhelm, personal communication, October 5th

, 2012) and host distributions can be equally

heterogeneous (Long and Azam, 2001), differences across the Bay may seem inconsequential.

However, given that significant virus heterogeneity can occur over centimeters, a better

understanding of the spatial and temporal microscale processes will elucidate a relevant scale

for future sampling (Breitbart, 2012). Obviously, attempting to sample everything, everywhere,

at all times is not feasible for any environmental study, but rather taking small steps towards

better understanding the ecological role of algal viruses is a worthy goal for research on aquatic

ecosystems. This seminal Bay of Quinte algal virus study supports past observations suggesting

that viruses are omnipresent in freshwater ecosystems, and that genetically similar algal viruses

(e.g., based on polB, MCP and sheath protein genes) can be detected in oceans and freshwater

lakes all over the world. This study brings into sharp relief the apparent complexities in the

ecology of algal viruses, and the potential intricacies of Bay of Quinte phytoplankton ecology.

Predicting and modeling phytoplankton dynamics in the Bay and determining the factors that

are contributing to the modern emergence of harmful algal blooms is not trivial. Efforts such as

this to dive below the surface of phytoplankton ecology are returning more questions than

answers and are finding these waters deeper than were ever imagined.

Page 55: Diversity and Dynamics of Algal Viruses in the Bay of Quinte · Diversity and Dynamics of Algal Viruses in the Bay of Quinte Robin Marie Rozon Master of Science ... and Heather Niblock

44

Future Directions

With the confirmation of the presence of algal viruses in the Bay of Quinte, this study

demonstrated the surprisingly unique seasonality of individual virus types, suggesting that

predictions of virus seasonality cannot be based on virus identity, or taxonomic affiliation.

Future algal studies in the Bay of Quinte should compare results from a more ‘typical’ sampling

season to these results, as well as to other freshwater viruses from the same and different water

basins. Increasing the number of sequences in a clone library would provide a more in depth

examination of the algal virus diversity in the Bay of Quinte. Quantitative analysis of other

viruses as well as confirmed host-virus pairs from the Bay of Quinte would help better

understand virus-host interactions during periods of algal blooms and long term ‘seed-bank’

persistence (Short and Short, 2009). Continued modeling efforts to predict and forecast harmful

algal blooms (in the Bay of Quinte and around the world) must take into account the effect

viruses have on their host species by looking at virus-algae interactions. Questions related to

how viruses affect the ecology of their host algae, or which environmental parameters are the

best predictor of algae population growth need to be addressed. Future work with multivariate

statistics could make use of the abundant environmental and phytoplankton data from the past

40 years of Project Quinte and compare observations to hypothesize which environmental

parameters best predict hosts population oscillations, and eventually virus population growth

and decline. Ecology of the microscale is on the cusp of being properly understood, and with

conclusions like the ones from this study limiting the scope of overarching statements related to

viral dynamics, proper modeling of algal virus communities will be a laborious yet rewarding

campaign.

Page 56: Diversity and Dynamics of Algal Viruses in the Bay of Quinte · Diversity and Dynamics of Algal Viruses in the Bay of Quinte Robin Marie Rozon Master of Science ... and Heather Niblock

45

Literature Cited

Bellec, L., Grimsley, N. & Desdevises, Y. 2010. Isolation of Prasinoviruses of the Green

Unicellular Algae Ostreococcus spp. on a Worldwide Geographical Scale. Applied and

Environmental Microbiology, 76 (1): 96-101.

Bergh, O., Borcheim, K. Y., Bratbak, G. & Heldal, M. 1989. High abundance of viruses found

in aquatic environments. Nature, 340: 467-468.

Bratbak, G., Egge, J. K. & Heldal, M. 1993. Viral mortality of the marina alga Emiliania huxleyi

(Haptophyceae) and termination of algal blooms. Marine Ecology Progress Series, 93: 39-48.

Breitbart, M. 2012. Marine Viruses: Truth or Dare. Annual Review of Marine Science, 4 (1):

425-448.

Breitbart, M. & Rohwer, F. 2005. Here a virus, there a virus, everywhere the same virus? Trends

in microbiology, 13 (6): 278-284.

Brown, R. M. 1972. Algal Viruses. Advances in Virus Research, 17: 243-277.

Chen, F. & Suttle, C. A. 1995. Amplification of DNA polymerase gene fragments from viruses

infecting microalgae. Applied and Environmental Microbiology, 61 (4): 1274-1278.

Chen, F. & Suttle, C. A. 1996. Evolutionary relationships among large double-stranded DNA

viruses that infect microalgae and other organisms as inferred from DNA polymerase genes.

Virology, 219: 170-178.

Chen, F., Suttle, C. A. and Short, S. M. 1996. Genetic diversity in marine algal virus

communities as revealed by sequence analysis of DNA polymerase genes. Applied and

Environmental Microbiology 62: 2869-2874.

Clasen, J. L., Brigden, S. M., Payet, J. P. & Suttle, C. A. 2008. Evidence that viral abundance

across oceans and lakes is driven by different biological factors. Freshwater Biology, 53 (6):

1090-1100.

Clasen, J. L. & Suttle, C. A. 2009. Identification of freshwater Phycodnaviridae and their

potential phytoplankton hosts, using DNA pol sequence fragments and a genetic-distance

analysis. Applied and Environmental Microbiology, 75 (4): 991-997.

Committee, R. C. 1990. Stage 1: Environmental Setting + Problem Definition. The Bay of

Quinte Remedial Action Plan. Environment Canada. 302p.

Committee, R. C. 1993. Stage 2: Time to Act. The Bay of Quinte Remedial Action Plan.

Environment Canada. 271p.

Cottrell, M. T. & Suttle, C. A. 1995. Dynamics of a lytic virus infecting the photosynthetic

marine picoflagellate Micromonas-pusilla. Limnology and Oceanography, 40: 730-739.

Cullen, J. J., Doolittle, W. F., Levin, S. A. & Li, W. K. W. 2007. Patterns and Predictions in

Microbial Oceanography. Oceanography, 20 (2): 34-46.

Page 57: Diversity and Dynamics of Algal Viruses in the Bay of Quinte · Diversity and Dynamics of Algal Viruses in the Bay of Quinte Robin Marie Rozon Master of Science ... and Heather Niblock

46

Delaroque, N., Müller, D. G., Bothe, G., Pohl, T., Knippers, R. & Boland, W. 2001. The

Complete DNA Sequence of the Ectocarpus siliculosus Virus EsV-1 Genome. Virology, 287

(1): 112-132.

Deng, L. I. & Hayes, P. K. 2008. Evidence for cyanophages active against bloom-forming

freshwater cyanobacteria. Freshwater Biology, 53 (6): 1240-1252.

Derelle, E., Ferraz, C., Escande, M. L., Eychenié, S., Cooke, R., Piganeau, G., Desdevises, Y.,

Bellec, L., Moreau, H. & Grimsley, N. 2008. Life-cycle and genome of OtV5, a large virus of

the pelagic marine unicellular green alga Ostreococcus tauri. PLoS One, 28 (5): e2250

Dorak, M. T. 2006. Real-time PCR, New York, Taylor & Francis. 362p.

Freeman, N. G. & Prinsenberg, S. J. 1986. Exchange Flows in the Adolphus Reach/North

Channel. In: Minns, C. K., Hurley, D. A. & Nicholls, K. H. (eds.) Project Quinte : point-source

phosphorus control and ecosystem response in the Bay of Quinte, Lake Ontario. Ottawa. 277p.

Gouvêa, S. P., Melendez, C., Carberry, M. J., Bullerjahn, G. S., Wilhelm, S. W., Langen, T. A.

& Twiss, M. R. 2006. Assessment of Phosphorus-microbe Interactions in Lake Ontario by

Multiple Techniques. Journal of Great Lakes Research, 32: 455-470.

Graves, M. V. & Meints, R. H. 1992. Characterization of the Major Capsid Protien and Cloning

of its Gene from Algal Virus PBCV-1. Virology, 188: 198-207.

Hall, T. A. 1999. BioEdit: a user-friendly biological sequence alignment editor and analysis

program for Windows 95/98/NT. Nucleic Acids Symposium Series, 41: 95-98.

Johnson, G. M. & Hurley, D. A. 1986. Overview of Project Quinte – 1972-82. In: Minns, C. K.,

Hurley, D. A. & Nicholls, K. H. (eds.) Project Quinte : point-source phosphorus control and

ecosystem response in the Bay of Quinte, Lake Ontario. Ottawa. 277p.

Jones, D. T., Taylor, W. R. & Thornton, J. M. 1992. The rapid generation of mutation data

matrices from protein sequences. Computer Applications in the Biosciences, 8: 275-282.

King, A. M. Q., Adams, M. J., Carstens, E. B. & Lefkowitz, E. J. 2012. Virus Taxonomy: Ninth

Report of the International Committee on Taxonomy of Viruses, London, Elsevier. 1338p.

La Scola, B., Audic, S., Robert, C., Jungang, L., De Lamballerie, X., Drancourt, M., Birtles, R.,

Claverie, J.-M. & Raoult, D. 2003. A Giant Virus in Amoebae. Science, 5615: 2033.

Larsen, J. B., Larsen, A., Bratbak, G. & Sandaa, R. A. 2008. Phylogenetic analysis of members

of the Phycodnaviridae virus family, using amplified fragments of the major capsid protein

gene. Applied and Environmental Microbiology, 74 (10): 3048-3057.

Leisti, K. E. & Doka, S. E. 2008. Water temperature monitoring in the Bay of Quinte, 2006.

Project Quinte Annual Report 2006. Kingston, Ontario: Project Quinte Members. 148p.

Liengme, B.V. 2008. A Guide to Microsoft Excel 2007 for Scientists and Engineers, London,

Academic Press. 326p.

Page 58: Diversity and Dynamics of Algal Viruses in the Bay of Quinte · Diversity and Dynamics of Algal Viruses in the Bay of Quinte Robin Marie Rozon Master of Science ... and Heather Niblock

47

Long, R. A. & Azam, F. 2001. Microscale patchiness of bacterioplankton assemblage richness

in seawater. Aquatic Microbial Ecology, 26: 103-113.

Maranger, R. & Bird, D. F. 1995. Viral abundance in aquatic systems: a comparison between

marine and fresh waters. Marine Ecology Progress Series, 121: 217-226.

Matteson, A. R., Loar, S. N., Bourbonniere, R. A. & Wilhelm, S. W. 2011. Molecular

enumeration of an ecologically important cyanophage in a Laurentian Great Lake. Applied and

Environmental Microbiology, 77 (19): 6772-6779.

Mayer, J. A. & Taylor, F. J. R. 1979. A virus which lyses the marine nanoflagellate Micromonas

pusilla. Nature, 281: 299-301.

McCombie, A. M. 1967. A recent study of the phytoplankton of the Bay of Quinte, 1963–1964.

Procreedings of the 10th Conference on Great Lakes Research, International Association Great

Lakes Research, 37-62.

Minns, C. K., Hurley, D. A. & Nicholls, K. H. 1986. Project Quinte: Point-Source Phosphorus

Control and Ecosystem Response in the Bay of Quinte, Lake Ontario, Ottawa, Ontario, Canada,

Department of Fisheries and Oceans. 277p.

Minns, C. K., Munawar, M., Koops, M. A. & Millard, E. S. 2011. Long-term ecosystem studies

in the Bay of Quinte, Lake Ontario, 1972–2008: A prospectus. Aquatic Ecosystem Health &

Management, 14 (1): 3-8.

Monier, A., J.M. Claverie, and H. Ogata. 2008a. Taxonomic distribution of large DNA viruses

in the sea. Genome Biology, 9:R106.

Monier, A., Larsen, J. B., Sandaa, R-A., Bratbak, G., Claverie, J-M. and Ogata, H. 2008b.

Marine mimivirus relatives are probably large algal viruses. Virology Journal, 5(12).

Mullis, K., Faloona, F., Scharf, S., Saiki, R., Horn, G. & Erlich, H. 1986. Specific Enzymatic

Amplification of DNA in vitro: The Polymerase Chain Reaction. Cold Spring Harbor Symposia

on Quantitative Biology, 51 (1): 263-273.

Munawar, M., Fitzpatrick, M., Niblock, H. & Lorimer, J. 2011. The relative importance of

autotrophic and heterotrophic microbial communities in the planktonic food web of the Bay of

Quinte, Lake Ontario 2000–2007. Aquatic Ecosystem Health & Management, 14 (1): 21-32.

Munawar, M. & Munawar, I. F. 1982. Phycological studies in lakes Ontario, Erie, Huron, and

Superior. Canadian Journal of Botany, 60: 1837-1858.

Nagasaki, K., Shirai, Y., Tomaru, Y., Nishida, K. & Pietrokovski, S. 2005. Algal viruses with

distinct intraspecies host specificities include identical intein elements. Applied and

Environmental Microbiology, 71 (7): 3599-607.

Nicholls, K. H. 1999. Effects of Temperature and Other Factors on Summer Phosphorus in the

Inner Bay of Quinte, Lake Ontario: Implications for Climate Warming. Journal of Great Lakes

Research, 25 (2): 250-262.

Page 59: Diversity and Dynamics of Algal Viruses in the Bay of Quinte · Diversity and Dynamics of Algal Viruses in the Bay of Quinte Robin Marie Rozon Master of Science ... and Heather Niblock

48

Nicholls, K. H. 2001. Assessment of the State of Impairment of Beneficial Uses for the Bay of

Quinte Remedial Action Plan: I. Phytoplankton. Environment Canada and the Department of

Fisheries and Oceans. Sunderland, Ontario. 83p.

Nicholls, K. H. 2010. Total phosphorus, nitrogen, chlorophyll and phytoplankton in the Bay of

Quinte, 2008. Project Quinte Annual Report 2008. Kingston, Ontario: Project Quinte Members.

133p.

Nicholls, K. H. & Carney, E. C. 1979. The taxonomy of Bay of Quinte phytoplankton and the

relative importance of common and rare taxa. Canadian Journal of Botany, 57: 1591-1608.

Nicholls, K. H. & Carney, E. C. 1986. Nitrogen and Phosphorus Limitation to Phytoplankton in

the Bay of Quinte and Implications for Phosphorus Loading Controls. In: Minns, C. K., Hurley,

D. A. & Nicholls, K. H. (eds.) Project Quinte: Point-Source Phosphorus Control and Ecosystem

Response in the Bay of Quinte, Lake Ontairo. Ottawa. 277p.

Nicholls, K. H. & Carney, E. C. 2011. The phytoplankton of the Bay of Quinte, 1972–2008:

point-source phosphorus loading control, dreissenid mussel establishment, and a proposed

community reference. Aquatic Ecosystem Health & Management, 14 (1): 33-43.

Nicholls, K. H., Carney, E. C., Beaver, J. L. & Middleton, D. 1986. Some Effects of Phosphorus

Loading Reductions on Phytoplankton in the Bay of Quinte, Lake Ontario. In: Minns, C. K.,

Hurley, D. A. & Nicholls, K. H. (eds.) Project Quinte: point-source phosphorus control and

ecosystem response in the Bay of Quinte, Lake Ontario. Ottawa. 277p.

Nicholls, K. H. & Heintsch, L. 1986. A Comparison of the Net Phytoplankton in the lower Bay

of Quinte Near Indian Point, 1945 and 1981. In: Minns, C. K., Hurley, D. A. & Nicholls, K. H.

(eds.) Project Quinte: point-source phosphorus control and ecosystem response in the Bay of

Quinte, Lake Ontario. Ottawa. 277p.

Nicholls, K. H., Heintsch, L. & Carney, E. C. 2002. Univariate Step-trend and Multivariate

Assessments of the Apparent Effects of P Loading Reductions and Zebra Mussels on the

Phytoplankton of the Bay of Quinte, Lake Ontario. Journal of Great Lakes Research, 28 (1): 15-

31.

Nicholls, K. H., Heintsch, L. & Carney, E. C. 2004. A multivariate approach for evaluating

progress towards phytoplankton community restoration targets: Examples from eutrophication

and acidification case histories. Aquatic Ecosystem Health & Management, 7 (1): 15-30.

Nicholls, K. H. & Millard, E. S. 2005. Nutrients and Phytoplankton. Project Quinte Annual

Report 2003. Kingston, Ontario: Project Quinte Members. 138p.

Noble, R. T. & Fuhrman, J. A. 1998. Use of SYBR Green I for rapid epifluorescence counts of

marine viruses and bacteria. Aquatic Microbial Ecology, 14: 113-118.

Park, Y., Lee, K., Lee, Y. S., Kim, S. W. & Choi, T. J. 2011. Detection of diverse marine algal

viruses in the South Sea regions of Korea by PCR amplification of the DNA polymerase and

major capsid protein genes. Virus Research, 159 (1): 43-50.

Page 60: Diversity and Dynamics of Algal Viruses in the Bay of Quinte · Diversity and Dynamics of Algal Viruses in the Bay of Quinte Robin Marie Rozon Master of Science ... and Heather Niblock

49

Proctor, L. M. & Fuhrman, J. A. 1990. Viral mortality of marine bacteria and cyanobacteria.

Nature, 343 (6253): 60-62.

Project Quinte. 2012. Project Quinte Annual Report 2010 (#21). Monitoring Report. Kingston,

Ontario, Canada. 149p.

Raoult, D., Audic, S., Robert, C., Abergel, C., Renesto, P., Ogata, H., La Scola, B., Suzan, M. &

Claverie, J.-M. 2004. The 1.2-Megabase Genome Sequence of Mimivirus. Science, 5700: 1344-

1350.

Reynolds, C. 2006. Ecology of Phytoplankton, Cambridge, Cambridge University Press. 550p.

Robinson, G. W. 1986. Water Quality of the Bay of Quinte, Lake Ontario, Before and After

Reductions in Phosphorus Loading. In: Minns, C. K., Hurley, D. A. & Nicholls, K. H. (eds.)

Project Quinte : point-source phosphorus control and ecosystem response in the Bay of Quinte,

Lake Ontario. Ottawa. 277p.

Roux, S., Enault, F., Robin, A., Ravet, V., Personnic, S., Theil, S., Colombet, J., Sime-Ngando,

T. & Debroas, D. 2012. Assessing the Diversity and Specificity of Two Freshwater Viral

Communities through Metagenomics. PLoS ONE, 7 (3): 1-12.

Safferman, R. S. & Morris, M.-E. 1963. Algal Virus: Isolation. Science, 140 (3567): 679-680.

Safferman, R. S. & Morris, M.-E. 1964. Control of algae with viruses. Journal of American

Water Works Association, 56: 1217-1224.

Schroeder, D. C., Oke, J., Hall, M., Malin, G. & Wilson, W. H. 2003. Virus Succession

Observed during an Emiliania huxleyi Bloom. Applied and Environmental Microbiology, 69 (5):

2484-2490.

Schwalbach, M. S., Hewson, I. & Fuhrman, J. A. 2004. Viral effects on bacterial community

composition in marine plankton microcosms. Aquatic Microbial Ecology, 34 (2): 117-127.

Seymour, J. R., Seuront, L., Doubell, M., Waters, R. L. & Mitchell, J. G. 2006. Microscale

patchiness of virioplankton. Journal of the Marine Biological Association of the United

Kingdom, 86: 551-561.

Short, C. M., Rusanova, O. & Short, S. M. 2011a. Quantification of virus genes provides

evidence for seed-bank populations of Phycodnaviruses in Lake Ontario, Canada. The ISME

Journal, 5 (5): 810-21.

Short, S. M. 2012. The ecology of viruses that infect eukaryotic algae. Environmental

Microbiology, 14 (9): 2253-2271.

Short, S. M., Chen, F. & Wilhelm, S. W. 2010. The construction and analysis of marker gene

libraries. In: Wilhelm, S. W., Weinbauer, M. G. & Suttle, C. A. (eds.) Manual of Aquatic Viral

Ecology. 82-91.

Short, S. M., Rusanova, O. & Staniewski, M. A. 2011b. Novel Phycodnavirus genes amplified

from Canadian freshwater environments. Aquatic Microbial Ecology, 63 (1): 61-67.

Page 61: Diversity and Dynamics of Algal Viruses in the Bay of Quinte · Diversity and Dynamics of Algal Viruses in the Bay of Quinte Robin Marie Rozon Master of Science ... and Heather Niblock

50

Short, S. M. & Short, C. M. 2008. Diversity of algal viruses in various North American

freshwater environments. Aquatic Microbial Ecology, 51: 13-21.

Short, S. M. & Short, C. M. 2009. Quantitative PCR reveals transient and persistent algal

viruses in Lake Ontario, Canada. Environmental Microbiology, 11 (10): 2639-2648.

Short, S. M. & Suttle, C. A. 2002. Sequence Analysis of Marine Virus Communities Reveals

that Groups of Related Algal Viruses Are Widely Distributed in Nature. Applied and

Environmental Microbiology, 68 (3): 1290-1296.

Short, S. M. & Suttle, C. A. 2003. Temporal dynamics of natural communities of marine algal

viruses and eukaryotes. Aquatic Microbial Ecology, 32: 107-119.

Suttle, C. A. 2005. Viruses in the sea. Nature, 437 (7057): 356-61.

Suttle, C. A., Chan, A. M. & Cottrell, M. T. 1990. Infection of phytoplankton by viruses and

reduction of primary productivity. Nature, 347: 467-469.

Takashima, Y., Yoshida, T., Yoshida, M., Shirai, Y., Tomaru, Y., Takao, Y., Hiroishi, S. &

Nagasaki, K. 2007. Development and Application of Quantitative Detection of Cyanophages

Phylogenetically Related to Cyanopahge Ma-LMM01 Infecting Microcystis aeruginosa in Fresh

Water. Microbes and Environments, 22 (3): 207-213.

Tamura, K., Peterson, D., Peterson, N., Stecher, G., Nei, M. & Kumar, S. 2011. MEGA5:

molecular evolutionary genetics analysis using maximum likelihood, evolutionary distance, and

maximum parsimony methods. Molecular Biology and Evolution, 28 (10): 2731-9.

Thingstad, T. F. 2000. Elements of a theory for the mechanisms controlling abundance,

diversity, and biogeochemical role of lytic bacterial viruses in aquatic systems. Limnology and

Oceanography, 45 (6): 1320-1328.

Thyrhaug, R., Larsen, A., Thingstad, F. & Bratbak, G. 2003. Stable coexistence in marine algal

host-virus systems. Marine Ecology Progress Series, 254: 27-35.

Thomas, R., Grimsley, N., Escande, M.L., Subirana, L., Derelle, E., and Moreau, H. 2011.

Acquisition and maintenance of resistance to viruses in eukaryotic phytoplankton populations.

Environmental Microbiology, 13: 1412-1420.

Tucker, A. 1948. The Phytoplankton of the Bay of Quinte. Transactions of the American

Microscopical Society 67: 365-383.

Van Etten, J. L., Burbank, D. E., Xia, Y. & Meints, R. H. 1983. Growth Cycle of a Virus,

PBCV-1, that infects Chlorella-like Algae. Virology, 126: 1983.

Van Etten, J. L., Graves, M. V., Muller, D. G., Boland, W. & Delaroque, N. 2002.

Phycodnaviridae - large DNA algal viruses. Archives of Virology, 147 (8): 1479-1516.

Page 62: Diversity and Dynamics of Algal Viruses in the Bay of Quinte · Diversity and Dynamics of Algal Viruses in the Bay of Quinte Robin Marie Rozon Master of Science ... and Heather Niblock

51

Wilhelm, S. W., Carberry, M. J., Eldridge, M. L., Poorvin, L., Saxton, M. A. & Doblin, M. A.

2006. Marine and freshwater cyanophages in a Laurentian Great Lake: evidence from infectivity

assays and molecular analyses of g20 genes. Applied and Environmental Microbiology, 72 (7):

4957-63.

Wilhelm, S. W. & Smith, R. E. H. 2000. Bacterial carbon production in Lake Erie is influenced

by viruses and solar radiation. Canadian Journal of Fisheries and Aquatic Sciences, 57: 317-

326.

Wilson, W. H., Van Etten, J. L. & Allen, M. J. 2009. The Phycodnaviridae: The Story of How

Tiny Giants Rule the World. Current Topics in Microbiology and Immunology, 328: 1-42.

Winter, C., Bouvier, T., Weinbauer, M. G. & Thingstad, T. F. 2010. Trade-offs between

competition and defense specialists among unicellular planktonic organisms: the “killing the

winner” hypothesis revisited. Microbiology and Molecular Biology Reviews, 74 (1): 42-57.

Wommack, K. E. & Colwell, R. R. 2000. Virioplankton: Viruses in Aquatic Ecosystems.

Microbiology and Molecular Biology Reviews, 64 (1): 69-114.

Yoshida, T., Takashima, Y., Tomaru, Y., Shirai, Y., Takao, Y., Hiroishi, S. & Nagasaki, K.

2006. Isolation and characterization of a cyanophage infecting the toxic cyanobacterium

Microcystis aeruginosa. Applied and Environmental Microbiology, 72 (2): 1239-47.

Zavarzina, N. B. & Protsenko, A. E. 1958. The lysis of the cultures of Chlorella pyrenoidosa

PRINGS. Doklady Akademii Nauk SSSR, 122: 840-843.

Page 63: Diversity and Dynamics of Algal Viruses in the Bay of Quinte · Diversity and Dynamics of Algal Viruses in the Bay of Quinte Robin Marie Rozon Master of Science ... and Heather Niblock

52

Tables

Table 1. Limnological divisions of the Bay of Quinte.

Delineationa Length Width Depth Range

Upper Bay Trenton – Deseronto 35 km Various 4-8 m

Middle Bay Deseronto – Glenora 13 km 0.8-5.6 km 6-17 m

Lower Bay Glenora – Lake Ontario 16 km 3 km 17-52 m

a Map with the delineations indicated is shown in Figure 1 (Minns et al., 1986; Leisti and Doka, 2008).

Page 64: Diversity and Dynamics of Algal Viruses in the Bay of Quinte · Diversity and Dynamics of Algal Viruses in the Bay of Quinte Robin Marie Rozon Master of Science ... and Heather Niblock

53

Table 2. Sample locations.

Site Latitude Longitude Notes

Belleville (B) 44° 09’ 13.201” N 77° 20” 44.002” W PQ-U

B2 44° 09’ 15.480” N 77° 14’ 59.748” W SS-U

BQ6 44° 09’ 20.376” N 77° 20’ 00.744” W SS-U

BQ7 44° 08’ 35.844” N 77° 21’ 33.552” W SS-U

BQ8 44° 11’ 42.792” N 77° 01’ 46.056” W SS-U

BQ9 44° 09’ 56.772” N 77° 03’ 32.112” W SS-U

Hay Bay (HB) 44° 05’ 35.999” N 77° 04’ 18.001” W PQ-M

HB4 44° 06’ 52.549” N 77° 01’ 08.764” W SS-M

HB5 44° 08’ 58.560” N 76° 58’ 06.593” W SS-M

Napanee (N) 44° 10’ 49.001” N 77° 02’ 22.801” W PQ-U

NR1 44° 11’ 58.524” N 77° 00’ 56.556” W SS-U

P 44° 02’ 42.000” N 77° 07’ 00.001” W SS-M

Table adapted from data provided by DFO-GLLFAS (2012), with the GPS coordinates of the

historical Project Quinte (PQ) stations and spatial survey (SS) stations used in this research

project, each with limnological classification (U – Upper Bay, M – Middle Bay; No Lower Bay

stations were included in this study).

Page 65: Diversity and Dynamics of Algal Viruses in the Bay of Quinte · Diversity and Dynamics of Algal Viruses in the Bay of Quinte Robin Marie Rozon Master of Science ... and Heather Niblock

54

Table 3. Reagents used in PCR reactions.

Primer Total

vol.

10x

buffer MgCl2

dNTP

each Forward Reverse

pTaq

Invitrogen Viral

Template

AVS 50 µl 5 µl 1.5 mM 0.2 mM AVS1

a

400 nM

AVS2a

600 nM 1 unit 5 µl

MCP 25 µl 2.5 µl 1.5 mM 0.2 mM MCPf

b

500 nM

MCPrb

500 nM 0.5 units 2 µl

Sheath 25 µl 2.5 µl 1.5 mM 0.2 mM SheathF2

c

400 nM

SheathR2c

400 nM 0.5 units 2 µl

a Sequence described in Chen and Suttle (1995)

b Sequence described in Larsen et al. (2008)

c Sequence described in Takashima et al. (2007)

Page 66: Diversity and Dynamics of Algal Viruses in the Bay of Quinte · Diversity and Dynamics of Algal Viruses in the Bay of Quinte Robin Marie Rozon Master of Science ... and Heather Niblock

55

Table 4. Thermocycling parameters for PCR reactions.

Cycle name Denaturation

step

# of

cycles Step 1 Step 2 Step 3

Final

Extension Hold

AVS-40 2m @ 95°C 40 30s @

95°C

1m @

50°C

45s @

72°C

30m @

72°C

∞ @

8°C

AVS-25 2m @ 95°C 30 30s @

95°C

1m @

50°C

45s @

72°C

30m @

72°C

∞ @

8°C

MCP50-35 2m @ 94°C 35 30s @

95°C

30s @

50°C

30s @

72°C

20m @

72°C

∞ @

4°C

Sheath01 2m @ 94°C 35 30s @

94°C

30s @

56°C

1m @

72°C

20m @

72°C

∞ @

4°C

Page 67: Diversity and Dynamics of Algal Viruses in the Bay of Quinte · Diversity and Dynamics of Algal Viruses in the Bay of Quinte Robin Marie Rozon Master of Science ... and Heather Niblock

56

Table 5. Quantitative Primers and Probes.

Continued...

Target Name Probe Primer

252-M5.13

(Mimivirus-Pras 252)

ACT GTC TCA TGC GTG

CCT CG

F CTT GGT CTA GAT GCG

ACC

R CTC GAA GTC CAG GTT

AAT G

356-M5.3

(Mimivirus-Pras 356)

CGT AAC CTC GCC TGG

TCC AA

F CAC GAA GTC AAG ATC

AAC

R CAG TGT CAA GGT AGA

TGT A

399-M5.4

(Mimivirus-Prym 399)

CTT CCG CTT ACG CTC

AGT CTC T

F TCC GGT GAT ACA AAG

GTA

R CGT AGT CAA CAT AGA

GAG AAG

356-M5.14

(Mimivirus-Prym 356)

CAC ATC TGG AAC AAC

TTG ACT CTC C

F GCG TAT TGA TCG CCA

GTA

R GCC AAC CAT AGCATA

GTA AC

253-SH

(Sheath 253)

AAA CCT CTA TTA GCA

GTG TTA TCG TTT

F TCT CGT TTT ACG TAA

CGG

R GTG ACT AAT GCA GGA

CTA G

282-SH

(Sheath 282)

CGT AGA GGC AAC TGA

TAA CAC CAC TA

F CTG GGT CTA TCA GCA

ATC

R CAG AGT AGT AGT GAC

TAA TG

LO.20May09.33b

(Chlorovirus)

TGT CCA CAG TTC CGT

CCT CT

F GAT ACA GAT TCC GTT

ATG GT

R CAT CTT GAA GTG TGC

CTC

Page 68: Diversity and Dynamics of Algal Viruses in the Bay of Quinte · Diversity and Dynamics of Algal Viruses in the Bay of Quinte Robin Marie Rozon Master of Science ... and Heather Niblock

57

Table 5. Quantitative Primers and Probes. (continued)

a Previously described by Short and Short (2009)

b Previously described by Short et al (2011a)

Target Name Probe Primer

LO1b-49a

(Prasinovirus49)

CGA CAA TCT TCC AGG

TG

F TGT TAC TCA ACT CTG TA

CTT G

R GCG AAC TTG TAA GTC

CTA CC

LO Jul.16.20b

(Prasinovirus16.20)

TGG AAC GCA AGG

CAA CAT ACC

F CAG TTG GCC TAC AAG

ATT

R CCT TCA TGG TGA CAG

TTG

LO1a-68a

(Prasinovirus68)

TTG CTA CTC ATC CCT

CG

F GTT GTA TCC ATC TAT

TAT GAT TGC

R GAA AGT CTC ATA GGT

GAT GC

Page 69: Diversity and Dynamics of Algal Viruses in the Bay of Quinte · Diversity and Dynamics of Algal Viruses in the Bay of Quinte Robin Marie Rozon Master of Science ... and Heather Niblock

58

Table 6. Reagents used in quantitative PCR.

Total

Volume

10X

Buffer MgCl2

dNTP

each Forward Reverse Probe

ROX

dye

pTaq

Invitrogen Template

Group A 20 µL 2 µL 5 mM 0.2 mM 0.25 µM 0.25 µM 0.1 µM 30 nM 0.5 units 2 µL

Group B 25 µL 2.5 µL 5 mM 0.2 mM 0.4 µM 0.4 µM 0.2 µM 30 nM 0.625 units 2 µL

Group C 20 µL 2 µL 5 mM 0.2 mM 0.5 µM 0.5 µM 0.25 µM 30 nM 0.5 units 2 µL

Group D 20 µL 2 µL 1.8 mM 0.2 mM 0.4 µM 0.4 µM 0.2 µM 30 nM 0.5 units 2 µL

Page 70: Diversity and Dynamics of Algal Viruses in the Bay of Quinte · Diversity and Dynamics of Algal Viruses in the Bay of Quinte Robin Marie Rozon Master of Science ... and Heather Niblock

59

Table 7. Test of Primer and Probe Specificity.

Target

Sequence

Closest

non-target

relative

Percent

nucleotide

identity

Mismatches Ct at 10

7

target

Ct at 107

non-target Forward

primer

Reverse

primer Probe

252-M5.13 (Mimivirus-Pras 252)

252-M5.1 78 6 2 4 16.9 28.7

356-M5.3 (Mimivirus-Pras 356)

252-M5.5 73 1 14 9 14.8 31.5

399-M5.4 (Mimivirus-Prym 399)

399-M5.1 94 3 3 3 15.5 31.2

356-M5.14 (Mimivirus-Prym 356)

399-M5.14 74 5 6 6 12.8 28.2

253-SH (Sheath 253)

282-SH 92 3 1 5 12.7 23.0

282-SH (Sheath 282)

253-SH 92 1 2 6 13.0 No Ct

LO.20May09.33b

(Chlorovirus) UTM.09jun09.47 69 6 7 13 12.3 33.1

LO1b-49a

(Prasinovirus49) LO1a-42 73 6 4 4 14.9 33.3

LO Jul.16.20b

(Prasinovirus16.20) LO1a-68 81 2 4 4 13.9 31.5

LO1a-68a

(Prasinovirus68) LO1b-53 83 6 3 4 14.3 39.6

a Previously described by Short and Short (2009)

b Previously described by Short et al (2011a)

Page 71: Diversity and Dynamics of Algal Viruses in the Bay of Quinte · Diversity and Dynamics of Algal Viruses in the Bay of Quinte Robin Marie Rozon Master of Science ... and Heather Niblock

60

Table 8. Friedman analysis of virus abundances between stations.

Virus abundance

between stations χ²(2) p-value

Post-hoc: Significant

differences between..

Sheath 282 12.667 < .01**

HB-B

Sheath 253 5.600 .06 --

Chlorovirus 0.600 .74 --

Prasinovirus49 7.789 .02* No significant differences

Prasinovirus68 15.440 < .01**

HB-N; HB-B

Prasinovirus16.20 18.727 < .01**

HB-N; HB-B

Mimivirus-Pras 356 4.667 .10 --

Mimivirus-Pras 252 0.444 .80 --

Mimivirus-Prym 399 2.387 .30 --

Mimivirus-Prym 356 0.000 1.0 --

* Indicates significance at p < .05

** Indicates significance at p < .01

Page 72: Diversity and Dynamics of Algal Viruses in the Bay of Quinte · Diversity and Dynamics of Algal Viruses in the Bay of Quinte Robin Marie Rozon Master of Science ... and Heather Niblock

61

Figures

Figure 1. Map of the Bay of Quinte. Note the locations of the stations sampled during the 2011 sampling season. The historical

Project Quinte stations (B, HB, N) are marked in larger font than the spatial stations. Note the approximate delineations between the

Upper Bay, Middle Bay and Lower Bay. Image taken at an altitude of 53.6 km (Google Earth; USA, 2012).

Page 73: Diversity and Dynamics of Algal Viruses in the Bay of Quinte · Diversity and Dynamics of Algal Viruses in the Bay of Quinte Robin Marie Rozon Master of Science ... and Heather Niblock

62

Figure 2. Neighbor joining phylogeny of inferred amino acid sequences from DNA polymerase

gene fragments (polB) of NCLDVs. Virus sequences obtained from the Bay of Quinte are

shown in bold, whereas sequences from GenBank are shown in italicized font with accession

numbers in parentheses (Short et al., 2011a). For Bay of Quinte sequences, OTUs are

represented by a single sequence and are followed in parentheses by the number of redundant

sequences observed in clone libraries; singleton sequences are those without numbers in

parentheses. Values at nodes indicate bootstrap support as a percent of 500 replicates. The scale

bar represents the number of substitutions per site. **Indicates sequences used for primer and

probe design (see Table 5 for quantitative nomenclature).

Page 74: Diversity and Dynamics of Algal Viruses in the Bay of Quinte · Diversity and Dynamics of Algal Viruses in the Bay of Quinte Robin Marie Rozon Master of Science ... and Heather Niblock

63

Figure 3. Neighbor joining phylogeny of inferred amino acid sequences from major capsid

protein (MCP) gene fragments of NCLDVs. Virus sequences obtained from the Bay of Quinte

are shown in bold, whereas sequences from GenBank are shown in italicized font with accession

numbers in parentheses (Larsen et al., 2008; Park et al., 2011). For Bay of Quinte sequences,

OTUs are represented by a single sequence and are followed in parentheses by the number of

redundant sequences observed in clone libraries; Singleton sequences are those without numbers

in parentheses. Values at nodes indicate bootstrap support as a percent of 500 replicates. The

scale bar represents the number of substitutions per site. **Indicates sequences used for primer

and probe design (see Table 5 for quantitative nomenclature).

Page 75: Diversity and Dynamics of Algal Viruses in the Bay of Quinte · Diversity and Dynamics of Algal Viruses in the Bay of Quinte Robin Marie Rozon Master of Science ... and Heather Niblock

64

Figure 4. Abundances of individual virus genes plotted against time. Note that the y-axis is

drawn in log scale to highlight the range of values, and the X-axis intercept, “bd”, indicates that

the genes were below detection limits of the assay. Also note that Napanee abundances begin

one sampling event after Belleville and Hay Bay.

Page 76: Diversity and Dynamics of Algal Viruses in the Bay of Quinte · Diversity and Dynamics of Algal Viruses in the Bay of Quinte Robin Marie Rozon Master of Science ... and Heather Niblock

65

Figure 5. Virus gene abundances at each station plotted against time. Note the log scale of the Y

axis, and the X-axis intercept, “bd”, indicates that the genes were below detection limits of the

assay.

Page 77: Diversity and Dynamics of Algal Viruses in the Bay of Quinte · Diversity and Dynamics of Algal Viruses in the Bay of Quinte Robin Marie Rozon Master of Science ... and Heather Niblock

66

Figure 6. Virus gene abundance. The plot shows the median, first and third quartiles of each

individual virus’ abundance, with whiskers showing the range of values. Note the log scale of

the Y axis, and the X-axis intercept, “bd”, to illustrate a range from below detection limits of the

assay.

Page 78: Diversity and Dynamics of Algal Viruses in the Bay of Quinte · Diversity and Dynamics of Algal Viruses in the Bay of Quinte Robin Marie Rozon Master of Science ... and Heather Niblock

67

Figure 7. Regression of the sum of virus gene abundances on chlorophyll a. Note that at Hay

Bay on July 19th

, the total abundance peaked at 15x higher than any other abundance from that

station, therefore that point was included above the axis, labeled with sum of virus abundance in

parentheses. The dotted regression line at Hay Bay is the regression including July 19th

, while

the solid regression line is the regression excluding July 19th

.

Page 79: Diversity and Dynamics of Algal Viruses in the Bay of Quinte · Diversity and Dynamics of Algal Viruses in the Bay of Quinte Robin Marie Rozon Master of Science ... and Heather Niblock

68

Figure 8. Cluster analysis of virus abundance. Dendrograms created by UPGMA clustering of a distance matrix created from pair-

wise Pearson Correlations of abundance (i.e., 1 – the correlation coefficient). The upper left dendrogram was created using all the

biweekly samples for all ten quantitative assays. Note that Prasinovirus16.20 and Prasinovirus68 were excluded from the Belleville

dendrogram as these viruses were consistently below the detection limit at this location.

Page 80: Diversity and Dynamics of Algal Viruses in the Bay of Quinte · Diversity and Dynamics of Algal Viruses in the Bay of Quinte Robin Marie Rozon Master of Science ... and Heather Niblock

69

Figure 9. A comparison of different proximity measures. A) Dendrogram based on a distance

matrix calculated from a Pearson Correlation, and grouping patterns created by the UPGMA

clustering method. B) Dendrogram based on a distance matrix calculated using a Bray-Curtis

dissimilarity matrix, clustered using the same UPGMA method.

Page 81: Diversity and Dynamics of Algal Viruses in the Bay of Quinte · Diversity and Dynamics of Algal Viruses in the Bay of Quinte Robin Marie Rozon Master of Science ... and Heather Niblock

70

Figure 10. Cluster analysis of biweekly stations. Dendrogram based on a distance matrix

created from a Pearson Correlation, and an UPGMA clustering method using a composite of all

quantitative abundance data. Each station designation is followed by the sample date during the

2011 Bay of Quinte sampling season.

Page 82: Diversity and Dynamics of Algal Viruses in the Bay of Quinte · Diversity and Dynamics of Algal Viruses in the Bay of Quinte Robin Marie Rozon Master of Science ... and Heather Niblock

71

Figure 11. Cluster analysis of spatial stations. Dendrogram created based on a distance matrix

created from a Pearson Correlation, and an UPGMA clustering method using the abundance

data from all ten virus assays during the spatial survey. The limnological divisions of each

station are indicated in parentheses.