+ All Categories
Home > Documents > Journal of Great Lakes Research - University of...

Journal of Great Lakes Research - University of...

Date post: 27-Jun-2020
Category:
Upload: others
View: 1 times
Download: 0 times
Share this document with a friend
10
Anthropogenic inuences on the sedimentary geochemical record in western Lake Superior (1800present) Molly D. O'Beirne a, , Ladislaus J. Strzok a,b , Josef P. Werne a,c , Thomas C. Johnson a,d , Robert E. Hecky a,e a Large Lakes Observatory, University of Minnesota Duluth, Duluth, MN 55812, USA b Minnesota Pollution Control Agency, St. Paul, MN 55155, USA c Department of Geology and Planetary Science, University of Pittsburgh, Pittsburgh, PA 15260, USA d Department of Earth and Environmental Sciences, University of Minnesota Duluth, Duluth, MN 55812, USA e Department of Biology, University of Minnesota Duluth, Duluth, MN 55812, USA abstract article info Article history: Received 13 September 2013 Accepted 23 September 2014 Available online 6 December 2014 Communicated by Gerald Matisoff Index words: Sedimentary organic carbon Carbon isotopes n-Alkane biomarkers Lake Superior Productivity Anthropogenic impacts The sediments of western Lake Superior hold a record of environmental changes that have accompanied the set- tlement and urbanization of the surrounding watershed. Organic carbon concentrations are low (1.5%) with little variation in stable isotope composition (-26.5 ± 0.5) prior to 1900. Organic carbon and nitrogen concentra- tions begin to rise after 1900, as increased anthropogenic disturbance led to enhanced inputs of terrigenous matter as well as nutrients (i.e., nitrogen and phosphorus) from the watershed. An episode of enhanced aquatic productivity from 1900 to 1970 is recorded in the sediments by the 13 C-enrichment of bulk organic carbon as well as the observed correlation between the bulk and aquatic molecular δ 13 C records, coinciding with the major de- velopmental period of the Duluth-Superior harbor region. Decreasing organic carbon accumulation after 1925, prior to regulatory implementation of municipal discharges to the lake, is likely due to the construction of hydro- power dams along the St. Louis River and a decrease in forest harvest within the immediate watershed. Recentshort-lived decreases in the accumulation rate of organic matter can be attributed to the implementation and operation of water treatment plants, but the 13 C-enrichment observed in the last ~ decade remains enigmat- ic, though we hypothesize that it may be attributable to climate change impacts on primary production. © 2014 International Association for Great Lakes Research. Published by Elsevier B.V. All rights reserved. Introduction Lakes are intimately connected with their surrounding environment, making them sensitive indicators of perturbations within their water- sheds. Organic geochemical proxies preserved in lake sediments can pro- vide a history of ecosystem change (e.g., carbon cycling). These proxies enable us to gauge the historical impacts of regional development and subsequent regulations within lake systems while providing insight to guide possible management action to address any emerging issues. The stable carbon isotope composition of sedimentary organic car- bon (δ 13 C org ), a proxy for lacustrine productivity (Meyers, 1994), is often used to track the response of lakes to historic anthropogenic activ- ities (i.e., nutrient loading) and subsequent regulation. This approach is based on the premise that the δ 13 C of organic carbon produced photo- synthetically responds to the rate of primary productivity in the water column (Schelske and Hodell, 1991; Meyers, 2003). In general, increases in lacustrine productivity should lead to increases in the deposition of total organic carbon (TOC) and total nitrogen (TN) concentrations along with changes in their stable isotope compositions (δ 13 C org and δ 15 N, respectively). δ 13 C org and δ 15 N are typically 13 C- and 15 N-enriched (depleted) with increasing (decreasing) in-lake primary production (Mckenzie, 1985; Meyers, 1997, 2003). Thus, the bulk iso- tope compositions as well as accumulation rates of C org and TN can be used as a proxy for lacustrine productivity. One caveat to this general approach is that sedimentary C and N are derived from both terrigenous and aquatic sources, thus increases or de- creases cannot be unequivocally allocated to one source or the other. Similarly, the isotopic composition of sedimentary TOC is inuenced by both terrigenous and aquatic sources, making it useful to look at mo- lecular biomarkers (e.g., n-alkanes) to separate contributions from these sources. The two principal sources of hydrocarbons to lake sedi- ments are those of photosynthetic algae and bacteria, dominated by the odd-numbered short chain n-alkanes(n-C 17 + n-C 19 + n-C 21 ) (Cranwell et al., 1987; Giger et al., 1980) and those contributed from vascular land plants, which contain large proportions of the odd- numbered long chain n-alkanes(n-C 27 + n-C 29 + n-C 31 ) that are pro- duced in their epicuticular waxy coatings (Cranwell, 1981; Eglinton and Hamilton, 1963, 1967; Rieley et al., 1991). In order to circumvent the issue of source, the stable carbon isotope composition of n-alkanes (δ 13 C n-alkane ), when compared to the bulk δ 13 C org record, can provide an indication of the relative contributions of terrigenous and aquatic Journal of Great Lakes Research 41 (2015) 2029 Corresponding author at: Department of Geology and Planetary Science, University of Pittsburgh, 200 Space Research Coordination Center, 4107 O'Hara Street Pittsburgh, PA 15260-3332. Tel.: +1 412 624 8780. E-mail address: [email protected] (M.D. O'Beirne). http://dx.doi.org/10.1016/j.jglr.2014.11.005 0380-1330/© 2014 International Association for Great Lakes Research. Published by Elsevier B.V. All rights reserved. Contents lists available at ScienceDirect Journal of Great Lakes Research journal homepage: www.elsevier.com/locate/jglr
Transcript
Page 1: Journal of Great Lakes Research - University of Pittsburghjwerne/uploads/3/0/1/0/30101831/45.obeirneetal.jglr1… · Anthropogenic influences on the sedimentary geochemical record

Journal of Great Lakes Research 41 (2015) 20–29

Contents lists available at ScienceDirect

Journal of Great Lakes Research

j ourna l homepage: www.e lsev ie r .com/ locate / jg l r

Anthropogenic influences on the sedimentary geochemical record inwestern Lake Superior (1800–present)

Molly D. O'Beirne a,⁎, Ladislaus J. Strzok a,b, Josef P. Werne a,c, Thomas C. Johnson a,d, Robert E. Hecky a,e

a Large Lakes Observatory, University of Minnesota Duluth, Duluth, MN 55812, USAb Minnesota Pollution Control Agency, St. Paul, MN 55155, USAc Department of Geology and Planetary Science, University of Pittsburgh, Pittsburgh, PA 15260, USAd Department of Earth and Environmental Sciences, University of Minnesota Duluth, Duluth, MN 55812, USAe Department of Biology, University of Minnesota Duluth, Duluth, MN 55812, USA

⁎ Corresponding author at: Department of Geology andPittsburgh, 200 Space Research Coordination Center, 41015260-3332. Tel.: +1 412 624 8780.

E-mail address: [email protected] (M.D. O'Beirne).

http://dx.doi.org/10.1016/j.jglr.2014.11.0050380-1330/© 2014 International Association for Great Lak

a b s t r a c t

a r t i c l e i n f o

Article history:Received 13 September 2013Accepted 23 September 2014Available online 6 December 2014

Communicated by Gerald Matisoff

Index words:Sedimentary organic carbonCarbon isotopesn-Alkane biomarkersLake SuperiorProductivityAnthropogenic impacts

The sediments of western Lake Superior hold a record of environmental changes that have accompanied the set-tlement and urbanization of the surroundingwatershed. Organic carbon concentrations are low (1.5%)with littlevariation in stable isotope composition (−26.5 ± 0.5‰) prior to 1900. Organic carbon and nitrogen concentra-tions begin to rise after 1900, as increased anthropogenic disturbance led to enhanced inputs of terrigenousmatter as well as nutrients (i.e., nitrogen and phosphorus) from the watershed. An episode of enhanced aquaticproductivity from1900 to 1970 is recorded in the sediments by the 13C-enrichment of bulk organic carbon aswellas the observed correlation between the bulk and aquatic molecular δ13C records, coinciding with the major de-velopmental period of the Duluth-Superior harbor region. Decreasing organic carbon accumulation after 1925,prior to regulatory implementation ofmunicipal discharges to the lake, is likely due to the construction of hydro-power dams along the St. Louis River and a decrease in forest harvest within the immediate watershed.Recentshort-lived decreases in the accumulation rate of organic matter can be attributed to the implementationand operation of water treatment plants, but the 13C-enrichment observed in the last ~ decade remains enigmat-ic, though we hypothesize that it may be attributable to climate change impacts on primary production.

© 2014 International Association for Great Lakes Research. Published by Elsevier B.V. All rights reserved.

Introduction

Lakes are intimately connected with their surrounding environment,making them sensitive indicators of perturbations within their water-sheds. Organic geochemical proxies preserved in lake sediments can pro-vide a history of ecosystem change (e.g., carbon cycling). These proxiesenable us to gauge the historical impacts of regional development andsubsequent regulations within lake systems while providing insight toguide possible management action to address any emerging issues.

The stable carbon isotope composition of sedimentary organic car-bon (δ13Corg), a proxy for lacustrine productivity (Meyers, 1994), isoften used to track the response of lakes to historic anthropogenic activ-ities (i.e., nutrient loading) and subsequent regulation. This approach isbased on the premise that the δ13C of organic carbon produced photo-synthetically responds to the rate of primary productivity in the watercolumn (Schelske andHodell, 1991;Meyers, 2003). In general, increasesin lacustrine productivity should lead to increases in the deposition oftotal organic carbon (TOC) and total nitrogen (TN) concentrations

Planetary Science, University of7 O'Hara Street Pittsburgh, PA

es Research. Published by Elsevier B

along with changes in their stable isotope compositions (δ13Corg

and δ15N, respectively). δ13Corg and δ15N are typically 13C- and15N-enriched (depleted) with increasing (decreasing) in-lake primaryproduction (Mckenzie, 1985; Meyers, 1997, 2003). Thus, the bulk iso-tope compositions as well as accumulation rates of Corg and TN can beused as a proxy for lacustrine productivity.

One caveat to this general approach is that sedimentary C and N arederived fromboth terrigenous and aquatic sources, thus increases or de-creases cannot be unequivocally allocated to one source or the other.Similarly, the isotopic composition of sedimentary TOC is influencedby both terrigenous and aquatic sources, making it useful to look atmo-lecular biomarkers (e.g., n-alkanes) to separate contributions fromthese sources. The two principal sources of hydrocarbons to lake sedi-ments are those of photosynthetic algae and bacteria, dominated bythe odd-numbered short chain n-alkanes(n-C17 + n-C19 + n-C21)(Cranwell et al., 1987; Giger et al., 1980) and those contributed fromvascular land plants, which contain large proportions of the odd-numbered long chain n-alkanes(n-C27 + n-C29 + n-C31) that are pro-duced in their epicuticular waxy coatings (Cranwell, 1981; Eglintonand Hamilton, 1963, 1967; Rieley et al., 1991). In order to circumventthe issue of source, the stable carbon isotope composition of n-alkanes(δ13Cn-alkane), when compared to the bulk δ13Corg record, can providean indication of the relative contributions of terrigenous and aquatic

.V. All rights reserved.

Page 2: Journal of Great Lakes Research - University of Pittsburghjwerne/uploads/3/0/1/0/30101831/45.obeirneetal.jglr1… · Anthropogenic influences on the sedimentary geochemical record

21M.D. O'Beirne et al. / Journal of Great Lakes Research 41 (2015) 20–29

sourced organic carbon to the bulk δ13Corg record (Bourbonniere andMeyers, 1996; Meyers, 2003).

The Laurentian Great Lakes have experienced dramatic changes in re-sponse to anthropogenic loadings and management activities. Studies oforganic geochemical proxies preserved within sediment cores fromLakes Erie and Ontario have shown periods of increased lacustrine pro-ductivity in correspondence with watershed/regional development.Schelske and Hodell (1995) and more recently, Lu et al. (2010), foundthat increases in lacustrine productivity (increases in the δ13Corg) of LakeErie were related to historic changes in anthropogenic phosphorus load-ing during the development of the watershed and subsequent decreaseswere in response to the implementation of phosphorus abatement pro-grams in the mid-1970s. A study by Hodell and Schelske (1998) in LakeOntario shows similar correspondence between productivity and anthro-pogenic activity as inferred from the sedimentary geochemical record.One study from Lake Superior, measuring sedimentary biogenic silica(BSi), provides evidence of low-level eutrophication associated with thelate 1800s European habitation and subsequent development of the wa-tershed (Schelske et al., 2006). To date, these findings based on BSi inLake Superior have not been corroborated by other proxies.

Here, we report temporal trends in productivity over the last 150years, through the use of bulk organic and stable isotope compositionsof carbon and nitrogen as well as n-alkane molecular biomarker abun-dances and their stable carbon isotope compositions, as contained inthe sediments of the western arm of Lake Superior. We compare twocores from the same location taken 6 years apart which allow us to de-termine not only the reproducibility of the sedimentary record but alsoevaluate the effect of early sedimentary diagenesis on the preservationof biomarker signals through time.

Regional setting and background

Lying along boundary of the United States (US) and Canada, LakeSuperior (Fig. 1), the largest of the Laurentian Great Lakes and the

Fig. 1. (A)Mapof LaurentianGreat Lakes, (B)mapofwestern Lake Superiorwith core locations, Ito right, Scanlon Dam, Thomson Dam, and Fond du Lac Dam).

world's largest freshwater lake by surface area, is often regaled as themost “pristine” of the Great Lakes. This is due, in part, to the smallamount of surface runoff Lake Superior receives in relation to itsvolume, as well as to the fact that most of the 127,700 km2 watershed(considered small, in comparison to its surface area with a terrigenouscatchment area to lake area ratio of ~1.5:1) is sparsely populated andheavily forested with little agriculture. As a result, Lake Superior hasnot suffered from the high nutrient loadings or industrial pollution tothe same extent as other Great Lakes (e.g., Lake Erie; Bourbonniereand Meyers, 1996; Schelske et al., 2006).

Although Lake Superior is considered the least altered of the Lauren-tian Great Lakes, has been classified as ultra-oligotrophic andmostly re-tains its pristine water quality conditions (Munawar and Munawar,1978; Weiler, 1978), changes in its physical, biological and chemicalprocesses have been documented. Observations of rising nitrate con-centrations since the 1900s (Sterner et al., 2007; Weiler, 1978), aswell as increasing water temperatures over the past two centuries(Austin and Colman, 2007, 2008) and changes in primary productivity(Urban et al., 2005; Urban, 2007; Vollenweider et al., 1974) indicatethat anthropogenic activities can impact Lake Superior.

The highest population density within the Lake Superior watershedis located in the St. Louis River drainage basin, which is also the secondlargest inflow to Lake Superior. The St. Louis River basin has been heavi-ly impacted by anthropogenic changes and has been designated as oneof 43 areas of concern (having impaired beneficial uses due to pollution)within the Laurentian Great Lakes by the International Joint Commis-sion (IJC). The area of concern was later extended to include theDuluth-Superior Harbor region (Fig. 1).

The objective of this study is to define changes in the primary pro-ductivity of the western end of Lake Superior (Fig. 1) over the past150 years in order to ascertain what the important drivers of changemay have been and to assess the efficacy of regulatory action to addressthose changes. If anthropogenically induced changes in the trophic stateof western Lake Superior have occurred, we should be able to track

JC designatedArea of Concern (shaded area) and the St. Louis riverwith dam locations (left

Page 3: Journal of Great Lakes Research - University of Pittsburghjwerne/uploads/3/0/1/0/30101831/45.obeirneetal.jglr1… · Anthropogenic influences on the sedimentary geochemical record

22 M.D. O'Beirne et al. / Journal of Great Lakes Research 41 (2015) 20–29

these changes in the sedimentary geochemical record, as has been donesuccessfully in lakes Erie and Ontario (Hodell and Schelske, 1998; LuandMeyers, 2009; Lu et al., 2010; Schelske andHodell, 1995). Therefore,we hypothesized that increases in productivity should coincide withincreases in the δ13C of both bulk and molecular aquatic biomarkerproxies and such increases should be expected to occur within thetimeframe of regional development (1900–present). Conversely, de-creases in productivitywould be expected to be observed after regulato-ry actions were put in place in the mid-1970s.

Methods

Sediment cores

Two sediment multicores, designed to capture a well-preserved sed-iment water interface, were collected aboard the R/V Blue Heron fromthe westernmost arm of Lake Superior (Fig. 1), one in 2003 (BH03-1;46° 48.51′N, 91° 51.95′W; water depth 44 m) and one in 2009 (BH09-5; 46° 49.00′N, 91° 51.9′W;water depth 48m). Core siteswere identifiedusing a Knudsen 12 kHz Hi-Res and CHIRP seismic reflection profilersonboard the R/V Blue Heron. The coring sites were chosen based on evi-dence of uninterrupted stratigraphy indicating continuous sedimenta-tion characteristic of similar locations in western Lake Superior. Whilestorm wave activity can periodically disrupt sedimentation in the rela-tively shallow depths of these core sites, the sedimentation rate in thispart of Lake Superior is unusually high and we assume that it greatly ex-ceeds the occasional erosional effect of a passing storm. This assumptionis confirmed by 210Pb results, discussed below. Core BH09-5MC was se-lected to give an indication of the reproducibility of results, acting as asister location to core BH03-1MC as well as to evaluate possible diage-netic effects on the developing record and recent trends.

Initial core description and splitting were performed at the NationalLacustrine Core Repository (LacCore). Subsamples were taken fromeach core at 0.5 cm intervals to 10 cm and every 1 cm thereafter tothe base of the core. Sediment samples were subsequently frozen andfreeze-dried for geochemical analyses.

210Pb geochronology

A total of 24 sub-samples from core BH03-1, and 20 sub-samplesfrom BH09-5 were taken for 210Pb geochronology. Intervals at greatestsample depth were included to determine the background (supported)210Pb activity. Unsupported 210Pb activity is expected to be negligible atthese depths, where sediment age is likely to exceed 150 years (N6half-lives).

Bulk and stable isotopic analysis of carbon and nitrogen

Bulk sediment samples were analyzed for weight percent organic C(TOC) and total N (TN) concurrently with isotope ratio determinations.All concentrations were determined on acid fumigated (decarbonated)samples, eliminating variability in total C content caused by any carbon-atewithin samples. N2 and CO2 peak areas (Isodat v3.0)were convertedto weight percent compositions using response factors generated fromstandards of known composition (acetanilide, caffeine, B-2153, B-2159and urea), whichwere run between every ten samples. Analytical preci-sion, based on replicate standard runs for bulk measurements, was typ-ically better than±0.80% (1σ, n=24) for carbon andnitrogenweight %values and better than ±0.20‰ for carbon and ±0.25‰ for nitrogenisotope values. Reproducibility between duplicate samples was betterthan ±0.05% for both carbon and nitrogen weight % values; isotopicvalues of carbon and nitrogen between duplicate samples were betterthan ±0.05‰ and ±0.09‰, respectively. Analyses were performed atthe Large Lakes Observatory (LLO) Stable Isotope Lab using a CostechElemental Analyzer coupled with a ThermoFinnigan DeltaPlusXP stableisotope ratio monitoring mass spectrometer (EA-IRMS).

Lipid analysis for molecular biomarkers

The sediment sampling frequency for lipid analysis was half that forbulk analysis (i.e., ten samples from each core were analyzed for molec-ular biomarkers). The total lipid extract (TLE) was obtained from sedi-ments using a DIONEX Accelerated Solvent Extractor (ASE) 350, witha solvent ratio of 9:1 dichloromethane/methanol (DCM/MeOH). TLEsfrom core BH03-1 were separated into sub-fractions using Alltech®aminopropyl bond elute columns. The neutral fraction, containing then-alkane biomarkers of interest, was eluted first with 9:1 DCM/MeOH.After sulfur removal by activated (reduced) copper beads, TLEs fromcore BH09-5 (not exceeding 5 mg in weight) and neutral fractionsfrom core BH03-1 were separated by activated (150 °C for 2 h)alumina column chromatography using solvents of increasing polarity(9:1 hexane/DCM, 2:1 DCM/MeOH and MeOH, for apolar, aromaticandpolar compounds, respectively). The n-alkanes were present in thefirst (apolar) fraction; this fraction was next passed through a Ag+ im-pregnated silica pipette column to separate saturated and unsaturatedhydrocarbons. The saturated fraction was eluted first with n-hexane.Each fraction was dried under a stream of N2 gas.

The saturated fraction was quantified using an Agilent 6890 GCequipped with a flame ionization detector (FID). An internal standard(androstane)was added to each sample prior to injection for quantifica-tion of individual n-alkanes. Identification of n-alkanes was accom-plished using an Agilent 6890A Series GC System equipped with anAgilent 5973 single quadrupole mass spectrometer (GC/MS).

Stable carbon isotope compositions of individual n-alkanes were de-termined with a ThermoFinnigan DeltaPlusXP isotope ratio mass spec-trometer, using a modified GCC III interface (GC-C-IRMS). Calibrationwas performed using a mixture of n-alkanes (C16 to C30, Schimmelman“Mix A”) of known isotopic value, which was analyzed multiple timesdaily. Based on these replicate measurements, typical precision ofindividual δ13Cn-alkane measurements was better than ±0.50‰ (1σ,n=24). Squalene was co-injected with each sample as an internal iso-topic standard. Samples were run in duplicate, with reproducibility bet-ter than ±0.50‰. In all samples the C17n-alkane was not present inadequate abundance to obtain an accurate isotopic measurement. Incore BH09-5, theC19n-alkane in thefirst two samples co-elutedwith an-other compound and those isotopic values have been excluded from thedata set. Values are expressed in conventional delta notation as per mil(‰) deviations from the Vienna Pee Dee Belemnite (VPDB) standard.

Correction for the Suess Effect

To account for the change in the δ13C of atmospheric CO2 from an-thropogenic fossil fuel burning (the Suess effect) over the past two cen-turies, it is necessary to correct bulk and molecular δ13C values.According to Verburg (2007), lakes that are in isotopic equilibriumwith the atmosphere (e.g., lakes N1000 km2Bade et al. (2004); LakeSuperior is ~82,000 km2) will reflect the isotopic signature in atmo-spheric CO2 and the Suess effect. Verburg (2007) additionally suggeststhat the Suess correction be applied with caution in extremely hetero-trophic lakeswhere the source of sedimentary organicmatter is primar-ily allochthonous. Atilla et al. (2011) have shown that Lake Superior is inequilibriumwith CO2atm and that the lake, overall, is not respiring largeamounts of terrigenous CO2 (terrigenous carbon inputs do not exceed17% of the total carbon budget; Zigah et al., 2011). Zigah et al. (2011)also report bulk Δ14CDIC measurements in Lake Superior trackatmospheric radiocarbon levels, with a calculated residence time fordissolved inorganic carbon (DIC) of ~3 years. These observations sub-stantiate our application of the Suess correction to bulk and molecularδ13C values in Lake Superior sediments. Therefore, all δ13C values werecorrected using the equation of Verburg (2007), as it encompasses theentire time period of the two cores (2009 to 1700 AD) covered by ourage model.

Page 4: Journal of Great Lakes Research - University of Pittsburghjwerne/uploads/3/0/1/0/30101831/45.obeirneetal.jglr1… · Anthropogenic influences on the sedimentary geochemical record

23M.D. O'Beirne et al. / Journal of Great Lakes Research 41 (2015) 20–29

The Pearson test (two-tailed)was conducted using SigmaPlot v. 11.0to examine bivariate correlations among different paleoenvironmentalproxies. The α-level was set at 0.05.

Results

Sedimentation rate and age model

The age–depth relationship of the two cores was estimated fromsemi-log plots of excess 210Pb activity versus accumulated sedimentmass (Fig. 2) using methods described in Johnson et al. (2012). Theslopes of the straight-line segments below the uppermost centimeterin the core are proportional to sediment mass accumulation rates(MARs), which are applied to arrive at sediment age at each horizon.A bioturbation zone is apparent in the uppermost cm, exhibits thesteepest slope observed (Edgington and Robbins, 1990; Robbins andEdgington, 1975) and precludes application of a Constant Rate of Supply(CRS)model to derive sediment age profiles in the core. TheMAR in thebioturbated zone is assumed to be constant and equal to that derivedfrom the slope of the line segment immediately below it. The 210Pbdata revealed a bioturbated zone of approximately one cm in eachcore, which is equivalent to ~3 years of sedimentation.

The profiles of excess 210Pb vs. accumulated sediment mass are re-markably similar in the two cores (Fig. 2). The MARs in Lake Superiorhave not been constant over the periods of depositional history record-ed in these two cores.We approximate the agemodels by selecting twointervals of constant MAR in each core, determined by a least squareslinear fit through the data in each interval (Fig. 2). While the profilescould have been split intomore than 2 straight-line segments to achievemore refined age models, the improved temporal resolution, if anymore accurate, would not impact the conclusions of our study. The cal-culatedmass accumulation rates in both cores are analogous, showing a

Fig. 2. Left: excess 210Pb vs. accumulated sediment mass in each core. The sediment mass accumdata illustrated in the linear portions of the two cores from western Lake Superior. Right: age v

decrease in rates around 1950. The MAR for the first 12 cm (2002–1948 A.D.) in BH03-1 is 0.086 g/cm2/yr and 0.094 g/cm2/yr belowthat. Core BH09-5 shows a mass sedimentation rate of 0.090 g/cm2/yrfor the top 14 cm (2008–1950 A.D.) and 0.110 g/cm2/yr below.

Bulk geochemical data

Patterns of accumulation of the bulk proxy data are similar in bothBH03-1 and BH09-5. Most notable are the increases in organic carbonmass accumulation rate (TOCMAR) and total nitrogen mass accumula-tion rate (TNMAR) beginning in 1850, reaching a maximum in 1930(Fig. 3). Values tend to decrease after 1930 showing an additional in-crease beginning after 1990. The Corg:N record tracks changes in boththe TOCMAR and TNMAR records, with the highest values occurringwhen TOCMAR (Pearson r=0.850, p b 0.0001, n=28) and TNMAR (Pear-son r= 0.458, p b 0.01, n= 28) are the highest, before 1940. Corg:N re-main high after TOCMAR and TNMAR decrease after 1940, beforebeginning a continuous steady decline since 1970.

The general pattern observed in bulk sediment carbon isotopicvalues (δ13Corg) for both cores is a period of 13C-enrichment from1900 to 1960 followed by a leveling off and slight decline of δ13Corg

values in the 1970s (Fig. 3). The most striking increase is observed be-tween 1900 and 1960, where δ13Corg values increase almost 2‰, fromb−26‰ to−24.5‰. A second period of increasing δ13Corg begins post1990, with the most enriched δ13C signatures occurring at the top ofthe cores. Unlike δ13Corg values, δ15N values start their increase later,ca. 1925, with values increasing from 1.5‰ to 3‰ by 1975. δ15N valuesalso decrease more significantly from 1980 to 2000. Similar to the earli-er record, the δ15N values again increase but later than δ13C beginning in2000 and increasing to the top of the core, following the observedtrends in TOC and TN MARs.

ulation rates (MARs) presented correspond to the linear segments that were fitted to thes. depth plots with associated error.

Page 5: Journal of Great Lakes Research - University of Pittsburghjwerne/uploads/3/0/1/0/30101831/45.obeirneetal.jglr1… · Anthropogenic influences on the sedimentary geochemical record

Fig. 3. Temporal variation in bulk organic geochemical proxies within two corresponding cores from western Lake Superior. Shaded areas indicate periods of inferred enhanced produc-tivity based on the multiple proxies.

24 M.D. O'Beirne et al. / Journal of Great Lakes Research 41 (2015) 20–29

Molecular biomarkers

The flux of both aquatic (C17 + C19 + C21) and terrigenous (C27 +C29 + C31) n-alkane biomarkers show an overall increasing trend upto 1925, after which values decrease until 1975 and level off thereafter;with a minor increase around 2000 (Fig. 5). Variations in the weightedisotopic averages of both aquatic (i.e., δ13C19 and δ13C21 expressed asδ13Caquatic) and terrigenous (i.e., δ13C27, δ13C29 and δ13C31, expressed asδ13Cterrigenous) n-alkane biomarkers are used to examine changes inprimary productivity as observed in the bulk δ13Corg record (Fig. 5).δ13Caquatic values display an increase (−34‰ to −30‰) from 1825 to1960, decreasing slightly thereafter, before increasing again in the mid-1990s. δ13Cterrigenous values show a decreasing trend between 1825 and1925, after which values remain steady. Generally, aquatic biomarkerδ13C valuesfluctuate between−35‰ and−30‰, while terrigenous bio-marker δ13C values fluctuate between−33‰ and −30‰ (Fig. 6).

Discussion

Effects of diagenesis on the sedimentary geochemical record

It is important when using organic geochemical indicators to recon-struct the recent past (the period in which degradation rates are thegreatest) to consider the role of diagenesis in altering the sedimentarygeochemical record. Organic carbon in excess of that which can be at-tributed to either steady-state degradation (Li et al., 2012) or changingsediment MARs is present in the two cores deposited since 1900. This

conclusion is supported by downcore increases in TOCMAR profiles andelemental Corg:N ratios in sections of the two cores (Fig. 3), which arenot typical of diagenetic profiles present in Lake Superior sediments(cf. Li et al., 2012). Our data indicate that TOCMAR is a reliable indicatorof productivity changes in these cores, especiallywhen used in conjunc-tion with other proxies (i.e., δ13Corg, δ15N and Corg:N).

Period of 13C-enrichment (1900–1970)

The Lake Superior watershed witnessed its first European settle-ments in the early 1800s, with the beginning of the fur trade.Subsequent population increases occurred with the further develop-ments of logging, mining, shippingand processing of raw materials,until the Duluth-Superior Harbor region reached its historic maximumof 100,000 people in 1922. A series of water quality studies, conductedby the United States Geological Survey (USGS) and Minnesota Board ofHealth in the 20s, 40sand then 60s concluded that the water qualitywithin the St. Louis River had steadily declined with regional develop-ment (U.S. Department of the Interior, 1969 and references therein).The cumulative impacts of this cultural and economic developmenteventually resulted in the International Joint Commission (IJC) identify-ing the region as an Area of Concern under the Great LakesWater QualityAgreement of 1987 because of multiple beneficial use impairments.

We use the time period, 1800–1900, as a baseline on which to com-pare later changes in the bulk proxy data. Prior to 1900, we expect littleanthropogenic disturbance in the watershed as European settlementand development of the region was slow. Prior to 1900, both cores ex-hibit δ13Corg, δ15Nand Corg:N values (Figs. 3 and 4) that indicate organic

Page 6: Journal of Great Lakes Research - University of Pittsburghjwerne/uploads/3/0/1/0/30101831/45.obeirneetal.jglr1… · Anthropogenic influences on the sedimentary geochemical record

25M.D. O'Beirne et al. / Journal of Great Lakes Research 41 (2015) 20–29

matter derived primarily from aquatic algae growing within a cold, oli-gotrophic environment (Meyers and Ishiwatari, 1993), such as LakeSuperior—although a slight increase in TOCMAR and Corg:N could be re-lated to logging in the region. The mass accumulation rates of TOC andTNdiffer little between the two cores and appear to be highly correlatedtemporally, suggesting dominant controls from in-lake primary produc-tion. This inference is supported by Corg:N ratios, with values fallingmostly near the lacustrine algae range rather than within C3 or C4 landplants (Meyers, 1994; Fig. 4), although an increasing contributionfrom terrigenous C3 land plants to the sediment organic matter pool isobserved between 1900 and 1970.

We interpret the observed increases in δ13Corg and δ13Caquatic valuesbetween 1900 and 1960 to reflect a period of enhanced lacustrine pro-ductivity (Figs. 3 and 5). Increasing δ13Corg values are consistentwith in-creasing lacustrine primary production from 1900 to 1960 and coincidewith increased anthropogenic disturbances during the development ofthe St. Louis Riverwatershed. The observed 13C-enrichment of sedimen-tary organic matter could also be due to the preferential mineralizationof isotopically light carbon compounds during early diagenesis. Thispossibility seems unlikely; in low organic content (b2–3%) sedimentssuch as those in Lake Superior, isotopic discrimination associated withmineralization appears to have aminor influence on the δ13Corg(Meyers,1994). Additionally, Hodell and Schelske (1998) found that in Lake On-tario post-burial diagenesis reduced the mass of organic carbon buriedwithin sediment cores taken 6 years apart but did not change the

Fig. 4. Cross-plot of atomic Corg:N and δ13Corg from two corresponding multicores. Datafrom the two cores demonstrate the relative source contributions to bulk sediment organ-icmatter inwestern Lake Superior. Top: time periods of differing source contributions arenoted; magnified version of bottom plot. Bottom: generalized atomic Corg:N ratios andδ13Corg values of major sources of organic matter to lake sediments (after Meyers, 1994).

Fig. 5. Temporal variation in the flux of both aquatic and terrigenous n-alkane biomarkerswithin two cores from western Lake Superior. Shaded areas indicate periods of inferredenhanced productivity based on the multiple proxies.

δ13Corg values; a condition which we also observe in the two coresfrom western Lake Superior presented in this study (Fig. 3).

Carbon isotopic values of algal-derived molecules (δ13Caquatic)should not be influenced by terrigenous carbon inputs, and thereforethey can provide a better indication of trophic state than that of bulkδ13Corg and n-alkane abundances (Meyers, 2003). The trends in theδ13Caquatic signature agree well with the bulk organic carbonδ13Corgsignature, suggesting that aquatic productivity drives the isoto-pic composition of the sedimentary geochemical record (Fig. 6). Abso-lute n-alkane abundances for both aquatic and terrigenous sourcesshow similar trends down core. Increased fluxes from 1920 to 1950 cor-respond with increases in aquatic productivity and transport of terrige-nous material to the lake, consistent with the development of thewatershed (Fig. 5).

The variability in the δ15N record and the temporal shifts in both di-rections—especially evident in core BH03-1(Fig. 3)—indicate thatmulti-ple sources of N with differing isotope compositions are influencing the

Page 7: Journal of Great Lakes Research - University of Pittsburghjwerne/uploads/3/0/1/0/30101831/45.obeirneetal.jglr1… · Anthropogenic influences on the sedimentary geochemical record

Fig. 6. Comparison of temporal variations in molecular δ13Caquatic, δ13Cterrigenous, and bulk δ13Corg values between two cores from western Lake Superior. Shaded areas indicate periods ofinferred enhanced productivity based on the multiple proxies.

26 M.D. O'Beirne et al. / Journal of Great Lakes Research 41 (2015) 20–29

final δ15N signature preserved in the sediment core records. The morerapid temporal variations between 1900 and 1970 are likely due to wa-tershed disturbances causing the export of varying amounts of soil N tothis part of the lake (soil NO3

−: 3–12‰, Hodell and Schelske, 1998). Thisinterpretation is supported by increases in the Corg:N ratio between1900 and 1970 (Fig. 4). However, the most plausible explanation forthe overall long term trend of increasing δ15N between 1900 and 1970in the two cores is the gradual increase in primary production, as evi-denced by increasing δ13Corg and δ13Caquatic values during the sametime period (Fig. 6).

The observation that δ15N starts increasing after the observed in-crease in productivity (increases in TOCMAR, TNMAR and δ13Corg) indi-cates that nitrogen limitation is not the factor determining the amountof primary production (i.e., δ13Corg begins increasing around 1900,whereas δ15N increases starting at ~1920; Fig. 3). Thus, some other lim-iting nutrient must be responsible for the increases in primary produc-tion from 1900 to 1960. Historical measurements of phosphorus withinLake Superior are nonexistent until the 1970s; therefore, we can onlysurmise that the nutrient responsible for the historical increase in pro-ductivity is phosphorus.

P-availability is known to control productivity within present-day Lake Superior with Fe playing a co-limiting role at least duringartificially stimulated primary production (Ivanikova et al., 2007;Sterner et al., 2007). Schelske et al. (2006) offer evidence of anthro-pogenically driven eutrophication in western Lake Superior,reporting enhanced biogenic silica accumulation between 1900 and1970 from which they infer increased historic phosphorus supply(diatom growth in Lake Superior is phosphorous limited; Schelskeet al., 1986). This is consistent with data presented by Munawarand Munawar (1978) who report the dominant phytoplankton spe-cies in the western arm of Lake Superior during 1973 was the diatomspecies Melosira granulata (now identified as Aulacoseira granulata),which is often associated with eutrophication/perturbed environ-ments (Holland, 1968). The observed increases in productivity pre-sented in the current study are consistent with increasedanthropogenic phosphorus loading beginning in 1900, which alsocorrelates well with the increases in productivity, in response toP loading, reported in Lakes Erie and Ontario during the same timeperiod (Hodell and Schelske, 1998; Lu and Meyers, 2009; Lu et al.,2010; Schelske and Hodell, 1995).

Page 8: Journal of Great Lakes Research - University of Pittsburghjwerne/uploads/3/0/1/0/30101831/45.obeirneetal.jglr1… · Anthropogenic influences on the sedimentary geochemical record

27M.D. O'Beirne et al. / Journal of Great Lakes Research 41 (2015) 20–29

Lake Superior is the largestmanaged freshwater body in theworld. Ithas been regulated (levels and flow) through the International JointCommission (IJC) since the signing of the Boundary Waters Treaty in1909. It is clear from observations of the sedimentary geochemical re-cord that anthropogenic disturbances correlated with increasinghuman populations led to increased lacustrine productivity, most likelydriven by P loading, during the major developmental period of the wa-tershed, at least in the Duluth-Superior harbor region.

Declines in proxy data beginning in 1925

Decreases in MARs of organic C (with no accompanying change inδ13Corg) and total N and fluxes of both aquatic and terrigenous n-alkanebiomarkers began post 1940, years before the strict regulatory P controlactions of themid-1970s as required after the signing of the Great LakesWater Quality Agreement in 1972. Three large hydropower dams werebuilt on the St. Louis River between 1900 and 1925 (Fig. 1); the Thom-son Water Project (a series of dams) in 1907, the Scanlon Dam in 1922and the Fond du Lac Dam in 1924. During this same time period, thelumber industry began its decline due to depletion of high value lumberproducts from the forests and increased competition; only one offifteen original lumber mills was still operating in 1925 (White andMladendoff, 2004; Minnesota Historical Society). The construction ofthese dams and the decrease in lumbering are themost plausible causesof the observed declines in TOCMAR and TNMAR in both cores, as well asthe delivery of aquatic and terrigenous n-alkane biomarkers in BH09-5and to a lesser extent in BH03-1(Figs. 3 and 5). Corg:N ratios show thatterrigenous contribution to the sedimentary record is still influentialin this time period as values trend towards C3 land plants between1900 and 1970 (Fig. 4).

An additional, although comparatively short, decrease in TOCMAR

and TNMAR is observed between 1975 and 1990 (Fig. 3). In 1971,-Minnesota State Legislature approved the creation of the WesternLake Superior Sanitary District (WLSSD), which then became operation-al in 1978. This could account for the observed decrease in sedimentorganic matter, as the tertiary plant combined virtually all of theMinnesota industrial and municipal discharges in the area and reducedanthropogenic P loading to the western end of the lake. Decreases inCorg:N ratios (Fig. 4) indicate that algal derived organic matter becomesa larger component of sediment organic matter around 1975, eventhough δ13Corg shows little variation (Fig. 3).

During this same time period (1975–1990), δ15N values decreasewhile TNMAR remains relatively stable (Fig. 3), which is counter toexpectation if aquatic productivity alone is driving the observed δ15Nsignal. Therefore, this period of 15N-depletion is a likely reflectionof continually increasing atmospheric deposition of isotopically lightN to the lake from anthropogenic fossil fuel combustion and syntheticfertilizers that is subsequently incorporated during primary production.The contribution of fertilizers to the depletion of δ15N is unlikely to be ofsignificance in Lake Superior, given the small watershed with littleagricultural development. Large increases in NOx emissions wereobserved in the United States (EPA, 2000) and Canada(Chen et al.,2000) between 1940 and 1980, after which emissions remain relativelyconstant. Increasing δ15N values after 1990 in the cores may be aresponse to constant NOx emissions, whereby δ15N values begin to re-flect the influence of primary production rather than anthropogeniclight N inputs.

The latest increase in productivity (1990–present)

The latest increase in δ13Corg values, beginning near 1990, is enig-matic but may be attributable to increasing water temperatures andlength of the stratified season. Magnuson et al. (1997) predicted thatincreased epilimnetic temperatures are unlikely to affect rates of phyto-plankton primary production directly, but increases in annual produc-tion should be expected within the Laurentian Great Lakes because of

longer open water growing seasons. This phenomenon has been ob-served in the history of Lake Baikal, a high-latitude large lake similarto Lake Superior (Hampton et al., 2008; Moore et al., 2009). In contrast,Lehman (2002) predicted a longer stratified season causing nutrientlimitation, resulting in a reduction in rates of productivity within LakeSuperior, which is already the most oligotrophic of the LaurentianGreat Lakes. Lake Tanganyika in East Africa has experienced an en-hanced thermal gradient between the epilimnion and hypolimnion,resulting in inefficient mixing (nutrient turnover) of the two layersand a reduction in primary production (Verburg et al., 2003; O'Reillyet al., 2003; Tierney et al., 2010). The length of ice-free season hasbeen predicted to be proportional to phytoplankton production in theCanadian Shield Lakes (Fee et al., 1992). Lake Superior was noted asan exception to this due to its large areas of open water during mostwinters. However, Lake Superior is not completely ice-free for 12months of the year, and a shift to longer ice-free and summer stratifiedperiods has been noted in the past 20 years (Austin and Colman, 2007,2008). Quay et al. (1986) showed that the δ13C of dissolved inorganicC increased as the stratified season lengthened in Lake Washingtonwhich would be reflected in the δ13Corg photosynthetically producedfrom the DIC. A recent three-dimensional model of Lake Superior, in-cluding dynamic and thermodynamic ice cover, shows an increasingtrend in total annual primary production (for the simulation period of1985–2008) that correlates with decreasing winter ice cover, as wellas increasing annual average surface water temperatures (White et al.,2012). Our results, showing increasing bulk organic C and N contentwithin sediments as well as enriched δ13Corg and δ13Caquatic values, areconsistent with the modeled results and the results of Quay et al.(1986).

Implications

Although our data clearly indicate that anthropogenic disturbancesin the St. Louis Riverwatershed impacted lacustrine productivitywithinthe western arm of Lake Superior, we do not assume this to be the casefor the whole of Lake Superior due to its large size and increasing depthoff shore. In general, we would expect impacts such aswe observe to belimited to relatively shallow, near shore areas in large lakes, where theloading of nutrients and productivity would be highest. Thus, theseare constrained to the western arm of Lake Superior, the area mostheavily influenced by disturbances in the watershed of the second larg-est inflow (i.e., the St. Louis River) to Lake Superior. However, we notethat the human impacts described could potentially influence the restof Lake Superior, or any lake system, given a large enough anthropogen-ic input relative to lake size.

On the other hand, the influence of increased atmospheric deposi-tion on the nitrogen record (i.e., 15N-depletion) is likely to be a lakewide occurrence. Lake Superior has the largest surface area for anylake in the world, and due to the fact that atmospheric depositionisnot localized, wewould expect lakewide incorporation of atmospher-ically deposited, isotopically light N during primary production. The ob-servation of 15N-depletion, as recorded in the sediments, and attributionto atmospheric sources is oneof thefirst of its kind for Lake Superior andlarge lakes in general, but has been previously observed in many small,remote lakes (Holtgrieve et al., 2011). While atmospheric deposition ofisotopically light N is the most plausible scenario relating to the N re-cord,more detailed studies are needed in order to confirm our interpre-tation. Our observations regarding the most recent increases inproductivity post 1990 are the first direct sedimentary evidence of a re-cent trend of increasing productivity within Lake Superior. This trend isconsistent with increasing temperatures stimulating primary produc-tivity through a positive feedback related to the length of the ice-freeperiod, although further investigation should be undertaken to confirmthis inference and whether or not it is a lake wide or localizedphenomenon.

Page 9: Journal of Great Lakes Research - University of Pittsburghjwerne/uploads/3/0/1/0/30101831/45.obeirneetal.jglr1… · Anthropogenic influences on the sedimentary geochemical record

28 M.D. O'Beirne et al. / Journal of Great Lakes Research 41 (2015) 20–29

Conclusions

Sediments of western Lake Superior provide a historic record ofchanges in the source and flux of sedimentary organic matter; these,in turn, provide a record of historical ecosystem disturbances/changeswithin the western arm of the lake and its watershed. Early settlementand subsequent industrial development produced an increase in aquaticproductivity, most likely driven by P loading associated with urbaniza-tion of the watershed. Increases in aquatic primary productivity aremarked by enhanced organic carbon delivery to the sediment as wellas the enrichment of δ13Corg values of sedimentary organic matter.Further evidence of aquatic productivity being themajor influence driv-ing the changes seen in δ13Corg is observed in the increase of aquatic bio-marker δ13Caquatic values during the same time period. The most recentdecrease in bulk δ13Corg, δ15N andCorg:N valuesmay have begun as earlyas 1925 and were largely accomplished prior to establishment ofWLSSD. Nevertheless, δ13Corg and δ15N show their strongest decreaseafter 1970, which indicates that improved waste collection and treat-ment did reduce anthropogenic P loading, thereby reducing algal pro-ductivity. Observations of recent increased primary productivity areenigmatic but may be attributable to increasing water temperaturesstimulating primary productivity through a positive feedback relatedto the length of the ice-free period.

Acknowledgements

Support for this research was provided by Minnesota Sea Grant toRLH, TCJ and JPW. The 210Pb analyses were carried out by alpha spec-trometry in the Department of Soil Science, University of Manitoba,under the direction of Dr. Paul Wilkinson. The authors give thanks tothe following: the crew of the R/V Blue Heron for core collection,Sarah Grosshuesch and Koushik Dutta for their laboratory expertise,Julia Halbur for helpwith sample preparation and Jillian Votava. The au-thors also thank the associate editor and two anonymous reviewerswhose valuable comments considerably improved this manuscript.

References

Atilla, N., McKinley, G.A., Bennington, V., Baehr, M., Urban, N., DeGrandpre, M., Desai, A.R.,Wu, C., 2011. Observed variability of Lake Superior pCO2. Limnol. Oceanogr. 56,775–786.

Austin, J.A., Colman, S., 2007. Lake Superior summer water temperatures are increasingmore rapidly than regional air temperatures: a positive ice-albedo feedback. J.Geophys. Lett. 34.

Austin, J.A., Colman, S., 2008. A century of temperature variability in Lake Superior.Limnol. Oceanogr. 53, 2724–2730.

Bade, D.L., Carpenter, S.R., Cole, J.J., Hanson, P.C., Hesslein, R.H., 2004. Controls of δ13C-DICin lakes: geochemistry, lake metabolism, and morphometry. Limnol. Oceanogr. 49,1160–1172.

Bourbonniere, J.A., Meyers, P.A., 1996. Sedimentary geolipid records of historical changesin watersheds and productivities in Lakes Ontario and Erie. Limnol. Oceanogr. 41,352–359.

Chen, J., Chen, W., Liu, J., Cihlar, J., Gray, S., 2000. Annual Carbon Balance of Canada'sforests during 1895–1996. Glob. Biogeochem. Cycles 14, 839–850.

Cranwell, P.A., 1981. Diagenesis of free and bound lipids in terrestrial detritus depositedin a lacustrine sediment. Org. Geochem. 3, 79–89.

Cranwell, P.A., Eglinton, G., Robinson, N., 1987. Lipids of aquatic organisms as potentialcontributors to lacustrine sediments-II. Org. Geochem. 11, 513–527.

Edgington, D.N., Robbins, J.A., 1990. Time scales of sediment focusing in large lakes as re-vealed bymeasurement of fallout Cs-137. In: Tilzer, M., Serruya, S. (Eds.), Large Lakes:Ecological Structure and Function. Brock/Springer Series in Contemporary Bioscience.Springer Verlag, Berlin, pp. 210–223.

Eglinton, G., Hamilton, R.J., 1963. The distribution of alkanes. In: Swain, T. (Ed.), ChemicalPlant Taxonomy. Academic Press.

Eglinton, G., Hamilton, R.J., 1967. Leaf epicuticular waxes. Science 156, 1322–1335.Environmental Protection Agency (EPA), 2000. National air pollution emission trends,

1900–1998. Rep. EPA-454/R-00-002, (238 pp., Washington D.C.).Fee, E.J., Shearer, J.A., DeBruyn, E.R., Schindler, E.U., 1992. Effects of lake size on phyto-

plankton photosynthesis. Can. J. Fish. Aquat. Sci. 49, 2445–2459.Giger, W., Schaffner, C., Wakeham, S.G., 1980. Aliphatic and olefinic hydrocarbons in re-

cent sediments of Greifensee, Switzerland. Geochim. Cosmochim. Acta 44, 119–129.Hampton, S.E., Izmest'eva, L.R., Moore, M.V., Katz, S.L., Dennis, B., Silow, E.A., 2008. Sixty

years of environmental change in the world's largest freshwater lake—Lake Baikal,Siberia. Glob. Chang. Biol. 14, 1–12.

Hodell, D.A., Schelske, C.L., 1998. Production, sedimentation, and isotopic composition oforganic matter in Lake Ontario. Limnol. Oceanogr. 43, 200–214.

Holland, R.E., 1968. Correlation of Melosira species with trophic conditions in LakeMichigan. Limnol. Oceanogr. 13, 555–557.

Holtgrieve, G.W., Schindler, D.E., Hobbs, W.O., Leavitt, P.R., Ward, E.J., Bunting, L., Chen, G.,Finney, B.P., Gregory-Eaves, I., Holmgren, S., Lisac, M.J., Lisi, P.J., Nydick, K., Rogers, L.A.,Saros, J.E., Selbie, D.T., Shapley, M.D., Walsh, P.B., Wolfe, A.P., 2011. A coherent signa-ture of anthropogenic nitrogen deposition to remote watersheds of the NorthernHemisphere. Science 334, 1545–1548.

Ivanikova, N.V., McKay, R.M.L., Bullerjahn, G.S., Sterner, R.W., 2007. Nitrate utiliza-tion by phytoplankton in Lake Superior is impaired by low nutrient (P, Fe) avail-ability and seasonal light limitation—a cyanobacterial bioreporter study. J.Phycol. 43, 475–484.

Johnson, T.C., Van Alstine, J.D., Rolfhus, K.R., Colman, S.M., Wattrus, N.J., 2012. A high res-olution study of spatial and temporal variability of natural and anthropogenic com-pounds in offshore Lake Superior sediments. J. Great Lakes Res. 38, 673–685.

Lehman, J.T., 2002. Mixing patterns and plankton biomass of the St. Lawrence Great Lakesunder climate change scenarios. J. Great Lakes Res. 28, 213–250.

Li, J., Crowe, S.A., Miklesh, D., Kistner, M., Canfield, D.E., Katsev, S., 2012. Carbon mineral-ization and oxygen dynamics in sediments with deep oxygen penetration, LakeSuperior. Limnol. Oceanogr. 57, 1634–1650.

Lu, Y., Meyers, P.A., 2009. Sediment lipid biomarkers as recorders of the contaminationand cultural eutrophication of Lake Erie, 1909–2003. Org. Geochem. 912–921.

Lu, Y., Meyers, P.A., Eadie, B.J., Robbins, J.A., 2010. Carbon cycling in Lake Erie during cul-tural eutrophication over the last century inferred from the stable carbon isotopecomposition of sediments. J. Paleolimnol. 43, 261–272.

Magnuson, J.J., Webster, K.E., Assel, R.A., Bowser, C.J., Dillon, P.J., Eaton, J.G., Evans, H.E.,Fee, E.J., Hall, R.I., Mortsch, L.R., Schindler, D.W., Quinn, F.H., 1997. Potential effectsof climate changes on aquatic systems: Laurentian Great Lakes and precambrianshield region. Hydrol. Process. 11, 825–871.

McKenzie, A.J., 1985. Carbon Isotopes and Productivity in the Lacustrine and MarineEnvironment. In: Stumm, Werner (Ed.), Chemical Processes in Lakes. Wiley-Interscience, New York, pp. 99–118.

Meyers, P.A., 1994. Preservation of elemental and isotopic source identification of sedi-mentary organic matter during and after deposition. Chem. Geol. 144, 289–302.

Meyers, P.A., 1997. Organic geochemical proxies of paleoocenagraphic, paleolimnologic,and paleoclimatic processes. Org. Geochem. 27, 213–250.

Meyers, P.A., 2003. Applications of organic geochemistry to paleolimnologicalreconstructions: a summary of examples from the Laurentian Great Lakes. Org.Geochem. 34, 261–289.

Meyers, P.A., Ishiwatari, R., 1993. Lacustrine organic geochemistry—an overview of indicatorsof organicmatter sources and diagenesis in lake sediments. Org. Geochem. 20, 867–900.

Moore, M.V., Hampton, S.E., Izmest'eva, L.R., Silow, E.A., Peshkova, E.V., Pavlov, B.K., 2009. Cli-mate change and theWorld's “Sacred Sea”—Lake Baikal, Siberia. Bioscience 59, 405–417.

Munawar, M., Munawar, I.F., 1978. Phytoplankton of Lake Superior 1973. J. Great LakesRes. 4, 415–442.

O'Reilly, C.M., Alin, S.R., Plisnier, P.D., Cohen, A.S., Mckee, B.A., 2003. Climate change decreasesaquatic ecosystem productivity of Lake Tanganyika, Africa. Nature 424, 766–768.

Quay, P.D., Emerson, S.R., Quay, B.M., Devol, A.H., 1986. The carbon cycle for LakeWashington-A stable isotope study. Limnol. Oceanogr. 31, 596–611.

Rieley, G., Collier, R.J., Jones, D.M., Eglinton, G., Eakin, P.A., Fallick, P.A., 1991. Sources ofsedimentary lipids deduced from stable isotope analyses of individual compounds.Nature 352, 425–427.

Robbins, J.A., Edgington, D.N., 1975. Determination of recent sedimentation rates in LakeMichigan using Pb-210 and Cs-137. Geochim. Cosmochim. Acta 39, 285–304.

Schelske, C.L., Hodell, D.A., 1991. Recent changes in productivity and climate of LakeOntario detected by isotopic analysis of sediments. Limnol. Oceanogr. 36, 961–975.

Schelske, C.L., Hodell, D.A., 1995. Using carbon isotopes of bulk sedimentary organic mat-ter to reconstruct the history of nutrient loading and eutrophication in Lake Erie.Limnol. Oceanogr. 40, 918–929.

Schelske, C.L., Fahnenstiel, G.L., Haibach, M., 1986. Phosphorus enrichment, silicautilization, and biogeochemical silica depletion in the Great Lakes. Can. J. Fish.Aquat. Sci. 43, 407–415.

Schelske, C.L., Stoermer, E.F., Kenney, W.F., 2006. Historic low-level phosphorous enrich-ment in the Great Lakes inferred from biogenic silica accumulation in sediments.Limnol. Oceanogr. 51, 728–748.

Sterner, R.W., Anagnostou, E., Brovold, S., Bullerjahn, G.S., Findlay, J.C., Kumar, S., McKay,R.M.L., Sherrel, R.M., 2007. Increasing stoichiometric imbalance in North America'slargest lake: nitrification in Lake Superior. Geophys. Res. Lett. 34.

Tierney, J.E., Mayes, M.T., Meyer, N., Johnson, C., Swarzenski, P.W., Cohen, A.S., Russell, J.M.,2010. Late-twentieth-century warming in Lake Tanganyika unprecedented since AD500. Nat. Geosci. 3, 422–425.

U.S. Department of the Interior, 1969. Federal Water Pollution Control Administra-tion—Great Lakes Region. An Appraisal of Water Pollution in the Lake SuperiorBasin.

Urban, N.R., 2007. Nutrient cycling in Lake Superior: a retrospective and update. In:Munawar, M., Heath, R. (Eds.), The State of Lake Superior. Ecovision World Mono-graph Series.

Urban, N.R., Auer, M.T., Green, S.A., Lu, X., Apul, D.S., Powell, K.D., Bub, L., 2005. Carboncycling in Lake Superior. J. Geophys. Res. 110, C06S90. http://dx.doi.org/10.1029/2003JC002230.

Verburg, P., 2007. The need to correct for the Suess effect in the application of d13C in sed-iment in autotrophic Lake Tanganyika, as a productivity proxy in the anthropocene. J.Paleolimnol. 37, 591–602.

Verburg, P., Hecky, R.E., Klins, H., 2003. Ecological consequences of a century of warmingin Lake Tanganyika. Science 301, 505–507.

Page 10: Journal of Great Lakes Research - University of Pittsburghjwerne/uploads/3/0/1/0/30101831/45.obeirneetal.jglr1… · Anthropogenic influences on the sedimentary geochemical record

29M.D. O'Beirne et al. / Journal of Great Lakes Research 41 (2015) 20–29

Vollenweider, R.A., Munawar, M., Stadelmann, P., 1974. A comparative review of phyto-plankton and primary production in the Laurentian Great Lakes. J. Fish. Res. BoardCan. 31, 739–762.

Weiler, R.R., 1978. The chemistry of Lake Superior. J. Great Lakes Res. 4, 370–385.White, M.A., Mladendoff, D.J., 2004. Old-growth forest landscape transitions from pre-

European settlement to present. Landsc. Ecol. 9, 191–205.

White, B., Austin, J., Matsumoto, K., 2012. A three-dimensional model of Lake Superiorwith ice and biogeochemistry. J. Great Lakes Res. 38, 61–67.

Zigah, P.K., Minor, E.C., Werne, J.P., McCallister, S.L., 2011. Radiocarbon and stable isotopicinsights into provenance and cycling of carbon in Lake Superior. Limnol. Oceanogr.56, 867–886.


Recommended