ArticlePDF Available

Seed dispersal distances: A typology based on dispersal modes and plant traits

Authors:

Abstract

Vittoz P. and Engler R. 2007. Seed dispersal distances: a typology based on dispersal modes and plant traits. Bot. Helv. 117: 109–124. The ability of plants to disperse seeds may be critical for their survival under the current constraints of landscape fragmentation and climate change. Seed dispersal distance would therefore be an important variable to include in species distribution models. Unfortunately, data on dispersal distances are scarce, and seed dispersal models only exist for some species with particular dispersal modes. To overcome this lack of knowledge, we propose a simple approach to estimate seed dispersal distances for a whole regional flora. We reviewed literature about seed dispersal in temperate regions and compiled data for dispersal distances together with information about the dispersal mode and plant traits. Based on this information, we identified seven “dispersal types” with similar dispersal distances. For each type, upper limits for the distance within which 50% and 99% of a species’ seeds will disperse were estimated with the 80th percentile of the available values. These distances varied 5000-fold among the seven dispersal types, but generally less than 50-fold within the types. Thus, our dispersal types represented a large part of the variation in observed dispersal distances. The attribution of a dispersal type to a particular species only requires information that is already available in databases for most Central European species, i.e. dispersal vector (e.g. wind, animals), the precise mode of dispersal (e.g. dyszoochory, epizoochory), and species traits influencing the efficiency of dispersal (e.g. plant height, typical habitats). This typology could be extended to other regions and will make it possible to include seed dispersal in species distribution models.
1
Vittoz and Engler (2007) Botanica Helvetica 117: 109-124
Seed dispersal distances: a typology based on dispersal modes and plant traits
Pascal Vittoz
1
and Robin Engler
2
1
Faculty of Geosciences and Environment, University of Lausanne, Bâtiment Biophore, CH-1015
Lausanne, Switzerland; e-mail: pascal.vittoz@unil.ch (corresponding author)
2
Department of Ecology and Evolution, University of Lausanne, Bâtiment Biophore, CH-1015
Lausanne, Switzerland; e-mail: robin.engler@unil.ch
Vittoz P. and Engler R. 2007. Seed dispersal distances: a typology based on dispersal modes
and plant traits. Botanica Helvetica, 117 (2), 109–124. DOI: 10.1007/s00035-007-0797-8
Abstract
The ability of plants to disperse seeds may be critical for their survival under the current constraints
of landscape fragmentation and climate change. Seed dispersal distance would therefore be an
important variable to include in species distribution models. Unfortunately, data on dispersal
distances are scarce, and seed dispersion models only exist for some species with particular
dispersal modes. To overcome this lack of knowledge, we propose a simple approach to estimate
seed dispersal distances for a whole regional flora. We reviewed literature about seed dispersal in
temperate regions and compiled data for dispersal distances together with information about the
dispersal mode and plant traits. Based on this information, we identified seven "dispersal types"
with similar dispersal distances. For each type, upper limits for the distance within which 50% and
99% of a species' seeds will disperse were estimated with the 80
th
percentile of the available values.
These distances varied 5000-fold among the seven dispersal types, but generally less than 50-fold
within the types. Thus, our dispersal types represented a large part of the variation in observed
dispersal distances. The attribution of a dispersal type to a particular species only requires
information that is already available in databases for most Central European species, i.e. dispersal
vector (e.g. wind, animals), the precise mode of dispersal (e.g. dyszoochory, epizoochory), and
species traits influencing the efficiency of dispersal (e.g. plant height, typical habitats). This
typology could be extended to other regions and will make it possible to include seed dispersal in
species distribution models.
Key words: Anemochory, anthropochory, autochory, hydrochory, plant migration, zoochory.
Introduction
Plant dispersal has attracted scientists since long ago (Darwin 1859; Schmidt 1918; Ridley
1930; Müller-Schneider 1983) and is particularly relevant with relation to human-driven
environmental changes. For example, the survival of plant metapopulations in fragmented
landscapes strongly depends on their dispersal potential (Fischer et al. 1996; Couvreur et al. 2004;
Soons and Ozinga 2005), and the predicted global warming will require considerable migration
rates for plant species to remain under similar climatic conditions (Malcolm et al. 2002).
Nevertheless, most models attempting to predict future plant distributions did not include dispersal,
considering it as unlimited (Guisan and Theurillat 2000; Thuiller et al. 2005). Even without
constraints on seed dispersal, these models already predict local extinctions, e.g. for isolated
populations in mountains (Guisan and Theurillat 2000; Dirnbock et al. 2003; Thuiller et al. 2005).
The actual extinction rates might be even higher if plant species cannot keep pace with rapid climate
change due to limited dispersal. A more precise assessment of plant species extinction risk thus
2
Vittoz and Engler (2007) Botanica Helvetica 117: 109-124
calls for the incorporation of plant dispersal potential (Pitelka et al. 1997; Davis et al. 1998; Ronce
2001).
Many studies have measured or estimated dispersal distances of plants in the field (Schneider
1935; Stöcklin and Bäumler 1996; Jongejans and Telenius 2001), and several mathematical models
have been developed to estimate these distances (Tackenberg et al. 2003; Mouissie et al. 2005a;
Nathan et al. 2005; Soons and Ozinga 2005). However, all of these studies have considered only a
limited number of species or dispersal vectors. No dispersal distance data exist for a complete
regional flora. Müller-Schneider (1986) reviewed dispersal vectors for the entire flora of
Graubünden (East of Switzerland), but his work includes only few dispersal distances, most of
which stem from anecdotal observations. Likewise, Bonn and Poschlod (1998) and Bonn (2004)
wrote important syntheses on seed dispersal in Central Europe, but dispersal distances were only
provided for a few dispersal vectors, mainly from anecdotal observations. It is thus currently
impossible to conduct an assessment of the extinction risk of plant species under landscape
fragmentation or global warming that would take dispersal into account.
The distance over which plants disperse seeds depends on plant traits as well as environmental
conditions and varies strongly in time and space. This variability can be represented by a dispersal
curve (dispersal kernel), which gives the proportion of seeds reaching a given distance (Mouissie et
al. 2005a). However, it would be highly time consuming, if not impossible, to determine dispersal
kernels for each species of a region. Thus, a simplified approach is needed to estimate dispersal
distances for a whole regional flora. For example, if dispersal curves could be classified into a
limited number of types with similar dispersal distances, and if plant species could be attributed to
these "dispersal types" based on generally available plant traits, it would be possible to estimate
dispersal kernels for all of them.
In this paper, we develop such an approach for the Swiss flora based on an extensive review of
seed dispersal literature. We propose a typology of dispersal curves that can be applied to most
Swiss and Central European plants. This typology could be extended to other regions and could be
used to account for dispersal distances in species distribution models, enabling refined extinction
risk assessments to be made for large numbers of species.
Methods
Plant dispersal is generally achieved through seeds. These can be enclosed in fruits or larger
structures (usually called "diaspores"), but for the sake of simplification, the term “seed” will be
used here as a general denomination.
Data for seed dispersal distances were compiled by reviewing a large proportion of available
literature from Switzerland and other European countries, including monographies (Müller-
Schneider 1983, 1986), reviews and research articles. Swiss species or close relatives were
considered first priority, since our aim was to develop a typology for this region. However, other
species were included when data available for Swiss species were insufficient to assess dispersal
distances for certain dispersal modes (see below). The complete data set (ca. 300 values) is
presented in Appendix 1. Species nomenclature follows Aeschimann et al. (1996) for the Swiss
species.
The data set proved to be very heterogeneous. A small proportion of the distances had been
determined through experiments, detailed field observations of seed or seedling distributions, or
mathematical models. In such cases, it is often possible to calculate a dispersal kernel. However,
most of the available data represent isolated and often anecdotal observations, from which a precise
dispersal kernel cannot be derived. Some of these isolated observations clearly represented long-
distance dispersal events (LDD), i.e. extreme values reached only by a very small minority of seeds.
We therefore classified the data into three categories: (1) mean, mode or median values, (2)
maximum values (99th percentiles of distribution kernels) and (3) values for LDD (clearly above the
potential dispersal of 99% of the seeds). LDD values were excluded from the further analysis of the
data.
3
Vittoz and Engler (2007) Botanica Helvetica 117: 109-124
Our typology of dispersal curves was based on the dispersal modes recognised by Müller-
Schneider (1983). The English translation of Müller-Schneider's German terminology generally
follows Bonn et al. (2000). Müller-Schneider's (1983) classification of dispersal modes is primarily
based on the dispersal vector (wind, water, animals, etc.), with additional subdivisions for the
differing ways in which seeds are released and transported (e.g. on the fur or after ingestion by
animals). Additional subdivisions were made for dispersal modes whose efficiency clearly depends
on supplementary factors: plant height, pappus efficiency and environing vegetation structure for
anemochory, and vector size for zoochory. Of the numerous possible subdivisions, only those
considered most relevant were retained for our classification, as explained in the next section. This
yielded a total of 21 refined dispersal modes (Tab. 1).
Tab. 1. Dispersal distances for seven dispersal types, estimated as the upper limits of the distances within which 50%
and 99% of the seeds of a plant population are dispersed. Note that actual dispersal distances will usually be lower than
those given here (cf. Fig. 1). The dispersal distances were estimated from the 80
th
percentile of the data compiled in Fig.
1 as well as additional qualitative information as explained in the text ('Dispersal modes and evaluation of published
dispersal distances'). The dispersal modes included in each dispersal type are indicated; they are based on dispersal
vectors (categories in parentheses) and plants traits that influence the efficiency of dispersal.
Type Corresponding dispersal modes
50% 99%
1 0.1 1 Blastochory (autochory)
Boleochory (anemochory) for species < 30 cm
Ombrochory (hydrochory)
2 1 5 Ballochory (autochory)
Cystometeorochory (anemochory)
Chamaechory (anemochory) for fruits in grassland
Boleochory (anemochory) for species > 30 cm
3 2 15 Pterometeorochory (anemochory) for herbs
Myrmecochory (zoochory)
Cystometeorochory (anemochory) ferns, Orchidaceae, Pyrolaceae, Orobanchaceae in forest
Trichometeorochory (anemochory) in forest or little efficient plumes
Epizoochory (zoochory) for small mammals
4 40 150 Chamaechory (anemochory) for seeds on snow or dry inflorescence
Pterometeorochory (anemochory) for trees
Dyszoochory (zoochory) for seeds not stocked and dispersed by small animals
5 10 500 Trichometeorochory (anemochory) in openland with efficient plumes
Cystometeorochory (anemochory) ferns, Orchidaceae, Pyrolaceae, Orobanchaceae in openland
6 400 1500 Dyszoochory (zoochory) for seeds stocked by large animals
Endozoochory (zoochory) for seeds eaten by birds and large vertebrates
Epizoochory (zoochory) by large mammals
7 500 5000 Agochory (anthropochory)
Dispersal distances [m]
Each dispersal distance in our data set was attributed to a dispersal mode, which was either the
mode for which the distance had been determined (if mentioned in the original study) or the main
dispersal mode of the species according to Müller-Schneider (1986). For species with more than one
dispersal mode, distances that could not be clearly related to one of the modes were excluded from
further data analysis. Dispersal types were then defined by grouping together dispersal modes with
similar dispersal distances. This was done graphically by plotting the mean and maximal distances
for each dispersal mode and identifying modes for which distances were in the same order of
magnitude (Fig. 1).
Finally, we estimated upper limits of the distances, within which 50% and 99% of the seeds
would disperse, by using the 80
th
percentiles of the available mean, mode or median values and of
the maximum values. Results were rounded to one significant digit to reflect their approximate
nature. Our aim was to provide a conservative estimate of the dispersal constraint experienced by
most species belonging to a dispersal type. Therefore we did not take the average of the published
values (Fig. 1), but rather the 80th percentile of the distribution, as this allowed us to exclude the
most extreme values. In some cases, a comparison of the results with qualitative information from
the literature or with the authors' experience indicated that the available data were not quite
4
Vittoz and Engler (2007) Botanica Helvetica 117: 109-124
representative for a certain dispersal type; values were then adjusted to obtain more realistic
estimates. Such decisions are explained in the next section of the text for the individual dispersal
modes.
Fig. 1. Distribution of the dispersal distances found in literature (Appendix 1) for each dispersal mode,
and subdivision of the data set into seven dispersal types. Diamonds are for mean, median or mode
values, and crosses for 99% or maximum values (without long-distance dispersal). Four retained
maximum values of type 5 are outside of the graph: 1714 m, 2112 m, 2194 m and 3673 m. See Table 1
for definitions of the dispersal modes.
Dispersal modes and evaluation of published dispersal distances
Autochory
Autochorous plants disperse seeds without the help of an external vector. As a result, dispersal is
limited to very short distances.
In blastochory, the stem of the plant grows or crawls on the ground to deposit the seeds as far as
possible from the mother plant (e.g. Cymbalaria muralis, Polygonum aviculare, Veronica
hederifolia; Müller-Schneider 1983). No data were found in the literature, but since the dispersal
5
Vittoz and Engler (2007) Botanica Helvetica 117: 109-124
distance corresponds to the length of the stem, although species-specific, it is mostly very short and
blastochory can hence be classified as type 1 (Tab. 1). This dispersal mode is, however, frequently
completed by another one (Müller-Schneider 1986).
In ballochory, the explosion of the fruit ejects the seeds (ballistichory, ballistic dispersal). This
explosion may be due to the turgescence of tissues (Impatiens sp., Cardamine sp.) or the tension
between cells or different cell layers when the fruit is drying (Viola sp., Vicia sp., Lotus sp.).
Published values are scattered and very variable (maximum 0.89-6.2 m; Fig. 1). Ballochory is
classified in dispersal type 2 (Tab. 1).
Two further dispersal modes are barochory (seeds fall from the plant) and herpochory (seeds
creep on the soil by the movement of organs in a succession of dry and wet conditions). However,
since these strategies are not very efficient and always combined with other dispersal modes
(Müller-Schneider 1986), they were not retained here.
Anemochory
Anemochorous seeds are dispersed by wind, often with the help of specific organs. This
dispersal vector is the most studied as it is easily observable and measurable, at least over short
distances (e.g. Bullock and Clarke 2000; Jongejans and Telenius 2001). Moreover, it relies on
physical processes that can be translated into models (Tackenberg et al. 2003; Nathan et al. 2005;
Soons and Ozinga 2005). Anemochory is subdivided according to the organs used to slow down the
falling of seeds.
An air filled structure lightens small seeds in cystometeorochory (balloon-like). This dispersal
mode is little studied. Maximum calculated distances are below 2 m (Soons and Ozinga 2005), but
extreme values were measured up to 80 m for Calluna vulgaris (Bullock and Clarke 2000). This
mode is certainly less efficient in forests, as wind is weaker, but it seemed useless to subdivide these
already low values and thus cystometeorochory as a whole was attributed to type 2.
The tiny seeds of Orchidaceae, Pyrolaceae and Orobanchaceae also have a low falling velocity
(0.2-0.31 m/s for Orchidaceae; Müller-Schneider 1986). But only a calculated dispersion distance is
available (median 0.95 m and 99-percentile 14.7 m for Cephalanthera damasonium, Soons and
Ozinga 2005). However, because it is thought that very light seeds (<0.05 mg), even without
corresponding adaptation for anemochory, are as efficient in wind as plumed seeds (Bonn and
Poschlod 1998; Greene and Calogeropoulos 2002), we decided to classify these plant families with
trichometeorochory in type 5 in openland but decreased to type 3 for forest species (Tab. 1). Fern
spores can be included in cystometeorochory as well, but no data exist on their dispersal capacity
except a calculated distance of 330 km for Lycopodium sp. based on its very low falling velocity
(1.8 cm/s; Schmidt 1918). This value seems exaggerated and in the absence of a more precise value,
we attributed the ferns to the same types as orchids.
Plumed seeds are more efficient for wind dispersal. In trichometeorochory, seeds are
completed with a hairy structure (e.g. pappus) to reduce falling velocity. These organs have very
variable efficiency, however, with falling velocity varying from 8 cm/s for Epilobium angustifolium
to 165 cm/s for Pulsatilla alpina (Müller-Schneider 1986). With an arbitrary separation at 30 cm/s,
on the basis of our own observations, we distinguished species with less efficient plumes from those
with efficient plumes (long plumes for small seeds). The first group has maximum distances
between 1-15.7 (36) m, corresponding to type 3, and the second mainly between 20 and 179 m (Fig.
1). However, because some species have much higher calculated potentiality (e.g. up to 3600 m for
Typha latifolia; Soons and Ozinga 2005), we retained intermediate values and assigned
trichometeorochory to type 5. Forest species were classified with trichometeorochory for less
efficient plumes (type 3).
In pterometeorochory (or pterochory), seed dispersal is improved through wings. Trees are
frequent in this category, but herbs are present as well, with a generally higher falling velocity.
Because tree seeds are often large and easy to find, many available maximum dispersal distances are
to be classified as LDD (e.g. Müller-Schneider 1983). Reviewed maximum distances ranged mainly
between 80-314 m for trees and 1-12 m for herbs (Fig. 1). Pterometeorochory was thus classified as
dispersal type 3 for herbs and type 4 for trees (Tab. 1).
6
Vittoz and Engler (2007) Botanica Helvetica 117: 109-124
A much less studied dispersal mode is chamaechory, with diaspores rolling on the ground
pushed by the wind. This diaspore can be either a circular-shaped fruit (Colutea arborescens,
Astragalus alpinus), the fruit with calyx (Anthyllis vulneraria) or the complete, dry, inflorescence
(synaptospermie of Eryngium campestre, Carlina acaulis). Chamaechory is especially common and
efficient in steppes where nothing hampers dispersal (Müller-Schneider 1983), but it also occurs in
mountains, with small seeds on snow (e.g. Saxifraga bryoides, S. exarata, Sempervivum
montanum). The only available data are from Greene and Johnson (1997), who observed Betula
alleghaniensis seeds and calculated a possible dispersion of 38 m for spherical 1mg-seeds on snow.
Dispersal is usually restricted because seeds get stuck in irregularities. For chamaechory, we
retained dispersal distance type 2 for fruits in grassland and type 3 for seeds on snow or carried by
dry inflorescences (Tab. 1), but supplementary data would be necessary to get more precise values.
Boleochory (semachory) is another mode used by anemochorous plants. The small seeds
without particular features are spread when the fruit is shaken by wind. At maturity, the stem of
such plants is often rigid but elastic and sways in the wind, acting like a catapult. As animals or
others may shake the capsules as well, some classify this mode independently (semachory; Bonn et
al. 2000). Although small, the seeds are dense and have a high falling velocity (1.2-5 m/s; Müller-
Schneider 1983; Tackenberg 2001). Consequently, Soons and Ozinga (2005) calculated very short
dispersal distances, generally <0.5 m, but without considering the catapult effect. Yet, this effect is
certainly important, as measured distances sometimes exceed 10 m and are always higher than
calculations for the same or close species. Since the catapult effect strongly depends on the stem
size, we distinguished small species (<30 cm) whose seeds rarely go beyond 1 m (type 1) from taller
species (>30 cm), whose seeds may reach up to 3-5 m (type 2).
Hydrochory
Water can disperse seeds in various ways. In wetland plants, seeds are often light enough to
float and move on rivers, lakes or ponds (nautochory of Alisma plantago-aquatica, Carex flava, C.
elata, Iris pseudacorus, Sparganium sp.). Some seeds can float and survive for one year or more
(Müller-Schneider 1983). Similarly, running water may carry many different types of seeds with
heavy rains (bythisochory), sometimes to rivers and down to lowland areas. Bythisochory is
complementary to other dispersal modes and randomly affects many different species dwelling on
slopes. It is through this vector that high mountain species are frequently observed on gravel areas
along rivers (Bill et al. 1999). Although the dispersal distances may be important, we did not
attribute dispersal types to hydrochorous dispersal modes because distances are highly unpredictable
and never documented. Moreover, nautochory is geographically limited and the bythisochory
downslope restricted.
Rain may contribute to disperse seeds through the shock generated by the rain droplets hitting
the fruits (ombrochory). Some species (e.g. Caltha palustris, Veronica serpyllifolia, Prunella
vulgaris, Thlaspi perfoliatum) have developed fruit shapes and elastic fruitstalks in order to use this
energy to eject seeds. Very few measurements are published for ombrochory, but they are all below
or close to 1 m (type 1).
Zoochory
Animals are frequent and efficient vectors of dispersal, either voluntary when foraging or
involuntary when carrying seeds on their fur or in their guts. Even though zoochory has often been
observed and studied, estimating dispersal distances nevertheless remains difficult, as they highly
depend on the disperser's behaviour. Zoochory can be split into four subcategories.
Many seeds are foraged as food by animals, which sometimes hide them as stock for the winter
and forget about them, or lose them during transport (dyszoochory or dysochory). Vectors are
mainly rodents or birds, and the dispersal distance is thus strongly dependent on the vector size.
Small rodents, like voles or mice (Clethrionomys sp., Microtus sp., Apodemus sp.), generally
disperse seeds less than 30 m (Cain et al. 1998; Xiao et al. 2004), and squirrels (Sciurus sp.) a little
farther. In most cases, small birds disperse seeds by chance when feeding, for example when tits or
woodpeckers are looking for a convenient place to break a nut. The rare available data do not
7
Vittoz and Engler (2007) Botanica Helvetica 117: 109-124
exceed 60 m. However, some larger species are more efficient dispersers by hiding fruits for winter
stocks. The most famous examples are the nutcracker (Nucifraga carcyocatactes; Müller-Schneider
1986; Mattes 1992) and the jays (Garrulus glandarius; Müller-Schneider 1949; Kollmann and
Schill 1996). The literature contains different data, but those are unfortunately too often extreme
values (Mattes 1992), and most of the seeds are probably hidden within a few hundred meters. We
thus retained type 4 when the vector of dyszoochory is a small animal and type 6 when seeds are
stocked by a large animal (Tab. 1).
A particular case of dyszoochory is myrmecochory, or dispersal by ants. Generally interested
by the elaiosome, a fatty appendix of the seeds, ants transport the seeds before eating the elaisome
but leaving the rest of seed untouched and still able to germinate. Seeds may be used as building
material for their nest as well, without loosing their germination potential (Müller-Schneider 1963;
Cherix 1981). This dispersal mode has been extensively studied in the world, but only rarely are
distances available for European plants, and they rarely exceed 10 m. Some exceptional
observations nevertheless give values up to 70 m (Müller-Schneider 1983; Bonn and Poschlod
1998), and myrmecochory was hence classified as type 3.
Animals are important dispersal vectors when eating fruit or even the complete plant (Janzen
1984), and seeds go undamaged through their gut (endozoochory). Many authors have studied the
survival of the seeds through vertebrate guts and the importance of this vector (see Janzen 1984;
Pakeman 2001). As the consumer can be anything from a worm to a snail, mammal or bird,
dispersal distance is very dependent on its size and mobility. No data exist for small animals and
they are scarce and mostly anecdotal for larger ones such as birds or foxes. Models based on seed-
retention time is a possibility for getting dispersal distance estimates, but they are still rare (e.g.
Hickey et al. 1999; Vellend et al. 2003), and seed-retention time depends on seed and animal
species (Bonn 2004; Mouissie et al. 2005b). Moreover, these models usually calculate linear
distance, but animals generally live in a limited territory and do not move linearly. We chose type 6
to translate potential dispersal by large mammals or birds (Tab. 1).
Seeds are also frequently transported by animals in fur (epizoochory). This is partly the result
of specific structures, with seeds or fruits bearing hooks or glandulous hairs (Galium aparine,
Arctium sp., Saxifraga tridactylites,...) but seeds without an appendix can attach to fur as well (e.g.
Fischer et al. 1996; Mouissie et al. 2005a; Römermann et al. 2005). Observations in natural
conditions are rare and most of the data are from retention time measurements with sheep, cattle or
dummies (e.g. Fischer et al. 1996; Mouissie et al. 2005a). Although small rodents may disperse
seeds as well (Kiviniemi and Telenius 1998), the most efficient epizoochory is obtained with taller
animals. The maximum distance calculated by models based on seed retention time in fur is
between 435-1242 m, but can be longer with sheep whose long and curled wool is particularly
efficient at retaining seeds (Fischer et al. 1996; Mouissie et al. 2005a). The habitual dispersal
distance is thus estimated as type 6 for epizoochory with large animals, but much longer distances
can occasionally be achieved by sheep during transhumance (Fischer et al. 1996).
Anthropochory
Seed dispersal by humans certainly always occurred, but it strongly increased during the last
centuries, and became particularly important a few decades ago with the market globalisation and
the intercontinental transport of goods (e.g. Hodkinson and Thompson 1997; Tinner and
Schumacher 2004).
Müller-Schneider (1983, 1986) distinguished three modes of anthropochory: plants or seeds
being sold for agriculture and gardening (ethelochory), seeds being involuntarily mixed with the
previous ones (speirochory), or seeds travelling hidden in goods, cars, soil under soles, with hay,
etc. (agochory). All three means can potentially lead to very long dispersal distances and are, for
example, responsible for the advent of neophytes in Switzerland and Europe. But while ethelochory
and speirochory mostly concern urban and cultivated areas, agochory is probably more important in
natural or semi-natural ecosystems. Seed dispersal distance through anthropochory is strongly
dependent on the type of human activity but, in general, agricultural activities are the most
susceptible to spreading seeds in semi-natural ecosystems due to movements between fields or
8
Vittoz and Engler (2007) Botanica Helvetica 117: 109-124
meadows (McCanny and Cavers 1988). We can thus limit most of the dispersal distance to the
approximate size of a farming property (type 7).
Dispersal types and estimated dispersal distances
Despite the heterogeneous origin of the data compiled here, dispersal distances for individual
dispersal modes proved to be rather consistent, mostly belonging to the same order of magnitude.
Across the entire data set, maximal dispersal distances ranged between 0.09 and 6300 m (LDD
excluded), corresponding to a factor of 70'000 between the highest and lowest value. After
classification into dispersal types, this variation was reduced to a factor of 10 for type 1, 40 for type
2, 70 for type 3, 20 for type 4, 1700 for type 5, 200 for type 6, and 1 for type 7 (very few data). This
variability within types may still seem considerable, but it is small compared to the 5000-fold
difference in dispersal distances between types 1 and 7. Furthermore, the high value for type 5
(trichometeorochory with efficient plumes) reflects the high variability of pappus efficiency in this
category and the high variability found within species (e.g. Taraxacum officinale). The typology
presented here thus expresses a large part of the variation in seed dispersal distances. Accordingly,
attributing species to dispersal types makes it possible to describe interspecific variation in dispersal
capacity.
The estimated distances in Table 1 do of course not represent the dispersal kernel of one single
plant, nor even the mean pluri-annual dispersal kernel of a particular plant population. They were
estimated as the upper limits (80
th
percentile) of the dispersal distance values (Fig. 1), meaning that
they represent the dispersal potential of the plant species grouped into a dispersal type. Most plant
populations will disperse over smaller distances than those indicated in Table 1, but data and
models indicate that they could potentially disperse 50% or 99% of their seeds inside the retained
distances. Estimating upper limits to dispersal, rather than average distances, is justified when
dispersal is included as a possible constraint to species survival in predictive models of species
distributions. In this case, upper dispersal limits yield a constraint that holds for all species of a
certain dispersal type. This ensures that dispersal constraints will not be overestimated.
Alternative dispersal modes
Multiple dispersal vectors
About 40% of the species considered by Müller-Schneider (1986) have two or more dispersal
modes. The species can either use them alternatively depending on the available vector (Picea abies
is anemochorous or dyszoochorous with the red squirrel or some birds) or on its phenology (Urtica
dioica is anemochorous and avoided by animals when green but grazed and endozoochorous once
dry), or it can rely on them successively to improve dispersal (Leucojum vernum is firstly
blastochorous and lately myrmecochorous; Müller-Schneider 1983).
If the most obvious dispersal mode can often be inferred from the seed or fruit morphology,
finding out what the alternative dispersal modes of a species are generally requires precise
observations. For example, Campanula rotundifolia and Primula elatior are considered
endozoochorous by Müller-Schneider (1986) but not Campanula scheuchzeri or Primula veris,
which are only described as boleochor species. This difference, probably incorrect, strongly affects
their dispersal potential, as endozoochory is much more efficient than boleochory (Tab. 1), and
shows the gaps in our attainments.
Recent results showed that this problem appears with other dispersal modes too. Tackenberg et
al. (2003) modelled wind dispersion of seeds on the basis of their falling velocity and release height.
They concluded that some species normally not considered as anemochorous could be as efficient as
species traditionally thought-of as wind dispersed. Another example is given by Higgins et al.
(2003), who demonstrated that a 7.8 g Carya glabra nut is able to disperse 647 m if uplifted by
strong winds. Similarly, epizoochory concerns more species than what diaspore morphology
9
Vittoz and Engler (2007) Botanica Helvetica 117: 109-124
indicates, and many plumed seeds for anemochory or smooth seeds are transported as well (Fischer
et al. 1996; Couvreur et al. 2004; Mouissie et al. 2005a; Römermann et al. 2005).
When multiple vectors are recognized, it is logical to classify the species into the dispersal
distance type corresponding to the most efficient one (e.g. dyszoochory for Picea abies or
endzoochory for Campanula rotundifolia). But this can not consider the unsuspected supplementary
vectors.
Long-distance dispersal (LDD) and Reid's paradox
The inadequacy between the dispersal potential of plants and their post-glacial recolonisation,
also known as “Reid’s paradox” (Clark et al. 1998), is an issue that has been recognized for a long
time (Reid 1899; Skellam 1951; Cain et al. 1998). Lang (1994) calculated the migration rate of
anemochorous trees through Europe and found per-generation travel distances of 0.5-5 km for Tilia
sp, 1.2-9 km for Abies alba, 10-20 km for Acer sp. or 15-60 km for Pinus sylvestris. This is much
higher than the 200 m considered in table 1 for 99
th
percentile. Similarly, dyszoochorous species
with an estimated potential dispersal of 1 km (Tab. 1) showed post-glacial colonisation rate of 2.2-
15 km per generation for Quercus sp. or 7-14 km for Fagus sylvatica (Lang 1994). However, as was
recently found for Fagus sylvatica (Magri et al. 2006), it is possible that those recolonisation rates
are overestimated because some glacial refugia remain yet unknown (Clark et al. 1998; Stewart and
Lister 2001; Pearson 2006).
Recent data for invasive species show similar high rates of spread for many species. Pyšek and
Hulme (2005) listed 16 species with colonisation superior to 1 km/y for long-distance dispersal,
with a maximum of 167 km yr
-1
. They showed that the rate of spread may be similarly high for
wind, water or animal dispersed plants. But the landscape structure and human activity influence
this spreading, with higher rates found in densely inhabited or particularly economically active
regions (Williamson et al. 2005).
A solution to resolve this discrepancy between estimated dispersal distances and observed
migration rates is to consider that dispersal vectors indicated by seed morphology mainly explain
the short dispersal distances, with the rare events responsible for LDD relying on other vectors
(Cain et al. 1998; Higgins et al. 2003). For example, 78% of the plants that arrived on Surtsey island
(Iceland) were transported by water when only one quarter of those taxa were morphologically
adapted for water dispersal (Higgins et al. 2003). Birds can transport seeds in mud sticking to their
feet (Carlquist 1967), ingest some anemochorous seeds (Wilkinson 1997) or use them to build their
nest (Salix sp. or Clematis vitalba; Müller-Schneider 1983; Dean et al. 1990). Seed plumes or
pappus are not only very efficient for wind dispersal (anemochory), but also for fixing on animal fur
(Fischer et al. 1996; Couvreur et al. 2004). Finally, humans also are efficient involuntary dispersal
vectors nowadays (e.g. Hodkinson and Thompson 1997), but were also vectors during the post-
glacial recolonisation, like for Corylus avellana or agricultural weeds (Braun-Blanquet 1970; Lang
1994; Clark et al. 1998).
Taking into account the influence of LDD on plant migration in a better manner can possibly be
achieved by improving the models fitted on dispersal observations (Kot et al. 1996; Clark et al.
1998; Higgins and Richardson 1999). These improved models would help to propose dispersal
distance values for the remaining 1% of the seeds (Tab. 1). But up to now, the necessary values for
this improvement are missing for most species and dispersal modes, and hence we cannot propose
realistic values for our dispersal distance types. Yet, even though this improvement in LDD
estimation would be achieved, it could explain only a part of LDD, as the randomness of
unconventional dispersal vectors cannot be standardised for all species. The importance of these
accidental dispersions is not known in nature. It may be an important factor for colonising large,
new areas (Higgins et al. 2003) or disturbed areas (Bergelson et al. 1993; Williamson et al. 2005),
but it is certainly less frequent in closed, natural vegetation. Takahashi and Kamitani (2004)
observed the colonisation of native herbaceous species in an artificial pine forest. They found that
the distances dispersed by species using various dispersal vectors were similar to what we proposed
for our dispersal types (Tab. 1), thus indicating that the unconventional dispersal vectors were
certainly not predominant.
10
Vittoz and Engler (2007) Botanica Helvetica 117: 109-124
Conclusions
Although the data compiled in this paper are certainly incomplete, they are the most
comprehensive data set currently available for the Central European flora. Our method for
estimating dispersal distances based on dispersal types is less precise than the calculation of species-
specific dispersal models. On the other hand, our typology can be applied to almost all European
plant species, which is not the case of a species specific model. As discussed above, our typology is
able to represent a large fraction of the interspecies variation in dispersal distances as long as long-
distance dispersal is ignored.
Future research on dispersal mechanisms as well as the inclusion of our estimates in species
distribution models will show whether the use of this typology leads to predicted migration rates
that are close to the observed ones. If differences prove to be important for some of the dispersal
types, our typology could be improved by adjusting the corresponding dispersal distances.
Alternatively, if observed migration rates are consistently underestimated by the use of our
typology, this would suggest that long-distance dispersal is much more important for long-term
plant displacements than the dispersal modes presented here. Our typology is therefore certainly not
the final one, but an important basis for improving predictive models of species distributions.
Résumé
La capacité des plantes à disperser est un facteur important à leur survie dans un paysage
fragmenté ou sous l'influence des changements climatiques. Il est donc important de pouvoir tenir
compte des distances de dispersion dans les modèles de répartition des espèces, mais les valeurs
existantes, mesurées ou calculées, sont rares. Nous proposons donc une approche simple permettant
d'estimer ces distances pour l'ensemble d'une flore régionale. Nous avons recherché dans la
littérature les données disponibles pour la flore des régions tempérées (avant tout pour les espèces
suisses) et associé les distances de dispersion trouvées avec le mode de dispersion et des traits
biologiques. Sept types de dispersion ont pu être identifiés sur la base de ces informations, chaque
type regroupant des espèces avec des distances de dispersion proches. Les distances à l'intérieur
desquelles 50 % et 99 % des graines sont dispersées ont été estimées sur la base du 80
e
percentile
des valeurs disponibles au sein de chaque type. Ces distances varient d'un facteur 5000 entre les sept
types de dispersion, alors que les valeurs à disposition pour chaque type ne dépassent généralement
pas un facteur de 50. Nos types de dispersion conservent donc une large part de la variation
existante dans la dispersion des graines. L'attribution d'une espèce à un type de dispersion ne
nécessite que des informations couramment disponibles, comme le vecteur de dispersion (vent,
animaux, …), le mode précis de la dispersion (dyszoochorie, épizoochorie, …) et des traits
biologiques influençant la dispersion (hauteur de la plante, habitat, …). Cette typologie pourrait être
étendue à d'autres régions et permet d'inclure la dispersion des graines dans les modèles de
répartition des espèces.
This research has been partly supported by the Federal Office for the Environment. We are grateful to Christophe
Randin and Antoine Guisan for useful advice and discussion. We thank S. Güsewell, J. Kollmann and an anonymous
reviewer for their useful comments on an earlier version of the manuscript.
Bibliographie
Aeschimann D., Heitz C., Palese R., Perret P. et Moser D.M. 1996. Index synonymique de la Flore de Suisse et
territoires limitrophes (ISFS). CRSF, Genève.
Bergelson J., Newman J.A. and Floresroux M.E. 1993. Rates of weed spread in spatially heterogeneous environments.
Ecology 74: 999-1011.
Bill H.-C., Poschlod P., Reich M. and Plachter H. 1999. Experiments and observations on seed dispersal by running
water in Alpine floodplain. Bull. Geobot. Inst. ETH 65: 13-28.
11
Vittoz and Engler (2007) Botanica Helvetica 117: 109-124
Bonn S. 2004. Dispersal of plants in the Central European landscapes – dispersal processes and assessment of dispersal
potential exemplified for endozoochory. PhD thesis, Universität Regensburg.
Bonn S. und Poschlod P. 1998. Ausbreitungsbiologie der Pflanzen Mitteleuropas. Quelle and Meyer, UTB, Wiesbaden.
Bonn S., Poschlod P. and Tackenberg O. 2000. "Diasporus" a database for diaspore dispersal. Concept and
applications in case studies for risk assessment. Z. Ökologie u. Naturschutz 9: 85-97.
Braun-Blanquet J. 1970. Associations messicoles du Languedoc. Leur origine, leur âge. Melhoramento 22: 55-75.
Bullock J.M. and Clarke R.T. 2000. Long distance seed dispersal by wind: measuring and modelling the tail of the
curve. Oecologia 124: 506-521.
Cain M.L., Damman H. and Muir A. 1998. Seed dispersal and the Holocene migration of woodland herbs. Ecol.
Monogr. 68: 325-347.
Carlquist S. 1967. The biota of long distance dispersal. V. Plant dispersal to Pacific islands. Bull. Torrey Bot. Club 94:
129-162.
Cherix D. 1981. Contribution à la biologie et à l’écologie de Formica lugubris Zett. (Hymenoptera, Formicidae). Le
problème des super-colonies. Thèse de doctorat, Université de Lausanne.
Clark J.S., Fastie C., Hurtt G., Jackson S.T., Johnson C., King G.A., Lewis M., Lynch J., Pacala S., Prentice C., Schupp
E.W., Webb T. and Wyckoff P. 1998. Reid's paradox of rapid plant migration (dispersal theory and interpretation
of paleoecological records). Bioscience 48:13-24.
Couvreur M., Christiaen B., Verheyen K. and Hermy M. 2004. Large herbivores as mobile links between isolated nature
reserves through adhesive seed dispersal. Appl. Veg. Sci. 7: 229-236.
Darwin C.R. 1859. The origin of species. John Murray, London.
Davis A.J., Jenkinson L.S., Lawton J.H., Shorrocks B. and Wood S. 1998. Making mistakes when predicting shifts in
species range in response to global warming. Nature 391: 783-786.
Dean W.R.J., Milton S.J. and Siegfried W.R. 1990. Dispersal of seeds as nest material by birds in semiarid karoo
shrubland. Ecology 71: 1299-1306.
Dirnbock T., Dullinger S. and Grabherr G. 2003. A regional impact assessment of climate and land-use change on alpine
vegetation. J. Biogeogr. 30: 401-417.
Fischer S.F., Poschlod P. and Beinlich B. 1996. Experimental studies on the dispersal of plants and animals on sheep in
calcareous grassland. J. Appl. Ecol. 33: 1206-1222.
Greene D.F. and Calogeropoulos C. 2002. Measuring and modelling seed dispersal of terrestrial plants. In: Bullock
J.M., Kenward R.E. and Hails R.S. (eds) Dispersal ecology. The 42
nd
Symposium of the British Ecological Society
held at the University of Reading 2-5 April 2001. Blackwell, Oxford, 3-23.
Greene D.F. and Johnson E.A. 1997. Secondary dispersal of tree seeds on snow. J. Ecol. 85: 329-340.
Guisan A. and Theurillat J.-P. 2000. Assessing alpine plant vulnerability to climate change: a modeling perspective.
Integrated Assessment 1: 307-320.
Hickey J.R., Flynn R.W., Buskirk S.W., Gerow K.G. and Willson M.F. 1999. An evaluation of a mammalian predator,
Martes americana, as a disperser of seeds. Oikos 87: 499-508.
Higgins S.I., Nathan R. and Cain M.L. 2003. Are long distance dispersal events in plants usually caused by nonstandard
means of dispersal? Ecology 84: 1945-1956.
Higgins S.I. and Richardson D.M. 1999. Predicting plant migration rates in a changing world: the role of long-distance
dispersal. Am. Nat. 153: 464-475.
Hodkinson D.J. and Thompson K. 1997. Plant dispersal: The role of man. J. Appl. Ecol. 34: 1484-1496.
Janzen D.H. 1984. Dispersal of small seeds by big herbivores: Foliage is the fruit. Am. Nat. 123: 338-353.
Jongejans E. and Telenius A. 2001. Field experiments on seed dispersal by wind in ten umbelliferous species
(Apiaceae). Plant Ecology 152: 67-78
Kiviniemi K. and Telenius A. 1998. Experiments on adhesive dispersal by wood mouse: seed shadows and dispersal
distances of 13 plant species from cultivated areas in southern Sweden. Ecography 21: 108-116.
Kollmann J. and Schill H.P. 1996. Spatial patterns of dispersal, seed predation and germination during colonization of
abandoned grassland by Quercus petraea and Corylus avellana. Vegetatio 125: 193-205.
Kot M., Lewis M.A. and van den Driessche P. 1996. Dispersal data and the spread of invading organisms. Ecology 77:
2027-2042.
Lang G. 1994. Quartäre Vegetationsgeschichte Europas. Methoden und Ergebnisse. Fischer Verlag, Jena.
Magri D., Vendramin G.G., Comps B., Dupanloup I., Geburek T., Gomory D., Latalowa M., Litt T., Paule L., Roure
J.M., Tantau I., van der Knaap W.O., Petit R.J. and de Beaulieu J.L. 2006. A new scenario for the Quaternary
history of European beech populations: palaeobotanical evidence and genetic consequences. New Phytol. 171: 199-
221.
12
Vittoz and Engler (2007) Botanica Helvetica 117: 109-124
McCanny S.J. and Cavers P.B. 1988. Spread of proso millet (Panicum miliaceum L.) in Ontario, Canada. II. Dispersal
by combines. Weed Res. 28: 67-72.
Malcolm J.R., Markham A., Neilson R.P. and Garaci M. 2002. Estimated migration rates under scenarios of global
climate change. J. Biogeogr. 29: 835-849.
Mattes H. 1982. Die Lebensgemeinschaft von Tannenhäher, Nucifraga caryocatactes (L.), und Arve, Pinus cembra L..
Ber. Eidg. Anst. Forstl. Versuchswes. 241: 1-74.
Mouissie A.M., Lengkeek W. and van Diggelen R. 2005a. Estimating adhesive seed-dispersal distances: field
experiments and correlated random walks. Funct. Ecol. 19: 478-486.
Mouissie A.M., van der Veen C.E.J., Veen G.F. and van Diggelen R. 2005b. Ecological correlates of seed survival after
ingestion by fallow deer. Funct. Ecol. 19: 284-290.
Müller-Schneider P. 1949. Unsere Vögel als Samenverbreiter. Orn. Beob. 46: 120-123.
Müller-Schneider P. 1963. Neue Beobachtungen über Samenverbreitung durch Ameisen. Ber. Schweiz. Bot. Ges. 73:
153-160.
Müller-Schneider P. 1983. Verbreitungsbiologie (Diasporologie) der Blütenpflanzen. 3. Aufl. Veröff. Geobot. Inst. ETH
Stiftung Rübel Zürich 61: 1-226.
Müller-Schneider P. 1986. Verbreitungsbiologie der Blütenpflanzen Graubündens. Veröff. Geobot. Inst. ETH Stiftung
Rübel Zürich 85: 1-263.
Nathan R., Sapir N., Trakhtenbrot A., Katul G.G., Bohrer G., Otte M., Avissar R., Soons M.B., Horn H.S., Wikelski M.
and Levin S.A. 2005. Long-distance biological transport processes through the air: can nature's complexity be
unfolded in silico? Diversity Distrib. 11: 131-137.
Pakeman R.J. 2001. Plant migration rates and seed dispersal mechanisms. J. Biogeogr. 28: 795-800.
Pearson R.G. 2006. Climate change and the migration capacity of species. Trends Ecol. Evol. 21: 111-113.
Pitelka L.F., Gardner R.H., Ash J. Berry S., Gitay H., Noble I.R., Saunders A., Bradshaw R.H.W., Brubaker L., Clark
J.S., Davis M.B., Sugita S., Dyer J.M., Hengeveld R., Hope G., Huntley B., King G.A., Lavorel S., Mack R.N.,
Malanson G.P., McGlone M., Prentice I.C. and Rejmanek M. 1997. Plant migration and climate change. Am. Sci.
85: 464-473.
Pyšek P. and Hulme P.E. 2005. Spatio-temporal dynamics of plant invasions: linking pattern to process. Ecoscience 12:
302-315.
Reid C. 1899. The origin of the British flora. Dulau, London.
Ridley H.N. 1930. The dispersal of plants throughout the world. Reeve, Ashford.
Römermann C., Tackenberg O. and Poschlod P. 2005. How to predict attachment potential of seeds to sheep and cattle
coat from simple morphological seed traits. Oikos 110: 219-230.
Ronce O. 2001. Understanding plant dispersal and migration. Trends Ecol. Evol. 16: 663.
Schmidt W. 1918. Die Verbreitung von Samen und Blütenstaub durch die Luftbewegung. Oesterr. Bot. Z. 67: 313-328.
Schneider S. 1935. Untersuchungen und Samenschleudermechanismen verschiedener Rhoeadales. Jahrb. Wiss. Botanik
81: 663-704.
Skellam J.G. 1951. Random dispersal in theoretical populations. Biometrika 38: 196-218.
Soons M.B. and Ozinga W.A. 2005. How important is long-distance seed dispersal for the regional survival of plant
species? Diversity Distrib. 11: 165-172.
Stewart J.R. and Lister A.M. 2001. Cryptic northern refugia and the origins of the modern biota. Trends. Ecol. Evol. 16:
608-613.
Stöcklin J. and Bäumler E. 1996. Seed rain, seedling establishment and clonal growth strategies on a glacier foreland. J.
Veg. Sci. 7: 45-56.
Tackenberg O. 2001. Methoden zur Bewertung gradueller Unterschiede des Ausbreitungspotentials von Pflanzenarten.
PhD thesis, Philipps-Universität Marburg.
Tackenberg O., Poschlod P. and Bonn S. 2003. Assessment of wind dispersal potential in plant species. Ecol. Monogr.
73: 191-205.
Takahashi K. and Kamitani T. 2004. Effect of dispersal capacity on forest plant migration at landscape scale. J. Ecol.
92: 778-785.
Thuiller W., Lavorel S., Araújo M.B., Sykes M.T. and Prentice I.C. 2005. Climate change threats to plant diversity in
Europe. Proc. Natl. Acad. Sci. U.S.A. 102: 8245-8250.
Tinner U. and Schumacher H. 2004. Flora auf Bahnhöfen der Nordostschweiz. Bot. Helv. 114: 109-125.
Vellend M., Myers J.A., Gardescu S. and Marks P.L. 2003. Dispersal of Trillium seeds by deer: implications for long-
distance migration of forest herbs. Ecology 84: 1067-1072.
Wilkinson D.M. 1997. Plant colonization: are wind dispersed seeds really dispersed by birds at large spatial and
temporal scales? J. Biogeogr. 24: 61-65.
13
Vittoz and Engler (2007) Botanica Helvetica 117: 109-124
Williamson M., Pyšek P., Jarošík V. and Prach K. 2005. On the rates and patterns of spread of alien plants in the Czech
Republic, Britain, and Ireland. Ecoscience 12: 424-433.
Xiao Z., Zhang Z. and Wang Y. 2004. Dispersal and germination of big and small nuts of Quercus serrata in a
subtropical broad-leaved evergreen forest. Forest Ecol. Manage. 195: 141–150.
Appendix
App. 1. Literature data on seed dispersal distances. The examples are mainly from the Swiss flora,
except when data were insufficient for certain dispersal modes. Some supplementary species were
thus added, mostly from temperate regions. Asterisks indicate values that were considered to
represent long-distance dispersal and were therefore excluded from data analysis.
This Appendix can be downloaded freely from http://www.birkhauser.ch/BH, "Electronic
supplementary material".
... The dispersal distance should depend on not only dispersal vectors, but also on plant traits that attract and manipulate the position and behaviour of pollinators. There has been substantial progress in investigating traits that affect seed dispersal (Vittoz and Engler 2007;Thomson et al. 2011;Côrtes and Uriarte 2013;Tamme et al. 2014), but we remain largely ignorant of how pollen dispersal distance varies among dispersal vectors and for different plant trait values, likely flew longer distances after visiting flowers with a low quantity of nectar, presumably implying greater pollen dispersal distances (Waddington 1981). In hummingbird-pollinated Ipomopsis aggregata, mean pollen dispersal distance, estimated using pollen dyes, was found to increase with the variance but decrease with the mean stamen length, whereas it was independent of flowering date, number of flowers, and corolla size (Campbell and Waser 1989). ...
... Both male and bisexual flowers are predominantly visited by flies, including houseflies and syrphid flies (Chen and Pannell 2022). Ripe fruits (technically achenes) have elongated pappus hair and are dispersed by wind in early autumn (Vittoz and Engler 2007). After fruit dispersal, above-ground vegetative parts senesce, but individuals persist underground over winter. ...
... However, effective seed dispersal is expected to break down spatial patterns created by limited pollen dispersal (Meirmans et al. 2011). We quantified only the pollen dispersal kernel of P. alpina, but its seeds are dispersed by wind from elongated floral stalks (Chen and Pannell 2022), with maximum dispersal distances estimated to be of the order of 80 m (Vittoz and Engler 2007). In addition, the very high inbreeding depression estimated for P. alpina (i.e., 0.93) also means that progeny produced by mating among neighbouring relatives should largely be removed from the population in early life stages, further reducing any potential for the build-up of fine-scale genetic structure, as has been found for Rhododendron brachycarpum (Hirao 2010). ...
Article
Full-text available
Pollen-mediated gene flow and spatial genetic structure have rarely been studied in alpine plants that are pollinated by dipteran insects. In particular, it is not clear how different floral traits, such as floral gender, phenology, and ancillary traits, may affect pollen dispersal distance within alpine plant populations. In this study, we conducted a paternity analysis to track pollen flow in a population of Pulsatilla alpina, an andromonoecious alpine herb producing male and bisexual flowers. We found that the pollen was dispersed over short distances (mean = 3.16 m), with a dispersal kernel following a Weibull distribution. Nonetheless, spatial genetic structure was weak in the population (Sp statistic = 0.013), pointing to effective seed dispersal and/or high inbreeding depression. The pollen dispersal distance was independent of the gender of the flower of origin but depended positively on floral stalk height and negatively on flowering date and tepal length. Although male siring success did not correlate with pollen dispersal distance, selection may favour traits that increase the pollen dispersal distance as a result of reduced bi-parental inbreeding. Our study not only provides new insights into the nature of pollen dispersal of alpine plants, but also reveals the effects of floral traits on a component of male reproductive success.
... We examined how entomophilous plant species assemblages responded to connectivity by analyzing three combined groups of species based on their growth form and longevity (: annual/biennial herbs, perennial herbs and shrubs). These groups of species have characteristic life-history traits, such as seed size and plant height, which are associated with dispersal modes that could be used to explain their persistence in the landscape (Soons et al., 2005;Lindborg & Eriksson, 2005;Vittoz & Engler, 2007;Lindborg et al., 2012). ...
... As landscape connectivity depends on the dispersal ability of the different species, plant species were classified according to their dispersal types proposed by Vittoz and Engler (2007), based on dispersal modes and plant traits that influence the efficiency of dispersal. Each dispersal type was estimated as the upper limits of the distances within which 50% and 99% of the seeds of a plant species population are dispersed (Appendix S2). ...
... Following the typology described in the previous section (Vittoz & Engler, 2007), five dispersal distances were included in the connectivity analysis (i.e., 15, 150, 500, 1500 and 5000 m) to reflect the variation on seed dispersal capacity. Thus, by using a range of dispersal distances we can examine how the landscape may be connected for a wide range of entomophilous plant species corresponding to different typology (Appendix S2). ...
Article
Aims Functional connectivity is crucial for conserving biodiversity and ecosystem services in fragmented landscapes. Linear landscape elements play an important role in providing refuge for plants and pollinators, as well as serving as biological corridors. The aim of this study was to assess the importance of road verges in maintaining the ecological connectivity of plant species assemblages in the Tandilia mountain system. Location South America, Southern Pampa region of Argentina, Tandilia mountain system. Methods Using graph theory, landscape connectivity was quantified both at landscape scale and at the patch scale using the probability connectivity index (PC num ), taking into account the dispersal ability of observed species. The contribution of each landscape element to habitat availability and connectivity was evaluated using different fractions of the PC num metric. Likewise, the influence of connectivity variables in explaining species assemblages grouped by functional categories was investigated using canonical ordination techniques. Results The inclusion of road verges led to an 81% increase in overall connectivity at a threshold distance of 500 m. Connectivity significantly explained the assemblages of entomophilous plant species for all dispersal distances. Annual species assemblage variation was associated with intra‐patch connectivity. Variation in perennial and shrub species assemblages was explained by intra‐patch and inter‐patch connectivity. Conclusions Preservation and restoration of these linear landscape elements play a pivotal role in transitioning toward more sustainable agroecosystems. This ecological and practical information may help prioritize plant species and the sites used to restore an essential but neglected ecosystem.
... Both male and bisexual owers are predominantly visited by ies, including house ies and syrphid ies (Chen and Pannell, 2022). Ripe fruits (technically achenes) with elongated pappus hair are dispersed by wind in early autumn (Vittoz and Engler 2007). After fruit dispersal, above-ground vegetative parts senesce, but individuals persist underground over winter. ...
... In general, high spatial genetic structure is expected for species pollinated by insects, especially those with predominant sel ng, and those with a herbaceous growth form and seed dispersed by gravity (Vekemans and Hardy 2004). In contrast, seed dispersal by wind in P. alpina is likely to be e cient, especially as the species produces elongated oral stalks up to 70 centimeters above ground from which the bisexual owers disperse their seeds (Chen and Pannell 2022), with a maximum dispersal distance of 80 meters (Vittoz and Engler 2007). Of course, we quanti ed only the pollen dispersal kernel in this study, so that estimates of both pollen and seed dispersal kernels together will be needed to provide a more complete picture of the mechanisms shaping the spatial genetic structure of P. alpina (e.g., García et al., 2007;Krauss et al., 2008). ...
Preprint
Full-text available
Pollen-mediated gene flow and spatial genetic structure have rarely been studied in alpine plants pollinated by Dipteran insects. Furthermore, it is not clear how different floral traits, such as floral gender, phenology, and ancillary traits, may affect pollen dispersal distance within a population. In this study, we conducted a paternity analysis to track pollen flow in a population of Pulsatilla alpina , an andromonoecious alpine herb producing male and bisexual flowers. We found that the pollen was dispersed over short distances (mean = 3.16 meters) with a dispersal kernel of Weibull distribution. Nonetheless, spatial genetic structure was weak in the population ( Sp statistic = 0.013), pointing to effective seed dispersal. The pollen dispersal distance was independent of the gender of the flower of origin but depended positively on floral stalk height and negatively on flowering date and tepal length. Although male siring success did not correlate with pollen dispersal distance, selection may favor traits increasing pollen dispersal distance as a result of reduced bi-parental inbreeding. Our study has not only provided new insights into the nature of pollen dispersal, especially of alpine plants, but has also revealed the effects of floral traits on an important component of male reproductive success.
... A c c e p t e d M a n u s c r i p t subalpine to alpine climate) (Kattge et al. 2020;Ellenberg et al., 1991), and are generally poor dispersers, with seeds typically dispersing ~1 m from the parent plant (Vittoz & Engler, 2007). Populations at Lake Superior are considered highly disjunct, with a mean distance from Lake Superior to species" core ranges of 514 km. ...
Article
Full-text available
Climate refugia can serve as remnant habitat or stepping stones for species dispersal under climate warming. The largest freshwater lake by surface area, Lake Superior, USA and Canada, serves as a model system for understanding cooling-mediated local refugia, as its cool water temperatures and wave action have maintained shoreline habitats suitable for southern disjunct populations of arctic-alpine plants since deglaciation. Here we seek to explain spatial patterns and environmental drivers of arctic-alpine plant refugia along Lake Superior’s shores, and assess future risk to refugia under moderate (+3.5 °C) and warmest (+5.7 °C) climate warming scenarios. First, we examined how the interactive effects of summer surface water temperatures and wind affected onshore temperatures, resulting in areas of cooler refugia. Second, we developed an ecological niche model for presence of disjunct arctic-alpine refugia (pooling 1253 occurrences from 58 species) along the lake’s shoreline. Third, we fit species distribution models for 20 of the most common arctic-alpine disjunct species and predicted presence to identify refugia hotspots. Finally, we used the two climate warming scenarios to predict changes in presence of refugia and disjunct hotspots. Bedrock type, elevation above water, inland distance, July land surface temperature from MODIS/Terra satellite, and near-shore depth of water were the best predictors of disjunct occurrences. Overall, we predicted 2,236 km of the shoreline (51%) as disjunct refugia habitat for at least one species under current conditions, but this was reduced to 20% and 7% with moderate (894 km) and warmest (313 km) climate change projections.
... Although no dispersal studies have been conducted for the genus, the presence of fruits with winged valves in these herbs (fruits formed by schizocarpic siliques that break at maturity into two 1-seeded indehiscent mericarps, Salariato et al., 2014) suggests possible wind dispersal (pterometeorochory), which is associated with short to moderate distances (approx. 2-15 m, Vittoz and Engler, 2007). Therefore, for model projection in each species was used the same accessible area utilized for calibration (the 50km-MCP of each species). ...
Article
Climate change impact on species can be heterogeneous depending on their environments, exposure, and intrinsic characteristics. Likewise, global warming may have an uneven effect on lineages, depending on whether phylogenetic conservatism or divergence of ecological niches predominates during clade diversification, imposing a higher risk to species groups from certain regions, habitats and lineages. This study evaluates the impact of future climate change on Menonvillea, a genus with 24 species distributed along the Andes and contiguous regions of the Southern Cone. The impact on the main phylogenetic, ecological and biogeographic groups is evaluated, also analyzing the effect on its richness and phylogenetic diversity. Results show a strongly negative impact on most species of the genus. However, the greatest pressure seems to be recovered for high Andean species, mainly from the southern portion of the Southern Andes (between 34°S–53S°), and mostly included in Menonvillea sect. Cuneata. Richness appears to be more impacted in high Andean regions, and the loss of phylogenetic diversity is greater than expected at random. These results highlight the strong negative impact that climate change can induce on lineages distributed in the Andean-Patagonian region, and that show patterns of phylogenetic niche conservatism.
... In species distribution models, two contrasting dispersal scenarios are typically considered: unlimited dispersal, where a species can occupy any suitable area, and no dispersal, where it cannot follow its preferred environmental conditions [121], though the reality is often somewhere in between [122]. The choice between these scenarios, often made due to the simplicity of application or lack of detailed ecological data (e.g., lifecycle, seed characteristics, dispersal vectors, and methods) [123], is usually based on fundamental species characteristics such as the level of endemicity and life form. Without incorporating dispersal limits and a comprehensive set of environmental variables that closely mirror the species' niche into our analysis, Quercus alnifolia might be expected to thrive in a warmer and drier world. ...
Article
Full-text available
Endemic island species face heightened extinction risk from climate-driven shifts, yet standard models often underestimate threat levels for those like Quercus alnifolia, an iconic Cypriot oak with pre-adaptations to aridity. Through species distribution modelling, we investigated the potential shifts in its distribution under future climate and land-use change scenarios. Our approach uniquely combines dispersal constraints, detailed soil characteristics, hydrological factors, and anticipated soil erosion data, offering a comprehensive assessment of environmental suitability. We quantified the species’ sensitivity, exposure, and vulnerability to projected changes, conducting a preliminary IUCN extinction risk assessment according to Criteria A and B. Our projections uniformly predict range reductions, with a median decrease of 67.8% by the 2070s under the most extreme scenarios. Additionally, our research indicates Quercus alnifolia’s resilience to diverse erosion conditions and preference for relatively dry climates within a specific annual temperature range. The preliminary IUCN risk assessment designates Quercus alnifolia as Critically Endangered in the future, highlighting the need for focused conservation efforts. Climate and land-use changes are critical threats to the species’ survival, emphasising the importance of comprehensive modelling techniques and the urgent requirement for dedicated conservation measures to safeguard this iconic species.
... Their dispersal is only successful if habitat patches are sufficiently connected 147,148 , or if it is facilitated by suitable landscape features 149 . Indeed, the ability of plants to disperse in fragmented landscapes also depends on their dispersal strategy 150,151 . Overall, biodiversity loss largely depends on the temporal patterns of environmental changes, e.g., when biological response cannot match the rate of soil-use changes 152 . ...
Article
Full-text available
This study investigated the germination capacity (endogenous factor) of Petagnaea gussonei (Spreng.) Rauschert, an endemic monospecific plant considered as a relict species of the ancient Mediterranean Tertiary flora. This investigation focused also on the temporal trends of soil-use, climate and desertification (exogenous factors) across the natural range of P. gussonei. The final germination percentage showed low values between 14 and 32%, the latter obtained with GA3 and agar at 10 °C. The rising temperatures in the study area will further increase the dormancy of P. gussonei, whose germination capacity was lower and slower at temperatures higher than 10 °C. A further limiting factor of P. gussonei is its dormancy, which seems to be morpho-physiological. Regarding climate trends, in the period 1931–2020, the average temperature increased by 0.5 °C, from 15.4 to 15.9 °C, in line with the projected climate changes throughout the twenty-first century across the Mediterranean region. The average annual rainfall showed a relatively constant value of c. 900 mm, but extreme events grew considerably in the period 1991–2020. Similarly, the land affected by desertification expanded in an alarming way, by increasing from 21.2% in 2000 to 47.3% in 2020. Soil-use changes created also a complex impacting mosaic where c. 40% are agricultural areas. The effective conservation of P. gussonei should be multilateral by relying on germplasm banks, improving landscape connectivity and vegetation cover, and promoting climate policies.
Article
Full-text available
The plant diversity of riparian zones has significantly deteriorated due to human disturbances, making their restoration a key focus in research. However, restoration efforts often fail due to the lack of available propagules. Hydrochory, i.e., plants dispersed via water, which plays a critical role in transporting propagules downstream and is pivotal for riparian habitat recovery. Although hydrochory is closely related to the establishment and restoration of downstream riparian vegetation, most previous studies have concentrated mainly on the species richness and the number of propagules dispersed by water, but have overlooked the functional diversity. We explored the temporal variations in functional traits composition and functional diversity of hydrochorous propagules in three upland rivers within the upper Yangtze River catchment, situated in a humid subtropical monsoon climate. We find that during the high flows of summer, these rivers transport more species and exhibit higher functional diversity. This underscores the critical role of high summer flows for breaking the dispersal bottleneck of species with limited dispersal ability, and emphasizes the importance of hydrochory during the flood season for the recovery of riparian vegetation and maintaining high river flows is a critical strategy for restoration in the era of global flow regulation.
Preprint
Ecological connectivity is key to maintaining a coherent and resilient network of protected areas in the EU. The EU Biodiversity Strategy for 2030 has identified the unhindered movement of species, nutrients and ecological processes across connected landscapes as a key feature of a coherent Trans-European Nature Network (TEN-N) of protected and conserved areas. However, to date, streamlined guidance on planning for and implementing connectivity measures specifically at the European scale has been limited. This report presents a coherent methodological framework and guidelines for mapping functional and structural connectivity at the European scale, as part of the Horizon Europe NaturaConnect project, which is supporting EU Member States in developing a coherent TEN-N of protected and conserved areas. It describes key ecological connectivity concepts and approaches; outlines methods and tools for estimating connectivity; presents an overview of connectivity projects across Europe; identifies connectivity priorities, gaps and challenges following a stakeholder consultation process; and provides practical and operational guidelines for implementing ecological connectivity for conservation projects ranging from regional to national and European levels. The guidelines present a strategic blueprint aimed at enhancing ecological connectivity across Europe, and address the specific challenges and opportunities related to planning ecological connectivity in the European context. This report has been written for practitioners and individuals involved in the management and administration of protected areas and ecological connectivity projects across Europe. This includes professionals working in TEN-N implementation at national or regional levels, others involved in spatial planning outside protected areas, and professionals engaged in the implementation of connectivity projects and protected area management.
Article
Four train stations of north eastern Switzerland were investigated during two years: Rorschach, St. Margrethen, Wattwil, and Wil. The complete list of vascular plants contains approximately 300 different species for each station, altogether 509 species. The plants are ecologically grouped. Typical train station species are extracted and their survival strategy is presented. Vascular plants on train stations are mainly therophytes and hemicryptophytes, which are well adapted for periodic interference by measures to combat weed. Train stations are well suited for alien plants. Some of them remain rare and adventitious, others become aggressive neophytes. Examples of rare adventitious species and aggressive neophytes are discussed.
Article
1. In-stream dispersal and floating capacity of 18 pioneer plants (nine chamaephytes, eight hemicryptophytes and one shrub) from alluvial floodplains on the upper Isar (Bavaria, Germany) were studied using field and laboratory experiments between 1994 and 1996. 2. Drift collection of seeds with nets at the water surface proved unsuitable because of the great amount of invertebrates drifting along the river, though it allowed an estimation of the amount of drifting seeds. In August and September we found over 9,000 drifting seeds per meter stream width and day or over 120,000 seeds over the whole stream width in 24 h. Sediment-filled baskets proved to be better traps but were often destroyed during flooding periods. Although drift collection was of limited value for sampling seeds in a highly dynamic floodplain, water transport of viable seeds could be demonstrated for nine of the 18 species. 3. In laboratory experiments, in which seeds were placed into a beaker, stirred up to five days and then used for germination tests, seeds of 16 of the 18 species showed a moderate to good floating capacity. Only two species had seeds that could not float longer than a few minutes (Gypsophila repens and Silene vulgaris ssp. glareosa). 4. We hypothesised that dispersability decreases with increasing seed mass, but the assumption that heavy seeds are not dispersed over great distances is valid only for environments without the possibility of in-stream dispersal.
Article
»Diasporus«, a database for diaspore dispersal characteristics, is presented. The database refers to a vector-based dispersal classification. It includes dispersal relevant traits of the parent plant as well as of the diaspore. Standard methods to ascertain indicator parameters for dispersal by wind, water and animals are proposed. Examples of the application of the database in the context of viability analysis and risk assessment for plant species are given.