EFFECTS OF ANTHROPOGENIC DISTURBANCES AND BIOTIC INTERACTIONS

ON STREAM BIOTA IN GULF COASTAL PLAIN STREAMS

A Dissertation

Presented in Partial Fulfillment of the Requirements for the Doctor of Philosophy in the Graduate School of The Ohio State University

By Archis Robert Grubh, M.S. * * * * * The Ohio State University 2006

Doctoral Examination Committee: Lance Williams, Advisor Charles Goebel Andrew Ward Richard Moore

Approved by

______Advisor

Graduate Program in Natural Resources

ABSTRACT

Stream organisms have a complex relationship with the habitats they occupy, and

their relationships can vary greatly across spatial, temporal, or taxonomic scales.

Anthropogenic modification of ecosystems generally causes alteration to the temporal

regime of natural variation and disturbance. The effects of timber harvesting on stream

systems have been studied extensively in the Pacific Northwest and Northeast of the

United States, and indicate long-term deleterious effects. These disturbances resulted in

increase in sediment, discharge, temperature, and decrease in dissolved oxygen levels.

Although stream ecosystems are a product of historic geologic and climatologic

attributes, large-scale human disturbances can change the landscape. Over the past 200

years the forestlands of the southeastern Gulf Coastal Plain ecoregion of the United

States has experienced extensive pressure primarily from timber harvesting. In spite of

the known deleterious effects of timber harvesting on lotic systems, this region has not

received much attention.

In this study, macroinvertebrates and fish were used to quantify the effects of

timber harvesting, and its major byproduct, road crossing, in headwater streams in west- central Louisiana. The comparative study of macroinvertebrate assemblages during the pre-, during-, and post- timber harvest years did not show any significant difference.

Neither was there a significant different in macroinvertebrate assemblage between the reference stream and streams with varying levels of timber harvest activities. Although

ii no significant difference was detected in macroinvertebrate taxa at the annual scale, a

significant seasonal difference was detected. Similarly, scaling down from stream level

to mesohabitat level, significant difference was detected at smaller scale. This is

suggestive of assemblages responding to smaller temporal and spatial scales structured to

the biogeographic history.

The effects of road crossings on fish movement were significant when compared

to natural reaches of the streams in Fort Polk, Louisiana. Movement through various

bridge designs did not indicate a difference, which was a result of lower recaptures. An

inverse relationship was detected between water depth and fish movement. Lack of

significant relationship of movement with current velocity is a result of greater variation in the stream hydrologic conditions in the headwater reaches.

iii

DEDICATION

To my beautiful daughter,

Nazerene Shammah Grubh

iv

ACKNOWLEDGEMENTS

I would like to first thank my advisor, Lance R. Williams for his support,

encouragement, and guidance throughout my study, and stay in Columbus, Ohio. I

would also like to thank my committee members, Charles Goebel, Andrew Ward, and

Richard Moore for their support and insightful suggestions. Like many studies, a major dissertation is not complete with help and suggestions from fellow students, and close friends.

I would like to thank the Fort Polk, Louisiana crew for assisting me in the field and providing technical suggestions. Danny Hudson was my primary source of help while in Louisiana. I would also like to thank Mr. Lynn Bennett, who was in charge of the bunk house on the army base, we shared many insightful stories. I would like to thank Mr. Jevanse for sharing his excellent culinary recipes, and was able to enjoy his sumptuous meals on occasion. On several of my trips to Louisiana, I was able to get ample help from graduate and undergraduate friends from the Ohio State University,

Sarah Beck, Marie Schrecengost, Andrea Shyjka, Erin Rothman, Lauren Glockner, Katie

LaFay, Marsha Williams, and John Foltz. I enjoyed my first spicy hot boiled crawfish, served Cajun Style in Alexandria, Louisiana. I would like to thank Tim Bonner, from

Texas State University, and his crew from San Marcos, for helping out in field and sharing some important data.

v

I would also like to thank my parents, Robert Grubh, and Shailaja Grubh, and my

brother Kumudan Grubh for supporting me and making long distance calls from India for encouraging me. I remember Dave Ferguson, and Charles Woods, who have been my

long term friends, and have done everything possible to make me feel at home in the U.S. during my initial stay. During my second year of study, I met Raquel, my beloved wife,

who has since been a constant source of encouragement, and joy in my life. I remember

several barbecues we had on the back porch; in San Antonio over the several weekends I

went to visit her, before our wedding. I would like to express my sincere gratitude to my

mother-in-law, Noemi Herrera, for staying with us on several occasions and helping out with babysitting our daughter, Nazerene, and giving me and Raquel a much needed

break.

Funding for this project was supported by the School of Environmental and

Natural Resources (Ohio State University), and the Environmental and Natural Resource

Management Division (Fort Polk, Louisiana). Many thanks go to the staff members in

201 Kottman Hall on the university campus who have helped me in several ways. And last but not the least, I am eternally grateful to my Lord and savior, Jesus Christ.

vi

VITA

Education 1996…………………….. B.Sc., Major: Zoology; Minor: Botany, Chemistry; Manonmaniam Sundaranar University, India

1998…………………….. M.Sc., Major: Marine Biotechnology, Manonmaniam Sundaranar University, India

2002…………………….. M.S., Wildlife & Fisheries Sciences, Texas A&M Unviersity, Texas

Publications 2003…………………….. Grubh, A.R. and W.J. Mitsch. 2003. Distribution and abundance of macroinvertebrates in created wetland ecosystems. Pages 105-116 in Mitsch, W.J., L.Zhang, and C. L. Tuttle editors. Olentangy River Wetland Research Park at The Ohio State University Annual Report 2003, Columbus, Ohio. 2004…………………….. Grubh, A.R. and K.O. Winemiller. 2004. Ontogeny of Scale Feeding in the Asian Glassfish, Chanda nama (Ambassidae). Copeia 2004:903- 907. 2005…………………….. Williams, L.R., M.G. Williams, A.R. Grubh, E.E. Swinehart, R.W. Standage. 2005. Food Habits of the Federally Threatened Leopard Darter (Percina pantherina). American Midland Naturalist: in press.

FIELDS OF STUDY

Major Field: Natural Resources

vii

TABLE OF CONTENTS

Page Abstract...... ……………………………………………………..... ii Dedication………………………………………………………………... ii Acknowledgements……………………………………………………..... iv Vita………...…………………………………………………………….. vii List of Tables…………………………………………………………….. xi List of Figures……………………………………………………………. xiii

Chapter 1 Introduction…………………………………………………………….. 1 Objectives……………………………………………………………. 3 Literature Review……………………………………………………… 4 Biotic Response to physical disturbance in Fort Polk, Louisiana.. 4 Impacts of road crossings across streams on fish movement in Fort Polk, Louisiana……………………………………………… 7 Successional patterns of macroinvertebrates in relation to presence/absence of predators…………………………………… 9 Background history on the military activities with special reference to road crossings………………………………………. 10 References…………………………………………………………….... 17 Chapter 2 Impacts of timber harvesting on macroinvertebrate assemblage dynamics in western Louisiana headwater streams……………………. 30 Introduction…………………………………………………………….. 30 Methods………………………………………………………………... 33 Study Area…………………………………………………………… 33 Sampling Methods…………………………………………………… 34 Field Techniques…………………………………………………. 34 Statistical Analyses………………………………………………. 36 Results………………………………………………………………….. 38 Discussion……………………………………………………………… 43 References……………………………………………………………… 49 Chapter 3 Impacts of road crossings on fish movement in Gulf Coastal Plain streams……………...... 75 Introduction……………………………………………………………. 75 Methods………………………………………………………………... 78 Study Area…………………………………………………………… 78 viii

Sampling Methods……………………………………………… 80 Statistical Analyses……………………………………………... 81 Results…………………………………………………………….. 82 Discussion………………………………………………………… 85 References………………………………………………………… 88 Chapter 4 Effects of fish and invertebrate predators on benthic invertebrates in headwater streams of west-central Louisiana………………….. 102 Introduction……………………………………………………….. 102 Methods…………………………………………………………… 105 Study Area………………………………………………………. 105 Design and Methods………………………………………… 106 Experiment 1. Effect of fish absence on macroinvertebrate assemblage structure……………………… 106 Experiment 2. Effects of different fish feeding guilds on invertebrate assemblage…………………... 107 Statistical Analyses……………………………………………... 107 Results…………………………………………………………….. 108 Experiment 1. Effect of fish absence on macroinvertebrate assemblage structure….. 108 Experiment 2. Effects of different fish feeding guilds on invertebrate assemblage…………………... 109 Discussion………………………………………………………… 111 References………………………………………………………… 115

Chapter 5 Conclusions……………………………………………………….. 128 References………………………………………………………… 132

Appendix A Habitat data collected by seasons in the years 2003, 2004, and 2005, in Little Sandy, Odom, and Tiger creeks in Fort Polk, Louisiana………………………………….…………….…………... 134 Appendix B Presence/Absence of macroinvertebrate Taxa with scientific Nomenclature and their feeding guilds, collected in Little Sandy, Odom, and Tiger creeks during 2003-2005…………….…………... 136

ix

Appendix C List of fish species with scientific nomenclature collected in Red River, Sabine, and Calcasieu drainages during Summer 2003 and Summer 2004.………………………………………………………. 140 Appendix D Habitat data collected at three segments, at crossing, and at natural reach in the three drainages in Fort Polk, Louisiana.…….…………. 142 Appendix E Abundance of fish captured in different drainages collected during Summer 2003, and Summer 2004 in Fort Polk, Louisiana..………... 144 Appendix F Total fish captured and recaptures during Summer 2003 and Summer 2004 in Fort Polk, Louisiana….…………………………... 146 Appendix G Environmental parameters measured at each Exclosure/Enclosure during the field experiment during Summer 2003 in Fort Polk, Louisiana……………………………………………………………. 149 Appendix H Abundance of macroinvertebrate taxa with scientific nomenclature and their feeding guilds, collected in the Enclosure Experiment (I) during Summer 2003………………………………………………... 151 Appendix I Abundance of macroinvertebrate taxa with scientific nomenclature and their feeding guilds, collected in the Enclosure Experiment (II) during Summer 2003……………………………………………….. 154

Bibliography………………………………………………………… 157

x

LIST OF TABLES

Table Page 2.1 Sampling site description in the catchment basin affected by clear-cutting and new roads, during the year of timber harvest disturbance (2004). Stream characteristics, and geomorphological feature data was collected during summer 2005……………………………...………………………...... 63 2.2 Mean physical and chemical water quality data, and stream habitat quality data, collected from Red River drainage streams in the years 2003, 2004, and 2005 represented by pre-, during-, 64 and post- timber harvest activities respectively………………... 2.3 Regression equations describing the Red River drainage relationships between the % EPT and % Diptera taxa and the stream physical and chemical parameters. ANOVA results of 65 the whole regression model are also displayed…...... 2.4 Renkonen Similarity Index of sites across years encompassing the years 2003, 2004, and 2005 that was represented by pre-, 66 during-, and post- timber harvest period……………………… 2.5 Percent variance of partial canonical correspondence analysis on macroinvertebrate assemblage through timber harvest regime (2003 – 2005) in Ft. Polk, Louisiana, as explained by year (pre-, during-, and post- timber harvest years), stream (3 streams), habitat (3 mesohabitat), and season (4 levels). All the % variance explained in significant (P = 0.001) based on Monte 67 Carlo randomization test. …………………………………...…

3.1 Recapture percentage, and proportional daily movement of fish in the studied streams in Fort Polk, Louisiana, during the 94 summers of 2003 and 2004.…………………………….……... 3.2 Summary of Chi Square results on generalized linear model (GLM) analysis for the variables depth, velocity, discharge, and crossing type (categorical variable; viz., arch, box, ford, and 95 natural reach) on fish movement………………………..……... 3.3 Correlation (Spearman’s) of environmental parameters of the sampled streams, with the fish species richness, abundance, and movement……………………………………… 96

xi

4.1 Family Richness main effects and interaction, Experiment 1……………………………………………………. 118 4.2 Family Abundance main effects and interaction, Experiment 1……………………………………………………. 118 4.3 Nested ANOVA with Day nested within the Stream block. Day was at seven levels (0, 2, 5, 10, 15, 20, 30), and Stream at three (Drakes, Whisky Chitto, and Birds Creeks; Experiment 1). ………………………………………………….. 119 4.4 Family Richness main effects and interaction, Experiment 2. …………………………………………………… 120 4.5 Family Abundance main effects and interaction, Experiment 2. ………………………………………………...… 120 4.6 Generalized linear model results of the effect of Fish Guilds on individual macroinvertebrate feeding guilds. Fish guilds were represented by Algivore, Invertivore, Mix, Piscivore, Open, and Control (no fish). The macroinvertebrate (Invert) feeding guilds were represented by Collector, Filterer, Piercer, Predator, Scraper, and Shredder. The Control, and Collector were held as intercepts in the interaction model……………………………… 121

xii

LIST OF FIGURES

Figure Page 1.1 Early bridge designed used to convey light weight vehicles ….…. 24 1.2 Simple concrete topped bridge with support beams ……………... 24 1.3 Single pipe culvert topped with metal road.………………. …….. 25 1.4 Single pipe culvert armored with interlocking blocks …………… 25 1.5 Ovoid-top culvert used for bridge construction ………………….. 26 1.6 Three 10” box culvert design blocked by plant and woody debris . 26 1.7 Pipe culvert crossing, also serving as wet fords during high rainfall events…………………………………………...... 27 1.8 PVC pipes installed below the road to allow cross flow during flood events………………………………………………………... 27 1.9 Ford crossing design with interlocking concrete blocks buried in the stream bed substrate…………………………………………… 28 1.10 Ford crossing showing erosion because of unpaved stream bed………………………………………………………………… 28 1.11 New crossings topped with poured concrete, withstanding heavy mechanized vehicles……………………………………….. 29 1.12 New Arch culvert design to reduce impacts by high variation in discharge………………………………………………………... 29

2.1 Study area in the Peason Ridge Training Area (PRTA) in Fort Polk, Louisiana, with three adjacent drainages. Sampling sites located on the streams in Red River drainage are – A is Little Sandy Stream, B is Tiger Stream, and C is Odom Stream……………………………………………. 68

xiii

2.2 Relationship between site locations, and physical and chemical conditions of the stream. Proportion of variance accounted for by the PCA axes: Axis 1, 28.946%, and axis 2, 14.57%. PCA axis I showed a linear bed substrate gradient, water quality conditions, and stream bank condition. PCA axis II showed a gradient along the stream bed dimensions. With increasing PCA II scores, stream environment changes from deeper to shallower and wider 69 areas……………………………….……………………………… 2.3 Macroinvertebrate assemblage composition represented by the mean relative abundance of Ephemeroptera, Plecoptera, and Trichoptera (EPT, open circles), and the relative abundance of Diptera (filled circles) within the total assemblage shown as a 70 function of pre-, during-, and post- timber harvest years………… 2.4 Macroinvertebrate assemblage patterns in mean values of a) family richness; b) family abundance (log transformed); c) Shannon Weiner index; and d) Evenness index, in the Red River drainage, Louisiana, across the of pre-, during-, and post- timber harvest 71 period …………………………………………………………….. 2.5 Proportion of functional feeding groups of macroinvertebrates based on percent abundance in Red River drainage streams in 72 Louisiana, across the years 2003, 2004, and 2005……………….. 2.6 Results from detrended correspondence analyses (DCA) based on macroinvertebrate assemblages addressing the spatial and temporal component of streams in the Red River Basin, Louisiana, from 2003 to 2005. Site scores coded by year did not show a distinct clustering (not shown). a) DCA site scores for mesohabitats showed a gradient along the first axis with pool, riffle, and run scores. Pool scores had lowest values along Axis 1, whereas Riffle scores were highest. b) The DCA site scores for the seasonal component showed separate grouping of the spring, summer, fall, and winter scores occupying each quadrant. A longitudinal gradient is detected along the first axis with Winter scores having lowest values, and Summer scores, the 73 highest……….……………………………………………………

xiv

2.7 Ordination biplot depicting the relationship between environmental parameters and site scores. a) Axis I shows a

water quality, and habitat features gradient. Axis II shows a stream bed substrate composition, and dimensions gradient. The site scores show a clustering pattern of summer plots with lower

scores, and winter plots with higher scores along Axis 1. b) Same ordination with site scores enveloped by stream type……... 74

3.1 Mark and recapture conducted on fish over road crossing across the streams belonging to the Red River drainage – A- Reaugaulle, B- Stagestand, C- Odom, and D- Squirrel branch; the Sabine River drainage – E-Martin; and Calcaseiu River drainage – F- Comrade. Crossing type on Stagestand, Squirrel branch, Martin, and Comrade streams was box culvert. Reaugaulle stream had ford type crossing, and Odom stream had arch type crossing over it……………………… 97 3.2 Description of the fish marking strategy across the road crossing, and the natural reach along the stream, with stream 98 flowing from top to the bottom of the page …………………...... 3.3 Maximum rainfall event for the month, and air temperature data for the Fort Polk region of Louisiana. ………………………. 99 3.4 Comparison of proportional daily movement of fish across different crossing types showing greater movement in the natural reach. The ANOVA results show no significant difference in movement between the crossing types compared to the natural 100 reach (df = 6 F = 0.09, p-value = 0.51 N = 40)………………... 3.5 Canonical correspondence analysis on relationship between the fish species, and the environmental factors they were exposed to, in the three drainages (viz., Red River, Sabine, and Calcasieu) in Fort Polk, Louisiana. The first two axes explained 60 % of the 101 variance significantly (p-value = 0.001)…………………………..

xv

4.1 Sampling conducted on streams in Fort Polk, Louisiana. A- Drake’s creek, B- Whisky Chitto creek, and C- Bird’s creek…….. 122 4.2 Mean species richness of macroinvertebrates sampled over time interval did not show statistically significant differences (F6, 0.48, P=0.81: Experiment 1)……………………………………. 123 4.3 Mean species abundance of macroinvertebrates sampled over time inverval did not show significant differences (F6, 2.23, P-0.06; Experiment 1)……………………………………. 123 4.4 Successional mean macroinvertebrate abundance from Day 0 to 124 Day 30, in absence of fish, based on feeding guilds……………… 4.5 Proportional abundance of top five abundant 125 macroinvertebrate taxa sampled over time interval………………. 4.6 Mean macroinvertebrate species richness in the various fish enclosures did not show significant difference (F6, 0.6, P=0.7; Experiment 2)……………………………………… 126 4.7 Mean macroinvertebrate taxa abundance in the various fish enclosures did not significant difference (F , P=0.3; 6, 1.3 126 Experiment 2)……………………………………………...……… 4.8 Proportional abundance of top five macroinvertebrate taxa 127 occuriing in the various Fish Guild enclosures……………………

xvi

CHAPTER 1

Introduction

In the eastern United States timber harvesting has had deleterious effects on

ecosystems over the past couple of centuries. Timber harvesting activities can greatly impact aquatic systems because of the tight linkages between streams, wetlands, lakes and the watersheds they drain (e.g., Hynes, 1975; Bormann and Likens, 1979). Major elements of timber harvesting include skidding, yarding, site preparation for replanting, road construction, and road crossings, all of which can have pervasive influence on streams. The deleterious effects of timber harvest on streams range from local (e.g., sediment release, alteration of riffle-pool morphology, etc.; Campbell and Doeg 1989) to regional effects (Williams et al. 2002).

Roads and bridges can alter the development of shorelines, stream channels, floodplains, and wetlands. Because of the energy associated with moving water, physical effects often propagate long distances from the site of a direct road incursion (Richardson et al. 1975). Alteration of hydrodynamics and sediment deposition can result in changes in stream geomorphology many kilometers away, both down- and up-stream of a road crossing. It is not always possible to predict the nature of such response to channel and shoreline alteration, and depends on the sequence of flood and sedimentation events after the alteration is done. Roads on floodplains can redirect water, sediment, and nutrients between streams and wetlands and their riparian ecosystems, to the detriment of water 1 quality and ecosystem health. Roads are among the many human endeavors that impair natural habitat development and woody debris dynamics in forested floodplain rivers

(Piegay and Landon 1997).

Road crossings commonly act as barriers to the movement of fishes and other

aquatic (Furniss et al. 1991). Because road crossings and other human developments fail to provide for fish passage, they serve as migration barriers to

headwater populations of salmonids fishes that are naturally migratory, and now only

exist as fragmented headwater isolates (Kershner et al. 1997; Rieman et al. 1997). Many riverine fishes, especially Salmonids, actively move into seasonal floodplain wetlands and small valley-floor tributaries to escape the stresses of main-channel flood events

(Copp 1989), but valley-bottom roads can destroy or block access to these seasonally important habitats (Brown and Hartman 1988). Persistent barriers may encourage local selection for behaviors that do not include natural migration patterns, potentially reducing both the distribution and productivity of a population.

For assessing the health of an aquatic ecosystem, organisms that live in these environments are used as indicators of degradation. Biological assessments have become a standard practice throughout the U.S. and the world (Karr and Chu 1999). Bioassessing impacts of timber harvesting activities on biota has a particularly long history in eastern

U.S. streams (Tebo 1955) and has become an essential part of stream monitoring programs for most states (Davis et al. 1996). The growing realization that maintaining water quality (i.e., water chemistry) standards alone does not necessarily protect biological integrity (Karr and Chu 1999) has led to the proliferation of stream bioassessment programs. The successful implementation of bioassessment for streams

2 and the growing recognition of the value of wetlands have led researchers to apply

bioassessment to wetlands over the last 10 years (Rader et al., 2001). Since the passing

and legal enforcement of the Clean Water Act, point-source pollution (i.e., that from a

defined location such as a pipe) has been ameliorated greatly. Today, however, aquatic

systems primarily face different threats, e.g., non-point-source pollution produced from land use practices such as agriculture, timber harvesting, and urban and suburban

development. Mitigation of detrimental effects associated with non-point-source pollution has come primarily in the form of best management practices (BMPs) that are

designed to reduce the amount of onsite damage to habitats, as well as limit off-site effects related to the export of harmful materials (e.g., sediment, toxics). BMPs are

implemented widely across the U.S. as part of forestry activities.

OBJECTIVES

To determine the effects of timber harvesting, and its byproduct (road crossings)

on lotic habitats, I used macroinvertebrates and fish as indicators of ecosystem health.

Although long-term studies have monitored the impacts of timber harvesting in high-

gradient streams of the Pacific Northwestern U.S., little is known about effects in

southern coastal streams (Williams et al. 2005). The timber harvesting activities by the

U.S. Army in Fort Polk, Louisiana provided an excellent opportunity to study its effects

on streams in the Gulf Coastal Plain. My specific objectives for this study were:

1. To assess habitat associations and macroinvertebrate assemblage structure on a

spatial and temporal scale, and to monitor the change in their assemblage structure

3 before, during, and after a large-scale logging project in the Red River drainage (Little

Sandy, Tiger, and Odom, streams) in Ft. Polk, Louisiana.

2. To assess the effects of different bridge designs on fish movement in relation to natural reach with no road crossings on streams in Ft. Polk, Louisiana.

3. To determine the effects of presence/absence of predators on the successional dynamics of macroinvertebrates.

LITERATURE REVIEW

Biotic response to physical disturbance in Ft. Polk, Louisiana

Stream ecologists have long recognized that physical environmental heterogeneity

influences patterns of species richness and abundance (Hynes 1975). The intensity,

frequency, and area of disturbance may act to structure assemblage of stream organisms

(Resh et al. 1988, Huston 1994, Townsend et al. 1997). Increasing disturbance intensity may remove more individuals, more species, and more of the food resources necessary for recolonization. If disturbance frequency, however, is greater than the rate of

competitive exclusion, diversity may be maintained at a high level (Huston 1979).

Increasing the extent of disturbance removes more individuals, thus reducing the local

pool of potential colonists. Although all three aspects of disturbance can affect species

richness (Sousa 1985), little is known about their interactions (Death and Winterbourn

1995).

Although riparian areas are subjected to a variety of natural disturbances (e.g., flooding, landslides, and wildfire) that alter habitat structure and biodiversity (Naiman

1998, Ilhardt et al. 2000), human disturbances, like timber harvesting, can alter riparian 4 forests in complex and often synergistic ways (Gregory et al. 1991, Williams et al. 2002).

Because the riparian zones play a major role in buffering watersheds, many government

agencies and private organizations, protect a riparian corridor around streams. The

importance of understanding how landscape scale disturbances influence the quality of

aquatic ecosystems becomes clear when one examines the importance of headwater

systems within the context of regional landscapes.

Aquatic communities have been used by government agencies as a measure of

water quality and watershed condition since the early 1980’s (Fausch et al. 1990). State

and federal entities are required to establish water quality standards to meet Clean Water

Act (CWA) objectives because the primary objective of the CWA is to restore and

maintain physical, chemical, and biological integrity of U.S. surface waters. Biological

integrity is the “capability of supporting and maintaining a balanced, integrated, adaptive

community of organisms having a species composition, diversity, and functional

organization comparable to that of natural habitats of the region” (Karr and Dudley

1981). A complete assessment of stream health involves the evaluation of fish and

macroinvertebrate assemblages, aquatic habitat and water quality, and adjacent riparian

condition.

Several studies indicate that the changes in stream macroinvertebrate assemblages

caused by forest harvesting may be long-term (Erman et al. 1977, Haefner and Wallace

1981, Silsbee and Larson 1983), but these studies have been conducted in Pacific

Northwest or in upland eastern streams. Very few studies have examined inputs of timber harvesting on streams in the southeastern U.S. (Williams et al. 2002, 2005).

Studies by Williams et al. (2002) on disturbances (U.S. Forest Service in Arkansas) failed

5 to detect effects of timber harvesting on biota as it was studied ten years after the regime

(probably because of the high resilience of biotic communities inhabiting these regions).

The Red River drainage in Ft. Polk, Louisiana, is fed by streams from the northern region of Peason Ridge Wildlife Management Area, and drains into the Kisatchie Bayou. The primary vegetation is a mixture of loblolly (Pinus taeda) and longleaf pine (P. palustris) with hardwood trees like bald cypress Taxodium distichum, white oak Quercus alba, water oak Quercus nigra, and magnolia Magnolia grandiflora. Studies on fish and macroinvertebrate assemblage structure conducted in this region (Williams et al., 2005) have been done prior to the timber harvesting regime. My study spanned throughout the whole timber harvesting regime (i.e., I included data collected before, during, and after timber harvesting) from Summer 2003 to Summer 2005 in Fort Polk, Louisiana. Thus, this study will be able to better elucidate the impacts of timber harvesting and resilience of the aquatic biota.

The seminal paper by Resh et al. (1988) has recently drawn a lot of attention to disturbance concepts applied to lotic ecosystems. This has helped recent work on stream ecology (Southwood 1988, Poff 1992, Townsend and Hildrew 1994, Williams et al.

2002, Lamouroux et al. 2004) to focus on the role of disturbance in structuring these systems at a landscape level. Disturbances are ecological events, but if their temporal distribution is predictable, ecological responses may be small because organisms and communities have adjusted to them in evolutionary terms. Differences within or among streams with respect to ecological responses to specified disturbance intensities can be hypothesized to reflect differences in disturbance regimes among sites, but more studies need to be done to test such hypotheses (Holling 1973; Poff 1992).

6

Impacts of road crossings across streams on fish movement in Ft. Polk, Louisiana

Small streams are a good model system to examine patterns of stability and

variability across spatial and temporal scales in assemblage structure (Wiens et al. 1986,

Grossman et al. 1990, Lammert and Allan 1999), as they are naturally subject to a high

level of disturbance regime (Peckarsky 1983, Matthews 1986, Matthews et al. 1988).

Systems experiencing environmental disturbances at irregular intervals would be

influenced differently (Wiens et al. 1986, Ward 1989). Disturbance is recognized

increasingly as playing an important role in the structure and dynamics of communities

(Sousa 1984).

Movement is a critical process that allows fishes to meet their resource needs in

spatially and temporally variable stream environments (Schlosser and Angermeier 1995;

Fausch and Young 1995; Fausch et al. 2002). Small scale movement allows individual

fish to occupy the most suitable habitats for survival and growth (Gowan and Fausch

2002), and at larger scale, permits migrations between habitats used by different life

history stages, exploitation of refugia from large-scale disturbances, gene flow for small

populations, and colonization or recolonization of unoccupied habitats (Brown and

Kodric- Brown 1977; Peterson and Bayley 1993; Northcote 1997). These larger-scale

movements are essential for persistence of individual populations and metapopulations

(Meffe and Sheldon 1990; Schlosser and Angermeier 1995). Thus, proper knowledge of

movement is necessary to improve our understanding of stream fish ecology. The need

for conservation efforts with respect to fish movement is necessary because modifications by humans may disrupt the movements (Warren et al. 1997; Wiens 2001). 7 One of the major anthropogenic disturbances to streams is timber harvesting, and it typically results in removal of riparian vegetation that stabilizes soils, holds water and prevents erosion into the stream channel. Building road crossings, for transporting timber on heavy-duty vehicles, is also another major byproduct of timber harvesting that causes habitat alteration. Barriers to movement such as dams and road crossings have an obvious impact on fish populations by preventing migrations, gene flow, and colonization

(Warren and Pardew 1998; Pringle et al. 2000). Fishes that readily move under certain ecological conditions may fail to do so in modified situations (Gowan et al. 1994).

Identifying ecological factors that drive dispersal-mediated life history events (e.g., spawning migrations) is a key step in predicting how fishes will respond to natural and anthropogenic changes in environmental conditions (Railsback et al. 1999). Managing fish populations is far more complicated when factors influencing movement vary among species. Diverse life histories and habitat requirements suggest that factors may also be species specific in fishes. For many stream-fish species, movement has been associated with physical structure (i.e., presence of cover; e.g., Aparicio and Sostoa 1999; Harvey et al. 1999; Gilliam and Fraser 2001). Identifying common factors that influence movement across species may facilitate management efforts that benefit entire assemblages.

Despite management implications, few studies have linked movement of stream fish to ecological factors (Gowan et al. 1994; Gilliam and Fraser 2001). Most studies that have examined ecological correlates of movement, have been limited in the number of factors and/or attributes of movement examined, which leads to an incomplete understanding of movement (Ims and Hjermann 2001). Furthermore, virtually no studies that have explicitly examined how ecological correlates of movement, vary within an

8 assemblage of fishes. The main objective of this part of my study was to identify factors

associated with the movement of fishes in a network of streams over a period of three years. In addition to annual variation in movement, this fish mark, release, and recapture study would also integrate additional pre- and post- road crossings installation sites. I will examine relationships between each attribute of movement and a suite of predictor variables related to reach specific habitat characteristics and attributes of individual fish species.

Successional patterns of macroinvertebrates in relation to presence/absence of predators

Of the many variables that influence the distribution and abundance of organisms in aquatic communities, predator-prey interaction can affect the structure in several ways (Connell 1975, Glasser 1979, Sih 1982). Predators can drastically reduce the density of individuals by indiscriminate feeding in proportions in which they are found in habitats. Similarly, absence of predators could cause an increase in their populations.

Thus, understanding the influence of predators and resource supply on food webs is important in community ecology (Hunter and Price 1992, Nystrom et al. 2003). Even after comparing theoretical and empirical studies there seems to be no conclusion whether ecosystem function is controlled by resource availability (bottom-up) or predation (top-down) (e.g., McQueen et al. 1989, Polis and Winemiller 1996, Williams et al. 2003). This could be attributed to immigration by macroinvertebrates (Flecker 1984), predator avoidance mechanism (Allan 1982), risks of predation-associated feeding, lack of competitive dominants (Allan 1982), and differences in sampling and habitats of individual studies (Dhal 1998). Thus the prey population structure will depend on the

9 ecological characteristics of organisms (predator-prey relationship), and the environment

(Power 1992).

Primary producers support most food webs in any ecosystem, but allochthonous

input plays a major role in the stream system (Vannote et al. 1980, Polis and Strong 1996,

Sabo and Power 2002). Increase in terrestrial detritus input in natural streams increases

abundance in all trophic levels (Nisbet et al. 1997, Wallace et al. 1997, Richardson 1991).

Allochthonous input from surrounding riparian forest may reduce the impact of drift- feeding fish on the abundance of benthic prey (Dhal and Greenberg 1996, Nakano et al.

1999). Predator-prey relationship is complex, in that if the predator is not selective, it

might still cause loss of diversity and local extinction of relatively rare species, especially seen in short food chain systems (Peckarsky 1984). In contrast, prey species diversity may increase if predators selectively remove keystone predators (dominant competitors),

thus allowing fugitive species (inferior competitors) to colonize the habitats.

Background history on the military activities with special reference to road crossings

The U.S. Army has maintained a military presence at Fort Polk from the

Louisiana Maneuvers of 1939 until the present. This region was chosen because the

historic logging had left the landscape resembling European battlefields during WWI and

facilitated the testing of new fast moving cavalry maneuvers. In order to ensure

uninhibited maneuverability, various stream and river crossing designs were tested and

implemented by the U.S. Army Corps of Engineers. Logistical sense dictated that

permanent stream and river crossings be constructed on obvious maneuver corridors that

10 were influenced by dominant terrain. This need lead to the construction of several types

of crossings on Fort Polk.

Prior to the establishment of Fort Polk as a military base, various logging companies placed small number of stream crossings during the course of timber harvesting activities. These crossings varied with stream size, but were predominantly

identified as ford and span type crossings. Fords were utilized for small first order crossings, with spans being limited to larger downstream sites. Unimproved ford

crossings were first implemented for horse drawn timber skidding. Stream banks were

graded and back sloped with no effort being made to harden the stream bottom.

Compacted sand substrates provided an acceptable level of stability for limited use. The

next evolution of ford crossings consisted of heavy sacrifice timbers that were laid parallel to stream flows at perpendicular crossing sites. These timbers were forced into the soft sand bottom substrates and created a stable platform capable of handling heavier skid loads as well as wheeled traffic. Some of these fords are still in existence today and are restricted to limited use. An interesting by product of timber-hardened fords is that the high silica content in the streams has “fossilized” the timbers through mineral replacement. This phenomenon was noted at one individual stream site prior to planned construction upgrades.

The early bridges were all wood constructions with support timbers driven into the stream bottom. The most prevalent design consisted of paired vertical supports with flat hewn lateral supports attached in an X fashion. Horizontal beams were affixed at the appropriate vertical grade and overtopped with heavy square beams running the length of

the span. Smaller 2’x12’ or smaller planking was then attached in order to create a

11 passable surface. Safety rails were sometimes limited to low linear planking, but the

remnants of handrails have been observed. This bridge type was probably intended only to support wagon or light wheeled traffic and differs from tram crossings (Fig. 1).

Industrial logging efforts were supported by small tramlines, which necessitated the construction of sturdier span crossings due to increased weight. Construction designs appeared to be similar to that described above, but with increased supports and larger timber size. Some remnant pilings have been observed with pitch or pine tar treatments, but this method of preservation cannot be reliably attributed to original construction and could be the result of post-construction maintenance efforts.

During the establishment and modernization of Fort Polk, major surfaced roadways were constructed that brought about the need for durable spans capable of handling heavy loads. The two primary thoroughfares were Artillery road and Lookout road, which run east west across the major land area of Fort Polk. Stream drainages are predominately north south and required the construction of several span crossings

capable of handling the rigors of military traffic. The primary design utilized creosote

timbers driven into the stream with heavy cross bracing. Planking consisted of heavy

8”X8” creosote timbers placed over the support structure and overlain with road material

Safety rails were also constructed of heavy creosote treated timbers bolted to the

bracings. Most of these bridges were upgraded with either heavier timber pilings or

concrete pilings with concrete top supports. Flat reinforced concrete slabs or U shaped

concrete beams were bolted in atop the pilings, concrete or metal safety rails were

installed, and the road surfaces were upgraded (Fig. 2).

12 Culverts are one of the most numerous crossing types on Fort Polk and vary

according to the size and purpose of the crossing. Round single culverts from 30 inches

to 6 feet are the most common. At intermittent stream crossings, the bed area is

excavated and replaced with washed gravel. The culvert is then placed at the stream grade and backfilled prior to overtopping with road material. Some single culvert crossings are backfilled with concrete prior to overtopping, with the culverts being cut to an angle equal to the bed slope (Fig. 3). These small sites are usually not armored to prevent erosive hydraulic flow, however, in limited instances they are either armored with riprap or interlocking block (Fig. 4). Hybrid squash culverts have also been used.

These culverts are typified by their flat bottoms and ovoid tops (Fig. 5). They are placed in a similar manner as described above. One example of a “triple threat” culvert is still in use on Fort Polk. This type of crossing is characterized by the use of three 30” culverts

placed below grade. Concrete is backfilled over the pipes, with low wing walls being

formed as well. The road surface is then placed over the top to a height of approximately

10 feet. This type of crossing is highly susceptible to debris damming and beaver

activity, causing periodic flooding even during low-grade precipitation events. Debris

must be regularly cleaned out to prevent road damage during storm events (Fig 6).

Another hybrid design is one that can only be considered as a culvert/ford. This

type of crossing utilizes three 30” concrete culverts emplaced slightly below stream grade

and overtopped with a low concrete crossing. During normal flows, water is confined to

passage through the culverts and underneath the road, creating a dry crossing. During

moderate rain events, the stream will overtop the crossing, but still allow passage as a wet

ford (Fig. 7).

13 Box culverts are used to great extent because of their ease of placement and high weight capacities. Larger box culvert crossings utilize culverts that are 4’-6’ height by

10’ width, arranged in a stack or “shotgun” fashion. Streams are diverted around the

construction site and the streambed is excavated and backfilled with gravel or poured

concrete. Poured concrete headwalls are constructed on the extreme sides of the expanse.

Culverts are then placed on the prepared bed surface and backfilled. Concrete is then

used as the crossing surface, with side sloping, expansion joints, and PVC drains being

installed. This crossing type varies from one to as many as four box culverts depending

upon drainage and is typically constructed at a 25-year flood height. Some crossings on

higher gradient streams are constructed to allow passage for 50-year floods. Approaches

are long concrete causeways with PVC piping installed at intervals to allow cross flows

during flood events (Fig. 8). Most of these crossings are installed at or slightly below

streambed grades. Some slight undercutting has been noted beneath the culverts at sights

that were installed at grade and debris pile-ups must be periodically removed at these

sites.

The more current span designs consist of square concrete pilings driven into the

streambed and adjacent approach slope. Square beams are attached to the top of the

pilings prior to the placement of U shaped reinforced concrete beams being placed

linearly for the length of the span. The roadbed usually consists of poured concrete with

expansion joints, but some spans have asphalt surfacing. These structures are also characterized by having concrete armored slopes running from the top of the approach slope to a point slightly above the normal stream gradient.

14 Ford crossing design has seen the most innovation at Fort Polk. Fords are the

most cost effective, feasible, and least intrusive construction method used. Original

designs called for bed excavation at the crossing site, followed by the placement of a wire

wrapped packet of gravel often referred to as the “burrito”. The intent of this activity was

to allow hyporheic flows that would normally be impeded by bed compaction. The

original content of the packet consisted of limestone gravel. After the bed packet was

emplaced, washed gravel was spread over the top prior to the placement of interlocking

concrete blocks that created a firm crossing substrate (Fig. 9). The blocks were placed at a grade height equal to the natural streambed prior to construction. Water quality and periphyton assemblages were monitored above and below the site in order to discern any detrimental effects. It was determined that the limestone packet was creating a higher pH gradient below the crossing (personal communication, Hudson 2004). Periphyton assemblages were also shown to be moderately effected by the shift in water quality

(personal communication, Hudson 2004). It was further determined that the pavestone blocks needed to be placed below grade, as the difference in roughness coefficients and slight grade miscalculations created an upstream pooling effect. Further observations concluded that the crossing approach should be hardened with pavestones for its full length in order to decrease erosion (Fig. 10). As a result of this data, washed gravel was used in the packet, the crossing grade was lowered to allow better flows, and approaches were hardened to prevent erosion. The interlocking blocks began to break apart after a few years of heavy mechanized traffic and required frequent replacement. Future crossings were modified with poured concrete instead of interlocking blocks (Fig. 11).

15 New crossings were also lowered to 6 inches below grade in order to allow a natural bottom substrate to be established at the crossing site.

Recent attempts to minimize impacts to natural hydrology have resulted in the construction of arch culverts (Fig. 12) and modified placement techniques for box type culverts. Round culverts and improperly graded box culverts have been shown to negatively effect natural passage of fish, even in low gradient systems. Round culverts greatly increase hydraulic flows to a level above the swim capability of small fish. This can cause passage barriers that limit immigration to upstream sites. Arch culverts maintain natural bottom substrates, limit hydraulic changes, and allow free immigration of stream biota. Debris accumulation is also minimal due to the lack of vertical flow impediments. New box culvert crossings are designed in such a fashion as to allow

(when possible) for the use of a single culvert placed well below grade and capable of allowing the passage of a near bank-full discharge. New corrugated steel set up culverts allow for a broader range of size selections tailored for each stream width. These designs have been discussed for future crossings, but have not been constructed to date.

Stream crossing design at Fort Polk is ever evolving and predicated on the capability of handling heavy military traffic. Environmental considerations in design seek to minimize stream impacts while still focusing on mission support. The end result is a team approach that involves military planners, engineers, and scientists coming together to accomplish a common goal. Much of the history of Fort Polk crossings was provided by J.D. Hudson, Environmental Biologist, Fort Polk, Louisiana.

Much of the information on history of stream studies have focused on habitat relationships, including impacts of disturbances (Allan 1995). To fully understand how

16 aquatic assemblages are structured, both biotic and abiotic components of ecosystems must be considered. This dissertation simultaneously examines effects of both abiotic and biotic interactions in structuring stream communities in Gulf Coastal Plain drainages of west-central Louisiana.

REFERENCES

Allan, J. D. 1982. The effects of reduction in trout density on the inverebrate communtiy of a mountain stream. Ecology 63:1444-1455.

Allan, J. D. 1995. Stream ecology: structure and function of running waters. Chapman and Hall, London.

Aparicio, E., and A. de Sostoa. 1999. Pattern of movements of adult Barbus haasi in a small Mediterranean stream. Journal of Fish Biology 55:1086-1095.

Bormann, F. H., and G. E. Likens. 1979. Pattern and Process in a Forested Ecosystem. Springer-Verlag New York Inc, NY.

Brown, J. H., and A. Kodric-Brown. 1977. Turnover rates in insular biogeography: effect of immigration on extinction. Ecology 58:445-449.

Brown, T. G., and G.F. Hartman. 1988. Contribution of seasonally flooded lands and minor tributaries to the production of coho (Oncorhynchus kisutch) salmon in Carnation Creek, British Columbia. Transactions of the American Fisheries Society 117:546–551.

Clark, J. 1991. Disturbance and population structure on the shifting mosaic landscape. Ecology 72:1119-1137.

Connell, J. H. 1975. Some mechanisms producing structure in natural communities: a model and evidence from field experiments. . Pages 460-490 in M. L. Cody and J. M. Diamond, editors. Ecology and Evolutions of Communities. Belknap, Cambridge.

Connell, K. W. 1978. Diversity in tropical rain forests and coral reefs. Science 199:1302- 1310.

Copp, G. H. 1989. The habitat diversity and fish reproductive function of floodplain ecosystems. Environmental Biology of Fishes 26:1–27.

17 Dahl, J. 1998. Effects of a benthivorous and a drift-feeding fish on a benthic stream assemblage. Oecologia 116:426-432.

Dahl, J., and L. Greenberg. 1996. Impact on Stream benthic prey by benthic vs. drift feeding predators: a meta-analysis. Oikos 77:177-181.

Davis, W. S., B. D. Snyder, J. B. Stribling, and C. Stoughton. 1996. Summary of state biological assessment programs for streams and wadeable rivers. EPA 230-R-96- 007, Office of Policy, Planning, and Evaluation, US Environmental Protection Agency, Washington, DC.

Death, R. G., and M. J. Winterbourn. 1995. Diversity patterns in stream benthic invertebrate communities: the influence of habitat. Ecology 76:1446-1460.

Erman, D. C., J. D. Newbold, and K. B. Reby. 1977. Evaluation of streamside bufferstrips for protecting aquatic organisms. Technical Completion Report No. 165, California Water Resources Center, University of California, Berkeley.

Fausch, K. D., C. Gowan, K. Richmond, and S. C. Riley. 1995. The role of dispersal in trout population response to habitat formed by large woody debris. Bulletin Francais de la Peche et de la Pisciculture 338:179-190.

Fausch, K. D., J. Lyons, J.R. Karr, and P.L. Angermeier. 1990. Fish communities as indicators of environmental degradation. American Fisheries Society Symposium 8:123-144.

Fausch, K. D., C. E. Torgersen, C. V. Baxter, and H. W. Li. 2002. Landscapes to riverscapes: bridging the gap between research and conservation of stream fishes. Bioscience 52:483-498.

Fausch, K. D., and M. K. Young. 1995. Evolutionary significant units and movement of resident stream fishes: a cautionary tale. Pages 360-370 in J. L. Nielsen, editor. American Fisheries Society, Symposium 17, Bethesda, Maryland.

Fisher, S. G. 1990. Recovery processes in lotic ecosystems: limits of successional theory. Environmental Management 14:725-736.

Flecker, A. S. 1984. The effect of predation and detritus on the structure of a stream insect community: a field test. Oecologia 64:300-305.

Furniss, M. J., T. D. Roelofs, and C. S. Yee. 1991. Road construction and maintenance. American Fisheries Society Special Publication 19:297-323.

Gilliam, J. F., and D. F. Fraser. 2001. Movement in corridors enhancement by predation threat, disturbance, and habitat structures. Ecology 82:258-273.

Glasser, J. W. 1979. The role of predation in shaping and maintaining the structure of communities. American Naturalist 113:631–641. 18 Gowan, C., and K. D. Fausch. 2002. Why do foraging stream salmonids move during summer? Environmental Biology of Fishes 64:139-153.

Gowan, C., M. K. Young, K. D. Fausch, and S. C. Riley. 1994. Restricted movement in resident stream salmonids: paradigm lost? Canadian Journal of Fisherriees and Aquatic Sciences 51:2626-2637.

Gregory, S. V., F. J. Swanson, W. A. McKee, and K. W. Cummins. 1991. An Ecosystem Perspective Of Riparian Zones. Bioscience 41:540-551.

Grossman, G. D., J. F. Dowd, and M. Crawford. 1990. Assemblage stability in stream fishes: a review. Environmental Management 14:661-671.

Haefner, J. D., and J. B. Wallace. 1981. Production and potential seston utilization by Parapsyche cardis and Diplectrona modesta (Trichoptera : Hydropsychidae) in two streams draining contrasting southern Appalachian watersheds. Environmental Entomology 10:433-441.

Harvey, B. C., R. J. Nakamoto, and L. J. White. 1999. Influence of large woody debris and a bankfull flood on movement of adult resident coastal cutthroat trout, Oncorhynchus clarki, during fall and winter. Canadian Journal of Fisherriees and Aquatic Sciences 56:2161-2166.

Holling, C. S. 1973. Resilience and stability of ecological systems. Annual Review of Ecology and Systematics 4:1-23.

Hunter, M. D., and P. W. Price. 1992. Playing chutes and ladders: heterogeneity and the relative roles of bottom-up and top-down forces in natural communities. Ecology 73:724-732.

Huston, M. A. 1979. A general hypothesis of species diversity. American Naturalist 113:81-99.

Huston, M. A. 1994. Biological diversity: the coexistence of species on changing landscapes. Cambridge University Press, Cambridge, UK.

Hynes, H. B. N. 1975. The stream and its valley. in Verhandlungen der internationale Vereinigung der theoretische und angewandte Limnologie.

Ilhardt, B. L., E. S. Verry, and B. J. Palik. 2000. Defining riparian areas. Pages 23-42 in E. S. Very, J. W. Hornbeck, and C. A. Doloff, editors. Riparian management in forests of the continental eastern United States, New York.

Ims, R. A., and D. O. Hjermann. 2001. Condition-dependent dispersal. Pages 600-621 in J. Clobert, E. Danchin, A. A. Dhondt, and D. J. Nichols, editors. Dispersal. Oxford University Press, New York.

19 Karr, J. R., and D. R. Dudley. 1981. Ecological perspective on water quality goals. Environmental Management 5:55-68.

Karr, J. R., and E.W. Chu. 1999. Restoring Life in Running Waters - Better Biological Monitoring. Island Press, Covelo, CA.

Kershner, J. L., C. M. Bischoff, and D. L. Horan. 1997. Population, habitat, and genetic characteristics of Colorado River cutthroat trout in wilderness and nonwilderness stream sections in the Uinta Mountains of Utah and Wyoming. North American Journal of Fisheries Management 17:1134–1143.

Lammert, M., and J. D. Allan. 1999. Assessing biotic integrity of streams: effects of scale in measuring the influence of land use/cover and habitat structure on fish and macroinvertebrates. Environmental Management 23:257-270.

Lamouroux, N., S. Doledec, and S. Gayraud. 2004. Biological traits of stream macroinvertebrate communities: effects of microhabitat, reach, and basin filters. Journal of North American Benthological Society 23:449-466.

Matthews, W. J. 1986. Fish faunal structure in an Ozark stream: stability, persistence and a catastrpohic flood. Copeia 1986:388-397.

Matthews, W. J., R. C. Cashner, and F. P. Gelwick. 1988. Stability and persistence of fish faunas and assemblages in three midwestern streams. Copeia 1988:945-955.

McQueen, D. J., M. R. S. Johannes, J. R. Post, T. J. Stewart, and D. R. S. Lean. 1989. Bottom-up and top-down impacts on freshwater pelagic community structure. Ecological Monographs 59:289–309.

Meffe, G. K., and A. L. Sheldon. 1990. Post-defaunation recovery of fish assemblage in southeastern blackwater streams. Ecology 71:657-667.

Moloney, K. A., and S. A. Levin. 1996. The effects of disturbance architecture on landscape level population dynamics. Ecology 72:375-394.

Naiman, R. J. 1998. Biotic Stream Classification. Pages 97-199 in R.J. Naiman and R.E. Bilby, editors. River Ecology and Management. Springer-Verlag, New York.

Nakano, S., H. Miyasaka, and N. Kuhara. 1999. Terrestrial-aquatic linages: riparian inputs alter trophic cascades in a stream food web. Ecology 80:2435- 2441.

Nisbet, R. M., E. McCauley, W. S. C. Gurney, W. W. Murdoch, and A. M. de Roos, editors. 1997. Simple representations of biomass dynamics in structured populations. Case Studies in Mathematical Modeling. Prentice Hall.

Northcote, T. J. 1997. Potamodromy in Salmonidae - living and moving in the fast lande. North American Journal of Fisheries Management 17:1029-1045. 20 Nystrom, P., A. R. McIntosh, and M. J. Winterbourn. 2003 Top-down and bottom-up processes in grassland and forested streams. Oecologia 136 596-608.

Pecarsky, B. L. 1983. Biotic interactions or abiotic limitations? A model of lotic community structure. in T. D. Fontaine III and S. M. Bartell, editors. Dynamics of lotic ecosystems. Ann Arbor Science, Ann Arbor, Michigan.

Peckarsky, B. L. 1984. Predator-prey interactions among aquatic . Pages 196-254 in V. Resh and D. M. Rosenberg, editors. The Ecology of Aquatic Insects. Praeger, New York.

Peterson, J. T., and P. B. Bayley. 1993. Colonization rates of fishes in experimentally defaunated warmwater streams. Transactions of the American Fisheries Society 122:199-207.

Piegay, H., and N. Landon. 1997. Promoting ecological management of riparian forests on the Drome river, France. Aquatic Conservation 7:287-304.

Poff, N. L. 1992. Why disturbances can be predictable: a perspective on the definition of disturbance in streams. Journal of the North American Benthological Society 11:86-92.

Polis, G. A., and D. R. Strong. 1996. Food web complexity and community dynamics. American Naturalist 147:813-846.

Polis, G. A., and K. Winemiller. 1996. Food Webs: Integration of Patterns and Dynamics. Chapman & Hall, New York.

Power, M. E. 1992. Habitat heterogeneity and the functional significance of fish in river food webs. Ecology 73:1675-1688.

Pringle, C. M., M. C. Freeman, and B. J. Freeman. 2000. Regional effects of hydraulic alterations on riverine macrobiota in the new world: tropical-temperate comparisons. Bioscience 50:807-823.

Rader, R. B., D. P. Batzer, and S. A. Wissinger, editors. 2001. Bioassessment and Management of North American Freshwater Wetlands. John Wiley and Sons, New York, NY.

Rahel, F. J. 1990. The hierarchical nature of community persistence: a problem of scale. American Naturalist 136:328-344.

Railsback, S. F., R. H. Lamberson, B. C. Harvey, and W. E. Duffy. 1999. Movement rules for individual-based models of stream fish. Ecological Modelling 123:73-89.

Resh, V. H., A. V. Brown, A. P. Covich, M. E. Gurtz, H. W. Li, G. W. Minshall, S. R. Riece, A. L. Sheldon, J. B. Wallace, and R. C. Wissmar. 1988. The role of 21 disturbance in stream ecology. Journal of North American Benthological Society 7:433-455.

Richardson, E. V., B. Simons, S. Karaki, M. Mahmood, and M. A. Stevens. 1975. Highways in the river environment: hydraulic and environmental design considerations training and design manual. U.S. Department of Transportation, Federal Highway Administration, Washington, D.C.

Richardson, J. S. 1991. Seasonal food limitation of detritivores in a montane stream: an experimental test. Ecology 72:873- 887.

Rieman, B. E., D.C. Lee, and R.F. Thurow. 1997. Distribution, status, and likely future trends of bull trout within the Columbia River and Klamath River Basins. North American Journal of Fisheries Management 17:1111-1125.

Sabo, J. L., and M. E. Power. 2002. River–watershed exchange effects of riverine subsidies on riparian lizards and their terrestrial. Ecology 83:1860-1869.

Schlosser, I. J., and P. L. Angermier. 1995. Spatial variation and demographic processes of lotic fishes: conceptual models, imperical evidence, and implications for conservation. Pages 392-401 in J. L. Nielsen, editor. Evolution in the aquatic ecosystem: defining unique units in population conservation.

Sih, A. 1982. Foraging strategies and the avoidance of predation by an aquatic insect, Notonecta hoffmanni. Ecology 63:786-796.

Silsbee, D. G., and G. L. Larson. 1983. A comparison of streams in logged and unlogged areass of Great Smoky Mountains National Park. Hydrobiologia 102:99-111.

Sousa, W. P. 1985. Disturbance and patch dynamics on rocky intertidal shores. Pages 101-124 in S. T. A. Pickett and P. S. White, editors. The ecology of natural disturbance and patch dynamics. Academic Press, San Diego.

Tebo, L. B. 1955. Effects of siltation resulting from improper logging on the bottom of a small trout stream in the southern Appalachians. Progressive Fish-Culture 17:64– 70.

Townsend, C. R., M. R. Scarsbrook, and S. Doledec. 1997. The intermediate disturbance hypothesis, refugia, and biodiversity in streams. Limnology and Oceanography 42:938-949.

Vannote, R. L., G. W. Minshall, K. W. Cummins, J. R. Sedell, and C. E. Cushing. 1980. The river continuum concept. Canadian Journal of Fisheries and Aquatic Sciences 37:130-137.

Wallace, J. B., S. L. Eggert, J. L. Meyer, and J. R. Webster. 1997. Multiple trophic levels of a forest stream linked to terrestrial litter inputs. Science 277:102–104. 22 Ward, J. V. 1989. The four-dimentional nature of lotic ecosystems. Journal of North American Benthological Society 8:2-8.

Warren Jr., M. L., P. L. Angermier, B. M. Burr, and W. R. Haag. 1997. Declines of a diverse fish fauna: patterns of imperilment and protection of the southeastern United States. Pages 105-164 in G. W. Benz and D. E. Collins, editors. Aquatic fauna in peril: the southeastern perspective. Lenz Design in Communications, Decatur, GA.

Warren Jr., M. L., and M. G. Pardew. 1998. Road crossings as barriers to small-stream fish movement. Transactions Of The American Fisheries Society 127:637-644.

Wiens, J. A. 2001. The landscape contents of dispersal. Pages 97-109 in J. Clobert, E. Danchin, A. A. Dhondt\, and J. D. Nichols, editors. Dispersal. Oxford University Press, New York.

Wiens, J. A., J. T. Rotenberry, and B. van Horne. 1986. A lesson in the limitation of field experiments: shrubsteppe birds and habitat alteration. Ecology 67:365-376.

Williams, L. R., T. H. Bonner, J. D. Hudson III, M. G. Williams, T. R. Leavy, and C. S. Williams. 2005. Interactive efects of environmental variability and military training on stream biota of three headwater drainages in Western Louisiana. Transactions of the American Fisheries Society 134:192-206.

Williams, L. R., C. M. Taylor, M. L. Warren Jr, and J. A. Clingenpeel. 2002. Large-scale effects of timber harvesting on stream systems in the Ouachita Mountains, Arkansas, USA. Environmental Management 29:76-87.

Williams, L. R., C. M. Taylor, M. L. Warren Jr., and J. A. Clingenpeel. 2003. Environmental variability, historical contingency, and the structure of regional fish and macroinvertebrate faunas in Ouachita Mountain stream systems. Environmental Biology of Fishes 67:203-216.

.

23 FIGURES

Figure 1.1 Early bridge designed used to convey light weight vehicles.

Figure 1.2 Simple concrete topped bridge with support beams. 24

Figure 1.3 Single pipe culvert topped with metal road.

Figure 1.4 Single pipe culvert armored with interlocking blocks.

25

Figure 1.5 Ovoid-top culvert used for bridge construction.

Figure 1.6 Three 10” box culvert design blocked by plant and woody debris.

26

Figure 1.7 Pipe culvert crossing, also serving as wet fords during high rainfall events.

Figure 1.8 PVC pipes installed below the road to allow cross flow during flood events.

27

Figure 1.9 Ford crossing design with interlocking concrete blocks buried in the stream bed substrate.

Figure 1.10 Ford crossing showing erosion because of unpaved stream bed.

28

Figure 1.11 New crossings topped with poured concrete, withstanding heavy mechanized vehicles.

Figure 1.12 New Arch culvert design to reduce impacts by high variation in discharge.

29

CHAPTER 2

Impacts of timber harvesting on macroinvertebrate assemblage dynamics in

western Louisiana headwater streams

INTRODUCTION

Stream organisms have a complex relationship with the habitats they occupy, and this relationship can vary greatly across spatial, temporal, or taxonomic scales (Schlosser

1987; Angermeir 1987; Lammert and Allan 1999). From studies conducted on a variety of systems, many authors have argued strongly that aquatic communities are structured by habitat variability (Gorman and Karr 1978; Gorman 1986; Capone and Kushlan 1991;

Jackson and Harvey 1993; Williams et al. 1996). While many past studies have focused on one group of organisms, more recent studies are trying to understand the role of habitat in structuring entire communities at broader spatial and temporal scales (Jackson and Harvey 1993; Williams et al. 2002).

Aquatic communities in headwater streams rely on allochthonous nutrient input from the riparian corridor as their primary source of energy (Vannote et al. 1980).

According to the River Continuum Concept (Vannote et al. 1980), trophic guilds of aquatic macroinvertebrates include predators, grazers, shredders, and collectors, and span habitats from headwater streams to lower reaches of rivers. The riparian corridor forms an ecotone, or transitional zone, between the stream and uplands, and supports 30 productivity and biotic diversity far greater than its relatively small area would indicate

(Erman et al. 1977, Frissell et al. 1986, Meyer et al. 2003). In low gradient streams, a range of taxa including plants, invertebrates, and vertebrates have evolved life-history

strategies that depend on floodplain wetlands. Some macroinvertebrates complete their

entire lifecycle in these habitats, persisting in seasonal wetlands in drought resistant

forms such as eggs (Wallace 1990; Poff and Ward 1990). Vertebrates (fishes,

amphibians, mammals, and birds) frequently make seasonal movements into the

floodplain, and key periods of their lifecycle (e.g., breeding, rear young or migration) are

timed to riparian zone flooding (Naiman and Decamps 1997). Especially in low-gradient

systems, disturbance to the riparian corridor could potentially affect structure and function of stream communities (Resh et al. 1988).

Anthropogenic modification of ecosystems generally causes alteration to the temporal regime of natural variation and disturbance, by rendering adaptive mechanisms less effective, redefining ecosystems from equilibrium to nonequilibrium, and reducing potential for species to evolve adaptive mechanisms (Urban et al. 1987). Cairns et al.

(1971) discussed the extent of anthropogenic disturbance on ecological processes and patterns in lotic systems. The extent of disturbance on ecological processes and patterns in lotic systems can be large and persistent, often having a lasting impact on terrestrial

(Pearson et al. 1998; Dupouey et al. 2002) and aquatic communities (Harding et al.

1998), even when the riparian habitat appears to have recovered (Carnis et al. 1971,

Turner et al. 2003). Changes in landuse practices result in various types of disturbances in the form of habitat fragmentation and deterioration of water quality leading to

increasing threats to biodiversity in headwater streams (Vuori and Joensuu 1996). In

31 lotic systems, anthropogenic disturbances (e.g., timber harvesting, road crossings,

hydroelectric dams, thermal constancy below hypolimnetic release dams, agricultural

effluent) alter temporal patterns of environmental variability, and may exceed tolerance

thresholds for certain biota. These disturbances result in high nutrient, and sediment input from timber harvesting causing exclusion of filter feeding organisms (Wallace and

Merritt 1980). Temperature regimes in streams change after timber harvesting that includes removal of trees and shrubs, up to and including the stream bank (Webster and

Waide 1982, Culp and Davies 1983). The destruction of protective vegetation and compaction of the soil surface reduces soil permeability to water, increasing erosive surface runoff.

Although stream ecosystems are a product of historic geologic and climatologic attributes, large-scale human disturbances can change the landscape (Hooke 1999). Over the past 200 years the forestlands of the southeastern Gulf Coastal Plain ecoregion of the

United States has experienced extensive pressures from agriculture, timber harvesting, and population growth (Omernik 1987, Frost 1993). Several studies indicate that the changes in stream macroinvertebrate assemblages caused by timber harvesting may be long-term (Erman et al. 1977, Haefner and Wallace 1981, Silsbee and Larson 1983,

Swank et al. 2001), but these studies have been conducted in the Pacific Northwest or in upland eastern streams. Extrapolation of the effects on communities from these systems with stony, high-gradient streams, to sandy, low-gradient streams might not be applicable

(Lenat and Crawford 1994, Feminella 2000, Maloney et al. 2005). Recent study by

Williams et al. (2002, 2005) on southeastern streams failed to detect effects of timber harvesting on biota as they were studied some years after the disturbance event (and also

32 probably because of the high resilience of biotic communities inhabiting these regions).

However, stream geomorphology and hydrology of streams in Georgia were affected by

military activities, with increased sedimentation, and unstable stream banks (Maloney et

al. 2005; also see Williams et al. 2005).

In this paper, I assess habitat associations and assemblage structure of stream

macroinvertebrates and monitor the change assemblage structure before, during, and after

a large-scale logging project in the Red River drainage Little Sandy, Tiger, and Odom

streams in Ft. Polk, Louisiana. This study builds on pre-disturbance patterns presented

by Williams et al. (2005) in the same systems. The main objective of this study was to do

a comparative study on the macroinvertebrate assemblage structure pre-, during-, and post- timber harvesting period, and the effects of timber harvesting activities on the physical and chemical environmental factors affecting the macroinvertebrate assemblages.

METHODS

Study Area

The headwater streams of the Red River drainage originate in the Peason Ridge

Training Area (PRTA), located on Fort Polk Army base, west-central Louisiana. The first-order headwater streams in the study area belong to the Red River drainage (Little

Sandy, Odom, and Tiger streams; Figure 2.1), and are fed by seepage from the northern

scarp of Peason Ridge, and drain to the Kisatchie Bayou, and eventually to the Red and

Mississippi Rivers (Williams et al. 2005). These streams are low gradient, with substrate

consisting of sandy-loam soils (Martin et al. 1990). No timber harvesting was conducted 33 around Little Sandy stream, but moderate to extensive timber harvest was conducted

along Odom and Tiger streams in the year 2004 (Table 2.1). The forests are dominated

by longleaf pine (Pinus palustris), interspersed with hardwoods (e.g., bald cypress

Taxodium distichum, white oak Quercus alba, water oak Quercus nigra, and magnolia

Magnolia randiflora) which are more common along the lowland reaches (Williams et al.

2005). Because longleaf pine is part of a ‘fire climax’ community, the army conducts a

regulatory prescribed burn cycle, on a 2-3 year rotation. This ecosystem supports a

population of federally endangered Red Cockaded Woodpecker (Picoides borealis).

Timber harvesting was started in 2003 in the upper reaches of the watershed for

constructing a range complex to facilitate combined arms training that is a part of Digital

Multipurpose Battle Area Course (DMPBAC). This would require creating new firing

lanes, and conveyance road systems and numerous stream crossings that are constructed

to allow transport of large equipment. Vegetation along the riparizone was completely removed without leaving any buffer along the firing lanes, and new roads constructed.

The zoogeography and history of army land use in PRTA is summarized by Williams et al. (2005).

Sampling Methods

Field Techniques

Macroinvertebrates were collected for four seasons from 2003 to 2006 in three streams in the northern region of Fort Polk, Louisiana (Little Sandy, Odom, and Tiger streams; only through Summer in 2006). Sampling was not conducted on Little Sandy

stream during Fall 2004 because of access restriction by military training schedule.

34 Sampling was defined at mesohabitat scale to obtain fair representation of habitat

diversity in the system (Williams et al. 2004), where three replicates of each mesohabitat

(runs, riffles and pools) were sampled when available. Macroinvertebrates were sampled at each mesohabitat using a Surber sampler and D-net. A five-minute Surber sample was taken at each locality in runs and riffles, and D-net samples along pools, runs, and riffles

from sampling sites selected as stated previously. Macroinvertebrates were preserved in

field with 95% ethanol and Rose-Bengal stain to facilitate picking from leaf-detritus.

Chemical environmental variables, including temperature (ºC), conductivity

(μS/cm), pH, dissolved oxygen (mg/l), and turbidity (NTU), were recorded using a YSI-

Model 600 multi-probe meter at each site. Physical environmental characters including

percent substrate composition (i.e., clay, silt, sand, gravel, cobble, and bedrock), water

depth, woody debris, and undercut bank, were recorded along the habitats sampled.

Mean water depth (m), and mean current velocity (m/s) were obtained using Marsh-

McBirney Flowmate from 3-4 cross-sectional profiles per mesohabitats unit.

Geomorphological measurements of the stream were obtained to determine the effects of land use on the streams studied (e.g., average width, bankful width, floodplain width; Table 2.1). At each site, measurements on the channel materials, dimension, pattern, and profile were recorded. Procedures used were consistent with the guidelines presented by Harrelson et al. (1994). Channel profile was surveyed on 100-300 m long reaches and included bed, water surface, top of bank and indicators of channel formation

(i.e., depositional surfaces such as benches, bars, and breaks in vertical grade of the bank). Azimuths of each profile distance defined planar form. Wolman Pebble Counts

(Ward et al., 2004) were used to measure bed material type.

35 Statistical Analyses

Principal components analysis is an unconstrained ordination technique (Pielou

1984) which partitions a resemblance matrix (variance-covariance or correlation) into a

set of orthogonal (perpendicular) axes or PCA components (Ludwig and Reynolds 1988).

Abiotic environmental variables (viz., bed substrate composition, water quality

parameters, physical characteristics of the stream bank, and channel dimensions) were

used in the PCA to determine their relationships. The first two PC axes explaining the largest variation in the data set were interpreted as gradients of overall habitat complexity

(Meffe and Sheldon 1988). % EPT and % Diptera from the Red River drainage (Little

Sandy, Odom, and Tiger creeks combined) were related to log transformed stream

physical and chemical factors using stepwise multiple linear regressions.

Biotic diversity was determined using Shannon-Weiner index (loge base + 1), and

evenness using Buzas and Gibon’s E (E = eH/S) were determined for each site across

seasons (Hayek and Buzas 1997). Similarities of macroinvertebrate families by stream

and year were determined by Renkonen Similarity Index (RSI; Krebs 1999). Similarity

trends among sites were further explored using detrended correspondence analysis (DCA;

Gauch 1982). DCA is a unimodal ordination technique that can be used to derive

important environmental gradients from species composition data. DCA arranges species

and sites across a set of orthogonal axes so that sites that are similar in terms of their species compositions are closer to one another in ordination space, and species with similar distributions across sites tend to group together in ordination space; it also

removes the ‘arch effect’, and rescales the axis, from correspondence analysis (ter Braak

and Smilauer 2002). Species are located at the mode of the unimodal distribution of the 36 individual species’ relative abundance along that axis. Sites are located along the DCA axes based on a weighted average of species scores. Analyses were conducted using

CANOCO software (ter Braak and Smilauer 2002).

I described the relationships between macroinvertebrate taxa and the physical, and chemical environmental variables, along spatial and temporal gradients using canonical correspondence analysis (CCA; ter Braak 1986). The spatial component of the stream was described by the three different mesohabitats, viz., riffle, run, and pool; whereas, the temporal component was at three levels represented by the pre-, during-, and post- timber harvest years. This ordination approach is a widely used direct ordination technique (ter

Braak and Smilauer 2002). Data on species distributions in sampling units is used to define the multivariate axes, and the multiple regression analyses is used to correlate factor loading scores for samples in species space with environmental predictor variables.

Abundances were square-root transformed, to dampen the effects of predominant taxa, and rare taxa were down-weighted (McCune and Mefford 1999).

Variance partitioning was used with a series of partial CCAs, similar to partial regression techniques (Quinn and Keough 2002), to examine the pure effect of anthropogenic disturbance in Ft. Polk, Louisiana, from the effects of stream, season, and mesohabitat (Williams et al. 2002). Because the response variable, macroinvertebrate assemblage, was studied in three different streams, with varying levels of timber harvest conducted on each (Table 2.2), I used stream as one of the factors contributing to the observed variation. This ecoregion experiences four distinct seasons (viz., spring, summer, fall, and winter), which could be an important factor explaining the variation in the macroinvertebrate assemblage. Within the streams, the habitats were categorized into

37 three mesohabitats (viz., riffle, run, and pool) which could contribute to the spatial

component of variation. Anthropogenic disturbance was represented by pre-, during-,

and post- timber harvest in the years 2003, 2004, and 2005, respectively. To assess the

pure effect of each of the four main variables, I used the other three factors as covariates

in the partial CCAs. For each partial CCA, a Monte Carlo test (1000 permutations) was

used to estimate the significance of each variable (ter Braak and Smilauer 2002).

Multivariate analyses based on randomization procedures (e.g., Monte Carlo tests)

represent patterns in community structure better than a series of univariate tests, because

they are not restricted in meeting several univariate assumptions (McCune and Grace

2002, Williams et al. 2005).

RESULTS

The undisturbed reference stream from the Ft. Polk area, Little Sandy was selected as there was no timber harvest activity in its watershed (Table 2.1). The land-

use characteristics of Odom, and Tiger streams were described by disturbance levels of

58% and 62% of their watershed area, respectively. All the streams had similar stream

characteristics, with floodplain widths ranging from 9-16 meters, and bankful widths

ranging from 5-10 meters. From the geomorphological features, all three streams were

classified as F type (Rosgen 1996). The lower bed slope (< 2%), deeply incised stream

bed, and the failing condition of the banks suggests a Type F stream channel system

(Ward 2004). The tractive force calculated (T=1000*mean water depth*bed slope)

suggests that the existing current velocity can move substrate size (48mm in Little Sandy,

38 66mm in Odom creek, and 117mm in Tiger creek) greater than that found in the stream bed.

Mean values of conductivity, temperature, velocity, percentage of sand, and percentage of woody debris showed a change during 2004, the year of timber harvest

(Table 2.2, Appendix A). Percentage of gravel showed a marked increase; whereas, percentage of cobble dropped in the year following timber harvest. The percentage of undercut bank increased in the year following timber harvest, which could be attributed to incision in the stream channel.

Using indirect ordination technique, the environmental variables from Table 2.2 explained 43.4 % of the variation among sites across years in the PCA analysis (Figure

2.2). The first PCA axis explained 28.9% overall variation and was representative of

substrate gradient (sandy to cobble), water quality, and physical characters. The second

PCA axis explained 14.5 % of the variation and represented the stream dimensions (from greater depth of the site to greater width and length of mesohabitats). Eigenvalues of the first two axes (axis 1 = 4.13, axis 2 = 3.6) were higher than the broken-stick eigenvalues

(axis 1 = 3.49, axis 2 = 2.49), and hence the total variation explained by the data were interpretable (Jackson 1993). There was not a distinct separation between the pre-,

during-, and post- timber harvest years. A weak trend showed a clumping of the year

2004 scores toward the left lower quadrant, which described the timber harvest year with greater sand composition and undercut bank percentage.

A total of 23,010 macroinvertebrates were represented by 94 genera (88 families, and 23 orders) from samples collected during the three year study period (Appendix B).

For statistical purposes data were examined at family level to minimize error because of

39 some genera could not be identified confidently. Also, this scale has been shown to be

appropriate for many other stream studies on similar topics. The most abundant families

were Chironomidae (Diptera), Baetidae, Caeniidae, Heptageniidae (Ephemeroptera), and

Perlidae (Plecoptera) with relative abundances greater than 5% each; and the first two

families greater than 10% each. The relative abundances of Ephemeroptera, Plecoptera,

and Trichoptera were lower in years 2003 and 2004, and increased in the post- timber harvest year 2005, but none showed significant trends because of higher variance (p- values ranged from 0.11 to 0.71; Figure 2.3). As EPT apparently increased, the relative abundance of larval Dipterans showed a decreasing trend with highest values post-harvest for all three streams. A separate regression analyses on the % EPT and % Diptera in the

Red River drainage (all three streams combined) shows the varying environmental effect on them (Table 2.3). Current velocity had a negative effect on % EPT, whereas pH, dissolved oxygen, percent undercut bank and percent woody debris had a positive relationship. % Diptera showed a different relationship with the environmental paramters, with conductivity, percent undercut bank, and percent woody debris, having a negative effect, and stream length, and water depth, having a positive effect.

Species richness, abundance, diversity, and evenness of macroinvertebrates in mesohabitats by year were highest, with low variance in Tiger Creek, and high variance in Odom Stream (Figure 2.4). Macroinvertebrates grouped by feeding guild based on

Merritt and Cummins (1996) showed a drop in the scraper guild, and an increase in the shredder guild during the year 2004, compared to the pre- and post- harvest years (p- values not significant) in Odom and Tiger streams (Figure 2.5). This trend was not observed in Little Sandy stream. Scraper abundance, mainly Heptageniidae (Stenonema

40 sp.) and Baetidae, decreased during the disturbance year, but increased in the following

year. Shredder abundance, Capniidae (Allocapnia sp.) increased slightly during the

disturbance year, but dropped back the following year to preexisting abundance.

Functional group distribution in the reference stream (Little Sandy) during the three years

was fairly constant. The functional guilds based on habit (position occupied on the

microhabitats in stream; Merritt and Cummins 1996) showed similar trends across

streams with a drop in numbers in each guild during the timber harvest year 2004 (Figure

2.5). Each of the guilds showed a recovery in their numbers in the following year.

Guilds in Tiger Stream however, did not show a marked change in numbers over the years.

The macroinvertebrate assemblages in Odom and Tiger streams showed greater similarities (RSI; Table 2.4) in the pre- and during- timber harvest years compared to the

post harvest period. This greater similarity between the pre- and during- years could possibly be because of a lag in effects of timber harvest on macroinvertebrate assemblages. The streams were fairly similar, with mean RSI of 0.6 (SE = 0.025) between Little Sandy and Odom streams, mean RSI = 0.56 (SE = 0.04) between Little

Sandy and Tiger streams, and mean RSI = 0.59 (SE = 0.015) between Odom and Tiger streams. The detrended correspondence analysis of species and sites did not reflect any

distinctiveness across the years (treatment effect of pre-, during-, and post- timber

harvest years; Figure 2.6), but the DCA with sites labeled by habitat and season showed a

pattern. The site scores of separate DCA by mesohabitats (riffle, run, and pool), and by

season (spring, summer, fall, and winter), were plotted and coded to illustrate assemblage

trends. The mesohabitat component of the stream system was described by the pool sites

41 occupying the left corner of axis 1, run sites in the center, and the riffle sites on the far

right. The seasonal component of the stream system was described by the DCA site

scores grouping separately based on the four different seasons during which sampling

was conducted.

The full CCA model of habitat variables, season, stream, and year, explained 43.8

% of the total variance in the macroinvertebrate assemblage (Table 2.5). Pure effects of

timber harvest disturbance (year 2004) explained very low amount of assemblage

variability (2.98 %; P = 0.001), which was very close to the variability explained by

differences in stream (3.05 %; P = 0.001). Pure effect of season explained the highest variability (13.97 %; P = 0.001), followed by the mesohabitat variable (9.16; P = 0.001).

Lack of greater variance explained by year can be attributed to harsh conditions experienced historically in this region (Williams et al. 2005). Variance that could not be attributed to measured environmental variables totaled 56.2% Table 2.5).

Environmental parameters on the first CCA axis explained 30.7 % of the variation with high positive loadings of turbidity, and velocity, and greater negative loadings represented by temperature, conductivity, pH, and dissolved oxygen. Habitat features

(i.e., % woody debris, and % undercut bank) had high positive loadings on axis 1. The second CCA axis explained 15.3 % of the variation showing a substrate gradient of gravel on the positive quadrant, and sand on the negative quadrant. A stream dimension gradient was observed with depth, and width having greater negative loadings, and length on the positive loading. The site scores did not show any strong trends based on year, stream, or habitat, but showed distinct clustering based on season (Figure 2.8). Thus, the most interesting pattern from these analyses was that macroinvertebrate taxa were not as

42 strongly affected by timber harvest as compared to shifts related to changing seasons.

This pattern contradicted our original prediction that macroinvertebrate assemblage

would be strongly affected by timber harvest activities.

DISCUSSION

Ecological communities are maintained and regulated by the biogeographic

history, spatial and temporal aspects of the environment, and species interactions (Pianka

1994). The effects of environmental complexity and its dynamics on streams is scale dependent (Lancaster et al. 1990), and is responsible for structuring macroinvertebrate

assemblages at stream reach scale (Richards & Host 1994, Clenaghan et al. 1998). The

regional and historical factors influencing assemblages at larger scales also act as filters

to the species assemblages at smaller scales (Hugueny et al. 1997, Li et al. 2001). These

physical and biological variables on a small spatial scale are, in turn, influenced by

variables on larger spatial scales according to the hierarchy theory (O’Neill et al. 1986).

Based on this hierarchy theory, a biogeographic framework is developed where stream

communities at a site can be seen as a product of a series of filters (e.g. historical,

regional, habitat) through which species occurring at a site have had to pass (Caley and

Schluter 1997, Poff 1997, Tonn 1990). Knowledge of history and larger scale filters is

important in understanding the abundance and distribution of regional communities.

Various types of land-use practices, including urbanization, agriculture, and

timber harvest, can cause instability of the streambed and streambank by increasing

discharge, and thereby increasing sediment release (Wates 1995, Sutherland et al. 2002,

Lorang and Hauer 2003). The effects of timber harvesting and military activities on 43 southeast Gulf Coastal streams have been studied previously (Williams et al 2002, 2005), but failed to detect a deleterious effect on stream biota. According to the Rosgen’s level

II stream classification (Rosgen 1996), the three streams studied are Type F streams. The streams were entrenched, with slopes less than 2 %. The stream banks were unstable, with higher percentage of exposed undercut bank, and trees fallen within the stream channel as a result of bank failure. The greater tractive force calculated from the observed stream parameters suggests its unstable stage. With the observed d50 of particle size much smaller than the calculated value, I conclude that the streambed is in degradation processes. Although Little Sandy was designated as the reference reach, it was a bit different from the other streams, in that it had a certain percentage of bedrock as bed material (Table 2.2). This characteristic could cause the widening of the stream (as a result of disturbance), rather than bed incision.

Recognizing the interaction between vegetation, sediment and geomorphology is important in understanding stream dynamics (Roberts 2000, Tabacchi et al. 2000).

Riparian vegetation plays a major role in modifying bank erosion, and deposition processes by altering bank hydrology, and bank hydraulic characteristics such as boundary shear stress (Abernethy and Rutherfurd 2001, Wood et al. 2001). Although large woody debris in streams is a result of bank failure (i.e., fallen trees), they are known to increase stream stability by reducing shear stress on the stream bed (Downs and Simon

2001), and also increasing sediment residence time by trapping sediment (Florsheim et al.

2001). The current condition of the Gulf Coastal Plain streams is largely a result of fluvial processes during the last drop in sea level about 14,000 to 18,000 years ago, resulting in high sediment deposition and relatively low discharges (Saucier 1994, Hupp

44 2000). The species composition and distribution of plants, fish, and terrestrial insects is a product of refuge from past glaciation, frequent fire disturbances, and colonization patterns of these highly variable stream habitats (Gilzenstein et al. 1995, Connor and

Suttkus 1986, Noonan 1988). Historical events like glaciation and more recent fire disturbances act as filters constraining species composition to the ones tolerant to small scale disturbances (Stewart et al. 1976, Surdick 1985).

A typical impact of timber harvesting activity on streams, related to loss of riparian shading, is an increase in water temperature (Brosofske et al. 1997). Removal of canopy cover from adjacent forest causes loss of windscreen, and thus leads to increase in temperature. Odom stream showed a marked increase in temperature because of the severe disturbance by timber harvest, and several new roads constructed along its length

(Table 2.1). Although the riparian vegetation immediately surrounding Little Sandy stream was not disturbed, it did show an increase in temperature during the year of timber harvest. However, a marked rise in water temperature in Tiger Stream was not observed, presumably because of ground-water discharge being the main source of flow, or because the slash insulating the stream channel from solar radiation. This could also be as a result of fewer roads constructed along Tiger Creek (Table 2.1).

Intensive forest management practices and their by-product, road constructions and road crossings, are linked to increased stream sedimentation (Burns 1972, Waters

1995). Sediment release in streams typically leads to increased substrate embeddedness, and eventually reducing habitat for macroinvertebrates (Rinne 1990). Inverse relationships have been observed between increased turbidity in streams and EPT taxa

(McClelland and Brusven 1980, Mebane 2001). The streams I studied show increased

45 turbidity during and following the timber harvest years, suggesting increased embeddedness of the substrate. The relative abundance of EPT taxa showed a drop in numbers during and after timber harvesting in Little Sandy and Odom streams, corresponding to the increase in turbidity in the stream channel. However, relative abundance of EPT taxa in Tiger stream did not show a significant shift, perhaps resulting from ground water discharge in the stream channel. In the current study, although turbidity was correlated to relative abundance of EPT taxa (r2 = 0.14 – 0.75) and

Dipteran larvae (r2 = 0.01 – 0.94), I did not detect statistically significant relationship.

A study conducted ten years post disturbance (timber harvesting) on streams of southwestern Arkansas showed that fish and macroinvertebrate assemblages were not affected by disturbance (Williams et al. 2002). On a similar study conducted in Ft. Polk,

Louisiana, Williams et al. (2005) found no significant impact of military training on fish and macroinvertebrate assemblages. In contrast to these studies, Maloney et al. (2005) found higher sedimentation resulting from military activities. The current study captured the pre-, during-, and post- timber harvest relationship between environmental variables, and macroinvertebrate assemblage. Despite failing to detect any significant effects of timber harvest activities on macroinvertebrates, these activities are known to influence biota at smaller spatial scales (Campbell and Deog 1989). Although this study did not detect effects of timber harvesting on the Gulf Coastal Plain streams, continuous impact over greater spatial and temporal scales can effectively alter habitat and regional distribution and abundance patterns of stream biota.

Stream macroinvertebrates have been found to vary with seasons of the year, and with mesohabitats across streams (Matthews et al. 1991). Thus, understanding the scale

46 of spatial and temporal disturbance in stream ecology is important (Minshall 1988), and

has direct consequences on the ability of biotic recovery in the system (Kelly and Harwell

1990; Figure 2.9). Although the macroinvertebrate assemblage did not show any

significant variation over three years (treatment, pre-, during-, post- timber harvest years), a seasonal shift had a higher influence in structuring macroinvertebrate

assemblage. This strong seasonal component driving the macroinvertebrate assemblage

is a combination result of degree of habitat specificity, relatively short life cycles, and

seasonal changes in environmental conditions (e.g., discharge, temperature; Linke et al.

1999; Williams et al. 2002). Similarly, anthropogenic effect (treatment, reference stream

compared to two disturbed streams) did not seem to have an effect on macroinvertebrate

assemblage structure, but at smaller scale of mesohabitat, there was a distinct pattern (Fig

2.4 a, b).

Comparison of functional guilds of macroinvertebrates among pre-, during-, and post- timber harvest years did not reveal significant shifts. Forest management activities typically shift the macroinvertebrate assemblage toward scraper guild (Stone and Wallace

1998). Because riparian vegetation limits algal production in headwater streams through light attenuation, clear-cutting activities would release this effect, causing the streams to become autochthonous by allowing algal periphyton growth (Gregory et al. 1991;

Murphy and Hall 1981). This increase in periphyton growth would increase the density

of scrapers (Silsbee and Larson 1983). One of the two distinct trends observed in feeding

guilds during the disturbance year was a decrease in scraper abundance, mainly

Heptageniidae (Stenonema sp.) and Baetidae, but it increased the following year. The

second was an increase in shredder abundance, Capniidae (Allocapnia sp.) during the

47 disturbance year (although not significant, because of high variance), but dropped back

the following year to preexisting abundance. Several studies have observed an increase

in mayfly taxa following forest management practice in response to increase in primary

production (Gurtz and Wallace, 1984, Wallace and Gurtz, 1986, Stone and Wallace

1998). The functional guild based on habit responded to timber harvest effect by a drop in their numbers (again, not significant). However, most of the guilds showed a recovery in numbers, if not greater than before, from the timber harvesting impact, which could be attributed to greater resilience of communities occupying historically harsh environments

(Peckarsky 1983).

Knowledge of spatial and temporal patterns in community assemblages is essential in predicting the effects of timber harvesting activities. The biotic assemblage structure in streams largely depends on the history, climate, geomorphology, and hydrology. The results of this study can be used in predicting timber harvesting effects on stream macroinvertebrates in similar ecoregions (west central Louisiana). At smaller spatial scales, local habitat variability becomes more important (Poff and Ward 1989;

Cooper et al. 1998). The macroinvertebrate assemblages showed stronger associations based on mesohabitats, as opposed to no significant distinction based on streams. The macroinvertebrate assemblage results at the time scale explained greater amount of variance based on seasonal level, compared to very little variance detected at the year scale. Thus macroinvertebrates in streams of the Red River drainage, in general, were primarily influenced by seasons, followed by microhabitat types, and were least influenced by timber harvesting activities. Because of their intrinsic differences, small streams are difficult to manage, especially for making long-term predictions; hence

48 detailed studies on larger spatial and temporal scales are necessary for proper comprehension (Poff and Ward 1989).

REFERENCES

Abernethy, B., and I. D. Rutherfurd. 2001. The distribution and strength of riparian tree roots in relation to riverbank reinforcement. Hydrological Processes 15:39-61.

Angermeier, P. L., A. P. Wheeler, and A. E. Rosenberger. 2004. A conceptual framework for assessing impacts of roads on aquatic biota. Fisheries 29:19-29.

Angermeir, P. L. 1987. Spatiotemporal variation in habitat selection by fishes in small Illinois streams. Pages 38-58 in W. J. Matthews and D. C. Heins, editors. Community and evolutionary ecology of North American stream fishes. University of Oklahoma Press, Norman, Oklahoma.

Aparicio, E., and A. de Sostoa. 1999. Pattern of movements of adult Barbus haasi in a small Mediterranean stream. Journal of Fish Biology 55:1086-1095.

Baker, C. O., and F. E. Votapka. 1990. Fish Passage Through Culverts, San Dimas, CA:USDA. FHWA-FL-90-006:1-67, Forest Service Technology and Development Center.

Bowman, D., and G. S. Boggs. 2006. Fire ecology. Progress in Physical Geography 30:245-257.

Bowman, M. F., and R. C. Bailey. 1997. Does taxonomic resolution affect the multivariate description of the structure of freshwater benthic macroinvertebrate communities? Canadian Journal of Fisherriees and Aquatic Sciences 54:1802- 1807.

Bridges, E. L., and S. L. Orzell. 1989. Longleaf pine communities of the Western Gulf Coastal Plain. Natural Areas Journal 9:246-262.

Brosofske, K. D., J. Chen, R. J. Naiman, and J. F. Franklin. 1997. Harvesting effects on microclimate gradients from small streams to uplands in western Washington. Ecological Applications 7:1188-1200.

Brown, J. H., and A. Kodric-Brown. 1977. Turnover rates in insular biogeography: effect of immigration on extinction. Ecology 58:445-449.

Burns, J. W. 1972. Some effects of logging and associated road construction on northern California streams. Transactions of the American Fisheries Society 101:1-17. 49 Cairns Jr., J., J. S. Crossman, K. L. Dickson, and E. E. Herricks. 1971. The recovery of damaged streams. Bulletin of the Association of Southeastern Biologists 18:79- 106.

Caley, M. J., and D. Schluter. 1997. The relationship between local and regional diversity. Ecology 78:70-80.

Campbell, I. C., and T. J. Deoeg. 1989. Impact of timber harvesting and production on streams: a review. Australian Journal of Marine and Freshwater Resources 40.

Capone, T. A., and J. A. Kushlan. 1991. Fish community structure in dry-season stream pools. Ecology 72:983-992.

Clark, J. 1991. Disturbance and population structure on the shifting mosaic landscape. Ecology 72:1119-1137.

Clenaghan, C., P. S. Giller, J. O'Halloran, and R. Hernan. 1998. Stream macroinvertebrate communities in a coniferr-afforested catchment in Ireland: Relationships to physico-chemical and biotic factors. Freshwater Biology 40:175- 193.

Conner, J. V., and R. D. Suttkus. 1986a. Zoogeography of freshwater fishes of the western Gulf slope. Pages 413-456 in C. H. Hocutt and E. O. Wiley, editors. The zoogeography of North American freshwater fishes. Wiley, New York.

Conner, J. V., and R. D. Suttkus. 1986b. Zoogeography of freshwater fishes of the western Gulf Slope of North America. Pages 413-456 in C. H. Hocutt and E. O. Wiley, editors. The Zoogeography of North American Freshwater Fishes. John Wiley and Sons, New York, NY.

Cooper, S. D., S. Diehl, and O. Sarnelle. 1998. Implications of scale for patterns and processes in stream ecology. Australian Journal of Ecology 23:27-40.

Culp, J. M., and R. W. Davies. 1983. An assessment of the effects of streambank clear- cutting on macroinvertebrate communities in a managed watershed. Canadian Technical Report of Fisheries and Aquatic Sciences No. 1208.

Downs, P. W., and A. Simon. 2001. Fluvial geomorphological analysis of the recruitment of large woody debris in the Yalobusha River network, Ventral Mississippi, U.S.A. Geomorphology 37:65-91.

Dupouey, J. L., E. Dambrine, J. D. Laffite, and C. Moares. 2002. Irreversible impact of past land use on forest soils and biodiversity. Ecology 83:2978–2984.

Erman, D. C., J. D. Newbold, and K. B. Reby. 1977. Evaluation of streamside bufferstrips for protecting aquatic organisms. Technical Completion Report No. 165, California Water Resources Center, University of California, Berkeley. 50 Fausch, K. D., C. E. Torgersen, C. V. Baxter, and H. W. Li. 2002. Landscapes to riverscapes: bridging the gap between research and conservation of stream fishes. Bioscience 52:483-498.

Fausch, K. D., and M. K. Young. 1995. Evolutionary significant units and movement of resident stream fishes: a cautionary tale. Pages 360-370 in J. L. Nielsen, editor. American Fisheries Society, Symposium 17, Bethesda, Maryland.

Feminella, J. W. 2000. Correspondence between stream macroinvertebrate assemblages and 4 ecoregions of the southeastern USA. Journal of the North American Benthological Society 19:442-461.

Florsheim, J. L., J. F. Mount, and L. T. Rutten. 2001. Effect of baselevel change on floodplain and fan sediment storage and ephemeral tributary channel morphology, Navarro River, California. Earth Surface Processes and Landforms 26:219-232.

Frissell, C. A., W. J. Liss, C. E. Warren, and M. D. Hurley. 1986. A hierarchical framework for stream habitat classification: viewing streams in a watershed context. Environmental Management 10:199-214.

Frost, C. C. 1993. Four centuries of changing landscape patterns in the longleaf pine ecosystem. in S. M. Hermann, editor. 18th Tall Timbers Fire Ecology Conference, Tallahassee, Florida.

Gauch, H. G. 1982. Multivariate analysis in community ecology. Cambridge University Press, Cambridge.

Gerking, S. D. 1959. The restricted movement of fish populations. Biological Review 34:221-242.

Gilliam, J. F., and D. F. Fraser. 2001. Movement in corridors enhancement by predation threat, disturbance, and habitat structures. Ecology 82:258-273.

Glasser, J. W. 1979. The role of predation in shaping and maintaining the structure of communities. American Naturalist 113:631–641.

Glitzenstein, J. S., W. J. Platt, and D. R. Streng. 1995. Effects of fire regime and habitat on tree dynamics in north Florida longleaf pine savannas. Ecological Monographs 65:441-476.

Gorman, O. T. 1986. Assemblage organization of stream fishes: the effect of rivers on adventitious steams. American Naturalist 128:611-616.

Gorman, O. T., and J. R. Karr. 1978. Habitat structure and stream fish communities. Ecology 59:507-515.

Gowan, C., and K. D. Fausch. 2002. Why do foraging stream salmonids move during summer? Environmental Biology of Fishes 64:139-153. 51 Gowan, C., M. K. Young, K. D. Fausch, and S. C. Riley. 1994. Restricted movement in resident stream salmonids: paradigm lost? CANADIAN JOURNAL OF FISHERIES AND AQUATIC SCIENCES 51:2626-2637.

Gregory, S. V., F. J. Swanson, W. A. McKee, and K. W. Cummins. 1991. An Ecosystem Perspective Of Riparian Zones. Bioscience 41:540-551.

Grossman, G. D., J. F. Dowd, and M. Crawford. 1990. Assemblage stability in stream fishes: a review. Environmental Management 14:661-671.

Gurtz, M. E., and J. B. Wallace. 1984. Substrate-mediated response of stream invertebrates to disturbance. Ecology 65:1556–1569.

Haefner, J. D., and J. B. Wallace. 1981. Production and potential seston utilization by Parapsyche cardis and Diplectrona modesta (Trichoptera : Hydropsychidae) in two streams draining contrasting southern Appalachian watersheds. Environmental Entomology 10:433-441.

Harcup, M. F., R. Williams, , and D. M. Ellis. 1984. Movements of brown trout, Salmo trutta L., in the River Gwyddon, South Wales. Journal of Fish Biology 24:415– 426.

Harding, J. S., E. F. Benfield, P. V. Bolstad, G. S. Helfman, and E. B. D. Jones III. 1998. Stream biodiversity: the ghost of land use past. Proceedings of the National Academy of Sciences USA 95:14843–14847.

Harrell, L. 1992. Handling euthanasia in production facilities. Pages 129 in D. O. Schaeffer, K. M. Kleinow, and L. Krulisch, editors. The care and use of amphibians, reptiles and fish in research. Scientists Center for Welfare, Bethesda, Md.

Harrelson, C. C., C. L. Rawlins, and J. P. Potyondy. 1994. Stream channel reference sites: An illustrated guide to field technique. USDA Forest Service, General Technical Report RM-245, Collins, Colorado.

Harvey, B. C. 1991. Interactions among stream fishes: predator-induced habitat shifts and larval survival. Oecologia 87:29-39.

Harvey, B. C., R. J. Nakamoto, and L. J. White. 1999. Influence of large woody debris and a bankfull flood on movement of adult resident coastal cutthroat trout, Oncorhynchus clarki, during fall and winter. Canadian Journal of Fisheries and Aquatic Sciences 56:2161-2166.

Hayek, L. A. C., and M. A. Buzas. 1997. Surveying natural populations. Columbia University Press, New York, New York, USA.

52 Hilderbrand, R. H., and J.L. Kershner. 2000. Movement patterns of stream-resident cutthroat trout in Beaver Creek, Idaho-Utah. Transactions of the American Fisheries Society 129:1160-1170.

Hooke, J. M. 1999. Decades of change: contributions of geomorphology to fluvial and coastal engineering and management. Geomorphology 31:373-389.

Hughes, N. F. 1998. Use of whole-stream patterns of age segregation to infer the interannual movements of stream salmonids: a demonstration with Arctic grayling in an interior Alaskan stream. Transactions of the American Fisheries Society 127:1067–1071.

Hughes, N. F., and J. B. Reynolds. 1994. Why do Arctic grayling (Thymallus arcticus) get bigger as you go upstream? Canadian Journal of Fisheries and Aquatic Sciences 51:2154-2163.

Hugueny, B., L. T. deMorais, S. Merigoux, B. deMerona, and D. Ponton. 1997a. The relationship between local and regional species richness: comparing biotas with different evolutionary histories. Oikos 80:583-587.

Hugueny, B., L. T. deMorais, S. Merigoux, B. deMerona, and D. Ponton. 1997b. The relationship between local and regional species richness: comparing biotas with different evolutionary histories. Oikos 80:583-587.

Hunter, M. D., and P. W. Price. 1992. Playing chutes and ladders: heterogeneity and the relative roles of bottom-up and top-down forces in natural communities. Ecology 73:724-732.

Hupp, C. R. 2000. Hydrology, geomorphology and vegetation of Coastal Plain rivers in the south-eastern U.S.A. Hydrological Processes 14:2991-3010.

Ims, R. A., and D. O. Hjermann. 2001. Condition-dependent dispersal. Pages 600-621 in J. Clobert, E. Danchin, A. A. Dhondt, and D. J. Nichols, editors. Dispersal. Oxford University Press, New York.

Jackson, D. A., and H. H. Harvey. 1993. Fish and benthic invertebrates: Community concordance and community-environment relationships. Canadian Journal of Fisheries and Aquatic Sciences 50:2641-2651.

Kaller, M. D., and W. E. Kelso. 2006. Effects of a small-scale clearing on habitat and macroinvertebrates of a southwestern Louisiana coastal bottomland stream. The Southwestern Naturalist 51:143-151.

Kaufman, P. R., P. Levine, E. G. Robison, C. Seeliger, and D. V. Peck. 1999. Quantifying physical habitat in wadeable streams. EPA/620/R-99/003, U.S. Environmental Protection Agency, Office of Research and Development, Environmental Monitoring and Assessment Program, Washington, DC. 53 Kelly, J. R., and M.A. Harwell. 1990. ndicators of ecosystem recovery. Environmental Management 14:527-545.

Kirk, D. J., and L. A. Smock. 2000. Interspecific and intraspecific interactions between crayfish (Cambarus longulus) and hellgrammites (Corydalus cornutus) and the influence of a predatory fish (Micropterus dolomieu). American Midland Naturalist 144:317–327.

Knaepkens, G., L. Bervoets, E. Verheyen, and M. Eens. 2004. Relationship between population size and genetic diversity in endangered populations of the European bullhead (Cottus gobio): implications for conservation. Biological Conservation 115:403–410.

Krebs, C. J. 1999. Ecological methodology. Addison Wesley Longman, New York.

Lammert, M., and J. D. Allan. 1999. Assessing biotic integrity of streams: effects of scale in measuring the influence of land use/cover and habitat structure on fish and macroinvertebrates. Environmental Management 23:257-270.

Lancaster, J., A. G. Hildrew, and C. R. Townsend. 1990. Stream-flow and predation effects on the spatial dynamics of benthic invertebrates. Hydrobiologia 203:177- 190.

Larson, G. L., R. L. Hoffman, and S. E. Moore. 2002. Observations of the distributions of five fish species in a small Appalachian stream. Transactions of the American Fisheries Society 131:791-796.

Lenat, D. R., and J. K. Crawford. 1994. Effects of land use on water quality and aquatic biota of three North Carolina Piedmont streams. Hydrobiologia 294:185–199.

Li, J., A. Herlihy, W. Gerth, P. Kaufmann, S. Gregory, S. Urquhart, and D. P. Larsen. 2001. Variability in stream macroinvertebrates at multiple spatial scales. Freshwater Biology 46:87-97.

Linke, S., R. C. Bailey, and J. Schwindt. 1999. Temporal variability of stream bioassessments using benthic macroinvertebrates. Freshwater Biology 42:575- 584.

Lonzarich, D. G., M. R. Lonzarich, and M. L. Warren Jr. 2000. Effects of riffle length on the short-term movement of fishes among stream pools. Canadian Journal of Fisheries and Aquatic Sciences 57:1508–1514.

Lorang, M. S., and F. R. Hauer. 2003. Flow competence evaluation of steambed stability: An assessment of the technique and limitations of application. Journal of the North American Benthological Society 22:475-491.

54 Lotrich, B. A., and W.H. Meredith. 1974. A technique and the effectiveness of various acrylic colors for subcutaneous marking of fish. Transactions of the American Fisheries Society 103:140-142.

Ludwig, J. A., and J. F. Reynolds. 1988. Statistical Ecology: a Primer on Methods and Computing. John Wiley & Sons, Inc., New York.

Macdonald, J. S., MacIsaac, E.A., and Herunter, H. 2003. The effects of variable- retention riparian buffer zones on water temperatures in small headwater streams in sub-boreal forest ecosystems of British Columbia. Canadian Journal of Forest Research 33:1371–1382.

MacDonald, L. H., and E. L. Huffman. 2004. Post-fire soil water repellency: Persistence and soil moisture thresholds. Soil Science Society of America Journal 68:1729- 1734.

Maloney, K. O., P. J. Mulholland, and J. W. Feminella. 2005. Influence of catchment- scale military land use on stream physical and organic matter variables in small southeastern plains catchments (USA) Environmental Management 35:677-691.

Martin, G., C. L. Butler, E. Scott, and J. E. Lyles. 1990. Soil survey of Natchitoches Parish, Louisiana. United States Department of Agriculture, Soil Conservation Service.

Matthews, G. B., R. A. Matthews, and B. Hachmoller. 1991. Mathematical analysis of temporal and spatial trends in the benthic macroinvertebrate communities of a small stream. Canadian Journal of Fisheries and Aquatic Sciences 48:2184-2190.

Matthews, W. J. 1986. Fish faunal structure in an Ozark stream: stability, persistence and a catastrpohic flood. Copeia 1986:388-397.

Matthews, W. J., R. C. Cashner, and F. P. Gelwick. 1988. Stability and persistence of fish faunas and assemblages in three midwestern streams. Copeia 1988:945-955.

McClelland, W. T., and M. A. Brusven. 1980. Effects of sedimentation on the behavior and distribution of riffle insects in a laboratory stream. Aquatic Insects 2:161-169.

McCullagh, P., and J. A. Nelder. 1989. Generalized Linear Models, 2nd edition. Chapman and Hall, London.

McCune, B., and J. B. Grace. 2002. Analysis of ecological communities. MjM Software, Gleneden Beach, Oregon.

McCune, B., and M. J. Mefford. 1999. Multivariate analysis of ecological data. Version 4. MjM Software Design, Gleneden Beach, OR.

55 McQueen, D. J., M. R. S. Johannes, J. R. Post, T. J. Stewart, and D. R. S. Lean. 1989. Bottom-up and top-down impacts on freshwater pelagic community structure. Ecological Monographs 59:289–309.

Mebane, C. A. 2001. Testing bioassessment metrics: macroinvertebrate, sculpin, and salmonid responses to stream habitat, sediment, and metals. Environmental Monitoring and Assessment 67:293–322.

Meffe, G. K., and A. L. Sheldon. 1988. The influence of habitat structure on fish assemblage composition in southeastern black water streams. American Midland Naturalist 120:225-240.

Meffe, G. K., and A. L. Sheldon. 1990. Post-defaunation recovery of fish assemblage in southeastern blackwater streams. Ecology 71:657-667.

Merritt, R. W., and K. W. Cummins. 1996. An introduction to the aquatic insects of North America, 3rd edition. Kendall Hunt Publishing Company, Iowa.

Meyer, J. L., L. A. Kalpan, D. Newbold, D. L. Strayer, C. J. Woltemade, J. B. Zedler, R. Beilfuss, Q. Carpenter, R. Semlitsch, M. C. Watzin, and P. H. Zedler. 2003. Where rivers are born: The scientific imperative for defending small streams and wetlands. American Rivers, Sierra Club, Washington DC.

Minshall, G. W. 1988. Stream ecosystem theory: global perspective. Journal of the North American Benthological Society 7:263-288.

Moloney, K. A., and S. A. Levin. 1996. The effects of disturbance architecture on landscape level population dynamics. Ecology 72:375-394.

Murphy, M. L., and J. D. Hall. 1981. Varied effects of clear-cut logging on predators and their habitat in small streams of the Cascade Mountains, Oregon. Canadian Journal of Fisheries and Aquatic Sciences 38:137–145.

Naiman, R. J., and H. Decamps. 1997. The ecology of interfaces: riparian zones. The Annual Review of Ecological and Systematics 28:621-658.

Naiman, R. J., and H. Decamps. 1997. The ecology of interfaces: riparian zones. The Annual Review of Ecological and Systematics 28:621-658.

Nisbet, R. M., E. McCauley, W. S. C. Gurney, W. W. Murdoch, and A. M. de Roos, editors. 1997. Simple representations of biomass dynamics in structured populations. Case Studies in Mathematical Modeling. Prentice Hall.

Noonan, G. R. 1988. Biogeography of North American and Mexican insects, and a critique of vicariance biogeography. Systematic Zoology 37:366-384.

Northcote, T. J. 1997. Potamodromy in Salmonidae - living and moving in the fast lande. North American Journal of Fisheries Management 17:1029-1045. 56 Nystrom, P., A. R. McIntosh, and M. J. Winterbourn. 2003 Top-down and bottom-up processes in grassland and forested streams

Oecologia 136 596-608.

O’Neill, R. V., D. L. DeAngelis, J. B. Waide, and T. F. H. Allen. 1986. A Hierarchical Concept of Ecosystems. Princeton University Press, Princeton, New Jersey.

Omernik, J. M. 1987. Ecoregions of the conterminous United States. Annals of the Association of American Geographers 77:118-125.

Palmer, M. W. 1993. Putting things in even better order: the advantages of canonical correspondence analysis. Ecology 74:2215-2230.

Pearson, S. M., A. B. Smith, and M. G. Turner. 1998. Forest fragmentation,land use, and cove-forest herbs in the French Broad River Basin. Castanea 63:382–395.

Pecarsky, B. L. 1983. Biotic interactions or abiotic limitations? A model of lotic community structure. in T. D. Fontaine III and S. M. Bartell, editors. Dynamics of lotic ecosystems. Ann Arbor Science, Ann Arbor, Michigan.

Peckarsky, B. L. 1984. Predator-prey interactions among aquatic insects. Pages 196-254 in V. Resh and D. M. Rosenberg, editors. The Ecology of Aquatic Insects. Praeger, New York.

Peckarsky, B. L., and M. A. Penton. 1990. Mechanisms of prey selection by stream- dwelling stoneflies. Ecology 70: :1203–1218.

Peterson, J. T., and P. B. Bayley. 1993. Colonization rates of fishes in experimentally defaunated warmwater streams. Transactions of the American Fisheries Society 122:199-207.

Pianka, E. R. 1994. Evolutionary ecology, 5th edition. Harper Collins, New York, NY.

Pielou, E. C. 1984. The Interpretation of Ecological Data: a Primer on Classification and Ordination. John Wiley & Sons, Inc., New York.

Poff, N. L. 1997. Landscape filters and species traits: Towards mechanistic understanding and prediction in stream ecology. Journal of the North American Benthological Society 16:391-409.

Poff, N. L., J. D. Allan, M. B. Bain, J. R. Karr, K. L. Prestegaard, B. D. Richter, R. E. Sparks, and J. C. Stromberg. 1997. The Natural Flow Regime - A paradigm for river conservation and restoration. BioScience 47:769-784.

Poff, N. L., and J.D. Allan. 1995. Functional organization in stream fish assemblages in relation to hydrologic variability. Ecology 76:606-627. 57 Poff, N. L., and J. V. Ward. 1989. Implications of streamflow variability and predictability for lotic community structure: a regional analysis of streamflow patterns. Canadian Journal of Fisheries and Aquatic Sciences 46:1805-1818.

Poff, N. L., and J. V. Ward. 1990. Physical habitat template of lotic systems: recovery in the context of historical pattern of spatiotemporal heterogeneity. Environmental Management 14:629-645.

Polis, G. A., and D. R. Strong. 1996. Food web complexity and community dynamics. American Naturalist 147:813-846.

Polis, G. A., and K. Winemiller. 1996. Food Webs: Integration of Patterns and Dynamics. Chapman & Hall, New York.

Pringle, C. M., M. C. Freeman, and B. J. Freeman. 2000. Regional effects of hydraulic alterations on riverine macrobiota in the new world: tropical-temperate comparisons. Bioscience 50:807-823.

Quinn, G. P., and M. J. Keough. 2002. Experimental design and data analysis for biologists. Cambridge University Press, Cambridge, UK.

Rader, R. B., and J. V. McArthur. 1995. The relative importance of refugia in determining the drift and habitat selection of predaceous stoneflies in a sandy- bottom stream. Oecologia 103:1–9.

Rahel, F. J. 1990. The hierarchical nature of community persistence: a problem of scale. American Naturalist 136:328-344.

Railsback, S. F., R. H. Lamberson, B. C. Harvey, and W. E. Duffy. 1999. Movement rules for individual-based models of stream fish. Ecological Modelling 123:73-89.

Resh, V. H., A. V. Brown, A. P. Covich, M. E. Gurtz, H. W. Li, G. W. Minshall, S. R. Riece, A. L. Sheldon, J. B. Wallace, and R. C. Wissmar. 1988. The role of disturbance in stream ecology. Journal of North American Benthological Society 7:433-455.

Richards, C., and G. Host. 1994. Examining land use influences on stream habitats and macroinvertebrates: a GIS approach. Water Resources Bulletin 30:729-738.

Richardson, J. S. 1991. Seasonal food limitation of detritivores in a montane stream: an experimental test. Ecology 72:873- 887.

Rinne, J. N. 1990. Status, distribution, biology, and conservation of two rare southwestern (U.S.A.) salmonids, the Apache trout, Oncorhynchus apache Miller, and the Gila trout, O. gilae Miller. . Journal of Fish Biology 37:189-191.

58 Roberts, J. 2000. The influences of physical and physiological characteristics of vegetation on their hydrological response. Hydrological Processes 14:2885- 28901.

Rosgen, D. 1996. Applied River Morphology. Wildland Hydrology Pagosa Springs, CO.

Ross, S. T. 2001. Inland fishes of Mississippi. University Press of Mississippi, Jackson.

Sabo, J. L., and M. E. Power. 2002. River–watershed exchange effects of riverine subsidies on riparian lizards and their terrestrial. Ecology 83:1860-1869.

Saucier, R. T. 1994. Geomorphology and quarternary geologic history of the Lower Mississippi Valley. U.S. Army Corps of Engineers, Vicksberg, Mississippi.

Schaefer, J. 2001. Riffles as barriers to interpool movement by three cyprinids (Notropis boops, Campostoma anomalum and Cyprinella venusta). Freshwater Biology 46:379-388.

Schaefer, J. F., E. Marsh-Matthews, D. E. Spooner, K. B. Gido, and W. J. Matthews. 2003. Effects of barriers and thermal refugia on local movement of the threatened leopard darter, Percina pantherina. Environmental Biology of Fishes 66:391-400.

Schlosser, I. J. 1987. A conceptual framework for fish communities in small warmwater streams. Pages 17-24 in W. J. Matthews and D. C. Heins, editors. Community and evolutionary ecology of North American stream fishes. University of Oklahoma Press, Norman, Oklahoma.

Schlosser, I. J. 1995. Dispersal, boundary processes, and trophic level interactions in streams adjacent to beaver ponds. Ecology 76:908-925.

Schlosser, I. J., and P. L. Angermier. 1995. Spatial variation and demographic processes of lotic fishes: conceptual models, imperical evidence, and implications for conservation. Pages 392-401 in J. L. Nielsen, editor. Evolution in the aquatic ecosystem: defining unique units in population conservation.

Shaw, E. A., and J.S. Richardson. 2001. Effects of fine inorganic sediment on stream invertebrate assemblages and rainbow trout (Oncorhynchus mykiss) growth and survival: implications of exposure duration. Canadian Journal of Fisheries and Aquatic Sciences 58:2213-2221.

Silsbee, D. G., and G. L. Larson. 1983. A comparison of streams in logged and unlogged areass of Great Smoky Mountains National Park. Hydrobiologia 102:99-111.

Smithson, E. B., and C. E. Johnston. 1999. Movement patterns of stream fishes in a Ouachita Highlands stream: an examination of the restricted movement paradigm. Transactions of the American Fisheries Society 128:847–853.

59 Sousa, W. P. 1984. The role of disturbance in natural communities. Annual Review of Ecology and Systematics 15:335-391.

Stewart, K. W., B. P. Stark, and T. G. Huggins. 1976. The stoneflies (Plecoptera) of Louisiana. Great Basin Naturalist 36:366-384.

Stock, J. D., and I. J. Schlosser. 1991. Short-term effects of a catastrophic beaver dam collapse on a stream fish community. Environmental Biology of Fishes 31:123– 129.

Stone, M., and J. Wallace. 1998. Long-term recovery of a mountain stream from clear-cut logging: the effects of forest succession on benthic invertebrate community structure. Freshwater Biology 39:151–169.

Surdick, R. F. 1985. Neartic genera of Chloroperlinae (Plecoptera: Chloroperlidae). Illinois Biological Monographs 54:1-146.

Sutherland, D. G., M. W. Hansler-Ball, S. J. Hilton, and T. L. Lisle. 2002. Evolution of a landslide-induced sediment wave in the Navarro River, California. Geological Society of America Bulletin 114:1036-1048.

Swank, W. T., J. M. Vose, and K. J. Elliot. 2001. Long term hydrologic and water quality responses following commercial clearcutting of mixed hardwoods on a southern Appalachian catchment. Forest Ecology and Management 143:163-178.

Tabacchi, E., L. Lambs, H. Guilloy, A.-M. Planty-Tabacchi, E. Muller, and H. Decamps. 2000. Impacts of riparian vegetation on hydrological processes. Hydrological Processes 14:2959-2976.

Taylor, C. M. 1997. Fish species richness and incidence patterns in isolated and connected stream pools: effects of pool volume and spatial position. Oecologia 110:560–566. ter Braak, C. J. F. 1986. Canonical correspondence analysis: a new eigenvector technique for multivariate direct gradient analysis. Ecology 67:1167-1179. ter Braak, C. J. F., and P. Smilauer. 2002. CANOCO reference manual and CanoDraw for Windows user's guide: software for canonical community ordination (version 4.5). Microcomputer Power, Ithaca, New York, USA.

Toepfer, C. S., W. L. Fisher, and J. A. Haubelt. 1999. Swimming performance of the threatened leopard darter in relation to road culverts. Transactions of the American Fisheries Society 128:155-161.

Tonn, W. M. 1990. Climate Change and Fish Communities - a Conceptual-Framework. Transactions of the American Fisheries Society 119:337-352.

60 Townsend, C. R. 1989. The patch dynamics concept of stream community ecology. Journal of the North American Benthological Society, 8:36-50.

Turner, M. G., S. M. Pearson, P. Bolstad, and D. N. Wear. 2003. Effects of land-cover change on spatial pattern of forest communities in the Southern Appalachian Mountains (USA). Landscape Ecology 18:449-464.

Urban, D. L., R. V. O'Neill, and H. H. Shugart Jr. 1987. Landscape ecology: a hierarchical perspective can help scientists understand spatial patterns. Bioscience 37:119-127.

Vannote, R. L., G. W. Minshall, K. W. Cummins, J. R. Sedell, and C. E. Cushing. 1980. The river continuum concept. Canadian Journal of Fisheries and Aquatic Sciences 37:130-137.

Vuori, K. M., and I. Joensuu. 1996. Impact of forest drainage on the macroinvertebrates of a small boreal headwater stream: do buffer zones protect lotic biodiversity? Biological Conservation 77:87-95.

Wallace, J. B. 1990. Recovery of lotic macroinvertebrate communities from disturbance. Environmental Management 14:605-620.

Wallace, J. B., and R. W. Merritt. 1980. Filter-feeding ecology of aquatic insects. Annual Review of Entomology 25:103-132.

Wallace, J. B., S. L. Eggert, J. L. Meyer, and J. R. Webster. 1997. Multiple trophic levels of a forest stream linked to terrestrial litter inputs. Science 277:102–104.

Ward, A. D., and S. W. Trimble. 2004. Environmental Hydrology, 2nd edition. A CRC Press Company, Washington D.C.

Ward, J. V. 1989. The four-dimentional nature of lotic ecosystems. Journal of North American Benthological Society 8:2-8.

Warren Jr, M. L., P. L. Angermier, B. M. Burr, and W. R. Haag. 1997. Declines of a diverse fish fauna: patterns of imperilment and protection of the southeastern United States. Pages 105-164 in G. W. Benz and D. E. Collins, editors. Aquatic fauna in peril: the southeastern perspective. Lenz Design in Communications, Decatur, GA.

Warren Jr, M. L., and M. G. Pardew. 1998. Road crossings as barriers to small-stream fish movement. Transactions of the American Fisheries Society 127:637-644.

Waters, T. F. 1995. Sediment in streams: sources, biological effects, and control. Bethesda, Maryland, USA.

Webster, J. R., and J. B. Waide. 1982. Within-stream factors affecting nutrient transport from forested and logged watersheds. Freshwater Biology 12:331-341. 61 Wheeler, A. P., P. L. Angermeier, and A. E. Rosenberger. 2005. Impacts of new highways and subsequent landscape urbanization on stream habitat and biota. Reviews in Fisheries Science 13:141-164.

Wiens, J. A. 2001. The landscape contents of dispersal. Pages 97-109 in J. Clobert, E. Danchin, A. A. Dhondt\, and J. D. Nichols, editors. Dispersal. Oxford University Press, New York.

Wiens, J. A., J. T. Rotenberry, and B. van Horne. 1986. A lesson in the limitation of field experiments: shrubsteppe birds and habitat alteration. Ecology 67:365-376.

Williams, L. R., T. H. Bonner, J. D. Hudson III, M. G. Williams, T. R. Leavy, and C. S. Williams. 2005. Interactive efects of environmental variability and military training on stream biota of three headwater drainages in Western Louisiana. Transactions of the American Fisheries Society 134:192-206.

Williams, L. R., C. M. Taylor, M. L. Warren Jr, and J. A. Clingenpeel. 2002. Large-scale effects of timber harvesting on stream systems in the Ouachita Mountains, Arkansas, USA. Environmental Management 29:76-87.

Williams, L. R., C. M. Taylor, M. L. Warren Jr, and J. A. Clingenpeel. 2004. The role of regional factors in structuring Ouachita Mountain stream assemblages. in J.M. Guldin, editor. Ouachita and Ozark Mountains symposium: ecosystem management research. USDA Forest Service, Southern Research Station. Gen Tech Rep. SRS-74, Asheville, North Carolina.

Williams, L. R., C. S. Toepfer, and A. D. Martinez. 1996. The relationship between fish assemblages and environmental gradients in an Oklahoma prairie stream. Journal of Freshwater Ecology 11:459-468.

Wondzell, S. M., and J. G. King. 2003. Postfire erosional processes in the Pacific Northwest and Rocky Mountain regions. Forest Ecology and Management 178:75-87.

Wood, A. L., A. Simon, P. W. Downs, and C. R. Thorne. 2001. Bank-toe processes in incised channels: the role of apparent cohesion in the entrainment of failed bank material. Hydrological Processes 15:39-61.

62 TABLES

Little Sandy Odom Tiger

Disturbance extent of the streams Total Area (hectares) 1015.87 1377.58 224.21 Clear-cut Area (hectares) 0 266.8 4.84 Thinned Area (hectares) 0 504.98 129.44 New Roads Constructed 0 34.13 4.08 (hectares) Percent Disturbed 0 58.5 61.7 Stream characteristics Floodplain width (m) 12 16 9 Bankfull width (m) 10 9 5 Mean water depth (m) 0.27 0.23 0.25 Max. depth (m) 0.039 0.029 0.041 Hydraulic radius 2.5 2.1 2.1 D50 (mm) 0.4 0.28 41 D84 (mm) 0.66 0.41 58 Tractive force (Kg force/m2) 4.86 6.67 11.75 Geomorphological features Entrenchment Ratio 1.23 1.68 1.6 Width to Depth Ratio 12.1 13.3 6.4 Sinuosity 1.1 1.8 2 Bed Slope (%) 0.18 0.29 0.47

TABLE 2.1 Sampling site description in the catchment basin affected by clear-cutting and new roads, during the year of timber harvest disturbance (2004). Stream characteristics, and geomorphological feature data was collected during summer 2005.

63

Little Sandy Creek Odom Creek Tiger Creek 2003 2004 2005 2003 2004 2005 2003 2004 2005

Conductivity (µS/cm) 116.50 (14.7) 32.23 (14.4) 67.33 (18.9) 90.0 (6.8) 38.33 (3.8) 48.0 (12.1) 107.25 (24.3) 59.5 (10.7) 50.33 (11.8) Dissolved Oxygen (mg/L) 9.2 (1.0) 4.62 (2.8) 8.84 (0.3) 8.22 (0.6) 8.25 (0.6) 9.13 (0.4) 7.93 (0.5) 7.81 (0.2) 8.65 (0.6) pH 7.17 (0.3) 7.63 (0.9) 6.34 (0.4) 7.0 (0.2) 6.22 (0.2) 6.33 (0.9) 7.39 (0.2) 7.59 (0.6) 6.68 (0.6) Temperature (ºC) 17.66 (3.3) 26.48 (2.3) 22.09 (2.7) 18.51 (4.7) 25.18 (3.5) 21.86 (3.0) 17.2 (3.1) 18.42 (1.8) 20.62 (2.2) Turbidity (NTU*) 11.68 (6.0) 51.0 (7.5) 12.93 (4.9) 8.53 (4.7) 66.13 (12.5) 14.1 (6.0) 21.98 (8.4) 32.03 (3.8) 56.03 (40.2) Mean Water Velocity (m/s) 0.08 (0.04) 0.17 (0.05) 0.14 (0.05) 0.16 (0.04) 0.32 (0.08) 0.07 (0.03) 0.09 (0.05) 0.11 (0.02) 0.04 (0.02) Mean Mesohabitat 64 Length (m) 99.14 (5.3) 104 (14.3) 81.77 (4.6) 236.3 (23.4) 294.7 (28.0) 128.1 (12.9) 62.82 (3.8) 61.97 (1.0) 101.3 (12.9) Mean Mesohabitat Width (m) 11.82 (0.4) 28.84 (4.1) 12.67 (0.5) 14.7 (0.5) 16.61 (0.2) 13.42 (1.5) 7.42 (0.2) 8.53 (0.1) 9.51 (0.1) Mean Water Depth (m) 0.17 (02) 0.6 (0) 0.15 (0.04) 0.11 (0.03) 0.12 (0.03) 0.52 (0.3) 0.17 (0.03) 0.17 (0.02) 0.12 (0.06)

Clay % 0.0 (0) 8.58 (7.6) 0.0 (0) 0.0 (0) 0.0 (0) 0.0 (0) 0.0 (0) 0.0 (0) 0.0 (0) Silt % 0.35 (0.4) 7.5 (7.3) 0.0 (0) 0.0 (0) 0.0 (0) 0.0 (0) 0.0 (0) 0.0 (0) 1.67 (1.7) Sand % 69.41 (7.4) 40.25 (3.1) 67.92 (3.1) 96.63 (3.1) 96.88 (3.1) 66.67 (33.3) 84.81 (8.6) 88.16 (7.7) 82.02 (6.4) Gravel % 1.25 (1.3) 3.57 (2.9) 5.21 (4.1) 0.0 (0) 0.0 (0) 0.0 (0) 8.08 (3.9) 5.56 (3.2) 15.09 (8.0) Cobble % 18.15 (2.9) 13.15 (7.2) 16.11 (2.2) 3.38 (3.1) 3.13 (3.1) 0.0 (0) 7.12 (5.2) 6.28 (5.5) 1.22 (0.9) Bedrock % 10.83 (7.9) 2.92 (2.8) 10.76 (5.7) 0.0 (0) 0.0 (0) 0.0 (0) 0.0 (0) 0.0 (0) 0.0 (0) Percent Undercut Bank 6.61 (2.2) 6.23 (3.8) 8.27 (1.7) 6.07 (3.0) 7.62 (6.3) 19.96 (7.1) 6.81 (2.8) 6.44 (1.7) 11.75 (2.9) Percent Woody Debris 6.12 (3.6) 2.53 (1.1) 2.23 (0.5) 5.15 (1.6) 5.56 (2.2) 7.52 (2.6) 5.39 (0.9) 9.07 (2.0) 8.04 (1.8) *Nephelometric turbidity units

TABLE 2.2 Mean physical and chemical water quality data, and stream habitat quality data, collected from Red River drainage streams in the years 2003, 2004, and 2005 represented by pre-, during-, and post- timber harvest activities respectively.

Source DF Sum of Squares Mean Square F Ratio Prob > F Model 6 2692.95 448.82 23.99 <.0001 % EPT Error 17 317.92 18.7 C. Total 23 3010.87 Regression Eq: % EPT = – 107.16 – 58.59 Current velocity + 3.44 Turbidity + 83.05 pH + 36.25 DO + 25.12 %Undercut bank + 17.16 %Woody debris

r2 = 0.89, Adj. r2 = 0.85, n = 24

Model 7 4810.42 687.2 7.41 0.0004 % Diptera Error 17 1574.88 92.64 C. Total 24 6385.3 Regression Eq: % Diptera = – 38.82 – 14.24 Conductivity + 141.49 DO – 28.31 %Undercut bank + 11.86 Stream length + 61.93 Water depth – 44.24 %Woody debris

r2 = 0.75, Adj. r2 = 0.65, n = 25

TABLE 2.3 Regression equations describing the Red River drainage relationships between the % EPT and % Diptera taxa and the stream physical and chemical parameters. ANOVA results of the whole regression model are also displayed.

65

Stream Little Sandy Odom Tiger 2003 2004 2005 2003 2004 2005 2003 2004 2005 2003 0.66 0.66 0.48 0.53 0.54 0.38 0.33 0.64 2004 0.84 0.59 0.64 0.62 0.61 0.60 0.60 2005 0.67 0.71 0.64 0.62 0.60 0.68 2003 0.92 0.62 0.64 0.60 0.59 2004 0.66 0.62 0.57 0.64 Year 2005 0.56 0.50 0.63 2003 0.70 0.50 2004 0.45 2005

TABLE 2.4 Renkonen Similarity Index of sites across years encompassing the years 2003, 2004, and 2005 that was represented by pre-, during-, and post- timber harvest period.

66

Factor Variance explained (%) Year 2.98 Stream 3.05 Habitat 9.16 Season 13.97 Full model 43.8

TABLE 2.5 Percent variance of partial canonical correspondence analysis on macroinvertebrate assemblage through timber harvest regime (2003 – 2005) in Ft. Polk, Louisiana, as explained by year (pre-, during-, and post- timber harvest years), stream (3 streams), habitat (3 mesohabitat), and season (4 levels). All the % variance explained in significant (P = 0.001) based on Monte Carlo randomization test.

67

FIGURES

Red River A B Drainage

C

Sabine River Drainage N Calcasieu River Drainage W E

0 51015S

Kilometers

FIGURE 2.1 Study area in the Peason Ridge Training Area (PRTA) in Fort Polk, Louisiana, with three adjacent drainages. Sampling sites located on the streams in Red River drainage are – A is Little Sandy Stream, B is Tiger Stream, and C is Odom Stream. Little Sandy was the reference stream that did not experience timber harvesting activities; whereas, the shaded protion of the Red River drainage experienced timber harvesting to varying degrees.

68

3 2003 2004 2005 2

1

0

2 (14.5%) Axis -1

Depth Width, Length Width, Depth -2

-3

Dimensions: -10123 Axis 1 (28.9%) Substrate: Sandy Cobble Water quality: Conductivity Turbidity Physical: Undercut bank % Woody debris %

FIGURE 2.2 Relationship between site locations, and physical and chemical conditions of the stream. Proportion of variance accounted for by the PCA axes: Axis 1, 28.946%, and axis 2, 14.57%. PCA axis I showed a linear bed substrate gradient, water quality conditions, and stream bank condition. PCA axis II showed a gradient along the stream bed dimensions. With increasing PCA II scores, stream environment changes from deeper to shallower and wider areas.

69

60 Little Sandy Odom Tiger

40

20 % EPT

% Diptera 0 003 003 05 2 2004 2005 2003 2004 2005 2 2004 20

FIGURE 2.3 Macroinvertebrate assemblage composition represented by the mean relative abundance (and Standard Error bars) of Ephemeroptera, Plecoptera, and Trichoptera (EPT, open circles), and the relative abundance of Diptera (filled circles) within the total assemblage shown as a function of the years 2003, 2004, and 2005 that was represented by pre-, during-, and post- timber harvest period

70

Little Sandy Odom Tiger

40

20

Family Richness Family 0

4 3 2

1

log Family Abundance 0

5

4 3 2 1 0 Shannon WeinerIndex

3

2

1

Evenness index Evenness 0 l l l l l l l l n o o e n e n e le n e o o ffl ffl un ff ffl un Poo Ru P Run P iffl Run Poo Ru Poo i Ru Poo i R Pool i Run Poo Ru Poo i R Riffle Riffle R Riffle R R R Riffle R 2003 2004 2005 2003 2004 2005 2003 2004 2005

FIGURE 2.4 Macroinvertebrate assemblage patterns in mean values of a) family richness; b) family abundance (log transformed); c) Shannon Weiner index; and d) Evenness index, in the Red River drainage, Louisiana, across the years 2003, 2004, and 2005 that was represented by pre-, during-, and post- timber harvest period (Little Sandy creek was the reference reach with no timber harvest activities.

71

Little Sandy Odom Tiger 100% Shredder

80% Scraper

60% Predator

40% Piercer Filterer

20% Collector

0%

2003 2004 2005 2003 2004 2005 2003 2004 2005

FIGURE 2.5 Proportion of functional feeding groups of macroinvertebrates based on percent abundance in Red River drainage streams in Louisiana, across the years 2003, 2004, and 2005.

72

5

4

3 Pool 2 Riffle Axis 2 1 Run a) 0

-1 -1012345

Axis 1 5

4

3 Winter Fall 2

Axis 2 Summer 1 Spring b) 0

-1 -1012345

Axis 1

FIGURE 2.6 Results from detrended correspondence analyses (DCA) based on macroinvertebrate assemblages addressing the spatial and temporal component of streams in the Red River Basin, Louisiana, from 2003 to 2005. Site scores coded by year did not show a distinct clustering (not shown). a) DCA site scores for mesohabitats showed a gradient along the first axis with pool, riffle, and run scores. Pool scores had lowest values along Axis 1, whereas Riffle scores were highest. b) The DCA site scores for the seasonal component showed separate grouping of the spring, summer, fall, and winter scores occupying each quadrant. A longitudinal gradient is detected along the first axis with Winter scores having lowest values, and Summer scores, the highest.

73

4 Summer

3 Fall 2

Gravel 1 Winter

0 2 (15.3) Axis -1

Length Depth Width, : Sand Sand : a) -2 Spring

-3

-3-2-101234

Dimensions: Substrate Axis 1 (30.7%)

Water quality: Temperature, Conductivity, Turbidity,

pH, Dissolved Oxygen Velocity Habitat features: % Woody debris, % Undercut bank

4

3 Odom Creek

2 Little Sandy Creek 1

Tiger Creek 0

(15.3) 2 Axis -1 Pre-Timber harvest b) -2 During-

Post- -3 -3 -2 -1 0 1 2 3 4 Axis 1 (30.7%)

FIGURE 2.7 Ordination biplot depicting the relationship between environmental parameters and site scores. a) Axis I shows a water quality, and habitat features gradient. Axis II shows a stream bed substrate composition, and dimensions gradient. The site scores show a clustering pattern of summer plots with lower scores, and winter plots with higher scores along Axis 1. b) Same ordination with site scores enveloped by stream type.

74

CHAPTER 3

Impacts of road crossings on fish movement in Gulf Coastal Plain streams

INTRODUCTION

Movement is a critical process that allows fish to meet their resource needs in

spatially and temporally variable stream environments (Schlosser and Angermeier 1995;

Fausch and Young 1995; Fausch et al. 2002). Fish movement, at small scale, allows

individual fish to occupy the most suitable habitats for survival and growth (Gowan and

Fausch 2002), and at larger scale, permits migrations between habitats used by different

life history stages, exploitation of refugia from large-scales disturbances, gene flow for small populations, and colonization or recolonization of unoccupied habitats (Brown and

Kodric- Brown 1977, Peterson and Bayley 1993, Northcote 1997). While smaller scale movements are necessary for access to prey or avoidance from predators (Harvey 1991), whereas, larger-scale movements are essential for persistence of individual populations and metapopulations (Meffe and Sheldon 1990, Schlosser and Angermeier 1995).

Because of the patchiness of resources, and its variability in quantity and quality over

time, stream organisms are forced to move longitudinally to fulfill their life history

obligations (Townsend 1989). Recent studies on stream fish movement have revealed

that typically only a few individuals in a population move long distances, while most

individuals tend to remain localized, making short runs exploring adjacent habitats 75 (Smithson and Johnston 1999, Schaefer 2001). Thus, proper knowledge of movement is necessary to improve our understanding of stream fish ecology. The need for conservation efforts with respect to fish movement is necessary because human modifications through road crossings or other barriers to dispersal may disrupt natural movement patterns (Warren et al. 1997; Wiens 2001).

Small streams are a good model system to examine patterns of stability and variability in assemblage structure across spatial and temporal scales (Wiens et al. 1986,

Grossman et al. 1990, Lammert and Allan 1999), as they are naturally variable environments and may experience a “harsh” natural disturbance regime (Pecarsky 1983,

Matthews 1986, Matthews et al. 1988). Disturbance is recognized increasingly as playing an important role in the structure and dynamics of communities (Sousa 1984,

Schlosser 1991). Timber harvesting, and its byproducts (roads, and road crossings) are major forms of anthropogenic disturbance to streams (Williams et al. 2002). Timber harvesting typically results in removal of riparian vegetation that serves to stabilize soils, hold water, and prevent erosion into the stream channel, and ultimately may lead to change in fish assemblage structure (Campbell and Doeg 1989). Streams in the U.S.

Pacific NW have been adversely affected by timber harvesting with effects lasting past five years (Macdonald et al. 2003, Shaw and Richardson 2001); however, the southeastern U.S. stream biota have shown greater resilience to small- (Kaller and Kelso

2006), medium- (Warren and Pardew 1998, Williams et al. 2005), and large- (Williams et al. 2002) scale timber harvesting activities, including impacts of roads and crossings.

Barriers to movement such as dams and road crossings have an obvious impact on fish populations by preventing migrations, gene flow, and colonization (Warren and

76 Pardew 1998; Pringle et al. 2000). Poorly designed bridges can inhibit movement

because of altered hydrology and geomorphology. Culverts especially can present

barriers to fish movement because of low water level within the culverts, lack of refuge

pools on both ends, and a drop from the outlet to the stream surface (Baker and Votapka

1990). Although round pipe and box culverts allow water to pass through, they

potentially reduce migration of fishes by concentrating discharge and altering the stream

cross-section fish use to move within the stream. A lab study by Toepfer et al. (1999)

found that the current velocity in both pipe and box culverts restrict movement of darters.

The rate of movement across natural and artificial barriers is different among stream fish

species based on their tolerance limits (Warren and Pardew 1998, Lonzarich et al. 2000).

Fishes that readily move under certain ecological conditions may fail to do so in modified

situations (Gowan et al. 1994). Identifying ecological factors that drive dispersal- mediated life history events (e.g., spawning migrations) is a key step in predicting how fishes will respond to natural and anthropogenic changes in environmental conditions

(Railsback et al. 1999).

Managing fish populations is far more complicated when factors influencing

movement vary among species. Diverse life histories and habitat requirements suggest

that factors may also be species specific in fishes. For many stream-fish species,

movement has been associated with physical structure (i.e., presence of cover; Aparicio

and Sostoa 1999; Harvey et al. 1999; Gilliam and Fraser 2001). Identifying common

factors that influence movement across species may facilitate management efforts that benefit entire assemblages. Despite the management implications, few studies have linked movement of stream fish to ecological factors (Gowan et al. 1994; Gilliam and

77 Fraser 2001). Most studies that have examined ecological correlates of movement, have been limited in the number of factors and/or attributes of movement examined, which

leads to an incomplete understanding of movement (Ims and Hjermann 2001).

Furthermore, virtually no studies have explicitly examined how ecological correlates of

movement vary within an assemblage of fishes. The main objective of my study is to

identify factors associated with the movement of fishes in a network of streams in

western Louisiana over a period of two years. In addition to annual variation in

movement, this fish mark, release, and recapture study integrates pre- and post- road

crossing installation sites. I examined relationships between each attribute of movement

and a suite of predictor variables related to reach specific habitat characteristics and

attributes of individual fish species.

METHODS

Study Area

The streams in the study area (Figure 3.1) belong to three drainages: Red River

drainage (Reaugaulle, Stagestand, Odom, and Squirrel Branch streams); Calcasieu River

drainage (Comrade stream); and Sabine River drainage (Martin stream). The forests are

dominated by longleaf pine (Pinus palustris), interspersed with hardwoods (e.g., bald cypress Taxodium distichum, white oak Quercus alba, water oak Quercus nigra, and magnolia Magnolia randiflora) more common along the lowland reaches (Williams et al.

2005). Because longleaf pine (P. palustris) is a ‘fire climax’ community, Fort Polk management conducts a regulatory prescribed burn cycle on a 2-3 year cycle. This

78 ecosystem supports a population of federally endangered Red Cockaded Woodpecker

(Picoides borealis).

The streams of Red River drainage are fed by seepage from the northern region of

Peason Ridge and drain into Kisatchie Bayou on the northeastern side. These streams are fairly high gradient, and have loamy to sandy soils (Williams et al. 2005). The southern drainages, Sabine and Calcasieu, are of low gradient because of the sediment deposits during the glaciation period (Conner and Suttkus 1986), and share common fish species composition (Douglas 1974). Timber harvesting began in the year 2003 in the upper reaches of the watershed for constructing a range complex that facilitates combined arms

training, which is a part of Digital Multipurpose Battle Area Course (DMPBAC). The

upper reaches of this region were clear-cut in late 2003 for creating firing lanes, road

systems to allow transport of large equipment, and numerous stream crossings. The

military primarily uses this region for live fire exercises and troop maneuvers using

heavy-duty vehicle.

I sampled road crossings on Reaugaulle, Stagestand, Odom, Squirrel branch,

Martin, and Comrade streams during summer of 2003 and summer 2004 (Martin and

Comrade streams were sampled only in summer 2003). The crossing types included were

box culvert and ford crossing. Odom and Reaugaulle streams were marked for road

crossing construction by the military personnel, which allowed sampling to be conducted

before (in 2003), and after (in 2004) road crossings were constructed. The substrate type

in these streams was predominantly sandy, but ranged from gravel to cobble in some

streams. Comrade, Martin, Squirrel branch, and Stagestand streams had box culvert type

crossings with four, five, six, and seven boxes for each, respectively. Arch culvert was

79 constructed on Odom stream, whereas, Ford crossing was constructed on Reaugaulle

Creek.

Sampling Methods

The road crossings built by the Army served as artificial barrier templates for this study. To understand the movement of fish through these road crossings, I selected a reach of 33 m upstream of the road crossing (segment 1), and 33 m downstream of the crossing (segment 2; Figure 3.2). For comparing the fish movement through road crossings and the stream itself, I assigned ‘natural reach’, spanning the distance of the road crossing. Segment 2 was located upstream of the ‘natural reach’, whereas segment 3 on the downstream side.

The streams were sampled over the summers of 2003 and 2004. The data collected during 2003 was used as a baseline comparison. During 2004 the two southern drainages were dropped from sampling because of very low water levels, and beaver dams at the road crossing sites. During each season, the fish were collected, marked with different colors (red, orange, and green) to differentiate between different segments, and released in the respective segments from where they were captured. After the initial marking, the sites were resampled two more times each season. Time interval between recaptures averaged 15 days in the year 2003, and 17 days in 2004; this was dependent on the Army schedule when the land was not being used by military personnel for training.

During the first recapture, the fish that had moved were remarked with a color respective to the segment in which they were recaptured, and unmarked fish were marked.

The fish were sampled using a backpack electroshocker, and seining through each segment. Block nets were placed at either end of the designated segments to prevent 80 escape. Collected fish were placed in MS 222 (tricaine methanesulfonate, 1:45,000

conc.) solution to prevent them from going into shock (Harrel 1992) before they were

identified and marked. The fish were marked sub-cutically by injecting a fluorescent

polymer dye near the dorsal peduncle region (Lotrich and Meredith 1974). Mean current velocity (m/s) was obtained using Marsh-McBirney Flowmate from 3-4 cross-sectional profiles per segment. Depths of crossings were averages of crossings at the upstream and

downstream opening. Depth at a cross-section was averaged from 4 readings, and

maximum depth was taken through a segment along the thalweg. Discharge was

calculated as the product of mean width, mean depth, and mean velocity.

Statistical Analysis

Fish movements through a crossing or a natural reach at a site was assessed as

proportional daily movement (PMD).

where, PMD = M x R-1x D-1

M is the number of fish that moved,

R is total number of recaptures in both segments,

D is the number of days since the first marking.

I tested for relationships among physical characteristics (viz., depth, velocity,

max. depth, and discharge) of a reach and fish movement by using Spearman’s

correlation. A generalized linear model (GLM) with Gaussian error structure and identity

link, and Chi square statistic was conducted on fish movement, as the response variable,

using depth, velocity, discharge, and crossing type as predictor variables (JMP 6;

McCullagh and Nelder 1989). The crossing type was a categorical variable with four

81 levels, viz., arch, ford, box culvert, and natural reach. Fish movement values were square

root transformed to achieve normality. The classical process in GLM is to build all

possible models using a step-down approach, sequentially implementing all the factors

and their interactions. However, because of my preset predictions, I tested for the limited

model with specific parameters.

The relationship between fish assemblage structure and the stream environmental

factors upstream and downstream of the crossings was described using canonical

correspondence analysis (CCA; ter Braak 1986). This direct gradient ordination

approach is a widely used analysis in community ecology (Palmer 1993). Data on

species distributions in sampling units is used to define the multivariate axes, and the

multiple regression analyses used to correlate factor loading scores for samples in species

space with environmental predictor variables. Abundances were square-root transformed,

to dampen the effects of predominant taxa, and rare taxa were down-weighted (McCune

and Mefford 1999).

RESULTS

In this study, I sampled and marked a total of 2,076 fishes (1,004 during the year

2003, and 1,072 during 2004) represented by 28 species and 10 families (Appendix C).

A total of 233 individuals (11.2%) were recaptured during the two year study period, with

81 individuals (8.06%) recaptured during 2003, and 152 individuals (14.17%) during

2004. Of the 233 recaptures, only 40 (or 17.16 % of recaptures) were movements through either the crossing, or the natural reach. The overall proportional daily

82 movement of fish across the culverts and natural reaches ranged from no movement to

highest of 0.057 seen in Reaugaulle Creek during 2004.

Maximum rainfall events per month for the two sampling years was plotted

(Figure 3.3) to include any major rainfall event that would affect stream discharge, which

in turn would affect fish assemblage. The first year of study (June and July of 2003)

experienced higher rainfall (average = 5.2 cm) then the following sampling year (July and

August 2004; average = 2.1 cm). This could explain the greater recapture rate during

2004.

Recapture percentage of marked fish ranged from 3.54% in Comrade Creek to

18.96% in Stagestand Creek (Table 3.1). Recapture percentage was higher in the year

2004 (14.18%), compared to 8% during 2003. Average proportional daily movement in

streams during 2003 was 0.025 with low variance, and was higher (0.032) during 2004 but varied more. Analysis of variance results showed the average current velocity in the upstream segment (0.09 m/s) of the crossing was significantly lower (p = 0.02) compared

to the downstream segment (0.17 m/s) over the two studied years. Average depth was

significantly (p = 0.0002) greater in the upstream segment (0.15 m) compared to the

downstream reach (0.096 m).

No movement across ford crossing was detected. Average movement through

Arch crossing was 0.007. The box culverts showed varying movement values with the 6

box culvert having a high of 0.013 and the 7 box culvert having a low of 0.005. Average

fish movement through natural segment across all streams was the highest (0.016; Figure

3.4). The generalized linear model results indicate a significant inverse relationship

between fish movement and depth (Table 3.2). Velocity and discharge did not have a

83 significant effect on fish movement. Fish movement in the natural reach (0.09) was

greater than that for the arch crossing (0.08), ford crossing (0.05), and box culvert (0.08).

Individual contrasts of crossing type (arch, ford, and box culvert) when compared against

the natural reach separately did not show significant differences. However, contrast of all

crossings put together compared against the natural reach showed significant effect on

movement (Prob>Chisq = 0.02). There was no pre- and post- crossing construction effect on fish movement through Odom, and Reaugaulle streams because there was no movement recorded through the reach marked for construction. This was because of lower percentage of recaptures in these reaches (Table 3.1).

The correlation analysis showed no significant relationship between the environmental parameters (viz., average depth, average velocity, and discharge) and movement (Table 3.3). Species richness and species abundance were not significantly associated with the tested environmental parameters. The first two axes of canonical correspondence analysis explained 75% of the variance in the fish assemblage composition (Figure 3.5). The site scores distributed themselves with the downstream scores along the positive side of Axis 1 and upstream scores along the negative side. The streams in the Calcasieu and Sabine drainages separated on the far left side of Axis 1 from the Red River drainage scores. The downstream section on the plot was characterized by higher current velocity, and discharge values; whereas, the upstream section was characterized by greater stream depth, and higher temperatures. The sunfish species (Centrarchidae) were clustered in the upper left quadrant of the CCA plot in the upstream reach where the stream channel was deeper, had lower discharge, and higher

84 temperatures. The Cyprinids were distributed along the downstream section of the CCA

plot, and the other species distributed over the whole plot.

DISCUSSION

Links between ‘source’ and ‘sink’ populations, colonization of new habitats, and

size and age distribution in a stream can be understood by closely following fish

movement in streams (Hughes and Reynolds 1994, Schlosser 1995, Taylor 1997, Hughes

1998). Stream fishes are a perfect candidate because the longitudinal movement can be

studied along a single dimension. Many recent studies have documented the effects of

road crossings on fish movement (Warren and Pardew 1998, Toepfer et al. 1999,

Schaefer et al. 2003, Angermeier et al. 2004, Wheeler et al. 2005). The results of my study indicate that fish movement was affected by road crossings when compared to the movement through a natural reach.

Movement through the ford crossing (Reaugaulle creek), and box culverts

(Comrade and Martin creeks) was completely absent. Although fish movement through box culverts on Squirrel Branch, and Stagestand creeks was detected, it was less than the movement detected through the natural reach (Figure 3.4). Similar ford and box crossings on southeastern Arkansas streams showed a much higher fish movement

(Warren and Pardew 1998) through the crossings. However, the mean movement across the natural reach in my study was much higher (0.016) than that found in those streams

(0.004) studied by Warren and Pardew (1998). Even though arch culvert did not physically alter the stream substrate composition, the movement was low because of greater length of the crossing itself. This could also be an artifact of the study being 85 conducted immediately after the installation of new crossing. The estimated coefficients of the transformed movement data can be used to make qualitative predictions in order to compare the different crossing types, without back-transformation. Back-transformed estimates would theoretically allow for making true movement predictions across different crossing types. However, because crossing itself can explain only a limited part of fish movement, the use of such coefficients must be limited to qualitative predictions, and that quantitative predictions (of movement) should not be performed from these back-transformed data, whether in management or in scientific research.

The results of this study revealed that a portion of the fish community in the headwater streams I examined exhibited movement across barriers (road crossings). The bidirectional movement of fish observed across the barrier indicates movement along the home range, and not complete out-migration. However, the lower number of recaptures was because fewer individuals were captured, and also possible low fish retention in the sampled reach, which is not unusual (Fausch and Young 1995). Stream geomorphology is typically altered by a scour immediately downstream of the crossing, thus displacing

the fishes (Stock and Schlosser 1991). Current velocity downstream of the culverts was

higher than the upstream reaches. This faster flow could be potential bidirectional

barriers to passages for sunfishes. I found topminnows had the highest number of

movement across crossing (almost same across crossing, and natural reach), which was in

contrast to the findings of Warren and Pardew (1998). Closely followed by the

topminnow, were striped shiner, blacktail shiner, and brown madtom in their movement across crossings. Movement studies on bullhead catfish indicate that they exhibit

‘switching behavior’, which is described as periods of site fidelity (sedentary),

86 punctuated by times of movement (Harcup et al. 1984, Hilderbrand and Kershner 2000,

Knaepkens et al. 2004). A possible explanation to few number of movements could be substantial emigration from the sampling area, but this could be examined with extensive survey over larger lengths of the stream.

Fish movement out of a stream reach typically occurs with increased presence of predators, is dependent on presence of cover, and decreases with higher current velocity and depth (Aparicio and Sostoa 1999 Schaefer 2001). Headwater streams generally exhibit greater flow variability and are influenced by day to day rainfall events (Poff and

Allan 1995). The higher rainfall events during 2003 (Figure 3.3) could have caused higher variability in stream flow leading to the lower recapture percentage during that year, which was reflected in the observed proportional daily movement (Table 3.2).

However, current velocity did not show significant effect on fish movement, in contrast to findings of Warren and Pardew (1998). In conjunction to the study by Aparicio and

Sostoa (1999), the fish movement in my study was inversely correlated to water depth.

The streams in Red River drainage compared to the Sabine and Calcaseiu drainages, showed a difference in fish species composition, and habitat features. The Red

River drainage streams typically are higher in gradient to than those in Sabine and

Calcaseiu drainages (Conner and Suttkus 1986, Williams et al. 2005). Interestingly, this was observed in the CCA plot with Calcaseiu a nd Sabine drainage scores (Appendix D) clustered around the low velocity, and low discharge area. Species composition averaged around 14 for Red River drainage, 21 in Sabine drainage (Martin creek), and 16 in

Calcaseiu drainage (Comrade creek). Sabine shiners and striped shiners were not found

in Sabine and Calcaseiu drainages, as these are associated with swifter flows (Ross 2001,

87 Williams et al. 2005). These drainages, however, had grass pickerel, pirate perch,

warmouth, and yellow bullhead catfish that Red River drainage did not have (Appendix

E, F).

In conclusion, the recapture percentage, and proportional daily movement of fish

in streams in Fort Polk, Louisiana, varied across years and crossing types. Overall fish

movement through the natural reach was higher than the movement detected through

various crossing types studied in the system. The fish species abundance was

significantly affected by stream discharge with an inverse relationship. The importance

of maintaining the natural flow regime is recognized by understanding the interaction

between environmental parameters and fish assemblages (Poff and Allan 1995, Poff et al

1997). Thus anthropogenic alteration to the stream systems by road crossing construction

even on small headwater streams may influence the rates of local immigration or

extinction.

REFERENCES

Angermeier, P. L., A. P. Wheeler, and A. E. Rosenberger. 2004. A conceptual framework for assessing impacts of roads on aquatic biota. Fisheries 29:19-29.

Aparicio, E., and A. de Sostoa. 1999. Pattern of movements of adult Barbus haasi in a small Mediterranean stream. Journal of Fish Biology 55:1086-1095.

Baker, C. O., and F. E. Votapka. 1990. Fish Passage Through Culverts, San Dimas, CA:USDA. FHWA-FL-90-006:1-67, Forest Service Technology and Development Center.

Brown, J. H., and A. Kodric-Brown. 1977. Turnover rates in insular biogeography: effect of immigration on extinction. Ecology 58:445-449.

Campbell, I. C., and T. J. Deoeg. 1989. Impact of timber harvesting and production on streams: a review. Australian Journal of Marine and Freshwater Resources 40.

88 Conner, J. V., and R. D. Suttkus. 1986. Zoogeography of freshwater fishes of the western Gulf Slope of North America. Pages 413-456 in C. H. Hocutt and E. O. Wiley, editors. The Zoogeography of North American Freshwater Fishes. John Wiley and Sons, New York, NY.

Fausch, K. D., C. E. Torgersen, C. V. Baxter, and H. W. Li. 2002. Landscapes to riverscapes: bridging the gap between research and conservation of stream fishes. Bioscience 52:483-498.

Fausch, K. D., and M. K. Young. 1995. Evolutionary significant units and movement of resident stream fishes: a cautionary tale. Pages 360-370 in J. L. Nielsen, editor. American Fisheries Society, Symposium 17, Bethesda, Maryland.

Gerking, S. D. 1959. The restricted movement of fish populations. Biological Review 34:221-242.

Gilliam, J. F., and D. F. Fraser. 2001. Movement in corridors enhancement by predation threat, disturbance, and habitat structures. Ecology 82:258-273.

Gowan, C., and K. D. Fausch. 2002. Why do foraging stream salmonids move during summer? Environmental Biology of Fishes 64:139-153.

Gowan, C., M. K. Young, K. D. Fausch, and S. C. Riley. 1994. Restricted movement in resident stream salmonids: paradigm lost? Canadian Journal of Fisheries and Aquatic Sciences 51:2626-2637.

Grossman, G. D., J. F. Dowd, and M. Crawford. 1990. Assemblage stability in stream fishes: a review. Environmental Management 14:661-671.

Harcup, M. F., R. Williams, , and D. M. Ellis. 1984. Movements of brown trout, Salmo trutta L., in the River Gwyddon, South Wales. Journal of Fish Biology 24:415– 426.

Harrell, L. 1992. Handling euthanasia in production facilities. Pages 129 in D. O. Schaeffer, K. M. Kleinow, and L. Krulisch, editors. The care and use of amphibians, reptiles and fish in research. Scientists Center for Animal Welfare, Bethesda, Md.

Harvey, B. C. 1991. Interactions among stream fishes: predator-induced habitat shifts and larval survival. Oecologia 87:29-39.

Harvey, B. C., R. J. Nakamoto, and L. J. White. 1999. Influence of large woody debris and a bankfull flood on movement of adult resident coastal cutthroat trout, Oncorhynchus clarki, during fall and winter. Canadian Journal of Fisheries and Aquatic Sciences 56:2161-2166.

89 Hilderbrand, R. H., and J.L. Kershner. 2000. Movement patterns of stream-resident cutthroat trout in Beaver Creek, Idaho-Utah. Transactions of the American Fisheries Society 129:1160-1170.

Hughes, N. F. 1998. Use of whole-stream patterns of age segregation to infer the interannual movements of stream salmonids: a demonstration with Arctic grayling in an interior Alaskan stream. Transactions of the American Fisheries Society 127:1067–1071.

Hughes, N. F., and J. B. Reynolds. 1994. Why do Arctic grayling (Thymallus arcticus) get bigger as you go upstream? Canadian Journal of Fisheries and Aquatic Sciences 51:2154-2163.

Ims, R. A., and D. O. Hjermann. 2001. Condition-dependent dispersal. Pages 600-621 in J. Clobert, E. Danchin, A. A. Dhondt, and D. J. Nichols, editors. Dispersal. Oxford University Press, New York.

Kaller, M. D., and W. E. Kelso. 2006. Effects of a small-scale clearing on habitat and macroinvertebrates of a southwestern Louisiana coastal bottomland stream. The Southwestern Naturalist 51:143-151.

Knaepkens, G., L. Bervoets, E. Verheyen, and M. Eens. 2004. Relationship between population size and genetic diversity in endangered populations of the European bullhead (Cottus gobio): implications for conservation. Biological Conservation 115:403–410.

Lammert, M., and J. D. Allan. 1999. Assessing biotic integrity of streams: effects of scale in measuring the influence of land use/cover and habitat structure on fish and macroinvertebrates. Environmental Management 23:257-270.

Larson, G. L., R. L. Hoffman, and S. E. Moore. 2002. Observations of the distributions of five fish species in a small Appalachian stream. Transactions of the American Fisheries Society 131:791-796.

Lonzarich, D. G., M. R. Lonzarich, and M. L. Warren Jr. 2000. Effects of riffle length on the short-term movement of fishes among stream pools. Canadian Journal of Fisheries and Aquatic Sciences 57:1508–1514.

Lotrich, B. A., and W.H. Meredith. 1974. A technique and the effectiveness of various acrylic colors for subcutaneous marking of fish. Transactions of the American Fisheries Society 103:140-142.

Macdonald, J. S., MacIsaac, E.A., and Herunter, H. 2003. The effects of variable- retention riparian buffer zones on water temperatures in small headwater streams in sub-boreal forest ecosystems of British Columbia. Canadian Journal of Forest Research 33:1371–1382.

90 Matthews, W. J. 1986. Fish faunal structure in an Ozark stream: stability, persistence and a catastrpohic flood. Copeia 1986:388-397.

Matthews, W. J., R. C. Cashner, and F. P. Gelwick. 1988. Stability and persistence of fish faunas and assemblages in three midwestern streams. Copeia 1988:945-955.

McCullagh, P., and J. A. Nelder. 1989. Generalized Linear Models, 2nd edition. Chapman and Hall, London.

McCune, B., and M. J. Mefford. 1999. Multivariate analysis of ecological data. Version 4. MjM Software Design, Gleneden Beach, OR.

Meffe, G. K., and A. L. Sheldon. 1990. Post-defaunation recovery of fish assemblage in southeastern blackwater streams. Ecology 71:657-667.

Northcote, T. J. 1997. Potamodromy in Salmonidae - living and moving in the fast lande. North American Journal of Fisheries Management 17:1029-1045.

Palmer, M. W. 1993. Putting things in even better order: the advantages of canonical correspondence analysis. Ecology 74:2215-2230.

Pecarsky, B. L. 1983. Biotic interactions or abiotic limitations? A model of lotic community structure. in T. D. Fontaine III and S. M. Bartell, editors. Dynamics of lotic ecosystems. Ann Arbor Science, Ann Arbor, Michigan.

Peterson, J. T., and P. B. Bayley. 1993. Colonization rates of fishes in experimentally defaunated warmwater streams. Transactions of the American Fisheries Society 122:199-207.

Poff, N. L., J. D. Allan, M. B. Bain, J. R. Karr, K. L. Prestegaard, B. D. Richter, R. E. Sparks, and J. C. Stromberg. 1997. The Natural Flow Regime - A paradigm for river conservation and restoration. BioScience 47:769-784.

Poff, N. L., and J.D. Allan. 1995. Functional organization in stream fish assemblages in relation to hydrologic variability. Ecology 76:606-627.

Pringle, C. M., M. C. Freeman, and B. J. Freeman. 2000. Regional effects of hydraulic alterations on riverine macrobiota in the new world: tropical-temperate comparisons. Bioscience 50:807-823.

Railsback, S. F., R. H. Lamberson, B. C. Harvey, and W. E. Duffy. 1999. Movement rules for individual-based models of stream fish. Ecological Modelling 123:73-89.

Ross, S. T. 2001. Inland fishes of Mississippi. University Press of Mississippi, Jackson.

91 Schaefer, J. 2001. Riffles as barriers to interpool movement by three cyprinids (Notropis boops, Campostoma anomalum and Cyprinella venusta). Freshwater Biology 46:379-388.

Schaefer, J. F., E. Marsh-Matthews, D. E. Spooner, K. B. Gido, and W. J. Matthews. 2003. Effects of barriers and thermal refugia on local movement of the threatened leopard darter, Percina pantherina. Environmental Biology of Fishes 66:391-400.

Schlosser, I. J. 1995. Dispersal, boundary processes, and trophic level interactions in streams adjacent to beaver ponds. Ecology 76:908-925.

Schlosser, I. J., and P. L. Angermier. 1995. Spatial variation and demographic processes of lotic fishes: conceptual models, imperical evidence, and implications for conservation. Pages 392-401 in J. L. Nielsen, editor. Evolution in the aquatic ecosystem: defining unique units in population conservation.

Shaw, E. A., and J.S. Richardson. 2001. Effects of fine inorganic sediment on stream invertebrate assemblages and rainbow trout (Oncorhynchus mykiss) growth and survival: implications of exposure duration. Canadian Journal of Fisheries and Aquatic Sciences 58:2213-2221.

Smithson, E. B., and C. E. Johnston. 1999. Movement patterns of stream fishes in a Ouachita Highlands stream: an examination of the restricted movement paradigm. Transactions of the American Fisheries Society 128:847–853.

Sousa, W. P. 1984. The role of disturbance in natural communities. Annual Review of Ecology and Systematics 15:335-391.

Stock, J. D., and I. J. Schlosser. 1991. Short-term effects of a catastrophic beaver dam collapse on a stream fish community. Environmental Biology of Fishes 31:123– 129.

Taylor, C. M. 1997. Fish species richness and incidence patterns in isolated and connected stream pools: effects of pool volume and spatial position. Oecologia 110:560–566.

ter Braak, C. J. F. 1986. Canonical correspondence analysis: a new eigenvector technique for multivariate direct gradient analysis. Ecology 67:1167-1179.

Toepfer, C. S., W. L. Fisher, and J. A. Haubelt. 1999. Swimming performance of the threatened leopard darter in relation to road culverts. Transactions of the American Fisheries Society 128:155-161.

Townsend, C. R. 1989. The patch dynamics concept of stream community ecology. Journal of the North American Benthological Society, 8:36-50.

92 Warren Jr, M. L., P. L. Angermier, B. M. Burr, and W. R. Haag. 1997. Declines of a diverse fish fauna: patterns of imperilment and protection of the southeastern United States. Pages 105-164 in G. W. Benz and D. E. Collins, editors. Aquatic fauna in peril: the southeastern perspective. Lenz Design in Communications, Decatur, GA.

Warren Jr, M. L., and M. G. Pardew. 1998. Road crossings as barriers to small-stream fish movement. Transactions of the American Fisheries Society 127:637-644.

Wheeler, A. P., P. L. Angermeier, and A. E. Rosenberger. 2005. Impacts of new highways and subsequent landscape urbanization on stream habitat and biota. Reviews in Fisheries Science 13:141-164.

Wiens, J. A. 2001. The landscape contents of dispersal. Pages 97-109 in J. Clobert, E. Danchin, A. A. Dhondt\, and J. D. Nichols, editors. Dispersal. Oxford University Press, New York.

Wiens, J. A., J. T. Rotenberry, and B. van Horne. 1986. A lesson in the limitation of field experiments: shrubsteppe birds and habitat alteration. Ecology 67:365-376.

Williams, L. R., T. H. Bonner, J. D. Hudson III, M. G. Williams, T. R. Leavy, and C. S. Williams. 2005. Interactive efects of environmental variability and military training on stream biota of three headwater drainages in Western Louisiana. Transactions of the American Fisheries Society 134:192-206.

Williams, L. R., C. M. Taylor, M. L. Warren Jr, and J. A. Clingenpeel. 2002. Large-scale effects of timber harvesting on stream systems in the Ouachita Mountains, Arkansas, USA. Environmental Management 29:76-87.

93

TABLES

Percentage of Recapture Proportional Daily Movement Streams 2003 2004 2003 2004 Odom 7.80 8.11 0.024 0.026 Reaugaulle 7.50 12.34 – 0.057 Squirrel Branch 10.84 14.94 0.024 0.019 Stagestand 10.00 18.96 0.027 0.026 Comrade 3.54 * – * Martin 8.78 * – * Average (S.E.) 8.07 (0.88) 14.18 (1.76) 0.025 (0.00073) 0.032 (0.0065) * sampling not conducted because of low water levels – no movement detected

Table 3.1 Recapture percentage, and proportional daily movement of fish in the studied streams in Fort Polk, Louisiana, during the summers of 2003 and 2004.

94

Term Estimate Standard Error Chi Square Prob>Chisq Intercept 0.091 0.031 7.441 0.006 Depth -0.202 0.097 4.098 0.042 Velocity -0.096 0.201 0.227 0.633 Discharge -0.274 0.408 0.447 0.503 Type[arch] -0.0044 0.031 0.021 0.885 Type[box] -0.0071 0.022 0.105 0.745 Type[ford] -0.035 0.031 1.313 0.251

Table 3.2 Summary of Chi Square results on generalized linear model (GLM) analysis for the variables depth, velocity, discharge, and crossing type (categorical variable; viz., arch, box, ford, and natural reach) on fish movement.

95

Average Depth Average Velocity Max. Depth Discharge Species Richness -0.11 -0.24 -0.05 -0.27 Species Abundance -0.28 -0.08 -0.28 -0.31* Fish Movement -0.26 -0.03 -0.08 -0.25 * significant with p-value = 0.05

Table 3.3 Correlation (Spearman’s) of environmental parameters of the sampled streams, with the fish species richness, abundance, and movement.

96

FIGURES

Red River C Drainage A D B

N Sabine River

Drainage E W E

F Calcasieu River Drainage S

0 5 10 15 Kilometers

Figure 3.1 Mark and recapture conducted on fish over road crossing across the streams belonging to the Red River drainage – A- Reaugaulle, B- Stagestand, C- Odom, and D- Squirrel branch; the Sabine River drainage – E-Martin; and Calcaseiu River drainage – F- Comrade. Crossing type on Stagestand, Squirrel branch, Martin, and Comrade streams was box culvert. Reaugaulle stream had ford type crossing, and Odom stream had arch type crossing over it.

97

Transect 1

33 m Red

Road-crossing

Transect 2

33 m Orange

Natural reach, the width of the road-crossing above

33 m Green

Figure 3.2 Description of the fish marking strategy across the road crossing, and the natural reach along the stream, with stream flowing from top to the bottom of the page.

98

Temperature

30 Rainfall 12

2003

20 8

10 4

0 0

Sampling period

30 12 2004 ) C )

º ( m

20 8 c e

( r

l u l t a a f r e n i p 10 4 a

m R e

T

0 0 n b r r y l g p t v c a e a p a un Ju u c o e J F M A M J A Se O N D

Figure 3.3 Maximum rainfall event for each month, and air temperature data for the Fort Polk region of Louisiana. (Weather data obtained from archives of www.weatherunderground.com)

99

0.025

t 0.02

0.015

0.01

Proportional Daily Movemen 0.005

0 Natural Ford Arc h 4 Boxes______5 Boxes 6 Boxes 7 Boxes Box Culverts with number of compartments

Figure 3.4 Comparison of mean proportional daily movement of fish across different crossing types showing greater movement in the natural reach. The ANOVA results show no significant difference in movement between the crossing types compared to the natural reach (df = 6, F = 0.09, p-value = 0.51, N = 40).

100

2 Upstream Downstream

Rain Velocity 1 Temperature Depth

Centrarchidae (upper left quadrant) 0

D-Max Axis2 (18 %) Length Discharge -1

Drainages

Calcasieu Cyprinidae (lower half) Width Sabine Red River -2 -1.5 -1.0 -0.5 0.0 0.5 1.0 1.5 2.0 Axis 1 (42 %)

Figure 3.5 Canonical correspondence analysis on relationship between the fish species, and the environmental factors they were exposed to, in the three drainages (viz., Red River, Sabine, and Calcasieu) in Fort Polk, Louisiana. The first two axes explained 60 % of the variance significantly (p-value = 0.001).

101

CHAPTER 4

Effects of fish and invertebrate predators on benthic invertebrates in

headwater steams of west-central Louisiana

INTRODUCTION

An aquatic ecosystem is generally supported by primary producers, but headwater stream systems are driven by allochthonous input (Vannote et al. 1980, Polis and Strong

1996, Sabo and Power 2002). Allochthonous input from surrounding riparian forest reduces the impact of drift-feeding fish on the abundance of benthic prey, and thereby increases abundance of all trophic levels (Richardson 1991, Dhal and Greenberg 1996,

Nisbet et al. 1997, Wallace et al. 1997, Nakano et al. 1999). Of the many variables that influence the aquatic invertebrate assemblage structures, predator-prey interaction has the potential to play an important role in several ways (Connell 1975, Glasser 1979, Sih

1982). The predator-prey relationship (between fish and macroinvertebrates) is complex, in that if the predator is not selective, it might cause loss of diversity and local extinction of relatively rare species, especially seen in short food chain systems (Peckarsky 1984).

In contrast, prey species diversity may increase if secondary predators selectively remove keystone primary predators (dominant competitors), thus allowing fugitive species

(inferior competitors) to colonize the habitats.

102 Predators can drastically reduce the density of individuals by indiscriminate feeding in proportions in which they are found in habitats (Hunter and Price 1992,

Nystrom et al. 2003). Similarly, absence of predators could cause an increase in macroinvertebrate populations. Even after comparing theoretical and empirical studies there seems to be no conclusion whether ecosystem function in streams is controlled by resource availability (bottom-up) or predation (top-down; e.g. McQueen et al. 1989, Polis and Winemiller 1996, Williams et al. 2003). Possible answers to this question lie in investigating predator avoidance mechanism, risks of predation-associated feeding (Allan

1982), lack of competitive dominants (Allan 1982), rates of macroinvertebrates immigration (Flecker 1984), and differences in sampling strategies and habitats of individual studies (Dhal 1998). Thus prey population structure is dependent on the ecological characteristics of organisms (predator-prey relationship), and the environments they inhabit (Power 1992).

Although a prey community in a natural system faces multiple predator species simultaneously, most studies look at the effect of one predator species on the prey assemblage (McIntosh and Peckarsky 1999, Eklov and Van Kooten 2001). To understand the dynamics of an invertebrate assemblage in a natural system, it is essential to understand the responses of prey to multiple predators, and the mechanisms determining the outcome of interactions of multiple predators, both fish and invertebrate

(McIntosh and Peckarsky 1999, Eklov and Van Kooten 2001). A prey’s risk of predation can be additive (Wilbur and Fauth 1990, Sokol-Hessner and Schmitz 2002), or non- additive (Soluk and Collins 1988, Sih et al. 1998). The degree of predation risk on prey depends on the availability of microhabitats, prey behavior, and the life history strategy

103 of prey species (see Peckarsky 1979, Vance-Chalcraft et al. 2004). Thus, experimental

studies involving manipulating the predator species composition in enclosure settings can help in evaluating the effects of predation and competition on distribution patterns

(Peckarsky 1979).

Fish predation, in general, has a direct effect on invertebrates, and indirect effect on primary producers (Kerfoot and Sih 1987). Studies on lotic systems involving higher trophic levels have found strong such interactions (Bronmark et al 1997). Pressure of predation on smaller invertebrates by predatory invertebrates is temporarily released if fish consume these predators. In fish-free settings, predatory effect is still present as the role is taken over by predatory invertebrates (Dahl and Greenberg 1997). In addition to the direct predatory effects on invertebrates, fish can also compete for periphyton resources (Dahl and Greenberg 1996, Dahl 1998). The indirect effects of fish predators on primary producers through trophic cascades are because of both direct consumption of grazer invertebrates, and altered prey behavior to presence of fish (Power 1990,

Resenfeld 1997, Dahl 1998). Thus, studying feeding habits of fish species can add useful information to the variation macroinvertebrate assemblage. Habitat complexity, omnivory, and prey defense mechanisms are also some of the factors found to affect the strength of trophic cascades (Power 1992, Diehl 1993). The strength of interaction between three trophic levels (viz., fish, grazers, and algae) is altered by the efficiency of predation on prey populations. In this study, I used a two part experiments with exclosures to determine the macroinvertebrate predator-prey relationship in absence of top predator, fish. The second part of the experiment was using enclosures to examine the invertebrate assemblage structure in response to varying predatory pressure in the

104 presence of different fish feeding guilds. With this experimental study, I specifically assess: 1) the successional pattern of invertebrate community in absence of fish predation, and 2) if the presence of various fish feeding guilds have different effects on macroinvertebrate assemblage structure.

METHODS

Study Area

The experiments were conducted on three streams located at Fort Polk Army Base and the USDA Kisatchie National Forest in Vernon Parish, west central Louisiana during summer 2003 (Figure 4.1). Fort Polk is a U.S. military training base established in

1940’s and encompasses 80,000 hectares in Vernon Parish. The streams are characteristic to the low gradient drainages of the western Gulf Slope, commonly known as the Gulf Coastal Plain (Conner and Suttkus 1986). Birds, Drakes, and Whisky Chitto

Streams belong to the Calcaseiu River drainage and empty further south in the Gulf of

Mexico. The low gradient streams in this drainage are characterized by loamy to clay soils (Martin et al. 1990). The forests are dominated by longleaf pine (Pinus palustris), interspersed with hardwoods (e.g., bald cypress Taxodium distichum, white oak Quercus alba, water oak Quercus nigra, and magnolia Magnolia randiflora) more common along the lowland reaches (Williams et al. 2005). Historically, prior to human prescribed fire intervention, the longleaf pine fire climax community was subject to seasonal lightning strikes (Bridges and Orzell 1989). Following the prescribed fire rotations in Fort Polk, the army personnel have maintained mature longleaf pine community over the past 50

105 years. This understory habitat devoid of vegetation is crucial for the existing endangered

Red-cockaded Woodpecker Picoides borealis population on the army base.

Design and Methods

Experiment 1: Effect of fish absence on macroinvertebrate assemblage structure

To examine the successional dynamics of macroinvertebrates in the absence of top predators, fish, I conducted an exclosure experiment on three streams during summer of 2003. The exclosure cages were made of heavy mosquito nets (< 1 mm mesh size) secured with wooden stakes along four corners (1 x 1 x 1 m) in the streambed. The bottom edge of the net was buried 10 cm under the stream bed substrate to prevent entry of fish, or bidirectional movement of invertebrates (> 1 cm). A total of 42 exclosures were installed on three streams (two replicates per treatment, with 14 exclosures on each stream), with a minimum gap of three meters between each. Invertebrates were sampled from the exclosures on day 0, 2, 5, 10, 15, 20, & 30, and sequentially the respective exclosures were dismantled one by one after each sampling day. Sampling was conducted with a D-net to collect suspended and perched invertebrates, while a Surber sampler was employed to collect invertebrates buried under the bed material. Sampled invertebrates were preserved in 70% ethanol in the field, and sorted and identified in the lab. Most invertebrates were identified to genus level when possible; to allow identification of feeding and habit guilds based on Merritt and Cummins (1996) classification. Current velocity (m/s), and water depth (m) in the exclosures was measured on the sampling days using a Marsh-McBirney Flowmate (Appendix G).

106

Experiment 2: Effects of different fish feeding guilds on invertebrate assemblage

To separate the effects of various fish feeding guilds on invertebrate prey, I conducted an enclosure experiment on the same three streams during summer of 2003.

The enclosures were installed with the same dimensions and precision as in the first experiment. A total of 36 enclosures were installed (two replicates, with 12 on each stream), and sampling was conducted on the first day, and 10 days later to allow sufficient time for predator-prey interactios. Six different kinds enclosures (two replicates) were installed each containing a different feeding guild of fish. 1) Algivore enclosure had two adult Black Redhorse; 2) Insectivore enclosure had five adult blacktail shiners, and five adult redfin shiners; 3) Piscivore enclosure had two medium sized largemouth bass; 4) Mixed enclosure had three adult blacktail shiners, three adult redfin shiners, and two adult longear sunfish; 5) Open enclosure – four stakes with no enclosure material around; and 6) Control enclosure, which was sampled at the beginning of the experiment. Environmental parameters and invertebrate data were collected and processed at the end of the experiment in similar way explained for the previous experiment (Appendix G).

Statistical Analyses

For both the experiments, the macroinvertebrates were categorized based on their feeding guilds according to Merritt and Cummins (1996). Least square regression analyses were performed on family richness and abundance (both values were square- root transformed to achieve normality), using the common predictor variables of current velocity, and water depth. This analysis was conducted in both the experiments to detect 107 any interaction as a result of water depth and current velocity, with the respective

treatments (Day, in case of experiment 1, and Guild, in case of experiment 2). In case of

experiment 1, the categorical variable used as a predictor was Days, with 7 levels (viz.,

days 0, 2, 5, 10, 15, 20, and 30). Separate nested design ANOVAs were performed on

the macroinvertebrate feeding guilds, with Day nested within the Stream block, to

determine the successional dynamics in a top-predator (fish) free environment.

For experiment 2, the categorical variable used was Guilds, with 6 levels (viz,

Control – initial sampling, Open – no enclosure around supporting stakes, Algivore,

Insectivore, Piscivore, and Mix). A generalized linear model (GLM) with Gaussian error

structure and identity link, and Chi-quare statistic was conducted on macroinvertebrate

abundance, with Fish Guild type and the macroinvertebrate feeding guilds as predictor

variables (JMP 6; McCullagh and Nelder 1989). The Fish Guild was a categorical

variable with 6 levels (Algivore, Invertivore, Mix, Piscivore, Open, and Control -no fish).

The macroinvertebrate feeding guilds was a categorical variable with 6 levels (Collector,

Filterer, Piercer, Predator, Scraper, and Shredder). I tested for main effects of Fish Guild and macroinvertebrate guilds, and their interactions.

RESULTS

Experiment 1: Effect of fish absence on macroinvertebrate assemblage structure

A total of 5,832 invertebrates were collected and identified, averaging 1,194 ±

455 individuals, representing 33 ± 1 families per stream (Appendix H). The family richness dropped from first day to Day 5, where it stabilized until Day 20 (no significant difference in analysis; Figure 4.2). Day 30 was characterized by an increase in the 108 species richness. Family abundance showed a similar downward trend from the initial

sampling, with Day 5 having the lowest abundance; although there was no statistical significance (Figure 4.3). Least square regression analysis of the environmental variables

(current velocity, and depth), and Day (categorical variable with 7 levels), showed a

significant effect (< 0.001) of current velocity on family richness (Table 4.1, 4.2).

Although there was no significant interaction effect, specific interaction of current

velocity and Day 5 had a positive significant effect on family richness. The nested design

ANOVAs run on individual feeding guilds, with Day nested within stream block showed

a significant difference in the predator guild (Table 4.3, Figure 4.4). Three feeding guilds

(Filterer, Predator, and Scraper) showed a significant difference among the stream block.

The top five abundant taxa of the macroinvertebrate assemblage were represented

by the taxa: Chironomidae, Caeniidae, Elmidae, Heptageniidae, and Oligochaeta

occurring in exclosures from the selected streams (Figure 4.5). The abundance

comparisons of these selected taxa from the whole assemblage showed the following

trends. Percent abundance of Caeniidae increased from Day 0 to Day 10, after started

dropping until Day 30. Chironomidae showed inverse relationship to Caeniidae, in that

their percent abundance dropped from Day 0 to Day 10, after which it started increasing.

Heptageniidae showed a marked increase in abundance on Day 10. Oligochaeta

abundance slowly increased from Day 0 to highest on Day 30 (Figure 4.5).

Experiment 2: Effects of different fish feeding guilds on invertebrate assemblage

Total number of invertebrates sampled in this experiment was 4,250 with 1,416 ±

180 per stream, and was represented by 30 ± 2 families (Appendix I). The family

109 richness did not significantly vary across different enclosure types even after 10 days of

exposure to specific fish guilds of fish (Figure 4.6). Family abundances had higher

values in the Open enclosure compared to other enclosures that had different fish enclosed (Figure 4.7). Insectivore enclosure had the lowest family richness, and lowest family abundance values. The Piscivore enclosure had lower family abundance compared to the Open enclosure, but was least affected compared to the other enclosure kinds. Least square regression analysis on the environmental parameters (current velocity, and depth), and Fish Guild (categorical variable with 6 levels) showed a significant effect of current velocity (p = 0.01) on family richness. There was no significant effect of interactions (Table 4.4). Family abundance was significantly affected by depth (p = 0.04; Table 4.5). The Fish Guild enclosures did not have a significant effect either on family richness, or family abundance.

The GLM results showed that Fish Guild did not have significant effect, but specific macroinvertebrate feeding guilds showed a significant response in their abundances, and a significant interaction was observed (Table 4.6). Control (from Fish

Guild), and Collector (from macroinvertebrate feeding guild) were represented by the

intercept, which showed a significant difference from the other effects. Within the Fish

Guild enclosures, the Open enclosure was significantly different. Within the

macroinvertebrate category, all the feeding guilds showed a significant effect (Table 4.6.

Specific interaction effects were detected in the Insectivore*Filterer, and the

Open*Scraper parameter estimates.

Top five abundant taxa of the macroinvertebrates collected in Experiment 2 were

Caeniidae, Chironomidae, Elmidae, Heptageniidae, and Hydropsychidae (Figure 4.8).

110 The abundance comparisons of these selected taxa from the whole assemblage showed the following trends. Oligochaeta was not represented in higher numbers, probably because of predation by fish in the enclosures. Caeniidae were found in greater numbers in the Algivore enclosure, and lowest in the Open enclosure. Chironomidae were in higher proportion compared to the other taxa, and did not show specific trends in relation to different enclosure types. Elmidae, Heptageniidae, and Hydropsychidae were in lower proportion in the enclosure types, and did not show any distinctive trends.

DISCUSSION

In the absence of a top predator (fish), the macroinvertebrates did not show a significant increase either in taxa richness, or taxa abundance (experiment 1). However, when decomposed or analysed by specific feeding guilds of macroinvertebrates, significant trends were observed. Successional trends detected in the macroinvertebrate feeding guilds in the absence of top predator, fish, were as follows. The predatory guild significantly increased in abundance by the end of the experiment (30 days later).

Scraper guild showed a significant downward trend (P=0.05, one way ANOVA). In experiment 2, the effects of specific Fish Guilds on macroinvertebrates are as follows.

Insectivore enclosure had lowest macroinvertebrate abundance suggesting a strong effect of this fish guild compared to other guilds. Oligochaete abundance was very low in the

Fish Guild enclosures, suggesting a strong fish predation on this taxa. These findings give a better understanding of the macroinvertebrate assemblage dynamics, and the predator-prey interaction.

111 In stream systems experiencing periodic disturbances in the form of fluctuating

discharges, colonization tactics play an important role in succession of biotic

communities (Diamond 1975, Robinson and Edgemon 1988, Drake et al. 1993). Because disturbance is an integral component of headwater streams, varying discharges can lead to periodic exclusion of fish predators from these reaches. Here, the colonization history becomes a critical factor in determining the development of present communities. The first experiment focuses on the dynamics of macroinvertebrates, in the absence of fish, in

the form of competitive interaction with other taxa. This competitive interaction can

influence the abundance and species composition in successional dynamics of the final

community (Gilpen and Case 1976, Drake 1991). In contrast to the findings of Flecker

(1992), and Williams et al. (2003), I did not detect a change in macroinvertebrate density, in response to removal of fish. This was because the exclosures hindered immigration/emmigration, at least to majority of the taxa > 1mm. Hydropsychids are known to aggressively compete for space and inhibit colonization potential of other taxa

(Englund 1991, 1993), but they did not occur in the top five abundant taxa in the current study. This was probably because the possible restriction in the current velocity. The lower current velocity, coupled with the absence of fish could have led to increase in

Oligochaete abundance. Similarly, Chironomids also showed an increase in their abundance over time, although in the initial few days their numbers had dropped. The competitive advantage of initial colonists can reduce the diversity of other taxa (Moyle and Light 1996). Although the predatory guild in my study showed a significant increase in numbers, none of the other guilds decreased in abundance, with the exception of shredder guild. This could again be an artifact of lack of food substrate, arising from leaf

112 material being effectively blocked out of the exclosure. In general, I did not detect an inhibitory effect of predators on the other taxa abundances. However, it is also possible that resources did not become limited in the interval of the experimental study.

Fish affect macroinvertebrate abundance by direct predation (Williams et al.

2003) and indirectly by competing for resources (Flecker 1992). The second experiment attempts to assess the effect of direct predation of individual feeding guild of fishes, and the indirect resource competing forces on the macroinvertebrate assemblage. The Open treatment had the highest taxa richness and abundance when compared to the treatments with various fish guilds enclosed. The macroinvertebrate abundance in the Insectivore enclosure was the lowest, but not significantly different from other treatments. Within the Insectivore enclosure, the filterer macroinvertebrate feeding guild was found to be significantly lower. Chironomids are typically the first choice of fish (Williams et al.

2003), but they maintained the highest proportion across treatments. Although stomachs of the experimental fish were not examined, it could be speculated that the Chironomids might have eluded the fish effectively, and thus maintained their higher numbers in the samples. Their higher numbers could be because of my digging deeper into the substrate than the fish. Indirect effects of grazing fishes (Algivores) on macroinvertebrate densities can be greater than direct effects by predation by fishes (Flecker 1992). The indirect competition effects of Algivorous fish on the scraper macroinvertebrate community has long been known (see Power 2000). The results of GLM (Table 4.6) of my study did not detect an interaction between the Algivore treatment and any of the macroinvertebrate feeding guild abundances. This lack of resource competition effect

113 could be a result of limited algal production in the canopy covered headwater reaches of the study sites.

In many aquatic systems, the survival of an individual depends on its ability to obtain refuge (Peckarsky and Penton 1989, Kirk and Smock 2000). Similarly, the prey species typically emigrates or drifts in presence of predators (Peckarsky and Penton 1989,

Rader and McArthur 1995). This combination of refuge availability, and prey active- movement and drift, reduces the risk of predation. However, in my experiment the prey could not preferably move away from the invertebrate predators because of the enclosure settings. Thus the lack of decrease in the overall abundance in my experiment suggests either successful predator avoidance by prey, or anti-predator defense mechanism at work. Specific macroinvertebrate predator-prey interaction was detected by the inverse relationship in the predator and shredder guilds.

The effects of invertebrate predators have been suggested to have a stronger effect on prey (Wooster 1994), but my results indicate fish to have greater effect (from comparing results of my two experiments). Different feeding guilds of fish had varying effects on macroinvertebrate assemblages. The advantages of this type of study compared to other experimental studies are: a) response of entire macroinvertebrate assemblage in absence of fish (experiment 1); b) experiments were conducted in field, as opposed to lab settings, and c) macroinvertebrates and fish >1mm were effectively excluded from both the experimental settings. One disadvantage of this field setting was that I could not control for differences in the initial abundances of macroinvertebrates across exclosures/enclosures in either of the experiments. This discrepancy could contribute to the differences in results obtained. Despite the confounding effect of

114 variability in initial macroinvertebrate assemblage, I was able to detect successional

trends in the feeding guilds of macroinvertebrates.

REFERENCES

Allan, J. D. 1982. The effects of reduction in trout density on the inverebrate communtiy of a mountain stream. Ecology 63:1444-1455.

Bridges, E. L., and S. L. Orzell. 1989. Longleaf pine communities of the Western Gulf Coastal Plain. Natural Areas Journal 9:246-262.

Bronmark, C., J. Dahl, and L. Greenberg. 1997. Complex trophic interactions in freshwater benthic food chains. Pages 55-88 in B. Streit, T. Stadler, and C.M. Lively, editors. Ecology and Evolution of freshwater animals. Birkhaeuser- Verlag, Basel, Switzerland.

Connell, J. H. 1975. Some mechanisms producing structure in natural communities: a model and evidence from field experiments. . Pages 460-490 in M. L. Cody and J. M. Diamond, editors. Ecology and Evolutions of Communities. Belknap, Cambridge.

Dahl, J. 1998. Effects of a benthivorous and a drift-feeding fish on a benthic stream assemblage. Oecologia 116:426-432.

Dahl, J., and L. Greenberg. 1996. Impact on Stream benthic prey by benthic vs. drift feeding predators: a meta-analysis. Oikos 77:177-181.

Dahl, J., and L. Greenberg. 1997. Foraging rates of a vertebrate and a inverebrate predator in stream enclosures. Oikos 78:459-466.

Diamond , J. M. 1975. Assembly of species communities. In Cody , M. > & J. M . Diamond (eds), Ecology and Evolution of Communities. Pages 342-444 in Harvard University Press, editor., Cambridge, Massechuttes.

Diehl, S. 1993. Effects of habitat structure on resource availability, diet and growth of benthivorous perch, Perca fluviatilis. Oikos 67:403-414.

Drake, J. A. 1991. Community-assembly mechanics and the structure of an experimental speciaes ensemble. American Naturalist 137:1-26.

Drake, J. A., G. H. Flum, T. Witteman, A. M. Voskuil, C. Hoylman, D. A. Creson, G. R. Kenny, C. S. Huxel, and Larue & J.R. Duncan. 1993. The construction and assembly of an ecological landscape. animal ecology 62:117-130. 115 Eklov, P., and T. Van Kooten. 2001. Facilitation among piscivorous predators: effects of prey habitat use. . Ecology 89:2486-2494.

Englund, G. 1991. Asymmetric resource competition in a filter-feeding stream insect. Hydropsyche siltalai: Trichoptera Fresh water Biology:26.

Englund, G. 1993. Interactions in a lake oulet stream community: direct and indirect effects of net-spinning caddis larvae. Oikos 66:431-438.

Flecker, A. S. 1984. The effect of predation and detritus on the structure of a stream insect community: a field test. Oecologia 64:300-305.

Flecker, A. S. 1992. Fish Trophic guilds and the structure of a tropical stream: weak direct vs. strong indirect effects. Ecology 73:927-940.

Gilpin, M. E., and T. J. Case. 1976. Multiple domains of attraction in competition commuinities. Nature 261:40-42.

Kerfoot, E. C., and A. Sih, editors. 1987. Predation: Direct and indirect impacts on aquatic communities. University Press of New England, Hanover, London.

Martin, G., C. L. Butler, E. Scott, and J. E. Lyles. 1990. Soil survey of Natchitoches Parish, Louisiana. United States Department of Agriculture, Soil Conservation Service.

McCullagh, P., and J. A. Nelder. 1989. Generalized Linear Models, 2nd edition. Chapman and Hall, London.

McIntosh, A. R., and B. L. Peckarsky. 1999. Criteria determining behavinoural reponses to multiple predators by a stream mayfly. Oikos 85:554-564.

Merritt, R. W., and K. W. Cummins. 1996. An introduction to the aquatic insects of North America, 3rd edition. Kendall Hunt Publishing Company, Iowa.

Moyle, P. B., and T. Light. 1996. Boilogical invasions of fresh water: empirical rules and assebly theory. Boilogical Conservation 78:149-161.

Nakano, S., H. Miyasaka, and N. Kuhara. 1999. Terrestrial-aquatic linages: riparian arthropod inputs alter trophic cascades in a stream food web. Ecology 80:2435- 2441.

Peckarsky, B. L. 1979. Biological interactions as determinants of destributions of benthic invertebrates within the substrates of stony streams. Limnology and Oceanography 24:59-68.

116 Poff, N. L., and J. V. Ward. 1989. Implications of streamflow variaablilty and predictability for lotic community structure: a regional analysis of streamflow patterns. Canadian Jouranal of Fisheries Aquatic & Sciences 46:1805-1818.

Power, M. E. 1990. Effects of fish in river food webs. Science 250:811-814.

Power, M. E. 1992. Habitat heterogeneity and the functional significance of fish in river food webs. Ecology 73:1675-1688.

Rahel, F. J., and W. A. Hubert. 1991. Fish assemblages and habitat gradient in a Rocky Mountain - Great Plains stream: biotic zonation and additive patterns of community change. Transations of Amercian Fisheries Society 120:319-332.

Robinson, J. V., and M. A. Edgemon. 1988. An experimental evaluation of the effect of invasion history on community structure. Ecology 69:1410-1417.

Rosenfeld, J. S. 1997. The influence of upstream predation of the expression of fish effects in sownstream patches. . Freshwater Biology 37:535-543.

Sih , A. 1982. Foraging strategies and the avoidance of predations by an aquatic insect, Notonecta hoffmani. Ecology 63:786-796.

Sih, A., G. Englund, and D. Wooster. 1998. Emergent impacts of multiple predators on prey. . Trends in Ecology and Evolution 13:350-355.

Sokol-Hessner, L., and O. J. Schmitz. 2002. Aggregate effects of multiple predator species on a shared prey. . Ecology 83:2367-2372.

Soluk, D. A., and N. C. Collins. 1988. Synergistic interactions between fish and stoneflies: Facilitation and interference among stream predators. . Oikos 52:94- 100.

Vannote, R. L., G. W. Minshall, K. W. Cummins, J. R. Sedell, and C. E. Cushing. 1980. The river continuum concept. Canadian Journal of Fisheries and Aquatic Sciences 37:130-137.

Wilbur, H. M., and J. E. Fuath. 1990. Experimental aquatic food webs: interactions between two predators and two prey. . American Naturalist 135:176-204.

Williams, L. R., C. M. Taylor, M. L. Warren Jr., and J. A. Clingenpeel. 2003. Environmental variability, historical contingency, and the structure of regional fish and macroinvertebrate faunas in Ouachita Mountain stream systems. Environmental Biology of Fishes 67:203-216.

Wooster , D. 1994. Predatore impacts on stream benthic prey. Oecologia 99:7-15.

117

TABLES

Source DF Sum of Squares F Ratio Prob > F Velocity 1 166.51080 14.3192 0.0009 Depth 1 0.00106 0.0001 0.9925 Day 6 66.33954 0.9508 0.4775 Velocity *Depth 1 0.36007 0.0310 0.8617 Day* Velocity 6 94.67088 1.3569 0.2702

Table 4. 1 Family Richness main effects and interaction, Experiment 1

Source DF Sum of Squares F Ratio Prob > F Velocity 1 10036.267 1.9286 0.1772 Depth 1 9518.340 1.8290 0.1883 Day 6 64096.388 2.0528 0.0958 Velocity *Depth 1 4383.551 0.8423 0.3675 Day* Velocity 6 15331.540 0.4910 0.8088

Table 4. 2 Family Abundance main effects and interaction, Experiment 1

118

Shredder 0.73 0.29

0.01

0.53 was at seven levels (0, 2, 5, 10, 15, 20, 0.01 0.05 0.01 33 1.17 1.02 0.36 1.18 0.75 1.60 0.31 ed within the Stream block. Day Chitto, and Birds Creeks; Experiment 1). Collector Filterer Predator Scraper Source (Day) Str. Stream (Day) Str. Stream (Day) Str. Stream (Day) Str. Stream (Day) Str. Stream Significant p-valuesSignificant in bold. DF Squares of Sum Ratio F 0.20 P-value 0.04 18 1.05 2. 0.46 2.10 2 0.15 1.27 0.30 18 5.74 2.17 2 6.95 18 0.96 2 5.47 0.75 18 1.31 2 18 2 30), and Stream at three (Drakes, Whisky Table 4. 3 Nested ANOVA with Day nest

119

Source DF Sum of Squares F Ratio Prob > F Velocity 1 98.695703 7.4643 0.0125 Depth 1 12.905146 0.9760 0.3344 guild 5 81.783561 1.2370 0.3272 Velocity *Depth 1 2.735004 0.2068 0.6539 Guild* Velocity 5 35.338494 0.5345 0.7478

Table 4. 4 Family Richness main effects and interaction, Experiment 2.

Source DF Sum of Squares F Ratio Prob > F Velocity 1 6471.598 1.8552 0.1876 Depth 1 16129.279 4.6238 0.0433 guild 5 31278.777 1.7934 0.1579 Velocity *Depth 1 76.497 0.0219 0.8837 Guild* Velocity 5 7244.151 0.4153 0.8327

Table 4. 5 Family Abundance main effects and interaction, Experiment 2.

120

Source DF ChiSquare Prob>Chisq Guild 5 7.7493 0.1706 a) Invert 5 400.5360 <.0001 Guild*Invert 25 38.9280 0.0375

Term Estimate Std Error ChiSquare Prob>Chisq Intercept 0.73 0.0179 458.81 <.0001 Guild[Alg] -0.04 0.0395 0.99 0.3191 b) Guild[Ins] -0.04 0.0426 1.09 0.2952 Guild[Mix] 0.04 0.0395 0.90 0.3411 Guild[Opn] 0.09 0.0395 5.6 0.0180 Guild[Pis] -0.02 0.0395 0.2 0.6499 Invert[filterer] -0.27 0.0401 43.26 <.0001 Invert[piercer] -0.72 0.0401 195.9 <.0001 Invert[predator] 0.25 0.0401 36.07 <.0001 Invert[scraper] -0.07 0.0401 3.77 0.0519 Invert[shredder] -0.39 0.0401 80.71 <.0001 Guild[Alg]*Invert[filterer] -0.06 0.0885 0.55 0.4557 Guild[Alg]*Invert[piercer] 0.03 0.0885 0.12 0.7252 Guild[Alg]*Invert[predator] 0.14 0.0885 2.67 0.1017 Guild[Alg]*Invert[scraper] -0.09 0.0885 1.26 0.2603 Guild[Alg]*Invert[shredder] -0.02 0.0885 0.08 0.7718 Guild[Ins]*Invert[filterer] 0.3 0.0952 9.99 0.0016 Guild[Ins]*Invert[piercer] 0.03 0.0952 0.14 0.7032 Guild[Ins]*Invert[predator] -0.12 0.0952 1.74 0.1865 Guild[Ins]*Invert[scraper] -0.13 0.0952 2.01 0.1560 Guild[Ins]*Invert[shredder] -0.04 0.0952 0.23 0.6244 Guild[Mix]*Invert[filterer] -0.07 0.0885 0.75 0.3850 Guild[Mix]*Invert[piercer] 0.004 0.0885 0.00 0.9632 Guild[Mix]*Invert[predator] 0.05 0.0885 0.36 0.5450 Guild[Mix]*Invert[scraper] -0.05 0.0885 0.33 0.5601 Guild[Mix]*Invert[shredder] 0.12 0.0885 1.99 0.1579 Guild[Opn]*Invert[filterer] -0.13 0.0885 2.36 0.1241 Guild[Opn]*Invert[piercer] -0.1 0.0885 1.34 0.2469 Guild[Opn]*Invert[predator] -0.09 0.0885 1.07 0.3003 Guild[Opn]*Invert[scraper] 0.43 0.0885 22.50 <.0001 Guild[Opn]*Invert[shredder] -0.11 0.0885 1.73 0.1874 Guild[Pis]*Invert[filterer] 0.03 0.0885 0.17 0.6751 Guild[Pis]*Invert[piercer] 0.009 0.0885 0.01 0.9135 Guild[Pis]*Invert[predator] 0.06 0.0885 0.53 0.4663 Guild[Pis]*Invert[scraper] -0.05 0.0885 0.36 0.5443 Guild[Pis]*Invert[shredder] -0.1 0.0885 1.4 0.2364 Table 4. 6 Generalized linear model results of the effect of Fish Guilds on individual macroinvertebrate feeding guilds. Fish guilds were represented by Algivore, Invertivore, Mix, Piscivore, Open, and Control (no fish). The macroinvertebrate (Invert) feeding guilds were represented by Collector, Filterer, Piercer, Predator, Scraper, and Shredder. The Control, and Collector were held as intercepts in the interaction model. a) Main effect tests, and b) Parameter estimates, and specific interactions.

121

FIGURES

C

A B

N

W E

S

0 5 10 15

Kilometers

Figure 4.1 Sampling conducted on streams in Fort Polk, Louisiana. A- Drake’s creek, B- Whisky Chitto creek, and C- Bird’s creek.

122

Species Richness

20

15

10

5

0 0 2 5 10 15 20 30

Figure 4.2 Mean species richness of macroinvertebrates sampled over time interval did not show statistically significant differences (F6, 0.48, P=0.81: Experiment 1)

Species Abundance

300

200

100

0 0 2 5 10152030

Figure 4.3 Mean species abundance of macroinvertebrates sampled over time inverval did not show significant differences (F6, 2.23, P=0.06; Experiment 1).

123

collector 2.5 filterer predator 2 scraper

shredder e

1.5

1 Abundanc log

0.5

0

0 2 5 10 15 20 30

Days

Figure 4.4 Successional mean macroinvertebrate abundance from Day 0 to Day 30, in absence of fish, based on feeding guilds.

124

100%

e 80% Oligochaeta Heptageniidae 60% Elmidae

40% Chironomidae

Invertebrat of Proportion 20% Caeniidae

0% 02510152030

Number of Days

Figure 4.5 Proportional abundance of top five abundant macroinvertebrate taxa sampled over time interval.

125

Species Richness

20

15

10

5

0 Control Open Algivore Insectivore Piscivore Mix

Figure 4.6 Mean macroinvertebrate species richness in the various fish enclosures did not show significant difference (F6, 0.6, P=0.7; Experiment 2).

Species Abundance

250

200

150

100

50

0 Control Open Algivore Insectivore Piscivore Mix

Figure 4.7 Mean macroinvertebrate taxa abundance in the various fish enclosures did not significant diffeence (F6, 1.3, P=0.3; Experiment 2).

126

100% Hydropsychidae s Heptageniidae 80% Elmidae

60%

40% Chironomidae 20%

Proportion of Invertebrate of Proportion Caeniidae 0%

pen Mix O Control lgivore iscivore A P Insectivore

Fish feeding guild enclosure type

Figure 4.8 Proportional abundance of top five macroinvertebrate taxa occuriing in the various Fish Guild enclosures.

127

CHAPTER 5

Conclusions

An important area in stream ecology is understanding biotic and abiotic

interactions, and their effects on fish and macroinvertebrate assemblages (Gorman and

Karr 1978, Richards and Host 1994, Vinson and Hawkins 1998). These relationships have been known to be highly dependent on the spatial and temporal scales of individual studies (Vinson and Hawkins 1998, Lammert and Allan 1999). A stream is closely tied to its adjacent riparian zone, and is a product of its biogeographic history (Hooke 1999).

The effect of anthropogenic land use on stream systems is prevalent in the Gulf Coastal

Plains ecoregion of the Southeastern United States, where extensive agriculture, timber

harvesting, and human population has reduced the forest cover over the past 200 years

(Omernik 1987, Frost 1993). Large scale disturbances arising from anthropogenic or

natural causes can alter the nutrient cycling, stream geomorphology, hydrologic patterns,

sedimentation rates, and woody debris input. In this dissertation, I examined the effects

of timber harvesting activities on the Southeastern Coastal Plain streams in Fort Polk,

Louisiana. I also examined impacts of biological interactions between fish and

macroinvertebrates under exclosure/enclosure experiments under field settings.

In the first chapter, timber harvest activities did not show a significant deleterious

effect on the macroinvertebrate assemblage. The comparative study of macroinvertebrate

assemblages during the pre-, during-, and post- timber harvest years did not show any 128 significant difference. The geomorphologic data on the system suggested high levels of

disturbance to the physical structure of the stream. Despite this, the lack of effect on

macroinvertebrate assemblages was consistent with studies of Williams et al. (2002,

2005), which suggests other higher order forces structuring the assemblages. The

biogeographic history predetermines the geology and hydrology of individual basins, and

if the degree of anthropogenic disturbance is lower than this level, it may not have

significant effect on stream communities. The effects of land-use activities have been

observed to have stronger effects on stream communities at local rather than regional

scales (Campbell and Deog 1989, Lammert and Allan 1999). Because the study was

conducted on streams within a drainage, the results were limited at that scale. Scaling

down from stream level to mesohabitat level, I was able to detect differences in

assemblages that were not detected across streams; with pools having the least, and the

runs having highest taxa richness. Similarly, with temporal scale studied covered three

years, with no significant difference. Although no difference was detected within years,

with timber harvest activity during one of the years, a strong seasonal shift in

macroinvertebrate assemblage was detected. This suggests assemblages that respond to smaller temporal scale at the seasonal level (Lohr and Fausch 1997). The streams in this

region historically experience harsh conditions, and thus would serve as a filter to the

intolerant species. Another reason for lack of timber harvesting effect could be the

relatively high spatial and temporal variations in stream communities within sites, which

can make it harder to detect changes resulting from natural or anthropogenic disturbance

at the stream-reach scale unless the disturbance is at a much larger scale.

129 The second chapter was an extension of the first chapter, which focused on the effects of road crossings, a byproduct of timber harvesting activities, on fish movement.

Poor road crossing designs can alter the stream geomorphology, and potentially reduce migration of fishes (Baker and Votapka 1990). I found negative effects of road crossings on fish movement, when compared to movement through natural reaches of the streams.

Movement through various bridge designs did not indicate a significant difference, which was a result of lower number of recaptures. Although water depth and current velocity both typically affect fish movement (Warren and Pardew 1998), I only detected an inverse relationship between water depth and fish movement, and no significant relationship with current velocity. Lack of significant relationship of movement with current velocity is a result of greater variation observed in the stream hydrologic conditions in the headwater reaches. Larger sample sizes including sampling during different seasons could give a better interpretation of fish movement, capturing the movement during different life history stages.

In the third chapter, I determined that there was a shift in macroinvertebrate assemblage based on the feeding guilds, in the absence of fish. The successional dynamics of the macroinvertebrate assemblage showed a significant increase in the

predatory guild, corresponding with a downward trend in the scraper guild. By removing

the fish predators I did not detect a significant increase in macroinvertebrate numbers,

because the experimental setting prevented entry or escape of organisms > 1mm. This

was an excellent setting for invertebrate predator-prey interactions, and competition

between guilds for resources. In a separate setting I also found significant effect of

Insectivore fishes on macroinvertebrate community, when compared against other

130 feeding guilds of fish. The effects of various Fish feeding guilds on macroinvertebrates

reduced with the Mix, and Piscivore enclosures in descending order. This was an

important study tied to the main study of the effects of disturbance, in that the variability

in water depths in the headwater reaches often results in isolation of small pools devoid

of fish. Thus, in the absence of top predator, fish, the predatory invertebrates would take

over an important role in structuring the community. Separating the effects of various

feeding guilds of fish on macroinvertebrates is useful in quantifying the predation risk of

macroinvertebrates (Vance-Chalcraft and Soluk 2005).

The results of this dissertation indicate that disturbances resulting from timber

harvesting may not significantly affect macroinvertebrate assemblages of the Gulf

Coastal Plain streams, but it is essential to consider the long-term effects of multiple, small-scale disturbances such as crossings, and road drainage, on these communities.

Because of the close interdependence of stream with the riparian vegetation structure, it is essential to understand the factors affecting the riparian zone. Although studies suggest strong negative effects of forest fragmentation on stream macroinvertebrate communities

(Resh et al. 1988, Allan et al. 1997, Cushing and Allan 2001), my results indicate otherwise. This lack of effect on macroinvertebrate assemblage suggests greater resilience of these communities in midst of disturbances. The effects of these disturbances cannot be comprehended unless the historic natural disturbance regime is taken into consideration. In view of this, I hope to have answered some of the questions of the complex processes operating on the stream system, rather than raising many more.

131 REFERENCES

Allan, J. D., D. L. Erickson, and J. Fay. 1997. The influence of catchment land use on stream integrity across multiple spatial scales. Freshwater Biology 37:149-161.

Baker, C. O., and F. E. Votapka. 1990. Fish Passage Through Culverts, San Dimas, CA:USDA. FHWA-FL-90-006:1-67, Forest Service Technology and Development Center.

Campbell, I. C., and T. J. Deoeg. 1989. Impact of timber harvesting and production on streams: a review. Australian Journal of Marine and Freshwater Resources 40.

Cushing, C. E., and J. D. Allan. 2001. Streams: Their ecology and life. Academic Press, San Diego, California.

Frost, C. C. 1993. Four centuries of changing landscape patterns in the longleaf pine ecosystem. in S. M. Hermann, editor. 18th Tall Timbers Fire Ecology Conference, Tallahassee, Florida.

Gorman, O. T., and J. R. Karr. 1978. Habitat structure and stream fish communities. Ecology 59:507-515.

Lammert, M., and J. D. Allan. 1999. Assessing biotic integrity of streams: effects of scale in measuring the influence of land use/cover and habitat structure on fish and macroinvertebrates. Environmental Management 23:257-270.

Lohr, S. C., and K. D. Fausch. 1997. Multiscale analysis of natural variability in stream fish assemblages of a Western Great Plains Watershed. Copeia 1997:706-724.

Omernik, J. M. 1987. Ecoregions of the conterminous United States. Annals of the Association of American Geographers 77:118-125.

Resh, V. H., A. V. Brown, A. P. Covich, M. E. Gurtz, H. W. Li, G. W. Minshall, S. R. Riece, A. L. Sheldon, J. B. Wallace, and R. C. Wissmar. 1988. The role of disturbance in stream ecology. Journal of North American Benthological Society 7:433-455.

Richards, C., and G. Host. 1994. Examining land use influences on stream habitats and macroinvertebrates: a GIS approach. Water Resources Bulletin 30:729-738.

Vance-Chalcraft, H. D., and D.A. Soluk. 2005. Multiple predator effects result in risk reduction for the prey across multiple prey densities. Oecologia 144:472-480.

132 Vinson, M. R., and C. P. Hawkins. 1998. Biodiversity of stream insects: variation at local, basin, and regional scales. Annual Review of Entomology 43:271-293.

Warren Jr., M. L., and M. G. Pardew. 1998. Road crossings as barriers to small-stream fish movement. Transactions Of The American Fisheries Society 127:637-644.

Williams, L. R., T. H. Bonner, J. D. Hudson III, M. G. Williams, T. R. Leavy, and C. S. Williams. 2005. Interactive efects of environmental variability and military training on stream biota of three headwater drainages in Western Louisiana. Transactions of the American Fisheries Society 134:192-206.

Williams, L. R., C. M. Taylor, M. L. Warren Jr., and J. A. Clingenpeel. 2003. Environmental variability, historical contingency, and the structure of regional fish and macroinvertebrate faunas in Ouachita Mountain stream systems. Environmental Biology of Fishes 67:203-216.

133

APPENDIX A

HABITAT DATA COLLECTED BY SEASONS IN THE YEARS 2003, 2004, AND 2005, IN LITTLE SANDY, ODOM, AND TIGER CREEKS IN FORT POLK, LOUISIANA.

134

Little Sandy Creek Odom Creek Tiger Creek Fall Spring Summer Winter Fall Spring Summer Winter Fall Spring Summer Winter Conductivity (µS/cm) 137.00 64.11 81.94 67.11 84.00 46.75 76.50 59.67 74.38 65.88 93.65 59.13 Dissolved Oxygen (mg/L) 6.40 9.37 8.32 10.04 8.09 9.23 7.75 9.83 6.76 9.44 7.93 7.97 pH 6.90 7.25 6.52 6.20 7.15 6.50 7.11 5.93 8.13 7.46 7.18 6.52 Temperature (ºC) 16.40 20.23 25.47 9.15 15.46 20.05 29.16 11.56 15.27 20.40 21.60 13.78 Turbidity (NTU*) 0.00 26.42 17.78 21.77 17.23 34.45 15.88 19.50 21.53 34.09 15.60 52.14 Mean Water Velocity (m/s) 0.04 0.14 0.06 0.20 0.10 0.23 0.16 0.21 0.06 0.09 0.05 0.15 Mean Mesohabitat Length (m) 41.75 29.46 29.48 28.54 68.35 85.73 77.00 26.89 20.89 19.53 21.77 16.61 Mean Mesohabitat Width (m) 3.11 4.37 3.48 4.49 3.97 6.03 3.72 3.90 2.33 2.93 2.25 2.56 Mean Water Depth (m) 0.16 0.16 0.09 0.21 0.13 0.36 0.10 0.20 0.16 0.20 0.12 0.19

135 Clay % 0.00 5.28 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 Silt % 1.40 0.11 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.75 0.00 Sand % 68.60 60.72 69.38 70.83 99.33 100.00 50.00 90.63 79.69 87.60 90.15 76.74 Gravel % 5.00 0.28 8.13 6.67 0.00 0.00 0.00 0.00 7.50 12.00 6.00 13.91 Cobble % 15.00 16.67 18.75 11.39 0.67 0.00 0.00 9.38 12.81 0.40 3.10 9.35 Bedrock % 10.00 16.94 3.75 11.11 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 Percent Undercut Bank 11.50 6.85 3.48 7.38 1.19 8.40 7.94 17.74 4.80 9.68 5.50 10.27 Percent Woody Debris 2.03 2.25 3.25 7.75 5.89 6.28 3.64 8.93 9.61 6.18 8.18 8.13

APPENDIX B

PRESENCE/ABSENCE OF MACROINVERTEBRATE TAXA WITH SCIENTIFIC NOMENCLATURE AND THEIR FEEDING GUILDS, COLLECTED IN LITTLE SANDY, ODOM, AND TIGER CREEKS DURING 2003-2005.

136

Feeding Little Sandy Odom Tiger Order Family Genus Guild Creek Creek Creek Amphipoda Amphipoda Crangonyx collector 1 0 0 Arachnida Pisauridae Dolomedes predator 1 0 1 Bivalve Bivalve filterer 1 1 1 Coleoptera Chironomidae collector 0 0 1 Chrysomelidae shredder 1 0 1 Dryopidae Helicus scraper 1 0 1 Dytiscidae Celina predator 0 0 1 Coptotomus predator 1 0 0 Cybister predator 0 1 0 Desmopuchria predator 1 1 1 hydroporus predator 1 0 0 Hygrotus predator 1 0 0 Laccophilus predator 1 0 1 Nebrioporus predator 0 1 0 Oreodytes predator 1 0 1 Elmidae Ancryonyx collector 1 0 0 Cylloepus collector 1 0 1 Dubaraphia collector 1 1 0 Macronychus collector 1 1 1 Stenelmis collector 1 1 1 Gyrinidae Dineutus predator 1 0 0 Gyretes predator 1 1 1 Halipilidae shredder 1 0 0 Hydrochidae Hydrochus shredder 0 0 1 Hydrophilidae collector 0 1 1 Psephenidae Ectopria scraper 1 1 1 Scirtidae Cyphon scraper 0 0 1 Collembola Sminthuridae collector 1 1 1 Collembolla Sminthuridae collector 1 1 1 Copepoda Copepoda collector 1 1 1 Crawfish Crawfish shredder 1 1 1 Daphnia Daphnia filterer 1 0 0 Diptera Ceratopogonidae predator 1 1 1 Chaoboridae piercer 1 0 1 Chironomidae collector 1 1 1 Culicidae collector 1 1 1 Deuterophlebiidae scraper 1 0 0 Dixidae collector 1 1 1 Empididae predator 1 1 1 Ephydridae collector 0 1 1 Nymphomyiidae scraper 0 1 1 Simuliidae collector 1 1 1 Tabanidae piercer 1 1 1 Thaumaleidae scraper 0 0 1 Tipulidae shredder 1 1 1

137

Ephemeroptera Baetidae Acentrella collector 1 0 0 Acerpenna collector 1 0 0 Baetis collector 1 0 1 Labiobaetis collector 1 0 1 Procloeon collector 1 0 0 Baetiscidae Baetisca collector 1 1 1 Caenidae Brachycercus collector 1 1 1 Caenis collector 1 1 1 Ephemerellidae collector 1 1 1 Ephemeridae Hexagenia collector 1 1 1 Pentagenia collector 0 0 1 Heptageniidae Stenacron scraper 1 1 1 Stenonema scraper 1 1 1 Isonychiidae Isonychia collector 1 1 1 Leptophlebidae Leptophlebia collector 1 1 1 Limnophlebidae shredder 1 0 0 Neoephemeridae Neoephemera collector 1 0 0 Tricorythidae Tricorythodes collector 1 0 1 Corixidae Trichocorixa piercer 0 0 1 Gelastocoridae Gelastocorids piercer 1 0 0 Gerridae Gerris piercer 0 0 1 Limnoporous piercer 0 0 1 Metrobates piercer 1 0 1 Rheumatobates piercer 0 1 1 Trepobates piercer 1 1 1 Hebridae Hebrus piercer 1 0 1 Lipogomphus piercer 1 1 1 Hydrometridae Hydrometra piercer 1 1 1 Mesoveliidae Mesovelia piercer 0 1 1 Naucoridae Pelocoris piercer 0 0 1 Nepidae Curicta piercer 0 1 0 Noteridae Hydrocanthus predator 1 0 0 Notonectidae Notonecta piercer 1 1 1 Veliidae Microvelia piercer 1 1 1 Raghovelia piercer 1 1 1 Hydrachnida Hydrachnida predator 1 1 1 Isopoda Asellidae Caecidotea collector 1 0 0 collector 1 1 1 shredder 1 1 1 Megaloptera Cordulegasteridae Corydalus predator 0 0 1 Corduliidae Corydalus predator 0 1 0 Corydalidae Corydalus predator 1 1 1 Neohermes predator 0 0 1 Sialidae Sialis predator 1 0 0 Nematoda Nematoda collector 0 0 1

138

Odonata Aeshnidae Anax predator 1 1 1 Boyera predator 1 1 1 Boyeria predator 1 0 1 Coryphaeschna predator 0 0 1 Caenidae Caenis collector 1 0 0 Calopterygidae Calopteryx predator 1 1 1 Coenogrionidae Argia predator 1 1 1 Cordulegasteridae Cordulegaster predator 1 0 0 Corduligaster predator 1 1 1 Corduliidae Epitheca predator 0 0 1 Helocordulia predator 0 0 1 Somatochlora predator 0 0 1 Gomphidae Dromogomphus predator 1 1 1 Progomphus predator 1 1 1 Libuellulidae Libellula predator 0 1 1 Macromiidae Macromia predator 1 1 1 Oligochaeta Oligochaeta collector 1 0 0 Orthoptera Tetrigidae shredder 1 1 1 Plecoptera Capniidae Allocapnia shredder 1 1 1 Capnia shredder 1 1 1 Nemouridae Amphinemura shredder 1 1 0 Perlidae Acroneuria predator 1 1 1 Neoperla predator 0 0 1 Perlesta predator 1 1 1 Taeniopterygidae Taeniopteryx shredder 1 1 0 Ticks Ticks predator 0 0 1 Trichoptera Helicopsychidae Helicopsyche scraper 0 0 1 Hydropsychidae Cheumatopsyche collector 1 1 1 Hydropsyche collector 1 1 1 Hydroptilidae Agraylea piercer 1 1 0 Hydroptila piercer 1 0 0 Ochrotrichia piercer 1 1 1 Lepidosomatidae shredder 0 0 1 Leptoceridae collector 1 1 1 Molannidae scraper 1 1 1 Philopotamidae Chimarra collector 1 0 1 Phryganeidae shredder 0 0 1 Polycentropodidae Cernotina predator 1 0 1 Cyrnellus predator 0 0 1 Psychomyiidae Lype collector 1 0 0

139

APPENDIX C

LIST OF FISH SPECIES WITH SCIENTIFIC NOMENCLATURE COLLECTED IN RED RIVER, SABINE, AND CALCASIEU DRAINAGES DURING SUMMER 2003 AND SUMMER 2004.

140

Family Genus Species Common name Aphredoderidae Aphredoderus sayanus pirate perch Catostomidae Minytrema melanops spotted sucker Centrarchidae Lepomis cyanellus green sunfish Centrarchidae Lepomis gulosus warmouth sunfish Centrarchidae Lepomis macrochirus bluegill Centrarchidae Lepomis marginatus dollar sunfish Centrarchidae Lepomis megalotis longear sunfish Centrarchidae Lepomis punctatus spotted sunfish Centrarchidae Micropterus punctulatus spotted bass Centrarchidae Micropterus salmoides largemouth bass Cyprinidae Notropis atrocaudalis blackspot shiner Cyprinidae Notropis chrysocephalus striped shiner Cyprinidae Notropis fumeus ribbon shiner Cyprinidae Notropis sabinae sabine shiner Cyprinidae Notropis umbratilis redfin shiner Cyprinidae Notropis venustus blacktail shiner Cyprinidae Sematilus atromaculatus creek chub Cyprinodontidae Fundulus olivaceus top minnow Elassomatidae Elassoma zonatum pigmy sunfish Esocidae Esox niger chain pickerel Ictaluridae Ictalurus natalis yellow bullhead Ictaluridae Noturus nocturnus freckled madtom Ictaluridae Noturus phaeus brown madtom Percidae Etheostoma chlorosomum bluntnose darter Percidae Etheostoma gracile slough darter Percidae Etheostoma whipplei redfin darter Percidae Percina sciera dusky darter Poeciliidae Gambusia affinis mosquito fish

141

APPENDIX D

HABITAT DATA COLLECTED AT THREE SEGMENTS, AT CROSSING, AND AT NATURAL REACH IN THE THREE DRAINAGES IN FORT POLK, LOUISIANA

142

Data Segment Odom Reaugalle Squirrel Stagestand Comrade Martin Branch 1 0.06 0.17 0.12 0.20 0.40 0.09 2 0.06 0.09 0.08 0.06 0.23 0.15 3 0.06 0.12 0.05 0.05 0.41 0.56

Depth (m) Depth Crossing 0.07 0.12 0.03 0.03 0.10 0.04 Mean Water Mean Natural 0.07 0.19 0.06 0.04 0.74 0.39 1 0.20 0.08 0.07 0.04 0.04 0.10 2 0.19 0.18 0.21 0.20 0.05 0.16 3 0.18 0.10 0.20 0.20 0.01 0.03 Crossing 0.22 0.13 0.16 0.21 0.06 0.13 Mean Current Current Mean Velocity (m/s) Natural 0.18 0.08 0.21 0.17 0.01 0.06 1 3.33 3.86 5.53 3.92 1.95 3.13 2 3.24 2.46 2.28 2.12 4.25 1.91

(m) 3 3.59 2.75 2.98 2.44 9.23 3.22 Crossing ------Mean Width Width Mean Natural 3.24 3.58 2.93 2.82 3.88 2.92

143

APPENDIX E

ABUNDANCE OF FISH CAPTUREDIN DIFFERENT DRAINAGES COLLECTED DURING SUMMER 2003, AND SUMMER 2004 IN FORT POLK, LOUISIANA

144

Species composition of fishes in the drainages in Fort Polk, Louisiana Species Red River* Calcasieu Sabine

Blackspot shiner 13 0 45 Blacktail shiner 178 0 3 Bluegill 0 7 17 Bluntnose darter 2 0 2 Brown madtom 171 0 0 Chain pickerel 0 6 3 Creek chub 98 10 17 Dollar sunfish 21 36 2 Dusky darter 1 0 0 Freckled madtom 0 0 4 Green sunfish 30 2 2 Largemouth bass 0 2 0 Longear sunfish 21 19 26 Mosquito fish 0 12 0 Pigmy sunfish 0 1 0 Pirate perch 0 4 9 Redfin darter 70 0 5 Redfin shiner 164 32 97 Ribbon shiner 4 0 3 Sabine shiner 83 0 0 Slough darter 0 0 1 Spotted bass 20 2 1 Spotted sucker 0 0 1 Spotted sunfish 2 2 5 Striped shiner 346 0 0 Top minnow 392 58 11 Warmouth 0 4 5 Yellow bullhead 0 1 3

* Data from Red River drainage is the sum of species found in Odom, Reaugalle, Squirrel Branch, and Stagestand creeks

145

APPENDIX F

TOTAL FISH CAPTURED AND RECAPTURES DURING SUMMER 2003 AND SUMMER 2004 IN FORT POLK, LOUISIANA

146

Abundance of fish species collected during 2003

Calcasieu Sabine Red River Comrade Martin Odom Reaugalle Squirrel Branch Stagestand Species A B A B A B A B A B A B Blackspot shiner 0 0 45 2 0 0 1 0 2 0 2 0 Blacktail shiner 0 0 3 0 14 2 2 1 11 0 0 0 Bluegill 7 1 17 2 0 0 0 0 0 0 0 0 Bluntnose darter 0 0 2 0 0 0 0 0 0 0 0 0 Brown madtom 0 0 0 0 19 1 7 0 25 2 13 3 Chain pickerel 6 0 3 0 0 0 0 0 0 0 0 0 Creek chub 10 0 17 0 18 1 2 0 8 0 3 0 Dollar sunfish 36 3 2 0 0 0 7 0 3 0 3 0 Freckled madtom 0 0 4 0 0 0 0 0 0 0 0 0 Green sunfish 2 0 2 0 6 0 2 0 2 1 6 0 Largemouth bass 2 0 0 0 0 0 0 0 0 0 0 0 Longear sunfish 19 0 26 2 1 0 2 0 0 0 1 0 Mosquito fish 12 0 0 0 0 0 0 0 0 0 0 0 Pigmy sunfish 1 0 0 0 0 0 0 0 0 0 0 0 Pirate perch 4 0 9 0 0 0 0 0 0 0 0 0 Redfin darter 0 0 5 0 5 0 4 0 5 0 12 1 Redfin shiner 32 1 97 16 0 0 18 1 14 0 37 4 Ribbon shiner 0 0 3 0 0 0 0 0 0 0 0 0 Sabine shiner 0 0 0 0 3 0 3 0 10 1 3 1 Slough darter 0 0 1 0 0 0 0 0 0 0 0 0 Spotted bass 2 0 1 0 0 0 3 0 4 0 2 0 Spotted sucker 0 0 1 0 0 0 0 0 0 0 0 0 Spotted sunfish 2 0 5 0 0 0 0 0 0 0 1 0 Striped shiner 0 0 0 0 47 2 19 2 62 5 5 0 Top minnow 58 2 11 1 28 5 10 2 57 13 32 3 Warmouth 4 0 5 0 0 0 0 0 0 0 0 0 Yellow bullhead 1 0 3 0 0 0 0 0 0 0 0 0 Total 198 7 262 23 141 11 80 6 203 22 120 12

(A – total fish captured; B – total marked recaptures – B)

147

Abundance of fish species collected during 2004

Red River Drainage

Odom Reaugalle Squirrel Branch Stagestand Species A B A B A B A B Blackspot shiner 1 0 5 0 0 0 2 0 Blacktail shiner 48 6 0 0 103 13 0 0 Bluntnose darter 1 0 1 0 0 0 0 0 Brown madtom 40 2 8 0 32 8 27 3 Creek chub 19 1 16 0 18 2 14 2

Dollar sunfish 0 0 1 0 1 0 6 0 Dusky darter 0 0 1 0 0 0 0 0 Green sunfish 2 0 2 0 5 0 5 0 Longear sunfish 1 0 11 1 1 0 4 0 Redfin darter 10 0 4 0 16 0 14 0 Redfin shiner 2 0 10 0 42 9 41 10

Ribbon shiner 1 0 0 0 3 0 0 0

Sabine shiner 10 2 1 0 44 9 9 3 Spotted bass 0 0 4 1 0 0 7 1 Spotted sunfish 0 0 0 0 1 0 0 0 Striped shiner 26 2 38 8 140 12 9 1 Top minnow 24 2 52 9 116 25 73 20

Total 185 15 154 19 522 78 211 40

(A – total fish captured; B – total marked recaptures – B)

* Rivers from Calcasieu and Sabine drainages were not sampled during 2004 because of very low water conditions

148

APPENDIX G

ENVIRONMENTAL PARAMETERS MEASURED AT EACH EXCLOSURE/ENCLOSURE DURING THE FIELD EXPERIMENT DURING SUMMER 2003 IN FORT POLK, LOUISIANA

149 Experiment 1 – Environmental parameters measured at each exclosures during Summer 2003 (two replicates per exclosure)

Birds Drakes Whisky Chitto Day Velocity Depth Velocity Depth Velocity Depth 0 0.03 0.3 0.31 0.37 0.11 0.2 0 0.08 0.25 0.38 0.13 0.11 0.25 2 0.13 0.33 0.22 0.16 0.21 0.4 2 0.11 0.29 0.27 0.31 0.03 0.42 5 0.04 0.32 0.25 0.36 0.15 0.5 5 0.12 0.3 0.27 0.22 0.15 0.42 10 0.11 0.24 0.23 0.12 0.19 0.12 10 0.02 0.9 0.29 0.22 0 0.09 15 0.07 0.3 0.24 0.2 0 0.1 15 0.07 0.26 0.16 0.17 . . 20 0.04 0.4 0.1 0.5 0.01 0.26 20 0.05 0.4 0.2 0.71 0.42 0.2 30 0.1 0.45 0.19 0.49 0 0.25 30 0.06 0.5 0.1 0.48 0.03 0.7

Experiment 2 – Environmental parameters measured at each enclosures during Summer 2003 (two replicates per exclosure)

Birds Drakes Whisky Chitto Type Velocity Depth Velocity Depth Velocity Depth Algivore 0.03 0.5 0.33 0.2 0.18 0.42 Algivore 0.02 0.45 0.3 0.18 0.1 0.22 Control 0.03 0.3 0.31 0.37 0.11 0.2 Control 0.08 0.25 0.28 0.13 0.11 0.25 Insectivore 0.08 0.31 0.16 0.25 0.09 0.16 Insectivore 0.12 0.38 0.26 0.22 . . Mix 0.21 0.39 0.26 0.38 0.05 0.38 Mix 0.16 0.31 0.14 0.24 0.08 0.45 Open 0.03 0.3 0.31 0.37 0.11 0.2 Open 0.08 0.25 0.28 0.13 0.11 0.25 Piscivore 0.2 0.32 0.28 0.2 0.23 0.18 Piscivore 0.15 0.26 0.26 0.22 0.28 0.16

* Units of Velocity in meters/sec, and Depth in meters. 150

APPENDIX H

ABUNDANCE OF MACROINVERTEBRATE TAXA WITH SCIENTIFIC NOMENCLATURE AND THEIR FEEDING GUILDS, COLLECTED IN THE EXCLOSURE EXPERIMENT (I) DURING SUMMER 2003

151

Days Order Family Genus Feeding 0 2 5 10 15 20 30 Guild Amphipoda Amphipoda Amphipoda collector 2 1 11 Bivalve filterer 17 12 6 11 30 412 Coleoptera Chrysomelidae shredder 1 1 Dytiscidae Coptotomus predator 5 1 5 1 2 2 Laccodytes predator 1 1 Elmidae Ancryonyx collector 13 2 10 13 10 11 8 Dubaraphia collector 2 7 2 3 11 19 Macronychus collector 5 4 2 1 Stenelmis scraper 43 14 3 2 7 4 1 Gyrinidae Dineutus predator 1 Gyretes predator 3 1 Gyrinus predator 1 Psepheridae Ectopria scraper 1 1 Collembolla collector 1 11 Copepoda collector 3 3 4 Crayfish shredder 8 5 7 10 2 2 Diptera Ceratopogonidae collector 17 53 22 22 61 50 45 Chironomidae collector 54 58 15 22 26 26 48 0 5 9 1 1 3 6 Culicidae collector 2 4 1 4 2 Ephydridae collector 1 1 Simuliidae collector 19 2 2 1 Tabanidae predator 2 5 3 3 6 2 7 Tanyderidae scraper 1 Thaumaleidae scraper 6 5 8 9 5 Tipulidae shredder 3 1 3 2 1 4 Ephemeroptera Baetidae collector 51 9 27 4 6 7 7 Caeniidae Brachycercus collector 1 24 12 22 24 16 30 Caenis collector 68 57 33 64 46 16 16 Ephemeridae Hexagenia collector 3 1 3 2 Ephemerillidae collector 9 10 1 7 3 3 Heptageniidae Stenacron collector 49 17 17 70 17 15 17 Stenonema scraper 37 3 11 1 Isonychitidae Isonychia collector 8 3 1 Leptophelibidae Choroterpes collector 3 1 Neochoroterpes collector 1 Polymitarcyidae collector 1 Trichorichtidae collector 1 1 1 4 Fishing Spider predator 1 1 Gastropoda filterer 3 1 1 1 Hemiptera Geriidae Trepodates predator 2 1 Hebridae Hebrus predator 1 Lipogomphus predator 1 2 1 Hydrometridae Hydrometra predator 1

152

Mesoveliidae Mesovelia predator 1 Veliidae Microvelia predator 3 7 Raghovelia predator 6 10 7 9 3 6 1 Hydrachnida predator 4 1 2 Isopoda collector 1 Leech predator 2 Megaloptera Corydalidae predator 1 Nematoda collector 1 Nematodea collector 1 Neuroptera Sisyra predator 1 Odonata Aeshnidae Boyeria predator 3 2 Epiaeschna predator 4 Calyopterigidae Calyopteryx predator 14 6 13 3 2 1 2 Coenagrionidae Argia predator 17 20 2 11 17 27 45 Cordulidae Cordaliinae predator 1 1 Macromia predator 1 Somatochlora predator 2 2 1 predator 2 1 1 Gomphidae Dromogomphus predator 3 12 2 6 2 5 7 Erpetogomphus predator 5 Gomphus predator 1 Hagenius predator 2 Progomphus predator 5 1 3 Libellulidae predator 8 Macromeia predator 1 Macromiidae Macromiinae predator 1 Petaluridae Tachopteryx predator 5 Oligochaeta collector 97 112 71 75 75 131 360 Palaemonidae collector 6 10 5 13 9 27 29 Plecoptera Perlidae Perlesta predator 3 Ticks predator 1 4 1 Trichoptera Hydropsychidae Cheumatopsyche collector 35 Hydropsyche collector 56 3 1 2 Hydroptiidae piercer 2 3 Leptoceridae Trianodes shredder 1 Trianoides shredder 1 Molannidae scraper 3 3 5 1 2 4 2 Polycentropodidae Cernotina predator 12 1 2 3 2 1 4 Neureclipsis collector 1 Paranyctiophylax predator 2 1 1 Psychomyiidae predator 2 1

153

APPENDIX I

ABUNDANCE OF MACROINVERTEBRATE TAXA WITH SCIENTIFIC NOMENCLATURE AND THEIR FEEDING GUILDS, COLLECTED IN THE ENCLOSURE EXPERIMENT (II) DURING SUMMER 2003

154

Enclosure Type Insecti- Pisci- Order Family Genus Algivore Control Mix Open vore vore

Amphipoda Amphipoda 1 1 2 1 Bivalve 11 11 34 11 27 12 Coleoptera Dryopidae Helicus 1 Dytiscidae Coptotomus 4 5 2 2 4 2 Elmidae Ancryonyx 5 13 1 7 19 6 Dubaraphia 3 2 1 2 Macronychus 1 13 3 Stenelmis 4 1 2 3 22 7 Copepoda 1 1 Crayfish 2 10 3 16 5 4 Diptera Ceratopogonidae 24 22 23 24 11 8 Chironomidae 249 221 171 304 497 328 Culicidae 1 3 1 Ephydridae 1 1 1 Sciomyzidae 1 Simuliidae 1 Tabanidae 8 3 6 7 1 1 Tanyderidae 1 Thaumaleidae 3 5 2 7 2 Tipulidae 4 2 1 1 Ephemeroptera Baetidae 12 4 11 5 12 7 Caeniidae Brachycercus 19 22 17 20 5 16 Caenis 84 64 16 66 61 50 Ephemeridae Hexagenia 3 5 1 1 Ephemerillidae 2 7 3 7 6 Heptageniidae Stenacron 23 70 5 17 31 45 Stenonema 1 15 45 6 Isonychitidae Isonychia 3 13 1 Leptophlebidae 1 Tricorithidae Tricorythoides 6 Gastropoda 2 3 2 Hemiptera Corixidae Graptocorixa 1 Hebridae Lipogomphus 2 2 1 4 Mesoveliidae Mesovelia 1 Veliidae Microvelia 2 Raghovelia 1 9 3 5 10 13 Hydrachnida Arrenuroidae 1 Isopoda 1 Leech 1 Megaloptera Corydalidade Corydalus 1 2

155

Odonata Aeshnidae Boyeria 2 1 2 2 Epiaeschna 1 Calyopterigidae Calyopteryx 8 3 1 4 4 4 Coenagrionidae Argia 8 11 11 24 12 27 Corduligastridae Corduligaster 1 Gomphidae Dromogomphus 19 6 14 19 5 18 Gomphus 2 1 Progomphus 1 Stylurus 1 Macromidae Macromia 3 1 1 Macromiidae 1 Macromiinae 1 1 Oligochaeta 82 75 136 105 59 93 Palaemonidae 4 13 4 17 6 8 Plecoptera Perlidae Neoperla 1 Perlesta 6 Ticks 4 5 Trichoptera Hydropsychidae Cheumatops 2 10 19 4 yche Hydropsyche 15 2 47 17 Parapsyche 1 2 8 Hydroptilidae 1 Leptoceridae Oecetis 1 Trianoides 1 1 1 1 Molanidae 2 1 1 4 1 Odontoceridae 4 Philopotamidae Chimarra 3 Polycentropodidae Cernotina 4 3 3 4 7 4 Cyrnellus 2 Paranyctiophylax 1 1 Polycentropus 1 Psychomyiidae 11

156 BIBLIOGRAPHY

Abernethy, B., and I. D. Rutherfurd. 2001. The distribution and strength of riparian tree roots in relation to riverbank reinforcement. Hydrological Processes 15:39-61.

Allan, J. D. 1982. The effects of reduction in trout density on the inverebrate communtiy of a mountain stream. Ecology 63:1444-1455.

Allan, J. D. 1995. Stream ecology: structure and function of running waters. Chapman and Hall, London.

Allan, J. D., D. L. Erickson, and J. Fay. 1997. The influence of catchment land use on stream integrity across multiple spatial scales. Freshwater Biology 37:149-161.

Angermeier, P. L., A. P. Wheeler, and A. E. Rosenberger. 2004. A conceptual framework for assessing impacts of roads on aquatic biota. Fisheries 29:19-29.

Angermeir, P. L. 1987. Spatiotemporal variation in habitat selection by fishes in small Illinois streams. Pages 38-58 in W. J. Matthews and D. C. Heins, editors. Community and evolutionary ecology of North American stream fishes. University of Oklahoma Press, Norman, Oklahoma.

Aparicio, E., and A. de Sostoa. 1999. Pattern of movements of adult Barbus haasi in a small Mediterranean stream. Journal of Fish Biology 55:1086-1095.

Baker, C. O., and F. E. Votapka. 1990. Fish Passage Through Culverts, San Dimas, CA:USDA. FHWA-FL-90-006:1-67, Forest Service Technology and Development Center.

Bormann, F. H., and G. E. Likens. 1979. Pattern and Process in a Forested Ecosystem. Springer-Verlag New York Inc, NY.

Bowman, D., and G. S. Boggs. 2006. Fire ecology. Progress in Physical Geography 30:245-257.

Bowman, M. F., and R. C. Bailey. 1997. Does taxonomic resolution affect the multivariate description of the structure of freshwater benthic macroinvertebrate communities? Canadian Journal of Fisherriees and Aquatic Sciences 54:1802- 1807.

Bridges, E. L., and S. L. Orzell. 1989. Longleaf pine communities of the Western Gulf Coastal Plain. Natural Areas Journal 9:246-262.

157 Bronmark, C., J. Dahl, and L. Greenberg. 1997. Complex trophic interactions in freshwater benthic food chains. Pages 55-88 in B. Streit, T. Stadler, and C.M. Lively, editors. Ecology and Evolution of freshwater animals. Birkhaeuser- Verlag, Basel, Switzerland.

Brosofske, K. D., J. Chen, R. J. Naiman, and J. F. Franklin. 1997. Harvesting effects on microclimate gradients from small streams to uplands in western Washington. Ecological Applications 7:1188-1200.

Brown, J. H., and A. Kodric-Brown. 1977. Turnover rates in insular biogeography: effect of immigration on extinction. Ecology 58:445-449.

Brown, T. G., and G.F. Hartman. 1988. Contribution of seasonally flooded lands and minor tributaries to the production of coho (Oncorhynchus kisutch) salmon in Carnation Creek, British Columbia. Transactions of the American Fisheries Society 117:546–551.

Burns, J. W. 1972. Some effects of logging and associated road construction on northern California streams. Transactions of the American Fisheries Society 101:1-17.

Cairns Jr., J., J. S. Crossman, K. L. Dickson, and E. E. Herricks. 1971. The recovery of damaged streams. Bulletin of the Association of Southeastern Biologists 18:79- 106.

Caley, M. J., and D. Schluter. 1997. The relationship between local and regional diversity. Ecology 78:70-80.

Campbell, I. C., and T. J. Deoeg. 1989. Impact of timber harvesting and production on streams: a review. Australian Journal of Marine and Freshwater Resources 40.

Capone, T. A., and J. A. Kushlan. 1991. Fish community structure in dry-season stream pools. Ecology 72:983-992.

Clark, J. 1991. Disturbance and population structure on the shifting mosaic landscape. Ecology 72:1119-1137.

Clenaghan, C., P. S. Giller, J. O'Halloran, and R. Hernan. 1998. Stream macroinvertebrate communities in a coniferr-afforested catchment in Ireland: Relationships to physico-chemical and biotic factors. Freshwater Biology 40:175- 193.

Connell, J. H. 1975. Some mechanisms producing structure in natural communities: a model and evidence from field experiments. . Pages 460-490 in M. L. Cody and J. M. Diamond, editors. Ecology and Evolutions of Communities. Belknap, Cambridge.

158 Connell, K. W. 1978. Diversity in tropical rain forests and coral reefs. Science 199:1302- 1310.

Conner, J. V., and R. D. Suttkus. 1986a. Zoogeography of freshwater fishes of the western Gulf slope. Pages 413-456 in C. H. Hocutt and E. O. Wiley, editors. The zoogeography of North American freshwater fishes. Wiley, New York.

Conner, J. V., and R. D. Suttkus. 1986b. Zoogeography of freshwater fishes of the western Gulf Slope of North America. Pages 413-456 in C. H. Hocutt and E. O. Wiley, editors. The Zoogeography of North American Freshwater Fishes. John Wiley and Sons, New York, NY.

Cooper, S. D., S. Diehl, and O. Sarnelle. 1998. Implications of scale for patterns and processes in stream ecology. Australian Journal of Ecology 23:27-40.

Copp, G. H. 1989. The habitat diversity and fish reproductive function of floodplain ecosystems. Environmental Biology of Fishes 26:1–27.

Culp, J. M., and R. W. Davies. 1983. An assessment of the effects of streambank clear- cutting on macroinvertebrate communities in a managed watershed. Canadian Technical Report of Fisheries and Aquatic Sciences No. 1208.

Cushing, C. E., and J. D. Allan. 2001. Streams: Their ecology and life. Academic Press, San Diego, California.

Dahl, J. 1998. Effects of a benthivorous and a drift-feeding fish on a benthic stream assemblage. Oecologia 116:426-432.

Dahl, J., and L. Greenberg. 1996. Impact on Stream benthic prey by benthic vs. drift feeding predators: a meta-analysis. Oikos 77:177-181.

Dahl, J., and L. Greenberg. 1997. Foraging rates of a vertebrate and a inverebrate predator in stream enclosures. Oikos 78:459-466.

Davis, W. S., B. D. Snyder, J. B. Stribling, and C. Stoughton. 1996. Summary of state biological assessment programs for streams and wadeable rivers. EPA 230-R-96- 007, Office of Policy, Planning, and Evaluation, US Environmental Protection Agency, Washington, DC.

Death, R. G., and M. J. Winterbourn. 1995. Diversity patterns in stream benthic invertebrate communities: the influence of habitat. Ecology 76:1446-1460.

Diamond , J. M. 1975. Assembly of species communities. Pages 342-444 in M. Cody and J. M . Diamond, editors. Ecology and Evolution of Communities. Harvard University Press, Cambridge, Massechuttes.

159 Diehl, S. 1993. Effects of habitat structure on resource availability, diet and growth of benthivorous perch, Perca fluviatilis. Oikos 67:403-414.

Downs, P. W., and A. Simon. 2001. Fluvial geomorphological analysis of the recruitment of large woody debris in the Yalobusha River network, Ventral Mississippi, U.S.A. Geomorphology 37:65-91.

Drake, J. A. 1991. Community-assembly mechanics and the structure of an experimental speciaes ensemble. American Naturalist 137:1-26.

Drake, J. A., G. H. Flum, T. Witteman, A. M. Voskuil, C. Hoylman, D. A. Creson, G. R. Kenny, C. S. Huxel, and Larue & J.R. Duncan. 1993. The construction and assembly of an ecological landscape. animal ecology 62:117-130.

Dupouey, J. L., E. Dambrine, J. D. Laffite, and C. Moares. 2002. Irreversible impact of past land use on forest soils and biodiversity. Ecology 83:2978–2984.

Eklov, P., and T. Van Kooten. 2001. Facilitation among piscivorous predators: effects of prey habitat use. . Ecology 89:2486-2494.

Englund, G. 1991. Asymmetric resource competition in a filter-feeding stream insect. Hydropsyche siltalai: Trichoptera Fresh water Biology:26.

Englund, G. 1993. Interactions in a lake oulet stream community: direct and indirect effects of net-spinning caddis larvae. Oikos 66:431-438.

Erman, D. C., J. D. Newbold, and K. B. Reby. 1977. Evaluation of streamside bufferstrips for protecting aquatic organisms. Technical Completion Report No. 165, California Water Resources Center, University of California, Berkeley.

Fausch, K. D., C. Gowan, K. Richmond, and S. C. Riley. 1995. The role of dispersal in trout population response to habitat formed by large woody debris. Bulletin Francais de la Peche et de la Pisciculture 338:179-190.

Fausch, K. D., J. Lyons, J.R. Karr, and P.L. Angermeier. 1990. Fish communities as indicators of environmental degradation. American Fisheries Society Symposium 8:123-144.

Fausch, K. D., C. E. Torgersen, C. V. Baxter, and H. W. Li. 2002. Landscapes to riverscapes: bridging the gap between research and conservation of stream fishes. Bioscience 52:483-498.

Fausch, K. D., and M. K. Young. 1995. Evolutionary significant units and movement of resident stream fishes: a cautionary tale. Pages 360-370 in J. L. Nielsen, editor. American Fisheries Society, Symposium 17, Bethesda, Maryland.

160 Feminella, J. W. 2000. Correspondence between stream macroinvertebrate assemblages and 4 ecoregions of the southeastern USA. Journal of the North American Benthological Society 19:442-461.

Fisher, S. G. 1990. Recovery processes in lotic ecosystems: limits of successional theory. Environmental Management 14:725-736.

Flecker, A. S. 1984. The effect of predation and detritus on the structure of a stream insect community: a field test. Oecologia 64:300-305.

Flecker, A. S. 1992. Fish Trophic guilds and the structure of a tropical stream: weak direct vs. strong indirect effects. Ecology 73:927-940.

Florsheim, J. L., J. F. Mount, and L. T. Rutten. 2001. Effect of baselevel change on floodplain and fan sediment storage and ephemeral tributary channel morphology, Navarro River, California. Earth Surface Processes and Landforms 26:219-232.

Frissell, C. A., W. J. Liss, C. E. Warren, and M. D. Hurley. 1986. A hierarchical framework for stream habitat classification: viewing streams in a watershed context. Environmental Management 10:199-214.

Frost, C. C. 1993. Four centuries of changing landscape patterns in the longleaf pine ecosystem. in S. M. Hermann, editor. 18th Tall Timbers Fire Ecology Conference, Tallahassee, Florida.

Furniss, M. J., T. D. Roelofs, and C. S. Yee. 1991. Road construction and maintenance. American Fisheries Society Special Publication 19:297-323.

Gauch, H. G. 1982. Multivariate analysis in community ecology. Cambridge University Press, Cambridge.

Gerking, S. D. 1959. The restricted movement of fish populations. Biological Review 34:221-242.

Gilliam, J. F., and D. F. Fraser. 2001. Movement in corridors enhancement by predation threat, disturbance, and habitat structures. Ecology 82:258-273.

Gilpin, M. E., and T. J. Case. 1976. Multiple domains of attraction in competition commuinities. Nature 261:40-42.

Glasser, J. W. 1979. The role of predation in shaping and maintaining the structure of communities. American Naturalist 113:631–641.

Glitzenstein, J. S., W. J. Platt, and D. R. Streng. 1995. Effects of fire regime and habitat on tree dynamics in north Florida longleaf pine savannas. Ecological Monographs 65:441-476.

161 Gorman, O. T. 1986. Assemblage organization of stream fishes: the effect of rivers on adventitious steams. American Naturalist 128:611-616.

Gorman, O. T., and J. R. Karr. 1978. Habitat structure and stream fish communities. Ecology 59:507-515.

Gowan, C., and K. D. Fausch. 2002. Why do foraging stream salmonids move during summer? Environmental Biology of Fishes 64:139-153.

Gowan, C., M. K. Young, K. D. Fausch, and S. C. Riley. 1994. Restricted movement in resident stream salmonids: paradigm lost? Canadian Journal of Fisheries and Aquatic Sciences 51:2626-2637.

Gregory, S. V., F. J. Swanson, W. A. McKee, and K. W. Cummins. 1991. An Ecosystem Perspective Of Riparian Zones. Bioscience 41:540-551.

Grossman, G. D., J. F. Dowd, and M. Crawford. 1990. Assemblage stability in stream fishes: a review. Environmental Management 14:661-671.

Gurtz, M. E., and J. B. Wallace. 1984. Substrate-mediated response of stream invertebrates to disturbance. Ecology 65:1556–1569.

Haefner, J. D., and J. B. Wallace. 1981. Production and potential seston utilization by Parapsyche cardis and Diplectrona modesta (Trichoptera : Hydropsychidae) in two streams draining contrasting southern Appalachian watersheds. Environmental Entomology 10:433-441.

Harcup, M. F., R. Williams, , and D. M. Ellis. 1984. Movements of brown trout, Salmo trutta L., in the River Gwyddon, South Wales. Journal of Fish Biology 24:415– 426.

Harding, J. S., E. F. Benfield, P. V. Bolstad, G. S. Helfman, and E. B. D. Jones III. 1998. Stream biodiversity: the ghost of land use past. Proceedings of the National Academy of Sciences USA 95:14843–14847.

Harrell, L. 1992. Handling euthanasia in production facilities. Pages 129 in D. O. Schaeffer, K. M. Kleinow, and L. Krulisch, editors. The care and use of amphibians, reptiles and fish in research. Scientists Center for Animal Welfare, Bethesda, Md.

Harrelson, C. C., C. L. Rawlins, and J. P. Potyondy. 1994. Stream channel reference sites: An illustrated guide to field technique. USDA Forest Service, General Technical Report RM-245, Collins, Colorado.

Harvey, B. C. 1991. Interactions among stream fishes: predator-induced habitat shifts and larval survival. Oecologia 87:29-39.

162 Harvey, B. C., R. J. Nakamoto, and L. J. White. 1999. Influence of large woody debris and a bankfull flood on movement of adult resident coastal cutthroat trout, Oncorhynchus clarki, during fall and winter. Canadian Journal of Fisheries and Aquatic Sciences 56:2161-2166.

Hayek, L. A. C., and M. A. Buzas. 1997. Surveying natural populations. Columbia University Press, New York, New York, USA.

Hilderbrand, R. H., and J.L. Kershner. 2000. Movement patterns of stream-resident cutthroat trout in Beaver Creek, Idaho-Utah. Transactions of the American Fisheries Society 129:1160-1170.

Holling, C. S. 1973. Resilience and stability of ecological systems. Annual Review of Ecology and Systematics 4:1-23.

Hooke, J. M. 1999. Decades of change: contributions of geomorphology to fluvial and coastal engineering and management. Geomorphology 31:373-389.

Hughes, N. F. 1998. Use of whole-stream patterns of age segregation to infer the interannual movements of stream salmonids: a demonstration with Arctic grayling in an interior Alaskan stream. Transactions of the American Fisheries Society 127:1067–1071.

Hughes, N. F., and J. B. Reynolds. 1994. Why do Arctic grayling (Thymallus arcticus) get bigger as you go upstream? Canadian Journal of Fisheries and Aquatic Sciences 51:2154-2163.

Hugueny, B., L. T. deMorais, S. Merigoux, B. deMerona, and D. Ponton. 1997a. The relationship between local and regional species richness: comparing biotas with different evolutionary histories. Oikos 80:583-587.

Hugueny, B., L. T. deMorais, S. Merigoux, B. deMerona, and D. Ponton. 1997b. The relationship between local and regional species richness: comparing biotas with different evolutionary histories. Oikos 80:583-587.

Hunter, M. D., and P. W. Price. 1992. Playing chutes and ladders: heterogeneity and the relative roles of bottom-up and top-down forces in natural communities. Ecology 73:724-732.

Hupp, C. R. 2000. Hydrology, geomorphology and vegetation of Coastal Plain rivers in the south-eastern U.S.A. Hydrological Processes 14:2991-3010.

Huston, M. A. 1979. A general hypothesis of species diversity. American Naturalist 113:81-99.

Huston, M. A. 1994. Biological diversity: the coexistence of species on changing landscapes. Cambridge University Press, Cambridge, UK. 163 Hynes, H. B. N. 1975. The stream and its valley. in Verhandlungen der internationale Vereinigung der theoretische und angewandte Limnologie.

Ilhardt, B. L., E. S. Verry, and B. J. Palik. 2000. Defining riparian areas. Pages 23-42 in E. S. Very, J. W. Hornbeck, and C. A. Doloff, editors. Riparian management in forests of the continental eastern United States, New York.

Ims, R. A., and D. O. Hjermann. 2001. Condition-dependent dispersal. Pages 600-621 in J. Clobert, E. Danchin, A. A. Dhondt, and D. J. Nichols, editors. Dispersal. Oxford University Press, New York.

Jackson, D. A., and H. H. Harvey. 1993. Fish and benthic invertebrates: Community concordance and community-environment relationships. Canadian Journal of Fisheries and Aquatic Sciences 50:2641-2651.

Kaller, M. D., and W. E. Kelso. 2006. Effects of a small-scale clearing on habitat and macroinvertebrates of a southwestern Louisiana coastal bottomland stream. The Southwestern Naturalist 51:143-151.

Karr, J. R., and D. R. Dudley. 1981. Ecological perspective on water quality goals. Environmental Management 5:55-68.

Karr, J. R., and E.W. Chu. 1999. Restoring Life in Running Waters - Better Biological Monitoring. Island Press, Covelo, CA.

Kaufman, P. R., P. Levine, E. G. Robison, C. Seeliger, and D. V. Peck. 1999. Quantifying physical habitat in wadeable streams. EPA/620/R-99/003, U.S. Environmental Protection Agency, Office of Research and Development, Environmental Monitoring and Assessment Program, Washington, DC.

Kelly, J. R., and M.A. Harwell. 1990. ndicators of ecosystem recovery. Environmental Management 14:527-545.

Kerfoot, E. C., and A. Sih, editors. 1987. Predation: Direct and indirect impacts on aquatic communities. University Press of New England, Hanover, London.

Kershner, J. L., C. M. Bischoff, and D. L. Horan. 1997. Population, habitat, and genetic characteristics of Colorado River cutthroat trout in wilderness and nonwilderness stream sections in the Uinta Mountains of Utah and Wyoming. North American Journal of Fisheries Management 17:1134–1143.

Kirk, D. J., and L. A. Smock. 2000. Interspecific and intraspecific interactions between crayfish (Cambarus longulus) and hellgrammites (Corydalus cornutus) and the influence of a predatory fish (Micropterus dolomieu). American Midland Naturalist 144:317–327.

164 Knaepkens, G., L. Bervoets, E. Verheyen, and M. Eens. 2004. Relationship between population size and genetic diversity in endangered populations of the European bullhead (Cottus gobio): implications for conservation. Biological Conservation 115:403–410.

Krebs, C. J. 1999. Ecological methodology. Addison Wesley Longman, New York.

Lammert, M., and J. D. Allan. 1999. Assessing biotic integrity of streams: effects of scale in measuring the influence of land use/cover and habitat structure on fish and macroinvertebrates. Environmental Management 23:257-270.

Lamouroux, N., S. Doledec, and S. Gayraud. 2004. Biological traits of stream macroinvertebrate communities: effects of microhabitat, reach, and basin filters. Journal of North American Benthological Society 23:449-466.

Lancaster, J., A. G. Hildrew, and C. R. Townsend. 1990. Stream-flow and predation effects on the spatial dynamics of benthic invertebrates. Hydrobiologia 203:177- 190.

Larson, G. L., R. L. Hoffman, and S. E. Moore. 2002. Observations of the distributions of five fish species in a small Appalachian stream. Transactions of the American Fisheries Society 131:791-796.

Lenat, D. R., and J. K. Crawford. 1994. Effects of land use on water quality and aquatic biota of three North Carolina Piedmont streams. Hydrobiologia 294:185–199.

Li, J., A. Herlihy, W. Gerth, P. Kaufmann, S. Gregory, S. Urquhart, and D. P. Larsen. 2001. Variability in stream macroinvertebrates at multiple spatial scales. Freshwater Biology 46:87-97.

Linke, S., R. C. Bailey, and J. Schwindt. 1999. Temporal variability of stream bioassessments using benthic macroinvertebrates. Freshwater Biology 42:575- 584.

Lohr, S. C., and K. D. Fausch. 1997. Multiscale analysis of natural variability in stream fish assemblages of a Western Great Plains Watershed. Copeia 1997:706-724.

Lonzarich, D. G., M. R. Lonzarich, and M. L. Warren Jr. 2000. Effects of riffle length on the short-term movement of fishes among stream pools. Canadian Journal of Fisheries and Aquatic Sciences 57:1508–1514.

Lorang, M. S., and F. R. Hauer. 2003. Flow competence evaluation of steambed stability: An assessment of the technique and limitations of application. Journal of the North American Benthological Society 22:475-491.

165 Lotrich, B. A., and W.H. Meredith. 1974. A technique and the effectiveness of various acrylic colors for subcutaneous marking of fish. Transactions of the American Fisheries Society 103:140-142.

Ludwig, J. A., and J. F. Reynolds. 1988. Statistical Ecology: a Primer on Methods and Computing. John Wiley & Sons, Inc., New York.

Macdonald, J. S., MacIsaac, E.A., and Herunter, H. 2003. The effects of variable- retention riparian buffer zones on water temperatures in small headwater streams in sub-boreal forest ecosystems of British Columbia. Canadian Journal of Forest Research 33:1371–1382.

MacDonald, L. H., and E. L. Huffman. 2004. Post-fire soil water repellency: Persistence and soil moisture thresholds. Soil Science Society of America Journal 68:1729- 1734.

Maloney, K. O., P. J. Mulholland, and J. W. Feminella. 2005. Influence of catchment- scale military land use on stream physical and organic matter variables in small southeastern plains catchments (USA) Environmental Management 35:677-691.

Martin, G., C. L. Butler, E. Scott, and J. E. Lyles. 1990. Soil survey of Natchitoches Parish, Louisiana. United States Department of Agriculture, Soil Conservation Service.

Matthews, G. B., R. A. Matthews, and B. Hachmoller. 1991. Mathematical analysis of temporal and spatial trends in the benthic macroinvertebrate communities of a small stream. Canadian Journal of Fisheries and Aquatic Sciences 48:2184-2190.

Matthews, W. J. 1986. Fish faunal structure in an Ozark stream: stability, persistence and a catastrpohic flood. Copeia 1986:388-397.

Matthews, W. J., R. C. Cashner, and F. P. Gelwick. 1988. Stability and persistence of fish faunas and assemblages in three midwestern streams. Copeia 1988:945-955.

McClelland, W. T., and M. A. Brusven. 1980. Effects of sedimentation on the behavior and distribution of riffle insects in a laboratory stream. Aquatic Insects 2:161-169.

McCullagh, P., and J. A. Nelder. 1989. Generalized Linear Models, 2nd edition. Chapman and Hall, London.

McCune, B., and J. B. Grace. 2002. Analysis of ecological communities. MjM Software, Gleneden Beach, Oregon.

McCune, B., and M. J. Mefford. 1999. Multivariate analysis of ecological data. Version 4. MjM Software Design, Gleneden Beach, OR.

166 McIntosh, A. R., and B. L. Peckarsky. 1999. Criteria determining behavinoural reponses to multiple predators by a stream mayfly. Oikos 85:554-564.

McQueen, D. J., M. R. S. Johannes, J. R. Post, T. J. Stewart, and D. R. S. Lean. 1989. Bottom-up and top-down impacts on freshwater pelagic community structure. Ecological Monographs 59:289–309.

Mebane, C. A. 2001. Testing bioassessment metrics: macroinvertebrate, sculpin, and salmonid responses to stream habitat, sediment, and metals. Environmental Monitoring and Assessment 67:293–322.

Meffe, G. K., and A. L. Sheldon. 1988. The influence of habitat structure on fish assemblage composition in southeastern black water streams. American Midland Naturalist 120:225-240.

Meffe, G. K., and A. L. Sheldon. 1990. Post-defaunation recovery of fish assemblage in southeastern blackwater streams. Ecology 71:657-667.

Merritt, R. W., and K. W. Cummins. 1996. An introduction to the aquatic insects of North America, 3rd edition. Kendall Hunt Publishing Company, Iowa.

Meyer, J. L., L. A. Kalpan, D. Newbold, D. L. Strayer, C. J. Woltemade, J. B. Zedler, R. Beilfuss, Q. Carpenter, R. Semlitsch, M. C. Watzin, and P. H. Zedler. 2003. Where rivers are born: The scientific imperative for defending small streams and wetlands. American Rivers, Sierra Club, Washington DC.

Minshall, G. W. 1988. Stream ecosystem theory: global perspective. Journal of the North American Benthological Society 7:263-288.

Moloney, K. A., and S. A. Levin. 1996. The effects of disturbance architecture on landscape level population dynamics. Ecology 72:375-394.

Moyle, P. B., and T. Light. 1996. Boilogical invasions of fresh water: empirical rules and assebly theory. Boilogical Conservation 78:149-161.

Murphy, M. L., and J. D. Hall. 1981. Varied effects of clear-cut logging on predators and their habitat in small streams of the Cascade Mountains, Oregon. Canadian Journal of Fisheries and Aquatic Sciences 38:137–145.

Naiman, R. J. 1998. Biotic Stream Classification. Pages 97-199 in R.J. Naiman and R.E. Bilby, editors. River Ecology and Management. Springer-Verlag, New York.

Naiman, R. J., and H. Decamps. 1997. The ecology of interfaces: riparian zones. The Annual Review of Ecological and Systematics 28:621-658.

Naiman, R. J., and H. Decamps. 1997. The ecology of interfaces: riparian zones. The Annual Review of Ecological and Systematics 28:621-658. 167 Nakano, S., H. Miyasaka, and N. Kuhara. 1999. Terrestrial-aquatic linages: riparian arthropod inputs alter trophic cascades in a stream food web. Ecology 80:2435- 2441.

Nisbet, R. M., E. McCauley, W. S. C. Gurney, W. W. Murdoch, and A. M. de Roos, editors. 1997. Simple representations of biomass dynamics in structured populations. Case Studies in Mathematical Modeling. Prentice Hall.

Noonan, G. R. 1988. Biogeography of North American and Mexican insects, and a critique of vicariance biogeography. Systematic Zoology 37:366-384.

Northcote, T. J. 1997. Potamodromy in Salmonidae - living and moving in the fast lane. North American Journal of Fisheries Management 17:1029-1045.

Nystrom, P., A. R. McIntosh, and M. J. Winterbourn. 2003. Top-down and bottom-up processes in grassland and forested streams. Oecologia 136 596-608.

O’Neill, R. V., D. L. DeAngelis, J. B. Waide, and T. F. H. Allen. 1986. A Hierarchical Concept of Ecosystems. Princeton University Press, Princeton, New Jersey.

Omernik, J. M. 1987. Ecoregions of the conterminous United States. Annals of the Association of American Geographers 77:118-125.

Palmer, M. W. 1993. Putting things in even better order: the advantages of canonical correspondence analysis. Ecology 74:2215-2230.

Pearson, S. M., A. B. Smith, and M. G. Turner. 1998. Forest fragmentation,land use, and cove-forest herbs in the French Broad River Basin. Castanea 63:382–395.

Pecarsky, B. L. 1983. Biotic interactions or abiotic limitations? A model of lotic community structure. in T. D. Fontaine III and S. M. Bartell, editors. Dynamics of lotic ecosystems. Ann Arbor Science, Ann Arbor, Michigan.

Peckarsky, B. L. 1979. Biological interactions as determinants of destributions of benthic invertebrates within the substrates of stony streams. Limnology and Oceanography 24:59-68.

Peckarsky, B. L. 1984. Predator-prey interactions among aquatic insects. Pages 196-254 in V. Resh and D. M. Rosenberg, editors. The Ecology of Aquatic Insects. Praeger, New York.

Peckarsky, B. L., and M. A. Penton. 1990. Mechanisms of prey selection by stream- dwelling stoneflies. Ecology 70: :1203–1218.

Peterson, J. T., and P. B. Bayley. 1993. Colonization rates of fishes in experimentally defaunated warmwater streams. Transactions of the American Fisheries Society 122:199-207. 168 Pianka, E. R. 1994. Evolutionary ecology, 5th edition. Harper Collins, New York, NY.

Piegay, H., and N. Landon. 1997. Promoting ecological management of riparian forests on the Drome river, France. Aquatic Conservation 7:287-304.

Pielou, E. C. 1984. The Interpretation of Ecological Data: a Primer on Classification and Ordination. John Wiley & Sons, Inc., New York.

Poff, N. L. 1992. Why disturbances can be predictable: a perspective on the definition of disturbance in streams. Journal of the North American Benthological Society 11:86-92.

Poff, N. L. 1997. Landscape filters and species traits: Towards mechanistic understanding and prediction in stream ecology. Journal of the North American Benthological Society 16:391-409.

Poff, N. L., J. D. Allan, M. B. Bain, J. R. Karr, K. L. Prestegaard, B. D. Richter, R. E. Sparks, and J. C. Stromberg. 1997. The Natural Flow Regime - A paradigm for river conservation and restoration. BioScience 47:769-784.

Poff, N. L., and J.D. Allan. 1995. Functional organization in stream fish assemblages in relation to hydrologic variability. Ecology 76:606-627.

Poff, N. L., and J. V. Ward. 1989a. Implications of streamflow variaablilty and predictability for lotic community structure: a regional analysis of streamflow patterns. Canadian Jouranal of Fisheries Aquatic & Sciences 46:1805-1818.

Poff, N. L., and J. V. Ward. 1989b. Implications of streamflow variability and predictability for lotic community structure: a regional analysis of streamflow patterns. Canadian Journal of Fisheries and Aquatic Sciences 46:1805-1818.

Poff, N. L., and J. V. Ward. 1990. Physical habitat template of lotic systems: recovery in the context of historical pattern of spatiotemporal heterogeneity. Environmental Management 14:629-645.

Polis, G. A., and D. R. Strong. 1996. Food web complexity and community dynamics. American Naturalist 147:813-846.

Polis, G. A., and K. Winemiller. 1996. Food Webs: Integration of Patterns and Dynamics. Chapman & Hall, New York.

Power, M. E. 1990. Effects of fish in river food webs. Science 250:811-814.

Power, M. E. 1992. Habitat heterogeneity and the functional significance of fish in river food webs. Ecology 73:1675-1688.

169 Pringle, C. M., M. C. Freeman, and B. J. Freeman. 2000a. Regional effects of hydraulic alterations on riverine macrobiota in the new world: tropical-temperate comparisons. Bioscience 50:807-823.

Pringle, C. M., M. C. Freeman, and B. J. Freeman. 2000b. Regional effects of hydraulic alterations on riverine macrobiota in the new world: tropical-temperate comparisons. Bioscience 50:807-823.

Quinn, G. P., and M. J. Keough. 2002. Experimental design and data analysis for biologists. Cambridge University Press, Cambridge, UK.

Rader, R. B., D. P. Batzer, and S. A. Wissinger, editors. 2001. Bioassessment and Management of North American Freshwater Wetlands. John Wiley and Sons, New York, NY.

Rader, R. B., and J. V. McArthur. 1995. The relative importance of refugia in determining the drift and habitat selection of predaceous stoneflies in a sandy- bottom stream. Oecologia 103:1–9.

Rahel, F. J. 1990. The hierarchical nature of community persistence: a problem of scale. American Naturalist 136:328-344.

Rahel, F. J., and W. A. Hubert. 1991. Fish assemblages and habitat gradient in a Rocky Mountain - Great Plains stream: biotic zonation and additive patterns of community change. Transations of Amercian Fisheries Society 120:319-332.

Railsback, S. F., R. H. Lamberson, B. C. Harvey, and W. E. Duffy. 1999. Movement rules for individual-based models of stream fish. Ecological Modelling 123:73-89.

Resh, V. H., A. V. Brown, A. P. Covich, M. E. Gurtz, H. W. Li, G. W. Minshall, S. R. Riece, A. L. Sheldon, J. B. Wallace, and R. C. Wissmar. 1988. The role of disturbance in stream ecology. Journal of North American Benthological Society 7:433-455.

Richards, C., and G. Host. 1994. Examining land use influences on stream habitats and macroinvertebrates: a GIS approach. Water Resources Bulletin 30:729-738.

Richardson, E. V., B. Simons, S. Karaki, M. Mahmood, and M. A. Stevens. 1975. Highways in the river environment: hydraulic and environmental design considerations training and design manual. U.S. Department of Transportation, Federal Highway Administration, Washington, D.C.

Richardson, J. S. 1991. Seasonal food limitation of detritivores in a montane stream: an experimental test. Ecology 72:873- 887.

170 Rieman, B. E., D.C. Lee, and R.F. Thurow. 1997. Distribution, status, and likely future trends of bull trout within the Columbia River and Klamath River Basins. North American Journal of Fisheries Management 17:1111-1125.

Rinne, J. N. 1990. Status, distribution, biology, and conservation of two rare southwestern (U.S.A.) salmonids, the Apache trout, Oncorhynchus apache Miller, and the Gila trout, O. gilae Miller. . Journal of Fish Biology 37:189-191.

Roberts, J. 2000. The influences of physical and physiological characteristics of vegetation on their hydrological response. Hydrological Processes 14:2885- 28901.

Robinson, J. V., and M. A. Edgemon. 1988. An experimental evaluation of the effect of invasion history on community structure. Ecology 69:1410-1417.

Rosenfeld, J. S. 1997. The influence of upstream predation of the expression of fish effects in sownstream patches. . Freshwater Biology 37:535-543.

Rosgen, D. 1996. Applied River Morphology. Wildland Hydrology Pagosa Springs, CO.

Ross, S. T. 2001. Inland fishes of Mississippi. University Press of Mississippi, Jackson.

Sabo, J. L., and M. E. Power. 2002. River–watershed exchange effects of riverine subsidies on riparian lizards and their terrestrial. Ecology 83:1860-1869.

Saucier, R. T. 1994. Geomorphology and quarternary geologic history of the Lower Mississippi Valley. U.S. Army Corps of Engineers, Vicksberg, Mississippi.

Schaefer, J. 2001. Riffles as barriers to interpool movement by three cyprinids (Notropis boops, Campostoma anomalum and Cyprinella venusta). Freshwater Biology 46:379-388.

Schaefer, J. F., E. Marsh-Matthews, D. E. Spooner, K. B. Gido, and W. J. Matthews. 2003. Effects of barriers and thermal refugia on local movement of the threatened leopard darter, Percina pantherina. Environmental Biology of Fishes 66:391-400.

Schlosser, I. J. 1987. A conceptual framework for fish communities in small warmwater streams. Pages 17-24 in W. J. Matthews and D. C. Heins, editors. Community and evolutionary ecology of North American stream fishes. University of Oklahoma Press, Norman, Oklahoma.

Schlosser, I. J. 1995. Dispersal, boundary processes, and trophic level interactions in streams adjacent to beaver ponds. Ecology 76:908-925.

Schlosser, I. J., and P. L. Angermier. 1995. Spatial variation and demographic processes of lotic fishes: conceptual models, imperical evidence, and implications for

171 conservation. Pages 392-401 in J. L. Nielsen, editor. Evolution in the aquatic ecosystem: defining unique units in population conservation.

Shaw, E. A., and J.S. Richardson. 2001. Effects of fine inorganic sediment on stream invertebrate assemblages and rainbow trout (Oncorhynchus mykiss) growth and survival: implications of exposure duration. Canadian Journal of Fisheries and Aquatic Sciences 58:2213-2221.

Sih, A. 1982a. Foraging strategies and the avoidance of predation by an aquatic insect, Notonecta hoffmanni. Ecology 63:786-796.

Sih , A. 1982b. Foraging strategies and the avoidance of predations by an aquatic insect, Notonecta hoffmani. Ecology 63:786-796.

Sih, A., G. Englund, and D. Wooster. 1998. Emergent impacts of multiple predators on prey. . Trends in Ecology and Evolution 13:350-355.

Silsbee, D. G., and G. L. Larson. 1983. A comparison of streams in logged and unlogged areass of Great Smoky Mountains National Park. Hydrobiologia 102:99-111.

Smithson, E. B., and C. E. Johnston. 1999. Movement patterns of stream fishes in a Ouachita Highlands stream: an examination of the restricted movement paradigm. Transactions of the American Fisheries Society 128:847–853.

Sokol-Hessner, L., and O. J. Schmitz. 2002. Aggregate effects of multiple predator species on a shared prey. . Ecology 83:2367-2372.

Soluk, D. A., and N. C. Collins. 1988. Synergistic interactions between fish and stoneflies: Facilitation and interference among stream predators. . Oikos 52:94- 100.

Sousa, W. P. 1984. The role of disturbance in natural communities. Annual Review of Ecology and Systematics 15:335-391.

Sousa, W. P. 1985. Disturbance and patch dynamics on rocky intertidal shores. Pages 101-124 in S. T. A. Pickett and P. S. White, editors. The ecology of natural disturbance and patch dynamics. Academic Press, San Diego.

Stewart, K. W., B. P. Stark, and T. G. Huggins. 1976. The stoneflies (Plecoptera) of Louisiana. Great Basin Naturalist 36:366-384.

Stock, J. D., and I. J. Schlosser. 1991. Short-term effects of a catastrophic beaver dam collapse on a stream fish community. Environmental Biology of Fishes 31:123– 129.

172 Stone, M., and J. Wallace. 1998. Long-term recovery of a mountain stream from clear-cut logging: the effects of forest succession on benthic invertebrate community structure. Freshwater Biology 39:151–169.

Surdick, R. F. 1985. Neartic genera of Chloroperlinae (Plecoptera: Chloroperlidae). Illinois Biological Monographs 54:1-146.

Sutherland, D. G., M. W. Hansler-Ball, S. J. Hilton, and T. L. Lisle. 2002. Evolution of a landslide-induced sediment wave in the Navarro River, California. Geological Society of America Bulletin 114:1036-1048.

Swank, W. T., J. M. Vose, and K. J. Elliot. 2001. Long term hydrologic and water quality responses following commercial clearcutting of mixed hardwoods on a southern Appalachian catchment. Forest Ecology and Management 143:163-178.

Tabacchi, E., L. Lambs, H. Guilloy, A.-M. Planty-Tabacchi, E. Muller, and H. Decamps. 2000. Impacts of riparian vegetation on hydrological processes. Hydrological Processes 14:2959-2976.

Taylor, C. M. 1997. Fish species richness and incidence patterns in isolated and connected stream pools: effects of pool volume and spatial position. Oecologia 110:560–566.

Tebo, L. B. 1955. Effects of siltation resulting from improper logging on the bottom of a small trout stream in the southern Appalachians. Progressive Fish-Culture 17:64– 70. ter Braak, C. J. F. 1986. Canonical correspondence analysis: a new eigenvector technique for multivariate direct gradient analysis. Ecology 67:1167-1179. ter Braak, C. J. F., and P. Smilauer. 2002. CANOCO reference manual and CanoDraw for Windows user's guide: software for canonical community ordination (version 4.5). Microcomputer Power, Ithaca, New York, USA.

Toepfer, C. S., W. L. Fisher, and J. A. Haubelt. 1999. Swimming performance of the threatened leopard darter in relation to road culverts. Transactions of the American Fisheries Society 128:155-161.

Tonn, W. M. 1990. Climate Change and Fish Communities - a Conceptual-Framework. Transactions of the American Fisheries Society 119:337-352.

Townsend, C. R. 1989. The patch dynamics concept of stream community ecology. Journal of the North American Benthological Society, 8:36-50.

Townsend, C. R., M. R. Scarsbrook, and S. Doledec. 1997. The intermediate disturbance hypothesis, refugia, and biodiversity in streams. Limnology and Oceanography 42:938-949. 173 Turner, M. G., S. M. Pearson, P. Bolstad, and D. N. Wear. 2003. Effects of land-cover change on spatial pattern of forest communities in the Southern Appalachian Mountains (USA). Landscape Ecology 18:449-464.

Urban, D. L., R. V. O'Neill, and H. H. Shugart Jr. 1987. Landscape ecology: a hierarchical perspective can help scientists understand spatial patterns. Bioscience 37:119-127.

Vance-Chalcraft, H. D., and D.A. Soluk. 2005. Multiple predator effects result in risk reduction for the prey across multiple prey densities. Oecologia 144:472-480.

Vannote, R. L., G. W. Minshall, K. W. Cummins, J. R. Sedell, and C. E. Cushing. 1980. The river continuum concept. Canadian Journal of Fisheries and Aquatic Sciences 37:130-137.

Vinson, M. R., and C. P. Hawkins. 1998. Biodiversity of stream insects: variation at local, basin, and regional scales. Annual Review of Entomology 43:271-293.

Vuori, K. M., and I. Joensuu. 1996. Impact of forest drainage on the macroinvertebrates of a small boreal headwater stream: do buffer zones protect lotic biodiversity? Biological Conservation 77:87-95.

Wallace, J. B. 1990. Recovery of lotic macroinvertebrate communities from disturbance. Environmental Management 14:605-620.

Wallace, J. B., and R. W. Merritt. 1980. Filter-feeding ecology of aquatic insects. Annual Review of Entomology 25:103-132.

Wallace, J. B., S. L. Eggert, J. L. Meyer, and J. R. Webster. 1997. Multiple trophic levels of a forest stream linked to terrestrial litter inputs. Science 277:102–104.

Ward, A. D., and S. W. Trimble. 2004. Environmental Hydrology, 2nd edition. A CRC Press Company, Washington D.C.

Ward, J. V. 1989. The four-dimentional nature of lotic ecosystems. Journal of North American Benthological Society 8:2-8.

Warren Jr, M. L., P. L. Angermier, B. M. Burr, and W. R. Haag. 1997. Declines of a diverse fish fauna: patterns of imperilment and protection of the southeastern United States. Pages 105-164 in G. W. Benz and D. E. Collins, editors. Aquatic fauna in peril: the southeastern perspective. Lenz Design in Communications, Decatur, GA.

Warren Jr, M. L., and M. G. Pardew. 1998. Road crossings as barriers to small-stream fish movement. Transactions of the American Fisheries Society 127:637-644.

174 Warren Jr., M. L., P. L. Angermier, B. M. Burr, and W. R. Haag. 1997. Declines of a diverse fish fauna: patterns of imperilment and protection of the southeastern United States. Pages 105-164 in G. W. Benz and D. E. Collins, editors. Aquatic fauna in peril: the southeastern perspective. Lenz Design in Communications, Decatur, GA.

Warren Jr., M. L., and M. G. Pardew. 1998. Road crossings as barriers to small-stream fish movement. Transactions of the American Fisheries Society 127:637-644.

Waters, T. F. 1995. Sediment in streams: sources, biological effects, and control. Bethesda, Maryland, USA.

Webster, J. R., and J. B. Waide. 1982. Within-stream factors affecting nutrient transport from forested and logged watersheds. Freshwater Biology 12:331-341.

Wheeler, A. P., P. L. Angermeier, and A. E. Rosenberger. 2005. Impacts of new highways and subsequent landscape urbanization on stream habitat and biota. Reviews in Fisheries Science 13:141-164.

Wiens, J. A. 2001. The landscape contents of dispersal. Pages 97-109 in J. Clobert, E. Danchin, A. A. Dhondt\, and J. D. Nichols, editors. Dispersal. Oxford University Press, New York.

Wiens, J. A., J. T. Rotenberry, and B. van Horne. 1986. A lesson in the limitation of field experiments: shrubsteppe birds and habitat alteration. Ecology 67:365-376.

Wilbur, H. M., and J. E. Fuath. 1990. Experimental aquatic food webs: interactions between two predators and two prey. . American Naturalist 135:176-204.

Williams, L. R., T. H. Bonner, J. D. Hudson III, M. G. Williams, T. R. Leavy, and C. S. Williams. 2005. Interactive efects of environmental variability and military training on stream biota of three headwater drainages in Western Louisiana. Transactions of the American Fisheries Society 134:192-206.

Williams, L. R., C. M. Taylor, M. L. Warren Jr, and J. A. Clingenpeel. 2002. Large-scale effects of timber harvesting on stream systems in the Ouachita Mountains, Arkansas, USA. Environmental Management 29:76-87.

Williams, L. R., C. M. Taylor, M. L. Warren Jr, and J. A. Clingenpeel. 2004. The role of regional factors in structuring Ouachita Mountain stream assemblages. in J.M. Guldin, editor. Ouachita and Ozark Mountains symposium: ecosystem management research. USDA Forest Service, Southern Research Station. Gen Tech Rep. SRS-74, Asheville, North Carolina.

Williams, L. R., C. M. Taylor, M. L. Warren Jr., and J. A. Clingenpeel. 2003. Environmental variability, historical contingency, and the structure of regional

175 fish and macroinvertebrate faunas in Ouachita Mountain stream systems. Environmental Biology of Fishes 67:203-216.

Williams, L. R., C. S. Toepfer, and A. D. Martinez. 1996. The relationship between fish assemblages and environmental gradients in an Oklahoma prairie stream. Journal of Freshwater Ecology 11:459-468.

Wondzell, S. M., and J. G. King. 2003. Postfire erosional processes in the Pacific Northwest and Rocky Mountain regions. Forest Ecology and Management 178:75-87.

Wood, A. L., A. Simon, P. W. Downs, and C. R. Thorne. 2001. Bank-toe processes in incised channels: the role of apparent cohesion in the entrainment of failed bank material. Hydrological Processes 15:39-61.

Wooster , D. 1994. Predatore impacts on stream benthic prey. Oecologia 99:7-15.

176