non-pollen palynomorphs as indicators of water quality in lake simcoe, ontario, canada

16
This article was downloaded by: [Eindhoven Technical University] On: 19 October 2014, At: 15:31 Publisher: Taylor & Francis Informa Ltd Registered in England and Wales Registered Number: 1072954 Registered office: Mortimer House, 37-41 Mortimer Street, London W1T 3JH, UK Palynology Publication details, including instructions for authors and subscription information: http://www.tandfonline.com/loi/tpal20 Non-pollen palynomorphs as indicators of water quality in Lake Simcoe, Ontario, Canada Donya C. Danesh a , Francine M.G. McCarthy a , Olena Volik a & Matea Drljepan a a Department of Earth Sciences , Brock University , St Catharines , ON , Canada , L2S 3A1 b Department of Biology , Queens University , Kingston , ON , Canada , K7L 3N6 Accepted author version posted online: 13 Mar 2013.Published online: 01 Nov 2013. To cite this article: Donya C. Danesh , Francine M.G. McCarthy , Olena Volik & Matea Drljepan (2013) Non-pollen palynomorphs as indicators of water quality in Lake Simcoe, Ontario, Canada, Palynology, 37:2, 231-245, DOI: 10.1080/01916122.2013.782366 To link to this article: http://dx.doi.org/10.1080/01916122.2013.782366 PLEASE SCROLL DOWN FOR ARTICLE Taylor & Francis makes every effort to ensure the accuracy of all the information (the “Content”) contained in the publications on our platform. However, Taylor & Francis, our agents, and our licensors make no representations or warranties whatsoever as to the accuracy, completeness, or suitability for any purpose of the Content. Any opinions and views expressed in this publication are the opinions and views of the authors, and are not the views of or endorsed by Taylor & Francis. The accuracy of the Content should not be relied upon and should be independently verified with primary sources of information. Taylor and Francis shall not be liable for any losses, actions, claims, proceedings, demands, costs, expenses, damages, and other liabilities whatsoever or howsoever caused arising directly or indirectly in connection with, in relation to or arising out of the use of the Content. This article may be used for research, teaching, and private study purposes. Any substantial or systematic reproduction, redistribution, reselling, loan, sub-licensing, systematic supply, or distribution in any form to anyone is expressly forbidden. Terms & Conditions of access and use can be found at http:// www.tandfonline.com/page/terms-and-conditions

Upload: matea

Post on 26-Feb-2017

214 views

Category:

Documents


1 download

TRANSCRIPT

Page 1: Non-pollen palynomorphs as indicators of water quality in Lake Simcoe, Ontario, Canada

This article was downloaded by: [Eindhoven Technical University]On: 19 October 2014, At: 15:31Publisher: Taylor & FrancisInforma Ltd Registered in England and Wales Registered Number: 1072954 Registered office: Mortimer House,37-41 Mortimer Street, London W1T 3JH, UK

PalynologyPublication details, including instructions for authors and subscription information:http://www.tandfonline.com/loi/tpal20

Non-pollen palynomorphs as indicators of water qualityin Lake Simcoe, Ontario, CanadaDonya C. Danesh a , Francine M.G. McCarthy a , Olena Volik a & Matea Drljepan aa Department of Earth Sciences , Brock University , St Catharines , ON , Canada , L2S 3A1b Department of Biology , Queens University , Kingston , ON , Canada , K7L 3N6Accepted author version posted online: 13 Mar 2013.Published online: 01 Nov 2013.

To cite this article: Donya C. Danesh , Francine M.G. McCarthy , Olena Volik & Matea Drljepan (2013) Non-pollenpalynomorphs as indicators of water quality in Lake Simcoe, Ontario, Canada, Palynology, 37:2, 231-245, DOI:10.1080/01916122.2013.782366

To link to this article: http://dx.doi.org/10.1080/01916122.2013.782366

PLEASE SCROLL DOWN FOR ARTICLE

Taylor & Francis makes every effort to ensure the accuracy of all the information (the “Content”) containedin the publications on our platform. However, Taylor & Francis, our agents, and our licensors make norepresentations or warranties whatsoever as to the accuracy, completeness, or suitability for any purpose of theContent. Any opinions and views expressed in this publication are the opinions and views of the authors, andare not the views of or endorsed by Taylor & Francis. The accuracy of the Content should not be relied upon andshould be independently verified with primary sources of information. Taylor and Francis shall not be liable forany losses, actions, claims, proceedings, demands, costs, expenses, damages, and other liabilities whatsoeveror howsoever caused arising directly or indirectly in connection with, in relation to or arising out of the use ofthe Content.

This article may be used for research, teaching, and private study purposes. Any substantial or systematicreproduction, redistribution, reselling, loan, sub-licensing, systematic supply, or distribution in anyform to anyone is expressly forbidden. Terms & Conditions of access and use can be found at http://www.tandfonline.com/page/terms-and-conditions

Page 2: Non-pollen palynomorphs as indicators of water quality in Lake Simcoe, Ontario, Canada

Non-pollen palynomorphs as indicators of water quality in Lake Simcoe, Ontario, Canada

Donya C. Danesha,b*, Francine M.G. McCarthya, Olena Volika and Matea Drljepana

aDepartment of Earth Sciences, Brock University, St Catharines, Ontario, Canada L2S 3A1; bDepartment of Biology,Queens University, Kingston, Ontario, Canada K7L 3N6

The distribution of non-pollen palynomorphs (NPP) in a core from Cook’s Bay, Lake Simcoe, Ontario, Canadashows a response to changes in water quality accompanying agriculture, urbanization, and industrialization. Lowconcentrations of nutrients in sediments with little non-arboreal pollen (NAP) record low disturbance prior toEuropean settlement around the 1850s. These sediments are rich in desmids such as Cosmarium spp., Euastrum spp.,and Staurastrum spp., an assemblage indicative of oligotrophic conditions. A decline in desmids, together with anincrease in dinoflagellate cysts and thecamoebians up-core is consistent with increased nutrients. Abundant phytolithsin sediments that are relatively rich in Poaceae and other NAP records the draining of the Holland Marshes. A sharpincrease in nutrient levels, together with a transition from high nitrite (NO2) to high nitrate (NO3) concentrations,records a sudden increase in biological oxygen demand leading to depletion of dissolved oxygen associated with thecreation of polders in the 1920s and 1930s. A second influx of phytoliths immediately preceded the sharp rise inAmbrosia, recording rapid land clearing accompanying the five-fold post-World War II population boom in theCook’s Bay watershed. These Ambrosia-rich sediments are rich in metals and have high total phosphorus and NO3,with abundant Pediastrum spp. and Peridinium spp., notably Peridinium willei and Peridinium volzii, recordingeutrophication. The abundance of the ciliate Codonella cratera and the difflugiid thecamoebians Cucurbitella tricuspisand Difflugia protaeiformis in palynological preparations, as well as in washed thecamoebian samples from the upperpart of the core, records low dissolved oxygen associated with continued eutrophication of Cook’s Bay.

Keywords: non-pollen palynomorphs; water quality; eutrophication; Lake Simcoe; Canada; polders

1. Introduction

1.1. Environmental setting

Lake Simcoe is the largest lake in southern Ontario,

Canada after the Laurentian Great Lakes with a sur-

face area of 722 km2 (Winter et al. 2007; LSEMS

2008). Thirty-five tributaries originating mostly alongthe Oak Ridges Moraine flow north before discharging

into Lake Simcoe (Singer et al. 2003), which is part of

the Trent-Severn Waterway that connects Georgian

Bay (Lake Huron) to Lake Ontario. Lake Simcoe has a

single outflow at Atherley Narrows in the north

(LSRCA 2009; OMOE 2010a), and a residence time of

approximately 11 years (Johnson and Nicholls 1989;

Helm et al. 2011; Palmer et al. 2011; Winter et al.2011). In addition to the main basin, Lake Simcoe con-

sists of two bays, Kempenfelt Bay and Cook’s Bay

(Young et al. 2010) (Figure 1). Lake Simcoe makes up

approximately 20% of the watershed (Singer et al.

2003) that incorporates 23 municipalities, including

Barrie, Newmarket and Aurora, which have the fastest

growing populations in the region (LSRCA 2007).

1.2. Cultural eutrophication of Cook’s Bay

The Lake Simcoe region has been adversely impacted

by anthropogenic activities over the last two centuries,

beginning with the establishment of York County

by Governor John Graves Simcoe in the 1790s. Thiscontinued with the construction of Yonge Street

(Highway 11) north from Toronto to Lake Simcoe

along the Iroquois trails that connected Lake Huron to

Lake Ontario, which played a fundamental role in the

planning and layout of Upper Canada (LSRCA 2000).

Several communities were established along the route,

including the two largest in the East Holland subwa-

tershed – Newmarket (population 74,295; StatisticsCanada 2012) and Aurora (population 47,629;

Statistics Canada 2012). Two periods of rapid popula-

tion growth were recorded (LSRCA 2000), (1) in the

1850s when the Ontario, Simcoe and Huron Railway

was completed, and the combined population of

Newmarket and Aurora rose from �600 in 1841 to

�3350 in 1871, and (2) after World War II (WWII),

when the combined population rose to �32,550 in1971 from �6750 in 1941. While natural sources of

*Corresponding author. Email: [email protected]

� 2013 AASP – The Palynological Society

Palynology, 2013

Vol. 37, No. 2, 231–245, http://dx.doi.org/10.1080/01916122.2013.782366

Dow

nloa

ded

by [

Ein

dhov

en T

echn

ical

Uni

vers

ity]

at 1

5:31

19

Oct

ober

201

4

Page 3: Non-pollen palynomorphs as indicators of water quality in Lake Simcoe, Ontario, Canada

eutrophication (e.g. atmospheric and terrestrial runoff

and plant debris; Maier et al. 2009) contribute approxi-

mately 30% of the phosphorus (P) loading to Lake

Simcoe annually (Palmer et al. 2011). The remaining

70% is attributed to cultural eutrophication from activi-

ties such as urbanization, agriculture, surface runoff and

sewage treatment plants (OMOE 2010a). Anthropogenicactivities are thus much more significant causes of eutro-

phication in this region (Smith & Schindler 2009).

Urban development is one of the main causes of the

recent increase in P loading to Lake Simcoe. There is

little infiltration of storm water runoff and 15 municipal

water treatment plants, of which seven facilities assimi-

late wastewater directly into the lake, and the remaining

eight facilities discharge into tributaries draining into the

lake (LSPP 2009; Young et al. 2010). The five main sour-

ces of P input to Lake Simcoe are: (1) tributaries (urban

and non-urban), (2) polders (East/West Holland Marsh),

(3) sewage treatment plants, (4) septic systems and (5)

the atmosphere (LSRCA 2007).

Water quality issues, such as harmful and excessive

algal blooms leading to beach closures and contamina-tion of drinking water, began to present a problem in

the 1970s due to a substantial increase in total phos-

phorus (TP) loading (Evans et al. 1996; Winter et al.

2007; Palmer et al. 2011) from the natural background

load of 32 tonnes per year during pre-European settle-

ment to current TP loading of 72 tonnes per year

(LSRCA 2009; Ginn 2011; Winter et al. 2011). In tem-

perate lakes, P is considered the limiting nutrient

Figure 1. Map of the Lake Simcoe basin, its surrounding subwatersheds and its location relative to the Great Lakes (fromOMOE 2010). The towns of Newmarket and Aurora are located within the East Holland subwatershed. The inset shows thelocation of the core and dissolved oxygen contours (from Stantec Consulting Inc. 2006).

232 D.C. Danesh et al.

Dow

nloa

ded

by [

Ein

dhov

en T

echn

ical

Uni

vers

ity]

at 1

5:31

19

Oct

ober

201

4

Page 4: Non-pollen palynomorphs as indicators of water quality in Lake Simcoe, Ontario, Canada

needed for primary production (Gloterman et al. 1975;

Schindler 1977; Wetzel 2001), and the highest TP con-

centrations in Lake Simcoe are in the southernmost

part of Cook’s Bay, ranging from �14 mg/L to�48 mg/L (Winter et al. 2002; LSRCA 2009; Young

et al. 2010; Ginn 2011). This is because Cook’s Bay, a

relatively small (surface area 44 km2) and shallow

(depths up to 15 m) arm of Lake Simcoe, receives

�20 tonnes of P per year (roughly 22% of the total TP

loading into all of Lake Simcoe) into a small volume of

water (LSRCA 2007, 2009). The high TP loads to

Cook’s Bay come primarily from surface runoff via itstributaries from the East and West Holland River sub-

watersheds. As phosphates tend to bind to most soils

and sediments (Correll 1998), this reflects intense agri-

cultural activity (Holland Marsh is the largest culti-

vated marsh area in Ontario) and urbanization

(Newmarket and Aurora) (Winter et al. 2002; Palmer

et al. 2011).

Dissolved oxygen (DO) concentrations (Figure 1)have been observed to be as low as 1.3mg/L in shallow

regions of Cook’s Bay near the outlets of the East and

West Holland Rivers (Stantec Consulting Ltd. 2006;

Young et al. 2010). A substantial decrease in DO lev-

els in the hypolimnion due to oxygen consumption

during bacterial decay (biochemical oxygen demand –

BOD) and macrophyte respiration (Petr 2000) is one

of the chief undesirable effects of high nutrient levelsin eutrophic lakes. The dense growth of macrophytes

found within Cook’s Bay (average maximum plant

biomass �1118.5 g/m2) (Ginn 2011) alters the compo-

sition of benthic communities (Kilgour et al. 2008;

Young et al. 2010). Their respiration and decomposi-

tion produces DO levels of < 3 mg/L (Young et al.

2010) which is well below the recommended minimum

for cold-water biota in fresh water: 9.5 mg/L for theearly life stages and 6.5 mg/L for other life stages (Ca-

nadian Council of Ministers of the Environment

2007). This adversely impacts cold-water fish species,

including lake trout and whitefish (Smith & Des-

vousges 1986; Winter et al. 2007). As a result, the pop-

ular cold-water fishery (primarily ice fishing) that

generates over $200 million per year (Young et al.

2010; Palmer et al. 2011) has been sustained in recentyears because stocks of fish are added each season

(OMOE 2010a). Due to the declining ecological

health of the lake, the Lake Simcoe Protection Plan

(2009) was set in place by the Ontario Ministry of the

Environment targeting Lake Simcoe as a key site for

studying lake management issues (OMOE 2010b).

Moreover, from 1980 to 2010 the Ontario Ministry of

the Environment, in partnership with the Lake SimcoeRegion Conservation Authority, monitored water

quality throughout Lake Simcoe biweekly during the

ice-free seasons, thus contributing to one of the most

extensive monitoring records of water chemistry in

Canada (Winter et al. 2011).

1.3. Non-pollen palynomorphs as proxies ofeutrophication

Measuring the concentration of major nutrients (e.g.

P or nitrites) and the products of photosynthesis and bio-

mass (e.g. chlorophyll a in lake water) provides valuable

insights into ecosystem health (Carlson 1977). However,

these synoptic assessments present only a brief snapshot

at any particular time (Detenbeck et al. 1996; Bradshawet al. 2002; Torbick et al. 2008). Although Lake Simcoe

has extensive monitoring data, these data do not reflect

pre-disturbance conditions that could prove beneficial

for future management strategies. Therefore, biological

proxy indicators (organisms that leave a fossil record in

the sediment) reflect environmental conditions over an

extended period of time (Yoder & Rankin 1998).

Particularly when integrated with geochemical analysisof the sediment, time series data made available by

pollen and fossil plankton records in cores can demon-

strate the long-term pattern of cultural eutrophication

(Dale 2009).

A variety of microfossil proxies has been studied in

cores to document cultural eutrophication in Southern

Ontario lakes, including diatoms (Dixit & Smol 1994;

Ramstack et al. 2003; Ekdahl et al. 2004, 2007; Kiretaet al. 2007), thecamoebians (Reinhardt et al. 2005;

McCarthy et al. 2012), and a variety of non-

pollen palynomorphs (NPP) (e.g. Turton & McAn-

drews 2006), including dinoflagellate cysts (Burden

et al. 1986; McCarthy et al. 2011; McCarthy &

Krueger forthcoming). NPP are being used increas-

ingly to determine long-term environmental impacts.

The recent publication of special volumes edited byvan Geel (2006) and Haas (2010) have helped to high-

light the potential of NPP for paleoenvironmental and

geoarcheological studies.

This study uses a multi-proxy approach to compare

the distribution of NPP, focusing on algal and proto-

zoan microfossils in a core from Cook’s Bay, with the

well-documented historical record of human impact in

the Lake Simcoe region. The pollen record serves as achronological tool and as a proxy of land-use changes,

with metal concentrations as chemical proxies of water

quality measured from sediments in the same core.

2. Methods

A 102 cm-long sediment core was collected from

Cook’s Bay (44�10’31”N, -79�30’16”W) on 28 August28 2010 in a region showing trends of low DO

(Figure 1; Stantec Consulting Ltd. 2006). A DO read-

ing of 12 mg/L was recorded at midday using an YSI

Palynology 233

Dow

nloa

ded

by [

Ein

dhov

en T

echn

ical

Uni

vers

ity]

at 1

5:31

19

Oct

ober

201

4

Page 5: Non-pollen palynomorphs as indicators of water quality in Lake Simcoe, Ontario, Canada

556 MPS Dissolved Oxygen Meter. Coring employed a

technology developed by EnviroFix Corporation that

is currently patent pending. This technology uses a

pneumatic system to recover undisturbed sedimentcores. The corer is located on the platform of a pon-

toon boat, and the pontoon boat is anchored for stabil-

ity. The 5 cm-diameter aluminium tubes are lined with

clear plastic tubes to facilitate initial visual analysis of

the core. The advantage of this technology is that the

operator can control the force exerted and the rate of

penetration during drilling. However, as this technol-

ogy is in the early stages of testing, coring was re-stricted to a maximum water depth of 5.9 m for

sediment penetration.

Subsamples of 2 cm3 were taken every 10 cm and

submitted with a chain-of-custody to a Canadian As-

sociation for Laboratory Accreditation (CALA) Cer-

tified laboratory for chemical analysis of nitrite

(NO2), nitrate (NO3), TP, and heavy metals. Analysis

followed procedures laid out in the Standard Methodsfor the Analysis of Water and Wastewater (Rice et al.

2012) using a Thermo-Fisher iCAP 6300 ICP Spec-

trometer. Subsamples of 2.5 cm3 were taken every

5 cm for palynological analysis in the Palynology

Laboratory at Brock University, using a slightly mod-

ified procedure from Faegri and Iversen (1975); muds

were disaggregated using a weak base (0.02% Calgon),

and no acetolysis treatment was performed. Carbo-nates were dissolved using warm 10% hydrochloric

acid (HCl), and warm hydrofluoric acid (HF) (48%)

was used to dissolve silicates. A tablet containing a

known number of Lycopodium clavatum spores was

added during HCl treatment in order to quantify the

concentrations of palynomorphs, following Stock-

marr (1971). The elimination of the use of potassium

hydroxide (KOH) and acetolysis, both of which arestandard in the palynological processing of freshwater

sediments, is in keeping with the recommendations of

Mertens et al. (2009), and although a hot water

(90 �C) bath was used during acid treatment, the

exposures were relatively short (< 30 minutes). Resi-

dues were sieved using 10-mm Nitex mesh and

mounted on slides using glycerine jelly. The slides

were examined using a light microscope at 400 �mag-nification, and pollen and embryophyte spores were

identified following McAndrews et al. (1973). Rela-

tive abundance of pollen and spores was calculated

based on a total sum of at least 200 pollen grains.

NPP were identified to the lowest possible taxonomic

level at 400 � magnification using a Leica DMLB mi-

croscope using the following references: Beyens and

Meisterfeld 2001; Wehr and Sheath 2003; Coesel andMeesters 2007; Kramer et al. 2010; Mudie et al. 2010.

Palynomorphs were photographed using a Leica EC3

Digital Imaging Camera.

Fourteen subsamples were taken at the same depths

as the subsamples taken for palynological analysis for

loss on ignition (LOI) analysis following the method

presented in Heiri et al. (2001) to estimate the organic,carbonate and silicate content of the lake sediments.

Samples were dried in an oven at 100 �C for 24 hours

to determine the dry weight of the sediment. They were

then placed in a muffle furnace and heated to 550 �Cfor 12 hours to combust the organic matter. Once the

samples were cooled, they were placed again in the

muffle furnace at 1000 �C to combust the calcium car-

bonate, leaving only silicates. Samples were weighedbefore and after each step to determine the weight of

organics and calcium carbonate, which in turn pro-

vided the sediment profile of the core.

Because tests of amoebae were relatively common

in our palynological preparations, six subsamples of

2.5 cm3 were taken from the core and prepared for

conventional (non-chemical) thecamoebian analysis at

Brock University in order to compare thecamoebianassemblages in palynological preparations with con-

ventional thecamoebian preparations. Sediments were

sieved to retain the > 45-mm fraction, and in order to

allow comparison with a variety of published studies

(McCarthy et al. 1995; Scott et al. 2001; Patterson &

Kumar 2002; Reinhardt et al. 2005); the 45–63-mm

fraction was analysed separately from the > 63-mm

fraction. Thecamoebians were examined in a Petri dishat 100 � using a Leica ZOOM 2000 and identified pri-

marily using the key of Kumar and Dalby (1998) and

the monograph of Medioli and Scott (1983).

3. Results

3.1. Loss on ignition (LOI) and chemical analysis

LOI results show three intervals of increased relative

abundance of organic matter at approximately 80 cm,

45 cm and 6 cm (Figure 2) and relative abundance of

silicates increases slightly up-core. TP concentrations

increase rapidly up-core from 98.5 mg/kg at 85 cm to

409 mg/kg at 44 cm, remain somewhat stable to 24 cm

(424 mg/kg), and rise again to 716 mg/kg at 6 cm

(Figure 2). NO2 concentrations rise sharply from0.63 mg/kg at 85 cm to 9.29 mg/kg at 44 cm, and de-

cline sharply to 0.52 mg/kg at 6 cm. NO3 concentrations

remain steady around 0.3mg/kg from 85 cm to 64 cm,

then peak to 18.7 mg/kg at 24 cm and reduce to

14.9 mg/kg at 6 cm. Metal analysis demonstrated a

trend up-core: very low concentrations at the base of the

core, an approximate doubling in concentration between

35 cm and 25 cm, then again from 25 cm to 15 cm, andfinally reaching peak concentrations at 14 cm followed

by a slight decline in concentrations of chromium (Cr),

lead (Pb), and arsenic (As) at 6 cm (Figure 2).

234 D.C. Danesh et al.

Dow

nloa

ded

by [

Ein

dhov

en T

echn

ical

Uni

vers

ity]

at 1

5:31

19

Oct

ober

201

4

Page 6: Non-pollen palynomorphs as indicators of water quality in Lake Simcoe, Ontario, Canada

3.2. Chronology

Chronology was inferred based on the well-established

pollen zonation of southern Ontario by McAndrews

(1981); McAndrews and Boyko-Diakonow (1989);

Campbell and McAndrews (1991); Yu and McAn-

drews (1994). These pollen zones were established

using several sites with 210Pb-dated sediment cores inorder to establish calendar years that corresponded

with changes in pollen assemblages (McAndrews

1988). Although sufficient sediment was not available

for radiometric dating of the Cook’s Bay core, dates

from a 210Pb-dated core from a recently published

study from Cook’s Bay (Hawryshyn et al. 2012) taken

near this study site corroborate the chronology deter-

mined by the pollen diagram (Figure 3). Constrainedcluster analysis (CONISS) was used on the relative

abundance of the pollen taxa found in the Cook’s Bay

core in order to ensure independent pollen zonation

(Figure 3). With the use of the well-established pollen

zonation for regional vegetation of southern Ontario

(McAndrews 1981; McAndrews & Boyko-Diakonow

1989; Campbell & McAndrews 1991; Yu & McAn-

drews 1994), and CONISS, two distinct pollen zoneswere identified: pollen zone 3 and pollen zone 4

(Figure 3). The latter (indicating disturbance) is deter-

mined by a distinct increase in Ambrosia (ragweed) and

Poaceae (grass) pollen referred to as the ‘ragweed zone’

(McAndrews 1994). Moreover, these herbaceous plantsare found in disturbed environments caused by defor-

estation and are established to show direct association

with European settlement. Early European settlement

(late eighteenth century to nineteenth century) is thus

recorded by an initial rise in Ambrosia and Poaceae at

�75 cm. CONISS also helped further determine two

pollen subzones at: (1) �45 cm, recording recent paly-

nological events, like the regional decline in Ulmus

resulting from Dutch Elm Disease which devastated

elm populations all across North America during the

mid 1900s (Newhouse et al. 2007; Solheim et al. 2011),

and (2) �24 cm, recording the population boom

through the 1950s as determined by a second sharp rise

in Ambrosia and Poaceae, further supporting our chro-

nology. This is consistent with the 210Pb dates from

Hawryshyn et al. (2012) and the regional pollen chro-nology of southern Ontario (McAndrews 1988, 1994).

3.3. Palynological analysis

Relative abundance of Pinus strobus (white pine)

decreases slightly as Tsuga (hemlock) and Betula

(birch) increase sharply between 100 cm and 80 cm,

while Fagus (beech), Acer saccharum (sugar maple),and Quercus (oak) remain relatively stable during this

period, and Picea (spruce) abundance is very low. An

initial increase in Ambrosia is evident between 80 cm

Figure 2. Loss On Ignition (LOI) and metal concentrations scaled by depth from a sediment core taken within Cook’s Bay,Lake Simcoe, Ontario, Canada. Values for nitrate and nitrite are almost the same in the bottom three samples, and nitrate valuesbegin to increase at approximately 65 cm. This figure shows organic, marly mud with a slight increase in silicates and negligibleconcentrations of metals until the mid-1990s. The solid horizontal lines delineate the two significant Ambrosia rises. CaCO3 ¼calcium carbonate.

Palynology 235

Dow

nloa

ded

by [

Ein

dhov

en T

echn

ical

Uni

vers

ity]

at 1

5:31

19

Oct

ober

201

4

Page 7: Non-pollen palynomorphs as indicators of water quality in Lake Simcoe, Ontario, Canada

and 70 cm in the core (Figure 3). There is a sharp de-cline in Tsuga around 75 cm, and Fagus, Acer saccha-

rum, Quercus and Betula rise in relative abundance.

Tsuga increases again in relative abundance above

55 cm, when Quercus begins to decline. Tsuga, Fagus,

and especially Acer saccharum show a peak in relative

abundance in the sample from 44 cm while the abun-

dance of Ulmus (elm) is comparatively low. The rela-

tive abundance of Poaceae and Ambrosia shows twodistinct peaks, rising initially to over 7% and 8% (re-

spectively) of the pollen sum at 44 cm before rising

again to a peak value of nearly 16% and 10% (respec-

tively) at 14 cm in the core. Large numbers of phyto-

liths (highly resistant silica structures found in and

around plant cell walls; Lu & Liu 2003; Morris et al.

2009, 2010) representing Cerealia (Plate 1), are associ-

ated with the earlier grass peak and with the base ofthe second Ambrosia rise around 30 cm (Figures 3 and

4). Fagus and Acer saccharum remain common until

24 cm and there is a resurgence of Pinus strobus in the

upper 24 cm of the core, accompanied by a rise in

Picea. Other non-arboreal (herb) pollen (NAP), domi-

nantly Artemisia (sage) and Chenopodiinae (cheno-

pods), are found throughout the core, but are most

abundantly associated with the grass peaks. The pollenof emergent and submerged aquatic taxa (excluding

grasses, which cannot be discriminated), such as

Cyperaceae, Typha latifolia (common cattail), Typha

angustifolia (narrow cattail), Potamogeton (pondweed), and Nymphaea (water lily), are always rare, but

do seem to correlate with NAP (Figure 3). Relative

abundance and concentrations of all pollen taxa were

high throughout the core; however, pollen concentra-

tions were relatively low (< 90,000 grains/cm3) in the

NAP-rich sediments (upper 20 cm), which may be due

to less sediment compaction in the upper layer.

The most abundant NPP identified were algal, in-cluding conjugated green algae (Division Charophyta

Class Zygnematophyceae of the Order Desmidiales),

such as Staurastrum spp., Cosmarium spp., and Euas-

trum spp., colonial green algae (Division Chlorophyta,

Class Chlorophyceae, Order Chlorococcales) such as

Pediastrum (Plate 1, figure 4), and dinoflagellate cysts

assigned to the genera Peridinium and Parvodinium

(Division Dinoflagellata, Class Dinophyceae, OrderPeridiniales) (Plate 1, figure 5). Protozoans were also

seen in palynological preparations comprising

thecamoebians/testate amoebae (Phylum Amoebozoa,

Class Lobosa, Order Arcellinida), primarily Centro-

pyxis spp., and the ciliate Codonella cratera (Phylum

Ciliophora, Class Spirotrichea, Order Tintinnida)

(Figures 4 and 6; Plate 1).

The three dominant desmid genera show similartrends, with highest abundances in the lower part of

the core (below the lower grass peak �44 cm) and low-

est concentrations between 54 cm and 14 cm (Figure 4).

Figure 3. Percent relative abundance of pollen taxa scaled by depth from a sediment core taken within Cook’s Bay, LakeSimcoe, Ontario, Canada. Local pollen assemblage zones were determined independently using CONISS and the three majorzones are highlighted with chronological indications. Arboreal pollen is represented starting at the left of the plots followed byherbaceous pollen to the right. Relative abundance for total NAP and aquatics are represented at the end followed by a concen-tration of total number of grains. McAndrews (1994) pollen zones 3 and 4 are illustrated on the right-hand side and delineated bya solid horizontal black line. The stipple highlights the sparse organic and silicate-rich samples (see Figure 2) with unusually highabundances of Poaceae and Ambrosia pollen, thought to represent the deliberate draining of the Holland Marshes to createpolders in the 1920s and 1930s. The dashed line highlights the second rapid rise in Ambrosia and other herbaceous taxa duringthe post-WWII population boom in the watershed.

236 D.C. Danesh et al.

Dow

nloa

ded

by [

Ein

dhov

en T

echn

ical

Uni

vers

ity]

at 1

5:31

19

Oct

ober

201

4

Page 8: Non-pollen palynomorphs as indicators of water quality in Lake Simcoe, Ontario, Canada

Plate 1. Figures represent specimens found in palynologically processed samples from a sediment core taken within Cook’s Bay,Lake Simcoe, Canada. Scale bars represent 10 mm, with the exception of figure 4 in which the scale bar represents 20 mm.Figure 1. Euastrum sp. (95–96 cm) mid-view. Figure 2. Cosmarium sp. (95–96 cm) high view. Figure 3. Staurastrum sp.(95–96 cm) mid-view. Figure 4. Centropyxis constricta (95–96 cm) mid-view. Figure 5. Pediastrum (84–85 cm) mid-view. Figure6. Codonella cratera (24–25 cm) mid-view. Figure 7. Saccharum sp. (30–31 cm) mid-view. Figure 8. Parvodinium inconspicuum(14–15 cm). Figure 9. Peridinium wisconsinensis (14–15 cm). Figure 10. Peridinium willei (29–30 cm). Figure 11. Peridinium volzii(29–30 cm).

Palynology 237

Dow

nloa

ded

by [

Ein

dhov

en T

echn

ical

Uni

vers

ity]

at 1

5:31

19

Oct

ober

201

4

Page 9: Non-pollen palynomorphs as indicators of water quality in Lake Simcoe, Ontario, Canada

Figure 4. NPP concentrations scaled by depth from a sediment core taken within Cook’s Bay, Lake Simcoe, Ontario, Canada.Local NPP zones were determined independently using CONISS. Independent zonation of NPP assemblages shows the same ma-jor zones seen in Figure 3. Desmids are found in high concentrations in pollen zone 3 (delineated by a solid horizontal black line).The stipple highlights a sharp decrease in desmids and the presence of phytoliths, which record the possible draining of theHolland Marshes to create polders in the 1920s and 1930s. The dashed line highlights the rapid rise in Ambrosia (see Figure 3)and other NPP during the post-WWII population boom in the watershed.

Figure 5. Dinoflagellate cyst concentrations scaled by depth from a sediment core taken within Cook’s Bay, Lake Simcoe,Ontario, Canada. Local dinoflagellate cyst zones were determined independently using CONISS. Independent zonation showsthe same three major zones seen in Figure 3. The first appearance of dinoflagellate cysts Peridinium volzii, P. willei and P. wiscon-sinense occurs at the base of pollen zone 4, which is associated with the first appearance of Ambrosia, suggesting eutrophication,followed by the appearance of Parvodinium inconspicuum. The stipple highlights a sharp increase in Peridinium spp., corroborat-ing the nutrient flux to Cook’s Bay associated with the creation of polders in the 1920s and 1930s. The dashed line marks a secondincrease in dinoflagellate cyst concentrations attributed to the post-WWII population boom in the watershed.

238 D.C. Danesh et al.

Dow

nloa

ded

by [

Ein

dhov

en T

echn

ical

Uni

vers

ity]

at 1

5:31

19

Oct

ober

201

4

Page 10: Non-pollen palynomorphs as indicators of water quality in Lake Simcoe, Ontario, Canada

Desmids made up almost half the NPP sum in the sam-ples from 95 cm through 54 cm (mean 43.1%, range

19.8–57%). The total concentration of desmids remains

low from 50 cm through 24 cm (mean 9.7%, range 6.1–

19.8%), but Staurastrum spp. and Cosmarium spp. re-

cover slightly in the upper 20 cm (mean 17.9%, range

10–25%). Species-level identifications of desmids were

not attempted in this study, but initial observations

suggest that different species of Staurastrum spp. andCosmarium spp. are present in the upper and lower

parts of the core.

Pediastrum spp. show the opposite trend: an ini-

tial rise at 70 cm, but the highest concentrations in

the upper 35 cm, peaking at 24.2% of NPP in the

sample from 14 cm (Figure 4). The first dinoflagel-

late cysts were seen in the sample from 75 cm where

cysts of Peridinium wisconsinense were the most com-mon from 70 cm to 50 cm. Peak concentrations (>11,000 cysts/cm3) were found in the samples from

45 cm and 40 cm and in the upper 20 cm (Figure 5),

where cysts of Peridinium volzii and Peridinium willei

increased sharply in abundance at 45 cm and remained

dominant until the uppermost sample, when a resur-

gence in Peridinium wisconsinense was noted. Cysts of

Parvodinium inconspicuum are present in the upper60 cm of the core, peaking at 45 cm and then again at

20 cm, but these cysts are never as abundant as those

of the Peridinium spp.

Various species of thecamoebian (testate amoeba)occur in most palynological preparations throughout

the core, peaking in the sample from 20 cm, where they

made up > 17% of the NPP (Figure 6). Organic-rich

tests of the genus Centropyxis were most commonly

found, although a few specimens of Arcella, Cucurbi-

tella, and even of the coarsely agglutinated Difflugia,

were also identified. There is surprisingly little similar-

ity in thecamoebian and ciliate protozoan assemblagesidentified in traditionally processed (washed/sieved)

microfossil samples (bar graphs in Figure 6) and in pal-

ynologically processed (palynological slides) microfos-

sil samples (NPP-shadow diagrams in Figure 6).

Thecamoebian tests were most abundant in washed/

sieved samples from 40 cm and 30 cm, where virtually

no tests were preserved in palynological slides, whereas

high thecamoebian concentrations were estimatedbased on specimens seen in palynological slides from

60 cm, where fewer than 20 tests were found in each

cm3 of sediment washed/sieved. The tintinnid ciliate

Codonella cratera was only identified (palynological

slides) in the upper 24 cm of the core, peaking at the

surface, where the taxon made up �5% of the NPP,

but they were most abundant in washed/sieved samples

from 30 cm and 40 cm. The scarcity of difflugiid theca-moebians, which are relatively coarsely agglutinated

and thus likely to be destroyed by palynological proc-

essing, in the NPP counts was not surprising, but the

Figure 6. Thecamoebian concentrations scaled by depth from a sediment core taken within Cook’s Bay, Lake Simcoe, Ontario,Canada. The chronology representing the three major events in the basin are taken from Figure 3. There is surprisingly little similar-ity in thecamoebian and ciliate protozoan assemblages identified in washed microfossil samples (represented by the bar graphs) andin palynological preparations (represented by the two shadow graphs at the end of the figure). Interpretations based on the tradition-ally washed samples (retaining the > 45-mm fraction) show an increase in thecamoebian abundance and diversity up-core, peakingin the sample from 30 cm and 40 cm, while there are virtually none recorded in the palynological preparations of the same intervals.

Palynology 239

Dow

nloa

ded

by [

Ein

dhov

en T

echn

ical

Uni

vers

ity]

at 1

5:31

19

Oct

ober

201

4

Page 11: Non-pollen palynomorphs as indicators of water quality in Lake Simcoe, Ontario, Canada

lack of correlation between the more organic tests of

Centropyxis and Arcella spp. in the two datasets was

disconcerting and difficult to explain. As a result, we

did not pursue the comparison with conventionallyprocessed thecamoebians beyond the six exploratory

samples, although further study should be undertaken.

3.4. Conventional thecamoebian analysis

Low thecamoebian populations made up exclusively ofCentropyxis aculeata were identified in the sample

from 95 cm and, although the population size in-

creased at 80 cm, the diversity remained very low, with

Centropyxis aculeata and Centropyxis constricta almost

completely dominating the assemblage (bar graphs in

Figure 6). The number and diversity of thecamoebian

tests in washed samples increased sharply above the

pollen zone 4 boundary at 75 cm, and by 30 cm difflu-giid thecamoebians (especially Cucurbitella tricuspis)

dominated the assemblage. Other common difflugiid

taxa present in the upper 30 cm were Difflugia corona,

Difflugia oblonga, and Difflugia protaeiformis. Tests of

the ciliate tintinnid Codonella cratera were present in

the upper three washed samples examined (40 cm,

30 cm and 1 cm), and they dominated the surface sam-

ple, where the most common thecamoebian was Difflu-

gia protaeiformis.

4. Discussion

Over the past two centuries Lake Simcoe has been im-

pacted by anthropogenic activities, which have been

recorded by shifts in fossil pollen and NPP assemblages.

In Cook’s Bay, palynological analysis shows threedistinct events in the smaller basin of Lake Simcoe.

These are: (1) regional vegetation shifts coinciding with

early European settlement and subsequent land clear-

ance throughout the nineteenth century, (2) abrupt

shifts in arboreal pollen, NAP, and NPP assemblages

indicating canal construction and damming of the East

and West Holland marshes by the Dutch settlers during

the early twentieth century, and (3) pronounced shifts inregional vegetation and NPP assemblages, along with

increases in metal concentrations consistent with the

five-fold population boom and extensive urbanization

around the Cook’s Bay basin after WWII.

4.1. Initial European colonisation

Pollen zone 3 (Figure 3) is indicative of pre-settlement

or background environments and is initially dominated

by Tsuga and Fagus but ends with a Tsuga minimumand rise in Pinus strobus. This has been interpreted as a

regional event recording the Little Ice Age (Campbell

& McAndrews 1991; Yu & McAndrews 1994; Munoz

& Gajewski 2010). Desmids indicative of oligotrophic

lacustrine environments, such as Staurastrum spp.,

Cosmarium spp., and Euastrum spp. (Wehr & Sheath

2003) were found in highest abundances in pollen zone3 (Figure 4) indicating a low nutrient environment.

Other algae generally associated with eutrophic condi-

tions (Shubert 2003; McCarthy et al. 2011), such as

most Pediastrum spp. and Peridinium spp., were rare at

the bottom of the core, confirming relatively low nutri-

ent levels (Figures 4 and 5). The relative abundance of

Poaceae pollen is slightly higher than expected during

this period by comparison with other sites in southernOntario (McAndrews 1988, 1994) and may be associ-

ated with the establishment of York County in the

Lake Simcoe area by Governor Simcoe in the 1790s,

which preceded the arrival of Europeans (LSRCA

2000).

The base of pollen zone 4 results from the increase in

relative abundance of Ambrosia and other NAP record-

ing land disturbance (McAndrews 1994) approximatingthe mid-nineteenth century. This coincides with the first

population boom in the 1850s when the Ontario, Sim-

coe and Huron Railway was completed and the com-

bined population of Newmarket and Aurora (in the

East and West Holland Marsh subwatersheds) rose

from �600 in 1841 to �3350 in 1871 (LSRCA 2000;

Eimers et al. 2005; Hawryshyn et al. 2012). The de-

crease in abundance and diversity of desmids towardsthe top of pollen zone 3 and base of pollen zone 4

(Figure 4) also appears to record early European settle-

ment around Cook’s Bay around the 1850s.

The first appearance of dinoflagellate cysts associ-

ated with the first appearance of Ambrosia suggests

that eutrophication began with the first land clearing

of the region (�1850s). The abundance of Peridinium

wisconsinense in the lower part of the core (nineteenthcentury) is consistent with the observations of Burden

et al. (1986) and McCarthy et al. (2011) that this spe-

cies was common prior to settlement of the Severn

Sound (Penetanguishene-Midland) region by both ab-

original Wendat and Europeans. McCarthy and

Krueger (forthcoming) also found this species more

common prior to the Iroquois settlement of Crawford

Lake. The very low thecamoebian abundances and Cen-

tropyxis-dominated assemblages are consistent with oli-

gotrophic conditions (McCarthy et al. 1995).

The more or less steady increase in TP in marly

organic muds of up-core records indicates increased

nutrient flux to Cook’s Bay since the initial Euro-

pean Settlement of the region (Figure 2). The grad-

ual increase in nitrite and nitrate concentrations

tracks the TP increase below 44 cm in the core, butmetal concentrations remain negligible, recording

little to no industrial activity in the Cook’s Bay re-

gion during this time.

240 D.C. Danesh et al.

Dow

nloa

ded

by [

Ein

dhov

en T

echn

ical

Uni

vers

ity]

at 1

5:31

19

Oct

ober

201

4

Page 12: Non-pollen palynomorphs as indicators of water quality in Lake Simcoe, Ontario, Canada

4.2. Dutch settlement and the draining of the HollandMarshes

The relative abundance of herbaceous taxa is low from

the base of pollen zone 4 (Figure 3) until the sharp rise

in Ambrosia and Poaceae (between �55 cm and 35 cm)

that is consistent with extensive deforestation and land

clearing for agriculture and settlement. During the

1920s and 1930s, Dutch settlers extensively drained

and then dammed the East and West Holland Marshes

in order to use the fertile soil to produce rich agricul-tural land (Johnson & Nicholls 1989; LSEMS 1994;

LSRCA 2000). The creation of these polders during

the early–mid-1900s is confirmed with the presence of

phytoliths (Figure 4) representative of Cerealia during

this time period due to the release of phytoliths into

the sediment through the decay of dead plants (Rovner

1971; Fredlund 2001) and its close association with

land clearing (Morris et al. 2010). The sharp increasein the abundances of Peridinium willei and Peridinium

volzii (Figure 5), represent dinoflagellate cysts associ-

ated with cultural eutrophication in other southern

Ontario lakes (Burden et al. 1986; McCarthy et al.

2011; McCarthy & Krueger forthcoming). This is also

associated with the high influx of phytoliths postulated

to have resulted from the development of polders in

the Holland Marshes and the ensuing flushing ofnutrients into Cook’s Bay during the early twentieth

century. This chronology of events is consistent with a

rise in Poaceae pollen as well as Ambrosia and other

NAP, and with the anomalous decline in Ulmus at

�44 cm (Figure 3) that records the onset of Dutch Elm

Disease. The reported cause, due to infected logs

brought over by the Dutch (French et al. 1980; Hubbes

1999) is also consistent with the Dutch settlement ofthe Holland Marshes. The increase in relative abun-

dance of organic and silicate sediment (Figure 2) fur-

ther supports the creation of polders as it indicates

increased nutrient input and mineral sedimentation

into Cook’s Bay, which is consistent with soil erosion

expected from land clearing and agricultural activities

(McAndrews 1988; LSRCA 2000). The overall decline

in the abundance of desmids, most notably Cosmarium

spp., also occurred during this time (Figure 4) and is

consistent with a shift from oxygen-rich, oligotrophic

waters to oxygen-poor, mesotrophic/eutrophic waters

in Cook’s Bay (Wehr & Sheath 2003). This depletion in

oxygen resources was caused by the excessive amounts

of nutrients, plant debris, and sediments consistent

with soil erosion that would be expected from draining

a wetland into a lake basin in order to create fertile ag-ricultural land (McAndrews 1988; LSEMS 1994;

LSRCA 2000).

The shift from high NO2 to high NO3 associated

with the grass and phytolith peak from �50 to 40 cm

(Figure 2) records a sudden decline in DO, probably in-

duced by high biological oxygen demand (BOD) as

large quantities of organic matter were drained into the

basin to create the polders. This is consistent with thepeak in organic matter measured by LOI. This increase

in BOD is one of the major causes in the decrease in

DO available to the aquatic organisms found within

Cook’s Bay (Wetzel 2001; LSEMS 2008; LSSAC 2008)

and can be considered the beginning of the deteriora-

tion of ecosystem health in Cook’s Bay.

4.3. Post-World War II urbanization andindustrialization

The rapid rise in Ambrosia between 30 cm and 24 cm,

immediately following a second sharp rise in Poaceae

and phytolith abundance (Figures 3 and 4), is attributed

to the five-fold post-WWII population boom (when the

combined population rose to �32,550 in 1971 from

�6750 in 1941) and the increasing urbanization and in-dustrialization, especially in Newmarket and Aurora,

both located in the East Holland Marsh subwatershed,

in the mid to late 1900s (LSRCA 2000). Concentrations

of desmids remain low, while abundant Pediastrum,

Peridinium, thecamoebians, and Codonella cratera in pal-

ynological preparations from the upper 24 cm (Figures 4

and 5) of the core from Cook’s Bay indicate a shift to-

wards eutrophic environments. Barbieri and Orlandi(1989) noted that dense populations of Codonella were

found in poorly oxygenated, muddy bottom waters in a

eutrophic reservoir in Brazil. The fact that Codonella cra-

tera is only present in the top 24 cm of the Cook’s Bay

core may confirm substantial decreases in dissolved oxy-

gen levels in Cook’s Bay over the last few decades

(LSRCA 2007, 2009). The sharp increase in Ambrosia,

associated with rapid urbanization following WWII,also saw a second increase in dinoflagellate cyst abun-

dance, with Peridinium volzii and Peridinium willei peak-

ing from 15 cm to 6 cm. Interpretations based on the

traditionally washed microfossil samples (retaining the

> 45-mm fraction) show an increase in thecamoebian

abundance and diversity up-core, peaking in the sample

from 30 cm, where Cucurbitella tricuspis strongly domi-

nates, recording eutrophic conditions (Figure 6).The increase in TP and NO3 concentrations and in

metals such as zinc, chromium, copper, lead, nickel, ar-

senic and cadmium to levels above what is considered

desirable records the addition of sewage treatment

plants and septic systems, and increased construction

of roads, dwellings, and industrial plants, all associated

with this post-war boom. The eutrophication of Lake

Simcoe, and Cook’s Bay in particular, has been con-firmed in recent studies carried out by the Lake Simcoe

Region Conservation Authority, and has impacted the

cold-water fishery in such a way that every year the fish

Palynology 241

Dow

nloa

ded

by [

Ein

dhov

en T

echn

ical

Uni

vers

ity]

at 1

5:31

19

Oct

ober

201

4

Page 13: Non-pollen palynomorphs as indicators of water quality in Lake Simcoe, Ontario, Canada

populations need to be stocked (LSRCA 2007; LSEMS

2008; LSSAC 2008; LSPP 2009; LSRCA 2009; Young

et al. 2010).

The resurgence of Cosmarium spp. and Staurastrum

spp. toward the top of the core, together with a decline

in Pediastrum spp., Cucurbitella tricuspis, and dinofla-

gellate cyst taxa relative to Peridinium wisconsinense as

well as in chromium, lead, and arsenic concentrations,

is attributed to P abatement programs over the last sev-

eral decades, and may record a slight recent improve-

ment in ecosystem health. This is consistent with the

findings of Eimers et al. (2005), Winter et al. (2007)and Hawryshyn et al. (2012) describing a minor in-

crease in ecosystem health in the Lake Simcoe water-

shed. The decline in Cucurbitella tricuspis at the top of

the core suggests a decrease in nutrient availability, but

the abundance of Difflugia protaeiformis as well as the

ciliate Codonella cratera record continued heavy metal

pollution and low DO levels (Moore 1977; Barbieri &

Orlandi 1989; Patterson & Kumar 2002; Reinhardtet al. 2005). Species-level investigations of desmids and

Pediastrum spp. may clarify this, since a few species of

these dominantly oligotrophic to mesotrophic algae

are found in eutrophic environments.

5. Conclusions

The distribution of NPP shows a clear correlation withchemical proxies of water quality and metal concentra-

tions measured in a core from Cook’s Bay, illustrating

their potential in paleoenvironmental studies. Species-

level identifications of desmids and Pediastrum spp.

should be attempted in future studies in order to

improve paleoenvironmental reconstructions. Pollen

chronology allowed the microfossil and chemical data

to be compared with the well-documented historical re-cord in the Cook’s Bay region, and three significant an-

thropogenic events in the watershed are evident. These

are: (1) initial European colonization in the late eigh-

teenth century, marked by initial low abundances of

Ambrosia and a steady increase in nutrients (TP, NO2);

(2) the inception of intensive agriculture during the

1920s and 1930s when Dutch settlers drained the East

and West Holland Marshes to create polders (markedby the influx of abundant Poaceae and phytoliths to

Cook’s Bay, and a shift from high nitrate to high nitrite

concentrations in response to the sudden increase in

BOD); and (3) a population boom and urbanization

and industrialization in the Cook’s Bay watershed fol-

lowing the WWII (marked by an initial peak in phyto-

liths and a sharp rise in Ambrosia and other NAP, and

high concentrations of heavy metals like zinc, chro-mium, lead, copper, arsenic, nickel and cadmium).

Desmids (particularly Cosmarium spp.) decreased

sharply in abundance in response to initial European

colonization. Whereas other algae, like Peridinium

spp., appear to have responded positively to the nutri-

ent influx, as did thecamoebians, particularly difflugiid

taxa like Cucurbitella tricuspis, and various species ofDifflugia. Difflugiid thecamoebian taxa are typically

not well represented in palynological preparations, and

were primarily found in washed samples, but there was

surprisingly little agreement even between the centro-

pyxid thecamoebians, which have a higher preservation

potential due to their organic-rich tests. A sudden in-

flux of unknown NPP dominates the phytolith-rich

sediments attributed to the creation of polders duringthe early twentieth century and the base of the sharp

Ambrosia rise attributed to rapid urbanization in the

Cook’s Bay watershed (notably the towns of Newmar-

ket and Aurora). Abundant Pediastrum spp., Peridi-

nium willei, Peridinium volzii, Difflugia protaeiformis,

and Codonella cratera in Ambrosia-rich sediments de-

posited since WWII are consistent with the docu-

mented eutrophication and low DO that have stressedthe cold-water sport fishery that is important to the lo-

cal economy, as well as heavy metal contamination of

Cook’s Bay.

Acknowledgements

We thank EnviroFix Corporation for generously collecting thecore used for this research and E3 Laboratories for providingus with analytical results from the Cook’s Bay core sediments.The assistance of Mike Lozon at Brock University with draft-ing is also greatly appreciated. We would like to thankDr. Brian F. Cumming for providing valuable insights and wewould also like to acknowledge the encouraging comments ofDr. Bas van Geel on an earlier draft of this manuscript.

Author biographies

DONYA C. DANESH is currently anM.Sc. student under the supervision ofDr Brian Cumming at the Paleoecologi-cal, Environmental Assessment and Re-search Laboratory (PEARL) at Queen’sUniversity, Canada. She is researchingthe relationships between climate-relatedchanges in vegetation, water quality andfire history in the boreal region of north-

west Ontario using palaeoenvironmental indicators. Donya re-ceived her undergraduate degree in Environmental Sciencesfrom Brock University. She completed her undergraduate the-sis under the supervision of Dr Francine McCarthy, whosparked Donya’s interest in non-pollen palynomorphs. Donyahas a multi-disciplinary background in environmental engi-neering, environmental sciences and biological sciences. Herprevious work experience has contributed to her broad per-spective on water-related issues. Donya has always had an in-terest in organising conferences, and is currently the co-Chairof the Organizing Committee for the First Annual NationalWater Research Centre Student Conference at Queen’sUniversity.

242 D.C. Danesh et al.

Dow

nloa

ded

by [

Ein

dhov

en T

echn

ical

Uni

vers

ity]

at 1

5:31

19

Oct

ober

201

4

Page 14: Non-pollen palynomorphs as indicators of water quality in Lake Simcoe, Ontario, Canada

FRANCINE M.G. MCARTHY is a Professor of EarthSciences at Brock University in Canada’s Niagara Region.She and her students work with a number of microfossilgroups including pollen, dinoflagellate cysts, othernon-pollen palynomorphs (NPP), planktonic foraminiferaand thecamoebians, in marine, lacustrine and wetland sedi-ments of Miocene to Holocene age. Most of this research hasfocused on sea-levels, lake-levels, and palaeoclimates, but hasalso included providing a palaeoenvironmental context forarchaeological occupations, tracing pollutant migration inthe Great Lakes, and evaluating environmental remediationoptions in the Oil Sands of Alberta. Current projects includestudies of the Miocene palynology of the New Jersey shallowshelf to evaluate the impact of eustasy on the stratigraphicalrecord (International Ocean Drilling Program Expedition313), and of the use of freshwater dinoflagellate cysts andother NPP in studies of cultural eutrophication.

OLENA VOLIK completed her under-graduate and graduate studies at Terno-pil National Pedagogical University inUkraine, and was awarded her MEd in2003. After receiving a Ph.D. in Geogra-phy specialising in palaeogeography andgeomorphology in 2006, Olena took up aposition of Assistant Professor at the De-partment of Physical Geography at Ter-

nopil National Pedagogical University. Her research includedthe palaeoenvironmental conditions for travertine formation inwestern Ukraine, caves and other karst features of the Podillyaregion, Ukraine, and the management of nature reserves. In2011, Olena joined Dr Francine McCarthy’s group at BrockUniversity, Canada, and since then she has been pursuing herM.Sc. in Earth Sciences. Her main research focus is onnon-pollen palynomorphs and thecamoebians as proxies forenvironmental and anthropogenic changes.

MATEA DRLJPAN completed her un-dergraduate studies at Brock Universitywith a degree in Environmental Geo-sciences. She completed her undergrad-uate thesis with Dr Francine McCarthyin freshwater and marine palynofaciesstudies. Matea is currently working onher M.Sc. in Earth Sciences at BrockUniversity with Dr Francine McCarthy.

She is studying the micropalaeontology and palynology ofSluice Pond, Massachusetts, a lake affected by long-term in-dustrial activities.

References

Barbieri SM, Orlandi MJLG. 1989. Ecological studies on theplanktonic protozoa of a eutrophic reservoir (RioGrande Reservoir – Brazil). Hydrobiologia 183:1–10.

Beyens L, Meisterfeld R. 2001. Protozoa: testate amoebae.In: Smol JP, Birks JB, Last WM, editors. Tracking envi-ronmental change Using Lake Sediments. Volume 3: Ter-restrial, Algal, and Siliceous Indicators. Dordrecht (TheNetherlands): Kluwer Academic Publishers.

Bradshaw EG, Anderson NJ, Jensen JP, Jeppesen E. 2002.Phosphorus dynamics in Danish lake and the

implications of diatom ecology and palaeoecology.Freshwater Biol. 47:1963–1975.

Burden ET, McAndrews JH, Norris G. 1986. Palynology ofIndian and European forest clearance and farming inlake sediment cores from Awenda Provincial Park,Ontario. Can J Earth Sci. 23:43–54.

Campbell ID, McAndrews JH. 1991. Cluster analysis of lateHolocene pollen trends in Ontario. Can J Bot. 69:1719–1730.

Canadian Council of Ministers of the Environment. 2007.Canadian water quality guidelines for the protection ofaquatic life: Summary table. Updated December, 2007,in Canadian Environmental Quality Guidelines, 1999.Winnipeg (MA): Canadian Council of Ministers of theEnvironment.

Carlson RE. 1977. A trophic state index for lakes. LimnolOceanogr. 22:361–368.

Coesel PFM, Meesters KJ. 2007. Desmids of the lowlands:mesotaeniaceae and desmidiaceae of the EuropeanLowlands. Amsterdam (The Netherlands): KNNVPublishing.

Correll DL. 1998. The role of phosphorus in the eutrophica-tion of receiving waters: a review. J Environ Qual.27:261–266.

Dale B. 2009. Eutrophication signals in the sedimentaryrecord of dinoflagellate cysts in coastal waters. J Sea Res.61:103–113.

Detenbeck NE, Taylor DL, Lima A, Hagley C. 1996. Tempo-ral and spatial variability in water quality of wetlands inthe Minneapolis/St. Paul, MN metropolitan area: impli-cations for monitoring strategies and designs. EnvironMonit Assess. 40:11–40.

Dixit SS, Smol JP. 1994. Diatoms as indicators in the envi-ronmental monitoring and assessment program – surfacewaters. Environ Monit Assess. 31:275–306.

Eimers MC, Winter JG, Scheider WA, Watmough SA,Nicholls KH. 2005. Recent changes and patterns in thewater chemistry of Lake Simcoe. J Great Lakes Res.31:322–332.

Ekdahl EJ, Teranes JL, Guilderson TP, Turton CL, McAn-drews JH, Wittkop CA, Stoermer EC. 2004. Prehistoricalrecord of cultural eutrophication from Crawford Lake,Canada. Geology 32:745–748.

Ekdahl EJ, Teranes JL, Wittkop CA, Stoermer EF, ReavieED, Smol JP. 2007. Iroquoian and Canadian eutrophica-tion of Crawford Lake, Ontario, Canada: Ecosystem andlake resilience following two periods of cultural distur-bance. J Paleolimnol. 37:233–246.

Evans DO, Nicholls KH, Allen YC, McMurtry MJ. 1996.Historical land use, phosphorus loading, and loss of fishhabitat in Lake Simcoe, Canada. Can J Fish Aquat Sci.53 (SI1):194–218.

Faegri K, Iversen J. 1975 Textbook of pollen analysis. 3rd ed.Copenhagen (Denmark): Munksgaard.

Fredlund GG. 2001. Inferring vegetation history from phyto-liths. In: Egan D, Howell EA, editors. The historical ecol-ogy handbook: A restorationist’s guide to referenceecosystems. Waginshton (DC): Island Press; p. 335–362.

French DW, Ascerno ME, Stienstra WC. 1980. The Dutchelm disease. USA: Minnesota Extension Service, Univer-sity of Minnesota.

Ginn BK. 2011. Distribution and limnological drivers ofsubmerged aquatic plant communities in Lake Simcoe(Ontario, Canada); utility of macrophytes as

Palynology 243

Dow

nloa

ded

by [

Ein

dhov

en T

echn

ical

Uni

vers

ity]

at 1

5:31

19

Oct

ober

201

4

Page 15: Non-pollen palynomorphs as indicators of water quality in Lake Simcoe, Ontario, Canada

bioindicators of lake trophic status. J Great Lakes Res.37:83–89.

Gloterman HL, Clymo KE, Clymo RS. 1975. Physiologicallimnology: an approach to the physiology of lake ecosys-tems. Amsterdam (The Netherlands): Elsevier ScientificPublishing Company.

Haas JN, editor. 2010. Fresh insights into the palaeoecologi-cal and palaeoclimatological value of Quaternary non-pollen palynomorphs. Veg Hist Archaeobot. 19:5–6.

Hawryshyn J, Ruhland KM, Quinlan R, Smol JP. 2012.Long-term water quality changes in a multiple-stressorsystem: a diatom-based paleolimnological study ofLake Simcoe (Ontario, Canada). Can J Fish Aquat Sci.69:24–40.

Heiri O, Lotter AF, Lemcke G. 2001. Loss on ignition as amethod for estimating organic and carbonate contentin sediments: reproducibility and comparability ofresults. J Paleolimnol. 25:101–110.

Helm PA, Milne J, Hiriart-Baer V, Crozier P, Kolic T, LegaR, Chen T, MacPherson K, Gewurtz S, Winter J, MyersA, Marvin CH, Reiner EJ. 2011. Lake-wide distributionand depositional history of current- and past-use persis-tent organic pollutants in Lake Simcoe, Ontario, Canada.J Great Lakes Res. 37:132–141.

Hubbes M. 1999. The American elm and Dutch elm disease.Forest Chronicle 75:265–273.

Johnson MG, Nicholls KH. 1989. Temporal and spatial vari-ability in sediment and phosphorus loads to Lake Sim-coe, Ontario. J Great Lakes Res. 15:265–282.

Kilgour B, Clarkin C, Morton W, Baldwin R. 2008. Influenceof nutrients in water and sediments on the spatial distri-bution of benthos in Lake Simcoe. J Great Lakes Res.34:365–376.

Kireta AR, Reavie ED, Danz NP, Axler RP, Sgro GV, King-ston JC, Brown TN, Hollenhorst, T. 2007. Coastal geo-morphic and lake variability in the Laurentian GreatLakes: implications for a diatom-based monitoring tool.J Great Lakes Res. 33 (SI3):136–153.

Kramer A, Herzschuh U, Mischke S, Zhang C. 2010. LateQuaternary environmental history of the south-easternTibetan Plateau inferred from the Lake Naleng non-pollen palynomorph record. Veg His Archaeobot.19:453–468.

Kumar A, Dalby AP. 1998. Identification Key for Holocenelacustrine arcellacean (thecamoebian) taxa: Paleontolo-gia Electronica, v.1, Art. 1.14 Available from: http://palaeo-electronica.org/1998-1/dalby/ issue1.htm

[LSEMS] Lake Simcoe Environmental Management Strat-egy. 1994. Lower Holland river erosion control study,Technical Report No. Imp.A1. Lake Simcoe Conserva-tion Authority. ON (Canada): The Queen’s Printer.

[LSEMS] Lake Simcoe Environmental Management Strategy.2008. Lake Simcoe basin wide report. Lake Simcoe conser-vation authority. ON (Canada): The Queen’s Printer.

[LSPP] Lake Simcoe Protection Plan. 2009. Prepared and ap-proved under the Lake Simcoe Protection Act, 2008. ON(Canada): The Queen’s Printer.

[LSRCA] Lake Simcoe Region Conservation Authority.2000. State of the watershed report: East Holland riversubwatershed. Lake Simcoe Conservation Authority.ON (Canada): The Queen’s Printer.

[LSRCA] Lake Simcoe Region Conservation Authority.2007. Report on the phosphorus loads to Lake Simcoe2004 – 2007. Lake Simcoe Conservation Authority. ON(Canada): The Queen’s Printer.

[LSRCA] Lake Simcoe Region Conservation Authority.2009. Watershed Report Card: A Summary of Lake Sim-coe Watershed and Ecosystem Health. ON (Canada):The Queen’s Printer.

[LSSAC] Lake Simcoe Science Advisory Committee. 2008.Lake Simcoe and its watershed: a report to the Ministerof the Environment. ON (Canada): The Queen’s Printer.

Lu H, Liu K. 2003. Phytoliths of common grasses in thecoastal environments of southeastern USA. EstuarCoastal Shelf Sci. 58:587–600.

Maier RM, Pepper IL, Gerba CP. 2009. Environmental mi-crobiology. 2nd ed. San Diego (CA): Elsevier Inc.

McAndrews JH. 1981. Late Quaternary climate of Ontario:temperature trends from the fossil pollen record. In:Mahaney WC, editor. Quaternary paleoclimate. Norwich(UK): Geo Abstracts Ltd.; p. 319–333.

McAndrews JH. 1988. Human disturbance of North Ameri-can forests and grasslands: the fossil pollen record.In: Huntley B, Webb III T, editors. VegetationHistory. Utrecht (The Netherlands): Kluwer AcademicPublishers.

McAndrews JH. 1994. Pollen diagrams for southern Ontarioapplied to archaeology. In: MacDonald R, editor. Greatlakes archaeology and paleoecology: exploring interdisci-plinary initiatives for the Nineties. ON (Canada): Quater-nary Sciences Institute, University of Waterloo; p. 179–196.

McAndrews JH, Berti AA, Norris G. 1973. Key to theQuaternary pollen and spores of the Great Lakes Region.Royal Ontario Museum Publications in Life Sciences.Toronto (ON): UofT Press.

McAndrews JH, Boyko-Diakonow M. 1989. Pollen analysisof varved sediment at Crawford Lake, Ontario: evidenceof Indian and European farming. In: Fulton RJ, editor.Quaternary Geology of Canada and Greenland. Vol. 1.Ottawa: Geological Survey of Canada; p. 528–530.

McCarthy FMG, Collins ES, McAndrews JH, Kerr HA, ScottDB, Medioli FS. 1995. A comparison of postglacial arcel-lacean (“thecamoebian”) and pollen succession in AtlanticCanada, illustrating the potential of arcellaceans for paleo-climatic reconstruction. J Paleontol. 69:980–993.

McCarthy FMG, Mertens KN, Ellegaard M, Pospelova V,Ribeiro S, Vertauteren D. 2011. Resting cysts of freshwa-ter dinoflagellates in southeastern Georgian Bay (LakeHuron) as proxies of cultural eutrophication. Rev Palae-obot Palynol. 166:46–62.

McCarthy FMG, Tiffin SH, Sarvis AP, McAndrews JH,Blasco SM, 2012. Early Holocene brackish closed basinconditions in Georgian Bay: microfossil evidence. J Pale-olimnol. 47:429–445.

McCarthy FMG, Krueger AM. Forthcoming. Freshwaterdinoflagellates in paleolimnological studies: Peridiniumcysts as proxies of cultural eutrophication in the south-eastern Great Lakes region of Ontario, Canada. SpecialPublication, The Geological Society.

Medioli FS, Scott DB. 1983. Holocene Arcellacea (thecamoe-bians) from eastern Canada: Cushman Foundation forForaminiferal Research, Special PublicationNo. 21; p. 63.

Mertens KN, Verhoeven K, Verleye T, Louwye S, AmorimA, Ribeiro S, Deaf AS, Harding IC, De Schepper S,Gonz�alez C, et al. 2009. Determining the absolute abun-dance of dinoflagellate cysts in recent marine sediments:the Lycopodium marker-grain method put to the test.Rev Palaeobot Palynol. 157:238–252.

Moore JW. 1977. Some factors influencing the density of sub-arctic populations of Bosmina longirostris, Holopedium

244 D.C. Danesh et al.

Dow

nloa

ded

by [

Ein

dhov

en T

echn

ical

Uni

vers

ity]

at 1

5:31

19

Oct

ober

201

4

Page 16: Non-pollen palynomorphs as indicators of water quality in Lake Simcoe, Ontario, Canada

gibberum, Codonella cratera and Ceratium hirundinella.Hydrobiologia 56:199–207.

Morris LR, West NE, Baker FA, Van Migroet H, Ryel RJ.2009. Developing an approach for using the soilphytolith record to infer vegetation and disturbance re-gime changes over the past 200 years. Quatern Int.193:90–98.

Morris LR, West NE, Ryel RJ. 2010. Testing soil phytolithanalysis as a tool to understand vegetation change in thesagebrush steppe and pinyon-juniper woodlands of theGreat Basin Desert, USA. Holocene 20:267–709.

Mudie PJ, Marret F, Rochon A, Aksu AE. 2010. Non-pollenpalynomorphs in the black sea corridor. Veg His Archae-obot. 19:531–544.

Munoz SE, Gajewski K. 2010. Distinguishing prehistoric hu-man influence on late-Holocene forests in southernOntario, Canada. Holocene 20:967–981.

Newhouse AE, Schrodt F, Liang H, Maynard CA, PowellWA. 2007. Transgenic American elm shows reducedDutch elm disease symptoms and normal mycorrhizalcolonization. Plant Cell Report. 26:977–987.

[OMEO] Ontario Ministry of the Environment [Internet].2010. ON (Canada): Queen’s Printer for Ontario; [cited2011 April 8] Available from: http://www.ene.gov.on.ca/environment/en/local/lake_simcoe_protection/index.htm

[OMOE] Ontario Ministry of the Environment. 2010a. Impli-menting key elements of the Lake Simcoe ProtectionPlan. Toronto (ON): Land and Water Policy Branch.

[OMOE] Ontario Ministry of the Environment. 2010b. LakeSimcoe Protection. Toronto (ON): Queen’s Printer forOntario; [cited July 16, 2010]. Available from: http://www.ene.gov.on.ca/environment/en/local/lake_simcoe_protection/index.htm

Palmer ME, Winter JG, Young JD, Dillon PJ, Guilford SJ.2011. Introduction and summary of research on LakeSimcoe: research, monitoring, and restoration of a largelake and its watershed. J Great Lakes Res. 37:1–6.

Patterson RT, Kumar A. 2002. A review of current testaterhizopod (thecamoebian) research in Canada. Palaeogeo-graph Palaeoclimatol Palaeoecol. 180:225–251.

Petr T. 2000. Interactions between fish and aquatic macro-phytes in inland waters: a review. FOA Fisheries Techni-cal Paper No 396. Rome (Italy): FAO.

Ramstack JM, Fritz SC, Engstrom DR, Heiskary SA. 2003.The application of diatom-based transfer functions toevaluate regional water-quality trends in Minnesota since1970. J Paleolimnol. 29:79–94.

Reinhardt EG, Donato S, Little M, Findlay D, Krueger A,Clark C, Boyce J. 2005. Arcellacean (thecamoebian)evidence of land-use change and eutrophication inFrenchman’s Bay, Pickering, Ontario. Environ Geol.47:729–739.

Rice EW, Baird RB, Eaton AD, Clesceri LS, Bridgewater L.editors. 2012. Standard methods for the analysis of waterand wastewater. 22nd ed. Washington (DC): AmericanPublic Health Association.

Rovner I. 1971. Potential of opal phytoliths for use in paleo-ecological reconstruction. Quaternary Res. 1:343–359.

Schindler DW. 1977. Evolution of phosphorus limitation inlakes: natural mechanisms compensate for deficiencies ofnitrogen and carbon in eutrophied lakes. Science195:260–262.

Scott DB, Medioli FS, Schafer CT. 2001. Monitoring in coastalenvironments using foraminifera and thecamoebian indica-tors. Cambridge (UK): Cambridge University Press; p. 177.

Shubert E. 2003. Non-motile coccoid and colonial green al-gae, In: Freshwater Algae of North America: Ecologyand Classification. San Diego (CA): Elsevier Science.

Singer SN, Cheng CK, Scafe MG. 2003. The hydrogeology ofSouthern Ontario. ON, Canada: Environmental Monitor-ing and Reporting Branch, Ministry of the Environment.

Smith VK, Desvousges WH. 1986. Measuring water qualitybenefits. Boston (MA): Kluwer-Nijhoff Publishing.

Smith VH, Schindler DW. 2009. Eutrophication science: wheredo we go from here? Trend Ecology Evol. 24:201–207.

Solheim H, Eriksen R, Hietala AM. 2011. Dutch elm diseasehas currently a low incidence on wych elm in Norway.Forest Pathol. 41:182–188.

Stantec Consulting Inc. 2006. Benthic macro-invertebratessampling and analysis of Lake Simcoe. ON (Canada):Lake Simcoe Conservation Authority.

Statistics Canada: Focus on geography series, 2011 Census[Internet]. Ottawa (ON): Statistics Canada Catalogue no.98-310-XWE2011004, Analytical Products; [cited 2012Sep 5]. Available from: http://www12.statcan.gc.ca/census-recensement/2011/as-sa/fogs-spg/Facts-csd-eng.cfm?Lang = Eng&TAB = 1&GK = CSD&GC = 3519048

Stockmarr J. 1971. Tablets with spores used in absolute pol-len analysis. Pollen Spore. 13:615–621.

Torbick N, Hu F, Zhang J, Qi J, Zhang H, Becker B. 2008.Mapping chlorophyll-a concentrations in West Lake,China using Landsat 7 ETMþ. Int Assoc Great LakesRes. 34:559–565.

Turton CL, McAndrews JH. 2006. Rotifer loricas in secondmillennium sediment of Crawford Lake, Ontario, Can-ada. Rev Paleobot Palynol. 141:1–6.

van Geel B. 2006. ‘Quaternary non-pollen palynomorphs’ de-serve our attention. Rev Palaeobot Palynol. 141:vii–viii.

Wehr JD, Sheath,RG. 2003. Freshwater algae of NorthAmerica: Ecology and Classification. San Diego (CA):Elsevier Academic Press.

Wetzel RG. 2001. Limnology: Lakes and River Ecosystems.3rd ed. San Diego (CA): Elsevier Academic Press.

Winter JG, Dillon PJ, Futter MN, Nicholls KH, ScheiderWA, Scott LD. 2002. Total phosphorus budgets andnitrogen loads: Lake Simcoe, Ontario (1990 to 1998).J Great Lakes Res. 28:301–314.

Winter JG, Eimers MC, Dillon PJ, Scott LD, Scheider WA,Willox CC. 2007. Phosphorus inputs to Lake Simcoe from1990 to 2003: declines in tributary loads and observationson lake water quality. J Great Lakes Res. 33:381–396.

Winter JG, Young JD, Landre A, Stainsby E, Jarjanazi H.2011. Changes in phytoplankton community composi-tion of Lake Simcoe from 1980 to 2007 and relationshipswith multiple stressors. J Great Lakes Res. 37:63–71.

Yoder CO, Rankin ET. 1998. The role of biological indica-tors in a state water quality management process. Envi-ron Monit Assess. 51:61–88.

Young J, Landre A, Winter J, Jarjanazi H, Kingston J. 2010.Lake Simcoe water quality update. Environmental Moni-toring and Reporting Branch, Ministry of the Environ-ment. ON (Canada): The Queen’s Printer for Ontario.

Yu Z, McAndrews JH. 1994. Holocene water levels at RiceLake, Ontario, Canada: sediment, pollen and plant-macrofossil evidence. Holocene 4:141–152.

Palynology 245

Dow

nloa

ded

by [

Ein

dhov

en T

echn

ical

Uni

vers

ity]

at 1

5:31

19

Oct

ober

201

4