Sei sulla pagina 1di 12

Science of the Total Environment 646 (2019) 770–781

Contents lists available at ScienceDirect

Science of the Total Environment

journal homepage: www.elsevier.com/locate/scitotenv

Irrigation canals are newly created streams of semi-arid agricultural regions


Erick A. Carlson a,b,⁎, David J. Cooper a,b, David M. Merritt a,c, Boris C. Kondratieff a,b,d, Reagan M. Waskom e
a
Graduate Degree Program in Ecology, Colorado State University, Fort Collins, CO 80523, United States of America
b
Department of Forest and Rangeland Stewardship, Colorado State University, Fort Collins, CO 80523, United States of America
c
USFS National Watershed, Fish and Wildlife Staff, Fort Collins, CO 80526, United States of America
d
Bioagricultural Sciences and Pest Management, Colorado State University, Fort Collins, CO 80523, United States of America
e
Colorado Water Institute, Colorado State University, Fort Collins, CO 50523, United States of America

H I G H L I G H T S G R A P H I C A L A B S T R A C T

• Irrigation canals are more prevalent in


some semi-arid agricultural landscapes.
• Riparian vegetation of irrigation canals
showed similarities to local streams.
• Aquatic macroinvertebrates of irrigation
canals were similar to local streams.
• Land-use affected similarities between
canals and streams.

a r t i c l e i n f o a b s t r a c t

Article history: The natural hydrologic processes that create and maintain the diversity of aquatic and riparian habitats along the
Received 30 March 2018 World's streams and rivers have been profoundly altered by humans. Diversion of surface water to support produc-
Received in revised form 16 July 2018 tion agriculture in arid and semi-arid regions has degraded ecosystems but also created potential habitat along and
Accepted 17 July 2018
in canals specifically designed to transport water. The prevalence of canals and the immense amount of water used
Available online 29 July 2018
for agriculture have created these new artificial stream systems. This study demonstrates the potential for irrigation
Editor: D. Barcelo canals to support riparian and aquatic communities similar to natural streams in urban/residential and agricultural
landscapes. We examined the hydrological and ecological characteristics of streams and irrigation canals in urban
Keywords: and agricultural landscapes in northeastern Colorado, typical of regions dominated by irrigation-supported agricul-
Riparian vegetation ture. Flow patterns in canals depended on their size and had a range of patterns with potential ecological conse-
Aquatic macroinvertebrates quences such as rapidly rising and falling water stage, intermittent dry periods, and delayed peak and base flows
Irrigation canal compared to natural streams. Despite these hydrologic differences, the taxonomic and functional composition of
Community ecology riparian plant and aquatic macroinvertebrate communities indicated that ecological similarities exist between
streams and canals, but are dependent, in part, on their landscape setting with stronger similarities in agricultural
areas. We also tested the influence of characterizing taxa by functional groups using physiology, ecology and life
history traits to explore attributes of habitats including woody canopy structure and water quality. We used a
Habitat Quality Index (HQI) that combined physical and biological measures into a single index. Streams scored
higher on average within agriculture and urban/residential settings compared to canals; however, one third of
urban canals scored above the average of agricultural streams. This multidisciplinary study shows that irrigation
canals can be valuable riparian and aquatic habitat, especially in regions with severely degraded streams.
© 2018 Published by Elsevier B.V.

⁎ Corresponding author at: Department of Biology and Environmental Sciences, Marietta College, Marietta, OH 45750, United States of America.
E-mail address: eac005@marietta.edu (E.A. Carlson).

https://doi.org/10.1016/j.scitotenv.2018.07.246
0048-9697/© 2018 Published by Elsevier B.V.
E.A. Carlson et al. / Science of the Total Environment 646 (2019) 770–781 771

1. Introduction exchange (Fig. 1). However, because streams are the source of water
for canals, they create connectivity allowing riparian plant and aquatic
Streamflow reductions for hydropower generation and municipal macroinvertebrate propagule introductions (Ernegger et al., 1998;
and agricultural use affect rivers of all sizes across the Earth Koetsier and McCauley, 2015). Canals have long been recognized as sus-
(Graf, 1999; Kingsford, 2000). Production agriculture in arid and ceptible to the invasion of non-native species through hydrochory
semi-arid regions requires water to be diverted from streams or (Egginton and Robbins, 1920), as well as facilitating the introduction
pumped from the ground to irrigate crops (Khan et al., 2006). The ef- of non-native, weedy species to natural streams thorough agricultural
fects of dewatering streams on aquatic and riparian biota have been return flows.
relatively well studied (Bunn and Arthington, 2002; Malmqvist and Aquatic macroinvertebrates are a key component of aquatic, riparian
Rundle, 2002; Poff and Zimmerman, 2010), however in many regions and some upland food webs (Polis et al., 1997; Nakano et al., 1999; Leigh
canals specifically designed to transport water to agricultural land sup- et al., 2013) and are highly sensitive to flow variation (Nakano and
port little known aquatic and riparian ecosystems (Patten, 1998). Some Murakami, 2001; Miller et al., 2007; Muehlbauer et al., 2014). Their pop-
researchers and managers have suggested that some canals may sup- ulations are influenced by channel physical characteristics, food type
port species-rich plant and aquatic invertebrate communities like and availability, and hydrologic patterns at relatively small scales
those of natural streams (Chester and Robson, 2013), but the character- (Williams and Feltmate, 1992). Aquatic macroinvertebrates often use
istics of these communities is not well known. seasonal and hydrologic cues for their life history development
Riparian ecosystems are highly productive and biologically diverse (Butler, 1984). Flow variability in canals can include rapid drawdowns
relative to surrounding uplands in most regions (Naiman et al., 1993). and multiple high and low flow events per year creating unnatural
Floodplains and riparian zones function to store, transform and cycle timing that can affect aquatic macroinvertebrate growth and reproduc-
nutrients, organic matter, sediment and water (Jacobs et al., 2007; tion (Kennedy et al., 2016). Drainage ditches in agricultural areas are
Tank et al., 2010). Hydrologic and geomorphic processes shape the populated by biota with life history adaptations to highly variable flow
physical landscape, and the disturbance regime controls the coloniza- regimes including short life cycles, desiccation resistant life stages and
tion and persistence of organisms (Shafroth et al., 2002; Katz et al., winged adults (Whatley et al., 2015). Seasonal and random disturbance
2009) and laterally connect terrestrial and aquatic ecosystems at local events combined with the physical setting can influence colonization
scales and longitudinally connect river networks at landscape scales and competition and the formation of communities (Petraitis et al.,
(Harvey and Gooseff, 2015). 1989; Hobbs and Huenneke, 1992). The presence of aquatic plants in ar-
Streams and riparian areas throughout the world have been directly tificial temporary channels and wetlands further can support regional
modified by floodplain development and channelization and indirectly diversity local food webs in agricultural watersheds (Bolpagni and
through watershed modification, stream flow alteration, and land use Piotti, 2016).
changes (Patten, 1998; Nilsson and Svedmark, 2002; Bunn and In natural watersheds, streams are organized from small headwa-
Arthington, 2002). Human activities have altered the physical structure ter channels to larger main stem rivers, while canal systems are the
of rivers and their floodplains through dam and levee construction, opposite, with the largest canals fed from rivers, and branching in
canal diversion structures, channel straightening and bank stabilization. dendritic patterns downgradient to reach the complex network of
Impoundments have changed the historic sediment regime and re- smaller canals that deliver water to fields. Canals add extensive com-
duced flood intensity, frequency and duration, which have altered geo- plexity and channel length to stream networks and significantly
morphic processes and the structure, composition and functioning of change watershed scale riparian habitat from its pre-settlement con-
riparian plant communities (Merritt and Cooper, 2000). For instance, dition. Although canals have been critical to the foundation of human
Populus spp. dominated riparian forests rely on natural flood distur- societies in many regions of the world, research on their biological
bance for health and regeneration with reduced growth, dieback and characteristics has been limited to a few studies in Chile, Spain,
mortality under altered flow regimes or water table depths (Cooper France, and the United States (Habit et al., 1998; Fernald et al., 2007;
et al., 1999; Williams and Cooper, 2005). Decades of diversion from nat- Lopez-Pomares et al., 2015; Aspe and Jacque, 2015). The prevalence of
ural rivers has decreased aquatic insect and fish populations (Miller canals can be easily overlooked, as their vegetation may resemble natu-
et al., 2007; Kingsford, 2000), homogenized streamflow, habitats and bi- ral streams (Fig. 2).
otic communities (Bunn and Arthington, 2002; Leigh and Sheldon, Colorado has a 150-year history of irrigation that has facilitated agri-
2008), and in extreme cases created rivers that resemble artificial chan- cultural production and provides support for human population growth
nels with engineered bed and banks (Urban and Rhodes, 2003; Gregory, (Evans and Evans, 1991). Stream water is diverted into canals and ap-
2006). The natural hydrologic processes that create and maintain a di- plied directly to crops or stored in off-channel reservoirs to extend the
versity of aquatic and riparian habitats along the World's streams and agricultural irrigation season and supply year-round water to cities.
rivers have been profoundly reduced by human actions to guard against Water diversions from the Cache la Poudre River (a tributary to the
floods, develop floodplains for agriculture and divert water for irrigation South Platte River in the Missouri River basin) have occurred since
(Malmqvist and Rundle, 2002; Miller et al., 2007; Poff and Zimmerman, 1864 (Strange et al., 1999) with 85% of the mean annual flow being re-
2010). moved from the river for agricultural use. Thousands of kilometers of ca-
Human activities completely shape the physical structure and flow nals have been constructed to deliver water and most remain in use. The
characteristics of canals that influence their riparian and aquatic ecosys- ongoing conversion of agricultural land to residential and industrial
tems. The metrics commonly used to characterize stream geomorphic uses continues to alter the timing and amount of water flowing in canals
processes including overbank flooding, sinuosity, and channel migra- and the management of vegetation along stream banks. Maintenance
tion, are rarely applicable for highly engineered water conveyance ca- along canals varies from no activities to burning, herbicide spraying
nals that are designed to minimize turbulent flow and maximize and mechanical removal of sediment and vegetation. The range of
conveyance. In addition, the lack of hydrologic variability and spatial canal sizes and flows, along with varied maintenance, supports a wide
heterogeneity limits fluvial landform creation (Swamee, 1995). Canal variety of riparian vegetation and channel characteristics through the
management, including removal of sediment and woody debris, weed labyrinth of irrigation canals.
suppression and controlled burns, may simplify and homogenize ripar- The contribution of canal riparian vegetation and aquatic inverte-
ian and aquatic ecosystems (Andersen and Nelson, 1997; Meany et al., brates to local and regional biodiversity is poorly documented and
2003; Casas et al., 2011). understood (Paul and Meyer, 2001; Herzon and Helenius, 2008;
Canals are largely decoupled from the surrounding landscape by Vermonden et al., 2009; Verdonschot et al., 2011). In this paper, we ad-
constructed berms that limit water, sediment and organic matter dress the following questions: (1) do canals support riparian vegetation
772 E.A. Carlson et al. / Science of the Total Environment 646 (2019) 770–781

Fig. 1. Cross section of typical stream and canal with two geomorphic surfaces (bank and top/floodplain).

and aquatic macroinvertebrate communities similar to natural streams 2. Methods


of similar size, (2) are functional groups of plants and aquatic macroin-
vertebrates similar between canals and streams, (3) is there agreement 2.1. Study area and data collection
between species and functional group data, and (4) do riparian and
aquatic indices support species and functional group patterns between The study was conducted in north-central Colorado (Fig. 3) a semi-
canals and streams? Canals could support riparian and aquatic ecosys- arid region with extensive irrigated agriculture. Precipitation averages
tems comparable to natural streams as plants and animals have colo- 250 mm during the summer and 135 mm during winter (www.
nized this new water distribution system. usclimatedata.com). Total annual precipitation is insufficient to support

Fig. 2. Example streams and canals in agricultural and residential landscapes: A) agricultural stream, Willow Creek, Weld County, Colorado B) agricultural canal, Fort Collins, Larimer
County, Colorado. C) residential/urban canal, Fort Collins, Larimer County, Colorado. D) agricultural stream, Lone Tree Creek, Weld County, Colorado.
E.A. Carlson et al. / Science of the Total Environment 646 (2019) 770–781 773

desirable crops such as corn, beans, and vegetables. Water for irrigation remaining sites were classified as herbaceous unless the bank and chan-
is diverted from rivers fed by melting snow in the Rocky Mountains and nel bottom were concrete-lined.
applied to crops from April through September. Each canal site was selected from a randomly generated set of points
The study area includes irrigation ditch companies in the South distributed proportionally across the canal network using the total
Platte River Basin Water District 3 that divert water from the Cache la length of canals in each canopy cover class to determine the frequency
Poudre River primarily through earthen canals. The canal network in- of sampling. Stream sites were selected to minimize hydrologic modifi-
cludes 2014 km of canals, providing irrigation water to 54,529 ha of cations within the reach and maximize undeveloped floodplain width
crops within the study area (CDSS). Canals ranged in size from b1 to to represent the least impacted natural stream reaches and are consid-
18 m wide and up to 3 m deep. Vegetation management along canals ered reference sites. Vegetation was sampled at 47 canals and seven
varies from frequent and intense where crops were being grown, to streams during the summer of 2013. Each site had three transect lengths
minimal in urban and residential areas. We separated sites as agricul- oriented perpendicular to the channel evenly spaced five bank full chan-
tural (Ag.) or residential (Res.) using the dominant land use within nel widths apart. Aquatic macroinvertebrates were sampled every three
100 m of each canal segment. The western region (Larimer County) in- weeks from May through August in 2015 for a maximum of six samples
cluded more residential area while the eastern region (Weld County) per site for 20 sites representing all substrate types, channel sizes, and
was primarily agricultural. Flow of the Cache la Poudre River is altered flow regimes.
by numerous low-head dams that divert water into irrigation canals,
and by the addition of water from trans-basin diversions (Evans and 2.3. Environmental and hydrologic variables
Evans, 1991). The floodplain has also been constrained by gravel min-
ing, roads and agriculture (Strange et al., 1999; Wohl, 2001). Channel and floodplain physical characteristics, land use, network,
and hydrologic variables were selected and modified from existing ri-
2.2. Site selection and layout parian and stream assessment methods (Gonzalez del Tanago and
Garcia de Jalon, 2011) (details in Appendix A). Stream flow data were
To quantify vegetation cover and structure along canals, we strati- obtained from two USGS gages on the Cache la Poudre River for the pe-
fied the vegetation into five categories: heavy tree cover, light tree riod 1999–2015 (Fort Collins #06752260, and Greeley #06752500).
cover, shrub cover, herbaceous cover, and concrete-lined, using sub- Streamflow records for Spring Creek were only available for
meter resolution aerial and satellite imagery in ArcGIS v.10. Heavy 2013–2015 from the City of Fort Collins flood warning system. Owl
tree was defined as N50% cover of woody plants over 3 m tall visually es- Creek and Willow Creek with additional stream sites lack stream flow
timated. Light tree had 25–50% tree cover. Shrub sites had woody vege- records. Flow in irrigation canals was recorded at the point of diversion
tation b3 m in height covering N50% of the bank vegetation. All from the Cache la Poudre River. Streams and canals were separated by

Fig. 3. Study area in north-central Colorado.


774 E.A. Carlson et al. / Science of the Total Environment 646 (2019) 770–781

bank full width into large (N4 m) and small (b4 m) for some hydrologic flowing too swiftly for safe access. To account for this difference, abun-
comparisons. Flow depth was measured and compared to bankfull dance counts of taxa were averaged for each site.
depth as a ratio (observed/bankfull). We calculated the traditional EPT index (Ephemeroptera [mayflies],
Plecoptera [stoneflies], and Trichoptera [caddisflies] for taxa richness
2.4. Riparian vegetation (EPTr) as well as the proportion of EPT taxa using abundance (EPTa)
(Kerans and Karr, 1994; Wallace et al., 1996; Rosenberg et al., 2008).
Canopy cover of each species was visually estimated for each plot. The EPTr index was calculated for each site as the number of EPT taxa di-
Five plots were sampled on the top/floodplain and bank surfaces. vided by the total number of taxa present, the EPTa index replaced rich-
These were oriented perpendicular to a transect that spanned both ness with abundance in the proportional calculation. Physiological and
sides of the canal or stream. This sampling scheme of 20 plots per tran- ecological traits of genera and families were used to create the func-
sect was repeated 3 times at each site for a total of 60 plots per site. tional aquatic dataset (Poff et al., 2006). Counts of individuals from
Cover was estimated using modified Braun-Blanquet cover classes taxa within a functional group were added together to produce an
(trace, b1%, 1–2%, 2–5%, 5–10%, 10–25%, 25–50%, 50–75%, 75–95%, abundance for the functional group.
N95%) (Bonham, 2013) and averaged for each site to capture site level
vegetation characteristics. Plant species nomenclature follows Weber 2.6. Statistical analyses
and Wittman (2012). Height of species was placed into classes (b1 m,
1–2 m, 2–5 m, 5–10 m, N10 m) (Merritt and Bateman, 2012). Bare Community analyses for vegetation and aquatic macroinvertebrates
sediment was analyzed using the same cover classes as plant species. were performed using Primer v.7 software (Clarke et al., 2014). Non-
Diversity metrics including total richness, native species richness, metric Multi-Dimensional Scaling (nMDS) was used to analyze the
Shannon-Wiener (log e) and Simpson's index (1-λ) were calculated overall structure of communities, and principal components analysis
using all species. Species present in fewer than 5% of sites were removed for physical characterization of riparian and aquatic habitats. For all sta-
(McCune and Grace, 2002) reducing the number of plant species used in tistical tests an alpha b0.05 was used to indicate a significant result.
nMDS and cluster analysis from 251 to 126. Diversity metrics, stream biotic and habitat indices were tested for dif-
We developed a habitat quality index (HQI) that adds quantitative ferences between all groups using ANOVA and between channels con-
metrics for site level hydrologic, geomorphic and vegetation character- trolling for land use, for example canal vs. stream, using Student's t-test.
istics (detailed in the Appendix A) to the riparian quality index (RQI) of
Gonzalez del Tanago and Garcia de Jalon (2011). Functional redundancy
2.6.1. Environmental variables
(FR) (Bruno et al., 2016) was used to identify diversity in functional
Each environmental variable was normalized by subtracting the
groups and is used as a component of an ecosystem's resilience to veg-
mean and dividing by the standard deviation (Clarke et al., 2014).
etation disturbance. FR was calculated as species richness divided by
Euclidean distance was used to calculate a similarity matrix for environ-
functional group richness using the full plant species list.
mental variables. Environmental variables were related to both species
Plant species were classified into 22 functional groups (Table B.2.). A
and functional group datasets for vegetation and aquatic macroinverte-
priori groups were used because little is known about plant species re-
brates using distance-based linear modeling with the AICc selection
sponses to fluvial disturbance including inundation, burial and physical
criteria and by Spearman Rank correlation of environmental and biotic
damage for many of the observed species. Groups were created with
similarity matrices (Anderson et al., 2008).
three ecological characteristics: origin (native or introduced), growth
form (graminoid, forb, shrub, tree (USDA Plants), and National Wetland
Inventory wetland indicator status (Lichvar, 2012). Cacti and vines were 2.6.2. Ecological variables
a small proportion of total species and included in the forb category. Abundance data for plants and macroinvertebrates were square-
Three wetland indicator groups were used; wetland (obligate and facul- root transformed to down weight the few abundant taxa and included
tative wetland), facultative (facultative, facultative upland) and upland. the predictive value of infrequent species, and a Bray-Curtis similarity
Similar attributes have been suggested for use in riparian plant guild matrix was created (Clarke et al., 2014). An nMDS was calculated to vi-
classification (Merritt et al., 2010). Species were replaced by functional sualize the data and multivariate dispersion was tested using the
groups and cover summed to create a guild dataset. The guilds were PermDISP routine in Primer with distances measured to the centroid
treated as pseudo species for statistical analysis. of the group. Permutational multivariate analysis of variance
(PerMANOVA) was used to test the effects of land use, channel type,
2.5. Aquatic macroinvertebrates and their interaction on vegetation and aquatic macroinvertebrate com-
position. Due to the unbalanced sampling design, Type III sum of
Aquatic macroinvertebrates were collected at each study site every squares and unrestricted permutation of the raw data were used in
three weeks from May–August 2015 using a standard D-frame kick the calculation with 999 permutations. Comparisons that resulted in
(Merritt et al., 2008) net with a 500-μm capture net. One kicknet sample fewer than 100 permutations were reported with Monte Carlo
was collected from each site on each sampling date by disturbing the corrected p-values (Anderson et al., 2008). These statistical analyses
top 5 cm of sediment and submerged vegetation for three minutes. A were performed on the species/taxonomic and functional datasets for
composite kicknet of micro-habitats including riffles, pools, banks, sub- vegetation and macroinvertebrates.
merged and aquatic vegetation were collected as one sample. A sweep-
ing motion under the water using the D-frame net was used to simulate 3. Results
flow in stagnant pools. Presence of submerged vegetation and aquatic
macrophytes was recorded. Forty-seven canal sites and seven stream sites were included in this
All aquatic macroinvertebrates were removed from samples and study with a range of physical attributes (see Appendix B). Land use dif-
preserved in 80% ethanol. Volume based subsampling was used in five fered markedly between two regions in the study area. The western re-
samples few cases where more than several hundred individuals oc- gion is 33.1% agricultural and 29% residential/urban. The eastern region
curred in a sample (Hickley, 1975). Individuals in most groups were is 70% agricultural and only 1.6% residential/urban. The length of open
identified to genus, worms and leaches to family. A total of 97 samples irrigation canals in the more densely populated western region was
were collected, five of which contained zero individuals. The number 370 km compared to 250 km of streams, for a canal/stream length
of samples collected at each site varied between three and seven be- ratio of 1.5 even as portions of canals are directed into pipes under-
cause some canals were dry during collection dates and others were ground in urban areas. Open channels were more common in eastern
E.A. Carlson et al. / Science of the Total Environment 646 (2019) 770–781 775

agricultural areas with 1644 km of canals to 487 km of streams for a agricultural landscape, however, streams were not significantly differ-
canal/stream length ratio of 3.4. ent between land uses and in the residential areas, streams and canals
were on border of significance (t = 1.344, p = 0.05). Stream and
3.1. Hydrology canal vegetation did not separate into distinct groups in non-metric
multidimensional scaling (Fig. 5). Canopy type influenced some mixing
Flow varied by two orders of magnitude across channel types. Re- of groups, for example concrete lined canals, a recently reconstructed
cords from stream gages indicate that precipitation driven peak flows canal and a naturally intermittent stream were grouped in the lower
were most prevalent on small streams while snowmelt runoff was left of the ordination space with significant bare sediment in Panel A.
most prevalent in the large stream (Table 1). Small channels had two The same intermittent stream had a high proportion of upland grasses
patterns: 1) a short early summer peak and low late summer flows, drawing the point to the upper left of Panel B. A trend of sites with
and 2) a consistent flow level through the entire summer (Fig. 4). Ob- woody species (upper half) to those dominated by grasses and forbs
served flow depths indicated that canals in residential areas were (lower half) creates additional mixing between groups. The percent
flowing close to bankfull conditions (higher Ratio) in June and decreas- cover of woody vegetation (trees and shrubs) was highest for residen-
ing over the summer, while agricultural canals had the opposite trend, tial sites while grasses were more common in the agricultural areas
with the most flow occurring in August as crop irrigation demand peaks. (Fig. 6). The stream sites on the far right have the most diverse assembly
Flow variability and short-term changes assessed as daily coefficient of functional groups including grasses, forbs, shrubs and trees not
of variation (CV) (Table 1) indicated that most streams had greater var- matched by any canal sites.
iability than canals. The exception was Dry Creek whose watershed has Vegetation composition of canals and streams were different across
several storm water retention structures. The date of peak flow was both land-use categories (F = 2.018, p = 0.017, df = 53) with compa-
22 days later for the large canal and 51 days earlier for small canals rable variance (F b 0.01, p = 0.991, df = 53). Land-use influenced
than similar sized streams. The large canal and large stream had a sim- species composition with a significant difference (F = 4.248, p =
ilar number of low flow events and days, while small canals had more 0.001, df = 53) between riparian vegetation in agricultural and residen-
than small streams. The average number of days with flow from April tial areas. Pairwise PerMANOVA using the two by two matrix of channel
1–September 30 (183 days) was 174 (SE = 3.1) for the large canal, type and land-use indicated that 3 of 4 comparisons had significantly
47.8 (6.1) for small canals and 183 (0) for streams, all of which were different species composition (Table 3a) with streams similar in urban
perennial. and agricultural areas.

3.2. Environmental characteristics 3.4. Community composition: aquatic macroinvertebrates

Canals varied from 1 to 22 m wide and 0.4 to 2.5 m deep at bank full Aquatic macroinvertebrates were relatively diverse across the study
flow. Streams had a similar range, 1.5 to 30 m wide and 0.6 to 2.5 m area with 147 typical regional taxa identified (Table B.3). Thirty-one
deep at bank full flow. Mean water temperature in canals in agricultural taxa were found only in streams including seven EPT taxa that suggest
areas was 5–6 °C warmer than residential streams and canals, and agri- good water quality such as Seratella micheneri (Traver), and Pteronarcella
cultural streams were also 5–6 °C warmer than residential streams. badia (Hagen). Thirty-one taxa were collected only from canals and in-
Temp in canals was more positively correlated with month (r = 0.47, cluded mayfly in the genera Drunella, Ameletus and Rhithrogena. The
p b 0.001) than distance to the point of diversion (r = 0.11, p = 0.28) most diverse orders found were Diptera with 64 taxa represented, Cole-
and was weakly negatively correlated to percent woody canopy optera with 20, and Ephemeroptera with 18. Mean site level richness
(r = −0.15, p = 0.14). was highest for residential streams (37 ± 11.3), followed by agricultural
streams (28 ± 4.2), residential canals (20.1 ± 11.3), and agricultural ca-
3.3. Community composition: plants nals (18.9 ± 3.8). Many taxa were infrequently collected, and 25% were
collected only once. The variance was similar between sites measured as
A total of 3247 plots were analyzed for vegetation composition and multivariate dispersion, (F = 1.59 p = 0.833, df1 = 3, df2 = 16). There
cover at 54 sites (see Appendix B for complete list). Forbs comprised was overlap between site groups in Fig. 7 for both functional and species
50% of species present, graminoids 32%, shrubs and trees 14%, and datasets with PerMANOVA pairwise tests confirming similarity between
vines and cacti 4%. Native species provided 51% of the total richness comparisons in Table 3b.
but contributed only 36% of total plant cover. Two exotic species,
Bromopsis inermis (Leysser) and Phalaris arundinacea (L.), and the native 3.5. Functional community composition: plants
Carex emoryi (Dewey) dominated canal and stream riparian vegetation
in both the urban and agricultural regions, accounting for 49% of total A permutational distance-based test for homogeneity of multivari-
cover. Plant diversity metrics and vegetation indices were not signifi- ate dispersion (PERMDisp) indicated that functional community
cantly different between canals and streams (Table 2). Composition, in- composition of riparian vegetation was similar between channel types
cluding cover had mixed results with significant differences between (F = 0.09, p = 0.832) and land-use categories (F = 0.007, p = 0.944).
canals in both land uses and between streams and canals in the The number of functional plant groups observed at canal and stream

Table 1
Hydrologic metrics used to characterize flow regimes of canals and streams calculated from April 1 to Sept 30 including coefficient of variation (CV), date of peak flow (in Julian days),
number of high flow events (flows exceeding 90%), number of low flow events (flows b10%), number of days below 10%, and number of days with zero flow. Detailed definitions are
in the Appendix A. 1 located in Larimer County, Colorado; 2 located in Weld County, Colorado.

Cache la Poudre River Cache la Poudre River Spring Larimer and Weld New Mercer Larimer #2
(USGS gage 6752260)1 (USGS gage 6752500)2 Creek1 Canal1 Canal1 Canal1

CV 1.77 1.6 0.89 1 0.81 0.79


Date of peak 150 155 186 174 135 128
# High flow events 1.6 1.3 5 3 1.6 2.3
# Low flow events 4.3 3.3 2 3.7 3 3.3
# Low flow days 18 18 17.6 30 13.3 34
Zero flow days 0 0 0 12.3 2.6 10.3
776 E.A. Carlson et al. / Science of the Total Environment 646 (2019) 770–781

Fig. 4. Mean daily streamflow (cms) averaged for three irrigation seasons (April-Sept) 2013–2015 in one small stream (Spring Creek), and three small canals. All canal data are from gages
at the point of diversion from the Cache la Poudre River, Larimer County, Colorado.

sites averaged 9.8 and was statistically similar between channel 3.7. Comparison of taxonomic and function data sets
types (t = 1.68, p = 0.361). The number of functional plant groups
at each site did not change after removing rare species. Introduced Comparisons of species and functional group composition for ripar-
upland forb was the most common functional group and occurred ian vegetation similarity showed agreement when holding channel type
at 98% of sites, and introduced wetland grasses comprised 24% of constant (Table 3a). The two datasets disagreed when holding land-use
all vegetation cover, the most of any functional plant group. Forbs constant and comparing channel types. Aquatic macroinvertebrate tax-
were a relatively small proportion of total cover (18.5%), although onomic and functional group data agreed for all comparisons (Table 3b).
they accounted for the highest diversity of any group with 142 spe- The similarities in vegetation and macroinvertebrates (Table 3c) indi-
cies, or 60% of all species in the study area. cated more agreement between the two habitats when using functional
groups.

3.6. Functional community composition: aquatic macroinvertebrates


3.8. Environmental conditions and landscape and biotic indices
Sixty-eight aquatic macroinvertebrate taxa (Table B.4), accounting
for 71% of collected individuals, were classified by functional traits The average Habitat Quality Index (HQI) score for canals was 2.9
into six groups by kR means cluster analysis (R = 0.94), results in while for streams it was 4.4, a significant difference (t = −2.93, p =
Table 4. The number of functional macroinvertebrates groups was sig- 0.01, df = 7). Channels in the agricultural landscape had a lower aver-
nificantly lower (t = −2.95, p = 0.006) for canals with an average of age score of 2.7 compared to residential channels, 3.5 on average, also
4.1 of a possible six groups compared to streams with 5.25 observed. significantly different (t = −2.22, p = 0.02, df = 42). Residential canals
Group A was commonly absent from canals collected at only 25% of and agricultural streams were similar (t = 1.00, p = 0.19, df = 4). Other
sites. All sites had some members of Group D, but this was the only func- environmental variables including water temperature and estimated
tional group represent at all canals and stream sites. The abundance of flow were poor predictors of species and functional group composition
each functional group indicated a similarly between canals and streams for riparian vegetation.
(F = 1.54, p = 0.2, df = 19). Multivariate dispersion was similar across The EPTr index was similar between site groups but generally higher
sites (F = 2.5776, p = 0.225, df1 = 3, df2 = 16). in residential areas (Table 2). The proportional abundance of EPT taxa,

Table 2
Summary riparian vegetation values for diversity metrics and indices of wetland prevalence, conservative species cover, native richness and percent cover. Aquatic macroinvertebrate taxa
richness (at the level of genus), mean total abundance, mean EPT richness (EPTr) and mean proportion of abundance of EPT (EPTa) taxa groups. ANOVA results of comparison of four land-
scape X channel type groups (F-statistic and p-value) reported for each metric.

ANOVA F-statistic Riparian vegetation Aquatic macroinvertebrates


(p-value)
Species Native Shannon Simpson's Wetland Cover weighted Native Taxa Mean EPTa EPTr
richness richness diversity (1-λ) PI mean C cover % richness abundance

0.38 (0.77) 0.94 (0.43) 0.34 (0.80) 0.49 (0.69) 0.29 1.12 (0.35) 0.85 (0.47) 3.92 4.85 0.62 0.30
(0.83) (0.03) (0.01) (0.61) (0.83)

CANALS 22.2 10.9 1.6 0.7 2.87 1.61 37.3 19.4 78.8 19.7 19
Residential 23.5 12.8 1.61 0.71 2.90 1.63 37.7 20.1 46.5 22.5 22.4
Agricultural 21.3 9.6 1.59 0.69 2.84 1.59 37 18.9 104.1 17.5 16.4
STREAMS 23.8 14 1.44 0.62 3.21 1.11 26.1 32.5 170 39.7 19.7
Residential 26.3 14.5 1.56 0.61 3.05 1.88 39.5 37 181.8 39.4 22.8
Agricultural 22 13.3 1.35 0.63 3.33 0.53 16 28 158.4 40.1 16.7
E.A. Carlson et al. / Science of the Total Environment 646 (2019) 770–781 777

irrigation and other calls for water by downstream users. The length
and prominence of these man-made aquatic and riparian ecosystems
in arid and semi-arid regions of the world highlights their importance.
We found that canals supported similar riparian plant and aquatic mac-
roinvertebrate species and functional groups as natural streams in our
study region. Riparian vegetation was influenced by surrounding land-
uses while aquatic macroinvertebrate communities were similar be-
tween and within land-use categories. Agricultural canals had similar
macroinvertebrate taxa, functional group composition, and plant func-
tional group composition, but different plant species composition
when compared to agricultural streams. Residential channels had simi-
lar macroinvertebrate communities yet distinct plant species and func-
tional groups. Habitat quality index (HQI) scores were lower for canals
than natural streams, yet residential canals had only slightly lower
scores than agricultural streams suggesting that irrigation canals could
be important ecosystem elements (albeit functioning at a lower level
than natural aquatic and riparian ecosystems) in human-dominated
landscapes.
Previous studies of the ecological attributes of constructed water-
ways have focused on drainage ditches in agricultural regions (Herzon
and Helenius, 2008; Leslie and Lamp, 2017) and transportation canals
(Harvolk et al., 2014). These artificial waterways have been shown to
support diverse biotic communities (Chester and Robson, 2013). Fish
utilizing irrigation canals were larger, had access to more and diverse
prey items while maintaining synchronous reproduction like fish in riv-
ers (Habit et al., 2005). In many regions artificial channels and wetlands
are temporary habitats but maintain biodiversity and ecological func-
tioning in agricultural landscapes (Bolpagni and Piotti, 2016).
Processes associated with natural streams in the study area such as
overbank flooding, vertical and lateral erosion and sediment deposition
rarely, or do not occur at all under normal canal construction, operation,
and maintenance (Depeweg and Mendez, 2002). The design and con-
struction of irrigation canals limits groundwater supported base flows
and surface water inputs characteristic of a stream's interaction with
its watershed, although inadvertent water leakage from canals can sup-
Fig. 5. nMDS of vegetation averaged by site (n = 54). Panel A – species level data with
vectors for selected species with a Pearson r N 0.5, 2-D stress was 0.22. Panel B – functional port extensive wetlands (Sueltenfuss et al., 2013).
group data with vectors for functional groups with Pearson r N 0.5, (IUG = introduced up- Daily streamflow had distinct inter- and intra-annual variability
land grass, IUF = introduced upland forb, IWG = introduced wetland grass, NMW = na- with peak flows driven by snowmelt and rain events. Peak flow in
tive mesic woody, IMG = introduced mesic grass), 2-D stress was 0.21.
small canals occurred earlier than in small streams and the opposite
was true for larger streams. A striking difference was that during the
however, was on average more than double for streams (Table 2). Chan- 183-day April–September study period canals averaged only 147 days,
nel size was a major contributor to EPT for streams where the large 80%, with measurable flows, while all study streams had flow on all
Cache la Poudre River sites had EPTr and EPTa of approximately 32 and 183 days.
80, while smaller streams were 13 and 3. The mayfly Tricorythodes
explicatus (Eaton) is relatively tolerant of poor water quality and higher 4.1. Riparian vegetation and aquatic macroinvertebrate communities
water temperatures (Relyea et al., 2000) and comprised 45% of the EPT
abundance in agricultural streams but was not common in agricultural In constructed channels, such as canals, stream flow patterns and
canals. Residential canals and streams shared many EPT taxa including variance can be analogous to natural streams, yet mechanical channel
the mayflies Baetis, Ephemerella, and Heptagenia. The HQI was highly maintenance and vegetation management can limit the formation of
correlated with the EPTa (r = 0.69, t = 4.05, p b 0.001) and moderately aquatic habitats and in-channel food resources. The hydrologic and geo-
with the EPTr indices (r = 0.47, t = 2.28, p = 0.04). Pearson correlation morphic differences between streams and canals measured in this study
for distance along canal network from the point of diversion from the produced no differences in species richness, Shannon diversity, or
river (Dist) to EPTr and EPTa was significantly negative while bankfull Simpson's index for riparian vegetation between natural and con-
width was positive. structed channels. Aquatic macroinvertebrate communities in agricul-
Stream vegetation was not hydrophytic according to the wetland tural and residential landscapes were also similar to streams. Habit
prevalence index as scores were N3.05, while canal vegetation on aver- et al. (1998) found similar species and biomass in irrigation canals and
age did have predominantly wetland plants with scores b2.90 (Table 2). streams in Chile, as did Koetsier and McCauley (2015) in Idaho, U.S.A.
Cover weighted mean C was between 1.59 and 1.63 for canals Other man-made channels have been reported to support diverse
surrounded by different land use while streams had a much larger aquatic and riparian communities along agricultural drainage ditches
range between 0.53 for agricultural streams and 1.88 for residential in temperate regions of the northern hemisphere (Herzon and
streams (Table 2). Helenius, 2008; Verdonschot et al., 2011).

4. Discussion 4.2. Functional groups of riparian plants and aquatic macroinvertebrates

Irrigation canals are artificial waterways that are simple in physical Plant functional groups or guilds have been used to characterize en-
form with strict flow schedules determined by the timing of crop vironmental filters in riparian ecosystems, and wildlife habitat (Merritt
778 E.A. Carlson et al. / Science of the Total Environment 646 (2019) 770–781

Fig. 6. Box plots of average percent cover of major plant growth forms for each channel type. Endpoints indicate maximum and minimum values, the box indicates 25th and 75th
percentiles with the median as the center line.

et al., 2010; Merritt and Bateman, 2012; Bruno et al., 2016). Aquatic between canals and streams. Decades of flow alteration in streams
macroinvertebrates can be grouped using physiological traits, life his- and intermittent flow patterns in canals could have reduced the re-
tory, and physical habitat requirements and are informative about the gional pool of taxa through environmental and trait based filters leading
type and condition of aquatic habitats (Poff et al., 2006). For plants to a higher proportion of generalist taxa that are tolerant of or able to
and aquatic invertebrate biota, functional groups indicated similarities avoid difficult conditions through adult colonization, rapid maturity,
and desiccation resistance (Mackay, 1992; Miller and Golladay, 1996;
Bogan and Boersma, 2012; Whatley et al., 2015).
Table 3 Functional plant attributes differed between residential and agricul-
Selected pairwise comparisons riparian vegetation (A), aquatic macroinvertebrates
tural landscapes. Fewer sites were characterized by woody plants along
(B) composition of landuse (Ag. – agricultural, Res. – residential) and channel types using
PERMANOVA. The agreement of significant differences found between streams and canals agricultural streams and canals compared to those in residential areas.
using the vegetation and aquatic macroinvertebrates datasets using a threshold p-value of This is likely due to vegetation management, particularly vegetation
0.05 (C). t-statistic (t-stat) and p-values (p-val) reported. 1 indicates Monte Carlo clearing, burning and herbicide application as well as a potentially
corrected p-value. more variable and erratic stream flows along canals and limited protec-
A) Species dataset Functional dataset tion of streams in agricultural landscapes. Vertical structure and species
T-stat p-val T-stat p-val
diversity and wetland plant cover were also significantly lower along
1
agricultural streams and canals. This has implications for wildlife that
Ag. stream vs. res. stream 0.948 0.449 1.118 0.2571
require structurally diverse riparian zones to provide cover and food
Ag. canal vs. Res. canal 2.052 0.001 1.768 0.008
Ag. canal vs. Ag. stream 1.083 0.027 1.008 0.392 (Meany et al., 2003; Lopez-Pomares et al., 2015).
Res. canal vs. Res. stream 1.344 0.05 1.752 0.013

B) Taxonomic dataset Functional dataset 4.3. Comparison of taxonomic and function data sets
T-stat p-val T-stat p-val
Comparisons of species composition among sites are commonly
Ag. stream vs. Res. stream 0.78 0.6571 0.68 0.7121 used to identify differences in biological communities (Petchey and
Ag. canal vs. Res. canal 1.35 0.025 1.38 0.029
Gaston, 2002). Grouping species into physiologically and ecologically
Ag. canal vs. Ag. stream 0.86 0.6441 0.70 0.8571
Res. canal vs. Res. stream 1.40 0.0981 1.41 0.1131 similar groups can add information about physical structure, food re-
sources and adaptations of communities to disturbance that are valu-
C) Species or taxonomic dataset Functional dataset able in understanding current or potential ecological functioning
Ag. stream vs. Res. stream Agree Agree (Merritt and Bateman, 2012; Petchey and Gaston, 2002).
Ag. canal vs. Res. canal Agree Agree We found statistical similarity between riparian plant species and
Ag. canal vs. Ag. stream Disagree Agree functional group composition for some comparisons of channel type/
Res. canal vs. Res. stream Disagree Disagree
land-use categories. A disagreement between the two datasets would
E.A. Carlson et al. / Science of the Total Environment 646 (2019) 770–781 779

ecosystems (Pozo et al., 1997; Moore et al., 2005). These can be even
more important in agricultural landscapes where riparian buffers are
being increasingly utilized to improve water quality and aquatic habitat
(Wooster and DeBano, 2006). This could improve the aquatic food web
and linkages with the surrounding riparian and terrestrial ecosystems
(Nakano et al., 1999; Nakano and Murakami, 2001; Ballanger and
Lake, 2005). Not all riparian areas have woody canopies, and prairie
streams in the region were likely dominated in some areas by herba-
ceous vegetation in the pre-settlement period due to frequent fires
that hindered survival of woody species. Herbaceous canals may act as
modern analogs (Dodds et al., 2004; Vandermyde and Whiles, 2015).
Distance along a canal from the point of stream diversion was hy-
pothesized to negatively impact macroinvertebrate communities, mak-
ing them dissimilar to regional streams. Small lateral canals at terminal
positions in the irrigation network had fewer taxa and were dominated
by chironomids and Simulium, likely in response to their adaptations to
poor water quality and short life history suitable for channels with long
dry periods, short flow duration, and turbid water. However, distance
was poorly correlated with overall community composition.

5. Conclusions

A holistic view of canals used to transport water for agricultural and


urban use, as integrated elements of the landscape is necessary to recog-
nize their contribution to and support of local and regional ecosystems
and biota. The prevalence of canals in semi-arid landscapes around the
world makes them important aquatic habitat to analyze and understand
in the context of river conservation and integrated river basin manage-
ment. The volume of water diverted into canals may exceed that re-
maining in natural streams, and in many regions, canals may be the
dominant lotic ecosystems. The landscape setting influences the quality
of streams and canals and their potential similarity. We found that ca-
nals supported similar riparian vegetation and aquatic macroinverte-
brate communities to streams in agricultural landscapes. These results
support the concept that canal networks have created habitat typical
of streams. This does not indicate that all irrigation canals create habi-
Fig. 7. nMDS of macroinvertebrate samples (n = 97). Panel A – taxa level data with vectors
of selected taxa showing correlation to points, 2-D stress was 0.24. Panel B – functional
tats comparable to those of natural streams, because adjacent land-use
group data, 2-D stress was 0.15. affected the similarities. Furthermore, the similarities between streams
and canals could be attributed to the degraded conditions of streams
and not to the development of diverse and high functioning riparian
and aquatic ecosystems.
indicate that a group of species provide similar ecological value or re-
The physical and biotic attributes of these artificial ecosystems sup-
spond to similar habitat characteristics, thus several species could be
port many rare species (Meany et al., 2003), habitat connectivity, tro-
present to provide the ecological functions. The similarity of canals
phic subsidies and regional biodiversity (Fernald et al., 2007; Lopez-
and streams in residential areas was indicated by an inconclusive statis-
Pomares et al., 2015). The importance of irrigation canals as ecosystems
tical result. Aquatic macroinvertebrate taxonomic and functional group
is further supported by the prevalence of artificial waterways in agricul-
composition were similar across all streams and canals.
tural and residential landscapes, the degradation of natural streams,
point and non-point source pollution, water extraction, and channeliza-
4.4. Environmental conditions and landscape and biotic indices tion. Irrigation canals have been critical to the development and support
of agriculture in arid and semi-arid regions throughout history in all re-
The HQI developed for this study was well correlated to EPTr and gions of the world. These byproducts of agricultural systems have cre-
EPTa indices as the physical and chemical characteristics of aquatic eco- ated riparian and aquatic habitats in sufficient amounts and with
systems such as water temperature, food resources, and in-channel similar biological communities as natural steams and should be consid-
woody debris are partly dependent on inputs from adjacent riparian ered integral components of agro-ecosystems.

Table 4
Description of functional macroinvertebrate groups determine through cluster analysis of physiological, ecological and life history traits from Poff et al. (2006).

Functional group Description of key characters Representative taxa

A Desiccation resistant predators/collector gatherers Isoperla, Optioservus, Dubiraphia, Atherix pachypus


B Large predators with long lived, winged adults Agabus, Claassenia sabulosa, Aeshna
C Collector gatherers of erosional habitats with synchronized life histories Brachycentrus, Hydropsyche, Simulium
D Small, tolerant collector gatherers of depositional habitats Chironomus, Cricotopus, Orthocladius, Tanytarsus
E Aquatic adults, strong swimmers, predators/herbivores Trichocorixa, Belostoma flumineum, Rhagovelia distincta
F Free moving collector gatherers with synchronized life histories Baetis, Siphlonurus occidentalis, Tricorythodes explicatus, Heptagenia
780 E.A. Carlson et al. / Science of the Total Environment 646 (2019) 770–781

Acknowledgements Habit, E., Bertran, C., Arevalo, S., Victoriano, P., 1998. Benthonic fauna of the Itata river and
irrigation canals (Chile). Irrig. Sci. 18, 91–99.
Habit, E., Victoriano, P., Campos, H., 2005. Trophic ecology and reproductive aspects of
The authors would like to thank the National Science Foundation Trichomycterus areolatus (Pisces, Trichomycteridae) in irrigation canal environments.
IGERT, I-WATER: Integrated Water, Atmosphere, Ecosystem Education Rev. Biol. Trop. 53, 195–210.
Harvey, J., Gooseff, M., 2015. River corridor science: hydrologic exchange and ecological
and Research Program at Colorado State University [DGE-0966346] consequences from bedforms to basins. Water Resour. Res. 51, 6893–6922.
and Colorado Water Institute (2014CO296B) for financial support and Harvolk, S., Symmank, L., Sundermeier, A., Otte, A., Donath, T.W., 2014. Can artificial wa-
Colorado State University for equipment. Permission to access irrigation terways provide a refuge for floodplain biodiversity? A case study from north west-
ern Germany. Ecol. Eng. 73, 31–44.
canals through private land was necessary at every site and deeply ap- Herzon, I., Helenius, J., 2008. Agricultural drainage ditches, their biological importance
preciated. Volunteers aided in collecting and processing plant and mac- and functioning. Biol. Conserv. 141, 1171–1183.
roinvertebrate samples, a very tedious task. This manuscript benefited Hickley, P., 1975. An apparatus for subdividing benthos samples. Oikos 26, 92–96.
Hobbs, R.J., Huenneke, L.F., 1992. Disturbance, diversity, and invasion: implications for
from theoretical and technical editing by a talented writing group at
conservation. Conserv. Biol. 6, 324–337.
Colorado State University. Jacobs, S.M., Bechtold, J.S., Biggs, H.C., Grimm, N.B., Lorentz, S., McClain, M.E., Naiman, R.J.,
Perakis, S.S., Pinay, G., Scholes, M.C., 2007. Nutrient vectors and riparian processing: a
Appendices review with special reference to African semiarid savanna ecosystems. Ecosystems
10, 1231–1249.
Katz, G.L., Stromberg, J.C., Denslow, M.W., 2009. Streamside herbaceous vegetation re-
Supplementary data to this article can be found online at https://doi. sponse to hydrologic restoration on the San Pedro River, Arizona. Ecohydrology 2,
org/10.1016/j.scitotenv.2018.07.246. 213–225.
Kennedy, T.A., et al., 2016. Flow management for hydropower extirpates aquatic insects,
undermining river food webs. Bioscience 66, 561–575.
Kerans, B.L., Karr, J.R., 1994. A benthic index of biotic integrity (B-IBI) for rivers of the Ten-
References nessee Valley. Ecol. Appl. 4, 768–785.
Khan, S., Tariq, R., Yuanlai, C., Blackwell, J., 2006. Can irrigation be sustainable? Agric.
Andersen, D.C., Nelson, S.M., 1997. Vegetation characteristics and butterfly use of unlined Water Manag. 80, 87–99.
and PVC-lined reaches of an irrigation delivery canal, Government Highline Canal, Kingsford, R.T., 2000. Ecological impacts of dams, water diversions and river management
Colorado, U.S.A. J. Arid Environ. 35, 747–764. on floodplain wetlands in Australia. Austral. Ecol. 25, 109–127.
Anderson, M.J., Gorley, R.N., Clarke, K.R., 2008. PERMANOVA+ for PRIMER: Guide to Soft- Koetsier, P., McCauley, L.M.M., 2015. An irrigation canals as a lotic mesocosm: examining
ware and Statistical Methods. PRIMER-E Plymouth, U.K. the relationship between macroinvertebrate benthos and drift. West. North Am. Nat.
Aspe, C., Jacque, M., 2015. Agricultural irrigation canals in southern France and new urban 75, 259–270.
territorial uses. Agric. Agric. Sci. Procedia 4, 29–39. Leigh, C., Sheldon, F., 2008. Hydrological changes and ecological impacts associated with
Ballanger, A., Lake, P.S., 2005. Energy and nutrient fluxes from rivers and streams into ter- water resource development in large floodplain rivers in the Australian tropics.
restrial food webs. Mar. Freshw. Res. 57, 15–28. River Res. Appl. 24, 1251–1270.
Bogan, M.T., Boersma, K.S., 2012. Aerial dispersal of aquatic invertebrates along and away Leigh, C., Reis, T.M., Sheldon, F., 2013. High potential subsidy of dry-season aquatic fauna
from arid-land streams. Freshwat. Sci. 31, 1131–1144. to consumers in riparian zones of wet-dry tropical rivers. Inland Waters 3, 411–420.
Bolpagni, R., Piotti, A., 2016. The importance of being natural in a human-altered Leslie, A.W., Lamp, W.O., 2017. Taxonomic and functional group composition of
riverscape: role of wetland type in supporting habitat heterogeneity and vegetation macroinvertebrate assemblages in agricultural drainage ditches. Hydrobiologia 787,
functional diversity. Aquat. Conserv. Mar. Freshwat. Ecosyst. 26, 1168–1183. 99–110.
Bonham, C.D., 2013. Measurements for Terrestrial Vegetation. John Wiley & Sons, West Lichvar, R.W., 2012. The National Wetland Plant List. Cold Regions Research and Engineer-
Sussex. ing Laboratory, Hanover NH.
Bruno, D., Gutierrez-Canovas, C., Velasco, J., Sanchez-Fernandez, D., 2016. Functional re- Lopez-Pomares, A., Lopez-Iborra, G.M., Martin-Cantarino, C., 2015. Irrigation canals in a
dundancy as a tool for bioassessment: a test using riparian vegetation. Sci. Total En- semi-arid agricultural landscape surrounded by wetlands: their role as a habitat for
viron. 566, 1268–1276. birds during the breeding season. J. Arid Environ. 118, 28–36.
Bunn, S.E., Arthington, A.H., 2002. Basic principles and ecological consequences of altered Mackay, R.J., 1992. Colonization by lotic macroinvertebrates: a review of processes and
flow regimes for aquatic biodiversity. Environ. Manag. 30, 492–507. patterns. Can. J. Fish. Aquat. Sci. 49, 617–628.
Butler, Malcolm G., 1984. Life histories of aquatic insects. In: Resh, Vincent H., Rosenberg, Malmqvist, B., Rundle, S., 2002. Threats to running water ecosystems of the world. Envi-
David M. (Eds.), The Ecology of Aquatic Insects. Praeger Publishers, New York, ron. Conserv. 29, 134–153.
pp. 24–55. McCune, B., Grace, J.B., 2002. Analysis of Ecological Communities. MjM Software,
Casas, J.J., Sánchez-Oliver, J.S., Sanz, A., Furné, M., Trenzado, C., Juan, M., Paracuellos, M., Gleneden Beach, Oregon, USA www.pcord.com.
Suárez, M.D., Fuentes, F., Gallego, I., Gil, C., Ramos-Miras, J.J., 2011. The paradox of Meany, C.A., Ruggles, A.K., Ludbow, B.C., Clippinger, N.W., 2003. Abundance, survival and
the conservation of an endangered fish species in a Mediterranean region under ag- hibernation of Preble's meadow jumping mice (Zapus hudsonius preblei) in Boulder
ricultural intensification. Biol. Conserv. 144, 253–262. County, Colorado. Southwest. Nat. 48, 610–623.
Chester, E.T., Robson, B.J., 2013. Anthropogenic refuges for freshwater biodiversity: their Merritt, D.M., Bateman, H.L., 2012. Linking streamflow and groundwater to avian habitat
ecological characteristics and management. Biol. Conserv. 166, 64–75. in a desert riparian system. Ecol. Appl. 22, 1973–1988.
Clarke, K.R., Gorley, R.N., Somerfield, P.J., Warwick, R.M., 2014. Change in Marine Commu- Merritt, D.M., Cooper, D.J., 2000. Riparian vegetation and channel change in response to
nities: An Approach to Statistical Analysis and Interpretation. 3rd Ed. Bretonside river regulation: a comparative study of regulated and unregulated streams in the
Copy, Plymouth, UK. Green River Basin, USA. Regul. Rivers Res. Manag. 16, 543–564.
Cooper, D.J., Merritt, D.M., Andersen, D.A., Chimner, R., 1999. Factors controlling Fremont Merritt, R.W., Cummins, K.W., Resh, V.H., Batzer, P., 2008. Sampling aquatic insects: col-
cottonwood seedling establishment on the upper Green River, Colorado and Utah. lection devices, statistical considerations, and rearing procedures. In: Merritt, R.W.,
Regul. Rivers Res. Manag. 15, 419–440. Cummins, K.W., Berg, M.B. (Eds.), An Introduction to the Aquatic Insects of North
Depeweg, H., Mendez, V.N., 2002. Sediment transport applications in irrigation canals. America, 4th edition Kendall Hunt, Dubuque Iowa, pp. 15–37.
Irrig. Drain. 51, 167–179. Merritt, D.M., Scott, M.L., Poff, N.L., Auble, G.T., Lytle, D.A., 2010. Theory, methods and tools
Dodds, W.K., Gido, K., Whiles, M.R., Fritz, K.M., Matthews, W.J., 2004. Life on the edge: the for determining environmental flows for riparian vegetation: riparian vegetation-
ecology of Great Plains prairie streams. Bioscience 54, 205–216. flow response guilds. Freshw. Biol. 55, 206–225.
Egginton, G.E., Robbins, W.W., 1920. Irrigation water as a factor in the dissemination of Miller, A.M., Golladay, S.W., 1996. Effects of spates and drying on macroinvertebrate as-
weed seeds. The Agricultural Experiment Station of the Colorado Agricultural College semblages of an intermittent and perennial prairie stream. J. N. Am. Benthol. Soc.
Bulletin. 253, pp. 3–25. 15, 670–689.
Ernegger, T., Grubinger, H., Vitek, E., Csekits, C., Eitzinger, J., Gaviria, S., Kotek, D., Krisaf, H., Miller, S.W., Wooster, D., Li, J., 2007. Resistance and resilience of macroinvertebrates to ir-
Nachtnebel, H.P., Pritz, B., Sabbas, T., Sschmutz, S., Schreiner, P., Stephan, U., Unfer, G., rigation water withdrawals. Freshw. Biol. 52, 2494–2510.
Wychera, U., Neudorfer, W., 1998. A natural stream created by human engineering: Moore, R.D., Spittlehouse, D.L., Story, A., 2005. Riparian microclimate and stream temper-
investigations on the succession of the Marchfeld Canal in Austria. Regul. Rivers ature response to forest harvesting: a review. J. Am. Water Resour. Assoc. 41,
Res. Manag. 14, 119–139. 813–834.
Evans, H.E., Evans, M.A., 1991. Cache la Poudre: The Natural History of a Rocky Mountain Muehlbauer, J.D., Collins, S.F., Doyle, M.W., Tockner, K., 2014. How wide is a stream? Spa-
River. University Press of Colorado, Niwot, Colorado. tial extent of the potential “stream signature” in terrestrial food webs using meta-
Fernald, A.G., Baker, T.T., Guldan, S.J., 2007. Hydrologic, riparian, and agroecosystem func- analysis. Ecology 95, 44–55.
tions of traditional acequia irrigation systems. J. Sustain. Agric. 30, 147–171. Naiman, R.J., DeCamps, H., Pollack, M., 1993. The role of riparian corridors in maintaining
Gonzalez del Tanago, M., Garcia de Jalon, D., 2011. Riparian Quality Index (RQI): a meth- regional biodiversity. Ecol. Appl. 3, 209–212.
odology for characterizing and assessing the environmental conditions of riparian Nakano, S., Murakami, M., 2001. Reciprocal subsidies: dynamic interdependence between
zones. Limnetica 30, 235–254. terrestrial and aquatic food webs. Proc. Natl. Acad. Sci. 98, 166–170.
Graf, W.L., 1999. Dam nation: a geographic census of large American dams and their hy- Nakano, S., Miyasaka, H., Kuhara, N., 1999. Terrestrial–aquatic linkages: riparian arthro-
drologic impacts. Water Resour. Res. 35, 1305–1311. pod inputs alter trophic cascades in a stream food web. Ecology 80, 2435–2441.
Gregory, K.J., 2006. The human role in changing river channels. Geomorphology 79, Nilsson, C., Svedmark, M., 2002. Basic principles and ecological consequences of changing
172–191. water regimes: riparian plant communities. Environ. Manag. 30, 468–480.
E.A. Carlson et al. / Science of the Total Environment 646 (2019) 770–781 781

Patten, D.J., 1998. Riparian ecosystems of semi-arid North America: diversity and human Swamee, P.K., 1995. Optimal irrigation canal sections. Irrig. Drain. Eng. 121, 467–469.
impacts. Wetlands 18, 498–512. Tank, J.L., Rosi-Marshall, E.J., Griffiths, N.A., Entrekin, S.A., Stephen, M.L., 2010. A review of
Paul, M.J., Meyer, J.L., 2001. Streams in the urban landscape. Annu. Rev. Ecol. Syst. 32, allochthonous organic matter dynamics and metabolism in streams. J. N. Am. Benthol.
333–365. Soc. 29, 118–146.
Petchey, O.L., Gaston, K.J., 2002. Functional diversity (FD), species richness and commu- Urban, M.A., Rhodes, B.L., 2003. Catastrophic human-induced change in stream-channel
nity composition. Ecol. Lett. 5, 402–411. planform and geometry in an agricultural watershed, Illinois, USA. Ann. Assoc. Am.
Petraitis, P.S., Latham, R.E., Niesenbaum, R.A., 1989. The maintenance of species diversity Geogr. 93, 783–796.
by disturbance. Q. Rev. Biol. 64, 393–418. Vandermyde, J.M., Whiles, M.R., 2015. Effects of experimental forest removal on macroin-
Poff, N.L., Zimmerman, J.K.H., 2010. Ecological responses to altered flow regimes: a liter- vertebrate production and functional structure in tallgrass prairie streams. Freshwat.
ature review to inform the science and management of environmental flows. Freshw. Sci. 34, 519–534.
Biol. 55, 194–205. Verdonschot, R.C.M., Keizer-Vlek, H., Verdonschot, P.F.M., 2011. Biodiversity value of agri-
Poff, N.L., Olden, J.D., Vieira, N.K.M., Finn, D.S., Simmons, M.P., Kondratieff, B.C., 2006. cultural drainage ditches: a comparative analysis of the aquatic invertebrate fauna of
Functional trait niches of North American lotic insects: trait-based ecological applica- ditches and small lakes. Aquat. Conserv. Mar. Freshwat. Ecosyst. 21, 715–727.
tions in light of phylogenetic relationships. J. N. Am. Benthol. Soc. 25, 730–755. Vermonden, K., Leuven, R.S.E.W., van der Velde, G., van Katwijk, M.M., Roelofs, J.G.M.,
Polis, G.A., Anderson, W.B., Holt, R.D., 1997. Toward an integration of landscape and food Hendrisk, A.J., 2009. Urban drainage systems: an undervalued habitat for aquatic
web ecology: the dynamics of spatially subsidized food webs. Ann. Rev. Ecol. 28, macroinvertebrates. Biol. Conserv. 142, 1105–1115.
289–316. Wallace, J.B., Grubaugh, J.W., Whiles, M.R., 1996. Biotic indices and stream ecosystem pro-
Pozo, J., Gonzalez, E., Diez, J.R., Molinero, J., Elosegui, A., 1997. Inputs of particulate organic cesses: results from an experimental study. Ecol. Appl. 6, 140–151.
matter to streams with different riparian vegetation. J. N. Am. Benthol. Soc. 16, Weber, W.A., Wittmann, R.C., 2012. Colorado Flora: Eastern Slope. 4thd Edition. Univer-
602–611. sity Press of Colorado, Boulder.
Relyea, C.D., Minshall, G.W., Danehy, R.J., 2000. Stream insects as bioindicators of fine sed- Whatley, M.H., Vonk, J.A., van der Geest, H.G., Admiraal, W., 2015. Temporal abiotic vari-
iment. Proc. Water Environ. Fed. (6), 663–686. ability structures invertebrate communities in agricultural drainage ditches.
Rosenberg, D.M., King, R.S., Resh, V.H., 2008. Use of aquatic insects in biomonitoring. Limnologica 52, 20–29.
Chapter 7. In: Merritt, R.W., Cummins, K.W., Berg, M.B. (Eds.), An Introduction to
Williams, C.A., Cooper, D.C., 2005. Mechanisms of riparian cottonwood decline along reg-
Aquatic Insects of North America, 4th Edition 2008. Kendal Hunt (pp).
ulated rivers. Ecosystems 8, 382–395.
Shafroth, P.B., Stromberg, J.C., Patten, D.T., 2002. Riparian vegetation response to altered
Williams, D., Feltmate, B., 1992. Aquatic Insects. CAB International Wallingford, U.K.
disturbance and stress regimes. Ecol. Appl. 12, 107–123.
Wohl, E.E., 2001. Virtual Rivers: Lessons from the Mountain Rivers of the Colorado Front
Strange, E.M., Fausch, K.D., Covich, A.P., 1999. Sustaining ecosystem services in human-
Range. Yale University Press, New Haven.
dominated watersheds: biohydrology and ecosystem processes in the South Platte
Wooster, D.E., DeBano, S.J., 2006. Effect of woody riparian patches in croplands on stream
River Basin. Environ. Manag. 24, 39–54.
macroinvertebrates. Arch. Hydrobiol. 165, 241–268.
Sueltenfuss, J.P., Cooper, D.J., Knight, R.L., Waskom, R.M., 2013. The creation and mainte-
nance of wetland ecosystems from irrigation canal and reservoir seepage in a semi-
arid landscape. Wetlands 33, 799–810.

Potrebbero piacerti anche