species positions and extinction dynamics_ferenc.j.theo.biol

Upload: angstron328

Post on 30-May-2018

217 views

Category:

Documents


0 download

TRANSCRIPT

  • 8/14/2019 Species Positions and Extinction Dynamics_Ferenc.J.theo.Biol

    1/8

    J. theor. Biol. (2002) 215, 441448

    doi:10.1006/jtbi.2001.2523, available online at http://www.idealibrary.com on

    Species Positions and Extinction Dynamics in Simple Food Webs

    Ferenc Jordan*wz, Istvan Scheuringy and GaborVidaw

    nCollegium Budapest, Institute for Advanced Study, Budapest H-1014, Szenth !aroms !ag u. 2., Hungary

    wDepartment of Genetics, Evolutionary Genetics Research Group, E.otv .os University, Budapest H-1117,P !azm !any P. s. 1/c, Hungary and y Department of Plant Taxonomy and Ecology, Research Group of

    Ecology and Theoretical Biology, E.otv .os University, Budapest H-1117, P !azm !any P. s. 1/c, Hungary

    (Received on 22 June 2001, Accepted in revised form on 18 December 2001)

    Recent investigations on the structure of complex networks have provided interesting resultsfor ecologists. Being inspired by these studies, we analyse a well-defined set of small modelfood webs. The extinction probability caused by internal LotkaVolterra dynamics iscompared to the position of species. Simulations have revealed that some global properties ofthese food webs (e.g. the homogeneity of connectedness) and the positions of species therein(e.g. interaction pattern) make them prone to modelled biotic extinction caused bypopulation dynamical effects. We found that: (a) homogeneity in the connectedness structureincreases the probability of extinction events; (b) in addition to the number of interactions,their orientations also influence the future of species in a web. Since species in characteristicnetwork positions are prone to extinction, results could also be interpreted as describing the

    properties of preferred states of food webs during community assembly. Our results maycontribute to understanding the intimate relationship between pattern and process inecology.

    r 2002 Elsevier Science Ltd. All rights reserved.

    Introduction

    Interesting ideas in ecology frequently come

    from distant fields of science. The conclusions of

    recent studies on the structure of complex

    networks (Watts & Strogatz, 1998; Barab!asi &

    Albert, 1999; Albert et al., 2000; Jeong et al.,2000; Amaral et al ., 2000; Williams et al .,

    submitted), giving information about how meta-

    bolic and neuronal networks seem to be orga-

    nized, may also be useful for ecologists. An

    interesting result is that metabolic networksfollow a power-law connectivity distribution

    indicating robust functional insensitivity against

    random errors (Jeong et al., 2000).

    The dynamical behaviour of a well-defined set

    of small sink webs (Cohen, 1978) has been

    studied (Jord!an & Moln!ar, 1999; Jord!an et al.,

    submitted), giving rise to the problem of how thenetwork position of frequently extinct species

    can be characterized. Since the large-scale

    structure of biological networks depends highly

    on internal dynamical processes and constraints

    (Sugihara, 1984), we analyse how the internal

    extinction dynamics (based on LotkaVolterra

    models) in these webs is related to the topology

    of trophic interactions. We think that these basic

    equations, even if their applicability is limited,

    are good as a starting point for further analyses.zAuthor to whom correspondence should be addressed.

    E-mail: [email protected]

    0022-5193/02/$35.00/0 r 2002 Elsevier Science Ltd. All rights reserved.

  • 8/14/2019 Species Positions and Extinction Dynamics_Ferenc.J.theo.Biol

    2/8

    Methods

    We compared the extinction dynamics of

    species constituting 25 model sink webs (i.e. food

    webs with a single top-predator, see Fig. 1). This

    set of webs contains all of the 25 topologicallydifferent model sink food webs possibly built up

    from five species (graph nodes) and five trophic

    interactions (links). Since the connectedness of

    these webs equals uniformly C 0.5, we are able

    to analyse the effect of link pattern on population

    dynamics, distinguished from the effects of

    changing connectedness. Food webs and indivi-

    dual species are characterized by the following

    structural and dynamical properties.

    The degree (D) of a graph node is the number

    of its neighbours (both prey and predators in the

    food web graph). Because all these webs contain

    five links, the sum of D values (sumD) equals

    10 for every web. However, the D values of

    individual species differ; this difference is mea-

    sured by the standard deviation of D within a

    web (stdD).

    The orientation of neighbours is characterized

    by D0DinDout (where Din is the in-degree: the

    number of links directed to a particular node,

    and Dout is the out-degree: the number of links

    coming from a particular node). Low and high

    values of D0 characterize multiple exploitedproducers and generalist consumers, respec-

    tively, in the directed graph.

    The concept of net status (Harary, 1959,

    1961) has been modified slightly in order to be

    ecologically more reasonable (see Jord!an et al.,

    1999; Jord!an, 2000, 2001). The modified index

    was constructed to give a quantitative measure

    of the importance of species (species of

    outstanding importance are called keystone

    species, Paine, 1969; Power et al., 1996). This

    is the positional keystone index (K) which

    considers not only neighbours in the web (like

    D) but their neighbours as well, thus providing

    more information about network structure. If

    bottom-up and top-down forces are considered

    to be of equal importance, then let Ki Kib

    Kit ; where K

    ib and K

    it refer to the bottom-up

    1

    1111

    1

    11

    111

    1

    1

    111

    1

    1

    11

    1

    2

    2222

    2

    22

    22

    22

    2

    2

    2

    2

    22

    22

    22

    2

    2

    2

    3

    333

    3

    3

    3

    3

    3

    3

    33

    33

    3 3

    3

    333

    3

    3

    3

    3

    3

    4

    444

    4

    4

    4

    4

    44

    4

    44

    44

    4

    4

    44

    44

    4

    4

    55

    55 5

    55

    5

    555

    5555

    5

    555

    55

    55

    1 2 3

    7 8 9 1011 12 13

    14 1516

    1718 19 20

    21

    22

    2324

    25

    4 5 6

    1 1

    4 4

    1 1

    5 5

    Fig. 1. Twenty-five model sink webs containing five species (nodes) and five trophic links. Both species and webs arenumbered for further reference. Higher species consume lower ones (species #1 is always the top-predator; the direction oflinks is not shown for simplicity). For data on structural (D, D0, K) and dynamical (E) characterization of species, consulthttp://falco.elte.hu/Bjordanf/data2.prn.

    F. JORDA N ET AL.442

  • 8/14/2019 Species Positions and Extinction Dynamics_Ferenc.J.theo.Biol

    3/8

    and top-down effects. These indices count the

    number of disconnected species after the re-

    moval of species i. If species i is in an important

    network position it has a rich system of trophic

    interactions, then deletion will lead to more

    disconnections in both directions. The effects ofspecies deletion on the disconnection of both

    bottom-up material flows and top-down flows of

    trophic regulatory effects are quantified by these

    indices of positional importance. The Kb value of

    the i-th species can be calculated as

    Kib

    Xnc1

    1

    dc

    1

    dcK

    cb

    ; 1

    where n is the number of its predators, dc is the

    number of prey eaten by its c-th predator, andKc

    b is the bottom-up keystone index of its c-th

    predator. Thus, the bottom-up index should be

    calculated first for top species. Kt is calculated

    similarly after turning the web upside down. It

    should be emphasized that K refers to keystone

    species only in a trophic sense (i.e. importance in

    maintaining trophic network flows). K charac-

    terizes food web pattern at an intermediate scale,

    between the local (measured by D and D0) and

    global (measured by stdD) properties of webs.

    Nevertheless, in our small webs K seems to be

    a rather global index. Both the sum and thestandard deviation of K-values within webs

    (sumK and stdK) were also calculated for

    analyses.

    We are interested in whether these structural

    indices are possibly as informative and interest-

    ing as traditional food web statistics having been

    used extensively in the literature (e.g. trophic

    height, compartmentalization, the level of

    omnivory, etc., see Pimm, 1980).

    The behaviour of species is characterized by

    dynamical LotkaVolterra models. Population

    dynamics is described by the following set of

    equations:

    xi aiXnj1

    bijxj

    !xi; i 1;y; n: 2

    Here, xi is the population density of species i,

    ai is the per capita rate of reproduction (for

    growth, ai40, for decay, aio0). The parameter

    bij indicates the per capita effect of species

    j on the per capita reproduction rate of species

    i. Here, ai is a uniformly distributed random

    variable in the range (0 1) if species i is a basal

    species, while ai is selected randomly from (1 0)

    if species i is intermediate or top species.

    Similarly, bij falls into (0 1) if i is the pre-dator of j, and bji is selected from (1 0) if i

    eats j. If ij, bij is selected from (1 0) in case

    of producers but equals zero for higher species

    (May, 1973; Pimm, 1982). The dynamics of

    system (2) is simulated with 2000 random

    parameter sets selected in the way described

    above. We considered a species to be extinct

    if its density fell below a critical level (here,

    d1015; results do not differ qualitatively

    for other values of d) after a given time (here,t 200 steps). The extinction value (E) o f a

    species gives the number of extinctions out of

    2000 simulations (we suggest that our approach

    to stability is more general than to study stable

    fix points).

    Results

    The distribution of the basic structural indices

    (D and K) of species in these model webs is

    shown in Fig. 2: the majority of species have two

    neighbours (D 2), while K is distributed evenly

    (especially if sink species are not considered).The most frequently extinct nodes can be

    characterized by an intermediate degree (D 2,

    see Fig. 3). The keystone index itself does not

    seem to be in correlation with the probability of

    extinction (Fig. 3; but seemingly extinctions are

    not frequent with very low K). However, the sum

    of extinction events within a web is clearly higher

    if the sum of keystone indices is higher (Fig. 3).

    The fact that Kcorrelates with Eonly at the level

    of whole webs reflects how keystone index is

    calculated: it also quantifies indirect effects,which are of more global nature than locally

    understandable direct interactions. In order to

    study the effects of both D and K on extinction

    dynamics, they are plotted together (Fig. 4). If

    D 1, species go extinct with high probability if

    K 4; however, the probability of extinction is

    low at smaller K indices. If D 2, extinction can

    be very probable only with low Kvalues. In case

    of a dangerous number of neighbours (i.e. if

    D 2), additional indirect effects (e.g. K 4) may

    EXTINCTION DYNAMICS IN FOOD WEBS 443

  • 8/14/2019 Species Positions and Extinction Dynamics_Ferenc.J.theo.Biol

    4/8

    decrease the probability of dynamical extinction.

    If D 3 or 4, extinction is less probable and

    is less sensitive to K. Thus, measuring both

    exclusively direct (D) and direct plus indirect (K)

    effects gives an interesting extinction probability

    landscape.

    Fig. 2. The frequency distribution for the degree of nodes is a discrete, unimodal distribution (on the left). K values areevenly distributed (on the right; note that for all nodes of the last column K 4: 25 out of 47 nodes represent the sink species;thus, basal and intermediate species are distributed much more evenly).

    Fig. 3. Relationships between structural and dynamical properties: top left, the number of extinction events concerningeach species (E, out of 2000 simulations) is plotted against the number of their neighbours (D, ( ) show the average; theMannWhitney test on a distribution of 105 randomly generated samples shows significance at po0.0003, po0.0114, and

    po0.0025 for differences between [D 1, 2], [D 2, 3], and [D 3, 4] pairs of categories, respectively); top right, the numberof extinction events (E) is plotted against species keystone indices (K); bottom, the sum of extinction events (sumE) withineach web is plotted against the sum of their species keystone indices (sumK).

    F. JORDA N ET AL.444

  • 8/14/2019 Species Positions and Extinction Dynamics_Ferenc.J.theo.Biol

    5/8

  • 8/14/2019 Species Positions and Extinction Dynamics_Ferenc.J.theo.Biol

    6/8

    nodes). These positional properties of nodes may

    characterize either extinction risk (as far as

    randomly chosen interaction parameters are

    regarded as perturbations) or preferences in

    community assembly (as far as possible struc-

    tures created by invaders are compared). If

    positional properties (D, D0, K) influence extinc-

    tion dynamics (E), species do not die out

    randomly, but rather in an internally directed

    way (alike attacks are defined as directed node

    deletions in large networks, see Albert et al.,

    2000). For this set of small sink webs, it can

    be concluded that the presence of species in key

    positions can be dynamically either advanta-

    geous [if only direct interactions are considered

    (see stdD in Fig. 6)] or disadvantageous [if

    both direct and indirect interactions are

    taken into account (see Fig. 3)] against these

    internal attacks [in contrast with Albert et al.

    (2000), where non-random deletions are more

    Fig. 5. The number of species extinction events (E) is plotted against species D0 indices ( ) show the average; seeexplanation in text): not only the number but also the orientation of direct trophic links influence extinction dynamics. The

    MannWhitney test on a distribution of 105

    randomly generated samples shows significant difference between [D0

    1, 0] atpo0.00001. [D00, 1] and [D01, 2] show different average E at po 0.1. D02 and 3 differ significantly at po0.003, while[D0 2, 1] and [D03, 4] do not differ significantly.

    Fig. 6. The sum of extinction events (sumE) for each web is plotted against the standard deviation of degrees ( stdD,( )) and the standard deviation of keystone indices (stdK, ( )) of species. Living in a homogeneous web is moredangerous.

    F. JORDA N ET AL.446

  • 8/14/2019 Species Positions and Extinction Dynamics_Ferenc.J.theo.Biol

    7/8

    dangerous in heterogeneous networks]. The

    description of extinction-prone network posi-

    tions is suggested to identify unpreferred struc-

    tural arrangements during web assembly

    (cf. Sugihara, 1984): the success of invasion

    may depend characteristically on the particularposition of the invader within the community

    network (and this may hold also in case of

    speciation).

    The main mechanisms behind the simulated

    network dynamics include direct (e.g. predation,

    food supply) and indirect (e.g. exploitative

    competition, trophic cascade) trophic interac-

    tions among species (see a catalogue of indirect

    interactions in Menge, 1995). Given the topology

    of a trophic network, one may predict either the

    nature of a particular interaction or the fate of a

    particular species, but the dynamics of whole

    webs remain unclear until simulated. Only joint

    dynamical equations can give higher-level pre-

    dictions (e.g. the fate of a whole web). Our

    simulation results show typical patterns in the

    dynamics of food webs and their species. For

    example, let us consider food web #12 in Fig. 1.

    This web can be considered typical as for the

    frequency of extinctions (1760 extinction events

    occurred out of 2000 simulations). The extinc-

    tion frequency of the top-predator (#1, E482)

    is close to the average. Its position is neither toorisky nor very safe. One of its prey (#2, E832)

    is prone to extinction, as are species with D00

    in general. They have a single source to consume

    and their fate depends strongly on their pre-

    dators behaviour. Its other prey (#4, E241)

    has a high extinction probability compared to

    other producers. It is caused by the following: if

    the top-predators density increases, it exploits

    species #4 both directly (by predation) and

    indirectly (by having a negative effect on species

    #2, which is good for species #3, which is badagain for species #4). Species #3 (E 87) is well

    saved by topology: it has a single predator and

    two preys. Finally, there is a producer in this

    community (#5, E 118) which goes to

    extinction less frequently than the other one

    (#4), because its position is better as for the

    number of predators.

    We can conclude that dynamically caused

    species extinctions are less frequent if the

    following conditions are satisfied: at the local

    scale, (1) species have either a single or many

    (3 or 4) direct trophic links (see, for example,

    species #5 and #1 of web #1, Fig. 1), (2) the

    neighbours are clustered in either bottom-up or

    top-down direction (D0a0, see species #2 in web

    #3); at an intermediate scale, (3) they have eithera single neighbour and a low K-index or two

    neighbours and a high K-index (see species #5 in

    web #7 and species #5 in web #21), (4) the sum

    of K-indices within a web is low (see web #3),

    and, at the global scale, (5) whole webs have

    differently connected nodes (heterogeneity, stdD,

    see web #2). We have to emphasize that the

    simplest dynamical model was analysed: asym-

    metries in interaction strength (bij parameters)

    may lead to qualitative differences in our

    conclusions (Jord!an et al., submitted).

    If trophic control in a particular community is

    considered of primary importance, i.e. food web

    studies have strong predictive power, then

    linking trophic structure to species dynamicscould provide useful information for conserva-

    tionists. The quest for keystone species (Poweret al., 1996) is a big challenge for community

    ecologists. Quantitative approaches to identify-

    ing keystone species in ecosystems (e.g. based on

    food web position) could increase the predictive

    power of studies on the relative importance of

    species within communities. In addition, if key-stones really make communities less prone to

    extinction, we can get closer to explaining their

    outstanding importance. Thus, the prediction

    of extinction dynamics by food web analysis

    could give a useful tool for conservationists

    (cf. Simberloff, 1998). If biological diversity can

    be expressed in terms of positional diversity of

    species in trophic networks, then the diverse

    population dynamics of species can also be

    approached quantitatively.

    We focused only on a particular but well-defined set of small food webs. Thus, possible

    effects of scale- and context-dependence were

    neglected. First, it should be clarified whether

    our results based on the analysis of small webs

    would also be applicable for large networks (see

    Bersier & Sugihara, 1997). Second, the problem

    of how the dynamics of community webs differs

    from that of sink webs is to be studied. The

    behaviour of persistent nodes following pertur-

    bations is to be analysed in detail (some

    EXTINCTION DYNAMICS IN FOOD WEBS 447

  • 8/14/2019 Species Positions and Extinction Dynamics_Ferenc.J.theo.Biol

    8/8

    preliminary results on whole webs are given in

    Jord!an et al., submitted). Further, the manner in

    which certain network properties affect basic

    organizational principles, similar to the ageing

    and cost of nodes in the study ofAmaral et al.

    (2000), should be considered. Nevertheless, wethink that this study can contribute to the

    understanding of the pattern and process

    problems in ecology. Future studies on the

    homogeneity of ecological networks may reveal

    important aspects of the organization of food

    webs, and, in general, the ecological principles of

    community assembly and biotic extinctions.

    We are grateful to J!anos Podani for his help instatistics and correcting English. F. J. also thanksProfessor E .ors Szathm!ary for enabling a fruitful stay

    at Collegium Budapest, Mauro Santos and EliasZintzaras for some methodical help. Two anonymousreviewers are kindly acknowledged for helpful com-ments on the manuscript. F. J. is supported by twogrants of the Hungarian Scientific Research Fund,OTKA F 029800 and OTKA F 035092. I. Sch. issupported by the grant OTKA T 029789. Both F. J.and I. Sch. are recipients of the Bolyai ResearchGrant of the Hungarian Academy of Sciences.

    REFERENCES

    Albert, R., Jeong, H. & Baraba si, A.-L. (2000). Errorand attack tolerance of complex networks. Nature 406,378381.

    Amaral, L. A. N., Scala, A., Barthe le my, E . &Stanley, H. E. (2000). Classes of small-world networks.Proc. Natl Acad. Sci. U.S.A. 97, 11 14911 152.

    Baraba si, A.-L. & Albert, R. (1999). Emergence ofscaling in random networks. Science 286, 509512.

    Bersier, L. F., & Sugihara, G. (1997). Scaling regions forfood web properties. Proc. Natl Acad. Sci. U.S.A. 94,12471251.

    Cohen, J. E. (1978). Food Webs and Niche Space.Princeton, NJ: Princeton University Press.

    Harary, F. (1959). Status and contrastatus. Sociometry 22,2343.

    Harary, F. (1961). Who eats whom? Gen. Syst. 6, 4144.Jeong, H., Tombor, B., Albert, R., Oltvai, Z. N . &

    Baraba si , A.-L. (2000). The large-scale organization ofmetabolic networks. Nature 407, 651654.

    Jorda n, F. (2000). Is the role of trophic control larger ina stressed ecosystem? Commun. Ecol. 1, 139146.

    Jorda n, F. (2001). Trophic fields. Community Ecol. 2,181185.

    Jorda n, F . & Molna r, I. (1999). Reliable flows andpreferred patterns in food webs. Evol. Ecol. Res. 1,591609.

    Jorda n, F., Taka cs-sa nta, A . & Molna r, I. (1999).A reliability theoretical quest for keystones. Oikos 86,453462.

    Jorda n, F., Scheuring, I. & Molna r, I. Persistence, flowreliability in simple food webs. Submitted.

    May, R. M. (1973). Stability and Complexity in ModelEcosystems. Princeton, NJ: Princeton University Press.

    Menge, B. A. (1995). Indirect effects in marine rockyintertidal interaction webs: patterns and importance.Ecol. Monogr. 65, 2174.

    Paine, R. T. (1969). A note on trophic complexity andcommunity stability. Am. Nat. 103, 9193.

    Pimm, S. L. (1980). Properties of food webs. Ecology 61,219225.

    Pimm, S. L. (1982). Food Webs. London: Chapman & Hall.Power, M. E., Tilman, D., Estes, J. A., Menge, B. A.,

    Bond, W. J., Mills, L. S., Daily, G., Castilla, J. C.,Lubchenco, J. & Paine, R. T. (1996). Challenges in thequest for keystones. BioScience 46, 609620.

    Simberloff , D. (1998). Flagships, umbrellas, andkeystones: is single-species management pass!e in the

    landscape era? Biol. Conserv. 83, 247257.Sugihara, G. (1984). Graph theory, homology and foodwebs. Proc. Symp. Appl. Math. 30, 83101.

    Watts, D . J . & Strogatz, S. H. (1998). Collectivedynamics of small-world networks. Nature 393,440442.

    Williams, R. J., Martinez, N. D., Berlow, E. L.,Dunne, J . A . & Barab !Asi , A.-L. Two degrees ofseparation in complex food webs. Proc. Natl. Acad. Sci.U.S.A. (In press).

    F. JORDA N ET AL.448