102
Taking Action On Lake Erie IJC Science Advisory Board TAcLE Work Group Science Summary Report Dr Sue Watson (Canadian Co-Chair) Dr David Carpenter (US Co-Chair) Prepared as part of the Lake Erie Ecosystem Priority April 2013 This document is a Science Summary Report prepared for the IJC’s Lake Erie Ecosystem Priority (LEEP) by the Science Advisory Board workgroup and is not for citation. Views expressed are solely those of the authors. See the draft LEEP report for findings and recommendations from the IJC.

Taking Action On Lake Erie - International Joint Commission · reduce external inputs of total phosphorus (TP) and in particular, the rising proportion of the most bioavailable form

  • Upload
    others

  • View
    1

  • Download
    0

Embed Size (px)

Citation preview

Page 1: Taking Action On Lake Erie - International Joint Commission · reduce external inputs of total phosphorus (TP) and in particular, the rising proportion of the most bioavailable form

Taking Action On

Lake Erie

IJC Science Advisory Board TAcLE Work Group

Science Summary Report

Dr Sue Watson (Canadian Co-Chair) Dr David Carpenter (US Co-Chair)

Prepared as part of the Lake Erie Ecosystem Priority

April 2013

This document is a Science Summary Report prepared for the IJC’s Lake Erie Ecosystem Priority (LEEP) by the Science Advisory Board workgroup and is not for citation. Views expressed are solely those of the authors. See the draft LEEP report for findings and recommendations from the IJC.

Page 2: Taking Action On Lake Erie - International Joint Commission · reduce external inputs of total phosphorus (TP) and in particular, the rising proportion of the most bioavailable form

i

Executive Summary

Lake Erie (LE) is the most vulnerable of the North American Great Lakes. It the shallowest, warmest and the most susceptible to eutrophication and the effects of climate change. The recent accelerating decline of this lake, manifest as impaired water quality, massive, summer-long algal blooms, hypoxia and fish kills, has focussed attention on the need for rapid action to reduce external inputs of total phosphorus (TP) and in particular, the rising proportion of the most bioavailable form of P, soluble reactive P (SRP) which is seen as a primary cause of this decline. The Taking Action on Lake Erie (TAcLE) Task Force initiated a series of workgroups to evaluate the current status and threats, the effectiveness of current management measures and the potential impacts of climate change. This was undertaken as a series of white papers, which were tabled for review and discussion at an expert workshop. This report summarizes the major results of this effort, with focus on the following key issues Current External and Internal P Loading The bulk of the external P loading to LE is delivered to the West Basin (WB) from six major tributaries. In order to guide prioritised management actions, annual external TP and SRP loading estimates were updated to 2011 using a mass balance approach; however these are provisional on the inclusion of loads from Canadian tributaries and a more rigorous evaluation of Detroit River inputs. US tributary inputs were further resolved to sub-basin inputs by SWAT-based modelling to identify areas contributing the highest P and SRP from within the 6 drainage basins (Maumee, Sandusky, Grand, Raisin, Cuyahoga and Vermillion). Annual P loads are greatest from the Maumee and Sandusky Rivers, which also have high P and SRP concentrations. Inputs from the Detroit River also contribute a significant fraction of the total annual P loads, which are discharged at much lower concentrations but weighted by a very high volume of flow. In general, algal growth responds most immediately to the concentration of biologically-available P, and it is proposed that the WB blooms develop primarily in response to spring SRP inputs from the Maumee, while other loading (e.g., the Detroit River) contributes to longer term processes that drive Central Basin (CB) hypoxia. SWAT model outputs were of mixed success, with a high level of uncertainty around simulated results particularly for SRP. SWAT loading estimates were most reliable for basins dominated by agricultural inputs (Maumee, Sandusky), but unable to simulate urban-dominated inputs, and failed to capture the all-important inputs during extreme events, which can represent up to half of a total annual load. Atmospheric sources of P to Lake Erie include wet and dry deposition and are likely influenced by basin use, with potentially higher inputs from basins with high agricultural and urban development. Estimates are that SRP may account for up to 12% of the TP in wet deposition from some areas. Current estimates of atmospheric sources as ~6% of the total external TP input are likely underestimates, as they are based only on wet and not bulk deposition. There is a significant need for more detailed direct monitoring of both wet and dry fractions of atmospheric deposition; however efforts to track atmospheric P are lacking in existing monitoring programs such as the Integrated Atmospheric Deposition Network and the Great Lakes Precipitation Network. Coordination with existing nutrient monitoring efforts is an alternative, and a binational nutrient strategy approach could be considered similar to the existing Great Lakes Binational Toxics Strategy. Environment Canada’s Great Lakes Nutrient Initiative is the most recent effort supporting advancements in research and management for nutrients in the Great Lakes, and

Page 3: Taking Action On Lake Erie - International Joint Commission · reduce external inputs of total phosphorus (TP) and in particular, the rising proportion of the most bioavailable form

ii

inclusion of atmospheric P measurements in such an initiative would provide a much needed opportunity to expand our understanding of the phosphorus cycle. Internal loading occurs within the lake itself, and includes inter-basin exchange, biological processing and sediment regeneration. The external loading delivered to WB is eventually transferred to the central and deeper East Basin. Biological processing from uptake, sequestering and exchange across the food web recycle or transform P between particulate and dissolved fractions. Invasive dreissenid mussels have profoundly altered the lake’s response to nutrient inputs by altering the efficiency of internal P cycling and inshore-offshore exchange, and trapping nutrients and material in the warmer and shallower nearshore zones. Sediment SRP release under hypoxic conditions (CB) and aerobic decomposition of organic matter from the sediment-water interface and resuspended material (WB) may be significant. Evidence suggests that hypoxic sediment P regeneration in the CB does occur, but its significance and frequency is unclear and should be examined in detail. Current monitoring may have failed to capture this process adequately, which has yet to be measured directly. None of these internal loading processes have been well characterised in LE, which sometimes unduly casts doubt on the rationale for management actions, because this mechanism may delay response to expensive nutrient controls. Based on the lake’s recovery seen in the 1980s it is reasonable to expect that LE’s response would not be unduly delayed past 10-15 years, but regime shifts, changes in the nature and timing of the external loading and climate change may necessitate more stringent management targets and a re-evaluation of expected response times. The potential effects of climate change on external loading are of particular concern, since this has implications both for setting and achieving nutrient targets. This issue was also examined using SWAT models, which predicted that in-stream sediment and nutrient yields would generally increase under alternative climate scenarios, but these yields decreased in several instances. Overall, across the six watersheds the models predicted increasing flow, sediment and N (total & dissolved) with moderate and severe climate change. Total annual flows and in stream sediment increased by up to 17-22% under severe climate change, while more modest increases were predicted in nutrients. The response of N and P, and TP and SRP, were different. At the watershed level, in-stream flow, sediment and nutrient yields from the six Lake Erie watersheds responded very differently to alternative climate scenarios; management strategies aimed at TP management, therefore, may not address SRP. Effects of Agricultural and Urban Best Management Practices (BMPs) on Total P and SRP Loads to Lake Erie P loading from urban sources is often disregarded, underrated but may contribute ~20% of total basin input. Urban loading comes from construction activities, stormwater runoff, lawn and garden activities, leaves from deciduous trees and pet waste. Urban BMPs include installation of porous pavement, media filters, filter strips, green roofs, bioretention, detention basins, and wetlands. However the effectiveness of some of these methods is not high and their installation may be expensive. Urban BMP efficiency varies greatly and, because they were not designed for P removal, some BMPs actually increase P loadings. Bioretention ponds and wetland basins are quite effective, but may be relatively expensive to install and maintain. Better management of leaves and pet waste, street sweeping and use of native plants may also reduce urban phosphorus inputs. There has been major progress in reducing P loads from urban waste water sources, but there are still problems with combined sewer overflows from several sites.

Page 4: Taking Action On Lake Erie - International Joint Commission · reduce external inputs of total phosphorus (TP) and in particular, the rising proportion of the most bioavailable form

iii

Agriculture is a major source on nonpoint inputs of P, N and sediments, with P the issue of greatest concern. A comprehensive review shows that there has been no rigorous assessment of agricultural BMPs specific to the Lake Erie watersheds. BMP assessments have focused on TP and sediment reduction based on the traditional idea that the majority of P losses occur as particulate P attached to sediments. The range of BMP effectiveness greatly varies, and results are often conflicting. Major issues that need to be addressed are: a) scaling up of BMP effects to watershed scale from plot scale; and b) model reliability. Since most BMPs are used in combination with at least one other BMP, assessing the effectiveness of a single BMP is difficult and caution should be exercised in extrapolating effectiveness to other sites. There is no silver bullet to solve non-point source pollution, particularly SRP. The use of a suite of BMPs (i.e., a toolbox) is currently recommended. A major challenge is how to evaluate the synergy of BMP effectiveness. Other challenges include soil-P stratification and the tradeoffs in controlling SRP and sediment/erosion, e.g., no-till vs. conventional-till vs. nutrient management. There is a need of an inventory of implemented agricultural BMPs (a BMP clearing house) in Lake Erie watersheds and in the Great Lakes in general. This inventory is essential in assessing the potential effects of BMP implementation. Overall, a review of over 240 primary sources resulted in the following findings: 1) very few studies have quantified P load reductions by urban or agricultural BMPs within the Lake Erie watershed; 2) It is not possible to determine BMP cost-effectiveness due to costs rarely being reported; 3) BMP effectiveness, both urban & agricultural, vary greatly and are often contradictory; BMPs designed to control other issues (e.g., flow) may actually increase P loads; 4) Most methods commonly used to quantify BMP performance are ineffective; 5) there is a need to move beyond TP measurements as the only metric used to quantify P, assessing speciation is necessary to advance BMP performance; 6) improved models are required to accurately predict treatment efficiency of BMPs under a variety of conditions and climates; 7) while some databases exist, a central data repository is critically needed to synthesize data collected and improve understanding of BMP effectiveness Hypoxia Hypoxia, especially in the Central Basin, is an annual and natural event and one that probably precedes current urban and agricultural development. Hypoxia is driven by algal production, which then sinks and results in depletion of dissolved oxygen in the hypolimnion during the summer months. Of concern is the fact that in recent years the extent of central basin hypoxia has grown. This is a direct consequence of the increased algal blooms discussed below. Wind events are important in creating movement of hypoxic waters, resulting in fish kills. Hypoxia has major effects on food webs, and impacts animal life at all levels. Harmful and Nuisance Algal Blooms (HNABs) in Lake Erie: Planktonic harmful cyanobacterial blooms (cHABs) Beginning in the mid-1990s, increases in highly bioavailable SRP loading to LE have coincided with a second substantial decline in water quality and a resurgence of cHABs. Blooms of Microcystis aeruginosa and other cyanobacteria have formed annually in the western basin, and now are appearing elsewhere at inshore sites In the last decade, the blooms have developed earlier and extended later than in the past; the 2011 Lake Erie bloom was visible from satellites until mid-October. Blooms of Microcystis have generally been reported from the WB, but have recently been developing along the north shore of the Central and Eastern Basins. Potentially toxic species of Planktothrix and Anabaena have also been observed with

Page 5: Taking Action On Lake Erie - International Joint Commission · reduce external inputs of total phosphorus (TP) and in particular, the rising proportion of the most bioavailable form

iv

increasing frequency. While cyanobacteria are known to produce a number of toxins, microcystins are of particular concern. Early detection of cHAB formation is critical to formation of a proper response and available detection methods, including remote sensing, have greatly improved in sensitivity and speed in recent years. Molecular methods have tracked historical existence of Microcystis dating to the 1970s in Lake Erie, and show that at specific sites the current population is genetically indistinguishable from the historical population. This suggests that environmental or anthropogenic influences have resulted in a surge in the Microcystis population in recent years, not the invasion of a new strain. There is strong evidence that P is a key factor driving cyanobacterial biomass and dominance in many freshwater systems, including Lake Erie, but recent work suggests an important role for other nutrients (i.e., nitrogen), trace elements, climate change, and the co-occurring heterotrophic community. An understanding of the collective impact of these factors will lead to more accurate mathematical modelling and prediction of future cHABs. Collective field evidence suggests that while P may control total phytoplankton biomass, evidence of N limitation can be observed and the availability of particular N species may influence algal community structure. Climate models predict temperature increases that would favour cyanobacterial dominance and could contribute to an extension of annual bloom duration. Nuisance Algal Blooms (NABs) in Lake Erie: Benthic algae Over the past 10 – 15 years, studies on benthic algae in Lake Erie have been largely limited to two species that presently form annual nuisance blooms; Cladophora in the eastern basin and, more recently, Lyngbya in the western basin. These blooms foul recreational beaches, clog municipal and industrial water intakes, impair water quality and pose potential microbial health risks to wildlife and humans. Although the ecology of Cladophora is generally well documented in Lake Erie, far less information is available for Lyngbya. Conclusions on temporal trends in degree of severity of nuisance blooms of Cladophora and Lyngbya are hindered by the lack of standardized data collection protocols and reports, and infrequent sampling. Heterogeneity in environmental conditions relevant to the growth of benthic algae adds additional uncertainty. Recent applications of remote sensing technology and mobile survey technology have successfully documented spatial patterns in the coverage and attached biomass of Cladophora and offer new approaches to expand the spatial scope of future research. Differences in substrate availability and light climate appear to be major determinants of Lyngbya and Cladophora abundance in LE. In the more turbid WB, Lyngbya is often found in turbid shallow water associated with sand and dreissenids or their shell debris over a limited depth range (1.5 – 3.5 m), while Cladophora in the more transparent eastern basin is found attached to dreissenids, rocks and other hard substrate at depths between 0.5 – 10 m. The advent of dreissenid mussel populations in Lake Erie has likely contributed to nuisance blooms of Cladophora by improving water clarity, supplying nutrients, and providing substrate for filament attachment. The influence of dreissenid mussels on Lyngbya is less clear and merits further exploration. Increased SRP loading from major tributaries is the most parsimonious explanation for the recent development of Lyngbya blooms in the WB. High abundance and biomass of the

round goby may modify dreissenid‐Cladophora dynamics on a local scale, either through P

excretion or grazing relaxation; however, there is only circumstantial support for these mechanisms at present. Ideally, these interactions should be evaluated in habitats where mussels, Cladophora and gobies co-occur. Calibrated growth models for Cladophora show promise for representing its growth response

Page 6: Taking Action On Lake Erie - International Joint Commission · reduce external inputs of total phosphorus (TP) and in particular, the rising proportion of the most bioavailable form

v

under a variety of environment forcing conditions. Models for other benthic algae such as Lyngbya have not been developed. Hindcast simulations indicate that increased water clarity (attributed primarily to dreissenid mussel filtering) is largely responsible for the increased shoreline fouling by Cladophora since the late 1990s and early 2000s. Cladophora models are data intensive and require highly resolved spatial and temporal resolution of light, nutrient, thermal and disturbance regimes. Integration of Cladophora models into complex 3D hydrodynamic lake models may provide a means to evaluate responses on lake-wide to local scales. Climate change has the potential to greatly influence these interrelationships, as a consequence of changes in precipitation and temperature. The ultimate response will depend on the interactive effects of nutrients, light and disturbance regimes and a longer growing season. Impacts of changing circulation are also potentially important, given the ability of near shore hydrodynamics to impact the thermal, nutrient, light and disturbance regime. The issue of nuisance benthic algal blooms in Lake Erie merits sustained programs of integrated research and monitoring because the symptoms (both anthropocentric such as beach fouling, and ecological such as botulism outbreaks) of impairment to coastal regions in Lake Erie are so major that they cannot feasibly be ignored. It is clear that the synergistic impacts of human activity, invasive species and climate change present an extraordinary challenge in developing management objectives for nuisance blooms of benthic algae in Lake Erie. Much of the information regarding nuisance benthic algal blooms in the Great Lakes has been limited to site-specific assessments, sometimes supplemented with experimentation and simulation modeling. We now know that there are a number of important factors that influence the dynamics of benthic algal blooms in nearshore waters of the Great Lakes. Hydrodynamic and circulation of water masses shape the interaction of lake water with land-based runoff and tributary discharges, and strongly influence the nutrient, light, temperature and disturbance regimes in the near-shore. In addition, we now have a stronger sense of the ability of filter feeding organisms such as dreissenid mussels to act as a benthic – pelagic coupling mechanism, which may attenuate or exacerbate conditions suitable for the growth of benthic algae. What is lacking at present is a comprehensive understanding of how these various factors work together to create the attendant conditions associated with nuisance blooms of Cladophora and Lyngbya. This understanding is crucial for the sound development of management activity. Lake Erie Fish Populations: Status and impact of Climate Change Lake Erie fisheries have very large recreational and commercial value. Each species of fish has preferred prey and temperature ranges, and all depend upon adequate oxygen. In general, the shallow, warm and productive WB is currently dominated by species that are tolerant of high turbidity and warm temperatures. The EB, which is the deepest, coldest and least productive, is dominated by deepwater fish that prefer cold temperatures, high dissolved oxygen and clear waters. The CB, being intermediate, is dominated by cool-water species like yellow perch and walleye. Changes in lake temperature may dramatically alter this distribution, even to the loss of cold water species from the lake. Seasonal hypoxia will influence fish production both directly and indirectly, at multiple temporal and spatial scales. There are direct physiological costs for aerobic organisms occupying hypoxic habitats. Further, organisms that can avoid hypoxic regions (through vertical or horizontal migrations) may be forced to occupy inferior thermal and optical habitats, immediately constraining growth. Such behavioural migrations may alter the spatiotemporal overlaps, growth efficiencies, and vulnerabilities of predators and prey, leading

Page 7: Taking Action On Lake Erie - International Joint Commission · reduce external inputs of total phosphorus (TP) and in particular, the rising proportion of the most bioavailable form

vi

to long-term changes to food-web structure and energy flow. There would be coincident shifts in invertebrate and fish community composition. In general the interactive effects of climate change and nutrient loading are expected to promote a eutrophication-tolerant fish community. Visual feeding, cold-water and hypoxia-sensitive fish will decline while warm water tolerate fish will increase. However, the complex interactions between hypoxia, reduced water clarity, harmful algal blooms and altered prey base have the potential to directly and indirectly mediate population trajectories in ways not yet well understood. Modelling Lake Parameters in Relation to Nutrient Loading in Lake Erie Models are essential if we are to anticipate the changes in Lake Erie as a consequence of changes in nutrient loading, harmful and noxious algal blooms, climate change and management actions. Both empirical and process-based models can and have been used to study these effects, with varying degrees of success. Several general principles must apply. The selection of mathematical equations and parameter values should not be based on subjective or arbitrary criteria, but must be ecologically defensible and tightly linked to our contemporary understanding of the system. Model uncertainly analysis is critical, and is too often missing. However, the existence of various models with different strengths and weaknesses offers a unique opportunity for synthesis and improvement of modelling practice, and for predicting the impact of our management actions. For example, a good model can provide us with information as to what level of annual total phosphorus load must be achieved to reduce Central Basin hypoxia. Response curves were developed to predict levels of HABs, chlorophyll and hypoxia as a function of P-loading, to provide some guidance towards setting management targets. HABs response (modelled as an index of as planktonic cHABs in the WB) was predicted as a function of spring TP loading from the Maumee, postulated by the authors as the primary driver of these blooms. The results indicate that the mean TP load would have to be reduced by 24% to minimise or eliminate the risk of cHABs, which will take considerably stronger BMP implementation that is current in place. However this would not be sufficient for reducing the CB hypoxic area to a desired level. Response curves for CB hypoxia were developed as a function of total P loading to both the WB and CB, and predicted that, for example, to achieve an average hypoxic area < 2 km2, the annual TP target load would need to be reduced by ~70% from 11,000 MT/year to 3200MT/yr. Summary and Recommendations A critical role for science within Lake Erie is to inform both lake and land management practices. Efforts to improve water quality and ecological integrity are underway at local, state, national, and international levels in both Canada and the United States. Current focus on reduction of non-point source loading of TP and SRP remains a priority for national and state level organizations. The ongoing debate for nutrient management with a focus on P or both N and P should not delay action. Efforts to resolve this debate will advance both management practices and the science of freshwater cHABs. Continued engagement in monitoring and scientific endeavours along with monitoring and modelling to track changes in nutrients and algal blooms in Lake Erie is a high priority for provincial, state and national organizations. Overall, the key results of the TAcLE workgroups are: 1. Current knowledge justifies immediate and targeted effort to reduce external loading of

nutrients and sediment to Lake Erie from both agricultural and urban sources. Phosphorus

Page 8: Taking Action On Lake Erie - International Joint Commission · reduce external inputs of total phosphorus (TP) and in particular, the rising proportion of the most bioavailable form

vii

and especially SRP is of primary concern; but the roles of N and climate in modifying the lake’s response to P need to be resolved. The highest priority for management action should be the Maumee and Detroit River basins, followed by the Sandusky basin. Updated and revaluated loading estimates for the Detroit River and Canadian tributaries represent a second important priority need. Nevertheless, it is important to work quickly towards a cooperative a whole basin approach for long-term sustainable recovery.

2. To understand and manage expectations for recovery, there is a critical need for - urban & farm-based land use information (currently restricted by privacy legislation) to identify high-input sub-basin areas and reduce the high level of uncertainly in current modelled basin inputs - direct measures of internal P loading: current estimates are that that recovery will take 10-15 yrs even if external inputs are reduced in the near future – and longer, if management actions are further delayed

3. BMPs play a critical role in effective nutrient management yet current practices in the Erie basin have not been rigorously evaluated or applied. Most BMPs are used in combination with at least one other BMP and many are not designed for nutrient removal and may actually increase nutrient inputs. There is no single solution to non-point source pollution, particularly SRP and the use of a suite of BMPs is recommended along with more rigorous testing (especially for TP/SRP), standard metrics, and a central database. In view of the lack of information specific to Lake Erie, it is also recommended that a further BMP assessment could derive examples from other analogous water basins e.g. Lake Winnipeg, Lake Ontario.

4. Anoxia is a major issue in the central basin (CB) where it may exacerbate internal nutrient loading and present a threat to fish and other elements of the foodweb. The areal extent and severity of anoxia has been linked with overall TP inputs to the WB and CB, particularly the Detroit River (~ 50% of the total load).

5. Phosphorus is a key factor driving harmful algal/cyanobacterial blooms in Lake Erie, but it is clear that the severity and frequency of these events are significantly modified by other factors (notably N, climate and foodweb) which need to be incorporated into management models and actions. Evidence indicates: - Toxic cHABs in the west basin (WB) have been largely attributed to agricultural SRP inputs from WB tributaries (notably the Maumee but the direct / indirect causes of the increasing occurrence of toxic cHABs in other areas (e.g. North central, east basins) requires resolution. - Shoreline benthic algal impairment (HNABS) show two distinct patterns: the (non toxic) cyanobacteria Lyngbya is problematic in the WB where it responds to low light, high P and DOC; Cladophora is most severe in the less eutrophic central and east basins where its resurgence is most clearly linked with increased water transparency and shoreline /mussel nutrient inputs. In both cases, more science is urgently required to develop effective management actions.

6. Models have the potential to provide evidence-based goals for reduction of nutrient inputs to Lake Erie but to be of value, their limitations and level of uncertainly must be explicitly stated and critically evaluated.

Page 9: Taking Action On Lake Erie - International Joint Commission · reduce external inputs of total phosphorus (TP) and in particular, the rising proportion of the most bioavailable form

viii

7. Much research is required to fully understand the complex interactions between nutrients and effects on Lake Erie, and funds to support this are essential. This is particularly true given the range of possible changes in climate, including changes in ice-cover, precipitation patterns, lake levels and temperature and resulting effects on fish populations which have shown a significant shift in species composition and incidents of large-scale fishkills.

8. Overall, there is an urgent need for prompt action to reduce TP/DRP inputs to Lake Erie. Modelled response curves were generated to provide two achievable interim targets to guide management actions, pending improved data. These show that a - 23% reduction in Maumee River spring TP input is required to minimise WB cHABs - 53% reduction in Lake Erie annual TP load is required to minimise other cHABs and CB anoxia

9. It is further recognized that to allow for effective basin wide restoration and evaluation of

success, there is a critical need for - a standardized, targeted and adequate monitoring system which would incorporate the newly revised Indicators (SAB/IJC) as measures of progress. - public engagement through communication and workshops

April 10th 2013 Sue Watson, Canadian co-chair David Carpenter, US co-chair TAcLE Workgroup Members: Sue Watson, David Carpenter, John Casselman, Carol Miller, Mike Murray, George Arhonditsis, Nate Bosch, Greg Boyer, Murray Charlton, Remegio Confesor Jr, Dave Depew, Dave Dolan, Joe DePinto, Scott Higgins, Tomas Hook, Todd Howell, Stu Ludsin, Shawn P. McElmurry, Chitra Gowda, Peter Richards, Don Scavia, Ralph Smith, Morgan Steffen, Steve Wilhelm, Weitao Zhang GLRO Support: Raj S. Bejankiwar

Page 10: Taking Action On Lake Erie - International Joint Commission · reduce external inputs of total phosphorus (TP) and in particular, the rising proportion of the most bioavailable form

ix

Page 11: Taking Action On Lake Erie - International Joint Commission · reduce external inputs of total phosphorus (TP) and in particular, the rising proportion of the most bioavailable form

x

Table of Contents

Executive Summary i The IJC Lake Erie Ecosystem Priority Science Initiative: Taking Action on Lake Erie 1 Introduction 1 Taking Action on Lake Erie (TAcLE) 3 1. Current External, Internal P Loadings 4 1.1 External Loads 4 1.2 Current Status: Updated Lake Erie Phosphorus Loading Estimates: 2008-2011 4 1.3 Prioritizing Action: SWAT-based prediction of sub-basin loading sources within major 6 water basins 1.4 Atmospheric Deposition of Phosphorus 11 1.5 Internal Loading: Mechanisms and current understanding of its importance in Lake Erie 15 1.6 Lake Management Implications of Internal Loading/Recycling 18 1.7 Climate Change Implications on External TP/SRP Loadings 19 2. Effects of Agricultural and Urban Best Management Practices (BMPs) in TP/SRP 21 Loads to Lake Erie 2.1 Urban BMPs 21 2.1.1 Regulatory Framework 22 2.1.2 Selection of Urban BMPs 23 2.1.3 Non-Structural (Alternative Behaviour/Management) BMPs 23 2.1.4 Non-Point Source Structural or Engineered BMPs 24 2.1.5 Evaluation of Structural BMPs’ Treatment Efficiency 25 2.1.6 Cost of Structural BMPs 27 2.1.7 Urban Point Sources 29 2.1.8 Monitoring Urban BMPs 29 2.1.9 Urban BMPs’ Performance Metrics 29 2.2 BMPs in Agricultural and Rural Environments 30 2.2.1 Agricultural P-Source BMPs 30 2.2.1.1 Farm Gate Regulation 31 2.2.1.2 Managing P Application to Soil 33 2.3.1 P- Transport BMPs 34 2.3.1.1 Residue and Tillage Management (Conservation Tillage) 35 2.3.1.2 Conservation Cropping 35 2.3.1.3 Conservation Buffers 35 2.3.1.4 Wetlands (Constructed Wetlands) 36 2.3.1.5 Grassed Waterways 37 2.3.1.6 Drainage Water Management 37 2.3.1.7 Emerging Technologies 37 2.4 Effects of Extreme Weather 37 2.5 Focus on Management of SRP vs. PP 37 2.6 Current Policy and Implications 40 3. Hypoxia 40 3.1 Causes of Hypoxia 40 3.2 P Abatement Programs 41

Page 12: Taking Action On Lake Erie - International Joint Commission · reduce external inputs of total phosphorus (TP) and in particular, the rising proportion of the most bioavailable form

xi

3.3 Climate Change 41 4. Harmful Algal Blooms (HABs) in Lake Erie: Status, causes, and controls 42 4.1 HABs: Planktonic cyanobacteria 42 4.2 Current Status of cHABs in Lake Erie 42 4.3 Nutrients and Lake Erie HABs 43 4.4 Abiotic Factors, Climate Change, and cHABs 44 5. HNABs: Attached Algae 44 5.1 Overview 44 5.2 Abiotic and Biotic Factors Associated with NABs in Lake Erie 46 5.3 Western Basin (Lyngbya) 48 5.4 Eastern Basin (Cladophora) 49 5.5 Modelling in Support of Management 51 5.6 Potential Impacts on NABs Under Climate Change 52 5.7 NABs: Summary, conclusions, and recommendations for future work 55 6. Lake Erie Fish Populations: Status and impact of climate change 56 6.1 Current Status 56 6.2 Effects of Hypoxia 57 6.3 Effects of Water Clarity 59 6.4 Historical Patterns 60 6.5 Theoretical and Empirical Studies 61 6.6 Ecological Effects of Basal Prey Production 61 6.7 Future Interactional Effects of Climate Change 62 6.8 Conclusions 63 7. Lake Erie Coastal Wetlands and Implications for Nutrient Loading 64 7.1 Overview 64 7.2 Nutrient Cycling in Freshwater Wetlands 64 7.3 The Importance of Wetlands to Lake Erie Nutrient Loading 64 8. Modelling Lake Parameters in Relation to Nutrient Loading in Lake Erie 65 8.1 Past and Current Modelling Efforts: A critical review 65 8.2 Response Curves 68

8.2.1 Predictors for cHABs in Western Lake Erie 68 8.2.2 Response Curves Expressing the Relationship between TP/SRP Loads 71 and Central Basin Hypoxia and Chlorophyll Concentration 8.2.2.1 Model Calibration and Testing 72 8.2.2.2 Response Curves 75

List of Figures 1. Lake Erie: Annual external TP loading 4 2. SWAT-based estimated average annual sub-basin P yields from two West Basin watersheds 10 3. Average estimated phosphorus loadings for the Lake Erie basin (1996-2002) 15 4. Spectrum of urban structural BMPs 24 5. Treatment efficiency for 6 classes of urban structural BMPs 27 6. Average total facility costs in 2012 28 7. Estimated average annual maintenance cost 28

Page 13: Taking Action On Lake Erie - International Joint Commission · reduce external inputs of total phosphorus (TP) and in particular, the rising proportion of the most bioavailable form

xii

8. Locations where blooms of Cladophora or Lyngbya wollei have been recorded since 1995 45 9. Lake Erie substrate types 47 10. Kriged surface of turbidity along a site in Lake Erie near Wainfleet 50 11. Modelled seasonality of cumulative Cladophora biomass under scenarios of climate warming 53 12. Bloom intensity from 2002 to 2011 compared to TP spring load 68 13. A comparison of the TPq model presented against the Q model 69 14. Comparison of CI observations to the TP load model 69 15. CI estimates 70 16. The required level of BMP implementation to achieve load reduction from the Maumee River 71 17. Conceptual diagram of simple eutrophication model 71 18. Epilimnetic Chlorophyll and TP calibration results 72 19. Dissolved oxygen calibration results in Lake Erie Central basin 73 20. Water-column depletion rates compared to earlier estimates 74 21. Comparison of hypoxic area and average hypolimnetic DO concentration 74 22. August-September means and standard deviations of area 75 23. Modelled response curves as a function of annual Lake Erie TP load (DO & hypoxic area) 75 24. Modelled response curves as a function of annual Lake Erie TP load (DO & hypoxic days) 76

List of Tables 1. Outline of issues addressed under LEEP and associated workgroups and outcomes 2 2a. Estimated annual external TP loads to Lake Erie 5 2b. Dissolved Reactive Phosphorus Loads to Lake Erie 5 3a. Calibration and validation for monthly TP loads for 6 modelled watersheds 7 3b. Calibration and validation for monthly SRP loads 7 4. Average P loading (1998-2005) and % SRP/TP for 7 major Lake Erie drainage basins 1988-2005 8 5. Potential sources of atmospheric phosphorus deposition to Lake Erie 12 6. Climate change scenarios used in SWAT modelling 19 7. Modelled climate change influence on Lake Erie rivers 20 8. Actions associated with P-source BMPs 31 9. Nutrient reduction efficiencies of animal waste system BMPs 32 10. Summary of nutrient management efficiencies in reducing pollutant loss 35 11. Summary of measured warming rates offshore surface waters and air temperatures 52 12. Summary of modelled scenarios to evaluate response of LE Cladophora to climate warming 53 13. Bloom categories and associated Maumee spring TP loading levels 70

Page 14: Taking Action On Lake Erie - International Joint Commission · reduce external inputs of total phosphorus (TP) and in particular, the rising proportion of the most bioavailable form

1

The IJC Lake Erie Ecosystem Priority Science Initiative: Taking Action on Lake Erie Introduction The Laurentian Great Lakes are one of the world’s most valuable natural resources, representing some 20% of the global freshwater supplies, a resource that is expected to become increasingly valuable in the near future (Schottler and Eisenreich 1994). Lake Erie alone provides millions of dollars in revenue each year from tourism and fishery industries (Box 1). It is the smallest and shallowest of the five Laurentian Great Lakes and currently the most severely threatened by human activities. Over the past decade, Lake Erie (LE) has undergone a significant decline in nearshore and offshore water quality, with intensified hypoxia in the bottom waters of the central basin1, extensive blooms of harmful and nuisance algae (HNABS)2 and shifts in the composition and resilience of the foodweb. These issues have had severe and costly socioeconomic impacts for example on drinking water supplies, recreational and tourist industries and property values. It has become apparent that there is an urgent need to develop bi-national policy to arrest this decline and work towards sustained restoration and management - and apply the lessons learned to the other Great Lakes, many of which are also showing signs of decline. However, despite a significant body of research and monitoring on this and other Great Lakes, some fundamental knowledge gaps preclude the development of robust policy that would define a clear action plan to address these issues. In recognition of the urgency and importance of this issue and the potential cost of further delay, in 2012 the IJC commissioned the Lake Erie Ecosystem Priority (LEEP) to evaluate current conditions, identify knowledge gaps and develop modelled scenarios to provide guidance for developing policy and management targets in the face of climate change. In keeping with the need for immediate and rigorous action, the desired goals are in three years, to have i) measurably reduced dissolved reactive phosphorus loads and HNABs in Lake Erie, ii) a better scientific understanding of causes and controls of these and related impairments, iii) an adequate monitoring system in place, iv) improved coastal resiliency and governance, and v) better public understanding and support.

1 some degree of central basin hypoxia is natural; see section.

2 ‘Harmful’ refers to blooms with the potential to produce (lethal or non-lethal) toxins that affect human and animal

health, which in the Great Lakes and most other freshwaters are exclusively produced by certain species of largely

planktonic cyanobacteria ’Nuisance’ refers to a broader subset of [algae and cyanobacteria blooms] which are non

toxic to humans but which cause ecological and socioeconomic harm; these more frequently refer to attached algae

like Cladophora and Lyngbya

Page 15: Taking Action On Lake Erie - International Joint Commission · reduce external inputs of total phosphorus (TP) and in particular, the rising proportion of the most bioavailable form

2

The LEEP strategy is aligned with the IJC Nearshore Framework, i.e. i) research and assessment of conditions; ii) diagnosis of problems; iii) prioritization of restoration and stewardship actions; iv) adaptive management and reporting based on ecosystem response to actions. It recognizes the importance of climate change, and the need to identify the major drivers of climate change (e.g. temperature, precipitation) and how these will affect the ecosystem response to management actions. The overall workplan combines Science and Socioeconomic themes into a comprehensive treatment of the issue, with each component addressed by a series of working groups led by different IJC Advisory Boards, Councils, staff and management teams (Table 1).

Table 1: Outline of issues addressed under LEEP and associated workgroups and outcomes.

Theme Workgroup Topic Outcome

1. Science SAB: TAcLE Current external, internal P loading Draft review papers

Effects of climate change on P loading, HNABs, wetlands & fish

Draft review papers

Effectiveness of agricultural and urban Best Management Practices

Draft review papers

P and HAB management models and targets

Draft review papers

CGLM Effectiveness of LE monitoring Draft review papers

2. Socio Economic

WQB Legislative/Regulatory Framework Draft review papers

IJC staff & management Costs/ Benefits/ Barriers/ Incentives of Inaction or Action(s)

Draft review paper

Recommended Social/Economic Solutions

Draft review paper

Public/stakeholder outreach & engagement

Open houses/ workshops/ printed and electronic media etc

The IJC Science Advisory Board (SAB) was tasked with all but one of the issues under the Science theme. These have been addressed under framework of the Taking Action on Lake Erie (TAcLE) by four teams, composed of SAB members and contracted and consulted experts

Box 1: Annual Lake Erie Related Revenue Estimates

Tourism $7.4 billion

Seaports $1 billion

Sports fisheries $520 million

Commercial fisheries $25.8 million

Drinking water supply to 3127 water treatment plants

(e.g. USDA NRCS 2005 Western Lake Erie Basin Water Resources Protection Plan; Kranzberg G and

DeBoer C. 2006 A Valuation of Ecological Services in the Great Lakes Basin Ecosystem to Sustain

Health Communities and a Dynamic Economy OMNR report.)

Page 16: Taking Action On Lake Erie - International Joint Commission · reduce external inputs of total phosphorus (TP) and in particular, the rising proportion of the most bioavailable form

3

in the field, and has produced a series of white papers. Initial drafts of these reports were reviewed by the SAB and a broader group of expertise at a two-day LEEP workshop in Windsor, in February 2013. This final report synthesizes the key finding of this effort to provide the IJC with the best available knowledge on which to make recommendations for policy and action. Taking Action on Lake Erie Lake Erie is the southernmost, warmest and most vulnerable of the Laurentian Great Lakes. With a large proportion of its surface area shallower than 15m, LE consistently reaches temperatures above 25° C during summer months (Burns et al. 2005; Stumpf et al. 2012). Despite decades of international efforts to reduce nutrient loading, LE not only continues to receive substantial nutrient inputs from agricultural and urban/industrial sources, but these have shifted from point-source (‘end of pipe’) to diffuse (e.g. edge of field) in nature, which are far more difficult to both identify and control. External phosphorous (P) loading to the western basin (WB), primarily via the Maumee River, has been identified as the largest controllable cause of planktonic cHABs and exacerbated hypoxia (e.g. Stumpf et al 2012 and references therein). Importantly, although total P (TP) inputs have not risen significantly in the past few decades, a documented increase in the proportion of bioavailable phosphorus in this loading (SRP3) has been ear-marked as a major factor driving the density and extent of the annual summer planktonic blooms in the west basin. Nitrogen is also a major component of these external loadings which can influence the species composition (and potential toxicity) of the blooms, and perhaps increase algal biomass, but from a management perspective is considered of secondary importance for action. Nevertheless, as discussed below (section 5), a constrained focus on west basin P inputs is unlikely to address the second significant type of algal impairment, from attached (i.e. benthic) nuisance algal blooms (NABs). The influence of climate change is already manifest as reduced ice and snow cover, warmer waters, changes in runoff, mixing and circulation, and increases in severe storm events, all of which have a profound influence on the ecosystem integrity and its resilience to high external loadings and other human impacts. Continued population growth, basin development and pressure on the ecosystem services provided by this and the other great lakes and In response to the critical need for prompt action, based on the best available information on how these inputs will change with time, how to develop the most effective policies to reduce these inputs, and how the lake will respond in terms of water quality, nutrient levels, hypoxia, HNABs and other important foodweb components.

3 referred to interchangeably as soluble reactive P (SRP) or dissolved reactive P (DRP)

Page 17: Taking Action On Lake Erie - International Joint Commission · reduce external inputs of total phosphorus (TP) and in particular, the rising proportion of the most bioavailable form

4

1. Current external, internal P loading 1.1 External Loads

To facilitate prioritized management action, the goals of this LEEP component were to provide

the best available updated estimates of the phosphorus (total and SRP)

tributary/watershed/basin load for LE through 2011, and evaluate the relative contributions of

these sources to total loading. This and the response curve modelling work (section D4) was

carried out as part of a larger EcoFor initiative4

1.2 Current status: updated Lake Erie Phosphorus Load Estimation: 2008 – 2011 TP and SRP loads for LE from 2003-2008 were updated through to 2011 using data sources and methods outlined in the full report, and draft annual loads were derived subject to update with additional data (Table 2). Based on these estimates, monitored tributaries contributed the largest fraction

4 http://sitemaker.umich.edu/ecoforelake.erie/the_forecasting_models

Box 2A: External TP Loading Annual Mass Balance (2003-2011)

Monitored tributaries contributed the largest fraction (53%) of the total annual P load.

Unmonitored areas and direct point sources each accounted for ~16%.

Other sources each accounted for between 4-6%.

Annual variability in total load is largely driven by variability in monitored tributary input (± 24%).

Page 18: Taking Action On Lake Erie - International Joint Commission · reduce external inputs of total phosphorus (TP) and in particular, the rising proportion of the most bioavailable form

5

Table 2B. Dissolved Reactive Phosphorus Loads to Lake Erie (metric tons per year)

Year CY 2005 CY 2007 WY 2008 WY 2009 WY 2010 WY 2011 Mean (%CV)

Direct Point Source 975 1067 1047 1027 839 884.65 949 (11%)

Indirect Point Source 344 324 333 338 403 373.8 361 (9%)

Atmospheric 127 151 173 173 173 172.55 173 (0.1%)

Lake Huron Input 189 155 155 155 155 155 155 (0.1%)

Tributary Monitored Less Indirect

801 1139 1253 785 380 802.32 805 (44%)

Adjustment for Unmonitored Area

244 278 395 272 372 272 303 (20%)

Total: 2679.108 3555.95 3355.95 2749.16 2221.65 2660.32 2747 (17%)

Notes: WY= Water year; CY= Calendar year. Average calculated from WY 2008-2011; estimates subject to update

from data; no estimates for 2006.

(53%) of the total annual P load, unmonitored areas and direct point sources each accounted for ~16% and other sources each accounted for on average between 4-6%. No long term trends were evident over the period examined (Figure: Box 2A). Average TP and SRP concentrations in the Detroit River5 are low compared to those measured in other major tributaries and it has been concluded that this river does not contribute greatly to the WB HABs. In contrast, the P concentrations in the rivers draining the agricultural areas of the WB (e.g. Table 2-3), particularly the Maumee, are generally much higher and may have a

5 Estimated by summing all point sources, tributaries, and unmonitored areas upstream of the Detroit River mouth

along with an estimated inflow from Lake Huron (Chapra and Dolan, 2012).

Table 2A. estimated annual external TP loads to Lake Erie (metric tons / yr)

Water Year 2003 2004 2005 2006 2007 2008 2009 2010 2011 Mean (± %CV)

Direct Point Source 1257 1410 1445 1475 1518 1496 1467 1199 1263 1392 (± 8.6%)

Indirect Point Source 432 484 466 492 487 476 483 575 534 492 (± 8.3%)

Atmospheric 813 511 363 632 432 493 493 493 493 525 (± 24.6%)

Lake Huron 366 364 354 335 325 321 321 321 321 336 (± 5.8%)

Tributary Monitored 3970 4427 5162 4476 7066 6196 4498 2550 4609 4773 (± 24.0%)

Unmonitored Area 1132 1736 1451 1396 2080 1693 1229 683 1297 1411 (± 28.4%)

TOTAL 7970 8932 9241 8806 11908 10675 8491 5821 8517 8929 (± 19%)

Notes: draft annual loads lack the following data: 2009: atmospheric, ON Industrial Pt Sources; 2010: atmospheric,

ON Industrial Pt Sources, Michigan monitored tributaries; 2011: atmospheric, ON Industrial Pt Sources, Michigan & Ontario monitored tributaries;

Page 19: Taking Action On Lake Erie - International Joint Commission · reduce external inputs of total phosphorus (TP) and in particular, the rising proportion of the most bioavailable form

6

more direct influence on the development of HABs. Nevertheless, the Detroit River represents the largest hydrological loading to Lake Erie and, by virtue of it flow, contributes significantly to the total loading (estimated ~ 40-50%) which may be a significant factor in the annual hypoxia. 1.3 Prioritizing action: SWAT based prediction of sub-basin loading sources within major WB

The above mass balance estimates, derived only for TP, show non-point sources as the most important external inputs. In order to prioritize management actions it is important to take this a step further and identify the highest contributors of both TP and SRP, given the currently held belief that HABs are primarily fuelled by the more biologically available dissolved fraction. This was carried out for water years 1998-2005 using SWAT6 models to estimate the sub-basin contributions of TP and SRP for six major US catchment areas draining to WB of LE: the Maumee, Sandusky, Grand, Raisin, Cuyahoga, Vermillion and Detroit, the latter representing a combined load from Lake Huron and the mixed urban-agri watersheds. Modelling efforts have several sources of uncertainty which need to be considered when interpreting results and deciding on management implications. The modelling efforts using SWAT in the current work likewise have quantified and unquantified sources of uncertainty that are discussed here: the input data used to parameterize the models (see Bosch et al. 2011) (e.g. precipitation, temperature, land use classes, soil types, topography, atmospheric deposition, tile drain coverage, current agriculture land management, etc) ii) field sampling and lab analysis used in calibration of TP and SRP iii) quantified uncertainty of the model as evaluated during the calibration and validation steps of model development (Table 3). Finally, there is uncertainty associated with the climate scenarios chosen for the current modelling based on climate model uncertainty and diverse future climate predictions.

6 process based temporally resolved hydrologic models widely applied in the USA by y the EPA and other

US agencies for e.g. TMDL analyses; see http://swat.tamu.edu/

Box 2B: External Loading Modelling Study – Highlights

Key Results: modeled outcome suggests i) “hot spot” source areas within the major

watersheds which contribute a large proportion of the P inputs; ii) climate change impacts

difference parameters and watersheds differently and in often sporadic (difficult to explain)

ways

Recommendations: i) target BMP implementation programs and nutrient source reduction

efforts in high loading sub-basins; ii) consider watersheds and parameters individually when

considering policy for climate change adaptation

Emerging Science Questions: i) how much would targeted implementation of BMPs impact

the load at the river mouth? Ii) Why do we see these sporadic effects of climate change?

Lack of data means that the models cannot answer these questions

Page 20: Taking Action On Lake Erie - International Joint Commission · reduce external inputs of total phosphorus (TP) and in particular, the rising proportion of the most bioavailable form

7

Yields calculated by SWAT represent material transfer from land to stream channel, not the levels discharged at the tributary mouth, and may include a considerable margin of uncertainty. SWAT models are ‘data hungry; but because finely resolved data are often lacking (as is the case for much of the LE watershed), output is generated from a combination of available and extrapolated data. The modelled output is influenced by the scale of data resolution which can misrepresent high-low flow events, soil heterogeneity, etc (e.g. Beven 2006; 2007). With this uncertainly in mind, the results of the models should be interpreted with caution. The SWAT modelling exercise delivered a series of maps and tables (detailed in TAcLE report) showing the highest P-yielding sub watersheds within each drainage basin, two of which (Maumee and Sandusky, which have the highest total load) are shown in Figs. 2-3. Major outcomes were: i) Overall, the best results were obtained from agriculturally dominated watersheds (i.e. Maumee, Sandusky), for which SWAT models were originally designed. Performance for more

Table 3A. Calibration and validation for monthly TP loads (Mg P) for 6 modeled watersheds. Coefficient of

determination (R2) and Nash–Sutcliffe simulation efficiency (NSE) used as evaluators of model performance.

Statistics in bold type categorized as satisfactory or better.From Bosch et al. 2011.

(a) Calibration (b) Validation

Watershed

Obs. mean (Mg P)

Simul. mean (Mg P) R

2 NSE

Obs. mean (Mg P)

Simul. mean (Mg P) R

2 NSE

Huron 1 1 0.53 -2.55 1 3 0.16 -46.12

Raison 11 8 0.52 0.47 9 8 0.50 0.49

Maumee 149 149 0.72 0.72 182 137 0.86 0.74

Sandusky 23 24 0.66 0.64 43 30 0.87 0.77

Cuyahoga 14 14 0.41 0.11 29 16 0.68 0.39

Grand 6 7 0.71 0.68 13 9 0.30 0.28

Table 3B. Calibration and validation for monthly SRP loads (Mg P) ,

(a) Calibration (b) Validation

Watershed

Obs. mean (Mg P)

Simul. mean (Mg P) R

2 NSE

Obs. mean (Mg P)

Simul. mean (Mg P) R

2 NSE

Huron 1 1 0.37 -1.12 1 3 0.75 -182.45

Raison 11 8 0.81 0.79 9 8 0.70 0.70

Maumee 149 149 0.65 0.61 182 137 0.69 0.29

Sandusky 23 24 0.43 0.44 43 30 0.45 0.35

Cuyahoga 14 14 0.05 -38.26 29 16 0.11 -34.31

Grand 6 7 0.53 0.50 13 9 0.75 -182.45

Notes: R2: fraction of variance in data explained by model; is biased by outliers; NSE=1: perfect fit. NSE>0.5:

satisfactory to good fit; NSE<0: observed value is a better predictor than simulated value i.e. unacceptable model performance (Moriasi et al 2007).

Page 21: Taking Action On Lake Erie - International Joint Commission · reduce external inputs of total phosphorus (TP) and in particular, the rising proportion of the most bioavailable form

8

urbanized LE basins was poor, particularly for Huron where the modelled results were worse that the observed data. ii) The model performance varied considerably among the six LE basins and was particularly weak for the dissolved phosphorus fraction (SRP) (as was also the case for dissolved N7). Based on NSE values, only the Maumee and Sandusky sub-basin modelled TP output showed an acceptable level of uncertainty, with the Raison as marginal (Table 3A). For SRP, a reasonable level of confidence could only be assigned to the modelled output for the Raison (Table 3B). iii) SWAT models were able to simulate variance in flow and sediment reasonably well, but failed to capture the dynamic range of the nutrient export - particularly for extreme episodic events (Figs 3 and 5 in Bosch et al. 2011). This is of critical importance, since these high flow events can contribute a significant proportion of the total annual export, accounting, for example, for up to 50% of the TP exported into Hamilton Harbour at the west end of Lake Ontario (Kellershohn and Tsanis 1999). For example up to 50% of P inputs into these Erie tributaries may come from tile drainage systems which are not parameterized by SWAT modelling, which deals only with surface runoff.

Table 4: Average P loading (1998-2005) and % SRP/TP for 7 major Lake Erie drainage basins 1988-2005 estimated

as edge of field output

Total Loadings (Kg P/ ha/yr)

Tributary # Subwater-sheds

P [fraction] Average load (%) SRP/TP

± stdev Max load

Maumee 203 TP 372.0 60.4 489.2

SRP 4.8 (1.3%) 1.4 7.1

Sandusky 39 TP 51.6 19.7 84.9

SRP 0.5 (1%) 0.2 0.9

Cuyahoga 23 TP 45.9 12.5 64.9

SRP 2.1 (4.5%) 0.6 3.2

Grand 22 TP 35.7 8.1 50.4

SRP 0.5 (1.4%) 0.1 0.7

Raison 32 TP 34.7 7.4 50.6

SRP 3.5 (10.1%) 0.6 4.3

Huron 31 TP 13.5 5.0 21.4

SRP 1.1 (8.2%) 0.4 1.8

iv) Average and maximum total TP and SRP load estimates are summarised in Table 4, which shows by far the greatest areal and total TP loading from the Maumee (372 ± 60.4 kg P/ha/yr) which also showed the highest SRP loading. It should be noted, however, that while the

7 TN and nitrate SWAT modelling results not provided for this report but presented in the full paper from which this

work was derived (Bosch 2011)

Page 22: Taking Action On Lake Erie - International Joint Commission · reduce external inputs of total phosphorus (TP) and in particular, the rising proportion of the most bioavailable form

9

average estimated total P loads from other watersheds were significantly lower, the fraction of this load delivered as SRP was ~3-7 times higher in 3 out of 5 of these basins, notably the Raison, which had estimated annual SRP loadings equivalent to 70% those of the Maumee (Table 5). The timing of these loadings is important, since this has a significant influence on the response of the biological community and risk of HABs. Summarised results for average annual yields from each watershed (provided in more detail in the full reports) were: i.) Huron: SRP and TP yields of up to 0.099 and 1.38 kg P ha-1 yr-1, resp.; high-loading areas

generally located in the central part for TP, but in the lower regions for SRP. ii.) Raison: SRP and TP sub-basin yields range between 0.016 - 0.231 and 0.10-3.00 kg P ha-1

yr-1, respectively, with high-loading areas generally located in the central part of the watershed for both SRP and for TP.

iii.) Maumee watershed SRP and TP yields range between 0.006 - 0.050 and 0.18 - 4.20 kg P ha-1 yr-1, respectively. The SWAT model predicts higher TP loads from areas in the upper Maumee watershed (north-eastern and southern areas) while in contrast, high SRP-loading regions occur irregularly throughout the watershed (Fig. 2).

iv.) Sandusky SRP and TP yields between 0.009 - 0.025 and 0.10 - 2.57 P ha-1 yr-1 respectively show different spatial patterns with no clear geographical zonation (Fig 2).

v.) Cuyahoga SRP and TP yields between 0.026 - 0.145 and 0.63 - 3.85 kg P ha-1 yr-1 respectively with high-loading sub-watersheds mostly located in the upper part of the watershed for both TP and SRP.

vi.) Grand TP and SRP yields range from 0.23 - 2.69 and 0.011 - 0.034 kg P ha-1 yr-1, respectively; high-loading subwatersheds for TP are identified in upper parts of the watershed, and the central portion of the watershed for SRP.

In particular, two of these outcomes should be more fully investigated, and point to the need for better data and alternative approaches: i) the modelled SRP and TP yields vary considerably within and among each of the watersheds; ii) in many cases, there was little or no spatial overlap in regions delivering high TP and SRP loading. For example, high SRP yielding sub-basins in the upper reaches of a watershed may have less immediate impact on Lake Erie than those in the lower draining areas, due to biological uptake and transformation in stream channels. The effects of particulate loading will depend on the form it takes (suspended sediment and living/dead organic material) and flow/sedimentation/ transformation during stream transit.

Page 23: Taking Action On Lake Erie - International Joint Commission · reduce external inputs of total phosphorus (TP) and in particular, the rising proportion of the most bioavailable form

10

Figure. 2. SWAT-based estimated average annual sub-basin P yields for A: SRP, B:TP from two of the six

major West Basin watersheds modelled for Lake Erie: Maumee (top) and (Sandusky (bottom)

Page 24: Taking Action On Lake Erie - International Joint Commission · reduce external inputs of total phosphorus (TP) and in particular, the rising proportion of the most bioavailable form

11

1.4 Atmospheric Deposition of Phosphorus Atmospheric deposition to lakes is difficult to quantify and measure but may represent a significant contribution to the total external load. Phosphorus can find its way from the airshed into lake ecosystems via inputs to the watershed from 1) rain or snowfall, known as wet deposition; and (2) wind transported particles, or dry deposition (Anderson & Downing, 2006; Zhai et al., 2009). Wet and dry nutrient loadings to aquatic ecosystems, particularly phosphorus, have increased over the years as a direct result of anthropogenic activities (Herut et al., 1999; Zhai et al., 2009; Winter et al., 2007). This study comprised a review of available literature on atmospheric phosphorus deposition to obtain a more complete description of the sources and transfer mechanisms. Lakes of all sizes were included though relevance to the Great Lake Basin, especially Lake Erie, was a primary consideration. Research examining both wet and dry deposition was rare largely because direct measurements of wet deposition of phosphorus are especially difficult to collect and prone to method contamination. A limited number of studies identified levels of SRP within the wet or dry fractions of atmospheric phosphorus delivered to aquatic systems, but comprehensive studies are not available for the Great Lakes basin. This synthesis was undertaken to better describe current understanding of atmospheric phosphorus inputs to Lake Erie and its potential impact on the occurrence of harmful algal blooms within the Great Lakes. Dust can travel great distances from its source location depending on its particle size and eventually deposited into water bodies. In addition to phosphate rock particulates, numerous natural and anthropogenic sources of P are available for wind transport (Table A1). Due to the difficulty in measuring atmospheric P, the significance of specific sources can only be estimated (Brown et al., 2011; Dolislager et al., 2012). Such estimates are highly context dependent, with local sources of P having a pronounced influence than those transported from outside of a watershed (Dolislager et al., 2012). In one example in Lake of the Woods and the Rainy River, it was recognized that the largest source of atmospheric P was dust aerosols (Hargan et al., 2011), while anthropogenic urban, industrial, and agricultural sources were found to be the dominant atmospheric sources to Lake Michigan (Eisenreich et al. 1977). Natural sources generally contribute less to total atmospheric load than anthropogenic sources (Hargan et al., 2011). Generally, P originating from outside a basin is attributed to true atmospheric deposition while sources from within the basin are viewed as contamination or local recycling rather than true deposition (Ahn & James, 2001). The importance of these local versus external sources to atmospheric deposition to lakes is dependent on air circulation patterns between the sources relative to the lake of interest.

Weather conditions can produce significant fluctuations in atmospheric P levels. Significant regional, annual and seasonal variability has been observed in atmospheric particulate P (PP) concentrations (Tsukuda et al., 2004; Ahn & James, 2001). A maximum loading in the summer and spring periods and lowest rates in the winter season were found at several locations, including Lake Simcoe (Brown et al., 2011), South Central Ontario (Hargan et al., 2011), and Lake Michigan ( Eisenreich et al., 1977). While agricultural activities and mining sites have attracted the most interest, the EPA has identified unpaved roads, aggregate storage piles, and heavy construction as additional potential sources of particulate matter ( US EPA, 2000).

Page 25: Taking Action On Lake Erie - International Joint Commission · reduce external inputs of total phosphorus (TP) and in particular, the rising proportion of the most bioavailable form

12

Table 5: Potential sources of atmospheric phosphorus deposition to Lake Erie. Estimates of relative

significance are based on presence/absence of a source in the watershed, anthropogenic sources recognized as most dominant.

Source Significance to Lake Erie

Reference(s) (see full report for bibliography)

Natural

Erosive process generating dust aerosols

High Herut et al., 1999

Micorbial decomposition of sewage sludge, landfill and compost heaps

High Winter et al., 2002

Pollen Moderate Winter et al., 2002; Report on the Phosphorus Loads to Lake Simcoe 2004-2007, 2009; Blake & Downing, 2009

Vegetation detritus Moderate Dolislager et al., 2012

Biomass burning due to lighting Low Grimshaw & Dolske, 2002

Insects and their guano Low Tsukuda et al., 2004

Ocean aerosols Low Ahn & James, 2001; J. Luo et al., 2011

Volcanic activity Low Tsukuda et al., 2006, pdf, n.d.; Winter et al., 2002

Anthropogenic

Coal combustion High Winter et al., 2002; Hartmann et al, 2008; Report on the Phosphorus Loads to Lake Simcoe 2004-2007, 2009;

Biomass burning High Morales et al., 2001; Winter et al., 2002; Zhang et al., 2002; Tamatamah et al., 2005; Tsukuda et al., 2006; Blake & Downing, 2009; Dolislager et al., 2012

Aeolian re-suspension of phosphorus from distributed soils, including quarries, deforested areas, agricultural fields, soil tiling and ploughing, and unpaved roads

High Herut et al., 1999; Morales et al., 2001; Grimshaw & Herut et al., 1999; Winter et al., 2002; Koelliker et al, 2004; Tsukuda et al., 2006; Report on the Phosphorus Loads to Lake Simcoe 2004-2007, 2009; Blake & Downing, 2009; Brown et al., 2011; Dolislager et al., 2012

Urban inputs primarily associated with automobile combustion emissions exhaust

High Conor et al., 1992; L Luo et al., 2007; Dolislager et al., 2012

Industrial activities such as fertilizer production and coal combustion

High Grimshaw & Dolske, 2002; Koeliker et al., 2004; Tsukuda et al., 2006; Brown et al., 2011; L. Luo et al., 2011

Atmospheric P deposition is derived mainly from particulate and aerosol materials found in dry deposition ( Zhai et al., 2009), and is generally reported to be 4 -10 times greater than wet deposition ( Migon & Sandroni, 1999; Ahn & James, 2001). Contact between raindrops and particulates or aerosols results in the dissolution of some of the particulate inorganic P as highly

Page 26: Taking Action On Lake Erie - International Joint Commission · reduce external inputs of total phosphorus (TP) and in particular, the rising proportion of the most bioavailable form

13

bioavailable phosphate (PO48), which may account for a significant fraction of the TP in wet

deposition (e.g. up to 13% ; Ridame & Guieu, 2002). Anthropogenic sources of PP release a greater proportion of PO4 into water than natural sources ( Mackey et al., 2012), particularly from areas devoted to agriculture (Anderson & Downing, 2006). Data for both wet and dry atmospheric phosphorus deposition is not available for the Lake Erie basin or elsewhere in the Great Lakes basin; however, the land surrounding Lake Erie is the most highly urbanized and highly farmed of all the Great Lakes. Inferences on the relative importance of atmospheric phosphorus deposition in Lake Erie rely on studies that predominantly measured bulk deposition, or a single pool of atmospheric P (wet or dry). While measurements of bulk atmospheric deposition encompasses both wet and dry pools, the relative abundance of bioavailable P in each pool cannot be determined conclusively based on bulk deposition measurements alone. The contribution of atmospheric P to Lake Erie has been generally attributed limited importance. Recent results from Lake Simcoe in Ontario indicate that there is reason to revisit such assumptions, given the shifts in population and land use surrounding the Lake Simcoe. As part of the Great Lakes Basin, Lake Simcoe is one of the largest lakes in Canada outside the Laurentian Great Lakes. Although it has a much smaller volume, Lake Simcoe is similar to Lake Erie in its proximity to major urban and farming centres, implying that its airshed is subject to similar impacts from human activity. Recent direct monitoring of atmospheric P has provided measurements of bulk TP deposition and shown that this may contribute up to 25-50% of the TP loadings to Lake Simcoe (Winter et al., 2007). The available information for Lake Erie indicates that atmospheric P deposition is a small but not inconsequential contribution to total P in the system (Table A1). Previous efforts to quantify regional P loading into Lake Erie have monitored TP rather than specifically targeting SRP9, and found that atmospheric deposition contributed higher levels of TP than direct industrial discharge for 1996-2002 (Dolan & McGunagle, 2005). Annual loads for the atmospheric fractions of TP were considered a minor component for the years for which data are available between 1974 and 2007 (Figure 2). However, unlike the most recent measurements of bulk deposition for Lake Simcoe, these estimates were developed based on reported precipitation (rain and snow) in the area and therefore represent only a fraction of bulk deposition, namely wet deposition of TP. Using the same methods, the most recent estimates of % contribution of atmospheric wet deposition of TP to Lake Erie ranged from < 10% for 1996-2002 (Dolan & McGunagle, 2005) and 7.1% for 2002-2006 (Baker, 2011). While previous work indicates this same approach would characterize a significant fraction of SRP (from atmospheric wet deposition), these estimates omit dry deposition from reported values and thus do not encompass a total measure of bioavailable atmospheric P. A better understanding of the sources of particulates to the Great Lakes, and their role in nutrient budgets for the Lakes is needed to quantify atmospheric contributions of P. Furthermore, the potential influence of farming practices and other basin uses on atmospheric particles should be evaluated. For example, higher P retention in agricultural surface soils as a result of no-till practices could provide a source of high-nutrient particulates to the atmosphere that could travel great distances and impact the Great Lakes basin. Delivery pathways and

8 Usually considered synonymous with SRP

9 PO4 is more difficult to monitor due to its rapid turnover and loss from collector samples held over extended periods

Page 27: Taking Action On Lake Erie - International Joint Commission · reduce external inputs of total phosphorus (TP) and in particular, the rising proportion of the most bioavailable form

14

constituents of atmospheric P deposition may have special relevance for supporting primary production in oligotrophic interior portions of other Great Lakes, while the contribution of atmospheric P to Lake Erie’s TP load is currently understood to be relatively minor. However, while inputs of atmospheric phosphorus to Lake Erie may be low, they are not negligible. In summary, the sources of atmospheric P in the Great Lakes are not well understood, with wet deposition more thoroughly monitored and reported than dry deposition. There is a significant need for more detailed direct monitoring of both wet and dry fractions of atmospheric deposition. Binational coordination of monitoring efforts is valuable from both a scientific and budgetary perspective, however efforts to track atmospheric P are lacking in existing monitoring programs such as the Integrated Atmospheric Deposition Network and the Great Lakes Precipitation Network. These existing monitoring networks could be augmented to include measurements of wet and dry atmospheric P deposition. Coordination with existing nutrient monitoring efforts is an alternative, and a binational nutrient strategy approach could be considered similar to the existing Great Lakes Binational Toxics Strategy. The October 2012 notification of Environment Canada’s GLNI is the most recent effort supporting advancements in research and management for nutrients in the Great Lakes, and inclusion of atmospheric P measurements in such an initiative would provide a much needed opportunity to expand our understanding of the P cycle. The question surrounding levels of bioavailable P within both wet and dry fractions is also unresolved. Concerns related to seasonal sampling frequency and measurement techniques need to be addressed, and again binational coordination would be essential to establishing regional capacity to expand monitoring programs to include an atmospheric component. Given the importance of developing effective management frameworks for phosphorus inputs to the Great Lakes that reflect the current state of human impacts on nutrient cycling, phosphorus budgets and nutrient models should be examined and improved across the region, especially as such models are modified to examine regional effects of climate.

Page 28: Taking Action On Lake Erie - International Joint Commission · reduce external inputs of total phosphorus (TP) and in particular, the rising proportion of the most bioavailable form

15

. 1.5 Internal Loading – Mechanisms and Current Understanding of its Importance in Lake Erie There are three main types of internal P cycling in Lake Erie: i) inter-basin exchange: Much of the external P load is delivered to the WB, where a portion is stored but much is transported to the Central and East basins (CB,EB). There is currently insufficient information on inter-basin transfers in LE to evaluate the magnitude of these exchanges and whether they have changed. Dreissenid mussel veligers represent a relatively new sedimentation vector in LE that might affect inter basin transfers although the net effects of the new species may not be easy to discriminate from other recent changes (see below).

Figure 3. Average estimated phosphorus loadings (MTA), for the Lake Erie basin (1996-2002); from Dolan and

McGunagle (2005). Data was obtained from government databases, updated based on monitoring site results from

New York, Pennsylvania, Ohio, Indiana, Michigan, and Ontario. Breakdown of TP loading to Lake Erie by sub-

basin is available in Dolan and McGunagle (2005), with the Western Basin generally contributing 75% of TP

loadings each year. Data includes estimates for contributions from Lake Huron and unmonitored areas. Direct

sources are point sources that discharge into Lake Erie or St. Clair, the Detroit or St. Clair Rivers, or unmonitored

areas.

Page 29: Taking Action On Lake Erie - International Joint Commission · reduce external inputs of total phosphorus (TP) and in particular, the rising proportion of the most bioavailable form

16

ii) biological processing: Uptake and release recycles the bioavailable fraction (SRP and dissolved organic P) between the biota and the water column. This involves numerous interconnected processes which are almost impossible to characterise individually, e.g., uptake and excretion by bacteria, algae, aquatic plants, zooplankton, fish, birds and macrobenthos and microbenthos; death, decay, sedimentation and bioturbation. Early empirical models (e.g. Vollenweider, 1968) used correlations between TP and algal biomass (usually expressed as chlorophyll-a (chla)) to derive P management targets. The unknown myriad details of internal P recycling were integrated as a ‘sedimentation coefficient’ and the models were largely successful, albeit with rather large uncertainty around predictions. However, fundamental changes in the trophic structure of the LE biological community since then has caused major changes in biological sequestration and exchange. The recent resurgence of HNABs and other issues have stimulated interest in resolving major underlying processes of P biological recycling, to allow a comparison between the effects of load variations and internal processes. Perhaps one of the most widely acknowledged impacts is seen from the widespread colonization by dressenid mussels, which have profoundly altered the light and nutrient regimes, and the foodweb. Dreissenids have changed internal P recycling and altered the inshore-offshore exchange of materials and nutrients, trapping these in the warmer and shallower nearshore zones by what is termed the ‘nearshore shunt’ (Hecky et al 2004). Zhang et al (2011) modeled dressenid P uptake and excretion and concluded that the lake may be more sensitive to P inputs than during the pre-mussel period. Mussel grazing can be as much as that by zooplankton but the mussels can also utilize detritus that would otherwise form sediment. Mussel excretion can, therefore, return more P to the water than would occur otherwise. But despite the apparently altered biological recycling regime, algal productivity is “ultimately" regulated by external P loads. ii) Regeneration of sediment P to the water column. This includes sediment P release (as PO4) during periods of hypolimnion hypoxia (central basin), aerobic decomposition of organic matter from the sediment-water interface and resuspension of sedimented material by water movement and benthic animals. Most interest has been in hypoxic P release, because the degree of hypoxia is influenced by productivity, which relates in turn to external P loading, one of the few factors controllable by management action. Under oxic bottom water conditions, P release is

Box 4: Internal Loading

Internal loading represents a time-delayed recycling of accumulated external loads, which may

both stimulate algal blooms and delay the response of the lake to management actions.

Externally loaded P is subject to a complex series of in-lake processes, including physical

transport, biological uptake/recycling and chemical transformations before it is exported at the

outflow or deposited to the sediment.

Depending on the basin morphometry and mixing regimes, sedimented P can be regenerated

from the sediment back into the overlying water by direct resuspension, microbial degradation

of organic matter, benthic invertebrate and fish activity (including mussels), and by mineral

and redox reactions.

As a result of this internal P cycling, many lakes exhibit a delayed response to reduced

external loading. Understanding and quantifying these recycling processes is critical to

establishing P loading targets and managing expectations for recovery.

Page 30: Taking Action On Lake Erie - International Joint Commission · reduce external inputs of total phosphorus (TP) and in particular, the rising proportion of the most bioavailable form

17

influenced by organic matter deposition and biologically or physically enhanced transport of free PO4 from deeper anoxic sediments. The relative importance of these processes10 is as yet unknown, but they can play a significant and often overlooked role in sediment P exchange Evidence suggests that hypoxic sediment P regeneration in the CB does occur, but its significance and frequency is unclear and should be examined in detail. High SRP concentrations in hypolimnetic water have been observed intermittently, but not commonly reported. Burns (1976) found about half of the CB in 1970 was covered by an anoxic hypolimnion with up to 93 µg/L SRP; long term monitoring data collected since then suggest that this degree of P release does not occur with high frequency (e.g. Environment Canada and EPA databases; 1970-present). However current monitoring and research may simply have failed to capture this process adequately, which has yet to be measured directly using in situ probes or other methods. Field studies may not sample frequently or deep enough (most samples do not go deeper than 1-2 m above the sediments), while P release may vary on a diurnal basis and/or may be rapidly dissipated into the water column by the bottom currents which are documented by other studies (Rao et al 2008). Recycled P from hypoxic sediment can appear in the epilimnion11 by upward diffusion and water exchange between the epilimnion and hypolimnion layers. Incorporation of epilimnetic water into the hypolimnion represents an oxygen flux in addition to oxygen that may be contributed by algae in the hypolimnion (Burns, 1976). Low oxygen conditions result in the release of PO4 from iron-phosphorus complexes in sediment which can re-precipitate on encountering the oxygenated metalimnion and epilimnion, or wtih DO flux into the hypolimnion with the advent of early fall de-stratification.

10 whereby vertically migrating algae can harvest P in or near the thermocline and transport this into the epilimnion, or

sedimented live algal cells and resting stages (e.g. large diatoms, cyanobacteria) uptake P at the sediment surface and subsequently move this cell-bound P back up into the water column

11 Upper oxygenated mixed layer

Page 31: Taking Action On Lake Erie - International Joint Commission · reduce external inputs of total phosphorus (TP) and in particular, the rising proportion of the most bioavailable form

18

Thus, the probability that P regenerated from sediment stimulates surface algal populations is unknown, although blooms often develop during or shortly after periods where there is maximum probability of this release. As noted above, the significance of biological P translocation is also unknown and difficult to demonstrate. Spatially and temporally detailed studies by Burns (1976) do not show significant evidence to support this mechanism. Pore-water concentrations of SRP in shallow sediments can be over 1000 times that in lake water (Azcue et al. 1996) which may be why re-suspended P can be up to 40% soluble in nature (Burns, 1976) and may be more responsible for fall blooms than hypoxic regeneration. The lake productivity, however, responds to P concentration. Regardless of whether SRP is regenerated during hypoxia or by resuspension, excessive accumulation of P in Lake Erie sediments should be avoided and the only way to do this is to limit external nutrient loads. 1.6. Lake Management Implications of Internal Loading/Recycling Internal load is a mechanism that would delay the response to expensive nutrient controls, casting doubt on the rationale for nutrient controls. Once there is excessive P in a lake the size of Lake Erie, there are no management options to eliminate it except to wait for natural processes to take their course. Burns et al. (1976) estimated that ~92% of the P entering LE was retained in the lake. Most of the external hydrological and nutrient loading enters into the WB and is exported to the CB and EB. Nevertheless, the fraction P retained in the WB can be a source of recycled P that may delay recovery. This statement was as true in the 1970s as it is today, but what is the likely extent of the delay? As noted earlier, the GLWQA of 1973 reduced TP loading led to reduced concentrations despite internal cycling, but regime shifts, changes the nature and timing of the external loading (below) and climate change may require more stringent management targets and a re-evaluation of expected response times. A recent review of responses in European shallow lakes found that most reached equilibrium with reduced P loads in 10 to 15 years (Jeppesen et al. 2007) and it is reasonable to expect that Lake Erie’s

Box 5: Internal Loading in Lake Erie – Major results and recommendations

There are three main types of internal P cycling in Lake Erie i) inter-basin exchange; ii)

biological processing; iii) regeneration from the sediment to the water column. Most focus is

on the last process, but more attention should be paid to the second (biological recycling).

Evidence suggests that hypoxic sediment P regeneration in the CB does occur, but its

significance and frequency is unclear and should be examined in detail.

Current monitoring may simply have failed to capture hypoxic loading adequately, which has

yet to be measured directly.

The probability that P regenerated from sediment stimulates surface algal populations in

unknown.

Internal loading or recycling is sometimes discussed as a mechanism that would delay

response to expensive nutrient controls which needlessly casts doubt on the rationale for

nutrient controls.

It is reasonable to expect that LE’s response to P management actions would not be unduly

delayed past 10-15 years, but regime shifts, changes in the nature and timing of the external

loading, and climate change may require more stringent management targets and a re-

evaluation of expected response times.

Page 32: Taking Action On Lake Erie - International Joint Commission · reduce external inputs of total phosphorus (TP) and in particular, the rising proportion of the most bioavailable form

19

response would not be unduly delayed. 1.7. Climate Change Implications on External TP/SRP Loadings

Climate change can have significant impacts on the delivery of nutrients from agricultural watersheds. The critical variables are amount and seasonal distribution of precipitation, and changes in temperature. Calibrated SWAT models were run for six major WB watersheds to predict relative and absolute changes in average annual loads of P and other important parameters (flow, sediment, TP, SRP, Total N, and nitrate) as a function of anticipated future temperatures and precipitation patterns over 1998-2005. Three climate scenarios were considered, baseline (current), moderate, and severe climate change conditions (Table 5). SWAT models predicted that in-stream sediment and nutrient yields would generally increase under alternative climate scenarios, but yields decreased in several instances. Overall, across the six watersheds, the models predicted increasing flow, sediment and N (total and dissolved) with moderate and severe climate change as follows (summarised in Table 6) i) total annual flows by 2-9% and 5-17% ii) in-stream sediment yields by 5% and 22% iii) a more modest increase in nutrient yields, with differences in the direction and of changes in

tributary N and P: − minor average increases in N with climate change severity by 5% and 6% for TN

and 12 and 14% for dissolved inorganic N (mostly as NO3) with moderate and severe change, respectively.

− differences in predicted TP and SRP yields: TP would increase more under moderate (average 12%) than under severe climate change (9%); SRP yields actually decreased (average 4%) under moderate climate and increased only slightly under severe climate change

Table 6: Climate change scenarios (temperature and precipitation)

used in SWAT modelling

Moderate Severe

Season Temp (C) Precip (%) Temp (C) Precip (%)

Winter +2 +5

Spring +11 +29

Summer +4 +7

Fall -7

Page 33: Taking Action On Lake Erie - International Joint Commission · reduce external inputs of total phosphorus (TP) and in particular, the rising proportion of the most bioavailable form

20

At the watershed level, however, in-stream sediment and nutrient yields from the six Lake Erie watersheds responded very differently to alternative climate scenarios (Table 6). For example, in the Huron watershed, sediment yield decreased by 5% and increased by 30% under moderate and severe climate scenarios, respectively. The other five watersheds showed consistent increases in sediment yield under future climate scenarios, and increases were larger under more the extreme climate scenario. Modelled P yields for the Raisin differed from all other watersheds, with yield declines for both TP and SRP yields under alternative climate conditions. SWAT models found increases in all parameters in the Maumee and Sandusky watersheds under both climate scenarios. The Cuyahoga and Grand watersheds showed the

Table 7. Modelled climate change influence on Lake Erie rivers (flow, suspended sediments, nutrients) based on

scenarios in Table 4

River

Climate change

Flow Sediment TP SRP Total N Nitrate

(mm/y) (mg/km2) (kg P/km

2) (kg N/km

2)

Huron Baseline 206 1.2 15.5 4.9 220 195

Moderate 209 1.1 22.0 4.2 244 196

Severe 216 1.5 18.7 5.4 243 214

Raisin Baseline 263 20.7 32.8 12.2 1656 1346

Moderate 287 23.3 35.3 11.2 1734 1440

Severe 307 26.0 32.4 11.3 1898 1628

Maumee Baseline 328 52.6 101.5 26.0 2377 1995

Moderate 346 56.9 101.0 26.3 2441 2081

Severe 364 69.6 109.5 27.4 2590 2201

Sandusky Baseline 313 20.5 81.2 25.1 2593 2405

Moderate 334 22.0 84.5 27.2 2786 2611

Severe 351 23.7 91.0 30.1 3043 2863

Cuyahoga Baseline 511 35.2 110.3 80.7 2078 1658

Moderate 522 35.6 126.6 79.4 2121 1649

Severe 542 38.1 118.8 83.6 2099 1658

Grand Baseline 411 52.7 40.0 5.0 669 374

Moderate 427 56.2 41.3 4.6 729 413

Severe 447 62.2 42.4 4.7 815 460

Page 34: Taking Action On Lake Erie - International Joint Commission · reduce external inputs of total phosphorus (TP) and in particular, the rising proportion of the most bioavailable form

21

least change in yields of all parameters for both climate scenarios compared to the other four watersheds.

2. Effectiveness of Agricultural and Urban Best Management Practices (BMPs) in TP/SRP Loads to Lake Erie This review provides an overview of urban and rural BMPs that are employed to reduce P loads, that are likely to be considered for implementation within the LE basin to reduce P discharges in stormwater. BMPs are often designed to reduce an array of pollutant loads. Most commonly they are designed to reduce peak flow and total suspended solids, particularly in urban environments. This review specifically focuses on BMPs that have been evaluated using scientific methods for P reduction. A secondary focus was to highlight BMPs that have been implemented within the LE watershed, or at least in the Great Lake region. 2.1 Urban BMPs Phosphorus (P) loads to aquatic environments from urban sources are often underestimated. For example urban sources constitute only 3% of the land within the Lake Champlain watershed, but have been estimated to contribute 18% of the estimated P load (Meals and Budd 1998). The largest source of P from urban areas is associated with construction activities, which can temporarily generate even larger P loads per area than agricultural row crops (Burton and Pitt 2001). Due to the multitude of land uses within urban watersheds it is often difficult to pinpoint P loads from specific urban land cover. Given the significant loading from urban environments, there is a clear need to also consider urban sources of P if surface waters are to be managed appropriately. Importantly, in order to mitigate the diffuse urban inputs, equally disperse BMPs will likely have to be implemented. In one example, Winter and Duthie (2000) determined that P removal would have to be applied to all water entering the watershed from urban areas in order to have an appreciable reduction in stream P concentrations.

Box 6: Climate Change Influence on External TP/SRP Loadings – Major results and

recommendations

1. SWAT models predicted that in-stream sediment and nutrient yields would generally increase

under alternative climate scenarios, but these yields decreased in several instances.

2. Overall, across the six watersheds, the models predicted increasing flow, sediment and N (total

and dissolved) with moderate and severe climate change

Total annual flows and in stream sediment increased by up to 17-22% with severe climate

change.

More modest increases in nutrients, with differences in the response and magnitude of

changes in tributary N and P.

Differences in the climate-based response patterns predicted for TP and SRP yields.

3. At the watershed level, in-stream flow, sediment, and nutrient yields from the six Lake Erie

watersheds responded very differently to alternative climate scenarios.

4. Management strategies aimed at TP management, therefore, may not address SRP.

Page 35: Taking Action On Lake Erie - International Joint Commission · reduce external inputs of total phosphorus (TP) and in particular, the rising proportion of the most bioavailable form

22

2.1.1 Regulatory Framework

Discharges of P from urban sources in the U.S. are primarily controlled by Phase I and II of the U.S. EPA stormwater program. Phase I rules were promulgated in 1990 under the Clean Water Act and utilizes the National Pollutant Discharge Elimination System (NPDES). Phase I NPDES permits cover (USEPA 2005): i) “medium” and “large” municipal separate storm sewer systems (MS4s) generally serving populations of 100,000 or greater, ii) construction activity which disturbs 5 acres of land or more, and ii) 10 categories of industrial activity. Phase II requires MS4s and operators of small construction sites to implement programs and practices (e.g. BMPs) to reduce pollutant loads in stormwater runoff. The Phase II program again utilizes NPDES permits. However, most P loadings are derived from twelve (1.7%) wastewater treatment plants (WWTPs) that discharge > 15 MGD. It should be noted, that loads generated from WWTPs do not include discharges from CSO facilities. Michigan has twice enacted legislation limiting the amount of P in cleaning agents, the second of these in 1977 restricting the amount of P in household laundry detergents to 0.5% or less by weight (USEPA 2009). Together with improvements to the Detroit Water and Sewerage Department (DWSD) WWTP, this has generally resulted in effluent TP below 1 mg/L since the early 1980s.

Box 7: Effectiveness of Agricultural and Urban BMPs in TP/SRP Loads to Lake Erie

Review of over 240 primary sources has resulted in the following findings:

1. Very few studies have quantified P load reductions by urban or agricultural BMPs within the Lake

Erie watershed;

2. It is not possible to determine BMP cost-effectiveness due to costs rarely being reported;

3. BMP effectiveness, both urban and agricultural, vary greatly and are often contradictory; BMPs

designed to control other issues (e.g. flow) may actually increase P loads;

4. Most methods commonly used to quantify BMP performance are ineffective;

5. There is a need to move beyond total P measurements as the only metric used to quantify P,

assessing speciation is necessary to advance BMP performance;

6. Improved models are required to accurately predict treatment efficiency of BMPs under a variety

of conditions and climates;

7. While some databases exist, a central data repository is critically needed to synthesize data

collected and improve understanding of BMP effectiveness.

Box 8: Recommendations – Non-structural urban BMPs

Regardless of the empirical approach used to select BMPs or quantify P loads, whether

qualitative or quantitative, more effective controls are needed.

In order to increase BMP effectiveness across large regions and a variety of urban flow

conditions it is necessary to base future BMP designs and selection on mechanistic

understanding.

Page 36: Taking Action On Lake Erie - International Joint Commission · reduce external inputs of total phosphorus (TP) and in particular, the rising proportion of the most bioavailable form

23

2.1.2 Selection of Urban BMPs The selection of urban BMPs for P removal is often based on general classifications of perceived BMP utility, some which may be founded on empirical evidence (e.g. Gibb et al. (1999) and rarely on mechanistic models that can accurately describe P reductions. Unfortunately, models capable of describing the fate and transport of P in systems similar to some structural BMPs are not effective at modelling the P removal (Roy-Poirier et al. 2010). Export coefficients for P loads at the watershed scale have also been developed (Winter and Duthie 2000).

2.1.3. Non-structural (Alternative Behaviour/Management) BMPs Educational campaigns focused on changing residents’ behaviour have been found to result in only modest changes, with some BMP practices adopted more readily than others. A common non-structural BMP often considered by communities facing P related problems in surface waters is reducing P loads from lawn fertilizers. Loadings are reduced considerably if fertilization is based on soil tests rather than routine practice (Erickson et al. 2005). Alternatively, composted manure used as a source of slow release P reduces total P loadings to urban streams compared to conventional commercial turf-grass sod imported and maintained with inorganic P fertilizer (Richards et al. 2008). Significant reductions in TP and a trend of dissolved P reduction followed a municipal ordinance limiting the application of lawn fertilizers containing P in Ann Arbor MI (Lehman et al. 2008). Dietz et al. (2004) found 82% of residents began to leave lawn clippings in place while only 11% applied fertilizer after soil tests. Unfortunately, these changes were not found to result in a significant change in P loadings. Other non-structural changes include better management of leaves, pet waste, street sweeping and the use of native plants. For example, leaves from deciduous trees (e.g. oak, poplar) are reported to leach 54-230μg P/g, most (~85%) of which is soluble reactive (Cowen and Lee 1973). This release was nearly tripled when leaves were cut, as during mulching. Pet waste accounted for 84% of P inputs to urban runoff in the Minneapolis-Saint Paul, (Minn.) area (Fissore et al. 2012). Regular street sweeping may remove 40-70% of total P non-CSO loading (NVPDC and ESI 1992) although Hurley and Forman (2011) report a much lower removal (~11-14%) with combined street sweeping and structural BMPs in the Charles River although this combination is more likely to enhance water quality during and immediately after storms when waters are most impaired. Finally, the use of low maintenance plants that are indigenous to the eco-region may reduce stormwater P transport (Hipp et al. 1993).

Box 9: Non-Structural Urban BMPs

Based on these findings, the following non-structural BMPs are suggested:

Remove leaves immediately; do not mulch

Removal of pet waste immediately

Fertilizing based on soil tests rather than routine maintenance practice

Utilize native plants

Page 37: Taking Action On Lake Erie - International Joint Commission · reduce external inputs of total phosphorus (TP) and in particular, the rising proportion of the most bioavailable form

24

Based on these findings, the following non-structural BMPs are suggested (Box 8): i) remove leaves immediately (Cowen and Lee 1973); ii) removal of pet waste immediately (Fissore et al. 2012); iii) Fertilizing based on soil tests rather than routine maintenance practice (Erickson et al. 2005; Hipp et al. 1993); iv) Utilize native plants (Hipp et al. 1993)

2.1.4. Non-Point Source Structural or Engineered BMPs Traditionally, stormwater infrastructure was designed to mitigate flooding and move water as rapidly as possible to nearby water bodies and more recently at reducing peak flows, sediment loads, and turbidity during runoff events. Both objectives ignore other factors such as nutrient loads that play a more significant role in water quality impairments (EPA 2009). As a result, new BMPs are evolving to be more holistic and sustainable, with the aim of reducing pollutant loads (Batroney et al. 2010). Urban structural BMPs are best thought of as a spectrum of approaches rather than specific types (fig 5). Structural BMPs (engineered systems) typically employ filtration, detention – which allows for settling of sorbed material – or a combination of both. Likewise they can be designed as artificial systems and/or utilize/mimic natural processes and include the following (reviewed in more detail in the full TAcLE BMP report): Porous Pavements: reports are inconclusive although TP removal of 60-71% has been reported through the use of porous pavements (Hogland et al. 1987; Young et al. 1996). Media Filters appear to be a promising BMP to reduce P loads in stormwater runoff. Subsurface sand filters are reported to remove 43-82% TP and an unspecified amount of SRP in stormwater exiting media filters versus concentrations entering (e.g. Maniquiz et al. 2010; Leisenring et al. 2010). Filter Strips/Bioswales: level-spreader-grassed filter strips (LSGFS) along highways appear to result in significant reductions in P loadings in stormwater runoff (48%) (Horner et al., 1994; MMS, 1992; Reeves, 1994). When the majority of total P in stormwater is particulate, level-

Figure 4: Spectrum of urban structural BMPs

Page 38: Taking Action On Lake Erie - International Joint Commission · reduce external inputs of total phosphorus (TP) and in particular, the rising proportion of the most bioavailable form

25

spreaders perform about as well as retention basins and permeable pavements (Winston et al. 2011). However, data reported to the International Stormwater BMP Database (19 studies, 293 outflow data points) indicated a statistical significant increase in the median total, ortho- and total dissolved P concentration in stormwater runoff (Leisenring et al. 2010). Green Roofs and Filter Boxes have many positive attributes, such as reducing the peak flow generated from urban roof tops, providing added insulation, etc.; however initially they may contribute more P than they absorb as a result of leaching from material used to construct the green roof. Limited data suggests differences in performance in the short- versus long-term suggesting a need for more rigorous long-term monitoring (Berndtsson 2010).

Bioretention Basins include rain gardens, filter boxes and all other vegetative basins designed to increase infiltration and evapotranspiration. Removal efficiencies of P by bioretention basins are reported to be as high as 97% depending on the composition of soils used to construct bioretention cells (Carpenter and Hallam 2010). Detention and Retention Basins Treatment efficiencies vary considerably, ranging from 20% to 90% removal, depending on their design Retention basins ( “wet ponds”), remove ~ 47% TP and 51- 60% SRP. Designs with more natural features (longer flow paths, the use of native wetland plants, etc.) may increase P retention Wetlands Removal efficiencies of constructed wetlands vary widely and have been found to remove between 30% and 70% TP loads with some evidence of reduction in dissolved P. However removal in both subsurface flow and open surface wetlands are hampered by low oxygen conditions that can result in the release of previously sequestered P (Van de Moortel et al. 2009). Ultimately, a better understanding of the dynamic geochemical processes within wetland systems is required. Commercial Devices: Oil and grit separators have been found to be relatively ineffective (<10% removal efficiency) in reducing total P loads. Another type of commercial device, a type of subterranean concrete detention basin designed to remove settled solids, similar to septic systems (i.e. StormvaultTM), was found to remove approximately 50% of the P loads (Zhang et al. 2010). 2.1.5. Evaluation of Structural BMPs Treatment Efficiency Over 6,000 records from the International Stormwater BMP Database (www.bmpdatabase.org) were used to evaluate the treatment efficiency of structural BMPs. For this analysis we relied on the EMCs as one of the most accurate means for estimating pollutant loads. Details of the methods and model used are provided in the report. When the concentration of P leaving the BMP structure is less than that entering, the treatment efficiency is <1 and the structure is removing P (i.e. the BMP is effective). When the concentration of P leaving the BMP structure is greater than the concentration entering the treatment efficiency is > 1 and the structure is

The importance of selecting the appropriate material is exemplified in studies where media

high in P have resulted in increased P loads being discharged from BMP structures e.g.

bioretention, Hunt et al. (2006); bioswales and green roofs, Leisenring et al. (2010).

Page 39: Taking Action On Lake Erie - International Joint Commission · reduce external inputs of total phosphorus (TP) and in particular, the rising proportion of the most bioavailable form

26

adding P (i.e. the BMP contributes to P loading). In this analysis, the general BMP classifications used by the International Stormwater BMP Database were used.

Results of this analysis highlight the importance of understanding the different forms of P. Consider the results observed for detention basins (Figure 5). For these systems total P removal was observed in ~ 66% of all samples. However, only 45% of the dissolved P samples demonstrated removal. Detention basins, biofilters and wetland channels were all found to have clearly different removal efficiencies for total versus dissolved P. The removal of different forms of P by structural BMPs tended to follow the order particulate > total > dissolved. It is important to note that the EMC for each form of P are often not equal. For instance, the average EMC for TP entering detention basins was 0.48 mgP/L while the average EMC for dissolved P was only 0.17 mgP/L. Therefore, detention basins are likely to still reduce overall P loads.

Box 10: Analysis of Structural BMPs Treatment Efficiency – Key results

Only 43% of the samples demonstrated P removal.

Bioretention ponds and wetlands basins were shown to be the most effective urban BMPs, with

removal by ~82% and 75% of the paired total P EMCs.

Alternatively, grass biofilters (e.g. bioswales) were generally found to be ineffective at removing

P.

Page 40: Taking Action On Lake Erie - International Joint Commission · reduce external inputs of total phosphorus (TP) and in particular, the rising proportion of the most bioavailable form

27

2.1.6 Cost of Structural BMPs Little reliable data is available on the cost of structural urban BMPs. Data from the International Stormwater BMP Database were broken down into total facility costs – which generally include the cost of design, construction, excavation, landscaping, etc. – and maintenance costs (Figure 6) and data provide a general indication of costs. Engineered infiltration basins are the most expensive, detention basins and infiltration trenches the cheapest. However, these cost

Figure 5: Treatment efficiency for 6 classes of urban structural BMPs (see text)

Page 41: Taking Action On Lake Erie - International Joint Commission · reduce external inputs of total phosphorus (TP) and in particular, the rising proportion of the most bioavailable form

28

estimates are based on (1) a small sample size, (2) a diversity of specific BMPs included within broad categories, and (3) do not account for size of watersheds and facilities.

Figure 7. Estimated average annual maintenance cost (US/Canadian Dollars)

Figure 6. Average total facility costs in 2012 US/Canadian Dollars (note log scale).

Page 42: Taking Action On Lake Erie - International Joint Commission · reduce external inputs of total phosphorus (TP) and in particular, the rising proportion of the most bioavailable form

29

2.1.7. Urban Point Sources Structural and non-structural BMPs for point sources are primarily responsible for large reductions in P loads to LE. Lake Erie receives the largest municipal load of P of the Great Lakes, but large-scale WWTPS have been nearly 100% compliant since the 1990s. However combined sewer overflows (CSOs) deliver ~ 90.4 MTA TP to LE from Ohio alone (Ohio EPA 2010). Nineteen CSOs discharge untreated sewage directly into Lake Erie and 107 other CSOs to receiving waters that empty into Lake Erie which include Mill Creek, the Cuyahoga River, Rocky River, and Big Creek (Gomberg 2007). Reducing P loads before they enter sewer systems is likely more cost effective than reducing P discharges from WWTPs and CSOs. Efforts should be made to identify all major sources of P that could be eliminated; e.g. one potential source of P that is not typically considered in urban point source load is the addition of P to drinking water systems for corrosion prevention. 2.1.8. Monitoring Urban BMPs There is a general lack of coordination in evaluating urban, as well as rural, BMPs in the Lake Erie basin, as for the Great Lakes region. However, urban BMPs have been implemented, some specifically to reduce P loads to surface waters:

Multiple BMPs – Laurel Creek, Grand River in Southern Ontario

Bioretention - Detroit storm-sewer-shed, Southfield, Michigan

Openwater Wetland – Swift Run Wetland, Huron River Watershed Ann Arbor, MI

Detention Basin – Traver Creek Detention Basin, Huron River Watershed Ann Arbor, MI

Retention Basin – Pittsfield Retention Basin, Huron River Watershed, Ann Arbor, MI

Bioretention Basins and surface wetlands – Lake St. Clair Metropark, Mt. Clements, MI (expected to be completed 2013)

2.1.9 Urban BMPs Performance Metrics BMP performance can vary dramatically depending on the metric used (Lenhart and Hunt 2011). It is inadvisable to evaluate BMPs based on concentrations alone because performance varies during and between stormwater runoff events (Lenhart and Hunt 2011). Particularly problematic is the simple %removal metric because it is dependent on the initial concentration of pollutant (e.g. Zhang et al. 2010) and does not account for background water quality, eco-region differentiation, and background, or "irreducible," concentrations. Additionally, it inherently assumes an association between influent and effluent pollutant concentrations which in fact may not exist (McNett et al. 2011). Regardless of the type of BMP, three main mechanisms are responsible for P removal in stormwater: biouptake, sorption and precipitation. Ultimately P is retained via physical processes, either by attaching to material within BMPs (e.g. sorption to wetland plants) or by settling out - directly as a precipitate or indirectly while associated with biological material or suspended solids. Of these mechanisms, sorption reactions are the most common mechanism employed by most BMPs. On average, ~ 70% of P in stormwater is removed by the elimination of particles > 20μm in diameter; 90% P is associated with dissolved particles (>0.45 μm) (WERF 2003). However, because P partitioning between particulate and soluble forms can vary widely depending on amount and type of solids present and can convert rapidly, improving BMP

Page 43: Taking Action On Lake Erie - International Joint Commission · reduce external inputs of total phosphorus (TP) and in particular, the rising proportion of the most bioavailable form

30

performance “will also likely need to address dissolved P in order to achieve high and/or consistent pollutant removal” (Leisenring et al. 2010). This need for more advanced analysis of phosphorus is a common theme throughout urban and agricultural BMPs. Stormwater runoff events are not independent of each other and therefore all storm volumes and their chemical composition cannot be considered equal (Strecker et al. 2001). Consider wetlands or detention basins with permanent pools where runoff from one storm event inevitably mixes with water retained from previous storm events where sediments deposited previously may be resuspended later during high flow events. Sorption reactions may be reversed due to changes in chemical conditions (O2 concentrations) resulting in an increased dissolution of P previously retained on soil, sediment or other material. Both of these processes, resuspension and desorption, can result in increased P loadings from physical structures installed as BMPs (e.g. Leisenring et al. 2010) and effluent quality may be a better estimate of BMP effectiveness than %removal (Strecker et al. 2001). Overall much more rigorous sampling analysis protocols are needed to assess the effectiveness of urban BMPs. Sampling schemes must take into account the variability inherent to these dynamic systems. Grab sampling is not effective in assessing water quality and flow weighted event mean concentrations are the preferred measurement technique for estimating loadings. If systems are large enough or contain sufficient vegetation (e.g. constructed wetlands) sampling will likely also need to account for diurnal and seasonal variation. An important example is seen in the high spatiotemporal variance in water quality and nutrients measured in the Detroit River by a joint EC_USEPA focussed sampling study, which suggests a far higher contribution to the LE P-loading than currently represented in mass balance models (e.g. Burniston et al. 2009). 2.2 BMPs in Agricultural and Rural Environments In the following sections, agricultural BMPs recommended by the Ohio Lake Erie Phosphorus Task Force (OH-LEPFT) for P, N and sediment exports to Lake Erie are presented (OH-EPA 2010). Several reviews have addressed BMP effectiveness; most evaluated a suite of BMPs implemented simultaneously together (e.g. Bishop et al. 2005) These studies were done or focused outside of Lake Erie watersheds, but some these BMPs maybe already implemented or are applicable in Lake Erie watersheds. BMP effectiveness is very site-specific and depends on local topography, climate, cropping systems, maintenance, selection, and installation (Alfera and Weismiller 2002). The BMP effectiveness tool (available online (Gitau 2013) created by Merriman et al. (2009) with the addition of recent references was used to tabulate the effectiveness of the BMPs discussed. 2.2.1 Agricultural P-source BMPs Agriculture is a major source of nonpoint inputs of P, N and sediments. Agricultural systems have evolved from being net P sinks, where crop production is P-limited, to P sources where there is net P excess on farms. The control of agricultural P losses should be directed towards the long-term goal of increasing farm P-use efficiency, achieved by practices that balance P inputs and outputs within a watershed and improve the management of soil, manure, and mineral fertilizer at the farm, watershed, or regional scales. The effectiveness of agricultural BMPs is better understood by identifying the P forms in the soil, their transformation, transport, and pathways (i.e., the P cycle). Soil P occurs in organic or inorganic (mineral) form. Organic P associated with plant residues, organic matter, microbial

Page 44: Taking Action On Lake Erie - International Joint Commission · reduce external inputs of total phosphorus (TP) and in particular, the rising proportion of the most bioavailable form

31

biomass, etc., makes up a large portion of P in the soil. This is not available for plant uptake until the organic materials are mineralized and P is released. SRP is usually sorbed in aluminium, iron, or calcium compounds, depending on soil pH, and only present at trace amounts in the soil solution (Espinoza et al. 2005). During storm events, particulate P (PP, organic and inorganic) and total dissolved P (TDP; organic and SRP) are transported by runoff water. The amount of P in runoff water is directly correlated with the available P in the soil (e.g. Schroeder et al. 2004). TDP maybe leached with the downward vertical water movement through the soil profile, and is an issue in P-saturated soils and where there is macropore and bypass flow to tile drains (Sharpley et al. 2006). Tile drains are an important source of P to streams and there is a considerable increase in dissolved and especially particulate P concentrations in tile-drains with increasing discharge (Tan and Zhang 2011). Table 8: Actions associated with P-source BMPs (From Sharpley et al., 2006 )

1. Balance P inputs with outputs at farm or watershed scale 2. Minimize P in livestock feed 3. Test soil and manure to maximize P management 4. Physically treat manure to separate solids from liquid 5. Chemically treat manure to reduce P solubility (i.e., alum, fly ash, and water treatment residuals) 6. Biologically treat manure (i.e., microbial enhancement) 7. Calibrate fertilizer and manure spreaders 8. Apply proper application rates of P 9. Use proper method for P application (i.e., broadcast, plowed in, injected, subsurface placement, or

banding) 10. Carefully time P application to avoid imminent heavy rainfalls 11. Implement remedial management of excess P areas (spray fields and disposal sites) 12. Compost or pelletize manures and waste products to provide alternate use 13. Mine P from high-P soils with certain crops and grasses 14. Manage urban P use (lawns and gardens)

BMPs can be divided according to source and transport of different forms of P. In the US, most of these BMPs are established according to the United States Department of Agriculture’s National Resources Conservation Service (USDA-NRCS) standards12. Source BMPs minimize the potential of P as a pollution at the origin, i.e., before P is transported from the soil by water movement. Transport BMPs are mostly structures and methods that reduce the transport of P with water and sediments 2.2.1.1 Farm Gate Regulation Fertilizer management Economic pressure and extension activities have resulted in efficient fertilizer management; thus, the unnecessary influx and over-application of fertilizers into farms and agricultural soils is not usually considered a major cause of nonpoint source pollution (Sharpley et. al., 2006). However farmers may apply two years’ worth of fertilizer at the start of a crop rotation although this practice is now being discouraged Efficient fertilizer management is a component of the

12 These standards and their description are listed in the following URL (accessed on December 10, 2012:

http://www.nrcs.usda.gov/wps/portal/nrcs/detailfull/national/technical/nra/nri/?&cid=nrcs143_026849)

Page 45: Taking Action On Lake Erie - International Joint Commission · reduce external inputs of total phosphorus (TP) and in particular, the rising proportion of the most bioavailable form

32

“4R” nutrient stewardship principle and further discussed below Animal feed management In animal-based agriculture, feed mass balance has become an evolving and important BMP. Animal farms have decreased in numbers but their capacity (herd size and animal densities) has increased. As a consequence, net nutrient influxes and net nutrient excess occurs in most of these farms (Sims, 1977). There is a direct relationship between dairy cow P-intake and P-excretion (Dou et al. 2002). Feed management (NRCS 592 standard) is defined as “manipulating and controlling the quantity and quality of available nutrients, feedstuffs, or additives fed to livestock and poultry” (USDA-NRCS 2011). The objectives are: 1) Improve feeding efficiency to facilitate and contribute to the conservation of natural resources; 2) Reduce the quantity of nutrients (e.g., N, P, S, salts, etc.) excreted in the manure; 3) Reduce pathogens in manure, and 4) Reduce odor, particulate matter, and greenhouse gas (GHG) emissions from animal feeding operations (USDA-NRCS 2011). Effectiveness: Decreasing P in feeds is the best method to mitigate P loss from faeces. Manure total P reductions with feed management range from 16% to 33%. Manure Management Manure export from the farm is not a viable management option because of hauling costs and off-farm land application options are generally restricted to the nearest neighbours (Sharpley et. al., 2006). In most areas, waste storage, composting, and land applications are the most viable options for manure management. Table 9:. Nutrient reduction efficiencies of animal waste system BMPs (%).

Study Method Study scale Remarks/State TP DRP PP TN SS Reference

Field studies Field Poultry litter alum 70 - - - - (Lory 1999)

Field Studies Field Swine manure alum - 84 - - - (Smith et al. 2001)

Plot Studies Plot/rainulator

Poultry litter alum - 87 - - - (Shreve et al. 1995)

Literature Small watershed

Poultry litter alum 72 75 - - - (Moore et al. 1999)

Literature Field Animal waste system 90 - - 80 60 (Cestti et al. 2003)

Modeling Large watershed

Waste Storage Facility 27 - - 29 - (Mostaghimi et al. 1997)

Literature Field Waste Storage Facility 90 - - 75 - (Randall et al. 1987)

Literature Field Waste Treatment Lagoon 0-90 - - - - (Gilliam 1995)

Field Field Waste Treatment Lagoon 34 - - 43 - (Hubbard et al. 2004)

Animal waste system Waste storage facility, treatment lagoon, and waste treatment (NRCS codes 313, 359, 591). This impoundment is made from an embankment and/or pit or dugout, or a fabricated structure (USDA-NRCS 2003), “to temporarily store wastes such as manure, wastewater, and contaminated runoff as a storage function component of an agricultural waste management system.” Effectiveness: See Table 8 Composting facility (NRCS code 317). The goals are: 1) reduce the pollution potential and improve the handling characteristics of

Page 46: Taking Action On Lake Erie - International Joint Commission · reduce external inputs of total phosphorus (TP) and in particular, the rising proportion of the most bioavailable form

33

organic waste solids, and 2) produce a soil amendment that adds organic matter and beneficial organisms and provides slow-release plant-available nutrients, and improves soil condition. Effectiveness: reported to reduce runoff SRP levels from field sites by 19- 23%. (Bekele et al. 2006) Vegetative treatment area (VTA, NRCS code 635) There are 4 kinds: 1) VFS that matches crop N uptake with estimated N in runoff and requires sheet flow across the filtering slope; 2) constructed wetlands; 3) discharging or nondischarging infiltration basin systems with berms; and 4) overflow and/or cascading terraces that are similar to infiltration basins (Koelsch et al. 2006). In 2008, the USDA-NRCS released updated and specific guidelines and standards for vegetative treatment areas (USDA-NRCS 2008). Effectiveness: Most of the literature identified VTAs as VFS, constructed wetlands, and infiltration/ponding basins or a combination. Effectiveness is discussed below 2.2.1.2 Managing P Applications to Soil Nutrient management (NRCS code 590) The USDA-NRCS standards for nutrient (fertilizer and manure) management entails managing the amount (rate), source, placement (application method), and timing of plant nutrients and soil amendments. Nutrient management is devised to: 1) budget, supply, and conserve nutrients for plant production; 2) minimize agricultural nonpoint source pollution of surface and groundwater resources; 3) properly utilize manure or organic by-products as a plant nutrient source; 4) protect air quality by reducing odours, nitrogen emissions (ammonia, oxides of nitrogen), and the formation of atmospheric particulates; and 5) maintain or improve the physical, chemical, and biological condition of soil (USDA-NRCS 2012).

The “4R” stewardship framework (right source, right rate, right time, and right place) nutrient stewardship framework is jointly promoted by The Fertilizer Institute (TFI), International Plant Nutrition Institute (IPNI), the International Fertilizer Industry Association, and the Canadian

Box 11: The “4R” Stewardship Framework is based on:

1. Right Fertilizer Source: matching appropriate fertilizer source & product with soil properties &

crop needs. Nutrient interactions should be accounted & nutrients should be balanced according

to crop needs & soil tests.

2. Right Rate: matching application rates with crop requirements. Excessive fertilizer application

may lead to nutrient loss to the environment with no additional gain crop yield & quality.

3. Right Time: making the nutrient available when the crops need them; influenced by pre-plant or

split application timing, controlled release technologies, stabilizers, & inhibitors.

4. Right Place: placing & keeping nutrients where the crop can efficiently use them. The method of

fertilizer application is critical. The most appropriate placement method is determined by the

crop, cropping systems, & soil properties. Injection or incorporation is the preferred method but

soil disturbance needs to be balanced with erosion-control BMPs such as conservation tillage,

buffer strips, cover crops, & irrigation management.

Page 47: Taking Action On Lake Erie - International Joint Commission · reduce external inputs of total phosphorus (TP) and in particular, the rising proportion of the most bioavailable form

34

Fertilizer Institute (CFI). The role of fertilizer distributors, retailers, and certified crop advisors (CCAs) is vital in managing fertilizer at the farm. The implementation of nutrient-management related BMPs has environmental as well as economic benefits; off-farm nutrient transport is an investment loss for the farmers (Mullen et al., 2009). Among the specific practices are:

i). Soil, Manure, and Tissue Sampling and Laboratory Analyses. Nutrient planning must be based on recent soil, manure, and tissue tests and the analysis should be according to a land-grant university guidance or a university-recognized industry practice (USDA-NRCS 2010). Soil testing is the most cost effective and environmentally sound practice a producer can implement (Mullen et al., 2009). Fertilizer recommendations are usually available for each state in the U.S. and application rates can be calculated based from the soil tests and the crop demands.

ii). Timing and applications. According to the (USDA-NRCS 2010), the timing and placement of nutrients must correspond with crop demand and account for nutrient source, cropping system limitations, soil properties, weather conditions, drainage system, soil biology, and nutrient risk assessment results. Furthermore, nutrients must not be applied on: 1) frozen and/or snow-covered soils, and 2) when the top 2 inches of soil are saturated from rainfall or snow melt.

Effectiveness: (see Table 9) Nutrient management and related practices (e.g., soil and tissue test, fertilizer rates calculation, variable rate application, precision agriculture, etc.) in crop-based agriculture were primarily geared towards efficient agronomic output but not necessarily environmental quality. Concern with the latter seems to be more prevalent in animal-based agricultural production and not well-studied in purely crop-based production agriculture. The method of nutrient application is related to tillage methods (Andraski et al. 1985). The effect of fertilizer application rate on P loss at a farm scale is directly related to application method, the hydrologic soil group, and crop type. Nutrient management in combination with tillage and erosion practices may reduce total P loads by more than 80% but in some cases may increase the loads (Cestti et al. 2003). 2.3.1 P-Transport BMPs P Transport BMPs are aimed at erosion control and total P reduction (Box 11). These practices leave at least 30% of the soil surface covered with crop residue following tillage and planting to reduce soil erosion (Galloway et al. 1981). They are reviewed in detail in the full report and include a range of different BMPs which fall under the following general categories

Page 48: Taking Action On Lake Erie - International Joint Commission · reduce external inputs of total phosphorus (TP) and in particular, the rising proportion of the most bioavailable form

35

2.3.1.1 Residue and Tillage Management (Conservation Tillage) These are management practices that leave at least 30% of the soil surface covered with crop residue following tillage and planting to reduce soil erosion (Galloway et al. 1981). Effectiveness: the plant residue left on the soil surface increases infiltration thereby decreases runoff but not necessarily runoff P-concentration (Nicolaisen et al. 2007). In general, conservation tillage reduces TP loads up to 60- 80% when in conjunction with nutrient management and a farm plan and was primarily implemented for erosion and TP control. A few studies which evaluated SRP reduction reported a wide range of efficiency (-390% to 91%). 2.3.1.2 Conservation Cropping (Includes crop rotation, cover crops conservation cover and strip cropping) Effectiveness: Bosch et al. (2009) observed that post-BMP loading of SRP decreased by 74% and nitrate (NO3) by 73 - 88%. The implemented BMPs included a mix of crop rotations with tillage practices, etc. Jiao et al. (2011) also reported that double cropping systems reduced runoff volume and losses of total dissolved P (TDP), PP, and TP as compared with a wheat-fallow system. 2.3.1.3 Conservation Buffers Generally designed to create or improve habitat, reduce sediment, organic material, nutrients and pesticides in surface runoff and shallow ground water flow. They include contour buffer strips (narrow strips of permanent, herbaceous vegetative cover around landscape contours,

Table 10: Summary of nutrient management (NM) efficiencies (%) in reducing pollutant loss.a

Method Scale Remarks/Location TP DRP PP TN SS Reference

Plot Studies

Plot/rainulator Fertilizer and tillage treatments, VA

nab na na 96 95

(Mostaghimi et al. 1991)

Literature review

Plot, Field, watershed

Modeling and field app. across the US

35 na na 15 Na (USEPA 1993)

Literature review

Plot, Field, watershed

NM & a suite of BMPs, Cheasapeake Bay

-10c

to 88

na na -2 to 84

Na (Cestti et al. 2003)

Modeling Field/farm Manure and Tillage systems, PA

46 -62 na na 56 (Sedorovich et al. 2007)

Field Small watershed NM & 4 other BMPs, manure based, NY

69 74 na 70 Na (Lewis and Makarewicz 2009)

Field Small watershed NM with 4 other BMPs, Manitoba, Canada

38 41 42 41 Na (Li et al. 2011)

Modeling Small watershed Manure nutrient management, TX

-5 to 12

na na -3 to 11

-22 to 12

(Rossi et al. 2012)

a Negative values indicate percent increase in nutrient loss instead of nutrient loss reduction.

b na: not

applicable; not calculated.c as total dissolved phosphorus (TDP).

Page 49: Taking Action On Lake Erie - International Joint Commission · reduce external inputs of total phosphorus (TP) and in particular, the rising proportion of the most bioavailable form

36

riparian forest buffers (areas dominated by trees or shrubs adjacent to and up-slope of watercourses or water bodies) and filter strips or areas of herbaceous vegetation. No evaluation has been carried out.

2.3.1.4 Wetlands (constructed wetlands) Designed to treat wastewater and runoff primarily from agricultural processing, livestock, and aquaculture facilities and also used to improve storm runoff quality or other water flows lacking specific water quality discharge criteria. Effectiveness: the majority of the wetlands reduce N and P loading; however, increased nutrient loading may result in elevated soluble N and P species over time. Hoffmann et al. (2012) observed a high N reduction but a net P release in two restored riparian wetlands. Rogers et al. (2009) observed that a disturbed wetland exported 50% more sediments and 30% more TP above that which was present in the inputs.

Box 12: Actions Associated with Transport BMPs (from Sharpley et al., 2006)

1. Minimize erosion, runoff, and leaching 2. Use cover crops to protect soil surface from erosion

3. Terrace to minimize runoff and erosion

4. Practice strip cropping to minimize runoff and erosion

5. Practice contour farming to minimize runoff and erosion

6. Manage irrigation to minimize runoff and erosion

7. Practice furrow management to minimize runoff and erosion

8. Install filter strips and other conservation buffers to trap eroded P and disperse runoff

9. Manage riparian zones to trap eroded P and disperse runoff

10. Install grass waterways to trap eroded P and disperse runoff

11. Manage wetlands to trap eroded P and disperse runoff

12. Manage drainage ditch to minimize erosion

13. Stabilize streambank to minimize erosion

14. Fence streambank to keep livestock our of water course

15. Protect wellhead to minimize bypass flow to ground water

16. Install and maintain impoundments to trap sediment and P

17. Retain crop residues to minimize erosion and runoff

18. Consider reduced tillage systems to minimize erosion and runoff

19. Manage grazing (pasture and range) to minimize erosion and runoff

20. Restrict animals from certain sites

21. Install and maintain manure handling systems (houses and lagoons)

22. Manage barnyard storm water

23. Install and maintain milk-house waste filtering systems

24. Practice comprehensive nutrient management planning (CNMP)

25. Install and maintain tail-water return flow ponds

Page 50: Taking Action On Lake Erie - International Joint Commission · reduce external inputs of total phosphorus (TP) and in particular, the rising proportion of the most bioavailable form

37

2.3.1.5 Grassed Waterways (NRCS code 412) Grassed waterways are shaped or graded channels established with appropriate vegetation to carry surface water at a non-erosive velocity to a stable outlet. Grassed waterways are designed to 1) convey runoff from terraces, diversions, or other water concentrations without causing erosion or flooding, 2) reduce gully erosion, and 3) protect and improve water quality (USDA-NRCS 2010). 2.3.1.6 Drainage Water Management Manages the discharge water from surface and/or subsurface agricultural drainage systems to: i) reduce nutrient, pathogen, and/or pesticide loading from drainage systems into downstream receiving waters; ii) improve productivity, health, and vigour of plants; III) reduce oxidation of organic matter in soils; iv) reduce wind erosion or particulate matter (dust) emissions provide seasonal wildlife habitat Effectiveness: may sometimes cause water quality third-party impacts. Studies report annual NO3 load reduction of 15 to 97% while a combination of BMPs, including drainage water management, results in a 20% reduction of edge-of-field N loss 2.3.1.7 Emerging Technologies Examples of current and emerging technologies that need more research for the reduction of P loadings from agricultural areas include: 1) two-stage ditches (Powell et al. 2007); 2) controlled drainage (Kroger et al. 2011; Nistor and Lowenberg-DeBoer 2007); 3) focus on overall soil quality/health; 4) hydrologic attenuation; 5) nutrient management education as a BMP; 6) treatment of tile outlets with bioreactors, filters, etc. (McDowell et al. 2008). Lastly, more edge of field research and monitoring under real world farming systems and real world climatic events should be done (Steve Davis, personal communication). 2.4 Effects of Extreme Weather BMP effectiveness is confounded by weather patterns, and effectiveness therefore should be assessed over a long period of time to account for these uncertainties. For example Chaubey et al. (2010) reported that pollutant losses are greater under certain extreme weather conditions than the pollutant reductions caused by BMP implementation in a watershed. The NCWQR data (http://www.heidelberg.edu/academiclife/distinctive/ncwqr; unpublished) also showed that the SRP loads in the Sandusky and Maumee watersheds for spring 2012 were less than 3% of the SRP loads in spring 2011. Spring of 2011 was among the wettest on record, while the spring of 2012 was among the driest. It is noteworthy that BMPs were similar between these two periods. 2.5 Focus on Management of SRP vs. PP Traditionally, TP was considered as 23-33% bioavailable (Baker 2010), however, Seo et al. (2005) measured SRP as 70% of TP in runoff from a no-tilled and broadcast fertilizer field. Sweeney et al. (2012) also observed that P loss from a field applied with turkey litter was mainly in soluble form, and annual losses tended to increase with larger annual flow volumes. The effects of no-till in the fall on SRP losses are frequently negative, as observed by Ulen et al. (2010) where SRP losses were increased by four times in a no-till system compared to conventional tillage in field experiments. Tiessen et al. (2010) observed that conversion to

Page 51: Taking Action On Lake Erie - International Joint Commission · reduce external inputs of total phosphorus (TP) and in particular, the rising proportion of the most bioavailable form

38

conservation tillage increased P concentrations and exports, with soluble P comprising the majority of the P export, especially during snowmelt. Kleinman et al. (2011) also discussed the tradeoffs between fertilizer and manure management vs. no-till systems. They showed that TP increased by 12% and PP decreased by 37% in a no-till watershed compared with conventional-tilled watershed. The increase in TP was attributed to the increase in dissolved P due to severe soil stratification. BMPs that lower the accumulation of soil-P and plant residue at the soil surface should be considered in areas where dissolved P is a major concern (Tiessen et al. 2010).

Data gathered by the National Center for Water Quality Research (NCWQR), Heidelberg University, show that from the mid-1970s to date, the sediment and PP exports to Lake Erie have been reduced and are still decreasing while TP load remains relatively constant. These trends imply that conservation practices to control sediments and PP were successful (Richards et al. 2010). However, from the mid-1990s the SRP load has been rapidly increasing in the monitored tributaries (Figure 8). Daloglu et al. (2012) used the Soil and Water Assessment Tool (SWAT) watershed model to demonstrate that the increasing DRP trend after the mid-1990’s was driven by increasing storm events, changes in fertilizer application timing and rate, and management practices that enhanced P-soil stratification. A summary of BMPs (BMP toolbox) focused on controlling DRP was developed as a part of a project funded by the Great Lakes Protection Fund (GLPF, Crumrine 2011).

Box 13: BMPs – Key points to consider

1. Pet waste was found to be responsible for 84% of P inputs in the Minneapolis-Saint Paul,

Minnesota metropolitan area.

2. The amount of P in runoff water is directly correlated with the available P in the soil

3. Tile drains are a major source of P to streams.

4. The effects of no-till in the fall on SRP losses are frequently negative, as observed by Ulen et al.

(2010) where SRP losses increased 4x in a no-till system compared to conventional tillage in

field experiments.

5. Tiessen et al. (2010) observed that conversion to conservation tillage increased P concentrations

and exports with soluble P compromising the majority of the P export.

6. BMPs that lower the accumulation of soil-P and plant residue at the soil surface should be

considered in areas where dissolved P is a major concern.

7. Daloglu et al. (2012) used SWAT models to demonstrate the increasing SRP trend in LE after the

mid-1990s was driven by increasing storm events, changes in fertilizer application timing and

rate, and management practices that enhanced P-soil stratification.

Page 52: Taking Action On Lake Erie - International Joint Commission · reduce external inputs of total phosphorus (TP) and in particular, the rising proportion of the most bioavailable form

39

The recurrence of severe algal blooms in Lake Erie in the mid-1990s coincided with this increase in SRP loads. A combination of several factors may have caused the increase in SRP export from agricultural lands (OH-NRCS 2012):

Conservation practices (e.g., reduced- and no-till cropping systems) implemented since the early 1990s in predominantly agricultural northwest Ohio have mainly focused on reduction of sediment and TP; these practices are less useful for controlling dissolved P.

Farming equipment has become larger and the producers broadcast fertilizer onto the soil surface, rather than banding it.

Large-equipment traffic may have caused soil compaction resulting in decreased infiltration and increased runoff.

Increasingly, fertilizer is applied in the fall instead of spring.

The application of two years’ worth of fertilizer in one year for a corn-corn or corn-soybean crop sequence saves money, time, and labour for the producers but results in higher rates and amounts of fertilizer available for export out of the cropland into the streams.

The maximization of crop yields through fertilizer application and the use of conservation tillage may have also increased soil P levels, particularly at the soil surface (soil stratification) over a long period of time.

Box 14: Summary of Findings and Research Needs – Agricultural BMPs

1. There is no silver bullet to solve non-point source pollution, particularly SRP. The use of a

suite of BMPs (or toolbox) is currently recommended. A major challenge is how to evaluate

the synergy of BMP effectiveness.

2. Most studies were done or focused outside of Lake Erie and there was no rigorous

assessment of agricultural BMPs specific to the Lake Erie watersheds; some of the BMPs

maybe already implemented or are applicable in the Lake Erie watersheds.

3. Most BMP effectiveness assessments were focused on TP and sediment reduction. The

traditional idea was that the majority of P losses occur as particulate P attached to sediments.

4. The range of BMP effectiveness greatly varies and results from numerous studies of BMP

effectiveness are often conflicting.

5. BMP assessments were done at different scales (plot, field, and watershed scales). Methods

are either field studies and applications or simulation modelling. Most BMP assessment at

watershed scales were done with modelling or assessment of general trend changes in WQ

parameters at the watershed outlet. Major issues that need to be addressed are: a) scaling up

of BMP effects to watershed scale from plot scale and b) model reliability.

6. Assessing the effectiveness of a single BMP is complicated and difficult since most BMPs are

in combination with at least another BMP. Caution should be done in extrapolating

effectiveness to other sites and in modelling.

7. Other challenges include soil-P stratification and the trade-offs in controlling SRP and

sediment/erosion, e.g., no-till vs. conventional-till vs. nutrient management.

8. There is a need of an inventory of implemented agricultural BMPs (a BMP clearing house) in

Lake Erie watersheds and in the Great Lakes in general. This inventory is essential in

assessing programmatic and the overall effects of BMP implementation.

Page 53: Taking Action On Lake Erie - International Joint Commission · reduce external inputs of total phosphorus (TP) and in particular, the rising proportion of the most bioavailable form

40

2.6 Current Policy and Implications A critical role for science within Lake Erie is to inform both lake and land management practices. Efforts to improve lake water quality and address trophic status are underway at local, state, national, and international levels in both Canada and the United States. Current focus on reduction of nonpoint source loading of phosphorus and in escalating SRP levels remains a priority for national and state level organizations (EPA 2010). An ongoing debate within the scientific community as to the best method for nutrient management has yet to be resolved. Evidence exists to support a focus on P (Schindler 2012) and a dual-nutrient management strategy for both N and P) (Lewis et al. 2011; see also section 4 on Harmful Algal Blooms in this report). Further efforts to resolve this debate will advance both management practices and the science of freshwater cHABs. Continued sponsorship of monitoring and scientific endeavours to track changes in the status of cHAB events in Lake Erie and gain insight into the forces contributing to events is also an important goal of provincial, state and national organizations. It should remain a priority; however, to continue to engage in experimental science along with monitoring and modelling efforts as such efforts create context and validation.

3. Hypoxia 3.1 Causes of Hypoxia Bottom hypoxia (oxygen concentrations < 4 mg/l) recurs in a diversity of freshwater, brackish, and marine ecosystems. In brief, within a thermally or chemically stratified water body, low photosynthetic (O2-generating) activity in the bottom layer, coupled with low, density-impeded dissolved O2 (DO) replenishment from the oxygenated top layer, can lead to bottom hypoxia. In dimictic waters, this phenomenon is generally seasonal, with hypoxia developing after an extended period of stratification and ceasing with water-column turnover. In Lake Erie, seasonal hypoxia is a natural event, which may have occurred in the CB for hundreds to thousands of years (Delorme, 1982). In fact, the recurrence of seasonal hypoxic events in the central basin subsequent to GLWQA based nutrient reductions suggests that these events are not due solely to human-induced eutrophication (Charlton et al. 1993). A major reason for hypoxia formation in central Lake Erie relates to its bathymetry (Rosa and Burns 1987). Because the CB is deep enough to stratify (mean depth (zm ) = 18.3 m), but shallow enough that the thermocline is established relatively close to the lake bottom (typically < 6 m from the bottom; Rosa and Burns 1987), the volume of hypolimnetic water and associated DO reserve that is isolated from surface DO replenishment is small and can be depleted before fall turnover, thus leading to bottom hypoxia. By contrast, hypolimnetic volume of the deeper EB (zm = 24.4 m) is relatively large and does not become hypoxic before fall re-mixing (turnover), whereas the shallow (zm = 7.3 m) WB typically does not become hypoxic owing to wind-driven circulation and storm events that keep the WB well mixed during summer. While hypoxia is a natural phenomenon in central Lake Erie, the DO depletion rate and areal extent of hypoxia can be modified by human activities (Rosa and Burns 1987, Bertram 1993). For example, owing to excessive P inputs that stimulated algal production, DO depletion rates during summer increased during the mid-1900s, producing a hypoxic area as large as 11,000 km2 (Beeton 1963). While hypoxic conditions in the central basin of Lake Erie were first detected in the late 1950s (Charlton 1987), anthropogenic nutrient loading likely contributed to hypoxia much earlier. During the height of cultural eutrophication, even the shallow WB of Lake Erie

Page 54: Taking Action On Lake Erie - International Joint Commission · reduce external inputs of total phosphorus (TP) and in particular, the rising proportion of the most bioavailable form

41

could become hypoxic during windless periods in summer (Hartman 1972). In fact, by 1963, even a 5 day period of hot, calm weather could cause 50% of the west basin to become hypoxic (Hartman 1972). 3.2 P Abatement Programs Initiated as part of the 1972 GLWQA are attributed to a decline in bottom hypoxia in both western and central Lake Erie through the early 1990s (Charlton et al. 1993), but since the late 1990s, the extent of bottom hypoxia has increased to levels on par with those during the height of cultural eutrophication (Hawley et al. 2006). The exact causal mechanisms for this increase are not fully understood, although it does coincide with altered precipitation patterns, warmer water temperatures, increased non-point nutrient inputs and extensive algal blooms. 3.3 Climate Change Climate change is predicted to influence hypoxia formation in the Lake Erie ecosystem in several ways. Predictions made for other temperate freshwater ecosystems indicate that continued climate change will exacerbate the magnitude, duration and frequency of hypoxia (Kling et al. 2003, Ficke et al. 2007, Fang and Stefan 2009, Jiang et al. 2012). Most directly, warmer future conditions are expected to facilitate a longer stratified period during summer, with earlier establishment of thermal stratification and turnover occurring later in the year. Thereby, bottom DO depletion will initiate earlier and hypoxic conditions are likely to persist over an extended time period (Fang and Stefan 2009). Expected reductions in water levels could further exacerbate bottom hypoxia. For example, lower water levels may interact with warmer temperatures and allow the thermocline to develop closer to the bottom (i.e., result in a thinner hypolimnion). Since a large proportion of hypolimnetic oxygen demand comes from the sediment (Rucinski et al. 2010), DO would be depleted more rapidly from a thinner hypolimnion. While uncertainty surrounding future regional precipitation patterns is greater than future regional temperatures, it is plausible that precipitation patterns will be characterized by less frequent, but more intense, precipitation events (Kling et al. 2003, Kunkel et al. 1999). Such intense events could lead to high nutrient runoff from agricultural and urban lands, and in the absence of dramatic changes in land use, lead to increased overall nutrient loads to Lake Erie. Depending on the timing of runoff, future nutrient loading, coupled with warmer epilimnetic temperatures, could lead to greater overall phytoplankton production and ultimately exacerbate decomposition and oxygen depletion rates in the hypolimnion. Potential changes to future wind patterns have not received the same amount of attention as temperature and precipitation (e.g., Kling et al. 2003), but by affecting thermal stratification, also have the potential to alter hypoxia patterns. Specifically, intense wind events could contribute to mass movement of water, including seiches and potential influx of hypoxic bottom waters into nearshore zones. Moreover, strong wind events could: a) facilitate vertical mixing and both delay stratification in the late spring and bring about earlier turnover in the fall (i.e., decrease the period of hypolimnetic oxygen depletion) and b) favour a deeper thermocline (i.e., lead to a thinner hypolimnion with more rapid oxygen depletion). In short, while future wind patterns will likely impact hypoxia patterns, the magnitude (and even direction) of such effects is unclear. The effects of hypoxia on foodwebs (particularly invertebrate and fish communities), and how this may be influenced with climate change, are addressed below in section D.

Page 55: Taking Action On Lake Erie - International Joint Commission · reduce external inputs of total phosphorus (TP) and in particular, the rising proportion of the most bioavailable form

42

4. Harmful Algal Blooms (HABs) in Lake Erie: Status, causes and controls 4.1 Harmful Algal Blooms: Planktonic cyanobacteria Cyanobacteria (‘blue-green algae’) are bacteria, not true algae, and include a diversity of species which range in size and colour. Many are extremely beneficial and the smaller (<2 µm) pico-cyanobacteria are essential foodweb components of the more oligotrophic areas of the Great Lakes including the eastern basin of Lake Erie. The primary concern, however, is with the uncontrolled growth of the more noxious cyanobacteria, which respond to excessive nutrient levels and other environmental imbalances, , as many of these harmful cyanobacterial blooms (cHABs) have the potential to produce toxins. The formation of large swaths of cyanobacteria across Lake Erie is not a new phenomenon. Beginning in the early 20th century, a marked increase in phytoplankton biomass and decline in dissolved O2 were attributed to excessive P loading via point sources into the system (Charlton et al. 1993). While largely comprised of diatoms, the phytoplankton community also included cyanobacteria (Nicholls et al. 1977) which developed a fall peak in biomass referred to as the “autumnal maximum, (numerically13) dominated by N-fixing (Aphanizomenon, Anabaena) and non-fixing taxa (Lyngbya, Oscillatoria14, Microcystis; Davis 1954). Although some of these taxa may have been potentially toxic, cyanobacterial toxins were at that time not widely recognised as a threat and the focus remained on reduction of P, nuisance algal biomass and hypoxia. Restoration under the GLWQA effectively halved the P loading into the lake within 10 years and resulted in a reduction in total phytoplankton biomass of up to 89% in some areas of the lake (Charlton et al. 1993, Makarewicz 1993). Beginning in the mid-1990s, however, increases in highly bioavailable SRP loading to the lake were observed and a resurgence of cyanobacterial blooms (Conroy et al. 2005b; Michalak et al 2013). The identification of a toxic bloom of the colonial cyanobacteria Microcystis aeruginosa in Lake Erie in 1995 (Brittain et al. 2000) coincided with the onset of a second substantial decline in Lake Erie water quality and increasing annual occurrences of extensive cHABs. Since this first report, blooms of Microcystis aeruginosa and other cyanobacteria blooms have formed annually in the western basin, and are appearing elsewhere at inshore sites across the lake. In the last decade, the blooms have developed earlier and extended later than in the past; the 2011 Lake Erie bloom was visible from satellites until mid-October. 4.2 Current Status of cHABs in Lake Erie Microcystis aeruginosa is a colonial, bloom-forming cyanobacterium that has been identified in freshwater systems worldwide. Blooms of Microcystis have generally been relegated to the western basin of Lake Erie, but have recently been developing along the North shoreline of the Central and Eastern Basins (Watson et al unpublished data). Potentially toxic species of the cyanobacteria Planktothrix and Anabaena have also been observed with increasing frequency

13 It is important to note that abundance, used in earlier papers, inflates the relative importance of smaller celled taxa

(e.g. ‘ants vs elephants’) which are often more abundant but represent a smaller biomass than the larger

(filamentous, colonial) taxa. Current reporting uses ‘wet weight or biomass, calculated from estimated cell volumes

14 likely to have been same species that are now reclassified as Planktothrix

Page 56: Taking Action On Lake Erie - International Joint Commission · reduce external inputs of total phosphorus (TP) and in particular, the rising proportion of the most bioavailable form

43

in the west basin (Davis et al. 2012; Saxton et al. 2012). While cyanobacteria are known to produce a number of toxins, the group of hepatotoxins15 called microcystins are of particular concern. These toxins were first identified in 1959 from a Microcystis bloom in Saskatchewan (Bishop et al 1959) but have since been found in other cyanobacteria which often co-occur with Microcystis in Lake Erie, notably Anabaena and Planktothrix The genes encoding for microcystin synthesis (mcyA-J) have been well-characterized and are often used to identify toxin producers in the environment (e.g. Rinta-Kanto and Wilhelm 2006). Microcystins have been involved in numerous animal mortalities, and in a very limited number of cases, humans (Azevedo et al. 2002; Jochimsen et al. 1998) and have yet to be reported in the Laurentian Great Lakes (Watson et al 2008). Potential carcinogenic properties of microcystins are thought to be caused by their inhibitory effect on DNA repair mechanisms During the last decade, Microcystis and other potential toxic cyanobacteria have been identified in Lake Erie using a variety of microscopic, molecular, and imaging techniques. Early detection of cHAB formation is critical to formation of a proper response, and available detection methods have greatly improved in sensitivity and speed in recent years. Sensitivity and predictive ability of remote sensing data has extended warning periods before large scale blooms in Erie (Becker et al. 2009; Binding et al. 2012). Molecular methods have determined that Microcystis is not localized to near shore areas, but can be found in each basin across the lake (Dyble et al. 2008; Ouellette et al. 2006). Gene sequencing allows for the identification of potentially toxic cyanobacteria within a mixed bloom of toxic and non toxic cells and the source of high microcystin levels during bloom events (Ha et al. 2008). These techniques have tracked historical existence of Microcystis dating to the 1970s in Lake Erie, and show that at specific sites the current population is genetically indistinguishable from the historical population surveyed. This suggests that environmental or anthropogenic influences have resulted in a surge in the Microcystis population in recent years, not an invasion of a distinct population (Rinta-Kanto et al. 2009). As technology continues to develop, detection methods become more comprehensive. Next generation sequencing techniques are currently being employed to examine the structure of bloom communities, in addition to identifying the function of bloom community members in Lake Erie (Steffen et al. 2012). 4.3 Nutrients and Lake Erie cHABs Eutrophication has long been acknowledged as a driver for freshwater phytoplankton biomass. Phosphorus has historically been accepted as the key limiting nutrient in freshwater systems, and there is strong evidence that P is a key factor driving cyanobacterial biomass and dominance in many freshwater systems including Lake Erie (e.g. Downing et al 2001; Michalak et al 2013), but recent work suggests an important role for other nutrients (i.e., nitrogen), trace element availability, climate change, and the co-occurring heterotrophic community. An understanding of the collective impact of these factors will lead to more accurate mathematical modelling and prediction of future cHABs. Collective field evidence suggests that while P may control total phytoplankton biomass, the communities often show short-term N limitation and furthermore, the chemical speciation of N may influence algal community structure (DeBruyn et al. 2004; Dolman et al. 2012). Additional evidence suggests that urea, increasingly used in fertilisers, may facilitate Microcystis blooms and influence their toxicity (Davis et al. 2010; Finlay

15 Liver toxins

Page 57: Taking Action On Lake Erie - International Joint Commission · reduce external inputs of total phosphorus (TP) and in particular, the rising proportion of the most bioavailable form

44

et al. 2010). This has, in some circles, led to debate over dual nutrient management (Conley et al. 2009; Schindler 2012). Further study will be required to characterise the function of both N and P in governing phytoplankton biomass and toxicity in Lake Erie and other important freshwater systems. Additional consideration should be given to the impact of filter feeding zebra mussels which facilitate cHABs by increasing light penetration, recycling nutrients and selective filtration and digestion. For example, live Microcystis have been found in mussel pseudofaeces, while other forms of phytoplankton are digested. 4.4 Abiotic Factors, Climate Change and cHABs Microcystis and other cHAB species are favoured by warm temperatures (Paerl and Huisman 2009). IPCC climate models predict temperature increases that would favour cyanobacterial dominance in freshwaters and may contribute to an extension of annual bloom duration. Warmer temperatures have already resulted in an earlier thaw in ice cover on Lake Erie, with recent data showing above 0º C temperatures for the entire month of March 2012. How shorter ice seasons will affect cyanobacterial proliferation remains an unanswered question: shifts in community succession and nutrient cycles will no doubt occur and need to be considered in future studies. Climate change is projected to affect not only temperature trends, but also precipitation patterns, irradiance, and the number of storm events, all of which may potentially impact the success of cHABs in Lake Erie. The science of toxic blooms in Lake Erie has evolved from detection and identification to more comprehensive study of the factors influencing bloom success and toxicity. While current trends mirror a historic focus on the role of nutrients, the importance of climate change and total microbial community contribution are now recognized. Toxic bloom communities are complex systems comprised of microbes other than cyanobacteria; the role of heterotrophic bacteria in the success of HAB species has recently been highlighted in Lake Erie as well as aquatic systems worldwide (Sher et al. 2011; Steffen et al. 2012). Because heterotrophs shape nutrient dynamics and thus may influence the toxicity and physiology of cyanobacteria, it is becoming apparent that the entire bloom community must be examined (Dziallas and Grossart 2012; Steffen et al. 2012). This “systems biology” approach to studying bloom events moves beyond the study of a single organism, and will provide insight on the relative contributions of biotic and abiotic driving forces behind massive bloom events. This in turn will allow for a more comprehensive approach to modelling and forecasting efforts and may eventually lead to novel management practices in the field.

5. HNABS: Attached Algae 5.1 Overview Owing to its southern location, shallow depth, and variety of substrate types, LE provides favourable habitat for a diverse array of benthic algal taxa (Stewart and Lowe 2008). While over 200 species have been recorded, focus has been directed on those primarily responsible for nuisance benthic algal blooms (NABs), largely dominated by the filamentous green algae Cladophora, and more recently, large blooms of the cyanobacterium Lyngbya wollei. Negative impacts associated with these blooms include the fouling of recreational beaches, clogging of municipal and industrial water intakes, impaired water quality and potential microbial health risks to wildlife and humans. There are likely additional ecological impacts which are less well

Page 58: Taking Action On Lake Erie - International Joint Commission · reduce external inputs of total phosphorus (TP) and in particular, the rising proportion of the most bioavailable form

45

understood and much less well documented than those of anthropocentric origin. Nuisance blooms of Cladophora (and probably other benthic algae) were common in the 1950s to the early 1970s and a major reason for the implementation of phosphorus (P) controls under the GLWQA.

While there is evidence that these mitigation programs were successful in reducing the occurrence of noxious Cladophora blooms, the heightened focus of monitoring lake responses in offshore waters left much of the near shore un-monitored outside of a few site-specific locations impacted by P sources that were specifically reduced under the GLWQA. The return of blooms of Cladophora (and more recently, the mass appearance of Lyngbya) to the near-shore areas of Lake Erie in the past 10 – 15 years has coincided with attainment of offshore TP targets and other indicators of trophic status (i.e. chla), a shift in the dominant source of P loading (point vs. non-point) and pervasive ecological disturbance by invasive species, notably the filter-feeding zebra (Dreissena polymorpha) and quagga mussels (D. bugensis). Both species have demonstrated an ability to achieve significant biomass in near-shore and off-shore regions of all the Great Lakes (except Superior) and extensive research indicates that dreissenid mussels may enhance conditions for benthic algal growth through direct and indirect means. In addition to the direct and obvious impacts, both Cladophora and L. wollei are considered to be poor food resources for other aquatic biota. Thus the impacts of these blooms may have far reaching ecological consequences that are presently poorly characterized or quantified (but see Hudon et al. 2012). Because these changes have occurred more or less concurrently, it remains challenging to evaluate their relative roles in the resurgence of NABs. Questions remain as to whether past P

Figure 8. Locations where blooms of Cladophora (circles) or Lyngbya wollei (squares) have been recorded since

1995. Cladophora recorded as present (green circles), below (yellow circles) or above (black circles) the nuisance

biomass threshold of 50 g m-2

DW of Auer et al. (2010). Lyngbya wollei denoted as present (black squares) or absent

(white squares) as a nuisance threshold has not yet been defined.

Page 59: Taking Action On Lake Erie - International Joint Commission · reduce external inputs of total phosphorus (TP) and in particular, the rising proportion of the most bioavailable form

46

control programs sufficiently reduced P loading to control NABs, whether the shift to a dominance of diffuse non-point sources with high spatial and temporal variability contributes to these blooms, or whether the resurgence is simply a consequence of high biomasses of benthic filter feeders (i.e. dreissenid mussels) that have fundamentally altered the structure of the near-shore littoral zone to one that resembles the highly productive reefs found throughout oligotrophic regions of the world’s oceans.

5.2 Abiotic and Biotic Factors Associated with NABs in Lake Erie Cladophora requires hard substrate for attachment and generally achieves optimal photosynthesis and growth rates in slightly alkaline waters (pH ~ 8.0) with moderate turbulent energy and temperatures (15 – 22 °C), and high light availability (> 100 μmol m-2 s-1) and P (reviewed in Higgins et al. 2008b and references therein). In contrast, optimal conditions for L. wollei appear to occur in relatively sheltered and low energy environments (i.e. embayments) characterized by soft substrates, with comparatively lower light levels (< 50 μmol m-2 s-1), warmer temperatures (> 20°C) and high concentrations of phosphorus (Vis et al. 2008). In addition to factors such as light availability and nutrient levels (discussed below), the substrate distribution in Lake Erie (Figure 4) likely plays an important role confining blooms of Cladophora to the eastern basin (EB) and Lyngbya to the WB given the dominance of bedrock and glacial till in the near shore of the EB, and a mixture of till, sand, gravel and mud in the near shore of the WB (Figure 9).

Box 15: Benthic Algal Blooms – Overview of Issues

Over the past 10-15 years, study on benthic algae in Lake Erie has been largely limited to two

species that presently form annual nuisance blooms; Cladophora in the eastern basin, and more

recently, Lyngbya in the western basin.

Although the ecology of Cladophora is generally well documented in the Great Lakes (including

Erie) and elsewhere, far less information is available for Lyngbya.

Conclusions on temporal trends in degree of severity of nuisance blooms of Cladophora or

Lyngbya are hindered by the lack of standardized data collection protocols, reporting styles and

infrequent sampling. Heterogeneity in environmental conditions relevant to the growth of benthic

algae adds additional uncertainty.

Recent applications of remote sensing technology and mobile survey technology have

successfully documented spatial patterns in the coverage and attached biomass of Cladophora

and offer new approaches to expand the spatial scope of future research.

Page 60: Taking Action On Lake Erie - International Joint Commission · reduce external inputs of total phosphorus (TP) and in particular, the rising proportion of the most bioavailable form

47

Figure 9. Lake Erie substrate types. Reproduced from Haltuch et al. (2000) and used with permission. Copyright

(2000) by the association for the sciences of Limnology and Oceanography, Inc.

The common biotic and abiotic factors among basins of particular relevance to benthic algal blooms include; changing nutrient loads and near shore nutrient levels, the optical environment (light climate), and the abundance and biomass of exotic species such as dreissenid mussels and round gobies. The relevance of nutrient levels and light availability to benthic algae is relatively clear; an increase in nutrient levels or light availability will almost always lead to an increase in benthic algal production and biomass. The potential impact of dreissenid mussels on benthic algae h has been developed as a conceptual model that highlights their ability to alter the spatial and temporal dynamics of nutrient and energy cycling. This construct, termed “the near-shore shunt” is proposed to explain many of the observed ecological changes in the lakes, including nuisance blooms of Cladophora (Hecky et al. 2004). A number of studies have demonstrated the ability of both zebra and quagga mussels to excrete soluble forms of N and P at ecologically meaningful concentrations and rates (Table 3), improve water transparency and reduce phytoplankton biomass (Higgins and Vander Zanden 2010). Hecky et al. (2004) also suggested that in addition to providing suitable substrate for attachment of Cladophora, decomposition of faecal matter may supplement the C requirements benthic algae through increased CO2 efflux. Ultimately, the ability of mussels to translate these abilities into significant ecosystem impacts is predicated on the achievement and maintenance of sufficient biomass. Evidence of significant impacts of goby predation on both dreissenid and non-dreissenid prey (e.g. amphipods) have been documented in the EB of Lake Erie (Barton et al. 2005) and elsewhere. The relevance of such impacts to benthic algal blooms is not well characterized and may well be negative – excessive consumption of mussels may reduce their ability to beneficially alter conditions for benthic algae, or, the impacts may be positive - depending on excretion rates, consumption of mussels and other invertebrates may increase the amount of P available to algae above and beyond what would be provided by mussels in the absence of predation (Bunnell et al. 2005). Alternatively, selective consumption of herbivorous invertebrates may allow benthic algae to outgrow grazing losses sufficiently (Kipp and Ricciardi 2012).

Page 61: Taking Action On Lake Erie - International Joint Commission · reduce external inputs of total phosphorus (TP) and in particular, the rising proportion of the most bioavailable form

48

5.3 Western Basin (Lyngbya) The WB is the most turbid basin in LE owing to a combination of shallow depths which increase the susceptibility to wind driven re-suspension, a large suspended sediment load from tributaries such as the Maumee River (Richards et al. 2001) and elevated nutrients (above) that lead to high algal production. Satellite imagery and long-term assessment of monitoring records indicate that water clarity in the western basin has actually declined since the late 1990s (Barbiero and Tuchman 2004, Binding et al. 2007). The precise causes of these changes are difficult to ascertain, however it is likely that they reflect initial declines in phytoplankton biomass (Nicholls and Hopkins 1993) followed by a subsequent increase in later years as P loads increased (Conroy et al. 2005b), and/or reflect the dominance of non-phytoplankton turbidity owing to loadings of suspended sediments and re-suspension of mineral particles by wind events (Barbiero and Tuchman 2004, Peng and Effler 2010). Similar reversals of water clarity trends have been observed in other shallow, well-mixed systems invaded by mussels (e.g., Saginaw Bay; Pillsbury et al. 2002). The trend toward increasing turbidity may actually favour the development of blooms of L. wollei because it reportedly achieves optimal photosynthetic rates at ~ 50 μmol m-2 s-1 (Pinowska et al. 2007). In western LE, Lyngbya is found at depths of 1.5 – 3.5 m. Based on the average light attenuation this corresponds to ~ 1 – 4 % of surface irradiance and while apparently suitable to sustain Lyngbya growth, is far too low to support extensive production of Cladophora, and would restrict the location of Cladophora largely to depths <1 m, which is where it has largely been found in recent years (Ross 2006). Recent studies on L. wollei in the Saint Lawrence River also indicate that this species is positively associated with high dissolved organic carbon (DOC) concentrations (Levesque et al. 2012). Although counter to expectations for benthic autotrophs, the capacity for heterotrophic growth exists in some Lyngbya species (Burkholder et al. 2008), but the importance to Lake Erie strains is unknown at present. The potential impact of dreissenid mussels on Lyngbya blooms is relatively unknown, although it is commonly associated with mixed sand/mussel shell and live mussel substrates. Recent USGS surveys indicate that dreissenid populations are relatively stable with evidence of a gradual replacement of zebra mussels by quagga mussels (cited in Nalepa et al. 2012). The lower biomass and abundance of dreissenid mussels reported in the WB as compared to the EB is directly linked to the much smaller sizes of mussels sampled in the WB (Patterson et al. 2005). This implies that these are largely newly recruited mussels and the dominance of small mussels may be related to the dominance of softer and more unstable substrates around much of the littoral region that can be disturbed by wind and wave action (Patterson et al. 2005). The potential impact of round gobies on Lyngbya is also unknown. There is little information on what if any impact Lyngbya has on invertebrate abundance so it is unknown if gobies would forage within Lyngbya mats. Bunnell et al. (2005) used an individual-based and mass-balance modeling approach to estimate the potential impacts on P availability achieved by consumption of dreissenid mussels by the round goby. Despite the apparent minor impact on dreissenid population biomass, the P excreted by gobies as a result of dreissenid consumption was an average of 2.8 times more that the potential P excretion that would have occurred had the mussels not been consumed. The relative importance of this P to Lyngbya will ultimately depend on the spatial associations of Dreissena and Lyngbya, as well as gobies and their prey. Lyngbya are generally considered to be unpalatable to most benthic invertebrate for Lygnbya, toxin production (Lajeunesse et al. 2012) and a structurally thick sheath around the filaments reduces palatability to herbivorous grazers, consequently, grazing losses of Lyngbya are likely

Page 62: Taking Action On Lake Erie - International Joint Commission · reduce external inputs of total phosphorus (TP) and in particular, the rising proportion of the most bioavailable form

49

to be negligible, particularly for avoiding accumulations of excessive biomass. L. wollei, as a N-fixing cyanobacterium (Philips et al. 1992), will thrive in conditions of suboptimal N:P ratios. Depletion of NO3

- has been suggested to be important in shifting N:P to ratios where Lyngbya has a competitive advantage (Vis et al. 2008, Levesque et al. 2012), but the evidence collected to date indicates P loads as a likely culprit. Increased sediment re-suspension and blooms of planktonic algae such as Microcystis that are also stimulated by high P loads in the western basin (Conroy et al. 2005b, Bridgeman et al. 2012) will further contribute to the maintenance of a highly turbid, chronically light-limited benthic environment which Lyngbya is well suited to exploit 5.4 Eastern Basin (Cladophora) It is difficult to ascertain the status and long-term trends in nutrient concentrations in the near shore of eastern Lake Erie in large part because there has been no regular or consistent effort to monitor what are operationally defined as near shore waters (Kelly 2009). Although offshore data indicate that nutrient levels in the EB responded in a similar fashion to P controls in the 1970s, because the eastern basin is much deeper than the WB, it has a more clearly defined near-shore zone and conditions in the offshore may not be a suitable proxy. Unlike the WB, non-point sources of P are estimated to contribute ~20 % of the total load to the EB (WB is ~71 % non – point source; Baker et al. 2009). However, there is little information on the status or trends of non-point source P loads other than its relative contribution to the total loads which have been estimated to be generally stable through to 2002 (Dolan and McGunagle 2005). Most of the tributaries on the northern shore of LE routinely have TP concentrations above the 30 μg L-1 provincial water quality guideline (OME, 2002). Evidence has also emerged documenting a persistent increase in proportion of P exported from watersheds around LE since the 1990s (Han et al. 2012). Therefore, increased non-point source loading to the EB cannot be ruled out. Despite the high degree of uncertainty, much of the open near shore has been characterized as having nutrient levels similar to that of the off-shore environment since the late 1990s (North et al. 2012) and indicate that conditions have been much improved since the 1960s and 70s (Gregor and Rast 1979, 1982) when much of the Canadian near shore frequently exceeded 15 μg L-1 of TP. However, high-intensity sampling along shorelines impacted by tributary discharge (e.g., the Grand River) show significant variability in P concentrations that are a function of discharge, time of year, and plume behaviour which is dependent on the mixing and flow regime in the near shore, as well as the temperature of the interacting water masses (Howell et al. 2012). The disparity between vessel- and shore-based sampling is likely due to the depth limitations of vessel sampling in very near-shore environments.

Page 63: Taking Action On Lake Erie - International Joint Commission · reduce external inputs of total phosphorus (TP) and in particular, the rising proportion of the most bioavailable form

50

Figure 10. Kriged surface of turbidty (FTU) along a site in Lake Erie near Wainfleet Twp. Showing localized

reductions in turbidity and secchi depth (m) over bedrock areas (purple outline) consistent with a dreissenid filtration effect. Note increased turbidity near localized tributary input

Although the relative improvements in near-shore nutrient levels and the occurrence of widespread Cladophora blooms (Higgins et al. 2005a) are counterintuitive, the high and variable biomass of dreissenid mussels in the EB, particularly along the rocky northern shoreline may obscure patterns or trends in nutrient levels through filtering activity (Hecky et al. 2004). Such an effect is predicted as a consequence of the ‘near-shore shunt’ and observations of low nutrient levels over open near shore areas in conjunction with widespread Cladophora blooms are not inconsistent with this mechanism (Depew et al. 2011). It is also clear that near-shore regions in the vicinity of tributaries are subjected to variable nutrient and light regimes depending on the interacting nature of tributary plumes and lake currents (Rao and Schwab 2007). The importance of these transient features to Cladophora growth in the EB is not well studied. On one hand, there is limited evidence that tributaries had a limited and localized influence on Cladophora growth prior to the advent of dreissenid populations (Painter and McCabe 1987). On the other hand, the widespread growth of Cladophora that is now a characteristic feature in the EB (e.g. Higgins et al. 2005a) even in the absence of tributaries suggests that ambient conditions are sufficient to support nuisance blooms (e.g. Fig 11). However, it is not yet clear if the shunt mechanism facilitates greater particle and nutrient trapping efficiency of plume associated nutrients and thus has increased the relative importance of these features for supporting nuisance Cladophora blooms. The largest increases in water clarity in eastern Lake Erie generally occur in spring months when dreissenids would have access to the entire water column (Binding et al. 2007) and may be very important for the development of nuisance Cladophora blooms, because this is also a time of generally high nutrient concentrations, both on a lake wide and localized scale due to high discharges from snowmelt. Although there remains some uncertainty around the trajectory of nutrient levels in the near shore, the impact that dreissenid mussels have had on water clarity in the eastern basin is compellingly strong (Howell et al. 1996, Barbiero and Tuchman 2004, Binding et al. 2007). On

Box 16: NABs: Western Lyngbya and Eastern Cladophora blooms

Differences in substrate availability and light climate appear to be major determinants of Lyngbya and Cladophora abundance in LE. In the more turbid western basin, Lyngbya is often associated with sand and crushed and live dreissenids over a limited depth range (1.5 – 3.5 m), while Cladophora in the more transparent eastern basin is found attached to dreissenids, rocks and other hard substrate at depths between 0.5 – 10 m.

The advent of dreissenid mussel populations in Lake Erie has likely contributed to nuisance blooms of Cladophora by improving water clarity, supplying nutrients, and providing substrate for filament attachment. The degree of influence of dreissenid mussels on Lyngbya is less clear and merits further exploration.

Increased SRP loading from major tributaries in the western basin is the most parsimonious explanation for the recent development of Lyngbya blooms in the western basin. Evidence for similar trends in tributaries from the eastern basin is lacking, but the increase in proportional P exports observed in Lake Erie watersheds in recent years suggests that this cannot be ruled out at present.

High abundance and biomass of the round goby may modify dreissenid-Cladophora dynamics on a local scale, either through P excretion, grazing relaxation, however there is only circumstantial support for these mechanisms at present. Ideally, these interactions should be evaluated in habitats where mussels, Cladophora and gobies co-occur.

Page 64: Taking Action On Lake Erie - International Joint Commission · reduce external inputs of total phosphorus (TP) and in particular, the rising proportion of the most bioavailable form

51

the basis of satellite observations, secchi depths were estimated to have ranged from 2 – 4.5 m in the eastern basin during 1979 – 1985 and from 3 - 7 m during 1998 – 2005 (Binding et al. 2007). There is some debate as to whether the increased water clarity observed in the eastern basin and in other deeper parts of the Great Lakes is mediated by dreissenid filtration or a reduction in whitings or CaCO3 precipitation events (Barbiero and Tuchman 2004, Barbiero et al. 2006). Even so, dreissenids have been invoked as a possible mechanism contributing to a decline in whitings by sequestering large amounts of Ca2+ in shell material (Barbiero et al. 2006). However, the largest increases in water clarity in eastern Lake Erie are generally associated with spring months of April and May, a time period when dreissenids would have access to the entire water column (Binding et al. 2007). The exceptional increase in spring-time clarity may be very important for the development of nuisance Cladophora blooms, because this is also a time of generally high nutrient concentrations, both on a lake wide and localized scale due to high discharges from snowmelt. The importance of dreissenid mussels to the development and generation of nuisance Cladophora blooms is reasonably well supported by in situ and modeling studies (Higgins et al. 2005a,b, 2006, Tomlinson et al. 2010). Studies in Lakes Michigan (Dayton 2011) and Ontario (Martin 2010) indicate that the water directly above mussels can become significantly enriched with SRP (2 – 8 fold higher) when compared to typical surface-water samples. If mussels are indeed a major factor then changes in biomass and abundance of mussels, or changes in the depth distribution of mussels may alter bloom dynamics. Potential changes to dreissenid population structure or depth distribution is important. Gobies appear to have nearly eliminated dreissenids between ~ 4 -14 mm in size, leaving only newly recruited and much older mussels (Barton et al. 2005). The EB is less productive than the western or central basins and the consumption of mussels may well be higher here (Barton et al. 2005). Based on correlations between Cladophora biomass and round goby biomass and abundance, it has also been suggested that gobies may relax grazing pressure on benthic algae such as Cladophora (Kipp and Ricciardi 2012). Like Lyngbya, Cladophora is generally considered to be unpalatable to most benthic invertebrates, either due to a lack of nutritional value relative to benthic diatoms or difficulty ingesting large filaments (Dodds and Gudder 1992). The degree to which herbivorous grazer control may (or may not) operate in eastern Lake Erie is unclear and has not been examined in detail. However of the invertebrate taxa that remain in the near-shore in the post-dreissena period, only Gammarus fasciatus has been observed to graze Cladophora in laboratory experiments (Szabo 2004, Malkin 2009). However, the abundances and reported grazing rates are thought to be insufficient to remove much biomass in the field (Malkin 2009). 5.5 Modeling in support of management Mathematical models were developed in the 1970s to characterize the response of Cladophora growth to point source P pollution (e.g., Auer et al. 1982) and were largely successful in predicting the response of Cladophora growth to subsequent P controls (Canale and Auer 1982a). In recent years, the Canale and Auer (1982b) model has been redeployed with some adjustments as the “Cladophora Growth Model” (CGM; Higgins et al. 2005b), and the “Great Lakes Cladophora Model” (GLCM; Tomlinson et al. 2010). The primary focus of these model revisions was to evaluate proximate causes for the apparent resurgence of Cladophora in Lakes Erie, Ontario and Michigan (Higgins et al. 2006, 2012, Malkin et al. 2008, Tomlinson et al. 2010) primarily in response to ecosystem changes related to dreissenid mussel invasion. Brief but detailed summaries for each study are provided in the full report. In summary:

Calibrated growth models for Cladophora show promise for representing the growth

Page 65: Taking Action On Lake Erie - International Joint Commission · reduce external inputs of total phosphorus (TP) and in particular, the rising proportion of the most bioavailable form

52

response across a variety of locations in the Great Lakes under a variety of environment forcing conditions. Models for other benthic algae such as Lyngbya have not been developed specifically for the Great Lakes.

Hindcast simulations have indicated that increased water clarity (attributed primarily to dreissenid mussel filtering) is largely responsible for the increased shoreline fouling by Cladophora since the late 1990s and early 2000s.

The models for Cladophora are data intensive and require inputs with sufficient spatial and temporal characterization of light, nutrient, thermal and disturbance regimes. Integration of Cladophora models into complex 3D hydrodynamic lake models offer a means to simulate forcing conditions to evaluate responses on lake –wide to local scales.

Because the models are highly sensitive to relatively small changes in SRP, it is imperative that P kinetics and the utility of water column SRP as measures of P availability be re-evaluated under present conditions. Additional modeling of P flux from dreissenids are now possible based on experimental and theoretical approaches and are being integrated with lake and Cladophora models. The potential importance of other forms of available P (i.e. DOP) to Cladophora may merit further investigation.

5.6 Potential Impacts on NABs under Climate Change

Temperature: Air temperatures in the Great Lakes region are projected to increase 1.4 0.6 °C

over the near term (2010 – 2039), 2.0 0.7 °C under lower and 3.0 1.0 °C under higher

emissions by mid-century (2040 – 2069) and by 3.0 1.0 °C under lower and 5.0 1.2 °C under higher emissions by end-of-century (2070 – 2099) relative to 1961 – 1990 (Hayhoe et al. 2010).Lake waters are expected to warm faster than regional air temperatures because the loss of winter ice cover facilitates the earlier onset of radiant energy absorption and stratification (Austin and Coleman 2007). (Table 10).

Cladophora growth models indicate that the primary factor responsible for the increased production and subsequent nuisance accumulation of Cladophora along shorelines is increased light transparency in the post-dreissenid era (Auer et al. 2010). Although there remain legitimate uncertainties in the models at present, they may still provide useful information for forecasting potential responses to changes in forcing variables due to climate warming. We improved earlier modelling efforts to determine the effects of climate on Cladophora growth by extending the modelling period to cover the longer growing season predicted. Details of the model are presented in the full report.

Table 11. Summary of measured warming rates offshore surface waters (SW Rate; °C yr-1

) and air

temperatures (AT Rate; °C yr-1

) Data from Austin and Coleman 2007, Dobiesz and Lester 2009

Basin SW Rate (°C yr-1

) AT Rate (°C yr-1

) Period

Erie – wholeb

0.024 0.037 1968 – 2002

Erie – Eastb

0.021 0.037 1968 – 2002

Erie – Centralb

0.014 0.037 1968 – 2002

Erie – Westb

0.072 0.037 1968 – 2002

Erie – Westa

0.010 0.02 0.011 0.02 1979 – 2006

Page 66: Taking Action On Lake Erie - International Joint Commission · reduce external inputs of total phosphorus (TP) and in particular, the rising proportion of the most bioavailable form

53

Two permutations were employed to model the response of Cladophora under these scenarios; the first (B and C) assumed that other biota (phytoplankton/zooplankton) would also follow a similar seasonal advancement trajectory and the SRP and kPAR profiles were advanced to the same degree as temperatures. The second (B1 and C1) assumed that only the water temperature change occurred, and SRP and kPAR levels remain seasonally as in 2002. These scenarios are summarized in Table 11.

Table 12. Summary of modelled scenarios to evaluate response of Lake Erie Cladophora to climate warming

Scenario Model Year Days of spring

advancement

Total warming added (C)

Days of fall extension

A 2002 0 0 0 B 2050 8 1.01 0

B1 2050 8 1.01 0 C 2100 18 2.06 0 C1 2100 18 2.06 0

Model output indicated that under projected climate change, the seasonality of Cladophora growth is likely to advance as water temperatures warm in the spring, with peak biomass attainment achieved around 15 to 18 days earlier (Figure 11). This is consistent with observed inter-annual patterns in Lake Ontario between warm and cool years (Malkin et al. 2008) and observations of early growth in the vicinity of the thermal discharges from OPG Nanticoke in Lake Erie (Kirby and Dunford 1982). Both the peak depth integrated attached seasonal biomass (Figure 11) and total cumulative production (attached + sloughed) (Figure 11) are lower (22 – 25 % and 15 – 17 %) respectively under the warming scenarios outlined. If the vast majority of re-growth in the spring arises from over-wintering filaments rather than zoospores (Mason 1965), then the modelled projections may underestimate spring growth if a lack of ice scour in winter months fails to remove filaments. There is some evidence for generation of multiple growth cohorts under scenarios B and C but these are amplified under B1 and C1 (Figure 11). This likely represents the response to higher

Figure 11. Modelled seasonality of cumulative Cladophora biomass (0.5 – 10 m depth) under scenarios of climate

warming. Note the increased peaks of re-growth after the main detachment event in scenarios B1 and C1

Page 67: Taking Action On Lake Erie - International Joint Commission · reduce external inputs of total phosphorus (TP) and in particular, the rising proportion of the most bioavailable form

54

SRP concentrations in scenarios B1 and C1 because the SRP profile remained seasonally consistent with Scenario A rather than advance with water temperatures. Precipitation: Climate models predict increases in winter and spring precipitation of up to 20 % under lower and 30 % under higher warming scenarios by 2070 (Hayhoe et al. 2010). However, predictions for summer months are generally inconsistent. The competing effects between shifting precipitation patterns and warmer temperatures (greater evaporative losses) do not

suggest that major changes in water levels are in store for the Great Lakes at present. Using climate model projections, Hayhoe et al. (2010) estimate that changes in mean water levels at the end of century will range from -0.2 m (Lake Superior) to -0.6 m (Lake Michigan/Huron). Lake Erie is expected to be, on average, -0.3 m lower in 2099 than present (Hayhoe et al. 2010). While this will undoubtedly result in the loss of some suitable growth habitat in very shallow waters (0.3 m or less), the corresponding gain in suitable habitat from deeper waters is not readily calculated owing to the 1 m resolution of lake-wide bathymetry. Unless water level change dramatically differs from projected, it seems that fluctuating water levels will not have a strong effect on Cladophora or other attached algae in the near future. In addition to shifting precipitation patterns, the intensity and frequency of precipitation events is also projected to change. It has recently been suggested that the linkage of extreme events (precipitation, heat waves, droughts) is related to a relaxation of poleward gradients in temperature and pressure in the upper atmosphere that reduce the rate of Rossby wave progression across the upper atmosphere (Francis and Vavrus 2012). Consequently, weather patterns in mid-latitudes will move more slowly and often persist for longer periods of time in a given location. This reduction in speed at which weather systems transit an area will increase the probability that flooding may result from rainfall events or drought from dry periods. However, rainfall events may not always be translated into loads, particularly if land use changes (i.e. catch basin installation in urban areas, recovery of agricultural lands to shrubland/forest) modify the amount of overland runoff reaching waterways. A more appropriate metric to evaluate is the discharge from gauged tributaries. An analysis of discharge events from gauged Ontario tributaries to the east and central basins of Lake Erie (Table 7) suggests that the number of days in a year when extreme discharge events are observed (extreme event

defined as discharge that exceeds the 3 level of probability) varies between 2 and 10, but appears to have declined in the most recent decade. The percentage of extreme discharge events occurring during prime algal growth periods (May – October) is also declining in 3 of the 4 tributaries examined, which is consistent with a shift to more winter – spring precipitation.

Page 68: Taking Action On Lake Erie - International Joint Commission · reduce external inputs of total phosphorus (TP) and in particular, the rising proportion of the most bioavailable form

55

Reasons for these changes are not clear at present but may be related to land use change within the selected catchments. The importance of these events to Cladophora (and also perhaps to Lyngbya) are not well understood. On one hand, increased event intensity may significantly increase loadings (Han et al. 2012) and it is generally well accepted that further increases in P loading (particularly as SRP) will almost certainly result in a worsening of conditions. The relevance of these events (or lack thereof) to Cladophora can be estimated with the modeling approach; however there remains no satisfactory method by which to derive an estimate of SRP loading unless suitably monitored. Moreover, if the bulk of the P load is indeed in the particulate form, then the immediate availability to Cladophora or Lyngbya may be questionable unless it can be effectively captured and remineralized. Based on the structure of the CGM, the potential impacts of severe discharge events that transiently elevate P concentrations can be estimated. For example, changes at times of high biomass will exert proportionately more impact on total biomass when biomass crop is high (summer) than when it is low (early spring). Because of the effect of SRP within the model framework on net specific growth rates (multiplicative), transient increases in SRP concentrations during the summer will be more relevant to the accumulation of Cladophora biomass than those in the spring when biomass is low. Circulation: Due to the increased relative importance of non-point source loads to the total Lake Erie P load (Dolan and McGunagle 2005), understanding the average and variability of large scale circulation patterns in Lake Erie may be of critical importance for Cladophora and Lyngbya blooms because the mixing, circulation and temperature effects will structure the light and nutrient regime in the near shore and affect the timing of productive periods (both pelagic and benthic). Wind speeds above the lake surface are estimated to be increasing on the order of 0.05 m sec-1 yr-1 (Austin and Coleman 2007). Given that the average wind speed is 4 – 6 m sec-

1, this is not insignificant. The increasing wind speed is largely attributed to a destabilized atmospheric boundary layer as a consequence of a reduced density gradient near the lake surface (Desai et al. 2009). Momentum from wind is therefore more readily transferred to the water, and in theory, should increase current speeds, decreasing mixing time, lead to deeper mixed layer depths and a stronger stable stratification. Inter-annual variability in current flows within a large body of water such as Lake Erie is difficult to project. Models indicate that such inter-annual variation is probably small (Bennington et al. 2010), thus it may be possible to define specific near shore physical regimes for areas or regions requiring management action. On the other hand, shifting wind trajectories may make such predictions difficult (e.g., Waples and Klump 2002). 5.7 NABS: Summary, conclusions and recommendations for future work The issue of nuisance benthic algal blooms in Lake Erie merits sustained programs of integrated research and monitoring, because the symptoms (both anthropocentric such as beach fouling, and ecological such as botulism outbreaks) of impairment to coastal regions in Lake Erie are so pervasive that they cannot feasibly be ignored. It is clear that the synergistic impacts of human activity, invasive species and climate change present an extraordinary challenge in developing management objectives for nuisance blooms of benthic algae in Lake Erie. At this point, much of the information regarding nuisance benthic algal blooms in the Great Lakes in the past (and in more recent years) has been limited to site-specific assessments, sometimes supplemented with experimentation and simulation modeling. We now know that

Page 69: Taking Action On Lake Erie - International Joint Commission · reduce external inputs of total phosphorus (TP) and in particular, the rising proportion of the most bioavailable form

56

there are a number of important factors that influence the dynamics of benthic algal blooms in near shore waters of the Great Lakes. Hydrodynamic and circulation of water masses shape the interaction of lake water with land-based runoff and tributary discharges, and strongly influence the nutrient, light, temperature and disturbance regimes in the near-shore. In addition, we now have a stronger sense of the ability of filter feeding organisms such as dreissenid mussels to act as a benthic – pelagic coupling mechanism, which may attenuate or exacerbate conditions suitable for the growth of benthic algae. What is lacking at present is a comprehensive understanding of how these various factors work together to create the attendant conditions associated with nuisance blooms of Cladophora and Lyngbya. This understanding is crucial for the sound development of management activity.

6. Lake Erie Fish Populations: Status and impact of Climate Change 6.1 Current Status While much research has explored the impacts of climate change on aquatic ecosystems, the expected impact(s) on the fisheries of large freshwater ecosystems such as the Great Lakes remains largely under-studied, particularly in respect to how climate change might indirectly influence fish populations and production through nutrient pathways. To address this gap, we evaluate mechanisms by which climate change (including both warming and altered precipitation patterns) can modify nutrient loading patterns and in-lake factors to influence Great Lakes fishes, focusing on Lake Erie. Lake Erie has a long history of responding rapidly to variation in nutrient inputs from the watershed; it supports a diversity of ecologically, recreationally, and economically important fish and fisheries that have different physical (e.g., temperature), chemical (e.g., DO), and biological (e.g., prey) requirements, and it has been experiencing warming trends and altered precipitation patterns that are typical of changes observed in other north-temperate ecosystems (including other North American Great Lakes basins). The Lake Erie ecosystem supports a species-rich and diverse fish assemblage, with more than 130 species being documented. In addition to their important ecological roles, several species also support large recreational and commercial fisheries. For example, walleye (Sander vitreus) supports the lake’s most valued recreational fishery and yellow perch (Perca flavescens),

Box 16: NABs – Impact of Climate Change

Projected growth dynamics of Cladophora under scenarios of climate warming indicate that the

seasonal growth dynamics may be appreciably altered, but the ultimate response will depend

largely on the interactive effects of nutrient, light and disturbance regimes in a longer

photoperiod (earlier seasonal growth).

Impacts on Cladophora in relation to changing precipitation patterns and increased non-point

source P loadings are not known, but based on the structure of the CGM will likely be more

important during periods of elevated biomass than low biomass.

Impacts of changing circulation are also poorly characterized but may be important given the

ability of near shore hydrodynamics to impact the thermal, nutrient, light and disturbance regime

in the near shore where Cladophora and Lyngbya grow.

Page 70: Taking Action On Lake Erie - International Joint Commission · reduce external inputs of total phosphorus (TP) and in particular, the rising proportion of the most bioavailable form

57

walleye and several other species support large commercial fisheries. Owing to the economic, ecological, and cultural import of fishes in the Lake Erie basin, fisheries management agencies have been working closely with the Great Lakes Fishery Commission to both protect—and when appropriate—rehabilitate native fish populations in the basin (Ryan et al. 2003). Lake Erie western, central, eastern basins vary in terms of a) water volume and depth (both increase from west to east), b) temperature and lower food web biomass and production (both decrease west to east), and c) fish community structure (Leach and Nepszy 1976, Sly 1976, Ludsin et al. 2001, Ryan et al. 2003). In general, the shallow, warm, and productive WB is dominated by species that are tolerant of high turbidity (i.e., low light levels) and warm temperatures, including temperate basses (Morone spp.), ictalurids, centrarchids, and cyprinids, with cool-water species such as yellow perch and walleye also being high in abundance during most of the year. By contrast, the EB, which is the deepest, coldest and least biologically productive basin, is typified by a deepwater fish community (e.g., salmonines, coregonines, and smelt) that prefers cold temperatures, high dissolved oxygen levels, and non-turbid waters. The CB, which falls intermediate in terms of temperature, productivity, and depth, correspondingly is dominated by cool-water species such as yellow perch and walleye, with both warm-water and cold-water species also present to a lesser degree. 6.2 Effects of Hypoxia Seasonal hypoxia may influence fishery production both directly and indirectly, at multiple temporal and spatial scales. There are direct physiological costs for aerobic organisms occupying hypoxic habitats. Further, organisms that can readily avoid hypoxic regions (through vertical or horizontal migrations) may be forced to occupy inferior thermal and optical habitats, immediately constraining growth. Such behavioural migrations may alter the spatiotemporal overlap, efficiencies, and vulnerabilities of predators and prey, leading to long-term changes to food-web structure and energy flow. The most direct effect of hypoxia is mortality due to sudden or prolonged exposure to low ambient oxygen concentrations (Burnett and Stickle 2001). In addition, hypoxic events can trigger various sub-lethal behavioral and physiological responses, which can cause dramatic indirect effects on aquatic food webs (e.g., Prepas et al. 1997; Rabalais and Turner 2001; Taylor et al. 2007, Ludsin et al. 2009). For example, several researchers have documented that fish (e.g., Aku et al. 1997) and zooplankton (e.g., Field and Prepas 1997; Qureshi and Rabalais 2001) adjust their vertical and horizontal distributions in response to low oxygen events. Such behavioural responses can impact the spatial overlap (and thus encounter rates) of predators and prey, and may lead individual organisms to concentrate in confined areas (i.e., leading to compensatory density-dependent effects; Eby and Crowder 2002) and/or occupy inferior thermal and optical habitats. Physiological responses to low oxygen (e.g., shift to anaerobic respiration, reduced activity levels, and decreased disease resistance) also can indirectly impact trophic interactions by affecting foraging abilities and prey vulnerabilities (Burnett and Stickle 2001). Ultimately, the combined direct and indirect effects of hypoxic events may have far-reaching trophic consequences, impacting the vital rates (growth and mortality) of multiple populations (both benthic and pelagic). While the entire suite of hypoxic impacts on the biota of Lake Erie is unknown, several lines of evidence suggest responses at individual, population, and community levels of biological organization: a) Coincident shifts in invertebrate/fish community composition with historical patterns of hypoxia. b) Laboratory trials and field observations demonstrate that many species

Page 71: Taking Action On Lake Erie - International Joint Commission · reduce external inputs of total phosphorus (TP) and in particular, the rising proportion of the most bioavailable form

58

found in Lake Erie are directly affected by low DO and will adjust vertical/horizontal distributions and diets in the presence of hypolimnetic hypoxia. c) Mechanistic models suggesting that habitat quality, growth rates and population biomasses may all respond to hypolimnetic hypoxia in central Lake Erie. Laboratory studies demonstrate that many Lake Erie zoobiota respond to direct exposure to low DO. Survival capacity under hypoxia varies among species and life stages. For example, relatively tolerant fish species like yellow perch are able to survive at lower DO compared to more sensitive species, like rainbow smelt. Similarly, while some Lake Erie zooplankton species experience poor survival under hypoxia (e.g., Daphnia mendotae; Goto et al. 2012), other taxa are seemingly able to survive under prolonged hypoxic conditions. Such taxa-specific direct effects of hypoxia are not limited to mortality. Casselman (1978) reviewed the environmental requirements of northern pike, a cool-water fish, and reported that any decrease in oxygen saturation had a direct negative effect on growth. Roberts et al. (2011) demonstrated that both consumption and growth rates of yellow perch decline in low oxygen conditions. In contrast, growth rates (as indexed by RNA:DNA ratios) of tolerant Chironomidae show no response to extremely low oxygen concentrations (Höök and Roberts unpublished data). While direct hypoxia exposure has the potential to affect survival, consumption and growth rates of many Lake Erie zoobiotic taxa, for mobile species such effects may never manifest as species simply adjust their location to avoid hypoxia. Vanderploeg et al. (2009a) used an optical plankton counter and hydroacoustics to evaluate vertical and horizontal distributions of collective zooplankton and fish biomass in central Lake Erie before, during and after development of hypolimnetic hypoxia. While a fraction of mesozooplankton utilized hypoxic areas, the majority of zooplankton and fish biomasses appeared to entirely avoid hypoxic waters (Vanderploeg et al. 2009a). While taxaon-specific spatial distributions are generally consistent with these patterns, certain fish and zooplankton utilize hypoxic waters. Vanderploeg et al. (2009b) evaluated taxa-specific vertical distributions and documented a diversity of responses: 1) some epilimnetic species remained in the epilimnion regardless of hypoxia and were unaffected (e.g., Leptodora); 2) some migrated from the epilimnion to hypolimnion under normoxic conditions, but avoided the hypolimnion under hypoxia (e.g., Bythotrephes); and 3) some continued to move into the hypolimnion even during hypoxia (e.g., Bosmina). Similarly, species-specific responses are evident for fishes. For example, some species, e.g., emerald shiners (Notropis atherinoides), primarily occupy the epilimnion before hypoxia development and remain in this top layer during the hypoxic period. In contrast, other species entirely avoid hypoxic waters. The obligate demersal16 round goby is present in the offshore bottom waters of central Lake Erie prior to hypoxia, but subsequently migrates horizontally and is entirely absent from this region once bottom hypoxia establishes (Höök, Pothoven and Ludsin unpublished data). Intolerant, cold-water rainbow smelt also entirely avoid hypoxic waters and respond by either a) moving up in the water column (to a thin metalimnetic zone with moderate temperatures and sufficient oxygen) or b) migrating horizontally to other regions of Lake Erie (e.g., eastern basin), where suitable thermal and oxygen conditions overlap. Finally, Lake Erie yellow perch display a complex response to hypoxia. While a portion of the yellow perch population appears to move horizontally away from the hypoxic region, many yellow perch remain in this region, but adjust their vertical distributions to zones above the hypoxic waters

16 live and feed on or near the bottom

Page 72: Taking Action On Lake Erie - International Joint Commission · reduce external inputs of total phosphorus (TP) and in particular, the rising proportion of the most bioavailable form

59

(e.g., Roberts et al. 2009 and 2012). Moreover, some of these yellow perch remaining in the hypoxic region undertake forays into the hypoxic zone; evidently diving into the hypoxic zone to feed on benthic invertebrates and then returning to the oxygenated epilimnion or metalimnion (Roberts et al. 2012). It is attractive to hypothesize that hypolimnetic hypoxia should lead to negative indirect effects on fish growth. Specifically, a) shifts in diets away from preferred prey, b) increased metabolic costs through occupation of higher, warmer zones in the water column, c) activity costs of long-distance horizontal migrations, and d) compensatory density-dependent effects through vertical and horizontal compression should collectively lead to reduced growth and population production rates. However, documenting such effects is not straightforward given that hypoxia in Lake Erie persists for a relatively short time period (i.e., 1 week to 2 months) and its effect on growth can be difficult to distinguish from a variety of other seasonal processes. Moreover, while nutrient additions can exacerbate hypoxia, such loadings can also increase system productivity, leading to greater prey densities and increasing growth and production rates. This positive effect may be most beneficial for epi- and meta-limnetic species that may feed at higher rates when their prey is concentrated into higher portions of the water column. For example, emerald shiners may benefit when zooplankters are forced up in the water column (Pothoven et al. 2009) and walleye may benefit when their piscine prey is concentrated in the metalimnion (Brandt et al. 2011). Further, these effects may not only increase prey vulnerabilities but also prey production rates, as many zooplankton taxa can express greater individual growth rates higher in the water column (i.e., where temperatures, light levels and phytoplankton are all greater; Goto et al. 2012). 6.3 Effects of Water Clarity Increases in temperature and high-flow tributary events are expected to occur with climate change and are likely to influence foraging and growth environments for fishes, especially those species that rely on vision to detect and capture prey (i.e., all commercially and recreationally important planktivores and piscivores). In general, water clarity will likely decrease in Lake Erie with continued climate change, owing to an increased frequency of precipitation-driven storm events during winter and spring that increase sediment and limiting nutrient (P) runoff from the watershed. These reductions in water clarity also are likely to be magnified by water warming and a longer growing season that stimulate phytoplankton production, especially positively buoyant, inedible species such as cyanobacteria (Paerl and Huisman 2008). Specifically, we expect that water clarity will decrease during spring primarily due to tributary loading of inorganic sediments from largely denuded agricultural landscapes, whereas during summer and fall, water clarity will be reduced primarily by biogenic turbidity such as noxious cyanobacteria (e.g., Microcystis spp.) that thrive under high temperatures and have the ability to store P for prolonged periods of time through luxury consumption (Elrifi and Turpin 1985, Paerl and Huisman 2008). Water clarity is a highly dynamic habitat attribute of aquatic ecosystems that has been shown to modify fish species behaviour (e.g., activity, foraging, mate selection and inter-species interactions (e.g., competition, predator-prey). Prominent studies have documented impacts of water clarity (or turbidity; i.e., suspended particles that cause “cloudiness” in water) on interactions involving visual foragers such as zooplanktivorous and piscivorous fishes. Reduced water clarity decreases prey detection by visual predators through low ambient light intensity and diminished contrast between prey and its background (Lythgoe 1979). Thereby, in the absence of other factors, interactions between visual foragers and their prey are lessened when

Page 73: Taking Action On Lake Erie - International Joint Commission · reduce external inputs of total phosphorus (TP) and in particular, the rising proportion of the most bioavailable form

60

turbidity increases. Such an alteration of predator-prey interactions can influence fish recruitment by affecting early life growth and survival (e.g., turbidity can provide a refuge from predation; Reichert et al. 2010), mediate trophic interactions and energy flow through aquatic food webs (e.g., weaken trophic cascades, increase trait-mediated indirect interactions; Pangle et al. 2012), and cause shifts in community composition (e.g., from vertebrate to invertebrate planktivore assemblages, from piscivorous to planktivorous assemblages). Spatiotemporal variability in water clarity is driven by multiple factors. Water clarity is primarily driven by the amount of suspended particles in the water column (i.e., turbidity), including both inorganic sediments and biogenic materials such as phytoplankton. Thus, those factors that influence turbidity hold great potential to indirectly affect fishes by altering their ability to consume prey and evade predators. Two such external factors include weather (e.g., precipitation) and watershed land use, which govern the timing, magnitude, and duration of sediment and nutrient inputs into downstream ecosystems through river inflow or overland runoff. Internal turbidity controls also exist, including biological controls (e.g., sessile filter-feeders such as invasive zebra and quagga mussels), whose water-filtering abilities have contributed to enhanced water clarity, and physical controls (e.g., wind-driven upwelling and water-mixing events), which can reduce water clarity through the resuspension of nutrients and sediments from bottom waters. The impact of water clarity on biological interactions involving fishes also can be influenced by multiple factors, including attributes of both the forager and foraging environment. Each species has a unique ability to tolerate turbidity. In turn, these species-specific tolerances of turbidity have helped explain variation in fish community composition among ecosystems. For example, in Europe ruffe (Gymoncephalus cernuus) tend to dominate the fish biomass in highly productive, turbid lakes of low light penetration, owing to sensory organelles that allow it to continue to feed under dim light conditions (Persson et al. 1991). Conversely, European perch (Perca fluviatilis) out-compete ruffe for prey in low productivity (low turbidity) ecosystems, and hence, typically dominate there. The trophic position of the forager also can be important, with piscivores being more negatively impacted by turbidity than planktivores (De Robertis et al. 2003). This finding may help to explain why planktivore to piscivore ratios tend to increase in highly eutrophic, turbid ecosystems (De Robertis et al. 2003). With respect to the role of the foraging environment, Wellington et al. (2010) showed experimentally that the negative effect of biogenic (algal) turbidity on planktivory by intermediate consumers (i.e., larval and juvenile fishes) was greater than the effect of inorganic (sediment) turbidity, with biogenic turbidity impairing foraging of larvae and juveniles more than sediment turbidity, likely owing to the reflective properties of sediments increasing the contrast of prey relative to algae. As is explained more fully below, perceived predation risk by the intermediate consumer (e.g., planktivorous fishes) also can be important, with different foraging rate response curves in relation to turbidity expected under conditions without predation risk (negative, linear) than in its presence (unimodal) (Pangle et al. 2012). 6.4 Historical Patterns While several long-term studies have evaluated fish community change in Lake Erie during the past half century, none has explicitly investigated or clearly identified an impact of changing water clarity on fish populations. Ludsin et al. (2001) found that a long-term (1969-1996) shift in western Erie from a fish species assemblage tolerant of eutrophic conditions (e.g., reduced water clarity) to one less tolerant of eutrophy coincided with a reduction in system productivity, including an increase in water clarity and speculated that reduced water clarity provided a

Page 74: Taking Action On Lake Erie - International Joint Commission · reduce external inputs of total phosphorus (TP) and in particular, the rising proportion of the most bioavailable form

61

foraging advantage to species intolerant of eutrophy that allowed them to out-compete more tolerant ones. However, because other factors besides water clarity changed simultaneously with oligotrophication during this time period (e.g., P inputs declined, several invasive species established populations, bottom hypoxia declined), Ludsin et al. (ibid) could not definitively identify the role of water clarity change and it remains an open question whether water clarity can affect fish community composition in Lake Erie. 6.5 Theoretical and Empirical Studies Despite no definitive evidence to demonstrate that water clarity has driven long-term patterns in Lake Erie fish assemblages, two field investigations have demonstrated a true impact of water clarity on larval yellow perch foraging and recruitment to the juvenile stage. Reichert et al. (2010) and Ludsin et al. (2011) showed that turbid, open-lake plumes produced by the Maumee River during spring (2006-2009) provided a recruitment advantage to larvae residing inside the plume over those outside of it. Further, Ludsin et al. (2011) used field habitat data, larval otolith growth information, predator abundance and diet data, and an individual-based model to clearly demonstrate that the positive effect of the Maumee River plume on yellow perch larvae emanated more from low water transparency (high turbidity) that reduced predation risk than from bottom-up driven effects on zooplankton availability and larval yellow perch foraging and growth. The importance of turbid plume production in western Lake Erie is fully highlighted by Ludsin et al. (2010), who found a very strong correlation (r = 0.99) between Maumee River plume size during spring and an index of juvenile yellow perch abundance in August, i.e., a strong predictor of the numbers of adults that recruit to the fishery at age-2 (Ludsin et al. 2001). 6.6 Ecological Effects of Basal Prey Production Food web structure and dynamics dictate the flow of energy through natural systems. Food web constituents are co-responsive such that alterations of a single component can cascade to affect divergent aspects of the network through both bottom-up and top-down processes. Therefore an appreciation of processes throughout the entire food web is necessary if one is to anticipate the consequences of stressors such as eutrophication. Over the past century, Lake Erie fish and invertebrate assemblages have shifted in response to eutrophication and other stressors. Future climate change (and resulting physical changes) will undoubtedly interact with existing stressors to further modify not only biotic composition and biomasses but also nutritional makeup of the prey base. By affecting physicochemical, biomolecule and phytoplankton conditions, increased nutrient loading and resulting eutrophication can lead to a suite of impacts to lower trophic level consumers. Hypolimnetic hypoxia will favour benthic invertebrates that can tolerate low oxygen conditions (e.g., Oligochaeta and Chironomidae) over more sensitive organisms (e.g., Hexagenia spp.). Similarly, some zooplankters are able to survive under low oxygen and may even use a hypolimnetic hypoxic zone as a refuge from predators (e.g., Vanderploeg et al. 2009b) and thereby alter planktivore-zooplankton interactions (Pothoven et al. 2012). Moreover, zooplankton taxa differ in their ability to feed under low light and to exploit various types of phytoplankton. Thus, low light and high cyanobacteria densities may lead to emergence of a eutrophic zooplankton assemblage (e.g., high biomass of cyclopoid copepods, cladocerans and low biomass of calanoid copepods; Gannon and Stemberger 1978). Historical patterns of lower trophic levels in Lake Erie are generally consistent with expected responses to eutrophication. For example, the elimination and subsequent recovery of sensitive

Page 75: Taking Action On Lake Erie - International Joint Commission · reduce external inputs of total phosphorus (TP) and in particular, the rising proportion of the most bioavailable form

62

burrowing mayflies (H. limbata and H. rigida) during the 1950s and 1980s, respectively, tracked changes in water quality conditions. In contrast, several tolerant benthic invertebrate taxa declined with decreased nutrient loads from the 1970s to 2000s (Soster et al. 2011). Similar to benthic invertebrates, a variety of specific zooplankton taxa seem to have responded to changes in trophic conditions (Johansson et al. 1999, 2000). At a coarser taxonomic level, changes in zooplankton assemblages from the 1970s through the 1990s were consistent with a response to reduced eutrophic conditions: i.e., during this time period a zooplankton index of trophic status (ratio of calanoids to cyclopoids + cladocerans) increased in all three Lake Erie basins (Johansson et al. 1999; Barbiero et al. 2001). In addition to shifts in taxonomic composition, nutrient loading affects system-wide production and lower trophic level biomass in Lake Erie. Such shifts in total biomass may ultimately have a stronger impact on fish stocks than shifts in taxa composition. In fact, despite dramatic temporal changes in zooplankton taxa composition, zooplankton size spectrum distributions in Lake Erie changed minimally from 1991 to 1997 (Sprules 2008); demonstrating that similar-sized prey were available notwithstanding taxonomic changes. Following reductions in nutrient loading, phytoplankton biomass in western Lake Erie declined by ~65%, and crustacean zooplankton biomass declined by as much as 60% (Makarewicz 1993a). Such reductions in planktonic production may impact not only planktivorous fishes, but also detritivores, since reduced water column production will directly translate to reduced settlement of detritus to the bottom. In fact, benthic densities of several Lake Erie benthic invertebrates may have declined with reductions in nutrient loading (Soster et al. 2011). Moreover, given that many eutrophic tolerant fish species are facultative detritivores, changes in abundance of such species may be linked to reduced detrital mass (availability) or its nutritional quality (Ludsin et al. 2001). Such reductions in detrital mass have been linked to reduced species diversity (e.g., species richness) in other systems (Gascon and Leggett 1977, Henderson and Crampton 1997). While high nutrient loads can increase overall system productivity, this can also result in shifts in physicochemical conditions and alter the makeup of primary producers. Dense phytoplankton blooms can lead to hypoxic conditions, decreased water clarity and limited available nitrogen (relative to P). Such factors can contribute to phytoplankton assemblages dominated by cyanobacteria that can tolerate low light and compete under both N and P limitation. Phytoplankton taxa differ not only in size, shape and buoyancy, but also chemical composition and nutritional value for consumers, for example, their fatty acid composition. As consumers cannot synthesize many essential fatty acids and are dependent on primary producers for such macro-molecules (Dalsgaard et al. 2003), shifts in phytoplankton assemblages will also lead to food web-wide shifts in fatty acid patterns. Ultimately, eutrophication-induced shifts in nutritional content of lower trophic levels can have important implications for fish growth, survival and reproductive success by affecting factors such as overwinter survival and offspring quality (e.g., Johnson and Evans 1990; Henderson et al. 1996). 6.7 Future Interactive Effects of Climate Change The impacts of climate change on lower trophic levels of Lake Erie will likely not be straightforward, and future climatic conditions will undoubtedly interact with nutrient loading and invasive species in structuring lower trophic level assemblages. As described above, warmer temperatures, lower lake levels and increased frequency of intense precipitation events all have the potential to enhance phytoplankton blooms, reduce water clarity and exacerbate future hypoxia. These expected responses would favour eutrophic tolerant invertebrate taxa. That is, zooplankton and benthic invertebrate taxa would likely increase in abundance if they can a)

Page 76: Taking Action On Lake Erie - International Joint Commission · reduce external inputs of total phosphorus (TP) and in particular, the rising proportion of the most bioavailable form

63

tolerate relatively warm temperatures, b) effectively consume cyanobacteria, c) feed under low light, and d) utilize a low oxygen zone as a refuge from predation. Moreover, with increased primary production and subsequent detrital deposition, total biomasses of both pelagic and benthic consumers could swell. On their own, expected climate change impacts could lead to invertebrate assemblages trending towards patterns observed during the 1950s and 1960s (i.e., at the height of eutrophication). Some important caveats regarding climate change effects on Lake Erie invertebrate assemblages relate to the mediating effects of land use and invasive species. 1) Aggressive changes in land-use practices have the potential to dramatically reduce nutrient loads. 2) Dreissenid mussels have contributed to system benthification, sequestering much carbon and production in benthic regions. Thereby, future increases in nutrient loads may not readily translate to higher planktonic production, lower water clarities and increased hypoxia. 3) Over the coming decades, additional non-indigenous species will undoubtedly establish in Lake Erie. In fact, several assessments speculate that changing climatic conditions will facilitate increased frequencies of invasion events worldwide. While many taxa have been identified as having a high probability of invading the Laurentian Great Lakes (e.g., Ricciardi and Rasmussen 1998), the identity of future invaders and their impacts on the ecosystem are unknown. Nonetheless, several recent assessments have focused on potential invasion and impacts of Asian carp (silver carp Hypophthalmichthys molitrix and bighead carp H. nobilis) on Great Lakes ecosystems; raising concerns that this large-bodied voracious planktivore could negatively affect the availability of lower trophic level prey for native fishes (Cooke and Hill 2010; Hansen 2010). While all dire predicted future effects of Asian carp may not materialize, as evidenced by dreissenid mussels, invasive species have the potential to dramatically alter ecosystem processes and one or more future invasive species could dramatically affect how climatic changes mediate eutrophic conditions in Lake Erie. 6.8 Conclusions Projections of warmer atmospheric temperatures and more intense precipitation events should lead to warmer water temperatures, a longer stratified period, and greater sediment and nutrient loading. In turn, these changes are expected to favor a more eutrophic state, with enhanced hypoxia, lower water clarity, frequent nuisance algal blooms and a less tolerant biotic community. Thus, we expect that in the absence of other changes, future climatic conditions and anthropogenic nutrient loading will promote a Lake Erie fish community resembling the community present during the height of eutrophication. That is, we expect cold-water species and species sensitive to low oxygen and reduced water clarity (e.g., salmonines, coregonines) to decline in abundance, while warm-water and tolerant species will expand. In general, the interactive effects of climate change and nutrient loading are expected to promote a eutrophic-tolerant fish community; with visual feeding, cold-water and hypoxia-sensitive fish declining, and warm-water, tolerant detritivores increasing (i.e., trending towards the right in Figure 1). However, beyond such generalizations, species-specific, population-level responses may be more nuanced and difficult to predict. By influencing population vital rates (survival, growth rate, and reproduction), mechanisms such as hypoxia, reduced water clarity, harmful algal blooms and altered prey base have the potential to directly and indirectly mediate population trajectories. For example, while reduced water clarity may compromise foraging success, and hence growth rates, of visual planktivores, certain warm-water planktivores (e.g., emerald shiners) may actually benefit indirectly from bottom hypoxia forcing their zooplankton prey up into the meta- and epilimnion. Similarly, population-level reproductive success

Page 77: Taking Action On Lake Erie - International Joint Commission · reduce external inputs of total phosphorus (TP) and in particular, the rising proportion of the most bioavailable form

64

(recruitment) of several Great Lakes fish populations appears to respond positively to increased spring-summer annual temperatures and high river discharge; suggesting that warmer conditions and high spring precipitation will enhance recruitment. However, ovarian development for many fish species occurs over winter, and relatively short, warm winters may not allow for adequate gonadal development, thereby compromising recruitment potential. Consideration of such counteracting processes may be necessary to predict future population-specific responses.

7. Lake Erie Coastal Wetlands and Implications for Nutrient Loadings 7.1 Overview Coastal wetlands are important features of the ecosystem and can affect nutrient cycling in Lake Erie. In addition to their significance as habits for fish and wildlife, in stabilizing water supplies (both during floods and drought), wetlands affect water quality as downstream recipients of water from and natural sources and reflect effects of human activities. Lake Erie coastal wetland can be categorized into three groups: lacustrine, riverine and barrier-protected, based on the dominant water source, geomorphic position and hydraulic connectivity to the lake. Each of these types can be further divided based on geomorphic type, with examples in each case being open lacustrine, drowned river-mouth and barrier beach lagoon. Lacustrine systems have a direct connection with Lake Erie waters, and are affected by water level fluctuations, seiches (wind- or storm-induced short-term water level fluctuations), currents and ice scour. Riverine wetlands occur along rivers flowing into the Lake Erie. Though flow rate, water quality and sediment inputs are controlled largely by the drainage, levels and fluvial processes and can also be affected by downstream coastal process (e.g., via seiche events). Barrier-protected wetlands (resulting from either riverine or coastal processes) include a barrier that has separated the nearshore and onshore processes from the lake. These wetlands are protected from wave action, but may still have openings (sometimes ephemeral) in the barrier allowing water movement, and thus affecting water levels and other aspects of the wetland. 7.2 Nutrient Cycling in Freshwater Wetlands Wetlands serve as sources, sinks or transformers of chemicals (including nutrients), depending on wetland type, hydrologic conditions, and extent of chemical loading. Wetlands can become saturated with nutrients, in particular if loading rate is high. Characteristic seasonal patterns of nutrient uptake/release are often observed (e.g. increased retention during the growing season). Wetlands need to be considered in the context of adjacent ecosystems, receiving nutrient inputs from upstream and producing benefits downstream. Wetlands can have a range of levels of productivity. Nutrient cycling in wetlands differs from other aquatic and terrestrial systems in that an increased fraction of nutrients in wetlands are associated with sediments and peat. However as in other ecosystems, wetlands have been affected significantly by human activities, and they have limited capacities to assimilate wastes and excessive nutrients. 7.3 The Importance of Wetlands to Lake Erie Nutrient Loading Restoration of coastal and other wetlands has significant potential to help reduce nutrient loadings to Lake Erie. Factors that need to be considered in assessing the reduction potential of a wetlands complex includes its geomorphic setting/location in the landscape, hydrology (including seasonal and annual changes), the nature of connection to riverine or lacustrine

Page 78: Taking Action On Lake Erie - International Joint Commission · reduce external inputs of total phosphorus (TP) and in particular, the rising proportion of the most bioavailable form

65

water bodies, the hydraulic residence time in the wetland, and nutrient loading magnitude and patterns of phosphorus. Several issues should be considered in wetland restoration as a tool to reduce nutrient loadings. These include the potential for initial increased nutrient export early in the restoration phase, nutrient saturation of the wetland, the potential for climate change to alter individual wetland function, the need to situate nutrient reduction amongst the broader suite of wetland services and identification of values for which management efforts may be directed. Further research is needed in this area, including on identifying priority locations within the Lake Erie watershed where wetland restoration for nutrient reduction goals might be targeted, further exploration through fields studies of factors leading to higher nutrient reduction from coastal wetland, identification of potential approaches to avoid or address phosphorus saturation in wetlands, potential for climate change to complicate nutrient reduction efforts via wetlands and how wetland restoration efforts in the basin can meet multiple objectives, including nutrient reduction.

8. Modelling Lake Parameters in Relation to Nutrient Loading in Lake Erie 8.1 Past and Current Modelling Efforts: A critical review Environmental modelling has been an indispensable tool for addressing lake eutrophication. A variety of data-oriented and process-based models have been used to examine the impact of nutrient loads to ecosystem integrity and to set water quality goals. Data oriented (or empirical) models are mainly steady-state, mass-balance approaches that predict lake TP as a function of lake morphometric/hydraulic characteristics, such as the areal phosphorus loading rate, mean depth, fractional phosphorus retention, and areal hydraulic loading, which are then associated with the chla and/or hypolimnetic DO concentrations (Brett and Benjamin, 2008; Cheng et al., 2009). Despite their successful application in predicting average whole-lake TP concentrations and DO levels in smaller inland lakes on the Canadian Shield (Dillon and Molot, 1996), these models have significant drawbacks when applied to larger systems: i) their fundamental assumption that the lake is thoroughly mixed with uniform concentrations throughout is profoundly violated in systems with larger size and complex shape; ii) their capacity in predicting the impact of episodic events (storm events, upwelling, climate change, and invasive species) is very limited. An alternative to these empirical strategies are process-based models, which can be used to understand ecological processes, to predict aquatic ecosystem responses to external nutrient loading changes, to evaluate management alternatives, and to support the policy making process (Reckhow and Chapra, 1999). This modelling framework uses a general set of equations to describe key physical, chemical, and biological processes with site-specific parameters, initial conditions, and forcing functions which are then used to reproduce real-world dynamics, to gain insights into the ecosystem functioning, and to project future system response under significantly different external conditions (e.g., nutrient enrichment, climate change). In Lake Erie, a variety of statistical (data-driven) and mathematical (process-based) models have been developed to understand ecological interactions, to gain insights into the role of specific facets of the ecosystem functioning (internal loading, dreissenids), and to predict the response of the lake to external nutrient loading reductions. As previously mentioned, the basic premise for the use of mechanistic models is to mimic the role of individual processes through

Page 79: Taking Action On Lake Erie - International Joint Commission · reduce external inputs of total phosphorus (TP) and in particular, the rising proportion of the most bioavailable form

66

proper mathematical description and realistic parameterization, which can then collectively offer a faithful depiction of ecosystem dynamics. In this section, we undertook a critical evaluation of these models. Important findings of this review are summarised below: Overall, there is considerable inconsistency in the parameterization of the existing modelling constructs, e.g., the same biological growth rates and nutrient uptake kinetics differ by one or two orders of magnitude among the different models. (i) Different mathematical expressions have been used to simulate processes that are

critical for projecting lake dynamics under alternative management conditions, i.e., the algal growth rate has been modelled whether as a single function of the ambient nutrient concentrations, or as a two-step process that first considers the nutrient uptake rate in relation to the ambient supply and subsequently the growth rate as a function of their intracellular storage capacity.

(ii) Because the differences in the structure and parameterization of competing models can significantly alter the predicted outcome, the selection of mathematical equations and parameter values should not be based on subjective (and often arbitrary) criteria, but must be ecologically defensible and tightly linked to our contemporary understanding of the system. This is particularly important in Lake Erie, given that nearly all the existing mechanistic modelling work involves very complex models (CE-QUAL-W2, ELCOM-CAEDYM), in which there is an inconceivably high number of calibration vectors that can reproduce the observed patterns equally well. [The same problem holds true for the SWAT model applications.]

(iii) Of equal importance is the fact that none of the existing models in LE has presented any type of uncertainty analysis associated with these overparameterized modelling efforts. Model uncertainty analysis is an excellent means to obtain rigorous assessment of the expected consequences of different management actions, to optimize the sampling design of monitoring programs, to align with the policy practice of adaptive management, and to express model outputs as probability distributions. As noted earlier, aside from the data used for their parameterization, uncertainty can result from the model structure and input error. Model structure error is mainly associated with the selection of the appropriate state variables for reproducing ecosystem behavior.

(iv) The selection of the suitable equations from a variety of mathematical possibilities to represent key ecological processes; and the fact that our models are based on relationships derived individually in controlled laboratory environments may not collectively yield an accurate picture of the real world dynamics.

(v) Because of the still poorly understood ecology, we do not have robust group-specific parameterizations that can support predictions in a wide array of spatiotemporal domain. There is little doubt that the capacity of the current generation of eutrophication models has not been proven yet, and therefore sceptical views in the modelling literature characterize all the efforts to simulate the dynamics of individual species or genera as attempts to "run before we can walk" (Anderson, 2005). For example, preliminary efforts to incorporate phytoplankton functional types into biogeochemical models were based on speculative parameterization and –not surprisingly- resulted in unreliable predictions.

Page 80: Taking Action On Lake Erie - International Joint Commission · reduce external inputs of total phosphorus (TP) and in particular, the rising proportion of the most bioavailable form

67

(vi) The effect of dreissenids on water quality is another critical aspect of the ecosystem functioning that requires robust modelling, yet existing dreissenid submodels lack the ability to predict changes in spatiotemporal densities or their distributions among different age groups

(vii) Sediment diagenetic modelling along with the associated data (e.g., pore water analysis, phosphorus fractionation, organic matter profiles) is still missing from Lake Erie (Dittrich et al., 2013). The sediments represent another important controlling factor of the water quality in Erie, acting as a source (or sink) for a wide variety of chemicals. Close to the sediment-water interface, intensive microbiological, geochemical and physical processes determine the fraction of organic matter, nutrients, and pollutants released into the overlying water. Detailed knowledge of the processes occurring in the top few cm of the sediment is essential for the assessment of water quality, the understanding of the manifestation of hypoxia, and the management of surface waters. In cases where measurements are impossible or expensive, diagenetic modelling is an indispensable tool to investigate the interplay among the sediment processes, to verify concepts, and to potentially predict system behaviours (Boudreau, 1997). The goal of such a modelling exercise is to improve our quantitative understanding of the mechanisms of phosphorus mobilization in the sediments and to identify process controls under a variety of conditions. The knowledge obtained will allow addressing research questions, such as: Can P retention in lake sediments be predicted based on the sediment mineralogy, sedimentation substance inputs, catchment type, and other characteristics? How sediment retention capacity with respect to P may respond to changes caused by human activities and/or climate change?

(viii) On a final note, we believe that the presence of various models with different strengths and weaknesses offers a unique opportunity for synthesis and improvement of the contemporary modelling practice. In the context of ecological process-based modelling, this approach should not be viewed solely as a framework to improve our predictive devices, but rather as an opportunity to compare alternative ecological structures, to challenge existing ecosystem conceptualizations, and to integrate across different (and often conflicting) paradigms.

Despite their simplicity, statistical based (empirical) models offer straightforward cause-effect relationships coupled with uncertainty estimates (e.g., response curves). Since they are founded on the available data from the system, they offer a less risky choice to move forward and indeed offer a pragmatic means to obtain insight about the response of the system. Nevertheless we have major reservations about their capacity to guide predictions outside the range associated with the dataset used. As an alternative, existing mechanistic models have significant heuristic value and potential to be used for predictive purposes; once rigorously evaluated. To this end, the presence of various models in Lake Erie with different strengths and weaknesses offers a unique opportunity for synthesis and improvement of the contemporary modelling practice (ensemble modelling). This approach should not be viewed solely as a framework to improve our predictive devices, but rather as an opportunity to compare alternative ecological structures, to challenge existing ecosystem conceptualizations, and to integrate across different (and often conflicting) paradigms. Lake Erie can be an excellent case study, where a wide range of mathematical and statistical models can be integrated to guide the decision making process.

Page 81: Taking Action On Lake Erie - International Joint Commission · reduce external inputs of total phosphorus (TP) and in particular, the rising proportion of the most bioavailable form

68

8.2 Response Curves 8.2.1 Predictors for cHABs in Western Lake Erie Stumpf et al. (2012) examined the role of phosphorus loading on the severity of the summer cyanobacterial bloom in western Lake Erie. Bloom severity was quantified from satellite with a “cyanobacteria index” (CI). CI and area are linearly related, with a CI of 1.0 being approximately equivalent to 300 km2 of bloom > 106 cells ml-1. The authors found that loads and discharge from the Maumee River for spring (March -June) explained the variation in CI between years from 2002 to 2011 (Figure 12). While the blooms peaked in late summer, loads in July and August had no bearing on the blooms. The result constrains the loading problem to spring, as years with large loads in July or in the previous winter showed no influence of those loading events. As a model, the most effective predictor was the discharge (Q). In years without extreme variations in June loads (not shown), June total phosphorus (TP) load could be a good predictor. However June TP load was not a good primary predictor, because of the inability to account for the load during the rest of the spring. TP and soluble reactive phosphorus (SRP) load, while showing good correlations, had poor skill in predicting the annual bloom severity. For comparison, the Q model had an error of 0.58 CI, while TP had an error of 1.75 CI, and SRP had an error of 3.1 CI, nearly equal to the magnitude of the blooms.

Figure 12: (left) Bloom intensity from 2002 to 2011 compared to TP spring load (m.tons). CI of 1.0 is nominally

equivalent to 300 km2 of bloom > 10

6 cells ml

-1 . (right) Spring discharge (m

3 s

-1) from the Maumee used for the Q

model. Data presented in Stumpf et al. (2012).

The importance of Q may reflect both the P load and dispersion of nutrients across the western basin. However, from the perspective of developing strategies for reducing the blooms, Q alone does not allow for assessing strategies in controlling nutrient loads. Therefore, the data set was re-examined to see if a model including phosphorus as a primary term could produce a model with skill. One model was:

Page 82: Taking Action On Lake Erie - International Joint Commission · reduce external inputs of total phosphorus (TP) and in particular, the rising proportion of the most bioavailable form

69

log10(CI) = -0.165 + 0.102*(TPq), with TPq = 5.126 + 0.0141*TP_marjun - 4.796*(TP_marjun/Q_marjun) Where TP_marjun is the March-June total load in m.tons, and Q_marjun is the average discharge over the four months in m3 s-1.

The root mean square error (RMS) for 2002 to 2010 is 0.56 CI, equivalent to the Q based model (0.58 CI). The observed and modeled CI values from the Q model and the TPq model for 2002-2010 are shown in Figure 13. The relative failure in the TPq model is that it severely overestimates 2011, giving a prediction of 37 CI vs. the observed 14.2 CI, whereas the Q model predicts CI of 14.8. Given the extreme loading in 2011, there is a limit to the TP concentration under extreme discharge. As an alternative, we explored an even simpler model relating CI only to the March-June TP load (Figure 14): CI = 0.260e0.0018 R2 = 0.83

Figure 13: A comparison of the TPq model (black squares) presented here against the Q model (open circles)

from Stumpf et al., (2012). Modeled CI (x-axis) vs observed CI (y-axis). Line shows 1:1 match. The far upper

right is the 2011 value for the Q model. The TPq model overestimates 2011 (modeled CI = 37, not on graph).

Figure 14: Comparison of CI observations to the TP load model.

Page 83: Taking Action On Lake Erie - International Joint Commission · reduce external inputs of total phosphorus (TP) and in particular, the rising proportion of the most bioavailable form

70

To test this model, given the strength of impact of the 2011 data, we constructed a regression with only 2002-2010 data for comparison (Figure 17). The resulting model is similar to the original one, but with a lower R2 of 0.68, and greater uncertainty (0.7 vs 0.37 for the Stumpf et al. (2012) model). However, this model accurately predicted the 2011 bloom. All models tested to date underestimate the 2012 bloom. Also, included on the graph are predictions for 1991-2001 when CI estimates were not available, highlighting the two years from that set (1995, 1998) with qualitatively reported blooms. Because prediction errors in these exponential models become quite large at higher loads and uncertainty around estimates in some of the bloom years, we are not comfortable with specific numerical predictions or annual forecasts using this model. Instead, it represents a scenario approach for exploring the relationship between loads and bloom categories. We suggest the following categorization and associated March-June TP load targets for the Maumee River: Table 13: Bloom categories and associated Maumee spring TP loading levels

CI Condition Maumee TP March-June Load (MT)

<1 no or mild bloom (2002, 2005, 2006, 2007) < 750 1-2.4 moderate bloom (2003, 2004, 2012) 750-1250 2.4-6 severe bloom (2008, 2009, 2010) 1250-1750 >6 extreme bloom (2011) >1750

The March-June TP Maumee River load varied between 387 - 2240 kg between 1991 and 2012, with a mean and standard deviation of 984 +/- 436 kg. Thus, from this analysis, the mean load would have to be reduced by 24% to avoid blooms. According to Bosch et al (in review), reducing Maumee loads by 24% will take considerably stronger BMP implementation than is current in place (Figure 16).

Figure 15: CI estimates (diamonds), regression model with 2002-2011 observations (black line), regression

model with 2002-2010 observations (red line), 2012 (triangle) and 2011 (box) predictions with the 2002-2010

regression, 1995-2001 hindcasts (dot), 1995 and 1998 qualitative bloom observations (circle), 2011 CI

estimate (orange triangle). Boundaries between bloom categories are represented as horizontal lines.

Page 84: Taking Action On Lake Erie - International Joint Commission · reduce external inputs of total phosphorus (TP) and in particular, the rising proportion of the most bioavailable form

71

Figure 16: from Bosch et al (in review) showing the required level of BMP implementation to achieve

increasing levels of load reduction from the Maumee River watershed.

8.2.2 Response Curves Expressing the Relationship between TP/SRP loads and Central Basin Hypoxia and Chlorophyll Concentrations A model, calibrated to observations in the Central Basin of LE, was used to develop response curves relating hypoxia and chlorophyll concentrations to phosphorus loads. The model is driven by a 1D hydrodynamic model that provides temperature and vertical mixing profiles (Rucinski et al. 2010). The biological portion of the coupled hydrodynamic-biological model incorporates P and C loading, internal phosphorus cycling, C cycling (in the form of algal biomass and detritus), algal growth and decay, zooplankton grazing, oxygen consumption and production processes, and sediment interactions (Figure 17).

Figure 17: Conceptual diagram of simple eutrophication model.

Dissolved

Oxygen

Available

Phosphorus

Basin

LoadingPhytoplankton

Detrital

Carbon

Sediments

Atmosphere

Unavailable

Phosphorus

Zooplankton uptake

respiration photosynthesis

reaeration

grazing

death

mineralization

SODSettling

Settling Settling

Death/grazing

Page 85: Taking Action On Lake Erie - International Joint Commission · reduce external inputs of total phosphorus (TP) and in particular, the rising proportion of the most bioavailable form

72

While this is a standard, relatively simple eutrophication model, a critical new feature was added to accommodate simulating scenario-based loading regimes. We adjusted the sediment oxygen demand (SOD) term in the model with a relationship developed by Borsuk et al. (2001) that relates SOD to deposited organic carbon. Deposited organic carbon in the model is a function of nutrient loading. Therefore, the SOD term can be related to phosphorus loading as:

, with the values of the maximum SOD (SODmax) and the half saturation term (KSOD) estimated via regression on the observations in Borsuk et al (2001). The results outlined in this report are being developed more fully in Rucinski et al (in prep). 8.2.2.1 Model Calibration and Testing The model was calibrated for several properties for 1987-2005 for the central basin and results for epilimnetic chlorophyll and TP concentrations are shown in Figure 18, and epilimnetic, metalimnetic, and hypolimnetic DO concentrations are shown in Figure 19.

Figure 18: Epilimnetic Chlorophyll (left) and TP (right) calibration results

Page 86: Taking Action On Lake Erie - International Joint Commission · reduce external inputs of total phosphorus (TP) and in particular, the rising proportion of the most bioavailable form

73

Figure 19: Dissolved oxygen calibration results for epilimnion, metalimnion and hypolimnion in Lake Erie Central

basin

The slopes of the DO concentrations represent the hypolimnetic DO depletion rate for a given year (Figure 20), a useful parameter for this sparsely sampled system. In an earlier analysis, Rucinski et al (2010) compared depletion rates determined directly from observations (black diamonds) with those required to match observations once vertical mixing was accounted for in a simple 1D model (blue diamonds and solid line). The depletion rates estimated from this eutrophication model (red boxes) are similar in magnitude and generally follow the same temporal patterns.

Page 87: Taking Action On Lake Erie - International Joint Commission · reduce external inputs of total phosphorus (TP) and in particular, the rising proportion of the most bioavailable form

74

While the model reproduces the oxygen profiles well, this 1-dimensional representation of the vertical profile in the central basin is not capable of deriving a key management property – hypoxic area. However, a recent geostatistical analysis (Zhou et al. 2012) provides both quantitative estimates and associated uncertainties of hypoxia area, as well as a relationship between hypoxic area and bottom-water dissolved oxygen concentration (Figure 21). Using this latter relationship, we convert modeled bottom water concentration estimates to area to both compare our area estimates with those of Zhou et al (2012) (Figure 21) and to general hypoxic area response curves.

We provide several measures of hypoxic area in the comparison, including both individual cruise statistics and monthly and summer mean statistics of both empirical and model results because no single observation measure is adequate to capture the full summer dynamics. It is of note that while, in most years, the model statistics capture the interannual variability in the observation estimates, there is a tendency for the model to generate larger areas in the early 2000s even though it captures the hypolimnetic seasonal depletion accurately (Figure 18).

Figure 20: Water-column depletion rates compared to earlier estimates from a 1D physical model and

observations (Rucinski et al. 2010)

Figure 21. Comparison of hypoxic area and average hypolimnetic DO concentration from Zhou et al.

(2012)

Page 88: Taking Action On Lake Erie - International Joint Commission · reduce external inputs of total phosphorus (TP) and in particular, the rising proportion of the most bioavailable form

75

Figure 22: Light symbols with dotted error bars are 95% confidence intervals from individual cruises (Zhou et al.

2012). Red symbols and error bars are monthly means and std of those estimates. Dark black line and bars represent model Aug-Sept means and std., and the shaded region represents model output encompassing August and September means.

The combined comparisons of the seasonal concentration dynamics and the secondary approximations to hypoxic area (23) support use of the model for developing associated response curves. 8.2.2.2 Response Curves With this calibrated model, we explored response curves for bottom water DO concentrations (Figure 23A), hypoxic-days (Figure 23B i.e. number of days per year with hypolimnetic DO <2 mg/l), hypolimnetic depletion rates (Figure 24A), and hypoxic area (Figure 24B) and as a function of annual TP load, March-June total phosphorus load, annual and March-June total phosphorus loads from the Maumee River, and the counterpart SRP loads. The total loads used in the response curves are the sum of the observed loads to the Western and Central basin.

Figure 23: Modelled response curves (black lines) as a function of annual Lake Erie TP load (MTA) for (left) bottom

dissolved oxygen (DO, mg/l) and (right) hypoxic area (in 1,000 km2). For both figures, calibration set includes data

from years in which the model was calibrated (1987-2005). Verification set includes additional data from years outside of the calibration period.

0

1

2

3

4

5

6

7

8

9

0 5000 10000 15000 20000 25000

Bo

tto

m L

aye

r D

O (m

g∙L-1

)

Total Annual TP Load

Model Average Verification Set Calibration Set

0

2

4

6

8

10

12

0 5000 10000 15000 20000 25000

Hyp

oxi

Are

a (k

m2 )

Total Annual TP Load

Calibration Set Model Average Verification Set

Page 89: Taking Action On Lake Erie - International Joint Commission · reduce external inputs of total phosphorus (TP) and in particular, the rising proportion of the most bioavailable form

76

Figure 24: Modelled response curves (black lines) as a function of annual Lake Erie TP load (MTA) for (left)

dissolved oxygen depletion rate (DO, mg/l/d) and (right) hypoxic days. Regression rates correspond to simple linear regression of the DO observations.

0.00

0.02

0.04

0.06

0.08

0.10

0.12

0 5000 10000 15000

De

ple

tio

n R

ate

(mg∙

L-1∙d

-1)

Total Annual TP Load

Regression Rates Model Average

0.0

10.0

20.0

30.0

40.0

50.0

60.0

70.0

0 5000 10000 15000

Hyp

oxi

c D

ays

Total Annual TP Load

Model Average

Implications for New Loading Targets:

Once new policies are set for acceptable levels of hypoxia, one can use Figure 27 to frame new target

loads. For example, if the desired outcome is for average hypoxic area to be below 2,000 km2 the

annual TP target load would be approximately 4200 MT/year, a 46% reduction from the 2003-2011

average loads and 54% below the current target.

It is also important to note that the potential target levels implied by the response curves developed for

Western Basin cHABs would not be sufficient to significantly reduce hypoxic area. For example, it was

suggested above and by the Ohio Lake Erie Phosphorus Task Force that the March-June Maumee River

target loads to eliminate blooms should be less than 800 MT, a 31% reduction from the 2005-2011

average of 1160 MT (Richards, pers. com). If all Central and Western Basin non-point sources were

reduced by the same 31% and applied across the full year, the resulting annual load would be reduced

to 6073 MT, considerably higher than the 4200 MT target identified above. So, in setting targets, it is

important to recognize HAB and hypoxia endpoints likely require separate considerations.

Page 90: Taking Action On Lake Erie - International Joint Commission · reduce external inputs of total phosphorus (TP) and in particular, the rising proportion of the most bioavailable form

References Aku, P. M. K., L. G. Rudstam, and W. M. Tonn. 1997. Impact of hypolimnetic oxygenation on vertical

distribution of cisco in Amisk Lake, AB. Can J of Fish Aqua Sci. 54:2182-2195. Alfera L.K., Weismiller R.A.. 2002. U.S. Experiences With Nutrient Management and Good Management

Practices to Control Non-Point Pollution from Agriculture. Washington DC: ESSD Unit, The World bank.

Ahn, H. & James, R.T. 2001. Variability, uncertainty, and sensitivity of phosphorus deposition load estimates in South Florida. Water, Air, & Soil Pollution 126: 37–51.

Anderson TR. 2005. Plankton functional type modelling: running before we can walk? J Plankton Res. 27:1073-1081.

Anderson, K.A. & Downing, J.A. 2006. Dry and wet atmospheric deposition of nitrogen, phosphorus and silicon in an agricultural region. Water Air Soil Pollut, 176: 351–374.

Andraski, B. J., Mueller, D. H., and Daniel, T. C. 1985. Phosphorus losses in runoff as affected by tillage. Soil Sci Soc Am J, 49(6), 1523-1527.

Auer, M.T., Canale, R.P. 1982. Ecological studies and mathematical modeling of Cladophora in Lake Huron : 2. Phosphorus uptake kinetics. J. Gt Lakes Res.. 84-92.

Austin, J.A. and Coleman, S.M. 2007. Lake Superior summer water temperatures are increasing more rapidly than regional air temperatures: A positive ice-albedo feedback. Geophys.l Res. Lett.. 34: L06604, 5pp. doi: 10.1029/2006GL029021.

Azcue, J.M., Rosa, F., and Mudroch, A. 1996. Distribution of major and trace elements in sediments and pore water of Lake Erie. J. Great Lakes Res. 22(2):389-402

Azevedo SM, Carmichael WW, Jochimsen EM, Rinehart KL, Lau S, Shaw GR, Eaglesham GK. 2002. Human intoxication by microcystins during renal dialysis treatment in Caruaru—Brazil. Toxicol. 181–182: 441-446.

Baker, D. B. 2010. Trends in Bioavailable Phosphorus Loading to Lake Erie. Lake Erie Protection Fund Grant 315-07 Final Report. Ohio Lake Erie Commission. http://lakeerie.ohio.gov/LinkClick.aspx?fileticket=OxJaciMDMoQ%3d&tabid=61."

Barbiero, RP; Little, RE; Tuchman, ML. 2001. Results from the US EPA's biological open water surveillance program of the Laurentian Great Lakes: III. Crustacean zooplankton. J Great Lakes Res. 27:167-184.

Barbiero, R., Tuchman, M.L. 2004. Long-term dreissenid impacts on water clarity in Lake Erie. J Great Lakes Res. 30: 557 -565.

Barbiero, R.P., Tuchman, M.L., Millard, E.S. 2006. Post-dreissenid Increases in transparency during summer stratification in the offshore waters of Lake Ontario: Is a reduction in whiting events the cause? . J Great Lakes Res. 32: 131 – 141.

Barton, D.R., Johnson, R.A., Campbell, L.M., Petruniak, J., Patterson, M. 2005. Effects of round gobies (Neogobius melanostomus) on dreissenid mussels and other invertebrates in Eastern Lake Erie, 2002 – 2004. J Great Lakes Res. 31: 252 – 261.

Batroney, T., Wadzuk, B. M., and Traver, R. G. 2010. Parking Deck's First Flush. Journal of Hydrologic Engineering, 15(2), 123-128.

Becker, R. H., M. I. Sultan, G. L. Boyer, M. R. Twiss, and E. Konopko. 2009. Mapping Cyanobacterial Blooms in the Great Lakes Using MODIS. J Great Lakes Res 35: 447-453.

Beeton, A. M. 1963. Limnological survey of Lake Erie in 1959 and 1960. Great Lakes Fishery Commission Technical Report 6:1-32.

Bekele, A., McFarland, A. M. S., and Whisenant, A. J. 2006. Impacts of a manure composting program on stream water quality. Trans. ASABE, 49(2), 389-400.

Bennington, V., McKinley, G.A., Kimura, N., Wu, C.H. 2010. General circulation of Lake Superior: mean, variability, and trends from 1979 – 2006. Journal of Geophysical Research. 115: C12015. 14pp.

Berndtsson, J. C. 2010. Green roof performance towards management of runoff water quantity and quality: A review. Ecol. Eng., 36(4), 351-360.

Beven, K.J. 2006: A manifesto for the equifinality thesis. J. Hydrol. 320, p.18–36 Beven, K.J. 2007: Towards integrated environmental models of everywhere: uncertainty, data and

modelling as a learning process. Hydrology and Earth System Sciences, 11(1), p. 460–467.

Page 91: Taking Action On Lake Erie - International Joint Commission · reduce external inputs of total phosphorus (TP) and in particular, the rising proportion of the most bioavailable form

Binding, C. E., T. A. Greenberg, and R. P. Bukata. 2012. An analysis of MODIS-derived algal and mineral turbidity in Lake Erie. J Great Lakes Res 38: 107-116.

Binding, C.E., Jerome, J.H., Bukata, R.P., Booty, W.G. 2007. Trends in water clarity in Lake Erie. J Great Lakes Res. 33: 828 – 841.

Bishop CT, Anet EFLJ, Gorham PR. 1959. Isolation and identification of the fast-death factor in Microcystis aeruginosa NRC-1. Canadian Jour Biochem and Physiol 37: 453–471.

Bishop, P. L., Hively, W. D., Stedinger, J. R., Rafferty, M. R., Lojpersberger, J. L., and Bloomfield, J. A. 2005. "Multivariate analysis of paired watershed data to evaluate agricultural best management practice effects on stream water phosphorus." J. Environ. Qual., 34(3), 1087-1101.

Boudreau, B.P. 1997. Diagenetic Models and Their Implementation. Modelling Transport and Reactions in Aquatic Sediments. Berlin, Heidelberg, New York, London, Paris, Tokyo, Hong Kong: Springer-Verlag. 414 pp.

Bosch, I., Makarewicz, J. C., Lewis, T. W., Bonk, E. A., Finiguerra, M., and Groveman, B. 2009. Management of agricultural practices results in declines of filamentous algae in the lake littoral. J Great Lakes Res. 35, 90-98.

Brandt, S.B., M. Costantini, S.E. Kolesar, S.A. Ludsin, D.M. Mason, C.M. Rae, and H. Zhang. 2011. Does hypoxia reduce habitat quality for Lake Erie walleye (Sander vitreus)? A bioenergetics perspective. Can J Fish Aquat Sci. 68:857-879.

Brett, M.T., and Benjamin, M.M. 2008. A review and reassessment of lake phosphorus retention and the nutrient loading concept. Freshw. Biol. 53:194–211.

Bridgeman, T.B., Chaffin, J.D., Kane, D.D., Conroy, J.D., Panek, S.E., Armenio, P.A. 2012. From river to lake: Phosphorus partitioning and algal community compositional changes in western Lake Erie. J Great Lakes Res. 38: 90 – 97.

Brittain, S., Z. A. Mohamed, J. Wang, V. K. Lehmann, W. W. Carmichael, and K. L. Rinehart. 2000. Isolation and characterization of microcystins from a river nile strain of Oscillatoria tenuis Agardh ex Gomont. Toxicon 38: 1759-1771.

Brown, L.J., Taleban, V., Gharabaghi, B. & Weiss, L. 2011. Seasonal and spatial distribution patterns of atmospheric phosphorus deposition to Lake Simcoe, ON. J. Gt Lakes Res. 37: 15–25.

Bunnell, D.B., Johnson, T.B., Knight, C.T. 2005. The impact of introduced round gobies (Neogobius melanostomus) on phosphorus cycling in central Lake Erie. Can J Fish Aquat Sci. 62: 15 – 29.

Burkholder, J.M., Gilber, P.M., Skelton, H.M. 2008. Mixotrophy, a major mode of nutrition for harmful algal species in eutrophic waters. Harmful Algae. 8: 77 – 93.

Burnett, L.E., and W.B. Stickle. 2001. Physiological responses to hypoxia. Pages 101-114 in N.N. Rabalais and R.E. Turner, editors. Coastal hypoxia: consequences for living resources and ecosystems.

Burniston, D., McCrea, R., Klawunn, P., Ellison, R., Thompson, A., and Bruxer, J. 2009. Detroit River Phosphorus Loading Determination. Environment Canada, Burlington, Ontario.

Burns, N.M. 1976. Temperature, oxygen, and nutrient distribution patterns in Lake Erie, 1970. J. Fish. Re. Board. Can. 33: 485-511.

Burns, N. M., D. C. Rockwell, P. E. Bertram, D. M. Dolan, and J. J. H. Ciborowski. 2005. Trends in temperature, Secchi depth, and dissolved oxygen depletion rates in the central basin of Lake Erie, 1983-2002 J Great Lakes Res. 31:35-49.

Burton, G. A. J., and Pitt, R. 2001. Stormwater Effects Handbook: A Tool Box for Watershed Managers, Scientists, and Engineers, CRC Press, Boca Raton, FL.

Canale, R.P., Auer, M.T. 1982a. Ecological studies and mathematical modeling of Cladophora in Lake Huron: 7. Model verification and system response. J Great Lakes Res. 8: 134 – 143.

Canale, R.P., Auer, M.T. 1982b. Ecological studies and mathematical modeling of Cladophora in Lake Huron : 5. Model development and calibration. J Great Lakes Res. 8: 112 – 125.

Carpenter, D. D., and Hallam, L. 2010. Influence of Planting Soil Mix Characteristics on Bioretention Cell Design and Performance. J Hydrol Eng, 15(6), 404-416.

Casselman, J. M. 1978. Effects of environmental factors on growth, survival, activity and exploitation of northern pike. American Fisheries Society Special Publication 11: 114-128.

Cestti, R., Srivastava, J., and Jung, S. 2003. Agriculture non-point source pollution control : good management practices--the Chesapeake Bay experience, World Bank, Washington, D.C.

Page 92: Taking Action On Lake Erie - International Joint Commission · reduce external inputs of total phosphorus (TP) and in particular, the rising proportion of the most bioavailable form

Chapra, S.C., Dolan, D. M. 2012. Great lakes total phosphorus revisited: 2. mass balance modeling. J Great Lakes Res. 38(4), 741-754

Charlton, M.N., LeSage, R., Milne, J.E. 1999. Lake Erie in transition : the 1990’s. In: Munawar, M., Edsall, T., Munawar, I.F. {Eds}. State of Lake Erie (SOLE) – Past, Present and Future. Ecovision World Mongraph Series. Backhuys Publishers, the Netherlands. Pp. 97 – 123.

Charlton, M. N., J. E. Milne, W. G. Booth, and F. Chiocchio. 1993. Lake Erie offshore in 1990: restoration and resilience in the central basin. J. Gt Lakes Res. 19:291–309.

Chaubey, I., Chiang, L., Gitau, M. W., and Mohamed, S. 2010. Effectiveness of best management practices in improving water quality in a pasture-dominated watershed. J. Soil Water Conserv, 65(6), 424-437.

Cheng, V., G. Arhonditsis, and M. Brett. 2009. A revaluation of lake-phosphorus loading models using a Bayesian hierarchical framework. Ecol. Res. 25: 59-76.

Conley DJ, Paerl HW, Howarth RW, Boesch DF, Seitzinger SP, Havens KE, Lancelot C, Likens GE. 2009. Ecology. Controlling eutrophication: nitrogen and phosphorus. Science 323: 1014-1015.

Conroy, J.D., W.J. Edwards, R.A. Pontius, D.D. Kane, H. Zhang, J.F. Shea, and D.A. Culver 2005a. Soluble nitrogen and phosphorus excretion of exotic freshwater mussels (Dreissena spp.):potential impacts for nutrient remineralization in western Lake Erie. Freshwater Biol. 50:1146–1162.

Conroy, J.D., Kane, D.D. ,Dolan, D.M., Edwards, W.J., Charlton, M.N., Culver, D.A. 2005b. temporal trends in Lake Erie plankton biomass: Roles of external phosphorus loading and dreissenid mussels J Great Lakes Res. 31: 89 – 110.

Cooke, S.L., and W. R. Hill. 2010. Can filter‐feeding Asian carp invade the Laurentian Great Lakes? A bioenergetic modelling exercise. Freshwater Biology. 55: 2138-2152.

Cowen, W. F., and Lee, G. F. 1973. Leaves as source of Phosphorus. Environ Sci Technol, 7(9), 853-854.

Crumrine, J. P. 2011. A BMP Toolbox for Reducing Dissolved Phosphorus Runoff From Cropland to Lake Erie. NCWQR, Heidelberg University. http://www.heidelberg.edu/academiclife/distinctive/ncwqr/p#bmp.

Daloglu, I., Cho, K. H., and Scavia, D. 2012. Evaluating Causes of Trends in Long-Term Dissolved Reactive Phosphorus Loads to Lake Erie. Environ Sci Technol, 46(19), 10660-10666.

Dalsgaard, J., M. St. John, G. Kattner, D. Muller-Navarra, and W. Hagen. 2003. Fatty acid trophic markers in the pelagic marine environment. Advances in Marine Biology. 46: 225-340.

Davis, T. W., F. Koch, M. A. Marcoval, S. W. Wilhelm, and C. J. Gobler. 2012. Mesozooplankton and microzooplankton grazing during cyanobacterial blooms in the western basin of Lake Erie. Harmful Algae 15: 26-35.

Davis, T. W. M. J. Harke, M. Marcoval, J. Goleski, C. Orano-Dawson, D.L. Berry, C. J. Gobler. 2010. Effects of nitrogenous compounds and phosphorus on the growth of toxic and non-toxic strains of Microcystis during cyanobacterial blooms. Aquat Microb Ecol 61: 149-162.

Dayton, A.I. 2011. Cladophora, mass transport, and the near shore phosphorus shunt. MSc thesis, Michigan Technological Univeristy, Houghton, MI.

De Robertis, A., C. H. Ryer, A. Veloza, and R. D. Brodeur. 2003. Differential effects of turbidity on prey consumption of piscivorous and planktivorous fish. Can J Fish Aquat Sci. 60:1517-1526.

Debruyn, J. M., J. A. Leigh-Bell, R. M. L. Mckay, R. A. Bourbonniere, and S. W. Wilhelm. 2004. Microbial Distributions and the Impact of Phosphorus on Bacterial Activity in Lake Erie. J Great Lakes Res 30: 166-183.

Delorme, L. D. 1982. Lake Erie oxygen: the prehistoric record. Can J Fish Aquat Sci. 39:1021-1029. Depew, D.C., Houben, A.J., Guildford, S.J., Hecky, R.E. 2011. Distribution of nuisance Cladophora in the

lower Great Lakes: Patterns with land use, near shore water quality and dreissenid abundance. J Great Lakes Res. 37: 656 – 671.

Desai, A.R., Austin, J.A., Bennington, V., McKinley, G.A. 2009. Stonger winds over a large lake in response to a weakening air-to-lake temperature gradient. Nature Geoscience. 2: 855 – 858.

Dietz, M. E., Clausen, J. C., and Filchak, K. K. 2004. Education and changes in residential nonpoint source pollution. Environ. Manage. 34(5), 684-690.

Dittrich, M., Chesnyuk, A., Gudimov, A. McCulloch, J., Quazi, S., Young, J., Winter, J., Stainsby, E., and

Page 93: Taking Action On Lake Erie - International Joint Commission · reduce external inputs of total phosphorus (TP) and in particular, the rising proportion of the most bioavailable form

Arhonditsis, G.B. 2013. Phosphorus retention in a mesotrophic lake under transient loading conditions: Insights from a sediment phosphorus binding form study. Wat. Res. 47:1433-1447.

Dobiesz, N.E. and Lester, N.P. 2009. Changes in mid-summer water temperature and clarity across the Great Lakes between 1968 and 2002. J Great Lakes Res. 35: 371 – 384.

Dodds, W.K., Gudder D.A. 1992. The ecology of Cladophora. J Phycol. 35: 42 – 53. Dolan, D.M., McGunagle, K.P. 2005. Lake Erie total phosphorus loading analysis and update: 1996 –

2002. J Great Lakes Res. 31: 11 – 22. Dolislager, L.J., VanCuren, R., Pederson, J.R., Lashgari, A. & McCauley, E. 2012. A summary of the Lake

Tahoe Atmospheric Deposition Study (LTADS). Atmos. Environ 46: 618–630. Dolman A.M., J. Rücker, F.R. Pick, J. Fastner, T. Rohrlack, U. Mischke, C. Wiedner. 2012. Cyanobacteria

and Cyanotoxins: The Influence of Nitrogen versus Phosphorus. PLoS ONE 7: e38757. Dou, Z. X., Knowlton, K. F., Kohn, R. A., Wu, Z. G., Satter, L. D., Zhang, G. Y., Toth, J. D., and Ferguson,

J. D. 2002. Phosphorus characteristics of dairy feces affected by diets. J Environ. Qual., 31(6), 2058-2065.

Dyble, J., G. L. Fahnenstiel, R. W. Litaker, D. F. Millie, and P. A. Tester. 2008. Microcystin concentrations and genetic diversity of Microcystis in the lower Great Lakes. Environ Toxicol 23: 507-516.

Dziallas, C., and H.-P. Grossart. 2012. Microbial interactions with the cyanobacterium Microcystis aeruginosa and their dependence on temperature. Mar Biol 159:2389–2398

Eby, L.A., Crowder, L.B. 2002. Hypoxia-based habitat compression in the Neuse River estuary: context-dependent shifts in behavioral avoidance thresholds. Can J Fish Aquat Sci. 59:952-965.

Elrifi, I. R. and D. H. Turpin. 1985. Steady-state luxury consumption and the concept of optimum nutrient ratios: a study with phosphate and nitrate limited Selenastrum minutum (Chlorophyta). J Phycol. 21:592-602.

EPA. 2009. National Water Quality Inventory: Report to Congress. 2004 Reporting Cycle. EPA 841-R-08-001, United State Environmental Protection Agency, Washington, DC.

Erickson, J. E., Cisar, J. L., Snyder, G. H., and Volin, J. C. 2005. Phosphorus and potassium leaching under contrasting residential landscape models established on a sandy soil. Crop Sci., 45(2), 546-552.

Eisenreich, S.J., Emmling, P.J. & Beeton, A.M. 1977. Atmospheric loading of phosphorus and other chemicals to Lake Michigan. J. Gt Lakes Res. 3: 291–304.

Espinoza, L., Daniels, M., Norman, R., and Slaton, N. 2005. The nitrogen and phosphorus cycle in soils. University Arkansas Cooperative extension service printing services: Agriculture and Natural Resources. Fact Sheet No. FSA2148‐2M‐10‐05N.

Field, K.M. and E.E. Prepas. 1997. Increased abundance and depth distribution of pelagic crustacean zooplankton during hypolimnetic oxygenation in a deep, eutrophic Alberta lake. Can J Fish Aquat Sci. 54:2146-2156.

Fiener, P., and Auerswald, K. 2006. Seasonal variation of grassed waterway effectiveness in reducing runoff and sediment delivery from agricultural watersheds in temperate Europe. Soil Till Res, 87(1), 48-58.

Finlay, K., A. Patoine, D. B. Donald, M. J. Bogard, and P. R. Leavitt. 2010. Experimental evidence that pollution with urea can degrade water quality in phosphorus-rich lakes of the Northern Great Plains. Limnol Oceanogr 55: 1213-1230.

Fissore, C., Hobbie, S. E., King, J. Y., McFadden, J. P., Nelson, K. C., and Baker, L. A. 2012. The residential landscape: fluxes of elements and the role of household decisions. Urban Ecosyst., 15(1), 1-18.

Francis, J.A. and Vavrus, S.J. 2012. Evidence linking Arctic amplification to extreme weather in mid-latitudes. Geophysical Research Letters. 39: L06801. doi: 10.1020/2012GL051000, 2012.

Galloway, H. M., Griffith, D. R., and Mannering, J. V. 1981. Adaptability of various tillage-planting systems to Indiana soils. Coop. Ext. Serv. Bulletin AY210. West Lafayette, Ind.: Purdue University.

Gannon, J.E., and R.S. Stemberger. 1978. Zooplankton (especially crustaceans and rotifers) as indicators of water quality. T Am Microsc Soc;. 16-35.

Gascon, D., and W.C. Leggett. 1977. Distribution, abundance and resource utilization of littoral zone fishes in response to a nutrient-production gradient in Lake Memphremagog. J. Fish. Res..Bd Can. 34: 1105-1117.

Page 94: Taking Action On Lake Erie - International Joint Commission · reduce external inputs of total phosphorus (TP) and in particular, the rising proportion of the most bioavailable form

Gentry, L. E., David, M. B., Royer, T. V., Mitchell, C. A., and Starks, K. M. 2007. Phosphorus transport pathways to streams in tile-drained agricultural watersheds. J Environ Qual., 36(2), 408-415.

Gibb, A., Kelly, H., Schueler, T., Horner, R., Simmler, J., and Knutson, J. 1999. Best Management Practices Guide for Stormwater. Greater Vancouver Sewerage and Drainage District, Durnaby, B.C.

Gilliam, J. W. 1995. Phosphorus control strategies. Ecol. Eng., 5(2-3), 405-414. Gitau, M. W. 2013. LMAV BMP Tool II. http://www.box.net/shared/xuvt811rqa. Gomberg, A. 2007. Sewage overflow billions of gallons of sewage contaminate Lake Erie. Environment

Ohio and Environmental Ohio Research and Policy Center, Columbus, Ohio, 18. Goto, D., K. Lindelof, D.L. Fanslow, S.A. Ludsin, S.A. Pothoven, J.J. Roberts, H.A. Vanderploeg, A.E.

Wilson, and T.O. Höök. 2012. Indirect consequences of hypolimentic hypoxia on freshwater zooplankton growth in a large eutrophic lake. Aquatic Biol. 16: 217–227.

Gregor, D.J.and Rast. W. 1979. Trophic characterization of the US and Canadian near shore zones of the Great Lakes. International Joint Comission.

Gregor, D.J., and Rast, W. 1982. Simple trophic state classification of the Canadian near shore waters of the Great Lakes. J. Am. Wat. Res. Assoc. 18: 565 – 573.

Ha, J. H., T. Hidaka, and H. Tsuno. 2008. Quantification of Toxic Microcystis and Evaluation of Its Dominance Ratio in Blooms Using Real-Time PCR. Environ Sci Technol 43: 812-818.

Haltuch, M.A., Berkman, P.A., Garton, D.W. 2000. Geographic Information System (GIS) analysis of ecosystem invasion: exotic mussels in Lake Erie. . Limnol. Oceanogr. 45: 1778 – 1787.

Han, H., Allan, D.J., Bosch, N.S. 2012. Historical pattern of phosphorus loading to Lake Erie watersheds. J Great Lakes Res. 38: 289 – 298.

Hansen, M.J. 2010. The Asian car threat to the Great Lakes. Report to the House Committee on Transportation and Infrastructure.

Hargan, K.E., Paterson, A.M. & Dillon, Peter J. 2011. A total phosphorus budget for the Lake of the Woods and the Rainy River catchment. J. Gt Lakes Res. 37: 753–763.

Hartman, W. L. 1972. Lake Erie: effects of exploitation, environmental changes and new species on the fishery resources. Journal of the Fisheries Research Board of Canada 29:899-912.

Hawley, N., T. H. Johengen, Y. R. Rao, S. A. Ruberg, D. Beletsky, S. A. Ludsin, B. J. Eadie, D. J. Schwab, T. E. Croley, and S. B. Brandt. 2006. Lake Erie Hypoxia Prompts Canada-U.S. Study. EOS Transactions 87(32):313-324.

Hayhoe, K., VanDorn, J., Croley, T., Schlegal, N., Wuebbles, D. 2010. Regional climate change projections for Chicago and the US Great Lakes. J Great Lakes Res. 36: 7 – 21.

Hecky, R.E., Smith, R.E.H., Barton, D.R., Guildford, S.J., Taylor, W.D., Charlton, M.N., Howell, E.T. 2004. The nearshore phosphorus shunt: a consequence of ecosystem engineering by dreissenids in the Laurentian Great Lakes. Can J Fish Aquat Sci. 61: 1285 – 1293.

Henderson, P.A., and W.G.R. Crampton. 1997. A comparison of fish diversity and abundance between nutrient-rich and nutrient-poor lakes in the Upper Amazon. J, Tropical Ecol. 13: 175-198.

Henderson, B.A., J.L. Wong, and S.J. Nepszy. 1996. Reproduction of walleye in Lake Erie: allocation of energy. Can J Fish Aquat Sci. 53: 127-133.

Herut, B., Krom, M.D., Pan, G. & Mortimer, R. 1999. Atmospheric input of nitrogen and phosphorus to the Southeast Mediterranean: Sources, fluxes, and possible impact. Limnol. Oceanogr 1683–1692.

Higgins, S.N., Howell, E.T., Hecky, R.E., Guildford, S.J., Smith, R.E.H. 2005a. The wall of green: the status of Cladophora glomerata on the northern shores of Lake Erie’s eastern basin, 1995 – 2002. J Great Lakes Res. 31: 547 – 563.

Higgins, S.N., 2005b. Modeling the growth, biomass, and tissue phosphorus concentration of Cladophora glomerata in eastern Lake Erie: model description and field testing. J Great Lakes Res. 31: 439 – 455.

Higgins, S.N., Hecky, R.E., Guildford, S.J., Smith, R.E.H. 2006. Environmental controls of Cladophora growth dynamics in eastern Lake Erie: Application of the Cladophora Growth Model. J Great Lakes Res. 32: 629 – 644.

Higgins, S.N., Pennuto, C.M., Howell, E.T., Lewis, T., Makarewicz, J.C. 2012. Urban influences on Cladophora blooms in Lake Ontario. Great Lakes Res. 38: Suppl 4, 116–123J.

Higgins, S.N., Vander Zanden, M.J. 2010. What a difference a species makes: A meta analysis of

Page 95: Taking Action On Lake Erie - International Joint Commission · reduce external inputs of total phosphorus (TP) and in particular, the rising proportion of the most bioavailable form

dreissenid mussel impacts on freshwater ecosystems. Ecol Mon. 80: 179 – 196. Hipp, B., Alexander, S., and Knowles, T. 1993. Use of Resource-Efficient Plants to Reduce Nitrogen,

Phosphorus, and Pesticide Runoff in Residential and Commercial Landscapes. Water Sci. Technol., 28(3-5), 205-213.

Hoffmann, C. C., Heiberg, L., Audet, J., Schonfeldt, B., Fuglsang, A., Kronvang, B., Ovesen, N. B., Kjaergaard, C., Hansen, H. C. B., and Jensen, H. S. 2012. Low phosphorus release but high nitrogen removal in two restored riparian wetlands inundated with agricultural drainage water. Ecol. Eng., 46, 75-87.

Hogland, W., Niemczynowice, J., and Wahlan, T. (1987). The Unit Superstructure during the Construction Period. The Science of the Total Environment, 59 411-424.

Horner, R. R., and Horner, C. R. 1995. Design, Construction, and Evaluation of a Sand Filter Stormwater Treatment System, Part II, Performance Monitoring. Report to Alaska Marine Lines, Seattle, WA.

Howell, E.T., Chomicki, K.M., Kaltenecker, G. 2012. Tributary discharge, lake circulation and lake biology as drivers of water quality in the Canadian near shore of Lake Ontario. J Great Lakes Res. 38:47-62.

Howell, E.T., Marvin, C.H., Bilyea, R.W., Kauss, P.B., and Somers, K. 1996. Changes in environmental conditions during Dreissena colonization of a monitoring station in eastern Lake Erie. J Great Lakes Res. 22: 744–756.

Hubbard, R. K., Gascho, G. J., and Newton, G. L. 2004. Use of floating vegetation to remove nutrients from swine lagoon wastewater. T Asae, 47(6), 1963-1972.

Hudon, C., Cattaneo, A., Tourville Poirier, A-M., Bordeur, P., Dumont, P., Mailhot, Y., Amyot, J-P., Despatie, S-P., de Lafontaine, Y. 2012. Oligotrophication from wetland epuration alters the riverine trophic network and carrying capacity for fish. Aquat Sci.. 74: 495 – 511.

Hurley, S. E., and Forman, R. T. T. 2011. Stormwater ponds and biofilters for large urban sites: Modeled arrangements that achieve the phosphorus reduction target for Boston's Charles River, USA. Ecol. Eng., 37(6), 850-863.

Jeppesen, E., Søndergaard, M., Meerjoff, M., Lauridsen, T. L., Jensen, J. P. 2007. Shallow lake restoration by nutrient loading reduction – some recent findings and challenges ahead. Hydrobiologia 584:239-252.

Jiao, P. J., Xu, D., Wang, S. L., and Zhang, T. Q. 2011. Phosphorus loss by surface runoff from agricultural field plots with different cropping systems. Nutr. Cycl. Agroecosyst., 90(1), 23-32.

Jochimsen EM, Carmichael WW, An JS, Cardo DM, Cookson ST, Holmes CE, et al. 1998. Liver failure and death after exposure to microcystins at a hemodialysis center in Brazil. N Engl J Med. 338: 873-878.

Johansson, O.E., C. Dumitru, and D.M.Graham. 1999. Estimation of zooplankton mean length for use in an index of fish community structure and its application to Lake Erie. J Great Lakes Res. 25:179-186

Johansson, O.E., R. Dermott, D.M. Graham, J.A. Dahl, E.S. Millard, D.D. Myles, and J. LeBlanc. 2000. Benthic and pelagic secondary production in Lake Erie after the invasion of Dreissena spp. with implications for fish production. J Great Lakes Res. 26:31-54.

Johnson, T. B., and D.O. Evans.1990. Size-dependent winter mortality of young-of-the-year white perch: climate warming and invasion of the Laurentian Great Lakes. Trans.Am.Fish Soc. 119:301-313.

Kellershohn, DA and Tsanis IK 1999 3D Eutrophication Modeling of Hamilton Harbour: Analysis of Remedial Options J. Great Lakes Res. 25(1):3–25

Kipp, R., Ricciardi, A. 2012. Impacts of the Eurasian round goby (Neogobius melanostomus) on benthic communities in the upper St. Lawrence River. Can J Fish Aquat Sci. 69: 469-486.

Kirby, M.K., Dunford, W.E. 1982. Attached algae of the lake Erie shoreline near Nanticoke Generating Station. J. Gt Lakes Res.. 7: 249 – 257.

Kleinman, P. J. A., Sharpley, A. N., McDowell, R. W., Flaten, D. N., Buda, A. R., Tao, L., Bergstrom, L., and Zhu, Q. 2011. Managing agricultural phosphorus for water quality protection: principles for progress. Plant Soil, 349(1-2), 169-182.

Koelsch, R. K., Lorimor, J. C., and Mankin, K. R. 2006. Vegetative treatment systems for management of open lot runoff: Review of literature. Appl. Eng. Agric., 22(1), 141-153.

Kroger, R., Moore, M. T., Farris, J. L., and Gopalan, M. 2011. Evidence for the Use of Low-Grade Weirs

Page 96: Taking Action On Lake Erie - International Joint Commission · reduce external inputs of total phosphorus (TP) and in particular, the rising proportion of the most bioavailable form

in Drainage Ditches to Improve Nutrient Reductions from Agriculture. Water Air Soil Pollut., 221(1-4), 223-234.

Kunkel, K., Andsager, K., Easterling, D. 1999. Long – term trends in extreme precipitation events over the conterminous United States and Canada. J. Climate. 12: 2515 – 2527.

Lajeunesse, A., Segura, P.A., Gelinas, M., Hudon, C., Thomas, K., Quilliam, M.A., Gagnon, C. 2012. Detection and confirmation of saxitoxin analogues in freshwater benthic Lyngbya wollei algae collected in the St. Lawrence River (Canada) by liquid chromatography – mass spectrometry. J. Chromatog A. 1219: 93 – 103.

Leach, J. H. and S. J. Nepszy. 1976. Fish community in Lake Erie. J. Fish Res Bd Can 33:622-638. Lehman, J., Bell, D., and McDonald, K. 2008. Evidence for Reduced River Phosphorus Following

Implementation of a Lawn Fertilizer Ordinanc. Ecology and Evolutionary Biology Natural Science Building, University of Michigan, 7.

Leisenring, M., Clary, J., Stephenson, J., and Hobson, P. 2010. International Stormwater Best Management Practices (BMP) Database Pollutant Category Summary: Nutrients. Geosyntec Consultants, Wright Water Engineers, International Stormwater BMP Database.

Lenhart, H. A., and Hunt, W. F. 2011. Evaluating Four Storm-Water Performance Metrics with a North Carolina Coastal Plain Storm-Water Wetland. Journal of Environmental Engineering-Asce, 137(2), 155-162.

Lewis, T. W., and Makarewicz, J. C. 2009. Winter application of manure on an agricultural watershed and its impact on downstream nutrient fluxes. J Great Lakes Res. 35, 43-49.

Lewis, W. M., W. A. Wurtsbaugh, and H. W. Paerl. 2011. Rationale for Control of Anthropogenic Nitrogen and Phosphorus to Reduce Eutrophication of Inland Waters. Environ Sci Technol 45: 10300-10305.

Li, S., Elliott, J. A., Tiessen, K. H. D., Yarotski, J., Lobb, D. A., and Flaten, D. N. 2011. The effects of multiple beneficial management practices on hydrology and nutrient losses in a small watershed in the Canadian prairies. J Environ Qual., 40(5), 1627-1642.

Lory, J. A. 1999. Managing Manure Phosphorus to Protect Water Quality. Publication No. G09182. University of Missouri-Columbia Extension.

Ludsin, S. A., M. W. Kershner, K. A. Blocksom, R. L. Knight, and R. A. Stein. 2001. Life after death in Lake Erie: nutrient controls drive fish species richness, rehabilitation. Ecological Applications 11:731-746.

Ludsin, S., K. Pangle, L. Carreon-Martinez, N. Legler, J. Reichert, D.D. Heath, B. Fryer, T. Johnson, J. Tyson, T. Höök, G. Leshkevich, D. Mason, D. Bunnell, C. Mayer, T. Johengen, H. Vanderploeg, and A. Drelich. 2011. River discharge as a predictor of Lake Erie yellow perch recruitment. Final completion report, Great Lakes Fishery Commission, Fisheries Research Program, Ann Arbor, MI. 166 pp

Ludsin, S. A., X. S. Zhang, S. B. Brandt, M. R. Roman, W. C. Boicourt, D. M. Mason, and M. Costantini. 2009. Hypoxia-avoidance by planktivorous fish in Chesapeake Bay: Implications for food web interactions and fish recruitment. J Exper Mar Biol Ecol 381:S121-S131.

Lythgoe, J. N. 1979. The ecology of vision. Clarendon Press Oxford. Makarewicz, J. C. 1993a. A lakewide comparison of zooplankton biomass and its species composition in

Lake Erie, 1983-1987. J Great Lakes Res. Pages 19: 275-290 Mackarewicz, J., Lewis, T.W., Bertram, P. 1999. Phytoplankton composition and biomass in the offshore

waters of Lake Erie: Pre – and post-Dreissena introduction (1983 – 1993). J. Gt Lakes Res.. 25: 135 – 148.

Malkin, S.Y. 2009. The Ecology of the Nuisance Macroalga, Cladophora glomerata, and its Resurgence in Lake Ontario. PhD Thesis, University of Waterloo, Waterloo, Ontario, Canada.

Malkin, S.Y., Guildlford, S.J., Hecky, R.E. 2008. Modeling the growth response of Cladophora in a Laurentian Great Lake to the exotic invader Dreissena and to lake warming. Limnol. Oceanogr.. 53: 1111 – 1124.

Maniquiz, M. C., Lee, S. Y., and Kim, L. H. 2010. Long-term monitoring of infiltration trench for nonpoint source pollution control. Water Air Soil Pollut., 212(1-4), 13-26.

Mankin, K. R., Ngandu, D. M., Barden, C. J., Hutchinson, S. L., and Geyer, W. A. 2007. Grass-shrub riparian buffer removal of sediment, phosphorus, and nitrogen from simulated runoff. J. Amer.

Page 97: Taking Action On Lake Erie - International Joint Commission · reduce external inputs of total phosphorus (TP) and in particular, the rising proportion of the most bioavailable form

Water Res. Assoc., 43(5), 1108-1116. Mason, C.P. 1965. Ecology of Cladophora in farm ponds. Ecology. 46: 421 – 429. McDowell, R. W., Dou, Z., Toth, J. D., Cade-Menun, B. J., Kleinman, P. J. A., Soder, K., and Saporito, L.

2008. A comparison of phosphorus speciation and potential bioavailability in feed and feces of different dairy herds using IT nuclear magnetic resonance spectroscopy. J Environ Qual., 37(3), 741-752.

Meals, D. W., and Budd, L. F. 1998. Lake Champlain Basin nonpoint source phosphorus assessment. J. Amer. Water Res. Assoc, 34(2), 251-265.

Merriman, K. R., Gitau, M. W., and Chaubey, I. (2009). "Tool for estimating best management practice effectiveness in Arkansas. Appl. Eng. Agric., 25(2), 199-213.

Michalak, A., Anderson, E Beletsky D, et al. 2013. Record-setting algal bloom in Lake Erie caused by agricultural and meteorological trends consistent with expected future conditions PNAS 2013 ; published ahead of print April 1, 2013, doi:10.1073/pnas.1216006110

Migon, C. & Sandroni, V. 1999. Phosphorus in rainwater: Partitioning inputs and impact on the surface coastal ocean. . Limnol. Oceanogr 44: 1160–1164.

Moore, P. A., Daniel, T. C., and Edwards, D. R. 1999. Reducing phosphorus runoff and improving poultry production with alum. Poultry Sci, 78(5), 692-698.

Mostaghimi, S., Park, S. W., Cooke, R. A., and Wang, S. Y. 1997. Assessment of management alternatives on a small agricultural watershed. Water Res., 31(8), 1867-1878.

Mostaghimi, S., Younos, T. M., and Tim, U. S. 1991. The Impact of Fertilizer Application Techniques on Nitrogen Yield from 2 Tillage Systems. Agr Ecosyst Environ, 36(1-2), 13-22.

Mullen, R., Diedrick, K., and Henry, D. 2009. Best Management Practices for Mitigating Phosphorus Loss from Agricultural Soils. FACT SHEET, Agriculture and Natural Resources. AGF-509-09. The Ohio State University Extension.

Nalepa, T.F., Dermott, R., Madenjian, C., Schloesser, D.W. State of the Great Lakes 2012 Draft Report. Nicholls, K.H., Hopkins G.J. 1993. Recent changes in Lake Erie (north shore) phytoplankton: cumulative

impacts of phosphorus loading reductions and the zebra mussel introduction. J Great Lakes Res. 19: 637 – 647.

Nicholls, K. H., D. W. Standen, G. J. Hopkins, and E. C. Carney. 1977. Declines in the Near-Shore Phytoplankton of Lake Erie's Western Basin Since 1971. J Great Lakes Res 3: 72-78.

Nicolaisen, J. E., Gilley, J. E., Eghball, B., and Marx, D. B. 2007. Crop residue effects on runoff nutrient concentrations following manure application. Trans. ASABE, 50(3), 939-944.

Nistor, A. P., and Lowenberg-DeBoer, J. 2007. The profitability factor of controlled drainage implementation. J. Soil Water Conserv., 62(6), 156a-156a.

North, R.L., Smith, R.E.H., Hecky, R.E, Depew, D.C., Leon, L.F., Charlton, M.N., Guildford, S.J. 2012. Distribution of seston and nutrient concentrations in the eastern basin of Lake Erie pre- and post-dreissenid mussel invasion. J Great Lakes Res. 38:463-476.

NVPDC, and ESI. 1992. Northern Virginia BMP Handbook: A Guide to Planning and Designing Best Management Practices in Northern Virginia. Northern Virginia Planning District Commission and Engineers and Surveyors Institute, Fairfax, VA.

OH-EPA. 2010. Ohio Lake Erie Phosphorus Task Force Final Report. Division of Surface Water, Ohio Environmental Protection Agency, Columbus, Ohio.

Ontario Ministry of Environment, 2002. Surface water monitoring and assessment 1998. Lake Erie report featuring a summary of tributary and near shore conditionsand trends for the Lake Erie basin. Environmental Monitoring and Reporting Branch, Etobicoke, ON.

Ouellette, A. J. A., S. M. Handy, and S. W. Wilhelm. 2006. Toxic Microcystis is widespread in Lake Erie: PCR detection of toxin genes and molecular characterization of associated cyanobacterial communities. Microb Ecol 51: 154-165.

Paerl, H. W. and J. Huisman. 2008. Blooms like it hot. Science. 320:57. Paerl, H. W., and J. Huisman. 2009. Climate change: a catalyst for global expansion of harmful

cyanobacterial blooms. Environmental Microbiology Reports 1: 27-37. Painter, D.S., McCabe, .K.J. 1987. The influence of the Grand River on eastern Lake Erie Cladophora.

NWRI Publication No. 87 – 74. Pangle, K. L., T. D. Malinich, D. B. Bunnell, D. R. DeVries, and S. A. Ludsin. 2012. Context-dependent

Page 98: Taking Action On Lake Erie - International Joint Commission · reduce external inputs of total phosphorus (TP) and in particular, the rising proportion of the most bioavailable form

planktivory: interacting effects of turbidity and predation risk on adaptive foraging. Ecosphere 3:art114.

Patterson, M.W.R., Ciborowski, J.J.H., Barton, D.R. 2005. The distribution and abundance of Dreissena species (Dreissenidae) in Lake Erie, 2002. J Great Lakes Res. 31: 223 – 237.

Peng, F., Effler, S.W. 2010. Characterization of individual suspended mineral particles in western Lake Erie: Implications for light scattering and water clarity. J Great Lakes Res. 36: 686 – 698.

Persson, L., S. Diehl, L. Johansson, G. Andersson, and S. F. Hamrin. 1991. Shifts in fish communities along the productivity gradient of temperate lakes – patterns and the importance of size-structured interactions. J. Fish Biol. 38:281-293.

Pillsbury, R.W., R.L. Lowe, Y.D. Pan, and J.L. Greenwood. 2002. The indirect effects of zebra mussels (Dreissena polymorpha) on benthic algal resources in Saginaw Bay, Lake Huron. JNABS 21: 238-252.

Pinowska, A., Stevenson, R.J., Sickman, J.O., Albertin, A., Anderson, M. 2007. Integrated interpretation of survey for determining nutrient thresholds for macroalgae in Florida springs. Florida Department of Environmental Protection, Tallahassee, FL.

Pothoven, S.A., H.A. Vanderploeg, T.O. Höök, and S. A. Ludsin. 2012. Hypoxia modifies planktivore-zooplankton interactions in Lake Erie. Can J Fish Aquat Sci. 69: 2018-2028.

Pothoven, S.A., H.A. Vanderploeg, S.A. Ludsin, T.O. Höök, and S.B. Brandt. 2009. Feeding ecology of rainbow smelt and emerald shiner in Lake Erie’s central basin. J. Gt Lakes Res.. 35: 190-198

Powell, G. E., Ward, A. D., Mecklenburg, D. E., and Jayakaran, A. D. 2007. Two-stage channel systems: Part 1, a practical approach for sizing agricultural ditches. J. Soil Water Conserv., 62(4), 277-286.

Prepas, E.E., K.M. Field, T.P. Murphy, W.L. Johnson, J.M. Burke, and W. M. Tonn. 1997. Introduction to the Amisk Lake Project: oxygenation of a deep, eutrophic lake. Can J Fish Aquat Sci. 54:2105-2110.

Qureshi, N.A., and N.N. Rabalais. 2001. Distribution of zooplankton on a seasonally hypoxic continental shelf. Pages 61-76 in N.N. Rabalais and R.E. Turner, editors. Coastal hypoxia: consequences for living resources and ecosystems.

Rabalais, N.N., and R.E. Turner. 2001. Hypoxia in the northern Gulf of Mexico: description, causes and change. Pages 1-36 in N.N. Rabalais and R.E. Turner, editors. Coastal hypoxia: consequences for living resources and ecosystems.

Randall, C. W., Krome, E. C., and Randall, A. A. 1987. Available technology for the control of nutrient pollution in the Chesapeake Bay watershed. Medium: X; Size: Pages: 116.

Rao, Y. R., Schwab D.J. 2007. Transport and mixing between the coastal and offshore waters in the Great Lakes: A review. J Great Lakes Res. 33:202-218.

Rao, Y. R., N. Hawley, M. N. Charlton and W. M. Schertzer. 2008. Physical processes and hypoxia in the central basin of Lake Erie, Limnol. Oceanogr. 52 (5) 2007-2020

Reckhow, K.H., and Chapra, S.C. 1999. Modelling excessive nutrient loading in the environment. Environ Pollut. 100: 197–207.

Reichert, J. M., B. J. Fryer, K. L. Pangle, T. B. Johnson, J. T. Tyson, A. B. Drelich, and S. A. Ludsin. 2010. River-plume use during the pelagic larval stage benefits recruitment of a lentic fish. Can J Fish Aquat Sci. 67:987-1004.

Richards, R.P., Baker, D.B., Kramer, J.W., Ewing, D.E., Merryfield, B.J., Miller, N.L. 2001. Storm discharges, loads and average concentrations in northwest Ohio rivers, 1975 – 1995. J. Amer. Water Res. Assoc. 37: 423 – 438.

Richards, R. P., Baker, D. B., Crumrine, J. P., and Stearns, A. M. 2010. Unusually large loads in 2007 from the Maumee and Sandusky Rivers, tributaries to Lake Erie. J. Soil Water Conserv., 65(6), 450-462.

Richards, C. E., Munster, C. L., Vietor, D. M., Arnold, J. G., and White, R. 2008. Assessment of a turfgrass sod best management practice on water quality in a suburban watershed. J Environ Manage., 86(1), 229-245.

Ridame, C. & Guieu, C. 2002. Saharan input of phosphate to the oligotrophic water of the open western Mediterranean Sea. Limnol. Oceanogr. 856–869.

Rinta-Kanto, J. M., and S. W. Wilhelm. 2006. Diversity of microcystin-producing cyanobacteria in spatially isolated regions of Lake Erie. Appl Environ Microb 72: 5083-5085.

Page 99: Taking Action On Lake Erie - International Joint Commission · reduce external inputs of total phosphorus (TP) and in particular, the rising proportion of the most bioavailable form

Rinta-Kanto, J. M., M.A. Saxtona, J.M. DeBruynb, Juliette L.S., C.H. Marvine, K.A. Kriegerf, G.S. Saylerb, G.L. Boyer, S.W. Wilhelma. 2009. The diversity and distribution of toxigenic Microcystis spp. in present day and archived pelagic and sediment samples from Lake Erie. Harmful Algae 8: 385-394.

Roberts, J.J. T.O. Höök, S.A. Ludsin, S.A. Pothoven, H.A. Vanderploeg, and S.B. Brandt. 2009. Effects of hypolimnetic hypoxia on foraging and distributions of Lake Erie yellow perch J Exp Mar Biol Ecol. 381: S132-S142.

Roberts, JJ., S.B. Brandt, D. Fanslow, S.A. Ludsin, S.A. Pothoven, D. Scavia, and T.O. Höök. 2011. Effects of hypoxia on consumption, growth, and RNA:DNA ratios of young yellow perch. Trans. Am. Fish. Soc.. 40: 1574–1586.

Roberts, J.J., P.A. Grecay, S.A. Ludsin, S.A. Pothoven, H.A. Vanderploeg, and T.O. Höök. 2012. Evidence of hypoxic foraging forays by yellow perch (Perca flavescens) and potential consequences for prey consumption. Freshwater Biol.. 57: 922-937.

Rogers, J. S., Potter, K. W., Hoffman, A. R., Hoopes, J. A., Wu, C. H., and Armstrong, D. E. 2009. Hydrologic and Water Quality Functions of a Disturbed Wetland in an Agricultural Setting. J. Amer. Water Res. Assoc, 45(3), 628-640.

Rosa, F., and N. M. Burns. 1987. Lake Erie central basin oxygen depletion changes from 1929-1980. J Great Lakes Res. 13:684-696.

Ross, S. 2006. Molecular phylogeography and species discrimination of freshwater Cladophora (Cladophorales, Chlorophyta) in North America. M.Sc. Thesis, University of Waterloo, Waterloo, ON., Canada.

Rossi, C. G., Dybala, T. J., Amonett, C., Arnold, J. G., and Marek, T. 2012. Manure nutrient management effects in the Leon River Watershed. J. Soil Water Conserv., 67(3), 147-157.

Roy-Poirier, A., Champagne, P., and Filion, Y. 2010. Bioretention processes for phosphorus pollution control. Environ. Rev., 18, 159-173.

Rucinski, D.K., Beletsky, D., DePinto, J.V., Schwab, D.J. and Scavia, D. 2010. A simple 1-dimensional, climate based dissolved oxygen model for the central basin of Lake Erie. J Great Lakes Res. 36:465-476.

Ryan, P., R. Knight, R. MacGregor, G. Towns, R. Hoopes, and W. Culligan. 2003. Fish-community goals and objectives for Lake Erie. Special Publication, Great Lakes Fishery Commission 3:56.

Saxton, M. A., R. J. Arnold, R. A. Bourbonniere, R. M. Mckay, and S. W. Wilhelm. 2012. Plasticity of total and intracellular phosphorus quotas in Microcystis aeruginosa cultures and Lake Erie algal assemblages. Frontiers in Microbiology 3.

Schindler, D. W. 2012. The dilemma of controlling cultural eutrophication of lakes. Proc. Royal Soc. B: DOI: 10.1098/rspb.2012.1032

Schottler, S. P., and S. J. Eisenreich. 1994. Herbicides in the Great Lakes. Environ Sci Technol 28: 2228-2232.

Schroeder, P. D., Radcliffe, D. E., and Cabrera, M. L. 2004. Rainfall timing and poultry litter application rate effects on phosphorus loss in surface runoff. J. Environ. Qual., 33(6), 2201-2209.

Sedorovich, D. M., Rotz, C. A., Vadas, P. A., and Harmel, R. D. 2007. Simulating management effects on phosphorus loss from farming systems. Trans. ASABE, 50(4), 1443-1453.

Seo, Y., Lee, J., Hart, W. E., Denton, H. P., Yoder, D. C., Essington, M. E., and Perfect, E. 2005. Sediment loss and nutrient runoff from three fertilizer application methods. T Asae, 48(6), 2155-2162.

Sharpley, A. N., Daniel, T., Gibson, G., Bundy, L., Cabrera, M., Sims, T., Stevens, R., Lemunyon, J., Kleinman, P., and Parr, R. 2006. Best Management Practices To Minimize Agricultural Phosphorus Impacts on Water Quality. U.S. Department of Agriculture, Agricultural Research Service. ARS–163. 50 pp.

Sher, D., J. W. Thompson, N. Kashtan, L. Croal, and S. W. Chisholm. 2011. Response of Prochlorococcus ecotypes to co-culture with diverse marine bacteria. ISME J 5:1125-1132.

Shreve, B. R., Moore, P. A., Daniel, T. C., Edwards, D. R., and Miller, D. M. 1995. Reduction of Phosphorus in Runoff from Field-Applied Poultry Litter Using Chemical Amendments. J. Environ. Qual., 24(1), 106-111.

Sly, P. G. 1976. Lake Erie and its basin. Journal of the Fisheries Research Board of Canada 33:355-370.

Page 100: Taking Action On Lake Erie - International Joint Commission · reduce external inputs of total phosphorus (TP) and in particular, the rising proportion of the most bioavailable form

Smith, D. R., Moore, P. A., Griffis, C. L., Daniel, T. C., Edwards, D. R., and Boothe, D. L. 2001. Effects of alum and aluminum chloride on phosphorus runoff from swine manure. J. Environ. Qual., 30(3), 992-998.

Soster, F.M., P.L. McCall, and K.A. Herrmann. 2011. Decadal changes in the benthic invertebrate community in western Lake Erie between 1981 and 2004. J Great Lakes Res. 37:226-237.

Sprules, W.G. 2008. Ecological change in Great Lakes communities - a matter of perspective. Can J Fish Aquat Sci. 65:1-9

Steffen, M. M., Z. Li, T. C. Effler, L. J. Hauser, G. L. Boyer, and S. W. Wilhelm. 2012. Comparative Metagenomics of Toxic Freshwater Cyanobacteria Bloom Communities on Two Continents. PLoS ONE 7: e44002.

Stewart, T.W., Lowe, R.L. 2008. Benthic algae of Lake Erie (1865 – 2006): A review of assemblage composition, ecology and causes and consequences of changing abundance. Ohio J.Sci. 108: 82 – 94.

Strecker, E. W., Quigley, M. M., Urbonas, B. R., Jones, J. E., and Clary, J. K. 2001. Determining urban storm water BMP effectiveness. J. Water Resour. Plan. Manage.-ASCE, 127(3), 144-149.

Stumpf, R.P., Wynne, T.T., Baker, D.B. and Fahnenstiel, G.L. 2012. Interannual variability of cyanobacterial blooms in Lake Erie. PLoS ONE. 7, e42444.

Sweeney, D. W., Pierzynski, G. M., and Barnes, P. L. 2012. Nutrient losses in field-scale surface runoff from claypan soil receiving turkey litter and fertilizer. Agr Ecosyst Environ, 150, 19-26.

Tan, C. S., and Zhang, T. Q. 2011. Surface runoff and sub-surface drainage phosphorus losses under regular free drainage and controlled drainage with sub-irrigation systems in southern Ontario. Can J Soil Sci, 91(3), 349-359.

Tan, C. S., Zhang, T. Q., Drury, C. F., Reynolds, W. D., Oloya, T., and Gaynor, J. D. 2007. Water quality and crop production improvement using a wetland-reservoir and drainage/subsurface irrigation system. Can.Wat.Res.J., 32(2), 111-118.

Taylor, J. C., P. S. Rand, and J. Jenkins. 2007. Swimming behavior of juvenile anchovies (Anchoa spp.) in an episodically hypoxic estuary: implications for individual energetics and trophic dynamics. Mar. Biol. 152:939-957.

Tiessen, K. H. D., Elliott, J. A., Yarotski, J., Lobb, D. A., Flaten, D. N., and Glozier, N. E. 2010. Conventional and Conservation Tillage: Influence on Seasonal Runoff, Sediment, and Nutrient Losses in the Canadian Prairies. J. Environ. Qual., 39(3), 964-980.

Tomlinson, L.M., Auer, M.T., Bootsma, H.A., Owens, E.M. 2010. The Great Lakes Cladophora Model: Development, testing and application to Lake Michigan. J Great Lakes Res. 36: 287 – 297.

Tsukuda, S., Sugiyama, M., Harita, Y. & Nishimura, K. 2004. A methodological re-examination of atmospheric phosphorus input estimates based on spatial microheterogeneity. Water Air Soil Pollut 152: 333–347.

Ulen, B., Aronsson, H., Bechmann, M., Krogstad, T., Oygarden, L., and Stenberg, M. 2010. Soil tillage methods to control phosphorus loss and potential side-effects: a Scandinavian review. Soil Use Manage, 26(2), 94-107.

US EPA. 2000. AP-42 Compilation of Air Pollutant Emission Factors, Chapter 13: Miscellaneous Sources, 5th ed. U.S Environmental Protection Agency.

USDA-NRCS. 2008. National Handbook of Conservation Practices, Conservation Practice Standard Code 554 (Drainage Water Management). USDA Natural Resources Conservation Service, Washington, DC.

USDA-NRCS. 2008. National Handbook of Conservation Practices, Conservation Practice Standard Code 635 (Vegetated Treatment Area). Washington, DC: USDA Natural Resources Conservation Service.

USDA-NRCS 2003. National Handbook of Conservation Practices, Conservation Practice Standard Code 313 (Waste Storage Facility). Washington, DC: USDA Natural Resources Conservation Service.

USDA-NRCS. 2003. National Handbook of Conservation Practices, Conservation Practice Standard Code 317 (Composting Facility). Washington, DC: USDA Natural Resources Conservation Service.

USDA-NRCS. 2010. National Handbook of Conservation Practices, Conservation Practice Standard Code 332 (Contour Buffer Strips). USDA Natural Resources Conservation Service, Washington,

Page 101: Taking Action On Lake Erie - International Joint Commission · reduce external inputs of total phosphorus (TP) and in particular, the rising proportion of the most bioavailable form

DC. USDA-NRCS. 2010. National Handbook of Conservation Practices, Conservation Practice Standard

Code 391 (Riparian Forest Buffers). USDA Natural Resources Conservation Service, Washington, DC.

USDA-NRCS. 2010. National Handbook of Conservation Practices, Conservation Practice Standard Code 393 (Filter Strip). USDA Natural Resources Conservation Service, Washington, DC.

USDA-NRCS. 2010. National Handbook of Conservation Practices, Conservation Practice Standard Code 412 (Grassed Waterway). USDA Natural Resources Conservation Service, Washington, DC.

USDA-NRCS. 2011. National Handbook of Conservation Practices, Conservation Practice Standard Code 327 (Conservation Cover). USDA Natural Resources Conservation Service, Washington, DC.

USDA-NRCS. 2011. National Handbook of Conservation Practices, Conservation Practice Standard Code 328 (Crop Rotation). USDA Natural Resources Conservation Service, Washington, DC.

USDA-NRCS. 2011. National Handbook of Conservation Practices, Conservation Practice Standard Code 329 (No Till/Strip Till/Direct Seed). Washington, DC: USDA Natural Resources Conservation Service.

USDA-NRCS. 2011. National Handbook of Conservation Practices, Conservation Practice Standard Code 345 (Mulch-till). Washington, DC: USDA Natural Resources Conservation Service.

USDA-NRCS. 2011. National Handbook of Conservation Practices, Conservation Practice Standard Code 346 (Ridge Till). USDA Natural Resources Conservation Service, Washington, DC.

USDA-NRCS. 2011. National Handbook of Conservation Practices, Conservation Practice Standard Code 592 (Feed Management). Washington, DC: USDA Natural Resources Conservation Service.

USDA-NRCS. 2011. National Handbook of Conservation Practices, Conservation Practice Standard Code 606 (Subsurface Drain)." USDA Natural Resources Conservation Service, Washington, DC.

USDA-NRCS. 2011. National Handbook of Conservation Practices, Conservation Practice Standard Code 607 (Surface Drain, Field Ditch). USDA Natural Resources Conservation Service, Washington, DC.

USDA-NRCS. 2011. National Handbook of Conservation Practices, Conservation Practice Standard Code 608 (Surface Drain, Main and Lateral). USDA Natural Resources Conservation Service, Washington, DC.

USDA-NRCS. 2012. National Handbook of Conservation Practices, Conservation Practice Standard Code 590 (Nutrient Management). USDA Natural Resources Conservation Service, Washington, DC.

USEPA. 1993. Guidance Specifying Management Measures for Sources of Nonpoint Pollution in Coastal Waters. EPA 840-B-92-002., U.S. Environmental Protection Agency, Washington, DC.

USEPA. 2009. INDICATOR: Phosphorus Discharges from Detroit Wastewater Treatment Plant. Detroit River-Western Lake Erie Basin Indicator Project, <http://www.epa.gov/med/grosseile_site/indicators/dwwtp.html>. (November 15, 2012).

USEPA. 2005. Stormwater Phase II Final Rule. Office of Water, ed., U.S. Environmental Protection Agency, Washington, D.C.

Van de Moortel, A. M. K., Rousseau, D. P. L., Tack, F. M. G., and De Pauw, N. 2009. A comparative study of surface and subsurface flow constructed wetlands for treatment of combined sewer overflows: A greenhouse experiment. Ecol. Eng., 35(2), 175-183.

Vanderploeg, H.A., S.A. Ludsin, S.A. Ruberg, T.O. Höök, S.A. Pothoven, S. B. Brandt, G.A. Lang, J.R. Liebig, and J.F. Cavaletto. 2009a. Hypoxia affects spatial distribution of pelagic fish, zooplankton, and phytoplankton in Lake Erie. Journal of Experimental Marine Biology and Ecology. 381: S92-S107.

Vanderploeg, H.A., S.A. Ludsin, J.F. Cavaletto, T.O. Höök, S.A. Pothoven, S. B. Brandt, J.R. Liebig, and G.A. Lang. 2009b. Hypoxic zones as habitat for zooplankton in Lake Erie: refugees from predation or exclusions zones? Journal of Experimental Marine Biology and Ecology. 381: S108-S120.

Vis, C., Cattaneo, A., Hudon, C. 2008. Shift from chlorophytes to cyanobacteria in benthic microalgae

Page 102: Taking Action On Lake Erie - International Joint Commission · reduce external inputs of total phosphorus (TP) and in particular, the rising proportion of the most bioavailable form

along a gradient of nitrate depletion. J. Phycol.. 44: 38 -44. Vollenweider, R.A., 1968. Scientific fundamentals of the eutrophication of lakes and flowing waters, with

particular reference to nitrogen and phosphorus as factors in eutrophication. OECD Tech. Rep. Wellington, C. G., C. M. Mayer, J. M. Bossenbroek, and N. A. Stroh. 2010. Effects of turbidity and prey

density on the foraging success of age 0 year yellow perch Perca flavescens. J. Fish Biol. 76:1729-1741.

WERF. 2003. Metals Removal Technologies for Urban Stormwater. Report # 97-IRM-2. Water Environ. Fed., Alexandria, VA, London.

Winston, R. J., Hunt, W. F., Osmond, D. L., Lord, W. G., and Woodward, M. D. 2011. Field Evaluation of Four Level Spreader-Vegetative Filter Strips to Improve Urban Storm-Water Quality. J. Irrigation and Drainage Engineering-Asce, 137(3), 170-182.

Winter, J. G., and Duthie, H. C. 2000. Export coefficient modeling to assess phosphorus loading in an urban watershed. J. Am. Wat..Res. Assoc. 36(5), 1053-1061.

Winter, J.G., Eimers, M.C., Dillon, P. J., Scott, L.D., Scheider, W.A. & Willox, C.C. 2007. Phosphorus inputs to Lake Simcoe from 1990 to 2003: declines in tributary loads and observations on lake water quality. J. Gt Lakes Res. 33: 381–396.

Young, G. K., Stein, S., Cole, P., Kammer, T., Graziano, F., and Bank, F. (1996). "Evaluation and Management of Highway Runoff Water Quality. ." Federal Highway Administration Report No. PD-96-032, U.S. Department of Transportation, Washington, DC.Zhai, S., Yang, L. & Hu, W. 2009. Observations of Atmospheric Nitrogen and Phosphorus Deposition During the Period of Algal Bloom Formation in Northern Lake Taihu, China. Environ. Man. 44: 542–551.

Zhang, R., Zhou, W. B., Li, J., and Yu, S. L. 2010. Field evaluation of an innovative stormwater treatment device-the Stormvault(TM) system. Environ. Monit. Assess., 169(1-4), 113-123.

Zhou, Y., Obenour, D.R., Scavia, D., Johengen, T.H. and Michalak, A.M., 2013. Spatial and Temporal Trends in Lake Erie Hypoxia, 1987-2007. Environ Sci Technol. 47, 899-905.