ArticlePDF Available

Effects of habitat area, isolation, and landscape diversity on plant species richness of calcareous grasslands

Authors:

Abstract and Figures

Calcareous grasslands harbour a high biodiversity, but are highly fragmented and endangered in central Europe. We tested the relative importance of habitat area, habitat isolation, and landscape diversity for species richness of vascular plants. Plants were recorded on 31 calcareous grasslands in the vicinity of the city of Gttingen (Germany) and were divided into habitat specialist and generalist species. We expected that habitat specialists were more affected by area and isolation, and habitat generalists more by landscape diversity. In multiple regression analysis, the species richness of habitat specialists (n = 66 species) and habitat generalists (n = 242) increased with habitat area, while habitat isolation or landscape diversity did not have significant effects. Contrary to predictions, habitat specialists were not more affected by reduced habitat area than generalists. This may have been caused by delayed extinction of long-living plant specialists in small grasslands. Additionally, non-specialists may profit more from high habitat heterogeneity in large grasslands compared to habitat specialists. Although habitat isolation and landscape diversity revealed no significant effect on local plant diversity, only an average of 54% of habitat specialists of the total species pool were found within one study site. In conclusion, habitat area was important for plant species conservation, but regional variation between habitats contributed also an important 46% of total species richness.
Content may be subject to copyright.
Effects of habitat area, isolation, and landscape
diversity on plant species richness of calcareous
grasslands
JOCHEN KRAUSS*, ALEXANDRA-MARIA KLEIN,
INGOLF STEFFAN-DEWENTER and TEJA TSCHARNTKE
Department of Agroecology, University of Go
¨
ttingen, Waldweg 26, D-37073 Go
¨
ttingen, Germany;
*Author for correspondence (e-mail: j.krauss@uaoe.gwdg.de; fax: +49-551-398806)
Received 18 November 2002; accepted in revised form 15 May 2003
Key words: Conservation, Generalists, Habitat fragmentation, Specialists, Species density, Species–area
relationships
Abstract. Calcareous grasslands harbour a high biodiversity, but are highly fragmented and endangered
in central Europe. We tested the relative importance of habitat area, habitat isolation, and landscape
diversity for species richness of vascular plants. Plants were recorded on 31 calcareous grasslands in the
vicinity of the city of Go
¨
ttingen (Germany) and were divided into habitat specialist and generalist
species. We expected that habitat specialists were more affected by area and isolation, and habitat
generalists more by landscape diversity. In multiple regression analysis, the species richness of habitat
specialists (n ¼ 66 species) and habitat generalists (n ¼ 242) increased with habitat area, while habitat
isolation or landscape diversity did not have significant effects. Contrary to predictions, habitat spe-
cialists were not more affected by reduced habitat area than generalists. This may have been caused by
delayed extinction of long-living plant specialists in small grasslands. Additionally, non-specialists may
profit more from high habitat heterogeneity in large grasslands compared to habitat specialists. Although
habitat isolation and landscape diversity revealed no significant effect on local plant diversity, only an
average of 54% of habitat specialists of the total species pool were found within one study site. In
conclusion, habitat area was important for plant species conservation, but regional variation between
habitats contributed also an important 46% of total species richness.
Introduction
Habitat loss and habitat fragmentation of natural and semi-natural habitats are
considered as major threats to biodiversity. Semi-natural calcareous grasslands
belong to the most species-rich plant and insect habitats in central Europe (Van
Swaay 2002; WallisDeVries et al. 2002). Habitat loss of calcareous grasslands in
Germany reached locally up to 60% in the last 100 years. Thereby, 1% of the
plant species restricted to this habitat type became extinct and 42% are threa-
tened in Germany (WallisDeVries et al. 2002). Calcareous grasslands are highly
fragmented, endangered, and nowadays protected by law in Germany (Riecken
et al. 1994). Interest in conservation is high (Beinlich and Plachter 1995;
WallisDeVries et al. 2002), leading to several studies on plant and insect com-
munity structure and extinction risks (e.g. Fischer and Sto
¨
cklin 1997; Bruun
2000; Zschokke et al. 2000; Krauss et al. 2003). However, a recent review
# 2004 Kluwer Academic Publishers. Printed in the Netherlands.
Biodiversity and Conservation 1427–1439, 2004.
13:
emphasises the lack of well-replicated community level studies on calcareous
grasslands (Steffan-Dewenter and Tscharntke 2002).
Habitat fragmentation is a major threat for plant species richness (e.g. Ouborg 1993;
Grashof-Bokdam 1997; Honnay et al. 1999; Bruun 2000). The inuence of landscape
context in habitat fragmentation studies is often ignored (Wiens 1997; Hanski 1999),
although it is essential to understand plant community dynamics (De Blois et al. 2002).
Habitat isolation might be seen as a simple measure of landscape context, as it de-
scribes the distance to neighbouring habitats or the proportion of the habitats of a
certain type within a landscape (see Moilanen and Nieminen 2002). But landscape
context includes more factors that may affect local communities, as there is the pro-
portion of other habitat types or the diversity of habitats in the landscape matrix
(Steffan-Dewenter et al. 2002). Similar habitats and high habitat diversity in the sur-
roundings might enhance species richness of a local site. Landscape context might
affect gene ow via pollen and seed dispersal, the pressure of herbivory, and the
functional connectivity between habitat patches (Ricketts 2001). Recently published
studies started to focus on the complex effects of landscape context on local plant
species richness (Kollman and Schneider 1999; Metzger 2000; So
¨
derstro
¨
m et al. 2001).
Not all species depend on habitat area, isolation and landscape context equally
(Tscharntke et al. 2002). (1) Habitat specialists are more affected by habitat loss
than generalists, (Warren et al. 2001). (2) The surrounding landscape is inhabitable
for habitat specialists, but at least partly habitable for generalists, supporting the
prediction that habitat isolation affects habitat specialists more than generalists (see
Jonsen and Fahrig 1997). (3) High landscape diversity in the surrounding matrix
provides more different habitat types for generalists or species with other habitat
preferences, supporting the prediction that landscape diversity enhances the number
of generalists, especially at edges, but hardly specialists (see Jonsen and Fahrig
1997). These predictions have rarely been tested for plant communities in the
context of spatial dynamics. True forest species showed a stronger response to
habitat area than other species, but no response to habitat isolation (Honnay et al.
1999), and shade tolerant species were more sensitive to habitat fragmentation than
shade intolerant species (Metzger 2000). Numerous functional groups, including
specialist grassland species, were positively related to habitat area and incon-
sistently to isolation (Bruun 2000).
Speciesarea relationships can be explained by two ecological hypotheses.
(1) The habitat heterogeneity hypothesis predicts higher species numbers because
of higher habitat heterogeneity. (2) The area per se or equilibrium hypothesis
considers colonisationextinction dynamics to cause increasing species numbers
with increasing habitat area independent of habitat heterogeneity (Rosenzweig
1995). Habitat area and habitat heterogeneity are often closely correlated (Kohn
and Walsh 1994; Rosenzweig 1995; Ko
¨
chy and Rydin 1997). Equal sample sizes
should reduce the habitat heterogeneity effect, making it possible to test separately
for area per se effects (Kelly et al. 1989; Ko
¨
chy and Rydin 1997).
In this study we tested the impact of the three parameters habitat area, habitat
isolation, and landscape diversity on the species richness of habitat specialist and
generalist vascular plants on calcareous grasslands.
1428
We tested the following predictions:
(1) Plant species richness increases with increasing habitat area, decreasing habitat
isolation and increasing landscape diversity.
(2) Species richness of habitat specialists is more sensitive to habitat area and
isolation, whereas species richness of generalists is more sensitive to landscape
diversity.
(3) Species density, that is, plant species richness per area in samples of equal size
on all study sites, is related to habitat area, isolation and landscape diversity.
Materials and methods
Study region and study sites
Altogether 31 calcareous grasslands in the vicinity of the city of Go
¨
ttingen in
Lower Saxony (Germany) were studied. The grasslands were located in the Leine-
Weser Mountain lands (Gauss-Kru
¨
ger: R ¼ 5695, H ¼ 3550=R ¼ 5724, H ¼ 3579)
and belong to the plant association Gentiano-Koelerietum. Low impact manage-
ment to stop succession and to remove woody bushes was carried out on the study
sites mainly from late summer to winter. The average rainfall in the area around
Go
¨
ttingen is 635 mm per annum, with an average temperature of 6.8 8C (Deutscher
Wetterdienst 2001). The landscape is structurally rich with a mosaic of diverse
habitat types. Calcareous grasslands can be sharply delimited from the surrounding
landscape with little or no ambiguity and cover 0.26% of the study region.
Habitat area, isolation, landscape diversity
The area of the 31 calcareous grasslands was measured with a differential GPS
GEOmeter 12L (GEOsat GmbH 1998) and ranged from 314 to 51395 m
2
.Area
covered with shrubs was excluded from the measurement.
Habitat isolation was measured as an index (I) of each study site (i) from edge to
edge on the basis of all known calcareous grasslands in a radius of 8 km around our
study sites using the following formula:
I ¼ e
dij
A
j
where A
j
is the size (in m
2
) of neighbouring calcareous grasslands and dij the
distance (in km) from the neighbouring grassland j to the study site i. The formula
is based on Hanski et al. (1994). Larger values of I indicate lower isolation than
smaller values. Additionally, habitat isolation was measured as distance from edge
to edge to the nearest calcareous grassland (551894 m). This isolation distance and
the isolation index were always log
10
transformed. Both were correlated
(r ¼0.445, n ¼ 31, P ¼ 0.012) and showed similar, always non-signicant results.
Therefore we show only the results from the isolation index.
1429
Landscape diversity was analysed using modied digital thematic maps
(ATKIS
1
-DLM 25=1 Landesvermessung and Geobasisinformationen Niedersachsen
19911996, and ATKIS
1
-DLM 25/2 Hessisches Landesvermessungsamt 1996).
Eleven land-use types were dened, including arable land (42.15% of the study re-
gion), forest (36.80%), grassland (12.14%), built-up area (6.24%), other habitats
(1.48%), garden land (0.31%), hedgerow (0.30%), calcareous grassland (0.26%),
orchard meadow (0.20%), plantation (0.06%), and fen (0.05%). We used the Shannon
Wiener index to calculate landscape diversity for each of the 31 grassland fragments
using a nested set of 12 circles with a radius ranging from 0.25 to 3.00 km in 0.25 km
steps:
Hs ¼p
i
ln p
i
where p
i
is the proportion of each of the 11 different land-use types (Krebs 1989).
Further we pooled the plant species-rich habitat types, grassland, garden land,
hedgerow, calcareous grassland, orchard meadow, and fen around the study sites.
We related the different plant species richness data with the proportion of this
pooled habitat type data set. We never found any positively signicant relation
between both, and the pooled habitat types were highly correlated with landscape
diversity (250 m scale: r ¼ 0.685, P<0.0001). Therefore we only show results from
landscape diversity.
Due to some missing ATKIS data we could test the landscape data for a radius of
0.25 and 0.50 km for all 31 study sites, while 0.752.00 km scales were tested for
only 30, 2.25 km for 29, and 2.503.00 km for 28 sites. For each landscape analysis,
the habitat area of the central study site was excluded.
Plants
The complete survey of vascular plants (Spermatophyta plus Equisetum arvense L.)
for all of the 31 calcareous grasslands was compiled from four independent data
sets from 1996, 2000 and 2001 to achieve a total list of plant species per study site.
In 1996 and 2000 the vegetation was mapped in May=June and again in August in
randomised plots of 25 m
2
according to Braun-Blanquet. These plots were the same
for both records in 1996, while they changed in 2000. In 1996 one plot was mapped
for small (>1500 m
2
), two plots for medium (15005000 m
2
), three plots for
medium to large (5000 1 0000 m
2
) and four plots for large grasslands (>10 000 m
2
)
(see Steffan-Dewenter and Tscharntke 2000). In 2000 one plot was mapped on
small grasslands (<1500 m
2
), two plots on medium grasslands (150010000 m
2
),
and three plots on large grasslands (>10000 m
2
). To complete these two Braun-
Blanquet surveys, in 2000 between April and August we mapped ve times the
plant species in ower of each grassland on randomised transects. Each transect
covered from 11.4% (largest habitat) to 100.0% (small habitats) of the total habitat
area (average: 62.7 0.4%). In a fourth survey in May 2001, again plants were
mapped on the 31 grasslands to complete the species list.
1430
We used the total number of species per grassland for statistical analyses. To
compare equal sample sizes (species density) we also tested all correlations using
only one 25 m
2
plot (mapped in spring and summer) of the 2000 data set.
All vascular plant species recorded as species restricted to oligotrophic grasslands
in Von Drachenfels (1994) were dened as habitat specialists (n ¼ 66) for calcareous
grasslands (see Appendix 1). These species mainly occur on calcareous grasslands in
the study region, but might inhabit other habitat types in other regions of Germany
and Europe. All other species with no habitat preferences or preferences for other
habitats were dened as generalist plants (242 species). Species identication and
nomenclature follow Ba
¨
ssler et al. (1999).
Statistical analyses
The statistical analyses of the data were performed using the software Statgraphics
Plus for Windows 3.0 (Statgraphics 1995). All data were tested on whether they
satisfy the assumption of normality. We calculated simple and multiple regressions,
Pearson correlations, and comparisons of regression lines (Sokal and Rohlf 1995).
We chose backward selections for stepwise multiple regressions. The independent
variables habitat area and isolation were always log
10
transformed. Species num-
bers in regressions were also log
10
transformed to calculate scale-independent
slopes (z-values) for comparison with other studies. Arithmetic means one
standard error are given in the text.
Results
Habitat area (0.03 5.14 ha, average: 0.90 0.23 ha) was not correlated with ha-
bitat isolation (index: 2051 85 978, average: 23 019 3368) (r ¼0.013,
P ¼ 0.944). Also landscape diversity (at a 250 m scale; ShannonWiener: 0.09
1.56, average: 1.09 0.05) was not correlated with habitat isolation (r ¼0.015,
P ¼ 0.936), but with habitat area (r ¼ 0.375, P ¼ 0.038).
Altogether 308 plant species were identied on all 31 calcareous grasslands with
pooled data. This was an average of 89.0 3.7 species per habitat with a minimum of
56 species and a maximum of 138 species. In Braun-Blanquet plots we found 192
species in 1996 and 234 species in 2000, resulting in a total of 278 species. Additional
recordings due to the transect walks in 2000 and the survey in 2001 resulted in a total
of 308 species. Nine of 66 (13.6%) specialist species (see Appendix 1) and seven out
of 242 (2.9%) habitat generalist species (Achillea millefolium L., Campanula ro-
tundifolia L., Daucus carota L., Leontodon hispidus L., Lotus corniculatus L.,
Plantago lanceolata L., and Plantago media L.) were found on each of the 31
calcareous grasslands. A minimum of 33% and a maximum of 67% of habitat spe-
cialists of the 66 specialists were found within one study site. On average, per study
site 54% of all specialist species was found, emphasising that 46% of total species
richness was due to between-habitat variation (see Veech et al. (2002) for calculations
of b-diversity).
1431
Species richness increased with increasing habitat area for both habitat specialist
and generalist species, leading to an increase of all species (Figure 1, Table 1). The
z-value (slope of loglog regressions) for specialists (z-value: 0.08) was lower than
Figure 1. Relationship between the number of plant species and grassland area (n ¼ 31 fragments). (A)
Specialist plant species (66 species): y ¼ 13.64 þ 6.32 log
10
x. (B) Generalist plant species (242 species):
y ¼1.40 þ 15.44 log
10
x. For statistics see Table 1. Comparison of regression lines: slopes: F ¼ 5.34,
P ¼ 0.024.
1432
Table 1. Simple regressions for the habitat specialists among the plant species (66 species), generalist species (242 species) and all plant species (308 species).
Relations between species numbers and species density with the three independent factors habitat area, habitat isolation, and landscape diversity (n ¼ 31). Habitat area
and habitat isolation are log
10
transformed.
Habitat area Habitat isolation Landscape diversity (radius: 250 m)
Fr P Fr PFr
P
Total species numbers
Habitat specialists 23.49 0.669 <0.0001 0.39 0.116 0.536 3.81 0.341 0.061
Habitat generalists 17.18 0.610 0.0003
0.95 0.178 0.337 0.16 0.074 0.692
All plants 24.67 0.678 <0.0001 1.04 0.186 0.316 0.76 0.160 0.390
Species density
Habitat specialists 0.09 0.054 0.773
0.39 0.116 0.535 0.00 0.008 0.965
Habitat generalists 0.02 0.028 0.879 0.00 0.012 0.950 0.00 0.004 0.981
All plants 0.06 0.046 0.807 0.11 0.061 0.744 0.00 0.000 0.999
1433
for generalists (z-value: 0.13), but did not differ signicantly (comparison of re-
gression lines F ¼ 2.16, P ¼ 0.148). The slope for not log-transformed species
numbers was even signicantly steeper for generalists (Figure 1). For all plant
species the z-value was 0.11. Habitat isolation showed no signicant effect on
species richness in simple regressions, while landscape diversity at the smallest
scale of 250 m was positively related to specialist species with marginal sig-
nicance (Table 1). We analysed the effects of landscape diversity at 12 nested
spatial scales. Correlations were highest at the smallest spatial scale of 250 m for
specialist, generalist and all plants (results for other spatial scales are not shown).
In multiple regressions with habitat area, isolation and landscape diversity
(250 m), habitat area was the only signicant factor explaining 44.8% of variation
for specialists, 37.2% for generalists and 46.0% for all plant species.
Considering only equal sample size with one plot per grassland, we found al-
together 199 plant species. This was an average of 46.5 1.4 species per habitat
with a minimum of 30 species and a maximum of 63 species. Neither in simple
(Table 1) nor in multiple regressions we found evidence for an impact of habitat
area, isolation, and landscape diversity on plant species density for specialists,
generalists or all plants on the studied 31 calcareous grasslands.
Discussion
Landscape diversity is known to increase generalist insect species numbers (Jonsen
and Fahrig 1997; Krauss et al. 2003), but did not affect the number of generalist
plants in our study. Increasing landscape diversity even tended to increase specialist
plant species numbers, but due to correlation with habitat area this effect was
eliminated in multiple regressions. We also could not detect an effect of landscape
on total plant species richness. This might be explained by the relatively complex
landscapes that surrounded our calcareous grasslands. In multiple tests for different
landscape parameters and numerous functional groups, inconsistent results of
correlations of plant species richness with landscape diversity were reported
(Kollman and Schneider 1999; Metzger 2000). To reduce multiple testing and
intercorrelations of landscape factors, Steffan-Dewenter et al. (2002) suggest fo-
cusing on landscape diversity and proportion of habitats in the surrounding matrix.
So
¨
derstro
¨
m et al. (2001) found lower plant species numbers on semi-natural pas-
tures with increasing proportion of arable elds in the surrounding landscape, but
landscape diversity was not studied. In general, landscape studies on plant species
composition are still rare, but essential to fully understand community dynamics
(De Blois et al. 2002).
Species numbers increased signicantly with increasing habitat area for both ha-
bitat specialist and habitat generalist plant species. Our results conrm the general
validity of speciesarea relationships, as shown before for plant communities (e.g.
Ouborg 1993; Grashof-Bokdam 1997; Honnay et al. 1999; Bruun 2000). Our high
sampling effort appeared to guarantee almost complete recordings. In addition, our
numbers of plant species on 27 of the grasslands were very similar to a complete
1434
survey made between 1980 and 1986 (Eggers 1986, unpublished data; B. Preuschoff,
personal communication). We did not measure habitat heterogeneity, but species
density, that is, species numbers measured with equal sample sizes on all grasslands,
was not affected by habitat area in our study, giving indirect evidence for the habitat
heterogeneity hypothesis as an explanation for speciesarea relations (Kelly et al.
1989; Holt et al. 1999). Habitat heterogeneity is assumed to be the main predictor for
plant species richness (Ko
¨
chy and Rydin 1997), as neither Ko
¨
chy and Rydin (1997)
nor Lawesson et al. (1998) found positive speciesarea relations for area-corrected
sample sizes.
Surprisingly, habitat specialists were not more affected by habitat area than gen-
eralists. This is in contrast to previous ndings for specialist and generalist plant
species, where the z-value was higher for true forest plant species (z ¼ 0.45) than for
species of edges and clearings (z ¼ 0.39) and for woody species and lianas (z ¼ 0.17)
in Belgium forests (Honnay et al. 1999). Also habitat specialist butteries on the
same calcareous grasslands were signicantly more affected by habitat loss than
generalists (Krauss et al. 2003). Large areas with high habitat heterogeneity can be
expected to offer a diverse mosaic of microhabitat types for species that are non-
specialists. This may explain the steeper speciesarea relationships for generalists in
our study. Habitat specialists are expected to have higher extinction rates than gen-
eralists, as shown for plants and butteries (Fischer and Sto
¨
cklin 1997; J. Krauss,
unpublished data). However, when these specialists are perennial plants, they may
persist for several years in small populations, thereby delaying their extinction
(Oostermeijer et al. 1994). An extinction debt for these species in small habitats keeps
the z-value lower than in equilibrium situations (Tilman et al. 1994; Gonzales 2000).
Hypotheses developed for mobile birds and butteries with short life cycles might be
not applicable to sessile plants with long life cycles (Eriksson 1996).
Contrary to our predictions, we could not nd effects of habitat isolation, neither
for specialist nor for generalist plant species. Isolation effects on plant species are
mainly found for plant species with low dispersal abilities (Van Ruremonde and
Kalkhoven 1991; Grashof-Bokdam 1997), while other species groups generally
show no or weak isolation effects (Ouborg 1993; Ko
¨
chy and Rydin 1997; Honnay
et al. 1999; Bruun 2000). Also for the more mobile buttery species no isolation
effects could be reported on the same calcareous grasslands (Steffan-Dewenter and
Tscharntke 2000; Krauss et al. 2003). As calcareous grasslands were better con-
nected decades ago (WallisDeVries et al. 2002), we might have found patterns that
show previous situations because of extinction debt of plants, and therefore no
isolation effect.
In conclusion, habitat area was the only predictor explaining vascular plant com-
munity structure for habitat specialists, generalists, and all plant species. Per study site
an average of 54% of all specialist species was found, indicating that 46% of total
species richness was due to between study site variation. Nevertheless, habitat iso-
lation and landscape diversity had no effect in our study region. These results stress the
point that habitat area and increasing habitat heterogeneity are the most important
basis of plant diversity. Several populations of specialist plant species on small habitat
fragments may be prone to extinction in the near future, showing the so-called
1435
extinction debt. Sheep ock migration might contribute to a better dispersal of seeds
(Poschlod and WallisDeVries 2002), to reduce potentially increasing specialist ex-
tinction in small and isolated habitats. For conservation we suggest to protect (a) the
largest grasslands with low extinction risk of habitat specialists, and (b) a series of
habitat fragments covering a sufcient range of geographical area to maximise re-
gional diversity.
Acknowledgements
We thank David Kleijn, Olivier Honnay, Sabine Gu
¨
sewell and one anonymous
referee for helpful comments on the manuscript and Bertram Preuschoff
(Untere Naturschutzbeho
¨
rde des Landkreis Go
¨
ttingen) for his expert advice. This
work was nancially supported by the German Science Foundation (Deutsche
Forschungsgemeinschaft).
Appendix 1
Habitat specialist plants on 31 calcareous grasslands in Southern Lower Saxony.
Taxon Number
of
occupied
habitats
Taxon Number
of
occupied
habitats
Agrimonia eupatoria L. 30 Koelaeria pyramidata (Lamk.) P.B. 29
Anemone sylvestris L. 3 Medicago falcata L. 7
Anthyllis vulneraria L. 20 M. lupulina L. 30
Astragalus glycyphyllos L. 10 Melampyrum arvense L. 3
Avenula pratensis (L.) Dum. 16 M. nemorosum L. 2
A. pubescens (HUDS.) Dum. 11 Onobrychis viciifolia Scop. 7
Brachypodium pinnatum (L.) P.B. 31 Ononis spinosa L. 24
Briza media L. 30 Ophris apifera Huds. 1
Bromus erectus Huds. 19 Ophrys insectifera L. em L. 16
Campanula rapunculoides L. 24 Orchis mascula (L.) L. 13
Carex caryophyllea Latourr. 15 O. militaris L. 3
C. flacca Schreber 30 O. purpurea Huds. 5
C. ornithopoda Willd. 3 Orchis tridentata Scop. 3
Carlina vulgaris L. 25 Origanum vulgare L. 3
Centaurea scabiosa L. 28 Pimpinella saxifraga L. 31
Cerastium arvense L. 11 Platanthera bifolia (L.) L.C. Richard 2
Cirsium acaule SCOP. 31 Polygala comosa Schkuhr 27
Clinopodium vulgare L. 15 Potentilla neumanniana Rchb. 31
Euphorbia cyparissias L. 14 Primula veris L. 30
Euphrasia stricta Wolff ex Lehm 2 Prunella grandiflora (L.) Scholler 11
Festuca ovina L. 31 Ranunculus bulbosus L. 31
Filipendula vulgaris Moench 1 Salvia pratensis L. 4
Fragaria viridis (Duchesne) Weston 18 Sanguisorba minor Scop. 31
Galium pumilum Murray 22 Scabiosa columbaria L. 31
1436
References
Ba
¨
ssler M., Ja
¨
ger E.J. and Werner K. 1999. Rothmaler, Exkursionsora von Deutschland, Bd. 2,
Gefa
¨
sspanzen: Grundband. Spektrum Akademischer Verlag, Heidelberg, Berlin, Germany.
Beinlich B. and Plachter H. 1995. Ein Naturschutzkonzept fu
¨
r Kalkmagerrasen der Mittleren
Schwa
¨
bischen Alb (Baden-Wu
¨
rttemberg): Schutz, Nutzung und Entwicklung. Beihefte zu den
Vero
¨
ffentlichungen fu
¨
r Naturschutz und Landschaftspege in BadenWu
¨
rttemberg 83: 1520.
Bruun H.H. 2000. Patterns of species richness in dry grassland patches in an agricultural landscape.
Ecography 23: 641650.
De Blois S., Domon G. and Bouchard A. 2002. Landscape issues in plant ecology. Ecography 25: 244256.
Deutscher Wetterdienst 2001. Klimadaten der Wetterstation Go
¨
ttingen. In: Stadt Go
¨
ttingen, Amt fu
¨
r
Statistik und Stadtforschung (eds), Statistikjournal 1/2001.
Eriksson O. 1996. Regional dynamics of plants: a review of evidence for remnant, source-sink and
metapopulations. Oikos 77: 248258.
Fischer M. and Sto
¨
cklin J. 1997. Local extinctions of plants in remnants of extensively used calcareous
grasslands 19501985. Conservation Biology 11: 727737.
GEOsat GmbH 1998. GEOlink Version 1. 46. Wuppertal, Germany.
Gonzales A. 2000. Community relaxation in fragmented landscapes: the relation between species
richness, area and age. Ecology Letters 3: 441448.
Grashof-Bokdam C. 1997. Forest species in an agricultural landscape in the Netherlands: effects of
habitat fragmentation. Journal of Vegetation Science 8: 2128.
Hanski I. 1999. Metapopulation Ecology. Oxford University Press, Oxford, UK.
Hanski I., Kuussaari M. and Nieminen M. 1994. Metapopulation structure and migration in the buttery
Melitaea cinxia. Ecology 75: 747762.
Holt R.D., Lawton J.H., Polis G.A. and Martinez N.D. 1999. Trophic rank and speciesarea relationship.
Ecology 80: 14951504.
Honnay O., Endels P., Vereecken H. and Hermy M. 1999. The role of patch area and habitat diversity in
explaining native plant species richness in disturbed suburban forest patches in northern Belgium.
Diversity and Distribution 5: 129141.
Jonsen I.D. and Fahrig L. 1997. Response of generalist and specialist insect herbivores to landscape
spatial structure. Landscape Ecology 12: 185197.
Kelly B.J., Bastow W.J. and Mark A.F. 1989. Causes of the speciesarea relation: a study of islands in
lake Manapouri, New Zealand. Journal of Ecology 77: 10211028.
Ko
¨
chy M. and Rydin H. 1997. Biogeography of vascular plants on habitat islands, peninsulas and mainlands
in an east-central Swedish agricultural landscape. Nordic Journal of Botany 17: 215223.
Appendix 1. (continued)
Taxon Number
of
occupied
habitats
Taxon Number
of
occupied
habitats
G. verum L. 25 Sedum sexangulare L. 6
Gentianella ciliata (L.) Borkh. 10 Silene nutans L. 3
G. germanica (Willd.) Bo
¨
rner 9 Thymus pulegioides L. 31
Gymnadenia conopsea (L.) R.Br. 26 Trifolium campestre Schreber 7
Helianthemum nummularium (L) 20 T. medium L. 18
Hieracium pilosella L. 29 T. montanum L. 2
Hippocrepis comosa L. 25 Veronica teucrium L. 17
Hypericum perforatum L. 30 Vincetoxicum hirundinaria Med. 3
Inula conyzae (Griesselich) Meikle 12 Viola hirta L. 26
1437
Kohn D.D. and Walsh D.W. 1994. Plant species richness the effect of island size and habitat diversity.
Journal of Ecology 82: 367377.
Kollmann J. and Schneider B. 1999. Landscape structure and diversity of eshy-fruited species at forest
edges. Plant Ecology 144: 3748.
Krauss J., Steffan-Dewenter I. and Tscharntke T. 2003. How does landscape context contribute to effects
of habitat fragmentation on diversity and population density of butteries? Journal of Biogeography
30: 889900.
Krebs C.J. 1989. Ecological Methodology. Harper Collins Publishers, New York.
Lawesson J.E., de Blust G., Grashof C., Firbank L., Honnay O., Hermy M., Hobitz P. and Jensen L.M.
1998. Species diversity and area-relationships in Danish beech forests. Forest Ecology and Manage-
ment 106: 235245.
Metzger J.P. 2000. Tree functional group richness and landscape structure in a Brazilian tropical frag-
mented landscape. Ecological Applications 10: 11471161.
Moilanen A. and Nieminen M. 2002. Simple connectivity measures in spatial ecology. Ecology 83:
11311145.
Oostermeijer J.G.B., Vant t Veer R. and Den Nijs J.C.M. 1994. Population structure of the rare, long-
lived perennial Gentiana pneumonanthe in relation to vegetation and management in the Netherlands.
Journal of Applied Ecology 31: 428438.
Ouborg N.J. 1993. Isolation, population size and extinction: the classical and metapopulation approaches
applied to vascular plants along the Dutch Rhine-system. Oikos 66: 298308.
Poschlod P. and WallisDeVries M.F. 2002. The historical and socioeconomic perspective of calcareous
grasslands lessons from the distant and recent past. Biology Conservation 104: 361376.
Ricketts T.H. 2001. The matrix matters: effective isolation in fragmented landscapes. The American
Nationalist 158: 8799.
Riecken U., Ries U. and Ssymank A. 1994. Rote Liste der gefa
¨
hrdeten Biotoptypen der Bundesrepublik
Deutschland. Kilda-Verlag, Greven, Germany.
Rosenzweig M.L. 1995. Species diversity in space and time. Cambridge University Press, Cambridge,
UK.
So
¨
derstro
¨
m B., Svensson B., Vessby K. and Glimska
¨
r A. 2001. Plants, insects and birds in semi-natural
pastures in relation to local habitat and landscape factors. Biodiversity and Conservation 10: 18391863.
Sokal R.R. and Rohlf F. 1995. Biometry. Freeman, New York.
Statgraphics 1995. Statgraphics plus for Windows, Version 3.0. Manugistics, Inc, Rockville, Mary land.
Steffan-Dewenter I. and Tscharntke T. 2000. Buttery community structure in fragmented habitats.
Ecology Letters 3: 449456.
Steffan-Dewenter I. and Tscharntke T. 2002. Insect communities and biotic interactions on fragmented
calcareous grasslands a mini review. Biological Conservation 104: 275284.
Steffan-Dewenter I., Mu
¨
nzenberg U., Bu
¨
rger C., Thies C. and Tscharntke T. 2002. Scale-dependent
effects of landscape context on three pollinator guilds. Ecology 83: 14211432.
Tilman D., May R.M., Lehman C.L. and Nowak M.A. 1994. Habitat destruction and the extinction debt.
Nature 371: 6566.
Tscharntke T., Steffan-Dewenter I., Kruess A. and Thies C. 2002. Characteristics of insect populations on
habitat fragments: a mini review. Ecological Research 17: 229239.
Van Ruremonde R.H.A.C. and Kalkhoven J.T.R. 1991. Effects of woodlot isolation on the dispersion of
plants with eshy fruits. Journal of Vegetation Science 2: 377384.
Van Swaay C.A.M. 2002. The importance of calcareous grasslands for butteries in Europe. Biological
Conservation 104: 315318.
Veech J.A., Summerville K.S., Crist T.O. and Gering J.C. 2002. The additive partitioning of species
diversity: recent revival of an old idea. Oikos 99: 39.
Von Drachenfels O. 1994. Kartierschlu
¨
ssel fu
¨
r Biotoptypen in Niedersachsen. Naturschutz und Land-
schaftspege Niedersachsen 4: 1192.
WallisDeVries M.F., Poschlod P. and Willems J.H. 2002. Challenges for the conservation of calcareous
grasslands in northwestern Europe: integrating the requirements of ora and fauna. Biological
Conservation 104: 265273.
1438
Warren M.S., Hill J.K., Thomas J.A., Asher J., Fox R., Huntley B. et al. 2001. Rapid responses of British
butteries to opposing forces of climate and habitat change. Nature 414: 6569.
Wiens J.A. 1997. Metapopulation dynamics and landscape ecology. In: Hanski I. andGilpin M.E. (eds)
Metapopulation Biology: Ecology, Genetics, and Evolution. Academic Press, London, pp. 4362.
Zschokke S., Dolt C., Rusterholz H.-P., Oggier P., Braschler B., Thommen G.H. et al. 2000. Short-term
responses of plants and invertebrates to experimental small-scale grassland fragmentation. Oecologia
125: 559572.
1439
... In fact, the chronic stress experienced in dog breeding centers, kennels, or shelters is mostly linked to changes in hierarchical relationships and social interactions between subjects, confined environments with spatial restrictions, repeated pregnancy, and changes in animal behavior. This situation predisposes dogs to intestinal disorders and inflammation [19][20][21][22]. ...
... In fact, it has been reported in the literature that dog-kennel conditions may result in a higher susceptibility to gastrointestinal disorders compared to dogs that live with their owners. This is likely due to the chronic stress to which they are subjected, which is linked to changes in hierarchical relationships and social interactions between subjects, as well as confinement to restricted environments with repeated pregnancies and subsequent changes in animal behavior [19][20][21][22]. Because of the gut-brain axis, stress can alter the composition of the intestinal microbiota, thereby enhancing the susceptibility of the gastrointestinal tract to inflammatory stimuli, causing dysbiosis, and eliciting immunosuppressive effects [18]. ...
Article
Full-text available
Oxidative stress causes several pathological conditions in humans and animals, including gastrointestinal disorders. The aim of this study was to analyze the antioxidant capacity of three natu- ral powdered raw materials containing quercetin, bromelain, and Lentinula edodes and develop a new feed supplement for dogs using a combination of them. The total phenolic content (TPC), antioxidant activity, DPPH (2,2-diphenyl-1-picrylhydrazyl), and ABTS (2,2′-Azino-bis (3-ethylbenzothiazoline-6- sulfonic acid) diammonium salt) of the extracts, either individually or in combination, were evaluated colorimetrically. The effects of this supplement on healthy adult dogs’ nutritional, inflammatory, and stress status were evaluated. American Staffordshire Terrier adult female dogs (n = 30) were randomly assigned to a control (n = 15) or a treated (n = 15) group. The supplement was added as powder to the food of the treated dogs once daily for 28 days. There was no significant difference in the body weight and body condition scores between the initial and final phases of the experiment. At the end of our study, a significant decrease in fecal calprotectin, cortisol, indole/skatole, and N-methylhistamine and a significant increase in short-chain fatty acids were observed as compared to the control group. In conclusion, this natural feed supplement can be used to improve gastrointestinal health and psycho-physical conditions in dogs. Citation: Atuahene, D.; Costale, A.; Martello, E.; Mannelli, A.; Radice, E.; Ribaldone, D.G.; Chiofalo, B.; Stefanon, B.; Meineri, G. A Supplement with Bromelain, Lentinula edodes, and Quercetin: Antioxidant Capacity and Effects on Morphofunctional and Fecal Parameters (Calprotectin, Cortisol, and Intestinal Fermentation Products) in Kennel Dogs. Vet. Sci. 2023, 10, 486. https://doi.org/10.3390/ vetsci10080486
... The effect of habitat on species richness has been demonstrated through various indicators (Menéndez et al. 2007;Liira et al. 2008;Gao et al. 2018), including habitat area (Krauss et al. 2004). In this study, hillside forest cover was used as a proxy of habitat availability, verifying a positive linear relationship with hillside tree richness. ...
Article
Full-text available
Background The Rio de la Plata grassland region is dominated by temperate grasslands, with the scarce natural forests, influenced floristically by adjacent biogeographical provinces. Uruguay represents the southern limit for many tree species of the Paranaense Province, several of which inhabit the hillside forests. With many species shifting poleward due to climate change, we do not yet know how current environmental factors, particularly climatic ones, are linked to the tree diversity of this flora nowadays. The aim of this study is to understand the geographic pattern of tree richness in the hillside forests of Uruguay, evaluating the water–energy and the environmental heterogeneity hypotheses. The distribution of the hillside forest trees was obtained by compiling and updating the herbaria database and distribution maps of woody plants of Uruguay. The presence/absence of each species, and then the species richness, were georeferenced over a grid that covers Uruguay with 302 cells (660 km ² ). Over the same grid were compiled environmental variables associated with climate and environmental heterogeneity. The relationship between richness and environmental variables was studied by applying general linear models (GLM). As a strong autocorrelation was detected, a residuals auto-covariate term was incorporated into the GLM, to take into account the species richness spatial structure. Results The tree flora of the hillside forest was composed mainly by Paranaense species that show a latitudinal gradient, with two high richness cores, in the east and northeast of Uruguay. The final model including the environmental variables and the spatial term explained 84% of the variability of tree richness. Species richness showed a positive relationship with precipitation, forest cover, potential evapotranspiration and productivity, while a negative effect of temperature variation was found. The spatial component was the primary predictor, accounting for a 30% of spatial pattern of tree richness. Conclusions This study accounts for a large proportion of the environmental and spatial variations of the tree richness pattern of the Paranense flora in its southernmost portion. It brings support to both water–energy and environmental heterogeneity hypotheses, emphasizing the role of climate and its variation and the habitat availability on the hillside forest diversity.
... The phylogenetic diversity-area relationship is little studied within aquatic ecosystems, so we had to move to terrestrial studies (Morlon et al., 2011) to find a similar result. The phylogenetic overdispersion found associated with larger patches could suggest that actively dispersing taxa are therefore not limited by their capacity to disperse at these scales and can select larger habitats that, in turn, may provide greater habitat heterogeneity that can host more diverse phylogenies (Heino, 2000;Krauss et al., 2004;Rosenzweig, 1995). ...
Article
Full-text available
The application of graph theory to metacommunity ecology allows a deeper analysis of the effect of network structure on diversity patterns. Here, we set out to test the role of network centrality metrics and environmental characteristics in diversity patterns of pond macroinvertebrate metacommunities. We tested two approaches to construct the networks: one used the percolation distance, whereas the other was based on a community-contingent distance. The role of each patch within the network was then analyzed using its centrality value. Later, we analyzed the relationships between the macroinvertebrate diversity and centrality metrics for four study sites. The calculated diversity metrics cover different facets of biodiversity at two scales: pond and pondscape. Environmental characteristics of the studied ponds were also included. All relationships were tested considering the entire macroinvertebrate dataset, but also differentiated by dispersal mode (i.e., active vs. passive) and considering the two types of network approaches analyzed. The results were mostly consistent when comparing the network approaches used. Centrality metrics tended to be positively related to alpha and negatively to beta diversity. Environmental uniqueness showed a positive effect on beta diversity metrics, regardless of the dispersal mode. We only observed a weak negative relationship between eutrophication and species richness of active dispersers. Pond size showed a positive effect on both alpha and beta diversity, but was detected more frequently on alpha diversity metrics. We could not find evidence for a clear negative effect of habitat degradation on diversity. We found a greater importance of environmental characteristics versus the centrality metrics for both alpha and beta diversity of active dispersers, while a combination of their contributions for passive dis-persers. An unexpected importance of centrality was observed for alpha diversity of passive dispersers. Using empirical data, we demonstrate that the centrality of a patch in an undirected network affects diversity regardless of the approach used to construct the networks, with a higher influence at local scale regardless of the dispersal mode. This study broadens knowledge of the relationships
... Economic, socio-cultural and political factors, such as intensification and abandonment, have led to the disappearance of extensive grassland management systems across Europe since the mid-twentieth century (Bakker and Berendse 1999;MacDonald et al. 2000;. As a result, the habitat mosaic of the cultural landscapes has changed, with fragmentation and disappearance (scrub encroachment, afforestation) of species-rich semi-natural grasslands having occurred, while the biodiversity of these habitats has declined (Eriksson et al. 2002;Krauss et al. 2004;Flynn et al. 2009;Laliberte et al. 2010). ...
Article
Full-text available
A significant proportion of Europe’s species-rich grasslands are semi-natural habitats. They have a long history of traditional management. Several studies have been carried out to conserve them, resulting in the establishment of subsidised conservation management schemes. On the other hand, many of these conservation management schemes have failed to provide locally adaptive solutions to maintain the diversity and functional status of species-rich grasslands. In addition, few studies have compared the conservation effectiveness of different levels of management complexity. The levels of management complexity in our study are based on how different management types (e.g. grazing and mowing etc.) and how different herbage removal intensities (e.g. lower and higher grazing intensities) are combined within and between years. To investigate this, we compared the overall effects of management complexity, herbage removal intensity and management type on plant diversity, plant functional type dominance relationships and plant physiognomy. Our field sampling was carried out in the sandy meso-xeric grasslands of the Turján Region of the Great Hungarian Plain (Central Hungary). We sampled nine 2 m × 2 m plots per grassland site (n = 12), recorded all the rooted plant species and estimated their percentage cover in each plot. High level of management complexity had significant positive effects on plant diversity, grazing had positive effects on plant diversity and phanerophyte density, while the studied levels of herbage removal intensity had no effect on diversity, plant functional types or plant physiognomy. In parallel, mowing and/or low levels of management complexity had some negative effects on conservation value (e.g. lower Shannon and Simpson diversity). In this landscape, the dominance of grazing and the more complex management is more optimal than relatively homogeneous mechanical mowing. The choice of management type and intensity is an important tool in the conservation management system of this landscape, but so too is its appropriate application in space and time. Through a detailed analysis of the effects of management complexity levels compared to management types and herbage removal intensity levels, we provide a new opportunity to make grassland management practices more effective for conserving biodiversity in this region, but it would be important to investigate these in different landscapes and conditions.
... The chemical constituents differed to some extent among the variants, which was not due to the treatment but also reflected the heterogeneity of the substrate. The reason for this could be the sampling, which reflects the heterogeneity of the substrate due to the different composition of the plants in the meadow [74]. ...
Article
Full-text available
The aim of this study was to use landscape grass from species-rich orchards for biogas production, thus preserving these very valuable areas for future generations. Since these grass clippings have high lignocellulose content, the substrate has to be pretreated before being fed into the biogas digester. In this study, three different mechanical treatment processes (cross-flow grinder, ball mill and a mounted mower) were investigated and compared with untreated grass clippings. Chemical composition, specific methane yield, degradation kinetics and microscopic images were analyzed. In order to derive recommendations, the harvesting and pretreatment processes were examined in terms of energy demand, additional methane yield, and suitability of the substrate for use in biogas plants, taking into account conservation aspects. Within the pretreatment process, ball milling leads to the highest significant increase in specific methane yield of up to 5.8% and the fastest gas formation kinetics (lag time λBM: 0.01 ± 0.0 d; duration to reach half of total gas production ½M(x)BM: 5.4 ± 0.2 d) compared to the untreated variant (λUT: 1.02 ± 0.2 d; ½M(x)UT: 6.5 ± 0.2 d). A comparison of the energy required for the mechanical disintegration of the substrates with the increased yield of methane during the digestion process shows that the mechanical processing of these substrates appears to be useful. A positive energy balance was achieved for the cross-flow grinder (12.3 kWh tVS−1) and the ball mill (21.4 kWh tVS−1), while the Amazone Grasshopper left a negative balance (−18.3 kWh tVS−1), requiring more energy for substrate pretreatment than was generated as methane surplus. In summary, the pretreatment of landscape management grass is a suitable approach for utilizing agricultural residues efficiently in a biogas plant and thus contributing to sustainable energy production.
... While the evenness outweighed the diversity of landscape in affecting the richness of wetland plants. To some extent, the richness of species has increased with the area of wetlands (Krauss et al., 2004). From landscape perspective, the number and evenness were two principal factors that influenced both diversity and richness of wetland plants. ...
... SARs have been studied for almost all taxa, for example vascular plants (Krauss et al. 2004;Powell et al. 2013;Patiño et al. 2014;D'Antraccoli et al. 2019;Dengler et al. 2020) and bryophytes (Weibull & Rydin 2005;Silva et al. 2018;Yu et al. 2020), but have been poorly investigated in lichens. The positive effect of increasing area size on lichen richness has been highlighted in some papers, but with these mainly considering the area of suitable habitat. ...
Article
Full-text available
The species–area relationship (SAR) states that species richness increases with the increase of the sampled area, although other factors can influence the pattern. SARs have been tested on many different organisms, but only rarely on lichens. We aimed to test the SAR, across a wide range of area sizes, for three main substratum-related guilds of lichens, namely epiphytic, epilithic and epigaeic. The test was performed using data from lichen inventories carried out in 44 protected areas of various sizes across Italy. We found a positive correlation of species richness with area size for all three guilds, better fitted by the logarithmic function for epilithic lichens and by the power function for epiphytic and epigaeic lichens. Our results support the fundamental role of area size as the main driver for lichen diversity, suggesting that in an area-based conservation framework, larger protected areas are fundamental to support high lichen species richness. However, finer scale investigations are also required to better elucidate whether and how other environmental factors could interact with area size and modify SAR patterns. Exhaustive lichen inventories could be useful information sources to more robustly test such relationships, and therefore better inform conservation practices.
... In general, landscape multiscale analysis has been extensively studied mainly focusing on explaining patterns of species occurrence, abundance, and biological diversity (e.g., [24][25][26]). However, species are not isolated entities, and their biotic interactions regulate all ecosystem attributes from primary productivity to population dynamics and, therefore, are fundamental to understanding the organization and maintenance of biodiversity [27••]. ...
Article
Full-text available
Purpose of Review Despite the potential and increased use of multiscale spatial analysis in landscape ecology, the theoretical and empirical information available in the literature generally focuses on species habitat modeling and the effects of landscape modification on biotic interactions have not been sufficiently explored. In this study, we provide a current and comprehensive overview on where we are and where we need to go when considering spatial scaling and the complexity of biotic interactions. Recent Findings Although the accumulated knowledge on scale dependency of biotic interactions has increased in the last decades, the use of an ecological framework involving the effects of spatial scaling on species interaction networks is still too limited. Generally, the few studies available in the literature analyzed the landscape structure at a single level or are limited to a set of a few spatial extents. Additionally, the little information available in the literature is related to some mechanisms involving species–area relationships (i.e., combining local networks into metawebs to increase the spatial sampling scales from small to regional levels) and the scale of effect (i.e., the spatial scale at which landscape structure best predicts a network structural descriptor). Summary We highlight the current status and some potential research directions on spatial scaling involving the complexity of biotic interactions, which should be considered in future studies to ensure precise, robust, and accurate interpretations of the organization of species interaction networks and help us test specific ecological hypotheses, identify potential universalities, and propose effective strategies of monitoring and management for biodiversity conservation.
Article
Full-text available
Evaluating the impacts of farming systems on biodiversity is increasingly important given the need to stem biodiversity loss, decrease fossil fuel dependency, and maintain ecosystem services benefiting farmers. We recorded woody and herbaceous plant species diversity, composition, and abundance in 43 wetland‐adjacent prairie remnants beside crop fields managed using conventional, minimum tillage, organic, or perennial cover (wildlife‐friendly) land management in the Prairie Pothole Region. We used a hierarchical framework to estimate diversity at regional and local scales (gamma, alpha), and how these are related through species turnover (beta diversity). We tested the expectation that gamma richness/evenness and beta diversity of all plants would be higher in remnants adjacent to perennial cover and organic fields than in conventional and minimum tillage fields. We expected the same findings for plants providing ecosystem services (bee‐pollinated species) and disservices (introduced species). We predicted similar relative effects of land management on alpha diversity, but with the expectation that the benefits of organic farming would decrease with increasing grassland in surrounding landscapes. Gamma richness and evenness of all plants were highest for perennial cover, followed by minimum tillage, organic, and conventional sites. Bee‐pollinated species followed a similar pattern for richness, but for evenness organic farming came second, after perennial cover sites, followed by minimum tillage and conventional. For introduced species, organic sites had the highest gamma richness and evenness. Grassland amount moderated the effect of land management type on all plants and bee‐pollinated plant richness, but not as expected. The richness of organic sites increased with the amount of grassland in the surrounding landscape. Conversely, for conventional sites, richness increased as the amount of grassland in the landscape declined. Our results are consistent with the expectation that adopting wildlife‐friendly land management practices can benefit biodiversity at regional and local scales, in particular the use of perennial cover to benefit plant diversity at regional scales. At more local extents, organic farming increased plant richness, but only when sufficient grassland was available in the surrounding landscape; organic farms also had the highest beta diversity for all plants and bee‐pollinated plants. Maintaining native cover in agroecosystems, in addition to low‐intensity farming practices, could sustain plant biodiversity and facilitate important ecosystem services.
Article
Full-text available
Eighty-five patches of semi-natural grassland of varying size scattered in a agricultural landscape were investigated for their flora of vascular plants. Relationships between species richness and patch area, spatial isolation and local habitat conditions including heterogeneity were examined. Differences between single species and among groups of species defined by life-history traits were also investigated.Area was shown to be an important determinant of species richness irrespective of habitat heterogeneity. Isolation in space and habitat heterogeneity also play significant roles. These results are consistent with results from a multitude of studies on fragments of ancient deciduous woodland in northern Europe. They are, however, contradictory to results from previous studies in grasslands within the same region.Seed mass and dispersal syndrome were poor predictors of the degree to which the species were affected by isolation of grassland patches. Seed mass deviation from community median could explain a small percentage of the variation in regional abundance. Logistic regression on species occurrences showed that few species are associated with large patches, and less than half seem to avoid isolated patches.
Article
Full-text available
Most ecological processes and interactions depend on scales much larger than a single habitat, and therefore it is important to link spatial patterns and ecological processes at a landscape scale. Here, we analyzed the effects of landscape context on the distribution of bees (Hymenoptera: Apoidea) at multiple spatial scales with respect to the following hypotheses: (1) Local abundance and diversity of bees increase with increasing proportion of the surrounding seminatural habitats. (2) Solitary wild bees, bumble bees, and honey bees respond to landscape context at different spatial scales. We selected 15 landscape sectors and determined the percentage of seminatural habitats and the diversity of habitat types at eight spatial scales (radius 250-3000 m) by field inspections and analyses of vegetation maps using two Geographic Information Systems. The percentage of semi- natural habitats varied between 1.4% and 28%. In the center of each landscape sector a patch of potted flowering plants (four perennial and two annual species) was placed in the same habitat type, a grassy field margin adjacent to cereal fields. In all, 865 wild bee individuals and 467 honey bees were observed and an additional 475 individuals were caught for species identification. Species richness and abundance of solitary wild bees showed a close positive correlation with the percentage of seminatural habitats at small scales up to 750 m, whereas bumble bees and honey bees did not respond to landscape context at these scales. In contrast, honey bees were correlated with landscape context at large scales. The densities of flower-visiting honey bees even increased with decreasing proportion of seminatural habitats at a radius of 3000 m. We are not aware of any empirical studies showing contrasting foraging patterns related to landscape context at different spatial scales. We conclude (1) that local landscape destruction affects solitary wild bees more than social bees, possibly changing mutualistic plant-pollinator and competitive wild bees- honey bees interactions and (2) that only analyses of multiple spatial scales may detect the importance of the landscape context for local pollinator communities.
Article
1 The objective of this study was to explore the interrelationship between island area, number of species, and habitat diversity. 2 A survey of dicotyledonous plant species was carried out on 45 uninhabited, unimproved, small islands off Shetland Mainland, plus two similar mainland head-lands treated as islands. In addition, species were counted within 50-cm square quadrats randomly placed on island vegetation. The largest island surveyed was 100 ha; 81 plant species in all were found. 3 A total of 14 physical (abiotic) habitat types were classified. The number of habitats on each island was counted, and the habitat types characteristic of each plant species were recorded. Island areas were determined from Ordnance Survey maps. 4 There are close-fitting positive correlations between species number, island area, and the number of habitat types on an island. 5 Data on species number within quadrats of standard area reveal an increase in small-scale species richness on islands of increasing size--evidence for an effect of island area alone on species total. 6 Habitat types containing fresh water were largely absent from islands of less than one hectare in size. Species primarily associated with fresh-water habitats were generally also missing from these smaller islands--evidence for an effect of addition of habitat types on species total. 7 Path analysis confirms that island area contributes to species number both directly and indirectly, through habitat diversity, and that while the direct effects of area and habitats on species are roughly equal in magnitude, the total effect of area is nearly twice that of habitats. 8 Presence and absence of particular habitat types may be a function of island size.
Article
A data set of 143 sites along the Dutch Rhine system is compared between 1956 and 1988. The multiple regression analysis of the 1956 distribution of total species number (classical approach) revealed significant contributions of site area and environmental variance within sites (26% and 5.7% of total variation, respectively). Isolation explained an additional 4.2% of total variation and is relatively unimportant. Logistic regression analysis of population extinction between 1956-1988 (the metapopulation approach) demonstrated that for some species large populations with close neighbours had a higher chance of persistence. For other species only population size was important for extinction and yet other species were only influenced by isolation. For Eryngium campestre, Medicago falcata and possibly Plantago media both extinction and colonization were influenced by isolation suggesting that populations of these species were integrated in a metapopulation structure. This structure must be mostly of the equal sized source-recipient type. -from Author
Article
(1) The species-area relation consistently found among islands has given rise to the Random Placement hypothesis, the Habitat Diversity hypothesis and the Equilibrium hypothesis. (2) In an attempt to test the Equilibrium hypothesis, twenty-three islands in Lake Manapouri, New Zealand were sampled for richness of vascular plant species. Sampling was restricted to two vegetation types (beech forest and manuka scrub) and was with a fixed-size quadrat (100 m2}). Fixed-quadrat size sampling should eliminate the effect of island area on observed species richness if either the Random Placement or the Habitat Diversity hypothesis is correct, but should leave an effect of island area if the Equilibrium hypothesis is correct. (3) In fact, the percentage of variation in species richness explained by island area was reduced from the 92% found in an earlier study by whole-island sampling to 17% and 10%, respectively, for the two vegetation types--both representing non-significant upward trends. However, the overall manuka scrub relation showed a significant quadratic relation, and examination of species-area relations within subgroups of quadrats selected for greater habitat uniformity showed one significant linear regression. (4) Possible explanations, not normally considered in species-area studies, are the Incidence Function hypothesis, the Small Island Effect hypothesis, and the Small Island Habitat hypothesis--that there are inevitable habitat differences between small and large islands.
Article
Despite a long tradition in plant ecology of studies of patch dynamics, recent developments of models for large scale dynamics in source-sink and metapopulations have largely focused on animals. In contrast to mobile unitary animals, many plants resist extinction, even under conditions where only a part of the life cycle can be maintained. This model of remnant population dynamics adds to the two commonly recognized source-sink and metapopulations dynamics. A review of the literature suggests that all three types of dynamics are common in plants. Regional dynamics are related to life-cycle characteristics determining dispersal ability and longevity of life cycle stages. Short-lived or highly habitat specialized plants with good dispersal tend to build up metapopulations, i.e. systems of local populations and non-occupied but potentially suitable sites, interconnected with dispersal, and in a continuous flux of local colonization and extinction. At the other extreme, long-lived plants with clonal propagation, or plants with extensive seed banks, tend to build up remnant population systems, in which many local populations persist over periods long enough to bridge unfavourable phases of successional development, intervening periods of favourable conditions. Source-sink populations are a special case of metapopulations, in the sense that they comprise both persistent refuge populations and ephemeral populations maintained by dispersal. It is suggested that a concept of local population inertia in remnant population systems, scales to higher level phenomena of vegetation inertia, and to community stabilization (through enhanced recovery after perturbations). Such an inertia may contribute to explain cases of exceptionally high species diversity, and lack of pronounced mass extinctions of plants in the fossil record.
Article
In the last decade, we have seen the emergence and consolidation of a conceptual framework that recognizes the landscape as an ecological unit of interest. Plant ecologists have long emphasized landscape-scale issues, but there has been no recent attempt to define how landscape concepts are now integrated in vegetation studies. To help define common research paradigms in both landscape and plant ecology, we discuss issues related to three main landscape concepts in vegetation researches, reviewing theoretical influences and emphasizing recent developments. We first focus on environmental relationships, documenting how vegetation patterns emerge from the influence of local abiotic conditions. The landscape is the physical environment. Disturbances are then considered, with a particular attention to human-driven processes that often overrule natural dynamics. The landscape is a dynamic space. As environmental and historical processes generate heterogeneous patterns, we finally move on to stress current evidence relating spatial structure and vegetation dynamics. This relates to the concept of a landscape as a patch-corridor-matrix mosaic. Future challenges involve: 1) the capacity to evaluate the relative importance of multiple controlling processes at broad spatial scale; 2) better assessment of the real importance of the spatial configuration of landscape elements for plant species and finally; 3) the integration of natural and cultural processes and the recognition of their interdependence in relation to vegetation management issues in human landscapes.