You are on page 1of 12

Community Structure, Distribution and Population Dynamics of Entomobryidae

(Collembola)
Author(s): J. J. Vegter, E. N. G. Joosse and G. Ernsting
Source: Journal of Animal Ecology, Vol. 57, No. 3 (Oct., 1988), pp. 971-981
Published by: British Ecological Society
Stable URL: https://www.jstor.org/stable/5105
Accessed: 09-11-2019 21:47 UTC

JSTOR is a not-for-profit service that helps scholars, researchers, and students discover, use, and build upon a wide
range of content in a trusted digital archive. We use information technology and tools to increase productivity and
facilitate new forms of scholarship. For more information about JSTOR, please contact support@jstor.org.

Your use of the JSTOR archive indicates your acceptance of the Terms & Conditions of Use, available at
https://about.jstor.org/terms

British Ecological Society is collaborating with JSTOR to digitize, preserve and extend access
to Journal of Animal Ecology

This content downloaded from 200.24.17.12 on Sat, 09 Nov 2019 21:47:59 UTC
All use subject to https://about.jstor.org/terms
Journal of Animal Ecology (1988), 57, 971-981

COMMUNITY STRUCTURE, DISTRIBUTION


AND POPULATION DYNAMICS OF
ENTOMOBRYIDAE (COLLEMBOLA)
BY J. J. VEGTER, E. N. G. JOOSSE AND G. ERNSTING

Department of Ecology and Ecotoxicology, Vrije Universiteit, 1007 MC Amsterdam,


The Netherlands

SUMMARY

(1) Entomobryid Collembola were sampled over 4 years in eight woodlan


differing in soil and litter type.
(2) Ranked abundance curves for each community were approximately l
semi-log scale. Evenness was highest in the youngest woodland which was
polder. In older mainland habitats the first three species in the sequence show
extreme pattern of dominance.
(3) Population density of hygric Tomocerus species was lowest in woodlands
heterogeneous moisture regimes whereas population density of the droug
species Orchesella cincta was high in these areas.
(4) The degree of spatial aggregation of a species appeared to be unrela
drought sensitivity or to spatial heterogeneity in moisture regime of the hab
general depended on population density.
(5) The intensity of temporal fluctuations in population number was density
and unrelated to environmental heterogeneity.

INTRODUCTION

Soil humidity and humus type are generally considered to be the most i
determining community composition and population dynamics in m
and litter arthropods, such as Collembola and mites (Gisin 1943; P
Wauthy & Lebrun 1980). A number of small-scale field and laborat
shown that habitat selection of various Collembola species depends
sensitivity (Verhoef& Witteveen 1980; Verhoef& Van Selm 1983). It is un
however, whether differences in drought sensitivity alone can fully exp
distributional patterns (Vegter 1983).
Since short-term studies in a given habitat focus on spatial heterogeneit
of occasional dry periods, the effects of long-term variability in soi
remain to be explored. Temporal variability might even be more import
heterogeneity since unfavourable patches can easily be avoided by d
organisms, but survival of unfavourable periods requires specific ad
1981; Verhoef & Li 1983). To study the relation between temporal v
moisture, community structure and population dynamics of litter-d
(Entomobryidae), a sampling programme was set up covering eight d
habitats, which were sampled over 4 years. A preliminary report of the
programme has been given by Joosse (1981). The present paper will give
971

This content downloaded from 200.24.17.12 on Sat, 09 Nov 2019 21:47:59 UTC
All use subject to https://about.jstor.org/terms
972 Population dynamics of Collembola

analysis of spatial patterns, fluctuations in numerical abundance and community


structure of entomobryid Collembola in these habitats. The analysis aims at testing the
hypothesis that drought-sensitive species show a higher degree of spatial aggregation,
larger fluctuations in numbers and lower abundance than drought-tolerant species.

MATERIALS AND METHODS

Study areas and sampling


The eight woodland areas covered a range of soil and litter types commonly
encountered in the Netherlands. Typical situations were: (i) deciduous woodland on
sandy soil with a well developed litter and humus layer of the mor type; (ii) dry and open
structured coniferous woodland on sandy soil with mor humus; and (iii) deciduous
woodland on clay, with a mull humus type. Two woodlands were selected in each
category. The remaining two woodlands were to some extent atypical: wet birch
woodland on sandy soil on the Dutch Wadden Island Schiermonnikoog, and a very
young poplar plantation on a former sand bank in a polder.
All woodlands were sampled in February, April, July, October and December in 1973-
76. On each sampling occasion, twenty-four samples of litter, humus and top layer of the
soil were taken at random in each woodland area with a soil corer (diameter 10 cm,
internal surface 72 54 cm2). The samples were transported in closed boxes and extracted in
a modified Berlese-Tullgren apparatus, described by van Straalen & Rijninks (1982), after
determination of their fresh weight. After the extraction period of 1 week the samples were
re-weighed to establish their water content. The extracted animals were stored in a
mixture of alcohol 95% (750 parts), ether (250 parts), acetic acid (30 parts) and formalin
40% (3 parts) and counted after identification. The taxonomy is based on Gisin (1960,
1964).

Analysis of the data


Temporal variability in population numbers in a given woodland area is measured by
the variance of log (x + 1) transformed total counts (x) for each species between sampling
dates. Preliminary analysis of the data indicated that this transformation effectively
normalizes the data, except at very low densities (cf. Gerard & Berthet 1966). The
geometric means of the total counts (obtained by back transformation) are used as
measures of abundance.
At a given mean density (arithmetic) the degree of spatial aggregation of a population is
reflected by the variance of the counts between soil cores. Methods to estimate degree of
aggregation without regard to density abound (Pielou 1977). Given the structure of the
sampling programme, where several estimates for means and spatial variances are
available for each population, the most convenient approach is the one advocated by
Taylor, Woiwod & Perry (1978). This method examines the relation between mean and
variance using the formula: S2 = a mb, where S2 is the variance, m is the mean and a and b
are parameters to be estimated by linear regression of log-transformed means and
variances. The parameter a measures spatial aggregation at a specified mean density
m = 1, whereas b measures the degree of density dependence in spatial pattern (Taylor &
Taylor 1977).
The average spatial variability (SV) in soil moisture of a woodland can be measured by
the geometric mean of the standard deviations (antilog log Sy) in moisture content (Sy) of

This content downloaded from 200.24.17.12 on Sat, 09 Nov 2019 21:47:59 UTC
All use subject to https://about.jstor.org/terms
J. J. VEGTER, E. N. G. JOOSSE AND G. ERNSTING 973

TABLE 1. Habitat characteristics and moisture regime of the woodland areas studied

Average
moisture
Study area content (% Spatial Temporal
(abbrev.) of dryweight) heterogeneity heterogeneity Humus type Litter type
Leuterveld (LE) 57 0-29 0-23 Mull Deciduous Clay Oak, beech
Odijk (OD) 69 0-36 0-28 Mull Deciduous Clay Willow, reed
Duivelsberg (DU) 124 0-52 0-32 Mor Deciduous Sandy Beech,
loam Chestnut
Berkenbos (BE) 99 0-29 0-33 Mor Deciduous Sand Birch
Roggebotse bos (RO) 48 0-44 0-34 Deciduous Sand Poplar
Zeist (ZE) 111 0-42 0-37 Mor Deciduous Sand Oak, beech
Dennenbos (DE) 73 0-48 0-40 Mor Coniferous Sand Pine
Rechteren (RE) 71 0-57 0 50 Mor Coniferous Sand Pine, heather

the twenty-four soil cores from each sampling date. But as standard deviat
a positive linear relation with the mean (y) in this type of data, the use o
coefficient of variation is more appropriate:

_ antilog log Sy_ Geometric mean of standard deviations


antilog log y Geometric mean of the mean moisture conte

Temporal variability in soil moisture is measured by an ordinary coeffici


the quotient of the standard deviation of mean moisture content of soil
sampling occasion and the grand mean.

RESULTS

Soil moisture regime

Average moisture content and the coefficients of variation measuring the


spatial and temporal heterogeneity in moisture of litter and humus layer are s
Table 1, together with other important characteristics of each study area.
The highest values for average moisture content are observed in woodlands w
type humus layer and the lowest in typical mull humus types on clay and in th
poplar plantation where hardly any litter is present. These differences between
mainly reflect the different amounts of soil organic matter which has a large w
capacity. As the relation between moisture availability for soil animals and
content depends rather critically on physical and chemical properties of litter,
soil, only similar litter/soil types can be compared. They appear to have fai
average moisture contents (Table 1).
More interesting differences are observed in spatial and temporal hetero
Variability in space and time is highest in coniferous woodlands and tends to b
deciduous woodlands on clay. The two variables describing heterogeneity i
regime are correlated (Fig. 1). The woodlands DU and BE show the largest
from the main trend indicated by the dashed line. The fairly large value
heterogeneity in DU can be explained by the location of this study area on the
hill. The deviation of BE might be due to its location in a moist dune valley

This content downloaded from 200.24.17.12 on Sat, 09 Nov 2019 21:47:59 UTC
All use subject to https://about.jstor.org/terms
974 Population dynamics of Collembola

0*6-

, 0.5- RE,
o /
- 0.4- DEA
_ ZEo,/
ol BEe /ORO DU
E 0.3- /

LE /

0.2 0.3 0-4 0.5 0.6

Spatial heterogeneity

FIG. 1. Relation between spatial and temporal heterogeneity (coef


content. Similarity in symbols indicates similarity in hu

TABLE 2. Species composition of the study ar


Woodland

Species Abbreviations LE OD DU ZE DE RE BE RO
Tomocerus minor (Lubbock) Tm + + + + +
Tomocerusflavescens (Tullberg) Tf + + + + +
Tomocerus longicornis (Muller) T1 + +
Heteromurus nitidus (Templeton) Hn + + +
Lepidocyrtus lignorum Fabricius LI + + + + + + + +
Lepidocyrtus cyaneus Tullberg Lc +
Lepidocyrtus violaceus Lubbock Lv +
Orchesella cincta (Linne) Oc + + + + + + + +
Orchesella villosa (Geoffroy) Ov + +
Orchesellaflavescens (Bourlet) Of + +
Entomobrya nivalis (Linne) En + + + + + + + +
Entomobrya corticalis (Nicolet) Ec +
Entomobrya albocincta (Templeton) Ea +

Community structure
Species composition of the study areas is given in Ta
displayed in log abundance versus rank plots (Fig
entomobryid species found at each site, these grap
diversity, i.e. evenness, which is visualized by the slop
abundance to rank. The slopes of these regressions are
the pattern of ranked abundance clearly deviates from
a split-line, fitted according to the method of Perry
similar to the one observed in the other five communi
Whereas community structure does not seem to d
neity in moisture regime, species abundance does (Fig
according to drought sensitivity, which decreases
(Verhoef & Witteveen 1980; Vegter 1983, Vegte

This content downloaded from 200.24.17.12 on Sat, 09 Nov 2019 21:47:59 UTC
All use subject to https://about.jstor.org/terms
J. J. VEGTER, E. N. G. JOOSSE AND G. ERNSTING 975

(a) (b) (e) Cf

320C
LL

160C

80C )- \\ITM Leuterveld


(-040)
40C

20C

IOC

5C - \\. ov
25
oc
12-5 HN

(c)
E 320C
160C )- ILL
80C ') ']\ Duivelsberg
(-0.47)
40C
TF
20C

IOC

5C

25
I- \
oc

12-5

6-3 I- ._ TL
3.

1.5

1 234567

Rank

FIG. 2. Community ranked abundance. Geometr


intervals on log scale against the rank of the species
is given.

(a) Tomocerus (b ) (c ) . cincta


Lepidocyrtus
0

400 0 3200- A 400


A
w 200 0
1600 -
200 A A A
0 A A
C (C 100
800 -
- 50 50 A
- 0 400-
< 20 0 25
200- - A A
10 10
100- A A
-?1 I III I I
0/ I II I I K// I I I I I I I

o 0 00 o ocd 0
cL ) '/~)A, 0
0 0 0 ,0 0 o
dc;d I C) C I
Temporal heterogeneity

FIG. 3. Relation between abundance (numbers m-2) on log-scale and temporal heterogeneity.
Total number of: (a) T. minor and T.flavescens; (b) L. lignorum and L. cyaneus (in DE only); (c) 0.
cincta.

This content downloaded from 200.24.17.12 on Sat, 09 Nov 2019 21:47:59 UTC
All use subject to https://about.jstor.org/terms
976 Population dynamics of Collembola

0.8 -
U

.0 - 0-6 -

U
Ci)0.
044 -

a 0
F-0
E
0*2
6? t A

a)
m* * 4A A
,,,
- AI *A
I I Am
I I A

50 200 800 3200

Abundance (N m-2)

FIG. 4. Temporal variability in animal numbers (var


against geometric mean numbers m-2. (0) T. minor;
cincta; (-) E. nivalis. Populations with very low d

TABLE 3. Statistics of the power function r


population densities of the more import
E. nivalis 0. cincta L. lignorum T. flaves, cens T. minor

a b r2 a b r2 a b r2 a b r2 a b r2

Leuterveld 3-76 1-48 0-97 1-07 1-72 0-83 2-10 1-36 0-97 2-15 1-71 0-93
Berkenbos 3-14 1-44 0'91 3-26 1-30 0-86 4-01 1-74 0-91 4-03 1-60 0-93
0-96 1-08 1-87 0-94 2-29 1-56 0-86 3-24 1-49 0-96
Odijk 3-31 1-53
Zeist 3-25 1 48 0-95 0-87 1-98 0-87 2-40 1-30 0-93
Dennebos 2-02 1-73 0-84 3-50 1-70 0-92 5-21 1-63 0-91

Roggebotse
bos - 3-54 151 0-82 3-59 1-85 0-84 - 296 202 0-87
Duivelsberg ? ? ?- - - 1-41 1-73 0-94 2-15 1-50 0-89 -
Rechteren 5-42 1-61 0-98 4-51 1-67 0-87 4-15 1-66 0-84 -

Insufficient data.

(a) (b)

5 A
0

A 1-9
A

?A A A 1-7
A A

. * * A - At0
a3 b 0

1.5
8
2
0

1.3 - 0 A
_A
A A
A
I

50 200 800 3200 50 200 800 3200

Abundance (N m-2)

FIG. 5. (a) Parameter a of the function S2 =a m2 for sp


numbers m-2. (b) Same for parameter b: (0) T. minor; (
cincta; (?) E. nivalis. Populations with very low dens

This content downloaded from 200.24.17.12 on Sat, 09 Nov 2019 21:47:59 UTC
All use subject to https://about.jstor.org/terms
J. J. VEGTER, E. N. G. JOOSSE AND G. ERNSTING 977

abundance of the two drought-sensitive Tomocerus species appears to decrease with the
degree of temporal heterogeneity of the habitat (product-moment correlations coefficient
r = - 092, P < 0-01). Abundance of the drought-tolerant species Orchesella cincta is much
higher in the more heterogeneous habitats than in the least heterogeneous ones
(r= +0-73, P<0-05) whereas the reverse seems to apply to Lepidocyrtus (r= -061,
N.S.).
The 95% confidence intervals for abundance of each species, (Fig. 2) are similar in most
cases and by no means larger in more heterogeneous habitats, except perhaps in OD
where larger fluctuations in animal numbers are observed. Therefore, the intensity of
fluctuations in animal numbers in collembolan populations is unrelated to environmental
heterogeneity but appears to be related to population density: the more abundant the
species, the less it tends to fluctuate in number (Fig. 4). Table 3 gives the estimated
parameters (a and b) of the power function relating the spatial variation in animal
numbers to the mean for the more common species. The proportion of the total variance
of log transformed spatial variances explained by a linear regression on log mean is
measured by r2. The values of this parameter show that the power function is an adequate
description of the relation between means and spatial variances.
The species in Table 3 are listed in order of increasing drought sensitivity and habitats
are listed in order of increasing spatial heterogeneity. Values for parameter a are expected
to be higher in the right lower part of the table. But they are not. Parameter b appears to be
unrelated to spatial heterogeneity as well. In contrast, the data for all species combined
show a relation with population density both for parameter a (Fig. 5a; r = -0-67,
P< 001) and parameter b (Fig. 5b; r=0-64, P<0-01). For parameter a the negative
relation is also apparent within the species Entomobrya nivalis (r= -0-77, N.S.),
Lepidocyrtus lignorum (r= -0-88, P < 001) and Tomocerus minor (r= -0-94, P< 0-05).
For parameter b the positive relation results from positive trends within the species E.
nivalis (r=0-78, N.S.), L. lignorum (r=0-37, N.S.), Tomocerusflavescens (r=0.46, N.S.)
and T. minor (r = 0 58, N.S.) but especially from the relation with overall mean abundance
of species (r=0-86, P< 010).

DISCUSSION

The use of ranked abundance in the analysis of community structure in


originally intended to be a convenient way of presenting abundance data
interesting questions arise when the shape of these plots is compared
expectations, as presented by Whittaker (1975) and May (1975). In
relations between log abundance and rank correspond approximately
geometric series distribution, or its statistically more realistic counterp
distribution. These types of distribution are often observed in pion
(Whittaker 1975; Southwood, Brown & Reader 1979; Meijer 1980). A
species-abundance distribution is also predicted by Caswell's (1976) n
community structure, a class of models assuming no interaction betw
successional stages show more even distributions, such as the log-n
statistical model not reflecting any features of community biolog
MacArthur's (1957) uniform Broken Stick distribution (see, for exa
van Loon 1982). All these distributions have roughly linear or slightl
abundance plots on a semi-log scale. Unfortunately, statistical testing of
models to real data is problematical (Pielou 1975), but gross comparis

This content downloaded from 200.24.17.12 on Sat, 09 Nov 2019 21:47:59 UTC
All use subject to https://about.jstor.org/terms
978 Population dynamics of Collembola

obtained in the present study with this theoretical framework leads to a hypothesis that
might be a starting point for future research. The entomobryid community from the
youngest woodland, the new poplar stand, showed the highest degree of evenness, and the
communities of older deciduous woodlands yielded non-linear ranked-abundance plots,
in both of which the first three species exhibit an extreme degree of dominance. These
findings are different from expectations based on the theoretical models of community
structure mentioned above. The results suggest that in later successional stages a few
species, better adapted to local resource supply, might partition the available resource
spectrum in a strongly hierarchical manner, and force other species into secondary roles in
the community. Similar ideas were expressed by Caswell (1978) and more recently by
Hansky (1982) in the discussion of his core and satellite species hypothesis.
When community composition is compared between habitats, drought-sensitive
Tomocerus species appear to have lower population densities in the more heterogeneous
environments, which might be due to a lack of suitable microhabitats for these species.
The high densities of Orchesella cincta in the more heterogeneous woodlands is easily
explained by quite a number of adaptations (Joosse 1981; Verhoef & Li 1983). Whether
the lower abundance of this species in homogeneous areas must be explained by
physiological limitations or by competition from drought-sensitive species is at present
unknown.
Environmental heterogeneity does not seem to have a large influence on temporal
fluctuations or spatial aggregation of populations. Similar conclusions were arrived at by
Taylor, Woiwod & Perry (1978) and Taylor & Woiwod (1980) in their studies on spatial
pattern and temporal stability of a taxonomically diverse array of organisms. Concerning
temporal fluctuations, Leigh (1981) suggested that a stable environment might allow
species to specialize, which renders them more susceptible to environmental fluctuations,
and that therefore no relations between environmental heterogeneity and fluctuations in
population numbers are to be expected. Wolda (1978) observed that annual fluctuations
in insect populations in the tropics are indeed as pronounced as in presumably less stable
temperate environments. However, insect populations from dryer and more heteroge-
neous areas, either tropical or temperate, tend to fluctuate somewhat more. Whether
density in these more heterogeneous environments is lower in Wolda's study as well
remains unclear.
In Collembola, the intensity of temporal fluctuations is correlated with population
density. These fluctuations are determined to a large extent by the occurrence of seasonal
birth waves (Joosse 1969; Gregoire-Wibo 1979; Takeda 1979; Petersen 1980; Huhta &
Mikkonen 1982; Mertens, Coesens & Blancquaert 1982; Leinaas & Bleken 1983; Verhoef
& van Selm 1983). Moreover, the high birth and death rates characterizing the life cycles
of these animals (van Straalen 1983) lead to a population-size structure in which newly
hatched individuals constitute by far the largest fraction. Therefore, a lack of favourable
periods for recruitment leads to low abundance as well as to high temporal variability.
The analysis of spatial aggregation with Taylor's power function approach has been
criticized by Iwao (1970) and Todd (1978) as being too insensitive to detect small
differences. It should be noted, however, that other measures of spatial aggregation, such
as Lloyd's index of patchiness (Lloyd 1967) used by these authors, are simple functions of
means and variances as well, and therefore contain no additional information.
If animal numbers in a given patch are the result of variable birth and death rates, the
distribution of animal numbers in space and time is more likely to be log-normal
(MacArthur 1960; May 1975; Tuljapurkar & Orzack 1980). Since the variance of the log-

This content downloaded from 200.24.17.12 on Sat, 09 Nov 2019 21:47:59 UTC
All use subject to https://about.jstor.org/terms
J. J. VEGTER, E. N. G. JOOSSE AND G. ERNSTING 979

normal increases with the square of the mean, the expected value of b in Taylor's power
law will be 2. Displacement of animals may change the pattern towards 'randomness' with
b= 1 in case of Poisson distributions. Most b values published by Taylor, Woiwod &
Perry (1978) lie between 1 and 2. This suggests that Poisson and log-normal distributions
can be considered as two limiting cases.
The results from this study and from Joosse (1970) and Verhoef & van Selm (1983)
indicate that entomobryid Collembola are more aggregated than expected when
compared to a Poisson distribution. In the present study, b increases with species-specific
abundance approaching 2 in L. lignorum. This suggests that in a low-density species the
effect of displacement of animals on their distribution is relatively more important than in
a high-density species. The positive trend of b with habitat-specific population density
observed in most species likewise suggests that displacement contributes relatively more
to the observed pattern in unfavourable habitats than in favourable ones. As to the
relation between clumping and population density Usher et al. (1979) conclude that in
Collembola the number of aggregations increases with population density rather than the
number of individuals per aggregation, which is in complete agreement with the results of
the present paper if one is willing to accept that at high densities soil cores will mainly
reflect the spatial and temporal dynamics of Collembola within aggregations, whereas at
low overall densities samples will reflect mainly spatial and temporal variations outside
aggregations. Processes underlying the dynamics within aggregations are inherently
multiplicative birth and death rates, whereas outside aggregations animal numbers are
the result of more or less random dispersal. This conclusion is based on a limited number
of species; further research in this field might prove whether the interpretation of spatial
and temporal dynamics presented in this paper applies to other Collembola and to soil
organisms in general.

ACKNOWLEDGMENTS

The authors wish to thank Dr W. Slob for advice and comments on earlier drafts of this
paper and S. C. Verhoef for assistance in field- and laboratory work.

REFERENCES

Boomsma, J. J. & van Loon, A. J. (1982). Structure and diversity of ant communities in su
valleys. Journal of Animal Ecology, 51, 957-974.
Caswell, H. (1976). Community structure: A neutral model analysis. Ecological Monogra
Caswell, H. (1978). Predator-mediated coexistence: a non-equilibria model. American Natur
Gerard, G. & Berthet, P. (1966). A statistical study of microdistribution of Oriba
transformation of the data. Oikos, 17, 142-149.
Gisin, H. (1943). Oekologie und Lebensgemeinschaften der Collembolen im schweizerischen
Basels. Revue Suisse de Zoologie, 50, 131-224.
Gisin, H. (1969). Collembolenfauna Europas. Musee d'Histoire Naturelle, Geneve.
Gisin, H. (1964). Collemboles d'Europe VII. Revue Suisse de Zoologie, 71, 649-678.
Gregoire-Wibo, C. (1979). Cycle ph6nologique de Folsomia quadrioculata en foret, (Ins
Annales de la Societe royale de Zoologie Belge, 109, 43-65.
Hansky, I. (1982). Dynamics of regional distribution: the core and satellite species hypoth
221.
Huhta, V. & Mikkonen, M. (1982). Population structure of Entomobryidae (Collembola) in a mature spruce
stand and in a clear-cut reforested area in Finland. Pedobiologia, 24, 231-240.
Iwao, S. (1970). Problems of a spatial distribution in animal population ecology. Random Counts in Scientific
Work, Vol. 2 (Ed. by G. P. Patil), pp. 117-149. Pennsylvania State University Press, University Park and
London.

This content downloaded from 200.24.17.12 on Sat, 09 Nov 2019 21:47:59 UTC
All use subject to https://about.jstor.org/terms
980 Population dynamics of Collembola
Joosse, E. N. G. (1969). Population structure of some surface dwelling collembola in a coniferous forest
Netherlands Journal of Zoology, 19, 621-634.
Joosse, E. N. G. (1970). The formation and biological significance of aggregations in the distributi
Collembola. Netherlands Journal of Zoology, 20, 299-314.
Joosse, E. N. G. (1981). Ecological strategies and population regulation of Collembola in heterogene
environments. Pedobiologia, 21, 346-356.
Leigh, E. G. Jr. (1981). The average lifetime of a population in a varying environment. Journal of theo
Biology, 90, 213-239.
Leinaas, H. P. & Bleken, E. (1983). Egg diapause and demographic strategy in Lepidocyrtus lignorum Fabr
(Collembola, Entomobryidae). Oecologia, 58, 194-199.
Lloyd, M. (1967). 'Mean crowding'. Journal of Animal Ecology, 36, 1-30.
MacArthur, R. H. (1957). On the relative abundance of bird species. Proceedings of the National Academ
Science, 43, 293-296.
MacArthur, R. H. (1960). On the relative abundance of species. American Naturalist, 94, 25-26.
May, R. M. (1975). Patterns of species abundance and diversity. Ecology and Evolution of Communities (Ed
M. L. Cody & J. M. Diamond), pp. 91-120. Belknap Press, Cambridge, Massachusetts.
Mertens, J., Coesens, R. & Blancquaert, J. P. (1982). Population structure of Orchesella cincta (Collembol
the field based on instar-determination. Pedobiologia, 23, 9-20.
Meijer, J. (1980). The development of some elements of the arthropod fauna of a new polder. Oecologia,
220-235.
Perry, J. N. (1982). Fitting split-lines to ecological data. Ecological Entomology, 7, 421-435.
Petersen, H. (1980). Population dynamic and metabolic characterization of Collembola species in a beech fores
ecosystem. Soil Biology as Related to Land Use Practices (Ed. by D. L. Dindal), pp. 806-833. Office of
Pesticides and Toxic substances, EPA, Washington.
Pielou, E. C. (1975). Ecological Diversity. Wiley & Sons, New York.
Pielou, E. C. (1977). Mathematical Ecology. Wiley & Sons, New York.
Ponge, J. F. (1980). Les bioc6noses des collemboles de la foret de S6nart. Actualites d'ecologieforestiere (Ed. b
P. Pesson), pp. 151-176. Gauthier-Villars, Paris.
Ponge, J. F. (1983). Les collemboles, indicateurs du type d'humus en milieu forestier. R6sultats obtenus au Sud
de Paris. Acta Oecologica/Oecologia Generalis, 4, 359-374.
Southwood, T. R. E., Brown, V. K. & Reader, P. M. (1979). The relationships of plant and insect diversities in
succession. Biological Journal of the Linnean Society, 12, 327-348.
van Straalen, N. M. (1983). Recruitment, body-growth and mortality in populations of forest floor collembola
Vergelijkende demografie van springstaarten. Ph.D. thesis, Vrije Universiteit, Amsterdam.
van Straalen, N. M. & Rijninks, P. C. (1982). The efficiency of Tullgren apparatus with respect to interpreting
seasonal changes in stage structure of soil arthropod populations. Pedobiolgia, 24, 197-209.
Takeda, H. (1979). Ecological studies on collembolan populations in a pine forest soil. 3. Life cycles and
population dynamics of some surface dwelling species. Pedobiologia, 19, 34-47.
Taylor, L. R. & Taylor, R. A. J. (1977). Aggregation, migration and population mechanics. Nature, 265, 415-
420.
Taylor, L. R. & Woiwod, I. P. (1980). Temporal stability as a density-dependent species characteristic. Journal o
Animal Ecology, 49, 209-224.
Taylor, L. R., Woiwod, I. P. & Perry, J. N. (1978). The density-dependence of spatial behaviour and the rarity o
randomness. Journal of Animal Ecology, 47, 383-406.
Todd, C. P. (1978). Changes in spatial pattern of an intertidal population of the nudibranch mollusc Onchidoris
muricata in relation to life cycle, mortality and environmental heterogeneity. Journal of Animal Ecology, 47,
189-203.
Tuljapurkar, S. D. & Orzack, S. H. (1980). Population dynamics in variable environments. I. Long-run growth
rates and extinction. Theoretical Population Biology, 18, 314-342.
Usher, M. B., Davis, P. R., Harris, J. R. W. & Longstaff, B. C. (1979). A profusion of species? Approaches
towards understanding the dynamics of the populations of the micro-arthropods in decomposer
communities. Population Dynamics (Ed. by R. M. Anderson, B. D. Turner & L. R. Taylor), pp. 359-384.
Symposium 20 of the British Ecological Society, Blackwell Scientific Publications, Oxford.
Vegter, J. J. (1983). Food and habitat specialization in coexisting springtails (Collembola, Entomobryidae).
Pedobiologia, 25, 253-262.
Vegter, J. J. & Huyer-Brugman, F. A. (1983). Comparative water relations in Collembola: transpiration,
desication tolerance and effects of body size. New Trends in Soil Biology (Ed. by Ph. Lebrun, H. M. Andre,
A. de Medts, C. Gregoire-Wibo & G. Wauthy), pp. 411-416. Proceedings of the VIII international
Colloquium of Soil Zoology, Louvain-la-Neuve (Belgium).
Verhoef, H. A. & Li, K. W. (1983). Physiological adaptations to the effect of dry summer periods in Collembola.
New Trends in Soil Biology (Ed. by Ph. Lebrun, H. M. Andre, A. de Medts, C. Gr6goire-Wibo and G.
Wauthy), pp. 345-346. Proceedings of the VIII international Colloquium of Soil Zoology, Louvain-la-
Neuve (Belgium). Diew-Brichart, Louvain-la-Neuve.

This content downloaded from 200.24.17.12 on Sat, 09 Nov 2019 21:47:59 UTC
All use subject to https://about.jstor.org/terms
J. J. VEGTER, E. N. G. JOOSSE AND G. ERNSTING 981
Verhoef, H. A. & van Selm, A. J. (1983). Distribution and population dynamics of Collembola in relation t
moisture. Holarctic Ecology, 6, 387-394.
Verhoef, H. A. & Witteveen, J. (1980). Water balance in Colembola and its relation to habitat selection: cut
water loss and water uptake. Journal of Insect Physiology, 26, 201-208.
Wauthy, G. & Lebrun, Ph. (1980). Synecology of forest soil oribatid mites of Belgium. 1. The zoosociolog
classes. Soil Biology as Related to Land Use Practices (Ed. by D. L. Dindal), pp. 795-805. Offic
Pesticides and Toxic Substances, EPA, Washington.
Whittaker, R. H. (1975). Communities and Ecosystems. Macmillan, New York.
Wolda, H. (1978). Fluctuations in abundance of tropical insects. American Naturalist, 112, 1017-1045.

(Received 13 April 1987)

This content downloaded from 200.24.17.12 on Sat, 09 Nov 2019 21:47:59 UTC
All use subject to https://about.jstor.org/terms

You might also like