spatial patterns in stage-structured populations with...

114
Spatial Patterns in Stage-Structured Populations with Density Dependent Dispersal Item Type text; Electronic Dissertation Authors Robertson, Suzanne Lora Publisher The University of Arizona. Rights Copyright © is held by the author. Digital access to this material is made possible by the University Libraries, University of Arizona. Further transmission, reproduction or presentation (such as public display or performance) of protected items is prohibited except with permission of the author. Download date 08/07/2018 23:27:36 Link to Item http://hdl.handle.net/10150/194472

Upload: truongnguyet

Post on 12-Jun-2018

220 views

Category:

Documents


0 download

TRANSCRIPT

Spatial Patterns in Stage-StructuredPopulations with Density Dependent Dispersal

Item Type text; Electronic Dissertation

Authors Robertson, Suzanne Lora

Publisher The University of Arizona.

Rights Copyright © is held by the author. Digital access to this materialis made possible by the University Libraries, University of Arizona.Further transmission, reproduction or presentation (such aspublic display or performance) of protected items is prohibitedexcept with permission of the author.

Download date 08/07/2018 23:27:36

Link to Item http://hdl.handle.net/10150/194472

Spatial Patterns in Stage-Structured

Populations with Density Dependent Dispersal

by

Suzanne Lora Robertson

A Dissertation Submitted to the Faculty of the

Graduate Interdisciplinary Programin Applied Mathematics

In Partial Fulfillment of the RequirementsFor the Degree of

Doctor of Philosophy

In the Graduate College

The University of Arizona

2 0 0 9

2

THE UNIVERSITY OF ARIZONAGRADUATE COLLEGE

As members of the Final Examination Committee, we certify that we have read thedissertation prepared by Suzanne Lora Robertson entitledSpatial Patterns in Stage-Structured Populations with Density

Dependent Dispersal

and recommend that it be accepted as fulfilling the dissertation requirement for theDegree of Doctor of Philosophy.

Date: 24 April 2009

J. M. Cushing

Date: 24 April 2009

Michael Tabor

Date: 24 April 2009

Joseph Watkins

Final approval and acceptance of this dissertation is contingent upon the candidate’ssubmission of the final copies of the dissertation to the Graduate College.

I hereby certify that I have read this dissertation prepared under my direction andrecommend that it be accepted as fulfilling the dissertation requirement.

Date: 24 April 2009

Dissertation Director: J. M. Cushing

3

STATEMENT BY AUTHOR

This dissertation has been submitted in partial fulfillment of requirements for anadvanced degree at The University of Arizona and is deposited in the UniversityLibrary to be made available to borrowers under rules of the Library.

Brief quotations from this dissertation are allowable without special permission,provided that accurate acknowledgment of source is made. Requests for permissionfor extended quotation from or reproduction of this manuscript in whole or in partmay be granted by the head of the major department or the Dean of the GraduateCollege when in his or her judgment the proposed use of the material is in the interestsof scholarship. In all other instances, however, permission must be obtained from theauthor.

SIGNED: Suzanne Lora Robertson

4

Acknowledgments

I would like to thank my advisor, Jim Cushing, for being a wonderful mentor. Hehas taught me so much and I am grateful for the opportunity to work with him.

Thanks to my committee members, Joseph Watkins and Michael Tabor, for all oftheir professional guidance and support during my time as a graduate student.

I am grateful to Bob Costantino for providing much of the biological data for thiswork and for his insight during our many research discussions.

I gratefully acknowledge Shandelle Henson, my undergraduate mentor, for inspiringme to pursue a graduate degree in Applied Mathematics.

Finally, I wish to thank my family and friends for their love and encouragement.

5

Dedication

For Mom

6

Table of Contents

List of Figures . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8

List of Tables . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11

Abstract . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12

Chapter 1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . 13

1.1. Life Cycle Stage Interactions and Spatial Structure . . . . . . . . . . 131.2. Life Cycle Stages and Spatial Patterns in Flour Beetles . . . . . . . . 15

1.2.1. Tribolium castaneum and Tribolium confusum . . . . . . . . . 151.2.2. Tribolium brevicornis . . . . . . . . . . . . . . . . . . . . . . . 16

1.3. Density Dependent Dispersal . . . . . . . . . . . . . . . . . . . . . . . 221.4. Spatial Models and Prior Investigations . . . . . . . . . . . . . . . . . 23

Chapter 2. A Bifurcation Analysis of Stage-StructuredDensity-Dependent Integrodifference Equations . . . . . . . . . 28

2.1. Definitions and Preliminaries . . . . . . . . . . . . . . . . . . . . . . . 292.2. Model Development and Existence of Equilibria . . . . . . . . . . . . 312.3. Equilibrium Stability . . . . . . . . . . . . . . . . . . . . . . . . . . . 37

2.3.1. Extinction Equilibrium . . . . . . . . . . . . . . . . . . . . . . 372.3.2. Stability and Direction of Bifurcation . . . . . . . . . . . . . . 38

2.4. Examples of Bifurcation Theory . . . . . . . . . . . . . . . . . . . . . 402.4.1. Example 1: Uniform dispersal . . . . . . . . . . . . . . . . . . 402.4.2. Example 2: Spatially dependent dispersal . . . . . . . . . . . . 442.4.3. Example 3: A more complex spatially dependent kernel . . . . 47

2.5. Appendix . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 512.5.1. The Krein-Rutman Theorem . . . . . . . . . . . . . . . . . . . 512.5.2. Relationship between n and λ0 . . . . . . . . . . . . . . . . . 512.5.3. Compactness of T + nΦ . . . . . . . . . . . . . . . . . . . . . 53

Chapter 3. Juvenile-Adult (Toy) Models . . . . . . . . . . . . . . . 55

3.1. Hostile Boundary Conditions . . . . . . . . . . . . . . . . . . . . . . . 563.1.1. Application of Theory . . . . . . . . . . . . . . . . . . . . . . 63

3.2. Mixed Boundary Conditions . . . . . . . . . . . . . . . . . . . . . . . 663.3. Role of Domain Size . . . . . . . . . . . . . . . . . . . . . . . . . . . 69

Table of Contents—Continued

7

Chapter 4. Case Studies Using Models for Flour Beetles (Tri-

bolium) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 73

4.1. Spatial Patterns in Tribolium castaneum . . . . . . . . . . . . . . . . 734.1.1. The LPA Model . . . . . . . . . . . . . . . . . . . . . . . . . 734.1.2. The Spatial LPA Model . . . . . . . . . . . . . . . . . . . . . 74

4.2. Spatial Patterns in Tribolium brevicornis . . . . . . . . . . . . . . . . 784.2.1. The SLPA Model . . . . . . . . . . . . . . . . . . . . . . . . 794.2.2. Parametrization . . . . . . . . . . . . . . . . . . . . . . . . . . 814.2.3. The Spatial SLPA Model . . . . . . . . . . . . . . . . . . . . 85

Chapter 5. Concluding Remarks . . . . . . . . . . . . . . . . . . . . . 100

5.1. Future Investigations . . . . . . . . . . . . . . . . . . . . . . . . . . . 105

References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 108

8

List of Figures

Figure 1.1. Data from Ghent [14]. 0 represents the surface of the vial and 1represents the bottom. . . . . . . . . . . . . . . . . . . . . . . . . . . . . 18

Figure 1.2. Cultures of T. brevicornis showing segregation of adults andother life stages. Photo 1.2b by R. F. Costantino. . . . . . . . . . . . . . 19

Figure 1.3. Aerial view of Tribolium castaneum adults on surface of rectan-gular box of flour. There are some adults near the edges of the box butno distinct pattern or segregation of life cycle stages. . . . . . . . . . . . 20

Figure 1.4. Aerial view of Tribolium brevicornis. Each of three rows showsegregation of adults and pupae, illustrating the pupal nest on the left.Each row was started with large larvae and adults on the left side. Theywere restrained in this subhabitat for 6 weeks, then a panel was removedand they were allowed to disperse throughout the entire row. This photowas taken a week after the door was opened. Photo by R. F. Costantino. 20

Figure 1.5. New callow adults emerging from nest in (b) indicate large larvaereturn to nest to pupate. Photos are 6 weeks apart. Photos by R. F.Costantino. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 21

Figure 1.6. Aerial view of Tribolium brevicornis. Doors are open at ends ofrows to allow beetles to move throughout entire box. Culture started with100 adults at upper right corner, immediately permitted to disperse. Nopupal nest is established. . . . . . . . . . . . . . . . . . . . . . . . . . . 21

Figure 3.1. Time-equilibrium attractor of Juvenile-Adult IDE model (3.5).b = 25, µL = µA = 0.2. There is no density dependence in the dispersalkernels: DJ = DA = 0 . . . . . . . . . . . . . . . . . . . . . . . . . . . . 58

Figure 3.2. Time-equilibrium attractor for Juvenile-Adult IDE model (3.5),showing spatial segregation of juveniles and adults. b = 25, µL = µA =0.2. DJ = 0.5, DA = 0. . . . . . . . . . . . . . . . . . . . . . . . . . . . 59

Figure 3.3. 2-cycle attractor of Juvenile-Adult IDE model (3.5), with tem-poral segregation of juveniles and adults. b = 25, µL = µA = 0.2. DJ = 0,DA = 0.5 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 59

Figure 3.4. 4-cycle attractor of Juvenile-Adult IDE model (3.5), with tem-poral segregation of juveniles and adults. b = 25, µL = µA = 0.2.DJ = DA = 0.5 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 60

Figure 3.5. Shown is a section of the DJ − DA parameter plane for b = 2.5.µJ = µA = 0.2. We look at 0 ≤ DJ ≤ 1 and 0 ≤ DA ≤ 1. Allcombinations of DJ and DA in this range result in equilibrium dynamicsfor model (3.5). Spatial segregation is possible for large enough DJ andsmall enough DA. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 61

List of Figures—Continued

9

Figure 3.6. Shown is a section of the DJ − DA parameter plane for b = 10.µJ = µA = 0.2. In addition to the model (3.5) attractors seen for b = 2.5,we can also see 2 − cycles and temporal segregation. . . . . . . . . . . . 61

Figure 3.7. Shown is a section of the DJ − DA parameter plane for b = 25.µJ = µA = 0.2. In addition to the model (3.5) attractors seen for b = 10,we can also see 4 − cycles and 4 − cycles with temporal segregation. . . 62

Figure 3.8. Shown is a Matlab simulation of model (3.20) with mixedboundary conditions. The left endpoint of the spatial domain, 0, rep-resents the surface of a vial of flour and is a no-flux boundary. The rightendpoint, π

2, represents the bottom of a vial of flour and is a hostile bound-

ary. Model parameters: b = 25, µJ = µA = 0.2, DJ = 0.5, DA = 0. Adultdensities peak at the surface and juvenile densities peak below the surface. 69

Figure 3.9. Simulation of Juvenile-Adult IDE model (3.22) with populationdynamics given by (3.2). b = 20, µJ = µA = 0.3, density independentkernel (3.23). There is no spatial segregation of juveniles and adults. . . 71

Figure 3.10. Simulation of Juvenile-Adult IDE model (3.22) with populationdynamics given by (3.2). b = 20, µJ = µA = 0.3, DJ = DA = 1. Juvenilesand adults are spatially segregated. . . . . . . . . . . . . . . . . . . . . . 72

Figure 3.11. Simulation of Juvenile-Adult IDE model (3.22) with populationdynamics given by (3.2). b = 20, µJ = µA = 0.3, DJ = DA = 0.1. Thereis no spatial segregation of juveniles and adults. . . . . . . . . . . . . . . 72

Figure 4.1. Simulations of the spatial LPA model (4.4) for various parametervalues in the dispersal kernels (4.2). All attractors shown are temporalequilibria. LPA parameter values given by: b = 10, cel = 0.01, cea =0.0175, µL = 0.435, cpa = 0.016, µA = 0.075. . . . . . . . . . . . . . . . . 77

Figure 4.2. Shown are time series data (solid lines) and one-step SLPA pre-dictions (dashed lines) for small larvae, for each of the four replicates ofthe T. brevicornis cultures. Parameters used for one-step predictions aregiven in Table 4.1. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 86

Figure 4.3. Shown are time series data (solid lines) and one-step SLPA pre-dictions (dashed lines) for large larvae, for each of the four replicates ofthe T. brevicornis cultures. Parameters used for one-step predictions aregiven in Table 4.1. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 87

Figure 4.4. Shown are time series data (solid lines) and one-step SLPA pre-dictions (dashed lines) for pupae, for each of the four replicates of the T.

brevicornis cultures. Parameters used for one-step predictions are givenin Table 4.1. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 88

List of Figures—Continued

10

Figure 4.5. Shown are time series data (solid lines) and one-step SLPA pre-dictions (dashed lines) for adults, for each of the four replicates of the T.

brevicornis cultures. Parameters used for one-step predictions are givenin Table 4.1. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 89

Figure 4.6. The decision function γL is shown for several values of DLA.γL(At(x)) gives the fraction of larvae dispersing from location x at time tas a function of adult density at location x at time t. . . . . . . . . . . . 91

Figure 4.7. Temporal equilibrium attractor of spatial SLPA model. Attrac-tor is spatially uniform with (Se, Le, Pe, Ae) = (12.21, 71.77, 2.59, 163.8).Parameter values used for SLPA model are maximum likelihood esti-mates: b = 11.41, µL = 0.134, µA = 0.0158, ces = 0.0135,cea = 0.0223,cel = 0.0169, ki = 0.0194. DAP = 1 and DLA = 0.05. General initial con-dition: L0 = 100M , A0 = 0 on the subinterval of domain [0, M

4], M > 0.

Here M = 1. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 95Figure 4.8. Temporal equilibrium attractor of spatial SLPA model, con-

sisting of two patches with (Se, Le, Pe, Ae)=(0.015, 13.93, 12.06, 0.0016)on [0, M

4] and (S∗

e , L∗e, P

∗e , A∗

e)=(6.58, 13.92, 0.08, 259.4) on [M4, M ]. Pa-

rameter values used for SLPA model are the maximum likelihood esti-mates: b = 11.41, µL = 0.134, µA = 0.0158, ces = 0.0135,cea = 0.0223,cel = 0.0169, ki = 0.0194. DAP = 1 and DLA = 0.05. General initial con-dition: A0 = 100M , L0 = 0 on the subinterval of domain [0, M

4], M > 0.

Here M = 1. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 96

11

List of Tables

Table 4.1. Maximum likelihood parameter estimates for the stochastic SLPAmodel (4.8). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 85

Table 4.2. Residual Analysis. Means and variances for the square root resid-uals for stages S, L, P and A. Untransformed residuals can be seen inFigures 4.2-4.5, as the difference between the model prediction and ob-served stage vector at each time-step. Maximum likelihood estimates forthe variances of the transformed residuals are given in Table 4.1. . . . . 90

Table 4.3. Comparison of population vectors and total population size forspatially uniform versus spatially segregated (patchy) attractors for vary-ing degrees of inhibition on the spatial domain [0, 1]. All other parametervalues used were: b = 11.41, µL = 0.134, µA = 0.0158, ces = 0.0135,cea =0.0223, cel = 0.0169, DAP = 1 and DLA = 0.05. Simulations with initialcondition a were started with 25 large larvae on the left 1

4of the spa-

tial domain (an initial distribution of L0(x) = 100, A0(x) = 0 on [0, 14]).

Each simulation started with this initial condition went to the spatiallyuniform attractor illustrated in Figure 4.7. Simulations with initial con-dition b were started with 25 adults on the left 1

4of the spatial domain

(an initial distribution of L0(x) = 0, A0(x) = 100 on [0, 14]). With the

exception of ki = 0, simulations with initial condition b led to the patchyattractor illustrated in Figure 4.8. For each value of ki, the total numberof small larvae, large larvae, pupae, adults and total population size (thesum of all stages) are calculated for each attractor (by integrating eachequilibrium distribution from 0 to 1 with respect to the spatial variablex). The greater number is shown in bold. . . . . . . . . . . . . . . . . . 94

12

Abstract

Spatial segregation among life cycle stages has been observed in many stage-structured

species, including species of the flour beetle Tribolium. Patterns have been observed

both in homogeneous and heterogeneous environments. We investigate density depen-

dent dispersal of life cycle stages as a mechanism responsible for this separation. By

means of mathematical analysis and numerical simulations, we explore this hypothe-

sis using stage-structured, integrodifference equation (IDE) models that incorporate

density dependent dispersal kernels.

In Chapter 2 we develop a bifurcation theory approach to the existence and sta-

bility of (non-extinction) equilibria for a general class of structured integrodifference

equation models on finite spatial domains with density dependent kernels. We show

that a continuum of such equilibria bifurcates from the extinction equilibrium when it

loses stability as the net reproductive number n increases through 1. We give several

examples to illustrate the theory.

In Chapter 3 we investigate mechanisms that can lead to spatial patterns in two

dimensional Juvenile-Adult IDE models. The bifurcation theory shows that such

patterns do not arise for n near 1. For larger values of n we show, via numerical

simulation, that density dependent dispersal can lead to the segregation of life cycle

stages in the sense that each stage peaks in a different spatial location.

Finally, in Chapter 4, we construct spatial models to describe the population

dynamics of T. castaneum, T. confusum and T. brevicornis and use them to assess

density dependent dispersal mechanisms that are able to explain spatial patterns that

have been observed in these species.

13

Chapter 1

Introduction

1.1 Life Cycle Stage Interactions and Spatial Structure

Spatial dispersal can be an important component that affects the dynamics of popula-

tions. There are many factors that may cause an organism to move between different

spatial locations, including quality of the environment, crowding and competition for

resources (between and within species). If space is not homogenous, mobile organisms

can take advantage of these spatial heterogeneities to develop their own spatial niche.

For example, similar species may prefer slightly different environments or types of

resources. The resulting non-uniform density distribution can allow similar species to

coexist through spatial segregation in situations where one species would otherwise

go extinct (competitive exclusion).

Interactions among life cycle stages may also promote non-uniform spatial struc-

ture. For example, spatial heterogeneities can give cannibalized (prey) stages an

opportunity to seek refuge from the cannibalistic (predator) stages, resulting in the

spatial segregation of life cycle stages. In some cannibalistic species, such as the

freshwater isopod Thermosphaeroma thermophilum [22], there is evidence that the

observed intraspecific habitat segregation is the result of an adaptation to cannibalism

as a stage-specific predation risk rather than other factors such as stage-specific re-

source preferences. Laboratory experiments have shown that juveniles actively avoid

the adults (both sexes, but more so the cannibalistic adult males) by taking advantage

14

of available refuges to keep from being eaten.

Even if spatial refuges are not available, cannibalized stages may still try to avoid

the cannibals. This phenomenon has been observed in the isopod Saduria entomon

[28, 43]. Larger isopods will cannibalize the smaller ones until they reach a critical

size of about 12 mm. Sparrevik and Leonardsson [43] experimentally demonstrated

that the smaller isopods will avoid swimming into an area of water that has passed

through a population of larger isopods, and the severity of the avoidance behavior

increases with the density of the larger conspecifics. Furthermore, a negative cor-

relation between densities of large and small isopods has been found in the field,

indicating that cannibalism (along with density dependent avoidance adaptations)

plays an important role in spatial structure as well as in density regulation for this

species.

The spatial separation of different age classes or life cycle stages has also been

documented in many other cannibalistic species, serving to reduce predation mortal-

ity. A study of the cave-dwelling mysids Hemimysis speluncola [40] found juveniles

near the cave entrance, while adults inhabited the innermost parts of the cave. Juve-

niles of the amphipod species Pallasea quadrispinosa tend to occupy shallow waters

while adults are mainly found in deeper, cooler waters. This is very similar to the

phototactic response noted by Hunte and Myers [20] in gammarid amphipods, which

they hypothesize is an evolutionary adaptation to reduce cannibalism of juveniles by

adults. These examples show heterogeneous environments, together with life cycle

stage interactions, can lead to differences in spatial distributions among conspecifics.

Life cycle stage interactions alone are sometimes sufficient for the formation of

non-uniform spatial distributions, as spatial patterns have also been observed in ho-

mogenous environments with no refuges available. Specifically, spatial segregation of

15

larvae and adults have been documented in the cannibalistic flour beetles Tribolium

confusum, T. castaneum, and T. brevicornis in homogeneous habitats of flour. While

cannibalism can be a mechanism that drives spatial segregation of life cycle stages,

other types of density dependent inter-stage interactions can also play a role. For

example, for one species of Tribolium, density related larval inhibition by adults can

enhance spatial patterns, as we will see in section 1.2 and in Chapter 4.

1.2 Life Cycle Stages and Spatial Patterns in Flour Beetles

1.2.1 Tribolium castaneum and Tribolium confusum

The flour beetle (genus Tribolium) progresses through four stages over the course of

its life cycle: egg, larva, pupa, and adult. Spatial patterns have been observed in

several species of flour beetles in which the larval and adult stages are mobile and can

travel great distances. In 1966, Arthur Ghent studied the depth distribution of the life

cycle stages of the cannibalistic species Tribolium castaneum and Tribolium confusum

in vials of flour [14]. For each species, he placed 200 adults on the surface of a

cylindrical fractionable-shell vial filled with flour. The vial was divided into five rings

each 10 mm high. 8 grams of flour filled this vial to 41 − 42 mm of flour so each of

the bottom four rings were filled and the last contained the surface layer made up

of the uppermost 1 to 2 mm of flour. Ghent placed this vial in an incubator for 26

days and counted the number of individuals in each stage in each of the four quarters

of flour and on the surface. For the two species of flour beetles studied, 26 days is

enough time to produce eggs, larvae, and pupae, but not new adults. Ghent found

adults of both species to be most abundant near the surface of the flour, while larval

densities peaked slightly below the surface in the upper quarter of flour. Figure 1.1

16

shows the observed depth distributions for various life stages of Tribolium castaneum

and Tribolium confusum. We note that surface patterns involving the segregation

of juveniles and adults have not been observed in T. castaneum or T. confusum,

although Neyman, Park and Scott [37] documented a general trend of increasing

density of adults towards the edges of a square container (and along edges towards

the corners).

1.2.2 Tribolium brevicornis

Further evidence of the spatial segregation of certain life cycle stages has been ob-

served in cultures of the cannibalistic species Tribolium brevicornis in biologist R.

F. Costantino’s lab at the University of Arizona. On the surface of a container of

flour, large numbers of adult beetles often aggregate together rather than spreading

out. Frequently they accumulate along one edge or corner of the box (shown in Fig-

ure 1.2a), but adults have also been observed clumping away from the boundaries

(Figure 1.2b). This pattern has been observed on the surface of round bottles as

well as rectangular boxes (Figure 1.2c). Segregation of life cycle stages on the sur-

face of a container of flour is the signature pattern for T. brevicornis, and has not

been observed for any of the other species in the genus Tribolium. This species has

several major biological differences from the stored-grain pests T. castaneum and

T. confusum, who typically form no discernible pattern on the surface of a tray of

flour (Figure 1.3). The most striking difference is that T. brevicornis is an inhibiting

species, meaning population numbers are regulated by delaying pupal metamorphosis

when adult densities are too great [21]. Inhibition and other biological traits of T.

brevicornis are discussed further in Chapter 4. The patterns shown in Figure 1.2 all

illustrate the non-uniform spatial distribution of adults as well as the segregation of

17

adults and other life stages. However, the specific location of the adults varies greatly

among cultures, even among containers of the same size and shape.

In one container, Costantino noticed that patterns were more reproducible than in

the containers shown in Figure 1.2. He used a box similar in size to the one in Figure

1.2b, but that could be subdivided into smaller rectangles by inserting removable

panels. Figure 1.4 shows three replicates of cultures in domains with about two-

thirds the length as those in Figure 1.2b but only about a quarter of the width. We

see the formation of a pupal “nest” on the left side of each of the three rows. Figure

1.5 provides evidence that this pupal nest persists over time. Figure 1.5b was taken

6 weeks after Figure 1.5a, and new callow (light brown) adults can be seen emerging

from the pupal nest, indicating large larvae return there to pupate. The pupal nest

is not seen in all cultures of T. brevicornis. Rather, the pattern formed depends on

the initial condition of the culture and whether the pupal nest has had a chance to

establish itself before widespread dispersal takes place. Figure 1.6 shows a culture of

T. brevicornis in which a pupal nest was never established and exhibits no surface

pattern.

18

0 1/4 1/2 3/4 10

20

40

60

80

100

120

Space

AdultsLarvaePrepupaePupae

(a) Depth distribution of larvae, pre-pupae, pupae and adults in Tribolium

castaneum.

0 1/4 1/2 3/4 10

10

20

30

40

50

60

70

80

90

Space

AdultsLarvae

(b) Depth distribution of larvae and adults in Tribolium confusum.

Figure 1.1: Data from Ghent [14]. 0 represents the surface of the vial and 1 representsthe bottom.

19

(a) Aggregation of Tribolium brevicornis

adults on surface of rectangular boxes.(b) Aggregation of Tribolium brevicornis

adults on surface of rectangular boxes,away from edges.

(c) Aggregation of Tribolium brevicornis

adults on surface of round bottle.

Figure 1.2: Cultures of T. brevicornis showing segregation of adults and other lifestages. Photo 1.2b by R. F. Costantino.

20

Figure 1.3: Aerial view of Tribolium castaneum adults on surface of rectangular boxof flour. There are some adults near the edges of the box but no distinct pattern orsegregation of life cycle stages.

Figure 1.4: Aerial view of Tribolium brevicornis. Each of three rows show segregationof adults and pupae, illustrating the pupal nest on the left. Each row was startedwith large larvae and adults on the left side. They were restrained in this subhabitatfor 6 weeks, then a panel was removed and they were allowed to disperse throughoutthe entire row. This photo was taken a week after the door was opened. Photo by R.F. Costantino.

21

(a)

(b) Six weeks after (a)

Figure 1.5: New callow adults emerging from nest in (b) indicate large larvae returnto nest to pupate. Photos are 6 weeks apart. Photos by R. F. Costantino.

Figure 1.6: Aerial view of Tribolium brevicornis. Doors are open at ends of rowsto allow beetles to move throughout entire box. Culture started with 100 adults atupper right corner, immediately permitted to disperse. No pupal nest is established.

22

1.3 Density Dependent Dispersal

Many stage-structured species have density dependent dispersal mechanisms. For

example, aphid larvae can develop into one of two adult morphs - winged or wingless.

The ability to fly aids in dispersal of the population. Studies have shown that the

proportion of adults having the winged morph is density dependent, depending on

the number of tactile encounters larvae have with other aphids [17]. Density may

also affect the proportion of winged adults with the ability to fly. Another example

of a polymorphism affecting dispersal ability is wing length in the brown planthopper

Nilaparvata lugens [24]. Nymphs developing under crowded conditions leads to a

greater fraction of long-winged adults.

The above examples provide evidence that density can affect the proportion of

a population able to disperse and the degree of dispersal ability. In sections 1.1

and 1.2 we noted some experimental evidence for the density dependent avoidance

of cannibals by the cannibalized stages in stage-structured species, showing density

can affect the locations to which individuals disperse. In this dissertation I will use

certain types of models for the dynamics of structured populations to explore the role

of density dependent dispersal as a mechanism for the spatial segregation of life cycle

stages. I will focus on observed spatial patterns in species of Tribolium for which we

believe density dependent dispersal is the main causal mechanism.

There is substantial evidence that density plays a role in the dispersal of both

T. castaneum and T. confusum [33, 48]. Naylor documented an inter-stage density

dependent avoidance response in T. confusum [34]. When presented with a choice

between unoccupied flour and flour occupied by medium-sized larvae, adult females

tended to choose the unoccupied flour. Furthermore, the number of adults found

23

in vials of flour decreased with increasing larval density. Naylor also observed larvae

leaving crowded vials of flour with high larval densities. The tendency of T. confusum

to move to locations of unoccupied flour over occupied flour was also observed by

Naylor [32].

Further evidence of the avoidance of larvae by adults comes from an experiment

conducted by Prus [38] to look at emigration ability and surface numbers of adult

beetles of different strains and sexes. All cultures were composed of a single sex

of adults, except for one replicate where several males accidentally were mixed in a

culture of females. This resulted in the appearance of larvae, and a notable increase in

the surface numbers and emigration ability of this replicate [38]. Later investigations

confirmed the effect of the presence of larvae in increasing adult surface numbers. In

adult only cultures, the intensity of emigration was found to depend on the relation

of the current density of beetles to the maximum possible density [49].

T. confusum and T. castaneum are major pests of stored flour and thus of great

economic importance. As a result, they have been extensively studied for over a

century. The species T. brevicornis is considered a minor pest. These beetles have

been found in the western United States, in nests of bees and decaying logs [42].

There is significantly less literature and laboratory work on T. brevicornis than on

T. castaneum and T. confusum, but some key characteristics have been documented

[21, 42] and will be discussed further in Chapter 4.

1.4 Spatial Models and Prior Investigations

Population dynamics of T. castaneum and T. confusum are very well described by the

three-dimensional stage-structured difference equation model known as the Larva −

24

Pupa − Adult, or LPA model (model (4.1)). The LPA model is one of the most

successful models in mathematical biology in the sense that it has been parameterized

and well-validated with laboratory data, and laboratory experiments have verified its

predicted dynamics, attractors and bifurcations (including a route to chaos) [10, 11].

Many modifications of the LPA model have also been successful over the last decade

in investigations of numerous other phenomena, such as competition between flour

beetle species [13] and population dynamics in a periodically fluctuating habitat [3].

A spatial extension of the LPA model is logical since adult beetles and larvae are very

mobile. For example, the studies of Prus [38] confirmed that both sexes of multiple

strains of T. castaneum and T. confusum are able and willing to emigrate from their

container given the slightest opportunity.

To take full advantage of the success of the LPA model, we base our spatial models

for structured populations on discrete time matrix models. We consider habitats

that are continuous in space, so the appropriate class of models are integrodifference

equations (first introduced in population ecology by Kot and Schaffer [26]) with stage-

structure. Since we are interested in the role of density dependent dispersal on spatial

patterns, we need to allow for density dependent dispersal kernels. We also allow for

density to affect the fraction of individuals undergoing dispersal, which results in an

extra non-spatial term being added to the integrodifference equations. This method of

incorporating density dependence into integrodifference equations was used by Dwyer

and Morris [12], who studied the effect of resource density on consumer dispersal and

invasion speeds by looking at traveling wave solutions on infinite spatial domains.

Many previous studies of integrodifference equation models focus on traveling wave

solutions on infinite domains, with applications to the speed of population invasions

[12, 25, 36, 45, 46]. Integrodifference equations on finite domains are considered by

25

Van Kirk and Lewis [44] to investigate when populations can persist in fragmented

habitats. In their study, dispersers may emigrate from the spatial domain, but they

may not return.

For long range dispersers, the form of the dispersal kernel may be derived from

first principles. An assumption of diffusive dispersal with a constant rate a of settling

leads to the Laplace kernel [35, 44]

k(x, y) =a

2exp(−a|x − y|) (1.1)

where the probability of moving to a location x depends on its distance from the

starting location y. In our applications, however, the domain will be small enough

for individuals to traverse the entire habitat in one time step.

Hassel, Comins and May [18] showed spatial segregation of hosts and parasitoids to

be a mechanism that promotes coexistence in uniform patchy environments with local

diffusive, density independent dispersal (results are not seen with global dispersal).

The spatial segregation can occur via spiral waves, chaotic spatial patterns, or a

temporally static “crystal lattice” pattern with patches of varying densities. These

results were later extended to include two competing host species in addition to the

parasitoid with similar results [19]. In order for these self-organizing patterns to

occur, individual patch dynamics must be unstable [2].

Levin [30] considered a system of two competing species and two patches in con-

tinuous time. Population dynamics were such that in any one patch, one species

would exclude the other depending on initial conditions. Allowing a limited amount

of dispersal, or migration, between the patches led to the existence of a coexistence

equilibrium with one species dominant in each patch. Density dependence was not a

26

factor in dispersal rates.

Shigesada et al. [41] also showed that (under certain conditions) spatial segre-

gation can allow for coexistence of two competing species in heterogeneous habitats,

even when both species had the same environmental preferences. Partial differential

equations were used with density dependent (cross-diffusion) dispersal. The system

considered by Shigesada et al. does not exhibit any diffusion-induced Turing insta-

bilities in the absence of cross-diffusion.

The systems we consider here are different from these previous studies in several

ways. First of all, we consider multiple stages of a single species (rather than hosts

and parasitoids or predators and prey, which involve multiple species). In our mod-

els, single individuals can transit between classes. (In contrast, a parasitoid never

becomes a host and a predator never becomes its prey.) Second, we focus on den-

sity dependent dispersal and spatially segregated stages rather than invasion speeds.

Finally, we consider finite spatial domains (heterogeneous and homogeneous) as op-

posed to infinite domains. Our boundary conditions will differ from those of Van Kirk

and Lewis [44] in that dispersers may not leave the spatial domain.

In his doctoral dissertation, M. Alzoubi [1] analyzed general stage-structured in-

tegrodifference equation models with density independent kernels. His analysis built

upon that of Van Kirk and Lewis [44], Hardin, Takac and Webb [16], and Kot and

Schaffer [26] by adding stage-structure. In Chapter 2 we extend and generalize the

theoretical results of Alzoubi so as to incorporate density dependent dispersal and

partial dispersal. In that chapter the focus is on a basic bifurcation theory for struc-

tured integrodifference equations that addresses the basic question of extinction ver-

sus (equilibrium) persistence. This theory generalizes the theory of Cushing [6] for

structured population dynamics in a non-spatial setting. In Chapter 3 we use the

27

theory (along with simulations) to examine some simple toy models that illustrate

mechanisms that can result in spatially segregated life cycle stages. In Chapter 4 we

study more complicated models designed to explain observed spatial patterns among

life cycle stages in Tribolium, as discussed above.

28

Chapter 2

A Bifurcation Analysis of Stage-Structured

Density-Dependent Integrodifference

Equations

In this chapter, we examine the existence and stability of the extinction equilibrium

and of positive equilibria for stage-structured density dependent integrodifference

equation (IDE) models. We do this using bifurcation theory based on the inherent

net reproductive number n. We relate the stability of positive equilibria near the

extinction state to the direction of bifurcation at the critical value n = 1. This work

extends and generalizes that of M. Alzoubi in his doctoral dissertation, A Dispersal

Model for Structured Populations [1]. Alzoubi looked at stage-structured integrodif-

ference equations, which are used to model the dynamics of structured populations

that have distinct reproduction and dispersal stages (of one or more classes). He

extended the modeling methodology of Cushing [6] to include a spatial component.

Here we improve upon Alzoubi’s results by extending the stage-structured integrod-

ifference equation models to allow dispersal to depend on the density of one or more

life cycle stages.

In our models, density may affect dispersal in two different ways: whether or

not an organism disperses, or given it does disperse, density may affect the distance

dispersed. The latter results in a density dependent kernel in the IDE model, while the

former results in an added non-integral term to the model (combining the modeling

29

methodologies of Cushing [6] and Alzoubi [1]).

2.1 Definitions and Preliminaries

In this section we gather together some preliminaries. The following definitions are

from Zeidler [47].

Definition 1. : Let E1 and E2 be Banach spaces, and F : D(F ) ⊆ E1 → E2 an

operator. F is called compact if and only if:

(i) F is continuous

(ii) F maps bounded sets into relatively compact (precompact) sets.

Definition 2. : Let E be a Banach space and let K be a subset of E. Then K is

called an order cone iff:

(i) K is closed, nonempty, and K 6= 0

(ii) a, b ∈ R, a, b ≥ 0, x, y ∈ K ⇒ ax + by ∈ K

(iii) x ∈ K and −x ∈ K ⇒ x = 0

We say x ≤ y iff y − x ∈ K, x < y iff x ≤ y and x 6= y, and x ≪ y iff y − x ∈ int(K).

An ordered Banach space refers to a Banach space together with an order cone.

Consider the operator equation where E is a Banach space:

F (µ, x) = 0, µ ∈ R, x ∈ E (2.1)

Definition 3. : The point (µ0, x0) is called a bifurcation point of equation (2.1) if

(i) F (µ0, x0) = 0

(ii) for i = 1, 2, . . . there are two sequences, (µi, xi) and (µi, yi), of solutions of

30

equation (2.1) which converge to (µ0, x0) as i → ∞. These are distinct sequences

with xi 6= yi for all i.

The equilibrium equations of the models we are interested in will take the general

form

x = A(λ, x) (2.2)

where λ ∈ R, E is a real Banach space with norm || · || and A : R×E → E is compact

and continuous. Let S be the closure of the set of nontrivial solution pairs (λ, x) of

equation (2.2). Assume G ⊆ E is an open subset that contains the closure of the

positive cone K. Note that 0 ∈ K. For each λ, A(λ, ·) is compact and continuous on

G. We assume the operator A can be written as

A(λ, x) = λLx + H(λ, x) (2.3)

where H(λ, x) is o(||x||) for x near 0 uniformly on bounded λ intervals and L is a

compact linear operator on E. We note that A(λ, 0) = 0 for all λ ∈ R. Solutions of

equation (2.2) of the form (λ, 0), λ ∈ R are called trivial solutions. We denote the set

of all reciprocals of the real nonzero eigenvalues of L by r(L) = µ ∈ R|ν = µLν

where ν ∈ E \ 0. If µ ∈ r(L), we call µ a characteristic value of L. All potential

bifurcation points from the trivial curve of solutions of x = A(λ, x) must be from the

set (µ, 0) : µ ∈ r(L). If µ is a characteristic value of odd (geometric) multiplicity,

then (µ, 0) is a bifurcation point [27].

Theorem 1. (Rabinowitz [39]) If µ is a characteristic value of odd (geometric) mul-

tiplicity, then S has a (maximal) subcontinuum Cλ ⊂ R × G such that (µ, 0) ∈ Cλ

and Cλ either

31

(i) meets ∂(R × G) or

(ii) meets (µ, 0), where µ 6= µ ∈ r(L).

In case (i), meeting ∂G includes the case of meeting infinity in R× E.

In many applications, the characteristic values of the linear operator L will be

simple (of multiplicity one). If µ is a simple characteristic value, and A(λ, x) is

Frechet differentiable in x near (µ, 0), then Cλ can be written as (λ(ǫ), x(ǫ)) = (µ +

o(1), ǫν + o(|ǫ|)) for ǫ ≈ 0, where ν ∈ K is an eigenvector corresponding to µ.

2.2 Model Development and Existence of Equilibria

Let Ω ⊆ Rn be a compact subset of Rn representing the spatial habitat of a species

that can be divided into distinct categories, or classes. These classes may be age

groups, size categories or different life-stages. We assume that population dynamics

(reproduction and class transitions) occur first, followed by dispersal. Dispersing

individuals cannot leave Ω. Let xi(t, s), i = 1, 2, . . . , m represent the density of

individuals at the location s ∈ Ω who are in the ith class at time t (unit of time

equal to dispersal period) and let ~x(t, s) = (x1(t, s), · · · , xm(t, s))T . Let tij(~x(t, ·), v)

be the expected fraction of individuals in class j at spatial position v who survive

and transfer to class i in one unit of time. This notation indicates (as it similarly

does in subsequent occurrences) that tij is a functional acting on ~x(t, s) as a function

of s. Surviving individuals might also disperse, and we let kij(s, v, ~x(t, ·)) denote the

dispersal kernel, or the fraction of individuals at position v at time t that settle at

position s by the end of the dispersal period. As indicated, these quantities may

depend on the density of any or all classes at any or all spatial locations.

Let fij(~x(t, ·), v) be the expected number of surviving i-class offspring at position

32

v per j-class individual per unit of time. Let the dispersal kernel lij(s, v, ~x(t, ·)) denote

the fraction of i-class offspring of a j-class individual at position v settling at position

s after one time unit. The total number of i-class individuals at position s at time

t + 1 is

xi(t + 1, s) =m

j=1

Ω

kij(s, v, ~x(t, ·))tij(~x(t, v), v)xj(t, v)dv

+

m∑

j=1

Ω

lij(s, v, ~x(t, ·))fij(~x(t, v), v)xj(t, v)dv

(2.4)

or

xi(t + 1, s) =

Ω

m∑

j=1

[kij(s, v, ~x(t, ·))tij(~x(t, v), v)

+ lij(s, v, ~x(t, ·))fij(~x(t, v), v)]xj(t, v)dv.

(2.5)

To be more general, we can also consider the case where only a fraction of the popula-

tion disperses at any given time. This fraction may be spatially or density dependent,

and we denote it by γij(~x(t, ·), v). The number of i-class individuals at spatial location

s at time t + 1 is now

xi(t + 1, s) =

Ω

m∑

j=1

[kij(s, v, ~x(t, ·))tij(~x(t, v), v)

+ lij(s, v, ~x(t, ·))fij(~x(t, v), v)]γij(~x(t, ·), v)xj(t, v)dv

+

m∑

j=1

(1 − γij(~x(t, ·), s))[tij(~x(t, s), s)xj(t, s) + fij(~x(t, s), s)xj(t, s)].

Using the m × m matrices T = (kijtijγij), F = (lijfijγij), T ∗ = (tij(1 − γij)) and

33

F ∗ = (fij(1 − γij)), we can write the above equation in matrix form:

~x(t + 1, s) =

Ω

[T (s, v, ~x(t, ·)) + F (s, v, ~x(t, ·))]~x(t, v)dv

+ [T ∗(s, ~x(t, ·)) + F ∗(s, ~x(t, ·))]~x(t, s).

(2.6)

The equilibrium equations are then given by

~x(s) =

Ω

[T (s, v, ~x(·)) + F (s, v, ~x(·))]~x(v)dv + [T ∗(s, ~x(·)) + F ∗(s, ~x(·))]~x(s). (2.7)

We are interested only in biologically relevant solutions of the equilibrium equation

(2.7). We want to look at solutions in the closure K+ of a cone K+ of positive valued

functions from a Banach space E of functions defined on Ω (such as C(Ω)). We will

assume the domain of the operator defined by the right hand side of the equilibrium

equation above is an open set G ⊆ E containing K+. In certain applications it is

appropriate to impose boundary conditions on our problem. Rather than working in

a restrictive subspace of E, we will carefully choose our kernel in these applications

so as to hold the boundary conditions invariant.

We denote the integral operators with kernels T (s, v, ~x(·)) and F (s, v, ~x(·)) by T

and F respectively. We denote the last two operators on the right side of equation

(2.7) by T ∗ and F ∗, respectively.

Assumption 1. Let E be an ordered Banach space that has a (positive) cone K+

and G be an open set G ⊆ E containing the closure K+. The operators T, F, T ∗, F ∗ :

G → G are continuous and Frechet differentiable.

34

Our goal is to write the equilibrium equations

~x = (T + T ∗)~x + (F + F ∗)~x (2.8)

in the form studied by Rabinowitz [39]. Expand T , F , T ∗ and F ∗ around ~x = ~0 and

rewrite the equilibrium equation as follows:

~x(s) − T ∗(~0)~x(s) −∫

Ω

T (s, v,~0)~x(v)dv =

Ω

F (s, v,~0)~x(v)dv + F ∗(~0)~x(s) + h(~x(s))

(2.9)

where h(~x) ≡ o(‖~x‖) for ~x ≈ ~0.

Define F (~y) =∫

ΩF (s, v,~0)~y(v)dv and T (~y) =

ΩT (s, v,~0)~y(v)dv. We make the

following assumptions:

Assumption 2. The operator (I − T ∗(~0) − T )−1 exists and is continuous on E.

Assumption 3. The operator (I − T ∗(~0) − T )−1(F + F ∗(0)) has a simple, positive,

strictly dominant eigenvalue n with a positive eigenvector ν ∈ K+. Furthermore, no

other eigenvalue corresponds to a nonnegative eigenvector ν ∈ K+.

Note 1. Following [1, 6] we called n the inherent net reproductive number. It is also

commonly denoted by R0.

Note 2. The Krein-Rutman Theorem (see appendix) can be invoked to verify As-

sumption 3 under certain conditions, namely when we have a strongly positive opera-

tor. An operator A is strongly positive if its kernel is of positive type, or if there exists

an integer n such that An maps any vector in the cone into the interior of the cone

[27]. It is not true in general that the linear operator L in our applications is strongly

positive. In fact, L cannot be strongly positive if we work on a space such as L2(Ω)

35

where the positive cone does not have a nonempty interior. However, Assumption 3 is

still often true in applications and can be shown directly in certain cases, eliminating

the need to use sufficiency theorems such as the Krein-Rutman Theorem. Examples

are given in section 2.4.

Note 3. Assumptions 2 and 3 are generalizations of those used in [6] for nonspatial

models.

Following [6], we choose n as our bifurcation parameter. To do this we use n to

normalize the fij . Let fij = nφij , so F = nΦ and F ∗ = nΦ∗. With this normalization

(I − T ∗(~0) − T )−1(Φ + Φ∗(~0)) has a dominant eigenvalue equal to one.

The equilibrium equations (2.9) can be written as

~x(s)−T ∗(~0)~x(s)−∫

Ω

T (s, v,~0)~x(v)dv = n

Ω

Φ(s, v,~0)~x(v)dv+nΦ∗(~0)~x(s)+h(n, ~x(s))

(2.10)

where h(n, ~x) ≡ o(‖~x‖) for ~x ≈ ~0 uniformly on bounded n intervals, or by Assumption

2 as

~x = nL~x + H(n, ~x) (2.11)

where

L~x = (I − T ∗(~0) − T )−1

Ω

Φ(~0)~x(u)du + (I − T ∗(~0) − T )−1Φ∗(~0)~x

and

H(n, ~x) = (I − T ∗(~0) − T )−1h(n, ~x).

Equation (2.11) has the form of the nonlinear eigenvalue problem studied by Rabi-

nowitz [39]. In order to apply Rabinowitz theory, we need the operator on the right

36

hand side of equation (2.11) to be compact and continuous on G. Sufficient for this

is the following:

Assumption 4. The operators T , F , T ∗ and F ∗ are compact on G.

We give theorems from Krasnoselsk’ii [27] in the appendix that may be used to

verify Assumptions 1 and 4 in applications when working in certain Banach spaces

(such as L2(Ω) or C(Ω)). For example, if all the terms tij, fij , etc. are continuous

functions of their arguments, then these assumptions hold on the Banach space E =

C(Ω). By Assumption 1, (I − T ∗(~0) − T )−1 is continuous, and it follows that L and

H are completely continuous (i.e. continuous and compact). Moreover, H(n, ~x) is

o(||~x||) near 0 uniformly on bounded n intervals. In summary, under Assumptions 1

- 4, equation (2.11) satisfies the conditions needed to apply the bifurcation theory of

Rabinowitz. The following theorem is a generalization to (2.11) of Alzoubi [1].

Theorem 2. Consider equation (2.11) under Assumptions 1 − 4. There exists a

continuum C+ of solution pairs (n, ~x) such that (1,~0) ∈ C+ and one of the following

alternatives holds:

1. C+ is unbounded in R × K+ (and thus contains only positive solutions ~x ∈ K+.

2. C+ contains a non-extinction solution (n∗, ~x∗) ∈ R × ∂K+, ~x∗ 6= ~0.

Proof. We have shown that equation (2.10) satisfies the conditions of Theorem 1

(Rabinowitz [39]) which guarantees the existence of the bifurcating continuum C+.

Suppose alternative 2 does not occur. Then C+ lies in R×K+ and, according to the

alternatives of Theorem 1, is either unbounded in R×K+ or connects to (i.e. contains

in its closure) a point (n,~0) where n 6= 1 is a characteristic value of L. But the latter

case implies n has an associated eigenvector ν ∈ K+ which contradicts Assumption

3. Thus, alternative 1 holds if alternative 2 does not hold.

37

In applications, one can often rule out alternative 2 of Theorem 2. For example if

we are working in a space with a solid positive cone, and can show L is a strongly

positive operator (maps vectors on the closure of the cone to the interior of the cone),

then there cannot be an equilibrium ~x 6= ~0 on the boundary of the positive cone,

∂K+.

2.3 Equilibrium Stability

2.3.1 Extinction Equilibrium

We have been looking at the integrodifference equation (2.6)

~x(t + 1, s) =

Ω

[T (s, v, ~x(t, ·)) + nΦ(s, v, ~x(t, ·)]~x(t, v)dv

+ [T ∗(s, ~x(t, s)) + nΦ∗(s, ~x(t, s))]~x(t, s).

(2.12)

The equilibrium equation is

~x(s) = A(~x(s), n)

where

A(~x(s), n) =

Ω

[T (s, v, ~x(·)) + nΦ(s, v, ~x(·)]~x(v)dv

+ [T ∗(s, ~x(s)) + nΦ∗(s, ~x(s))]~x(s).

(2.13)

~x = ~0 is a solution for all n. In this section we investigate the stability of ~x = ~0.

The Frechet derivative of A at ~x = ~0 has the following matrix form:

A′(~0, n)(~h)(s) =

Ω

(T (s, v,~0) + nΦ(s, v,~0))~h(v)dv + [T ∗(~0) + nΦ∗(~0)]~h(s). (2.14)

38

Assumption 5. The linear operator A′(~0, n) has a simple, positive, strictly dominant

eigenvalue λ0 associated with a unique and positive eigenvector ϕ0 ∈ G.

In certain applications, the Krein-Rutman theorem may be used to prove the

existence of λ0.

The extinction equilibrium is locally asymptotically stable if the dominant eigen-

value of A′(~0, n), λ0 < 1, and unstable if λ0 > 1 (this is the familiar linearization

principle; see [15]). We proved in the previous section that a continuous branch of

positive equilibria bifurcates from the extinction equilibrium at n = 1. In the ap-

pendix, we relate the inherent net reproductive number n to the dominant eigenvalue

λ0 of A′(~0, n) (Theorem 6). That theorem asserts n > 1 (n < 1) if and only if λ0 > 1

(λ0 < 1). So n = 1 if and only if λ0 = 1. From this relationship between n and λ0,

we obtain the following theorem.

Theorem 3. Under Assumptions 1-5, the extinction state is stable for n < 1 and

unstable for n > 1.

2.3.2 Stability and Direction of Bifurcation

In this section we examine the stability of the branch of positive equilibria whose

existence was proved in Theorem 2. Finding a formula for these positive equilibria

is in general impossible. However, it is usually possible to relate stability to the

direction of bifurcation, at least near the bifurcation point. That relationship follows

from the exchange of stability principle for transcritical bifurcations [23].

The following formula provides the relationship between the stability of equilibria

and direction of bifurcation near the bifurcation point (~x, n) = (~0, 1), where the

branch of nontrivial equilibria is parameterized by (x(ǫ), n(ǫ)), for |ǫ| small. Let λ(ǫ)

39

denote the dominant eigenvalue of the linearization at the nontrivial equilibrium.

Then the eigenvalue perturbation along the branch of equilibria is given by λ(ǫ), and

λ′(n0)n(0) = − ˙λ(0) (2.15)

where ′ = ddn

and · = ddǫ

(see [23], p. 27). In equation (2.15), n(0) determines the

direction of bifurcation. If n(0) > 0, then the bifurcation parameter n increases as ǫ

increases and we say the bifurcation is to the right (also called supercritical or forward

bifurcation). If n(0) < 0, n decreases as ǫ increases and the bifurcation is to the left

(also called subcritical or backward bifurcation). The sign of λ′(n0) tells us whether

the dominant eigenvalue λ0 increases or decreases through 1 as n increases through

the bifurcation point n = 1. The signs of these two quantities determines the sign of

˙λ(0) in (2.15).

The linearization of the system at the positive equilibrium (~x(ǫ), n(ǫ)) gives the

positive operator A′(~x(ǫ), n(ǫ))(h(v)) where A′(~x(ǫ), n(ǫ)) is the Frechet derivative of

the nonlinear operator A. λ(ǫ) is the dominant eigenvalue of A′(~x(ǫ), n(ǫ)) along the

bifurcating branch of positive equilibria. The sign of˙λ(0) tells us if the branch of

equilibria gains or loses stability as ǫ increases through 0. Since the extinction state

loses stability as n increases through 1, we get the following result from equation

(2.15) and the exchange of stability principle (see [23], p. 29).

Theorem 4. Assume Assumptions 1-5 hold. If the bifurcation is to the right (n(0) >

0), then the branch of positive equilibria are stable near the bifurcation point (˙λ(0) <

0). If the bifurcation is to the left (n(0) < 0), then the branch of positive equilibria

are unstable (˙λ(0) > 0).

40

2.4 Examples of Bifurcation Theory

In this section we look at three example integrodifference equation models with dif-

ferent dispersal kernels and apply the theory developed in this chapter. We consider

populations whose life cycles consist of two stages, and let ~x(t, s) = col(x1(s), x2(s)).

These stages can be considered to be juveniles and adults. We assume individuals are

born into stage x1 at a rate dependent on the density of x2 and let f21 = b1+x2

. Thus

b represents the maximum birthrate of the species. Individuals transit from stage x1

to stage x2 at a rate s1 (t12 = s1) and from stage x2 to x2 at a rate s2 (t22 = s2).

For each of the next three examples, we need to show that Assumptions 1-4

needed for Theorem 2 are satisfied. In all cases we use E = C(Ω) where Ω is a

finite interval [0, M ]. All terms fij , tij , etc. are continuous in their arguments. As

a result we need only verify Assumption 3, i.e. we need to show that the operator

(I − T ∗(0) − T )−1(F + F ∗(0)) has a simple, positive, strictly dominant eigenvalue

n with a positive eigenvector and no other eigenvalue corresponds to a nonnegative

eigenvector. In order to apply Theorems 3 and 4, we must also show that Assumption

5 is satisfied for each example, i.e. that the operator T+F+T ∗(0)+F ∗(0) has a simple,

positive, strictly dominant eigenvalue λ0 with an associated positive eigenvector.

2.4.1 Example 1: Uniform dispersal

For each stage (in the absence of the other) we take our dispersal kernel to be the

uniform distribution on the one dimensional interval Ω = [0, M ], with the entire

population dispersing every time step (γij = 1, 1 ≤ i, j ≤ 2). In addition, we

assume that the presence of one stage at a given location decreases the fraction of

individuals of the other stage who move to that location at the next time step. Then

41

k21 = k22 = exp(−D2x1(s))1M

and l12 = exp(−D1x2(s))1M

. Under these assumptions

the operators T and F are

T (s, v, ~x(t, ·)) =

0 0

s11M

exp(−D2x1(s)) s21M

exp(−D2x1(s))

, (2.16)

F (s, v, ~x(t, ·)) =

0 b1+x2(v)

1M

exp(−D1x2(s))

0 0

. (2.17)

All entries of T ∗ and F ∗ are zero since all individuals are dispersing at all times. Then

T (s, v,~0) =

0 0

s11M

s21M

(2.18)

and

F (s, v,~0) =

0 b 1M

0 0

(2.19)

We want to find the dominant eigenvalue n of (I − T )−1F . We start by calculating

the resolvent of T .

(I − T ) :

x1(s)

x2(s)

x1(s)

x2(s)

∫ M

0

0 0

s1

Ms2

M

x1(v)

x2(v)

dv.

To calculate the inverse of (I − T ) we need to solve the following for x1(s) and x2(s):

x1(s)

x2(s)

0

s1

M

∫ M

0x1(v)dv + s2

M

∫ M

0x2(v)dv

=

y1(s)

y2(s)

.

42

Clearly x1(s) = y1(s) and we need to solve the following integral equation for x2(s):

x2(s) −s1

M

∫ M

0

x1(v)dv − s2

M

∫ M

0

x2(v)dv = y2(s) (2.20)

Integrating both sides from 0 to M , we have

(1 − s2)

∫ M

0

x2(σ)dσ = s1

∫ M

0

y1(v)dv +

∫ M

0

y2(σ)dσ. (2.21)

Thus∫ M

0

x2(σ)dσ =s1

1 − s2

∫ M

0

y1(v)dv +1

1 − s2

∫ M

0

y2(σ)dσ, (2.22)

which, when substituted into equation (2.20) gives

x2(s) = y2(s) +1

M

[

s1

1 − s2

∫ M

0

y1(v)dv +s2

1 − s2

∫ M

0

y2(σ)dσ

]

.

Thus

(I − T )−1 :

y1(s)

y2(s)

y1(s)

y2(s) + 1M

[

s1

1−s2

∫ M

0y1(v)dv + s2

1−s2

∫ M

0y2(σ)dσ

]

and

F~x :

x1(s)

x2(s)

∫ M

0

0 bM

0 0

x1(v)

x2(v)

dv =

∫ L

0bM

x2(v)dv

0

.

43

Returning to our original problem, we see (I − T )−1F is given by

(I − T )−1F :

x1(s)

x2(s)

∫ M

0b

Mx2(v)dv

1M

bs1

1−s2

∫ M

0

∫ M

0x2(σ)dσdv

=

∫ M

0bM

x2(v)dv

bs1

1−s2

∫ M

0x2(σ)dσ

.

(2.23)

The eigenvalue problem (I − T )−1F x = nx is thus

∫ M

0b

Mx2(v)dv

bs1

1−s2

∫ M

0x2(σ)dσ

= n

x1(v)

x2(v)

.

We see the eigenvector must be a vector of constant functions, since the left hand

side of the system of equations is constant. Since x2 6= 0 (because x2 = 0 ⇒ x1 = 0),

the second equation implies n = bs1

1−s2. An eigenvector is

x1(s)

x2(s)

=

1

s1

1−s2

. (2.24)

The net reproductive number n = bs1

1−s2is a simple, strictly positive eigenvalue with

positive eigenvector (2.40), and Assumption 3 is satisfied.

Straightforward calculations show that the dominant eigenvalue of the linear op-

erator T + F is λ0 = s2

2+ 1

2

s22 + 4bs1, with associated eigenvector

x1(s)

x2(s)

=

1

s2

2b+ 1

2b

s22 + 4bs1

. (2.25)

44

Thus Assumption 5 is satisfied.

From Theorem 3 we know the extinction equilibrium ~x(s) = ~0 will be stable for

bs1

1−s2< 1 and unstable for bs1

1−s2> 1, and Theorem 2 guarantees that a continuum

of non-extinction equilibria bifurcates from the extinction equilibrium at bs1

1−s2= 1.

Theorem 4 tells us that the stability of this branch of equilibria is determined by the

direction of bifurcation.

2.4.2 Example 2: Spatially dependent dispersal

As an example of a kernel with spatial dependence, consider the density independent

kernel k(s, v) = l(s, v) = 12sin s on Ω = [0, π]. Movement of individuals of one

stage to any given location s is completely determined by its position in space in

the absence of the other stage. Starting position v does affect dispersal. Note that

the integral over s is equal to one, which indicates that individuals are not lost or

gained during dispersal, simply redistributed in space. We make the dispersal kernel

density dependent the same way as in Example 1. Let l12 = exp(−D1x2(s))sin s2C1

and k21 = k22 = exp(−D2x1(s))sin s2C2

, where C1 and C2 are normalization constants

to ensure the integral over s remains 1. As in Example 1, we assume the entire

population always disperses (γij = 1, 1 ≤ i, j ≤ 2). Under these assumptions the

operators T and F are

T (s, v, ~x(t, ·)) =

0 0

s1sin s2C2

exp(−D2x1(s)) s2sin s2C2

exp(−D2x1(s))

, (2.26)

F (s, v, ~x(t, ·)) =

0 b1+x2(v)

sin s2C1

exp(−D1x2(s))

0 0

. (2.27)

45

All entries of T ∗ and F ∗ are again zero since all individuals are dispersing at all times.

Then

T (s, v,~0) =

0 0

s1sin s

2s2

sin s2

(2.28)

and

F (s, v,~0) =

0 b sin s2

0 0

(2.29)

We again calculate the resolvent of T .

(I − T ) :

x1(s)

x2(s)

x1(s)

x2(s)

∫ π

0

0 0

s1 sin s2

s2 sin s2

x1(v)

x2(v)

dv.

To do this we solve the following for x1(s) and x2(s):

x1(s)

x2(s)

0

s1 sin s2

∫ π

0x1(v)dv + s2 sin s

2

∫ π

0x2(v)dv

=

y1(s)

y2(s)

.

Clearly x1(s) = y1(s) and we need to solve the following integral equation for x2(s):

x2(s) −s1 sin s

2

∫ π

0

x1(v)dv − s2 sin s

2

∫ π

0

x2(v)dv = y2(s). (2.30)

Integrate both sides from 0 to π. Since∫ π

0sin s

2ds = 1, we have

∫ π

0

x2(s)ds − s2

∫ π

0

x2(v)dv =

∫ π

0

y2(s)ds + s1

∫ π

0

y1(v)dv (2.31)

and∫ π

0

x2(s)ds =1

1 − s2

∫ π

0

y2(s)ds +s1

1 − s2

∫ π

0

y1(v)dv. (2.32)

46

We can now find an expression for x2 in terms of y1 and y2:

x2(s) = y2(s) +sin s

2

s1

1 − s2

∫ π

0

y1(v)dv +sin s

2

s2

1 − s2

∫ π

0

y2(s)ds.

It follows that:

(I − T )−1 :

y1(s)

y2(s)

y1(s)

y2(s) + sin s2

[

s1

1−s2

∫ π

0y1(v)dv + s2

1−s2

∫ π

0y2(s)ds

]

(2.33)

and since F~x =

∫ π

0b sin s

2x2(s)ds

0

we obtain

(I − T )−1F :

x1(s)

x2(s)

b sin s2

∫ π

0x2(v)dv

sin s2

s1

1−s2

∫ π

0b sin v

2

∫ π

0x2(σ)dσdv

(2.34)

Our eigenvalue equation is given by

b sin s2

∫ π

0x2(v)dv

sins2

s1

1−s2b∫ π

0x2(v)dv

= n

x1(s)

x2(s)

The system of equations decouples and we can solve for n from the second equa-

tion. Integrating both sides from 0 to π with respect to s, we see again that

n = bs1

1−s2

∫ π

0sin s

2ds = bs1

1−s2.1 The dominant eigenvalue n is simple and positive. Since

the left side of the eigenvalue equation is a multiple of sin s, the eigenvector must be

1In general, for these population dynamics, any dispersal kernel that is a function of s with∫ π

0k(s)ds = 1 will have this same eigenvalue.

47

a multiple of sin s. An eigenvector is given by

sin(s)

s1

1−s2sin(s)

This eigenvector is nonnegative on the spatial domain [0, π], satisfying Assumption

3.

Straightforward calculations show that the operator T + F has a unique dominant

eigenvalue given by λ0 = s2

2+ 1

2

s22 + 4bs1, with associated eigenvector

x1(s)

x2(s)

=

sin s

( s2

2b+ 1

2b

s22 + 4bs1) sin s

. (2.35)

Thus Assumption 5 is satisfied.

From Theorem 3, the extinction equilibrium ~x(s) = ~0 will be stable for bs1

1−s2<

1 and unstable for bs1

1−s2> 1, and from Theorem 2 we know a continuum of non-

extinction equilibria bifurcates from the extinction equilibrium at bs1

1−s2= 1. Theorem

4 asserts that the stability of this branch of equilibria is determined by the direction

of bifurcation.

2.4.3 Example 3: A more complex spatially dependent kernel

Next we consider a kernel for which dispersal of each stage (in the absence of the

other) depends on both starting location (v) and ending location (s). Specifically, let

k(s, v) =1

2sin s sin v +

1

8sin 2s sin 2v

48

on [0, π]. The coefficients 12

and 18

are chosen so that∫ π

0k(s, v)ds = 1, and k(s, v) ≥ 0

for all s, v ∈ [0, π]. We let the population dynamics be the same as for Examples

1 and 2. Here we assume that the first stage, x1, is sedentary and cannot disperse.

We assume x2 is mobile and the entire population disperses at all times. We in-

corporate density dependence to the dispersal kernel for x2, letting k21 = k22 =

1C

exp(−D2x1(s))(12sin s sin v + 1

8sin 2s sin 2v) where C is a normalization constant

(we assume dispersal related mortality is zero). The operator T is

T (s, v, ~x(t, ·)) =

0 0

s1

Ce−D2x1(s)(

1

2sin s sin v

+1

8sin 2s sin 2v)

s2

Ce−D2x1(s)(

1

2sin s sin v

+1

8sin 2s sin 2v)

(2.36)

F is zero since x1 does not disperse. All entries of T ∗ are also zero since all individuals

of x2 disperse, and F ∗ is

0 b1+x2(s)

0 0

(2.37)

Then

T (s, v,~0) =

0 0

s1(12sin s sin v + 1

8sin 2s sin 2v) s2(

12sin s sin v + 1

8sin 2s sin 2v)

(2.38)

and

F ∗(s, v,~0) =

0 b

0 0

(2.39)

49

We need to solve the following equation for x1 and x2:

x1(s)

x2(s)

0

sin s

2

∫ π

0

(sin v)(s1x1(v) + s2x2(v))dv

+sin 2s

8

∫ π

0

(sin 2v)(s1x1(v) + s2x2(v))dv

=

y1(s)

y2(s)

The first equation gives x1(s) = y1(s). The second equation gives

x2(s) =2s1

4 − s2πsin s

∫ π

0

(sin v)y1(v)dv +2s1

4 − s2πsin 2s

∫ π

0

(sin 2v)y1(v)dv + f(y2)

where f(0) = 0. Since newborns do not disperse, and

(I − T )−1F ∗(~0)~x = (I − T )−1

bx2(s)

0

,

we have

(I − T )−1F ∗(~0)~x =

bx2(s)

2s1

4−s2πsin s

∫ π

0(sin v)x1(v)dv + 2s1

4−s2πsin 2s

∫ π

0(sin 2v)x1(v)dv

Our eigenvalue equation is

bx2(s)

2s1

4−s2πsin s

∫ π

0(sin v)x1(v)dv + 2s1

4−s2πsin 2s

∫ π

0(sin 2v)x1(v)dv

= n

x1(s)

x2(s)

.

The second component of the eigenvector x2(s) must have the form x2(s) = c1 sin s+

c2 sin 2s with c1, c2 ∈ R. Substituting this back into the eigenvalue equation, we see

either c1 or c2 must be 0. If c1 = 0, then n = bs1π16−s2π

and the eigenvector is a multiple

50

of sin 2s. An eigenvector is given by

sin 2s

s1π16−s2π

sin 2s

.

If c2 = 0, n = bs1π4−s2π

and the eigenvector is a multiple of sin s. An eigenvector is given

by

sin s

s1π4−s2π

sin s

.

The dominant eigenvalue here is n = bs1π4−s2π

. This eigenvalue is also simple and

positive. The eigenvector corresponding to this dominant eigenvalue is the only non-

negative eigenvector on [0, π], so Assumption 3 is met.

Straightforward calculations show that the dominant eigenvalue of the linear op-

erator T + F ∗ is λ0 = s2π8

+ 18

(s2π)2 + 16bs1π, with associated eigenvector

x1(s)

x2(s)

=

sin s

( s2π8b

+ 18b

(s2π)2 + 16bs1π) sin s

, (2.40)

satisfying Assumption 5.

By Theorem 3, the extinction equilibrium is stable for bs1π4−s2π

< 1 and unstable

for bs1π4−s2π

> 1, and by Theorem 2 a continuum of non-extinction equilibria bifurcates

from the extinction equilibrium at bs1π4−s2π

= 1. Theorem 4 tells us the stability of these

equilibria is determined by the direction of bifurcation.

51

2.5 Appendix

2.5.1 The Krein-Rutman Theorem

Let E be a real Banach space with a total ordered cone K+, and L : E → E be a

linear operator. L is a strongly positive operator if Lx ≫ 0 whenever x > 0 ([47]).

That is, L maps all non-zero x ∈ K+ into the interior of the cone, int(K+). We

note that in order for an operator to be strongly positive, the cone K+ must have a

non-empty interior.

Theorem 5. (Krein-Rutman [47]) Let E be a real Banach space with an order cone

K+ having a non-empty interior. Then a linear, compact, and strongly positive

operator T : E → E has the following properties:

(1) T has exactly one eigenvector with x > 0 and ||x|| = 1. The corresponding

eigenvalue is the spectral radius ρ(T ) and this is algebraically simple. Furthermore,

x ≫ 0.

(2) The dual operator T ∗ has ρ(T ) as an algebraically simple eigenvalue with a strictly

positive eigenvector x∗.

Corollary 1. (Comparison Principle [47]) If S : E → E is a compact linear operator

with Sx ≥ Tx for all x ≥ 0, then ρ(S) ≥ ρ(T ). If Sx > Tx for all x > 0, then

ρ(S) > ρ(T ).

2.5.2 Relationship between n and λ0

For linear operators A, B : E → E on a Banach space E, we have the following

properties from Zeidler [47]:

(a) (AB)∗ = B∗A∗

52

(b) (A + B)∗ = A∗ + B∗

(c) (A−1)∗ = (A∗)−1

Theorem 6. [1, 8, 31] Let E be a real Banach space with an order cone K+ having

a non-empty interior and F, T : E → E are compact linear operators with I − T

invertible. Assume

1. F (I − T )−1 and its dual have a simple, positive, dominant eigenvalue n associated

with positive eigenvectors y > 0 and w∗ > 0.

2. T + cF is strongly positive for all real numbers c > 0.

Then one of the following holds:

1. r = n = 1

2. 1 < r < n

3. 0 < n < r < 1 where r = ρ(T + F ).

Proof. From assumption 1, (F (I − T )−1)∗w∗ = nw∗. Then (I − T ∗)−1F ∗w∗ = nw∗.

Apply I−T ∗ to both sides of this equation and divide by n. Then 1nF ∗w∗ = (I−T ∗)w∗

or (T ∗ + 1nF ∗)w∗ = w∗. This is equivalent to

(T +1

nF )∗w∗ = w∗. (2.41)

We see 1 is the dominant eigenvalue of the strongly positive operator (T + 1nF )∗ and

also T + 1nF . So ρ(F

n+ T ) = 1 We consider three cases.

1. Assume n = 1. Then 1 = ρ(Fn

+ T ) = ρ(F1

+ T ) = ρ(F + T ) = r.

2. Assume n > 1. Then using Corollary 1 we have T + Fn

< F + T < nT + F and

1 = ρ(Fn

+ T ) < ρ(F + T ) = r < ρ(nT + F ) = n. So 1 < r < n.

3. Assume 0 < n < 1. Then using Corollary 1 we have T + Fn

> F + T > nT + F

53

and 1 = ρ(Fn

+ T ) > ρ(F + T ) = r > ρ(nT + F ) = n. So 0 < n < r < 1.

2.5.3 Compactness of T + nΦ

The integral operator∫

ΩT (s, v, ~x(·))~x(v)+nΦ(s, v, ~x(·))~x(v)dv has the form of a non-

linear Urysohn operator. The following theorem of Krasnoselsk’ii [27] gives conditions

under which Urysohn’s operator is completely continuous in the space of continuous

functions E = C(Ω):

Theorem 7. (Krasnoselskii) Let the function K(s, v, ~x) be continuous with respect

to all the variables in the set s, v ∈ Ω, |x| ≤ a. Then the operator∫

ΩK(s, v, ~x)dv is

defined in the ball of radius a in the space C and is completely continuous.

If we want to work in the Banach space E = L2(Ω), we can use the following

theorem to satisfy Assumption 4 of Theorem 2 in an application.

Theorem 8. (Krasnoselskii) Assume

K(s, v, ~x) = T (s, v, ~x(·))~x(v) + nΦ(s, v, ~x(·)~x(v))

is continuous with respect to ~x and satisfies the inequality

|K(s, v, ~x)| ≤ R(s, v)(a + b|~x|α)

where s, v ∈ Ω, with∫

Ω

Ω|R(s, v)|α+1dsdv < ∞, where a, b > 0, α > 0. Then

Urysohn’s operator∫

ΩK(s, v, ~x)dv acts in the space Lα+1 and is a completely contin-

uous operator.

54

To work in L2(Ω), we need α = 1, and thus require a linear bound on |K(s, v, ~x)|

for the operator T + nΦ to be compact and continuous.

55

Chapter 3

Juvenile-Adult (Toy) Models

We hypothesize density dependent dispersal as a mechanism that potentially is re-

sponsible for the segregation of life cycle stages observed in populations of T. cas-

taneum and T. confusum. Before investigating more realistic models for Tribolium,

we study in this section some lower dimensional stage-structured IDE models with

density dependent dispersal kernels on closed domains, to explore if and when density

dependent dispersal can lead to such spatial patterns.

We consider Juvenile-Adult population dynamics models of the general form

Jt+1 = f(Jt, At)

At+1 = g(Jt, At)

(3.1)

Specifically we look at the following model:

Jt+1 =bAt

1 + At

At+1 = (1 − µJ)Jt + (1 − µA)At

(3.2)

We use a rational function, or Beverton-Holt, nonlinearity for density regulation

rather than an exponential, or Ricker, nonlinearity in order to keep population dynam-

ics tame. We want to isolate the effects of density dependent dispersal on the model

behavior. Complex population dynamics would make it more difficult to untangle ef-

fects of density dependent dispersal from non-spatial model behavior. In model (3.2),

56

b is the maximal birth rate, µJ is the juvenile death rate, and µA is the adult death

rate per unit time. The unique positive equilibrium (J∗, A∗) = (b− µA

1−µJ, b(1−µJ )

µA− 1)

is locally asymptotically stable for parameter values satisfying b(1−µJ )µA

> 1.

3.1 Hostile Boundary Conditions

We introduce (density independent) dispersal kernels KJ(x, y) and KA(x, y) that

describe the fraction of juveniles and adults, respectively, at location x at time t who

move to location y by time t + 1. For the first example model we take

KJ(x, y) = KA(x, y) =1

2sin x (3.3)

on the domain [0, π]. These kernels imply hostile boundary conditions, since the

probability of moving to position 0 or π is 0. We also note each stage views the

environment the same way in the absence of the other stage.

We build in density dependent dispersal by assuming that each stage avoids the

other. That is, the presence of adults at a given spatial location decreases the prob-

ability of a juvenile moving there, and visa versa. Specifically, we modify the kernels

as follows:

KJ(x, y, Jt(·), At(·)) =1

2(sin x) exp(−DJ((1 − µJ)Jt(x) + (1 − µA)At(x)))

KA(x, y, Jt(·), At(·)) =1

2(sin x) exp(−DA(

bAt(x)

1 + At(x))).

(3.4)

Thus, the fraction of juveniles dispersing to location x is decreased according to the

density of adults at that location, after population dynamics occur. Similarly, the

fraction of adults moving to a given location decreases as the number of juveniles at

57

that location increases. The coefficients DJ and DA can be thought of as dispersal

sensitivities.

The general theory developed in Chapter 2 allows for density to affect an in-

dividual’s decision to disperse as well as distance dispersed. Here we assume that

all individuals disperse every time step. We note that with kernels (3.4), a chance

of mortality accompanies dispersal. Individuals can be lost, but not gained, during

dispersal.

Our IDE, with population dynamics given by equation (3.2) and the density de-

pendent dispersal kernels (3.4), is

Jt+1(x) =

∫ π

0

1

2(sin x) exp(−DJ((1 − µJ)Jt(x) + (1 − µA)At(x)))

bAt(y)

1 + At(y)dy

At+1(x) =

∫ π

0

1

2(sin x) exp(−DA(

bAt(x)

1 + At(x)))((1 − µJ)Jt(y) + (1 − µA)At(y))dy.

(3.5)

If we set DA = DJ = 0, both dispersal kernels reduce to the density independent

kernels 12sin x. All sets of parameter values for b, µJ and µA result in a temporal

equilibrium with no segregation - the spatial distribution of both J and A mimic the

sine curve, with no individuals at the boundaries and maxima at the midpoint of the

domain, π2. A simulation illustrating the case where DJ = DA = 0, b = 25, and

µJ = µA = 0.2 appears in Figure 3.1.

If we increase DJ from 0 to 0.5, so juveniles “avoid” adults, and keep all other

parameters the same, we still have a temporal equilibrium. However, we now see

spatial segregation of juveniles and adults in the sense that the stages peak in different

spatial locations (see Figure 3.2). Adult numbers are still greatest in the center of

the domain, but juveniles have been “pushed” to the sides of the domain, exhibiting

58

0 0

20

40

60

80

100

120

140

160

Space

AdultsJuveniles

π

Figure 3.1: Time-equilibrium attractor of Juvenile-Adult IDE model (3.5). b = 25,µL = µA = 0.2. There is no density dependence in the dispersal kernels: DJ = DA = 0

a bi-modal distribution with peaks occurring near π4

and 3π4

and a dip between them

at π2.

If we bring DJ back to 0 and instead increase DA to 0.5, so adults “avoid” juveniles

(all other parameters remain the same), we see temporal segregation rather than

spatial segregation. We no longer have an equilibrium in time, but a temporal 2-cycle.

At any given time, one stage dominates at all spatial locations. The other stage is

present, but only in very small numbers. The dominant stage alternates between

juveniles and adults, each peaking in the center of the domain. An illustration of this

appears in Figure 3.3.

Next we look at DJ = DA = 0.5, keeping all population dynamics parameters the

same. The result is a temporal 4-cycle, again showing temporal segregation. Now

juveniles, when dominant, exhibit a bi-modal distribution. This is shown in Figure

3.4.

Figures 3.5-3.7 show the properties of the model attractor for various regions in the

59

0 0

1

2

3

4

5

6

7

8

Space

AdultsJuveniles

π

Figure 3.2: Time-equilibrium attractor for Juvenile-Adult IDE model (3.5), showingspatial segregation of juveniles and adults. b = 25, µL = µA = 0.2. DJ = 0.5,DA = 0.

0 0

5

10

15

20

25

30

35

40

Space

t = n−1

0 0

5

10

15

20

25

30

Space

t = n

AJ

π π

Figure 3.3: 2-cycle attractor of Juvenile-Adult IDE model (3.5), with temporal seg-regation of juveniles and adults. b = 25, µL = µA = 0.2. DJ = 0, DA = 0.5

60

00

1

2

3t = n−3

0 0

5

10t = n−2

0 0

5

10

Space

t = n−1

0 0

2

4

Space

t = n

AJ

π

π

π

π

Figure 3.4: 4-cycle attractor of Juvenile-Adult IDE model (3.5), with temporal seg-regation of juveniles and adults. b = 25, µL = µA = 0.2. DJ = DA = 0.5

DJ , DA parameter plane for increasing values of b (and therefore increasing values of

n). For n ≈ 1 there are no spatial patterns possible for any values of DJ or DA. This

is consistent with the mathematical results obtained earlier in Chapter 2, namely that

near the bifurcation point n = 1, distributions of J and A resemble the eigenvector,

which has sin x components.

As b is increased, more patterns become possible. For b = 2.5, there is a small

region of parameter space where we see spatial segregation of adults and juveniles.

For this to occur, DJ must be sufficiently large and DA must be sufficiently small.

All attractors are temporal equilibria.

Increasing b to 10, we can achieve temporal two-cycles as well as equilibria. If

DA is large enough, we see temporal segregation in the sense defined earlier. If b is

increased to 25, we add temporal 4-cycles to the list of potential attractors.

We have shown in this section (through simulation) that low dimensional stage-

structured IDE models can result in both spatial and temporal segregation of life

61

Figure 3.5: Shown is a section of the DJ − DA parameter plane for b = 2.5. µJ =µA = 0.2. We look at 0 ≤ DJ ≤ 1 and 0 ≤ DA ≤ 1. All combinations of DJ and DA

in this range result in equilibrium dynamics for model (3.5). Spatial segregation ispossible for large enough DJ and small enough DA.

Figure 3.6: Shown is a section of the DJ − DA parameter plane for b = 10. µJ =µA = 0.2. In addition to the model (3.5) attractors seen for b = 2.5, we can also see2 − cycles and temporal segregation.

62

Figure 3.7: Shown is a section of the DJ − DA parameter plane for b = 25. µJ =µA = 0.2. In addition to the model (3.5) attractors seen for b = 10, we can also see4 − cycles and 4 − cycles with temporal segregation.

cycle stages. Specifically, our study of this model suggests the following conclusions.

Increasing DJ (the sensitivity of juveniles to adult densities) tends to increase the

degree of spatial segregation, while increasing DA (the sensitivity of adults to juvenile

densities) tends to increase the degree of temporal segregation between the two life

cycle stages. Increasing the birth rate, or the net reproductive number, leads to larger

DJ − DA parameter regions where Juvenile-Adult spatial segregation occurs, and an

increase in the number of possible types of model behavior.

63

3.1.1 Application of Theory

Here we apply the theory developed in Chapter 2 to the Juvenile-Adult IDE model

(3.5). We let ~x(t, s) = (Jt(s), At(s)). Then

T (s, v, ~x(t, ·)) =

0 0

(1 − µJ) sin s2

exp(−DAJJt(s)) (1 − µA) sin s2

exp(−DAJJt(s))

(3.6)

and

F (s, v, ~x(t, ·)) =

0 b1+At(v)

sin s2

exp(−DJAAt(s))

0 0

. (3.7)

T (s, v,~0) =

0 0

(1 − µJ) sin s2

(1 − µA) sin s2

(3.8)

and

F (s, v,~0) =

0 b sin s2

0 0

(3.9)

so the operators T and F are the same as in Example 2.4.2 with s1 = (1 − µJ) and

s2 = (1 − µA).

The resolvent of T is the same as for Example 2.4.2 (equation (2.33)). The operator

(I − T )−1F is

(I − T )−1F :

J(s)

A(s)

b sin s2

∫ π

0A(v)dv

sin s2

(1−µJ )µA

∫ π

0b sin v

2

∫ π

0A(σ)dσdv

(3.10)

64

The eigenvalue equation is given by

b sin s2

∫ π

0A(v)dv

b sin s2

(1−µJ )µA

∫ π

0A(σ)dσ

= n

J(s)

A(s)

(3.11)

Solving for n, we find the net reproductive number is n = b(1−µJ )µA

with associated

eigenvector

J(s)

A(s)

=

sin s

1−µJ

µAsin s

. (3.12)

We also know from Example 2.4.2 that the operator T + F has a simple, positive,

strictly dominant eigenvalue with a positive associated eigenvector.

From Theorem 3 of Chapter 2, the extinction equilibrium

J(s)

A(s)

= ~0

will be stable for b(1−µJ )µA

< 1 and unstable for b(1−µJ )µA

> 1.

We can show that the bifurcation of non-extinction equilibria from the extinc-

tion equilibrium at n = 1 (guaranteed by Theorem 2) is a forward bifurcation (and

therefore the non-extinction equilibria are stable, by Theorem 4) by showing that

all density effects are deleterious, i.e. 0 ≤ P (s, v, ~x) ≤ P (s, v,~0) ∀s, v ∈ Ω, where

P = T + F and ~x = col(x1, x2) = col(J, A). As we will see, this implies that there

cannot be a stable positive equilibrium for n < 1.

65

For our model,

P (s, v, ~x(t, ·)) =

0 b1+x2(t,v)

sin s2

exp(−D12x2(t, s))

(1 − µ1)sin s2

exp(−DAJJt(s)) (1 − µ2)sin s

2exp(−D21x1(t, s))

(3.13)

where D12 = DJA, D21 = DAJ , µ1 = µJ and µ2 = µA.

P (s, v,~0) =

0 b sin s2

(1 − µ1)sin s

2(1 − µ2)

sin s2

(3.14)

Since 11+x2(t,v)

≤ 1, exp(−D12x2(t, s)) ≤ 1 and exp(−D21x1(t, s)) ≤ 1, it follows

that

0 ≤ P (s, v, ~x(t, ·)) ≤ P (s, v,~0) ∀s, v ∈ Ω.

Let

~y(t + 1, s) =

Ω

0 b sin s2

(1 − µJ) sin s2

(1 − µA) sin s2

~y(t, v)dv, ~y(0, s) = ~x(0, s). (3.15)

Assume ~x(t, s) ≤ ~y(t, s) ∀s ∈ Ω. Then

~x(t + 1, s) =

b sin s2

e−D12x2(s)∫

Ωx2(t,v)

1+x2(t,v)dv

sin s2

e−D21x1(s)∫

Ω(1 − µ1)x1(v) + (1 − µ2)x2(v)dv

b sin s2

Ωy2(t, v)dv

sin s2

Ω(1 − µ1)x1(v) + (1 − µ2)x2(v)dv

= ~y(t + 1, s)

(3.16)

By induction,

~x(t + 1, s) ≤ ~y(t + 1, s) ∀s ∈ Ω, ∀t (3.17)

66

and

|y| → 0 ⇒ |x| → 0. (3.18)

Let λ0 denote the dominant eigenvalue of P (s, v,~0). By Theorem 6, if n < 1 then

λ0 < 1 and |y| → 0 as t → ∞. Thus the extinction equilibrium ~x = ~0 is a global

attractor for n < 1 and the bifurcation cannot be to the left. This result is shown by

Cushing [6] for non-spatial models.

The equilibria of model (3.5) can be approximated by the eigenvector (3.12) for

values of n near 1. Since both components of the eigenvector are multiples of sin s,

there is no spatial segregation of juveniles and adults for n ≈ 1. Figures 3.5-3.7 show

the spatial patterns for model attractors in a region of DJ , DA parameter space for

n = 10,n = 40, and n = 100. The smallest n shown, n = 10 gives rise to spatial

segregation of juveniles and adults in a small region of parameter space. For n ≈ 1,

the only attractor seen is a temporal equilibrium with no spatial segregation. As

n increases, we found, however, that the effects of density dependence can cause

segregation of life cycle stages to occur.

3.2 Mixed Boundary Conditions

The spatial pattern observed by Arthur Ghent in T. castaneum [14] in a cylindrical

vial of flour can be looked at as one-dimensional in space, where the spatial dimension

is depth. Due to gravity, there is a gradient of increasing flour density from the surface

to bottom of the bottle [14]. Individuals can live on the surface but cannot pass above

it. We model this environment as a one-dimensional domain with mixed boundary

conditions, where the bottom of the bottle is a hostile boundary and the surface is a

no-flux boundary.

67

We are interested in the behavior of a Juvenile-Adult IDE model with these bound-

ary conditions. Ghent’s study was done with an organism best modeled with three

life-stages, but here we investigate what qualitative aspects of his observations can

be captured by a two life-stage model.

For this Juvenile-Adult IDE model with mixed boundary conditions, we use the

population dynamics model (3.2) and again build our boundary conditions into the

kernel. The kernels we use are

KJ(x, Jt(x), At(x)) = exp(−DJ ((1 − µJ)Jt(x) + (1 − µA)At(x))) cos x

KA(x, Jt(x), At(x)) = exp(−DA(bAt(x)

1 + At(x))) cosx.

(3.19)

Here KJ gives the fraction of juveniles at position y at time t who move to posi-

tion x by time t + 1. The spatial domain is the interval [0, π2]. The left endpoint,

0 (representing the surface of the bottle), is the no-flux boundary. The right end-

point, π2

(representing the bottom of the bottle), is the hostile boundary. Again this

kernel assumes that juveniles avoid adults and adults avoid juveniles. The juvenile-

adult integrodifference equation model (3.5) with these mixed boundary conditions

becomes:

Jt+1(x) =

∫ π2

0

1

2exp(−DJ((1 − µJ)Jt(x) + (1 − µA)At(x)))(cos x)

bAt(y)

1 + At(y)dy

At+1(x) =

∫ π2

0

1

2exp(−DA(

bAt(x)

1 + At(x)))(cos x)((1 − µJ)Jt(y) + (1 − µA)At(y))dy

(3.20)

Matlab simulations of model (3.20) for reasonable parameters for Tribolium show

model dynamics qualitatively similar to behavior observed by Ghent. For example

the simulation shown in Figure 3.8 shows juveniles moving to avoid high densities

68

of adults, and adult movement uninfluenced by the presence of juveniles. Adult

densities peak near the surface, and juveniles peak below the surface (closer to the

bottom of the bottle). While we don’t expect to be able to compare, accurately

or quantitatively, observations from Tribolium cultures with simulations from simple

Juvenile-Adult toy models, we note that lower dimensional IDE models with density

dependent dispersal can give rise to patterns qualitatively similar to those observed

in Tribolium. In Chapter 4 we will investigate a spatial extension of the biologically

more accurate three-dimensional LPA model and study what spatial patterns are

possible in that model.

69

0 0

1

2

3

4

5

6

7

8

Space

AdultsJuveniles

π

2

Figure 3.8: Shown is a Matlab simulation of model (3.20) with mixed boundaryconditions. The left endpoint of the spatial domain, 0, represents the surface of a vialof flour and is a no-flux boundary. The right endpoint, π

2, represents the bottom of

a vial of flour and is a hostile boundary. Model parameters: b = 25, µJ = µA = 0.2,DJ = 0.5, DA = 0. Adult densities peak at the surface and juvenile densities peakbelow the surface.

3.3 Role of Domain Size

We now consider a more complex kernel with several key differences from (3.1) and

(3.19). This dispersal kernel does not have any predetermined spatial dependence.

Rather, the probability of moving from location y to x is based on the distance

between location x and y and the density of each stage encountered along the way. The

distance a juvenile disperses is based on the total number of adults located between

the juvenile’s starting location and end location weighted by distance from starting

location, and visa versa for adult dispersal. The kernels are given by the following

equations, where the denominators of the kernels are normalization constants, so as

70

to reflect no mortality risk is associated with dispersal:

KJ(x, y, Jt(·), At(·)) =exp(−DJ

∫ x

y(s − y)g(Jt(s), At(s))ds)

∫ 1

0exp(−DJ

∫ z

y(w − y)g(Jt(w), At(w))dw)dz

KA(x, y, Jt(·), At(·)) =exp(−DA

∫ x

y(s − y)f(Jt(s), At(s))ds)

∫ 1

0exp(−DA

∫ z

y(w − y)f(Jt(w), At(w))dw)dz

(3.21)

where f(Jt(s), At(s)), g(Jt(s), At(s)) give the number of juveniles and adults, respec-

tively, at position s after population dynamics processes. The spatial Juvenile-Adult

model with kernel (3.21) is

Jt+1(x) =

∫ 1

0

exp(−∫ x

y(s − y)[DJg(Jt(s), At(s))]ds)

∫ 1

0exp(−

∫ z

y(w − y)[DJg(Jt(w), At(w))]dw)dz

f(Jt(y), At(y))dy

At+1(x) =

∫ 1

0

exp(−∫ x

y(s − y)[DAf(Jt(s), At(s))]ds)

∫ 1

0exp(−

∫ z

y(w − y)[DAf(Jt(w), At(w))]dw)dz

g(Jt(y), At(y))dy.

(3.22)

Consider the population dynamics f(Jt, At), g(Jt, At) given by equation (3.2) with

b = 20 and µJ = µA = 0.3.

We first want to know what the spatial attractor of the model is with density

independent dispersal. We cannot simply set DJ = DA = 0 in kernel (3.21), but

must consider the following kernel:

KJ(x, y, Jt(·), At(·)) =exp(−

∫ x

y(s − y)ds)

∫ 1

0exp(−

∫ z

y(w − y)dw)dz

KA(x, y, Jt(·), At(·)) =exp(−

∫ x

y(s − y)ds)

∫ 1

0exp(−

∫ z

y(w − y)dw)dz

.

(3.23)

With this kernel there is no spatial segregation between the two stages. A simulation

illustrating this appears in Figure 3.9.

71

0 0.2 0.4 0.6 0.8 115

20

25

30

35

40

45

50

Space

AdultsJuveniles

Figure 3.9: Simulation of Juvenile-Adult IDE model (3.22) with population dynamicsgiven by (3.2). b = 20, µJ = µA = 0.3, density independent kernel (3.23). There isno spatial segregation of juveniles and adults.

If we keep the parameter values for b, µL, and µA, but let DJA = DAJ = 1 in

model (3.22) with kernel (3.21), we see spatial segregation between juveniles and

adults. A simulation illustrating this spatial pattern appears in Figure 3.10. The

spatial domain for model (3.22) is set to be [0, 1]. The relative size of the domain

can be increased or decreased by changing the speed of dispersal, controlled by the

parameters DJ and DA. Decreasing these parameters so DJ = DA = 0.1 has the

effect of eliminating spatial segregation (see Figure 3.11), suggesting there may be a

minimum domain size necessary to sustain spatial segregation of life cycle stages.

72

0 0.2 0.4 0.6 0.8 115

20

25

30

35

40

45

50

55

Space

AdultsJuveniles

Figure 3.10: Simulation of Juvenile-Adult IDE model (3.22) with population dynam-ics given by (3.2). b = 20, µJ = µA = 0.3, DJ = DA = 1. Juveniles and adults arespatially segregated.

0 0.2 0.4 0.6 0.8 115

20

25

30

35

40

45

50

Space

AdultsJuveniles

Figure 3.11: Simulation of Juvenile-Adult IDE model (3.22) with population dynam-ics given by (3.2). b = 20, µJ = µA = 0.3, DJ = DA = 0.1. There is no spatialsegregation of juveniles and adults.

73

Chapter 4

Case Studies Using Models for Flour Beetles

(Tribolium)

4.1 Spatial Patterns in Tribolium castaneum

4.1.1 The LPA Model

As discussed in Chapter 1, one of the most heavily analyzed and well-validated mod-

els in mathematical ecology is the “Larva-Pupa-Adult” (LPA) model, a three dimen-

sional system of nonlinear difference equations designed to describe the population

dynamics of Tribolium castaneum and Tribolium confusum [5]:

Lt+1 = bA(t)e−celL(t)−ceaA(t)

Pt+1 = (1 − µL)L(t)

At+1 = Pte−cpaAt + (1 − µA)At

(4.1)

L(t) represents the number of larvae at time t, P (t) gives the number of individuals

in the “P stage” (which includes non-feeding larvae, pupae, and callow adults) at

time t, and A(t) gives the number of sexually mature adults at time t. Animals

transit to the next class in two weeks, and we use this as the time step for the model.

Upon reaching the adult class, animals remain there until death. Larval recruitment

in the absence of egg cannibalism occurs at an inherent rate b. Eggs must survive

74

cannibalism by larvae and adults in order to become larvae, and the exponential

exp(−celL(t) − ceaA(t)) represents the survival per unit time. cel ≥ 0 and cea ≥ 0

are cannibalism coefficients of eggs by larvae and eggs by adults, respectively. It

is assumed that cannibalism occurs as a result of random encounters of larvae and

adults with eggs, which accounts for the exponential nonlinearities used [7]. Larvae

die at a rate µL, 0 < µL < 1, and so a fraction (1 − µL) survive to become pupae.

The natural death rate of pupae is negligible, so there is no µP term included in the

model. Pupae must escape cannibalism by adults (cpa) to become adults. Adults die

at a rate µA, 0 < µA < 1, and so at each census, the fraction of surviving adults is

(1 − µA).

4.1.2 The Spatial LPA Model

Our goal in this section is to construct a spatial extension of the LPA model in

order to explain the spatial patterns in T. castaneum and T. confusum observed by

Ghent [14] and detailed in Chapter 1. In order to construct the spatial LPA model,

we need to make assumptions regarding the dispersal behavior of the two mobile

life-stages, larvae and adults (pupae are immobile). We assume adults prefer areas

with lower larval densities than their current location and larvae prefer areas with

lower adult densities than their current location. Thus the fraction of large larvae

that move from location y to location x depends on the difference in adult densities,

after population dynamics occur, between the two locations. Adult dispersal depends

similarly on density of large larvae. We take our spatial domain Ω to be the finite

interval [0, π2] representing depth in a cylindrical vial, with 0 corresponding to the

surface of a cylindrical and 1 corresponding to the bottom of the vial. We assume

both larvae and adults view the environment in the same way, preferring flour near

75

the surface over the more dense flour near the bottom, as described earlier in section

3.2. We model this with mixed boundary conditions, assuming a no-flux boundary at

the surface and a hostile boundary at the bottom. The kernels for larval and adult

spatial redistribution are given by the following, respectively:

KL =1

C1

e−DLA(e−cpaAt(x)Pt(x)+(1−µA)At(x))−(e−cpaAt(y)Pt(y)+(1−µA)At(y))(cosm x)

KA =1

C2

e−DAL(bAt(x)e−ceaAt(x)−celLt(x))−(bAt(y)e−ceaAt(y)−celLt(y))(cosm x)

(4.2)

where C1 and C2 are the following normalization constants:

C1 =

∫ π2

0

e−DLA(e−cpaAt(x′)Pt(x′)+(1−µA)At(x′))−(e−cpaAt(y)Pt(y)+(1−µA)At(y))(cosm x′)dx′

C2 =

∫ π2

0

e−DAL(bAt(x′)e−ceaAt(x′)−celLt(x

′))−(bAt(y)e−ceaAt(y)−celLt(y))(cosm x′)dx′

(4.3)

The term cosm x in the dispersal kernel incorporates density independent spatial

preference. As the parameter m increases, the undesirable region at the bottom

of the bottle grows in size.

After simplification, the spatial LPA model becomes

Lt+1(x) =

∫ π2

0

1

c1

e−DLA(e−cpaAt(x)Pt(x)+(1−µA)At(x))(cosm x)(bAt(y)e−celLt(y)−ceaAt(y))dy

Pt+1(x) =(1 − µL)Lt(x)

At+1(x) =

∫ π2

0

1

c2e−DAL(bAt(x)e−ceaAt(x)−celLt(x))(cosm x)(e−cpaAt(y)Pt(y)

+ (1 − µA)At(y))dy

(4.4)

76

where

c1 =

∫ π2

0

e−DLA(e−cpaAt(x′)Pt(x′)+(1−µA)At(x′))(cosm x′)dx′

c2 =

∫ π2

0

e−DAL(bAt(x′)e−ceaAt(x′)−celLt(x

′))(cosm x′)dx′.

(4.5)

Our initial conditions are triples (L0(x), P0(x), A0(x)) of continuous functions on [0, π2]

that satisfy (L0(π2), P0(

π2), A0(

π2)) = (0, 0, 0) and (L′

0(0), P ′0(0), A′

0(0)) = (0, 0, 0). It

is a feature of the kernels that the right sides of model (4.4) returns a triple of con-

tinuously differentiable functions that satisfies (Lt(π2), Pt(

π2), At(

π2)) = (0, 0, 0) and

(L′t(0), P ′

t(0), A′t(0)) = (0, 0, 0) for all subsequent time. We consider initial condi-

tions with the form (L0(x), P0(x), A0(x)) = (CL cos x, CP cos x, CA cos x) for model

simulations, where CL, CP and CA are real positive constants.

Figure 4.1 shows simulations of model (4.4) for various parameter values in the

dispersal kernel (4.2). Parameters for population dynamics were chosen so as to fall

in the maximum likelihood 95% confidence intervals calculated from control cultures

of the Desharnais experiment for LPA parameters [7]. The LPA model equilibrates

for this set of parameter values.

When DLA = DAL = 0 in the kernel, there is no spatial segregation between

stages in simulations of model (4.4) for any set of LPA parameter values. Spatial

distributions of all stages reflect that of the kernel, cosm x, reaching a maximum at

the no-flux boundary 0 and reaching their minimum (of zero) at the hostile boundary

π2. This can be seen in Figure 4.1a, where m = 1, and Figure 4.1c, where m = 2.

If the adults and larvae have equal dispersal sensitivities of DLA = DAL = 0.01,

the larvae are “pushed” towards the bottom of the vial along with the pupae, while

adult density peaks on the surface. This is shown in Figure 4.1b for m = 1 and 4.1d

77

for m = 2. Both of these results are qualitatively similar to Ghent’s observations

shown in Figure 1.1. As the parameter m is increased, the bottom of the habitat

becomes increasingly undesirable to both stages. The adults reach their maximum

density at the surface for all values of m, but the peak of the larval distribution moves

towards the surface as m increases.

00

20

40

60

80

100

120

140

160

Space

LarvaePupaeAdults

π

2

(a) m = 1, DLA = DAL = 0

00

50

100

150

200

250

Space

LarvaePupaeAdults

π

2

(b) m = 1, DLA = DAL = 0.01

00

20

40

60

80

100

120

140

160

180

Space

LarvaePupaeAdults

π

2

(c) m = 2, DLA = DAL = 0

00

50

100

150

200

250

300

350

Space

LarvaePupaeAdults

π

2

(d) m = 2, DLA = DAL = 0.01

Figure 4.1: Simulations of the spatial LPA model (4.4) for various parameter valuesin the dispersal kernels (4.2). All attractors shown are temporal equilibria. LPAparameter values given by: b = 10, cel = 0.01, cea = 0.0175, µL = 0.435, cpa = 0.016,µA = 0.075.

In general, spatial segregation of larvae and adults is observed for a range of

DLA > 0, even if DAL = 0. The effect is more severe when combined with values of

DAL > 0, but this is not necessary. However, the reverse is not true. It is necessary

that larvae move in response to adult densities; if DLA = 0 and only adults disperse in

78

response to larval densities, we do not see the segregation. Adults will not be pushed

to the bottom of the habitat for any values of DLA and DAL. Furthermore, if DLA

and/or DAL become too large, we begin to see temporal patterns as well.

We did not observe multiple attractors of the spatial LPA model for the range

of parameter values considered. Initial conditions of the form (L0(x), P0(x), A0(x))

= (CL cosm x, CP cosm x, CA cosm x), where CL, CP , CA and m are real positive con-

stants lead to the attractors in Figure 4.1. The boundary conditions can also be

satisfied by piecewise continuous step functions of the form (L0(x), P0(x), A0(x)) =

(CLIx<x∗, CP Ix<x∗, CAIx<x∗), where CL, CP , CA are real positive constants and Ix<x∗

is equal to 1 on 0 ≤ x ≤ x∗ and 0 on x∗ ≤ x ≤ π2. This set of initial conditions also

leads to the attractors in Figure 4.1.

4.2 Spatial Patterns in Tribolium brevicornis

As pointed out in Chapter 1, the flour beetle Tribolium brevicornis has major biolog-

ical differences from Tribolium castaneum that prevent the LPA model from being

used to describe its population dynamics accurately. The most glaring difference is

that T. brevicornis is an inhibitor. In the presence of high adult densities, larvae of

this species may fail to pupate. Whereas T. castaneum larvae always pupate after

spending two weeks as a larva, T. brevicornis larvae may remain in the larval stage

indefinitely. When local adult densities lower, T. brevicornis larvae will go on to pu-

pate. Secondly, the innate length of the larval stage for T. brevicornis in the absence

of adults is four weeks (rather than the two weeks for T. castaneum). During the

second two-week interval, larvae are noticeably larger and more mobile than in their

first two-weeks as small larvae. A third biological difference between these species is

79

the notable absence of pupal cannibalism by adults in T. brevicornis. Pupal canni-

balism is a mechanism that controls adult recruitment in other flour beetle species

such as T. castaneum and T. confusum, but T. brevicornis have the alternate control

mechanism of inhibition.

A question of considerable interest is that of the evolutionary or selective advan-

tage of inhibition. It is usually considered advantageous for a species to have high

numbers of reproductive adults, yet often laboratory cultures of T. brevicornis have

a larval population greater than the adult population. One can ask if being an in-

hibitor provides some kind of adaptive advantage for T. brevicornis. Of course, there

are many ways to define an advantage. We will look at one based on total population

size and the total numbers of animals in each class at equilibrium.

In order to model the spatial dynamics observed in T. brevicornis, we need a

non-spatial model to describe the population dynamics of the species. We modify

the LPA model to take into account the three main biological differences between T.

brevicornis and T. castaneum/T. confusum. We will refer to this new model as the

“SLPA′′ model.

4.2.1 The SLPA Model

In order to account for the longer larval stage, we split the “L” class of the LPA

model into two new classes and denote them by “S” and “L”. S(t) represents the

number of younger or “small larvae” at time t and L(t) now represents the number

of “large larvae” at time t. This class includes third week larvae as well as larvae

who have been inhibited for an indefinite amount of time. The time step of the

model remains two weeks. We model inhibition with a Ricker type, or exponential,

nonlinearity. This is appropriate given the assumption that inhibition is a result of

80

random encounters of larvae with adults at a rate ki and the fraction of larvae inhibited

increases with the density of adults. This is the same modeling methodology used

to describe cannibalism [7]. T. brevicornis eggs are subject to cannibalism by small

larvae as well as large larvae. Since large larvae have been observed to be more

voracious eaters than small larvae, we allow each larval stage its own cannibalism

rate. Thus ces is the cannibalism coefficient of eggs by small larvae, and cel is the

cannibalism coefficient of eggs by large larvae. The cannibalism coefficient of pupae

by adults that appears in the LPA model is set to zero, as observations show that

T. brevicornis pupae are not cannibalized. Non-cannibalistic pupal mortality is still

assumed to be zero. The SLPA model is given by the following equations:

St+1 = bAte−cesSt−celLt−ceaAt

Lt+1 = St + (1 − µl)(1 − e−kiAt)Lt

Pt+1 = (1 − µl)e−kiAtLt

At+1 = Pt + (1 − µA)At.

(4.6)

As in the LPA model, P (t) represents the number of non-feeding larvae, pupae, and

callow adults, and A(t) represents the number of sexually mature adults at time t.

We note that when the inhibition constant ki = 0, all large larvae who escape natural

mortality move through to the P stage after one time step. If either the inhibition

constant or the adult density is large, the fraction of large larvae pupating will be

small.

Before we derive a spatial version of the SLPA model, we obtain parameter values

for T. brevicornis.

81

4.2.2 Parametrization

Population data for T. brevicornis were collected by R. F. Costantino and R. A.

Desharnais in 1982 (R. F. Costantino, personal communication). Twenty adult beetles

were placed into a one-half pint bottle with 20 grams of medium (95% unbleached

wheat flour and 5% dried brewer’s yeast by weight). The beetles were incubated in

the dark at 33 ± 1 degree Celsius and 60 ± 5% relative humidity. Each such culture

was replicated four times.

Every two weeks the beetles were counted and the number of small larvae, large

larvae, pupae and adults were recorded in a vector (St, Lt, Pt, At). Eggs were not

counted. After the counting, live animals were placed back in the bottle with fresh

medium. Each culture was maintained for 36 weeks (18 time steps).

In order to connect the deterministic SLPA model with this data using the

methodology in [9, 11], we construct a stochastic version of the model. Given the

accuracy of the census counts, the two main types of noise to consider are envi-

ronmental and demographic. Environmental stochasticity is the result of external

changes to a system, whereas demographic stochasticity accounts for the variation

among individuals within the population in variables such as clutch size, probability

of dying, etc. Since the laboratory conditions are extremely well controlled, we con-

sider demographic noise to be more relevant to the conditions under which our data

were collected and will assume this type of noise when building our stochastic model.

(This was the case for T. castaneum in the extensive study of that species’ dynamics

in [6, 7].)

The stochastic SLPA model is constructed similarly to the stochastic LPA model

[9], namely, by taking advantage of the ability of binomial and poisson distributions

82

to model the aggregation of demographic events (such as birth and death) within

each life stage. Following the work of Dennis et al. [9], we can model the conditional

one-time-step distributions of St, Lt, Pt and At as independent, discrete distributions

of either Poisson, binomial or sums of Poisson and binomial random variables. For

each of these distributions, a square root transformation transforms the distribution

to one which is approximately normal with constant variance. We can then represent

our stochastic model as a nonlinear autoregressive (NLAR) model of the following

form [11]:

Wt = h(Wt−1) + Et (4.7)

where the distribution of Wt, given Wt−1, should be normally distributed with con-

stant variance σ2 that does not depend on wt−1, the state of the system at time

t − 1.

We propose the following stochastic version of the SLPA model for explaining

the observed time series data of T. brevicornis populations:

St+1 =√

bAte−cesSt−celLt−ceaAt + E1t

Lt+1 =√

St + (1 − µL)(1 − e−kiAt)Lt + E2t

Pt+1 =√

(1 − µL)e−kiAtLt + E3t

At+1 =√

Pt + (1 − µA)At + E4t

(4.8)

where Et = (E1t, E2t, E3t, E4t)′ is a random noise vector with a multivariate normal

distribution having mean (0, 0, 0, 0) and a diagonal variance-covariance matrix. We

denote the diagonal entries of this matrix by the vector Σ = (σ211, σ

222, σ

233, σ

244). The

assumed demographic nature of the noise implies that the off-diagonal covariances will

be equal to zero, and the Et will be uncorrelated through time [11]. The stochastic

83

model reduces to the deterministic SLPA model 4.6 when Σ = 0.

We can use (4.8) to obtain maximum likelihood parameter estimates. The adult

death rate, µA, is not included in the maximum likelihood parametrization but rather

calculated directly to be µA = 0.0158 based on recorded observations of the number of

dead adults each time step from each replicate. Thus we have 6 remaining unknown

parameters in our deterministic model equations (which we denote by a vector Θ)

and 4 more in the vector (Σ). The likelihood function is defined by

L(Θ, Σ) =

q∏

t=1

p(wt|wt−1) (4.9)

where wt = [√

st

√lt√

pt√

at]′ (observations at time t), gives the probability the state

of the system at time t is wt, given the state of the system at time t − 1 is wt−1.

q is the number of one-step observations in our time series. All one-step transitions

are assumed independent. For each one-step, p(wt|wt−1) has a multivariate normal

probability distribution with mean h(wt−1) and variance-covariance matrix Σ [9]:

p(wt|wt−1) = |Σ|−1/2(2π)−3/2exp[−(wt − h(wt−1))′Σ−1(wt − h(wt−1))/2]. (4.10)

Maximum likelihood parameter estimates for Θ and Σ can be found by maximizing

the above likelihood function. A mathematically equivalent and computationally

simpler approach is to maximize the log of the likelihood function:

ln L(Θ, Σ) =

q∑

t=1

ln p(wt|wt−1) =

− q(3

2ln(2π) − (

q

2) ln |Σ| − 1

2=

q∑

t=1

(wt − h(wt−1))′Σ−1(wt − h(wt−1)). (4.11)

84

Matlab is used for computations, specifically the Nelder-Mead algorithm2 is used

to minimize the negative log-likelihood function [9]. Maximum likelihood estimates

are given in Table 4.1.

The observed time series data for each replicate and each life-stage (S, L, P and

A) are plotted in Figures 4.2 - 4.5, respectively, as solid lines. Dashed lines connect

observed values at time t to stochastic SLPA model predicted values (open circles)

at time t +1. The model predictions are calculated based on the parameter values in

Table 4.1. Visually, we see from Figures 4.2 - 4.5 that the one-step model predictions

and the data agree fairly well. The model predictions are not consistently higher or

lower than the observed data points.

Residuals can be calculated at each time step by subtracting the square root of the

model prediction from the square root of the observed population size (since noise was

added on the square root scale in equations (4.8). A statistical analysis of square root

residual vectors serves as a check on the appropriateness of both the model skeleton

and the type of noise assumed. For the demographic stochasticity we assumed, we

expect residuals (on the square root scale) to look like uncorrelated observations from

a multivariate normal distribution with mean 0 and constant variances given in Table

4.1. A visual inspection of histograms of residuals for each stage confirms there is no

severe departure from a normal distribution. The means and variances calculated for

the square root residuals for each stage appear in Table 4.2. The means are close to

0 for all distributions. Furthermore, the variances in Table 4.2 agree well with the

parameterized values in Table 4.1.

We note that the deterministic SLPA model (Σ = 0) with the maximum likelihood

parameter estimates in Table 4.1 predicts a temporal equilibrium. The equilibrium

2Thanks to Shandelle Henson and Brian Dennis for Matlab code used in the model parametriza-tion.

85

Table 4.1: Maximum likelihood parameter estimates for the stochastic SLPA model(4.8).

Parameter Estimate

b 11.4096ces 0.0135cel 0.0169cea 0.0223µL 0.1339ki 0.0194µ∗

A 0.0158σ2

11 3.553σ2

22 2.8545σ2

33 4.3776σ2

44 0.4865

stage vector is (S∗, L∗, P ∗, A∗) = (12.21, 71.77, 2.59, 163.84).

Calculations show that as the inhibition parameter ki increases, all other param-

eters remaining fixed, the number of large larvae equilibrium stage vector increases

and eventually L is the dominant stage with L∗ > A∗.

4.2.3 The Spatial SLPA Model

To consider a spatial extension of the SLPA model for T. brevicornis, we need to first

define a spatial domain. Although the patterns observed in T. brevicornis (see Figures

1.2, 1.5, 1.4) have all been on a two-dimensional surface, we will take advantage of

an approximate cross-sectional symmetry in some patterns observed in T. brevicornis

and consider one spatial dimension.

We also need to consider what form the dispersal kernel should have. As the

observed patterns are on the surface of the flour container (unlike when we were

considering depth of flour in Ghent’s experiment) there are no inherent heterogeneities

86

0 2 4 6 8 10 12 14 16 180

100

200

Sm

all L

arva

e

0 2 4 6 8 10 12 14 16 180

100

200

Sm

all L

arva

e

0 2 4 6 8 10 12 14 16 180

100

200

Sm

all L

arva

e

0 2 4 6 8 10 12 14 16 180

100

200

Sm

all L

arva

e

time (2 weeks)

Replicate 2

Replicate 1

Replicate 4

Replicate 3

Figure 4.2: Shown are time series data (solid lines) and one-step SLPA predictions(dashed lines) for small larvae, for each of the four replicates of the T. brevicornis

cultures. Parameters used for one-step predictions are given in Table 4.1.

in the surface habitat in different spatial locations. The incubator is dark and all

locations are under the same temperature and humidity conditions. Furthermore,

animals can occupy space right up to the boundary, but cannot pass through the

boundary walls. To incorporate these conditions into our model, we use a kernel

that preserves no-flux boundary conditions at both endpoints and does not have any

spatial dependence in the absence of density effects. Note that a constant kernel

(uniform dispersal on a finite interval [0, M ]) satisfies these conditions.

Not all life-stages of T. brevicornis disperse. Like T. castaneum and T. confusum,

pupae are sedentary. Since younger larvae in their first two weeks are smaller and

slower than older larvae, we make the simplifying assumption that larvae in the S

class do not disperse. This is consistent with observations of T. brevicornis cultures.

87

0 2 4 6 8 10 12 14 16 180

100

200

Larg

e La

rvae

0 2 4 6 8 10 12 14 16 180

100

200

Larg

e La

rvae

0 2 4 6 8 10 12 14 16 180

100

200

Larg

e La

rvae

0 2 4 6 8 10 12 14 16 180

100

200

time (2 weeks)

Larg

e La

rvae

Replicate 1

Replicate 2

Replicate 3

Replicate 4

Figure 4.3: Shown are time series data (solid lines) and one-step SLPA predictions(dashed lines) for large larvae, for each of the four replicates of the T. brevicornis

cultures. Parameters used for one-step predictions are given in Table 4.1.

The remaining two stages, L and A, are dispersers. In Chapter 2 we constructed

general stage-structured integrodifference equation models that incorporated density

dependent dispersal in two ways.

For each stage, density could potentially affect an individual’s probability of dis-

persing from a given location x (determined by the decision function γ(Jt(·), At(·), x))

and/or the probability of moving from x to another location y, given dispersal oc-

curs (determined by the dispersal kernel K(Jt(·), At(·), x, y)). For our spatial SLPA

model of T. brevicornis, we assume that adult beetles always disperse (γA = 1 at

all spatial locations and at all times), and tend to prefer locations with lower pupal

densities than their starting location. This is biologically reasonable since pupae are

known to be better off if they have their own space to complete the pupation process

88

0 2 4 6 8 10 12 14 16 180

50

100

150

Pup

ae

0 2 4 6 8 10 12 14 16 180

50

100

150

Pup

ae

0 2 4 6 8 10 12 14 16 180

50

100

Pup

ae

0 2 4 6 8 10 12 14 16 180

50

100

150

time (2 weeks)

Pup

ae

Replicate 4

Replicate 3

Replicate 2

Replicate 1

Figure 4.4: Shown are time series data (solid lines) and one-step SLPA predictions(dashed lines) for pupae, for each of the four replicates of the T. brevicornis cultures.Parameters used for one-step predictions are given in Table 4.1.

and do not have other stages crawling over them. Recall T. brevicornis adults do

not cannibalize pupae and have no reason to actively seek them out. We incorporate

density dependent dispersal into the adult kernel by multiplying our constant density

independent kernel by an exponentially decreasing function of pupal density. Thus

we assume adult dispersal is governed by the following kernel, where KA gives the

fraction of adults that move from their starting location to location x:

KA =1

Cexp(−DAP ((1 − µL) exp(−kiAt(x))Lt(x)))

γA = 1

(4.12)

where C is a normalization constant. We also assume that large larvae have no reason

to avoid small larvae or pupae. However, if adult densities are too great, the large

89

0 2 4 6 8 10 12 14 16 180

100

200

Adu

lts

0 2 4 6 8 10 12 14 16 180

100

200

300

Adu

lts

0 2 4 6 8 10 12 14 16 180

100

200

Adu

lts

0 2 4 6 8 10 12 14 16 180

100

200

time (2 weeks)

Adu

lts

Replicate 1

Replicate 2

Replicate 3

Replicate 4

Figure 4.5: Shown are time series data (solid lines) and one-step SLPA predictions(dashed lines) for adults, for each of the four replicates of the T. brevicornis cultures.Parameters used for one-step predictions are given in Table 4.1.

larvae will be inhibited and unable to pupate. Therefore, we assume adult density

affects the fraction of large larvae dispersing at any given time and location. We model

the fraction of large larvae dispersing at each time step by an increasing function of

adult density, assuming the large larvae that do disperse redistribute uniformly over

the entire habitat. Under these assumptions, we describe the dispersal of large larvae

by the following kernels:

KL =1

M

γL = 1 − exp(−DLA(Pt(y) + (1 − µA)At(y))).

(4.13)

Here KL gives the fraction of dispersing larvae that move from their starting location

y to position x, and γL give the fraction of adults at position y that decide to disperse

90

Table 4.2: Residual Analysis. Means and variances for the square root residuals forstages S, L, P and A. Untransformed residuals can be seen in Figures 4.2-4.5, as thedifference between the model prediction and observed stage vector at each time-step.Maximum likelihood estimates for the variances of the transformed residuals are givenin Table 4.1.

Stage Mean Variance

S 0.158 3.576L 0.412 2.723P 0.194 4.402A -0.136 0.4746

at a given time. The decision function γL is plotted in Figure 4.6 for several values

of the parameter DLA. The fraction of larvae leaving a given location and time is a

function of the density of adults present at that location and time.

Combining the dispersal kernel, decision functions, and SLPA population dynam-

ics, we arrive at a spatial SLPA integrodifference equation model on the homogenous

spatial domain Ω = [0, M ]:

St+1(x) = bAt(x)e−cesSt(x)−celLt(x)−ceaAt(x)

Lt+1(x) =

∫ M

0

1

M(1 − e−DLA(Pt(y)+(1−µA)At(y)))×

(St(y) + (1 − µL)(1 − e−kiAt(y))Lt(y))dy

+ e−DLA(Pt(x)+(1−µA)At(x))(St(x) + (1 − µL)(1 − e−kiAt(x))Lt(x))

Pt+1(x) = (1 − µL)e−kiAt(x)Lt(x)

At+1(x) =

∫ M

0

1

Ce−DAP ((1−µL)e−kiAt(x)Lt(x))(Pt(y) + (1 − µA)At(y))dy.

(4.14)

To simulate this model, we must choose initial conditions that satisfy our no-flux

boundary conditions. An initial condition with a uniform spatial distribution will

91

0 10 20 30 40 50 60 70 80 90 1000

0.1

0.2

0.3

0.4

0.5

0.6

0.7

0.8

0.9

1

Adult Density

Fra

ctio

n of

Lar

vae

Dis

pers

ing

DLA

= 0.01

DLA

= 0.05

DLA

= 0.1

DLA

= 0.5

DLA

= 1

Figure 4.6: The decision function γL is shown for several values of DLA. γL(At(x))gives the fraction of larvae dispersing from location x at time t as a function of adultdensity at location x at time t.

remain a uniform spatial distribution for all time since there are no preferences for

different spatial locations built into the model. The fraction of individuals leaving or

settling at a given location depends only on the population density at that location.

There are many non-uniform initial distributions one could consider. We restrict

our investigation to a subset of initial conditions, namely a vector consisting of only

mobile stages (0, L0(x), 0, A0(x)). We can think of this initial condition as representing

a biological invasion of a new environment, and the only possible invaders are our

dispersing stages L and A.

In fact we subject our initial vector (0, L0(x), 0, A0(x)) to a uniform distribution

on a subinterval of the spatial domain [0, m]. Thus L0(x) = CL for 0 ≤ x ≤ m ≤ M

and L0(x) = 0 for m ≤ x ≤ M . This initial condition satisfies the no-flux boundary

conditions and the chosen dispersal kernels preserve them for all subsequent time.

Furthermore, these initial conditions can be easily produced experimentally.

Extensive numerical simulations show that with this set of initial conditions, un-

92

der the maximum likelihood SLPA model parameter estimates obtained above, the

spatial SLPA model admits multiple attractors. These attractors include a spatially

uniform distribution and a “patchy” distribution. The patchy attractor consists of a

spatially uniform equilibrium (Se, Le, Pe, Ae) on [0, m] and a spatially uniform equi-

librium (S∗e , L

∗e, P

∗e , A∗

e) on [m, M ]. For DAP large enough, adults are completely

restricted to either [0, m] or [m, M ] and are completely absent in the other patch.

Thus, the model predicts that pupae form a “nest,” i.e. a patch of pupae not occu-

pied by adults. As we will see below, these nests have been observed in experimental

cultures of T. brevicornis.

Initial conditions leading to each attractor vary with parameter values. For exam-

ple, the initial condition of L0(x) = 0, A0(x) = 100 on the left quarter of the domain2

[0, 14] leads to a patchy attractor for many values of ki, including ki = 0.01, 0.0194,

0.05, and 0.07. For these same values of ki, an initial condition of L0(x) = 100,

A0(x) = 0 for x ∈ [0, 14] leads to a spatially uniform attractor.

When inhibition is absent (ki = 0) the patchy distribution is either extremely

rare or never present. Most initial conditions lead to a spatially uniform distribution.

This suggests that inhibition is necessary for the segregation of life cycle stages in a

homogenous habitat.

These findings are illustrated in Figure 4.7 and Figure 4.8, which show the uniform

and patch attractors resulting from the same set of parameter values (given in figure

captions), but different initial conditions. In Figure 4.7 the initial conditions are

L0(x) = 100, A0(x) = 0 for x ∈ [0, 14] and the attractor is a spatially uniform

equilibrium with (Se, Le, Pe, Ae) = (12.21, 71.77, 2.59, 163.8). In Figure 4.8, the initial

conditions are L0(x) = 0, A0(x) = 100 for x ∈ [0, 14], and the result is a two patch

2Domain size M does not affect model attractors, only total population size. M = 1 and m = 1

4

were used for simulations in this chapter.

93

spatial distribution that equilibrates in time. The population density on the left patch

[0, 14] is given by (Se, Le, Pe, Ae) = (0.015, 13.92, 12.06, 0.0016) and the population

density on the right patch [14, 1] is given by (S∗

e , L∗e, P

∗e , A∗

e) = (6.58, 13.92, 0.08, 259.4).

A natural question to ask when comparing the spatially segregated attractor with

the spatially uniform attractor is how the total number of individuals in each life-

stage, or the total population size, differs between these distributions. Table 4.3

shows a comparison of the total number of individuals in each stage, along with the

total population size, for each attractor. For ki = 0 (no inhibition) a patchy attractor

could not be found. Table 4.3 shows that as the degree of inhibition ki increases (i.e.

it takes fewer adults to inhibit the same fraction of large larvae) the total population

size decreases for both attractors. Yet the difference in total population size between

the attractors also decreases, and for ki = 0.07, the total population size is greater

for the patchy distribution than the uniform distribution.

The relative density of each of the classes also changes as the degree of inhibition

increases. For each value of ki in Table 4.3, the attractor with the greater density of

each class size is in boldface (except for ki = 0). For ki = 0.01, we see that the uniform

attractor has a larger total density of all stages S, L, P and A compared to the patchy

attractor. For ki ≥ 0.0194 (the maximum likelihood parameter estimates for historical

T. brevicornis census data), SUniform > SPatch and LUniform > LPatch, but PPatch >

PUniform and APatch > AUniform. The two attractors in Figure 4.7 and Figure 4.8

are not the only possible attractors for this set of parameter values, but they are by

far the most common for the set of initial conditions we are investigating (a uniform

distribution of dispersing stages on a subinterval of the domain) and they are also

the two attractors seen by Costantino in the laboratory (personal communication). A

photo of the uniform attractor in T. brevicornis appears in Figure 1.6. The “patchy”

94

Table 4.3: Comparison of population vectors and total population size for spatiallyuniform versus spatially segregated (patchy) attractors for varying degrees of inhibi-tion on the spatial domain [0, 1]. All other parameter values used were: b = 11.41,µL = 0.134, µA = 0.0158, ces = 0.0135,cea = 0.0223, cel = 0.0169, DAP = 1 andDLA = 0.05. Simulations with initial condition a were started with 25 large larvaeon the left 1

4of the spatial domain (an initial distribution of L0(x) = 100, A0(x) = 0

on [0, 14]). Each simulation started with this initial condition went to the spatially

uniform attractor illustrated in Figure 4.7. Simulations with initial condition b werestarted with 25 adults on the left 1

4of the spatial domain (an initial distribution of

L0(x) = 0, A0(x) = 100 on [0, 14]). With the exception of ki = 0, simulations with

initial condition b led to the patchy attractor illustrated in Figure 4.8. For each valueof ki, the total number of small larvae, large larvae, pupae, adults and total popu-lation size (the sum of all stages) are calculated for each attractor (by integratingeach equilibrium distribution from 0 to 1 with respect to the spatial variable x). Thegreater number is shown in bold.

I.C. ki Attractor S L P A Total

a 0 Uniform 5.14 5.14 4.45 281.71 296.44b 0 Uniform 5.14 5.14 4.45 281.71 296.44

a 0.01 Uniform 8.51 38.21 3.51 222.02 271.34

b 0.01 Patch 4.73 11.89 3.14 198.30 218.02

a 0.0194 Uniform 12.21 71.77 2.59 163.84 250.41

b 0.0194 Patch 4.94 13.92 3.08 194.55 216.49

a 0.05 Uniform 16.69 114.39 1.36 85.82 218.26

b 0.05 Patch 4.97 14.17 3.07 194.14 216.34

a 0.07 Uniform 17.39 122 1.04 65.98 206.41b 0.07 Patch 4.97 14.17 3.07 194.14 216.34

95

0 M0

20

40

60

80

100

120

140

160

180

Space

Adults

Pupae

Large Larvae

Small Larvae

Figure 4.7: Temporal equilibrium attractor of spatial SLPA model. Attractor isspatially uniform with (Se, Le, Pe, Ae) = (12.21, 71.77, 2.59, 163.8). Parameter valuesused for SLPA model are maximum likelihood estimates: b = 11.41, µL = 0.134,µA = 0.0158, ces = 0.0135,cea = 0.0223, cel = 0.0169, ki = 0.0194. DAP = 1 andDLA = 0.05. General initial condition: L0 = 100M , A0 = 0 on the subinterval ofdomain [0, M

4], M > 0. Here M = 1.

96

0 M/4 M0

50

100

150

200

250

300

Space

Adults

Pupae

Large Larvae

Small Larvae

Figure 4.8: Temporal equilibrium attractor of spatial SLPA model, consist-ing of two patches with (Se, Le, Pe, Ae)=(0.015, 13.93, 12.06, 0.0016) on [0, M

4] and

(S∗e , L

∗e, P

∗e , A∗

e)=(6.58, 13.92, 0.08, 259.4) on [M4, M ]. Parameter values used for

SLPA model are the maximum likelihood estimates: b = 11.41, µL = 0.134,µA = 0.0158, ces = 0.0135,cea = 0.0223, cel = 0.0169, ki = 0.0194. DAP = 1and DLA = 0.05. General initial condition: A0 = 100M , L0 = 0 on the subinterval ofdomain [0, M

4], M > 0. Here M = 1.

97

attractor, and the establishment of a pupal nest by T. brevicornis, can be seen in

Figures 1.5 and 1.4.

The culture in Figure 1.6 was started with T. brevicornis adults in the upper

right corner of the box. The adults immediately spread out and no pupal nest was

ever established. To simulate this situation where adults are immediately allowed to

disperse, the spatial SLPA model needs to be started with an initial condition of

only L stage individuals. Since reproduction occurs before dispersal in the model,

L stage individuals will all pupate immediately (since no A stage individuals are

present to delay pupation) and emerge as adults. These adults will then disperse

according to equation (4.12), the adult dispersal kernel. The absence of pupae results

in the adults dispersing uniformly throughout the entire domain, matching what is

seen experimentally. Therefore, a laboratory initial condition of only adults who are

immediately permitted to disperse corresponds to a model simulation initial condition

of only large larvae. In two time steps this initial condition will result in a cohort of

dispersing adults, with no other stages present.

The three cultures in Figure 1.4 were started with L and A stage animals mixed

together on the left half of the domain. The movement of these animals was restricted

by a panel inserted to divide each row of flour in half. After 6 weeks the panel was

removed and animals were able to migrate into the right half of each row. This

resulted in the formation of a pupal nest. Such a patchy attractor can be predicted

by the model for initial conditions of adults only, or of both adults and large larvae.

Since reproduction occurs before dispersal in the model, the initial condition of A

only results in A and S stage individuals present at the time of dispersal. The small

larvae do not disperse, but the A spread out across the entire domain. The next

time step, small larvae become large larvae. These large larvae will not disperse

98

provided the adult density is low enough (this depends on the decision parameter

DLA) from the adults spreading out over the entire habitat. At the next time step,

these same large larvae will pupate in their original location if adult density is low

enough (this depends on the inhibition parameter ki). Once they do, a pupal nest

has been established and it will be avoided by the adults in subsequent time steps.

If the model is instead started with both L and A stage individuals present, a

situation similar to the one just described (for an initial condition of A only) occurs.

The initial density of A may be great enough to inhibit large larvae. If so, a fraction

of them, determined by γL, will disperse uniformly over the whole domain along with

the adults. The next time step, adults should be spread out enough to allow all large

larvae to pupate. Unless all L dispersed, the density of L should be greater in their

starting interval than the rest of the domain and this will result in a greater density

of pupae and mark the location of the pupal nest. The moment the door is opened

in the laboratory culture corresponds to halfway through a model time step - after

reproduction but right before dispersal.

Over time, the nest persists both in model simulations (since the patchy attractor

is a temporal equilibrium attractor) and laboratory cultures, in the location it was

originally established. This location does not have to be at the edge of the domain; it

can be an interval in the center of the domain as well. Adults can become very dense

outside of the nest, and this may provide a barrier to any invading species, including

those where adults cannibalize pupae. Costantino has hypothesized that the pupal

nest together with the adult barrier surrounding it may serve to protect the pupae

this species has invested so much time into developing (personal communication).

In summary, the spatial SLPA model (4.8) has been able to predict observed

spatial segregation in T. brevicornis. We were able to further connect model (4.8)

99

with experimental observations for select cases, providing experimental support for

the multiple spatial attractors predicted by the spatial SLPA model. The fact that

we were unable to find initial conditions leading to the patchy attractor when the

inhibition parameter ki = 0 suggests that the inhibition of large larvae is a necessary

condition for spatial segregation of life cycle stages. This is consistent with the absence

of surface patterns in non-inhibiting species such as T. castaneum and T. confusum.

Furthermore, the model predicts that for a species with the parameterized inhibition

level of T. brevicornis (ki = 0.0194), the spatial separation of life cycle stages can

affect the relative total population sizes of the stages. Specifically, the total number of

pupae and adults are higher for the patchy attractor relative to the uniform attractor.

Sexually mature adults become the dominant stage in the patchy attractor, whereas

the immature large larvae dominate for the spatially uniform attractor.

100

Chapter 5

Concluding Remarks

We have shown that density dependent dispersal can lead to the segregation of life

cycle stages in stage-structured populations (such as flour beetles), and furthermore

that these patterns can occur in both homogenous and heterogeneous spatial envi-

ronments. These results are not intuitively obvious. In the case of a juvenile-adult

species, even if juvenile avoid adults and adults avoid juveniles, juveniles will even-

tually become adults and adults will give rise to new juveniles, seemingly leading to

spatial mixing between stages.

In chapter 1 we reference spatial habitat segregation as a potential adaptation to

avoid cannibalism or predation. In Tribolium, larvae are not being directly cannibal-

ized by the adults but density dependent dispersal, and the avoidance of adults, may

still be a beneficial adaptation. T. castaneum and T. confusum pupae are cannibal-

ized by adults, so if larvae pupate in a location inhabited by a large number of adults,

they may not survive the pupal stage. T. brevicornis pupae, on the other hand, are

not eaten by adults. These larvae have a different reason to escape adults - inhibited

pupation caused by the presence of adults. Large larvae cannot complete their life

cycle if too many adults are nearby.

Based on our studies we also hypothesize about mechanisms that promote adult

dispersal. Eggs of both species are cannibalized by larvae, and the adults may want

to lay their eggs in flour where they have the greatest chance of hatching, and where

there is the greatest food supply once their young become larvae. This is our rationale

101

for adults avoiding larvae in the dispersal kernels in our model for T. confusum and T.

castaneum. For T. brevicornis, the adult avoidance of larvae does not lead to spatial

segregation of adults and other life stages. In this species, segregation is obtained

when the adults avoid the P stage, which includes pre-pupae. This species devotes

more time to the larval stage than T. confusum and T. castaneum, spending at least

four weeks as a larvae rather than only two. If larvae are inhibited, they may spend

months in the larval stage. After investing so much time in the larval stage, the

species takes great care of their pupae. Adults leave them alone, since crawling over

them can decrease the chances they will emerge as a healthy adult.

The toy models in Chapter 3 show that density dependent dispersal can result

in both spatial segregation of life cycle stages as well as temporal segregation (if

mortality is associated with dispersal). We stress that complex population dynamics

are not necessary for the formation of these spatial patterns and that our models do

not result in spatial segregation in the absence of density dependent dispersal (as long

as all mobile stages “view” the environment the same way).

In Chapter 4 we considered two biological applications of the spatial models devel-

oped in Chapters 2 and 3. First, we aimed to explain the uneven depth distribution of

larval and adult flour beetles of the species T. castaneum and T. confusum observed

by Ghent [14]. Due to gravity, flour is more dense at the bottom of a vial of flour

than at the top, and most likely less desirable an environment to beetles than the

less dense flour near the surface. We modeled this biological system assuming no-flux

boundary conditions at the surface and hostile boundary conditions at the bottom of

the bottle, as well as the density dependent dispersal of adults and larva (assuming

adults avoid larvae and visa versa). Even though larvae and adults have the same

spatial preferences, the adults end up with higher densities in the preferred flour.

102

Larvae and pupae appear to be “pushed” towards the bottom of the bottle. These

model predicted results agree qualitatively with what Ghent observed (Figure 1.1).

It is also of note that the lower dimensional juvenile-adult spatial model with mixed

boundary conditions considered in Chapter 3 also results in adults dominating near

the surface and juveniles reaching their maximum density at lower depths.

We do not expect quantitative agreements between our results and those of Ghent

for multiple reasons. First, parameter values for flour beetle populations can vary from

strain to strain, and we do not have parameter values for the strain used by Ghent.

Second, the vials used by Ghent held 8 grams of flour, while the typical bottles used

in experiments and parametrization of the LPA model held 20 grams of flour [7, 9].

Also, Ghent’s data was collected after 26 days in the vial, not long enough for new

adults to emerge. Our results are asymptotic model attractors.

As discussed in Chapter 1, there is literature to support the density dependent

dispersal of flour beetles, but the exact mechanisms and rates of dispersal are un-

known. Our results do show that our hypothesized density dependent interactions

and the resulting dispersal kernels are possible explanations for the observed spatial

distributions. The main message of this case study is that density dependent disper-

sal can produce a spatial pattern very similar to that observed by Ghent and may

have been a factor in its development.

The patterns observed in our second case study species, T. brevicornis, are more

striking than the subtle depth differences observed by Ghent. Adults clearly clump

together, a behavior not seen by Costantino or noted in the literature for any other

Tribolium species, even other inhibiting species such as T. freemani. However, T.

freemani larvae are inhibited by both larvae and adults, so escaping high densities of

adults alone would not help them pupate. In fact, laboratory cultures of this species

103

almost always result in a strong larval bottleneck with very few adults present. In

T. brevicornis, on the other hand, when larvae escape to areas of low adult density

they immediately pupate (provided they are old enough). Mathematically, inhibition

plays an important role in the formation of spatial segregation. The number of initial

conditions giving rise to the patchy attractor of the spatial SLPA model decreases as

the severity of inhibition decreases. In fact, the patchy attractor could not be found

when the inhibition parameter ki was set to zero.

A potential benefit of density dependent dispersal for inhibiting species lies in

population size and structure of the stage-class vector. All other parameters equal,

the model always predicts a non-inhibitor will have greater total population sizes

than an inhibiting species. This makes sense intuitively, since 100% of larva surviving

mortality go on to pupate in the absence of inhibition. Inhibition only decreases this

number and can only decrease total population size. Using this criteria, inhibition

appears to put a species at a disadvantage.

We also compared population numbers for an inhibiting species in two differ-

ent spatial structures. We found that if inhibition is strong enough, the spatially

segregated model attractor has a greater population size than the spatially uniform

attractor. This also makes biological sense. If a species is a strong inhibitor, very

few larvae will be able to pupate once an adult cohort has been established. Eggs

will still be laid but few new sexually mature adults will be established. The model

shows that separating the stages spatially and giving the larvae a refuge to pupate

results in higher population numbers, which is typically considered to be biologically

beneficial for the population.

For the parameterized inhibition level of T. brevicornis, spatial segregation does

not quite result in greater total population numbers, but it does shift the composition

104

of the equilibrium stage vector (S∗, L∗, P ∗, A∗) in favor of higher numbers of P ∗ and

A∗. Increasing the number of sexually mature adults in a population is also often

seen as biologically beneficial. If dispersal is really important in this species’ natural

habitat, adults may be the primary invaders of new colonies. Increasing the number

of adults could mean larger founding populations at their next location.

As noted above, the spatial SLPA model exhibits multiple attractors, including a

spatially uniform attractor and a “patchy” attractor with pupae and adults spatially

separated. These two attractors have been seen in experimental cultures of T. brevi-

cornis. The spatial segregation of adults and other life cycle stages has been observed

in many different sizes and shapes of T. brevicornis cultures, as seen in Figures 1.2,

1.5 and 1.4. However, the shape of the surface of the container used in Figures 1.5 and

1.4 is the closest to being one-dimensional and also produces the most reproducible

patterns. Vertical cross sections through each row yield an approximately uniform

distribution of beetles and so the pattern can be collapsed to one dimension more

easily than those in Figure 1.2.

Figures 1.5 and 1.4 clearly show the formation of the pupal nest that is predicted by

the patchy attractor of the spatial SLPA model, while Figure 1.6 shows the uniform

attractor of the spatial SLPA model. As discussed in Chapter 4.2, initial conditions

for the laboratory cultures are consistent with those used in model simulations. A

patchy attractor is reached if a pupal nest has a chance to be established. If a culture

is started with adults who immediately have the opportunity to disperse, they spread

out, taking advantage of the entire habitat. If a culture is started with large larvae

and adults who are contained in a subsection of the habitat, the adults inhibit the

large larvae and prevent them from pupating. Once the “door” is opened, allowing

them to access to the entire habitat, the adults disperse quickly. The large larvae

105

don’t get very far before sensing conditions are right to pupate, and a pupal nest is

established. Once the nest is established, it persists over time. New adults emerge and

leave the nest, while large larvae from outside the nest have been observed returning

to the nest.

5.1 Future Investigations

The integrodifference equation models we investigated thus far have been one di-

mensional in space. The observed patterns used to motivate our studies are higher

dimensional. While the patterns in Ghent’s experiments and in each of the rows in

Figure 1.4 can reasonably be collapsed to an approximately one dimensional distribu-

tion, the surface patterns shown in Figure 1.2 are definitely two dimensional surface

patterns. A two dimensional spatial model may be able to explain patterns such as

the “beetle butterfly” seen in Figure 1.2b. Mathematically, the models can easily be

extended to two dimensional spatial domains, and the bifurcation theory developed

in Chapter 2 applied in any spatial dimension. However, model simulations become

much more computationally intensive.

All our spatial LPA and SLPA model simulations were done on a one dimensional

domain consisting of a closed finite interval. An assumption in the formation of the

kernels for the spatial LPA and SLPA models was that all dispersing animals could

cover the entire habitat in one time step. We have not considered the role of domain

size in the formation of spatial patterns for these models. (We did, however, consider

domain size in the “toy” juvenile-adult models in Chapter 3, observed that spatial

segregation does not occur if the domain is too small.) Domain size and/or dispersal

speed may play an important role in the formation of spatial patterns of species

106

exhibiting density dependent dispersal on a larger domain. It would be interesting to

study larger domains which dispersers cannot cover in one time period.

Our results may have important implications for future multi-species competition

studies. Many experiments have been done on the subject of competition between

closely related species [29]. In cultures of T. confusum and T. castaneum, almost

all cultures saw one species exclude the other according to the principle of competi-

tive exclusion. The winning species depended on initial conditions. The LPA model

has had previous success modeling competition; an extension of the LPA model to a

competition model has led to potential counterexamples to the principle of competi-

tive exclusion, explaining prolonged coexistence between two species of closely related

flour beetles observed by Park [13].

Jillson and Costantino experimented with competition between T. brevicornis and

T. castaneum. Every culture resulted in competitive exclusion, with T. brevicornis

always being eliminated regardless of initial conditions [4, 21]. Inhibition of T. bre-

vicornis larvae is not species specific, so contact with T. castaneum adults will also

delay pupal metamorphosis [21]. Furthermore, T. castaneum adults will cannibalize

T. brevicornis pupae in addition to their own. Thus, T. brevicornis has a two-fold

disadvantage. Their larvae are inhibited by both species’ adults and their larvae that

do manage to pupate are now subject to cannibalism.

The nest of pupae established by T. brevicornis seen in Figure 1.5 does not contain

any nutritious flour. Instead the flour is so highly conditioned it is toxic. This makes

it a perfect place for pupae, as they do not need to eat, and it would provide safety

for T. brevicornis pupae in the presence of cannibalistic T. castaneum adults. As

discussed in Chapter 4.2, the high densities of T. brevicornis adults surrounding the

nest form a barrier for potentially cannibalistic invaders.

107

These factors suggest spatial structure may play an important role when con-

sidering competition between T. brevicornis and a non-inhibiting species such as T.

castaneum or T. confusum. If T. brevicornis has a chance to establish a pupal nest,

the species may be better able to resist an invasion by another flour beetle species.

108

References

[1] Alzoubi, M. Y. A Dispersal Model for Structured Populations. PhD thesis,

University of Arizona, 1999.

[2] Comins, H. N., and Hassell, M. P. Persistence of multispecies host-

parasitoid interactions in spatially distributed models with local dispersal. Jour-

nal of Theoretical Biology 183 (1996), 19–28.

[3] Costantino, R. F., Cushing, J. M., Dennis, B., Desharnais, R. A., and

Henson, S. M. Resonant population cycles in temporally fluctuating habitats.

Bulletin of Mathematical Biology 60, 2 (1998), 247–275.

[4] Costantino, R. F., and Desharnais, R. A. Population Dynamics and the

Tribolium Model: Genetics and Demography. Springer–Verlag, New York, 1991.

[5] Cushing, J. The LPA model. Fields Institute Communications 42 (2004),

29–55.

[6] Cushing, J. M. An Introduction to Structured Population Dynamics,

vol. 71. CBMS-NSF Regional Conference Series in Applied Mathematics, SIAM,

Philadelphia, 1998.

[7] Cushing, J. M., Costantino, R. F., Dennis, B., Desharnais, R. A.,

and Henson, S. M. Chaos in Ecology: Experimental Nonlinear Dynamics,

vol. 1 of Theoretical Ecology Series. Academic Press (Elsevier Science), New

York, 2003.

109

[8] Cushing, J. M., and Yicang, Z. The net reproductive value and stability in

matrix population models. Natural Resource Modeling 8 (1994), 297–333.

[9] Dennis, B., Desharnais, R. A., Cushing, J. M., and Costantino, R. F.

Nonlinear demographic dynamics: Mathematical models, statistical methods,

and biological experiments. Ecological Monographs 65 (1997), 261–281.

[10] Dennis, B., Desharnais, R. A., Cushing, J. M., and Costantino, R. F.

Transitions in population dynamics: equilibria to periodic cycles to aperiodic

cycles. Journal of Animal Ecology 66 (1997), 704–729.

[11] Dennis, B., Desharnais, R. A., Cushing, J. M., Henson, S. M., and

Costantino, R. F. Estimating chaos and complex dynamics in an insect

population. Ecological Monographs 71 (2001), 277–303.

[12] Dwyer, G., and Morris, W. F. Resource-dependent dispersal and the speed

of biological invasions. American Naturalist 167 (2006), 165–176.

[13] Edmunds, J., Cushing, J. M., Costantino, R. F., Henson, S. M., Den-

nis, B., and Desharnais, R. A. Park’s Tribolium competition experiments:

a non-equilibrium species coexistence hypothesis. Journal of Animal Ecology 72

(2003), 703–712.

[14] Ghent, A. W. Studies of behavior of the Tribolium flour beetles. II. Distri-

butions in depth of T. castaneum and T. confusum in fractionable shell vials.

Ecology 47 (1966), 355–367.

[15] Gil’, M. I. Difference Equations in Normed Spaces: Stability and Oscillations.

Elsevier, Amsterdam, 2007.

110

[16] Hardin, D. P., Takac, P., and Webb, G. Asymptotic properties of a

continuous-space discrete-time population model in a random environment. Jour-

nal of Mathematical Biology 26 (1988), 361–374.

[17] Harrison, R. G. Dispersal polymorphisms in insects. Annual Review of Ecology

and Systematics 11 (1980), 95–118.

[18] Hassell, M. P., Comins, H. N., and May, R. M. Spatial structure and

chaos in insect population dynamics. Letters to Nature 353 (1991), 255–258.

[19] Hassell, M. P., Comins, H. N., and May, R. M. Species coexistence and

self-organizing spatial dynamics. Letters to Nature 370 (1994), 290–292.

[20] Hunte, W., and Myers, R. A. Phototaxis and cannibalism in gammaridean

amphipods. Marine Biology 81 (1984), 75–79.

[21] Jillson, D. A., and Costantino, R. F. Growth, distribution, and competi-

tion of Tribolium castaneum and Tribolium brevicornis in fine-grained habitats.

The American Naturalist 116 (1980), 206–219.

[22] Jormalainen, V., and Shuster, S. M. Microhabitat segregation and can-

nibalism in an endangered freshwater isopod, Thermosphaeroma thermophilum.

Oecologia 111 (1997), 271–279.

[23] Kielhofer, H. Bifurcation Theory: An Introduction with Applications to

PDEs. Springer, New York, 2004.

[24] Kisimoto, R. Effect of crowding during the larval period on the determination

of the wing form of an adult plant-hopper. Nature 178 (1956), 641–642.

111

[25] Kot, M., Lewis, M. A., and van den Driessche, P. Dispersal data and

the spread of invading organisms. Ecology 77 (1996), 2027—2042.

[26] Kot, M., and Schaffer, W. Discrete-time growth-dispersal model. Mathe-

matical Biosciences 80 (1986), 109–136.

[27] Krasnosel’skii, M. A. Topological Methods in the Theory of Nonlinear Integral

Equations. Springer, Berlin, 1964.

[28] Leonardsson, K. Effects of cannibalism and alternative prey on population

dynamics of Saduria entomon (Isopoda). Ecology 72 (1991), 1273–1285.

[29] Leslie, P. H., Park, T., and Mertz, D. B. The effect of varying the initial

numbers on the outcome of competition between two Tribolium species. Journal

of Animal Ecology 37 (1968), 9–23.

[30] Levin, S. A. Dispersion and population interactions. American Naturalist 108

(1974), 207–228.

[31] Li, C. K., and Schneider, H. Applications of Perron-Frobenius theory to

population dynamics. Journal of Mathematical Biology 44 (2002), 450–462.

[32] Naylor, A. F. An experimental analysis of dispersal in the flour beetle, Tri-

bolium confusum. Ecology 40 (1959), 453–465.

[33] Naylor, A. F. Dispersal in the red flour beetle Tribolium castaneum (tenebri-

onidae). Ecology 42 (1961), 231–237.

[34] Naylor, A. F. Dispersal responses of female flour beetles, Tribolium confusum,

to presence of larvae. Ecology 46 (1965), 341–343.

112

[35] Neubert, M., Kot, M., and Lewis, M. Dispersal and pattern formation in

a discrete time predator-prey model. Theoretical Population Biology 48 (1994),

7–43.

[36] Neubert, M. G., and Caswell, H. Demography and dispersal: Calculation

and sensitivity analysis of invasion speed for structured populations. Ecology 81

(2000), 1613–1628.

[37] Neyman, J., Park, T., and Scott, E. Struggle for existence. The Tribolium

model: biological and statistical aspects. Proc. 3rd Berkeley Symp. Math. Statist.

Prob. 4 (1956), 41–79.

[38] Prus, T. Emigrational ability and surface numbers of adult beetles in 12 strains

of Tribolium confusum duval and T. castaneum herbst. Ekologia Polska Seria A

14 (1966), 548–588.

[39] Rabinowitz, P. H. Some global results for nonlinear eigenvalue problems.

Journal of Functional Analysis 7 (1971), 1291–1313.

[40] Ribes, M., Coma, R., Zabala, M., and Gili, J. Small scale spatial hetero-

geneity and seasonal variation in a population of a cave-dwelling Mediterannean

mysid. Journal of Plankton Research 18 (1996), 659–671.

[41] Shigesada, N., Kawasaki, K., and Teramoto, E. Spatial segregation of

interacting species. Journal of Theoretical Biology 1979 (1978), 83–99.

[42] Sokoloff, A., Albers, J., Cavataio, P., Mankau, S., McKibben, S.,

Mills, S., Perkins, R., Roberts, L., and Sandri, J. Observations on pop-

ulations of Tribolium brevicornis (Le conte)(Coleoptera, Tenebrionidae). I. Lab-

113

oratory observations of domesticated strains. Researches on Population Ecology

(1980).

[43] Sparrevik, E., and Leonardsson, K. Effects of large Saduria entomon

(isopoda) on spatial distribution of their small S. entomon and Monoporeia affinis

(amphipoda) prey. Oecologia 101 (1995), 177–184.

[44] Van Kirk, R. W., and Lewis, M. A. Integrodifference models for persistence

in fragmented habitats. Bulletin of Mathematical Biology 59 (1997), 107–137.

[45] Veit, R. R., and Lewis, M. A. Dispersal, population growth and the Allee

effect: Dynamics of the house finch invasion of eastern North America. American

Naturalist 148 (1996), 255 – 274.

[46] Wang, M.-H., Kot, M., and Neubert, M. G. Integrodifference equations,

Allee effects, and invasions. Journal of Mathematical Biology 44 (2002), 150–168.

[47] Zeidler, E. Nonlinear Functional Analysis and its Applications I. Springer,

Berlin, 1992.

[48] Ziegler, J. R. Evolution of the migration response: Emigration by Tribolium

and the influence of age. Evolution 30 (1976), 579–592.

[49] Zyromska-Rudzka, H. Abundance and emigrations of Tribolium in a labora-

tory model. Ekologia Polska Seria A 14 (1966), 491–518.