ha.“ h: .. It} ||| , 5.... . . , \. v... A In... an .ywzga , . .z an? v2 ‘flnfi... fl w. b . 3 2.. “no 4:: it . 1‘. Rut: .mefizwmm . Whammy“. . 1.5. rwé l a ,,.. -i. p... s. u :1 It: - A .! A. ‘ . é... . wiymlxwuss ‘ I l (4 1L . , VD;- - . I ifiQHIHmws fimvpmfi . 1. . 5. :11“. ‘ .‘ Waugh» 5 E3 k’g‘§.$ g'm. I.) A 5M ”1“) “Va n—r :-- This is to certify that the I RARY dissertation entitled Mlch 3r: State Ur versity Assessing wetlands and their restoration using algae presented by Lei Zheng has been accepted towards fulfillment of the requirements for the PhD. degree in Zoology/ EEBBirogram 7%,.Eim ,...._ Major Professor’s Signature 4314M /, 4005 \J Date MSU is an Affinnativo Action/Equal Opportunity Institution PLACE IN RETURN Box to remove this checkout from your record. To AVOID FINES return on or before date due. MAY BE RECALLED with earlier due date if requested. DATE DUE DATE DUE DATE DUE EP242007 r3240? 6/01 cJClRC/DateDue.p65—p.15 ASSESSING WETLANDS AND THEIR RESTORATION USING ALGAE By Lei Zheng A DISSERTATION Submitted to Michigan State University in partial fulfillment of the requirements For the degree of DOCTOR OF PHILOSOPHY Department of Zoology and the graduate program in Ecology, Evolutionary biology, and Behavior (EEBB) 2003 ABSTRACT ASSESSING WETLANDS AND THEIR RESTORATION USING ALGAE By Lei Zheng This dissertation develops a better understanding of relationship among a variety of ecological factors within wetlands, evaluates a number of valuable ecological attributes to develop indicators for wetland bioassessment and wetland restoration, and establishes biological and environmental criteria to protect wetland ecosystems. Ecological assessment in wetlands has to develop habitat or class specific attributes to reduce the spatial and temporal variability. I first evaluated spatial and temporal variation of environmental and algal attributes among habitats within wetlands and between restored and extant wetlands in Michigan. The results indicated non-diatom algal assemblages differed among habitats, but sediment and epiphytic diatom assemblages were not significantly different. Epiphytic assemblages in the restored and extant wetlands were relatively distinct, while both sediment and phytoplankton assemblages were not as different between restored and extant wetlands. The differences in algal assemblages among habitats within wetlands and between restored and extant wetlands were related to a number of environmental characteristics, but particularly nutrients and sediment organic matter. In my study of salt marsh restoration, organic matter in sediments and the similarity of diatom species composition between paired restored and reference salt marshes increased with age of restored marshes during spring and summer. Primary production by epiphytic and sediment algae in summer showed Site-specific changes and did not change consistently with marsh age. In this study, pairwise comparison of a restored marsh and a nearby reference marsh reduced regional variation among wetlands and helped identify the successional pattern of ecological development after restoration. Because nutrients were one of the most important factors regulating periphyton growth and species composition in wetlands, experimental and field surveys were conducted to determine the limiting factors in wetlands. Experiments indicated that algae were nitrogen limited and limitation was directly related to nutrient concentrations and nutrient ratios. Although individual diatom taxa responded to nutrient additions differently, the responses of indicator taxa in the nutrient bioassays were positively correlated with natural nutrient gradients among natural wetlands indicating variation in species composition among restored wetlands was related to N-limitation. Based on the establishment of a cause-effect and stressor-response relationships between nutrient stressors and algal attributes, I evaluated the response of a number of valued ecological attributes to nutrient stressors. The similarity in species composition to reference conditions and the number of native species decreased and trophic index calculated with species composition increased with increasing nutrient concentrations. Diatoms as indicators of wetland nutrient pollution and human disturbance were developed using paleoecological diatom assemblages to characterize reference conditions. Biological criterion was then established based on the historic trophic status indicated by diatom species composition, and a nutrient criterion to protect or restore the biological integrity of wetlands was then established from the relationship between trophic status index and nutrients and using the predicted nutrient concentration. My studies indicated that algal assemblages were valuable indicators for assessing biological condition of wetlands, monitoring restoration, and diagnosing stressors. This thesis is dedicated to My wife Jingfang Chen And My parents For their infinite love and constant inspiration iv ACHNOWLEDGEMENTS First, I would like to thank my major professor, Dr. R. J an Stevenson, for his constant support and encouragement throughout my graduate studies. His academic guidance was essential to the completion of this dissertation and has provided important insights to my scientific research. I would also like to express the gratitude for my committee members, Dr. Thomas Burton, Dr. Katherine L. Gross, and Dr. Donald Hall, for reading early drafts of this dissertation and providing valuable comments that helped to improve the presentation and contents of this dissertation. I am also grateful to all the people in the algae lab for their great help to accomplish this research. I would like to thank my wife Jenny Chen for her support and love during the dissertation writing. Her encouragement made this dissertation possible. My parents receive my deepest gratitude and love for their dedication and many years of support. TABLE OF CONTENTS List of Tables .......................................................................................... iv List of Figures ......................................................................................... xi Chapter 1. Algae as indicators of Biological Integrity in restored and natural wetland ecosystems 1.1 Wetland bioassessment... 1.2 Algal indicators... 1. 3 Wetland biogeochemistry and primary producers 1.4 Objectives of the dissertation... 1.5 References... bdbbb Chapter 2. Algal assemblages in multiple habitats of restored and extant wetlands 2.1 Abstract .......................................................................................................... 11 2. 2 Introduction... 13 2. 3 Methods... ....16 2.3.1 Site description 16 2. 3. 2 Sampling... 17 2.3.3 Samples analySis 18 2. 3. 4 Statistics ................................................................................................. 19 2. 4 Results... 21 2.4.1 Environmental vanables 1n reStored and extant wetlands... .......21 2. 4. 2 Non- diatom algal assemblage and diatom assemblages 1n restored and extant wetlands ................................................................................................ 26 2. 4. 3 Difference of taxonomic composition among different habitats and between restored and extant wetlands... 3O 2. 4. 4 Algal assemblages and env1ronmental vanables . . . . . . . ... ...33 2. 5 Discussion... ...37 2.5.1 Wetland restoration and env1ronmental change In reStored wetlands ........ 37 2. 5.2 Difference of algal species composition between restored and extant" ...38 2.5.3 Difference of algal species composition among different habitats ............ 41 2.6 References ...................................................................................................... 44 Chapter 3. Changes in benthic algal attributes during salt marshes restoration and use in bioassessment 3.1 AbstractSO 3.21ntroductlon52 3.3 Method554 3.3.1 Site location ........................................................................................... 54 3.3.2 Field samplmg55 3.3.3 Laboratory analysesS7 3. 3. 4 Statistical analyse359 3 .4 Results... 61 3.4.1 Biomass development 61 vi 3. 4. 2 Algal species diversity and composition.................................................64 3.4.3Productivity.... 69 3. 4. 4 Algal attribute development 1n Virginia salt marshes . . . . . . ... .69 3.5 Discussion” 72 3. 6 Acknowledgments ........................................................................................... 79 3. 7 References... 80 Chapter 4. Relations between nutrient conditions, algal growth, and species composition in the restored wetlands of southern Michigan 4.1 Abstract .......................................................................................................... 86 4. 2 Introduction... 87 4. 3 Methods... ..89 4.3.1 Sample collection 8.9 4. 3. 2 Nutrient diffusion experiment .................................................................. 91 4. 3. 3 Laboratory analyses ................................................................................ 92 4.3.4 Data analyses .......................................................................................... 94 4.4 Results ............................................................................................................ 95 4.4.1 Wetland habitats ...................................................................................... 95 4.4.2 Effects of nutrient enrichment and nutrient concentrations.......... . . . . . . . ......97 4.4.3 Algal species composition in natural and artificial substrata....... . . . . . . . .....101 4.4.4 Algal assemblage response to nutrient additions.....................................106 4.5 Discussion ..................................................................................................... 109 4.5.1 Algal nutrient limitation ......................................................................... 109 4.5.2 Algal species composition and nutrient concentrations.......................115 4.6 Conclusions ................................................................................................... 116 4.7 Acknowledgments118 4.8 References ..................................................................................................... 119 Chapter 5. Assessing wetland response to land use using diatoms and paleoecological characterization of reference condition 5.1 Abstracts ....................................................................................................... 127 5. 2 Introduction ................................................................................................... 128 5.3 Methods... .... 131 5.3.1 Material arid methods 131 5.3.2 Land use coverage ................................................................................. 133 5.3.3 Statistical analyses ................................................................................ 133 5.4 Results ...................................................................................... 135 5.4.1 Environmental variables ........................................................................ 135 5.4.2 Current diatom assemblages and environmental conditions............. . . . . ...139 5.4.3 Comparison between reference and current conditions.......... . . . . . ......143 5.5 Discussion ..................................................................................................... 147 5.6 ReferenceslSZ Chapter 6. Conclusion ............................................................................................. 158 vii Appendices Appendix 1 ....................................................................................... 164 Appendix 2 ....................................................................................... 165 viii LIST OF TABLES Table 2.1 Correlation matrix of environmental variables measured and principal component analysis (PCA) loadings of environmental variables in the 45 wetlands. When a correlation coefficient r >0.30, Dunn-Sidak P<0.05; when r>0.39, P<0.01. Correlation between age and environmental variables had a degree of freedom 24 and none of the correlation was significant ................ 23 Table 2.2 Mean alpha diversity (Number of species), Gamma (landscape) diversity, and beta diversity between habitats with a wetland and between restored and extant wetlands .................................................................................. 29 Table 2.3 MRPP comparisons of species composition in non-diatom algae and diatom between different habitats ..................................... - ........................ 31 Table 2.4 MRPP comparisons of species composition of non-diatom algae and diatoms between restored and extant wetlands ............................................... 32 Table 2.5 Pearson correlation matrix of dissimilarities in species composition of non- diatom and diatom between restored and extant wetlands in different habitats. A * indicates a P value <0.05. ...................................................... 33 Table 2.6 Regression by forward stepwise selection of independent variables. Dependent variable is dissimilarity among epiphytic, phytoplankton, and sediment non- diatom algae, and dissimilarity between epiphytic and sediment diatom assemblages, independent variables are uncorrelated environmental variables .................................................................................. 35 Table 2.7 Linear regressions by forward stepwise selection of independent variables. Dependent variable is average Bray-Curtis dissimilarity of algal species compositions in a restored wetland with extant wetlands. Independent variables are average dissimilarities of a restored wetland with all extant wetlands ...... 36 Table 3.1 Sampling sites and their properties during the summer sampling season ...... 56 Table 3.2 Results of regression analyses between algal attributes and environmental variables ................................................................................. 64 Table 3.3 Number of species found in all sites, habitats and seasons ....................... 64 Table 3.4 The 10 most dominant diatom species in all sites (sorted by relative abundance rank, habitat type, marsh type, and season) ........................................ 70 Table 4.1 Correlation matrix of ambient nutrients in 22 wetlands. A * indicates a significant correlation (Dunn-Sidak P<0.05) ....................................... 97 ix Table 4.2 Algal Nutrient Limitation and Corresponded Water Nutrient and sediment nutrients. Value in the parentheses are correlations between log-transformed value and algal nutrient limitation. A * indicates significant correlation (Dunn- Sidak P<0.05) ................................................................................. 100 Table 4.3 Weighted correlation matrix from Canonical Correspondent Analysis (CCA) .................................................................................. 103 Table 4.4 Twenty-six most dominant taxa and their responses to nutrient additions. The 4.digit number represents comparisons of the relative abundance of a diatom taxa in a treatment with control, N-enriched, P—enriched, and NP-enriched treatments. A significant change (P<0. l) of relative abundance between treatments was presented by 1, and insignificant effect was 0. * is empty when compared with itself .................................................................. 107 Table 4.5 Correlations between relative abundance of nutrient indicator species and nutrient concentration in different habitats ........................................ 108 Table 5.1 Pearson correlation matrix of environmental variables of 35 wetlands across Lower Peninsula of Michigan in 1994.995. A significant correlation (Dunn- Sidak Probabilities P<0.05) is indicated by * ...................................... 138 Table 5.2 The most dominant diatom taxa used in CCA and the correspondent taxa code ..................................................................................... 140 Table 5.3 Regression of observed environmental variables against predicted values based on diatom species composition. Simple regression and jackknifing cross- validation indicators based on weighted average (WA) and tolerance down- weighted WA used to compare the precision. MSE represents mean square error and RMSE represent residual of MSE ......................................... 142 Table 5.4 Correlation among community characteristics and environmental variables ................................................................................. 145 LIST OF FIGURES Figure 1.1 Stressors-response relationship between human disturbance and response of ecological valuable attributes .............................................. _ ........... 8 Figure 2.1 Wetland sampling locations in southern Michigan ................................ 16 Figure 2.2 Wetland physical and chemical characteristics and comparison between restored and reference wetlands. .................................................... 24 Figure 2.3 Wetland nutrient concentration and comparison between restored and extant wetlands ................................................................................. 26 Figure 2.4 Non-metric multidimentional scaling (NMDS) of the restored and extant wetlands based on non-diatom and diatom assemblages in different habitats. The circles are the sample ellipse which is centered on the sample means of the x and y variables. The size the ellipse is specified a probability value of 0.6827 .................................................................................... 30 Figure 2.5 Non-metric multidimentional scaling (NMDS) of the restored and extant wetlands based on non-diatom and diatom assemblages in different habitats. The circles are the sample ellipses which are centered on the sample means of the x and y variables. The size the ellipse is specified a probability value of 0.6827 .................................................................................... 31 Figure 3.1 Map of sampling sites between Morehead City and Wilmington along the North Carolina coast. Age (yr) of restored marshes is reported in parentheses .............................................................................. 54 Figure 3.2 Box-plot of dry mass (DM), ash free dry mass (AFDM), the organic matter proportion of sediments (0MP), algal biovolume, and Ch] a in reference marshes in surface sediments and on macrophytes during spring and summer 1998. .................................................................................... 61 Figure 3.3 Relationship between the attribute restoration ratio and age of restored marsh in surface sediment and on macrophytes during spring and summer 1998. The OMRR is the comparison of 0MP of sediments in a restored site with that of its reference marsh. The ACRR is the comparison of algal chlorophyll a concentration in a restored site with that of its reference marsh. The AVRR is the comparison of algal biovolume in a restored site with that of its reference marsh. The line represents a significant relationship (linear regression P<0.05). The correlation coefficients (R2) for each relationship are: 3a: 0.727; 3b, 0.931; 3c, 0.969; 3d, 0.235; 3e, 0.304; 3f, 0.647; 3g, 0.577; 3h, 0.749; 3i, 0.442; 3j, 0.741 ......................................................................... 63 xi Figure 3.4 Relationship between the percentage similarity index (PSI) of algal functional groups (4a, 4b) and diatom assemblages (4c, 4d) and the age of restored marshes in surface sediments and on macrophytes during spring and summer 1998. PSI was calculated for algal functional groups or diatom assemblages between restored and reference marshes. The line represents a significant relationship (linear regression P<0.05). The triangles represent sites in Virginia, and dots represent sites in North Carolina. ............................................ 67 Figure 3.5 Relationship between algal productivity ratio and age of restored marshes in summer 1998. Productivity ratio compared between algal GPP between restored and reference sites in surface sediments (A) and on macrophytes (B) ....................................................................................... 71 Figure 3.6 Conceptual model of algal structural and functional development during marsh restoration ............................................................................... 78 Figure 4.1 Map of Lower Peninsula of Michigan, showing the Upper Grand River and Maple River watersheds where sampling sites were located. .................. 90 Figure 4.2 Box-plot of environmental variables in the 22 wetlands ......................... 96 Figure 4.3 Algal biomass/production in artificial substrata of control, N-enriched, P- enriched, and NP-enriched treatments. They are represented by different attributes ................................................................................ 98 Figure 4.4 Algal cellular nutrient contents and molar N:P ratios in control, N-enriched, P- enriched, and NP-enriched treatments on artificial substrata .................... 99 Figure 4.5 Chl a: AFDM and Ch] azDM ratios in control, N-enriched, P-enriched, and NP- enriched treatments on artificial substrata ........................................ 100 Figure 4.6 Comparisons of algal N and P limitation in all 22 wetlands. Primary N or P limitation is defined as algal biomass of that nutrient treatment minus control treatment then weighted by control treatment. ................................. 101 Figure 4.6 The relationships between algal primary nutrient limitation and nutrient concentrations in the water column and the sediments. Primary nutrient limitation is defined as algal biomass of that nutrient treatment minus control treatment then weighted by control treatment ................................... 102 Figure 4.8 CCA biplots of algal assemblages and environmental variables. Diatom taxa on macrophytes, in sediments, and on artificial substrata in control treatments, and algal genera of phytoplankton were used for the CCA .................... 104 Figure 4.9 Box-plots of similarities of diatom assemblages between different habitats and treatments. E-epiphytes, S-sediment, C-control, N-N treatment, P-P treatment, xii NP- NP treatment. X-axis shows the comparison of two paired habitats. Y- axis shows the Percentage similarity between the two habitats .............. 105 Figure 5.1. Distribution of the 35 wetlands sampled in fall 1994 and summer 1995 in the lower peninsula of Michigan along a north-south and an east-west transect ................................................................................. l3 1 Figure 5.2. Box plot of environmental variables in 1994.1995 in the Michigan wetlands ............................................................................... 136 Figure 5.3 Relationship between average TP and % agricultural land use in sampled wetlands. Regression tree analysis was used to explore the relationship between TP and land use. A threshold effect occurred when agriculture occupied 57.9% land use around the wetland watershed ...................... 137 Figure 5.4 CCA biplots of sites and fall environmental variables from 35 palustrine wetlands in Michigan. (a) Dots show site scores in ordination space, (b) dots shows species scores in ordination space, and arrows indicate the strength and correlation between environmental factors and the ordination axes, The dotted circle shows sites with above 59% agricultural land cover. The dotted circle shows sites with above 57.9% agricultural land cover. Diatom species are indicated by eight-digits code (see Table 5.2 for complete taxon names)..140 Figure 5.5 Regression models of diatom inferred environmental variables against observed environmental variables. 5b shows the relationship between diatom inferred TP and agricultural land use ............................................. 143 Figure 5.6 Non-metric multidimensional scaling ordination diagrams of paleoecological diatom assemblages and current diatom assemblages. Reference (paleoecological diatom assemblages) and current (surface diatom assemblages) were indicated by different symbols. The centroids with 68.5% confident ellipse were plot around each region .................................. 144 Figure 5.7 Relationships of diatom assemblage similarity between current assemblages with nutrient concentration (TP) (a), and relationship of native taxa richness with nutrient concentration (TP) (b) in wetlands. S1 is the Bray-Curtis similarity of relative abundance between surface sediment and the centroid of presettlement samples ............................................................... 144 Figure 5.8. Relationship between van Dam trOphic index and palustrine wetland TP (a), and land use (b). Box—plot Shows paleoecological diatom trophic index from 14 palustrine wetlands to represent reference wetland condition. (a)Surface diatom trophic index (dots) respond linearly to TP concentration in wetlands (R2=0.407, P=0.001). The intercept of biological criteria and regression line represented nutrient criteria (stressor criteria=46 ug/L). (b). Dots represent change of trophic index with agricultural land use; a threshold effect was xiii observed when agriculture exceeds 56% total land use, with van Dam trophic index increasing from 2.46 to 4.85 ................................................ 146 xiv Chapter 1 Algae as indicators of biological integrity in restored and extant wetland ecosystems 1.1 Wetland bioassessment Wetlands have suffered serious impairment and loss since European settlement in North America. Approximately 53% of the wetlands in the United States have been lost since the late 17003 (Mitsch and Gosselink 1999), and a large proportion of wetlands have been identified as impaired by various pollutants such as nutrients and other environmental stressors. The US. Congress passed the Clean Water Act (CWA) in 1972 to preserve and improve the quality of the nation’s water resources, including wetlands. Over the past 30 years, wetland mitigation has been increasingly used to counteract loss of wetland habitat and the associated ecosystem functions. As a result, thousands of wetlands have been restored and created to fulfill the goal of the “no-net-loss” policy (Section 307 of the Water Resources Development Act). However, even for the elementary purpose of maintaining water quality, the success rates of wetland mitigation have been very low (USEPA 2000). To attain the stated goal of the CWA to “ restore and maintain physical, chemical, and biological integrity” of wetlands, a set of structural and functional criteria for extant and restored wetlands needs to be developed and evaluated. Traditional wetland assessment techniques evaluate physical and chemical conditions, especially chemical influences from point-source discharges. Increased human activities have contributed to nonpoint source runoff, introduced species, and other anthropogenic impacts that have imposed increasing damage to wetland ecosystems. New approaches based on structural and functional assessment of biological communities have been developed during the last two decades. Two of the most common approaches for assessing wetlands are the Hydrogeomorphic (HGM) approach, which focuses on functional assessment (Hauer & Smith 1998), and the Index of Biotic Integrity (IBI) approach of bioassessment (Danielson 2001). The HGM functional assessment approach is based on the hydorgeomorphic concepts for classifying wetlands and was specifically deve10ped for rapid wetland assessments in response to the requirements of “no overall net-loss” of wetland area and functions (Conservation Foundation, 1988). In contrast, the [BI approach of bioassessment was largely developed to meet water quality and biocriteria requirements and is based on the conventional Cowardin et a]. classification systems (Cowardin et a1. 1978). These two approaches both contrast and complement each other and both contribute to assessment of wetland structure and function (Stevenson and Hauer 2002). In the past 20 years, biological assessment has been developed to evaluate the integrity of aquatic systems. Biological integrity is a comprehensive reflection of physical and chemical factors and is defined as “the ability of an aquatic ecosystem to support and maintain a balanced, adaptive community of organisms having a species composition, diversity, and functional organization comparable to that of natural habitats within a region” (Karr and Dudley 1981). By evaluating the condition of one taxonomic community relative to physical and chemical conditions across a range of sites, biological assessment can identify stressors that impair ecological health and also diagnose the cause of environmental impairment. Bioassessment can also be used to evaluate wetland restoration and changes in ecosystem functions over time. Evaluation of wetland restoration and wetland impairment involves assessment of nutrient enrichment at multiple spatial and temporal scales and how they affect biological communities. Numerous studies have documented wetland structural and functional changes after restoration (Middleton 1999; Streever 1999). Most of these studies have focused on changes in biogeochemistry, macrophyte community structure, and invertebrate species composition. It is the general expectation that restored wetlands should increase in species diversity, nutrient retention, systems sustainability, and productivity over time. The direct impact of nutrient accrual and recycling during wetland development is a major regulator of wetland structure and function. 1.2 Algal indicators As pointed out by Landis and McLaughlin (2000), indicators have three basic purposes: to measure current status, to predict future changes in the system, and to predict progress toward the restoration of a system. The most common use of indicators is to establish the current status of a system. Algae are one of the most sensitive indicators for biological assessment of aquatic systems. The importance of algal assemblages and function has been summarized in a numbers of reviews (Dixit et al. 1992, Stevenson and Pan 1999, Stevenson 2001). Algal characteristics that contribute to their effectiveness in assessment, include great diversity, presence in almost all aquatic ecosystems, sensitivity to chemical and physical changes, and taxon-specific responses to environmental changes (Shubert 1984, Steneck & Dethier 1994, Stoermer 1990). Environmental stressors could lead to degradation of valued ecological attributes and thus the sensitive response of algal assemblages could be useful indicators of environmental degradation (Figure 1). Over the past several decades, evaluation of streams and lakes using algal attributes has proved successful (Lowe & Pan 1996). Although the biological communities and the underlying biogeochemical processes of wetlands are more complex than streams and lakes, the general approaches and methods for streams and lakes assessment can also be applied to wetland systems (McCormick and Stevenson 1998, Danielson 2001). Attempts to evaluate environmental gradients using algal attributes have been successful in Kentucky wetlands and the Everglades (Pan et al. 1996, McCormick 1998, Pan et a1. 2000) and have proved to be successful. 1.3 Wetland biogeochemistry and primary producers Wetland systems have complicated biogeochemical characteristics that vary both within and among different types of wetlands (Odum 1988, Mitsch and Gosselink 2000). Salt marshes differ from freshwater marshes because they receive chemicals from both oceanic and freshwater input. Anaerobic decomposition is predominated by sulfur reduction and fermentation in salt marshes. Large amounts of sulfur gases and C02, low amounts of methane, and unknown amounts of nitrous oxides are the main biogenic gases. Nitrogen is the major limiting factor for primary production in salt marsh ecosystems, while phosphorus is accumulated. Salt marshes have been shown to be both sources and sinks of nutrients, particularly nitrogen. The general model of salt marsh nutrient flux suggests that salt marshes act primarily as transformers of nutrients by importing dissolved oxidized inorganic forms and exporting dissolved and particulate reduced forms (ammonium, organic nitrogen, and phosphorous compounds). Salt marshes tend to be a net importer of nutrients during spring and summer and a net exporter in the autumn and winter (Odum 1988). The water and soil chemistry of freshwater marshes is dominated by a combination of factors associated with the underlying mineral soils, overlain with autochthonous inputs of organic matter from vegetation production (Mitsch and Gosselink 2000). The inflowing water has high amounts of dissolved materials and nutrients, derived from stream, groundwater, and surface water inflow. Organic matter in freshwater marshes can vary widely, but is generally higher compared to salt marshes (Mitsch and Gosselink 2000). Dissolved inorganic nitrogen and phosphorus vary seasonally from very low concentration in the summer to high concentration in the winter. Thus, nutrient limitation varies seasonally. Sulfur reduction is an important anaerobic decomposition process, but is of less significance in freshwater marshes than in salt marshes. Anaerobic decomposition is predominated by methanogenesis and fermentation in freshwater marshes. Large amounts of methane and C02, small amounts of sulfur gases, and unknown magnitude of nitrous oxides are the main biogenic gases. Most studies Show that freshwater wetlands are nutrient sinks and transformers. Nitrogen and phosphorus are generally limiting factors and are probably related to each other (Mitsch and Gosselink 2000). Both salt marshes and inland freshwater wetlands are among the most productive ecosystems in the world. Marsh grasses, periphyton, and phytoplankton compose the three major autotrophic units in salt marshes (Odum 1988). Periphyton production can be as high as or higher than that of the macrophyte community’s production. Vascular plant communities are always dominated by a few species, whereas algal species attain high species diversity and production. Freshwater marshes, on the contrary, have a much higher diversity of vascular plants but a relatively lower diversity of algae than salt marshes (Odum 1988). The plant productivity of inland marshes can be as high as salt marshes (Mitsch and Gosselink 2000). 1.4 Objectives of the dissertation The central goal of this dissertation was to apply algal biological assessment approaches that are broadly used in lakes and streams to evaluate ecological health of wetlands and to assess wetland restoration. Based on this objective, my research focused on how algal assemblages in both coastal salt marshes and inland freshwater wetlands respond to environmental changes and human disturbance. I conducted both field observations and experiments to delineate the pattern of algal responses along human disturbance gradients, especially nutrient gradients. I evaluated algal indicators for wetland nutrient enrichment and wetland restoration in both coastal salt marshes and freshwater wetlands. Both simple statistics and multivariate approaches were used to identify, develop, evaluate, and integrate a suite of algal indicators of ecological condition. Assessment of ecological conditions in least disturbed conditions and stressor- response relationships were used for development of biological criteria and corresponding stressor criteria for environmental monitoring and management (Fig. 1). This dissertation includes the following major components as separate chapters: Part 1 (chapter 2): Freshwater wetland restoration was evaluated by: examining the difference of environmental variables and algal assemblages between restored and extant wetlands; identifying what environmental variables causing the difference in algal assemblages between restored and extant wetlands; and comparing algal assemblages among different habitats and how similarities among habitats change with change of environmental conditions. Part H (chapter 3): The success of salt marsh restoration was assessed by: using the reference approach and pairing a restored wetland with a nearby extant wetland to minimize regional difference among wetlands; evaluating a suite of algal attributes and finding the best indicators for salt marsh restoration; and evaluating the structural and functional development of wetland ecosystem development over time. Part III (chapter 4): Nutrient enrichment experiments were conducted in a set of restored freshwater wetlands to determine: nutrient limitation, how algae respond to nutrient enrichment in artificial habitats, and how these responses are different or similar in experimental examination compared to natural environmental gradients, and thus establish indicators based on cause-effect relationship for wetland bioassessment. Part IV (chapter 5): A set of wetlands across Michigan were used to: determine stresses and stressor gradients through diatom ecological response in wetland ecosystems; investigate how these stressors are related to human disturbance gradients based on diatom ecological responses; and set biological and stressor criteria to protect restored wetlands. The ultimate goal of this research was to provide new tools for diagnosing and quantifying risks from pollution that threaten wetland ecosystems. The results of my research should support ecologically sensitive decision-making in land planning and wetland restoration. Specifically, I was able to define ecological thresholds based on stressor-response relationships and reference approaches. These results can be applied for watershed management by providing targets for decreasing anthropogenic pollution and thereby increasing the rate of successful restoration projects. I also developed easy and rapid biotic and abiotic indicators that serve as qualitative and quantitative criteria to assess water quality and ecosystem health. These studies help to identify ecosystems at risk due to excess nutrients in the environment. High 0, 5 Goal: 9.9. Number of Native spp. C o E s: 8 0.0.0.?0§0:000 ......OOOOOOOOOOOOOOO00...... : \ Response Criterion: 13’ : \ e.g. assemblage ‘>" E similarity to reference _ ' \ 8 3 x '55 3 ‘— Indicator 2 : Stressor Criterion \ O o \ f”) yset to protect goal Low I ° \ Low (Fief) High Stressor or Human Disturbance Gradient Figure 1.1 Stressors-response relationship between human disturbance and response of ecological valuable attributes 1.5 Reference Cowardin, L.M., V. Carter, F.C. Golet, E.T. LaRoe. 1979. Classification of wetlands and deep water habitats of the United States. US. Fish and Wildlife Service, Washington, DC. Conservation Foundation. 1988. Protecting America’s wetlands: An action agenda. Final Report of the National Wetlands Policy Forum. The Conservation Foundation, Washington, DC. Danielson, T. J. 2001. Methods for evaluating wetland condition: Introduction to wetland biological assessment. EPA 822-R-01-007a. US. Environmental Protection Agency, Office of Water, Washington, DC. Dixit, 8.8., J. P. Smol, J. C. Kingston, and D. F. Charles. 1992. Diatoms: powerful indicators of environmental change. Environmental Science and Technology. 25: 22- 33. Hauer, F. R., and RD. Smith. 1998. The hydrogeomorphic approach to functional assessment of riparian wetlands: evaluating impacts and mitigation on river floodplains in the USA. Freshwater Biology 40: 517-530. Karr, J. R., and D. R. Dudley. 1981. Ecological perspective on water quality goals. Environmental Management 4:55-68. Landis, W. G., and J. F. McLaughlin. 2000. Design criteria and derivation of indicators for ecological position, direction, and risk. Environ. Toxicol. Chem. 19: 1059-65. Lowe, R. L. and Y. Pan. 1996. Benthic algal communities as biological monitors. In R. J. Stevenson, M. L. Bothwell and R. L. Lowe [eds] Algal ecology: Freshwater benthic ecosystems. California Academic Press, San Diego, pp 705-739. McCormick, P. V., and R. J. Stevenson. 1998. Periphyton as a tool for ecological assessment and management in the Florida Everglades. J. Phycol. 34: 726-33. Middleton, B. 1999. Wetland restoration: Flood pulsing and disturbance dynamics. John Wiley & Sons, Inc. New York. pp6-7. Mitsch, W. J ., and J. G. Gosselink. 2000. Wetlands. John Wiley & Sons, Inc. New York. Odum, W. E.1988. Comparative ecology of tidal freshwater and salt marshes. Annual review of ecological system, 19: 147-176. Shubert, L. E. 1984. Algae as Ecological indicators. Academic Press, INC. Stevenson, R. J ., and Y. Pan. 1999. Assessing environmental conditions in rivers and streams with diatoms. In The diatoms: applications for the environmental and earth sciences. Edited by .E. F. Stoermer and J. P. Smol. Cambridge University Press, Cambridge, UK. pp 11-40. Stevenson, R. J ., R. P. Sweets, Pan Y., and R. E. Schultz. 1999. Diatom community patterns in wetlands of Michigan (USA) and their use as indicators of ecological conditions. Pages 517-527 in A. J. McComb, and J. A. Davis (editors). Proceedings of INTECOL’S Vth International Wetland Conference. Gleneagles Press, Adelaaide, Australia. Stevenson, R. J ., McCormick, P. V., Frydenborg, R. 2001. Methods for evaluating wetland condition: using algae to assess environmental conditions in wetlands. EPA 843-B-00—002k. US. Environmental Protection Agency; Office of Water, Washington, DC. Stevenson, R. J. and F. R. Hauer. 2002. Integrating Hydrogeomorphic and Index of Biotic Integrity approaches for environmental assessment of wetlands. J. N. Am. Benthol. Soc. Stoermer, E. F. 1990. Quantitative characteristics of phytoplankton assemblages. In Shubert L. E. [ed] Algae as ecological indicators. Academic Press, Inc. New York. Streever, W. 1999. An international perspective on wetland rehabilitation. Kluwer Academic Publishers. Boston. USEPA. 1992. Framework for ecological risk assessment. EPA 630/R-92/001. US. Environmental Protection Agency. Washington, DC. US. EPA. 2002. Methods for evaluating wetland condition, #11 Using algae to assess environmental conditions in wetlands. United States Environmental Protection Agency, Washington, DC, USA. EPA-822-R-02-021. 10 CHAPTER 2 ALGAL ASSEMBLAGES IN MULTIPLE HABITA TS 0F RES TORED AND EXTANT WETLANDS 2.1 Abstract The functional and structural attributes of algal assemblages were studied in 25 restored and 20 extant depressional, palustrine wetlands in southern central Michigan. Environmental conditions and algal assemblages were compared between restored and extant wetlands and among habitats within wetlands. Restored marshes generally had lower shading by macrophytes, nutrient concentrations, and organic matter in sediments. Relative biovolume of non-diatom algae was significantly different among habitats in restored wetlands, but did not differ between macrophytes and sediments in extant wetlands. Species composition of diatom assemblages was not significantly different between epiphytes and sediment in both restored and extant wetlands. The discovered differences in non-diatom algae could not be attributed to any environmental variables measured; however, diatom assemblage differences between habitats increased with light irradiance. Differences in sediment diatom assemblages were observed between restored and extant wetlands and were related to differences in nutrients, pH, and canopy cover. Differences were also observed between epiphytic diatom assemblages in restored and extant wetlands and they were related to light and dissolved oxygen. Key index words: diatoms; non-diatom algae; epiphyton; epipelon; phytoplankton; habitats; restored wetlands; extant wetlands; periphyton Abbreviations: DIN, dissolved inorganic nitrogen; DO, dissolved oxygen; LEC, light extinction coefficients; MANOVA, multivariate analysis of variances; MRPP, multi- ll response permutation procedure; NMDS, nonmetric multi-dimensional scaling; 0MP, organic matter proportion; PCA, principle component analyses; SRP, soluble reactive phosphorus; TN, total nitrogen; TP, total phosphorus 12 2.2 INTRODUCTION Wetland restoration has been increasingly used to mitigate loss and anthropogenic impact of wetland systems (McDonnell and Pickett 1990, Holland et a1. 1995, Detenbeck et al. 1999, Zedler 2000). Beginning in the early 1990's, a large effort was initiated in Michigan to compensate wetland losses by establishing wetland restoration programs (MDEQ 2001a). Hundreds of wetland restoration projects have been carried out since then. Most of the wetland restoration projects involve simple techniques, such as plugging agricultural ditches or breaking field tiles to restore the hydrology of an area, and then restore the biological community. Restored wetlands support aquatic and terrestrial habitats and sustain many plant, invertebrate, and wildlife species (NRC 1995, USEPA 2000). Ecological development after wetland restoration has been studied extensively (Whigham et a1. 1990, Patten 1994, Middle 1999, Keddy 1999), and many environmental variables change during transition from one successional state to another (Gosz 1992). Nutrient accumulation is one of the most noticeable changes in biogeochemical process affecting the biological community in wetlands. In human impacted landscapes, nutrient enrichment from human activities affects the wetland restoration process, and its detrimental effects on wetland systems have been frequently reported (Koerselman and Beltman 1988, Blair 1996, Magee et a1. 1999). Nutrient accumulation and storage from agricultural and urban runoff affect wetland productivity, species diversity, water quality, as well as the restoration rate of wetlands (Rader and Richardson 1992, Craft and Casey 2000). 13 Algae are one of the most important groups of primary producers in wetlands, and they play crucial roles in wetland biogeochemical cycles (Vymazal 1995). Algae, especially diatoms, are effective indicators of ecological conditions (Zedler 1993, Adamus 1996, Cottingham and Carpenter 1998, Detenbeck 2001, Stevenson et al. 2001) and have been used as indicators of ecological conditions in wetlands (Pan and Stevenson 1996, Stevenson et al. 1999, Pan et al. 2000, Stevenson 2001). Algae are also sensitive to nutrient concentration change (Tilman 1982). Soil organic C, N and P accumulation during wetland restoration can affect the biomass and species composition of the algal community (McCormick et al. 1996, Pan et al. 2000). Therefore, algal species assemblages could be indicators of restoration success because algal species composition and diversity would differ in low- and high-nutrient wetlands (Mayer and Galatowitsch 1999). Habitat differences, along with a number of other environmental factors, contribute to spatial and temporal variability of algal assemblages in aquatic ecosystems (Leland 1995, Stevenson 1996, Stevenson 1997). The extent and conditions in four major algal habitats, phytoplankton, epiphyton, epipelon, and metaphyton, vary temporally at different stages in wetland cycles (Goldsborough and Robinson 1996). A number of studies have reported that different habitats support different algal assemblages in aquatic systems (Stevenson and Hashim 1989, Pringle 1990, Burkholder 1996, McCormick et al. 1998, I-Iillebrand and Kahlert 2001, Lim et al. 2001). Thus the similarity of algal assemblages between different habitats should increase with the increase of wetland age and increasing similarity of habitat characteristics. 14 Environmental factors such as, substratatum, substratratum stability, and nutrient availability have been suspected as major determinants of species composition differences among habitats (Stevenson 1996, Burkholder 1996). Changes in environmental conditions of algal habitats during wetland development flowing creation or restoration may regulate heterogeneity in species composition among habitatis. The lack of substratum should distinguish the plankton from the two benthic habitats of all stages of wetland development, but heterogeneity between phytoplankton and benthic habitats could decrease if more benthic algae were suspended in the water during either early or late stages of development. Algae on plants and sediments get some nutrients direcetly from the substratum. Nutrient supply from sediments directly to epipelon may be available earlier during wetland development than the indirect supply from sediments, through plants, to epiphyton. Differences between epiphyton and epipelon probably decrease with wetland development as nutrient availabitliyt increases and differences between habitats decrease. To evaluate the success of wetland restoration and investigate the impact of human disturbance to both natural and restored wetland systems, I selected 45 (25 restored, 20 extant) freshwater depressional wetlands in the Maple and Upper Grand watersheds in southern Michigan. The environmental characteristics and algal assemblages of phytoplankton, epiphyton, and epipelon on different habitats in these wetlands were examined. I hypothesized that: environmental variables would differ between restored and extant wetlands; algal assemblages would differ among habitats and between restored and extant wetlands; differences in algal assemblages among habitats and between restored and extant wetlands would decrease with increasing age of restored wetlands; 15 and that decrease would be due to increasing similarity of environmental characteristics. I also expected to identify environmental factors that determine the distribution of algal assemblages and biodiversity among habitats and between restored and extant wetlands. 2.3 METHODS Site description. Forty-five (25 restored, 20 extant wetlands) depressional, palustrine wetlands (Cowardin et al. 1979) were selected in the Maple and Upper Grand River watersheds of Southern Michigan for this study (Figure 2.1). All restored wetlands were reestablished within the past 15 years as part of wetland mitigation program. All wetland sites were located in landscapes dominated by agricultural and urban land. The sampled wetlands were mostly Typha or Nuphar dominated marshes, but also included several swamps. Watershed land use data were obtained from the Michigan DNR spatial data library to determine land use around these wetlands. Wetland boundaries for each site were determined based on 100 100 200 Miles shifts from wetland to upland Figure 2.1. Wetland sampling locations in southern vegetation and changes in slope Michigan. 45 sites were selected in Maple and Upper Grand River watersheds. between the wetlands and the 16 adjacent upland. Duckweed covered at least some proportion of all wetlands during the sampling period. 2.3.] Sampling. The selected wetlands were sampled and assessed during mornings of July 2000. Water temperature and dissolved oxygen (DO) were measured with a YSI® DO meter (YSI Inc., Yellow Springs, Ohio, USA) at the beginning and at the end of the sampling period, allowing at least one hour between measurements. Wetland net production was calculated based on DO change over one hour period times the estimated average depth of the wetland. Conductivity and pH were measured using a YSI® model 33 SCT meter with a conductivity and pH electrode (Denver Instrument Company). Canopy cover by emergent vegetation was estimated with a spherical canopy densiometer held at water level and by averaging canopy cover from 4 directions. Duckweed coverage was estimated based on the proportional coverage of the wetland surface. Irradiance at the water surface (10) and 30 cm below the water surface ([2) were recorded with a LICOR light meter. Light extinction coefficients (LEC) were calculated as ln(Io)- ln(Iz)/0.3 m (Wetzel 2001). Water samples from open water area in each wetland were collected in 2 125-ml acid-washed polyethylene bottles, stored in ice, transported to the laboratory, and frozen until chemical analysis. Phytoplankton samples from 10 cm below the water surface were collected in 2 l-L polyethylene bottles for chlorophyll a and algal cell counts. Sediment samples were collected randomly from eight different locations in each wetland using hard plastic tubes (internal diameter = 3.8 cm, length = 20 cm). Only the top 1 cm of sediment was collected. All 8 sediment samples from a wetland were composited in a Whirl-pak® bag. Epiphytic algae were collected by cutting 4-5 macrophyte stems, brushing epiphytes from each macrophyte surface, and putting them 17 in a Whirl-pak® bag. Metaphyton were collected for qualitative identification. Algae were preserved with M3 (APHA 1998) after subsampling for chl a analysis. 2.3.2 Sample analysis. One 250-m] water sample from each wetland was analyzed for total nitrogen (TN) and total phosphorus (TP). Another bottle was filtered through Coleman® glass fiber filters (0.45um diameter) and the filtrate was analyzed for nitrate- nitrite (NO3'-N + NOz'-N) and soluble reactive phosphorus (SRP) using a Spectronic® GenesysTM 2 Spectrophotometer. TP and SRP were measured using the ascorbic acid method, while TN and nitrate and nitrite (NOx-N) were measured using cadium reduction methods (APHA 1998). Ammonia (NH4-N) was analyzed using a Wuick-Chem 8000 autoanalyzer at University of Michigan Biological Station (UMBS). Dissolved inorganic nitrogen (DIN) was calculated as the sum of ammonia and nitrate-nitrite concentrations. Sediment samples were rinsed with 200 ml of deionized water and homogenized using a Biospec® homogenizer. Appropriate dilution was made for sediment samples to measure sediment TN and TP concentration following the same methods as the water samples. Chl a in water sample was assayed by extracting in 90% buffered acetone overnight at 4°C, reading absorbance on a Spectronic® GenesysTM 2 Spectrophotometer, and calculating pheophytin-corrected chl a concentration (APHA 1998). Dry mass (DM) and ash free dry mass (AFDM) of periphyton were determined according to standard methods (APHA 1998). Organic matter proportion in sediment (OMP) was the AFDM/DM ratio of surface sediments. Phytoplankton, epiphytic algae, and sediment algae were diluted or condensed as necessary before counting. Algal densities and non-diatom species composition were determined using a Leica® microscope and a Palmer-Maloney (0.1 ml) counting chamber 18 under 400X and by counting 300 natural units. Algal taxa were identified to the lowest possible taxonomic level. Biovolume was estimated for at least 15 cells of each taxon by assigning them a geometric shape and measuring appropriate cell dimensions (Hillebrand et al. 1999, Charles et al. 2002). Diatom taxa and relative abundance were determined from permanent Naphax mounts of acid-cleaned diatoms (Stosch and Reimann 1970). At least 500 diatom valves were counted. Alpha diversity (species richness per locality), gamma diversity (total species diversity within landscape, all wetland sampled), and beta diversity (variation in species composition between 2 communities or habitats) were calculated for non-diatom algae and diatoms (Beilman 2001). 2.3.3 Statistics. To compare environmental variables between restored and extant wetlands, both individual variables and scores of principle component analyses (PCA) were used. Environmental variables were log-transformed to evenly distribute the variance before analysis, if necessary. The Student t-test was used to compare individual variables between extant and restored wetlands. PCA summarized major environmental patterns in this region. Multiple analysis of variance (MANOVA) was also used the test overall differences between restored and extant wetlands. The scores of the first three PCA axes were used as composite variables for MANOVA. All statistics were performed using SYSTAT® v. 10. Attributes of non-diatom algae and diatoms were analyzed separately. The relative biovolume of non-diatom algae and relative abundance of diatom species were calculated and used for comparisons. Only diatom species with a relative abundance 2 1% in a minimum of three sites or 2 5% in at least one site were included in the analysis. 19 Nonparametric multidimensional scaling (NMDS), a multivariate ordination technique (Clarke 1993), was used to summarize and illustrate patterns of non-diatom and diatom distribution among different habitats and between restored and extant wetlands. Multi-response permutation procedure (MRPP), a non-parametric procedure for testing the hypothesis of no difference between two or more groups (Biondini et al. 1985), was used to assess differences in taxonomic composition of non-diatom and diatom assemblages among different habitats and between restored and extant wetlands. A p-value was provided by the MRPP to characterize the significance of between group differences. An A value (0-1) characterizes within-group homogeneity compared to the random expectation. When all items are identical within groups, then A = 1; if heterogeneity within groups equals the expectation by chance, then A = 0. Both MRPP and NMDS were performed using PC-ORD ® v.4.0 (McCune and Mefford 1997). To relate differences in taxa composition to environmental variables, Bray-Curtis similarity matrics (srrrilarity in relative abundances or biovolumes) were calculated for species composition of non-diatom and diatom assemblages among different habitats and between restored and extant wetlands. Stepwise automatic linear regressions were used to explore the relationship between non-diatom or diatom assemblages and environmental variables. This was done by using similarity of non-diatom or diatom assemblages between habitats as separate dependent variables and a set of 19 uncorrelated variables as independent variables. Using these same procedures, similarities of species composition between restored and extant wetlands were calculated and expressed as a function of differences in environmental variables between wetlands. The dependent variables were the average Bray-Curtis similarities in species composition of a restored wetland with all 20 the extant wetlands. The Euclidean distance matrix for each environmental variable among different wetlands was calculated to indicate environmental differences between wetlands and the average Euclidean distances of individual environmental variables between a restored wetland with all extant wetlands were used as independent variables. Those variables that had the highest correlations with other variables were screened out, so only the most uncorrelated environmental variables were included for the regression analyses. For example, the distance of conductivity between restored and extant wetlands was positively correlated with the distance of sediment TP (1:0. 424, P=0.035), negatively correlated with pH distance (r=0.430, p=0.032), and related to several other variables. Therefore, it was excluded from the model. All these analyses were performed using Systat® v10. 2.4 RESULTS 2.4.1 Environmental variables in restored and extant wetlands. Land cover in the region was dominated by agriculture (78%) with smaller portions of urban (9%), forest (6%), and other land types. Most of the restored wetlands were Typha-dominated marshes. In contrast, extant marshes in this area were seldom dominated by one macrophyte Species. Phragmites, Typha, Nuphar, Ceratophyllum demersum, Myriophyllum, and Scirpus were co-dominant in many of the extant marshes. Three PCA scores accounted for 42% of environmental variance in the data set (Table 2.1). The first axis was positively correlated with major nutrients, such as TP (r=0.621), ammonium (r=0.784), and DIN (r=0.675) in water column, as well as light supplies (duckweed cover and light extinction coefficient). The second axis was negatively 21 correlated with N, P content in sediments and shading by vegetation, and positively correlated with conductivity. The third axis was positively correlated with SRP and DO. Overall, the scores of the first 3 axes were significantly different between restored and extant wetlands (MANOVA, Wilks’s lamba F3, 42: 5.938, P=0.002). Univariate F test indicated that PCA axis 1 score was significantly different between restored and extant wetlands (P=0.012), while the scores of PCA axis 2 and axis 3 were not different (P>0.1). Many variables did not differ between the two types of wetlands (P>0.1). Water temperature varied from 18.1 to 322°C among wetlands during the sampling period. Conductivity was highly variable (from 149 to 1003, 499i34 pmhos/cm). pH varied greatly among sites (from 6.8 to 9.6, 7.92i0.9 (Average :tS.E.)), but was not significantly different between restored and extant wetlands (P=0.110). D0 in the morning could be as high as 11 mg/L and as low as 0.7 mg/L in some of the wetlands (4.567i0.381mg/L). Wetland net production based on DO change ranged from -1.0 to 3.17 gom'z-h'l representing heterotrophic to autotrophic wetland types. Most restored wetlands had relatively low shading by vegetation (26.2i3.8%, mean i standard error). Extant wetlands had higher shading than restored sites (t-test, P=0.031) (Figure 2.2). Duckweed and other floating plants appeared in at least 2/3 of the sampled sites with an average coverage of 38.2% (i31.5%) of water surface during sampling period. Duckweed coverage was much higher in extant wetlands than restored ones (P=0.003). LEC at 0.3m depth in open water (if available) of restored wetlands (1.207 i012] umol m'3s‘1) was much lower than in extant wetlands (3.173i0.777 umol m'3s") (P=0.01) indicating that extant wetlands had much lower light avalaibility in the water column and at the bottom. 22 5m.0 NN.0- w0.0- $0- 20- m0.0 3.0- m~.0- v0.0 mad- 2.0- mmd 0N.0 52.0 Nv.0- 02.0 05.0 0m.0- 2N0- v0.0 mom vm.0- N00 00.0- 05.0- 0N0 2.0 2.0 2.0 2.0 2.0 2.0 20.0 $0 2&0- wvd w0.0- 2.0 02.0- 52.0 20.0- NOAH NN.0- No.0 w0.0 3.0 00.0 00.0 mmd- mud- 3.0 05.0 00.0 N00 and 00.0- 52.0- M500 mm.0- .500 00.0 3.0 2052 00.2 50.0 v0.0 00.0 N00 N00 N00 N00 N00 30- 30- no.0 02.0- 2.0 ~0.0- 02.0 50.0- 50.0 2.0 NN.0 ow< 00.2 NN.0- 20.0 00.0 No.0 3.0 00.0 00.0- 00.0- 2.0- 2.0- m0.0 50.0 w0.0 vmd 2N0. 3H0 02.0 00.0- $2 .Uom . 00.2 05.0 2.0 v0.0- 000- 00.0 2.0 2.0 02.0 V00 v0.0- 020- 52.0- 00.0- 50.0- 02.0 N0.0- 0N0 0.00m . 00.2 0N0 n0.0- 00.0 ~00- 00.0 m0.0- 20.0 50.0- 2.0- 2.0 0N0- 00.0- w0.0- 02.0 m0.0- wad 2.00m . 00.2 0V0 00.0- 2.0- 3.0 and N00 mud #00 20.0- 52.0- 00.0- 2.0- 3.0 and 52.0 #20 . 00.2 2.0- 2.0- 0N0 0~.0 2.0 mmd 2.0 00.0 :0. mad N00 wmd wm.0 80 «EU . 00.2 0nd 3.0- 00.0- 200 3.0- 2.0 2.0- 20.0 00.0 20.0- 00.0 3.0- 2.0- AFHZH . 00.2 mud- 0N.0 30 05:0- 2.0 N00- 020 00.0 00.0 02.0- 02.0- 0N.0- Amy—minn— . 00.2 05-0 mmd 0V0 5m.0 NN.0- 2.0- 20.0- 02.0 2.0 3.0 02.0 flaw 00.2 00.0 mmd 3.0 #00 50.0- 02.0- 200- 5m.0 20.0 3.0 .22 . 00.2 mN.0 vmd m0.0 5.0- M00- 0N0- wmd mvd 0V0 7:9 . 00.2 00.0 52.0 m0.0- 50.0 3.0- 5N.0 0Y0 2.0 at. . 00.— 52.0- 20.0- 00.0- 20.0- 50.0 00.0 00.0 2.5 . 00.2 3.0- 000 2.0 20.0 00.0- N00 20 . 00.— 2.0 02.0- 3.0- 2.0- 3.0- .950 . 00.2 2.0 50.0 52.0- 2.0 .095 00.2 who. 5m.0- 3.0- CD . 00. 2 00.0 02 .0 083x059 00.2 no.0 Um: . 8; macaw ow< muz dam 0 .com 2 dew E0 a EU AFHZP AEWHZHQ mam M22 ED “2. 2.5 In .950 .005 On— coospo—osn— Um: weaver—m .Hgocmcmmm we? coca—oboe 05 me one: 05 em 888.0 00 8&8 a 0m: 833:2, 28:08:83.8 0.3 own 5053 5022.50 .20.0vm .300 can? ”m0.0vm xaEm-ccsa .0m.0A e 220508 scum—oboe a :25, £28003 we 05 E 833:; 3:82.835 me $532 20% $3028 “5.59:8 3905.5 0.8 35.0.38 mpg—oat? 25:08:83.5 no 5.52: "802280 .2.N 033. 23 A100. , .- lm I 40 I I 5:3: : .- ° " r - 30- J §°40* ”£75 0 20" ' BSOF .1520" " >~ 1‘- 8- :3 L , g 025 T - 10- g - U 0 0 r l 0 1 1 O 1 1100 T I 10 I I 40 I r E - - t §9OOI I 9~ 3 ~53 t v - 3700' _‘ m 530 I T .. ' ‘ q.- 'a #2 3:500" ‘ E - « 820' - 8300- - 7' “E3 '2 ~ - 8100 ‘ ' 6 . 4 10 - . 12 l I 6 F i m: I i ‘ 10~ - is- - 60E g 8r- - 34L -§30- I 1 DD v 5 6* l 53- -§ 8 4+ 1 82- ~01” 1 . <1 2. J. .. . 1 0 1 O I l I R E R E R E Figure 2.2. Comparisons of physical and chemical characteristics between restored and reference wetlands. Nutrient concentrations in these two types of wetlands were different (Figure 2.3). TP (1.858i0.314 mg/L) and SRP (0.406i0.102 mg/L) were high in most of the wetlands, and had significantly higher concentrations in extant wetlands than in restored wetlands (P=0.007 and 0.027 respectively). TN (0.779zt0.069 mg/L overall) in extant wetlands was not different from restored wetlands (P>0.1); the major form of inorganic N was NHI, and N03 ‘was almost depleted in most of the wetlands (<0.010mg/L). DIN (005410.007) was significantly higher in extant wetlands (P=0.023) than restored wetlands. N:P and 24 DIN: SRP ratios were generally lower than 17. N: P ratio was higher in restored wetlands than in extant wetlands (P=0.002) while DIN:SRP ratio was not significantly different between these two types of wetlands (P=0.210). Most of wetlands had mineral soils with an average 0MP 17.4 (i2.3)%, which was higher in extant wetlands than in restored sites (P=0.035). TN (3.911i1.106 mg/L) and TP content (0.757i0. 186 mg/g) in soil surface was higher in extant wetlands than in restored wetlands (P=0.025 and 0.058 respectively), but N: P ratio (3.78:1.86) in sediment were not different (P>0.128). Many of the environmental variables related to trophic condition correlated with each other (Table 2.1). LEC at 30 cm depth positively correlated with duckweed cover, chl a, DIN, ammonia, TP concentration in the water colurrm, and sediment organic matter, and negatively correlated with D0. D0 in the water column negatively correlated with duckweed cover, LEC, and ammonia concentration. Most nutrient factors were also correlated with each other. TN in water column positively correlated with SRP, DIN, NH], and sediment TN. TP in water column positively correlated with SRP, ammonia, chl a. Sediment TN and TP content were positively correlated with each other. LEC, Duckweeds, SRP, and TP positively correlated with age of wetlands. Age of restored marshes was not a significant factor in differentiating the first 3 PCA axes. Environmental conditions between marshes less than 5-year-old and 5-10 year old were not significantly different (MANOVA, Wilks’s lamba F3,25=O.833, P=O.48) based on the three PCA axes. None of the environmental variables correlated with age of wetlands. 25 p—s C9. N:DM(mg/g) ‘2 h C? ———'-1 (mg/L) g L l '6 E { s_L _. T RE SRP y—s O V‘Vv V—T‘IVVVI'" *'_' m". y—a 9.. N:P__in sediment ‘2. t—b C? :’ l .r i a .F i. a :1: =' 1 E a i i F . E 1. E 3 F 1. R E Figure 2.3. Wetland nutrient concentration and comparison between restored and extant wetlands. 2.4.2 Non-diatom algal assemblage and diatom assemblages in restored and extant wetlands Of the 311 epiphytic algal taxa identified, diatoms had the highest taxa richness (157), followed by Chlorophyta (99 taxa) and cyanobacteria (34). Other groups (Euglenophyta, Dinophyta, and Chrysophyta) appeared but with less abundance. Chlorophyta were the most abundant, with an average of 48.2% (i6.9% SE) of the total cell density and 46.1% i5.0% of total biovolume. Stigeoclonium, Mougeotia, and 26 Oedogonium were the most abundant genera. The next abundant group in epiphytes was diatoms with 30.3% (i4.5%) of total density and 26.8% (i 4.3%) of total biovolume. Navicula minimum, Cocconeis placentula, Achnanthidium minutissimum, and Navicula cryptocephala were the most dominant diatom taxa. Cyanobacteria, Euglenophyta, Dinophyta, and Chrysophyta appeared but were low in density in epiphytic samples. Total number of algal taxa (gamma diversity) and average taxa richness (alpha diversity) were higher in restored wetlands (250 and 30.0 respectively) than in extant wetlands (189 and 25.8, respectively), mostly because of higher diatom gamma diversity in restored wetlands (Table 2.2). Beta diversity of diatoms was also higher in restored wetlands than in extant wetlands. Of the 235 epipelon taxa identified, diatoms had both the highest taxa richness (199) and abundance with an average of 54.4 % (i6.6% SE) of total cell density and 48.8% (:1: 6.5%) of total biovolume. Navicula minima, A. minutissimum, N. cryptocephala, and Nitzschia palea dominated diatom communities. Chlorophyta was the next dominant group with 19 taxa and 22.3 (i4.3%) of total density and 12.9% (i4.1%) of total biovolume because of dominance of Scenedesmus in some of the wetlands. Only 8 cyanobacteria taxa were found, but they remained in low density (13i2.3%) and biovolume (16.4i4.7%). Other groups (Euglenophyta, Dinophyta and Chrysophyta) appeared but with relatively low abundance. Although the alpha diversity of non-diatom algae and diatoms were pretty similar in restored (3.2 and 28.2 respectively) and extant wetlands (2.9 and 29.7 respectively), the gamma diversity of diatom taxa in restored wetlands (165) was much higher than in extant wetlands (129) and caused higher beta diversity in restored wetlands (Table 2.2). 27 Phytoplankton biomass was low in most sampled wetlands (0.030i0.009 mg chl a/L) and was not significantly different between restored and extant wetlands (P=0.113). The direct correlations of phytoplankton biomass with major nutrient gradients were weak (r=0.368 with TP, r=0. 126 with TN), but were significantly related to OMP in sediments (r=0.499, P=0.025). Phytoplankton taxa richness was relatively low compared to other habitats. A total of 153 non-diatom taxa were found in phytoplankton samples, but diatom density was very low in most wetlands. Most of the phytoplankton were Chlorophyta (42.3%) and cyanobacteria (33.1%) on the basis of cell density, only 15.2(:|:3.3%) of total algal abundance was diatom. The average relative biovolume was dominated by dinoflagellates (37.9i 4.7%), Englenophyta (26.5i3.7%), Chrysophyta (10.5i4.4%) and green algae (9.0i2.9%). Cyanobacteria (3.5i1.6%) and diatom biovolume (2.3:l:O.7%) were very low. Restored wetlands had higher alpha, beta, and gamma diversity than extant wetlands (Table 2.2). Planktonic diatoms were not used for quantitative analysis because of their extremely low density in half of the wetland sites. Metaphyton appeared in the majority of the wetlands and completely covered the basin of at least M: of wetlands. The most commonly found macroalgae were Stigeoclonium, Oedogonium, and Mougeotia. Because of the difficulty of quantifying their biomass, metaphyton were not used for further analysis. Non-diatom and diatom diversity among different habitats were quite different. Alpha, beta, and gamma diversity of non-diatom epiphyton and phytoplankton were 28 m6 M2: mmm :m WEN—m mgflfiwm 3w? 0.2 5Q 0.2 m2 mm #2 mmdfifiw Ndfiod odfld 8036-52 86 =< fie >6 9.: 52 @303 :Hmdm 89me wé ms 52 am: vdficdm fimfiwwm 8w? v; d ed Nd co mm aw modfios QNHQN v. 356 8036-52 mg» Nam 92 com Wmfihdm @392 SEED Efixm 3 2 § own 3%.: 33.8 8% 5.. ed“ mm 0.2 _ _ _ on w: ondfimd mdflmfi. m. 35w 8036-52 m6 fie mg mg wAHNwm 030.3 9:8me 62.38% 3:96:53 E0863 23:33 2:03:33 EoEEom oath—35m oEoE—EE EoEEom oflnmmdm Sum 2555 69m + owfiozwv 23¢. aficm—HOB uguxu fig GDHOHMOH GDQBHB Cam “g..—6:03 N 551$ 38an 5953 .0936 Son 98 56.33% Acmaowucws «8:30 A86 com 38on mo “38:75 flaccid «an? :32 .N.N 0331 29 higher than epipelon in all wetlands. Sediment diatoms had higher alpha diversity, higher gamma diversity, but lower beta diversity than epiphytic diatoms (Table 2.2). 2.4.3 Difference of taxonomic composition among habitats and between restored and extant wetlands Differences in algal assemblages were commonly observed among habitats within wetlands. Most comparisons of non-diatom algae among habitats were different in both restored and extant wetlands, except that the difference between epiphyton and epipelon assemblages was not significantly different in extant wetlands (P=0.113) (Table 2.3, Figure 2.4). Similarly, epiphytic diatom assemblages and sediment diatom assemblages were not significantly different in restored or extant wetlands (Table 2.3). The sample ellipse of epiphytic diatom assemblages (Figure 2.4) was larger than that of sediment diatoms, indicating a larger beta diversity for epiphytic diatom assemblages than sediment diatoms. Differences in species composition of algal assemblages between restored and extant wetlands revealed habitat specific patterns (Table 2.4, Fig. 2.5). Both epiphytic non- diatom algae and diatom assemblages in restored wetlands were significantly different from those in extant wetlands (Table 2.3, Fig. 2.5). However, sediment algae and phytoplankton were not different between these two types of wetlands. 30 Restored wetlands Extant wetlands 1.0 it Q 0.5 05 U) D L 2 0.0 0.0 - 2 -0.5 ' _10 -10 l .10 r 1 -1.0 -0.5 0.0 0.5 1.0 -l.0 -0.5 0.0 0.5 1.0 1.5 u 2.0 r u '0 1.0 r a 0.0 " ‘ ‘ ’1-0 h o . Epiphytes . ° A Phytoplankton '10 " - if Sediments ‘21) i 1 1 1 ‘2.0 J 4 -l.0 -0.5 0.0 0.5 1.0 1.5 -1 0 1 2 NMDSI NMDSI Figure 2.4. Non—metric multidimentional scaling (NMDS) of the restored and extant wetlands based on non-diatom and diatom assemblages in different habitats. The circles are the sample ellipse which is centered on the sample means of the x and y variables. The size of the ellipse is specified at a probability value of 0.6827. Table 2.3. MRPP comparisons of species composition in non-diatom algae and diatom between different habitats in restored and extant wetlands Comparisons Wetland T statistic A P tI/Pe Based on non-diatom algal biovolume Epiphytes vs. Sediments Restored -3.940 0.0145 0.002 Extant - l .241 0.0051 0.1 13 Phytoplankton vs. Epiphytes Restored -5.333 0.0174 0.000 Extant -4.876 0.0276 0.001 Phytoplankton vs. Sediments Restored -8.660 0.0371 0.000 Extant -6.944 0.0442 0.000 Based on diatom relative abundance Epiphytes vs. Sediments Restored -0.138 0.0006 0.389 Extant -1.1 18 0.0057 0.133 31 epiphytes 1.5 1 I r g P=0.023 - or) 1.0" '3 N a Eff . 8 .3 .0- , 1': 2 -O.5 ' -10 1 1 1 -2 -1 0 l 2 I l P=0.026 (D E O a .53 Q 3 NMDS] Figure 2.5. Non-metric multidimentional scaling (NMDS) of the restored and extant wetlands based on non-diatom and diatom assemblages in different habitats. The circles are the sample ellipse which is centered on the sample means of the x and y variables. The size of the ellipse is specified at a probability value of 0.6827. sediment phytoplankton 1.5 I I LUV I 3 fit 1. 0r P=0.762- - . 1 0'5 . 0.5” . " e 0.0’ . tr 0.0r - _0.5_ 1'. _ -0.5- - , P=O.594 , -10 ‘ -l.0 ‘ L ‘ “-2 -1 1 -1.0 -0.5 0.0 0.5 1.0 1'0 ' 110352 , " - - ' O Restored 0.5 i Q - X Extant 0.0 ' ' ) ' '05-05 0.0 0.5 1.0 NMDSl Table 2.4. MRPP comparisons of species composition of non-diatom algae and diatoms between restored and extant wetlands habitats T statistic A P Non-diatom Epiphytes -2.219 0.0070 0.023 algae Sediments 0.760 -0.0042 0.762 Phytoplankton 0.466 -0.0033 0.594 Diatom Epiphytes -2.322 0.0100 0.026 Sediment -0.279 0.0010 0.352 32 Table 2.5. Pearson correlation matrix of dissimilarities in species composition of non- diatom algae and diatoms between restored and extant wetlands in different habitats. A * indicates a P value <0.05. Epiphytic Sediment Epiphytic non- Planktonic Sediment non- diatom Diatom diatom algae non-diatom diatom algae Epiphytic 1.000 diatom Sediment 0613* 1.000 Diatom EP‘Ph-V‘” 0560* 0440* 1.000 Non-dratom Planktonic 0691* 0.377 0656* 1.000 non-dratom Sedfmem 0.394 0.163 0418* 0477* 1.000 non-diatom Differences in species composition of algal assemblages between restored and extant wetlands consistently showed the same pattern in different habitats (Table 2.5). The dissimilarities of non-diatom algae between restored and extant wetlands and the dissimilarity of diatoms between these two types of wetlands were correlated (P<0. 1) (Table 2.5). The only correlation not statistically significant was between the dis- similarities of sediment diatom and non-diatom algae. 2.4.4 Algal assemblages and environmental variables. Multiple linear regression of non-diatom algal similarities among habitats indicated that little variance could be explained by environmental variables (Table 2.6). P and sediment TP had very weak effect on algal assemblages among habitats. However, none of the regression models was significant. Bray-Curtis similarity between epiphytic and sediment diatoms, however, revealed significant regression model with shading by vegetation, LEC, and DO as predictors 33 (R2=0.346, P=0.001). With the increase of plant shading, LEC, and DO, similarity in diatom species composition between epiphytes and epipelon increased. Similar to inter-habitat comparison, similarity between restored vs. extant wetlands using non-diatom algae could not be as easily explained by environmental differences as diatom assemblages (Table 2.8). Only similarity of planktonic non-diatom algae between restored and extant wetlands slightly responded to canopy coverage (P=0.132). Non- diatom algae in other habitats did not respond to environmental variables. Similarity in diatom species composition between restored and extant wetlands in different habitats were related to different environmental factors (Table 2.8). Similarity of epiphytic diatom species composition between restored and extant wetlands could be negatively affected by LEC difference and positively to differences in plant shading. Similarity in sediment diatom species composition related to a number of factors, such as differences in pH, plant shading, and nutrient differences between restored and extant wetlands (R2=0.540, P=0.007). With the decrease of nutrient difference between restored and extant wetlands, similarity in diatom species composition increased. 34 Sod End 3 nommmocwom :2. $2- 485 um: 83 Ram. Rob 09228 386 35868 .3 maraam 03 o 95 N- Re C OD 2835 83 RE 23. “2%qu Sod woman 33 052286 god Sod- 0% Red $56 com; m Ed- m6 53626.80 8.886% .w> song—goinm ovflo Sod- In ....me 8036-52 3 3 Sad 23- RS 2 asses 5 a 35863. as 3.2%....— EE m8: 8wa 43¢. a. $5. 8.45392“. as 83%.QO :3. NE 2:? H E30580 z :oEmanoo mm assa> 935.": 868m .moBatg 3588560 622283: 03 333.3.» E65365 acme—988$ 8066 82:68 6.3 069360 :33qu b56636 6cm Jaw—a Eofi6éo= 82:68 65 .=o§§_ao§fi .osxnmmao wcoEm 5528? mm 03mg; “counomon— $2.662, 6065365 .6 868.8 83%.; 6.822 .3 :ommmocwom dd 033. 35 36 Sod 02 .o 8 8088mm 3 3 ~88 owed 8828 825 Rod 83- 83. um.— Sod 88.2 mm; 2588 8:83 Bed 9.2 8 80888 8585 wood :88 83- A; 82:88 God 23- 85 88260 888 Sod 83 80.0- 28 8. .o 83 :3- E Eco v8.7 23 :8 Sod 38.8 83 08080 35868 o 8 05568 83 2 S 82- 83 8 8828 >856 8883 ...8.“ 898.82 o 8 88880 m 0.838 m 020?“ 806F600 Z 0030.8; 3860: 80688800 8623 .6820? 88x0 :0 :65 68603 606808 s 800309 0608; a 6 0008306 0w80>0 80 00308; 80680806.: .6560? 80:0 63» 68603 606808 a 8 80688800 860% 6%? .6 58:88 6.8043}! 09808 8 038.8.» 8068082“ @0588; 806808068 6 8060060 083806 68308 .3 82000800 .8085 gum 030E 2.5 DISCUSSION Distinct ecological differences were evident between the restored and extant wetlands in southern-central Michigan. In addition to difference in shading by vegetation and duckweeds on water surface, nutrient concentrations and soil organic matter proportion also differed between these two types of wetlands. Differences in algal species composition between two types of wetlands depended on habitat. Epiphytic assemblages were more different between restored and extant wetlands than sediment and phytoplankton assemblages. Algal assemblages also differed among habitats. Abiotic environmental variables were related to many of the changes in algal assemblages among habitats in wetlands and between restored and extant wetlands. Age of a restored site was not a determinant factor for attributes in these wetlands. 2.5.1 Wetland restoration and environmental change in restored wetlands Althought the initial goal of restoration is usually to restore the original physical and chemical characteristics and biological communities of ecosystems, the actual consequences of wetland restoration and construction are variable (Mitsch and Wilson 1996, Kaiser 2001). Wetland restoration can recover some aspects of the lost function and structure of pristine wetlands (USEPA 2001), but most evidence (van der Valk 1981, Galatowitsch and van der Valk 1996, Magee et a1 1999) indicates that restored wetlands differ from extant wetlands in a number of ways. Restored wetlands generally have different hydrogeomorphology, more flood area with expanse of open water (Gwin et al. 1999), narrow borders of vegetation (Kentula et al. 1992, Bedford 1996), low floristic diversity, and different floristic composition (Galatowitsch and van der Valk 1996, 37 Magee et al. 1999) than natural wetlands. In this study, the extant wetlands in southern Michigan generally had complete emergent, submergent, and floating zones. Vegetation in these wetlands was more diverse, and vegetation shading was higher than restored wetlands. Restored wetlands were mainly composed of submergent zone with fewer sedge meadows. Restored vegetation was composed of one or fewer species, and canopy cover was lower. The higher canopy cover and duckweeds cover in extant wetlands indicated light limitation for algal growth. A number of studies have shown that extant wetlands in human impacted landscape had much higher nutrient concentration and sediment organic matter than restored wetlands, which has been related to long—term accumulation and wetland’s function in nutrient retention (Craft et al 1991, Holland et al. 1995, Craft and Richardson 1997). Studies (Bishel-Machung et al. 1996, Craft and Casey 2000) also found reference wetlands had higher C and N content in sediments than in constructed wetlands. Continuous saturation near the soil surface should limit decomposition of organic matter and retain N in natural wetlands (Stolt et al. 2000). My results also showed that nutrient concentration in the extant wetlands were well above average level (MDEQ 2001b) in southern Michigan, while restored marshes were at the beginning of succession and had relatively low nutrient concentration and organic accumulation. I also found significant differences in organic matter and nutrient concentrations between restored and extant wetlands, which may cause different biological communities. 2.5.2 Difference of algal species composition between restored and extant Presumably, algal assemblages in restored wetlands would increase its similarity with extant wetlands and eventually would have the same assemblages as the extant wetlands 38 due to an increasing similarity of environmental characteristics. However, the starting point of a restored wetland may be different from its original state. The physical, chemical, and biological conditions during succession could also differ from natural wetland succession. Human perturbation during succession could also significantly change the path of succession (Detenbeck et al. 1999). Thus, a restored wetland may not necessarily develop to a state similar to mature natural wetlands (Mitsch and Wilson 1996, Lockwood and Pimm 1999, Kaiser 2001). Although high variability in community composition among extant wetlands suggested a high degree of environmental heterogeneity in this study, both environmental conditions and epiphyton in restored wetlands still were significantly different from those in extant wetlands. None of the environmental variables or algal assemblage changes correlated with age of restored wetlands in this study. No linear relationship between age of wetland and organic carbon content in constructed wetlands was found. Previous studies (Bishel-Machung et al. 1996) also indicated that many environmental variables did not reach restoration goals within 10 years of initiating restoration. However, some authors have suggested that wetland nutrient content might recover after more than 15 years of restoration (Craft and Richardson 1998). According to MDEQ (2001a), only 29% of the mitigation projects in Michigan were considered biologically successful, and 26% of the mitigation sites in southern Michigan were observed to have poor water clarity. Mitsch and Wilson (1996) suggested that mitigation projects should be given as much as 15-20 years rather than 5 years before judging their success. The apparent difference of biogeochemical conditions between restored and extant wetlands make diatom assemblages great indicators of wetland conditions. 39 Generally, restored wetlands had higher diatom gamma diversity, beta diversity, and alpha diversity on epiphytes than observed in extant wetlands. Other studies have also observed that restored wetlands did not necessarily have less biodiversity of plant and invertebrates than reference sites (Scatolini and Zedler 1996, Bedford et al. 1999, Magee et al. 1999, Knutson 1999, Williams and Zedler 1999), and no evidence indicated a lower diatom species diversity in restored wetlands (Mayor et al. 1999). Two reasons might have caused low diversity in extant wetlands. First, eutrophication decreases algal species diversity and evenness (Detenbeck et al. 1999, Hillebrand and Sommer 2000). The significantly higher nutrient levels in extant wetlands may be the cause of lower alpha diversity. The other possible explanation was that higher spatial heterogeneity of restored wetlands at different stages of succession provided greater habitat variety for different algal species. The lower beta diversity of diatoms in extant wetlands suggested that diatom species dispersed evenly in extant wetlands, while the higher beta diversity suggested heterogeneous distribution of diatom assemblages among habitats in restored wetlands. The patterns in similarity of diatom species composition in restored wetlands with the extant wetlands were related to environmental factors that may regulate biological differences during the ecological succession. Differences in algal species composition on macrophytes between these two types of wetlands were complexly related to light variables, but the model only explained 20% of the variance of diatom similarity among wetlands. On the other hand, the similarity in sediment diatom species composition between extant and restored wetlands could be explained by difference in pH, nutrients, and light effects. This model was much stronger (explaining 54% of the variance in 40 diatom similarity) and indicated the importance of nutrient accumulation and light regime, particularly for regulating algal assemblages during wetland succession. 2.5.3 Algal species composition among different habitats within wetlands Phytoplankton communities were very different from benthic algal assemblages. As expected, phytoplankton and benthic algal communities had different growth forms which were adapted to planktonic and benthic habitats (Stevenson 1996). Phytoplankton communities contained few diatoms, despite their abundance in the benthos. The majority of southern Michigan wetlands in this study were in open and shelter states (Goldsbough and Robinson 1996), which should be dominated either by epiphyton or metaphyton. Phytoplankton density was generally very low. Duckweeds and macrophytes shading had not only caused light limitation for phytoplankton, but may also have competed with phytoplankton for nutrients as suggested in other studies (Herbst and Hartman 1981). Phytoplankton had high beta diversity, and the structural features of algal assemblages lacked a common pattern among sites. Grazing, allelopathy, light limitation, and alternating self-organization processes in these wetlands make phytoplankton dynamics more complex than in lakes due to the large number of interacting factors (Rojo 2000a, b). The composition of non-diatom algae differed much more between benthic habitats than diatom species composition. The difference of non-diatom algae between macrophytes and sediments in restored wetlands was probably due to the very complicated interactions of algae and their substrata, involving both physical and chemical processes (Burkholder 1996). The physical and chemical habitats for benthic 41 algae on organic substrata is very different than on firm substrata of macrophytes. Lower light due to denser macrophytes and duckweed cover may have constrained the development of differences between non-diatom epipelon and epiphyton growth in extant wetlands, the only comparison of non-diatom algae that was not significantly different among habitats. Diatom assemblages on macrophytes and in sediments, however, were not different in restored and extant wetlands. Lim et al. (2001) claimed that sediment diatom communities reflect changes in the diatom communities on epiphytic and rock substrates in response to limnological conditions. This study also demonstrated that sediment diatom assemblages contained the majority of epiphytic diatom taxa within a wetland, but included a wider range of species than epiphyton. Although epiphytic and sediment diatom did not differ in restored and extant wetlands, the regression model indicated that differences between these two habitats were a function of several different factors including light, DO, and nutrient ratio. Light and D0 are inter-correlated and may present different perspectives to diatom assemblages. The effect of light on diatom assemblages and productivity has been reviewed by Hill (1996). In addition to the demonstrated effects of conductance and pH on diatom assemblages, % open canopy was another important variable that determined patterns of algal taxa among sites in streams (Carpenter and Waite 2000). The direct effect of increased macrophyte shading and duckweed cover in Michigan wetlands could be providing light limited environment for epiphytes, thereby increasing the similarity of epiphytic algae to those in the limited-light environment of sediments. The indirect effect of decreasing light was to decrease DO concentrations. D0 is a comprehensive variable 42 and could be determined by algal productivity and plant and bacteria respiration (Kemp and Dodds 2002). DO also related to many other environmental variables. The negative relation between DO and NI-I4+ concentration may be related to high decomposition, animal excretion, nitrification of ammonia, and decomposition of high N organic matter. Duckweed cover might cause light limitation for algal productivity and low D0 in the water column, but algal biomass was not eliminated, which was indicated by the positive relation between duckweed cover and chl a concentration in the water column. Relatively higher DO environment for epiphytic diatoms and lower DO for sediment diatoms may be another regulating factor for autotrophic and heterotrophic diatom taxa and thus affect diatom species composition. In summary, the physical, chemical, and biological features of restored wetlands differed from extant wetlands in a number of ways. In particular, light reduction and nutrient concentrations, were distinctly greater in extant wetlands, and probably contributed to differences in algal communities among habitats within wetlands and between restored and extant wetlands. I found no direct evidence for increasing similarity in algal assemblages between restored and extant wetlands with increasing age of restored wetlands. However, response of algal assemblages to nutrient concentrations and light, two factors distinguishing restored and extant wetlands, indicated indirectly that successional processes in restored wetlands were regulating algal assemblages. Restored wetlands < 10 years old may need more time to develop mature qualities. 43 2.6 REFERENCES Adamus, PR. 1996. Bioindicators for assessing ecological integrity of prairie wetlands. US Environmental Protection Agency, Western ecological division, Corvallis, OR. USA. EPA/600/R-96/082. American Public Health Association (APHA). 1998. Standard Methods for the Examination of water and wastewater. 20th edition by LS. Clececeri, A.E. Greenberg, and AD. Eaton. Washington D. C., USA. Bedford, B. 1996. The need to define hydrologic equivalence at the landscape scale for freshwater wetland mitigation. Ecol. Appl. 6: 57-68. Beilman, D.W. 2001. Plant community and diversity change due to localized permafrost dynamics of bogs in western Canada. Can. J. Bot. -Rev. Can. Bot. 79: 983-93. Bishe-Machung, L., Brooks, R. A., Yates, S. S., & K. L. Hoover. 1996. Soil Properties of reference wetlands and wetland creation projects in Pennsyvania. Wetlands 16: 532- 41. Biondini, M. E., Bonham, C. D., & Redente, E. F. 1985. Secondary successional patterns in a sagebrush (Artemisia tridentata) community as they relate to soil disturbance and soil biological activity. Vegetatio 60: 25-36. Burkholder, J .M. 1996. Interactions of benthic algae with their substrata. In Stevenson, R.J., Bothwell, M.L., and Lowe, R.L. [eds] Algal Ecology: Freshwater benthic ecosystems. Academic Press, San Diego. pp3-30. Charles, D.F., Knowles, C., & Davis, RS. 2002. Protocols for the analysis of algal samples collected as part of the US. Geological Survey National water-Quality Assessment Program. Report No. 02-06: 215/299-1000. The Academy of Natural Sciences, Patrick Center for Environmental Research, Phycology Section, Philadephia, PA. Clarke, K. R. 1993. Non-parametric multivariate analyses of changes in community structure. Aust. J. Ecol. 18: 1 17-43. Cowardin, L., Carter, V., Golet, F. C., & LaRoe, E. T. 1978. Classification of wetlands and deepwater habitats of the United States. U. S. Fish and wildlife Serveice, Washington, DC, USA. FWS/OBS-79/31. Cottingham, K. L. & Carpenter, S. R. 1998. Population, community, and ecosystem variates as ecological indicators: Phytoplankton responses to whole lake enrichment. Ecol. Appl. 8: 508-30. Craft, C. B. & Richardson, C. J. 1997. Relationships between soil nutrients and plant species composition in everglades peatlands. J. Environ. Qual. 26: 224-32. Craft, C.B, Seneca, E. D., & Broome, SW. 1991. Poewater chemistry of natural and created marsh soils. J. Exp. Mar. Biol. Ecol. 152: 187-200. Craft, C.B., & Casey, WP. 2000. Sediment and nutrient accumulation in floodplain and depressional freshwater wetlands of Georgia, USA. Wetlands 20: 323-32. Detenbeck, N. E. 2001. Methods for evaluating wetland condition: Wetlands classification. EPA 822-R-01-007g. US. Environmental Protection Agency; Office of Water; Washington, DC. Detenbeck, N. E., Galatowitsch, S. M., Atkinson, J. & Ball, H. 1999. Evaluating perturbations and developing restoration strategies for inland wetlands In the Great Lakes Basin. Wetlands 19: 789-820. Galatowitsch, S. M. & van der Valk, A. G. 1996. The vegetation of restored and natural prairie wetlands. Ecol. Appl. 6: 102-112. Goldsborough, L. G., & Robinson, G. G. C. 1996. Pattern in wetlands. In R.J.Stevenson, M.L.Bothwell, and R.L.Lowe eds “Algal Ecology: freshwater benthic ecosystems. Academic Press, San Diego. Ppl21-48. Gosz, J .R. 1992. Gradient analysis of ecological changes in time and space: implications for forest management. Ecol. Appl. 2: 248-61. Gwin, S.E., Kentula, M.E., & Shaffer, P.W. I999. Evaluating the effects of wetland regulation through hydrogeomorphic classification and landscape profiles. Wetlands 19: 477-89. Herbst, R.P., & Hartman, RT. 1981. Phytoplankton distribution of a duckweed covered pond. J. F reshw. Ecol. 1:97-111. Hill, W. 1996. Effects of light. In R.J.Stevenson, Bothwell, ML, and Lowe R.L [eds.] Algal Ecology: freshwater benthic ecosystems. Academic Press, San Diego. Pp121- 48. Hillebrand, H., Durselen, C., Kirschtel, D, Pollingher, U., & Zohary, T. 1999. Biovolume calculation for pelagic and benthic microalgae. J. Phycol. 35:403-24. Hillebrand, H. & Sommer, U. 2000. Diversity of benthic microalgae in response to colonization time and eutrophication. Aquat. Bot. 67: 221-36. 45 Hillebrand, H. & Kahlert, M. 2001. Effect of grazing and nutrient supply on periphyton biomass and nutrient stoichiometry in habitats of different productivity. Limnol. Oceanogr. 46: 1881-98. Holland, C. C., Honea, J. E., Gwin, S. E., & Kentula, ME. 1995. Wetland degradation and loss in the rapidly urbanizing area of Portland, Oregon. Wetlands 15:336-345. John, J. 1993. The use of diatoms in monitoring the development of created wetlands at a sandmining site in western Australian. Hydrobiologia 269/270: 427-436. Kaiser, J. 2001. Recreated wetlands no match for original. Science 293: 25. Keddy, P. 1999. Wetland restoration: The potential for assembly rules in the service of conservation. Wetlands 19: 716-32. Kemp, M.J. & Dodds, W.K. 2002. The influence of ammonium, nitrate, and dissolved oxygen concentrations on uptake, nitrification, and denitrification rates associated with prairie stream substrata. Limnol. Oceanogr. 47: 1380-93. Kentula, M E., Brooks, R. P., Gwin, S. E., Holland, C. C., Sherman, A. D., & Sifness, J. C. 1992. Wetlands: An approach to improving decision making in wetland restoration and creation. Island Press, Washington, DC. 151.pp. Knutson, M. G., Sauer, J. R., Olsen, D. A., Mossman, M.J., Hemesath, L.M., & Lannoo, M.J. 1999. Effects of landscape composition and wetland fragmentation on frog and toad abundance and species richness in iowa and Wisconsin, USA. Conserv. Biol. 13: 1437-46. Koerselman, W., & Beltman, B. 1988. Evapotranspiration from fens in relation to penmans potential free-water evaporation (e0) and pan evaporation. Aquat. Bot. 31: 307-20. Leland, H. V. 1995. Distribution of phytobenthos in the Yahima River Basin, Washington, in relation to geology, land use, and other environmental factors. Can. J. Fish. Aqua. Sci. 52: 1108-29. Lockwood, J .L., & Pimm, S.L.1999. When does restoration succeed? In Weiher, E. & Keddy, P. [eds] Ecological Assembly Rules: Perspectives, advances, retreats. Pp363- 92. Magee, T. K., Ernst, T. L., Kentula, M.E., & Dwire, K. A. 1999. Floristic comparison of freshwater wetlands in an urbanizing environment. Wetlands 19: 517-34. Mayer, P. M. & Galatowitsch, S. M. 1999. Diatom communities as ecological indicators of recovery in restored prairie wetlands. Wetlands 19: 765-74. 46 McCormick, P. V., Rawlik, P.S., Lurding, K., Smith, E.P., & Sklar, F. H. 1996. Periphyton-water quality relationships along a nutrient gradient in the northern Florida Everglades. J. N. Am. Benthol. Soc. 15:433—49. McCormick, P. V., Shuford, R. B. E., Backus, J. G. & Kennedy, W. C. 1998. Spatial and seasonal patterns of periphyton biomass and productivity in the northern Everglades, Florida, USA. Hydrobiologia 362: 185-208. McCune, B. & Mefford, J. 1997. PC-ORD. Multivariate Analysis Of Ecological Data. Version 3.0 For Windows. Mjm Software Design, Gleneden Beach, OR, USA. McDonnell, M. J. & Pickett, S. T. A. 1990. Ecosystem Structure and Function Along Urban Rural Gradients - an Unexploited Opportunity for Ecology. Ecology 71: 1232- 37. Michigan Department of Environmental Quality (MDEQ). 2001a. Michigan wetland mitigation and permit compliance study, final report. Land and water management division, Michigan Department of environmental Quality, Lansing, MI. Michigan Department of Environmental Quality (MDEQ). 2001b. MDEQ wetland identification manual: a technical manual for identifying wetlands in Michigan. Land And Water Management Division, Michigan Department Of Environmental Quality. Middleton, B. 1999. Wetland restoration: flood pulsing and disturbance Dynamics. John Wiley & Sons, Inc. New York. Pp6-7. Mitsch, W.J., & Wilson., B. F. 1996. Improving the success of wetland creation and restoration with know-how, time, and self-design. Ecological applications 6:77-83. NRC. 1995. Wetlands: Characteristics and Boundaries. Nation Research Council. National Academy Press, Washington, DC, USA. Pan, Y., Stevenson, R. J ., Vaithiyanathan, P., Slate, J ., & Richardson, C. J. 2000. Changes In Algal Assemblages Along Observed And Experimental Phosphorus Gradients In A Subtropical Wetland, USA. F reshw. Biol. 44: 339-53. Pringle, C. M. 1990. Nutrient Spatial Heterogeneity - Effects on Community Structure, Physiognomy, and Diversity of Stream Algae. Ecology 71: 905-20. Rader, R. B., & Richardson, C.J. 1992. The effects of nutrient enrichment on algae and macroinvertebrates in the everglades: a review. Wetlands 12: 121-35. Rojo, C, Ortega-Mayagoitia, E., Alvarez-Cobelas, M. 2000a. Lack of pattern among phytoplankton assemblages. Or, what does the exception to the rule mean? Hydrobiologia 424: 133-39. 47 Rojo, C. E., Ortega-Mayagoitia, E., Rodrigo, M.A., & Alvarez-Cobelas, M. 2000b. Phytoplankton structure and dynamics in a semiarid wetland, the National Park "Las Tablas de Daimiel" (Spain). Archiv Fur Hydrobiologie 148: 397-419. Sayer, C., Roberts, N., Sadler, J ., David, C. & Wade, P. M. 1999. Biodiversity changes in a shallow lake ecosystem: a multi-proxy palaeolimnological analysis. J. Biogeogr. 26: 97-1 14. Scatolini, S.R. & Zedler, J .B. 1996. Epibenthic invertebrates of natural and constructed marshes of San Diego Bay. Wetlands 16: 24-37. Stevenson, R. J. 1996. An introduction to algal ecology in freshwater benthic habitats. In R.J.Stevenson, Bothwell, M.L, & Lowe, R.L [eds.] Algal Ecology: freshwater benthic ecosystems. Academic Press, San Diego. Pp3-30. Stevenson, R. J. & Hashim, S. 1989. Variation in Diatom Community Structure among Habitats in Sandy Streams. J. Phycol. 25: 678-86. Stevenson, R. J. 1997. Scale-dependent determinants and consequences of benthic algal heterogeneity. J. N. Am. Benthol. Soc. 16: 248-62. Stevenson, R. J ., Sweets, R. P., Pan, Y., & Schultz, R. E. 1999. Diatom community patterns in wetlands of Michigan (USA) and their use as indicators of ecological conditions. In McComb, A. J. & Davis, J. A.[eds.] Proceedings of INTECOL’s Vth International Wetland Conference. Gleneagles Press, Adelaaide, Australia. Pp. 517- 27. Stevenson, R. J ., McCormick, P V., & Frydenborg, R. 2001. Methods for evaluating Wetland Condition: Using Algae to assess environmental conditions in wetlands. EPA 843-B-00-002k. US. Environmental Protection Agency; Office of Water, Washington, DC. Stolt, M.H., Genthner, M.H., Daniels, W.L., Groover, V.A., Nagle, S., & Haering, KC. 2000. Comparison of soil and other environmental conditions in constructed and adjacent palustrine reference wetlands. Wetlands 20: 671-83. Stosch, H. A. von & Reimann, B.E.F. 1970. Subsilicea fragilarioide gen. Et spec. nov., eine Diatomee (Fragilariaceae) mit vorwiegend organischer membran. Nova Hedwigia Beih. 3121-36. Tilman, D. 1981. Tests of resource competition theory using four species of lake Michigan algae. Ecology 62: 802-15. USEPA, 2000. Principles for the Ecological Restoration of Aquatic Resources. EPA841- F-00-003. Office of Water (4501F), United States Environmental Protection Agency, Washington, DC. 4 pp. 48 van Dam, H. Mertens, A. & Sinkeldam, J. 1994. A coded checklist and ecological indicator values of freshwater diatoms from the Netherlands. Neth. J. Aquat. Ecol. 28:1 17-33. van Der Valk, AG. 1981. Succession in wetlands: a gleasonian approach. Ecology 62: 688-96. Vymazal, J. 1995. Algae and element cycling in wetlands. Lewis Publishers, Boca, Raton, FL. 689 pp. Whigham, D. F., Good, R. E., & Kvet, J. 1990. Wetland ecology and management: case studies. Kluwer Academic Publishers. Netherlands. Williams, G. D. & Zedler, J. B. 1999. Fish assemblage composition in constructed and natural tidal marshes of San Diego Bay: Relative influence of channel morphology and restoration history. Estuaries 22: 702-16. Wu, X. Y. & Mitsch, W. J. 1998. Spatial and temporal patterns of algae in newly constructed freshwater wetlands. Wetlands 18: 9-20. Zedler, J. B. 1993. Canopy architecture of natural and planted Cordgrass marshes: selecting habitat evaluation criteria. Ecol. Appl. 3: 123-38. Zedler, J. B. 2000. Progress in wetland restoration ecology. Trends Ecol. Evol. 15: 402- 07. 49 Chapter 3 CHANGES IN BENTHIC ALGAL ATTRIBUTES DURING SALT MARSH RESTORATION AND USE IN BIOASSESSMENT 3.1 Abstract: Although salt marsh restoration has been used to mitigate loss of habitat and ecosystem function and hundreds of restored marshes have been built along the US coast, few effective indicators have been developed to determine success of ecological restoration in salt marshes. To assess attributes of algal assemblages as indicators of salt marsh restoration, I chose 8 pairs of salt marshes in North Carolina, each pair with one restored marsh (from 1 to 28 year old) and a nearby reference salt marsh. Algae on Spartina altemifiora and sediments (sediment algae) were collected in each marsh during spring and summer 1998 for assaying biomass (dry mass (DM), ash free dry mass (AFDM), chl a content, algal biovolume), species composition and diversity, and gross primary production. An attribute restoration ratio was calculated by dividing attribute values from each restored marsh by values from the paired reference marsh. Controlling for regional variation in reference marshes substantially increased precision in relations between attributes and age of restored marshes. Organic matter restoration ratios of sediments increased with age of restored marshes in spring and summer. Algal biomass restoration ratios of epiphytes, calculated with algal biovolume and chl a, increased with restored marsh age in summer, but not during spring. In contrast, sediment algal biomass was not related to marsh age. The species diversity of sediment algae in summer showed an asymptotic relationship with sediment nutrient concentration. The similarity of diatom species composition between paired restored and reference sites increased with age of 50 restored marshes during both spring and summer. Primary production by epiphytic and sediment algae in summer showed site-specific changes and did not vary consistently with marsh age. Algal biomass, algal diversity, and diatom species assemblages during summer were positively correlated with sediment nitrogen and phosphorus concentration. Overall structural and functional development of restored wetlands, especially nutrient storage in sediments, regulates algal community structure and function, which can be used to evaluate marsh restoration success. Key Words: salt marsh, restoration, algal attributes, succession, algal diversity, diatoms, biomass, nutrients, environmental variation 51 3.2 INTRODUCTION Ecosystem restoration is one of the most pressing problems for applied ecologists (NRC 1992). With expanding urbanization and other human activities along coasts, many salt marshes have been degraded. A large effort has been mounted to stabilize these coastal ecosystems (Adam 1990). Along the US coast, hundreds of degraded or lost salt marshes have been restored in the past decades. Currently, few criteria exist to guide assessment of wetland restoration, so their evaluation is often superficial and subjective. Furthermore, very few reference data sets exist that can be used to compare restored habitats with natural habitats (Lockwood and Pimm 1999, Keer and Zedler 2002). The most commonly used criterion for judging the success of ecosystem restoration is whether or not the restored community resembles the original (Adamus 1988). Developing criteria to evaluate progress in restoration can be done by characterizing development of a salt marsh ecosystem following initial restoration and identifying attributes in that marsh that increase in similarity relative to mature, existing marshes. Alternatively, multiple restored marshes with different ages could be compared to reference marshes to relate successional processes in restored ecosystems to a reference system (Lockwood and Pimm 1999). Living and nonliving organic matter, species diversity, and nutrient retention increase during ecosystem development (Odum 1969). Previous studies of restored salt marshes suggest that some of these attributes develop very slowly (e.g., 15 years for soil nutrient pools, Craft et al. 1988), while others recover within 1-5 years (Craft et al. 1991). Slow changes in biogeocherrristry, water chemistry and plant and invertebrate species composition demonstrate the importance of long-terrn 52 research to characterize ecosystem succession (Hook 1988, Craft et al. 1991, Sacco et al. 1994, Scatolini and Zedler 1996). Algae are important primary producers in salt marsh habitats and affect secondary production, nutrient retention, and biogeochemical cycling (Pomeroy 1959, Leach 1970, Gallagher and Daiber 1974, Van Raalte 1976, Zedler 1980, Sullivan and Moncreiff 1988, 1990, Pinckney and Zingmark 1993a, b). Algal photosynthesis can account for a large proportion of primary production in salt marshes, ranging from 25% to as much as 70% in some coastal marshes (Adamus 1988). Teal (1962) estimated that 47 percent of the total net primary productivity in salt marshes was lost through respiration by microbes, however, algal debris still contributed to more than one third of soil organic matter. Algae are also a major energy resource and provide the primary diet for metazoans and other invertebrate grazers (Kreeger and Newell 2000). Finally, nitrogen fixed by cyanobacteria is a very important nutrient resource in nitrogen limited salt marsh ecosystems (Jones 1974, Raalte et al. 1976, Sage and Sullivan 1978). Algal species diversity, species composition, biomass, chemical composition, and productivity are most important attributes for assessing ecosystem structure and function in aquatic systems (Stevenson and Pan 1999). Algae, especially diatoms, have been widely used as indicators of ecological conditions in freshwater, marine, and terrestrial ecosystems (Shubert 1984, Steneck and Dethier 1994, Stevenson and Pan 1999, Sullivan and Currin 2000, Stevenson et a]. 2001), but they have never been used to assess salt marsh condition. Because of their species-specific response to biogeochemistry (Shubert 1984), changes in diatom community structure and function could be used to assess restoration of wetland structure and function. 53 In this project, algal attributes in restored salt marshes of different ages were compared with reference marshes to answer the following key questions: 1) what timescales are required to achieve full restoration of algal-related attributes of salt marsh structure and function; 2) which indicators can be used to determine whether full restoration of salt marsh structure and function has been achieved; and 3) can ecological succession theory apply to salt marshes restoration? I hypothesized that l) algae would rapidly colonize restored salt marshes and establish high primary production; 2) algal species diversity and organic matter on the surface of sediments Would increase slowly over time and became more similar to reference sites with greater age of restored marshes; and 3) diatom species assemblages and organic matter in sediments would be important indicators of successful restoration. 3.3 METHODS 3.3.1 Site Locations J; 30 80 90 120 150 Kilometers ; M an“ 7 $- , ‘ . . w E Morehead Clty 2 5 A OI ‘V / \\ ‘> .. *‘ Marine Lab Marsh (26) ~ Consultant Marsh (3) " W"'“"‘9‘°" . Port Marsh (3) . 4.2-5% , _ ' Dill’s Creek (13) . ‘ DOT Marsh (1) ‘ V \ (I Fine Knoll shores (24) 1 “ Swansboro (1 1) 4 Snow’s Cut (28) Figure 3.1. Map of sampling sites between Morehead City and Wilmington along the North Carolina coast. Age (yr) of restored marshes is reported in parentheses. 54 Salt marshes were sampled in the urbanizing area between Morehead City and Wilmington along the North Carolina coast (Fig. 3.1). The restored marshes examined were established for shoreline stabilization, mitigation of wetland loss, or research. The reference marshes were natural marshes near the restored marshes so that differences in water chemistry factors that affect algal assemblages, such as salinity and nutrient concentrations, would not vary due to regional factors. Each pair of wetlands received water from the same source. Eight pairs of salt marshes were selected and sampled in March 1998 (Figure 3.1): Long Beach (LB), Consultant Marsh (CS), Port Marsh (PT), Swansboro Marsh (SB), Dill Creek (DC), Pine Knoll (PN), Marine Lab (ML), and Snow’s Cut (SC). All restored marshes were planted with Spartina altemiflora Loisel at the time of restoration. Long Beach and Pine Knoll marshes were dropped from the study in July 1998 because of the poor macrophyte growth and an additional pair of salt marshes (Department of Transportation (DOT)) was added. To determine whether changes in developing salt marshes of North Carolina were similar to those in other regions, samples were collected from 3 pairs of restored and reference salt marshes in Chesapeake Bay at Virginia Beach, VA during summer 1999. The restored VA marshes were 2, 10, and 20 years old. These marshes represent very different habitats than those in North Carolina because they were far from the ocean, had salinity below 10 ppt, and were brackish to oligohaline. 55 3.3.2 Field Sampling An area referred to as "the front marsh" was sampled. This area was approximately 1- 2 meters from the levee away from the water and relatively low and flat. Sampling transects 30 to 50 m long were established parallel to the beach in this zone. Because some of these marshes did not have well-developed levee zones and several of them didn’t have Spartina in back marshes, only the front marsh area was included in this study. Spartina stem density was measured in 0.25 m2 plots (10 per marsh) in October 1998. Plant stem density in the front marsh zone of reference sites ranged from 201 to 316 stems/m2 except in Pine Knoll marsh, which had much higher stem density (855/m2) than any other reference sites. Restored sites had stem density ranging from 240 to 498 stems/m2 (Table 3.1). Salinity was measured in the field using a refractometer. Surface soils (0-10 cm, n=10 per marsh) were collected in June 1998 and analyzed for total N and P as described in Craft et a1. (1999). N and P concentrations in younger restored marshes were generally lower than its nearby reference marshes (Table 3.1). Table 3.1. Sampling sites and their properties during the summer sampling season (from Robert Freese and Christopher Craft, unpublished data). Salinity(ppt) Plant Stem density (lmz) Nitrogen (mg/g) in Phosphorus (mg/g) in surface sediment surface sediment Age Site Reference Restored Reference Restored Reference Restored 1 DOT 20-30 266 498 2. 1 0.2 0.264 0.21 l 3 CS 17-32 208 304 2 0.2 0.368 0.143 8 PT 18-30 255 342 1.2 0.3 0.955 0.263 1 1 SB 20-30 316 398 4.8 0.4 0.545 0.409 13 DC 14-33 280 240 3 1 0.358 0.236 24 PN 20-30 855 512 0.3 0.6 0.453 0.370 26 ML 20-30 201 423 2. 1 3 0.369 0.408 28 SC 5-20 271 297 4.4 3 0.492 0.371 56 Algal productivity on sediments and plants was measured during summer 1998. Eight sediment samples were collected from each transect by using plastic tubes (internal diameter = 3.8 cm, length = 20 cm) along each restored and reference site. Samples were taken early in the morning and brought back to the shore immediately. The sediments in the corers were covered with 30-ml of estuary water and then sealed air-tight with a thin, transparent plastic sheet and rubber bands. The cores from each transect were incubated in a light or dark chamber for 4 hours (4 cores in each treatment). Subsamples of 15 ml seawater were taken from each core to measure dissolved oxygen concentration (DO) using Winkler methods (APHA, 1998). The top l-cm of sediment from each core was sampled and stored on dry ice for later analysis of algal biomass. Epiphytic samples were collected from front marshes by cutting 1-3 randomly selected stems of Spartina from just above the ground and placing them into a zip-lock bag. Eight bags of stems were randomly collected in each zone of each marsh. All stems in a bag were washed using 200 ml of estuary water and removed from the bag. The epiphytes were left in the bag. All bags were then sealed underwater to avoid trapping air bubbles. The surface area of macrophytes from where the samples were taken was measured using a ruler. Four samples from each site were put in a light chamber, four were put in a dark chamber, and all were incubated for 4 hours. The experiments were done under light irradiance ranging from 800 to 2,200 umole quanta'm‘zs’l, 32 ppt salinity, and a water temperature of 30il°C. DO was measured using 60 ml subsamples before and after incubations. The remaining portions of samples were frozen on dry ice for later analysis of algal biomass. 57 3.3.3 Laboratory Analysis In the laboratory, the frozen algae sample was homogenized and split into different volumes for assay of organic mass (30 ml), chlorophyll (15 ml), and separate microscopic assays for all algae (10 ml) and diatoms (10 ml). 200 ml of filtered seawater were added to sediment samples to enable subsampling. Dry mass (DM) and ash free dry mass (AFDM) were analyzed using standard methods (APHA 1998) to estimate sediment and epiphytic mass and organic matter. The organic matter proportion of sediments (OMP) was then calculated based on the organic proportion of total sediment (AFDM/DM). Chlorophyll a (chl a) contents of sediment algae and epiphytes in summer were measured using a Spectronic GenesysZ® spectrophotometer (Spectronic Instruments) following standard methods (APHA 1998). Ch] a and phaeophytin were related to area of ground and plant surface sampled. Algal biomass and species composition of all algae and diatoms were determined by microscopically identifying and counting algae using two preparation techniques. After suitable dilution and settling, 300 natural units of algae in the chamber were identified and counted within a known portion of the samples. Non-diatom algae were recorded and identified to genus; diatoms were divided into 5 groups according to their length and width (less than 10, 11-20, 21-40, 41-60, and >60 um) and counted. Soft algal volume was determined by measuring at least 10 individuals or colonies. Average volumes of diatom size groups were estimated by measuring 30 cells for each group. Total algal biovolume was calculated to provide a second measure of algal biomass with chl a. Diatom samples were mounted in N aphrax® after digestion using HNO3 and KzCr207 58 (Stosch and Reimann 1970). Diatoms were identified to species and at least 500 valves were counted for each sample. Among all pairs of restored and nearby salt marshes, I determined that Snow’s Cut (28 year old) was a poor reference site for the nearby restored marsh, which was located in the middle of the Cape Fear River channel. Differences in algal assemblages between these marshes were probably related to large differences in salinity, nutrient input, and sedimentation. Therefore I decided to exclude this site from analyses that involved diatom species composition. 3.3.4 Statistical analyses I used SYSTAT® 10 to calculate summary statistics, t-test for means, F statistics for variance, and Analysis of variance (AN OVA) to compare attribute characteristics between restored and reference marshes. Analysis of Covariance (ANCOVA) was used to compare slopes of two different regression lines. Attribute restoration ratios were calculated to indicate relative condition in a restored site compared to the nearby mature reference marsh. A ratio equal to one meant that the attribute of the restored marsh was the same as in reference marsh. Thus, the organic matter restoration ratio (OMRR) was defined as (OMP in restored marsh)/(OMP in reference marsh). An algal chlorophyll restoration ratio (ACRR) was similarly calculated to relate chlorophyll concentrations in restored and reference marshes. The algal biovolume restoration ratio (AVRR) was defined as the ratio of total algal biovolume per unit surface area in a restored marsh vs. biovolume in the respective reference marsh and was calculated for both epiphytes and sediment algae. A primary production restoration 59 ratio (PRR) was defined as the ratio of gross primary production per unit surface area in a restored marsh vs. that in its reference marsh for both epiphytes and sediment algae. A percentage similarity index (PSI) was used to indicate the similarity of diatom species composition in restored and nearby reference marshes (Gauch 1982). The index was calculated as PSI=E minimum (pn, p2,), where PSI: percentage similarity between samples 1 and 2, p1, = percentage of the ith species in the restored marsh (i = 1,. . ..., k species), and pzi = percentage of species i corresponding reference marsh. PSI of algal functional groups was also calculated by comparing the biovolumes of functional groups in restored and reference marshes. Functional groups were defined by algal division and growth form (Steneck and Dethier 1994). All algae were divided into 9 functional groups: 5 diatom size groups, cyanobacteria, green algae, brown algae and red algae. To evaluate successional patterns of change of algal attributes, linear regression analyses (SYSTAT ®10) were used to relate algal attributes, restoration ratios, and diatom community similarities to age of restored marshes in both spring and summer. Nonlinear regression analyses were also performed to relate changes in attributes to environmental variables (SYSTAT® 10). Canonical correspondence analysis (CCA) was performed using the computer program CAN OCO (ter Braak, 1987) v4.02 to relate diatom species data to environmental variables. Sediment and epiphytic diatom taxa with relative abundances (taxa proportions of assemblages) greater than 1% in one or more sites during summer were included in our CCA with respect to the available environmental variables: sediment total nitrogen, total phosphorus, salinity, stem density, OMP, and chl a. A forward-selected CCA was then used to determine which environmental variables accounted for the greatest amount of variance in the distribution of the diatom taxa. The significance of each variable added in this fashion was tested using a Monte Carlo permutation test with 99 unrestricted permutations (P _<. 0.05) (ter Braak 1990). To determine the importance of regional factors, three pairs of restored and reference salt marshes from the Virginia coast were compared to the North Carolina summer sample set to determine how similar restoration patterns were between different regions. 3.4 RESULTS 3.4.1 Biomass Development 08 '- u u v i d 0 r u u u 06 I- ¢ .. 03 $ 1 l a? 0.4 “L? p B d E g .. . . . 0.02 . in 0.2 [F § 2 z .. . Q 0 E 0.01 . O r: $ I T I I 50 I t v 30 1 y . A1000 [ B . 1 b , a; 20 . 0 ¢ . NE : i o E 10 " E 10.0 f H g 1:; 30 - i E o g r i g 20 P ' "g: 5 p a g 1.0 I 1.33. a a 10' .43. 5 r ' 2 0.1 n l l O Q" ‘ . as" 59‘“ 59" 9“" ‘39 5““ $3 9““ 9““ 9°“ Epiphytes Sediment Epiphytes Sediment Epiphytes Sediment Figure 3.2. Box-plot of dry mass (DM), ash free dry mass (AFDM), the organic matter proportion of sediments (OMP), algal biovolume, and chl 3 samples from epiphytes and sediments in reference marshes during spring (8 samples) and summer 1998 (7 samples). 61 Characteristics of sediments (DM, AFDM, and OMP) and algal biomass (chl a and biovolume) in reference marshes differed between macrophytes and surface sediments (Figure 3.2). During spring, DM and algal biovolume on macrophytes were much lower than in surface sediments (t-test, P<0.05), AFDM on macrophytes was not different from sediment (P>0.05), but the OMP on macrophytes (27.3128 %) was higher than in sediment (14.2:l:3.7 %)(P<0.05). During summer, DM, AFDM, algal biovolume, and chl a on macrophytes were all significantly lower than in surface sediments (P<0.05), but OMP on macrophytes did not differ from sediment (P>0.05). Sediment mass (DM and AFDM), algal biovolume, OMP on macrophytes, and algal biomass in surface sediment were higher in spring than summer (P<0.05). Measures of sediment mass (DM and AFDM) were less in spring than summer on surface sediment (P<0.05) (Figure 3.2). DM, AFDM, and OMP in restored marshes were not related well to restored marsh age (P>0.05). However, the organic matter restoration ratio (OMRR) on macrophytes and in sediment during spring was linearly related to age of restored marshes (Figure 3.3). During summer, the OMRR on macrophytes was not related to age of restored marshes and the OMRR in sediments again significantly increased with age of restored marshes in a manner not statistically different from that observed in spring (ANCOVA, P=0.450). Algal biomass in young restored marshes was about half that observed in mature reference marshes. In general, algal biomass restoration ratios increased with age of restored marshes. AVRR was not related to restored marsh age during spring either on macrophytes and in sediments (P>0.05) (Figure 3.3e-f). During summer, sediment AVRR and sediment ACRR were not related to restored marsh age, but epiphytic AVRR and ACRR were significantly related to age of restored marshes (Figure 3.3g-j). The AVRR 62 patterns were statistically similar to the ACRR pattern for epiphytic and sediment algae in summer (ANCOVA, P=0.375 and P=0.453 respectively). OMRR AVRR ACRR 1.5 . . , , . 3e 1.0 , ..................................................... _ 0.5 ' ° . 0.0 :7 :' % If : 1.5 3f . 1.0 , ................. . 0.5 ° . . 00 ‘ l l l l l r r r 1 l 0 51015 20 25 30 0.00 51015 20 25 30 2.0 I I r T I 1.5 I T I I I 1.5 F I I I I O . 3C . 3g 31 I- II . E 1 5 / 1.0 ...; .................................................... ....“ 1.0 ¢ ' 0 g o .5 1 O " o . ran) ' . 0.5 - ° - 0.5 .o - E- 0.5 " ‘ . . ' 0.0 l l l 4 0-0 l A r 4 1 (’5) 2.0 r 1 I I T 0'0 v t v v 1.5 .V F I l I . 3d 1.5 m 1.5 " 7 v2, . 1.0 - 1-0‘ £1 1 0 - . 8‘ ° . ' 0-5 0.5 0.5 l‘ ‘ O O l 1 1 l I 00 1 l l l 1 0.0 O V = 5 : = ' 0 5 10 15 20 25 30 0 5 10 15 20 25 30 0 5 10 15 20 25 30 Age of restored marsh(yr) Age of restored marsh(yr) Age of restored marsh(yr) Figure 3.3. Relationship between the attribute restoration ratio and age of restored marsh in surface sediment and on macrophytes during spring and summer 1998. The OMRR is the comparison of OMP of sediments in a restored site with that of its reference marsh. The ACRR is the comparison of algal chlorophyll a concentration in a restored site with that of its reference marsh. The AVRR is the comparison of algal biovolume in a restored site with that of its reference marsh. The line represents a significant relationship (linear regression P<0.05). The correlation coefficients (R2) for each relationship are: 3a: 0.727; 3b, 0.931; 3c, 0.969; 3d, 0.235; 3e, 0.304; 3f, 0.647; 3g, 0.577; 3h, 0.749; 3i, 0.442; 3j, 0.741. 63 In summer, sample OMP increased significantly with increasing nitrogen and phosphorus concentration of surface sediments (Table 3.2). Sediment chl a concentration was asymptotically related to sediment nitrogen concentration and was barely correlated with surface phosphorus concentration (Table 3.2). Sediment algal biovolumes showed the same correlation with N and P as chl a. All three attributes were not significantly related with macrophyte stem density (Table 3.2). Table 3.2. Summary of regression analyses between algal attributes and environmental variables. Each marsh (both restored and reference marshes) was treated as one site. Attributes Environmental variables Adj R2 n Equation P(2 tail) Model 0.372 14 5.409+0.192X 0.012 linear OMP TP 0.531 13 -0.870+30.306X 0.003 linear Stem density 0.464 14 0.465 none TN 0.459 14 23.278 «1» (X /(0.311+ )0) 0.039 asymptotic Chlorophyll TP 0.232 13 5.501+33.551X 0.055 linear Stem density 0.119 14 8.558+0.029 X 0.227 linear TN 0.505 14 4.841+4.502X 0.003 linear Biovolume TP 0.564 13 -8.015+15.990X 0.002 linear Stem density 0.039 14 2.317+0.037X 0.241 linear TN 0.540 14 3.546 * (X / (X+0.024)) asymptotic Biodiversity TP 0.419 13 3.714 * (X / (X + 0.026)) asymptotic Stem density 14 none 3.4.2 Algal Species Diversity and Composition Table 3.3. Number of species found in all sites, habitats and seasons. Habitat No. of algal species No. of diatom taxa Season type Res. Ref. Total Common Res. Ref. Total Common Spring Sediment 204 218 292 l37(46.9%) 212 197 277 132(47.7%) Epiphyte 167 170 225 104(46.2%) 155 149 205 98(47.8%) Summer Sediment 164 159 205 1 17(57. 1 %) 154 156 198 1 14(57.6%) Epiphyte 131 118 161 85(52.8%) 112 123 150 81(54%) A total of 420 algal taxa, including 400 diatom taxa, were found on all substrates in the spring and summer samples. Of them, 362 algal taxa were found during spring and the majority (342 and 94.5%) were diatom taxa. Generally, more algal species were observed in sediment samples (292 taxa) than on plants (225 taxa) (Table 3.3). Fewer algal species were found in summer than in spring. Two hundred fifty-two algal taxa with 241 diatom taxa were found in all sampling sites in summer with more algal species (205 algal taxa) in sediments than on plants (161 species) (Table 3.3). Restored and reference marshes shared fewer algal taxa (about 47%) and fewer diatom taxa (about 48%) in spring than in summer (>52.8%)(Table 3.3). Shannon diversity indices of epiphytes varied from 2.02 to 3.10 in restored marshes and 2.59 to 3.00 in reference marshes in summer, Shannon indices of sediment algae ranged from 2.93 to 3.55 in restored marshes and from 3.36 to 3.58 in reference marshes. Generally, the index values for restored sites in summer were more variable than for reference marshes (F test, P=0.002 for epiphytes and P=0.016 for sediment algae). However, neither the epiphytic nor sediment diversity indices differed between restored sites and reference sites (P>0.05). Asymptotic relationships were observed between sediment algal diversity (H’ ) and sediment nitrogen and phosphorus concentrations (Table 3.2) in summer. Algal species were mostly composed of diatoms, filamentous green algae, and cyanobacteria. The most frequently observed soft algal genera were Enteromorpha spp., Rhizoclonium spp., Oscillatoria spp., and Schizothrix spp. Enteromorpha was distributed patchily in most marshes; however, because of the difficulty of quantifying the occurrence of macroalgal species such as Enteromorpha spp. and relative few numbers of 65 macroalgae taxa, they were not used to compare species composition of algal assemblages. The most common epiphytic algal group was diatoms with an average of 75% (il3% SE) of total algal biovolume in spring and 89% (i4.8%) in summer. Diatoms also were dominant in sediment assemblages, with an average of 73% (21:12%) of total biovolume in spring and 93% (i4%) in summer. Epiphytic cyanobacteria algae were 0.3% to 35.2% of total algal biovolume in restored marshes and from 0.04 to 22.3% in reference marshes. Sediment cyanobacteria were less than 10% of algal biovolume in both restored and reference marshes in spring. PSI of functional group biovolume between restored and its respective reference marshes did not change with restored marsh age in spring or summer (Figure 3.4a, b). Most of the paired sites had high similarity of functional group composition (larger than 50%) on both macrophytes and sediments. Diatom composition on macrophytes and in sediment shifted seasonally. The most dominant epiphytic species on macrophytes in spring were Melosira nummuloides Agardh, Denticula subtilis Grunow, Synedra fasiculata Grunow, and Navicula penninuta Grunow. During summer, M. nummuloides became much less abundant, while Nitzschia nana Grunow, Amphora coffeaeformis (Agardh) Kutzing, and Nitzschia frustulum Grunow became the most dominant epiphytic species. Dominance of sediment diatom assemblages changed little from spring to summer. With M. nummloides, A. cojfeaeformis, and Rhaphoneis surirella (Ehrenberg) Grunow being most common (Table 3.4). 66 Spring 4a Summer 4b 0.8 I I I I I 1.0 I I t j 0 o a, jg 0.6 - - 0-3 l . ' ' 8 g 0.6 ' ° .- 9 g 0.4 " ‘ 3’ a) ' 0.4 " ' g 02 II . .u ,9 ' 0.2 ' " E 0.0 . . . . . 0,0 : ; + . : § ‘0 0.8 I I I I I 1.0 . . . . g g 0.6 I- . II 0.8 .9- 0 0.6 ' ' " °' 13 0.4 - J J 0.4 'L' ' 0.2 r 0.2 . 0.0 L l n l n 0.0 n n n ' n n 0 5 10 15 20 25 30 0 5 10 15 20 25 30 4d 0.8 . . - --4L1 0,3 . . . . . E 0-5 " 0.6 “e’ -- 0.4 - - ,4 ($3 0 8 0.2 0.2 ‘9 00 32 - g 0.8 833 ‘9 06 06 a a ' E‘ 0.4 0.4 .9- 13 02 ' 02 '1 ' R2=0.65 ' R2=o.850 0.0 J L ‘ J 0.0 ' ‘ ‘ ‘ ‘ 0 5 10 15 20 25 30 0 5 1O 15 20 25 30 Age of restored sites (yrs) Age of restored sites (yrs) Figure 3.4. Relationship between the percentage similarity index (PSI) of algal functional groups (4a, 4b) and diatom assemblages (4c, 4d) and the age of restored marshes in surface sediments and on macrophytes during spring and summer 1998. PS1 was calculated for algal functional groups or diatom assemblages between restored and reference marshes. The line represents a significant relationship (linear regression P<0.05). The triangles represent sites in Virginia, and dots represent sites in North Carolina. PSI of diatom species composition between restored and reference marshes increased with age of restored salt marshes both on macrophytes and in sediments in the spring 67 (Figure 3.4 c). The rate of change in PSI between restored and reference marshes with age of restored marshes did not differ between epiphytic and sediment algae (ANCOVA, P=0.602). The PSI changed from 37% to 70% on epiphytes and from 30% to 50% in sediment. Similarity in epiphytic and sediment diatom communities between restored and reference marshes during summer also increased with increasing age of restored salt marshes (Figure 3.4d). The rate of change in PS1 between restored and reference marshes with age of restored marshes did not differ between epiphytic and sediment algae (ANCOVA, P=0. 160). The similarity changed from 30% to 73% on macrophytes and from 36% to 65% in sediment. Canonical correspondence analysis of epiphytic diatom assemblages showed that 26.8 of the total variance in the species data was explained by the first two axes. The first axis was correlated with OMP (r=0.769), nitrogen (r=0.495), chl a (r=0.524), and salinity (r = -0.467); the second axis was correlated with stem density (r=0.493), salinity (r = -0.474) and OMP(r=0.394). Similarly, CCA of sediment diatom assemblages and environmental variables showed that 26.2% of the species variance was explained by the first two axes. The first axis was negatively correlated with OMP (r=0.617), nitrogen (r=0.503), and chl a (r=0.490); the second axis was negatively correlated with salinity (r = -0.737). In epiphytic samples, Achnanthes exigua var. elliptica Hustedt, Actynoptychus splendens Ralfs, and Skeletonema costatum (Greville) Cleve dominated in high N concentration sites, while Navicula pygmaea Kutz., Nitzschia clausii Hantzsch, and Amphora proteus Gregory dominated in low N environments. In sediment samples, Plagiogramma pygmaeum Grev., Actinoptychus splendens (Shadbolt) Ralfs in Pritchard, Gyrosigma balticum (Ehrenberg) Rabenhorst, Diploneis papula (A. Schmidt) Cleve, Amphora gigantean Grunow ex A. 68 Schmidt were negatively correlated with nitrogen concentration. With an increase of nitrogen concentration, diatom species composition shifted to Nitzschia debilis (Amott) Grunow. 3.4.3 Productivity Gross primary production (GPP) of algae had a wide range during summer. The GPP of epiphytes ranged from 3.832 to 75.338 mg C rn'2 h’1 in reference marshes, and from 6.503 to 137.068 mg C m’2 h'1 in restored marshes. The sediment algal GPP showed less variation, ranging from 4.56 to 16.56 mg C m'2 h'1 in reference marshes and from 4.08 to 46.981 mg C m'2 h'1 in restored marshes. PRR did not change significantly with age of restored marshes either for epiphytes or sediment algae (Figure 3.5). PRR was generally higher in younger marshes. Great variation among different pairs of marshes was observed, but age of restored marshes was not a determinant of algal productivity in this study. 3.4.4 Algal Attribute Development In Virginia Salt Marshes DM, AFDM, OMP, and algal biovolume in Virginia salt marshes varied among different marshes and did not change with age of restored marshes. Sediment DM (0.182 1: 0.043 g/cmz) of Virginia reference marshes was lower than BM in North Carolina reference marshes (P<0.05); AFDM (0.033 i 0.010 g/cmz) and OMP in sediment (18.2 i 3.5%) were similar to North Carolina marshes. DM on plants (0.43 :l: 0.12mg/cm2), 69 30:80 000.000.0000.: 0.000802 30:80 000.000.0000.: 0.00.4002 30:80 30:80 0000:: 0.00302 2090: 0000.000 0.0000002 30:80 3:00:00an 0:00.200 000000002 30:80 000.000 0000000000. 8053— 0000000000000 0000.302 @853: 0000.20.00 0.0000002 30:80 000000.000. 0.0000002 30:80 00000000 00000000040 30:80 00.00000 0000000030 £80m< 0000000000000 00.20002 80: 30:80: 00%000 00000000 30:80 00.00000 00.00.000.000 30:80 00000000 0000000000. 885. 000000000000 0000.302 30:80 000.000.0000.: 0.0000002 30:80 30:80 000.000.0000.: 0.0000002 30:80 000000§030802 0>0_0 3000:0008 000.000 0.000% 3000:0008 0:00.000“. 0000000002 m8~8¥ 0000 0.000.302 835— 30:80 00.0300 00000000000 006.0 0088.6 .8. 0000.302 2000?: 00000000080 0000050 30:80 000.00 0000000030 8:00:85: 00000». 0.00002 8502 .w000 000000000v 0000008 30:80 0.0.0000 0000.000Q 30:80 0.0.0000 0000.000Q A5809: 00000000080 0000008 .8. 000.302 8820— 850M 30:80 000000000\ 0.0000002 30:80 0.0.0000 0000.000Q £0009: 0.800000%.00 00000000. 3000?: 0.800000%.00 00000050 30:80 0.0.0000 0000.000Q 30:80 0000 0.0000002 583.4. 000000000000 00.0.0002 £00090. 00000000000 00.0.0002 30:80 0.0.0000 0000.000Q .2 00 .m A5803: MENSM 000003.— 00000000 0.0000002 000000000 0000000000000 A€8w 0.10). P content in the P enriched treatments was marginally higher than in the N-enriched treatments (P=0.062). 50 r i r 1 .1 r l 40 1 l 1 4O : 3O .. 37°28 -?0 ~92” ' if, 510.r 1910 ' =10 3° 5 29*- :s 3 ‘2 E E : 1% Z (5.21? 12 "‘ : : C N P NP C N P NP C N P NP Figure 4.3. Algal cellular nutrient contents and molar N:P ratios in control, N-enriched, P-enriched, and NP-enriched treatments on artificial substrata. Low cellular N:P ratios on clay pots indicated that algal growth would be N limited (Figure 4.3). Molar cellular N:P ratio averaged 4.490 (21:0.686) in the control treatment, 11.148(:l:2.163) in the N-enriched treatment, 5.584(i1.614) in P-enriched treatment, and 9.971(i2.163) in NP treatment. Cellular N:P ratio in N- and NP-enriched treatments were significantly higher than control (P=0.004 and 0.016) and P-enriched (P=0.042 and 0.059) treatments. N:P ratios were not less in P-enriched than control treatments. 98 h Chl a (ug/cmz) 5 c3, 9 Algal densrty (lemz) 9%. .22» ..3 3 c - Z '——l:l:l—‘ AF‘DM (mg/erg?) a 5 o - z —1:1:1—- : t r r I l 1 103 ‘ l I C N P NP A C P NP C P NP 100 I r N I I r l I NA F 1&0"! 1. .. .. 5101'F i 5101: :A10 E 3 >1 : : u "I I” : 310-2- ‘ E s 35 0 I I 3 —3 h E : : O10 I IE 0- . :1 _; > : : 10 310'! . 1. .2 - N > 5 E mloir 15 .2104! E E 50 m : I O ' ‘ 10 L r r r ‘ E105!- r r r 1 10-1 1 r r - C N P NP 9 C N p NP C N P NP Figure 4.4. Algal biomass/production in artificial substrata of control, N-enriched, P- enriched, and NP—enriched treatments. They are represented by different attributes. Differences in algal biomass on nutrient diffusing substrata also indicated that nitrogen was overall the most important limiting nutrient in the 22 restored Michigan wetlands (Figure 4.4). Chl a, algal cell density, and algal biovolume were all significantly higher on N- and NP-enriched substrata than on control and P-enriched substrata (P<0.05). DM and AFDM responses to N enrichment were not as great as chl a, cell density, and biovolume. DM with N and NP additions was significantly greater over the control treatment (P=0.001 and 0.026 respectively) but not among other treatments. AFDM in the NP enriched treatments was higher than in control, P-enriched, and N- enriched treatments (P<0.01). No differences between the P-enriched treatments and control treatments or between NP- and N-enriched treatments were found (Figure 4.4). Little evidence of secondary P lirrritation was found. 99 Chl a: DM and Chl a: AFDM ratios were different among different treatments (Figure 4.5). N- and NP- enriched treatments had significantly higher Chl a :DM ratios than control and P-enriched treatments (P<0.05). N-enriched treatment also had higher Chl a: AFDM ratio than the control, P-enriched, and NP-enriched treatments. chl a : AFDM in the NP—enriched treatment did not differ from control treatment, but was higher than the P-enriched treatment. .. I i l I ‘ 100F 1 ’53 E =1: 5 E . w . t? 102' 2 E 1 a . . 52 a 1P i E E * E U i t . f - C N P NP l lllJ ..‘L ’67) E 1.0? l 1 °° E — 5 § : : a T I . 26 * ‘ 50.1:- '5 l I l - C N P NP Figure 4.5. Chl a: AFDM and Chl azDM ratios in control, N-enriched, P-enriched, and NP-enriched treatments on artificial substrata. Table 4.2. Algal nutrient limitation and corresponded water nutrients and sediment nutrients. Value in the parentheses are correlations between log-transformed value and algal nutrient limitation. A * indicates significant correlation (Dunn-Sidak P<0.05) Observed Correlation water nutrients sediment nutrients N P DlNzTP N:P N P N:P Primary N (N-C)/C 0.034 0612* 0.437 0.297 0.199 0.634- 0.378 lirrritation (0.122) (0592') (0546*) 00.601") (0.082) (0.150) (0.430) Primary P (P-C)/C 0.121 0.239 0.203 0.300 0.141 -0.089 0.137 Limitation (-0.069) (0.253) (0.237) (-0.282) (0.138) (0.043) (0.089) Secondary N (NP-P)/P 0.143 0.062 0.139 0.014 0.181 0.153 0.007 Linritation (0.028) (0.131) (0.049) (0.1 15) (-0019) (-0.106) (0.019) Secondary P (NP-N)/N 0.116 0.049 0.168 0.180 -0.064 0.239 0.309 Limitation (0.141) @049) (0.011) (0.115) (0.206) (0567*) (0.382) 100 The magnitude of N - and P-lirrritation in individual wetlands also showed that N limitation was consistently greater than any evidence of P limitation (Figure 4.6). Several wetlands showed some signs of P-linritation. However, the magnitude of N limitation, ranging from 0-13.5 units was much greater than the 0-2.5 units range in P-limitation. 15 I I I O 0 Figure 4.6. Comparisons of algal N and P O 10 ' . , ‘ limitation in all 22 wetlands. Primary N or P G . limitation is defined as algal biomass of that é 5 . : . nutrient treatment minus control treatment then 8 ‘ . weighted by control treatment. Primary nutrient 3:3: 5:. linritation is defined as algal biomass of that 'g 0 " 0‘ ' nutrient treatment minus control treatment then '23 weighted by control treatment. _5 I l I -5 0 5 10 15 P limitation (P-C/C) Primary limitation of algal accrual by N supply was related to nutrient concentrations and ratios both in the water column and sediments (Table 4.2, Figure 4.7). Direct correlation between N-lirrritation and DIN concentration was very weak. However, the magnitude of N-limitation in a wetland was positively correlated with TP both in the water and in the sediments, negatively correlated with DlNzTP and N: P ratios in the water column, and negatively correlated with N:P ratio in the sediments (Table 4.2, Figure 4.7). Secondary P limitation was negatively correlated with TP concentration in the sediments (Table 4.2). 4.4.3 Algal species composition in natural and artificial substrata Changes in species composition among wetlands indicated nutrients were an important factor regulating algae in epiphytic habitats. A total of 115 diatom taxa out of 213 algal taxa were found in epiphytic samples. Epiphytic non-diatom algae were 101 predominated by Oedogonium, Mougeotia, Scenedesmus, Pediastrum, and Trachemonas species, while Navicula minima, Achnanthes minutissima, and Navicula cryptocephala were the dominant diatom species. CCA based on epiphytic diatom assemblages and environmental variables showed that 31.4 % of the total variance of diatom assemblages were explained by the first three CCA axes (Table 4.3). Epiphytic diatom assemblages were closely correlated with pH (P=0.005). Conductivity (P=0.07), DIN (P=0.075), and TP (P=0.015) were the other important environmental factors with high eigenvalues explaining diatom distribution (Table 4.3, Figure 4.8). 30 I f 1 y 20 I I I I I 20 Pearson r=-0.199 Pearson 1:0608 20- ° - . ' o 10 h. 0 I 10 z. 0 1- C‘ :‘ 0 .0 a O h T O ‘E e g _10 r r 1 1 r r 1 r r E 0.0 0.5 1.0 1.5 2.0 25100 1 2 3 4 5 6'10 ' TN 1n water(mg/L) TP in water(mg/L Z 30 r ' ' ' 20 I I I I a 1 20 t: Pearson r=-0.034 Pearson 1:0.646 o 20- . d O . . ' 10 - o '. - 10 10’. . '1 0.0 . o . .0 ‘ 1? ' ' t . . " 0... J... 5 o .. O O - o 10 20 30 4o 50'100 10 20 3O 40 50 60 7010 TN in sediment (pg/mg) TP in sediment(ug/ mg) ' 1' "I'." V ' "U'VI' Pearson r: - 0.591 0.1 1.0 10.0 N :P in water I FTTI Pearson r=-0.454 1 LI]! 2 4 8 12 N:P in sediment Figure 4.7. Comparisons of algal N and P limitation in all 22 wetlands. Primary N or P limitation is defined as algal biomass of that nutrient treatment minus control treatment then weighted by control treatment. Changes in algal assemblages of sediment also revealed that nutrients regulated species composition. The majority of algal taxa (173) found in sediment samples were 102 diatom taxa (148). Non-diatom algae were mainly composed of Oscillatoria and Scenedesmus, while diatom communities were dominated by N. minima, A. minutissima, N. cryptocephala, and Achnanthes lanceolata var. frequentissima. CCA related changes in diatom assemblages to pH and N gradients (Figure 4.8). The first three ordination axes of CCA explained 27.4% of diatom taxa data. The first axis was correlated with pH (P=0.01), and the second axis was positively correlated with DIN (P=0.04)(Table 4.3). Table 4.3. Weighted correlation matrix from Canonical Correspondent Analysis (CCA). Epiphytes Sediment phytoplankton Artificial substrata Axisl AxisZ Axis3 Axisl Axis2 Axis3 Axisl AxisZ Axis3 Axisl AxisZ Axis3 Variance 14.3 9.2 7.9 11.5 8.2 7.7 9 8.5 7.9 14.3 9.0 8.0 pH 0.596 -0.453 0.128 -0.655 -0.026 -0.330 0.433 0.090 0.040 -0.509 -0.245 0.609 Conductivity -0.036 0.497 -0.208 -0.248 0.031 -0.482 -0.099 -0.076 -0.385 -0.244 0.746 0.261 Canopy Cov. -0.259 0.338 0.052 0.088 0.325 0.148 -0.075 0.338 -0.045 -0.073 -0.286 0.456 DIN -0.308 0.363 -0.520 -0.314 0.577 -0.511 -0.189 -0.044 0.084 -0.261 0.390 0.378 SRP -0.454 -0.094 0.135 0.123 -0.331 -0.156 -0.460 0.223 -0.250 -0.230 0.277 0.160 TN -0228 -0.201 -0.502 0.193 -0.250 -.0118 -0.353 -0.040 0.304 -0.230 0.277 0.160 TP -0.331 -0.008 -0.654 0.216 -0.034 0.296 -0.415 0.660 0.051 -0.115 0.104 0.492 TNzTP 0.289 0.374 -0.052 -0.066 -0.144 -0.115 -0.022 0.022 0.233 -0.125 -0.620 -0.421 Sed.N 0.138 -0.062 -0.117 0.013 -0.202 0.283 -0.456 0.005 0.304 -0.172 -0.132 0.250 Sed.P 0.218 0.335 0.152 -0.407 -0.l48 -0.416 -0.293 0.045 0.304 0.268 0.167 0.175 Phytoplankton assemblages were also affected by nutrient factors. Phytoplankton samples were very low in algal species and cell density in most wetlands. Only 99 non- diatom taxa were found. Swimming or floating genera, such as Scenedesmus, Pediastrum, Euglena, Trachelomonas, and Oscillatoria appeared in some of these sites, but none of them dominated in all wetlands. Diatom density was extremely low so they were not used for further analysis. CCA of non-diatom assemblages related changes in non-diatom species composition to TP (Figure 4.8). The first three ordination axes of CCA explained 25.4% of diatom taxa data. The only significant environmental variable was TP (P=0.03), which correlated with the second CCA axis (Table 4.3). 103 epiphytes DIN 3 sediment Conductivity DIN:SRP . CMMICROC . NAPUPREC NACAPITA DIN CABACILL GOCLAVAT . Sed. N ACLANCEO ' . ACMINUT ACHUNGAR ° ' ° ° Conductivity ° . .' 0 Axis L NALIBONE ' °-. '. Axisl ‘ V r . U . r ACEXIGUA - RPGIBBA ' NIPALEA NAERIFUG . . '- EUBILUNA o . . o NAVENE‘I‘A . EUBILUNA oooRACII.NALIBO PH FRCAPGRA O FRCAPMES .SYRUMPEN CI'H'HCROC ' DO NIPALDEB PH N o 24 C d “M 5 artificial substrata '11) AK AB :5 phytoplankton 0“ “ 9' CMMICROC NIRADICU ‘ .NIPALDEB NAVENETA . GOPARVUL . . TP' - ACHUNGARA“:l NASUBMIN . .0 0. . - NIPALEA . , ° ACLANCEO RPGIBBA CABACILL N:P Figure 4.8. CCA biplots of algal assemblages and environmental variables. Diatom taxa on macrophytes, in sediments, and on artificial substrata in control treatments, and algal genera of phytoplankton were used for the CCA. Codes for diatoms: ACEXIGUA- Achnanthes exigua, ACI-IUNGAR- Achnanthes hungrarica, ACLANCEO- Achnanthes lanceolata, ACMINUTI— Achnanthese minutissima, CABACILL- Caloneis bacillum, CCPLALIN- Coconeis placentula var. linearis, CMMICROC-Cymbella microcephala, E UBILUNA- Eunotia bilunaris, FRCAPGRA- F ragilaria capucina var. gracilis, FRCAPMES-Fragilaria capucina var. mesolopta, GOCLA VAT-Gomphonema clavata, GOPARVUL-Gomphonema parvulum, GOGRACIL—Gomphonema gracilis, NACAPITA- Navicula capitata, NACRYPTO- Navicula cryptocephala, NACRYTEN- Navicula cryptotenella, NAERIFUG- Navicula erifuga, NALIBONE- Navicula libonensis, NAMINMA- Navicula minima, NAPUPREC-Navicula pupula var. rectangularis, NA VENETA- Navicula veneta, NIFRUSTU-Nitzschiafrustulum, NIPALEA-Nitzschia palea, NIPALDEB-Nitzschia palea var. debilis, NIRADICU- Nitzschia radicula, RPGIBBA-Ropalodia gibba, S YRUMPEN-Synedra rumpens. Code for phytoplankton genus: AB-Anabaena, AK- Ankiostrodesmus, DI-diatoms, CE—Coelastrum, CF- Cladophora, CQ—Coelosphaerium, DB- Dinobryn, GL—Glenodinium, KR- Ceratium, MB -Micrasterias, OF- Ophiocytium, OS-Oscillatoria, PA- Pediastrum, PV- Pandorina, SC- Scenedesmus, XA- unknown algae. 104 Only 137 non-diatom taxa and 113 diatom taxa were found on the control artificial substrata. About half of the wetlands were dominated by Stigeoclonium, Oedgonium, and Mougeotia green macroalgae. The others were mostly diatom species, predominantly N. minima, A. minutissima, and N. cryptocephala. CCA indicated that 31.3% of variance in the diatom taxa data was explained by the first three CCA axes. Diatom assemblages in control substrata were correlated with pH (P=0.005), conductivity (P=0.015), N: P (P=0.06), and TP (P=0.065) (Table 4.3). 100 I 1 I 1 1 I Figure 4.9. Box-plots of 90 _ _ similarities of diatom assemblages between different 80 ' ‘ habitats and treatments. E- 70 __ __ epiphytes, S-sediment, C- L i control, N-N treatment, P-P g; 60 ' ‘ treatment, NP- NP treatment. X- ‘7,’ 50 _ _ axis shows the comparison of (L __ two paired habitats. Y-axis 40 ' 1" “ shows the Percentage similarity 30 .. '1' .. between the two habitats. :1: I - 1 0 1 1 1 1 1 1 E28 C:E C:S C:N C:NP C:P Substrate comparison Different substrata might have different effects on diatom species assemblages, however, the similarities in species composition on natural habitats, epiphyton (E) and sediments (S), was not significantly different than between natural habitats and control treatments (Figure 4.9). The average similarities between epiphytic and sediment diatoms, and epiphytic and control treatment diatoms, and sediment and control treatment diatoms (0.444(:|:0.053), 0.442 (21:0.038), and 0.474 (i0.034) respectively) were not 105 significantly different (P>0.05). Diatom species assemblages of all treatments on artificial substrata were more similar (P<0.05) to each other than with those on natural substrata (Figure 4.9). 4. 4.4 Algal assemblage response to nutrient additions Both diatom and non-diatom algae increased in density and biovolume in N -enriched treatments and NP-enriched treatments (P<0.01). The most dominant algal taxonomic group in control treatments was Chlorophyta (45.2 :1: 0.077%), followed by diatoms (21.9: 6.2%), Euglenophyta (15.7i 5.9%). Only a small proportion of the biovolume was cyanobacteria (5.6 i- 2.7%). N-enrichment did not change the relative biovolume of diatoms (29.6 i 6%) and filamentous green algae (48.51 7.6%) overall, but the relative biovolume of cyanobacteria (0.9: 0.6%) was significantly lower in the N-enriched treatment than in the control treatment (P<0.05). P-enrichment did not change the relative biovolume of any of the algal groups. NP-enrichment also decreased the relative biovolume of cyanobacteria (1.9i 0.9%). The most dominant algal genera, Oedogonium, Mougeotia, and Stigeoclonium, did not respond to nutrient addition based on relative biovolume (P>0.1). although they all increased in total biovolume and density with nitrate addition. The most abundant diatom taxa responded to nutrient additions in a variety of ways on artificial substrata (Table 4.4). The relative abundance of the most abundant taxa, such as N. minima, A. minutissima, A. exigua, C. placentula var. lineata, and N. frustulum did not change with nutrient addition. The relative abundance of G. parvulum, and N. cryptocephala increased with nitrogen addition (P<0.1), while the relative abundance of 106 N. palea and N. palea var. debilis decreased. The relative abundance of A. lanceolata and A. lanceolata v. frequentissima increased with P enrichment. Table 4.4 Twenty-six most dominant taxa and their responses to nutrient additions. The 4-digit number represents comparisons of the relative abundance of a diatom taxa in a treatment with control, N-enrichment, P-enrichment, and NP-enrichment treatments. A significant change (P<0.l) of relative abundance between treatments was presented by 1, and insignificant effect was 0. * is empty when compared with itself. Diatom taxa N-enriched P-enriched NP-enriched Indicator Achnanthes exigua Grun. 0*00 00*0 000* - Achnanthes hungarica (Grun.) I *00 00*0 000* N Grun Achnanthes lanceolata (Breb.) 0*00 10*0 100* P Grun. Achnanthes lanceolata v. 0*00 10*0 000* P frequentissima Lange-Bert. Achnanthes minutissima Ktltz. 0*00 00*0 000* - Amphora veneta Kutz. 0*00 00*0 000* - Cocconeis placentula v. lineate 0*00 00*0 000* - (Ehr.) V. H Cymbella microcephala Grun. 0*00 00*] 101 * N Epithemia sorex Kutz. 0*00 00*0 000* - Eunotia billularia (Ehr.) Mills 0*00 00*0 000* - F ragillaria construens(Ehr.) 0*00 00*] 001 * N Grun. F ragillaria capucina v. gracilis 0*00 00*0 000* - (08m) Hust. Gomphonema clavatum Ehr. I *01 11 *1 001 * N and P Gomphonema gracile Ehr. 0*00 00*0 000* - Emend. V.H. Gomphonema parvulum Kutz. 1*00 00* 1 001* N Navicula capitata Kutz 0*00 00*0 000* - Navicula cryptocephala Kutz I *00 00*0 000* N Navicula minima Grun. 0*00 00*0 000* - Navicula semilunum Grun 0*10 11 *1 001 * N Navicula trivialis Kutz 0*00 00*0 000* - Navicula veneta Kutz 0*00 00*0 000* - Nitzschia amphibia Grun. 1 *00 10*0 000* N Nitzschia frustulum (Kutz) 0*00 00*0 000* - Grun. Nitzschia palea v. debilis 1 *00 01 *0 000* P* (Ktltz.) Grun. Nitzschia palea (Kuetz.) Grun. I *00 00*0 000* P* Nitzschia perminata (Grun.) 0*00 00*] 001 * P Peragallo 107 A number of diatom species distributed along the nutrient gradients in CCA ordinations on artificial substrata, macrophytes, and sediments (Figure 4.8). On macrophytes, Cocconeis placentula dominated in high DIN environments and Nitzschia palea var. debilis, Synedra rumpens. and Navicula libonensis dominated in low DIN environments. Cymbella microcephala preferred high DIN: SRP environments but F ragilaria capucina var. mesolopta dominated in low DIN:SRP environments. Achnanthes exigua and Eunotia bilunaria were abundant in high TP environments. In sediment habitats, Navicula capitata, Gomphonema clavatum, Navicula pupula var. rectangularis appeared in high DIN environments while F ragilaria capucina var. gracilis and E. bilunaria dominated in low N conditions. On artificial substrata, Navicula cryptonella and Caloneis bacillum showed preference for high N: P environments while N. palea var. debilis for low N: P environments. Table 4.5. Correlations between relative abundance of nutrient indicator species and nutrient concentration in different habitats. * indicates significant correlation (P<0.05). DIN SRP TN TP Sed. N Sed. P N indicator on artificial substrata 0.310 0.236 0.034 -0.105 0.044 0.123 N indicator on macrophytes -0.004 0.041 0.339 -0.137 -0. 122 -0.350 N indicator in sediment 0.168 0.006 0.299 -0.068 -0.208 -0.079 P indicator on artificial substrata -0. 165 0.373 0137 0.562* -0.045 0.032 P indicator on macrophytes -0.241 0.330 -0.220 0.470* -0.163 -0.115 P indicator in sediments -0.019 0.282 -0.11 1 0.437 * -0.040 -0.076 Although individual taxa responded to a nutrient gradient differently, the indicator species (Table 4.4) delineated from experimental nutrient additions correlated with natural nutrient concentrations of wetlands (Table 4.5). Indicator diatom species for N and P were listed in Table 4.4. Relative abundance of N indicator species on artificial substrata weakly correlated with DIN concentration (r=0.331) in water column (Table 108 4.5). N indicator species on macrophytes and sediments weakly correlated with TN concentrations in water column. P indicators on all substrata were consistently correlated with TP concentrations in water column (Table 4.5). 4.5 Discussion 4.5.] Algal nutrient limitation Four lines of evidence indicated that N was the primary limiting nutrient for periphyton growth in restored wetlands in southern Michigan. First, ambient NO3-N concentration and N :P ratios were low. Second, N and NP additions caused increased periphyton cellular N content and cellular N:P ratios in the wetlands, which indicated a physiological N limitation for algal cells. Third, the highest algal biomass on N and NP enriched substrata revealed that N could stimulate algal biomass accrual in these wetlands. Fourth, this N limitation was 5 times stronger than any indication of P limitation in these systems. The low DIN concentration, especially depleted NO3'-N concentration (<1011g N03'- N/L), in most of the wetlands indicated strong N-lirrritation. Previous studies have found that periphyton growth could be limited when ambient levels were 50-90 11g NO3-N/L (Grimm and Fisher 1986, Lohman et al. 1992, Stelzer and Lamberti 2001), while Maberly et al. (2002) found that nitrogen limitation was more likely than phosphorus limitation where the DW was <65 pmol IL (7111g NIL) . The restored wetlands in southern Michigan were generally depleted in N03'-N, but not N Hf. Several possible biogeocherrrical processes might have led to the low NO3'-N concentrations in these wetlands. One possible cause was sequestering of nitrogen by duckweed, which can compete with phytoplankton for nutrients and deplete DIN supply (Pokomy and 109 Rejmankova 1983, Oscarson et al. 1989). Denitrification also can be a significant path of nitrogen loss, especially N031 In duckweed covered wetlands, low light irradiance and high decomposition rate led to low DO concentration. The low dissolved 02 could have limited nitrification rates and caused a relatively high NH]: N03 ratio. Although nutrient limitation is presumably related to the concentration of that limiting nutrient, the magnitude of nutrient limitation was not negatively correlated with DIN or TN concentrations. It was negatively correlated with TP concentrations in southern Michigan wetlands. Studies of nutrient addition experiments in New Zealand streams (Francoeur et al. 1999) indicated that the degree of nutrient limitation (mostly N limitation or NP co-limitation) was negatively correlated with NO3-N concentration when NOg' ranged from 3.2 —367p.g/L. NO3'-N concentrations in our wetlands were extremely low or depleted. The absence of a gradient of N concentrations could have caused the lack of a possible relationship between N and the magnitude of nutrient limitation. The increased TP concentration increased the difficulty for periphyton to absorb NO3-N from the water column and led to greater N-lirrritation with increased P gradient. In high ammonium wetlands, duckweeds covered the water surface and light may have been a limiting factor that prevented NH4+ utilization. Algal biomass accrual was barely detectable in these wetlands. N :P ratios in the water column and in sediments were much lower compared with the Redfield ratio, the optimal N:P ratio for periphyton growth (Redfield 1958, Tilman 1977, Fairchild 1985, Sommer 1993, Vymaza 1995, Hillebrand and Sommer 1999, Guildford and Hecky 2000, Stelzer and Lamberti 2001). In a review study on nutrient limitation of phytoplankton in lakes and oceans, Guildfold and Hecky (2000) concluded that the 110 absolute value of TN might not be a strong predictor of N limitation, the TN:TP ratio might be more useful determining the potentially limiting nutrient. However, Stelzer and Lamberti (2001) found that the overall abundance of lotic periphyton responded positively to increased total nitrogen concentration, but not to the N:P ratio. Apparently, nutrient concentration and nutrient ratios could both play important roles in different systems. Increased cellular N content and N :P ratio after N addition is a strong evidence for N limitation of algal accrual. Algal cellular N:P ratios in our control treatment were much lower than the Redfield Ratio. Increased N availability significantly increased cellular N content, but P addition did not change the P content which indicated that the ambient P concentration in the water column was saturated for algal growth. Previous studies showed that N and P enrichment would change N:P ratios and N and P contents in algal tissues (Vymazal et al. 1994, Vymaza] and Richardson 1995, Hillibrand and Sommer 1999, McCormick et al. 2001, Hillibrand and Kahlert 2001, Stelzer and Lamberti 2001). In the P-lirrrited Everglades, algal N:P ratio was higher than 100:1 (McCormick et a1. 2001) and increased P loading increased P content in the algal periphyton. Experimental manipulation of stream periphyton growth under different nutrient concentrations and N:P ratios revealed that periphyton N and P content increased with the N and P concentration of stream water (Stelzer and Lamberti 2001), and the cellular N: P ratio significantly followed the N:P ratio in the stream water. Algal chl a concentration, biovolume, and cell density increased with N and NP enrichments in restored Michigan wetlands, but DM and AFDM showed different response. This difference can be found by examining the chl a to AFDM and DM ratios. 111 My results showed that N-enriched periphyton was characterized by higher chl a to AFDM ratios than N-limited periphyton, and NP-enriched periphyton had lower chl a : AFDM than N enriched treatment. Previous studies also found the same results (Grirrun and Fisher 1986). Triska et al. (1983) found no difference in AFDM between N enriched and control treatments despite large differences in chl a because of higher chl a to AFDM ratios of N-enriched periphyton. Healey and Hendzel (1979) proposed that chl a to biomass ratio was a reasonably good indicator of N deficiency for several algal species because chl a synthesis often was reduced or ceased in N -deficient algal cells. Such cells tend to accumulate carbon storage compounds (Healey 1973). Grimm and Fisher (1986) attributed low chl a :AFDM ratio to earlier development of a thick layer of organic material (detrital accumulation) on the substratum surface in N-limited substrata. Weber (1973) proposed that the autotrophic index should be calculated based on estimates of periphyton chl a and AFDM. Apparently, N enrichment could stimulate high algal accumulation rather than high organic matter accumulation in P additions. Competition for nutrients in N-lirrrited substrata may have stimulated heterotrophic microbial processes (Tank and Webster 1998). Nutrient limitation is a complicated process and could be easily confounded by other factors. Algal nutrient limitation could be affected by many biotic factors (herbivory, resource competition) and abiotic factors (light, hydrology, and substrata) (Hill and Knight 1988, Munn et al. 1989, Hansson 1992, Cronk and Mitsch 1994, Toetz 1999, Lamed and Santos 2000, Rosemond et al. 2000). These factors sometimes confound change in periphyton growth with nutrient addition (Lowe et al. 1986, Hill and Knight 1988, Millard et al. 1996). The wetland systems in this study had high duckweed cover 112 and low phytoplankton density. Low algal accrual on control and nutrient enriched treatments indicated that light limitation and grazing were probably important in these wetlands. Despite the possibility that these confounding factors might have reduced the significance of algal nutrient limitation, nutrient diffusing bioassays still showed strong responses of algal biomass and algal assemblages with N additions. 4.5.2 Algal species composition and nutrient concentrations Although different habitats within wetlands and pH variation among wetlands affected algal species composition, the response of algal species to nutrient and nutrient ratios were consistently observed in all three CCA in our study. Previous studies in lakes and streams have indicated that substrata type (Pringle 1990, Cronk and Mitsch 1994, Tank and Webster 1998), pH (Dixit et al. 1992), and nutrients (Schindler 1987, Elser et al. 1990, Stevenson 1998, Leland et al. 2001) are three important factors affecting algal species composition. Algal communities on different substrata had different algal assemblages. However, the similarities of diatom communities between artificial substrata with sediment and epiphytic habitats were not different than the similarities of algal assemblages between natural (sediment and plant) habitats in this study. Although algal species composition in different habitats can respond to nutrients in different ways (Pringle 1990), pH was the most important factor regulating diatom assemblages on all three substrata in my study, and in a number of studies (Dixit et al. 1992, Lindstrom 1996, Stevenson and Pan 1999). Although algal response to nutrient gradients in wetlands was not as strong as the response to pH and conductivity (Pan et al. 1996, Stevenson et al. 1999), TP and N:P ratio are also important factors regulating algal assemblages in both regional surveys and experimental manipulation of nutrient additions (McCormick et al. 113 1996, 2001, Pan et al. 2000). Similarly, this studies indicated nutrient concentrations and ratios could both be important in determining diatom species composition in restored wetlands. One of the most noticeable changes of algal groups on artificial substrata was that cyanobacteria decreased in relative biovolume with N addition. The property of cyanobacteria fixing N and dominating in N deficient ecosystems have been noted by many studies (Rader and Richardson 1992, McCormick et al. 1996, 2001). Most studies indicated that N-enrichment could decrease bluegreen biomass, and low N environment stimulated cyanobacteria dominance. Marinho and Huszar (2002) reported that cyanobacteria abundance was negatively associated with NO3' and N/P ratio, and positively associated with temperature and light in a reservoir. My N- and NP-enriched treatments did decrease the abundance of cyanobacteria. However, our wetland system did not have high cyanobacteria biovolume. Green algae (primarily Stigeoclonium and Oedogonium) and diatoms dominated in the wetlands. Even though cyanobacteria were supposed to dominate to compensate N-linritation in N-limited systems (Smith 1986, Vitousek et al. 2002), low light levels and low temperatures in northern (high latitude) wetlands could have limited cyanobacteria dominance (Watermann et al. 1999, Sekar et al. 2002). Cyanobacteria abundance and N fixation could also have been constrained by trace-element limitation and/or grazing even when N:P ratios were low (Howarth et al. 1999). Individual diatom taxa responded to N and P additions on artificial substrata in a similar or different manner to their response along natural gradients. Some species showed exactly the same response to a nutrient addition as I observed on natural substrata 114 and controlled artificial substrata. For example, N. palea and N. palea var. debilis were found in low N and high P conditions on natural substrata as well as previous studies (Fairchild et al. 1985). This study proved that N enrichment decreased the relative abundance of these taxa and P addition did not change their abundance in P-enriched environments. G. clavatum and C. microcephala were also found in high DIN environments in natural habitats, and N addition also increased their relative abundance on artificial substrata. On the other hand, several species that appeared in the high or low end of nutrient gradients did not change significantly with nutrient addition. For example, Caloneis bacillum appeared in high DIN: SRP environments but N addition did not significantly increase its relative abundance. A. minutissima, which was reported to positively respond to N addition (Fairchild et al. 1985), did not respond to nutrient addition in this study. Pringle (1990) reported that A. minutissima was significantly less for all nutrient treatments relative to the control on sand-agar. A lack of pristine P environments may have caused insensitive response from species responding both to N and P additions and thus may have confounded the effect of N addition. Complicated wetland environments such as different pH among wetlands and grazing may have confounded individual species response to nutrient additions (John 1993, Rosemond et al. 1993). Little evidence has quantitatively related experimental response of diatom taxa to their response in natural habitats. Diatom species responses to natural nutrient gradients have been determined and indicator values have been assigned for many species (van Dam 1994, EPA 2002). Multispecies indicators of trophic status based on specis nutrients preferences proved to be applicable across geological regions (Kelly 1998). A few studies 115 have related natural distribution of diatom assemblages along nutrient gradients to experimental nutrient gradients (McCormick et al. 1996, 2001, Underwood et al. 1998, Pan et al. 2000). My studies indicated that despite variability among taxa, the overall response of the sensitive diatom taxa was similar for natural and experimental nutrient gradients. However, the response of N indicator species were not as strong as P indicator species. This may have been due to the lack of immediate N supply and N gradients in these wetlands resulting N indicator species not responding to N addition in a gradual way. Although the correlations between experimental results and natural gradients were weak, they provided a pioneering attempt to identify indicator species experimentally that could be used for wetlands. 4.6 Conclusions Algal growth in the restored wetlands of southern Michigan was limited by N supply. Cellular nutrient contents and N:P ratios, algal biomass accrual response to a nutrient addition, and the strong N limitation vs. weak P limitation proved that N was the primary limiting nutrient. Although the magnitude of limitation was not directly correlated with N concentration in the water sediments, its strong correlation with P concentration and N: P ratio in both the water column and the sediments indicate a strong limitation caused by N depletion. Algal species composition in natural habitats also responded to nutrient gradients. Experimental enrichment changed relative abundance of many dominant species. Overall, the relative abundance of species that increased with N or P enrichment on experimental substrata also increased as N or P concentrations increased in wetlands. 116 4.7 Acknowledgments Tom Eitniear from Michigan Department of Fisheries and Wildlife Service provided valuable site information. Christian Parker participated in some of the field work and sample collection. Kalina Manoylov assisted in diatom identifications. Julie Heinlein and Christian Parker provided valuable comments on an early draft of the manuscript. 117 4.8 References American Public Health Association (APHA). 1998. Standard methods for the examination of water and wastewater. 20‘h edition by LS. Clececeri, A.E. Greenberg, and AD. Eaton. Washington D. C., USA. Bayley, S. E., Zoltek, J ., Hermann, A. J ., Dolan, T. J., and Tortora, L.1985. Experimental manipulation of nutrients and water in a fresh-water marsh - effects on biomass, decomposition, and nutrient accumulation. Limnol. Oceanogr. 30:500-512. Bedford, B., Walbridge, M. R., and Aldous, A. 1999. Patterns in nutrient availability and plant diversity of temperate North American wetlands. Ecology 80: 2151-2169. Bothwell, M. L., and Stockner,J.G. 1980. Influence of secondarily treated kraft mill effluent on the accumulation rate of attached algae in experimental continuous-flow troughs. Can. J. Fish. Aquat. Sci. 37: 248-254. Carpenter, E. R., Caraco, N.F., Correll, D.L., Howarth, R. W., Sharpley, A. N., and Smith, V. H. 1998. Nonpoint pollution of surface waters with phosphorus and nitrogen. Ecol. Appl. 8: 559-568. Carrick, H.J., Schelske, C.L., Aldridge, F.J., and Coveney, M.F. 1993. Assessment of phytoplankton nutrient limitation in productive waters - Application of dilution bioassays. Can. J. Fish. Aquat. Sci. 50: 2208-2221. Chambers, P.A., Dale, A. R., Scrimgeour, G.J., and Bothwell, ML. 2000. Nutrient enrichment of northern rivers in response to pulp mill and municipal discharges. J. Aquat. Ecosyst. Stress and Recovery 8:53-66. Craft, C. B. and C. J. Richardson. 1997. Relationships between soil nutrients and plant species composition in everglades peatlands. J. Env. Qual. 26: 224-232. Craft, C. B., Vymazal, J ., and C]. Richardson. 1995. Response of everglades plant communities to nitrogen and phosphorus additions. Wetlands 15: 258-271. Cronk, J. K., and Mitsch, W. J. 1994. Periphyton productivity on artificial and natural surfaces in constructed fresh-water wetlands under different hydrologic regimes. Aquat. Botany 48: 325—341. Dixit, S. S., Smol, J .P., and Kinston, J .C., and Charles, D. F. 1992. Diatoms - powerful indicators of environmental-change. Environ. Sci.Tech. 26: 22-33. Downing, J .A., Osenberg, C.W., and Samelle, O. 1999. Meta-analysis of marine nutrient- enrichment experiments: variation in the magnitude of nutrient limitation. Ecology 80: 1157-1167. 118 Elser, J. J ., Marzolf, E. R., and Goldman, C. R. 1990. Phosphorus and nitrogen limitation of phytoplankton growth in the freshwaters of North America: a review and critique of experimental enrichments. Can. J. Fish. Aquat. Sci. 47 : 1468-1477. Elwood, J. W., Newbold, J. D., Trimble, A. F., and Stark, R. W. 1981. The limiting role of phosphorus in a woodland stream ecosystem - effects of P-enrichment on leaf decomposition and primary producers. Ecology 62:146-158. Emery, R. M., Moon, C.B., Welch, E. B. 1973. Enriching effects of urban runoff on the productivity of a mesotrophic lake. Water Res. 7: 1505-1516. Fahnenstiel, G. L., Stone, R. A., McCormick, M. J ., Schelske, C. L., and Lohrenz, S. E. 2000. Spring isothermal mixing in the Great Lakes: Evidence of nutrient limitation and nutrient-light interactions in a suboptimal light environment. Can. J. Fish. Aquat. Sci. 57: 1901-1910. Fairchild, G. W., Lowe, R.L., and Richardson, W. B. 1985. Algal periphyton growth on nutrient-diffusing substrata: An in situ bioassay. Ecology 66: 465-472. Francoeur, S. N., Biggs, B. J. F., Smith, R. A., and Lowe, R. L. 1999. Nutrient limitation of algal biomass accrual in streams: Seasonal patterns and a comparison of methods. J. N. Am. Benthol. Soc. 18: 242-260. Genter, R. B., Cherry, F. S., Smith, ER, and Cairns, J., Jr. 1987. Algal-periphyton population and community changes from zinc stress in stream mesocosms. Hydrobiologia 153: 261-275. Goldsborough, G., and Robinson, G.G.C. 1996. Pattern in wetlands. In Algal Ecology: Freshwater benthic ecosystems. Edited by R. J. Stevenson, M. L. Bothwell, and R. L. Lowe. Academic Press, San Diego. pp. 77-117. Grimm, N. B., Fisher, S. G., and Minckley, W. L. 1981. Nitrogen and Phosphorus Dynamics in Hot Desert Streams of Southwestern USA. Hydrobiologia 83:303-312. Grimm, N. B., and Fisher, S. G. 1986. Nitrogen limitation in a Sonoran Desert stream. J. N. Am. Benthol. Soc. 5: 2-15. Guildford, S. J ., and Hecky, R. E. 2000. Total nitrogen, total phosphorus, and nutrient limitation in lakes and oceans: Is there a common relationship? Limnol. Oceanogr.45:1213-1223. Hamilton, S. 2001. Nitrogen uptake and transformation in a Midwestern U.S. Stream: A stable isotope enrichment study. Biogeocherrristry 54: 297-340. 119 Hansson, LA. 1992. Factors regulating periphytic algal biomass. Limnol. Oceanogr. 37 : 322-328. Hart, D. D., and Robinson, C. T. 1990. Resource limitation in a stream community - phosphorus enrichment effects on periphyton and grazers. Ecology 71:1494-1502. Healey, F. P. 1973. Inorganic nutrient uptake and deficiency in algae. Crit. Rev. Microbiol. 3:69-113. Healey, F .P., and Hendzel, L. L. 1979. Indicators of phosphorus and nitrogen deficiency in five algae in culture. J. Fish. Res. Board Canada. 36: 1364-1369. Hecky, RE, and P. Kilham. 1988. Nutrient limitation of phytoplankton in freshwater and marine environments: A review of recent evidence on the effects of enrichment. Limnol. Oceanogr. 33: 796-822. Hill, W. R., and Knight, A.W. 1988. Nutrient and light limitation of algae in two northern California streams. J. Phycol. 24:125-132. Hillebrand, H., and Kahlert, M. 2001. Effect of grazing and nutrient supply on periphyton biomass and nutrient stoichiometry in habitats of different productivity. Limnol. Oceanogr. 46: 1881-1898. Hillebrand, H., Durselen, C. D., Kirschtel, D., Pollingher, U., and Zohary, T. 1999. Biovolume calculation for pelagic and benthic rrricroalgae. J. Phycol. 35:403-424. Hillebrand, H., and Sommer, U. 1999. The nutrient stoichiometry of benthic rrricroalgal growth: Redfield proportions are optimal. Limnol. Oceanogr. 44: 440-446. Hooper-Reid, N. M., and Robinson, G. G. C. 1978. Seasonal dynamics of epiphytic algal growth in a marsh pond - Productivity, standing crop, and community composition. Can. J. Bot. 56: 2434-2440. Howarth, R. W., Chan, F., and Marino, R. 1999. Do top-down and bottom-up controls interact to exclude nitrogen-fixing cyanobacteria from the plankton of estuaries? An exploration with a simulation model. Biogeochemistry 46: 203-231. Kelly, M. G. 1998. Use of the trophic diatom index to monitor eutrophication in rivers. Water Res.32:236-242. Larned, S.T., and Santos, SR. 2000. Light- and nutrient-limited periphyton in low order streams of Oahu, Hawaii. Hydrobiologia 432: 101-111. Leland, H. V., Brown, L. R., and Mueller, D. K. 2001. Distribution of algae in the San Joaquin River, California, in relation to nutrient supply, salinity and other environmental factors. Freshw. Biol. 46: 1139-1167. 120 Lindstrom, E-A. 1996. The hurrric lake acidification experiment (HUMEX): Impacts of acid treatment on periphyton growth and nutrient availability in Lake Skjervatjem, Norway. Environment International 22: 629-642. Lohman, K., Jones, J. R., and Perkins, B. D. 1992. Effects of Nutrient Enrichment and Flood Frequency on Periphyton Biomass in Northern Ozark Streams. Can. J. Fish. Aquat. Sci.49:l 198-1205. Lowe, R. L., Golladay, S.W., and Webster, J. R. 1986. Periphyton response to nutrient manipulation in stream cleaning clearcut and forest watersheds. J. N. Am. Benthol. Soc. 5: 221-229. Maberly, S. C., King, L., Dent, M. M., Jones, R. I., and Gibson, C. E. 2002. Nutrient limitation of phytoplankton and periphyton growth in upland lakes. Freshw. Biol.47: 2136-2152. Marinho, M. M., and Huszar, V. L. D. 2002. Nutrient availability and physical conditions as controlling factors of phytoplankton composition and biomass in a tropical reservoir (Southeastern Brazil). Arch. Hydrobiol.153: 443-468. McCormick, P. V., Rawlik, P.S., Lurding, K, Srrrith, E. P., and Sklar, F. H. 1996. Periphyton-water quality relationships along a nutrient gradient in the northern Florida Everglades. J. N. Am. Benthol. Soc. 15: 433-449. McCormick, P.B., O’Dell, M.B., Shuford III, R.B.E., Backus, J. G., and Kennedy, WC. 2001. Periphyton responses to experimental phosphorus enrichment in a subtropical wetland. Aquat. Bot. 71: 1 19-139. McDougal, R.L., Goldsborough, L. G., and Hann, B. J. 1997. Responses of a prairie wetland to press and pulse additions of inorganic nitrogen and phosphorus: Production by planktonic and benthic algae. Archiv Fur Hydrobiologie 140: 145-167. Millard, E.S., Myles, D.D., Johannsson, OE, and Ralph, KM. 1996. Seasonal phosphorus deficiency of Lake Ontario phytoplankton at two index stations: Light versus phosphorus limitation of growth. Can. J. Fish. Aquat. Sci. 53: 1112-1124. Mulholland, P. J ., Steinman, A. D., Palumbo, A. V., Elwood, J. W., and Kirschtel, D. B. 1991. Role of Nutrient Cycling and Herbivory in Regulating Periphyton Communities in Laboratory Streams. Ecology 72:966-982. Munn, M. D, Osborne, L. L., and Wiley, M. J. 1989. Factors influencing periphyton growth in agricultural streams of central Illinois. Hydrobiologia 174: 89-97. 121 Newman, 8., Grace, J. B., and Koebel, J. W. 1996. Effects of nutrients and hydroperiod on Typha, Cladium, and Eleocharis: Implications for Everglades restoration. Ecol. Appl. 6: 774-783. Oscarson, P., Larsson, C.M., and Ingemarsson, B. 1989. Growth and nitrate uptake properties of plants grown at different relative rates of nitrogen supply. 2. Activity and affinity of the nitrate uptake system in Pisum and Lemna in relation to nitrogen availability and nitrogen demand. Plant Cell Environ. 12:787-794. Pan, Y., Stevenson, R.J., Vaithiyanathan, P., Slate, J ., and Richardson, C]. 2000. Changes in algal assemblages along observed and experimental phosphorus gradients in a subtropical wetland, USA. Freshw. Biol. 44: 339-353. Perrin, CI, and Richardson, J S. 1998. N and P limitation of benthos abundance in the Nechako River, British Columbia. Can. J. Fish. Aquat. Sci. 54: 2574-2583. Peterson, B. J ., Hobbie, J. E., and Corliss, T. L. 1983. A continuous-flow periphyton bioassay: Tests of nutrient limitation in a tundra stream. Limnol. Oceanogr. 28: 583- 591. Pokomy, J., and Rejmankova, E. 1983. Oxygen regime in a fishpond with duckweeds (Lemnaceae) and Ceratophyllum. Aquatic Botany 17:125-137. Pringle, C.M., Paaby-Hansen, P., Vaux, P. D., and Goldman, CR. 1986. In situ nutrient assays of periphyton growth in a lowland Costa Rican stream. Hydrobiologia 134: 207-213. Pringle. C. M. 1990. Effects of water and substratum nutrient supplies on lotic periphyton growth: An integrated bioassay. Can.J.Fish.Aquat.Sci.4l:1247-1251. Qualls, R. G., and Richardson, C. J. 2000. Phosphorus enrichment affects litter decomposition, immobilization and soil microbial phosphorus in wetland mesocosms. Soil Sci. Soc. Am. J. 64: 799-808. Rader, R. B. and Richardson, CJ. 1992. The effects of nutrient enrichment on algae and macroinvertebrates in the everglades: A review. Wetlands 12: 121-135. Redfield, AC. 1958. The biological control of chemical factors in the environment. American Scientist 46: 205-221. Rosemond, A. D., Mulholland, P. J., and Elwood, J. W. 1993. Top-Down and Bottom-up Control of Stream Periphyton - Effects of Nutrients and Herbivores. Ecology 74: 1264- 1280. Rosemond, A. D., Mulholland, P. J ., and Brawley, SH. 2000. Seasonally shifting limitation of stream periphyton: Response of algal populations and assemblage 122 biomass and productivity to variation in light, nutrients, and herbivores. Can. J. Fish. Aquat. Sci. 57: 66-75. Schindler, D. W. 1987. Detecting ecosystem responses to anthropogenic stress. Can. J. Fish. Aquat. Sci.44: 6-25. Sekar, R., Nair, K. V. K., Rao, V. N. R., and Venugopalan, V. P. 2002. Nutrient dynamics and successional changes in a lentic freshwater biofilm. Freshw. Biol.47: 1893-1907. Smith, V. H. 1986. Light and nutrient effects on the relative biomass of blue- green-algae in lake phytoplankton. Can. J. Fish. Aquat. Sci.43: 148-153. Sommer, U. 1993. Phytoplankton competition in plusssee - A field-test of the resource- ratio hypothesis. Limnol. Oceanogr. 38: 838-845. Stelzer, R. S., and Lamberti, G. A. 2001. Effects of N : P ratio and total nutrient concentration on stream periphyton community structure, biomass, and elemental composition. Limnol. Oceanogr. 46: 356-367. Stevenson, R. J ., Singer, R., Roberts, D. A., and Boylen, C. W. 1985. Patterns of Epipelic Algal Abundance with Depth, Trophic Status, and Acidity in Poorly Buffered New- Hampshire Lakes. Can. J. Fish. Aquat. Sci.42:1501-1512. Stevenson, R. J. 1998. Diatom indicators of stream and wetland stressors in a risk management framework. Environ. Monit. Assess.51: 107-118. Stevenson, R. J ., Sweets, P.R., Pan, P., and Schultz, R. E. 1999. Diatom community patterns in wetlands of Michigan (USA) and their use as indicators of ecological conditions. In Proceedings of INTECOL’s Vth International Wetland Conference. Edited by A. J. McComb and J. A. Davis. Gleneagles Press, Adelaaide, Australia. pp. 5 17-527 . Stevenson, R. J ., and Pan, Y. 1999. Assessing environmental conditions in rivers and streams with diatoms. In The diatoms: applications for the environmental and earth sciences. Edited by .E. F. Stoermer and J. P. Smol. Cambridge University Press, Cambridge, UK. pp 11-40. Stosch, H. A. von, and Reimann, B.E.F. 1970. Subsilicea fragilarioide gen. Et spec. nov., eine Diatomee (Fragilariaceae) rrrit vorwiegend organischer membran Nova Hedwigia Beih. 31: 1-36. Svengsouk, L. J., and Mitsch, W. J. 2001. Dynamics of Mixtures of Typha latifolia and Schoenoplectus tabemaemontani in nutrient-enrichment wetland experiments. Am. Midl. Nat. 145: 309-324. 123 Tank J - L., and Webster, J. R. 1998. Interaction of substratum and nutrient availability on wood biofilm processes in streams. Ecology 79: 2168-2179. ter Braak, C.J.F., and Srrrilauer, P. 1997. Update Note: CAN OCO v 4.02. Ceter for biometry Wageningen, CPRO-DLO, Wageningen, The Netherlands. Tilman, D. 1981. Tests of resource competition theory using four species of lake Michigan algae. Ecology 62: 802-815. Tilman, D., Kiesling, R., Sterner, R., Kilham, SS, and Johnson, RA. 1986. Green, b1 uegreen and diatom algae: Taxonomic differences in competitive ability for phosphorus, silicon and nitrogen, Arch. Hydrobiol. 106: 473-485. Toetz, D. W. 1999. Multiple limiting nutrients in a subalpine stream, Colorado Front Range. J. Freshw. Ecol. 14: 349-355. ' Tl‘i Ska, F. J ., Kennedy, V. C., Avanzino, R.J., and Reilly, B.N. 1983. Effect of simulated canopy cover on regulation of nitrate uptake and primary production by natural periphyton communities. In Dynamics of lotic ecosystems. Edited by TD. Fontaine and S. M. Bartell. Ann Arbor Sci Publ., Ann Arbor, MI. pp 129-159. U - S - EPA. 2002. Methods for evaluating wetland condition, #11 Using algae to assess environmental conditions in wetlands. United States Environmental Protection Agency, Washington, DC, USA. EPA-822-R-02-021. Underwood, G. J. C., Phillips, J ., and Saunders, K. 1998. Distribution of estuarine benthic diatom species along salinity and nutrient gradients. Eur. J. Phycol. 33:173-183. Van Dam, H., Mertens, A., and Sinkeldam, I. 1994. A coded checklist and ecological indicator values of freshwater diatoms from the Netherlands. Neth. J. Aquat. Ecol. 28:117-33. Vitousek, P. M., Cassman, K., Cleveland, C., Crews, T., Field, C. B., Grimm, N. B., Howarth, R. W., Marino, R., Martinelli, L., Rastetter, E. B., and Sprent, J. I. 2002. Towards an ecological understanding of biological nitrogen fixation. Biogeochemistry 57:1-45. Vymazal, J, Craft, C. B., and Richardson, C]. 1994. Periphyton response to nitrogen and phosphorus additions in Florida Everglades. Archiv fuer Hydrobiologie 103: 75-97. VYmazal, J. 1995. Algae and element cycling in wetlands. Lewis Publisher, Boca Raton, FL. VYmazal, J ., and Richardson, C. J. 1995. Species composition, biomass, and nutrient content of periphyton in the Florida Everglades. J. Phycol. 31: 343-354. 124 Watermann, F., Hillebrand, H., Gerdes, G., Krumbein, W. E., and Sommer, U. 1999. Competition between benthic cyanobacteria and diatoms as influenced by different grain sizes and temperatures. Mar. Ecol-Frog. Ser.187:77-87 Weber. 1973. Recent developments in the measurement of the response of plankton and periphyton to changes in their environment. In Bioassay Techniques and Environmental chemistry. Edited by G. Glass. Ann Arbor Science publishers, Ann Arbor, Michigan. pp. 119-138. Wetzel, R.G. 2001. Limnology: Lake and River Ecosystems. Third Edition, Academic Press, San Diego. p 61. Wurtsbaugh, W. A., Gross, H.P., Budy, P., and Luecke, C. 2001. Effects of epilimnetic versus metalimnetic fertilization on the phytoplankton and periphyton of a mountain lake with a deep chlorophyll maxima. Can. J. Fish. Aquat. Sci. 58: 2156-2166. 125 Chapter 5 Assessing wetland response to land use using diatoms and paleoecological characterization of reference condition 5.1 Abstract. Diatom-based biological indicators were used to establish biological and stressor criteria for wetland protection and restoration and to indicate wetland ecosystem health. Physico-chemical characteristics and surface sediment diatom assemblages were studied in 35 palustrine wetlands in the lower peninsula of Michigan across a wide range of human disturbance gradients. Paleoecological diatom assemblages from cores and pre- settlement sediments were used to represent natural, reference conditions in wetlands. Stressor-response relations were used to delineate stressor criteria and the land use causing problems when biological conditions deviated from the "normal" range of natural conditions and when deviations were associated with stressors and land use. Canonical correspondence analysis (CCA) revealed that diatom species assemblages responded mainly to conductivity and total phosphorus (TP). Weighted average predictive models for these variables were generally successful. The similarity in species composition and number of native species in current wetlands decreased with increasing nutrient concentrations. van Dam's diatom trophic index was linearly correlated with the TP gradient. Both the weighted average TP model and van Dam's trophic index positively responded to agricultural land use around wetlands. These indices deviated from the "normal" natural condition when TP exceeded 46.6 ug/L. A threshold change in TP and biotic indices was detected when agricultural land exceeded 60% total land use. 126 5.2 Introduction Assessment of wetland systems has been increasingly considered important with the growing impact of human disturbance. About 53% wetlands have disappeared or been impaired by increasing human activities since European settlement, especially due to agricultural and urban development (Mitsch and Wilson 1996, Carpenter et al. 1998, Detenbeck et al. 1999, Freeland et al. 1999, USEPA 2000). Under most circumstances, wetlands are sinks and transformers of chemicals and nutrients. They buffer the impact of terrestrial nutrient pollution on aquatic systems (Richardson 2000) and maintain biodiversity. With the better understanding of large river, stream, and lake systems in the past decades, the importance of wetland systems has received increasing attention. Many states and environmental agencies have started to incorporate wetlands into their water quality program. The US. Environmental Protection Agency (USEPA) recently published a series of methods for evaluating the biological integrity of wetlands (e.g., van der Valk 2001). These methods recommend using a series of ecological parameters including landscape characterization, nutrient load estimates, hydrology, algae, vegetation, and soil and water chemistry. Ecological assessment should characterize condition of valued ecological attributes (VEA) by comparison to pre-defined criteria and diagnosis of stressors and human activities that threaten or impair valued attributes. Appropriate classification and definition of reference conditions could help to define the normal range of natural conditions and reduce natural variability among different systems. Stressor-response relation help to identify causes of impairment to VEA, and relate the stressors to human disturbance (land use). Combining the reference approach with stressor-response 127 relationships, appropriate biocriteria can be defined and related to stressors causing problems. Although both reference approaches and stressor-response relationships have been used for assessment of ecological conditions (Somlyody 1998, Stevenson et al. 2002), no studies have combined both reference and stressor-response relationship. Establishment of stressor criteria based on these approaches has never been reported. Reference conditions are usually defined by characterizing chemical and biological condition at least impacted sites, which are often selected based on surrounding land use. Unfortunatedly, most aquatic systems have suffered from eutrOphication by different sources, such as agriculture fertilizer and urban runoff, and finding pristine sites is unlikely (Hughes 2000). Historical records of aquatic ecosystems are well preserved in the paleoecological sediments. Shifts in species composition between historical and extant assemblages could reflect the change due to anthropogenic stresses (Wilson and Muhler 1983, Wilcox 1995). Models based on weighted average calibrations have been used by paleolimnologists to reconstruct the limnological history of lakes (Dixit et al. 1992, Cumming and Smol 1993, Hughes et al. 1993, Dixit et al. 1996, Smol et al. 1998). Thus, paleoecological records from species assemblages before European settlement could be used to establish reference conditions to indicate environmental changes before and after human disturbance. Comparison between assemblages in appropriate reference condition and impaired wetlands can help identify impaired conditions and degree of impairment in these systems. Stressor-response relationships are the foundation for evaluating ecosystems at risk and directing ecological management (Gordon and Majumder 2000). This model is based on an understanding of relationships between human disturbance, stressors, and VEAs of 128 biological communities. Assessment of aquatic communities identifies degradation of VEAs. By exploring the relationships between aquatic community and possible causes of degradation based on ecological response from indicator organisms, I can relate ecological responses of VEAs can be related to stressor gradients, and used to further refine human disturbance gradients (Lopez and Fennessy 2002). A variety of ecological attributes and indices could be used to define biological criteria when comparing with reference conditions (Hill et al. 2003). Based on pairing “stressor” variables with “response” variables, ecologically meaningful criteria can be established to protect and restore VEAs. Diatoms have been used as indicators of ecological conditions in streams and lakes (Charles 1985, Hall and Smol 1992, Dixie et al. 1996, Leland and Porter 2000), but relatively few studies have used diatoms as indicators of wetland conditions (Pan and Stevenson 1996, McCormick et al. 1996, Stevenson et al. 1999, Pan et al. 2000, Stevenson et al. 2001). Because of their great diversity and species-specific responses to varied chemical and physical conditions, diatoms should serve as valuable indicators of the heterogeneous environments in wetlands (Stevenson 1998, Stevenson et al. 2001). Diatom valves are usually well preserved in sediments because their cell walls are composed of resistant opaline silica. Therefore, diatom paleoecological assemblages represent past environmental conditions (Dixit et al. 1992). In recent years, diatom-based bioassessment has been improved and numerous indices have been developed in stream and lake habitat assessment (van Dam et al. 1994, Whitton and Kelly 1995, Hill et al. 2000, 2003). van Dam’s trophic index based on diatom autoecology (van Dam et al. 1994) is one of the most successful indices proven to be valid in inferring conditions of 129 aquatic systems in Europe. Although the diatom paleolimnological approach has rarely been used in wetlands (Gasse 1987), diatom-based inference models by weighted average calibration have been developed for conductivity and total phosphorus in Kentucky wetlands (Pan and Stevenson 1996). Diatoms thus have all the potentials for indicating wetland health. The goals of this research were to use diatoms to establish biocriteria and related nutrient criteria for wetland management. I used paleoecological information from pre- European settlement sediments to represent reference conditions in wetlands, and thus define biological criteria. Stressors were delineated by relating water chemistry to land use around wetlands. Stressor-response relations were used to identify nutrient criteria to protect the "normal" range of natural biological conditions. 5.3 Methods 5.3.1 Materials and methods Preliminary analysis (Stevenson et al. 1999) of diatom assemblages in Michigan wetlands indicated low species-environmental correlation due to the great variability among wetlands. I selected 35 palustrine wetlands (as defined by Cowardin et al 1979, Detenbeck 2001) in the lower peninsula of Michigan along a north-south and an east- west transect (Figure 5.1). Although there was seasonal and annual variation, water depth of all wetlands was less than 2 m, and all but one site had water throughout the year. 130 Figure 5.1. Distribution of the 35 wetlands sampled in fall 1994 and summer 1995 in the lower peninsula of Michigan along a north- south and an east-west transect. 80 O 80 160 Miles 55 All wetlands were visited once in fall 1994 and again in summer 1995. A sediment core was taken from a central location in each wetland. The coring methods used depended on water depth: a modified piston corer was used in waters deeper than 70 cm, and push corers (Dixit et al. 1996) were used in waters less than 70 cm. The cores, ranging in length from 20-50 cm, were cut into variable segments up to 4 cm in length. All the segments were preserved with M3 solution (APHA 1998). The top 1 cm of sediment depth was later used for analysis of surface diatom, and only fall sediment samples were included. One segment from a sediment core aged older than 200 years were used as presettlement wetland assemblages. Core samples from 14 wetlands were analyzed, and all taxa appearing in these samples were considered as native taxa. The dissected sediment samples from fall 1994 were digested for diatom enumeration using concentrated sulfuric acid and KCrO3 (Stosch and Reimann 1970). All 131 surface sediments and 14 core samples from fall 1994 were selected for analysis. Diatom samples were washed with distilled water and then mounted in Naphrax®. A minimum of 500 diatom valves were identified at lOOOx using a Nikon® Labophot-2 microscope. Diatom taxa were identified using the taxonomic monographs of Hustedt (1927-66), Kramer and Lange-Bertalot (1986-91), and Patrick and Reimer (1975). All water samples were collected from the middle of the water column. A number of water chemistry parameters were monitored in the field, including pH, specific conductance, total dissolved solids (TDS), apparent and true color, and temperature. Aerated pH was measured within 24 hours after sample collection. Water samples were preserved and analyzed on a Skalar® Sans-Plus automated water chemistry system according to APHA methods (1998) for total phosphorus (TP) and total nitrogen (TN), soluble reactive phosphorus (SRP), nitrate, ammonium, and chloride. In addition, a known volume of water was filtered through a 0.45-um Millipore® membrane filter for chlorophyll a (chl a) analysis on a Turner Designs® TOC-SOSOA fluorometer. Another water sample was also filtered for dissolved organic carbon (DOC) analysis on a Shimadzu® TOC-SOSOA. 5.3.2 Land Use coverage Land use coverage and digital soil data were obtained from the Spatial Data Library maintained by the Michigan Department of Natural Resource (MIDNR). These data are stored as Georef projections and as 1:250,000 scale maps from the US Geological Survey (USGS). Land use was grouped as urban, agriculture, rangeland, forest, water, wetland, and barren land. Michigan major land resource and soil types were categorized according to Michigan resource inventory as Indiana and Ohio Till Plain (1), Northern Michigan and 132 Wisconsin Sandy Drift (11), Southern Michigan and Northern Indiana Drift Plain (111), Southern Michigan Fruit and Truck Belt (IV), and Western Michigan and Northeastern Wisconsin Fruit belt (V) (1982). They were categorized into three regions as low erosive, medium erosive, and high erosive regions. The Lower Pennisula of Michigan was classified into two sub-ecoregion according to ecoregion map. Arcview 3.2 with the Spatial Analyst 1.1 extension, along with various extensions including Michigan projection extension and MIDNR Geometry Calculator Extension, were used to perform all spatial analyses. The watershed boundaries of wetlands were determined by selecting l to 2 km buffer areas around palustrine wetlands (according to size of wetlands) to determine surrounding land use (Berven and Grudzien 1990, Findlay and Lenton 2001). 5.3.3 Statistical analyses Diatom species relative abundances, species richness, and Shannon species diversity (H ’) were calculated from the species composition and enumeration data. van Dam diatom indices (van Dam et al. 1994) were calculated as the sum of products of species indicator values and proportional relative abundance (converted for proportion of species with indicator values) and used to indicate diatoms sensitive to nutrient enrichment change. All variables were standardized before analysis. Descriptive statistics, Pearson correlations, and regression analyses were conducted in Systat® v10. The Dunn-Sidak test was used to determine the significance of the correlation. Principal components analysis (PCA) was used to reduce and summarize variation of environmental variables in the region. The PCA (eigenvalue>l, 6 axes) sample scores were used as composite variables of chemistry to assess the differences among ecoregions or soil types. 133 Only diatom species with a relative abundance 2 1% in a minimum of three sites or _>. 5% in at least one site were included for comparison and gradient analysis. Nonmetric multidimentional scaling (NMDS) was used to identify diatom distribution patterns with relation to presettlement and current conditions. Euclidean distances were calculated for all samples based on diatom species composition, and then projected to 2-dimensional plots using NMDS (PCORD 4.0). Multiple response permutation procedures (MRPP), a nonmetric comparison method to discriminate differences among groups of assemblages, was used to determine the difference between reference and current diatom assemblages. Canonical correspondence analysis (CCA) was conducted to explore the relationship between diatom species distribution and environmental conditions (ter Braak 1990). All analyses were performed with the computer program CAN OCO v4.02 and PCORD v4.10 (McCune and Mefford 1999). Relative abundance data were square-root transformed and environmental variables were log transformed as necessary to obtain normal distributions. Whenever a variance inflation factor (VIF) was greater than 5, the least significant, collinear variables were dropped from the analyses .The environmental variables included all measured variables and the TN:TP ratio, DIN:SRP ratio, and wetland coordinates. Forward selection and Monte Carlo permutation (199 random permutations) were performed to determine statistical significance of a subset of environmental variables. The program CALIBRATE v0.6 (Juggins and ter Braak 1992) was used to perform weighted average regression and calibration. All models were based on square-root transformed diatom relative abundance data. J ackknifing was used to test the accuracy of the predictive models. Regression tree analysis (De'ath and Fabricius 134 2000) was performed using Systat® v10 to detect appropriate threshold values of environmental variables and community compositions with increasing agriculture land use. The Bray-Curtis similarity index (SI) was calculated as similarity of the relative abundance of surface diatom assemblage at one wetland with centroid of paeleological diatom assemblages to represent similarity with reference conditions. 5.4 Results: 5.4.1 Environmental variables The annual average of environmental variables varied among wetlands (Figure 5.2). Conductivity of these wetland waters ranged from 32.2 to 495.5 umol/cm (average :1: standard error (SE), 189.8i16.2 umol/cm). The pH was generally neutral (6.45:0.75). DOC averaged 17.9 mg/L (il.2 mg/L), DIN (137.2:t25.9 jig/L), and TN (1456.6:l:95.5 ug/L) concentrations were much higher than SRP (22.015.8ug/L) and TP (82.81: 16.3 ug/L) concentrations. Si and Cl averaged 3.9(i0.7) mg/L and 5.2(il.2) mg/L respectively. N :P (36.01 3.2) and DIN :SRP (26.0:I: 9.0) ratios were higher than 17 indicating potential P limitation. Environmental variables, especially nutrient variables were closely related to each other (Table 5.1). CI correlated with conductivity(r=0.668, P<0.001) and TDS (0.593, P<0.01). Si also correlated with conductivity and TDS (r=0.737, P<0.001), as well as NH4 (r=0.618,P<0.05), DIN (120.551, P<0.05), and SRP (r=0.547, P<0.05). Nutrient concentrations were tightly correlated with each other, and related to chl a concentration in the water column. TP correlated with TN(r=0.711, P<0.05), SRP(r=0.717, P<0.05), and N H4 concentrations (r=0.753, P<0.001). Chl a concentration in water column was 135 highly significantly (P<0.001) correlated with TP (m0.793), TN (r=0.805), and NH4 (r=0.850). O 100 200 300 400 500 5 6 7 8 9 0 100 200 300 Conductivitymmol/cm) pH Ch] a(mg/L) 4:13—— »« —Cl:l—— *4 1 l—* l l l ' W l l I l I l 0 10 20 30 1,0 10,0 \0 \‘3 '19 $53035 SI(mg/L) CL(mg/L) DOC(mg/L) l 2 3456 4,9 \QQ\59,5® 01020304050607080 N:P TN(mg/L) TP W l ‘ 'iJ l J l '1 100 1000 b.9 $9 (69 \9 '19 39W)“ DIN(mg/L) SRP(mg/L) DIN:SRP Figure 5.2. Box plot of environmental variables in 1994-1995 in the Michigan wetlands. The first three PCA axes accounted for 61.4% of variance in the environmental data set. The first axis was correlated with ionic concentrations (conductivity and pH) and the second axis correlated with total nutrients (TN and TP). Multiviate test of the first 3 axes revealed that environmental variables in wetlands were not different among sub- ecoregions (MANOVA, Wilks’s 2» F627 =l.784, P = 0.140) and among soil types (MANOVA, Wilks’s 7t F1154 =1.054, P = 0.386). Agricultural land use around wetlands correlated with average TP concentration in 1994 to 1995 (Table 5.3, Figure 5.3). Regression tree analysis was used to explore the relationship between agricultural land use and TP concentration. A threshold effect 136 occurred when agricultural land use exceed 58% and average TP increased from 46.7 to 207.3 ug/L. The regression tree model explained significantly more variance than a linear model (F test, P<0.05). 400 I I I I I f I I 300 l— . _ E: o 100 - O . . O - CI ’4. .. L...l. r ”m: r r 0.0 0.2 0.4 0.6 0.8 %Agriculture Figure 5.3. Relationship between average TP and % agricultural land use in sampled wetlands. Regression tree analysis was used to explore the relationship between TP and land use. A threshold effect occurred when agriculture occupied 58% land use around the wetland watershed. 137 ...de- 848- $2 $2... $8 and $3... 848 88 :3 83- 83 83 88.8 Ed- 233%? 83 :88- Pod 3.8- 48.8 58. ~88- 888- 5.8- 88.8 28.8 33- 23. M88 83- $8qu 83 $3 38- 83 ES 89c... 83 83- 88 33 08.8- 48.8- :3 ES flip 83 888- was 84.8 83- ~88- $8 :28 M88 088- 83 Ed 83- 25 83 83 888- 83... 838 388 3.8 Rod- 88 83 23 88.8 “am 88.3 :3 83 $8... 488 one? too 588 3.8 80¢ 88 £2 83 38.8- Rod- 83 3.8 98.8 was- 88 :3 88- ..oz 83 82 ~88 888- 888 88 83 W88 83- E. o8._ :3 58 $8 $3 :8 fine 28 E .83 3d :3 33 88.9.. $2 83- 6 83 83 $2.. 83... 82 89¢ a 88.3 83... 88.8- 83. 58 won 83 8.8 RS- 88 8.8 25 o8._ £3... 83- £82.28 83 38. an 83 deep $5qu $2 a :6 “am £2 52 E. E 6 a com 5ch £368.50 mm .85 .... .3 8565 a 8.3m 83385 032288 832:8 358%? <.W®O1VQQ— Cm Cam—Sou: MO N—Dmflmconm hogs mag—ON avg—~03 WM MO mD—DNE> §€OE€OHM>GD MO KEG—H— angman—0:00 COW—NOAH ..mm D~£Nr~t 138 5.4.2 Current diatom assemblages and environmental conditions A total of 269 diatom taxa were found in the surface sediment in the 35 palustrine wetlands during fall. Seventy-one diatom taxa (abundance > 1%, Table 5.2) from palustrine wetlands from surface sediment in fall were used in the CCA to explore the relationship between diatom assemblage and environmental variables. The first two axes explained 28.7% of cumulative variance in the taxonomical data. Stepwise permutation tests indicated that, of the 24 environmental variables, conductivity, TP, pH, and DIN: SRP ratio had significant canonical coefficients (P<0.05) with CCA axes and thus were selected in the final CCA. The four environmental variables alone in the first two axes captured 19.1% cumulative percentage variance of taxonomical data (Figure 5.4a, 4b). Conductivity and TP were the most significant environmental variables (P<0.005). The position of each taxon in the bioplot (Figure 5.4b) indicated its weighted-average optimum relative to other taxa and environmental variables. Epithemia adnata, Cyclotella nana, and Gomphonema mexicanum, among other taxa, indicated high TP concentration in the ordination space; while Cymebella microcephala, and Navicula leptostriata appeared most in low TP concentration. 139 A TN:TP NOx Ox N TN:TP EUPRAERU HAAMPHI o CMCESATI , NALEP’I‘OSO 0° 00. CMMICROCO ° O NASUBTIL ° 0‘50 oo° o M PIVIRIDI o ANSERIAN o PISTREP'I‘ : PIINTERR 0° PIMAIOR PIOSUBCAP ° EUPARALL O o FSRHOSAX EUPALUDO i .41 DIN:SRP 39 O 7033 tl I .. L GTN Conduc vty OCYGAMMA NAPRAETE DIN:SRP o P" FRCONBIN FRPINNOA . CSUBSAL O % o LGTN Conductivity a": NAGYSINLL 0 ° 0 ACEXIGUA ° ° . GOMEXICA o 0 0° 0 O CYNANA LG‘I'P oFRNITZSC . 0 O m , . . o EPADNATA GOCLAVAT 0.“ fl Figure 5.4. CCA biplot of sites and fall environmental variables from 35 palustrine wetlands in Michigan. (a) Dots show site scores in ordination space, (b) dots shows species scores in ordination space, and arrows indicate the strength and correlation between environmental factors and the ordination axes, The dotted circle shows sites with above 58% agricultural land cover. The dotted circle shows sites with above 58% agricultural land cover. Diatom species are indicated by eight-digits code (see Table 5.2 for complete taxon names). Table 5.2. The most dominant diatom taxa used in CCA and the correspondent taxa code Code Taxa name Code Taxa name Gomphonema gracile Ehr. emend. ACCONSPI Achnanthes conspicua A. Mayer GOGRACIL V. H. Gomphonema parvulum (Kiltz.) ACEXIGUA Achnanthes exigua Grun. GOPARVUL Kutz. ACHUNGAR Achnanthes hungarica (Grun.) Grun GOTRUNCA ACLANCEO Achnanthes lanceolata (Breb.) Grun. NACRYP'TC Gomphonema truncatum Ehr. Navicula cryptocephaloides Hust. 140 ACMINUTI ALPELLUC ANSERBRA ANVITREA AUCRENUL CCPLACEN CCPLALIN Achnanthes minutissima Kiltz. Amphipleura pellucida (Kutz.) Kiltz. Anomoeoneis serians var. brachysira (Bréb. ex Kiltz.) Hust. Anomoeoneis vitrea (Grun.) Ross Aulacoseira crenulata (Ehr.) Thwaites Cocconeis placentula Ehr. Cocconeis placentula var. lineata (Ehr.) V. H. CMLUNATA Cymbella lunata W. Sm. CMMICROC CMMINUTA CMSILESI DNKUETZI EPI'URGID EUARCUS EUBILMUC EUCURVAT EUEXIGUA EUFLEXUO EUNAEGEL EUPRAERU Cymbella microcephala Grun. C ymbella minuta Hilse ex Rabh Cymbella silesiaca Bleisch Denticula kuetzingii Grun. Epithemia turgida (Ehr.) Kiltz. Eunotia arcus Ehr. Eunotia bilunaris var. mucophila Lange-Bert. & Norpel Eunotia curvata (Kiltz.) Lagerst. NACRYPTO NACRYTEN NALEPTOS NAHARDII NAMEDIOC NAMINIMA NAPUPREC NARADIOS NASEMLUM N ASUBTIL NATANTUL NEIRIDIS NIAMPHIB NIGRACIL NIPALEA NIPERMIN Eunotia exigua (Bréb. ex Kutz.) Rabh NITROPIC Eunotiaflexuosa Bréb. ex Kiltz. Eunotia naegelii Migula Eunotia praerupta Ehr. EURHOMBO Eunotia rhomboidea Hust. FRBREVIS FRCAPGRA FRCAPMES FRCONSTU FRCONVEN FRCROTON FRPINNAT FSRHOCRA FSRHOMBO GOACUMIN GOAN GUST F ragilaria brevism'ata Grun. F ragilan'a capucina var. gracilis (Q8111) Hust. F ragilaria capucinavar. mesolepta Rabh. F ragilaria construens (Ehr.) Grun. F ragilaria construens var. venter (Ehr). Grun. F ragilan'a crotonensis Kitton F ragilaria pinnata Ehr. F rustulia rhomboides var. crassinervia (Bréb. ex W. Sm.) Ross F rustulia rhomboides (Ehr.) DeT. Gomphonema acuminatum Ehr. Gomphonema angustatum (Kutz.) Rabh. PIGIBBA PIINTERR PIMAIOR PIMICROS PISUBCAP PIVIRIDI SSANCEPS SSPHOENI SYNAN A SYRUMPEN SYTENERA SYULNA TAFLOCCU TAQUADRI 141 Navicula cryptocephala Klltz. Navicula cryptotenella Lange-Bert. Navicula leptostriata Jorgensen Navicula hardii Hust. Navicula mediocris Krasske Navicula minima Grun. Navicula papula var. rectangularis (Greg.) Grun. Navicula radiosa Kiltz. Navicula seminulum Grun. Navicula subtilissima Cl. Navicula tantula Hust. Neidium iridis (Ehr.) Cl. Nitzschia amphibia Grun. Nitzschia gracilis Hantz. ex Rabh. Nitzschia palea (Kiltz.) W. Sm. Nituchia perminata (Grun.) Peragallo Nitzschia tropica Hust. Pinnularia gibba Ehr. Pinnularia interrupta W. Sm. Pinnularia maior (Kutz.) Rabh. Pinnularia microstauron (Ehr.) Cl. Pinnularia subcapitata Greg. Pinnularia viridis (Nitz.) Ehr. Stauroneis anceps Ehr. Stauroneis phoenicenteron (Nitz.) Ehr. Synedra nana Meist. Synedra rumpens Kiltz. Synedra tenera W. Sm. Synedra ulna (Nitz.) Ehr. Tabellaria flocculosa (Roth) Kiltz Tabellaria quadriseptata Knud. Table 5.3 Regression of observed environmental variables against predicted values based on diatom species composition. Simple regression and jackknifing cross-validation indicators based on weighted average (WA) and tolerance down-weighted WA used to compare the precision. MSE represents mean square error and RMSE represent residual of MSE. Simple Cross-Validation MSE RMSE R2 MSE RMSE R2 Conductivity WA 4585 67.713 0.741 7132 86.693 0.578 T01 d/w WA 3672 60.598 0.792 1.23E+04 110.879 0.425 LGTP WA 0.0492 0.222 0.704 0.0958 0.310 0.435 T61 d/w WA 0.0496 0.223 0.702 0.123 0.351 0.334 pH WA 0.407 0.638 0.679 0.807 0.898 0.381 T61 d/w WA 0.299 0.547 0.764 0.757 0.870 0.446 DIN:SRP WA 133.8 11.568 0.634 282.1 16.795 0.255 T61 d/w WA 131 11.447 0.641 291 17.060 0.250 Diatom-based WA inference models for conductivity, TP, pH, and DIN: SRP were developed to predict environmental conditions for palustrine wetlands (Table 5.3). The WA inference model for conductivity produced a lower R2 (0.741) than the tolerance down-weighted WA model (WA(tol)); however, the simple WA model produced a much better model (R2=0.578, Figure 5.5a) after cross-validation with jackknifing than the tolerance down- weighted WA model (R2=0.425). ’Ihe WA TP inference models also showed better regression (R2=0.435) after cross-validation than the WA(tol) model (R2=O.334). The analysis of pH data indicated that tolerance down-weighted WA models were much better than simple WA models both before and after cross-validation. R2 values for WA models for DIN: SRP were low, especially after cross-validation jackknifing (1120.250). 142 8 I IIIIIII Predicted TP(ug/L) '5 I I IIIIII I IIIII‘ A LLJII. .31 lO Observed TP(ug/L) 100 300 I I I I I I I C I 200 . . . 0' O O 100 o '7 ' o C . ... . ..O‘ 1 L. 1 1 1 1 1 0102030405060708090 % Agricultural land Figure 5.5. Regression models of diatom inferred environmental variables against observed environmental variables. 5b shows the relationship between diatom inferred TP and agricultural land use. Based on diatom autecological data, predicted TP concentrations of wetlands along an agricultural land use were used to infer wetland trophic level changes (Figure 5.5a). In this model, diatom inferred TP showed a linear increase with increase of agricultural land use reached (R2=0.246, P=0.003) (Figure 5.5b). 5.4.3 Comparison between reference and current conditions A total of 169 diatom taxa were identified in 14 sediment core paleoecological samples, While 189 taxa were found in the corresponding surface sediment samples. An average of 55.4% (i2.4%) of the taxa in the surface samples were not found in the core samples in the same wetland, while 50% (i 4%) of the core taxa were not present in surface samples in the same site. Forty-one diatom taxa in paleoecological samples were not found in all 35 surface sediment. Diatom species compositions were significantly different between current and paleoecological samples. (MRPP, P<0.001, Figure 5.6). 143 2 1 Figure 5.6. Nonmetric o Paleoecological multidimensional scaling ordination ‘ diagrams of paleological diatom 1 ‘ Current assembalges and current diatom assemblages. Reference(paleological diatom assemblages) and current (surface diatom assemblages) were indicated by different symbols. The centroids with 68.5% confident ellipse were plotted around each region. N MDSZ o Native Species 0 l llJ 00 LII! 100 200400800 ' 100 200300400 TP(ug/L) TP(Mg/L) Figure 5.7. Relationship between richness of diatom native taxa with nutrient concentration (TP) (a), and relationship between similarity index (SI) of current diatom assemblage with pressettlement assemblages with nutrient concentration (TP) (b) in wetlands. Native taxa were defined as taxa found in the paleoecological samples. SI is the Bray-Curtis similarity of diatom relative abundance in surface sediment and the centroid of paeloecological samples. 144 Table 5.4. Correlation among community characteristics and environmental variables LGTN LGTP DIN SRP Chl a % agricultural land Species Richness -0471 -0513 -0434 -0427 0.359 0.130 Shannon H’ -0.534 -0420 -0.514 -0357 0.265 0.093 NO' 9f Name -0291 -0452 -0439 -0.302 0.195 0.015 Specres Similarity Index 0453 -0.357 -0395 -0.258 0.297 -0.062 Both the number of native taxa in surface sediments and similarity of diatom species composition with reference conditions decreased With increasing nutrient concentrations (Table 5.4, Figure 5.7). The number of native species in surface sediments was negatively correlated with increased nutrient concentrations, especially TP in wetlands. Diatom species composition in reference conditions showed large variation among different sites. The similarities ranged from 0.03 to 0.38 with a median similarity of 0.13. The similarities of current diatom relative abundances with paleoecological diatom species composition were also negatively correlated with TN and TP concentrations (r: —0.453 and —0.357 respectively). These similarities were generally low and varied Within the range of similarities among reference conditions. Thus a definition of biological criterion based on similarity change was unlikely. van Dam’s trophic indices defines the importance of trophic sensitive diatom species and trophic conditions in wetlands. I defined the 75th percentile of paleoecological diatom trophic indices in wetlands as the biological criterion of pristine conditions. van Dam’s trophic indices in reference conditions varied from 1.7 to 4.5 with a median value of 3.4. van Dam trophic indices in current diatom species composition was more viriable, especially at the higher value which represented higher relative abundance of trophic species in some wetlands now than in historic times. The values of indices showed a 145 linear increase with increase of TP concentration (R2=0.409, P=0.001, Figure 5.8a). The nutrient concentration at the intercept of the biological criterion with regression lines represented the nutrient criterion (stressor criterion=45.6 um/L). Change of van Dam trophic indices was also linearly correlated with agricultural gradient (R2=0.517, Figure 5.8b). However, a regression tree analysis also showed a changepoint at 56% agricultural land use where the van Dam trophic index increased from 3.46 to 4.85. :36 I I '0' I I 6 r l l r l . I r r 1 E "D‘ Paleoecological . :5 _ . . ... . .- 5- ' Current .. . °". .. E o ) g 0 o ...-”‘0 9 D4 1 o 9' ~-"' 04 - . . a ' . ............................ ,9- b 0 ° 0". 3 . 0'...“ 9. o 33 3. ----- "s '3 ' .r" :0 - ..."..v C -' Q -1 Q .0 o 5 2 - ' o - - > :1: 1 fl 1 1 1 11 I 1 I A 1 l l 1 l U «3% \90 $3090 0 10 20 30 40 50 60 70 80 90 log 10TP(ug/L) Agricultural land use (%) C Figure 5.8a. Relationship between van Dam trophic index and palustrine wetland nutrient concentration (TP). Box-plot shows paleological diatom trophic index from 14 palustrine wetlands to represent reference wetland condition. Surface diatom trophic index (dots) respond linearly to TP concentration in wetlands (R2=0.407, P=0.001). The intercept of biological criteria and regression line represented nutrient criteria (stressor criteria=46 ug/L). 8b. Relationship between van Dam trophic index and palustrine wetland land use. Box-plot shows paleological diatom trophic index from 14 palustrine wetlands representing reference wetland condition; dots represent change of trophic index with agricultural land use; a threshold effect was observed when agriculture exceeds 56% total land use, with van Dam trophic index increasing from 2.46 to 4.85. 5.5 Discussion 146 The goal of ecological assessment is to assess condition of ecosystems, often relative to an expected or desired condition, to determine the stressors that threaten or impair valued ecological attributes, and to determine the human activities producing those stressors. This has been challenging in wetlands due to spatial and temporal variability. I found that changes in algal species composition in wetlands were related to human activities and then developed biological and nutrient criteria that would protect the natural diatom assemblages of wetlands. I found that development of assessment criteria required classifying wetlands by type and restricting biological metrics to those most sensitive and directly related to stressors. Characterizing reference condition with paleoecological assemblages reduced doubt of the biological potential of the wetlands. Nutrient concentrations, especially total phosphorus concentration in the Michigan wetlands studied, were directly related to human activities, especially agricultural land use. Human activities are well-known causes of nutrient enrichment of aquatic systems (Vitousek et al. 1997, Carpenter et al. 1998). Land cover/land use indicators of human activities and related water quality models are commonly correlated with nutrient concentrations in surface waters (Soranno et al. 1996, Tufford et al. 1998, Freeland et al. 1999). As the receiver of storm runoff, overland flow, and ground water influx, palustrine wetlands received high nutrient input from agriculture (Thompson and Polet 2000). My study not only found that high phosphorus concentration in Michigan wetlands were correlated with agricultural land use surrounding wetlands, but also discovered a threshold effect when agricultural land use exceeded 57.9% total land cover. Similarly, Waltman et al. (unpublished) also found a threshold effect on nutrient concentrations 147 when the % agriculture in Kentucky stream basin exceeded 60%. This effect may represent saturation of the assimilative capacity of soils within watersheds. I found that changes in diatom species composition, number and relative abundance of individuals belonging to native species, and other valued ecological attributes changed with nutrient enrichment associated with human activities. TP, as a common cause of change in species composition, was an environmental stressor directly related to land use in Michigan wetlands. Examination of diatom taxa found in wetlands with 57.9 % agricultural land use showed that those wetlands were dominated by taxa requiring high nutrients, such as Epithemia adnata and Cyclotella nana, and had the lowest number of native taxa. TP enrichment was the possible cause of loss of native taxa. Both increased species diversity and decreased species diversity have been found in aquatic systems with increased nutrient concentrations (J utter et al. 1996, Hillebrand and Sommer 2000, Pan et al. 2000). I expected a decreasing native taxa richness because native taxa were less competitive in nutrient enriched environments. Competitive exclusion for other limiting resources, increased predation, or alteration of other abiotic factors through indirect effects of nutrients (e. g., dissolved oxygen concentration, microbial interaction, or habitat structure) may all decrease native taxa richness. The similarity of species composition to reference condition should be more sensitive than loss of native taxa (presence/absence of taxa), because relative abundance of taxa should change before they are lost from a habitat. However, my results only found a weak linear decrease in the similarity index (S1) with increasing nutrient concentration. The weak response of S1 is possibly because SI was affected by multiple natural variables, such as pH and conductivity, as well as nutrient enrichment. Therefore, S1 148 responses are complexly relate to a stressor and its interaction with other stressors (Lansid and McLaughlin 2000). Alternatively, van Dam’s trophic index represents sensitive response of diatom relative abundance specific to one environmental stressor, in this case, nutrients. Although it was developed in Europe, the strong relationship between diatom trophic indices and TP in Michigan wetlands suggests that this index could have broader geographic application. Reference condition is ideally based on the central tendency and variability in conditions in habitats with the least human disturbance. Paleoecological assessments derived from sediments deposited before European alterations of landscapes provide the ideal characterization of reference condition. When using S1 to describe reference condition, little difference was observed between reference condition and present condition in wetlands. SI was very low among paleoecological conditions in MI wetlands (usually 0.1-0.2) compared to present day variability among reference sites in the Everglades (0.7, Stevenson et al. 2002). This low similarity among reference wetlands was probably related to the great natural variability in pH, conductivity, and even development stage among wetlands. Greater differences between reference condition and present day condition of wetlands was observed when diatom assemblages were characterized using van Dam’s trophic indices, probably because van Dam's trophic index based on diatom autoecological information of individual taxa characterizes biological accumulation of trophic status in the system. Nutrient conditions were probably more consistently low before European alteration of landscapes than pH, conductivity, and other naturally varying factors that affect diatoms. Compared with S1, autoecological indices are usually more precise for assessment and diagnosis of aquatic 149 ecosystem problems (Stevenson et al. 2001). Thus, biocriteria for wetlands were based on van Dam's trophic index indicating the relative abundance of diatoms that were sensitive to a single stressor, nutrient enrichment. The biocriterion based on Van Dam's trophic index was established at the 75th percentile of its frequency distribution, which represents an upper bound of nutrient tolerant diatoms in the natural biological assemblages. The 75‘h percentile of reference conditions has been defined to indicate deterministic variation of algal assemblages and it has been used in other studies (Hill et a1. 2003). A variety of percentiles in reference conditions have been used by different agencies for different purposes. An ideal approach for defining biological criteria would be based on a threshold response along a stressor gradient. With the determination of a criterion for a valued ecological attribute, stressor criteria can be determined based on the VEA criterion and a stressor-response relationship. A cause-effect relationship has been established between nutrient enrichment and algal response in Michigan wetlands (Chapter 4). Stressor criteria should be no higher than the stressor level predicted to cause VEAs to be at the level of the biocriterion (Wang and Stevenson 2002, Stevenson et al. accepted). An alternative approach is to observe thresholds in VEA response along the stressor gradient. The TP criterion based on stressor-response relations and the biological criterion (van Dam trophic index) was 46 ug/L. This value was similar to the TP levels below the threshold response in TP with increasing agricultural land use. These complementary lines of evidence provide sound conceptual foundations for defining nutrient criterion. 150 The ultimate goal of assessing and managing aquatic ecosystems is to restore these systems to natural conditions. Thus, I first should estimate the natural condition should be estimated by direct assessment of reference condition with minimum human activity, contamination, and alteration. Second, cause-effect relationship should be determined between algal indicators and stressors by manipulated experimental approaches. Third, stressor criteria could be determined by integrating results from physical and chemical analyses, experimental and observational relation of stressor- response gradients, and bio- criteria based on reference approach. A suite of VEAs, such as native taxa richness, shifts in relative abundance, or sensitive taxa should be measured in relation to stressor gradients. Future research should explore ecological response of algae to nutrient concentrations in wetlands, and a nationwide database of individual diatom autoecological indices should be established. An understanding of benthic algal ecology is a prerequisite to implementing a biological approach to ecosystem management using algae. 151 5.6 References American Public Health Association (APHA). 1998. Standard Methods for the Examination of water and wastewater. 20th edition by LS. Clececeri, A.E. Greenberg, and AD. Eaton. Washington D. C., USA.Anderson, D. J ., and B. Vondracek. 1999. Insects as indicators of land use in three ecoregions in the prairie pothole region. Wetlands 19: 648-664. Anderson, D. J. and B. Vondracek. 1999. Insects as indicators of land use in three ecoregions in the prairie pothole region. Wetlands 19: 648-664. Berven, K. A., and T. A. Grudzien. 1990. Dispersal in the Wood Frog (Rana-Sylvatica) - Implications for Genetic Population-Structure. Evolution 44:2047—2056. Byron, ER, and C. R. Goldman, 1989. Land-use and water quality in tributary streams of lake tahoe. Califomia-Nevada. Journal of Environmental Quality 18:84-88. Carpenter, S. R., N. F. Caraco, D. L. Correll, R. W. Howarth, A. N. Sharpley, and V. H. Smith. 1998. Nonpoint pollution of surface waters with phosphorus and nitrogen. Ecological Applications 8:559-568. Carpenter, K. D., and I. R. Waite. 2000. Relations of habitat-specific algal assemblages to land use and water chemistry in the Willamette Basin, Oregon. Environmental Monitoring and Assessment 64: 247-257. Charles, D. F. 1985. Relationships between surface sediment diatom assemblages and Lake-water characteristics in Adirondack lakes. Ecology 66: 994-1011. Cowarding, L.M., V. Carter, F. C. Golet, and E. T. LaROe. 1979. Classification of wetlands and deepwater habitats of the United States. FWS/OBS-79/31, US. Fish and Wildlife Service, Washington, DC. 103 pp. De'ath, G., and K. E. Fabricius. 2000. Classification and regression trees: A powerful yet simple technique for ecological data analysis. Ecology 81:3178-3192. Detenbeck, N. E., S. M. Galatowitsch, J. Atkinson, and H. Ball. 1999. Evaluating perturbations and developing restoration strategies for inland wetlands In the Great Lakes Basin. Wetlands 19:789-820. Detenbeck, N. E. 2001. Methods for evaluating wetland condition: Wetlands classification. EPA 822-R-01-007g. US. Environmental Protection Agency; Office of Water; Washington, DC. Dixit, S.S., J. P. Smol, J. C. Kinston, and D. F. Charles. 1992. Diatoms: powerful indicators of environmental change. Environmental Science and Technology 26:23- 33. 152 Dixit, A. S., S. S. Dixit, and J. P. Smol. 1996. Long-term water quality changes in Ramsey Lake (Sudbury, Canada) as revealed through paleolimnology. Journal of Environmental Science and Health Part a- Environmental Science and Engineering & Toxic and Hazardous Substance Control 31:941-956. Findlay, C. S., J. Lenton, and L. G. Zheng. 2001. Land-use correlates of anuran community richness and composition in southeastern Ontario wetlands. Ecoscience 8:336-343. Freeland, J. A., J. L. Richardson, and L. A. Foss. 1999. Soil indicators of agricultural impacts on northern prairie wetlands: cottonwood lake research area, North Dakota, USA. Wetlands 10: 56-64. Gasse, F. 1987. Diatoms for reconstructing paleoenvironments and paleohydrology in tropical serrri-arid zones: examples of some lakes in Niger since 1200bp. Hydrologia 154:127-163. Gordon, S. I., and S. Majumder. 2000. Empirical stressor-response relationships for prospective risk analysis. Environmental Toxicology and Chemistry 19: 1106-1112. Hall, R.I., and J.P. Smol. 1992. A weighted averaging regression and calibration model for inferring total phosphorus concentration from diatoms in British-Columbia (Canada) lakes. Freshwater biology 27: 417-434. Heaney, S. 1., R. H. Foy, G. J. A. Kennedy, W. W. Crozier, and W. C. K. O'Connor. 2001. Impacts of agriculture on aquatic systems: lessons learnt and new unknowns in Northern Ireland. Marine and Freshwater Research 52:151-163. Hill, B. H., A. T. Herlihy, P. R. Kaufmann, R. J. Stevenson, F. H. McCormick, and C. B. Johnson. 2000. Use of periphyton assemblage data as an index of biotic integrity. Journal of the North American Benthological Society 19:50-67. Hill, B. H., A. T. Herlihy, P. R. Kaufmann, S.J. DeCelles, M. A. Vander Borgh. 2003. Assessment of streams of the eastern United States using a periphyton index of biotic integrity. Ecological Indicators 2: 325-338. Hillebrand, H., and U. Sommer. 2000. Diversity of benthic microalgae in response to colonization time and eutrophication. Aquatic Botany 67:221-236. Hughes, R. M., C.B. Johnson, S. S. Dixit, A.T. Herlihy, P.R. Jaufmann, W.L.Kinney, D.P.Larsen, P.A. Lewis, D.M. McMullen, A.K. Moors, R. J. O’Connor, S.G. Paulsen, R.S. Stemberger, S.A. Thiele, T.R. Whittier, and D. L. Kugler. 1993. Development of lake condition indicators for EMAP — l991n pilot. Pages 7-90 in D. P. Larsen and SJ. Christie. (editors). EMAP- surface waters 1991 pilot report. EPA-620-R-93-003. U.S.Environmental Protection Agency, Office of Research and Development, Corvallis, Oregon. 153 Hughes, R. M., S. G. Paulsen, and J. L. Stoddard. 2000. EMAP-Surface Waters: a multiassemblage, probability survey of ecological integrity in the USA. Hydrobiologia 422:429-443. McCune, B., and J. Grace. 2002. Analysis of Ecological community. MJM Software Design. Pp304. Gleneden Beach, Oregon. Hunsaker, C. T., D. A. Levine, S.P. Tmins, B. L. Jackson, and R. V. O’Neill. 1992. Landscape characterization for assessing regional water quality. Pages 997-1006 in D. H. McKenzie. D. E. Hyatt, and V. K .McDonald. (editors). Ecological indicators. Volume two. Elsevier Applied Science, New York, New York, USA. Hustedt, F. 1930-1966. Die kieselalgen Deutschlands, Osterreichs und der Schweiz unter Berucksichtigung der urtigen Lander Europas sowie der andrenzenden Meeresgebiete, Pages920pp,845 pp and 816 pp. in L. Rabenhorst editor. Kryptogamen-Flora von Deutschland, Osterreich und der Schweiz.Academische Verlagsgesellschaft Leipzig. Juggins, S., and C. J. F. ter Braak. 1992. CALIBRATE—a program for species- environment calibration by weighted averaging partial least squares regression. Environmental Change Research Center, University College, London. Juttner, I., H. Rothfritz, and S. J. Ormerod. 1996. Diatoms as indicators of river quality in the Nepalese Middle Hills with consideration of the effects of habitat-specific sampling. Freshwater Biology 36:475-486. Krammer, K., and H. Lange-Bertalot. 1986. Bacillariophyceae. 1. Teil: Naviculaceae.Pages 1-876 in H. Ettl, Gerloff, J ., Heynig, H., & Mollenhauer, D. editor. Sitsswasserflora von Mitteleuropa.Gustav Fisher Verlag, Germany. Krammer, K., and H. Lange-Bertalot. 1991. Bacillariophyceae.VEB Gustav Fischer Verlag Jena. Krammer, K., and H. Lange-Bertalot. 1997. Bacillariophyceae.VEB Gustav Fisher Verlag Jena. Landis, W.G., and J .F. McLaughlin. 2000. Design criteria and derivation of indicators for ecological position, derecetion, and risk. Environmental Toxicology and Chemistry: 19: 1059-1065. Leland, H. V. 1995. Distribution of Phytobenthos in the Yakima River Basin, Washington, in Relation to Geology, Land-Use, and Other Environmental-Factors. Canadian Journal of Fisheries and Aquatic Sciences 52: 1108-1129. 154 Leland, H. V., and S. D. Porter 2000. Distribution of benthic algae in the upper Illinois River basin in relation to geology and land use. Freshwater Biology 44: 279-301. Lopez, R. D., and M. S. Fennessy. 2002. Testing the floristic quality assessment index as an indicator of wetland condition. Ecological Applications 12:487-497. McCormick, P. V., P. S. Rawlik, K. Lurding, E. P. Smith, and F. H. Sklar. 1996. Periphyton-water quality relationships along a nutrient gradient in the northern Florida Everglades. Journal of the North American Benthological Society 15:433-449. McCune, B., and J. Mefford. 1997. PC-ORD. Multivariate Analysis Of Ecological Data. Version 3.0 For Windows. Mjm Software Design, Gleneden Beach, OR, USA. Mitsch, W. J ., and R. F. Wilson. 1996. Improving the success of wetland creation and restoration with know-how, time, and self-design. Ecological Applications?6:77-83. Osborne, L. L., and M. J. Wiley. 1988. Empirical relationships between land use/cover and stream water quality in an agricultural watershed. Journal of Environmental management 26:9-27. Pan, Y. D., and R. J. Stevenson. 1996. Gradient analysis of diatom assemblages in western Kentucky wetlands. Journal of Phycology 32: 222-232. Pan, Y., R. J. Stevenson, P. Vaithiyanathan, J. Slate, and C. J. Richardson. 2000. Changes in algal assemblages along observed and experimental phosphorus gradients in a subtropical wetland, USA. Freshwater Biology 44:339-353. Patrick, R., and C. W. Reimer. 1975. The Diatoms of the United States, exclusive of Alaska and Hawaii. The Academy of Natural Sciences of Philadelphia, Philadelphia. Richardson, C. J. 2000. Freshwater Wetlands. Pages 448-499 in MG. Barbour, and W.D. Billings (Editors). North American Terrestrial Vegetation. Cambridge University Press, New York. Smol, J. P., B. F. Cumming, A. S. Dixit, and S. S. Dixit. 1998. Tracking recovery patterns in acidified lakes: A paleolimnological perspective. Restoration Ecology 61318-326. Somlyody, L. 1998. Eutrophication modeling, management and decision making: The Kis-Balaton case. Water Science and Technology 37 : 165-175 . Stevenson, R. J. 1998. Diatom indicators of stream and wetland stressors in a risk management framework. Environmental Monitoring and Assessment 51: 107-118. Stevenson, R. J ., R. P. Sweets, Pan Y., and R. E. Schultz. 1999. Diatom community patterns in wetlands of Michigan (USA) and their use as indicators of ecological conditions. Pages 517-527 in A. J. McComb, and J. A. Davis (editors). Proceedings of 155 INTECOL’s Vth International Wetland Conference. Gleneagles Press, Adelaaide, Australia. Stevenson, R. J ., P V. McCormick, and R. Frydenborg. 2001. Methods for evaluating Wetland Condition: Using Algae to assess environmental conditions in wetlands. EPA 843-B-00-002k. US. Environmental Protection Agency; Office of Water, Washington, DC. Stevenson, R. J ., Y. Pan, and P. Vaithiyahathan. 2002. Ecological assessment and indicator development in wetlands: the case of algae in the Everglades, USA. Verh. Intemat. Verein. Limnol. 28: 1248-1252. Stosch, H. A. von and B.E.F. Reimann. 1970. Subsilicea fragilarioide gen. Et spec. nov., eine Diatomee (Fragilariaceae) mit vorwiegend organischer membran. Nova Hedwigia Beih. 3121-36. ter Braak, C.J.F. 1990. Update Note: CANOCO v 3.1. Agricultural Mathmatics Group, Wageningen, The Netherlands. Thompson, J. R., and G. Polet. 2000. Hydrology and land use in a Sahelian floodplain wetland. Wetlands 20: 639-659. Tufford, D. L., H. N. McKellar, Jr., and J. R. Hussey.l998. In-stream Nonpoint Source Nutrient Prediction with Land-use proximity and seasonality. Journal of Environmental Quality 27:100-111. United States Department of Agriculture, 1982. Michigan Data 1982 National Resource Inventory. United States Environment Protection Agency (USEPA). 2000. Principles for the Ecological Restoration of Aquatic Resources. EPA841-F-00-003. Office of Water (4501F), United States Environmental Protection Agency, Washington, DC. 4 pp. Van Dam, H., A. Mertens, and J. Sinkeldam. 1994. A coded checklist and ecological indicator values of freshwater diatoms from the Netherlands. Netherland Journal of Aquatic Ecology 28:117-133. Van der Valk, A. 2001. Methods for evaluating wetland condition: land-use characterization for nutrient and sediment risk assessment. EPA 822-R-01-007q. US. Environmental Protection Agency; Office of Water, Washington, DC. Vitousek, P. M., J. D. Aber, R. W. Howarth, Ge.E. Liken, P. A. Matson, D. W. Schindler, W. H. Schlesinger, and D. G. Tilman. 1997. Human alternation of the global nitrogen cycle: sources and consequences. Ecological Applications 7 2737-750. 156 Wang, Y-K., and R. J. Stevenson. 2002. An ecological assessment framework for watershed management. Journal of the Chinese Institute of Environmental Engineering 12: 209-215. Whitton, B. A., and M. G. Kelly. 1995. Use of algae and other plants for monitoring rivers. Australia. Journal of Ecology 20:45-56. Wilcox, D. A. 1995. Wetland and Aquatic Macrophytes as Indicators of Anthropogenic Hydrologic Disturbance. Natural Areas Journal 15:240-248. Wilson, M. V., and C.L.Mohler. 1983. Measuring compositional change along gradients. Vegetatio 54:129-141. 157 Chapter 6 Conclusions This dissertation develops a better understanding of relationship among a variety of ecological factors within wetlands, evaluates a number of valuable ecological attributes to deve10p indicators for wetland bioassessment and wetland restoration, and establishes biological and environmental criteria to protect wetland ecosystems. Because of the spatial and temporal variability within and among wetlands, ecological assessment in wetlands has to include development of habitat or class specific attributes for effective evaluation. My studies in Michigan wetlands indicated that the similarity of algal assemblages among habitats within wetlands and between restored and extant wetlands were related to a number of environmental characteristics. Local pairwise comparison based on reference conditions could reduce regional variation among wetlands and identify the successional pattern of ecological development after restoration with restored wetland age. Algal attributes in both freshwater and salt marshes were closely related to nutrients. Experimental determination indicated that algal biomass and algal assemblages were directly related to nutrient concentrations and nutrient ratio, and a cause-effect relationship was established between algal assemblages and nutrient gradients. Based on the stressor-response relationship between nutrient stressors and algal attributes, and by choosing paleoecological diatom assemblages as reference condition, I developed bio- criteria and nutrient criteria for wetland protection. I first evaluated spatial and temporal variation of environmental and algal attributes among habitats within wetlands and between restored and extant wetlands. Restored wetlands differed from extant wetlands in a number of ways in Southern Michigan. 158 Physical and chemical variables, especially plant shading and nutrient concentrations, were distinctly greater in extant wetlands and contributed to differences in algal communities among habitats within wetlands and between restored and extant wetlands. Whether algal assemblages in these two types of wetlands were similar or distinct depended on different habitats. Epiphytic assemblages in the restored and extant wetlands were relatively distinct, while neither sediment nor phytoplankton assemblages were as different between restored and extant wetlands. On the other hand, non-diatom algal assemblages from different habitats were mostly different, but sediment and epiphytic diatom assemblages were not significantly different. Ecological restoration is a long-term process and does not occur in a short period of time. I found no direct evidence for increasing similarity in algal assemblages between restored and extant wetlands with increasing age of restored wetlands. However, response of algal assemblages to nutrient concentration and light, two factors distinguishing restored and extant wetlands, indicated indirectly that succession processes in restored wetlands were regulating algal assemblages and restored wetlands < 10 years old may need more time to develop mature qualities. In the next chapter, I evaluated a suite of algal attributes of restored salt marshes based on paired reference and restored marsh approach and developed indicators for wetland development. I assessed algal biomass (dry mass (DM), ash free dry mass (AFDM), chl a content, algal biovolume), algal species composition and diversity, and gross primary production on both Spartina alterm'flora and sediments (sediment algae). Controlling for regional variation in reference marshes substantially increased precision in relations between attributes and the increase in age of restored marshes. The organic 159 matter restoration ratio of sediments increased with age of restored marshes in both spring and summer. The algal biomass restoration ratios of epiphytes, calculated with algal biovolume and chl a, increased with restored marsh age in summer but not during spring. Biomass of sediment algae was not related to marsh age. The similarity of diatom species composition between paired restored and reference sites increased with age of restored marshes during spring and summer. Primary production by epiphytic and sediment algae in summer showed site-specific changes and did not change consistently with marsh age. Algal biomass, algal diversity, and diatom species assemblages during summer were positively correlated with sediment nitrogen and phosphorus concentration. I concluded that overall structural and functional development of restored wetlands, especially nutrient storage in sediments, regulated algal community structure and function, which can be used to evaluate marsh restoration. I then determined the limiting nutrient in wetlands and established the cause-effect relationship between nutrient and algal assemblages because both the previous two chapters indicated that nutrients were one of the most important factors regulating periphyton growth and species composition in wetlands. Algal bioassays using nutrient diffusing substrata indicated that nitrogen (N) was the main limiting nutrient for periphyton growth in restored wetlands in southern Michigan. Algal N limitation positively correlated with phosphorus (P) concentration and negatively correlated with N:P ratios in wetland water and sediment. Canonical correspondence analysis (CCA) indicated that diatom species composition in natural habitats and control artificial habitats changed mainly along a pH gradient and secondarily along nutrient gradients indicated by total phosphorus (TP) and dissolved inorganic nitrogen (DIN). Individual diatom taxa 160 responded to nutrient addition differently. Overall, the responses of indicator species in nutrient bioassays were positively correlated with natural nutrient gradients among natural wetlands indicating a good match of diatom indicators between natural and experimental systems. Chapter 5 illustrated how the stressor-response relationship and reference approach could be used to develop stressor and biological criteria to protect wetland health. Diatom-based biological indicators were used to establish biological and stressor criteria for wetland protection and restoration and to indicate wetland ecosystem health. Paleoecological diatom assemblages from pre-settlement sediments were used to represent natural, reference conditions in wetlands. Stressor-response relations were used to delineate stressor criteria and human disturbance. Canonical correspondence analysis (CCA) revealed that diatom species assemblages responded mainly to conductivity and total phosphorus (TP). Similarity in species composition to reference condition and number of native species in current wetlands decreased with increase of nutrient concentrations in these wetlands. van Dam's diatom trophic index was imported to characterize wetland trophic status in North American wetlands and was linearly correlated with the TP gradient. Both the weighted average TP model and van Dam's trophic index positively responded to agricultural land use around wetlands. These indices deviated from the "normal" natural condition when TP exceeded 46 ug/L. A threshold change in TP and biotic indices was detected when agricultural land exceeded 57.9 % total land use. In summary, most extant wetlands have suffered a certain degree of nutrient impact from human disturbance. Wetland restoration has significantly decreased nutrient 161 concentrations and compensated for wetland loss. However, wetland restoration is a long- term process and it takes a long time for different attributes to develop to the status of reference condition. Continuous monitoring and assessment would lead to better understanding of ecosystem processes. My studies developed algal assemblages as indicators for diagnosing and quantifying nutrient pollution in wetlands and they identified ecosystems at risk due to excess nutrients in the environment. Based on the establishment of cause-effect relationship and stressor-response relationship between nutrient and algal attributes, ecological thresholds, algal biological criteria, and stressor criteria were developed for watershed management. The results of this research could be used to support ecologically sensitive decision-making in land planning and wetland restoration by providing targets for decreasing anthropogenic pollutions and thereby increasing the rate of successful restoration projects. 162 Appendices 163 Appendix A. Wetland sites sampled in Southern Michigan. Land use around wetland 100m radius were defined based on observation. RES: resident; AG: agricultural; CIT: constantly influenced by road; UND: undisturbed. Site Code Latitude Longitude Land County Sample Artificial Age Site code use Date substrata of property Res. 1 Dansla 42.56604 84.3968 RES INGHAM 8-18-2000 YES 6 Res. 2 JoIIy18 42.68161 84.4154 RES INGHAM 8-24-2000 YES 11 Res. 3 Jolly19 42.682 84.4182 RES INGHAM 8-24-2000 NO N/A Ext. 4 Green20 42.90778 84.5202 RES CLINTON 8-23-2000 NO N/A Ext. 5 Green25 42.81349 84.4527 RES CLINTON 8-23-2000 NO 8 Res. 6 Herbison2 42.8341 84.3962 AG CLINTON 8-1-2000 NO 7 Res. 7 27N 42.931 84.515 CIT CLINTON 8-23-2000 NO 8 Res. 8 278 42.9262 84.5159 CIT CLINTON 8-23-2000 NO 8 Res. 9 DansSl 42.55581 84.3642 RES INGHAM 8-18-2000 YES 6 Res. 10 Airport542 42.7905 84.6067 RES CLINTON 8-21-2000 YES 2 Res. 11 Alrport539 42.79212 84.6027 RES CLINTON 8—21-2000 NO N/A Ext. 12 Airport615 42.79212 84.6026 RES CLINTON 8-21-2000 NO N/A Ext. 13 Herbison6 42.8342 84.4007 AG CLINTON 8—1-2000 NO N/A Ext. 14 Alward 42.89237 84.4581 RES CLINTON 8-25-2000 NO N/A Ext. 15 Barnes 42.52287 84.3594 AG INGHAM 8-16-2000 NO N/A Ext. 16 CampusB 42.72074 84.4776 CIT INGHAM 8-25-2000 YES 9 Res. 17 CampusS 42.72079 84.4766 CIT INGHAM 8-25-2000 YES 11 Res. 18 Centerline 42.90436 84.4738 RES CLINTON 8-25-2000 NO N/A Ext. 19 Centeer 42.82354 84.4178 UND CLINTON 9-1-2000 NO N/A Ext. 20 CentraIPk 42.72789 84.4218 CIT INGHAM 8-22-2000 YES 2 Res. 21 Clark55 42.81374 84.4528 RES CLINTON 8-25-2000 NO N/A Ext. 22 CIark81 42.81422 84.4013 RES CLINTON 8-25-2000 NO N/A Ext. 23 Close 42.7564 84.5242 CIT INGHAM 8-21-2000 YES 11 Res. 24 College 42.63054 84.4852 CIT INGHAM 8-24-2000 YES 5 Res. 25 Cutler 42.8625 84.6893 UND CLINTON 8-19-2000 NO N/A Ext. 26 FarmL 42.70518 84.4804 CIT INGHAM 8-28-2000 YES 9 Res. 27 Forest 42.8621 84.6759 RES CLINTON 8-19-2000 NO N/A Ext. 28 Hawley 42.5187 84.3824 UND INGHAM 8-16-2000 NO N/A Ext. 29 Howe 42.84207 84.4137 RES CLINTON 8-19-2000 NO N/A Ext. 30 Krepps 42.9124 84.5073 AG CLINTON 8—23-2000 NO N/A Ext. 31 Loren 42.5207 84.3541 AG INGHAM 8-16—2000 YES 10 Res. 32 Marsh 42.7741 84.4124 CIT CLINTON 9-1-2000 YES 5 Res. 33 MasonPk 42.57729 84.4316 RES INGHAM 8-18-2000 NO N/A Ext. 34 Meijer 42.5674 84.3443 CIT INGHAM 8-18-2000 YES 12 Res. 35 Open 42.7564 84.5242 CIT INGHAM 8-21-2000 YES 7 Res. 36 Orchart 42.5838 84.4587 UND INGHAM 8-18-2000 YES 3 Res. 37 Pratt 42.90436 84.524 AG CLINTON 8-23-2000 YES 10 Res. 38 Price 42.9286 84.4418 RES CLINTON 8-25-2000 YES 5 Res. 39 Railway 42.72044 84.4788 CIT INGHAM 9-1-2000 NO N/A Ext. 40 Rosendom 42.77521 84.405 RES Clinton 9-1-2000 YES 2 Res. 41 Stall 42.79659 84.6052 RES CLINTON 8-21-2000 NO N/A Ext. 42 Tom 42.8543 84.7094 RES CLINTON 8—19—2000 YES 3 Res. 43 Upton 42.81598 84.4024 RES CLINTON 8-1-2000 NO N/A Ext. 44 VaughnE 42.5003 84.363 AG INGHAM 8-16-2000 YES 2 Res. 164 45 VaughnW 42.5003 84.363 AG INGHAM 8-16-2000 YES 2 Res. Appendix B. EMAP sampling location and their site property. Michigan subecoregion were classified into 6 types. Soil type are classified as: 1, Northern Michigan and Wisconsin Sandy Drift. 2, Western Michigan and Northeastern Wisconsin Fruit belt; 3, Southern Michigan and Northern Indiana Drift Plain; 4, Indiana and Ohio Till Plain; 5, Southern Michigan Fruit and Truck Belt. Agricultural land use around wetlands are based on 1 km radius around Site Code latitude longitude Ecoregion Soil Region 96 AGR MIWET03 44.39583 85.0242 2 1 0.166347 MIWETO4 44.43513 84.91 15 2 2 0 MIWE105 44.625 85.0552 2 1 0 MlWETl 0 45.36889 84.4081 1 1 0.049657 MIWETT 2 44.72083 84.4283 2 3 0.810269 MlWETl 8 43.26359 85.2996 3 3 0.776972 MIWET20 42.6521 7 85.5165 4 3 0 MIWET25 42.31386 84.1 737 5 4 0.842661 MIWET26 42.36821 84.2022 5 3 0.498948 MIWET31 42.57201 83.5506 5 4 0.518641 MIWET33 42.53601 83.6563 5 4 0.568018 MIWET35 42.33016 84.1287 5 3 0.646123 MIWET36 42.44701 83.9982 5 4 0.53634 MIWET39 42.31658 85.3603 4 3 0.276599 MIWET41 42.25543 85.3603 4 3 0.330213 MIWET42 41 .92935 85.8401 4 3 0.256192 MIWET43 41 .85054 85.75 4 3 0.383817 MIWET48 42.4021 7 85.3971 4 5 0.284899 MIWET49 41 .91848 85.7813 4 3 0.162268 MIWET50 42.14946 85.601 1 4 3 0.556 MIWET51 42.46603 85.3695 4 3 0.482928 MIWET52 42.49796 85.2206 4 3 0.605755 MIWET53 42.53533 86.0404 6 3 0.455534 MIWET54 42.64674 85.9228 6 5 0.56166 MIWET58 42.52446 86.0055 6 5 0.479886 MIWE159 42.54755 86.0037 6 5 0.338 MIWET60 42.46603 85.2243 4 3 0.144218 MIWET62 42.3913 84.977 4 3 0.671064 MIWET68 42.45788 83.9982 5 4 0.203087 MIWET69 42.45245 84.0055 5 4 0.002193 MIWET70 42.44973 84.0147 5 4 0.096715 MIWET71 43.77446 85.2132 2 3 0.224564 MIWET72 44.18278 84.8183 2 3 0 MIWET77 45.22147 84.3824 1 3 0.051222 MIWET78 45.55707 84.6531 1 3 0 MIWE180 45.71603 84.6725 1 2 0.519926 165