Skip to main content
Log in

The influence of host competition and predation on tick densities and management implications

  • ORIGINAL PAPER
  • Published:
Theoretical Ecology Aims and scope Submit manuscript

Abstract

Host community composition and biodiversity can limit and regulate tick abundance which can have profound impacts on the incidence and severity of tick-borne diseases. Our understanding of the relationship between host community composition and tick abundance is still very limited. Here, we present a novel mathematical model of a stage-structured tick population to study the influence of host behaviour and competition in the presence of heterospecifics and the influence of host predation on tick densities. We examine the influence of specific changes in biodiversity that modify the competition among and the predation on small and large host populations. We find that increasing biodiversity will not always reduce tick populations, but depends on changes in species composition affecting the degree and type competition among hosts, and the host the predation is acting on. With indirect competition, tick densities are not regulated by increasing biodiversity; however, with direct competition, increased biodiversity will regulate tick densities. Generally, we find that biodiversity will regulate tick densities when it affects tick-host encounter rates. We also find that predation on small hosts have a limited influence on reducing tick populations, but when the predation was on large hosts this increased the magnitude of tick population oscillations. Our results have tick-management implications: while controlling large host populations (e.g. deer) and adult ticks will decrease tick densities, measures that directly control the nymph ticks could also be effective.

This is a preview of subscription content, log in via an institution to check access.

Access this article

Price excludes VAT (USA)
Tax calculation will be finalised during checkout.

Instant access to the full article PDF.

Fig. 1
Fig. 2
Fig. 3
Fig. 4
Fig. 5
Fig. 6

Similar content being viewed by others

References

  • Barbour AG, Fish D (1993) The biological and social phenomenon of Lyme disease. Science 260(5114):1610–1616

    Article  CAS  PubMed  Google Scholar 

  • Bock R, Jackson L, De Vos A, Jorgensen W (2004) Babesiosis of cattle. Parasitology 129:S247–S269

    Article  PubMed  Google Scholar 

  • Brisson D, Dykhuizen DE, Ostfeld RS (2008) Conspicuous impacts of inconspicuous hosts on the Lyme disease epidemic. Proc R Soc London B 275:227–235

    Article  Google Scholar 

  • Brownstein JS, Holford TR, Fish D (2003) A climate-based model predicts the spatial distribution of the Lyme disease vector Ixodes scapularis in the United States. Environ Health Perspect 111(9):1152–1157

    Article  PubMed Central  PubMed  Google Scholar 

  • Brunner J, Ostfeld RS (2008) Multiple causes of variable tick burdens on small- mammal hosts. Ecology 89:2259–2272

    Article  PubMed  Google Scholar 

  • Brunner J, Duerr S, Keesing F, Killilea M, Vuong H, Ostfeld RS (2013) An experimental test of competition among mice, chipmunks, and squirrels in deciduous forest fragments. PLOS one 8(6): e66798

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  • Bull CM, Burzacott D (1993) The impact of tick load on the fitness of their lizard hosts. Oecologia 96:415–419

    Article  Google Scholar 

  • Caraco T, Gardner G, Maniatty W, Deelman E, Szymanski BK (1998) Lyme disease: Self-regulation and pathogen invasion. J Theor Biol 193(4):561–575

    Article  CAS  PubMed  Google Scholar 

  • Caraco T, Glavanakov S, Chen G, Flaherty JE et al (2002) Stage-structured infection transmission and a spatial epidemic: A model for Lyme disease. Am Nat 160(3):348–359

    Article  PubMed  Google Scholar 

  • Davidson DL, Morris DW (2001) Density-dependent foraging effort of deer mice Peromyscus maniculatus. Funct Ecol 15:575–583

    Article  Google Scholar 

  • Dobson A (2004) Population dynamics of pathogens with multiple host species. Am Nat 164:S64–S78

    Article  PubMed  Google Scholar 

  • Lyimo IN, Ferguson HM (2009) Ecological and evolutionary determinants of host species choice in mosquito vectors. Trends Parasitol 25(4):189–196

    Article  PubMed  Google Scholar 

  • Gaff HD, Gross LJ (2007) Modeling tick-borne disease: a metapopulation model. Bull Math Biol 69(1):265–288

    Article  PubMed  Google Scholar 

  • Giardina AR, Schmidt KA, Schauber EM, Ostfeld RS (2000) Modeling the role of songbirds and rodents in the ecology of Lyme disease. Can J Zool 78:2184–2197

    Article  Google Scholar 

  • Ghosh M, Pugliese A (2004) Seasonal population dynamics of ticks, and its influence on infection transmission: A semi-discrete approach. Bull Math Biol 66(6):1659–1684

    Article  PubMed  Google Scholar 

  • Gomes-Solecki MJC, Brisson DR, Dattwyler RJ (2006) Oral vaccine that breaks the transmission cycle of the Lyme disease spirochete can be delivered via bait. Vaccine 24:4440–4449

    Article  CAS  PubMed  Google Scholar 

  • Goodman JL, Dennis DT, Sonenshine DE (2005) Tick-borne diseases of humans. ASM Press, Washington, DC

    Book  Google Scholar 

  • Gratz NG (1999) Emerging and resurging vector-borne diseases. Annu Rev Entomol 44:51–75

    Article  CAS  PubMed  Google Scholar 

  • Hanincova K, Kurtenbach K, Diuk-Wasser M et al (2006) Epidemic spread of Lyme borreliosis, northeastern United States. Emerg Infectious Dis 12:604–611

    Article  Google Scholar 

  • Hersh MH, LaDeau SL, Previtali MA, Ostfeld RS (2014) When is a parasite not a parasite? Effects of larval tick burdens on white-footed mouse survival. Ecology 95(5):1360–1369

    Article  PubMed  Google Scholar 

  • Hobbs NT, Baker DL, Bear GD, Bowden DC (1996) Ungulate grazing in sagebrush grassland: Mechanisms of resource competition. Ecol Appl 6(1):200–217

    Article  Google Scholar 

  • Holt RD, Roy M (2007) Predation can increase the prevalence of infectious disease. Am Nat 169(5):690–699

    Article  PubMed  Google Scholar 

  • Irvine RJ (2006) Parasites and the dynamics of wildlife populations. Anim Sci 82:775–781

    Article  Google Scholar 

  • Jaenson TGT, Fish D, Ginsberg HS et al (1991) Methods for control of tick vectors of Lyme Borreliosis. Scand J Infectious Dis 4:151–157

    Google Scholar 

  • Jongejan F, Uilenberg G (2004) The global importance of ticks. Parasitology 129:S3–S14

    Article  PubMed  Google Scholar 

  • Jost L (2006) Entropy and diversity. Oikos 113:363–375

    Article  Google Scholar 

  • Keesing F, Holt RD, Ostfeld RS (2006) Effects of species diversity on disease risk. Ecol Lett 9:485–498

    Article  CAS  PubMed  Google Scholar 

  • Keesing F, Brunner J, Duerr S, Killilea M, LoGiudice K, Schmidt K, Vuong H, Ostfeld RS (2009) Hosts as ecological traps for the vector of Lyme disease. Proc R Soc B 276:3911–39119

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  • Labuda M, Nuttall PA (2004) Tick-borne viruses. Parasitology 129:S221–S245

    Article  CAS  PubMed  Google Scholar 

  • Lack D (1954) The natural regulation of animal numbers. Clarendon, Oxford

    Google Scholar 

  • Latham J (1999) Interspecific interactions of ungulates in European forests: an overview. For Ecol Manage 120(1-3):13–21

    Article  Google Scholar 

  • Li Y, Muldowney JS (1993) On Bendixson’s Criterion. J Differ Equ 106(1):27–39

    Article  Google Scholar 

  • Li MY, Muldowney JS (1996) Phase asymptotic semiflows, Poincare’s condition, and the existence of stable limit cycles. J Diff Equ 124(2):425–448

    Article  Google Scholar 

  • LoGiudice K, Ostfeld RS, Schmidt KA, Keesing F (2003) The ecology of infectious disease: Effects of host diversity and community composition on Lyme disease risk. PNAS 100(2):567–571

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  • LoGiudice K, Duerr S, Newhouse MJ, Schmidt K, Killilea ME, Ostfeld RS (2008) Impact of host community composition on lyme disease risk. Ecology 89(10):2841–2849

    Article  PubMed  Google Scholar 

  • Lou Y, Wu J (2014) Tick seeking assumptions and their implications for Lyme disease predictions. Ecol Comp 17:99–106

    Article  Google Scholar 

  • McKilligan NG (1996) Field experiments on the effect of ticks on breeding success and chick health of cattle egrets. Aust J Ecol 21:442–449

    Article  Google Scholar 

  • Mitchell WA, Abramsky Z, Kotler BP, Pinshow B, Brown JS (1990) The effect of competition on foraging activity in desert rodents - theory and experiments. Ecol 71(3):844–854

    Article  Google Scholar 

  • Munoz A, Bonal R (2007) Rodents change acorn dispersal behavior in response to ungulate presence. Oikos 116(10):1631– 1638

    Article  Google Scholar 

  • Murray JD (1989) Mathematical Biology, Springer, Biomathematics Vol. 19

  • Mwambi HG (2002) Ticks and tick-borne diseases in Africa: a disease transmission model. IMA J Math App Med Biol 19(4):275–292

    Article  Google Scholar 

  • Myers JH, Savoie A, van Randen E (1998) Eradication and pest management. Annu Rev Entomol 43:471–491

    Article  CAS  PubMed  Google Scholar 

  • Needham GR, Teel PD (1991) Off-host physiological ecology of Ixodid ticks. Annu Rev Entomol 36:659–681

    Article  CAS  PubMed  Google Scholar 

  • Norman R, Bowers RG, Begon M, Hudson PG (1999) Persistence of tick-borne virus in the presence of multiple host species: Tick reservoirs and parasite mediated competition. J Theor Biol 200(1):111–118

    Article  CAS  PubMed  Google Scholar 

  • Ogden NH, Bigras-Poulina M, O’Callaghanc CJ et al (2005) A dynamic population model to investigate effects of climate on geographic range and seasonality of the tick Ixodes scapularis. Int J Parasitol 35:375–389

    Article  CAS  PubMed  Google Scholar 

  • Ogden NH, Maarouf A, Barker IK (2006) Climate change and the potential for range expansion of the Lyme disease vector Ixodes scapularis in Canada. Int J Parasitol 36(1):63–70

    Article  CAS  PubMed  Google Scholar 

  • Ogden NH, Lindsay LR, Morshed M et al (2008) The rising challenge of Lyme borreliosis in Canada. Canada Communicable Dis Report 34:1–19

    CAS  Google Scholar 

  • Oliver JH (1989) Biology and systematics of ticks (Acari, Ixodida). Annu Rev Ecol Syst 20:397–430

    Article  Google Scholar 

  • Ostfeld RS, Keesing F (2000) Biodiversity and disease risk: the case of Lyme disease. Conserv Biol 14:722–728

    Article  Google Scholar 

  • Ostfeld RS, Holt RD (2004) Are predators good for your health? Evaluating evidence for top-down regulation of zoonotic disease reservoirs. Front Ecol Env 2(1):13–20

    Article  Google Scholar 

  • Ostfeld RS, Canham CD, Oggenfuss K et al (2006) Climate, deer, rodents, and acorns as determinants of variation in lyme-disease risk. PLoS Biol 4:e145–e155

    Article  PubMed Central  PubMed  Google Scholar 

  • Ostfeld RS (2011) Lyme disease: the ecology of a complex system. Oxford University Press, New York

    Google Scholar 

  • Perkins SE, Cattadori IM, Tagliapietra V et al (2006) Localised deer absence leads to tick amplification. Ecol 87(8):1981–1986

    Article  Google Scholar 

  • Randolph SE, Rogers DJ (1997) A generic population model for the African tick Rhipicephalus appendiculatus. Parasitology 115:265–279

    Article  PubMed  Google Scholar 

  • Randolph (2004) Tick ecology: Processes and patterns behind the epidemiological risk posed by Ixodid ticks as vectors. Parasitology 129:S37

    Article  PubMed  Google Scholar 

  • Rosá R, Pugliese A, Norman R, Hudson PJ (2003) Thresholds for disease persistence in models for tick-borne infections including non-viraemic transmission, extended feeding and tick aggregation. J Theor Biol 224 (3):359–376

    Article  PubMed  Google Scholar 

  • Sandberg S, Awerbuch TE (1992) A comprehensive multiple matrix model representing the life cycle of the tick that transmits the agent of Lyme disease. J Theor Biol 157:203–220

    Article  CAS  PubMed  Google Scholar 

  • Schmidt KA, Ostfeld RS (2001) Biodiversity and the dilution effect in disease ecology. Ecol 82(3):609–619

    Article  Google Scholar 

  • Schulze TL, Jordan RA, Hung RW, Schulze CJ (2009) Effectiveness of the 4-Poster Passive Topical Treatment Device in the Control of Ixodes scapularis and Amblyomma americanum (Acari: Ixodidae) in New Jersey. Vector-Borne and Zoonotic Dis 9:389–400

    Article  Google Scholar 

  • Sonenshine DE (2005) The Biology of Tick Vectors of Human Disease. in Goodman JLD, Sonenshine DE (eds) Tick-Borne Diseases of Humans. ASM Press. Washington DC:12–36

  • Sperling JLH, Sperling FAH (2009) Lyme borreliosis in Canada: biological diversity and diagnostic complexity from an entomological perspective. Can Entomol 141:521–549

    Article  Google Scholar 

  • Stanko M, Krasnov BR, Miklisova D, Morand S (2007) Simple epidemiological model predicts the relationships between prevalence and abundance in Ixodid ticks. Parasitology 134:59–68

    Article  CAS  PubMed  Google Scholar 

  • Van den Driessche P, Watmough J (2002) Reproduction numbers and sub-threshold endemic equilibria for compartmental models of disease transmission. Math Biosci 180(1-2):29–48

    Article  PubMed  Google Scholar 

  • Walker DH (1998) Tick-transmitted infectious diseases in the United States. Annu Rev Pub Health 19:237–269

    Article  CAS  Google Scholar 

  • White N, Sutherst RW, Hall N, Whish-Wilson P (2003) The vulnerability of the Australian beef industry to impacts of the cattle tick (Boophilus microplus) under climate change. Clim Change 61:157–190

    Article  Google Scholar 

  • Wikel S K (1996) Host immunity to ticks. Annu Rev Entomol 41:1–22

    Article  CAS  PubMed  Google Scholar 

  • Wonham MJ, Lewis MA, Renclawowicz J, Van Den Driessche P (2006) Transmission assumptions generate conflicting predictions in host-vector disease models: a case study in west Nile virus. Ecol Lett 9:706–725

    Article  PubMed  Google Scholar 

  • Wooton JT (1994) The nature and consequences of indirect effects in ecological communities. Annu Rev Ecol Syst 25:443–466

    Article  Google Scholar 

  • Yunger JA, Meserve PL, Guiterrez JR (2002) Small-mammal foraging behavior: Mechanisms for coexistence and implication for population dynamics. Ecol Monogr 71(4):561–577

    Article  Google Scholar 

Download references

Acknowledgments

This research is the direct result of the Pacific Institute for the Mathematical Sciences (PIMS) 11th Industrial Problem Solving Workshop held at the University of Alberta. The authors are grateful for the support given by PIMS and are particularly appreciative of the hard work put in by the local organizers. JT acknowledges the support of the PHARE training grant. CAC acknowledges the support of Royal Society grant TG090850 which funded a visit to work with JT. We also acknowledge the contributions during the initial development of the model from David Laferriere, Babak Pourziaei, Juan Ramirez, Marc D. Ryser, Wing Hung Sze, Hannah Dodd, Herb Freedman, and Ognjen Stancevic.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Christina A. Cobbold.

Appendices

Appendix A: Global stability of the extinction equilibrium and nonexistence of periodic orbits

Consider a system of differential equations d x/d t = f(x), where \(x=(x_{1},x_{2},x_{3})\in \mathbb {R}^{3}\) and x(t, x 0) is a solution of the equations which satisfies x(0, x 0) = x 0. We use a generalisation, to higher dimensions, of a criteria of Bendixson for the non-existence of invariant closed curves such as periodic or homoclinic orbits. The theory was developed by Li and Muldowney (1993, 1996) and shows that oriented infinitesimal line segments, y(t, y 0), evolve as solutions of

$$ \frac{dy}{dt}=\frac{\partial f}{\partial x}(x(t,x_{0}))y $$
(13)

and oriented infinitesimal areas, z(t, z 0) evolve as solutions of

$$ \frac{dz}{dt}=\frac{\partial f}{\partial x}^{[2]}(x(t,x_{0}))z $$
(14)

where \(\frac {\partial f}{\partial x}^{[2]}\) is the second additive compound matrix. For a general matrix A, the corresponding second additive compound matrix is given by A [2] as follows,

$$\begin{array}{@{}rcl@{}} A&=&\left[\begin{array}{ccc} a_{11} & a_{12} & a_{13} \\ a_{21} & a_{22} & a_{23} \\ a_{31} & a_{32} & a_{33} \end{array}\right],\\ A^{[2]}&=&\left[\begin{array}{ccc} a_{11}+a_{22} & a_{23} & -a_{13} \\ a_{32} & a_{11}+a_{33} & a_{12} \\ -a_{31} & a_{21} & a_{22}+a_{33} \end{array}\right]\;. \end{array} $$
(15)

Thus, for Eqs. (5)–(7), the second additive compound matrix is given by Eq. (16).

$$\phantom{r}\left(\frac{\partial f}{\partial x}\right)^{[2]}=\left(\begin{array}{ccc} -\mu_{1}-\mu_{2}-\frac{\gamma_{1}a_{1}}{(a_{1}+x_{1})^{2}}-\frac{\gamma_{2}a_{2}}{(a_{2}+x_{2})^{2}} & 0& -\frac{\alpha_{1}a_{3}}{(a_{3}+x_{3})^{2}} \\ \\ -\frac{\alpha_{3}a_{2}}{(a_{2}+x_{2})^{2}}&-\mu_{1}-\mu_{3}-\frac{\gamma_{1}a_{1}}{(a_{1}+x_{1})^{2}}-\frac{\gamma_{3}a_{3}}{(a_{3}+x_{3})^{2}}&0\\ \\ 0&\frac{\alpha_{2}a_{1}}{(a_{1}+x_{1})^{2}}&-\mu_{2}-\mu_{3}-\frac{\gamma_{2}a_{2}}{(a_{2}+x_{2})^{2}}-\frac{\gamma_{3}a_{3}}{(a_{3}+x_{3})^{2}} \end{array}\right)\;. $$
(16)

By Theorem 3.3 of Li and Muldowney (1993) if for each \(x_{0}\in \mathbb {R}^{3}_{+}\) (13) and (14) are uniformly asymptotically stable then all line segments collapse to the origin and we have global stability of (0, 0, 0) and there exists no invariant closed curves (periodic orbits, homoclinic or heteroclinic cycles) and the orbits converge to a single equilibrium.

Asymptotic stability of (13) and (14) is shown by constructing Lyapunov functions. Using the Lyapunov function V(x 1, x 2, x 3) = |x 1| + |x 2| + |x 3| and together with (13), we have

$$\begin{array}{@{}rcl@{}} \dot V(y)&=&(1,1,1)\cdot \frac{\partial f}{\partial x}=-\mu_{1}+\frac{a_{1}(\alpha_{2}-\gamma_{1})}{(a_{1}+x_{1})^{2}}\\ &-&\mu_{2}+\frac{a_{2} (\alpha_{3}-\gamma_{2})}{(a_{2}+x_{2})^{2}}-\mu_{3}+\frac{a_{3}(\alpha_{1}-\gamma_{3})}{(a_{3}+x_{3})^{2}} \end{array} $$

If \(\dot V(y)<0\), we have global stability of the zero solution of (13). Since γ 1α 2 and γ 2α 3, then a sufficient condition for \(\dot V(y)<0\) is μ 3 > (α 1γ 3)/a 3, condition (A) in Table 3. Showing that \(\dot V(y) = (1,1,1)\cdot \left (\frac {\partial f}{\partial x}\right )^{[2]}<0\) guarantees asymptotic stability of (14) and gives condition (C).

Table 3 Analytical criteria for tick eradication (A, B) and the absence of tick cycles (C, D) (see Appendix A for details of the derivations). Note that (A) and (B) are alternative criteria, only one of these needs to be satisfied and similarly for (C) and (D)

Alternatively, using the Lyapunov function V(x 1, x 2, x 3) = sup{|x 1|,|x 2|, |x 3|} gives stronger results (conditions B and D in Table 3).

Appendix B: \(\mathcal {R}_{0}\) and tick-borne disease dynamics

While ticks can feed on a variety of hosts, it is commonly believed that pathogens are associated with a particular host that acts as a disease reservoir that maintains the pathogen in the environment (Randolph 2004). For instance, the spirochete Borrelia burgdorferi s.l. is maintained mainly in deer mice: the spirochete is transferred to the tick when it feeds on an infected deer mouse; after which, the infected tick can transfer the disease to a human, causing Lyme disease, or to another deer mouse—thus maintaining the disease in the environment. If the tick feeds on an alternate small or large host that is not a disease reservoir (e.g., pocket mice, rabbits, humans), the pathogen will either be eliminated by the immune system, or lead to the death of the host, or not be transferred to another host, in all cases effectively acting as a dead end that removes the pathogen from the environment.

Larval ticks typically hatch free from infection and can acquire infection through a blood meal with an infected small host, at which point they molt to become infected nymphs. So larval ticks cannot transmit the disease. Infected nymphs can transmit the infection to the hosts they feed upon and the infection remains in the ticks when they molt to the adult stage. Adopting the approach of Lou and Wu (2014) we can extend our model in a simple way to capture the disease dynamics of Lyme disease by describing the disease status of the individuals in our model. The rate of change of infected small H 1 hosts \({H_{1}^{I}}(t)\), infected nymphs \({x_{2}^{I}}(t)\) and infected adult ticks \({x_{3}^{I}}(t)\) are given by Eqs. (19)–(20).

$$\begin{array}{@{}rcl@{}} \dot{{H_{1}^{I}}}&=&-\overbrace{\mu_{H_{1}} {H_{1}^{I}}}^{\text{\footnotesize{death}}}+\overbrace{\beta_{H}\frac{H_{1}-{H_{1}^{I}}}{H_{1}}\frac{\gamma_{2}^{\prime} {x_{2}^{I}}}{a_{2}+x_{2}}}^{\stackrel{\text{\footnotesize{infected nymphs transmitting}}}{\text{\footnotesize{disease to healthy hosts}}}}, \end{array} $$
(17)
$$\begin{array}{@{}rcl@{}} \dot{x}_{2}^{I}&=&-\overbrace{\mu_{2} {x_{2}^{I}}}^{\text{\footnotesize{death}}}+ \overbrace{\beta_{L}\frac{{H_{1}^{I}}}{H_{1}}\frac{\alpha_{2}^{\prime} x_{1}}{a_{1}+x_{1}}}^{\stackrel{\text{\footnotesize{larvae feeding on infected hosts}}}{\text{\footnotesize{molting to become infected nymph}}}} -\overbrace{\frac{\gamma_{2}^{\prime} {x_{2}^{I}}}{a_{2}+x_{2}}}^{\stackrel{\text{\footnotesize{infected nymphs molting}}}{\text{\footnotesize{to become infected adults}}}}, \end{array} $$
(18)
$$\begin{array}{@{}rcl@{}} \dot{{x_{3}^{I}}}&=&-\overbrace{\mu_{3} {x_{3}^{I}}}^{\text{\footnotesize{death}}}+\overbrace{\frac{\alpha_{3}^{\prime} (x_{2}-{x_{2}^{I}})}{a_{2}+x_{2}}}^{\stackrel{\text{\footnotesize{infected nymphs molting}}}{\text{\footnotesize{to become infected adults}}}} +\overbrace{\beta_{N}\frac{{H_{1}^{I}}}{H_{1}}\frac{\alpha_{3}^{\prime} (x_{2}-{x_{2}^{I}})}{a_{2}+x_{2}}}^{\stackrel{\text{\footnotesize{uninfected nymph feeding on infected hosts}}}{\text{\footnotesize{and molting to become infected adults}}}} -\overbrace{\frac{\gamma_{3} {x_{3}^{I}}}{a_{3}+x_{3}}}^{\text{\footnotesize{adults taking final blood meal}}}\;. \end{array} $$
(19)

We do not track infected large hosts as they can only transmit the infection to adult ticks which cannot pass the infection onto their offspring, so the large hosts are not acting as a reservoir for the disease the way that the small hosts are. We assume only the H 1 small hosts (e.g. deer mice) are a competent reservoir for the disease and that the H 2 small hosts are not (Ostfeld and Keesing 2000). β H , β L and β N are the transmission coefficients of the infection to H 1 hosts, larval ticks and nymphal ticks, respectively. \(\gamma _{i}^{\prime }\) is the contribution to γ i that comes from feeding on H 1 hosts only similarly for \(\alpha _{i}^{\prime }\). For example, \(\gamma _{2}^{\prime }=\sigma _{1}(p_{s})H_{1}\lambda _{2,1}\). Assuming the tick population are at equilibrium then, we can study the disease dynamics in isolation replacing x 1(t) and x 2(t) by their equilibrium values \(x_{1}^{*}\) and \(x_{2}^{*}\) and noting that the equation for infected adult ticks decouples. Hence, two equations form the epidemiological model,

$$\begin{array}{@{}rcl@{}} \dot{{H_{1}^{I}}}=&-\mu_{H_{1}} {H_{1}^{I}}+\beta_{H}\frac{H_{1}-{H_{1}^{I}}}{H_{1}}\frac{\gamma_{2}^{\prime} {x_{2}^{I}}}{a_{2}+x_{2}^{*}}, \end{array} $$
(20)
$$\begin{array}{@{}rcl@{}} \dot{x}_{2}^{I}=&-\mu_{2} {x_{2}^{I}}+ \beta_{L}\frac{{H_{1}^{I}}}{H_{1}}\frac{\alpha_{2}^{\prime} x_{1}^{*}}{a_{1}+x_{1}^{*}} -\frac{\gamma_{2}^{\prime} {x_{2}^{I}}}{a_{2}+x_{2}^{*}}. \end{array} $$
(21)

We can calculate the basic reproduction number for the disease using the next generation matrix method (see Van den Driessche and Watmough 2002). The transmission matrix and transition matrix are given by

$$ F= \left(\begin{array}{cc} 0 & \frac{\beta_{H}\gamma_{2}^{\prime}}{a_{2}+x_{2}^{*}} \\ \frac{\beta_{L}\alpha_{2}^{\prime}x_{1}^{*}}{H_{1}(a_{1}+x_{1}^{*})} & 0 \end{array}\right) \;\; \text{and}\;\; V =\left(\begin{array}{cc} \mu_{H_{1}} & 0 \\ 0 & \mu_{2}+\frac{\gamma_{2}^{\prime}}{a_{2}+x^{*}_{2}} \end{array}\right) $$
(22)

respectively. Together these yield the next generation matrix

$$ FV^{-1}= \left(\begin{array}{cc} 0 & \frac{\beta_{H}\gamma_{2}^{\prime}}{\mu_{2}(a_{2}+x_{2}^{*})+\gamma_{2}^{\prime}} \\ \frac{\beta_{L}\alpha_{2}^{\prime}x_{1}^{*}}{H_{1}(a_{1}+x_{1}^{*})\mu_{H_{1}}} & 0 \end{array}\right), $$
(23)

the dominant eigenvalue of which gives the basic reproduction number \(\mathcal {R}_{0}\) for the disease.

$$\begin{array}{@{}rcl@{}} \mathcal{R}_{0}&=&\sqrt{\frac{\beta_{H}\gamma_{2}^{\prime}}{\mu_{2}(a_{2}+x_{2}^{*})+\gamma_{2}^{\prime}} \frac{\beta_{L}\alpha_{2}^{\prime}x_{1}^{*}}{H_{1}(a_{1}+x_{1}^{*})\mu_{H_{1}}}}\\ &=&\sqrt{\frac{\beta_{H}\gamma_{2}^{\prime}}{\mu_{2}(a_{2}+x_{2}^{*})+\gamma_{2}^{\prime}} \frac{\beta_{L}\alpha_{2}^{\prime}}{H_{1}\alpha_{2}\mu_{H_{1}}}\left(\mu_{2}+\frac{\gamma_{2}}{a_{2}+x_{2}^{*}}\right)x_{2}^{*}} \end{array} $$
(24)

The unique endemic equilibrium is

$$\begin{array}{@{}rcl@{}} H_{1}^{I*}=H_{1}\left(1-\frac{1}{\mathcal{R}_{0}^{2}}\right) \end{array} $$
(25)
$$\begin{array}{@{}rcl@{}} x_{2}^{I*}= \beta_{L}\frac{\alpha_{2}^{\prime}}{\alpha_{2}}x_{2}^{*}\left(1+\frac{\gamma_{2}-\gamma_{2}^{\prime}}{\gamma_{2}^{\prime}+(a_{2}+x_{2}^{*})\mu_{2}}\right) \left(1-\frac{1}{\mathcal{R}_{0}^{2}}\right) \end{array} $$
(26)

Applying Theorem 2.1 from Lou and Jianhong (2014) shows that \(\mathcal {R}_{0}\) determines the global stability of the endemic equilibrium. Specifically, if \(\mathcal {R}_{0}>1\), the endemic equilibrium is globally asymptotically stable.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Cobbold, C.A., Teng, J. & Muldowney, J.S. The influence of host competition and predation on tick densities and management implications. Theor Ecol 8, 349–368 (2015). https://doi.org/10.1007/s12080-015-0255-y

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s12080-015-0255-y

Keywords

Navigation