ArticlePDF Available

Species traits and habitats in springtail communities: a regional scale study

Authors:
1
Species traits and habitatsin springtail communities: a 1
regional scale study 2
3
S. Salmon,J.F. Ponge 4
5
Muséum National d’Histoire Naturelle, CNRS UMR 7179, 4 avenue du Petit-Château, 6
91800 Brunoy France 7
8
Running title: Trait-habitat relationshipsin springtails 9
10
Corresponding author. Tel.: +33 6 78930133.
E-mail address:ponge@mnhn.fr (J.F. Ponge).
2
Abstract 1
Although much work has been done on factors patterning species trait assemblages in 2
emblematic groups such as plants and vertebrates, more remains to be done in 3
belowground invertebrate species. In particular,relationships between species traits and 4
habitat preferences are still a matter of debate. Springtails were sampled in a 5
heterogeneous landscape centered on the Sénart forest, near Paris (northern France), 6
embracingthe largest possible array of five environmental gradients (humus forms, 7
vegetation, moisture, vertical strata, and seasons) over which Collembola are known to 8
be distributed. Distances between samples varied from a few cm to several km. 9
Canonical correspondence analysis using species (128) as observations and species trait 10
attributes (30) and habitat indicators (82) as dependent and independent variables, 11
respectively, allowed to discern whether species habitats and species trait assemblages 12
were related and which trends could be found in trait/environment relationships. It was 13
concluded that, within the studied area, species habitatswere significantly associated 14
with species trait assemblages. The main gradient explaining the distribution of species 15
traits combined the vertical distribution of habitats (from the mineral soil to plant aerial 16
parts), and the openness of the environment, i.e. a complex of many ecological factors. 17
In the ecological traits of Collembola, this gradient corresponded to anincreasing 18
contribution of sensory and locomotory organs, bright color patterns, size and sexual 19
reproduction, all attributes associated with aboveground life under herbaceous cover. 20
Another important, although secondarycontrast concerned traits associated with habitats 21
far from soil but concealed (corticolous vs all other habitats). Soil acidity and water did 22
not contribute significantly to trait distribution, at least within the limits of our database. 23
Keywords:Collembola; species trait assemblages; habitats; trait-environment 24
relationships 25
3
Introduction 1
The indicative power of species trait assemblages has been intensively studied in 2
plants, birds and beetles and most species traits could be clearly related to habitat 3
preferences of species in these groups (Graves and Gotelli 1993; Ribera et al. 4
2001;Cornwell and Ackerly 2009; Mayfield et al 2009; Pavoine et al. 2011). 5
Surprisingly, although this is common sense and was reported for a long time in soil 6
zoology (Bornebusch 1930), few studies questioned whether the extraordinary diversity 7
of species traits which prevail in soil animal communities could be explained, and 8
potentially could have been selected, by differences in habitat use (Vandewalle et al. 9
2010; Decaëns et al. 2011; Bokhorst et al. 2011).Moreover, these studies focused either 10
on a restricted number of traits, or a restricted number of habitats which does not allow 11
providing general trends in relationships between species traits and habitat use. 12
The aim of our study was to determine trends that emerge from trait-13
environment relationships, i.e. how species traits vary along environmental gradients 14
(e.g. vegetation, soil, depth). 15
Among soil invertebrates, we selected springtails (Hexapoda, Collembola) as an 16
abundant and diversifiedmonophyletic group for which a great deal of work has been 17
devoted to the study of species/environment relationships at the community level (Poole 18
1962; Hågvar 1982; Ponge 1993; Chagnon et al. 2000; Auclerc et al. 2009). The Sénart 19
forest (Ile-de-France, northern France) and its vicinity were selected because they 20
display a great variety of soil and soil-related habitats (e.g. woodland, heathland, 21
grassland, ponds, paths, tree trunks) composing a little more than 3,000 ha of 22
heterogeneous landscape, now totally included in the Paris area. Data collected from 23
1973 to 1977, at a time when agriculture was still practiced both inside and outside the 24
4
forest, were revisited for a statistical analysis taking into account species 1
trait/environment relationships. The same pool of data (370 samples, 127 species) has 2
been already used in several studies dealing with species/environment relationships 3
(Ponge 1980, 1983, 1993) and was included in the COLTRAIT data base 4
[http://www.bdd-inee.cnrs.fr/spip.php?article51&lang=en], which also comprises data 5
about twelve morphological and life-history traits of more than 300 collembolan 6
species. 7
Materials and Methods 8
Site description 9
The Sénart state forest (3,000 ha) is located 20 km south-east of Paris on the 10
western border of the Brie plateau, delineated by a meander of river Seine and by a 11
tributary, the river Yerres, at an altitude ranging from 50 to 87 m a.s.l. At the time of 12
sampling it was mainly bordered by urbanized areas (communes of Quincy-sous-Sénart, 13
Boussy-Saint-Antoine, Brunoy, Yerres, Montgeron, Draveil) on its western and 14
northern parts, and by agricultural areas (communes of Soisy, Étiolles, Tigery, 15
Lieusaint, Combs-la-Ville) on its eastern and southern parts. Nowadays, the forest is 16
totally included in the metropolitan area of Paris. Private peripheral woods and 17
agricultural areas (cultures and meadows) were included in the study. Most of them 18
have now been incorporated into the state forest, to the exception of peripheral 19
agricultural areas which have been built or transformed into golf courses or other 20
recreational areas.A number of soil types can be observed in the Sénart forest, varying 21
according to the nature of quaternary deposits (loess or gravels) and permanent or 22
seasonal waterlogging resulting from clay migration (perched water tables) or 23
5
underlying impervious clay strata (permanent water tables). More details were given in 1
previously published papers (Ponge 1980, 1983, 1993). 2
Sampling procedure 3
Sampling took place from 15th October 1973 to 10th October 1977 in every 4
season and every kind of weather, our purpose being to embrace all climate conditions, 5
except when the soil was deeply frozen and could not be sampled at all. At each 6
sampling time, a point was randomly selected, around which all visible sitespotentially 7
available to springtails were investigated, from deep soil (leached mineral horizons) to 8
tree trunks two meters aboveground and to floating vegetation in water-filled ponds. No 9
effort was made to standardize sampling, the only requirement being to collect enough 10
litter (at all stages of decomposition), vegetation (aerial and subterranean parts), bark 11
(naked or covered with lichens or mosses) or soil (organo-mineral to mineral horizons) 12
to have enough animals as possible in each sample, the aim of the study being to know 13
which species were living together in the same micro-habitat and which species were 14
not.The volume sampled varied from 100 mL for moss cushions, which are particularly 15
rich in springtails (Gerson 1982) to 1 L for bleached mineral soil horizons which are 16
strongly impoverished in fauna (Hågvar 1983). Care was taken not to undersample 17
some poorly represented habitats. For that purpose some additional sampling was done 18
in agricultural areas, calcareous soils and dumping places. This procedure allows 19
environmental gradients to be better described (Gillison and Liswanti 2004). 20
Samples were taken with the help of a shovel for soil, and with fingersfor above-21
ground samples, care being taken not to lose too many jumping animals in particular 22
when sampling aerial parts of erected plants. No attempt was done to force a corer into 23
the soil. Samples were immediately put in plastic bags then transported to the nearby 24
6
laboratory, to be extracted on the same day. Extraction was done by the dry funnel 1
(Berlese) method over 10 days, using 25 W bulb lamps in order to avoid too rapid 2
desiccation of the samples, known to prevent slowly moving animals from escaping 3
actively the samples (Nef 1960). Animals were collected and preserved in 95% ethyl 4
alcohol in plastic jars. A total of 310 samples were collected and kept for the analysis. 5
Species identification 6
Animals were sorted in Petri dishes filled with ethyl alcohol then springtails 7
were mounted and cleared in chloral-lactophenol to be identified under a light 8
microscope at x 400 magnification. At the time of study the only key available for 9
European springtails was that of Gisin (1960), to which were added numerous detailed 10
published studies at family, genus or species level (complete list available upon 11
request), and miscellaneous (unpublished) additions by Gisin himself. Color patterns 12
were noted before animals were discolored in chloral-lactophenol. Young specimens, 13
when not identifiable to species level, were allocated to known species by reference to 14
adults or subadults found in the same sample, or in samples taken in the vicinity. For 15
instance in the genus Mesaphorura, where several species may cohabit and diagnostic 16
characters are not revealed in the first instar (Rusek 1980), unidentified juveniles were 17
proportionally assigned to species on the base of identified specimens found in the same 18
sample. Gisin’s nomenclature was updated using Fauna Europaea 2011 19
[http://www.faunaeur.org/]. A total of 128 species were found (Table 1). 20
Trait data 21
Twelve traits, mostly extracted from the COLTRAIT data base and collected 22
from numerous identification keys or synopses, describe morphology and reproductive 23
mode of the 128 species used in the analysis. Attributes of each trait (Table 3) were 24
7
considered as variables, and were coded as binary (dummy) variables, resulting in a list 1
of 30 attributes: mode of reproduction (parthenogenesis dominant, sexual reproduction 2
dominant), body size (small, medium, large), body form (cylindrical body, stocky body, 3
spherical body), body color (pale-colored, bright-colored, dark-colored), scales (absent, 4
present), antenna size (short, long), leg size (short, long), furcula size (absent or 5
vestigial, short, long), eyenumber (0, 1-5, > 5), pseudocella (absent, present), post-6
antennal organ (absent, simple, compound), and trichobothria (absent, present). 7
Antennae, eyes, post-antennal organsand trichobothria are supposed to play a sensory 8
role (Hopkin 1997). 9
Species habitat data 10
Field notes were used to classify habitat features (sensu lato, including micro-11
habitat and season) in 82 categories (Table 2). To each sample was thus assigned a set 12
of 82 habitat indicators which describe its main features at varying scales, from landuse 13
(heathland, grassland, woodland) to sampling plot (e.g. ditch, plain ground, pond, 14
vegetation, soil pH) then to within-plot scale (e.g. plant part, litter, earthworm casts, 15
mineral soil). Species presence was indicated by dummy variables (coded as 0 or 1) for 16
each of the 82 habitat categories. 17
Statistical treatment of the data 18
Canonical correspondence analysis was used to analyze trait-habitat 19
relationships (species as observations, species trait attributes as dependent variables, 20
species habitatsas constraining variables), permutation tests being used to test trait-21
habitat associations. 22
Rarefaction curves were calculated to estimate the exhaustiveness of our 23
sampling method.Rarefaction curves and jacknife estimators were calculated using 24
8
EstimateS (version 8.2.0).All other calculations were done using XLSTAT® 1
(Addinsoft®, Paris, France). 2
Results 3
The rarefaction curve of the 128 observed species showed that sampling had 4
approached an asymptote. Estimating the number of missing species according to Chao 5
(1987) put the expected total number of species for the Sénart forest to 133 and 6
indicated that the sampling was relatively exhaustive. 7
Canonical Correspondence Analysis (CCA)with species trait attributes as 8
explained variables and species habitats as explanatory variables showed that traits were 9
significantly explained by habitats (number or permutations = 500, pseudo-F = 0.94, P < 10
0.0001). Constrained variance (variance of species traits explained by species habitats) 11
represented 72.9% of the total variance. 12
The first two canonical components of CCA extracted 54% of the constrained 13
(explained) variance (40% and 14% for F1 and F2, respectively). The projection of trait 14
attributes and species in the F1-F2 plane is shown in Figures 1a and 1b, respectively. 15
Both species and trait attributes were distributed along three dimensions. Species with 16
pseudocella and post-antennal organ present (of compound type), parthenogenesis 17
dominant, regressed locomotory (furcula, legs) and sensorial organs (eyes, antennae, 18
thichobothria), and pale color were opposed to species displaying opposite attributes 19
along F1. According to principal coordinates of species habitats (Table 2) this 20
corresponded to opposite habitats: woodland vs grasslandand depth versus surface, from 21
negative to positive sides of F1. Heathland was in an intermediate position between 22
woodland and grassland (Table 2). Mineral soil, organo-mineral soil, humus (organic), 23
litter, plant aerial parts ranked in this order along F1. Sunlight was projected on the 24
9
positive side of F1 (open environments).The second canonical component F2 was more 1
specifically linked to corticolous microhabitats (trunks, wood and associated mosses 2
and lichens): associated trait attributes were short furcula, stocky and dark-colored 3
body, eyes present but in regressed number (1-5), post-antennal organ present but 4
simple. Acidity and humus type, as well as water, did not exhibit any pronounced 5
influence on species trait attributes. Partial CCA, allowing only water and soil acidity 6
(including humus type) to vary, showed that they did not influence the distribution of 7
trait attributes (pseudo-F = 0.17, P = 0.99). 8
Discussion 9
Previous studies showed that a limited number of ecological factors could 10
explain the distribution of collembolan species when collected in the same geographical 11
context, at a regional scale (Ponge 1993; Ponge et al. 2003). Vertical distribution is the 12
main gradient along which most springtail species are distributed (Hågvar 1983; Faber 13
and Joosse1993; Ponge 2000a), followed by the contrast between woodland and 14
grassland (Ponge et al. 2003), and other factors such as water availability (Verhoef and 15
Van Selm 1983) and soil acidity (Loranger et al. 2001). We showed that grassland and 16
epigeic habitats were mostly characterized by traits adapting species to surface life: big 17
size, high mobility, protection against desiccation by round shape or cuticular clothing 18
(Kaersgaard et al. 2004), avoidance of predation by flight and color signaling, and 19
sexual reproduction (Fig. 1, Table 2, F1 component, positive side). On the oppositeside, 20
woodland and endogeic habitats were mostly characterized by traits associated with 21
subterranean life: small size, small locomotory appendages, poor protection from 22
desiccation, avoidance of predation by toxic excreta (pseudocella), and parthenogenesis. 23
10
Much life in woodland is more concealed than in grassland: smaller forms, more 1
sensitive to environmental stress because of a higher surface/volume ratio (Kærsgaard et 2
al. 2004;Bokhorst et al. 2012), and less motile species (Auclerc et al. 2009), can find in 3
woodland better conditions for survival and reproduction. Mebes and Filser (1997) 4
showed that surface dispersal of Collembola was much more intense in agricultural 5
fields compared to adjoining shrubby fallows where litter began to accumulate, and 6
Alvarez et al. (1997, 2000) highlighted the role of hedgerows as temporary refuges for 7
species living at the surface of arable fields.Sexual reproduction needs easy-to-visit sites 8
for the deposition of spermatophores by males (Chahartaghi et al. 2006), and movement 9
in search of mating partners using olfactory or tactile clues (Chernova et al. 2010), 10
which is easier in surface than in depth, in the same sense as escape from predators 11
needs visual or tactile sensory organs to detect their presence (Baatrup et al. 2006) and 12
needs jumping movements (ensured by furcula acting as a spring) for fleeing away 13
(Bauer and Christian 1987). The fractionation of space within leaf or needle litter 14
horizons makes the forest floor improper to rapid surface movements (Bauer and 15
Christian 1987), while protecting soil-dwelling animals from surface predation by 16
carabids and vertebrates (Hossie and Murray 2010) and offering a variety of food 17
resources such as fungal colonies and animal excreta (Bengtsson et al. 1991; Salmon 18
and Ponge 2001). Other predators are subterranean and cannot be avoided through 19
active movements, hence the use of chemical repellents excreted by pseudocella 20
(Dettner et al. 1996; Negri 2004). 21
Despite clear trends of trait/habitat relationships exhibited by our results, 22
possible biases due to escape movements during sampling, in particular from the part of 23
big-size animals with long furcula, should not be overlooked. If such biases differ from 24
a habitat to another, this may flaw trait/habitat relationships. However, concerning the 25
11
association between big size and agricultural environments, which is novel to science, it 1
must be highlighted that it was less easy to collect vary motile specimens in the absence 2
of litter (i.e. in agricultural areas) than when litter was present (i.e. in forest areas), 3
stemming in a bias in quite opposite direction to the observed association. This made us 4
confident that such biases were not present in our dataset. 5
The second canonical component of trait-environment relationships (Fig. 1, 6
Table 2, F2 component) distinguishes traits associated with life in bark and associated 7
mosses and lichens: the combination of short furcula, dark color, stocky body, eyes 8
present but in limited number is an original adaptation to life in concealed environments 9
(hence small size and limited movements) but far from soil (hence the need to be 10
protected from UV radiation through pigmentation and possibilities offered by vision). 11
The structure of the post-antennal organ, opposing simple to compound structure (more 12
typical of edaphic habitat) is worthy of note, since no other studies considered its 13
ecological correlates. The exact role played by this organ is still unknown, but 14
anatomical observations on the innervation of these pitted porous plates located not far 15
from the protocerebrum point to sensory activity (Altner and Thies 1976). Differences 16
between simple and compound post-antennal organs concern the number of dendritic 17
branches, which are more numerous in compound organs (Altner and Thies 1976), 18
suggesting that compound post-antennal organs are more sensitive to chemical features 19
of the immediate environment. The higher sensitivity of the compound post-antennal 20
organ could be more adapted to deeper horizons by compensating the reduction or the 21
complete absence of other sense organs such as eyes. 22
The fact that we did not discern any association between traits and obvious 23
factors such as water and soil acidity (or humus type) does not preclude any further 24
scrutiny of such relationships. Two reasons could be invoked.First, that, in its present 25
12
state, our database did not cover the traits needed to establish this relationship. Ponge 1
(2000b) showed that acidophilic and acidophobic species cohabited within the same 2
lineage, pointing to corresponding traits as mainly based on physiology (mechanisms 3
counteracting oxidative stress) rather than on anatomy and reproduction mode. Traits 4
associated with aquatic life concern mainly the form and size of claws (Gisin 1960), and 5
of course physiology (resistance to desiccation), which were not considered here. 6
Second, in the particular case of the Senart forest, traits adapting species to habitats 7
varying in terms of water availability and/or soil acidity could be masked by landuse or 8
vertical stratification effects, pointing to the need for studying trait/habitat relationships 9
on a wider geographic scale, as suggested by Lepetz et al. (2009). 10
References 11
Altner, H., Thies,G., 1976. The post-antennal organ: a specialized unicellular sensory 12
input to the protocerebrum in apterygotan insects (Collembola). Cell Tissue Res. 13
167, 97-110. 14
Alvarez, T., Frampton, G.K., Coulson,D., 1997. Population dynamics of epigeic 15
Collembola in arable fields: the importance of hedgerow proximity and crop 16
type. Pedobiologia 41, 110-114. 17
Alvarez, T., Frampton, G.K., Coulson,D., 2000.The role of hedgerows in the 18
recolonisation of arable fields by epigeal Collembola.Pedobiologia 44, 516-526. 19
Auclerc, A., Ponge, J.F., Barot, S., Dubs,F., 2009. Experimental assessment of habitat 20
preference and dispersal ability of soil springtails. Soil Biol. Biochem. 41, 1596-21
1604. 22
13
Baatrup, E., Bayley, M.,Axelsen,J.A., 2006.Predation of the mite Hypoaspisaculeifer on 1
the springtail Folsomia fimetaria and the influence of sex, size, starvation, and 2
poisoning.Entomol. Exp. Appl. 118, 61-70. 3
Bauer, T., Christian,E., 1987.Habitat dependent differences in the flight behavior of 4
Collembola.Pedobiologia 30, 233-239. 5
Bengtsson, G., Hedlund, K., Rundgren, S., 1991. Selective odor perception in the soil 6
Collembola Onychiurusarmatus.J. Chem. Ecol. 17, 2113-2125. 7
Bokhorst, S., Phoenix, G.K., Bjerke, J.W., Callaghan, T.V., Huyer-Brugman, F., Berg, 8
M.P., 2012. Extreme winter warming events more negatively impact small rather 9
than large soil fauna: shift in community composition explained by traits not 10
taxa. Glob. Change Biol 18, 1152-1162. 11
Bornebusch, C.H.,1930. The fauna of forest soil.Forstl.Forsøksv.Danmark11, 1-158. 12
Chagnon, M., Hébert, C., Paré, D., 2000.Community structures of Collembola in sugar 13
maple forests: relations to humus type and seasonal trends. Pedobiologia 44, 14
148-174. 15
Chahartaghi, M., Scheu, S.,Ruess,L., 2006.Sex ratio and mode of reproduction in 16
Collembola of an oak-beech forest.Pedobiologia 50, 331-340. 17
Chao, A., 1987. Estimating the population size for capture-recapture data with unequal 18
catchability. Biometrics 43, 783-791. 19
Chernova, N.M., Potapov, M.B., Savenkova, Y.Y.,Bokova,A.I., 2010.Ecological 20
significance of parthenogenesis in Collembola.Entomol. Rev. 90, 23-38. 21
14
Cornwell, W.K., Ackerly, D.D., 2009. Community assembly and shifts in plant trait 1
distributions across an environmental gradient in coastal California. Ecol. 2
Monogr. 79, 109-126. 3
Decaëns, T., Margerie P., Renault, J., Bureau, F., Aubert, M., Hedde, M., 2011. Niche 4
overlap and species assemblage dynamics in an ageing pasture gradient in north-5
western France. ActaOecol. 37, 212-219. 6
Dettner, K., Scheuerlein, A.,Fabian, P.,Schulz, S.,Francke,W., 1996. Chemical defense 7
of giant springtail Tetrodontophorabielanensis (Waga) (Insecta: Collembola). J. 8
Chem. Ecol. 22, 1051-1074. 9
Faber, J.H., Joosse,E.N.G., 1993. Vertical distribution of Collembola in a Pinus nigra 10
organic soil.Pedobiologia 37, 336-350. 11
Gerson, U., 1982. Bryophytes and invertebrates.In:Smith, A.J.E. (Ed.), Bryophyte 12
Ecology Chapman and Hall, London, pp. 291-332. 13
Gillison, A.N., Liswanti,N., 2004. Assessing biodiversity at landscape level in northern 14
Thailand and Sumatra (Indonesia): the importance of environmental 15
context.Agr. Ecosyst. Environ. 104, 75-86. 16
Gisin, H., 1960. CollembolenfaunaEuropas. Muséum d’Histoire Naturelle, Geneva. 17
Graves, G.R., Gotelli, N.J., 1993.Assembly of avian mixed-species flocks in Amazonia. 18
P. Natl Acad. Sci. USA 90, 1388-1391. 19
Hågvar, S., 1982.Collembola in Norwegian coniferous forest soils. I. Relations to plant 20
communities and soil fertility. Pedobiologia 24, 255-296. 21
15
Hågvar, S., 1983.Collembola in Norwegian coniferous forest soils. II. Vertical 1
distribution. Pedobiologia 25, 383-401. 2
Hopkin, S.P., 1997. Biology of the Springtails (Insecta: Collembola). Oxford University 3
Press, Oxford. 4
Hossie, T.J., Murray, D.L., 2010. You can’t run but you can hide: refuge use in frog 5
tadpoles elicits density-dependent predation by dragonfly larvae. Oecologia 163, 6
395-404. 7
Kærsgaard, C.W., Holmstrup, M.,Malte, H., Bayley,M.,2004.The importance of 8
cuticular permeability, osmolyte production and body size for the desiccation 9
resistance of nine species of Collembola. J. Insect Physiol. 50, 5-15. 10
Lepetz, V., Massot, M., Schmeller, D.S., Clobert, J., 2009. Biodiversity monitoring: 11
some proposals to adequately study species’ responses to climate change. 12
Biodivers.Conserv. 18, 3185-3203. 13
Loranger, G., Bandyopadhyaya, I., Razaka, B., Ponge,J.F., 2001. Does soil acidity 14
explain altitudinal sequences in collembolan communities? Soil Biol. Biochem. 15
33, 381-393. 16
Mayfield, M.M., Boni, M.F., Ackerly, D.D., 2009. Traits, habitats, and clades: 17
identifying traits of potential importance to environmental filtering. Am. Nat. 18
174, E1-E22. 19
Mebes, K.H.,Filser,J., 1997. A method for estimating the significance of surface 20
dispersal for population fluctuations of Collembola in arable land.Pedobiologia 21
41, 115-122. 22
16
Nef, L., 1960. Comparaison de l’efficacité de différentes variantes de l’appareil de 1
Berlese-Tullgren. Z. Angew. Entomol. 46, 178-199. 2
Negri, I., 2004. Spatial distribution of Collembola in presence and absence of a 3
predator.Pedobiologia 48, 585-588. 4
Pavoine, S., Vela, E.,Gachet, S.,de Bélair, G.,Bonsall,M.B.,2011. Linking patterns in 5
phylogeny, traits, abiotic variables and space: a novel approach to linking 6
environmental filtering and plant community assembly.J. Ecol. 99, 165-175. 7
Ponge, J.F., 1980. Les biocénoses des collemboles de la forêt de Sénart. In Pesson, P. 8
(Ed.),Actualités d’Écologie Forestière.Gauthier-Villars, Paris, pp. 151-176. 9
Ponge, J.F., 1983. Les collemboles, indicateurs du type d’humus en milieu forestier: 10
résultats obtenus au sud de Paris. ActaOecol.Oec. Gen. 4, 359-374. 11
Ponge, J.F., 1993. Biocenoses of Collembola in atlantic temperate grass-woodland 12
ecosystems. Pedobiologia 37, 223-244. 13
Ponge, J.F., 2000a. Vertical distribution of Collembola (Hexapoda) and their food 14
resources in organic horizons of beech forests. Biol. Fert. Soils 32, 508-522. 15
Ponge, J.F., 2000b. Acidophilic Collembola: living fossils? Contr. Biol. Lab. Kyoto 16
Univ. 29, 65-74. 17
Ponge, J.F., Gillet, S., Dubs, F., Fedoroff, E., Haese, L., Sousa, J.P. Lavelle, P., 2003. 18
Collembolan communities as bioindicators of land use intensification. Soil Biol. 19
Biochem. 35, 813-826. 20
17
Poole, T.B., 1962. The effect of some environmental factors on the pattern of 1
distribution of soil Collembola in a coniferous woodland. Pedobiologia 2, 169-2
182. 3
Ribera, I., Doledec, S.,Downie, I.S. Foster,G.N.,2001.Effect of land disturbance and 4
stress on species traits of ground beetle assemblages. Ecology 82, 1112-1129. 5
Rusek, J., 1980. Morphology of juvenile instars in two Mesaphoruraspecies 6
(Collembola: Tullbergiinae). Rev. Ecol. Biol. Sol17, 583589. 7
Salmon, S., Ponge,J.F., 2001. Earthworm excreta attract soil springtails: laboratory 8
experiments on Heteromurus nitidus (Collembola: Entomobryidae). Soil Biol. 9
Biochem. 33, 1959-1969. 10
Vandewalle, M., de Bello, F.,Berg, M.P.,Bolger, T.,Dolédec, S.,Dubs, F.,Feld, 11
C.K.,Harrington, R.,Harrison, P.A.,Lavorel, S.,da Silva, P.M., Moretti, M., 12
Niemelä, J., Santos, P., Sattler, T., Sousa, J.P., Sykes, M.T., Vanbergen, A.J., 13
Woodcock,B.A., 2010. Functional traits as indicators of biodiversity response to 14
land use changes across ecosystems and organisms. Biodivers.Conserv. 19, 15
2921-2947. 16
Verhoef, H.A., Van Selm,A.J., 1983. Distribution and population dynamics of 17
Collembola in relation to soil moisture.Holarctic. Ecol. 6, 387-394. 18
19
18
1
Code Species name Abundance
Number of
samples
Code Spe cies name Abundance
Number of
samples
ACA Arrhopalites caecus 23 6 MKR Mesaphorura krausbaueri 813 69
AEL Anurida ellipsoides 12 4MMA Mesaphorura macrochaeta 2962 102
AFU Allacma fusca 1 1 MMI Megalothorax minimus 963 105
AGA Allacma gallica 5 2 MMS Mesaphorura massoudi 31 2
APR Arrhopalites principalis 9 7 MMT Micronychiurus minutus 1 1
APY Arrhopalites pygmaeus 13 7MMU Micranurophorus musci 5 1
ASE Arrhopalites sericus 24 8 MPY Micranurida pygmaea 829 72
BPA Brachystomella parvula 1036 33 MSE Micranurida sensillata 2 2
BVI Bourletiella viridescens 50 15 MYO Mesaphorura yosii 158 13
CAL Cyphoderus albinus 3 2 NDU Neonaphorura duboscqi 2 1
CBE Ceratophysella bengtssoni 436 4 NMU Neanura muscorum 115 53
CBI Cryptopygus bipunctatus 2 1 NNO Neonaphorura novemspina 1 1
CDE Ceratophysella denticulata 117 16 NRA Neotullbergia ramicuspis 28 2
CEX Cryptopygus exilis 4 3 OAM Onychiurus ambulans 2 2
CMA Caprainea marginata 9 2 OCI Orchesella cincta 1460 81
CSC Cryptopygus scapelliferus 22 3 OCR Oncopodura crassicornis 5 3
CTH Cryptopygus thermophilus 13 2 OPS Onychiuroides pseudogranulosus 347 13
DFL Deuterosminthurus flavus 7 5 OVI Orchesella villosa 167 44
DFI Deuteraphorura fimetaria 1 1 PAL Pseudosinella alba 279 51
DFU Dicyrtoma fusca 34 19 PAQ Podura aquatica 410 7
DJU Detriturus jubilarius 1 1 PAS Pseudachorutella asigillata 16 5
DMI Dicyrtomina minuta 56 30 PAU Protaphorura aurantiaca 740 24
DTI Desoria tigrina 1192 5 PCA Paratullbergia callipygos 430 53
EAL Entomobrya albocincta 120 17 PDE Pseudosinella decipiens 7 6
ELA Entomobrya lanuginosa 39 13 PLO Pogonognathellus longicornis 7 5
EMA Entomobrya multifasciata 166 11 PMA Pseudosinella mauli 430 48
EMU Entomobrya muscorum 17 15 PMI Proisotoma minima 156 25
ENI Entomobrya nivalis 74 8 PMU Proisotoma minuta 212 10
EPU Entomobryoides purpurascens 11 2 PNO Parisotoma notabilis 6095 180
FCA Folsomia candida 60 9PPA Pseudachorutes parvulus 229 35
FCL Friesea claviseta 67 11 PPE Pseudosinella petterseni 1 1
FMA Folsomia manolachei 6274 101 PPO Pseudosinella pongei 12 4
FMI Friesea mirabilis 109 13 PSE Pseudisotoma sensibilis 1464 12
FPA Folsomides parvulus 145 13 PSU Protaphorura subuliginata 193 20
FQU Folsomia quadrioculata 1810 45 SAQ Sminthurides aquaticus 1 1
FQS Fasciosminthurus quinquefasciatus 2 2 SAS Sminthurides assimilis 78 12
FTR Friesea truncata 361 57 SAU Sminthurinus aureus aureus 1054 75
GFL Gisinianus flammeolus 98 6 SDE Stenaphorurella denisi 32 5
HCL Heterosminthurus claviger 3 1 SEL Sminthurinus elegans 95 21
HIN Heterosminthurus insignis 33 7 SLA Superodontella lamellifera 4 3
HMA Heteromurus major 594 71 SMA Sminthurides malmgreni 591 43
HNI Heteromurus nitidus 28 18 SNI Sminthurus nigromaculatus 16 9
HPU Hypogastrura purpurescens 1 1 SPA Sminthurides parvulus 82 13
IAN Isotomurus antennalis 1 1 SPS Subisotoma pusilla 82 5
IMI Isotomiella minor 2136 116 SPU Sphaeridia pumilis 1566 107
IPA Isotomurus palustris 1483 101 SQU Stenaphorurella quadrispina 7 3
IPR Isotomodes productus 4 1 SSC Sminthurides schoetti 401 44
ISP Isotomodes sp. 2 2 SSE Schaefferia sexoculata 1 1
IVI Isotoma viridis 54 13 SSI Sminthurinus aureus signatus 2407 97
KBU Kalaphorura burmeisteri 30 5 STR Sminthurinus reticulatus 1 1
LCU Lepidocyrtus curvicollis 72 26 SVI Stenacidia violacea 6 2
LCY Lepidocyrtus cyaneus 889 35 TBO Tomocerus botanicus 35 9
LLA Lepidocyrtus lanuginosus 3399 160 TMI Tomocerus minor 312 45
LLI Lepidocyrtus lignorum 565 63 VAR Vertagopus arboreus 788 31
LLU Lipothrix lubbocki 15 6 WAN Willemia anophthalma 577 35
LPA Lepidocyrtus paradoxus 2 2 WBU Willemia buddenbrocki 5 3
LVI Lepidocyrtus violaceus 4 4 WIN Willemia intermedia 1 1
MAB Micraphorura absoloni 3 2 WNI Willowsia nigromaculata 3 1
MBE Mesaphorura betschi 12 6 WPO Wankeliella pongei 2 1
MGR Monobella grassei 32 14 XBR Xenylla brevisimilis 2 1
MHG Mesaphorura hygrophila 1 1 XGR Xenylla grisea 361 19
MHY Mesaphorura hylophila 633 42 XSC Xenylla schillei 18 5
MIN Megalothorax incertus 12 9 XTU Xenylla tullbergi 4673 68
MIT Mesaphorura italica 21 9XXA Xenylla xavieri 33 5
Table 1. Codes and species names of springtails collected in the Senart forest from 1973 to 1977, total abundance, and number of samples in
which the species was found. Species names according to Fauna Europaea 2011
19
1
2
Number of
samples
F1 F2
Number of
samples
F1 F2
Autumn 96 0.045 0.084 Hornbeam 42 -0.046 0.037
Winter 108 0.162 0.051 Linden 22 -0.012 0.001
Spring 88 0.091 0.035 Maple 8 0.053 0.049
Summer 46 0.109 0.030 Ash 8 0.011 0.026
Grassland 50 0.136 -0.021 Cherry 9 0.097 -0.066
Woodland 279 -0.124 0.006 Elm 3 0.170 0.057
Heathland 9 0.064 0.029 Elder 3 0.112 -0.012
Ditch/brook 44 0.106 0.059 Hazel 11 -0.027 -0.040
Pond 64 0.140 0.056 Pine 12 0.007 0.021
Plain ground 230 0.027 -0.011 Calluna 6 0.009 0.076
Water 107 0.078 0.023 Blackberry 5 0.124 -0.002
Sunlight 141 0.230 0.074 Ivy 4 0.013 0.036
pH < 5 32 0.030 0.102 Peat moss 18 0.022 0.071
pH 5-6 35 0.024 -0.003 Hair moss 5 0.183 0.008
pH > 6 32 -0.052 -0.069 Feathermoss 8 0.030 0.102
Limestone 48 0.002 -0.009 Liverwort 1 0.156 -0.041
Sand 20 -0.062 -0.009 Lichens 4 0.082 0.140
Pebbles 23 0.057 -0.004 Algae 3 0.155 0.023
Mull 57 -0.121 -0.036 Bracken 21 0.014 0.044
Moder 24 -0.090 0.046 Purple moor grass 21 0.108 0.039
Mor 2 0.086 -0.027 Hair-grass 5 0.084 -0.028
Hydromull 6 -0.019 -0.019 Fescue-like grass 8 0.201 -0.016
Hydromoder 3 -0.030 0.004 Rushes 6 0.219 -0.013
Hydromor 3 -0.017 0.021 Waterlilies 10 0.132 0.021
Trunk 33 0.108 0.143 Hawksbeard 1 0.121 -0.003
Herbs (aerial parts) 58 0.296 0.077 Sedges 4 0.078 0.016
Mosses (aerial parts) 74 0.163 0.146 Wood anemone 20 -0.037 0.045
Superficial soil 17 0.146 0.003 Bluebell 20 -0.037 0.045
Litter 80 0.213 0.068 Duckweed 1 0.121 0.007
Humus 41 0.136 0.043 Mustard 1 0.062 0.024
Organo-mineral soil 18 -0.049 -0.032 Chamomile 1 -0.013 -0.008
Mineral soil 68 -0.172 -0.111 Chickweed 9 0.053 -0.078
Mole hill 4 0.028 -0.007 Yarrow 4 0.003 -0.051
Vertebrate dung 3 0.181 0.010 Nettle 5 -0.046 -0.047
Garbage deposits 11 -0.045 0.058 Mercury 16 0.036 0.022
Wood 35 0.093 0.132 Solomon's seal 8 0.053 0.049
Earthworm casts 7 -0.036 -0.013 Wheat 7 -0.011 -0.051
Tree roots 5 0.067 0.053 Buttercup 1 0.142 -0.041
Herb roots 8 0.083 -0.006 Knotweed 1 0.091 0.037
Oak 142 -0.048 0.015 Clover 5 -0.091 -0.038
Birch 41 0.113 0.026 Mint 1 0.123 0.001
20
1
2
Trait Attribute
Number of
species
Mode of reproduction Parthenogenesis dominant 36
Sexual reproduction dominant 89
Body size Small 86
Medium 28
Large 14
Body form Slender 92
Stocky 6
Spheric 30
Body color Pale-coloured 60
Bright-coloured 30
Dark-coloured 38
Scales Absent 109
Present 19
Antenna size Short 65
Long 63
Leg size Short 61
Long 67
Furcula size Absent or vestigial 35
Short 25
Long 68
Eye number 0 42
1-5 24
> 5 62
Pseudocella Absent 105
Present 23
Post-antennal organ Absent 69
Simple 21
Compound 38
Trichobothria Absent 72
Present 56
Table 3. Trait attributes of the 128 springtail species collected in the Sénart
forest, and number of species where attributes were found
21
Figure legends 1
Figure 1.Canonical correspondence analysis of species trait attributes: projection of 2
traits (a) and habitat indicators (b) in the plane of the first two canonical factors 3
F1 and F2. 4
5
22
1
Fig. 1 2
Parthenogenesis dominant
Sexual reproduction
dominant
Small Medium
Large
Slender body
Stocky body
Spheric body
Pale-colored body Bright-coloured body
Dark-colored body
Scales absent
Scales present
Short antennae
Long antennae
Short legs
Long legs
Furcula absent or vestigial
Short furcula
Long furcula
Eyes absent
Eyes 1-5
Eyes > 5
Pseudocella absent
Pseudocella present
Post-antennal organ
absent
Post-antennal organ
simple
Post-antennal organ
compound
Trichobothria absent
Trichobothria present
F1
F2
a)
Autumn
Winter
Spring Summer
GRASSLAND
WOODLAND
HEATHLAND
Ditch/broo k Pond
Plain ground
Water
Sunlight
Limestone
Sand Pebbles
Trunk
Mole hill
Vertebrate dung
Garbage deposits
Superficial soil
Litter
Humus
Organo-miner al soil
Mineral soil
Wood
Earthworm casts
Tree roots
Herb roots
pH < 5
pH 5-6
pH > 6
Mull
Moder
Mor
Hydromull
Hydromoder
Hydromor
Oak
Birch
Hornbeam
Linden
Maple
Ash
Cherry
Elm
Willow
Elder
Hazel
Pine
Calluna
Blackberry
Ivy
Mosses (aerial parts)
Peat moss
Hair moss
Feathermoss
Liverwort
Lichens
Algae
Herbs (aerial parts)
Bracken Purp le moor grass
Hair-grass
Fescue-like grass
Rushes
Waterlilies
Hawksbeard
Sedges
Wood anemone
Bluebell
Duckweed
Mustard
Chamomile
Chickweed
Yarrow
Nettle
Mercury
Solomon' s seal
Wheat
Buttercup
Knotweed
Clover
Mint
F1
F2
b)
... Previous studies showed that moister micro-climate conditions provided by tree canopies affects positively Collembola richness and influences community structure towards a higher relative abundance of euedaphic species (Martins da Silva et al., 2012, 2016Salmon et al., 2014;Rossetti et al., 2015;Joimel et al., 2021). In contrast, the relative abundance of epigeics species is usually higher in open habitats, such as grasslands and open landscapes (Alvarez et al., 2000;Ponge et al., 2006;Martins da Silva et al., 2012;Salmon et al., 2014;Harta et al., 2021), as they are usually faster dispersers and more resistant to desiccation and disturbances in the upper soil layers (Verhoef and Van Selm, 1983;Alvarez et al., 1999;Ponge et al., 2006;Bokhorst et al., 2012;Salmon and Ponge, 2012). Yet, the findings above are focused on the Mediterranean, Temperate or Boreal regions. ...
... We expected that community traits of life-form would accurately indicate the influence of tree canopy on collembolan functional diversity. Based on previous observations in temperate and Mediterranean regions (e.g., Ponge et al., 2006;Salmon and Ponge, 2012;Salmon et al., 2014;Martins da Silva et al., 2016), we predicted that tree canopy influence collembolan richness and diversity in forested environments due to an increase in the relative abundance of euedaphics. Most euedaphics are expected to be particularly favored by local suitable microhabitat conditions promoted by closed canopy environments, as they are slower moving and are not equipped with traits protecting them against desiccation (Ponge et al., 2006;Bokhorst et al., 2012;Winck et al., 2017), such as larger body size or cuticular impermeability (Salmon and Ponge, 2012). ...
... Based on previous observations in temperate and Mediterranean regions (e.g., Ponge et al., 2006;Salmon and Ponge, 2012;Salmon et al., 2014;Martins da Silva et al., 2016), we predicted that tree canopy influence collembolan richness and diversity in forested environments due to an increase in the relative abundance of euedaphics. Most euedaphics are expected to be particularly favored by local suitable microhabitat conditions promoted by closed canopy environments, as they are slower moving and are not equipped with traits protecting them against desiccation (Ponge et al., 2006;Bokhorst et al., 2012;Winck et al., 2017), such as larger body size or cuticular impermeability (Salmon and Ponge, 2012). Also, hemi-and euedaphics can migrate to deeper soil layers to avoid desiccation, while epigeic collembolans are generally more adapted to the drier soil surface holding thinner litter layers compared to forested sites (Alvarez et al., 2000;Makkonen et al., 2011;Martins da Silva et al., 2012;Salmon et al., 2014;Meyer et al., 2021). ...
... The research on functional traits has attracted extensive attention in the research of soil animal ecology because it can intuitively reflect the adaptation mechanism of organisms to the environment [20]. It was found that a large body size is more adaptable to severe climate change in high-latitude areas, but there is no significant impact of climate change on the Collembola community based on species classification [21]. ...
... Concerning the relationship between the sensory ability of the springtail and the season as well as the soil depth, most springtails possess the PAO, indicating that the PAO can help to improve the springtail's sensory capacity [20]. It is also important to note that the soil substratum contains an increasing percentage of springtails without the PAO. ...
... Then, in such a situation, its highly developed sensory organ is advantageous [63]. In addition, the larger porosity, rather than the soil with low bulk density, can allow the huge springtail to pass through, explaining the negative association between the antenna/body, PAO, and soil bulk density [20]. ...
Article
Full-text available
The group of soil arthropods known as Collembola is characterized by its abundance and sensitivity to environmental changes. They are ideal an species for soil indicators. In order to clarify the effects of species invasion and inundation on the Collembola community in coastal mudflat wetlands, the correlation between the collembolan functional traits and environmental factors was studied in Shanghai Jiuduansha Wetland National Nature Reserve for the first time. Five sample plots, including three vegetations—Spartina alterniflora (an invasive species), Phragmites australis, and Zizania latifolia—were set up following the differences in vegetation types and between high and low tidal flats. Data on the diversity of the Collembolan species and their functional traits were collected and combined with the soil physicochemical properties and vegetation environment factors in different tidal flats. The key findings and conclusions of the study are as follows: a total of 18 species, four families, and three orders make up the obtained Collembola, two species of Proisotoma are dominant species that account for 49.59% and 24.91% of the total, respectively. The maintenance of the species diversity of Collembola is disturbed by the higher conversion efficiency of Spartina alterniflora rather than Phragmites australis with lower organic carbon (C) content and higher total nitrogen (N) content. The primary environmental variables influencing species distribution were the C/N ratio, total N, and bulk soil density. The bulk density of the soil impacts the movement and dispersal of the functional traits. The depth of the soil layer is related to the functional traits of the sensory ability. The analysis of the functional traits and environment is fairly helpful in exploring how species respond to their environment and offers a better explanation for the habitat selection of Collembola.
... The functional traits of springtails were measured directly viduals collected in the field. We selected traits related to dispersal ability, life fo habitat preference of springtails [20,42,43]. The selected traits are summarized in We used observed trait data to calculate the functional diversity (FD) indices prop Villéger et al. [44]. ...
... The functional traits of springtails were measured directly on individuals collected in the field. We selected traits related to dispersal ability, life form and habitat preference of springtails [20,42,43]. The selected traits are summarized in Table 1. ...
Article
Full-text available
We used trait-based approaches to reveal the functional responses of springtails communities to organic matter inputs in a rubber plantation in Côte d’Ivoire. Pitfall traps were used to sample springtails in each practice. The results showed that the total abundance of springtails increased significantly with the amount of organic matter (R0L0 < R2L1). Larger springtails (body length, furca and antennae) were observed in plots with high organic matter. Practices with logging residues and legume recorded the highest functional richness. The principal coordinate analysis showed different functional composition patterns between practices with logging residues (R1L1 and R2L1) and those without inputs (R0L0 and R0L1). This difference in functional composition (PERMANOVA analysis) was related to the effect of practices. These results highlight the pertinence of the functional trait approach in the characterization of springtail communities, a bioindicator of soil health, for organic matter management practice.
... Light may be a key cue for avoiding a dry habitat for terrestrial cave-dwelling arthropods such as spiders and carrion beetles. This may also explain why eyeless amphipods (9) and soil animals (39,40) remain sensitive to light. However, it is unnecessary for water beetles (14) to detect light to find humid environments. ...
Article
Full-text available
Subterranean animals living in perpetual darkness may maintain photoresponse. However, the evolutionary processes behind the conflict between eye loss and maintenance of the photoresponse remain largely unknown. We used Leptonetela spiders to investigate the driving forces behind the maintenance of the photoresponse in cave-dwelling spiders. Our behavioral experiments showed that all eyeless/reduced-eyed cave-dwelling species retained photophobic response and that they had substantially decreased survival at cave entrances due to weak drought resistance. The transcriptomic analysis demonstrated that nearly all phototransduction pathway genes were present and that all tested phototransduction pathway genes were subjected to strong functional constraints in cave-dwelling species. Our results suggest that cave-dwelling eyeless spiders still use light and that light detection likely plays a role in avoiding the cave entrance habitat. This study confirms that some eyeless subterranean animals have retained their photosensitivity due to natural selection and provides a case of mismatch between phenotype and genotype or physiological function in a long-term evolutionary process.
... Pseudocelli is related to the chemicals excretion that repel predators, in turn to compensate for the difficulty to escape through active movement in deep soil (Negri, 2004). In the same way, the presence of post antennal organ compensate for the absence of ocelli (Salmon and Ponge, 2012). These selection of euedaphic traits caused by the strong modifications of vegetation structure (i.e., plant height and canopy openness), diversity, and composition are in line with previous studies that detected similar patterns on Europe Salmon et al., 2014) and on Southern Brazil native grasslands . ...
Thesis
Full-text available
Grassland afforestation with fast-growing exotic species is the category of land use change that has grown the most in recent years in the Pampa biome. Studies around the world have shown negative effects of this type of land-use change on biodiversity. However, information about the effects on soil communities is still scarce. The organisms that live in the soil are extremely diverse and contribute to a wide range of ecosystem processes that underlies essential ecosystem services, such as carbon stock, regulation of the water cycle and food production. Therefore, soils with high diversity are healthy and present major functionality, promoting ecosystem functioning of and human welfare. At this context, in this dissertation we aim to investigate the effects of converting native grasslands of Pampa biome into Eucalyptus plantations on soil microbiota and fauna (model group: Collembola) which are considered indicators of soil functionality. The study was carried out in 20-paired sampling units in Pampa biome (10 grasslands and 10 Eucalyptus plantations) distributed in four municipalities in the state of Rio Grande do Sul. In each sampling unit, a 250 m transect was established where 10 points were sampled to collect soil in the 0-5 cm layer. The soil from five samples was used to determine soil pH and gravimetric moisture and the size and activity of the microbial community (soil basal respiration of and enzyme activity). Soil temperature and plant richness were measured in situ. The soil from the other five points was used for extraction (14 days) of collembolans by the Berlese-Tüllgren funnel method, which were identified and described by their functional attributes. The effects of land use change were tested using generalized mixed linear models. Our results demonstrate that grassland conversion led to a reduction in plant richness, which tends to decrease litter quality and diversity that enters the. This effect on soil trophic chains, associated with the a more acidic and drier soil may explain the drastic reduction observed in the size and activity of the microbial community that further can impair the functions related to carbon and nitrogen cycling, such as decomposition, soil fertility and carbon stock. The community of collembolans presented reduction of richness and had their taxonomic and functional composition altered, which can result in changes in the food webs and, consequently, in their functions. Our most unexpected result was the increase of functional diversity at the plantations by the increase on collembolans with forestry-adaptive traits. In addition, there was a decrease in functional redundancy, which tends to shape Eucalyptus plantations in a less stable state what may harm soil functionality towards environmental changes. However, the indicators of soil functionality were lower in Eucalyptus plantations. The results obtained in this study are a step forward to help fill this gap in ecological knowledge, supporting the elaboration of conservation policies in the Pampa grasslands with scientific- based information.
... Here, we sampled only one forest type (i.e., pure spruce forests) just 2 years after the disturbance, thus it is not surprising that geographic zone (i.e., related to historical patterns) is still the most important factor. On the other hand, small-scale variables, such as habitat type, still play a great importance as predictors for species composition, in agreement with previous studies (Salmon & Ponge, 2012;Sterzy nska & Skłodowski, 2018;Arribas et al., 2021). ...
Article
Windstorms and salvage logging lead to huge soil disturbance in alpine spruce forests, potentially affecting soil-living arthropods. However, the impacts of forest loss and possible interactions with underlying ecological gradients on soil microarthropod communities remain little known, especially across different environmental conditions. Here we used DNA metabarcoding approach to study wind-induced disturbances on forest communities of springtails and soil mites. In particular, we aimed to test the effect of forest soil disturbance on the abundance, richness, species composition, and functional guilds of microarthropods. We sampled 29 pairs of windfall-forest sites across gradients of elevation, precipitation, aspect and slope, 2 years after a massive windstorm, named Vaia, which hit NorthEastern Italy in October 2018. Our results showed that wind-induced disturbances led to detrimental impacts on soil-living communities. Abundance of microarthropods decreased in windfalls, but with interacting effects with precipitation gradients. Operative Taxonomic Units (OTU) richness strongly decreased in post-disturbance sites, particularly affecting plant-feeder trophic guilds. Furthermore, species composition analyses revealed that communities occurring in post-disturbance sites were different to those in undisturbed forests (i.e., stands without wind damage). However, variables at different spatial scales played different roles depending on the considered taxon. Our study contributes to shed light on the impacts on important, but often neglected arthropod communities after windstorm in spruce forests. Effects of forest disturbance are often mediated by underlying large scale ecological gradients, such as precipitation and topography. Massive impacts of stronger and more frequent windstorms are expected to hit forests in the future; given the response we recorded, mediated by environmental features, forest managers need to take site-specific conservation measures.
... However, most springtails in Eucalyptus plantations were eyeless, and lacked body pigmentation, which are functional traits commonly found in canopy-closed habitats. Moreover, most springtails in Eucalyptus plantations also had post antennal organ and pseudocelli, which are repugnatorial glands against predators, that compensate for the difficulty to escape through active movement (Negri 2004) and for the absence of ocelli (Salmon and Ponge 2012). This replacement from species with open canopy-adapted traits to those with closed canopy-adapted traits driven by afforestation is in line with previous studies conducted in Europe Salmon et al., 2014) and in southern Brazil (Winck et al., 2017). ...
Article
Full-text available
Afforestation of subtropical grasslands increased during the last decade and their impacts on soil biota and functions are not well known. Here we investigated the effects of grassland afforestation with Eucalyptus on soil Collembola communities in southern Brazil. By using paired transects (native grassland vs. plantation), we assessed Collembola taxonomic and functional composition and diversity and their relations with plant diversity and soil microbial and physical–chemical parameters. Eucalyptus plantations significantly decreased plant richness, soil pH, and soil temperature. Microbial biomass, soil basal respiration and enzyme activities were negatively affected by grassland afforestation, which may indicate an effect on decomposition, soil fertility, and carbon stock. At a regional scale, grassland afforestation diminished the richness, and at a local scale changed the taxonomic and functional composition of Collembola communities. An environmental filtering mechanism is suggested triggering trait turnover from open (grassland) to a closed canopy habitat (tree plantation). While grassland presented high abundance of larger, eyed, and pigmented Collembola, Eucalyptus favored smaller, blind, and non-pigmented Collembola with shadow-adapted sensorial structures. Our results on afforestation effects in subtropical grasslands are similar to those described for temperate zone, and may underpin the development of conservation strategies for land-use on subtropical grasslands.
... Secondly, a low score could arise from individuals not taken into account in the analysis if they were not identified to species level, notably juveniles. The proportion of juveniles (when not identifiable) in Collembola community analyses is a wellknown limitation (Salmon and Ponge, 2012), and applying the CWM metric using literature-based trait information does not account for this part of the community. This is problematic, as the proportion of Fig. 3. -Representation of two theoretical cases of thermal niche: species 1 (blue, left panel) presents a standard smooth curve of performance related to temperature and species 2 (orange, right panel) presents a heart-shaped pattern due to the alternance of its normal form (A) and ecomorphic form (B/C). juveniles might be a signal of primary importance in Mediterranean areas in which springtail populations display several phenotypical adjustments, such as different offspring optima or over-summering resistant eggs (Poinsot-Balaguer, 1984). ...
Article
Collembola are a widespread class of arthropods that live mostly in the soil or on its surface. Communities of Collembola have notably been used as bioindicators of several environmental factors such as pollution or land use. Recently, they have also opened perspectives for monitoring the effects of projected climate change on soil biodiversity, in particular through the responses of their traits. Collembola are known to exhibit several morphological variations throughout their lifecycle (other than growth and sexual dimorphism). One of these phenomena, ecomorphosis, has been described as a survival strategy mainly triggered by elevated temperature. This could be of interest in analysing collembolan adaptation to climate change, yet studies on this are - to date - sparse in international literature in English. To begin to address this gap, we conducted a literature review that enabled us to: (i) identify the concepts behind the ecomorphosis strategy, (ii) list the collembolan species known to display ecomorphosis, (iii) summarize its consequences on individuals; and (iv) analyse its theoretical implications for community ecology and functional ecology. We then discussed its potential use as a proxy for climate adversity. We thus suggested using the ability of a species to display ecomorphosis as a trait, and in the future test its responses in collembolan communities along climatic gradients. Considering the recent inputs of taking into account intraspecific trait variability in community ecology, we advocate for a better understanding of such eco-physiological strategies in order to improve our hypotheses-based approaches in trait-environment relationships. In a context of rapid global climate change, our findings may provide insights into functional responses to climatic gradients in Collembola, and hopefully contribute to stimulate discussion in other soil fauna biological models.
... Previous studies on the response of collembolans to land-use intensification and urbanization showed that their response was related to shifts in functional traits related to both life-history, behaviour and morphology (de Bello, 2012;Potapov et al., 2022;Santorufo et al., 2015). In our study, epedaphic species with patterned pigmentation, well-developed furca, large number of ocelli, and fast dispersal increased in open greenspace habitats, and this is consistent with findings of (da Silva et al., 2016Silva et al., , 2012Salmon and Ponge, 2012). By contrast, soil-dwelling euedaphic species with white pigmentation, absent furca and ocelli, and slow dispersal ability were more abundant in forest soils pointing towards feeding on resources deeper in the soil such as those associated with roots and indicating a more porous soil allowing to access these resources (Heiniger et al., 2015;Santorufo et al., 2014). ...
Article
Urban regions are rapidly expanding worldwide resulting in biotic homogenization and loss of ecological functions in urban ecosystems due to management practices targeting at satisfying aesthetic and health demands of urban residents. These practices also modify living conditions and food recourses of soil invertebrates thereby affecting the structure and functional diversity of soil animal communities including collembolans. Here, we assessed the response of the community composition and functional diversity of collembolans as a major component of soil food webs to urbanization (suburban vs urban region) and greenspace types (including forest and four park-associated greenspaces: lawn, lawn with shrubs, lawn with trees, and lawn with shrubs and trees). Our results highlight that both urbanization and greenspace type significantly affect soil properties and community structure of collembolans. The negative effect of urbanization and park-associated greenspaces on species and functional composition of collembolan communities were likely due to both changes in soil abiotic conditions and bacterial community composition, whereas the reduction of collembolan functional traits likely resulted from changes in soil abiotic conditions and fungal community composition. In park-associated greenspaces richness and diversity of bacterial communities were highest in lawns with trees and lowest in forests. By contrast, species richness and diversity of fungal communities were highest in lawns with shrubs, but, similar to bacteria, lowest in forests. Community composition and functional traits of collembolans were more homogeneous in urban than suburban greenspaces pointing to reduced functioning of collembolan assemblages in urban areas. Overall, our results suggest that changes in soil properties and bacterial communities caused by urbanization and greenspace type are important factors contributing to taxonomic homogenization of collembolan communities, while the loss of functional traits of collembolan communities in urban greenspaces is likely caused by changes in soil properties and fungal community composition.
Article
Full-text available
Atmobiotic species should escape by a single long distance, rapid jump that catapults them out of the range and field of view of an attacker. Litter-inhabitants execute short jumps but more jumps per unit time. The bark-inhabiting Entomobrya corticalis behaved differently. Even when touched it escapes by running quickly and very rarely employs a short jump. Tree inhabitants are characterized by greater resistance to desiccation while inhabitants of deeper litter layers show low resistance. Among species that live on or in the uppermost litter layer there is no conspicuous difference in resistance to dryness between long distance and short distance jumping species. -from Authors
Article
The average depth of most species was in the upper 6 cm in all soils. In podzols, only Tullbergia callipygos, T. quadrispina, Karlstejnia norvegica and Wankeliella mediochaeta had their mean depth consistently <6 cm (in the mineral layer). Except for the larger species restricted to the surface layers, many species showed considerable variation in depth distribution between both soils and seasons. A considerable overlap existed between hemiedaphic and euedaphic species. In most species, there was no consistent difference between mean depth in podzol and brown earth soils. A system of 5 depth levels, each with characteristic species, is described. A large number of factors probably influence the vertical position of a species. Each species continually adjusts its vertical distribution to optimize its life and reproductive conditions. This also implies that the community structure continually undergoes vertical changes. -from Author
Article
Recent studies have shown marked differences between Collembola communities in a field and a nearby fallow which had both been managed in a similar manner before management conversion. To investigate the extent to which these differences could be attributed to surface dispersal, small pitfall traps, provided with fences, were used to deter mine directions of dispersal. The dominant Collembola trapped in pitfalls were Isotomurus palustris, Lepidocyrtus cyaneus and Sminthuridae. Isotoma viridis and the Genus Folsomia were subdominant or seldom caught in pitfall traps. The number of Collembola trapped in pitfalls decreased from field to fallow, suggesting a lower activity under more favourable conditions in the fallow. The dispersal direction of Collembola depended on barrier position (field, borderline, or fallow). The most striking differences between the soil communities, within the Genus Folsomia, could not be assigned to surface dispersal, whereas the higher or increasing proportions of Isotomurus palustris and Sminthuridae in the fallow compared to the wheat field could partly be explained by migration.
Article
In this paper we test whether the morphology and life traits of species (in our case ground beetles of the family Carabidae) can be related to the main underlying axes of environmental variability of their habitats. Sites were selected a priori to maximize two gradients: land use as a general measure of disturbance characterized by an index of land management, and habitat adversity or stress as characterized by elevation and vegetation structure. The underlying environmental axes and the relationships of the morphology and life traits of the species with them were investigated using RLQ analysis, a multivariate ordination method able to relate a species trait table to a site characteristics table by way of a species abundance table. The first environmental axis was highly statistically significant and explained most of the variability. It was strongly negatively related to the intensity of land management, and positively related to increasing elevation and a set of variables reflecting vegetation stress. Two predictions were tested and found to be valid in the studied system: in highly managed lowland sites species were smaller, and the frequency of macropterous species (with better dispersal abilities) was higher. Other traits also showed significant relationships with the main environmental axis: in the intensively managed lowland sites species had broader bodies, longer trochanters, and wider femora (characters associated with plant eaters), were paler in color, overwintered only as adults, bred in spring or autumn, and were active in summer. We conclude that the ground beetle assemblages of the studied sites respond in a similar way to the same underlying environmental factors. This allows the precise definition of functional groups, which can be used to characterize functional diversity and its relationships with changes in land management.
Chapter
These thousand eyes have also been looking upon naturalists for quite a while, but only few have looked back. Usually they were zoologists interested in specific groups which live on, in or under bryophytes; in the role these animals play during initial land colonization by cryptogams; in freshwater associations and in diverse other aspects. Botanists have published far fewer observations, these dealing mainly with fertilization or spore dispersal by invertebrates. Although scattered and uneven, taken as a whole the compiled data offer suggestive insights into the relationships between bryophytes and invertebrates, especially in regard to their co-evolution.
Article
During spring 1996, the temporal and spatial dynamics of epigeic Collembola, in southern UK, were studied in fields planted with spring barley and vining peas. Composition of the fauna varied according to crop type and position within the field. Proximity of a hedgerow had a major impact upon population dynamics, with more species and a greater over all abundance near field edges. It was the major factor governing the abundance of Isotomurus palustris and Bourletiella hortensis, whereas crop type was most important in determining numbers of Jeannenotia stachi, Sminthurus viridis, Sminthurinus elegans and Lepidocyrus spp. Temporal variation in species abundance was also clear: abundances of J. stachi, B. hortensis, I. palustris and Lepidocyrtus spp. decreased through the sampling period while those of S. viridis and Isotoma viridis increased. Overall, the management practices associated with the two crops were an important determinant of collembolan abundance but not of species composition.