.- “lulu-"v .u "an un- . u M u m- ..; WWW»), ‘ w J , . ‘. . n “5:..1'391. ”-11 ., {£3 . " 12W)" m’l’l 1:3".3 .sJ:J:a1', ' {a J We ”1" [a ' ' ”immiur 1’” ‘ 45W“: 1.3‘. 35.1-51-9141‘?‘ =- u.““.§f".'1{:1fl“'fif ‘ ”gzlfixtlfi'fpsgpi? , ml‘ e4:':-L‘a'-H- 'W‘: 4521.1” . ,2. .-- . LEV-15:93 ; ‘ {émzfiignm n gt. 3.. ._- IQ..- ‘hli-‘V'J 1‘ O I ~ ”:03“ “"..£"“'3'JL." - . «0 ’um ~ 1.151313! 1' ’5‘ .- 41:! g: ' 1" 1 Hr. . .~ .6?» 33’5” ... 3-.“ :3? "‘9‘ V" .4.- e.-. '37: ii“. p '5 5“: :‘m: 4‘ My w. 3.":IL‘_11'.!5:'1‘ | .t 14.)}??? 1 1144‘?- . , '. “5:1 ‘4?!th . : ‘ I . ’Tta‘fi‘fi! . :::!"[h‘.ill .' ‘51:}?! '5‘ ,t‘. w ‘. 31,41-3 n-zw ., ”6711a" ,- .. «a... .m‘ f? x z..:' ‘ V . .. . . ....- -: , "'51:" , . 4:“. '0‘”. .._,__n c -‘ .‘. V ~ .« _—‘ i .A. 4% "7'..- 7.43 r (.1 , . -,.. _ v.42...” w .. ‘;.-’. .. ~ .-..._- . M... o r .¢- ,h a..- poo . '1 "‘ v1 .. WEN-.19 m “2‘1; 3 3369“,: 1.. "1.3 . A" ‘1. ‘1 g. 4} 'U’ 1'“ "‘l ‘t 3 up“! . ‘ - . t‘méiih i. ...I“ aim ' 4%sz 'l v 1".- K"! , ‘I' ~ ‘ I (I. . . ¢.... ,a;a=- .. w...\....._"\ 't. L ‘ I V . ' . ,".’_ ‘V..- «I -. .v... 7' -._..‘ F‘s -.. . .‘4- -no— ,2' .:‘ 15:.“ . 1 I 1x; 9 l .1 A ' .., . . .. . ‘. “m - f“ ... . i’é'éh .. .‘fi'l'.’ no 2:331 Fr 1.. -Av "c. c o _|- .4 “.2: i “ I. .n In ! my h I ' A ‘ I u,\ u' ' "I‘fiagm "w “H": . 1 - .. a .O . 4 ‘v _ a. r - '_ ' .n V . j g,‘_..a. . ‘ . ' , W-..«-._ 2 r.‘~ él' '.9 .',u‘ t. 5‘). , . _ . _ ,zépfi‘ffil' ‘ . I: g '. <. . , . .A , ,~ > .: '2‘ a ‘l‘, g ”t: 5: fI .‘ I 1“" 3d)” ‘ . ' ' L 0 . - . .‘ ‘ . . ‘fi‘ifilfirxflfitfifl \. 7.“?E‘1L'r'” ' 3 : lnu‘l 'L W” v y- . .4. ca ,. 4-,‘ ,__V .—‘. .v‘ uo 4 ,. h". ._: ,‘ | 4-33.. _ . 1 ' , qr". ‘ .W‘; 0 O l P. ..—. "43? w.’ 1'“ A. ‘ O THESIS It{I'lllljiflllil‘ilfllflllllfllljifiII LIBRARY Michigan State University This is to certify that the dissertation entitled Some effects of riparian habitat alteration on lotic invertebrate ecology presented by Roger Malcolm Strand has been accepted towards fulfillment of the requirements for Ph .D degree in Entomology Major prof ssor Date _lJ_Sep_t_emb_er_._l_9_9_6 MS U is an Affirmative Action/Equal Opportunity Institution 0-12771 «r—«r A — 4.. '—-—- PLACE It RETURN BOX to remove thte checkout from your record. TO AVOID FINES return on or before date due. DATE DUE DATE DUE DATE DUE MSU le An Affirmative Action/Equal Opportunity lnetltulon W ans-9.1 SOME EFFECTS OF RIPARIAN HABITAT ALTERATION ON LOTIC INVERTEBRATE ECOLOGY By Roger Malcolm Strand A DISSERTATION Submitted to Michigan State University in partial fulfillment of the requirements for the degree of DOCTOR OF PHILOSOPHY Department of Entomology 1996 ABSTRACT SOME EFFECTS OF RIPARIAN HABITAT ALTERATION ON LO'TIC INVERTEBRATE ECOLOGY By Roger Malcolm Strand Three studies were conducted to determine the effects of allogenic alteration of riparian habitat on benthic invertebrate ecology. Study 1. Community abundance, richness, and diversity were monitored during two radically different landscape alterations which occurred simultaneously in a northern Michigan watershed. This three-year study incorporated analyses of introduced substrata samples taken from an upstream forested reach undergoing the transition from unrestricted flow to beaver pond and from a perennially overgrazed pasture reach experiencing the initial effects of cattle exclusion. Beaver colonization and cattle exclusion both resulted in initial abundance increases. Recovery from overgrazing led to increased diversity while disturbance by beaver resulted in decreased diversity. Impoundment produced large changes in community structure and functional composition. In contrast, eattle exclusion did not meaningfully affect community composition. Disturbance by beaver fostered increased abundance of broadly distributed habitat generalists while recovering and undisturbed habitat harbored less common taxa with specialized habitat requirements. The forest reach supported many taxa not found in the overgrazed pasture reach, but the pasture reach contained few taxa not found in the forest reach. Therefore, losses of forest reaches that result from human activities may have relatively large impacts on watershed biodiversity. Study 2. The initial response by predacious stonefly nymphs (Paragnetina media) to intense sedimentation was examined in field trials with chambers designed to give nymphs the choice to stay or reposition in response to sediment additions. The results indicate that the often posited immediate-escape response to the onset of intense sedimentation may not be commonly enacted by stonefly nymphs or perhaps by other relatively immobile, lotic insects. Study 3. Net-spinning caddisflies often thrive in the heavily sedimented waters of midwestem agricultural catchments despite the presumed costs of sedimentation-induced net fouling. A laboratory experiment was conducted to determine whether larval growth and survival of Hydropsyche bettem' and Ceratopsyche spama are affected by intense, episodic sedimentation. Larvae of both species suffered increased mortality in sedimentation treatments relative to controls and taxa were differentially affected as H. betteni significantly outperformed C. spama. Sedimentation did not, however, alter the relative growth rate of either species. ACKNOWLEDGEMENTS D]. N guyen's consistency and amazing technical ability made the study presented in Chapter 2 far more accurate and interesting than it would have been had she not been dedicated to its completion. I thank Bill Morgan for helping me with all aspects of my research. Roger W. Strand worked tirelessly in the field during his annual visits to Carlson Creek, he assisted in the design and development of the apparatus used in the Chapter 3 study, and he also contributed some fantastic illustrations that I have used when presenting this research, one of which appears on page 13. I gratefully acknowledge the guidance provided by my advisory committee: Drs. W.T. Cooper, D]. Hall, L. Hedin, D. long, R.J. Stout, E.D. Walker and particularly that offered by my major advisor, Dr. Richard W. Merritt. Dr. Matt Ayres' much-appreciated consultation on mtters concerning data analysis, interpretation, and presentation vastly improved this dissertation. The Department of Entomology and the Center for Integrated Studies in Science provided me with several great opportunities to develop my teaching skills. The Center for Water Resources provided research funds. I thank the staffs of Department of Entomology the Center for Integrated Studies in Science for their professionalism and enthusiastic assistance. I also thank my colleagues from the Merritt Lab and Dr. Dan Herms for their advice and assistance in the field and laboratory. Finally, I am proud to have been part ofthe greatest ofall 1M sports franchises, the DINOSORES. TABLE OF CONTENTS List of Tables List of Figures Chapter 1. Benthic invertebrate ecology in Ameriea's pasture- and rangeland streams Literature Cited Chapter 2. Benthic invertebrate response to cattle exclusion and beaver colonization in a small, Michigan stream Overview Introduction Methods Results Discussion Summary Literature Cited Appendix 1. Forest-reach taxa organized by feeding guild and year of collection. letters after taxa names indicate sample type(s) from which animals were collected: R=rock; W=wood; L=leaf packs; D=drift ........ Appendix 2. Pasture-reach taxa organized by feeding guild and year of collection. Letters after taxa names indieate sample type(s) from which animals were collectect R=rock; W=wood; L=leaf packs; D=drift ----- Chapter 3. The behavioral response of the stonefly Parugnetina media (Plecoptera: Perliche) to the onset of intense sedimentation Introduction Methods Page iv H 14 14 15 20 24 31 36 52 53 55 55 57 Results Discussion Literature Cited Chapter 4. Effects of episodic sedimentation on the net-spinning caddisflies Hydropsyche betteni and Ceratopsyche spama (Trichoptera: Hydropsychidae) Introduction Methods Results Discussion Limture Cited Appendix 1. Contingency analyses testing whether Hydropsyche betteni and Ceratopsyche spama survival differed in sediment treatment and controls with pupae included Appendix 1. Record of deposition of voucher specimens Appendix 1.1. Voucher specimen data 59 59 61 68 68 7 l 73 74 77 87 88 89 LIST OF TABLES Chapter 2 Table 1. Baseline abiotic data collected from forest- and pasture-reach sites during 1991. Values indicate means i: 1 se Table 2. Carlson Creek invertebrate tolerance-index values (from Resh et al 1996). * = values estimated for this study ------------------ Table 3. AN OVA results comparing invertebrate community characteristics as estimated with rock, wood, and leaf pack samples Table 4. Contingency analyses testing whether the proportion of animals representing different functional feeding groups varied between reaches in rock, wood, and leaf-pack samples Chapter 3 Table 1. ANOVA results comparing the number of stonefly nymphs that selected each behavioral option Table 2. Particle-size distribution and chemical composition of bank sediments used in sediment treatment Chapter 4 Table 1. Particle—size distribution and chemieal composition of bank sediments used in sediment treatments Table 2. ANOVA results comparing relative growth rates of Hydropsyche betteni and Cenopsyche spama in two sediment treatments and two controls iv 41 42 43 65 81 82 Table 3. Contingency analyses testing whether Hydropsyche betteni and Ceratopsyche spama survival differed in sediment treatments and controls (two tanks each) 83 LIST OF FIGURES Chapter 1 Figure 1. Some effects on lotic ecology of grazing cattle 1n small- stream riparian zones Chapter2 Figuee 11.99Tgre pasture reach of Carlson Creek in (A) 1991, (B) 1992, and ( ) Figure 2.99Tg1e forest reach of Carlson Creek in (A) 1991, (B) 1992, and (C) 1 ~ Figure 3. Mean depth and current velocity :t 1 se in one forest—reach and two pasture-reach stations. Measurements taken during July, August, and September, 1991-1993 Figure 4. Suspended solid concentration from forest— and pasture-reach water samples taken monthly, July- September, 1991- 1993.... .... ....... Figure 5. Mean invertebrate abundance, richness, and diversity 1n rock, wood, and leaf pack sammes Figure 6. Community functional-fading group composition in forest-reach samples. Values indicate total number of animals collected. Functional-feeding groups that comprised less than 1% of total are not represented on pie-charts. These are: 3 herbivores in 1992 wood samples and 5 scrapers in 1992 leaf-packs Figure ’7. Community functional- —feeding group composition in pasture-reach samples. Values indicate total number of animals collected. Functional-feeding groups that comprised less than 1% of total are not represented on pie-charts These are: 2, 9, and 9 scrapers in 1991, 1992, and 1993 wood samples respectively; 1 herbivore in 1991 wood samples; and 6 scrapers in 1991 leaf-pack samples Chapter 3 Figure 1. A sedimentation-response chamber Figure 2. Mean number of stoneflies (2t 1 se) that chose each behavioral opfion. 13 45 46 47 48 49 50 51 67 Chapter 4 Figure 1. Mean NTU :tl se in two sediment-treatment and t wo control tanks 84 Figure 2. Figure 2. Box plots of mean RGR :l: 1 SE of Hydropsyche betteni and Ceratopsyche spama in two sedimented and two control tanks. Mean pre-trial dry mass (i 1 se) indieate relative size of fifth-instar H. bettem' (n=24) and C. spama (n=21) 85 Figure 3. Mean survivorship of larval Ceratopsyche spama and Hydropsyche betteni in two sedimented and two control which is 0.020 ... 86 BENTHIC INVERTEBRATE ECOLOGY TN AMERICA'S PASTURE- AND RANGELAND STREAMS Erosion eats into our hills like a contagion, and floods bring down the loosened soil upon our valleys like a scourge. Water, soil, animals, and plants - the very fabric of prosperity - react to destroy each other and us. Science can and must unravel those reactions, and government must enforce the findings of science (Leopold 1923). Seventy-three years have passed since Aldo Leopold issued this advisory to future ranchers, ecologists, and policy makers and today almost one-half of Earth's terrestrial space is grazed by domestic animals (Goude 1994). After more than a century of large- scale grazing in the United States, overgrazed pastures and range have become common features on the rural landscape (Fleischner 1994, Wissmar et al. 1994, Waters 1995). Cattle grazing is by far the most common way to use land the American west, where more than seventy percent of the landscape is grazed Most of this land is held in public trust and as evidence of ever-increasing habitat damage mounts, policy makers are being pressured by scientists and concerned citizens to restrict grazing in particularily sensitive habitats (Armour 1991, 1994; Fleischner 1994). There is general agreement among range managers and ecologists that many currently grazed habitats, for example alpine meadows (Kondolf 1994), can not sustain conflicting societal demands for maximized cattle production and optimized water- and wildlife-habitat quality ( Fleischner 1994, Li et al. 1994, Brown and McDonald 1995, Mosely et al. 1993). Despite this reality, grazing on public lands continues with little regulatory recognition of differential habitat sensitivity, even as restoration costs often far outweigh fiscal inputs from grazing fees (Minshall 1989, Fleischner 1994, Kondolf 1994). It is well-established that habitat types vary in nature and magnitude of response to cattle grazing (Milchunas and Lavenroth 1993, Li et a1. 1994). For example, plant 2 species richness in grasslands within the historic range of bison ofien increases with cattle grazing intensity, whereas introduction and intensification of cattle grazing in grasslands outside of the historical range of bison or other large herbivores is known to cause plant richness to sharply decline (Milchunas and Lavenroth 1993, Hofstede 1995). Cattle, unlike bison, spend a disproportionate amount of time feeding and wallowing in streams. This universal intensification, particularily during hot weather, renders riparian areas in all grazed habitat types especially vulnerable to overgrazing (Kauffman and Krueger 1984, Resh et al. 1988, Armour et al. 1991, 1994, Platts 1991, Fleischner 1994, Waters 1995). Most of the concern over grazing-induwd riparian and instream habitat damage has been centered on the consequent reductions of game-fish production that typically occur in heavily grazed catchments (Armour et al. 1991, 1994, Fleischner 1994). There is still, therefore, much left to learn about grazing effects on the numerically (Hynes 1970) and often energetically dominant animals (Waters 1984) that comprise lotic invertebrate communities. This chapter provides an overview of cattle-grazing effects on instream and riparian- zone habitat that are known or suspected to affect benthic invertebrate ecology. Subsequent chapters include an analysis of management-induced recovery of a perennially overgrazed, northern-Michigan stream and two studies conducted to fill voids in understanding of grazing-induced sedimentation effects on aquatic insect behavior, growth, and survival. General efiects of livestock grazing on stream ecology Overgrazing is a contemptuous term that remains undefined by range managers (Fleischner 1994), but has nonetheless been determined to be the cause a variety of negative effects in aquatic ecosystems (Minshall et al. 1989, Fleischner 1994, Waters 1995) (Figure 1). Overgrazing (sensu Ieopold 1923) in riparian areas is indicated by total vegetative denudation along banks and channels which results in increased erosion (Kauffman et a1. 1983 b, Gamougoun et al. 1984), sedimentation, eutrophieation (Odion et al. 1988, Waters 1995), and thermal regime variation (Li et al. 1994) (Figure l). Benthic invertebrates are sensitive to fluctuations in all of these environmental characteristics both directly through modification of physiological processes (Hynes 1966, 1970; Sweeny 1993) and indirectly through alteration of substrata characteristics (Kauffman and Krueger 1984, Waters 1995). Efiiects of riparian vegetation transformation Intense grazing pressure in riparian forests results in drastically diminished tree regeneration in aging pastures. This process typically ultimates in transformation of vegetative community from forest to grassland (Fausch and Bramblett 1991, Li et al. 1994). The consequently reduced inputs of abscised leaves and woody debris to streams causes the elimination food and shelter for a diverse array of benthic invertebrates (Merritt and Cummins 1996) and fish (Hynes 1970). Stream structural and biogeochemical characteristics are also altered by reduction of debris-dam habitat as habitat heterogeneity is lessened (Sedell 1990), and important sites for carbon (Meyer et al. 1988) and phosphorus dynamics (Munn 1989) are eliminated. The structural changes associated with riparian-zone denudation may also remove cues necessary for habitat recognization by aerial adult aquatic insects on mate-location or dispersion flights. The shift in dominance between two black fly species was demonstrated by Tim (1994) to be an example of this type of differential habitat selectivity by ovipositing females. After timber was cleared from a reach of a Rhine River tributary, the opportunistic Simulium omatum rapidly displaced S. vernum as the dominant black fly in the stream. There were no changes in substrata, water chemistry, or resource availability to account for the differences. However these species do behave differently when seeking oviposition sites. Simulium vemum prefers shaded riffles for egg deposition, whereas S. omatum only oviposits in riffle exposed to full sun. Because many other aquatic insects use physical cues when seeking oviposition sites (Wallace and Anderson 1996), it is probable that discovery awaits other such direct effects of grazing-mediated riparian transformation on benthic community composition. Efii'cts of increased erosion Bank sediments exposed by overgrazing erode into overland flow following spates and rapid snowmelt. These processes immediately elevate concentrations of suspended solids (Waters 1995), nitrogen, and phosphorus (Mosely et al. 1993) and eventually blanket stream beds with deposited sediments (Waters 1995). Substrata burial can be lethal to benthic invertebrates in extremely high levels (Thomas 1985) or in sensitive life stages such as immobile pupae (Rutherford and Mackay 1986). Benthic invertebrate communities often experience functional and taxonomic changes in response to persistent sedimentation. Typical transitions feature the replacement of animals that require solid substrata with those that thrive in thick deposits of soft substrata (Nuttall and Bielby 1973, Quinn et al. 1992, Waters 1995). For example, Strand and Merritt (Chapter 2) found that the relative proportional abundance of filter- feeders to omnivorous gatlrerers, a measure of particulate transport versus deposition and of the availability of solid substrata (Merritt and Cummins 1996), was far greater in an undisturbed reach of a northern Michigan stream than it was in an overgrazed reach. Habitat simplifieation through substrata burial also is likely to reduce hyporheic-zone quality. This sub—flow region is saturated with a mixture stream and ground water and is inhabited a variety of invertebrates, some of which are hyporheic enderrrics and others that forage and seek refuge from disturbance in hyporheic habitat (Williams 1984, Meyer et a1. 1988, Dahm and Valett 1996). As the habitat of occasional and permanent hyporheic residents is vanquished by interstitial filling, biogeochenrical activity at nutrient upwelling sites is also diminished (Grimm et al. 1991, Findlay 1995, Dahm and Valett 1996, Grimm 1996). This process imposes limits on stream metabolism through alteration of nutrient cycling rates (Grimm and Fisher 1984). Grazing-induced hyporheic burial is expected to be particularily destructive to the benthic ecology of temporary streams where population persistence for many invertebrates is dependant upon the survival of individuals that wait out no-flow periods in damp, subchannel interstitial spaces (Fisher and Gray 1983). Effects of excrement input Due to the periodic threat of bacterial pollution of drinking water, cattle excrement input to streams, perhaps more than any other grazing-related problem, concerns to humans who live downstream from grazed riparian areas. High loads of excrement input are also suspected to impose direct threats to aquatic life. Foremost among these is the potential of nutrient enrichment to affect instream biogeochemical processes (Mosely et al. 1993, Harris et a1 1994). In seasonally grazed streams, eutrophication that results from waste-mediated fertilization, principally through increased phosphorous loading (Allen et al. 1982, Mosely et al. 1993), can drastically reduce dissolved oxygen concentration (Harris et a1. 1994) and increase algal and macrophyte production (Fleischner 1994). Persistent 5 fertilization and selective herbivory by cows often result in the formation of dense instream accumulations of unpalatable plants such as Cladophora glommerata (Armour et al. 1991, 1994, Matthews et al. 1994), a widespread filamentous green alga, mats of which typically contain few, if any, benthic macroinvertebrates (personal observation). Substrata domination and benthic shading that result from the presence of thick algal growths may also reduce benthic secondary productivity by limiting the establishment of the more readily foraged-on five-kingdom amalgam known as biofilm or periphyton which coats most exposed substrata (Cumnrins 1973, Lamberti and Feminella 1996). Extremely high levels of excrement input in seasonally grazed streams may act to further slow instream secondary productivity through ammonia toxicity to invertebrates and fish. Although ammonia levels rarely exceed the tolerance threshold of most pasture-stream inhabitants (Overcash et al. 1983), particularily sensitive organisms may be displaced due to chronic, ammonia-induced damage to gill membranes (Hazel et al. 1979, DeGraeve et al. 1980). Acute ammonia toxicity (1.2-8 mg/l) (Hazel et al. 1979, DeGraeve et al. 1980) is also a potential problem associated with heavy grazing, especially during dry, hot weather when stream flow is low and cattle wallow to alleviate heat stress. Efikcts of thermal regime alteration Although cattle grazing does not universally cause pronounced elevations in stream temperature (Li et al. 1994), thermal regimes are known to vary more in grazed than in ungrazed streams (Kauffman and Krueger 1984). Li et al. (1994) reported that insolation increased with increasing grazing activity as did algal and invertebrate biomass. These conditions combined to render grazed habitat less amenable to trout as increased temperature caused a decline in the ratio of trout to invertebrate biomass. This relationship has been detected elsewhere and has been in part attributed to direct heat stress to trout (Platts and Rinne 1985, Platts and Nelson 1989, Minshall et. al. 1989) and partly to elevated temperature—mediated increases in the abundance of less-preferred invertebrate prey (Tait et al. 1994). Sweeny (1993) documented 2 to 4°C increases attributable to clearing of northeastern U.S. forest streams. This increase may seem trivial when compared to daily and seasonal temperature fluctuations in terrestrial realms; however, in the relatively constant lotic thermal environments, 2-4'C can be a large-enough difference to affect the physiology and alter distributions of sensitive species (Sweeny and Vannote 1986). In streams where grazing elevates insolation, consequent restrictions on intolerant species and concurrent increases in secondary production by thermally insensitive species (Murphy and Hall 1981, Murphy et al. 1981, Hawkins et al. 1983, Bilby and Bisson 1987) are therefore expected to result in changes in benthic community composition and bioenergetics. Riparian-forest clearing may also affect the metabolism and reproduction of aerial adult aquatic insects through thermal regime alteration. Tree removal eliminates shaded resting structure which is utilized extensively, and perhaps required by, adult aquatic insects. Most holometabolous aquatic insects rely on resources stored during the larval period to complete their life cycles (Sweeny 1993). Therefore, resource conservation prior to reproduction may have positive effects on fecundity. Reduction or elimination of cool, humid resting habitat could thus limit the reproductive potential of sensitive species. Aldo Ieopold warned us long ago that overgrazing of livestock can negatively affect ecological processes in pastured watersheds. Although there is still relatively little known about the specific effects of grazing on lotic animals, it is certain that the physical impacts of grazing and trampling can combine to alter channel hydrology, vegetative composition, and nutrient-retention capacity. The ultimate synergy of these effects produces simplified lotic ecosystems, less buffered from climatic disturbance, undergoing heightened stream-water losses of nutrients. This distressing scenario is far too common in American watersheds where short-term economic gains are commonly sought despite the glaring portent of ecological calamity. LITERATURE CITED Allen, L.H., Jr., J.M. Ruddell, G.J. Ritter, F.E. Davis, and P. Yates. 1982. Land use effects on Taylor Creek water quality. Proceedings of the Special Conference on Environ mentally Sound Water and Soil Management, Orlando, FL. American Society of Civil Engineering, New York. Armour, C.L., D.A. Duff, and W. Elmore. 1991. The effects of livestock grazing on riparian and stream ecosystems. Fisheries 16: 7-11. Armour, C.L., D.A. Duff, and W. Elmore. 1994. The effects of livestock grazing on western riparian and stream ecosystems 19: 9-12. Bilby, RE, and RA. Bisson. 1987. Emigration and production of hatchery coho salmon (Oncorhynchus kisutch) stocked in streams draining an old-grth and clearcut watershed. Canadian Journal of Fisheries and Aquatic Sciences 44: 1397-1407. Brown, J.H. and W. McDonald. 1995. Livestock grazing and conservation on southwestern rangelands. Conservation Biology 9: 1644-1647. Cummins, K.W. 1973. Trophic relations of aquatic insects. Annual Review of Entomology 18: 183-206. Dahm, C.N., and M. Valett. Hyporheic zones. Pages 107-119 in PR Hauer and GA. Lamberti. MW. New york: Academic Press. DeGraeve, G.M., R.L. Overcash, and H.L. Bergman. 1980. Toxicity of underground coal gasification condenser water and selected constituents to aquatic biota. Archives of Environmental Contamination Toxicology 9: 543- 555. Fausch, K.D., and R.G. Bramblett. 1991. Disturbance and fish communities in intermittent tributaries of a western Great Plains river. Copeia 1991: 659-674. Findlay, S. 1995. Importance of surface-subsurface exchange in stream ecosystems: the hyporheic zone. Limnology and Oceanography 40: 159-164. Fisher, 8.6., and L]. Gray. 1983. Secondary production and organic matter processing by collector macroinvertebrates in a desert stream. Ecology 64: 1217-1224. Fleischner, TL. 1994. Ecological costs of livestock grazing in western North America. Conservation Biology 8: 629-644. Gamougoun, N.D., R.P. Smith, M.K. Wood, and RD. Pieper. 1984. Soil, vegetation, and hydrologic responses to grazing management at Fort Stanton, New Mexico. Journal of Range Management 37: 538-541. Goude, A. 1994. The Hum Impact on the NM! Environmmgl, fomh edition. Cambridge, MIT Press. Grimm, N.B. 1996. Surface-subsurface interactions in streams. Pages 625-646 in ER. Hauer and GA. Lamberti. Methods in Stream Ecology. New york: Academic Press. Grimm, N.B., and S.G. Fisher. 1984. Exchange between interstitial and surface water: implications for stream metabolism and nutrient cycling. Hydrobiologia 111: 219-228. Grimm, N.B., H.M. Valett, E.H. Stanley, and S.G. Fisher. 1991. Contribution of the hyporheic zone to the stability of an arid-land stream. Verlranlungen der Internationalen Vereinigung filr Theoretische und Angewandte Limnologie 24: 1595-1599. Harris, W.G., H.D. Wang, and KR. Reddy. 1994. Dairy manure influence on soil and sediment composition: implications for phosphorous retention. Journal of Environmental Quality 23: 1071-1081. Hawkins, C.P., M.L. Murphy, N.H. Anderson, and M.A Wilzbach. 1983. Density of fish and salamanders in relation to riparian canopy and physical habitat in streams of the northwestern United States. Canadian Journal of Fisheries and Aquatic Sciences 40: 1173-1185. Hazel, R.H., C.E. Burkhead, and D.G. Higgins. 1979. The development of water quality criteria for ammonia and total residual chlorine for the protection of aquatic life in two Johnson County, Kansas streams. Office of Water Research and Technology, Department of the Interior, Washington, DC. Hofstede, RGM. 1995. Effects of livestock farming and recommendations for management and conservation of Paramo grasslands. Land degradation and rehabilitation 6: 133-147. Hynes, H.B.N. 1966. T11 Bi lo f Pollut W s. Liverpool: UK: Liverpool University Press. Hynes, H.B.N. 1970. W Toronto: University of Toronto Press. Kauffman, J .B., W.C. Krueger, and M. Vavra. 1983 (a). Effects of late season cattle grazing on riparian plant communities. Journal of Range Management 36: 685- 691. Kauffman, J.B., W.C. Krueger, and M. Vavra. 1983 (b). Impacts of cattle on streambanks in northeastern Oregon. Journal of Range Management 36: 683- 685. Kauffman, LB. and W.C. Krueger 1984. Livestock impacts on riparian ecosystems and streamside management considerations: a review. Journal of Range Management 37: 430-438. Kondolf, GM. 1993. Lag in stream channel adjustment to livestock exclosure, White Mountains, California. Restoration Ecology 4: 226-230. Lamberti, GA, and J.W. Ferrrinella. 1996. Plant-herbivore interactions. Pages 409- 430 in ER. Hauer and GA. Lamberti. Methods in Stream geology. New york: Academic Press. Ieopold, A. 1923. Some fundamentals of conservation in the southwest. Pages 86-97 in Flader, SJ. and J .B. Callicott. 1991 (eds.). The Rivg Of the Mothg of 99;; Madison: The University of Wisconsin Press. Ieopold, A. 1946. Erosion as a menace to the social and economic future of the Southwest. Journal of Forestry 44: 627-633. Li, H., G.A. Lamberti, T.N. Pearsons, C.K. Tait, J.L. Li, and LC. Buckhouse. 1994. Cumulative effects of riparian disturbances along high desert trout streams of the John Day Basin, Oregon. Transactions of the American Fisheries Society 123: 627-640. Matthews, B.W., L.E. Sollenberger, V.D. Nair, and CR. Staples. 1994. Impact of grazing management on soil nitrogen, phosphorous, potassium, and sulfur distribution. Journal of Environmental Quality 23: 1006- 1013. Merritt, R.W. and K.W. Cummins. 1996. Trophic relations of macroinvertebrates. Pages 453-474 in ER. Hauer and GA. Lamberti. Methods in 8min Elegy. New york: Academic Press. Meyer, J.L., W.H. McDowell, T.L. Bott, J .W. Elwood, C. Ishizaki, J.M. Melack, B.L. Pekarsky, BJ. Peterson and RA. Rublee. 1988. Elemental dynamics in streams. Journal of the North American Benthological Society 7: 410-432. Minshall, G.W., S.E. Jensen, and W.S. Platts. 1989. The ecology of stream and riparian habitats of the Great Basin region: a community profile. US Fish and Wildlife Service Biological Report 85 (7.24). Washington, DC. Milchunas, D.G. and W.K.Lavenroth. 1993. Quantitative effects of grazing on vegetation and soils over a global range of environments. Ecological Monographs 63: 327-366. Minshall, G.W., S.E. Jensen, and W.S. Platts. 1989. The ecology of stream and riparian habitats of the Great Basin region: a community profile. US Fish and Wildlife Service Biological Report 85. Mosely, J .C., T.A. Lance, J .W. Walker, D.E. Lucas, and CM. Falter. 1993. How does grazing effect water quality? Page 5 in D. Ortiz (ed.). 1199334111 Wm. Idaho Forest, Wildlife, and Range Experiment 10 Station, 1992 Annual Report. 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. Murphy, M.L., C.P. Hawkins, and NH. Anderson. 1981. Effects of canopy modification and accumulation of sediment on stream communities. Transactions of the American Fisheries Society 121: 427-436. Munn, N.N. 1989. The role of stream substrate and local geomorphology in the retention of nutrients in headwater streams, Ph.D. dissertation, University of Georgia, Athens, Georgia. Nutall, RM. and CH. Bielby. 1973. The effects of China-clay wastes on stream invertebrates. Environmental Pollution 5: 77-86. Odion, D.C., T.L. Dudley, and CM. D'Antonio. 1988. Cattle grazing in southeastern Sierran meadows: ecosystem change and prospects for recovery. Pages 277-292. in C.A.J. Hall and V. Doyle-J ones (eds.). Whig Mountain Symposium, Volume 3: Plant Biology of Easgm Califmia. University of California, Ios Angeles. Overcash, M.R., F.J. Humenik, and J .R. Miner. 1983. Livestmk Waste Mmement Volumes I and II. CRC Press, Boca Raton, Florida. Peters, J .C. 1962. The effects of stream sedimentation on trout embryo survival. Pages 275-279 in Tarzwell, C.M. (ed.). Biological Mlems in Wm Pollution. U.S. Public Health Service, Cincinnati, OH. Peters, J .C. 1967. Effects on a trout stream of sediment from agricultural practices. Journal of Wildlife Management 31: 805-812. Platts, W.S. 1991. Livestock grazing. Pages 389—423 in W.R. Meehan (ed.). Influences of Forest and Rmamrwmmmm Habita§. American Fisheries Society Special Publication 19: 389-423. Platts, W.S., and R.L. Nelson. 1989. Stream canopy and its relationship to salmonid biomass in the intermountain west. North American Journal of Fisheries Management 9: 446—457. Platts, W.S., and J .N. Rinne. 1985. Riparian and stream enhancement management and research in the Rocky Mountains. North American Journal of Fisheries Management 5: 113-125. Quinn, J.M., R.B. Williamson, R.K. Smith, M.L. Vickers. 1992. Effects of riparian grazing and channelisation on streams in Southland, New Zealand. 2. Benthic Invertebrates. New Zealand Journal of Marine and Freshwater Research 26: 11 259-273. Resh, V.H., A.V. Brown, A.P. Covich, M.E. Gurtz, H.W. Li, G.W. Minshall, S.R. Reice, A.L. Sheldon, J.B. Wallace and RC. Wissmar. 1988. The role of disturbance in stream ecology. Journal of the North American Benthological Society 7: 433-455. Rinne, J .N. 1988. Effects of livestock grazing exclosure on aquatic macroinvertebrates in a montane stream, New Mexico. Great Basin Naturalist 48: 146-153. Rutherford, J .E. and R]. Mackay. 1986. Patterns of pupal mortality in field rxnndadonsctlfwbmuuyehezuul(Hanananunyohe(Trkflunnenu Hydropsychidae). Freshwater Biology 16: 337-350. Sedell, J .R., G.H. Reeves, F.R. Hauer, J.A. Stanford. 1990. Role of refugia in recovery from disturbances: modern fragmented and disconnected river systems. Environmental Management 14: 711-724. Sweeny, B.W. 1993. Effects of streamside vegetation on macroinvertebrate communities of White Clay Creek in eastern North America. Proceedings of Natural Sciences of Philadelphia 144: 291-340. Sweeny, B.W. and R.L. Vannote. 1986. Growth and production of a stream stonefly: influences of diet and temperature. Ecology 67: 1396-1410. Tait, C.K., J.L. Li, G.A. Lamberti, T.N. Pearsons, and H.W. Li. 1994. Relationships between riparian cover and community structure of high desert streams. Journal of the North American Benthological Society 13: 45-56. Timm, T. 1994. Reasons for the shift in dominance between Simulium (N.) vemum and Simulium (S .) omatum (Diptera: Simuliidae) along the continuum of an unpolluted lowland stream. Archives fur Hydrobiologia 131: 199-210. Thomas, V.G. 1985. Experimentally determined impacts of a small, suction gold dredge on a Montana stream. North American Journal of Fisheries Management 5: 480-488. Wallace, J .B., and NH. Anderson. 1996. Habitat, life history, and behavioral adaptations of aquatic insects. Pages 41-73 in R.W. Merritt and K.W. Cummins. WWW Dubuque Iowa: Kendall Hunt. Waters, TE 1984. Annual production by Gammarus pseudolimnaeus among substrate types in Valley, Creek, Minnesota. American Midland Naturalist 112: 95-102. 12 Waters, TR 1995. Sediment in SMS. American Fisheries Society Monograph 7, Bethesda, Maryland. Webster, J.R., S.W. Golladay, E.F. Benfield, J.L. Meyer, W.T. Swank, and J .B. Wallace. 1992. Catchment disturbance and stream response: an overview of stream research at Coweeta Hydrologic Laboratory. Pages 231-253 in Boon, P.J., P. Calow, and GE. Petts (eds.). River Cenmgtien egg Muggemmt. John Wiley and Sons, New York. Williams, DD. 1984. The hyporheic zone as a habitat for aquatic insects and associated arthropods. Pages 430-455 in V.H. Resh and D.M. Rosenberg (eds.). The mlogy ef Aggtie lngts. New York: Praeger. Wissmar, R.C., J.E. Smith, B.A. McIntosh, H.W. Li, G.H. Reeves, and J .R. Sedell. 1994. A history of resource use and disturbance in riverine basins of eastern Oregon and Washington. Northwest Science 68: 1-35. 13 3:8 grant cavemafifim a. 0.38 wage» .0 30.08 0.8. :0 meet“. oEow .. oSwE uofi:.E..m 0&2:an 05.8.0». - 2.2.8... 23¢ 8.5.3.. .2353 {Av—«NECK «a: _Nm_< ,/,/, 523235.. ofiEEoo axe... mcgesm oozn.c_E.o «8.2a: 89:35 3.8 commute... .8868 822.8"... .atam ecoN 0.2.39»... 5:85.35 «Esaam 3.55m uofiEE=m 2.3.590 02:83 8328. 3.3%.; EoEEEEm .252... tonnage... 5.5.8:. 8555.... .2235 958m uouEm zo.h<...m0m> ¥zO ".0 Whomumm mEOw zo...<...m0m> >DOO>> ".0 Oz_m. .25 — -‘- 1 """""""" :1: ‘ ...... 8 eeeee 1 """""""" '6 2 ‘ m """"" """"" i -- 15 - " """" 5 1 L- ,_ 3 . 0 1 991 1 992 1 993 Figure 3. Mean depth and flow rate i 1 se in one forest-reach and two pasture-reach stations. Measurements taken during July, August, and September, 1991-1993. Suspended Solids (mg/L) 1 991 1 992 1 993 Figure 4. Suspended solid concentration from forest— and pasture- reach water sampless taken monthly, J uly—September, 1991-1993. 's91dums need-5291 pure ‘poom 61001 H; Amie/11p pue ‘ssauqop ‘aouepunqe armqanaAur meow 'g mar 49 Abundance log (inds. lday) log (indsJ m2) 0 p __. to o .1 our-tour-tcnromwocn-tmm 3661 166|- 866 l Richness log (taxa/day) log (taxa! 0.1 m2) 266 l 866 l 3661 1661- 866 l- SO FilterersO Gatherers . Scrapers edators . HerbivoresO Shredders. UNDISTURBED BEAVER YR. 1 BEAVER YR. 2 ROCKS 5°2 109° 495 3335 705 g 1727 Figure 6. Community functional-fading group composition in forest-reach samples. Values indicate total number of animals collected. Functional-feeding groups that comprised less than 1 % of total are not represented on pie-charts. These are: 3 herbivores in 1992 wood samples and 5 scrapers in 1992 leaf-packs. 51 Filterers. Gatherers . Scrapers. Predators . Herbivores O Shredders. OVERGRAZED YR. 1 RECOVERY YR. 2 RECOVERY LEAVES ,3, we. Figure 7. Community functional-feeding group composition in pasture-reach samples. Values indicate total number of animals collected. Groups that comprised less than 1% of total count are not represented on pie-charts. These are: 2, 9, and 9 scrapers in 1991-1993 wood samples respectively; 1 herbivore in 1991 wood samples; and 6 scrapers in 1991 leaf-pack samples. APPENDIX 1 52 Appendix 1. Forestjreach taxa organized by feeding guild and year of collection. letters after taxa names rndreate sample type(s) from which animals were collected: R=rock; W=wood; L=leaf packs; D=drift. FEEDING 1991 1992 1993 GUILD UNDISTURBED BEAVER 1N BEAVER YR. 2 Anhroplca R Ceratopryche D Cladocera W FILTERERS Bmchycenm L Cheumatopsyehe R W L D Copepoda W Cerawpuyche R W Hydropsyche R W D Cheumatopsyche R L D Pmsimuliwn R W L D Hydropsyche R W L D Neureclipsis W D Neurcclips'is W D Bivalvia R Prosimulim W L D Hydra R LD Bivalvia RWD Cladocera W L D Cladocera D Copepoda D Padura D Padura R W D Padum R GATHERERS Baetis RWLD Baetis RWLD Cami: W Cami: R W D Cami: R W Ephemerella W Pamleptophlebia L D Ephemerella R W Paralepcophlebla W Mystacides R Pmleptophlcbia W L D grammar W thomyi'a R W Mystacr'des W ironornidae R W Dubiraphia W Dubiraphia W D Dixella W Macromrchus W L D Macmnychw W D Oligochaeu R Opa'oservus L Optioservus W Acari R Stenelmis R Stenclnu's W L Chironomidae RWLD Chironomidae RWLD Nematoda R L Nematode R Acari R W D Acari R W D Macdwmoa R Nix: R Stenacmn W SCRAPERS Stenacron RD Stenacron RW LD Sma W Stenonema R W L D Stenonema R W L D Helicopsy‘he R D Helicopsyche R Gastropoda R D Gastropoda R W Amphiagn’an D Argia D Argia W PREDATORS Argia D Calopteou W Stalls W Bayen'a R D Isoperla L Oeccds W Chromagflon D Bayer-fa W Probezzia W Cordulegasm' R Chauh'odes L Tan inae R W Calopteoa R L D Nigronia R H ' R W Pmmogomphu: D Stalls R Nematomorpha W Isoperla L 0ecetis R W D Mgram'a R Nym'ophylax R W L Nyca'ophylax R Polycenu'opw R W 0ecetis R L D waua's D [’01me R Tanypodinae R W L D alaoborus D Hemerodromia R W L D Tanypodinae RWLD Hirudinea RWD Dicranota W Nematomorpha R W Probezzia R Heater-04mm W L Hirudinea R W Hydmpa'la R W HERBIVORES OWN w Peltodytes D Taenio W Taeniopteryx L Amphlpoda° W SHREDDERS ”mm; D ”40...... no Lepidostoma R LD Pycnopsyche R humming R W Trianodes W D Acemria D Amphipoda R W D Tipula. D Amphxpoda D IsopodaL 53 Appendix 2. Forest—reach taxa organized by feeding guild and year of collection. letters alter taxa names indicate sample type (s) from which animals were collected: R=rock; W=wood; L=leaf packs; D=drift. APPENDIX 2 54 FEEDING 1991 1992 1993 GUILD OVERGRAZED CATTLE EXCLUDED RECOVERY YR. 2 Chemnawps'yche R WD Brachycentrus R Brach centrus W FILTERERS Hydro psyc che RL Ceratopsyche R LD Chewzatopsyche RW Neureclipsi: D h p he RWLD HydropsycheW Prosimulium R L D Hydmpsyche R W L D Simuliidac R Bivalvia R W L D Neureclipsis Bivalvia R Prosimuliwn W L D Hyira R W Bivalvia R W D Hydra D Cladocera R L D Cepepoda D Daphuia RD Collembola W Collembola LD Podura R GATHERERS Ameletus D Podura D Baeu's RW Baetis-RWLD BaedsRWLD Caenis Caem‘sRLD CaenisRWLRDD ParaleptogvhlebiaRW Ephemerella W Ephemerella Myslacides W Paraleplophlebia L Paraleptophlebia R W L D sychomyia RW Hexagenia R Pswhomyia RW D Dubiraphia RW Siphloneuridae R Dubira aR L D Maeronychus W Psychomyia R Macronychu: W D Opaoservus RW Dubiraphia R L Stenelmis R WD S Inn's Macronychus W D Dixella RD Chironomidae RW Opfioservus L D Chironomidae R W L D Nematode R W Stenelmi: W Anlocha Acari RW DuellaLD Nematode RWLD CopepodaW Chironomidae R W L D Oligochaeta R NematodaRWLD Acari RWLD Oligochaeta L Copepoda RW Acari W Pseudocloeon R Heptagerua R Stenacron RW SCRAPERS Stenacron RL R Srenonema R Slenonema R L D Stenacron R D Helicopsyche R W Glossosoma D Steuonema R W LD D Psphenidae D Helicopsyvhe R Neophylax W Neophylax W a D Calopteryx D Aeschna D Boyer-in RW PREDATORS Polycentrapus W L Calopteryx D Calopte Dytiscida eLD Corixidae DD 0ecetis R Tanypodinae R W L D Nigrmia Polycentmpus W Probezzza D Oecen's I? W D Dineuu‘s R Hemerodromia R W L D Nyctiaphylax L Tanypodinae R W Hirudinea R LD Palycemropus W D limnophora R Agabus D Hememdromra RW Dineun‘ 115D Nematomorpha W Tanypodinae R W L D Hirudinea R W chranota R Hexawma R Hemerodromia R W L D Hirudinea R W LD Nematomorpha RW ' Hm‘laRWLD HydroptilawRW HERBIVORES ”mm W L 081.2 R w L D 0...)...” Pamponyx W Pampanyx RW Haliplus R Petroplu'la R Pelrodytes D Le ‘ stoma D [epidostoma D [epidoswma RW SI—IREDDERS Pygwdfpryche L D Pycnopsyche R W L D Pyvnopsyche Grensia D Necwpsyche RL HydawPhylax W 11thny L Trianodes RD Amphipoda RW Tipula D Hydatophylax D Amphipoda LD Tipulidae R Amphipoda R W L D R Decapoda Isopoda 3 THE BEHAVIORAL RESPONSE OF THE STONEFLY PARA GNE TINA MEDIA (PLECOPTERA: PERLIDAE) TO THE ONSET OF INTENSE SEDIMENTATION Abstract. The initial response by Paragnetz'na media (Plecoptera: Perlidae) nymphs to increased suspended sediment concentration was examined in field trials with chambers designed to give nymphs the choice to stay or reposition in response to experimental sediment additions. The results indicate that the often posited immediate-escape response to the onset of intense sedimentation may not be commonly enacted by stonefly nymphs or perhaps other relatively immobile, lotic insects. INTRODUCTION What do stream—dwelling animals do when they experience rapid increases in suspended sediment concentration? The answer to this question is fundamental to understanding the effects of the massive amount of sediment entering flowing water globally as a result of human activities (Waters 1995). However, except for the notable exception of several widely studied Salmonidae species (e.g. Tagert 1984, Reiser and White 1988), little is known about how, or if, high concentrations of suspended sediments affect aquatic organisms (Cordone and Kelly 1961, Reiser and White 1988, Servizi and Martens 1991, 1992, Waters 1995). The enhancement of suspended sediment levels associated with human agricultural, silvicultural, and industrial activities is widely suspected to force benthic invertebrates to flee impacted habitats (e.g. Hynes 1973, Newbury 1984, Williams and Feltmate 1992), however, virtually no empirical evidence has been presented to support this contention (Waters 1995). Extreme sedimentation events such as those that follow heavy rains in overgrazed river—bottom land (Tarzwell 1938), reservoir flushing (Gray and Ward 1982), road construction (Ogbeibu and Victor 1989), and riparian zone clear-cuts (Webster et al. 1992) definitely have the effect of removing benthic invertebrates. However, in most cases it is not known whether these responses were instantaneous reactions to stress created by elevated suspended solid concentrations or 55 56 an eventuality associated with habitat degradation (Waters 1995). Several field experiments have been conducted to determine the direct effects of heavy sedimentation on benthic invertebrate behavior (Brunskill et al. 197 3, Rosenberg and Snow 1975 a, b, Rosenberg and Wiens 1978, Culp et al. 1986). For example, Rosenberg and Wiens (1978) observed that introducing unsorted bank sediments to a Northwest Territories river caused a marked increase in drifting invertebrates five hours after the episode. They concluded that sediment addition at 30 mg/ I initially ”strips" benthos from substrata and that the most sensitive and most exposed organisms leave first. However, because samples were not taken immediately after sediment addition, initial stripping was not actually measured. After pouring sand into riffles of a British Columbia stream, Culp et. al ( 1986) also cited unmeasured scouring effects of saltating sediment particles as the causative factor of, in this case, instantaneous drift of invertebrates that inhabit stone surfaces. One technical problem common to traditional field studies of invertebrate response to sedimentation is that specific drift—initiating factors remain undetectable with standard drift—monitoring techniques (Waters 1995). After producing and amlyzing data such as that reviewed above, White and Gammon (1977) concluded that sedimentation- mediated drift may not be a consequence of stress caused by sediment (i.e. catastrophic drift) but rather to light-attenuation—induced behavioral drift (sensu Waters 1972, Mfiller 1974). Therefore, drift in response to rapid increases in suspended solid concentration may be interpreted as a response to local overpopulation or resource scarcity (density-dependent drift) that would not necessarily result in negative effects on benthic communities (Waters 1995). The primary objective of this study was to determine if nymphs of the relatively immobile, predacious stonefly Paragnetina media (Plecoptera: Perlidae) would immediately reposition in response to the onset of intense sedimentation. Exposure to sediments was limited to a brief period of time in order to focus on assessing the immediate, direct effects of sedimentation on stonefly behavior such as integument scouring and gill fouling. 57 METHODS Stoneflies Paragnetina media (Walker) nymphs are common inhabitants in eastern U. S. streams (Frison 1935, Stewart and Stark 1988). Characteristically associated with fast-flowing streams, their flattened body shape and gripping tarsal claws allow nymphs to crawl over relatively silt-free substrata and through accumulations of allochthonous debris in search of a wide variety of invertebrate prey (Stewart and Harper 1996). In mid—Michigan, P. media nymphs hatch from early summer-laid eggs and reach maturity by the spring of their second year (Heiman and Knight 1970). Tightly synchronous emergence (late May in mid-Michigan) enhances the probability of mating success of these weak—flying, short-lived adults (Feltmate and Pointing 1986) and consequently allows determination of nymphal age due to the great disparity in size when age classes overlap (Heiman and Knight 1970). Species determination was made with keys in Stewart and Stark (1988). Stoneflies in the autumn of their second year were used in experiments; they were collected and trials were conducted in a riffle of a second-order, mid-Michigan pasture stream that flowed over cobble and gravel substrata (Prairie Creek, Ionia Co., MI; 43°N, 85 °W). Experimental pmcedure Nymphs were collected with soft forceps and placed into a pan of aerated stream water that contained a small stone for them to cling to. When six nymphs were collected (five for one set of trials), the stone (with clinging nymphs) was removed and placed into a sedimentation-response chamber (Figure 1). Paragnaina media nymphs are relatively sessile once a position is established (Felmate and Pointing 1986). This allowed for trials to get underway before a move unrelated to the treatment was likely to be attempted (approximately 10 seconds after introduction). Sedimentation-response chambers are compartmentalized, 2-mm thick plastic boxes with l mm-diameter mesh walls (Figure 1). Chambers were designed to provide nymphs with four options in response to exposure to a sediment and water slurry delivered directly into the chamber from an upstream reservoir. Response options 58 include (I) maintain position in the (220 x 85 x 35 mm) overlying compartment (= stayed); (2) move down through 20 x 5 mm slots to a like-sized underlying chamber (=down); (3) move upstream into a 200-mm long, 0.5-mm diameter, mesh net (= upstream); and (4) move downstream into a into 200-mm long, 0.5 mm mesh net (= downstream) (see Table 1, Figure 2). The sediment-delivery system is composed of a 23 1, plastic pale fitted with 3 m of 40— mm diameter, plastic tube through which the sediment and water slurry (or just water for control trials) is delivered to a chamber from an upstream position. When the sediment and wate“ slurry (or water in control trials) had completely passed through the chamber (approximately 2 minutes after the onset of a trial), it was disconnected from the sediment-delivery system and removed from the stream. Nymphs were then collected from the nets and compartments and preserved in 70% ethanol. Trials were conducted on 19 September, 1993 and 23 September, 1994. A total of eight sediment and nine control trials were conducted. Sediments A slurry consisting of a 1.5 1 container of bank sediments and 23 l of stream water produced a level of sedimentation downstream from chambers that was similar to that measured during a heavy spate in Prairie Creek (24 mgll on 15 September, 1993); a level that is known to produce negative, yet sublethal effects on benthic invertebrates (Newcombe and MacDonald 1991). The actual suspended sediment level in chambers was approximately at 56 mgll as determined with a a Hach 2100 A NTU meter and a known mgll : NTU relationship. Trials were conducted on days when the stream was running relatively clear of suspended sediments. Therefore, control trials were conducted with low-sedirnent (> 1 mgll) water samples. Sediment particle-size and chemical composition was determined by the Michigan State University Plant and Soil Nutrient Laboratory (Table 2). Statistical analysis A single-factor, repeated-measures ANOVA model was used to test the effects of the two treatments on behavioral response. Response alternatives include maintaining 59 position, moving below the main compartment, moving upstream, or downstream (see Table 1). RESULTS Results presented in Figure 2 indicate that P. media nymphs do not immediately flee habitat in response to intense sedimentation. Nymphs were expected to reposition at a higher frequency in sedimentation trials, however they actually moved less in sediment trials than in controls (mean % 21 70.8 vs 50.7) (ANOVA: P > 0.053, Table 1, Figure 2). There were no significant differences between repositioning options within trials or between treatments (Table 1, Figure 2). DISCUSSION The experimental population of mid-Michigan P. media nymphs were apparently resistant to, or experimental conditions did not reproduce, gill-fouling (Hynes 1970) and integument shearing effects (Culp et al. 1986) believed to be associated with high levels of bank sediment in flow. Perhaps frequent exposure in the recent past has led to the establishment of the tendency to maintain position during high-sedimentation episodes due to the commonality of such incidents in the pastured experimental stream reach. The selective pressure against enacting movements unrelated to resource acquisition is likely reinforced in Prairie Creek by selection against relocation during daylight hours due to heightened risk of being eaten by visual-feeding insectivorous fish (Waters 1972). This pattern was also detected in unquantified preliminary trials conducted by Cheumatopsyche and Hydropsyche larvae (Trichoptera: Hydropsychidae) in a Minnesota stream with thriving insectivorous fish populations. The relatively immobile P. media, may have also acquired a position-maintenance strategy early in its evolutionary history. In order for an organism to successfully colonize a novel habitat, it must be able to respond to its most severe conditions. Almost universally, the behavior to deal with novel stress such as that which seemingly accompanies habitat shifts and rapidly changing environments, must be expressed prior to morphological adaptation (Mayr 1982 ). For example, protoplecoptems, terrestrially adapted animals believed to have secondarily invaded aquatic habitats some 300 mya 60 (Imms 1957, Illes 1965, Ross 1965), had to maintain position while exposed to high levels of flow and sedimentation events prior to acquiring lotic adaptations such as streamlined morphology and gripping tarsal claws (Wooten 1972). Therefore, throughout their history, P. media nymphs and their direct progenitors may have evolved a strategy of position maintenance in response to episodes of intense sedimentation. Evidence presented here indicates that one species with limited mobility, relatively long development time, and patchily distributed resources is more likely to endure sedimentation rather than attempt to avoid the stress of exposure by attempting to reposition. It is clear that benthic invertebrates do not respond to anthropogenic sedimentation in a consistent manner, thus reinforcing the exigency for clarification the roles of phylogeny, morphology, and regional ecology in influencing invertebrate behavioral response to the widespread problem of anthropogenic sedimentation in streams. 61 LITERATURE CITED Brunskill, G.J., D.M. Rosenberg, N.B. Snow, G.L. Vascotto, and R. Wagemann. 1973. Ecological studies of aquatic systems in the Mackenzie-Porcupine drainages in relation to proposed pipeline and highway developments, volume 1. Northern Pipelines Task Force, Environmental-Social Committee. Northern Oil Development Report 73-40, Winnipeg, Manitoba. Cordone, A.J. and D.W. Kelley. 1961. The influences of inorganic sediments on the aquatic life of streams. California Fish and Game 47: 189-228. Culp, J .M., P.J. Wrona, and R.W. Davies. 1986. Response of stream benthos and drift to fine sediment deposition versus transport. Canadian Journal of Zoology 64: 1345-1351. Felmate, B.W., R.L. Baker, and P.J. Pointing. 1986. Distribution of the stonefly nymph Paragnetina media (Plecoptera: Perlidae): influence of prey, predators, current speed, and substrate composition. Canadian Journal of Fisheries and Aquatic Sciences 43: 1582-1587. Frison, TH. 1935. The stoneflies, or Plecoptera, of Illinois. Bulletin of the Illinois Natural Survey 20: 281-471. Gammon, LR. 1970. The effect of inorganic sediment on stream biota. USEPA Water pollution Research Series 18050 DWC 12170. ‘ Gray, LI. and J .V. Ward. 1982. Effects of sediment releases from a reservoir on stream macroinvertebrates. Hydrobiologia 96: 177-184. Heiman, DR and A.W. Knight. 1970. Studies on growth and development of the stonefly Paragnetina media Walker (Plecoptera: Perlidae) American Midland Naturalist 84: 274-278. Hynes, H.B.N. 1973. The effects of sediment on the biota of running water. Pages 653-663 in W. Canadian Department of the Environment, Ottawa. Illes, J. 1965. Phylogeny and zoogeography of the Plecoptera. Annual Review of Entomology 10: 117-140. Imms, AD. 1957. W (revised by O.W. Richards and R.G. Davies). New York: E.P. Dutton. Mayr, E. 1982. W Cambridge: Harvard University Press. 62 Miiller, K. 1982. The colonization cycle of freshwater insects. Oecologia 52: 202- 207. Newbury, R.W. 1984. Hydrologic determinants of aquatic insect habitats. Pages 323-357 in V.H. Resh and D.M. Rosenberg (eds.). Ecology of Aquatic Iii—sects. New York: Praeger Newcombe, GP. and DD. MacDonald. 1991. Effects of suspended sediments on aquatic ecosystems. North American Journal of Fisheries Management 11: 72- 82. N utall, RM. and G.H. Bielby. 1973. The effects of china-clay wastes on stream invertebrates. Environmental Pollution 5: 77-86. Ogbeibu, A.E. and R. Victor. 1989. The effects of road and bridge construction on the bank-root macrobenthic invertebrates of a southern Nigerian streams. Environmental Pollution 56: 85-100. Reiser, D.W., and R.G. White. 1988. Effects of two sediment size-classes on survival of steelhead and Chinook salmon eggs. North American Journal of Fisheries Management 8: 432-437. Rosenberg, D.M., and A.P. Wiens. 1978. Effects of sediment addition on macrobenthic invertebrates in a northern Canadian river. Water Research 12: 753-763. Ross, H.H. 1965. A Textbook of Entomolggy, 3rd Edition. New York: John Wiley and Sons Servizi, J.A. and D.W. Martens. 1991. Effect of temperature, season, and fish size on acute lethality of suspen®d sediments to Coho Salmon (Oncorhynchus kisutch). Canadian Journal of Fisheries and Aquatic Sciences 48: 493-497. Servizi, J .A. and D.W. Martens. 1992. Sublethal responses of Coho Salmon (Oncorynchus kisutch) to suspended sediments. Canadian Journal of Fisheries and Aquatic Sciences 49: 1389-1395. Slaney, P.A., T.G. Halsey, and H.A. Smith. 1977. Some effects of forest harvesting on salmonid rearing habitat in two streams in the central interior of British Columbia. British Columbia Fish and Wildlife Branch, Fisheries Management Report 71, Victoria. Stewart, K.W. and RP. Harper. 1996. Plecoptera. Pages 217-261 in Merritt, R.W. and K.W. Cummins (eds.). An Introductign m the Aggtic Ingts of North America. Dubuque, IA: Kendall Hunt. 63 Stewart, K.W. and B.P. Stark. 1988. Nymphs of North American stonefly genera (Plecoptera). Thomas Say Foundation Series, Entomological Society of America 12: 1-460. Tagart, J .V. 1984. Coho salmon survival from egg deposition to fry emergence. Pages 173-182 in J .M. Walton and DB. Houston (eds.). Proceedings of the Olympic Wild Fish Conference. Peninsula College, Port Angeles, Washington. Tarzwell, CM. 1938. Factors influencing fish food and fish production in southwestern streams. Transactions of the American Fisheries Society 67 : 246:255. Tebo, L.B., Jr. 1955. Effects of siltation on trout streams. Proceedings of the Society of American Foresters 1956: 198-202. Wagener, S.M., and J .D. LaPierrere. 1985. Effects of placer mining on the invertebrate communities of interior Alaska streams. Freshwater Invertebrate Biology 4: 208-214. Waters, T.F. 1972. The drift of stream insects. Annual Review of Entomology 17: 253-272. Waters, T.F. 1995. Sediment in Streams. American Fisheries Society Monograph 7, Bethesda, Maryland. Wcislo, W.T. 1989. Behavioral environments and evolutionary change. Annual Review of Ecology and Systematics 20: 137-169. Webster, J .R., S.W. Golladay, E.F. Benfield, J.L. Meyer, W.T. Swank, and J.B. Wallace. Catchment disturbance and stream response: an overview of stream research at Coweeta Hydrological Laboratory. In Boon, P.J., P. Calow, and GE. Petts (eds.). WWW New York: John Wiley and Sons. White, D.S. and J .R. Gammon. 1977. The effect of suspended solids on macroinvertebrate drift in an Indiana creek. Proceedings of the Indiana Academy of Sciences 86: 182-188. Williams, DD. and B.W. Felmate. 1992. MW. Wallingford Oxen U.K.: C.A.B. International. Wootten, R]. 1972. The evolution of insects in freshwater ecosystems. Pages 69-82 in R.B. Clark and RJ. Wootten (eds.). W. Exeter: University of Exeter Press. 64 Table 1. ANOVA results comparing the number of stonefly nymphs that selected each behavioral option (see Figure 2). STAYED DOM U STRE W Source df MS F MS F MS F MS F Sed.Ievel 1 1709.68 4.427 1 0.079 1.931 0.013 3.529 0.000 0.007 Error 15 386.16 0.041 0.004 0.003 1‘ P: 0.053 65 Table 2. Particle-size distribution and chemical composition of bank sediments used in sedimentation treatments. Soil Type: Loam, pH 7.1 Mineral Component: 70.7% Organic Component: 29.3% 45.8 % Sand Na 258 ppm 32.7 % Silt Cl 620 ppm 21.4 % Clay N03 5.4 ppm NH4 74.7 ppm 66 Figure 1. A sedimentation-response chamber. Number Of Individuals 67 5.. - [3 Controls 4- Sediment Treatments 3— , 2- 1 _ 0 . I 13" 54v: Stayed Down Upstream D °Wnstream Figure 2. Mean number of stoneflies (i 1 se) that chose each behavioral option in 9 control and 8 sedimentation trials. 4 EFFECTS OF EPISODIC SEDIMENTATION ON THE NET-SPINNING CADDISFLIES H YDROPS YCHE BETTENI AND CERA TOPS YCHE SPARNA (TRICHOP'TERA: HYDROPSYCI-IIDAE) Abstract. Net-spinning caddisflies often thrive in the heavily sedimented waters of midwestern agricultural catchments despite the presumed costs of sedimentation- induced gill- and net-fouling. We conducted a laboratory experiment to determine whether larval growth and survival of two mid-Michigan Hydropsychidae species (Hydropsyche betteni and Ceratopsyche spama) are affected by daily exposure to high levels of sedimentation. Larvae of both species had a decreased likelihood of survival in sedimentation treatments relative to controls and taxa were differentially affected as H. bettem' significantly outperformed C. spama. Sedimentation did not alter the relative growth rate of either species although slight losses by H. betteni and gains by C. spama produced significant differences in relative growth rates between species. INTRODUCTION Sedimentation to North American streams, most of which is human generated (Waters 1995), is widely believed to harm stream-dwelling organisms through the effects of increased deposition and transport of sediments (Newcombe and MacDonald 1991, Waters 1995). High levels of sediment deposition and suspension have been demonstrated to negatively affect several game-fish species (principally salmonids) by burying eggs and fouling gills (Newcombe and MacDonald 1991; Sevizi and Martens 1991, 1992; MacDonald and Newcombe 1993; Waters 1995). The situation is not as clear, however, for the often numerically (Hynes 1970) and energetically dominant (Waters 1984) animals that comprise benthic invertebrate communities. In fact, the abundance of many benthic invertebrates typically increases in response to heavy sedimentation (Hamilton 1961, Hynes 1966, Gammon 1970, Learner et al. 1971, Nuttall and Bielby 1973, Waters 1995). 68 69 Invertebrate abundance increases that result from high levels of sediment deposition are typically experienced by relatively small, burrowing species adapted to exploit deposits of fine-particulate organic material. The conditions that favor these burrowing organisms often result in the exclusion of species that require solid substrata. Loss of the relatively large animals that inhabit exposed benthic habitat often results in overall invertebrate biomass declines (Waters 1995). Further biomass reductions also typically follow exclusion of exposed-substrata inhabitants as a result of secondary productivity declines in insectivorous fish populations (Tait et a1. 1994). Extremely high levels of sedimentation that result from acute disturbances, such as streambed-suction mining (Thomas 1985), riparian clearcutting (Webster et al. 1992), and road construction (King and Ball 1967) commonly have large, but impermanent effects on benthic communities. However, there is of yet only limited understanding of the effects of chronic sedimentation at levels moderate enough to increase suspended- solid concentration but preclude total habitat transformation through substrata burial (Waters 1995 ). In this study, a presumably agricultural-pollution tolerant species (Hydropsyche bettem) and a relatively intolerant species (Ceratopsyche spama) (as determined by Schmude and Hilsenhoff 1986) were exposed to daily episodes of heavy sedimentation to determine: ( 1) whether larval growth and survivorship are affected by episodic sedimentation (Fleischner 1994, Waters 1995), and (2) whether taxa would perform differentially as predicted by present distribution patterns (Ross 1944, Schmude and Hilsenhoff 1986). Efi’ects of sedimentation on filter-feeders Invertebrate species are known to be differentially affected by fluctuations in suspended sediment concentration. For example, Culp et al. (1986) found that sedimentation- mediated drift response was most pronounced in invertebrates that inhabit exposed substrata surfaces. Invertebrates are also expected to differ in response to suspended sediments as a result of relative exposure as dictated by feeding strategy. Due to their near uniform reliance on the availability of stable substrata, filter-feeding invertebrates 70 are suspected to be particularily sensitive to suspended-sediment increases (Hynes 1973). However, little evidence has been presented in support of this widely held view (Waters 1995). Filter feeders that utilize morphological filtration structures are apparently particularily sensitive to suspended sediment increases. For example, unionid clams (Aldridge et al. 1987), and cladocerans (McCabe and O'Brien 1983) are known to experience feeding limitation when exposed to high levels of suspended sediment. However, filter feeders that utilize external structures such as hydropsychid nets may not be as hindered by high levels of suspended sediment as are those that feed with anatomical filtration structures. Some Hydropsychidae species occupy both high- and low-sediment environments and are therefore potentially unaffected by moderate to heavy sedimentation (Schmude and Hilsenhoff 1986). learner et al. (1971) reported that Hydropsyche pellucidula larvae in clean reaches of a UK. stream were much larger than conspecifics in a reach receiving suspended solids from coal-mining operations. They concluded that development was delayed by sedimentation. However, because this hypothesis was not, and has not been tested (Waters 1995), it remains unknown if larvae in high- sedimentation environments grow, survive, and reproduce at rates similar to those in optimal habitat such as the low-sediment environment below many impoundments (Fremling 1960). Hydropsychidae nets trap sediment as well as a wide array of potential food items (Wallace and Sherberger 1974, 1975, Wallace and Merritt 1980). It thus seems probable that as sedimentation increases, nets require more maintenance and frequent replacement. The result of higher net-maintenance costs may be the expenditure of energy that would otherwise be used for respiration and growth. In addition, because silk production requires expenditure of lipid reserves critical to adult reproductive success (Wallace and Malas 1976, Petersson and Hasselrot 1994), more-frequent net replacement may affect caddisfly fitness. 71 METHODS Microcosms Four 45 -l aquaria were fitted with stream-flow simulators (BioQuip Products, Pasadena, CA) and surface-sanitized with 70% ethanol. A 10 x 10 x 40 cm, split-face concrete block, 5 g of ponderosa pine needles (used for retreat construction), 2 Kg aquarium sand, 1 Kg gravel, and 8.5 l stream water were added to each tank. All substrata were autoclave sterilized. An air compressor was used to generate flow. Water was allowed to circulate for 10 days prior to larval hydropsychid introduction in order to allow time for the current to create nearly uniform bed morphology between tanks. Caddisflies Final-instar (=fifth instar) Hydropsyche bettem' and Ceratopsyche spama (as determined by keys in Schmude and Hilsenhoff 1986) were collected from rocks removed from two second-order streams that flow through a combination of mid- Michigan agicultm'al and residential land (H. betteni: 21 November, 1994, Prairie Creek, Ionia Co., 43°N, 85 .W; C. spama 19 November,l994, Flint River, lapierre, Co., 43°N, 83° 15'W). Larvae were allowed to acclimate to tank temperature (mean = 20°C) from field-collection temperatures of (5°C) before they were introduced into tanks. Net-spinning activity of both species has been determined to be maximal in 20°C water (Fuller and Mackay 1980). Twenty-five individuals of each species were added to each tank. larvae were fed a daily diet comprised of 0.5 g powdered Tetra MinTM staple food and 0.1 g Red Jungle BrandTM Micro-food. Larvae of both species were observed feeding on this diet. Feeding was continued for two days following termination of sedimentation episodes to allow the passage of inorganic material so as to avoid nonfood items from affecting final weights. Surviving larvae were placed in 1.5 ml vials, stored at —20°C, thawed, identified, 72 cleansed of debris for 1 minute in a sonic cleaner, dried for 24 hours at 65°C, and weighed on a microbalance. Relative growth rate of each larva was determined with the equation: RGR = (1n final mass - 1n mean initial mass)/ time (days in captivity) Mean initial mass values were derived from measurements of a representative sample of each taxa (H. betteni n=24, C. spama n=21) which were frozen upon collection to be dried and weighed with the experimental animals. Sediments Sediment additions were initiated on 26 November, 1996 after all larvae had positioned themselves and constructed retreats. Sediment was collected from a bank deposit near the H. betteni collection site. Sediments were autoclave sterilized for one hour, thoroughly dried at 65°C, and sieved through a 0.6 mm—mesh sieve. Particle size array and chemical composition were determined by the Michigan State University Plant Nutrient Laboratory (see Table 1). Two tanks were randomly chosen to receive daily sedimentation. Each day (16 total), 11 g of sediment were added to experimental microcosms. The initial turbidity (mean = 23 nephelometric turbidity units (NTU)) is considered to be sufficient, if sustained, to cause adverse, but sub-lethal effects on benthic invertebrates such as reduced growth or forced abandonment (Newcombe and MacDonald 1991, MacDonald and Newcombe 1993). Suspended-solid level was measured with a Hach 2100 A NTU meter. NTU measurements were taken before each trial and several times throughout the first three hours after the onset of trials, after which levels in sediment-treatment tanks approximated those in control tanks (Figure 1). Abiotic conditions Measurements of water temperature and pH were taken daily. These parameters did not 73 vary between tanks more than the level of aceuracy of the meters (i 1°C, 0.5 pH) (mean temperature = 208°C, mean pH = 8.5). Current velocity was calibrated between tanks with a digital flow meter and was set to generate moderately turbid flow that did not sweep away larvae crawling on exposed substrata. Velocity 5 cm above retreats was ~5 cm/second. Statistical analyses An AN OVA model was used to test the full interaction of the effects of species (two species), tank (4 tanks), and sediment level (two levels) on relative growth rate (see Table 2). Survival-rate data were analyzed with a contingency analysis model (CATMOD procedure, SAS 1990) that tested the same interaction of effects (see Table 3). Survivorship data were analyzed with and without pupae. The addition of pupae to the analysis of survivorship did not meaningfully affect treatment effects (Appendix 1) and is therefore excluded from reported values. RESULTS Relative growth rate Sediment treatments had no effect on the relative growth rate of either species (Table 2, Figure 2). On average, C. spama grew and H. betteni lost mass in sediment treatments and controls (C. spama: 0.061 vs 0.059 mg dry mass gained per day and H. betteni: 0.061 vs 0.024 mg dry mass lost per day in control and sediment treatments respectively) (Figure 2). This interspecific difference is highly significant (Table 2). Survival Both species suffered higher mortality in sediment treatments than in controls, and as was the ease in relative growth rate measurements, taxa were differentially affected (Table 3, Figure 3). In contrast to the interspecific trend detected in relative growth rates, H. betteni vastly outperformed C. spama (Table 3, Figure 3). There was no 74 intraspecific, inter-tank variation with the exception of H. betteni in control tanks which differed by only one survivor. Individuals of each species pupated in both treatments (n = 2 H. betteni, 3 C. spama) prior to completion of the trial. One pupa of undetermined species was consumed by accidentally introduwd, larval Chironomidae (Eu/defiefiella sp.), a phenomenon also observed in microcosms by Rutherford (1986) and in nature by Rutherford and Mackay (1986). DISCUSSION Intraspecific effects Sediments were added at a level predicted to cause sub-lethal effects such as reduced growth or behavioral avoidance (sensu Newcombe and MacDonald 1991). However, although the mode of action is unclear, sediment treatments were lethal to a proportion of experimental populations of both species (Table 3, Figure 3). The two most—often posited deadly effects of high levels of sedimentation are critically lowered resilience to suspended solids and enhanced lethality eaused by substrata bmial (Waters 1995). Entrainment in thick deposits of sediment is known to be lethal to benthic invertebrates, but only when heavy sedimentation is prolonged enough to completely bury substrata (Thomas 1985) or when immobile forms like hydropsychid pupae are buried and ultimately suffocate (Rutherford and Mackay 1986). Either scenario occurring in a microcosm can be interpreted as representative of potentially negative effects on natural populations, but certainly not as absolutely lethal ones in habitats where successful behavioral avoidance is possible. It is also possible that each microcosm could, perhaps because of spatial limitation, only support 22 - 23 H. betteni and 10 C. spama (as in control tanks) and, through habitat denudation, sediment treatments reduced capacity to 15 H. betteni and 8 C. spama (as in both sediment tanks). Habitat simplification through substrata burial is known to cause eventual declines in benthic invertebrate populations (e. g. Tarzwell 1938, Hamilton 1961, Nuttall and Bielby 1973); but as insects were stocked at a small fraction 75 of field densities and many retreats were exposed and unoccupied at the termination of the 16-day experiment, burial seems unlikely to have been the only cause of larval death. Relative growth rates were similar intraspecifically and across treatments suggesting that surviving larvae were apparently unaffected by sedimentation (Table 2, Figure 2). Nets were observed to be clogged with sediment after exposure and were cleaned or replaced prior to the onset of the next trial. It seems reasonable, therefore, to presume that conditions in the sediment-treatment tanks required higher net maintenance costs relative to those in controls . However, because sediment treatments had no effect on relative growth rates (Figure 2), net maintenance costs over 16 days were probably negligible for survivors of sediment treatments. The possibility does exists that younger larvae would have responded differently due their typically higher growth rates (Cudney and Wallace 1980, Mackay 1979, 1984) and net-spinning activity (Fuller and Mackay 1980). However, other indirect measurements of growth rate have revealed that final-instar hydropsychids do gain mass (Cuffney and Minshall 1981). Intetspecific reflects The expectation that H. betteni are more tolerant to sediment treatments than C. spama (Schmude and Hilsenhoff 1986) was not corroborated by interspecific comparisons of relative growth rate and survivorship. Survivorship data suggest that larval mid- Michigan H. betteni are more manipulation-resistant than C. spama, which may indieate higher general resistance which is also evidenced by their distribution in Wisconsin (Schmude and Hilsenhoff 1986). However, although captivity was more lethal for C. spama larvae overall, survivors actually fared better than the average H. betteni larva which experienced mass loss. The overall higher mortality incurred by experimental C. spama populations could also have been in some way influenced by interspecific interactions, perhaps competition for high-quality retreat sites. Hydropsychid battles over retreat sites are common, and not surprisingly, the odds of intruder victory typically favor larger combatants (Jannsson and Vuoristo 1979). Therefore, because final-instar H. betteni are more than three 76 times larger than final-instar C. spama (Figure 2), the higher probability of winning battles over retreats may have influenced mortality in the spatially limited microcosm environments. In his excellent review of the effects of sedimentation in streams, Waters (1995: 60) proclaimed that ”on the basis of current knowledge, the direct effect of suspended sediment upon benthic invertebrates does not appear to be a significant influence upon stream invertebrate communities.” Results from this study, as well as those presented in Chapter 3, indicate that some benthic invertebrates are sensitive to, and potentially harmed by, high levels of sediment in suspension. Although the effects of sedimentation on invertebrate community composition are typically more the result of sediment deposition than transport, additional stress inflicted on benthos by elevated suspended sediment levels may impose chronic, low-level stress and therefore have indirect effects upon invertebrate communities. Given the enormity of the problem, it is conceivable that almost every lotic invertebrate population has been exposed to sedimentation that resulted from human activities. It seems certain, therefore, that widespread and chronic sedimentation-enhancing activities such as overgrazing of cattle in small stream riparian areas, have in aggregate, immense ecological consequences. 77 LITERATURE CITED Aldridge, D.W., B.S. Payne, and A.C. Miller. 1987. The effects of intermittent exposure to suspended solids and turbulence on three species of freshwater mussels. Environmental Pollution 45: 17-28. Cudney, MD. and J .B. Wallace. 1980. Life cycles, microdistribution, and production dynamics of six species of net-spinning eaddisflies in a large southeastern (USA) river. Holarctic Ecology 3: 169-182. Cuffney, T.F. and G.W. Minshall. 1981. Life history and bionomics of Arctopsyche grandis (Trichoptera) in a central Idaho stream. Holarctic Ecology 4: 252-262. Culp, J .M., F.J. Wrona, and R.W. Davies. 1986. Response of stream benthos and drift to fine sediment deposition versus transport. Canadian Journal of Zoology 64: 1345-1351. Fleischner, TL. 1994. Ecological costs of livestock grazing in western North America. Conservation Biology 8: 629-644. Fremling, CR. 1960. Biology of a large mayfly, Hexagenia bilineata (Say), of the Upper Mississippi River. Iowa Agricultural Experiment Station Research Bulletin 482: 842-852. Fuller, R.L. and RJ. Mackay. 1980. Field and laboratory studies of net—spinning activity by Hydropsyche larvae (Trichoptera: Hydropsychidae). Canadian Journal of Zoology 58: 2006-2014. Gammon, J .R. 1970. The effect of inorganic sediment on stream biota. USEPA Water Pollution Research Series 18050 DWC 12/70. Hamilton, J .D. 1961. The effect of sand-pit washings on stream fauna. Verh. int. Verein. theor. angew. Limnol. 14: 435-439. Hynes, H.B.N. 1966. W. Toronto: University of Toronto Press. Hynes, H.B.N. 1973. The effects of sediment on the biota of running water. Pages 653-663 in Fluvial Processes and Sedimentation. Canadian Department of the Environment, Ottawa. Jansson, A. and T. Vuoristo. 1979. Significance of stridulation in larval Hydropsychidae (Trichoptera). Behavior 71: 167-186. King, D.L., and RC. Ball. 1967. Comparative energetics of a polluted stream. Limnology and Oceanography 12: 27-33. Leamer, M.A., R. Williams, M. Harcup, and B.D. Hughes. 1971. A survey of the 78 macro-fauna of the River Cyon, a polluted tributary of the River Taff (South Wales). Freshwater Biology 1: 339-367. MacDonald, DD. and CF. Newcombe. 1993. Utility of the stress index for predicting suspended sediment effects: response to comment. North American Journal of Fisheries Management 13: 873-876. Mackay, R]. 1979. Life history patterns of of some species of Hydropsyche (Trichoptera: Hydropsychidae) in southern Ontario. Mackay, R]. 1984. Life history patterns of Hydropsyche bronta and H. morosa (Trichoptera: Hydropsychidae) in summer-warm rivers of southern Ontario. Canadian Journal of Zoology 62; 271-275. McCabe G.D. and J. O'Brien. 1983. The effects of suspended silt on feeding and reproduction of Daphnia pulex. The American Midland Naturalist 110: 324- 337. Newcombe, Cp. and DD. MacDonald. 1991. Effects of suspended sediments on aquatic ecosystems. North American Journal of Fisheries Management 11: 72- 82. Nutall, RM. and G.H. Bielby. 1973. The effects of china-clay wastes on stream invertebrates. Environmental Pollution 5: 77-86. Petersson, E. and A.T. Hasselrot. 1994. Mating and nectar feeding in the psychomyiid caddis fly Tinodes waeneri. Aquatic Insects 16: 1994: 177-187. Rosenberg, D.M., and A.P. Wiens. 1978. Effects of sediment addition on macrobenthic invertebrates in a northern Canadian river. Water Research 12: 753-763. Ross, H.H. 1944. The caddis flies, or Trichoptera, of Illinois. Bulletin of the Illinois Natural History Survey 23 1-326. Rutherford, J .E. 1986. Mortality in reared hydropsychid pupae (Trichoptera: Hydropsychidae). Hydrobiologia 132: 97-111. Rutherford, J .E. and RJ. Mackay. 1986. Patterns of pupal mortality in field populations of Hydropsyche and Cheumatopsyche (Trichoptera: Hydropsychidae). Freshwater Biology 16: 337-350. SAS. 1990. Version 6.04. SAS Institute, Cary, North Carolina, USA. Schmude, K.L. and W.L. Hilsenhoff. 1986. Biology, ecology, larval taxonomy, and 79 distribution of Hydropsychidae (Trichoptera) in Wisconsin. The Great lakes Entomologist 19: 123-145. Servizi, J .A. and D.W. Martens. 1991. Effect of temperature, season, and fish size on acute lethality of suspended sediments to Coho Salmon (Oncorhynchus kisutch). Canadian Journal of Fisheries and Aquatic Sciences 48: 493-497. Servizi, J .A. and D.W. Martens. 1992. Sublethal responses of Coho Salmon (Oncorynchus kisutch) to suspended sediments Canadian Journal of Fisheries and Aquatic Sciences 48: 49: 1389—1395. Sorenson, D.L., M.M. McCarthy, E.J. Middlebrooks, and DB. Porcella. 1977. Suspended and dissolved sediments effects on freshwater biota: a review. USEPA-600/3-77-042. Tait, C.K., J .L. Li, G.A. Lamberti, T.N. Pearsons, and H.W. Li. 1994. Relationships between riparian cover and community structure of high desert streams. Journal of the North Ameriean Benthological Society 13: 45-56. Thomas, V.G. 1985. Experimentally determined impacts of a small, suction gold dredge on a Montana stream. North American Journal of Fisheries Management 5: 480-488. Wallace, J .B. and D. Malas. 1976. The fine structure of eapture nets of larval Philopotamidae (Trichoptera) with special emphasis on Dolophilodes distinctus. Canadian Journal of Zoology 54: 1788-1802. Wallace, J.B. and R.W. Merritt. 1980. Filter-feeding ecology of aquatic insects. Annual Review of Entomology 25: 103-132. Wallace, J. B. and FF. Sherberger. 1974. The larval retreat and feeding net of Macmnema carolina Banks (Trichoptera: Hydropsychidae). Hydrobiologia 45: 177- 1 84. Wallace, J. B. and FF. Sherberger. 1975. The larval retreat and fading net of Macronema transversum Hagen (Trichoptera: Hydropsychidae). Animal Behavior 23: 592-596. Waters, T.F. 1984. Annual production by Gammarus pseudolimnaeus among substrate types in Valley, Creek, Minnesota. Ameriean Midland Naturalist 112: 95 -102. Waters, T.F. 1995. sew Ameriean Fisheries Society Monograph 7, Bethesda, Maryland. 80 Webster, J.R., S.W. Golladay, E.F. Benfield, J.L. Meyer, W.T. Swank, and J.B. Wallace. Catchment disturbance and stream response: an overview of stream research at Coweeta Hydrological laboratory. In Boon, P.J., P. Calow, and GE. Petts (eds.) River Conmag’on and M_a_n_agement. New York: John Wiley and Sons. 81 Table 1. Particle-size distribution and chemical composition of bank sediments used in sedimentation treatments. Soil Type: loam, pH 7.1 Mineral Component: 70.7 % Organic Component: 29.3% 45.8 % Sand Na 258 ppm 32.7 % Silt Cl 620 ppm 21.4 % Clay N03 5.4 ppm N H4 74.7 ppm 82 Table 2. ANOVA results comparing relative growth rates of Hydropsyche betteni and Certaopsyche spama in two sediment treatments and two controls (see Figure 2). Source df MS F P Species l 0.01 1 80.900 0.0001 Sed. Level 1 0.000 0.051 0.8221 Species x Sed. Level 1 0.000 0.291 0.5907 Tank (Sed. Level) 2 0.000 0.295 0.7452 Species x Tank (Sed. level) 2 0.000 1.968 0.1449 Error 103 0.000 83 Table 3. Contingency analyses testing whether Hydropsyche betteni and Ceratopsyche spama survival differed in sediment treatments and controls (two tanks each) (see Figure 3). Source df x2 P Species l 29.24 0.000 Sed. Level 1 9.54 0.002 Species x Sed. Level 1 4.38 0.036 Tank (Sed. Level) 2 0.16 0.923 Species x Tank (Sed. level) 2 0.16 0.923 84 25 ’ .... + Sediment Tanks + Control Tanks N O l 1 1 L 1 Nephelometric Turbidity Units <15 i“ A #1 A 1 I l o l I l I l l l l 0 1 45 60 90 120 180 240 360 Minutes After Onset of Sedimentation Figure 1. Mean NTU :1: 1 se in two sediment- treatment and two control tanks. 85 '8 - C, spama Initial Mace: ° 2.66 mg. i 0.25 'F‘ 0.061 : T (IL ill 0.02': '7'- J : El 4 .o .9 Relative Growth Rate (mg.dry mass/day) O 0.02; H. betteni lpl M f Inltlallhu: . 3.13 . 0.43 -o.04— "" i ~— 3 ._J -o.oe 1 Sedlment Treatments I Controls Figure 2. Box plots of mean RGR :i: 1 SE of Hydropsyche betteni and Ceratopsyche spama in two sedimented and two control tanks. Mean pre-trial dry mass (:1: 1 se) indicate relative size of fifth-instar H. betteni (n=24) and C. spama (n=2l). % Survival 86 Sediment Tanks Control Tanks C. spama H. betteni Figure. 3. Mean survivorship of larval Ceratopsyche spama and Hydropsyche betteni in two sedimented and two control tanks. Standard error = 0.000 for all but H. betteni control which is 0.020. 87 Appendix 1. Contingency analyses testing whether Hydropsyche betteni and Ceratopsyche spama survival differed in sediment treatments and controls with pupae included (see Table 3). Source df x2 P Species 1 34.84 0.001 Sed. level 1 6.76 0.000 Species x Sed. level 1 3.86 0.049 Tank (Sed. level) 2 0.32 0.850 Species x Tank (Sed. Level) 2 0.32 0.850 APPENDIX 1 88 APPENDIX 1 Record of Deposition of Voucher Specimens* The specimens listed on the following sheet(s) have been deposited in the named museum(s) as samples of those species or other taxa which were used in this research. Voucher recognition labels bearing the Voucher No. have been attached or included in fluid-preserved specimens. Voucher No.: 1996-8 Title of thesis or dissertation (or other research projects): Some effects of riparian habitat alteration on lotic invertebrate ecology Museum(s) where deposited and abbreviations for table on following sheets: Entomology Museum, Michigan State University (MSU) Other Museums: Investigator's Name (s) (typed) Roger Malcolm Strand Date 17 Septembe;l 1996 *Reference: Yoshimoto, C. M. 1978. Voucher Specimens for Entomology in Nerth America. Bull. Entomol. Soc. Amer. 24:141-42. Deposit as follows: Original: Include as Appendix 1 in ribbon copy of thesis or dissertation. Copies: Included as Appendix 1 in copies of thesis or dissertation. Museum(s) files. Research project files. This form is available from and the Voucher No. is assigned by the Curator, Michigan State University Entomology Museum. APPENDIX 1.1 89 APPENDIX 1.1 voucher Specimen Data Pages _'_of__'_ Page muen Houfi\mw. hufimuo>wep muwum cmwfinoHZo Mew ea ufimomce you muwfiaueen voumaa macaw one eo>aooem u mtomm. .oz noeo=o> cmm~ Honawummm ~_ muse vacuum aaooamz Howom Aeoekuv Amvuamz m.uouewaumo>eu Ahuwmmoooa ma muooan HMGOfiuweee omsv iversitfi Entomology Museum, Michigan State Un 4 (0.3 arna), 5 {H.betteni) sam_rwx-- .3.mm .z.ms. “m .xeouo mwufimum..oo mwaoH “Hz mmom wemuuen eaohmmouehm m emm_rwxlm_ .3.m_ omm .zome mw n“ uo>wm uefiam ..oo ucemma "H2 Ammomv meunmm osohmmoumueo m (_ _ Selig 5.3 5.3 (x as .xoouo mwuweum..oo mfieoH "Hz Auoxam3v eweea newumewnumm moor eouamoeoe was wen: we eeuuoHHoo nexeu ensue no mouuenm m e r r m m e .m u maoefioeeu new «use Henna e r o.d e .r .r a n. w s wm 3... m .w .m w m. a a ”M w.d.1 .U .A A" P. Nu.ru nu "we Honesz