Global Change Emmerson.etal.GCB

Embed Size (px)

Citation preview

  • 8/14/2019 Global Change Emmerson.etal.GCB

    1/12

    Global change alters the stability of food webs

    M A R K E M M E R S O N *, M A R T I J N B E Z E M E R w 1 , M A R K D . H U N T E R z and T . H E F I N J O N E S

    *Department of Zoology, University College Cork, Lee Maltings, Prospect Row, Cork, Ireland, wNetherlands Institute of Ecology

    (NIOO-KNAW), PO Box 40, 6666 ZG Heteren, The Netherlands, zInstitute of Ecology, University of Georgia, Ecology Building,

    Athens, GA 30602-2202, USA, Cardiff School of Biosciences, Cardiff University, PO Box 915, Cardiff, CF10 3TL, UK

    Abstract

    Recent research has generally shown that a small change in the number of species in afood web can have consequences both for community structure and ecosystem processes.However change is not limited to just the number of species in a community, but mightinclude an alteration to such properties as precipitation, nutrient cycling andtemperature. How such changes might affect species interactions is important, not justthrough the presence or absence of interactions, but also because the patterning ofinteraction strengths among species is intimately associated with community stability.Interaction strengths encompass such properties as feeding rates and assimilationefficiencies, and encapsulate functionally important information with regard to

    ecosystem processes. Interaction strengths represent the pathways and transfer ofenergy through an ecosystem. We review the best empirical data available detailing thefrequency distribution of interaction strengths in communities. We present theunderlying (but consistent) pattern of species interactions and discuss the implicationsof this patterning. We then examine how such a basic pattern might be affected givenvarious scenarios of change and discuss the consequences for community stability andecosystem functioning.

    Keywords: community, ecosystems, food webs, herbivore, persistence, plant, predators, prey, resilience,

    stability

    Received 9 September 2004; and accepted 15 October 2004

    Introduction

    Pioneering early ecologists (for example Odum, 1953;

    MacArthur, 1955; Elton, 1958) held that the dynamics of

    complex ecological systems should be more stable than

    simpler systems. They based this view on the premise

    that those ecological systems, which were more species

    rich, have more interspecies pathways along which

    energy can flow. These pathways can be depicted

    graphically as food web diagrams and energy flow

    may be characterized by the trophic interactions that

    take place among species within the food web. The

    biological strength of these trophic interactions details

    the conversion of abundance or biomass from one

    trophic level to another. In this paper, we will explore

    how increased productivity brought about by rising CO2

    levels and global warming might effect change in speciesinteractions at local scales. This is important because a

    disruption to the strength or arrangement of interspecific

    interactions in food webs can have consequences for the

    stability of those same systems. For instance, in the early

    1970s, Rosenzweig (1971) warned against the artificial

    enrichment of ecosystems (to increase productivity) in

    terms of increased nutrient and energy supply. He

    showed that for two trophic-level systems, either

    scenario would destroy steady states in simple models.

    Throughout this paper, we describe a scenario of global

    change (increased productivity) brought about by

    coincident global increases in temperature and CO2concentration. We explore this scenario of change using

    a simple three-species food web consisting of a basal

    resource (plant), primary consumer (herbivore) and

    secondary consumer (omnivorous predator).

    Although the study of species interactions has tradi-

    tionally been concerned with their effects on community

    stability (sensu May, 1973), recent research has focused

    more on how ecosystem processes such as productivity,

    nutrient flux and nutrient retention are affected by the

    Correspondence: Mark Emmerson, tel. 1 353 (0)21 490 4190.

    fax: 1353 (0)21 427 0562, e-mail: [email protected] address: Nature Conservation and Plant Ecology Group,

    Wageningen University and Research Centre, Bornsesteeg 69, 6708

    PD Wageningen, The Netherlands.

    Global Change Biology (2005) 11, 490501, doi: 10.1111/j.1365-2486.2005.00919.x

    490 r 2005 Blackwell Publishing Ltd

  • 8/14/2019 Global Change Emmerson.etal.GCB

    2/12

    loss of species (Chapin et al., 2000; McCann, 2000; Loreau

    et al., 2001; Madritch & Hunter, 2002). The transfer of

    energy (or carbon) through an ecosystem represents one

    such ecosystem process and interaction strengths provide

    a direct link between the diversity and ecosystem

    functioning of a community and its ecological stability

    (Duffy, 2002). Interaction strengths have been quantified

    in a number of ways and this general term has been usedvariously to describe both (i) the biological flux of

    material or energy from one trophic level to another and

    (ii) per-capita effects of a species on the growth rate of

    another. In this paper, we will focus on the latter use of

    the term interaction strength.

    Concerns over increased global rates of species

    extinction are driven by the urgent need to understand

    and predict the consequences of species loss for the

    stable and reliable provision of ecosystem services

    (Duarte, 2000; Engelhardt & Ritchie, 2001; Loreau

    et al., 2001; Lerdau & Slobodkin, 2002). To understand

    how communities might respond locally to global

    change (changes in productivity driven by increasedtemperature and CO2 concentrations), requires that we

    identify the basic arrangement of species interaction

    strengths in communities and investigate how one

    measure of global change might affect that pattern at

    the local ecological scale. To achieve this, we now

    review and detail measures of interaction strength, and

    present what we consider to be some of the best data

    currently available detailing the patterns of interaction

    strengths in ecological systems; this is necessary for our

    exploration of change. We then define some simple and

    biologically reasonable scenarios of ecological change

    and explore how the patterning of interaction strengthsmight be affected by such changes using a Lotka

    Volterra modelling framework. Our aim in using Lotka

    Volterra dynamics is to make qualitative forecasts

    regarding the consequences of change, not to make

    quantitative predictions. Our goal is to make a heuristic

    exploration of these scenarios using this simple model-

    ling approach. Finally, we will discuss the consequences

    of change for community and ecosystem stability.

    What is the pattern of interaction strengths incommunities?

    The effects of species interactions on community stability

    were first explored by May (1973) using linear stability

    analyses. May used LotkaVolterra models of the form:

    dXidt

    Xi ri Xnj1

    aijXj

    0@

    1A; 1

    where Xi and ri define the population density and

    intrinsic rate of increase of species i, to explore the

    dynamical behaviour of model communities close to

    equilibrium. The interaction between a predator and its

    prey is defined by the coefficient aij and represents the

    negative per-capita effect of a predator species j on a prey

    species i. The negative effects of species i on itself are

    denoted aii, while the positive effects of the prey species

    i on the predator j are given by aji. A system of such n

    linear equations describing the dynamics of a set ofspecies can be represented in matrix algebra terms so that

    the interaction coefficients (aij terms) are the elements of

    an n n matrix (here called A). Using matrix algebra, (1)

    above can be rewritten as:

    dX

    dt X r AX;

    where Xand rare vectors containing the densities (Xi,n)

    and intrinsic rates (ri,n) of those species present in the

    community or food web. In fact, May (1973) explored

    the stability of these model systems at equilibrium; he

    explored the dynamics of a matrix known as the

    Jacobian (C) whose elements (cij terms) were theproduct of the interaction coefficients contained in A

    and an n n matrix containing the vector of nontrivial

    equilibrium population densities (Xi where Xi does not

    equal zero) on the diagonal and zeros elsewhere (so

    that cij aijXi ). Essentially, the Jacobian matrix de-

    scribes the dynamics of a community close to an

    equilibrium point and details population-level interac-

    tions. May (1973) explored randomly constructed

    model communities by assigning interaction strengths

    (cij) from a uniform distribution to the elements of the

    Jacobian (C). He found that in systems with randomly

    assigned trophic interactions, an increased speciesrichness tended to decrease the stability of model

    communities. Pimm & Lawton (1977, 1978) investigat-

    ing simple model food webs subsequently showed that

    the patterning of species interactions (that is the

    presence or absence of omnivorous links and therefore

    food web structure alone) could have dramatic effects

    on the stability of such systems. Yodzis (1981) working

    with a compiled set of 40 real, as opposed to model,

    food webs found that the arrangement of species

    interaction strengths was essential for food web

    stability. When the detailed pattern of interaction

    strengths in stable food webs was disrupted the

    resulting webs were, on average, less stable.

    These classic studies using linear stability analyses, all

    made use of the Jacobian matrix for the determination of

    food web stability (May, 1973; Pimm & Lawton, 1977,

    1978; Yodzis, 1981). From these theoretical studies, there

    are two quantities that emerge as characterizing the

    interactions among species, aij the interaction coefficient

    and aijXi the element of the Jacobian matrix at equili-

    brium (it should be noted that the Jacobian exists

    G L O B A L C H A N G E A L T E R S T H E S T A B I L I T Y O F F O O D W E B S 491

    r 2005 Blackwell Publishing Ltd, Global Change Biology, 11, 490501

  • 8/14/2019 Global Change Emmerson.etal.GCB

    3/12

  • 8/14/2019 Global Change Emmerson.etal.GCB

    4/12

    Predictions suggest that atmospheric CO2 levels will

    rise from around 350700 ppm by the end of this

    century (IPCC, 2001). Associated with this increase in

    CO2 levels are: (1) direct effects of atmospheric carbon

    on plant photosynthesis and primary production; (2)

    increases in average global temperature (global average

    surface temperature has already increased by 0.6 1C in

    the 20th century, IPCC, 2001), and (3) changing patternsof precipitation (Aber & Melillo, 2001). Coupled atmo-

    sphereocean general circulation models provide us

    with the best predictions on how patterns of tempera-

    ture and precipitation will change as CO2 levels rise

    (IPCC, 2001). Most terrestrial regions will experience an

    increase in mean annual temperature between 3 and

    10 1C whereas rates of precipitation may either decrease

    or increase by up to 2 mm day1, depending on region.

    To begin integrating these climatic predictions into an

    understanding of species interactions, we begin our

    analysis with anticipated effects upon primary produc-

    tion. Then we will suggest how changes in primary

    production might influence the strength of trophicinteractions. Critical to this last step is distinguishing

    between changes in the rate of primary production and

    changes in the quality of the plant tissue that is

    produced. Whether plants react in a qualitative or

    quantitative way, or even both, is likely to have very

    different effects on species interactions.

    Direct effects of CO2 on primary production. It is generally

    accepted that photosynthetic rates are limited primarily

    by carbon (Drake et al., 1997; Norby et al., 1999; Aber &

    Melillo, 2001). Hence, most models predict that net

    primary production will increase as atmosphericconcentrations of CO2 increase. Various models

    predict changes that range from 0.7% to 1 32.4%

    (VEMAP, 1995). Overall, the average predicted increase

    in net primary production is around 20%. In many

    plant communities, primary production will increase

    below ground as well as above ground (Tate & Ross,

    1997) and elevated CO2 may increase fine root

    production of some forest trees by as much as 96%

    (King et al., 2001). For most plant species, increased

    rates of primary production are associated with

    increases in relative growth rate (Saxe et al., 1998). For

    example, in their study of 10 Acacia species, Atkin et al.

    (1999) reported that the relative growth rates of Acacia

    trees increased by an average of 10% over a 12-week

    period under elevated CO2 conditions.

    CO2-mediated changes in temperature and precipitation.

    While the prediction of increased productivity in

    response to rising concentrations of CO2 appears to be a

    relatively robust generalization, the effects of concomitant

    changes in temperature and precipitation are harder to

    predict (Aber & Melillo, 2001). Temperature effects are

    particularly difficult to predict because while photo-

    synthesis has a bell-shaped response to temperature,

    respiration increases exponentially (Larcher, 1995).

    Carbon gain relative to loss may therefore decline at

    high temperatures, resulting in a decline in production.

    The responses of various ecosystems to the

    interactive effects of changes in CO2, temperature andprecipitation will depend, in part, on their relative

    sensitivities to those environmental factors. Simply put,

    an ecosystem whose productivity is limited by

    precipitation is more likely to respond to a change in

    precipitation than one that is limited by temperature

    (Schloss et al., 1999). Broad generalizations that operate

    across many different ecosystems therefore, become less

    robust as the number of climatic variables considered is

    increased. Our ability to make predictions is also limited

    because models that include multiple interactive effects

    of global change are still relatively rare (Aber & Melillo,

    2001). One noteworthy study is that by Yu et al. (2002).

    This study incorporates predicted changes in CO2concentration, precipitation and temperature from

    seven general circulation models into predictions of

    ecosystem change in a forest transect in east China.

    While all seven general circulation models predicted

    increases in net primary productivity under elevated

    CO2, effects of precipitation were relatively weak

    whereas productivity was negatively correlated with

    temperature. Broadleaf forests were predicted to

    increase while conifer forests, shrubs and grasses were

    predicted to decrease. Overall, however, primary

    production was still seen to increase under all

    scenarios of elevated CO2 (Yu et al., 2002).Despite Yu et al. (2002) and predictions made from the

    bell-shaped relationship between photosynthesis and

    temperature (Aber & Melillo, 2001), increases in

    temperature under rising CO2 levels may still have an

    overall positive impact on net primary production,

    particularly if growing seasons are extended. A meta-

    analysis of available studies suggests that ecosystem

    warming should increase plant productivity by an

    average of 19%, with a 95% confidence interval of 15

    23% (Rustad et al., 2001). If true, then the effects of global

    warming on productivity may operate in exactly the

    same direction as the direct effects of CO2 on

    productivity.

    Quality vs. quantity. In a food chain context, the

    predicted increases in primary production may

    appear to favour herbivore populations; however,

    many empirical studies suggest that plant quality for

    consumers will decline as CO2 levels rise (Bezemer &

    Jones, 1998). This will have implications for the strength

    of per-capita trophic interactions between a consumer

    G L O B A L C H A N G E A L T E R S T H E S T A B I L I T Y O F F O O D W E B S 493

    r 2005 Blackwell Publishing Ltd, Global Change Biology, 11, 490501

  • 8/14/2019 Global Change Emmerson.etal.GCB

    5/12

    and its plant resource. Specifically, the accumulation of

    carbon under elevated CO2 dilutes concentrations of

    nitrogen in plant tissues by 1525% (Lincoln et al., 1993;

    Lindroth et al., 1995) thereby increasing C : N ratios

    (Ceulemans & Mousseau, 1994; Wilsey, 1996; Hughes &

    Bazzaz, 1997). Given that nitrogen is considered a

    limiting factor for many herbivore species (Mattson,

    1980), dilution of nitrogen under elevated CO2 mayplace further constraints on the growth rates of

    herbivores. In addition, the concentrations of carbon-

    rich secondary metabolites sometimes increase under

    elevated CO2 (Lindroth et al., 1995; Agrell et al., 2000).

    While there appears to be considerable variation in the

    responses of plant chemical defences to rising CO2levels (Hunter, 2001), some classes of tannin, such as

    condensed tannins, seem to respond most frequently.

    To summarize, in nearly every case examined to date,

    foliar nitrogen concentrations decline under elevated

    CO2 and, when present, foliar concentrations of

    condensed tannins increase (Fajer et al., 1989, 1991;

    Johnson & Lincoln, 1991; Lincoln et al., 1993; Lindrothet al., 1995; Jones & Hartley, 1999).

    Herbivores confronted with low-quality plant tissue

    would be expected to compensate by eating more. This

    prediction usually holds true, particularly for chewing

    insects (Lincoln et al., 1984, 1986, 1993; Fajer et al., 1989;

    Lindroth et al., 1993, 1995; Salt et al., 1995; Docherty

    et al., 1996; Kinney et al., 1997; Williams et al., 1997;

    Whittaker, 2000). Likewise, the area damaged by leaf-

    mining insects may also increase, for example, the area

    of leaf mines on Quercus myrtifolia increase by over 25%

    under elevated CO2, apparently because nitrogen

    concentrations fall by over 11% (Salt et al., 1995;Stiling et al., 2003). Nonchewing insects have been less

    well studied and in general patterns are yet to emerge.

    Bezemer et al. (1998), for example, showed that peach

    potato aphid (Myzus persicae) abundance was enhanced

    by elevated CO2 and temperature, but in a companion

    study, Bezemer et al. (1999) found that plant and aphid

    species significantly influenced the response.

    As well as separating quality and quantity effects, it

    is also crucial that we distinguish between overall

    levels of defoliation and per-capita consumption by

    insects. Increases in plant productivity and nitrogen-

    mediated declines in insect density can result in lower

    levels of defoliation on plants despite increases in per-

    capita consumption rates by herbivores (Hughes &

    Bazzaz, 1997, Stiling et al., 2003). A 2-year study of

    herbivore communities under open-topped CO2chambers in scrub oak forest has shown that all

    herbivore species decline in density under elevated

    CO2 (Stiling et al., 2003).

    While there is little information on how primary

    consumers (herbivores) respond, in the longer term, to

    changes in temperature and atmospheric CO2concentrations, there is even less information on the

    effects of climate change on secondary consumers

    (parasitoids and predators). In the few studies

    currently available on the direct effects of elevated

    CO2 on parasitoids, CO2 did not influence Cotesia

    melanoscela, a parasitoid of the gipsy moth (Lymantria

    dispar), although pre-enclosure mortality of theparasitoid was slightly increased when CO2 was

    elevated (Roth & Lindroth, 1995). Bezemer et al. (1998)

    also showed that parasitism rates of M. persicae

    remained unchanged in elevated CO2. Using

    mathematical models Hassell et al. (1991) have shown

    that hostparasitoid relationships are altered by

    environmental change; temperature elevation, for

    example, may differentially affect developmental rates

    of hosts and parasitoids. Such differences can

    potentially result in the breakdown of synchronization

    between the two populations, which, in turn, may have

    major effects on population dynamics.

    Models and methods

    Simulating the effects of change

    We use a LotkaVolterra framework, within which to

    explore the effects of change (defined here as an

    increase in productivity) on interaction strength. To

    investigate how increased productivity might affect

    species population sizes (Xi), per-capita interaction

    coefficients (aij) and, in turn, the elements of the

    Jacobian matrix (aijXi), we determined the period

    doubling bifurcations for a simple three species foodchain featuring omnivory in discrete time. The bifurca-

    tions occur when the stability of a system changes as a

    model parameter passes through a critical value.

    Essentially, the bifurcations describe how the dynamics

    of a species population change as a function of the

    species intrinsic rate of increase. The bifurcation

    diagram describes how a species population will

    change from a stable equilibrium, to limit cycles of

    varying amplitude and periodicity, to chaotic dynamics

    as the intrinsic rate of the basal species increases. In the

    simple food chain investigated here, Species 1 is basal,

    Species 2 is an herbivore and Species 3 is an omnivore

    feeding on Species 1 and 2 (see simple food web

    detailed in Fig. 1). Species population dynamics were

    determined using a discrete time version of the Lotka

    Volterra equation;

    X1;t1 X1;t X1;tb1 a11X1;t a12X2;t a13X3;t;

    X2;t1 X2;t X2;td2 a21X1;t a22X2;t a23X3;t;

    X3;t1 X3;t X3;td3 a31X1;t a32X2;t a33X3;t;

    2

    494 M . E M M E R S O N et al.

    r 2005 Blackwell Publishing Ltd, Global Change Biology, 11, 490501

  • 8/14/2019 Global Change Emmerson.etal.GCB

    6/12

    where birth and death processes have slightly different

    meanings for different trophic levels. b1 defines the

    positive per-capita birth rate of Species 1; basal species

    population growth is logistic so that b15 r1/K1, r1 being

    the intrinsic rate of increase of Species 1 and K1 the

    carrying capacity which is set equal to 1. For nonbasal

    species, d2 and d3 define the negative per-capita death

    rates of Species 2 and 3, set equal to 0.01 and 0.001,

    respectively (in the absence of food these species

    populations could not increase as for the basal auto-

    trophs. Consequently, they have a negative intrinsic

    population growth rate, which is offset by the food that

    they eat). The biological justification for this decline

    with increasing trophic height being that death rate is

    strongly correlated with a species body size. Body sizetends to increase with trophic height and so death rate

    would decline. Finally, t is time. Intraspecific terms also

    have different meanings for the different species

    present in the system. For the basal species a115 r1/K1,

    while for nonbasal species intraspecific competition is

    set equal to 0.1 (a22 and a33). Therefore, as r1 increases,

    the strength of basal species intraspecific competition

    increases relative to nonbasal species intraspecific

    competition.

    In this simple food chain, the secondary consumer

    (omnivorous predator) is capable of feeding on both the

    primary consumer (herbivore) and the basal resource

    (plant). A gradient of productivity was established byincrementing the intrinsic rate of increase (0.01 incre-

    ments), of the basal species r1 in the food chain. r1 was

    incremented over the interval [0,4]. To produce the

    bifurcations, at each value ofr1, the three discrete time

    equations of system (2) above, were iterated for 40 000

    time steps. Initial conditions for each run were set at

    X151, X25 0.5, X350.01, reflecting a biologically

    plausible pyramidal population structure (population

    density tends to decrease with increasing trophic height

    and is negatively correlated with increasing body size).

    The population density over the last 500 time steps of

    this series is represented in one dimension as a functionof the basal species intrinsic rate. Over this time interval,

    the system of equations converged either to an equili-

    brium value or onto stable attractors in the system.

    We explored three different scenarios to investigate

    how the persistence of this simple three species food

    chain might be affected given an increase in basal

    species productivity. These differing scenarios repre-

    sent situations where:

    (i) Interspecific and nonbasal species intraspecific

    interactions (a22 and a33) are considered constant over

    the range of basal species intrinsic rates. The per-capita

    effects of prey on predators (aji) are related to the per-

    capita effects of predators on prey (aij) by a predators

    conversion efficiency so that aji5 e aij, where e is 0.1.

    For the simulations presented here, the interaction

    coefficient matrix Awas parameterized in the following

    way:

    A

    rK 0:5 0:05

    0:05 0:1 0:30:0005 0:03 0:1

    24

    35:

    Fig1 Period doubling bifurcations for a three species omnivor-

    ous food web featuring (a) primary producer, (b) primary

    consumer and (c) omnivorous secondary consumer (for the foodweb diagram also shown filled circles indicate the position of the

    species in the food web). The omnivorous secondary consumer

    feeds on both the primary producer and the primary consumer.

    Three scenarios are represented: (i) Interspecific and consumer

    species intraspecific interactions are considered constant. (ii)

    Herbivorous interactions (aij) are considered a function of basal

    species intrinsic rates (r1), because as plant growth increases,

    plant quality declines and plant consumers must ingest more to

    compensate. (iii) Both herbivorous interactions and consumer

    assimilation efficiencies are considered to be functions of the

    primary producers intrinsic rate. This simulates increased plant

    quality at low levels of productivity and decreased quality at

    high levels of plant productivity (see accompanying text for

    further details).

    G L O B A L C H A N G E A L T E R S T H E S T A B I L I T Y O F F O O D W E B S 495

    r 2005 Blackwell Publishing Ltd, Global Change Biology, 11, 490501

  • 8/14/2019 Global Change Emmerson.etal.GCB

    7/12

    (ii) As the intrinsic rate of a plant increases, plant

    quality may decline. To compensate for this a herbivore

    must compensate and ingest more plant material to

    fulfil its nutritional requirements. Therefore, as the

    basal species intrinsic rate increases so too will the

    herbivores per-capita effects (aPH), where P and H refer

    to plant and herbivore, respectively. To simulate this

    scenario, we assumed herbivorous interactions (a12 anda13) in the simple food chain to be functions of the basal

    species intrinsic rate so that:

    aPH Fr1; 3

    where

    Fr1 l ur1

    b r1: 4

    Here, l and u are constants, approximately defining the

    lower and upper bounds to the interaction coefficient

    and were set at 0.2 and 0.8 for the primary consumer

    and 0.002 and 0.008 for the secondary consumer,

    respectively, b is a constant, set equal to unity. Forinteraction coefficients, this choice of l and u give a

    range from 0.2 to 1 for the primary consumer, and

    0.002 to 0.01 for the secondary consumer. Using this

    function, the interaction coefficient asymptotes as the

    intrinsic rate of the basal species increases. The

    biological justification for choosing this function is that

    herbivores are only capable of handling food at some

    maximal rate. This is essentially a Holling Type II

    functional response, but in this study, rather than

    ingestion being a function of prey density, it is a

    function of prey growth rate.

    (iii) At low levels of productivity, plant quality

    should be high, when prey intrinsic growth rates are

    low, the per-capita effects of herbivores on prey are

    consequently small. However, the benefits to predators

    should be large, because plant quality is high. It is

    difficult to represent this if it is assumed that the

    ecological efficiency of the predator remains constant.

    To simulate an increase in plant quality at low levels of

    plant productivity we assume that the ecological

    efficiency of a predator is also a function of plant

    intrinsic growth rate so that:

    e Fr1; 5

    where

    Fr1 u lr1

    b r1: 6

    As plant productivity increases, so the ability of a

    predator to convert plant biomass into predator

    biomass decreases asymptotically reflecting the fact

    that plant quality declines. Here, we set l and u both at

    0.3 this provides a biologically plausible range of 0.3

    0.06 for the ecological efficiency used in the herbivorous

    interactions of the primary and secondary consumers

    (Jonsson & Ebenman, 1998).

    Results

    Population size

    We constructed the period doubling bifurcations foreach of the scenarios ((i)(iii)) (Fig. 1). When interac-

    tion coefficients are independent of basal species

    intrinsic rate (Fig. 1, ac, i) the abundance of both

    primary and secondary consumer is low, being largely

    unresponsive to the behaviour of basal species popula-

    tion dynamics. However, the simple food web is not

    feasible when the intrinsic rate of the basal species is

    43 (there are no positive population densities). When

    herbivorous interaction coefficients are considered to be

    a function of basal species intrinsic rate (Fig. 1, ac, ii),

    basal species population dynamics remain largely

    unaffected. Nonbasal species populations, on the other

    hand, increase monotonically until the first basalspecies period doubling occurs. Subsequent increases

    in basal species intrinsic rate result in a slight decline in

    nonbasal species population size and an increased

    variability of each species population. When both,

    predator assimilation efficiency and interaction coeffi-

    cient are considered functions of basal species intrinsic

    rate (Fig. 1, ac, iii), the population size of nonbasal

    species both declines substantially and becomes more

    variable as basal species intrinsic rate increases. The

    decline in average population size and coincident

    increase in variability mean, that each population of

    these species may well be more prone to stochasticenvironmental perturbations as productivity increases.

    This is especially so when the scenarios that lead to a

    suggested increase in productivity also predict in-

    creased occurrence of extreme climatic events. Despite

    this, both scenarios ((ii) and (iii)) result in consumer

    population sizes that are still on average much larger

    than for Scenario (i) where interaction coefficients and

    assimilation efficiencies are independent of basal

    species productivity.

    Patterns of stability

    The mechanisms detailed above such as increased

    consumption with increased productivity, underlie

    herbivorous interactions and have implications for the

    stability of the simple three species food chain

    examined. Stability in the present context refers to the

    persistence of species in the food chain. For all three

    scenarios, persistence is determined largely by basal

    species population dynamics at high levels of produc-

    tivity. When interaction coefficients are insensitive to

    496 M . E M M E R S O N et al.

    r 2005 Blackwell Publishing Ltd, Global Change Biology, 11, 490501

  • 8/14/2019 Global Change Emmerson.etal.GCB

    8/12

    basal species productivity the chain does not persist

    when the basal species intrinsic rate exceeds 3 (r143).

    When only interaction coefficients are considered to be

    a function of basal species intrinsic rate, no species are

    lost from the food chain up to an intrinsic rate of 3.16

    (and the food chain persists). When both interaction

    coefficient and assimilation efficiency are considered as

    functions of r1, the chain does not persist abover15 3.11. At no point in the simulations do either

    species 2 or 3 become extinct, affecting the persistence

    of the entire food chain. This potentially could occur

    with either the primary consumer becoming extinct,

    leaving the omnivorous secondary consumer or the

    omnivorous secondary consumer itself could become

    extinct, leaving only the basal species and the primary

    consumer. Despite the three differing scenarios, we find

    that there is little qualitative difference in terms of

    species persistence. However, the differing scenarios do

    affect consumer population sizes differently. What,

    then, are the implications for the patterning of Jacobian

    elements (aijXi)?

    Patterning of interaction strengths

    To examine the distribution of predatory interaction

    coefficients or Jacobian elements for a simple three-

    species food chain is inappropriate, as the three

    predatory food chain interactions detailed here do not

    constitute a comprehensive distribution. Instead, we

    consider what happens to the Jacobian elements in the

    food chain as basal species productivity increases. Put

    simply, the patterning of Jacobian elements (here a12 see

    Fig. 2a) mirrors the patterning of each species popula-tion size. As basal species productivities increase, and

    each species population undergoes bifurcations, so the

    Jacobian elements become more variable. For Scenario

    (ii), Jacobian elements become increasingly negative

    (stronger) with increases in productivity. As interaction

    coefficients increase as a function of basal species

    intrinsic rate and assimilation efficiency remains con-

    stant, each consumer population size increases. For

    Scenario (iii), the Jacobian elements become weaker but

    still increase in variability. This decline in the strength

    of Jacobian elements is a consequence of the predatory

    interaction coefficients increasing as basal species

    productivities increase while the predators assimila-

    tion efficiency declines, resulting in a smaller predator

    population size overall. Given the scenarios we explore

    here, the patterning of the Jacobian elements seems to

    have little consequence for the persistence of the food

    chain, which remains feasible over the gradient of basal

    species productivities. At high productivities, the

    simple food chain and its constituent species popula-

    tions do not exist at a simple equilibrium (Fig. 1ac),

    but rather exist in proximity to some form of cyclic or

    chaotic attractor.

    Continuous time

    Until now, we have only considered the discrete time

    version of the LotkaVolterra equation ((2) above) and

    we have seen that the system converges to an

    equilibrium point or stable attractor in phase space. If

    we use the same parameter values for the interaction

    coefficients and birth and death processes using the

    continuous form of the LotkaVolterra equation (Eqn

    (1)), then a local stability analysis of the omnivorous

    food chain can be carried out. A stability analysis

    determines the ability of a community to return to

    equilibrium following a small perturbation that moves

    a community away from the equilibrium point. Such a

    stability analysis reveals that in close proximity to an

    attractor the food chain is always stable (see Fig. 2b),

    although if a perturbation moves a community a

    large distance from the attractor (which defines the

    Fig. 2 The consequences of Scenarios (i)(iii) (see text) for (a)

    Patterning of Jacobian elements (c12) as a function of primary

    producers intrinsic rate (r1) using LotkaVolterra dynamics in

    discrete time (Eqn (2)), and (b) Stability of three species food

    web measured as return time (1/Re(l1)) to equilibrium. For the

    determination of return time the model parameterizations

    suggested by Scenarios (i)(iii) are used to parameterize

    LotkaVolterra dynamics in continuous time (see Eqn (1)).

    G L O B A L C H A N G E A L T E R S T H E S T A B I L I T Y O F F O O D W E B S 497

    r 2005 Blackwell Publishing Ltd, Global Change Biology, 11, 490501

  • 8/14/2019 Global Change Emmerson.etal.GCB

    9/12

  • 8/14/2019 Global Change Emmerson.etal.GCB

    10/12

    temperature dependent and clearly scenarios of global

    warming also have implications for the strength of

    trophic interactions via a quite different mechanism

    than those we envisage here. The temperature depen-

    dence of basal metabolic rate clearly has implications

    for food web dynamics and ecosystem functioning.

    Linking the study of ecosystem dynamics through the

    study of interacting populations, metabolic theory andecological stoichiometry would be a fruitful area of

    research.

    In the present study, we chose a simple food web and

    a range of parameter values that produced feasible

    communities (where all species populations have

    positive densities) and have explored how changes to

    these parameter values affect community feasibility. At

    first, this might seem trivial but feasibility is of

    paramount importance. The exploration of complex

    food web dynamics is problematic mainly because

    finding parameter values that result in feasible com-

    munities is very difficult. We explored three scenarios

    that might affect a communitys coordinates in para-meter space and investigated whether predicted sce-

    narios of change might lead a community along some

    trajectory into an unfeasible region, where one or more

    species populations would decline or become extinct. In

    general for the scenarios we have explored, increased

    productivity does not result in decreased community

    persistence. However, we have investigated just one

    possible food web structure. There are many possible

    (simple) food web scenarios that remain to be explored,

    for example those that include apparent competition or

    intraguild predation.

    Our discussion of change has been limited to theeffects of increased productivity on the strength of a

    class of consumer resource interactions (plantherbi-

    vore). We have not considered other forms of change

    for example; changes in the size distribution of

    harvested populations, and how this might affect the

    strength or existence of size-based feeding relation-

    ships. We have not considered the effects of increased

    O3 on ultraviolet light radiation levels and how this

    might affect species in food webs detrimentally. There

    is a range of climatic factors that we have not

    considered and importantly, these remain to be ex-

    plored. We have not considered how species might

    adapt or respond evolutionarily to possible changes in

    the global biosphere and such scenarios remain an area

    ripe for research. Here, we show that simplistic models

    can provide important mechanistic insights into the

    causes of empirically observed patterns in real com-

    munities and the types of patterns we might expect to

    see as the global biosphere changes.

    Even within the limited scope of this study, we can

    suggest a number of consequences of change for

    communities. When change is envisaged as an

    increase in productivity: (1) herbivorous interaction

    coefficients (per-capita effects) will become stronger as

    consumers fulfil their dietary requirements, faced with

    a fast-growing poor-quality food resource. While per-

    capita interactions could increase, the strength of

    population-level effects (described by the Jacobian

    matrix) might decline if population sizes coincidentallydecrease. Neutel et al. (2002) using complex soil food

    webs have shown that the pattern of weak interactions

    in omnivorous loops is vital for food web stability. A

    disruption to the patterning of such interactions in

    more complex food webs than the simplistic web we

    explore here could have dramatic consequences for

    food web and ecosystem stability. (2) Average popula-

    tion size of consumers will likely decline as a

    consequence of poor-quality food resources; (3) Jaco-

    bian elements will become weaker as a consequence of

    reduced population size; and (4) species populations

    will become more variable and may take longer to

    recover from environmental or anthropogenic pertur- bations. This last point (4), is important as extreme

    stochastic environmental effects are predicted to in-

    crease with predicted scenarios of global warming. The

    combined effect of small population size, increased

    population variability and increased occurrence of

    stochastic environmental effects will likely result in an

    increased chance of extinction for the resulting popula-

    tions. In conclusion, empirical validation of our

    simulations is necessary but on the basis of this study

    communities may become inherently less stable, that is

    less resilient, smaller, more variable and hence more

    prone to extinction as they respond to predicted globalenvironmental change.

    Acknowledgements

    Special thanks to Bob Paine, Tim Wootton, Bill Fagan, DaveRaffaelli and Peter deRuiter for the provision of raw data. JohnPitchford for valuable comments. We thank all those thatparticipated in discussion during the Food webs in a changingworld meeting, Texel, the Netherlands. We also thank threeanonymous reviewers for constructive and useful comments.

    ReferencesAber JD, Melillo JM (2001) Terrestrial Ecosystems, 2nd edn.

    Academic Press, San Diego.

    Agrell J, McDonald EP, Lindroth RL (2000) Effects of CO2 and

    light on tree phytochemistry and insect performance. Oikos,

    88, 259272.

    Atkin OK, Schortemeyer M, McFarlane N et al. (1999) The

    response of fast- and slow-growing Acacia species to elevated

    atmospheric CO2: an analysis of the underlying components

    of relative growth rate. Oecologia, 120, 544554.

    G L O B A L C H A N G E A L T E R S T H E S T A B I L I T Y O F F O O D W E B S 499

    r 2005 Blackwell Publishing Ltd, Global Change Biology, 11, 490501

  • 8/14/2019 Global Change Emmerson.etal.GCB

    11/12

    Berlow EL, Navarrete SA, Briggs CJ et al. (1999) Quantifying

    variation in the strengths of species interactions. Ecology, 80,

    22062224.

    Bezemer TM, Jones TH (1998) Plantinsect herbivore interactions

    in elevated atmospheric CO2: quantitative analyses and guild

    effects. Oikos, 82, 212222.

    Bezemer TM, Jones TH, Knight KJ (1998) Long-term effects of

    elevated CO2 and temperature on populations of the peach

    potato aphid Myzus persicae and its parasitoid Aphidiusmatricariae. Oecologia, 116, 128135.

    Bezemer TM, Knight KJ, Newington JE et al. (1999) How general

    are aphid responses to elevated atmospheric CO2. Annals of the

    Entomological Society of America, 92, 724730.

    Ceulemans R, Mousseau M (1994) Tansley review no 71: effects

    of elevated atmospheric CO2 on woody plants. New Physiol-

    ogist, 127, 425446.

    Chapin F III, Zavelta E, Eviner V et al. (2000) Consequences of

    changing biodiversity. Nature, 405, 234242.

    Chase JM, Leibold MA, Downing AL et al. (2000) The effects of

    productivity, herbivory, and plant species turnover in grass-

    land food webs. Ecology, 81, 24852497.

    de Ruiter PC, Neutel A-M, Moore JC (1995) Energetics, patterns

    of interaction strengths, and stability in real ecosystems.

    Science, 269, 12571260.

    Docherty M, Hurst DK, Holopainen JK et al. (1996) Carbon

    dioxide induced changes in beech foliage cause female beech

    weevil larvae to feed in a compensatory manner. Global Change

    Biology, 2, 335341.

    Drake B, Gonzalez-Meler M, Long SP (1997) More efficient plants:

    a consequence of rising atmospheric CO2. Annual Review of

    Plant Physiology and Plant Molecular Biology, 48, 607637.

    Duarte CM (2000) Marine biodiversity and ecosystem services:

    an elusive link. Journal of Experimental Marine Biology and

    Ecology, 250, 117131.

    Duffy JE (2002) Biodiversity and ecosystem function: the

    consumer connection. Oikos, 99, 201219.Elton CS (1958) The Ecology of Invasions by Animals and Plants.

    Methuen, London.

    Emmerson MC, Raffaelli DG (2004) Predatorprey body size,

    interaction strength and the stability of a real food web. Journal

    of Animal Ecology, 73, 399409.

    Engelhardt KAM, Ritchie ME (2001) Effects of macrophyte

    species richness on wetland ecosystem functioning and

    services. Nature, 411, 687689.

    Fagan WF, Hurd LE (1994) Hatch density variation of a

    generalist arthropod predator: population consequences and

    community impact. Ecology, 75, 20222032.

    Fajer ED, Bowers MD, Bazzaz FA (1989) The effects of enriched

    carbon dioxide atmospheres on plantinsect herbivore inter-

    actions. Science, 243, 11981200.

    Fajer ED, Bowers MD, Bazzaz FA (1991) The effects of enriched

    CO2 atmospheres on the buckeye butterfly, Junonia coenia.

    Ecology, 72, 751754.

    Hassell MP, May RM, Pacala SW et al. (1991) The persistence of

    hostparasitoid associations in patchy environments. 1. A

    general criterion. The American Naturalist, 138, 568583.

    Hughes L, Bazzaz FA (1997) Effect of elevated CO2 on

    interactions between the western flower thrips, Frankliniella

    occidentalis (Thysanoptera: thripidae) and the common milk-

    weed, Asclepias syriaca. Oecologia, 109, 286290.

    Hunter MD (2001) Effects of elevated atmospheric carbon

    dioxide on insect-plant interactions. Agricultural and Forest

    Entomology, 3, 153159.

    Intergovernmental Panel on Climate Change (IPCC) (2001)

    Climate Change 2001: Synthesis Report. Cambridge University

    Press, New York.

    Johnson RH, Lincoln DE (1991) Sagebrush carbon allocationpatterns and grasshopper nutrition: the influence of CO2enrichment and soil mineral limitation. Oecologia, 87, 127134.

    Jones TH, Hartley SE (1999) A protein competition model of

    phenolic allocation. Oikos, 86, 2744.

    Jonsson T, Ebenman B (1998) Effects of predatorprey body size

    ratios on the stability of food chains. Journal of Theoretical

    Biology, 193, 407417.

    King J, Pregitzer K, Zak D et al. (2001) Fine-root biomass and

    fluxes of soil carbon in young stands of paper birch and

    trembling aspen as affected by elevated atmospheric CO2and trophospheric O3. Oecologia, 128, 237250.

    Kinney KK, Lindroth RL, Jung SM et al. (1997) Effects of CO2 and

    NO3 availability on deciduous trees: phytochemistry and

    insect performance. Ecology, 78, 215230.

    Larcher W (1995) Physiological Plant Ecology, 3rd edn. Springer

    Verlag, Berlin.

    Lerdau M, Slobodkin L (2002) Trace gas emissions and species-

    dependent ecosystem services. Trends in Ecology and Evolution,

    17, 309312.

    Lincoln DE, Couvet D, Sionit N (1986) Response of an insect

    herbivore to host plants grown in carbon dioxide enriched

    atmospheres. Oecologia, 69, 556560.

    Lincoln PE, Fajer ED, Johnson RH (1993) Plant-insect herbivore

    interactions in elevated CO2. Trends in Ecology and Evolution, 8,

    6468.

    Lindroth RL, Arteel GE, Kinney KK (1995) Responses of three

    saturniid species to paper birch grown under enriched CO2atmospheres. Functional Ecology, 9, 306311.

    Lindroth RL, Kinney KK, Platz CL (1993) Responses of

    deciduous trees to elevated atmospheric CO2: productivity,

    phytochemistry, and insect performance. Ecology, 74, 763

    777.

    Lincoln DE, Sionit N, Strain BR (1984) Growth and feeding

    response ofPseudoplusia includens (Lepidoptera: Noctuidae) to

    host plants grown in controlled carbon dioxide atmospheres.

    Environmental Entomology, 13, 15271530.

    Loreau M, Naeem S, Inchausti P et al. (2001) Biodiversity and

    ecosystem functioning: current knowledge and future chal-

    lenges. Science, 294, 804808.

    MacArthur RH (1955) Fluctuations of animal populations and a

    measure of community stability. Ecology, 36, 533536.

    Madritch MD, Hunter MD (2002) Phenotypic diversity influ-

    ences ecosystem functioning in an oak sand hills community.

    Ecology, 83, 20842090.

    Mattson WJ (1980) Herbivory in relation to plant nitrogen

    content. Annual Review of Ecology and Systematics, 11, 119

    161.

    May RM (1973) Stability and Complexity in Model Ecosystems, 2nd

    edn. Princeton University Press, Princeton.

    500 M . E M M E R S O N et al.

    r 2005 Blackwell Publishing Ltd, Global Change Biology, 11, 490501

  • 8/14/2019 Global Change Emmerson.etal.GCB

    12/12