26.12.2012 Views

Sponseller, R.A., E.F. Benfield, and H.M. Valett. 2001

Sponseller, R.A., E.F. Benfield, and H.M. Valett. 2001

Sponseller, R.A., E.F. Benfield, and H.M. Valett. 2001

SHOW MORE
SHOW LESS

Create successful ePaper yourself

Turn your PDF publications into a flip-book with our unique Google optimized e-Paper software.

Freshwater Biology (<strong>2001</strong>) 46,1409-1424<br />

APPLIED ISSUES<br />

Relationships between l<strong>and</strong> use, spatial scale<br />

<strong>and</strong> stream macroinvertebrate communities<br />

R. A. SPONSELLER, E. F. BENFIELD <strong>and</strong>H. M. VALETT<br />

Department of Biology, Virginia Polytechnic Institute <strong>and</strong> State University, Blacksburg, Virginia, U.S.A.<br />

SUMMARY<br />

1. The structure of lotic macroinvertebrate communities may be strongly influenced by<br />

l<strong>and</strong>-use practices within catchments. However, the relative magnitude of influence on<br />

the benthos may depend upon the spatial arrangement of different l<strong>and</strong> uses in the<br />

catchment.<br />

2. We examined the influence of l<strong>and</strong>-cover patterns on in-stream physico-chemical<br />

features <strong>and</strong> macroinvertebrate assemblages in nine southern Appalachian headwater<br />

basins characterized by a mixture of l<strong>and</strong>-use practices. Using a geographical information<br />

system (GIS)/remote sensing approach, we'quantified l<strong>and</strong>-cover at five spatial scales; the<br />

entire catchment, the-Tiparian corridojv<strong>and</strong> three riparian 'sub-corridors' extending 200,<br />

1000 <strong>and</strong> 2000 m upstream of sampling reaches.<br />

3. Stream water chemistry was generally related to features at the catchment scale.<br />

Conversely, stream temperature <strong>and</strong> substratum characteristics were strongly influenced<br />

by l<strong>and</strong>-cover patterns at the riparian corridor <strong>and</strong> sub-corridor scales.<br />

4. Macroinvertebrate assemblage structure was quantified using the slope of rankabundance<br />

plots, <strong>and</strong> further described using diversity <strong>and</strong> evenness indices. Taxon<br />

richness ranged from 24 to 54 among sites, <strong>and</strong> the analysis of rank-abundance curves<br />

defined three distinct groups with high, medium <strong>and</strong> low diversity. In general, other<br />

macroinvertebrate indices were in accord with rank-abundance groups, with richness <strong>and</strong><br />

evenness decreasing among sites with maximum stream temperature.<br />

5. Macroinvertebrate indices were most closely related to l<strong>and</strong>-cover patterns evaluated at<br />

the 200 m sub-corridor scale, suggesting that local, streamside development effectively<br />

alters assemblage structure.<br />

6. Results suggest that differences in macroinvertebrate assemblage structure can be<br />

explained by l<strong>and</strong>-cover patterns when appropriate spatial scales, are employed. In<br />

actdition, the influence of riparian forest patches on in-stream habitat features (e.g. the<br />

thermal regime) may be critical to the distribution of many taxa in headwater streams<br />

draining catchments with mixed l<strong>and</strong>-use practices.<br />

Keywords: geographical information system, l<strong>and</strong> use, macroinvertebrates, riparian corridor,<br />

temperature<br />

Introduction<br />

Maintenance of stream biodiversity in the face of<br />

•• encroaching human development has received much<br />

Correspondence: R. A. <strong>Sponseller</strong>, Department of Biology, attention in recent years (e.g. Allan & Flecker, 1993;<br />

Arizona State University, Tempe, Arizona 85287-1501, AZ, Harding et al., 1998). Furthermore, research has<br />

U.S.A. E-mail: rsponse@asu.edu<br />

© <strong>2001</strong> BlackweU Science Ltd 1409


1410 R.A. <strong>Sponseller</strong> et al.<br />

demonstrated that the conversion of forests to pastures<br />

<strong>and</strong>/or residential areas may influence<br />

in-stream habitat <strong>and</strong> macroinvertebrate communities<br />

in several ways. Loss of terrestrial vegetation (Swank,<br />

Swift & Douglas, 1988) <strong>and</strong> an increased area of<br />

impervious surfaces (Changnon & Demissie, 1996)<br />

can influence evapotranspiration <strong>and</strong> infiltration <strong>and</strong><br />

alter natural flow regimes (Poff et al., 1997). Many<br />

l<strong>and</strong>-use practices increase sediment inputs to<br />

streams, altering substratum characteristics <strong>and</strong> channel<br />

morphology, often reducing macroinvertebrate<br />

diversity (Lenat & Crawford, 1994; Waters, 1995;<br />

Quinn et al., 1997). Removal of streamside vegetation<br />

<strong>and</strong> subsequent increased solar radiation reaching the<br />

stream channel can increase temperature (Burton &<br />

Likens, 1973; Rutherford et al., 1997; Quinn et al,<br />

1997) <strong>and</strong> alter thermal regimes that are critical to<br />

the life history <strong>and</strong> ecology of macroinvertebrates<br />

(Vannote & Sweeney, 1980; Ward & Stanford, 1982;<br />

Quinn et al., 1994). In association with.altered catchment-<br />

hydrology <strong>and</strong> l<strong>and</strong> use, inputs of inorganic<br />

-nutrients from terrestrial sources (e.g. Omernik, 1977;<br />

Hunsaker & Levine, 1995; Johnson et al., 1997) may<br />

interact with increased light availability <strong>and</strong> stream<br />

temperature to enhance in-stream primary production<br />

(Webster et al., 1983), resulting in changes in the<br />

trophic structure of benthic communities (e.g. Gurtz<br />

& Wallace, 1984).<br />

The relative magnitude of l<strong>and</strong>-use effects may<br />

depend upon the spatial distribution of l<strong>and</strong> use in<br />

catchments (Allan & Johnson, 1997). Catchment<br />

hydrology, as well as the availability of inorganic<br />

nutrients in streams, is often related to processes that<br />

occur across the terrestrial l<strong>and</strong>scape (Omernik, 1977;<br />

Swank, Swift & Douglas, 1988; Hunsaker & Levine,<br />

1995). Conversely, the availability of light <strong>and</strong> organic<br />

carbon in streams is often related to processes restricted<br />

to the scale of streamside vegetation (Gregory<br />

et al., 1991). Several studies have shown that in-stream<br />

physico-chemical <strong>and</strong> biotic features may be constrained<br />

by catchment properties operating at different<br />

spatial scales. For example, Richards, Johnson &<br />

Host (1996) identified catchment-scale characteristics,<br />

particularly basin geology <strong>and</strong> the distribution of<br />

arable agriculture, as features mediating channel<br />

morphology <strong>and</strong> stream hydrology. At the same time,<br />

l<strong>and</strong>-cover in the riparian corridor had a strong<br />

influence on bank erosion <strong>and</strong> in-stream sedimentrelated<br />

variables. Further, they found that basin-wide<br />

features were the best predictors for macroinvertebrate<br />

assemblage structure. Conversely, Richards et al.<br />

(1997) showed that despite a close relationship<br />

between catchment-scale properties <strong>and</strong> channel<br />

structure <strong>and</strong> hydrology, macroinvertebrate species<br />

traits were correlated with local (reach-scale) features.<br />

In addition, spatial scales of catchment control may<br />

vary temporally. For example, Johnson et al. (1997)<br />

found that the strength of relationship between<br />

stream-water chemistry <strong>and</strong> catchment scale versus<br />

riparian scale l<strong>and</strong>-cover patterns varied among season<br />

<strong>and</strong> with dissolved chemical constituents in<br />

Michigan streams.<br />

The southern Appalachian region of the United<br />

States has been subject to a variety of l<strong>and</strong>-use<br />

changes during the last century (<strong>Benfield</strong>, 1995).<br />

Between 1880 <strong>and</strong> 1920, almost all of the timber in<br />

the states of Virginia, West Virginia, North Carolina<br />

<strong>and</strong> Tennessee was harvested (Yarnell, 1998). This<br />

•was followed by agricultural activity, which<br />

increased in importance through to the 1960s. Since<br />

1970 human population in the southern Appalachians<br />

has increased by nearly 30% <strong>and</strong> agricultural<br />

activity has declined by 31%, leading to a largescale<br />

conversion of agricultural l<strong>and</strong> to either<br />

forests or residential areas (SAMAB, 1996). This<br />

history has resulted in a mosaic of different l<strong>and</strong>use<br />

patches across the l<strong>and</strong>scape. Current l<strong>and</strong>cover<br />

patterns in many Southern Appalachian<br />

catchments are characterized by high gradient<br />

headwater areas that are difficult to develop <strong>and</strong><br />

are often forested, <strong>and</strong> low gradient areas subject to<br />

active agriculture or conversion to urban l<strong>and</strong><br />

(Wear & Bolstad, 1998).<br />

In this study, we investigated the effects of l<strong>and</strong>-use<br />

on benthic macroinvertebrate assemblages in southern<br />

Appalachian headwater streams. Specifically, we<br />

address how differences in l<strong>and</strong> use among catchments<br />

influences in-stream physico-chemical features<br />

<strong>and</strong> macroinvertebrate assemblage structure. In addition,<br />

we sought to determine how changes in<br />

in-stream features (both biotic <strong>and</strong> abiotic) relate to<br />

the spatial arrangement of l<strong>and</strong> use within catchments.<br />

To do this, we quantified the relationship<br />

between l<strong>and</strong>-cover <strong>and</strong> in-stream variables at multiple<br />

spatial scales. This approach allowed us to<br />

evaluate the ecological implications of l<strong>and</strong>-use practices<br />

that occur streamside versus those that occur<br />

across the catchment.<br />

© <strong>2001</strong> Bkckwell Science Ltd, Freshwater Biology, 46,1409-1424


Methods<br />

Site description<br />

The study was conducted in the Upper Roanoke River<br />

Basin (URRB) in south-western Virginia (Fig. 1) at the<br />

interface of the Appalachian Valley <strong>and</strong> Ridge <strong>and</strong><br />

Blue Ridge physiographic provinces. The URRB is<br />

topographically <strong>and</strong> geologically diverse, <strong>and</strong>.is characterized<br />

by Precambrian <strong>and</strong> Cambrian metamorphics<br />

<strong>and</strong> elastics at higher altitude <strong>and</strong> Cambrian <strong>and</strong><br />

Ordovician carbonates lower in the valley (Waller,<br />

1976). Nine catchments with second or third order<br />

streams were selected for study. Catchments ranged in<br />

altitude from 325 to 575 m above sea level, <strong>and</strong> varied<br />

from 278 to 1014 hectares. Six of nine catchments were<br />

characterized by some degree of human activity (e.g.<br />

agricultural or urban development); the remaining<br />

three had no such development. One 50-m stream<br />

North Fork<br />

Influences of l<strong>and</strong> use on stream macroinvertebrates 1411<br />

reach was selected within each catchment as our study<br />

site. Benthic habitat at all streams consisted of particles<br />

ranging from silt to cobbles, although the relative<br />

proportion of these size classes varied among sites.<br />

L<strong>and</strong> use quantification<br />

A geographical information system (GIS) (Arcview 3.1;<br />

ESRI, Redl<strong>and</strong>, CA, U.S.A.) was used to quantify l<strong>and</strong><br />

cover within each basin. Digital l<strong>and</strong>-cover information<br />

for the Roanoke Valley was obtained from a<br />

preliminary l<strong>and</strong>-cover map of Virginia, produced<br />

through the Virginia Gap Analysis Project (VAGAP)<br />

for 1992 (Morton, 1998). This statewide data set was<br />

generated from 14 L<strong>and</strong>sat thematic mapper scenes,<br />

<strong>and</strong> classified using both unsupervised <strong>and</strong> enhanced<br />

supervised methods (Morton, 1998). Pixels provided<br />

30 x 30 m resolution <strong>and</strong> included information from<br />

North<br />

10 km<br />

Fig. 1 Upper Roanoke River Basin (URRB), showing North <strong>and</strong> South Fork sub-basins, <strong>and</strong> selected study sites. Site names correspond<br />

as follows: (1) Powers Branch, (2) Sugar Run, (3) Purgatory Tributary, (4) Martins Creek, (5) Little Back Creek, (6) Greenbriar Branch,<br />

(7) Mudlick Creek, (8) Earnhardt Creek, (9) Franklin Creek.<br />

© <strong>2001</strong> BlackweU Science Ltd, Freshwater Biology, 46,1409-1424


1412 R.A. <strong>Sponseller</strong> et al.<br />

seven l<strong>and</strong>-use categories: deciduous forest, coniferous<br />

forest, mixed forest, shrub/scrubl<strong>and</strong>, herbaceous<br />

(mostly agriculture), open water <strong>and</strong> disturbed (areas<br />

lacking vegetation). We further categorized the data<br />

into forest <strong>and</strong> non-forest. The forest category included<br />

all forest types, but was dominated by the<br />

deciduous class. The non-forest category included<br />

both agricultural <strong>and</strong> urban/suburban areas.<br />

Morton (1998) used aerial videography to acquire<br />

reference data <strong>and</strong> assess the accuracy of the statewide<br />

l<strong>and</strong>-cover data set, which was found to be 81%.<br />

This level of accuracy may not hold for riparian<br />

corridors, which are more difficult to classify, particularly<br />

in mountainous areas. Our accuracy assessment<br />

of riparian corridors, from 25 locations across five<br />

study catchments with the greatest l<strong>and</strong>-cover heterogeneity,<br />

found approximately 75% of the streamside<br />

pixels were correctly classified. However, of the pixels<br />

classified as non-forest, 50% had' thin' ;strips (e.g.<br />

< 1 m) of riparian vegetation but otherwise lacked<br />

forested 'area. Incorrect classifications encountered in<br />

-the assessment generally occurred when pixels classified<br />

as forest had small residences but were otherwise<br />

well forested.<br />

Sampling reach<br />

Riparian subcorridor<br />

polygons<br />

Catchments were delineated using a watershed<br />

delineator (ESRI & Texas Natural Resource Conservation<br />

Commission, 1997) which uses neighbourhood<br />

functions with data from US Geologic Survey (USGS)<br />

digital elevation maps to quantify the surface area<br />

contributing to drainage through a given point. The<br />

outputs were converted to catchment polygons, which<br />

included the entire drainage area upstream of the<br />

sampling reach. L<strong>and</strong>-cover derived from all pixels<br />

within catchment polygons were used to characterize<br />

catchment scale l<strong>and</strong>-use patterns. Within each catchment<br />

polygon, a 60-m riparian corridor polygon was<br />

created with 30 m of width extending laterally,<br />

beginning at the sampling reach <strong>and</strong> continuing for<br />

the entire length of the stream. The influence of<br />

riparian l<strong>and</strong>-cover at specific distances upstream of<br />

the study reach was assessed by dividing riparian<br />

corridors longitudinally into sub-corridor polygons.<br />

We used five spatial scales for analysis: (1) the<br />

•catchment scale, including all of the area within its<br />

watershed, (2) riparian corridor scale, extending over<br />

the entire length of the stream, (3) 2000 m riparian<br />

sub-corridors, (4) 1000 m riparian sub-corridors <strong>and</strong><br />

(5) 200 m riparian sub-corridors (Fig. 2). Associated<br />

Catchment polygon<br />

Riparian corridor<br />

polygon<br />

Fig. 2 Examples of GIS polygons used to analyse l<strong>and</strong>-cover patterns at five spatial scales, including the entire catchment, riparian<br />

corridor <strong>and</strong> three riparian sub-corridors extending 200,1000 <strong>and</strong> 2000 m upstream from the sampling reach.<br />

© <strong>2001</strong> Blackwell Science Ltd, Freshwater Biology, 46,1409-1424


catchment, riparian corridor <strong>and</strong> sub-corridor polygons<br />

were overlaid onto the digital l<strong>and</strong>-cover data<br />

for each study basin. Percentages for l<strong>and</strong>-cover<br />

categories were quantified within all spatial polygons.<br />

Regression analysis was used to relate physicochemical<br />

<strong>and</strong> biological variables to l<strong>and</strong> cover at the<br />

catchment, riparian corridor <strong>and</strong> sub-corridor scale.<br />

Percent non-forest values were transformed .(arcsine<br />

square root) for statistical analyses.<br />

Physical <strong>and</strong> chemical characterizations<br />

Physical <strong>and</strong> chemical measurements were made for<br />

each stream from January to December 1999. Triplicate<br />

water samples were collected monthly from each site,<br />

passed through glass fibre filters (Whatman GFF,<br />

Gelman Type AE, Maidstone, U.K.), <strong>and</strong> kept frozen<br />

until analysed. Samples were analysed for ammonium-nitrogen<br />

(NH4-N) using the phenaj;e' ,'frtethod<br />

(Soloranzo, 1969) <strong>and</strong> nitrate-nitrogen, (NO3-N) by<br />

colormetric 'techniques following reduction by Cd<br />

(Wood, Armstrong & Richards, 1967) on a Technicon<br />

Auto-analyser (Bran & Luebbe, Buffalo Grove, IL,<br />

U.S.A.). Total inorganic nitrogen (TIN) was calculated<br />

as the sum of NH4-N <strong>and</strong> NO3-N. Ortho-phosphate<br />

(PO4-P) was analysed as soluble reactive phosphorus<br />

(SRP) using the molybdate colorimetric method<br />

(Murphy & Riley, 1962; Wetzel & Likens, 1991). Specific<br />

conductance was measured seasonally using a YSI<br />

Model 30/50- (YSI, Yellow Springs, OH, U.S.A.) conductivity<br />

meter. Alkalinity was determined once<br />

at each site by st<strong>and</strong>ard methods (APHA, 1998).<br />

Temperature was recorded monthly until continuous<br />

temperature data recorders (HOBO, Polasset, MA,<br />

U.S.A.) were installed in March 1999, after which it was<br />

recorded hourly through November 1999. Temperature<br />

variables used in analyses included the overall<br />

mean temperature (T, °C), <strong>and</strong> maximum summer<br />

temperature (Tmax, °C). Discharge was measured at<br />

least monthly at each site using velocity determined<br />

with an electronic flowmeter (FLOW-MATE, Marsh-<br />

McBirney, Frederick, MD, U.S.A.) <strong>and</strong> cross-sectional<br />

area (Gore, 1996). Mean substratum particle size (MPS)<br />

was determined using st<strong>and</strong>ard granulometry techniques<br />

(sensu Wolman, 1954) on 100 r<strong>and</strong>omly selected<br />

rocks at each site with a USGS gravelometer. We also<br />

quantified the percentage of all particles measured at<br />

each site small enough to be categorized as coarse<br />

gravel, s<strong>and</strong> <strong>and</strong> silt (% < 16 mm).<br />

© <strong>2001</strong> BlackweU Science Ltd, Freshwater Biology, 46,1409-1424<br />

Influences of l<strong>and</strong> use on stream macroinvertebrates 1413<br />

Benthic algal biomass<br />

Epilithic biomass (g m" 2 ) <strong>and</strong> chlorophyll a (mg m" 2 )<br />

were determined monthly at each site from the<br />

surfaces of five rocks collected along each 50 m reach.<br />

Upon collection, rocks were packed in ice, <strong>and</strong><br />

returned to the laboratory for processing. Upper<br />

surfaces were scrubbed <strong>and</strong> the resultant slurry subsampled<br />

<strong>and</strong> collected on preweighed glass fibre filters<br />

(Whatman GFF, Gelman Type AE). Filters were incubated<br />

in a 90% buffered acetone solution for 24 h (sensu<br />

Steinman & Lamberti, 1996). Photosynthetic pigments<br />

(chlorophyll a <strong>and</strong> phaeophytin) were measured using<br />

spectrophotometry (at 664, 665 <strong>and</strong> 750 nm) on a<br />

Shimadzu UV-1601 spectrophotometer. An additional<br />

subsample was taken to determine epilithic ash-free<br />

dry mass (AFDM) by gravimetric techniques (drying at<br />

50 °C for 48 h, followed by ashing at 550 °C for 1 h).<br />

Photo'synthetic pigment concentration (mg m" 2 ) <strong>and</strong><br />

^AFDM (g m~ 2 ) were normalized to surface area determined<br />

by covering scraped surfaces with aluminium<br />

foil of known mass per unit area.<br />

Invertebrate assemblages<br />

Benthic invertebrates were collected from all sites in<br />

April 1999. Five Hess samples (surface area,<br />

0.08 m 2 sample" 1 ; mesh size, 250 mm) were taken<br />

along each 50 m reach. Samples were preserved in<br />

80% ethanol, <strong>and</strong> individuals identified to genus<br />

following Merritt & Cummins (1996), Stewart & Stark<br />

(1993) <strong>and</strong> Wiggins (1996). Chironomid larvae from<br />

all samples within a site were combined, quantitatively<br />

subsampled, mounted <strong>and</strong> identified following<br />

Epler (1995) <strong>and</strong> Merritt & Cummins (1996).<br />

To address changes in benthic community structure<br />

among sites, we quantified traditional measures of<br />

richness <strong>and</strong> evenness, <strong>and</strong> a compositional index<br />

commonly used to indicate environmental stress.<br />

Indices used to assess diversity included taxon number<br />

(S), <strong>and</strong> the Shannon-Weiner Diversity Index:<br />

where<br />

Pi represents the proportion of individuals found<br />

in the ith taxon, with values summed across all taxa<br />

(S). We also used the associated evenness measure


1414 R.A. <strong>Sponseller</strong> et al.<br />

calculated from H'. This index assumes that maximum<br />

diversity occurs when all taxa are equally<br />

abundant (i.e. H' = H'max = hi S). Therefore, the ratio<br />

of H'/H'max represents a measure of evenness where<br />

an assemblage with an equal abundance of taxa<br />

would have a value of 1 (Pielou, 1969). As an<br />

additional measure of evenness we quantified the<br />

percentage of the total numbers accounted for by the<br />

five most dominant taxa at each site (% 5 Dominant)<br />

(Barbour et al., 1999). Finally, to assess compositional<br />

differences among sites, we quantified the taxonomic<br />

richness of commonly intolerant taxa (Ephemeroptera<br />

+ Plecoptera + Trichoptera, EFT), widely used as an<br />

indicator of disturbance to stream communities (Lenat<br />

& Crawford, 1994; Wallace et al., 1996).<br />

Rank-abundance curves (sensu; Hawkins, Murphy<br />

& Anderson, 1982; Marsh-Mathews & Mathews, 2000)<br />

were generated for invertebrate assemblages at each<br />

site. Curves consisted of lognormal;:'proportional<br />

abundance (pi) for each taxon plptted against the<br />

corresponding taxonomic abundance ranking. The<br />

-slope of this line (Alog p;/n, where n = number of<br />

taxa) was used as an integrative measure of taxon<br />

richness <strong>and</strong> evenness. Diversity is maximized when<br />

the slope of the curve approaches zero. Slopes become<br />

increasingly negative when the proportional abundance<br />

(pi) of the more dominant taxa increases<br />

(i.e. evenness decreases), or when taxon richness<br />

decreases. We used the absolute value of the slope<br />

for all analyses; therefore, larger values represent<br />

steeper rank-abundance curves, <strong>and</strong> thus, less diverse<br />

assemblages.<br />

Data analysis<br />

Bivariate regression analysis was used to relate l<strong>and</strong><br />

cover variables (e.g. % non-forest) at each spatial scale<br />

to the in-stream physico-chemical variables <strong>and</strong><br />

macroinvertebrate indices measured at all sites. In<br />

Site number<br />

Scale<br />

Catchment<br />

Entire riparian<br />

corridor<br />

Sub-corridors (m)<br />

200<br />

1000<br />

2000<br />

1<br />

7.5<br />

1.2<br />

0.0<br />

0.0<br />

0.0<br />

2<br />

11.7<br />

1.9<br />

0.0<br />

0.0<br />

0.8<br />

3<br />

13.9<br />

1.3<br />

0.0<br />

3.1<br />

1.5<br />

4<br />

20.3<br />

56.3<br />

75.0<br />

92.2<br />

77.6<br />

5<br />

8.4<br />

10.8<br />

23.1<br />

7.3<br />

3.7<br />

6<br />

34.2<br />

19.2<br />

14.3<br />

4.3<br />

22.8<br />

addition, macroinvertebrate indices were used as<br />

dependent variables in stepwise multiple regression<br />

analyses with in-stream physico-chemical parameters.<br />

Variables that failed to meet a 0.05 probability value<br />

were removed from multiple regression models.<br />

Regression analysis was used to test rank-abundance<br />

curves for linearity <strong>and</strong> the slope of the lognormal<br />

plots was compared among sites using methods<br />

described in Zar (1996) with a Bonferroni correction<br />

for multiple comparisons. All regression analyses<br />

were performed on SAS 7.0 (SAS Inc., Gary, NC,<br />

U.S.A.).<br />

Results<br />

The magnitude <strong>and</strong> spatial arrangement of nonforested<br />

l<strong>and</strong> varied considerably among drainage<br />

basins at the catchment, riparian corridor <strong>and</strong> subcpTridor<br />

scales (Table 1). In addition, the proportion<br />

•of non-forested area within any given riparian corridor<br />

often varied greatly when compared with that of<br />

the associated catchment. For example, in four basins<br />

(Table 1, sites 1, 2, 3 <strong>and</strong> 6), the percentage non-forest<br />

in the entire riparian corridor was approximately 25%<br />

of that observed at the scale of the entire catchment;<br />

indeed sites 2 <strong>and</strong> 3 were nearly an order of<br />

magnitude lower. At the same time, the percentage<br />

non-forest within the riparian corridor increased 2.8<br />

(site 4), 1.3 (site 5), 1.4 (site 8) fold at three other sites.<br />

Within riparian corridors, the distribution of nonforested<br />

area varied in proximity to the study reach<br />

(Table 1). In basins with low non-forest cover at both<br />

the catchment <strong>and</strong> total riparian corridor scales, nonforest<br />

cover in sub-corridors was also low. For<br />

example, at sites 1-3, six of nine riparian sub-corridors<br />

were entirely forested <strong>and</strong> a maximum of only 3%<br />

was non-forest (Table 1). For basins less forested at<br />

broader spatial scales, the percentage non-forest in<br />

sub-corridors often increased substantially. For the<br />

7<br />

41.4<br />

34.7<br />

85.7<br />

80.8<br />

49.3<br />

8<br />

19.3<br />

26.4<br />

92.9<br />

55.9<br />

51.1<br />

9<br />

19.7<br />

19.9<br />

84.6<br />

77.0<br />

54.6<br />

Table 1 L<strong>and</strong>-cover patterns for the nine<br />

study basins. Values are percent nonforested<br />

l<strong>and</strong> within each scale category.<br />

Sub-corridor values are the percent of<br />

non-forested l<strong>and</strong> within a polygon of<br />

60 m width <strong>and</strong> progressively greater<br />

distances upstream from the sampling<br />

reach<br />

© <strong>2001</strong> Blackwell Science Ltd, Freshwater Biology, 46,1409-1424


four basins with the least forest cover at the catchment<br />

scale,% non-forest in the sub-corridors averaged 58%<br />

<strong>and</strong> was as great as 93%. Finally, longitudinal<br />

changes in forest cover within riparian corridors<br />

varied greatly among basins. For instance, at site 9.<br />

20% of l<strong>and</strong> cover was non-forest at both the catchment<br />

<strong>and</strong> entire riparian corridor scales. However, if<br />

only the first kilometre of the riparian corridor was<br />

considered, 80% of the l<strong>and</strong>scape was non-forest.<br />

Conversely, site 6 was classified as 34% non-forest<br />

at the catchment scale, but only 19 <strong>and</strong> 4% was<br />

non-forest at the entire riparian corridor <strong>and</strong> 1000 m<br />

sub-corridor scales, respectively.<br />

Physical responses<br />

Variation in several in-stream physical properties was<br />

related to catchment features <strong>and</strong>/or l<strong>and</strong>-cover<br />

patterns at multiple spatial scales (Table;2). For<br />

example, stream discharge ranged from 15.3 L s" 1<br />

(site 3) to-55.4 L s" 1 (site 5), varied temporally to a<br />

similar extent among streams with annual CVs<br />

ranging from 46 to 83% <strong>and</strong> was closely related to<br />

basin area (r 2 = 0.85, n = 9, P - 0.004). Mean stream<br />

temperature varied by only 4.3 °C among sites, but<br />

maximum temperature varied by 6.7 °C, with the<br />

greatest mean temperature <strong>and</strong> highest maximum<br />

temperature both at site 7. Mean stream temperature<br />

decreased with increasing altitude (r 2 = 0.78, n = 9,<br />

P = 0.002), <strong>and</strong> multiple regression indicated that the<br />

combination of altitude <strong>and</strong> the percentage non-forest<br />

at the entire riparian corridor scale explained 93% of<br />

the variation in temperature (n = 9, P = 0.02). Maximum<br />

stream temperature increased with the percentage<br />

of non-forested l<strong>and</strong> at the entire riparian<br />

corridor scale (^^.79, n = 9, P = 0.001) <strong>and</strong> at the<br />

sub-corridor scales of 2000 <strong>and</strong> 1000 m (r 2 = 0.71,<br />

n = 9, P = 0.005; r 2 = 0.71, n = 9, P = 0.005, respectively).<br />

However, the strongest relationship between<br />

maximum stream temperature <strong>and</strong> the percentage<br />

non-forest was found at the 200 m riparian subcorridor<br />

scale (r 2 = 0.81, n = 9, P = 0.001). The mean<br />

size of coarse substratum ranged from 44.8 mm<br />

at site 8 to 72.4 mm at site 2, <strong>and</strong> decreased with<br />

the percentage of non-forested l<strong>and</strong> at the catchment<br />

(r 2 = 0.56, K = 9, P = 0.02), riparian corridor<br />

(r 2 = 0.74, n = 9,P= 0.003) <strong>and</strong> 2000,1000 <strong>and</strong> 200 m<br />

<strong>and</strong> sub-corridor (r 2 = 0.80, n - 9, P = 0.001; r 2 =<br />

0.61, n = 9, P = 0.013; ^ = 0.74, n = 9, P = 0.003,<br />

© <strong>2001</strong> Blackwell Science Ltd, Freshwater Biology, 46,1409-1424<br />

Influences of l<strong>and</strong> use on stream macroinvertebrates 1415<br />

S ON<br />

CO<br />

-H-H -H<br />

vo O<br />

-H -H<br />

Os O t*s<br />

^* o co<br />

-H-H -H<br />

O<br />

CM O<br />

•H -H<br />

oq<br />

co<br />

•H<br />

CM O O O rH IS<br />

•H-H-H-H-H-H<br />

CM O O O<br />

CM O O O r-t CO<br />

•H-H-H-H-H-H<br />

T-H O O O T-H<br />

t-< O O O I-H CO<br />

•H-H-H-H-H-H<br />

ON O O O i— CO<br />

•H-H-H-H-H-H<br />

ON O in I-H O O O rH. CO<br />

•H-H -H -H-H-H-H-H-H<br />

p i—i co ON oo o vo o in oo oo ON<br />

O *** r-t O Ift I-H<br />

o o o<br />

vq o<br />

co d<br />

-H -H<br />

^< (9 p<br />

co d in<br />

•H-H -H<br />

CO CM CO rn<br />

-s<br />

3<br />

-H-H •«<br />

'•O ON ts •* CO O<br />

VO<br />

ON p<br />

•H -H<br />

I-H O O O<br />

*O Cx O O rH<br />

-H -H -H -H -H -H -H<br />

=9<br />

•H<br />

t-< d d d<br />

CM<br />

O<br />

o=<br />

"5<br />

K o o O I-H<br />

•H -H -H -H -H<br />

O I-H<br />

g S •* o<br />

*>.<br />

•H<br />

O O<br />

-H -H •H -H<br />

00 CM ON ON<br />

-H-H<br />

p in<br />

•H -H


1416 R.A. <strong>Sponseller</strong> et al.<br />

respectively) scales. The percentage of particles<br />

< 16 mm ranged from 17.1 at site 3 to 44.4 at site 8,<br />

<strong>and</strong> was only weakly related to the percentage nonforest<br />

within 200 m sub-corridor polygons (r 2 = 0.52,<br />

n = 9, P = 0.03).<br />

Chemical responses<br />

Chemical variables were also related to catchment<br />

features or l<strong>and</strong>-cover patterns at the catchment<br />

scale. Alkalinity ranged from 16 mg IT 1 at site 1 to<br />

98 mg LT 1 at site 9 <strong>and</strong> decreased with catchment<br />

altitude (r 2 = 0.63, n = 9, P = 0.01). In addition,<br />

specific conductance (us cm" 1 ) varied among sites<br />

from 47.3 at site 1 to 261.8 at site 9, <strong>and</strong> was strongly<br />

related to alkalinity (r 2 = 0.95, n = 9, P = 0.0001) <strong>and</strong><br />

altitude (r 2 = 0.67, n = 9, P = 0.007). Total inorganic<br />

N varied among sites (CV = 80%), with minimum<br />

values of 0.105 mg IT 1 (site 5) <strong>and</strong> ma)drrnlm values<br />

of 0.932 mg LT 1 (site 6), <strong>and</strong> increased with the<br />

percentage of non-forested l<strong>and</strong> at the catchment<br />

-scale (r 2 = 0.59, n=9,P = 0.02). Soluble reactive<br />

phosphorus did not differ greatly among sites<br />

(CV = 38%) <strong>and</strong> did not correspond to any physical<br />

features of the catchment, or indeed l<strong>and</strong>-cover at<br />

any spatial scale.<br />

Biological responses<br />

Based on field observation, algal assemblages were<br />

generally dominated by diatoms, although filamentous<br />

algae (e.g. Cladophora sp.) were found in abundance<br />

in streams draining catchments with extensive<br />

agricultural or urban development (e.g. sites 8 <strong>and</strong> 9).<br />

Mean annual epilithic chlorophyll a concentration<br />

ranged from a"rnjnimum of just 7.0 mg m~ 2 at site 3 to<br />

121.1 mg m~ 2 at site 8. Chlorophyll a values were<br />

lowest for sites 1-3, where averages ranged from 7.02<br />

to 27.9 mg m~ 2 ; concentrations were higher at the<br />

remaining sites with averages ranging from<br />

54.8 mg m~ 2 (site 6) to 121.1 mg m~ 2 (site 8) (Table 2).<br />

Bivariate regression analysis did show a relationship<br />

between chlorophyll a concentration <strong>and</strong> mean stream<br />

temperature (r 2 = 0.66, n = 9, P = 0.008). However,<br />

multiple regression analysis incorporated only maximum<br />

temperature into the model, explaining 73% of<br />

the variability in chlorophyll a, (n = 9, P = 0.003).<br />

Epilithic biomass ranged from 5.3 g m~ 2 at site 3<br />

to 33.8 g rrf 2 at site 9, <strong>and</strong> was closely related to<br />

epilithic chlorophyll a concentration (r 2 = 0.80, n = 9,<br />

P = 0.001).<br />

Macroinvertebrate density varied among sites from<br />

6974 m" 2 at site 3 to 15 730 m~ 2 at site 9 (Table 3).<br />

Multiple regression analysis with density as the<br />

dependent variable incorporated only epilithic biomass<br />

into the model, accounting for 87% of the<br />

variability among sites (n = 9, P- 0.0003). A total of<br />

77 macroinvertebrate taxa were identified from the<br />

nine sites. Taxon richness ranged from 24 at site 8 to<br />

54 at site 3 with five EFT taxa at site 8 <strong>and</strong> 22 EFT taxa<br />

at sites 2 <strong>and</strong> 3 (Table 3). Results of the comparison of<br />

rank-abundance slopes indicated significant among<br />

site differences (P < 0.001, after Bonferroni correction),<br />

with three general groups emerging that represent<br />

high, medium <strong>and</strong> low diversity (Fig. 3). All<br />

rank-abundance curves fit linear models (r 2 > 0.9,<br />

P > 0.001) <strong>and</strong> slope values ranged from -0.0470 at<br />

bbth sites 1 <strong>and</strong> 2 to -0.1100 at site 7. Stepwise<br />

'multiple regression analyses between macroinvertebrate<br />

indices <strong>and</strong> in-stream physico-chemical variables<br />

generally incorporated a single independent<br />

variable into the models (Table 4). For example,<br />

the slope of rank-abundance curves increased with<br />

epilithic chlorophyll a concentration (r 2 = 0.64,<br />

n = 9, P = 0.009). Total taxon richness (S) <strong>and</strong> EFT<br />

richness decreased with maximum stream temperature<br />

(^ = 0.54, n=9, P = 0.02, 7^ = 0.75, n = 9,<br />

P = 0.003). Similarly, diversity (HO <strong>and</strong> evenness<br />

(H'/H'max) also decreased with maximum stream<br />

temperature (r 2 = 0.72, n = 9,P = 0.004; r 2 = 0.83,<br />

« = 9, P = 0.002, respectively). The percent 5 dominant<br />

taxa ranged from 55% at site 3 to 93% at site 8,<br />

<strong>and</strong> increased with increasing maximum stream<br />

temperature (r 2 = 0.75, « = 9, P = 0.003).<br />

The structure of macroinvertebrate assemblages<br />

corresponded to the variance in l<strong>and</strong>-cover patterns<br />

among catchments (Table 5). Macroinvertebrate density<br />

was most closely related to the percentage of<br />

non-forest l<strong>and</strong> at the 2000 m sub-corridor scale.<br />

Rank-abundance slopes increased with the percentage<br />

of non-forested l<strong>and</strong> at the entire riparian corridor,<br />

<strong>and</strong> at 2000 <strong>and</strong> 1000 m sub-corridor scales, but<br />

were most strongly associated with l<strong>and</strong>-cover patterns<br />

at the 200 m sub-corridor scale. In general, other<br />

measures of diversity, evenness <strong>and</strong> composition<br />

most closely corresponded to the percentage nonforest<br />

at the 200 m sub-corridor scale. Shannon's<br />

evenness decreased with the percentage non-forest,<br />

© <strong>2001</strong> Blackwell Science Ltd, Freshwater Biology, 46,1409-1424


aence Ltd, Freshwater Biology, 4<br />

\<br />

Table 3 Benthic macroinvertebrate indices for each study site. Density data are mean ± SE for five Hess samples from each 50 m study reach. Sites are presented in groups<br />

with other sites having statistically similar rank-abundance slopes<br />

Rank abundance<br />

diversity group<br />

Site number<br />

Density<br />

Taxonrichness (S)<br />

EFT taxa<br />

H'<br />

W'/H'max<br />

%5 Dominant<br />

W <br />

o '<br />

CO<br />

o '<br />

o<br />

1*<br />

o<br />

IO<br />

L*<br />

Jco<br />

IcD<br />

9<br />

15730 ± 1334<br />

40<br />

16<br />

1.98<br />

0.54<br />

83.3<br />

fe<br />

Low<br />

8<br />

13907 ± 671<br />

30<br />

11<br />

1.78<br />

0.52<br />

89.5<br />

7<br />

8923 ± 2664<br />

24<br />

5<br />

1.20<br />

0.38<br />

92.9<br />

Proportional abundance (P,)<br />

p p -» _* _k<br />

2 - ° S


1418 R.A. <strong>Sponseller</strong> et al.<br />

Table 5 Regression coefficients (r 2 ) for significant bivariate regressions (n = 9, P < 0.05) between macroinvertebrate indices <strong>and</strong><br />

the percentage non-forest at five spatial scales. R-A slope = absolute value of rank-abundance slopes, ns = Not significant<br />

Catchment<br />

Riparian corridor<br />

Sub-corridors (m)<br />

2000<br />

1000<br />

200<br />

Density<br />

ns<br />

ns<br />

0.57<br />

ns<br />

0.54<br />

R-A slope<br />

ns<br />

0.44<br />

0.51<br />

0.61<br />

0.82<br />

Taxon richness<br />

ns<br />

ns<br />

ns<br />

0.42<br />

0.64<br />

only at the 200 m sub-corridor scale. Taxon richness,<br />

diversity, the percent five dominant taxa <strong>and</strong> EPT taxa<br />

richness were most strongly related to l<strong>and</strong> cover<br />

within 200 m sub-corridors, but were also significantly<br />

related to l<strong>and</strong>-cover patterns in other subcorridors.<br />

In addition, EPT taxa richness was related<br />

to the percentage non-forest at catchment scale<br />

<strong>and</strong> was the only macroinvertebrate' irtdex that<br />

corresponded to l<strong>and</strong>-cover patterns at all spatial .<br />

scales evaluated.<br />

Discussion<br />

The observed variability in many of the chemical<br />

constituents studied in the URRB streams can be<br />

attributed to physical features (e.g. altitude) or l<strong>and</strong>cover<br />

patterns at the catchment scale. The relationships<br />

between alkalinity, specific conductance <strong>and</strong><br />

altitude probably reflect shifts in catchment geology<br />

(Waller, 1976). In this part of the URRB, catchments at<br />

high altitude are dominated by clastic materials (both<br />

granitic <strong>and</strong> sedimentary) <strong>and</strong> have relatively low<br />

alkalinity <strong>and</strong> specific conductance when compared<br />

with low altitude streams that are influenced more by<br />

soluble parenHithology. The three catchments nearest<br />

the town of Salem, Virginia (sites 7, 8 <strong>and</strong> 9) include<br />

both clastic <strong>and</strong> carbonate materials, <strong>and</strong> subsequently<br />

had the highest alkalinity <strong>and</strong> specific conductance<br />

concentrations among sites. Variation in<br />

stream-water TIN was only related to the percentage<br />

non-forest at the catchment scale. This is consistent<br />

with many studies that have shown l<strong>and</strong>-cover at the<br />

catchment scale to be a good predictor of in-stream<br />

nutrient concentration, particularly nitrate (Omernik,<br />

1977; Close & Davies-Colley, 1990; Hunsaker &<br />

Levine, 1995; Johnson et al, 1997).<br />

In-stream physical variables were also closely<br />

related to l<strong>and</strong>-cover patterns, particularly at the<br />

H'<br />

ns<br />

ns<br />

ns<br />

0.43<br />

0.64<br />

H'/Hmax<br />

ns<br />

ns<br />

ns<br />

ns<br />

0.60<br />

%5 Dominant<br />

ns<br />

ns<br />

0.44<br />

0.50<br />

0.76<br />

EPTtaxa<br />

0.46<br />

0.52<br />

0.45<br />

0.52<br />

0.68<br />

riparian corridor <strong>and</strong> sub-corridor scales. Mean<br />

stream temperature, although related to altitude,<br />

was also influenced by l<strong>and</strong>-cover patterns within<br />

riparian corridors. Moreover, maximum temperature<br />

was strongly related to the percentage non-forest at<br />

the scale of the 200 m riparian sub-corridor. High<br />

mean <strong>and</strong> maximum stream temperature may result<br />

from solar radiation reaching the stream channel<br />

'(Burton & Likens, 1973; Beschta & Taylor, 1988;<br />

Rutherford et al., 1997), <strong>and</strong>/or from run-off heated<br />

by impervious surfaces in residential areas (Galli,<br />

1991). The local influence of l<strong>and</strong>-cover on stream<br />

thermal regimes (i.e. at the 200 m sub-corridor scale)<br />

supports the idea that temperature can change quickly<br />

(with respect to longitudinal distance) when streamside<br />

vegetation is removed in headwater catchments.<br />

Burton & Likens (1973) found summer stream-water<br />

temperature fluctuations of 4-5 °C, alternating<br />

between 50 m reaches where riparian vegetation had<br />

been experimentally removed or left intact in Hubbard<br />

Brook streams. Similarly, Storey & Cowley (1997)<br />

showed that high temperature in pasture streams of<br />

New Zeal<strong>and</strong> return to 'forested control' levels after<br />

running through approximately 300 m of remnant<br />

riparian forests.<br />

Substratum characteristics were also influenced by<br />

catchment l<strong>and</strong> use. Mean substratum particle size<br />

was inversely related to the percentage non-forest at<br />

all spatial scales, but the strongest relationships were<br />

observed at the 2000 m sub-corridor scale. The<br />

percentage of particles


particularly when these occur adjacent to the channel<br />

(Lenat et al., 1981; Waters, 1995; Richards et al., 1996).<br />

Several- studies have also reported that intact<br />

streamside vegetation inhibits the delivery of sediments<br />

to streams. Peterjohn & Correll (1984) found<br />

that 19 m of riparian forests removed up to 90% of the<br />

particulate materials moving overl<strong>and</strong> from agricultural<br />

areas. Similarly, Robinson, Ghaffarzadeh &<br />

Cruse (1996) showed that the initial 3 m of riparian<br />

forests removed more than 70% of sediment run-off to<br />

Iowa streams.<br />

Among site variability in chlorophyll a <strong>and</strong> epilithic<br />

st<strong>and</strong>ing crop was also related to in-stream physicochemical<br />

changes induced by l<strong>and</strong>-cover patterns<br />

within riparian corridors. Chlorophyll a was closely<br />

related to thermal regimes in URRB streams, probably<br />

reflecting the influence of both stream temperature<br />

<strong>and</strong> light availability on benthic algae. Several studies<br />

have demonstrated a positive relationship-between<br />

stream temperature arid algal growth (Bothwell, 1988;<br />

Duncan &-Biinn, 1989; Suzuki & Takahashi, 1995) <strong>and</strong><br />

others have attributed greater algal biomass in forested<br />

regions to local increases in light availability<br />

(Lowe, Golladay & Webster, 1986; Hill & Harvey,<br />

1990; Quinn et al, 1997). Interestingly, while many<br />

studies have linked algal abundance <strong>and</strong> nutrient<br />

availability (Fairchild, Lowe & Richardson, 1985;<br />

Grimm & Fisher, 1986; Biggs, 1995), variability in<br />

algal biomass in this study was not related to<br />

in-stream concentrations of TIN or SRP. However,<br />

neither TIN nor SRP were low enough at our study<br />

sites to be considered limiting (Grimm & Fisher, 1986;<br />

Lohman, Jones & Baysinger-Daniel, 1991). The interplay<br />

between thermal regime <strong>and</strong> nutrient availability<br />

was demonstrated by Bothwell (1988); in the absence<br />

of nutrient limitation, stream temperature explained<br />

90% of the annual variability in benthic algal growth.<br />

Changes in algal cover among sites had a strong<br />

influence on the benthos, as macroinvertebrate density<br />

was highest at sites with the highest algal biomass<br />

<strong>and</strong> biofUm st<strong>and</strong>ing stocks. This relationship is most<br />

likely driven by differences in chironomid abundance,<br />

which ranged in two orders of magnitude among<br />

sites. Many studies have indicated that invertebrate<br />

density <strong>and</strong> production may increase with greater<br />

algal biomass (e.g. Dudley, Cooper & Hemphill, 1986;<br />

Behmer & Hawkins, 1986; Richards & Minshall, 1988;<br />

Death & Winterbourn, 1995). Furthermore, studies in<br />

impacted stream ecosystems (mostly agricultural)<br />

© <strong>2001</strong> Blackwell Science Ltd, Freshwater biology, 46,1409-1424<br />

Influences of l<strong>and</strong> use on stream macroinvertebrates 1419<br />

have documented high macroinvertebrate density,<br />

often associated with numerical dominance by environmentally<br />

tolerant taxa (Lenat & Crawford, 1994;<br />

Harding & Winterbourn, 1995; Quinn et al, 1997).<br />

High density at some sites in the URRB may also be<br />

related to the presence of filamentous algae. Dudley,<br />

. Hemphill & Cooper (1986) suggested that the abundance<br />

of Cladophora sp. might influence invertebrate<br />

density directly as a food source or by accumulating<br />

other food resources (e.g. detritus, smaller epiphytes),<br />

or indirectly by increasing habitat availability. In our<br />

study, highest algal st<strong>and</strong>ing crops, <strong>and</strong> macroinvertebrate<br />

density were measured at sites 8 <strong>and</strong> 9, where<br />

the epilithon was commonly dominated by filamentous<br />

green algae.<br />

The use of rank-abundance curves allowed us<br />

statistically to assess differences in macroinvertebrate<br />

assemblage structure among sites. We identified<br />

three groups of streams that differed in the<br />

number <strong>and</strong> relative proportion of benthic taxa. Sites<br />

""with low diversity <strong>and</strong> high dominance had the<br />

steepest rank-abundance slopes (sites 7 <strong>and</strong> 8) <strong>and</strong><br />

were characterized by large numbers of only a few<br />

chironomid taxa, particularly members of the Orthocladius/Cricocladius<br />

group. The group with intermediate<br />

rank-abundance slopes (sites 4, 5 <strong>and</strong> 9) had<br />

either slightly higher overall diversity, in conjunction<br />

with high dominance by Orthocladius/'Cricocladius<br />

chironomids (site 9), or had relatively diverse<br />

chironomid assemblages (sites 4 <strong>and</strong> 5). The group<br />

with rank-abundance slopes approaching zero had<br />

the most diverse macroinvertebrate assemblages,<br />

including a greater abundance of more sensitive<br />

taxa (e.g. EPT taxa), <strong>and</strong> frequently more than 10<br />

chironomid taxa. As was the case with macroinvertebrate<br />

density, rank-abundance slopes increased<br />

among sites with algal biomass, probably reflecting<br />

the numerical dominance of Orthodadius/Cricocladius<br />

chironomids at sites with the highest algal cover<br />

(e.g. sites 7 <strong>and</strong> 8).<br />

Other macroinvertebrate indices corresponded well<br />

to the three groups that emerged from the comparison<br />

of rank-abundance slopes, suggesting these curves<br />

adequately described assemblage structure in URRB<br />

streams. In addition, results from multiple regression<br />

analyses indicated that variation in index values<br />

among sites was most closely related to the thermal<br />

regimes, with diversity, evenness <strong>and</strong> EPT taxa<br />

richness being lowest at sites with the highest


1420 R.A. <strong>Sponseller</strong> et al.<br />

maximum stream temperature. Results from these<br />

analyses should be interpreted with some caution,<br />

however (see MacNally, 2000). Although each multiple<br />

regression incorporated only maximum stream<br />

temperature into the model, it is probable that other<br />

instream features (e.g. substratum size, flow regime,<br />

etc.) interact with the thermal regime to generate the<br />

habitat template (sensu Southwood, 1977; Poff &<br />

Ward, 1990) upon which macroinvertebrate assemblages<br />

are structured.<br />

The influence of temperature on the life history <strong>and</strong><br />

ecology of aquatic insects is well documented (Vannote<br />

& Sweeney, 1980; Ward & Stanford, 1982;<br />

Sweeney & Vannote, 1986; Quinn et al., 1994). Ward<br />

& Stanford (1982) suggested that large diel fluctuations<br />

in stream temperature might lead to increased<br />

macroinvertebrate diversity by generating acceptable<br />

thermal conditions, <strong>and</strong> the potential for niche segregation<br />

among a wide range of benthic-taxa. In the<br />

current study, this relationship between thermal<br />

regime <strong>and</strong> diversity may not hold if maximum<br />

-stream temperatures become too high, especially for<br />

thermally sensitive taxa such as many stoneflies<br />

(Vannote & Sweeny, 1980; Sweeney, Vannote &<br />

Dodds, 1986; Quinn et al, 1994). Sweeney (1993)<br />

suggested that changes in temperature of 3-5 °C in<br />

Piedmont streams lacking riparian cover is sufficient<br />

to deleteriously alter larval recruitment <strong>and</strong> growth<br />

for many mayfly taxa. In addition to this, reduced<br />

macroinvertebrate diversity has been attributed to<br />

high maximum temperatures in both large regulated<br />

rivers (Fraley, 1979) <strong>and</strong> New Zeal<strong>and</strong> pasture<br />

streams (Quinn & Hickey, 1990; Quinn et al, 1997;<br />

Storey & Cowley, 1997). In the URRB, both mean <strong>and</strong><br />

maximum temperature in streams draining forested<br />

catchments (sites 1-3) were ~3-6 °C lower than<br />

streams draining catchments with current l<strong>and</strong> use<br />

activities (sites 4-9). Moreover, sites 1-3 also had the<br />

most diverse <strong>and</strong> even macroinvertebrate assemblages<br />

among streams in this study.<br />

Macroinvertebrate diversity decreased among<br />

catchments with the percentage of non-forested l<strong>and</strong>,<br />

although only when appropriate spatial scales were<br />

used in analyses. Several studies comparing macroinvertebrate<br />

assemblages among basins characterized<br />

by different l<strong>and</strong>-use practices (e.g. native forest<br />

versus pasture) or parent geology have documented<br />

predictive ability at the catchment-scale (Lenat &<br />

Crawford, 1994; Harding & Winterbourn, 1995;<br />

Richards et al, 1996; Quinn et al, 1997). However, in<br />

our study, the relationship between l<strong>and</strong>-cover <strong>and</strong><br />

macro-invertebrate assemblage structure was strongest<br />

at the 200 m sub-corridor scale. This suggests that<br />

changes induced by local, near-stream development<br />

are sufficient to alter community structure, regardless<br />

of l<strong>and</strong>-cover patterns found further upstream. For<br />

example, the catchment of site 5 was over 90%<br />

forested when classified at the catchment scale, yet<br />

the first 200 m of riparian corridor upstream of the<br />

study reach was classified as ~25% non-forested as a<br />

result of recent (~5-15 year) residential development.<br />

Assemblage structure at site 5 had low taxon richness<br />

(40 taxa), <strong>and</strong> EFT taxa richness (15 taxa), suggesting<br />

that local l<strong>and</strong>-cover patterns had a strong influence<br />

on habitat quality. Similarly, studies investigating the<br />

importance of road construction (Lenat et al, 1981)<br />

<strong>and</strong> run-off from agricultural areas (Lemly, 1982) have<br />

documented the influence of local, streamside development<br />

on macroinvertebrate assemblages, including<br />

shifts in numeric dominance to intolerant taxa like<br />

chironomids. <strong>Sponseller</strong> & <strong>Benfield</strong> (<strong>2001</strong>) found that<br />

riparian l<strong>and</strong>-cover within 1 km of the study sites was<br />

the strongest l<strong>and</strong>scape predictor for leaf breakdown<br />

rate in URRB streams. Current results suggest that<br />

community structure may be more sensitive to local<br />

l<strong>and</strong>-use disturbances than ecosystem processes that<br />

incorporate both biotic <strong>and</strong> abiotic components, which<br />

appear to be organized at broader spatial scales. This<br />

supports the idea that functional <strong>and</strong> taxonomic<br />

responses to disturbance do not necessarily correspond<br />

to one another (e.g. Wallace, Vogel & Cuffney,<br />

1986).<br />

The spatial relationship between l<strong>and</strong>-cover <strong>and</strong><br />

macroinvertebrate indices also suggests that local,<br />

upstream patches of riparian forest may have an<br />

ameliorative effect on macroinvertebrate assemblages.<br />

For example, the stream at site 6 drains a catchment<br />

subject to a mixture of l<strong>and</strong>-use practices (both<br />

agriculture <strong>and</strong> residential), yet several hundred<br />

metres of riparian corridor remain forested upstream<br />

of the sampling site. The macroinvertebrate assemblage<br />

at site 6.was fairly diverse (51 taxa), <strong>and</strong> analysis<br />

of rank-abundance curves grouped this stream with<br />

those completely lacking urban or agricultural development.<br />

Storey & Cowley (1997) compared macroinvertebrate<br />

assemblages in upstream pasture reaches<br />

with downstream sites, after streams ran through<br />

approximately 600 m of native riparian forest. They<br />

© <strong>2001</strong> Blackwell Science Ltd, freshwater Biology, 46,1409-1424


1422 R.A. <strong>Sponseller</strong> et al.<br />

Close M.E. & Davies-Colley R.J. (1990) Baseflow chemistry<br />

in New Zeal<strong>and</strong> rivers 2. Influence of environmental<br />

factors. New Zeal<strong>and</strong> Journal of Marine <strong>and</strong><br />

Freshwater Research, 24,343-356.<br />

Death R.G. & Winterbourn M.J. (1995) Diversity patterns<br />

in stream benthic invertebrate communities: the influence<br />

of habitat stability. Ecology, 76,1446-1460. .<br />

Dudley T.L., Cooper S.D. & Hemphill N. (1986) Effects of<br />

macroalgae on a stream invertebrate community.<br />

Journal of the North American Benthological Society, 5,<br />

93-106.<br />

Duncan S.W. & Blinn D.W. (1989) Importance of physical<br />

variables on the seasonal dynamics of epilithic algae in<br />

a highly shaded canyon stream. Journal ofPhycology, 25,<br />

455-461.<br />

Epler J.H. (1995) Identification Manual for the Larval<br />

Chironomidae (Diptera) of Florida. Revised edition. Florida<br />

Department of Environmental Protection, Tallahassee,<br />

FL.<br />

Fairchild G.W., Lowe R.L. & Richardson W.B. (1985)<br />

Algal periphyton growth on nutrientTdiffusing substrates:<br />

an in situ bioassay. Ecology,.€6, 465-472. ___,,<br />

Fraley J.J. (1979) Effects of elevated stream temperatures<br />

- below a shallow reservoir on a cold water macroinvertebrate<br />

fauna. In: The Ecology of Regulated Streams<br />

(Eds J.V. Ward & J.A. Stanford), pp. 257-272. Plenum<br />

Press, New York.<br />

Galli J. (1991) Thermal Impacts Associated with Urbanization<br />

<strong>and</strong> Stormwater Management, Best Management Practices.<br />

Metropolitan Washington Council of Governments. Maryl<strong>and</strong><br />

Department of Environment, Washington, DC,<br />

188 pp.<br />

Gore J.A. (1996) Discharge measurements <strong>and</strong> streamflow<br />

analysis. In: Methods in Stream Ecology (Eds<br />

F.R. Hauer & G.A. Lamberti), pp. 53-74. Academic<br />

Press, San Diego, CA.<br />

Gregory S.V., Swanson F.J., Mckee W.A. & Cummins<br />

K.W. (1991) An ecosystem perspective of riparian<br />

zones. Bioscience, 41, 540-551.<br />

Grimm N.B. & Fisher S.G. (1986) Nitrogen limitation in a<br />

Sonoran Desert stream. Journal of the North American<br />

Benthological Society, 5, 2-15.<br />

Gurtz E.G. & Wallace J.B. (1984) Substrate-mediated<br />

response of stream invertebrates to disturbance. Ecology,<br />

65,1556-1569.<br />

Harding J.S., <strong>Benfield</strong> E.F., Bolstad P.V., Helfman G.S. &<br />

Jones E.B.D. (1998) Stream biodiversity: the ghost of<br />

l<strong>and</strong> use past. Proceedings of the National Academy of<br />

Science, 95,14843-14847.<br />

Harding J.S. & Winterbourn M.J. (1995) Effects of<br />

contrasting l<strong>and</strong> use on physico-chemical conditions<br />

<strong>and</strong> benthic assemblages of streams in a Canterbury<br />

(South Isl<strong>and</strong>, New Zeal<strong>and</strong>) river system. New<br />

Zeal<strong>and</strong> Journal of Marine <strong>and</strong> Freshwater Research, 29,<br />

479-192.<br />

Hawkins C.P., Murphy M.L. & Anderson N.H. (1982)<br />

The effects of canopy, substrate composition, <strong>and</strong><br />

gradient on the structure of macroinvertebrate communities<br />

in Cascade Range streams of Oregon. Ecology,<br />

63,1840-1856.<br />

Hill W.R. & Harvey B.C. (1990) Periphyton responses to<br />

higher trophic levels <strong>and</strong> light in a shaded stream.<br />

Canadian Journal of Fisheries <strong>and</strong> Aquatic Science, 47,<br />

2307-2314.<br />

Hunsaker C.T. & Levine D.A. (1995) Hierarchical approaches<br />

to the study of water quality in rivers. Bioscience,<br />

45,193-203.<br />

Johnson L.B., Richards C, Host G. & Arthur J.W. (1997)<br />

L<strong>and</strong>scape influences on water chemistry in Midwest<br />

stream ecosystems. Freshwater Biology, 37,193-208.<br />

Lemly D. (1982) Modification of benthic insect communities<br />

in polluted streams: combined effects of sedimentation<br />

<strong>and</strong> nutrient enrichment. Hydrobiologia, 87,<br />

/229-24S.<br />

Lenat D.R. & Crawford J.K. (1994) Effects of l<strong>and</strong> use on<br />

water quality <strong>and</strong> aquatic biota of three North Carolina<br />

Piedmont streams. Hydrobiologia, 294,185-199.<br />

Lenat D.R., Penrose D.L. & Eagleson K.W. (1981) Variable<br />

effects of sediment addition on stream benthos.<br />

Hydrobiologia, 79,187-199.<br />

Lohman K., Jones R.K. & Baysinger-Daniel C. (1991)<br />

Experimental evidence for nitrogen limitation in a<br />

northern Ozark stream. Journal of the North American<br />

Benthological Society, 10,14-23.<br />

Lowe R.L., Golloday S.W. & Webster J.R. (1986) Periphyton<br />

response to nutrient manipulation in streams<br />

draining clearcut <strong>and</strong> forested watersheds. Journal of<br />

the North American Benthological Society, 5,221-229.<br />

MacNally R. (2000) Regression <strong>and</strong> model building in<br />

conservation biology, biogeography <strong>and</strong> ecology: The<br />

distinction between - <strong>and</strong> reconciliation of - 'predictive'<br />

<strong>and</strong> 'explanatory' models. Biodiversity <strong>and</strong> Conservation,<br />

9, 655-671.<br />

Marsh-Mathews E. & Mathews W.J. (2000) Spatial variation<br />

in relative abundance of a widespread, numerically<br />

dominant fish species <strong>and</strong> its effect on fish<br />

assemblage structure. Oecologia, 125,283-292.<br />

Merritt R.W. & Cummins K.W. (1996) An Introduction to<br />

the Aquatic Insects of North America, 3rd edn. Kendall/<br />

Hunt Publishing Co., Dubuque, Iowa.<br />

Morton D.D. (1998) L<strong>and</strong> Cover of Virginia from L<strong>and</strong>sat<br />

Thematic Mapper Images. Master's Thesis. Virginia<br />

Polytechnic Institute <strong>and</strong> State University, Blacksburg,<br />

VA.<br />

Murphy J. & Riley J.P. (1962) Determination of phosphate<br />

in natural waters. Analytica Chimica Acta, 27,31-36.<br />

© <strong>2001</strong> Blackwell Science Ltd, Freshwater Biology, 46,1409-1424


Omernik J.M. (1977) Nonpoint source - stream nutrient<br />

level relationships: a nationwide study. US EPA - 600/<br />

3-77-105; Corvallis, Oregon.<br />

Peterjohn W.T. & Correll D.L. (1984) Nutrient dynamics<br />

in an agricultural watershed: observations on the role<br />

of a riparian forest. Ecology, 65,1466-1475.<br />

Pielou B.C. (1969) An Introduction to Mathematical Ecology.-<br />

Wiley, New York.<br />

Poff N.L., Allan J.D., Bain M.B., Karr J.R., Prestegaard<br />

K.L., Richter B.R., Sparks R.E. & Stromberg J.C. (1997)<br />

The natural flow regime. Bioscience, 47, 769-784.<br />

Poff N.L. & Ward J.V. (1990) Physical habitat template of<br />

lotic systems: recovery in the context of historical<br />

pattern of spatiotemporal heterogeneity. Environmental<br />

Management, 14, 629-645.<br />

Quinn J.M., Cooper A.B., Davies-Colley R.J., Rutherford<br />

J.C. & Williamson R.B. (1997) L<strong>and</strong> use effects on<br />

habitat, water quality, periphyton, <strong>and</strong> benthic invertebrates<br />

in Waikato, New Zeal<strong>and</strong> hill country streams.<br />

New Zeal<strong>and</strong> Journal of Marine <strong>and</strong> Freshwater Research,<br />

31,569-577. ' /•'*"<br />

Quinn J.M. & Rickey C.W. (1990) Characterization <strong>and</strong><br />

classification of benthic invertebrate communities in 88<br />

New Zeal<strong>and</strong> rivers in relation to environmental<br />

factors. New Zeal<strong>and</strong> Journal of Marine <strong>and</strong> Freshwater<br />

Research, 24,387-409.<br />

Quinn J.M., Steele G.L., Hickey C.W. & Vickers M.L.<br />

(1994) Upper thermal tolerances of twelve common<br />

New Zeal<strong>and</strong> stream invertebrate species. New Zeal<strong>and</strong><br />

Journal of Marine <strong>and</strong> Freshwater Research, 28,391-397.<br />

Richards C, Haro R.J., Johnson L.B. & Host G.E. (1997)<br />

Catchment <strong>and</strong> reach-scale properties as indicators of<br />

macroinvertebrate species traits. Freshwater Biology, 37,<br />

219-230.<br />

Richards C., Johnson L.B. & Host G.E. (1996) L<strong>and</strong>scapescale<br />

influences on stream habitats <strong>and</strong> biota. Canadian<br />

Journal of Fisheries <strong>and</strong> Aquatic Sciences, 53, 295-311.<br />

Richards C. & Minshall G.W. (1988) The influence of<br />

periphyton abundance on Baetis bicaudatus distribution<br />

<strong>and</strong> colonization in a small stream. Journal of the North<br />

American Benthobgical Society, 7, 77-86.<br />

Robinson C.A., Ghaffarzadeh M. & Cruse R.M. (1996)<br />

Vegetative filter strip effects on sediment concentration<br />

in cropl<strong>and</strong> runoff. Journal of Soil <strong>and</strong> Water Conservation,<br />

50, 227-230.<br />

Rutherford J.C., Blackett 5., Blackett C., Saito L. & Davies-<br />

Colley R.J. (1997) Predicting the effects of shade on<br />

water temperature in small streams. New Zeal<strong>and</strong><br />

Journal of Marine <strong>and</strong> Freshwater Research, 31,707-721.<br />

SAS Institute (1998) Version 7. Gary, NC.<br />

Soloranzo L. (1969) Determination of ammonium in<br />

natural waters by the phenolhypochlorite method.<br />

Limnology <strong>and</strong> Oceanography, 14,799-801.<br />

© <strong>2001</strong> BlackweU Science Ltd, Freshwater Biology, 46,1409-1424<br />

Influences of l<strong>and</strong> use on stream macroinvertebrates 1423<br />

Southern Appalachian Man <strong>and</strong> the Biosphere (SAMAB)<br />

(1996) The Southern Appalachian assessment Summary<br />

Report: Report 1 of 5. US Department of Agriculture,<br />

Forest Service, Southern Region, Atlanta.<br />

Southwood T.R.E. (1977) Habitat, the templet for ecological<br />

strategies? Journal of Animal Ecology, 46,337-365.<br />

<strong>Sponseller</strong> R.A. & <strong>Benfield</strong> E.F. (<strong>2001</strong>) Influences on l<strong>and</strong><br />

use on leaf breakdown in southern Appalachian<br />

headwater streams: a multiple scale analysis. Journal<br />

of the North American Benthological Society, 20,44-59.<br />

Steinman A.D. & Lamberti G.A. (1996) Biomass <strong>and</strong><br />

pigments of benthic algae. In: Methods in Stream Ecology<br />

(Eds F.R. Hauer & G.A. Lamberti), pp. 295-314.<br />

Academic Press, San Diego, CA.<br />

Stewart K.W. & Stark B.P. (1993) Nymphs of North<br />

American Stonefly Genera (Plecoptera). University of<br />

North Texas Press, Denton, Texas.<br />

Storey R.G. & Cowley D.R. (1997) Recovery of three New<br />

Zeal<strong>and</strong> rural streams as they pass through native<br />

forest remnants. Hydrobiologia, 353, 63-76.<br />

Suzuki Y. & Takahashi M. (1995) Growth resonses of<br />

, • several diatoms species from various environments to<br />

temperature. Journal ofPhycobgy, 31, 880-888.<br />

Swank W.T., Swift L.W. & Douglass J.E. (1988) Streamflow<br />

changes associated with forest cutting, species<br />

conversions, <strong>and</strong> natural disturbances. In: Ecological<br />

Studies 66. Forest Hydrology <strong>and</strong> Ecology at Coweeta (Eds<br />

W.T. Swank & D.A. Crossely), pp. 297-312. Springer-<br />

Verlag, New York, NY.<br />

Sweeney B.W. (1993) Effects of streamside vegetation on<br />

macroinvertebrate communities of White Clay Creek<br />

in eastern North America. Proceedings of the Academy of<br />

Natural Sciences of Philadelphia, 144,291-340.<br />

Sweeney B.W. & Vannote R.L. (1986) Growth <strong>and</strong><br />

production of a stream stonefly: influences of diet<br />

<strong>and</strong> temperature. Ecology, 67,1396-1410.<br />

Sweeney B.W., Vannote R.L. & Dodds P.J. (1986) The<br />

relative importance of temperature <strong>and</strong> diet to larval<br />

development <strong>and</strong> adult size of the winter stonefly,<br />

Soyedma carolinensis (Plecoptera: Nemouridae). Freshwater<br />

Biology, 16, 39-48.<br />

Vannote R.L. & Sweeney B.W. (1980) Geographic analysis<br />

of thermal equilibria: a conceptual model for<br />

evaluating the effect of natural <strong>and</strong> modified thermal<br />

regimes on aquatic insect communities. American<br />

Naturalist, 115, 667-695.<br />

Wallace J.B., Grubaugh J.W. & Whiles M.R. (1996) Biotic<br />

indices <strong>and</strong> stream ecosystem processes: results from an<br />

experimental study. Ecological Applications, 6,140-151.<br />

Wallace J.B., Vogel D.S. & Cuffney T.F. (1986) Recovery<br />

of a headwater stream from an insecticide-induced<br />

community disturbance. Journal of the North American<br />

Benthological Society, 5,115-126.


1424 R.A. <strong>Sponseller</strong> et al.<br />

Waller J.O. (1976) Geohydrology of the Upper Roanoke River<br />

Basin, Virginia. Virginia State Water Control Board -<br />

Planning Bulletin 302. Bureau of Water Control Management,<br />

Richmond, Virginia, 148 p.<br />

Ward J.V. & Stanford J.A. (1982) Thermal response in the<br />

evolutionary ecology of aquatic insects. AnnwlReview<br />

of Entomology, 27,97-117.<br />

Waters T.F. (1995) Sediment in Streams: Sources, Biological<br />

Effects <strong>and</strong> Control. American Fisheries Society Monograph<br />

7. American Fisheries Society, Bethesda, Maryl<strong>and</strong>.<br />

Wear D.N. & Bolstad P. (1998) L<strong>and</strong>-use changes in<br />

southern appalachian l<strong>and</strong>scapes: spatial analysis <strong>and</strong><br />

forecast evaluation. Ecosystems, \, 575-594.<br />

Webster J.R., Gurtz M.E., Hains J.J., Meyer J.L., Swank<br />

W.T., Waide J.B. & Wallace J.B. (1983) Stability of<br />

stream ecosystems. In: Stream Ecology (Eds J.B. Barnes<br />

& G.W. Minshall), pp. 355-394. Plenum Press, New<br />

York.<br />

Wetzel R.G. & Likens G.E. (1991) Limnokgical Analyses,<br />

2nd edn. Springer-Verlag, New York, NY:<br />

Wiggins G.B. (1996) larvae of the North American Caddisfly<br />

Genera (Trichoptera). University of Toronto Press Inc,<br />

Toronto, Canada.<br />

Wolman M.G. (1954) A new method of sampling coarse<br />

riverbed. Transactions of the American Geophysical Union,<br />

35, 951-956.<br />

Wood E.D., Armstrong F.A.J. & Richards F.A. (1967)<br />

Determination of nitrate in seawater by cadmiumcopper<br />

reduction to nitrate. Journal of the Marine Biology<br />

Association of the United Kingdom, 47, 23-31.<br />

Yarnell S.L. (1998) The Southern Appalachians: a History of<br />

the L<strong>and</strong>scape. General Technical Report SRS-18. U.S.<br />

Department of Agriculture, Forest Service, Southern<br />

Research Station, Asheville, NC, 45 pp.<br />

Zar J.H. (1996) Biostatistical Analysis, 3rd edn. Prentice<br />

Hall Inc., Upper Saddle River, NJ.<br />

(Manuscript accepted 5 January <strong>2001</strong>)<br />

© <strong>2001</strong> Blackwell Science Ltd, Freshwater Biology, 46,1409-1424

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!