journal of theoretical biology etal_a... · 2016. 10. 12. · a dynamical model for bark beetle...

13
A dynamical model for bark beetle outbreaks Vlastimil Křivan a,b,n , Mark Lewis c , Barbara J. Bentz d , Sharon Bewick e , Suzanne M. Lenhart e , Andrew Liebhold f a Institute of Entomology, Biology Centre, Czech Academy of Sciences, Branišovská 31, 370 05 České Budějovice, Czech Republic b Faculty of Sciences, University of South Bohemia, Branišovská 1760, 370 05 České Budějovice, Czech Republic c Department of Mathematical and Statistical Sciences, University of Alberta, Edmonton, Canada T6G 2G1 d USFS Rocky Mountain Research Station, 860 N. 1200 East, Logan, UT 84321, USA e National Institute for Mathematical and Biological Synthesis, 1534 White Avenue, Knoxville, TN 37996-1527, USA f USDA Forest Service, 180 Caneld St., Morgantown, WV 26505, USA HIGHLIGHTS A new epidemiological model for bark beetle outbreaks is developed. This model considers beetle aggregation dynamics and tree resistance to infestation. The resulting model is described by a differential equation with discontinuous right-hand side. Conditions that relate tree resistance, forest regeneration rate, rate of infestation by beetles, and immigration to the forest state are given. Analytical conditions when forest dies, recovers, or infestation becomes endemic are given. The case of infestation spread between patches is studied using a two stand system. article info Article history: Received 4 November 2015 Received in revised form 13 June 2016 Accepted 6 July 2016 Available online 7 July 2016 Keywords: Bistability Bark beetle Dendroctonus ponderosae Dispersal Filippov solution Hysteresis Population dynamics Stability SI models abstract Tree-killing bark beetles are major disturbance agents affecting coniferous forest ecosystems. The role of environmental conditions on driving beetle outbreaks is becoming increasingly important as global climatic change alters environmental factors, such as drought stress, that, in turn, govern tree resistance. Furthermore, dynamics between beetles and trees are highly nonlinear, due to complex aggregation behaviors exhibited by beetles attacking trees. Models have a role to play in helping unravel the effects of variable tree resistance and beetle aggregation on bark beetle outbreaks. In this article we develop a new mathematical model for bark beetle outbreaks using an analogy with epidemiological models. Because the model operates on several distinct time scales, singular perturbation methods are used to simplify the model. The result is a dynamical system that tracks populations of uninfested and infested trees. A limiting case of the model is a dis- continuous function of state variables, leading to solutions in the Filippov sense. The model assumes an extensive seed-bank so that tree recruitment is possible even if trees go extinct. Two scenarios are considered for immigration of new beetles. The rst is a single tree stand with beetles immigrating from outside while the second considers two forest stands with beetle dispersal between them. For the seed-bank driven re- cruitment rate, when beetle immigration is low, the forest stand recovers to a beetle-free state. At high beetle immigration rates beetle populations approach an endemic equilibrium state. At intermediate immigration rates, the model predicts bistability as the forest can be in either of the two equilibrium states: a healthy forest, or a forest with an endemic beetle population. The model bistability leads to hysteresis. Interactions between two stands show how a less resistant stand of trees may provide an initial toe-hold for the invasion, which later leads to a regional beetle outbreak in the resistant stand. & 2016 Elsevier Ltd. All rights reserved. 1. Introduction Tree-killing bark beetles (Coleoptera: Curculionidae, Scolyti- nae) are important disturbance agents affecting coniferous forest ecosystems, and population outbreaks have resulted in extensive, Contents lists available at ScienceDirect journal homepage: www.elsevier.com/locate/yjtbi Journal of Theoretical Biology http://dx.doi.org/10.1016/j.jtbi.2016.07.009 0022-5193/& 2016 Elsevier Ltd. All rights reserved. n Corresponding author at: Institute of Entomology, Biology Centre, Czech Academy of Sciences, Branišovská 31, 370 05 České Budějovice, Czech Republic. E-mail addresses: [email protected] (V. Křivan), [email protected] (M. Lewis), [email protected] (B.J. Bentz), [email protected] (S. Bewick), [email protected] (S.M. Lenhart), [email protected] (A. Liebhold). Journal of Theoretical Biology 407 (2016) 2537

Upload: others

Post on 30-Jan-2021

0 views

Category:

Documents


0 download

TRANSCRIPT

  • Journal of Theoretical Biology 407 (2016) 25–37

    Contents lists available at ScienceDirect

    Journal of Theoretical Biology

    http://d0022-51

    n CorrAcadem

    E-mmark.lesharon.baliebhol

    journal homepage: www.elsevier.com/locate/yjtbi

    A dynamical model for bark beetle outbreaks

    Vlastimil Křivan a,b,n, Mark Lewis c, Barbara J. Bentz d, Sharon Bewick e,Suzanne M. Lenhart e, Andrew Liebhold f

    a Institute of Entomology, Biology Centre, Czech Academy of Sciences, Branišovská 31, 370 05 České Budějovice, Czech Republicb Faculty of Sciences, University of South Bohemia, Branišovská 1760, 370 05 České Budějovice, Czech Republicc Department of Mathematical and Statistical Sciences, University of Alberta, Edmonton, Canada T6G 2G1d USFS Rocky Mountain Research Station, 860 N. 1200 East, Logan, UT 84321, USAe National Institute for Mathematical and Biological Synthesis, 1534 White Avenue, Knoxville, TN 37996-1527, USAf USDA Forest Service, 180 Canfield St., Morgantown, WV 26505, USA

    H I G H L I G H T S

    � A new epidemiological model for bark beetle outbreaks is developed.

    � This model considers beetle aggregation dynamics and tree resistance to infestation.� The resulting model is described by a differential equation with discontinuous right-hand side.� Conditions that relate tree resistance, forest regeneration rate, rate of infestation by beetles, and immigration to the forest state are given.� Analytical conditions when forest dies, recovers, or infestation becomes endemic are given.� The case of infestation spread between patches is studied using a two stand system.

    a r t i c l e i n f o

    Article history:Received 4 November 2015Received in revised form13 June 2016Accepted 6 July 2016Available online 7 July 2016

    Keywords:BistabilityBark beetleDendroctonus ponderosaeDispersalFilippov solutionHysteresisPopulation dynamicsStabilitySI models

    x.doi.org/10.1016/j.jtbi.2016.07.00993/& 2016 Elsevier Ltd. All rights reserved.

    esponding author at: Institute of Entomoy of Sciences, Branišovská 31, 370 05 České Bail addresses: [email protected] ([email protected] (M. Lewis), [email protected]@gmail.com (S. Bewick), lenhart@[email protected] (A. Liebhold).

    a b s t r a c t

    Tree-killing bark beetles are major disturbance agents affecting coniferous forest ecosystems. The role ofenvironmental conditions on driving beetle outbreaks is becoming increasingly important as global climaticchange alters environmental factors, such as drought stress, that, in turn, govern tree resistance. Furthermore,dynamics between beetles and trees are highly nonlinear, due to complex aggregation behaviors exhibited bybeetles attacking trees. Models have a role to play in helping unravel the effects of variable tree resistanceand beetle aggregation on bark beetle outbreaks. In this article we develop a new mathematical model forbark beetle outbreaks using an analogy with epidemiological models. Because the model operates on severaldistinct time scales, singular perturbation methods are used to simplify the model. The result is a dynamicalsystem that tracks populations of uninfested and infested trees. A limiting case of the model is a dis-continuous function of state variables, leading to solutions in the Filippov sense. The model assumes anextensive seed-bank so that tree recruitment is possible even if trees go extinct. Two scenarios are consideredfor immigration of new beetles. The first is a single tree stand with beetles immigrating from outside whilethe second considers two forest stands with beetle dispersal between them. For the seed-bank driven re-cruitment rate, when beetle immigration is low, the forest stand recovers to a beetle-free state. At high beetleimmigration rates beetle populations approach an endemic equilibrium state. At intermediate immigrationrates, the model predicts bistability as the forest can be in either of the two equilibrium states: a healthyforest, or a forest with an endemic beetle population. The model bistability leads to hysteresis. Interactionsbetween two stands show how a less resistant stand of trees may provide an initial toe-hold for the invasion,which later leads to a regional beetle outbreak in the resistant stand.

    & 2016 Elsevier Ltd. All rights reserved.

    logy, Biology Centre, Czechudějovice, Czech Republic.Křivan),(B.J. Bentz),.utk.edu (S.M. Lenhart),

    1. Introduction

    Tree-killing bark beetles (Coleoptera: Curculionidae, Scolyti-nae) are important disturbance agents affecting coniferous forestecosystems, and population outbreaks have resulted in extensive,

    www.sciencedirect.com/science/journal/00225193www.elsevier.com/locate/yjtbihttp://dx.doi.org/10.1016/j.jtbi.2016.07.009http://dx.doi.org/10.1016/j.jtbi.2016.07.009http://dx.doi.org/10.1016/j.jtbi.2016.07.009http://crossmark.crossref.org/dialog/?doi=10.1016/j.jtbi.2016.07.009&domain=pdfhttp://crossmark.crossref.org/dialog/?doi=10.1016/j.jtbi.2016.07.009&domain=pdfhttp://crossmark.crossref.org/dialog/?doi=10.1016/j.jtbi.2016.07.009&domain=pdfmailto:[email protected]:[email protected]:[email protected]:[email protected]:[email protected]:[email protected]://dx.doi.org/10.1016/j.jtbi.2016.07.009

  • V. Křivan et al. / Journal of Theoretical Biology 407 (2016) 25–3726

    landscape scale tree mortality events globally (Schelhaas et al.,2003; Meddens et al., 2012). In their native habitats, bark beetle-caused tree mortality, and its interactions with other dis-turbances including fire, play key roles in forest succession,species composition, and nutrient cycling (Hicke et al., 2013;Hansen, 2014). Recently, however, changing climate is alteringbark beetle outbreak dynamics indirectly, through effects to hosttrees (Chapman et al., 2012; Gaylord et al., 2013; Hart et al.,2014), and directly, by influencing beetle phenology, voltinismand the probability of survival (Bentz et al., 2010; Safranyik et al.,2010; Bentz et al., 2014; Weed et al., 2015). With continuedchanges in climate, trajectories of future forest succession will bealtered in ways that could have significant negative impacts onother native species as well as on biodiversity in general (Bentzet al., 2010; Fettig et al., 2013).

    The biology of tree killing bark beetles is complex and vari-able. Most species interact in mutualistic relationships withfungi, bacteria, mites and other organisms that provide protec-tion and nutrition, and help in detoxifying host plant chemicaldefenses (Boone et al., 2013; Hofstetter et al., 2015; Therrienet al., 2015). Native host tree species exhibit formidable con-stitutive and induced defenses that protect them from barkbeetle attacks when beetle population levels are low (Raffa et al.,2008). These defenses, however, can be overcome as beetlenumbers increase (Boone et al., 2011). As a result, many tree-killing bark beetle species have evolved chemically mediatedaggregative behaviors that depend on host tree chemicals, andallow them to attack en masse and at higher densities than wouldbe possible in the absence of coordination (Raffa et al., 2008). Incontrast, some other bark beetle species lack the feedback me-chanisms that facilitate mass attacks and instead colonize hosttrees that have reduced defenses due to a variety of stressorssuch as drought, fire or wind injury, and pathogens. The interplaybetween the threshold dependent colonization success andbeetle density, combined with the unique aggregative strategiesexhibited by many bark beetle species, leads to complex beetleoutbreak dynamics.

    The spatial and temporal dynamics of bark beetle populationoutbreaks will vary across the range of a given species, and alsowith the level of aggressiveness among species. Population out-breaks of those species without the feedback mechanisms thatdrive aggregative attacks are rare, and these species exhibit littleinter-annual variability in abundance. There are exceptions,however, including a large drought-driven outbreak of Ips speciesin the southwestern United States between 2002 and 2004(Santos and Whitham, 2010). When drought conditions subsided,so did the population outbreak. In contrast, species that exhibitfeedback mechanisms facilitating aggregation of large numbersof beetles in order to successfully colonize healthy trees (i.e.,aggressive species), including Dendroctonus ponderosae and D.frontalis, exhibit considerable temporal variability in abundance.Populations can exist at low levels for many years with oftenrapid eruptions to outbreak levels as a result of population in-dependent processes such as weather or delays, and non-linearities in density-dependent processes (Berryman, 1982;Martinson et al., 2013). Although the triggers for outbreaks ofthese aggressive species are varied and not well understood, treeresistance and weather can play large roles. The most resistanttrees often also have the greatest food resource for developingbeetles but require mass attacks to overwhelm the defenses.Compromised defenses through stressors that include drought(Anderegg et al., 2015) and pathogens (Goheen and Hansen,1993) can result in a tree being overwhelmed by fewer beetles.This can lead to build up of population in the less resistant treesand eventually becoming large enough to attack more vigoroustrees with greater food resources. Indeed, large scale outbreaks of

    aggressive species require large expanses of relatively vigoroushost trees (Fettig et al., 2014). In contrast, species that are in-capable of attacking vigorous trees are often found in areaswhere trees grow on marginal sites and stressed trees are com-monly available. For both aggressive and less aggressive beetlespecies, weather that is favorable for survival and seasonality ofbeetles and their associates is also required for outbreak initia-tion (Bentz et al., 2014; Addison et al., 2015; Weed et al., 2015).The complex interaction of tree resistance and weather can resultin considerable intra-range variation in population dynamics of agiven species as environmental conditions that influence hosttree resistance and beetle population dynamics vary temporallyand spatially. Low host tree resistance can influence the initiationof outbreaks of aggressive bark beetle species and can sustainoutbreaks of less aggressive species.

    To better understand the influence of aggressive attacks ontrees, we use a susceptible/infective (S/I) model to explore thelong-term dynamic interactions between forests ecosystems andbark beetle population dynamics. We assume that tree recruit-ment is not limited by seeds. We focus our analysis on the effect oftree resistance on the forest state. In particular, we show that,when resistance is low, the forest can be either beetle-free, or canhave an endemic beetle population depending on forest history,while, for high resistance, the forest will be beetle-free.

    1.1. Review of existing models

    Several models of bark beetle population dynamics alreadyexist. Here we review and compare the essential features of thesemodels in order to put our current study into context. Given theimportance of temperature in not only triggering but also sus-taining bark beetle outbreaks, several models have been devel-oped that incorporate temperature alone (Gilbert et al., 2004;Regniere and Bentz, 2007; Friedenberg et al., 2007) and thecombined effects of temperature and host trees (Powell and Bentz,2009, 2014) on bark beetle population success. For the purpose ofthis article, however, we restrict ourselves to consideration ofsimple, strategic models without the effects of climate that areamenable to mathematical analysis of general system properties.To facilitate comparison, we consider similarities and differencesin three structural aspects: the representation of forest structureand dynamics, the relationship between beetle density and treedeath, and the relationship between tree death and new beetleproduction.

    Because we are interested in the role of host resistance in long-term outbreak dynamics, sensible choices about the representa-tion of natural forest structure and regeneration are essential.Depending on the perspective and scenario under analysis, pre-vious modeling efforts have focused on different aspects of foreststructure. Berryman et al. (1984) and Økland and Bjørnstad (2006),for example, modeled a live forest class, and a transient, newlykilled tree class that they assumed was not resistant to beetles.Heavilin and Powell (2008) also allowed for two forest classes thatdiffered in their resistance, although, in this study, the less re-sistant class was allowed at least some level of resistance. Morerecently, Duncan et al. (2015) developed yet another two-classmodel. In this case, however, susceptible and resistant classeswere mechanistically linked to forest age structure. In reality, ofcourse, stands can be composed of many different types of treeswith varying resistance levels. Lewis et al. (2010) accounted forthis by allowing any possible distribution of vigor within a stand.Unfortunately, total generality comes at the expense of compli-cated and analytically intractable models. In the current study, wereturn to a simpler representation of internally homogeneousforest classes or cohorts consistent with the treatment by Heavilinand Powell (2008) and Duncan et al. (2015).

  • V. Křivan et al. / Journal of Theoretical Biology 407 (2016) 25–37 27

    The crucial mechanism of outbreak initiation is that beetledensity exceeds a threshold so that beetles can successfully attackthe dominant cohort of trees. One possibility is that resistance ofthe dominant cohort changes over time. For example, Berrymanet al. (1984) assumed that resistance decreases as live stem densityincreases. In this model, remaining trees regain resistance to at-tack when an outbreak thinned a stand. Although thinned standsmay be less susceptible to attack at low population levels, eventhinned stands can be heavily attacked during outbreaks (Fettiget al., 2014). However, in reality, old trees are more susceptible tomost tree-killing bark beetles than are young trees, regardless ofstand density. An alternative method of varying resistance overtime is to explicitly model the transition from highly resistantyoung trees to more susceptible old trees, as was done in Heavilinand Powell (2008).

    Although tree resistance changes dynamically through timedue to processes like aging and crowding, certain forest stands areinherently more or less resistant as a result of environmentalfactors, e.g., water stress (Anderegg et al., 2015). This spatial aspectof tree resistance has been less well studied from a modelingperspective. Nevertheless, the role of environmental conditions indriving beetle outbreaks will likely become increasingly importantas global climatic change alters environmental factors, for exampleby enlarging regions of drought stress. In this study, rather thanfocusing on aging and crowding as drivers of outbreak cycles, weinstead focus on how spatial and environmental drivers influencehost tree resistance and subsequent bark beetle outbreakdynamics.

    A final aspect of forest structure and dynamics is regeneration.One approach to modeling forest regeneration (Berryman et al.,1984; Økland and Bjørnstad, 2006) is a standard density-depen-dent growth model, where growth rate is proportional to theabundance of adult trees, and adult density increases to a carryingcapacity. However, pines, in particular, are characteristically shadeintolerant, and many species such as lodgepole pine, Pinus con-torta, are characterized by their maintenance of large seed banksof serotinous cones that do not germinate until after a stand re-placing disturbance (Johnson and Fryer, 1989). We therefore sug-gest that a model without recruitment limitation of the tree po-pulation may be a better representation of forest dynamics inmany of the systems susceptible to aggressive bark beetleoutbreaks.

    In practice, beetle population density is rarely monitored di-rectly. Instead, the number or proportion of infested trees is usedas a proxy for population density (sensu Meddens et al., 2012). Tobuild a simple model that can be compared to data, we followHeavilin and Powell (2008) and assume that the number of beetlesemerging from each successfully attacked tree is independent ofthe number of beetles that attacked the tree. This assumption ismet if the number of beetles required for successful attack isgreater than or equal to the number of beetles that can completelyexploit tree resources. Other models (Berryman et al., 1984; Powellet al., 1996; Økland and Bjørnstad, 2006; Lewis et al., 2010) haveexplicitly included intraspecific competition, thereby allowing amore complex relationship between attacking and emergingbeetles. Again, however, this detail comes at the expense of modeltransparency and tractability, thus we prefer the simpler for-mulation in Heavilin and Powell (2008) and leave more compli-cated relationships between beetle density and tree infestation fora future study.

    When stands are healthy, with a majority of trees that areresistant to beetle attack, it is difficult for low numbers ofbeetles to overcome tree resistance and colonize stands. Ag-gressive beetle species, however, are capable of killing trees inresistant stands following a trigger, as described above, andpopulation grows to the outbreak phase (Raffa et al., 2008). Our

    goal in this paper is to develop a qualitative understanding ofhow a population outbreak may be facilitated by a three stepprocess. First, there is successful colonization of highly stressedor compromised trees that have little resistance to bark beetles.Second, there is a build up of beetle densities as beetles exploitthese weakened trees and subsequent spread to surroundinghealthy trees. Third, these elevated populations of beetlesmoving into surrounding healthy trees exceed a threshold andthese trees therefore succumb, continuing to feed the expandingepidemic.

    Our approach starts by building a detailed mechanisticmodel for beetle behavior and reproduction and tree dynamicsin a single stand. This model is based on simple ideas fromepidemiology, extended to include nonlinear resistancethresholds and aggregation (Section 2). To analyze this model,we exploit the very different time scales for beetle behaviorand reproduction relative to tree growth. This allows us to usesingular perturbation arguments to show how beetle popula-tion dynamics exhibit properties such as bistability and hys-teresis. Analytical insight of the properties comes from a lim-iting case that relies on ideas from discontinuous dynamicalsystems. The three-step colonization process is then under-stood using a model that describes dynamics in two adjacentstands, one with higher resistance to beetles, and one withlower resistance (Section 3). Using this model, we give analy-tical conditions that can give rise to a regional outbreak in theresistant stand.

    2. One-stand models

    We begin by considering a single stand of trees with uniformresistance. We assume that the trees within this stand can beeither bark beetle free, and thus “susceptible”, (S), to infestation,or else already colonized by beetles, and thus productively “in-fected”, (I). In what follows we replace “infected” by “infested”which is a more appropriate term in this context. The movementof a tree from the susceptible class to the infested class is thenassumed to depend on a sequence of beetle-related events. First,the tree must be found by free-flying beetles, (B), that settleupon its surface, and begin to bore through the bark. Next, theseattacking beetles, (A), must effectively survive host tree de-fenses (e.g., resin) and gain access to the cambium layer. No-tably, when the number of beetles per tree is low, individualbeetles almost never surmount host defenses, and thus treesonly rarely become infested; however as the number of beetlesper tree increases, so too does the probability that host treedefenses will be overwhelmed. It is only after beetles havesuccessfully colonized a tree that we consider the tree to be inthe infested class. This leads to the following set of four coupleddifferential equations:

    σ β

    β σ

    λ μ

    λ β

    = ( ) − − ( )

    = ( ) − −

    = − − +

    = − − ( ) ( )

    dSdt

    G S I S A S S

    dIdt

    A S S I dI

    dBdt

    eI mB BS

    dAdt

    BS rA A S A

    , /

    /

    / 1

    where G is a function describing the rate of recruitment of new,susceptible trees within the stand, β is a function describingthe rate at which susceptible trees transition into the infestedclass, λ is the per beetle per tree rate at which beetles en-counter healthy trees, e is the rate at which beetles emerge

  • Table 1State variables.

    Symbol Units Dimension Definition

    S trees per hectare length�2 Density of susceptible (beetle free)trees

    I trees per hectare length�2 Density of infested (beetle in-fested) trees

    B beetles perhectare

    length�2 Density of free-flying beetles

    A beetles perhectare

    length�2 Density of attacking beetles

    R beetles per tree dimensionless Density of attacking beetles pertree

    V. Křivan et al. / Journal of Theoretical Biology 407 (2016) 25–3728

    from an infested tree, s is the natural mortality rate of healthytrees, d is the additional tree mortality that results from beetleinfestation, m is the mortality and/or emigration rate of free-flying beetles, r is the mortality rate of attacking beetles and μdescribes immigration of beetles from outside the stand. If atree becomes infested (which occurs at rate β), the number ofattacking beetles per tree (A/S) times the density of trees (S) isremoved from the beetle pool. This yields the last term in thelast equation.

    To simplify model analysis, we introduce a change of variablesby noting that model (1) can be conveniently expressed using thedensity of attacking beetles per susceptible tree, R¼A/S, ratherthan the absolute density of attacking beetles, A. When this isdone, the resulting set of ODEs becomes

    σ β

    β σ

    λ μ

    λ σ

    = ( ) − − ( )

    = ( ) − −

    = − − +

    = − ( ) + −( )

    ⎛⎝⎜

    ⎞⎠⎟

    dSdt

    G S I S R S

    dIdt

    R S I dI

    dBdt

    eI mB BS

    dRdt

    B RG S I

    Sr

    ,

    ,.

    2

    The state variables are summarized in Table 1.

    2.1. Tree recruitment, ( )G S I,

    To model tree recruitment within a conifer stand, we considerthe recruitment function ( ) = ( − − )G S I g K S I, , where K is the treecarrying capacity of the forest stand and g is a constant describingthe rate at which new, susceptible trees become available tobeetles. While it might be argued that such a recruitment model ispathological at S¼0 (as trees have a positive growth rate), it

    Table 2Parameter estimates for Eq. (2) used in this article.

    Symbol Definition Units

    g Rate of recruitment of new susceptible trees per day

    s Death rate of healthy trees per day

    d Tree death rate due to infestation per day

    e Per tree rate of beetle emergence beetles pm Death rate of free-flying beetles per dayr Death rate of attacking beetles per dayβ0 Maximum rate of infestation of new trees per dayλ Rate at which beetles find trees to attack hectaresK Tree carrying capacity trees peμ Immigration beetles pΓ Beetles per tree necessary for infestation beetles p

    should be noted that forests can, over a period of years, exhibitrecruitment in the absence of seed-producing adults as a result ofextensive seed-banks. This is true for conifer forests, since mosttree species are characterized by large, long-lived seed banks. As aresult, tree recruitment is rarely, if ever, limited by the availabilityof seed producing adults, although space (e.g., competition forlight) is still restrictive.

    2.2. Infestation rate, β ( )R

    In keeping with threshold-based mortality models, we assumethat the rate at which susceptible trees transition into the infestedclass, β ( )R , exhibits a nonlinear dependence on the number ofattacking beetles per susceptible tree (R). This nonlinearity is oneof the defining features of bark beetle dynamics, and arises fromthe fact that most host trees have natural defenses (e.g., resin) thatprotect against beetle infestation at low beetle densities, but be-come rapidly overwhelmed at high beetle densities. Accordingly,when beetles are scarce, tree infestation rates are depressed re-lative to what would be expected on the basis of mass action as-sumptions. To capture this depression mechanistically we assumea threshold number of attacking beetles per tree (typically de-pendent on tree resistance), θ, above which infestation succeedsand below which, infestation fails.

    We model infestation rate by the Hill function

    β βΓ

    βΓ

    ( ) =+

    =+ ( )−

    RR

    R R1 3

    n

    n n n n00

    where Γ roughly approximates tree resistance, or the thresholdnumber of beetles required for successful infestation and n is re-lated to the level of beetle aggregation. In particular, low values ofn represent high levels of aggregation, while high values of n in-dicate overdispersion (see Appendix C). To see this, consider thelimit → ∞n , wherein β ( )R defined by (3) becomes a step function.In this limit, an infinitely small increase in beetle density at Γ=Rleads to a sudden transition from a per tree infestation rate of zeroto a per tree infestation rate that is maximal for the system. Theabruptness of this transition implies that the addition of an ex-ceedingly small number of new beetles causes every tree to crossthe critical infestation threshold simultaneously, which will onlyhappen if beetles are uniform in their distribution over availabletrees (i.e., in the overdispersion limit).

    2.3. Model parameters

    Model parameters used in this article are summarized inTable 2. For tree population dynamics, we interpret g as reflectingthe rate at which new susceptible adult trees become available tobeetles per existing tree at carrying capacity. This parameter is

    Dimension Approximate values

    time�1 –− −10 104 3

    time�1 ( – ) × −0.5 5 10 5

    time�1 × −3 10 3

    er tree per day time�1 10–100time�1 0.05time�1 0.1time�1 0.003–0.07

    per tree per day length2 time�1 0.001r hectare length�2 100–10,000er hectare per day length�2 time�1 0–4000er tree dimensionless 30–3000

  • Table 3Non-dimensionalization scheme.

    Symbol Approximate value Symbol Approximate value

    λ˜ =S S m/ State variable μ λμ˜ = ( )em/ μ( – )0.0002 0.002

    λ˜ =I I m/ State variable ˜ =g g d/ 0.03–0.3˜ =R rR e/ State variable β β˜ = d/0 0 1–23

    λ˜ =B B e/ State variable σϵ = d/1 0.002–0.016˜ =t dt t0.003 ϵ = d m/2 0.06

    λ˜ =K K m/ 15–30 ϵ = d r/3 0.03

    Γ Γ˜ = r e/ 0.57–15

    V. Křivan et al. / Journal of Theoretical Biology 407 (2016) 25–37 29

    estimated to be approximately 0.05–0.5 years�1 for pine trees(Clark et al., 2001). Therefore we set = –− − −g 10 10 day4 3 1. Similarly,because tree species targeted by tree-killing bark beetles can havelifespans between 50 and 500 or more years, depending on thegeographic region and beetle species, we setσ = × – ×− − −5 10 5 10 day6 5 1. We assume that beetle infested trees,on the other hand, will produce beetles for approximately 1 year.Thus we set the rate of tree death due to beetle infestation at

    = × − −d 3 10 day3 1. The length of time before the susceptible treetransfers to the infested class is estimated to range between twoweeks and one year and hence β0 ranges from approximately0.003–0.07 day�1. The threshold for succumbing to attack forhealthy trees is approximately 10–100 beetles per m2 of bark area(Lewis et al., 2010). If a tree were between 10 and 20 m tall andhad an average diameter between 0.1 and 0.5 m, then its surfacearea would range between π and πm10 2. This would mean that thethreshold for succumbing to beetle attack would range between

    π10 and π1000 , i.e., 30–3000 beetles per tree. We assume that treecarrying capacity K is between 100 and 10,000 trees/ha for un-managed forests (Baker, 2009).

    2.4. Non-dimensionalization

    We can reduce the number of free parameters through non-dimensionalization. Using the non-dimensionalization schemesoutlined in Table 3 gives

    βΓ

    βΓ

    μ

    ˜˜ =

    ˜( ˜ ˜) − ϵ ˜ −˜ ˜ ˜

    ˜ + ˜˜˜ =

    ˜ ˜ ˜˜ + ˜

    − ϵ ˜ − ˜

    ϵ˜˜ =

    ˜ − ˜ − ˜ ˜ + ˜

    ϵ˜˜ =

    ˜ − ϵ˜ ˜( ˜ ˜)

    ˜ −˜ + ϵ ϵ ˜

    ( )

    dSdt

    G S I SSR

    R

    dIdt

    SR

    RI I

    dBdt

    I B BS

    dRdt

    BRG S I

    SR R

    ,

    ,4

    n

    n n

    n

    n n

    10

    01

    2

    33

    1 3

    where ˜ ( ˜ ˜)G S I, is the non-dimensionalized recruitment function˜ ( ˜ ˜) = ˜( ˜ − ˜ − ˜)G S I g K S I, . Table 3 gives values for dimensionlessparameters that correspond to those from Table 2.

    2.5. Pseudo-steady state approximation

    In general, the dynamics associated with beetle processes, in-cluding beetle mortality and tree death due to infestation, aresignificantly faster than natural tree dynamics. As a result, forrealistic parameter values (see, for example, Table 3), it will alwaysbe true that < ϵ ϵ ϵ ⪡0 , , 11 2 3 . This allows us to make a pseudo-steadystate approximation on (4). Specifically, taking the limit asϵ ϵ ϵ →, , 01 2 3 we find (in what follows we drop tildes for notationalsimplicity)

    μ= ++

    = ( )

    BI

    S

    R B

    15

    and

    β

    β

    = ( ) − ( )

    = ( ) − ( )

    dSdt

    G S I S I S

    dIdt

    S I S I

    , ,

    , 6

    where

    β β μμ Γ

    ( ) = ( + )( + ) + ( + ) ( )

    S II

    I S,

    1.

    7

    n

    n n n0

    2.6. Scenario one: uniform beetle distribution

    We begin our analysis by studying the behavior of themodel in the limit that beetles distribute uniformly overavailable trees (i.e., all trees are equally susceptible and there isno aggregating pheromone). To do this, we take → ∞n in (7), inwhich case, the infestation rate, β ( )S I, , becomes a step func-tion. Specifically,

    β

    μ Γ

    β μ Γ( ) =

    ++

    <

    ++

    >( )

    ⎧⎨⎪⎪

    ⎩⎪⎪

    S I

    IS

    IS

    ,0 if

    1

    if1

    .80

    From (8) we see that the minimum number of infested treesnecessary for beetle spread, Imin, can be expressed in terms ofthe number of susceptible trees, S, according to the expression

    Γ μ( ) = ( + ) − ( )I S S1 . 9min

    This threshold is shown by the solid line in Fig. 1C, E. It reflectsthe fact that, if beetles distribute uniformly over availabletrees, then when there are more susceptible trees, proportio-nately more beetles are needed to overcome the thresholdrequirement for infestation. Importantly, when the totalnumber of infested trees falls below the critical threshold forinfestation,

  • Fig. 1. These plots show trajectories of model (6) for uniform beetle distribution (β is given by (8)) (left panels) and aggregated beetle distribution (β is given by (7) withn¼10) (right panels). Panels A and B assume low tree resistance (Γ¼30), panels C and D assume intermediate resistance (Γ¼300), and panels E and F assume high resistance(Γ¼450). The solid line in panels C and E is the threshold Imin for infestation given by (9) above which beetle spread in the forest. The dotted line is the isocline for susceptibletrees and the dashed line is the isocline for infested trees. The black dot denotes a locally stable equilibrium, while the gray dot denotes an unstable equilibrium. Otheruntransformed parameters are g¼0.001, d¼0.003, m¼0.05, β = 0.010 , μ = 2000, K¼100. For simulations these parameters were non-dimensionalized following scheme inTable 3.

    V. Křivan et al. / Journal of Theoretical Biology 407 (2016) 25–3730

  • Table 4Locally stable equilibria for uniformly dispersing beetles and unlimited treerecruitment.

    Name Equilibrium Resistance Figure

    Endemic

    β β+ +β

    β β+ +

    ⎛⎝⎜

    ⎞⎠⎟

    gKg g

    ,gK

    g g0 0

    00 0

    Γ μ<+ K1

    1A

    Beetle-free/endemic

    ( )K , 0 ,

    β ββ

    β β+ + + +

    ⎛⎝⎜

    ⎞⎠⎟

    gKg g

    gKg g

    ,0 0

    0

    0 0

    μ Γ Γ+

    < < *K1

    a 1C

    Beetle-free ( )K , 0 Γ Γ* < 1E

    a Γ* is given by (13).

    Fig. 2. Dependence of the equilibrium infestation *I on the transformed standresistance. This figure documents hysteresis in the forest dynamics. Untransformedparameters: g¼0.001, d¼0.003, m¼0.05, β = 0.010 , μ = 2000, K¼1000. For simu-lations these parameters were non-dimensionalized following scheme in Table 3.

    V. Křivan et al. / Journal of Theoretical Biology 407 (2016) 25–37 31

    For ( ) = ( − − )G S I g K S I, , (11) has a single endemic equilibrium at

    β ββ

    β β( * *) =

    + + + + ( )

    ⎛⎝⎜

    ⎞⎠⎟S I

    gKg g

    gKg g

    , , .120 0

    0

    0 0

    Provided tree resistance is not too high and satisfies

    Γ Γβ μ β ββ β

    < * =+ ( + + )+ + + ( )

    gK g gg g gK

    ,13

    0 0 0

    0 0

    this equilibrium is in the part of the phase space where > ( )I I Smin ,and it is locally asymptotically stable there (see Appendix A).

    For stressed stands that are subject to a relatively large andconstant influx of beetles, i.e., Γ μ< ( + )K/ 1 , the endemicequilibrium ( * *)S I, (Fig. 1A) is the only locally asymptoticallystable equilibrium. However, when tree resistance is inter-mediate and satisfies μ Γ Γ( + ) < < *K/ 1 (we remark that for allparameter values μ Γ( + ) < *K/ 1 ) there are two locallyasymptotically stable equilibria: the beetle free equilibrium( )K , 0 and the endemic equilibrium ( * *)S I, (Fig. 1C). Conse-quently, whether the forest survives or not will depend on itshistory. Specifically, a fully grown forest with tree densitiesnearing the forest carrying capacity will be able to resistbeetle invasion, whereas a more sparsely populated forestwith tree densities well below carrying capacity will not.Notice that this is somewhat counterintuitive, since denseforests provide ample trees for beetles to attack. This, how-ever, is the problem. For uniformly distributing beetles, largenumbers of trees dilute the beetle population such that notree has sufficient beetle loads to become infested. In lessdense stands, the dilution effect is not so strong, and there areenough beetles per tree to mount successful attacks. In thissystem, a large perturbation to the beetle free equilibrium, forexample a significant but temporary influx of beetles, canmove the system across the line = ( )I I Smin (solid line inFig. 1C) that divides the two stable states. Ultimately, thismeans that a one-time influx of beetles can potentially causethe forest to evolve toward the beetle endemic state. Forsmaller beetle influxes, however, the system will not cross theseparatrix, thus once the influx has ceased, the system willreturn to its initial, beetle-free state. When the tree resistanceΓ exceeds Γ* the beetle free equilibrium ( )K , 0 is the onlyasymptotically stable equilibrium and the beetle-free standwill be immune to beetle invasions (Fig. 1E). These results aresummarized in Table 4. Importantly, intermediate stand re-sistance that results in bistability leads to a hysteresis loop(Fig. 2). When model parameters change slowly, which of thetwo locally stable equilibria the system finds itself in maydepend upon the path taken. For example, Fig. 2 considersdependence of the equilibrium infestation on the stand re-sistance. Let us assume that the stand resistance is high. Thenthe only equilibrium is the beetle-free forest. As the resistance

    decreases, the situation will continue to be the same until thelower critical threshold μ ( + )K/ 1 is reached. If the resistancekeeps decreasing, there is a sudden jump in the number ofinfested trees because for low resistance the endemic equili-brium is the only possible state of the forest. Now, let us as-sume that the resistance starts to increase. The forest will stayin the endemic state until resistance reaches the upperthreshold given by Γ*. For yet higher resistance the beetle-free forest is the only equilibrium.

    Because model (6) with uniform beetle distribution modeledby (8) is a differential equation with a discontinuous right-handside, solutions are defined in the Filippov sense (Filippov, 1988;Colombo and Křivan, 1993). To ensure existence of solutions, wemust analyze what happens along the switching line (9).Appendix B shows that there are two possibilities only. Eithertrajectories cross the switching line transversally, or trajectoriesmove away from the switching line in both directions (suchpoints are shown e.g., in Fig. 1C). In this latter case trajectories ofthe model are not uniquely defined. Thus, the so-called slidingregime does not occur and there are no additional equilibriaalong the switching line.

    2.7. Scenario two: aggregated beetle distribution

    To model a nonuniform distribution of beetles over availabletrees, we take n finite and not too large in (7). Most notably, thebasis of attraction for the endemic state at intermediate tree re-sistance shifts to the right (compare Fig. 1C with Fig. 1D). Thesuggestion is that beetles can attack and kill trees in forest standswith higher tree densities when they exhibit aggregative beha-vior. This, of course, makes intuitive sense. Aggregation coun-teracts beetle dilution across higher density stands. As a result,the beetle per tree threshold required for infestation is morelikely to be met by aggregating beetles, even in stands with largenumbers of trees.

    3. Two-stand model

    The goal in this section is to derive and analyze a model thatgives qualitative understanding of how a regional beetle outbreakmay be facilitated by a three step process: (i) infestation of highlystressed or compromised trees, who have little resistance to thebeetles; (ii) build up of beetle density in these trees and sub-sequent spread to surrounding healthy trees; (iii) increase in

  • V. Křivan et al. / Journal of Theoretical Biology 407 (2016) 25–3732

    beetle levels in surrounding healthy trees exceeding a thresholdand these trees succumbing to become part of the spreadingepidemic.

    We consider two forest stands coupled by beetle dispersal.Because we are interested in the role of beetle spillover betweenstands, we consider forest stands that only differ in terms of re-sistance, Γ, and beetle influx from other, more distant sources.Thus model (2) can be extended as follows:

    σ β

    β σ

    λ δ μ

    λ σ

    σ β

    β σ

    λ δ μ

    λ σ

    = ( ) − − ( )

    = ( ) − −

    = − − + ( − ) +

    = − ( ) + −

    = ( ) − − ( )

    = ( ) − −

    = − − + ( − ) +

    = − ( ) + −( )

    ⎛⎝⎜

    ⎞⎠⎟

    ⎛⎝⎜

    ⎞⎠⎟

    dSdt

    G S I S R S

    dIdt

    R S I dI

    dBdt

    eI mB B S B B

    dRdt

    B RG S I

    Sr

    dSdt

    G S I S R S

    dIdt

    R S I dI

    dBdt

    eI mB B S B B

    dRdt

    B RG S I

    Sr

    ,

    ,

    ,

    ,14

    11 1 1 1 1 1 1

    11 1 1 1 1

    11 1 1 1 2 1 1

    11 1

    1 1

    1

    22 2 2 2 2 2 2

    22 2 2 2 2

    22 2 2 2 1 2 2

    22 2

    2 2

    2

    where β β( ) =Γ +

    Ri iR

    R0in

    in

    in , ( ) = ( − − )G S I g K S I,i i i i i i i and we have as-

    sumed that all beetles dispersing from the first stand arrive at thesecond and vice versa (i.e., they are neither going to nor comingfrom additional stands) with dispersal rate δ > 0. In addition, therecan be stand specific immigration from outside of the two stands(μi). Notice that we have not assumed any seed rain between thestands, thus we are considering stands that are geographicallydistant enough that seed transfer is negligible, however not sodistant as to prevent beetles migrating from one stand to theother. Using a direct extension of the non-dimensionalizationscheme in Table 3, Eq. (14) can be rewritten as

    βΓ

    βΓ

    δ μ

    βΓ

    βΓ

    δ μ

    ˜˜ =

    ˜ ( ˜ ˜ ) − ϵ ˜ −˜ ˜ ˜

    ˜ + ˜

    ˜˜ =

    ˜ ˜ ˜˜ + ˜

    − ϵ ˜ − ˜

    ϵ˜˜ =

    ˜ − ˜ − ˜ ˜ + ˜( ˜ − ˜ ) + ˜

    ϵ˜˜ =

    ˜ − ϵ ˜˜ ( ˜ ˜ )

    ˜ −˜ + ϵ ϵ ˜

    ˜˜ =

    ˜ ( ˜ ˜ ) − ϵ ˜ −˜ ˜ ˜

    ˜ + ˜

    ˜˜ =

    ˜ ˜ ˜˜ + ˜

    − ϵ ˜ − ˜

    ϵ˜˜ =

    ˜ − ˜ − ˜ ˜ + ˜( ˜ − ˜ ) + ˜

    ϵ˜˜ =

    ˜ − ϵ ˜˜ ( ˜ ˜ )

    ˜ −˜ + ϵ ϵ ˜

    ( )

    dSdt

    G S I SS R

    R

    dIdt

    S R

    RI I

    dBdt

    I B B S B B

    dRdt

    B RG S I

    SR R

    dSdt

    G S I SS R

    R

    dIdt

    S R

    RI I

    dBdt

    I B B S B B

    dRdt

    B RG S I

    SR R

    ,

    ,

    ,

    ,15

    n

    n n

    n

    n n

    n

    n n

    n

    n n

    11 1 1 1 1

    0 1 1

    1 1

    1 0 1 1

    1 11 1 1

    21

    1 1 1 1 2 1 1

    31

    1 3 11 1 1

    11 1 3 1

    22 2 2 1 2

    0 2 2

    2 2

    2 0 2 2

    2 21 2 2

    22

    2 2 2 2 1 2 2

    32

    2 3 22 2 2

    22 1 3 2

    where δ δ˜ = m/ .Again, taking the limit as ϵ ϵ ϵ →, , 01 2 3 we find the following

    model under the pseudo-steady state approximation:

    δ δ δ μ δμδ δ

    δ δ δ μ δμδ δ

    β

    β

    β

    β

    =+ ( + + ) + ( + + ) +

    + + ( + ) + ( + + )

    =

    =+ ( + + ) + ( + + ) +

    + + ( + ) + ( + + )

    =

    = ( ) − ( )

    = ( ) −

    = ( ) − ( )

    = ( ) − ( )

    BI I S S

    S S S S

    R B

    BI I S S

    S S S S

    R BdSdt

    G S I S I S I S

    dIdt

    S I S I S I

    dSdt

    G S I S I S I S

    dIdt

    S I S I S I

    1 11 2 1

    1 11 2 1

    , , , ,

    , , ,

    , , , ,

    , , , 16

    12 1 2 2 1 2

    2 2 1 2

    1 1

    21 2 1 1 2 1

    2 2 1 2

    2 2

    11 1 1 1 1 1 2 2 1

    11 1 1 2 2 1 1

    22 2 2 2 2 2 1 1 2

    22 2 2 1 1 2 2

    where

    ( )

    ββ

    Γδ δ δ μ δμ

    δ δ

    ββ

    Γδ δ δ μ δμ

    δ δ

    ( ) =+

    + ( + + ) + ( + + ) ++ + ( + ) + ( + + )

    ( ) =+

    + ( + + ) + ( + + ) ++ + ( + ) + ( + + )

    ⎛⎝⎜

    ⎞⎠⎟

    ⎛⎝⎜

    ⎞⎠⎟ 17

    S I S II I S S

    S S S S

    S I S II I S S

    S S S S

    , , ,

    11 1

    1 2 1

    , , ,

    11 1

    1 2 1

    nn

    nn

    1 1 1 2 20

    12 1 2 2 1 2

    2 2 1 2

    2 1 1 2 20

    21 2 1 1 2 1

    2 2 1 2

    and tildes have been dropped for notational simplicity.We analyze the two-stand model by again studying model

    behavior in the limit that beetles distribute uniformly over avail-able trees ( → ∞n in (17)). As before, this leads to step functioninfestation rates, β ( )S I S I, , ,i i i j j , with

    ( )

    β

    δ δ δ μ δμδ δ

    Γ

    βδ δ δ μ δμ

    δ δΓ

    ( ) =

    + ( + + ) + ( + + ) ++ + ( + ) + ( + + )

    <

    + ( + + ) + ( + + ) ++ + ( + ) + ( + + )

    >

    ⎧⎨⎪⎪

    ⎩⎪⎪ 18

    S I S I

    I I S SS S S S

    I I S SS S S S

    , , ,0 if

    1 11 2 1

    if1 1

    1 2 1

    1 1 1 2 2

    2 1 2 2 1 2

    2 2 1 21

    02 1 2 2 1 2

    2 2 1 21

    and similarly for β2. From (18), the minimum number of infestedtrees necessary for beetle spread in stand 1, Imin,1, is calculatedfrom equation

    δ δ δ μ δμδ δ

    Γ+ ( + + ) + ( + + ) +

    + + ( + ) + ( + + )=

    I I S SS S S S1 1

    1 2 12 1 2 2 1 2

    2 2 1 21

    which yields

    Γ Γ δ μ δ μ μδ

    ( ) =( + )( + ) + ( + + ) − ( + ) − ( + + )

    + +I S S I

    S S S S S I

    S, ,

    1 1 2 11

    .min,1 1 2 21 2 1 1 2 1 2 1 2 1 2

    2

    Similar calculations for stand 2 give the critical threshold

    Γ Γ δ μ δ μ μδ

    ( ) =( + )( + ) + ( + + ) − ( + ) − ( + + )

    + +I S S I

    S S S S S I

    S, ,

    1 1 2 11

    .min,2 1 2 11 2 2 1 2 2 1 2 1 1 2

    1

    We observe that, due to dispersal, the minimum threshold forinfestation to spread in one stand depends on the state of theother stand, i.e., Imin,1 depends on S2 and I2 and, similarly, Imin,2depends on S1 and I1.

    To interpret stand dynamics we consider three possibilities:(a) beetle establishment does not occur in either stand ( I Imin2 ,2).

    In the first case, when the beetle population does not reachthreshold densities in either stand (

  • V. Křivan et al. / Journal of Theoretical Biology 407 (2016) 25–37 33

    = ( )

    = −

    = ( )

    = − ( )

    dSdt

    G S I

    dIdt

    I

    dSdt

    G S I

    dIdt

    I

    ,

    ,

    . 19

    11 1 1

    11

    22 2 2

    22

    The only stable equilibrium of (19) is the beetle-free equilibrium( * * * *) = ( )S I S I K K, , , , 0, , 01 1 2 2 1 2 . This will be a solution to the full system(16) (i.e., belongs to the part of the beetle-free–infested tree phase spacewhere ( ) > = *I K K I, , 0 0min i i, 1 2 , i¼1, 2) provided tree resistance in bothstands is high enough such that Γ Γ> a1 1 and Γ Γ> a2 2 (for definition ofthese and other thresholds below see the footnote of Table 5).

    We note that, without any immigration from outside (i.e., whenμ = 0i , i¼1, 2), the beetle-free state will always exist (as we as-sume that tree resistance is positive, i.e., Γ > 0i , i¼1, 2). Sufficientoutside immigration to either stand may cause the beetle-freestate to disappear in one or both of the stands.

    When only the first stand crosses the threshold for infestation( >I Imin1 ,1, ( * * *)I I S S I, ,min1 ,1 1 2 2 and* = < ( * * *)I I S S I0 , ,min2 ,2 1 2 1 ) provided Γ Γ< b1 1 and Γ Γ> .b2 2 In otherwords, (21) is an equilibrium provided tree resistance in stand 1 islow while tree resistance in stand 2 is high.

    Finally, in the case that both stands cross the threshold for

    Table 5Two-stand results for uniformly dispersing beetles.

    Name Equilibrium

    Beetle-free in both stands ( )K K, 0, , 01 2Stand 1 endemic, stand 2 beetle-free

    β β+ ( + )β

    β + (

    ⎛⎝⎜

    g Kg 1

    ,g K

    g1 1

    0 1 0

    0 10 1 1

    Stand 1 beetle-free, stand 2 endemic

    β β β+ ( + )

    ⎛⎝⎜ K

    g Kg

    , 0,1

    ,1 22

    0 2 0

    Two-stand endemic

    β ββ

    β+ ( + ) +

    ⎛⎝⎜

    g Kg g1

    ,1 1

    0 1 0

    0

    0 1

    a Γ = μ δ μ μδ δ

    ( + ) + ( + )+ +( + ) + ( + + )a

    KK K K K1

    1 2 1 1 21 2 2 2 1 1 2

    , Γ = μ δ μ μδ δ

    ( + ) + ( + )+ +( + ) + ( + + )a

    KK K K K2

    1 1 2 1 21 2 2 2 1 1 2

    .

    b Γ = β δ β β δ μ δμβ β β δ

    ( + + ) +( + + )(( + + ) + )( + )( + ( + + )) +( ( + + ) +( + )( + ) )b

    g K K g g KK g K g K K g K1

    1 1 0 1 2 1 0 1 0 1 2 1 21 2 0 1 1 1 0 1 2 1 2 1 1 2 2 0

    , Γ = β δ β β δμβ β+( + + )

    ( + )( + ( + + )bg K g g

    K g K21 1 0 1 0 1 0 11 2 0 1 1 1 0

    c Γ =δ

    β ββ

    β ββ δ β β δμ

    β β δ β δ β δμ

    δδ

    β ββ β δ β δ β δ

    β β β β

    + ++ + + +

    ++( + + )

    + + ( + + + + )+

    + +( + )

    + ++

    ( + + ( + + + + ))( + + )( + + ))

    ⎛⎝⎜⎜

    ⎞⎠⎟⎟

    ⎛⎝⎜⎜

    ⎞⎠⎟⎟

    c

    g Kg g

    g Kg g

    g K g gg K

    g Kg g

    g K g Kg g g g

    1

    1 2 22 0 2 0

    1 1 01 0 1 0

    2 2 0 2 0 2 0 20 0 2 1 2 0 0

    1

    1 2 2 21

    2 0 2 01 1 0 0 2 1 2 0 0

    1 0 1 0 2 0 2 0

    , Γ =δ

    β

    δ

    + ++

    + +

    ⎛⎝⎜⎜

    c

    gg

    2

    1 11 0

    1 2

    establishment ( >I Imin1 ,1, >I Imin2 ,2), model (16) becomes

    β

    β

    β

    β

    = ( ) −

    = −

    = ( ) −

    = − ( )

    dSdt

    G S I S

    dIdt

    S I

    dSdt

    G S I S

    dIdt

    S I

    ,

    ,

    . 22

    11 1 1 0 1

    10 1 1

    22 2 2 0 2

    20 2 2

    Again, the equilibria for (22) as well as their stability can be de-termined directly from results for the one-stand model. The en-demic equilibrium in both stands

    ( )β ββ

    β β β ββ

    β β

    ( * * * *)

    =+ ( + ) + ( + ) + ( + ) + ( + )

    ⎛⎝⎜

    ⎞⎠⎟ 23

    S I S I

    g K

    g

    g K

    g

    g K

    g

    g K

    g

    , , ,

    1,

    1,

    1,

    1.

    1 1 2 2

    1 1

    0 1 0

    0 1 1

    0 1 0

    2 2

    0 2 0

    0 2 2

    0 2 0

    will be a solution to the full system (16) (i.e., belongs to the part ofthe healthy–infested tree phase space where * > ( * * *)I I S S I, ,min1 ,1 1 2 2and * > ( * * *)I I S S I, ,min2 ,2 1 2 1 ) provided Γ Γ< c1 1 and Γ Γ< c2 2. In otherwords, (23) is an equilibrium provided tree resistance in bothstands is low. Table 5 summarizes these results.

    The effect of beetle dispersal between patches is shown inFig. 3. Here we focus on the following scenario: stand 1 has a lowerresistance when compared to stand 2, and there is external im-migration of beetles from outside of the system to stand 1 only.Thus, stand 2 can become infested only as a result of beetle dis-persal from stand 1, i.e., stand 1 serves as a springboard to infestpatch 2. Parameters are such that when immigration to stand 1 islow both stands are in a beetle-free state because resistance issufficiently high in stand 1 to prevent invasion of beetles. Thus,when immigration is low, we observe a stable equilibrium( )K K, 0, , 01 2 (Fig. 3A, B). As immigration to stand 1 increases, stand1 shifts to the endemic equilibrium while stand 2 stays beetle-free(Fig. 3C, D). For yet higher immigration rates to stand 1 bothstands shift to the endemic equilibrium (Fig. 3E, F).

    4. Discussion

    4.1. Summary

    This paper focuses on the formulation and analysis of a generalmodel for bark beetle outbreaks in continuous time. By capitaliz-ing on the fact that there are multiple time scales involved in thesystem, we are able to derive a simplified dynamical system that

    Tree resistance

    Γ Γ> a1 1a, Γ Γ> a2 2a

    β+ )

    ⎞⎠⎟K, , 0

    10

    2Γ Γ< b1 1b, Γ Γ> b2 2b

    ββ+ ( + )

    ⎞⎠⎟

    g Kg 10 2 2

    0 2 0

    Γ Γ> b1 1, Γ Γ< b2 2

    β β β( + ) + ( + )β

    β β+ ( + )

    ⎞⎠⎟

    g K g Kg1

    ,1

    ,g K

    g1 1

    0

    2 2

    0 2 0

    0 2 20 2 1 0

    Γ Γ< c1 1c, Γ Γ< c2 2c

    β β δ β δ β δ μβ δ

    +( + + ( + + + + ))) +( ( + + ) +( + )( + ) )

    g Kg K K g K

    0 0 1 1 1 0 0 21 2 1 2 1 1 2 2 0

    .

    ββ

    β ββ δ β β δμ

    β β δ β δ β δμ

    δβ β

    β β δ β δ β δβ β β β

    + + ++

    +( + + )+ + ( + + + + )

    +

    ( + )+ +

    +( + + ( + + + + ))

    ( + + )( + + ))

    ⎞⎠⎟⎟

    ⎛⎝⎜⎜

    ⎞⎠⎟⎟

    Kg

    g Kg g

    g K g gg K

    g Kg g

    g K g Kg g g g

    11 0

    2 2 02 0 2 0

    1 1 0 1 0 1 0 10 0 1 1 1 0 0

    2

    2 2 1

    2 0 2 01 1 0 0 2 1 2 0 0

    1 0 1 0 2 0 2 0

    .

  • Fig. 3. The springboard effect of stand 1 on the beetle outbreak in stand 2. These plots assume that beetles disperse between two forest stands and there is an allochthonousbeetle inflow to stand 1 (but not to stand 2). Stand 1 has a lower resistance (Γ = 301 ) when compared to stand 2 (Γ = 2002 ). When immigration of beetles to stand 1 fromoutside of the system is relatively small (panels A, B; μ = 1001 , μ = 02 ), both stands stay at the beetle-free state. For intermediate immigration rates to stand 1 (panels C, D;μ = 10001 , μ = 02 ), stand 1 shifts to the endemic equilibrium (12) while stand 2 stays at the beetle-free state. For high immigration rates to stand 1 (panels E, F; μ = 40001 ,μ = 02 ), both stands shift to the endemic equilibrium. Besides the above differences, both stands are assumed to be identical. The curve is a trajectory of model (16) whenbeetles distribute uniformly over available trees. The black dot denotes a locally stable equilibrium, while the gray dot denotes an unstable equilibrium. Untransformedparameters used for simulations: β = 0.010 λ λ= = 0.001,1 2 = =g g 0.001,1 2 = =r r 0.11 2 , = =m m 0.05,1 2 = =d d 0.003,1 2 = =K K 100,1 2 δ = 10, e¼10. For simulations theseparameters were non-dimensionalized following scheme in Table 3.

    V. Křivan et al. / Journal of Theoretical Biology 407 (2016) 25–3734

    describes the bark beetle population dynamics over long timescales. Further simplifications using ideas from piecewise dy-namics and Filippov dynamical systems (Filippov, 1988; Colomboand Křivan, 1993) allow us to mathematically deduce several keyproperties of the dynamical model. These include

    � Bistability of the forest dynamics arising from a threshold effectwith respect to beetle numbers. Here the beetle numbers mustexceed a critical value determined by tree resistance to infesthealthy trees. Although such threshold effects have been in-cluded in previous beetle models, ours is a mechanisticallyderived threshold, based on tree resistance. Most clearly this isseen in the case of the uniform beetle distribution where thethreshold for the invasion splits the forest phase space into twoparts, each with its own population dynamics (see the two re-gions separated by the solid line in Fig. 1C). In one region thebeetle-free forest is a locally stable equilibrium. In the other

    part of the phase space an endemic equilibrium is a locallystable equilibrium. The position of these equilibria with respectto the threshold value depends on parameters. However, forparameters that allow coexistence of the beetle-free forestequilibrium and the endemic equilibrium, we get bistability.Depending on the history, the forest can respond to a beetleimmigration event either by returning to the beetle-free state,or to move to the endemic state. Bistability carries over also tothe case where dispersing beetles show aggregative distribu-tion, modeled by a more gradual Hill function (cf. Fig. 1C and D).

    � Hysteresis: The model bistability naturally leads to hysteresiseffects. These are most easily understood in terms of changes inthe stand resistance as illustrated in Fig. 2. The lower thresholdvalue for resistance in the hysteresis loop is μ ( + )K/ 1 and thehigher value is Γ* (see Eq. (13)). These quantities can be directlyinterpreted in terms of the biological parameters describing theinteraction between trees and beetles (Table 2).

  • V. Křivan et al. / Journal of Theoretical Biology 407 (2016) 25–37 35

    � Interactions between multiple patches: Here multiple patchesthat are spatially linked can interact to produce new outcomes.For example, a less healthy patch of trees may provide abeachhead for the infestation process. Once established, thebeetles can then build up in numbers before progressing toneighboring healthy patches of trees, patches that wouldotherwise be unassailable, and causing them to succumb. Thesekinds of complex outcomes are illustrated in Fig. 3.

    � Our model with unlimited tree recruitment rate would be in-appropriate for forests that experience complete loss of adulttrees over periods longer than the viability of the seed-bank. Forthis reason we also analyzed the model with the logistic growth(results not showed here). On the contrary to the unlimitedrecruitment where the forest cannot completely die, the logistictree recruitment rate has also an extinction equilibrium. Inparticular, when a forest stand shows a low regeneration rate, asgiven by the ratio between the rate at which trees becomeavailable to beetles relative to the rate at which beetles removethe trees, the stand can go extinct.

    4.2. Model limitations

    In model (1) we have made several simplifying assumptionsthat, though reasonable in many outbreak contexts, will not holdunder all scenarios. First, we have taken the rate at which beetlesencounter trees, λ, as constant, implying that contact rates be-tween beetles and trees follow a simple mass action law. In reality,however, encounter rates likely exhibit some dependence on bothbeetle and tree density as well as beetle characteristics, includingspecies-specific search strategies and aggregation behaviors(Mitchell and Preisler, 1991; Safranyik et al., 2010; Powell andBentz, 2014). Second, we have assumed that the total number ofbeetles emerging from a tree is independent of the total number ofbeetles that infested the tree in the first place. More accurately, therate of emergence should be lower when the number of attackingbeetles is far from the carrying capacity of the tree (Light et al.,1983; Anderbrant et al., 1985). Third, we have assumed that therate at which beetles are killed by host tree defenses, r, is in-dependent of the number of attacking beetles per tree. Realisti-cally, however, the death rate of beetles on trees nearing thethreshold for infestation is probably lower than it is on trees withone or a few beetles (Raffa and Berryman, 1983). Furthermore, thethreshold, itself, is assumed to be a fixed number, describing theexact number of beetles per tree needed to mount a successfulattack. In reality, natural variation between trees would round offthis sharp threshold to something more gradual. Fourth, the ne-gative binomial model for beetle attacks necessarily oversimplifiesthe aggregation process. We are aware that other researchers havedeveloped spatially explicit model with a focus on determiningspecific attack locations (see, for example, Logan et al., 1998).However, we keep our model spatially implicit by using the ne-gative binomial probability mass function to provide a phenom-enological description. This approach has been used before as abaseline probability mass function for the attack density inmountain pine beetle (Chubaty et al., 2009). For generality, wehave chosen to model the dynamics in continuous time, althoughit may be that discrete-time models provide a more accurate de-scription of dynamics, particularly in scenarios where generationsare strongly non-overlapping (e.g., species or regions where bee-tles are univoltine). Additionally, factors such as environmentalstochasticity and interactions with tree signaling chemicals willplay a role in the outbreak dynamics.

    4.3. Model extensions

    Our modeling approach assumed that each stand comprised of

    a cohort of identical trees. Thus any variation between trees wasrelegated to the variation found between different stands situatedat different locales. In fact, stands are typically composed of sev-eral groups of trees, each group with different resistance to in-festation and different rate of beetle production. It would bepossible to extend the model to include such cases. This wouldallow us to evaluate the effect of stand structure on beetle out-break. Some initial attempts in this direction can be found in Lewiset al. (2010), Powell and Bentz (2014) and Duncan et al. (2015).Indeed, it is well known that factors influencing bark beetle in-festation are related to stand age and stage. When the character-istics of each group within a stand are determined by age or stage,it is necessary to include stage structure in the underlying dyna-mical model for the tree population (Koch et al., personal com-munication). Our analysis that focused on the simplified systemwhere beetles distribute uniformly leads to general insight thatalso may apply to the more complex system with clumped beetlepopulations. For clumped beetle populations the piece-wise linearanalysis applied here is not possible and numerical simulationswill be necessary to falsify our predictions.

    4.4. Concluding remarks

    In summary, our paper has focused on model development andanalysis for the dynamics of bark beetle infestation of trees, wheretree resistance and beetle aggregation have key roles to play in theinfestation outcomes. By carefully formulating a detailed model,and then using perturbation theory to distinguish between thedifferent time scales involved in the infestation process, we areable to derive a remarkably simple system of nonlinear ordinarydifferential equations for outbreak dynamics. These are furthersimplified in the limit associated with uniform dispersal of beetles,which gives rise to a Filippov-type dynamical system. Resultingbistable dynamics lead to hysteresis, and the multiple patch dy-namics lead to the possibility of less resistant tree populationsproviding a toe-hold for beetles, from which they build up andeventually outbreak, causing the healthier patch to succumb. Byestimating model parameters, based on beetle and tree biology,we are able to show that such behaviors fall within the range ofreasonable parameter values.

    Acknowledgments

    This work was conducted as a part of the Forest Insect PestsWorking Group at the National Institute for Mathematical andBiological Synthesis, sponsored by the National Science Founda-tion, the U.S. Department of Homeland Security, and the U.S. De-partment of Agriculture through NSF Award #EF-0832858, withadditional support from The University of Tennessee, Knoxville. V.K. acknowledges support provided by the Institute of Entomology(RVO:60077344). M.A.L. gratefully acknowledges support fromNSERC Discovery and Accelerator grants, a Canada Research Chairand a Killam Fellowship. This research was also supported by agrant to MAL from the Natural Science and Engineering ResearchCouncil of Canada (grant no. NET GP 434810-12) to the TRIANetwork, with contributions from Alberta Agriculture and For-estry, Foothills Research Institute, Manitoba Conservation andWater Stewardship, Natural Resources Canada – Canadian ForestService, Northwest Territories Environment and Natural Resources,Ontario Ministry of Natural Resources and Forestry, SaskatchewanMinistry of Environment, West Fraser and Weyerhaeuser.

    Appendix A. Local stability analysis

    Model (11) defines a linear system with matrix

  • V. Křivan et al. / Journal of Theoretical Biology 407 (2016) 25–3736

    ββ

    =− − −

    − ( )

    ⎛⎝⎜

    ⎞⎠⎟A

    g g

    1.

    A.1

    0

    0

    Because the trace of A is given by β( ) = − − −A gTr 1 0, while thedeterminant of A is given by β β( ) = + +A g gDet 0 0, both eigenva-lues of A have negative real parts for positive parameter values.Accordingly, the endemic equilibrium (12) will be asymptoticallystable.

    Appendix B. Behavior of trajectories of model (6) and (8) alongthe discontinuity line

    We study behavior of trajectories along the switching lineΓ μ= ( + ) −I S1 . The gradient vector to this line is Γ= { − }n , 1 .

    Let f1 denote the right-hand side of (10) and f2 denote the right-hand side of (11), respectively. Then

    β Γ〈 〉 = 〈 〉 + ( + )n f n f S, , 1 ,2 1 0

    where 〈 〉.,. denotes the scalar product. It follows that if 〈 〉 >n f, 01then 〈 〉 >n f, 0,2 or, similarly, if 〈 〉 | ¯ = } = − ( )X X R F RPr 1 , . Our assumption that trees are in-fested at rate β0 when θ>X , and that no trees become infestedwhen θ≤X gives the rate of infestation of new trees, β ( )R , interms of θ( )F R, as

    β β θ( ) = ( − ( )) ( )R F R1 , . C.10

    Beetle random dispersal is often described by a Poisson dis-tribution Pois(R). We do not include the specifics of active ag-gregation with respect to pheromones in the analysis. An approachpioneered for insects by Waters (1959) and popularized by May(1978) subsumes the spatial and behavioral complexities that leadto patterns of aggregation into the single phenomenological as-sumption that the net distribution of attacks upon hosts is of ne-gative binomial form. Although this was initially developed forparasitoids rather than bark beetles, the underlying modelingphilosophy is the same. In this case the Negative Binomial dis-tribution has mean ¯ =X R and dispersion parameter k, and is de-noted by ( )NB R k, . Unfortunately, because θ( )F R, is a complexcumulative distribution function, it creates difficulties in terms ofmodel analysis. We therefore replace θ( )F R, in (C.1) with a Hillfunction capable of caricaturing the cumulative distributionfunction (Fig. C1). We do not claim that the Hill function is aperfect approximation for the cumulative distribution function forthe negative binomial, only that it is an appropriate caricature for

    the degree of precision needed for the modeling at hand. Para-meter Γ in the Hill function (3) approximates the thresholdnumber of beetles required for successful infestation, θ, while nplays a role similar to the dispersion parameter, k. In particular,low values of n represent high levels of aggregation, while highvalues of n indicate overdispersion (see Fig. C1). If, for example, weassume that successful colonization of moderate size trees re-quires θ = 1000 beetles per tree, then by comparing the Hillfunction in Eq. (3) to the expression that it approximates in Eq.(C.1) and assuming either the Negative Binomial or Poisson dis-tribution for θ( )F R, , we can find the value of n and Γ that bestapproximates the beetle distribution. The right panel of Fig. 1shows the case when n¼10. Note that the qualitative behavior issimilar to that of the uniform beetle scenario although the quan-titative details differ.

    References

    Addison, A., Powell, J.A., Bentz, B.J., Six, D.L., 2015. Integrating models to investigatecritical phenological overlaps in complex ecological interactions: the mountainpine beetle–fungus symbiosis. J. Theor. Biol. 368, 55–66.

    Anderbrant, O., Schlyter, F., Birgersson, G., 1985. Intraspecific competition affectingparents and offspring in the bark beetle Ips typographus. Oikos 45, 89–98.

    Anderegg, W.R.L., Hicke, J., Fisher, R., Allen, C., Aukema, J., Bentz, B., Hood, S.,Lichstein, J., Macalady, A., McDowell, N., Pan, Y., Raffa, K., Sala, A., Shaw, J.,Stephenson, N., Tague, C., Zeppel, M., 2015. Tree mortality from drought, insectsand their interactions in a changing climate. New Phytol. 208, 674–683.

    Baker, W.L., 2009. Fire Ecology in Rocky Mountain Landscapes. Island Press, Wa-shington DC.

    Bentz, B., Vandygriff, J., Jensen, C., Coleman, T., Maloney, P., Smith, S., Grady, A.,Schen-Langenheim, G., 2014. Mountain pine beetle voltinism and life historycharacteristics across latitudinal and elevational gradients in the westernUnited States. For. Sci. 60, 434–449.

    Bentz, B.J., Régnière, J., Fettig, C.J., Hansen, E.M., Hayes, J.L., Hicke, J.A., Kelsey, R.G.,

    http://refhub.elsevier.com/S0022-5193(16)30192-8/sbref1http://refhub.elsevier.com/S0022-5193(16)30192-8/sbref1http://refhub.elsevier.com/S0022-5193(16)30192-8/sbref1http://refhub.elsevier.com/S0022-5193(16)30192-8/sbref1http://refhub.elsevier.com/S0022-5193(16)30192-8/sbref2http://refhub.elsevier.com/S0022-5193(16)30192-8/sbref2http://refhub.elsevier.com/S0022-5193(16)30192-8/sbref2http://refhub.elsevier.com/S0022-5193(16)30192-8/sbref3http://refhub.elsevier.com/S0022-5193(16)30192-8/sbref3http://refhub.elsevier.com/S0022-5193(16)30192-8/sbref3http://refhub.elsevier.com/S0022-5193(16)30192-8/sbref3http://refhub.elsevier.com/S0022-5193(16)30192-8/sbref3http://refhub.elsevier.com/S0022-5193(16)30192-8/sbref4http://refhub.elsevier.com/S0022-5193(16)30192-8/sbref4http://refhub.elsevier.com/S0022-5193(16)30192-8/sbref5http://refhub.elsevier.com/S0022-5193(16)30192-8/sbref5http://refhub.elsevier.com/S0022-5193(16)30192-8/sbref5http://refhub.elsevier.com/S0022-5193(16)30192-8/sbref5http://refhub.elsevier.com/S0022-5193(16)30192-8/sbref5http://refhub.elsevier.com/S0022-5193(16)30192-8/sbref6

  • V. Křivan et al. / Journal of Theoretical Biology 407 (2016) 25–37 37

    Negrón, T.P., Seybold, S.J., 2010. Climate change and bark beetles of the westernUnited States and Canada: direct and indirect effects. BioScience 60, 602–613.

    Berryman, A., Stenseth, N., Wollkind, D., 1984. Metastability of forest ecosystemsinfested by bark beetles. Res. Popul. Ecol. 26, 13–29.

    Berryman, A.A., 1982. Biological-control, thresholds, and pest outbreaks. Environ.Entomol. 11, 544–549.

    Boone, C.K., Aukema, B.H., Bohlmann, J., Carroll, A.L., Raffa, K.F., 2011. Efficacy of treedefense physiology varies with bark beetle population density: a basis for po-sitive feedback in eruptive species. Can. J. For. Res. – Rev. Can. Rech. For. 41,1174–1188.

    Boone, C.K., Keefover-Ring, K., Mapes, A.C., Adams, A.S., Bohlmann, J., Raffa, K.F.,2013. Bacteria associated with a tree-killing insect reduce concentrations ofplant defense compounds. J. Chem. Ecol. 39, 1003–1006.

    Chapman, T.B., Veblen, T.T., Schoennagel, T., 2012. Spatiotemporal patterns ofmountain pine beetle activity in the southern Rocky Mountains. Ecology 93,2175–2185.

    Chubaty, A.M., Roitberg, B.D., Li, C., 2009. A dynamic host selection model formountain pine beetle, Dendroctonus ponderosae Hopkins. Ecol. Model. 220,1241–1250.

    Clark, J.S., Lewis, M., Horvath, L., 2001. Invasion by extremes: population spreadwith variation in dispersal and reproduction. Am. Nat. 157, 537–554.

    Colombo, R., Křivan, V., 1993. Selective strategies in food webs. IMA J. Math. Appl.Med. Biol. 10, 281–291.

    Duncan, J.P., Powell, J.A., Gordillo, L.F., Eason, J., 2015. A model for mountain pinebeetle outbreaks in an age-structured forest: predicting severity and outbreak-recovery cycle period. Bull. Math. Biol. 77, 1256–1284.

    Fettig, C.J., Gibson, K.E., Munson, A.S., Negron, J.F., 2014. Cultural practices forprevention and mitigation of mountain pine beetle infestations. For. Sci. 60,450–463.

    Fettig, C.J., Reid, M.L., Bentz, B.J., Sevanto, S., Spittlehouse, D.L., Wang, T., 2013.Changing climates, changing forests: a western North American perspective. J.For. 111, 214–228.

    Filippov, A.F., 1988. Differential Equations with Discontinuous Righthand Sides.Kluwer Academic Publishers, Dordrecht.

    Friedenberg, N.A., Powell, J.A., Ayres, M.P., 2007. Synchrony's double edge: transientdynamics and the Allee effect in stage structured populations. Ecol. Lett. 10,564–573.

    Gaylord, M.L., Kolb, T.E., Pockman, W.T., Plaut, J.A., Yepez, E.A., Macalady, A.K.,Pangle, R.E., McDowell, N.G., 2013. Drought predisposes pinon-juniper wood-lands to insect attacks and mortality. New Phytol. 198, 567–578.

    Gilbert, E., Powell, J., Logan, J., Bentz, B., 2004. Comparison of three models pre-dicting developmental milestones given environmental and individual varia-tion. Bull. Math. Biol. 66, 1821–1850.

    Goheen, D., Hansen, E., 1993. Effects of pathogens and bark beetles on forests. In:Schowalter, T., Filip, G. (Eds.), Beetle–pathogen Interactions in Conifer Forests.Academic Press, London, UK, pp. 175–196.

    Hansen, E.M., 2014. Forest development and carbon dynamics after mountain pinebeetle outbreaks. For. Sci. 60, 476–488.

    Hart, S.J., Veblen, T.T., Eisenhart, K.S., Jarvis, D., Kulakowski, D., 2014. Drought in-duces spruce beetle (Dendroctonus rufipennis) outbreaks across northwesternColorado. Ecology 95, 930–939.

    Heavilin, J., Powell, J., 2008. A novel method for fitting spatio-temporal models todata, with applications to the dynamics of mountain pine beetle. Nat. Resour.Model. 21, 489–524.

    Hicke, J.A., Meddens, A.J.H., Allen, C.D., Kolden, C.A., 2013. Carbon stocks of treeskilled by bark beetles and wildfire in the western United States. Environ. Res.Lett. 8, 035032.

    Hofstetter, R., Dinkins-Bookwalter, J., Davis, T., Klepzig, K., 2015. Symbiotic asso-ciations of bark beetles. In: Vega, F.E., Hofstetter, R.W. (Eds.), Bark Beetles:

    Biology and Ecology of Native and Invasive Species. Academic Press, London,UK, pp. 209–246.

    Johnson, E.A., Fryer, G.I., 1989. Population dynamics in lodgepole pine-Engelmannspruce forest. Ecology 70, 1335–1345.

    Lewis, M.A., Nelson, W., Xu, C., 2010. A structured threshold model for mountainpine beetle outbreak. Bull. Math. Biol. 72, 565–589.

    Light, D.M., Birch, M.C., Paine, T.D., 1983. Laboratory study of intraspecific and in-terspecific competition within and between two sympatric bark beetle species,Ips pini and I. paraconfusus. Z. Angew. Entomol. 96, 233–241.

    Logan, J.A., White, P., Bentz, B.J., Powell, J.A., 1998. Model analysis of spatial patternsin mountain pine beetle outbreaks. Theor. Popul. Biol. 53, 236–255.

    May, R.M., 1978. Host-parasitoid systems in patchy environments: A phenomen-ological model. Journal of Animal Ecology 47, 833–843.

    Martinson, S.J., Ylioja, T., Sullivan, B.T., Billings, R.F., Ayres, M.P., 2013. Alternateattractors in the population dynamics of a tree-killing bark beetle. Popul. Ecol.55, 95–106.

    Meddens, A.J.H., Hicke, J.A., Ferguson, C.A., 2012. Spatiotemporal patterns of ob-served bark beetle-caused tree mortality in British Columbia and the westernUnited States. Ecol. Appl. 22, 1876–1891.

    Mitchell, R., Preisler, H., 1991. Analysis of spatial patterns of lodgepole pine at-tacked by outbreak populations of the mountain pine-beetle. For. Sci. 37,1390–1408.

    Økland, B., Bjørnstad, O., 2006. A resource-depletion model of forest insect out-breaks. Ecology 87, 283–290.

    Powell, J., Logan, J., Bentz, B., 1996. Local projections for a global model of mountainpine beetle attacks. J. Theor. Biol. 179, 243–260.

    Powell, J.A., Bentz, B.J., 2009. Connecting phenological predictions with populationgrowth rates for mountain pine beetle, an outbreak insect. Landsc. Ecol. 24,657–672.

    Powell, J.A., Bentz, B.J., 2014. Phenology and density-dependent dispersal predictpatterns of mountain pine beetle (Dendroctonus ponderosae) impact. Ecol.Model. 273, 173–185.

    Raffa, K.F., Aukema, B.H., Bentz, B.J., Carroll, A.L., Hicke, J.A., Turner, M.G., Romme,W.H., 2008. Cross-scale drivers of natural disturbances prone to anthropogenicamplification: the dynamics of bark beetle eruptions. Bioscience 58, 501–517.

    Raffa, K.F., Berryman, A., 1983. The role of host plant-resistance in the colonizationbehavior and ecology of bark beetles (Coleoptera, Scolytidae). Ecol. Monogr. 53,27–49.

    Regniere, J., Bentz, B., 2007. Modeling cold tolerance in the mountain pine beetle,Dendroctonus ponderosae. J. Insect Physiol. 53, 559–572.

    Safranyik, L., Carroll, A.L., Regniere, J., Langor, D.W., Riel, W.G., Shore, T.L., Peter, B.,Cooke, B.J., Nealis, V.G., Taylor, S.W., 2010. Potential for range expansion ofmountain pine beetle into the boreal forest of North America. Can. Entomol.142, 415–442.

    Santos, M.J., Whitham, T.G., 2010. Predictors of Ips confusus outbreaks during arecord drought in southwestern USA: Implications for monitoring and man-agement. Environ. Manag. 45, 239–249.

    Schelhaas, M., Nabuurs, G., Schuck, A., 2003. Natural disturbances in the Europeanforests in the 19th and 20th centuries. Glob. Change Biol. 9, 1620–1633.

    Therrien, J., Mason, C.J., Adams, A., Aukema, B.H., Currie, C.R., Erbilgin, N., 2015.Bacteria influence mountain pine beetle brood development through interac-tions with symbiotic and antagonistic fungi: implications for climate-drivenhost range expansion. Oecologia 179, 467–485.

    Waters, W.E., 1959. A quantitative measure of aggregation in insects. Journal ofEconomic Entomology 52, 1180–1184.

    Weed, A.S., Bentz, B.J., Ayres, M.P., Holmes, T.P., 2015. Geographically variable re-sponse of Dendroctonus ponderosae to winter warming in the western UnitedStates. Landsc. Ecol. 30, 1075–1093.

    http://refhub.elsevier.com/S0022-5193(16)30192-8/sbref6http://refhub.elsevier.com/S0022-5193(16)30192-8/sbref6http://refhub.elsevier.com/S0022-5193(16)30192-8/sbref6http://refhub.elsevier.com/S0022-5193(16)30192-8/sbref7http://refhub.elsevier.com/S0022-5193(16)30192-8/sbref7http://refhub.elsevier.com/S0022-5193(16)30192-8/sbref7http://refhub.elsevier.com/S0022-5193(16)30192-8/sbref8http://refhub.elsevier.com/S0022-5193(16)30192-8/sbref8http://refhub.elsevier.com/S0022-5193(16)30192-8/sbref8http://refhub.elsevier.com/S0022-5193(16)30192-8/sbref9http://refhub.elsevier.com/S0022-5193(16)30192-8/sbref9http://refhub.elsevier.com/S0022-5193(16)30192-8/sbref9http://refhub.elsevier.com/S0022-5193(16)30192-8/sbref9http://refhub.elsevier.com/S0022-5193(16)30192-8/sbref9http://refhub.elsevier.com/S0022-5193(16)30192-8/sbref10http://refhub.elsevier.com/S0022-5193(16)30192-8/sbref10http://refhub.elsevier.com/S0022-5193(16)30192-8/sbref10http://refhub.elsevier.com/S0022-5193(16)30192-8/sbref10http://refhub.elsevier.com/S0022-5193(16)30192-8/sbref11http://refhub.elsevier.com/S0022-5193(16)30192-8/sbref11http://refhub.elsevier.com/S0022-5193(16)30192-8/sbref11http://refhub.elsevier.com/S0022-5193(16)30192-8/sbref11http://refhub.elsevier.com/S0022-5193(16)30192-8/sbref12http://refhub.elsevier.com/S0022-5193(16)30192-8/sbref12http://refhub.elsevier.com/S0022-5193(16)30192-8/sbref12http://refhub.elsevier.com/S0022-5193(16)30192-8/sbref12http://refhub.elsevier.com/S0022-5193(16)30192-8/sbref13http://refhub.elsevier.com/S0022-5193(16)30192-8/sbref13http://refhub.elsevier.com/S0022-5193(16)30192-8/sbref13http://refhub.elsevier.com/S0022-5193(16)30192-8/sbref14http://refhub.elsevier.com/S0022-5193(16)30192-8/sbref14http://refhub.elsevier.com/S0022-5193(16)30192-8/sbref14http://refhub.elsevier.com/S0022-5193(16)30192-8/sbref15http://refhub.elsevier.com/S0022-5193(16)30192-8/sbref15http://refhub.elsevier.com/S0022-5193(16)30192-8/sbref15http://refhub.elsevier.com/S0022-5193(16)30192-8/sbref15http://refhub.elsevier.com/S0022-5193(16)30192-8/sbref16http://refhub.elsevier.com/S0022-5193(16)30192-8/sbref16http://refhub.elsevier.com/S0022-5193(16)30192-8/sbref16http://refhub.elsevier.com/S0022-5193(16)30192-8/sbref16http://refhub.elsevier.com/S0022-5193(16)30192-8/sbref17http://refhub.elsevier.com/S0022-5193(16)30192-8/sbref17http://refhub.elsevier.com/S0022-5193(16)30192-8/sbref17http://refhub.elsevier.com/S0022-5193(16)30192-8/sbref17http://refhub.elsevier.com/S0022-5193(16)30192-8/sbref18http://refhub.elsevier.com/S0022-5193(16)30192-8/sbref18http://refhub.elsevier.com/S0022-5193(16)30192-8/sbref19http://refhub.elsevier.com/S0022-5193(16)30192-8/sbref19http://refhub.elsevier.com/S0022-5193(16)30192-8/sbref19http://refhub.elsevier.com/S0022-5193(16)30192-8/sbref19http://refhub.elsevier.com/S0022-5193(16)30192-8/sbref20http://refhub.elsevier.com/S0022-5193(16)30192-8/sbref20http://refhub.elsevier.com/S0022-5193(16)30192-8/sbref20http://refhub.elsevier.com/S0022-5193(16)30192-8/sbref20http://refhub.elsevier.com/S0022-5193(16)30192-8/sbref21http://refhub.elsevier.com/S0022-5193(16)30192-8/sbref21http://refhub.elsevier.com/S0022-5193(16)30192-8/sbref21http://refhub.elsevier.com/S0022-5193(16)30192-8/sbref21http://refhub.elsevier.com/S0022-5193(16)30192-8/sbref22http://refhub.elsevier.com/S0022-5193(16)30192-8/sbref22http://refhub.elsevier.com/S0022-5193(16)30192-8/sbref22http://refhub.elsevier.com/S0022-5193(16)30192-8/sbref22http://refhub.elsevier.com/S0022-5193(16)30192-8/sbref23http://refhub.elsevier.com/S0022-5193(16)30192-8/sbref23http://refhub.elsevier.com/S0022-5193(16)30192-8/sbref23http://refhub.elsevier.com/S0022-5193(16)30192-8/sbref24http://refhub.elsevier.com/S0022-5193(16)30192-8/sbref24http://refhub.elsevier.com/S0022-5193(16)30192-8/sbref24http://refhub.elsevier.com/S0022-5193(16)30192-8/sbref24http://refhub.elsevier.com/S0022-5193(16)30192-8/sbref25http://refhub.elsevier.com/S0022-5193(16)30192-8/sbref25http://refhub.elsevier.com/S0022-5193(16)30192-8/sbref25http://refhub.elsevier.com/S0022-5193(16)30192-8/sbref25http://refhub.elsevier.com/S0022-5193(16)30192-8/sbref26http://refhub.elsevier.com/S0022-5193(16)30192-8/sbref26http://refhub.elsevier.com/S0022-5193(16)30192-8/sbref26http://refhub.elsevier.com/S0022-5193(16)30192-8/sbref27http://refhub.elsevier.com/S0022-5193(16)30192-8/sbref27http://refhub.elsevier.com/S0022-5193(16)30192-8/sbref27http://refhub.elsevier.com/S0022-5193(16)30192-8/sbref27http://refhub.elsevier.com/S0022-5193(16)30192-8/sbref27http://refhub.elsevier.com/S0022-5193(16)30192-8/sbref28http://refhub.elsevier.com/S0022-5193(16)30192-8/sbref28http://refhub.elsevier.com/S0022-5193(16)30192-8/sbref28http://refhub.elsevier.com/S0022-5193(16)30192-8/sbref29http://refhub.elsevier.com/S0022-5193(16)30192-8/sbref29http://refhub.elsevier.com/S0022-5193(16)30192-8/sbref29http://refhub.elsevier.com/S0022-5193(16)30192-8/sbref30http://refhub.elsevier.com/S0022-5193(16)30192-8/sbref30http://refhub.elsevier.com/S0022-5193(16)30192-8/sbref30http://refhub.elsevier.com/S0022-5193(16)30192-8/sbref30http://refhub.elsevier.com/S0022-5193(16)30192-8/sbref31http://refhub.elsevier.com/S0022-5193(16)30192-8/sbref31http://refhub.elsevier.com/S0022-5193(16)30192-8/sbref31http://refhub.elsevier.com/S0022-5193(16)30192-8/sbref277704http://refhub.elsevier.com/S0022-5193(16)30192-8/sbref277704http://refhub.elsevier.com/S0022-5193(16)30192-8/sbref277704http://refhub.elsevier.com/S0022-5193(16)30192-8/sbref32http://refhub.elsevier.com/S0022-5193(16)30192-8/sbref32http://refhub.elsevier.com/S0022-5193(16)30192-8/sbref32http://refhub.elsevier.com/S0022-5193(16)30192-8/sbref32http://refhub.elsevier.com/S0022-5193(16)30192-8/sbref33http://refhub.elsevier.com/S0022-5193(16)30192-8/sbref33http://refhub.elsevier.com/S0022-5193(16)30192-8/sbref33http://refhub.elsevier.com/S0022-5193(16)30192-8/sbref33http://refhub.elsevier.com/S0022-5193(16)30192-8/sbref34http://refhub.elsevier.com/S0022-5193(16)30192-8/sbref34http://refhub.elsevier.com/S0022-5193(16)30192-8/sbref34http://refhub.elsevier.com/S0022-5193(16)30192-8/sbref34http://refhub.elsevier.com/S0022-5193(16)30192-8/sbref35http://refhub.elsevier.com/S0022-5193(16)30192-8/sbref35http://refhub.elsevier.com/S0022-5193(16)30192-8/sbref35http://refhub.elsevier.com/S0022-5193(16)30192-8/sbref36http://refhub.elsevier.com/S0022-5193(16)30192-8/sbref36http://refhub.elsevier.com/S0022-5193(16)30192-8/sbref36http://refhub.elsevier.com/S0022-5193(16)30192-8/sbref37http://refhub.elsevier.com/S0022-5193(16)30192-8/sbref37http://refhub.elsevier.com/S0022-5193(16)30192-8/sbref37http://refhub.elsevier.com/S0022-5193(16)30192-8/sbref37http://refhub.elsevier.com/S0022-5193(16)30192-8/sbref38http://refhub.elsevier.com/S0022-5193(16)30192-8/sbref38http://refhub.elsevier.com/S0022-5193(16)30192-8/sbref38http://refhub.elsevier.com/S0022-5193(16)30192-8/sbref38http://refhub.elsevier.com/S0022-5193(16)30192-8/sbref39http://refhub.elsevier.com/S0022-5193(16)30192-8/sbref39http://refhub.elsevier.com/S0022-5193(16)30192-8/sbref39http://refhub.elsevier.com/S0022-5193(16)30192-8/sbref39http://refhub.elsevier.com/S0022-5193(16)30192-8/sbref40http://refhub.elsevier.com/S0022-5193(16)30192-8/sbref40http://refhub.elsevier.com/S0022-5193(16)30192-8/sbref40http://refhub.elsevier.com/S0022-5193(16)30192-8/sbref40http://refhub.elsevier.com/S0022-5193(16)30192-8/sbref41http://refhub.elsevier.com/S0022-5193(16)30192-8/sbref41http://refhub.elsevier.com/S0022-5193(16)30192-8/sbref41http://refhub.elsevier.com/S0022-5193(16)30192-8/sbref42http://refhub.elsevier.com/S0022-5193(16)30192-8/sbref42http://refhub.elsevier.com/S0022-5193(16)30192-8/sbref42http://refhub.elsevier.com/S0022-5193(16)30192-8/sbref42http://refhub.elsevier.com/S0022-5193(16)30192-8/sbref42http://refhub.elsevier.com/S0022-5193(16)30192-8/sbref43http://refhub.elsevier.com/S0022-5193(16)30192-8/sbref43http://refhub.elsevier.com/S0022-5193(16)30192-8/sbref43http://refhub.elsevier.com/S0022-5193(16)30192-8/sbref43http://refhub.elsevier.com/S0022-5193(16)30192-8/sbref44http://refhub.elsevier.com/S0022-5193(16)30192-8/sbref44http://refhub.elsevier.com/S0022-5193(16)30192-8/sbref44http://refhub.elsevier.com/S0022-5193(16)30192-8/sbref45http://refhub.elsevier.com/S0022-5193(16)30192-8/sbref45http://refhub.elsevier.com/S0022-5193(16)30192-8/sbref45http://refhub.elsevier.com/S0022-5193(16)30192-8/sbref45http://refhub.elsevier.com/S0022-5193(16)30192-8/sbref45http://refhub.elsevier.com/S0022-5193(16)30192-8/sbref660332http://refhub.elsevier.com/S0022-5193(16)30192-8/sbref660332http://refhub.elsevier.com/S0022-5193(16)30192-8/sbref660332http://refhub.elsevier.com/S0022-5193(16)30192-8/sbref46http://refhub.elsevier.com/S0022-5193(16)30192-8/sbref46http://refhub.elsevier.com/S0022-5193(16)30192-8/sbref46http://refhub.elsevier.com/S0022-5193(16)30192-8/sbref46

    A dynamical model for bark beetle outbreaksIntroductionReview of existing models

    One-stand modelsTree recruitment, G(S,I)Infestation rate, β(R)Model parametersNon-dimensionalizationPseudo-steady state approximationScenario one: uniform beetle distributionScenario two: aggregated beetle distribution

    Two-stand modelDiscussionSummaryModel limitationsModel extensionsConcluding remarks

    AcknowledgmentsLocal stability analysisBehavior of trajectories of model (6) and (8) along the discontinuity lineRelation between beetle aggregation and steepness of the Hill functionReferences