Skip to main content

Advertisement

Log in

The Impact of Recruitment on the Dynamics of an Immune-Suppressed Within-Human–Host Model of the Plasmodium falciparum Parasite

  • Special Issue: Mathematical Epidemiology
  • Published:
Bulletin of Mathematical Biology Aims and scope Submit manuscript

Abstract

A model is developed and used to study within-human malaria parasite dynamics. The model integrates actors involved in the development–progression of parasitemia, gametocytogenesis and mechanisms for immune response activation. Model analyses under immune suppression reveal different dynamical behaviours for different healthy red blood cell (HRBC) generation functions. Existence of a threshold parameter determines conditions for HRBCs depletion. Oscillatory dynamics reminiscent of malaria parasitemia are obtained. A dependence exists on the type of recruitment function used to generate HRBCs, with complexities observed for a more nonlinear function. An upper bound that delimits the size of feasible parasitized steady-state solution exists for a logistic function but not a constant function. The upper bound is completely characterized and is affected by parameters associated with HRBCs recruitment, parasitized red blood cells generation and the release and time-to-release of free merozoites. A stable density size for mature gametocytes, the bridge to invertebrate hosts, is derived.

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
Fig. 7
Fig. 8
Fig. 9
Fig. 10
Fig. 11

Similar content being viewed by others

Notes

  1. Note that the estimate in Bianconi et al. (2013) was for total cell count. However, the given range can be deduced based on the percentage of cells that are HRBCs.

  2. \(C_{1}^{2}-4C_{0}=\left( D_{2} R_{0}^{2}-D_{2} D_{1}R_{0}+D_{2} D_{0}\right) \), where \(D_{2}R_{0}^{2}=\frac{\left( 1-\sigma \right) ^{2}}{\beta ^{2}}\), \(D_{2} D_{1}R_{0}=\frac{2a_{0}}{a_{1}}\frac{\left( 1-\sigma \right) }{\beta }+\frac{2\left( 1-\sigma \right) }{\beta ^{2}a_{1}}\) and \(D_{2} D_{0}=\frac{a_{2}^{2}\left( \beta a_{0}+1\right) ^{2}+4\beta a_{0}a_{2}a_{3}}{\beta ^{2}a_{1}^{2}a_{2}^{2}}\).

  3. Here, \(V^{\prime } = -Aa_{0}\left( 1-r_{h}\right) ^{2}\frac{1}{r_{h}}+\left( -Aa_{1}+Ba_{1}-Ca_{2}\right) r_{h}m+\left( Ca_{2}\left( 1-\sigma \right) -B+Da_{4}\sigma \right) r_{p}\).

References

  • An G, Widness JA, Mock DM, Veng-Pedersen P (2016) A novel physiology-based mathematical model to estimate red blood cell lifespan in different human age groups. AAPS J 18(5):1182–1191

    Google Scholar 

  • Anderson RM (1998) Complex dynamic behaviours in the interaction between parasite populations and the host’s immune system. Int J Parasitol 28(4):551–566

    Google Scholar 

  • Anderson RM, May RM (1979) Population biology of infectious diseases: Part I. Nature 280:361–367

    Google Scholar 

  • Anderson RM, May RM (1991) Infectious diseases of humans: dynamics and control. Oxford University Press, Oxford

    Google Scholar 

  • Anderson RM, May RM, Gupta S (1989) Non-linear phenomena in hostparasite interactions. Parasitology 99(S1):S59–S79

    Google Scholar 

  • Aron JL (1983) Dynamics of acquired immunity boosted by exposure to infection. Math Biosci 64:249–253

    MATH  Google Scholar 

  • Aron JL (1988a) Acquired immunity dependent upon exposure in an sirs epidemic model. Math Biosci 88:37–47

    MathSciNet  MATH  Google Scholar 

  • Aron JL (1988b) Mathematical modelling of immunity to malaria. Math Biosci 90(1):385–396

    MathSciNet  MATH  Google Scholar 

  • Augustine AD, Hall BF, Leitner WW, Mo AX, Wali Tonu M, Fauci Anthony S (2009) Niaid workshop on immunity to malaria: addressing immunological challenges. Nat Immunol 10(7):673–678

    Google Scholar 

  • Baron S (1996) Medical microbiolgy—galveston (tx). University of Texas Medical Branch at Galveston

  • Baton LA, Ranford-Cartwright LC (2005) Spreading the seeds of million-murdering death: metamorphoses of malaria in the mosquito. Trends Parasitol 21(12):573–580

    Google Scholar 

  • Bianconi E, Piovesan A, Facchin F, Beraudi A, Casadei Raffaella, Frabetti Flavia, Vitale Lorenza, Pelleri Maria Chiara, Tassani Simone, Piva Francesco et al (2013) An estimation of the number of cells in the human body. Ann Hum Biol 40(6):463–471

    Google Scholar 

  • Bichara D, Cozic N, Iggidr A (2012) On the estimation of sequestered parasite population in falciparum malaria patients. [Research Report] INRIA, RR-8178:22

  • Bousema T, Drakeley C (2011) Epidemiology and infectivity of plasmodium falciparum and plasmodium vivax gametocytes in relation to malaria control and elimination. Clin Microbiol Rev 24(2):377–410

    Google Scholar 

  • Bousema T, Sutherland CJ, Churcher TS, Mulder B, Gouagna Louis C, Riley Eleanor M, Targett Geoffrey AT, Drakeley Chris J (2011) Human immune responses that reduce the transmission of plasmodium falciparum in african populations. Int J Parasitol 41(3):293–300

    Google Scholar 

  • Brännström BÅ, Sumpter DJT (2005) The role of competition and clustering in population dynamics. Proc R Soc B 272:2065–2072

    Google Scholar 

  • Brookhaven National Labortory (BNL) (2017) 56 Facts About Blood and Blood Donation. https://www.bnl.gov/hr/blooddrive/56facts.asp. Accessed April 2017

  • Chiyaka C, Garira W, Dube S (2008) Modelling immune response and drug therapy in human malaria infection. Comput Math Method Med 9(2):143–163

    MathSciNet  MATH  Google Scholar 

  • Cowman AF, Berry D, Baum J (2012) The cellular and molecular basis for malaria parasite invasion of the human red blood cell. J Cell Biol 198(6):961–971

    Google Scholar 

  • Cuomo MJ, Noel LB, White DB (2009) Diagnosing medical parasites: a public health officers guide to assisting laboratory and medical officers. Technical report, DTIC Document

  • Dean L, National Center for Biotechnology Information (U.S.) (2005) Blood groups and red cell antigens. NCBI

  • Eichner M, Diebner HH, Molineaux L, Collins WE, Jeffery GM, Dietz K (2001) Genesis, sequestration and survival of plasmodium falciparum gametocytes: parameter estimates from fitting a model to malariatherapy data. Trans R Soc Trop Med Hyg 95(5):497–501

    Google Scholar 

  • Gardiner DL, Trenholme KR (2015) Plasmodium falciparum gametocytes: playing hide and seek. Ann Transl Med 3(4):45

    Google Scholar 

  • Ginsburg H, Hoshen MB (2002) Is the development of falciparum malaria in the human host limited by the availability of uninfected erythrocytes? Malar J 1(1):18

    Google Scholar 

  • Ginsburg H, Stein WD (1987) New permeability pathways induced by the malarial parasite in the membrane of its host erythrocyte: potential routes for targeting of drugs into infected cells. Biosci Rep 7(6):455–463

    Google Scholar 

  • Gottlieb Y, Topaz O, Cohen LA, Yakov LD, Haber Tom, Morgenstern Abigail, Weiss Avital, Berman Karen Chait, Fibach Eitan, Meyron-Holtz Esther G (2012) Physiologically aged red blood cells undergo erythrophagocytosis in vivo but not in vitro. Haematologica 97(7):994–1002

    Google Scholar 

  • Gravenor MB, Kwiatkowski D (1998) An analysis of the temperature effects of fever on the intra-host population dynamics of plasmodium falciparum. Parasitology 117(02):97–105

    Google Scholar 

  • Gravenor MB, Lloyd AL (1998) Reply to: Models for the in-host dynamics of malaria revisited: errors in some basic models lead to large over-estimates of growth rates. Parasitology 117(05):409–410

    Google Scholar 

  • Gurarie D, Karl S, Zimmerman PA, King CH, Pierre Timothy G St, Davis Timothy ME (2012) Mathematical modeling of malaria infection with innate and adaptive immunity in individuals and agent-based communities. PLoS One 7(3):e34040

    Google Scholar 

  • Heffernan JM (2011) Mathematical immunology of infectious diseases. Math Popul Stud 18(2):47–54

    MathSciNet  MATH  Google Scholar 

  • Hellriegel B (1992) Modelling the immune response to malaria with ecological concepts: short-term behaviour against long-term equilibrium. Proc R Soc Lond B Biol Sci 250(1329):249–256

    Google Scholar 

  • Hethcote HW, Stech HW, van den Driessche P (1982) Periodicity and stability in epidemic models: a survey. In: Busenberg S, Cooke KL (eds) Differential equations and applications in ecology, epidemics, and population problems. Academic Press, San Diego, pp 65–82

    Google Scholar 

  • Hetzel C, Anderson RM (1996) The within-host cellular dynamics of bloodstage malaria: theoretical and experimental studies. Parasitology 113(01):25–38

    Google Scholar 

  • Hoffman SL, Crutcher JM (2017) Malaria, Chapter 83. Medical Microbiology, Galveston (TX): University of Texas Medical Branch at Galveston, 4th ed, 1996. Accessed March 2017

  • Hollowell JG, Van Assendelft OW, Gunter EW, Lewis BG, Najjar M, Pfeiffer C (2005) Hematological and iron-related analytes-reference data for persons aged 1 year and over: United states, 1988–94. Vital Health Stat Ser 11 Data Natl Health Surv 247(247):1–156

    Google Scholar 

  • Hoshen MB, Heinrich R, Stein WD, Ginsburg H (2000) Mathematical modelling of the within-host dynamics of plasmodium falciparum. Parasitology 121(03):227–235

    Google Scholar 

  • Iggidr A, Kamgang J-C, Sallet G, Tewa J-J (2006) Global analysis of new malaria intrahost models with a competitive exclusion principle. SIAM J Appl Math 67(1):260–278

    MathSciNet  MATH  Google Scholar 

  • Ingemar N (1985) Lecture notes in biomathematics. Springer, Berlin

    Google Scholar 

  • Janeway CA Jr, Travers P, Walport M, Shlomchik MJ (2001) Immunobiology: the immune system in health and disease, 5th edn. Garland Science, New York. Available from: https://www.ncbi.nlm.nih.gov/books/NBK10757/

  • Josling GA, Llinás M (2015) Sexual development in plasmodium parasites: knowing when it’s time to commit. Nat Rev Microbiol 13(9):573–587

    Google Scholar 

  • Kaushal DC, Carter R, Miller LH, Krishna G (1980) Gametocytogenesis by malaria parasites in continuous culture. Nature 286(5772):490–2

    Google Scholar 

  • Kirk K (2001) Membrane transport in the malaria-infected erythrocyte. Physiol Rev 81(2):495–537

    Google Scholar 

  • Kiszewski Anthony E (2010) Blocking plasmodium falciparum malaria transmission with drugs: the gametocytocidal and sporontocidal properties of current and prospective antimalarials. Pharmaceuticals 4(1):44–68

    Google Scholar 

  • Kuehn A, Pradel G (2010) The coming-out of malaria gametocytes. BioMed Res Int 21(4):683–696

    Google Scholar 

  • Landaw SA (1987) Factors that accelerate or retard red blood cell senescence. Blood Cells 14(1):47–67

    Google Scholar 

  • Langhorne J (2006) Immunology and immunopathogenesis of malaria. Current topics in microbiology and immunology. Springer, Berlin

    Google Scholar 

  • Langhorne J, Ndungu FM, Sponaas A-M, Marsh K (2008) Immunity to malaria: more questions than answers. Nat Immunol 9(7):725–732

    Google Scholar 

  • Li Y, Ruan S, Xiao D (2011) The within-host dynamics of malaria infection with immune response. Math Biosci Eng 8(4):999–1018

    MathSciNet  MATH  Google Scholar 

  • McKenzie EF, Bossert WH (1997) The dynamics ofplasmodium falciparumblood-stage infection. J Theor Biol 188(1):127–140

    Google Scholar 

  • Mitri C, Thiery I, Bourgouin C, Paul REL (2009) Density-dependent impact of the human malaria parasite plasmodium falciparum gametocyte sex ratio on mosquito infection rates. Proc R Soc Lond B Biol Sci 276(1673):3721–3726

    Google Scholar 

  • National Institute of Allergy and Infectious Diseases (NIAID) (2010) The life cycle of the malaria parasite. https://www.cdc.gov/malaria/about/biology/index.html. Accessed Jan 2018

  • Ngonghala CN, Ngwa GA, Teboh-Ewungkem MI (2012) Periodic oscillations and backward bifurcation in a model for the dynamics of malaria transmission. Math Biosci 240(1):45–62

    MathSciNet  MATH  Google Scholar 

  • Ngonghala CN, Teboh-Ewungkem MI, Ngwa GA (2015) Persistent oscillations and backward bifurcation in a malaria model with varying human and mosquito populations: implications for control. J Math Biol 70(7):1581–1622

    MathSciNet  MATH  Google Scholar 

  • Ngonghala CN, Teboh-Ewungkem MI, Ngwa GA (2016) Observance of period-doubling bifurcation and chaos in an autonomous ode model for malaria with vector demography. Theor Ecol 9(3):337–351

    Google Scholar 

  • Ngwa CJ, de Rosa A, Thiago F, Pradel G (2017) The Biology of Malaria Gametocytes, chapter Current Topics in Malaria. InTech, 2016. Accessed March 2017

  • Ngwa GA, Teboh-Ewungkem MI (2016) A mathematical model with quarantine states for the dynamics of ebola virus disease in human populations. Comput Math Method Med, Vol 2016, Article ID 9352725, 93 pp

  • Okrinya A (2015) Mathematical modelling of malaria transmission and pathogenesis. PhD thesis, Loughborough University

  • Pearl R (1925) The biology of population growth. Alfred A. Knopf, New York

    Google Scholar 

  • Perlmann P, Troye-Blomberg M (2002) Malaria immunology, chemical immunology and allergy. Karger, Basel

    Google Scholar 

  • Rothman KJ, Greenland S, Lash TL (2008) Modern epidemiology. Lippincott Williams & Wilkins, Baltimore

    Google Scholar 

  • Sackmann E (1995) Biological membranes architecture and function. Struct Dyn Membr 1:1–63

    Google Scholar 

  • Shemin D, Rittenberg D (1946) The life span of the human red blood cell. J Biol Chem 166(2):627–636

    Google Scholar 

  • Sinden RE (1982) Gametocytogenesis of plasmodium falciparum in vitro: an electron microscopic study. Parasitology 84(01):1–11

    Google Scholar 

  • Sompayrac LM (2015) How the immune system works. John Wiley & Sons, New York

    Google Scholar 

  • Talman AM, Domarle O, McKenzie FE, Ariey F, Robert Vincent (2004) Gametocytogenesis: the puberty of plasmodium falciparum. Malar J 3(1):24

    Google Scholar 

  • Tavares JC (2013) Malaria. Colloquium series on integrated systems physiology: from molecule to function. Biota Publishing, Princeton

    Google Scholar 

  • Teboh-Ewungkem MI, Wang M (2012) Male fecundity and optimal gametocyte sex ratios for plasmodium falciparum during incomplete fertilization. J Theor Biol 307:183–192

    MathSciNet  MATH  Google Scholar 

  • Teboh-Ewungkem MI, Yuster T (2010) A within-vector mathematical model of plasmodium falciparum and implications of incomplete fertilization on optimal gametocyte sex ratio. J Theor Biol 264(2):273–286

    MathSciNet  MATH  Google Scholar 

  • Teboh-Ewungkem MI, Yuster T (2016) Evolutionary implications for the determination of gametocyte sex ratios under fecundity variation for the malaria parasite. J Theor Biol 408:260–273

    MathSciNet  MATH  Google Scholar 

  • Teboh-Ewungkem MI, Podder CN, Gumel AB (2010) Mathematical study of the role of gametocytes and an imperfect vaccine on malaria transmission dynamics. Bull Math Biol 72(1):63–93

    MathSciNet  MATH  Google Scholar 

  • Teboh-Ewungkem MI, Ngwa GA, Ngonghala CN (2013) Models and proposals for malaria: a review. Math Popul Stud 20(2):57–81

    MathSciNet  MATH  Google Scholar 

  • Tewa J-J, Fokouop R, Mewoli B, Bowong S (2012) Mathematical analysis of a general class of ordinary differential equations coming from within-hosts models of malaria with immune effectors. Appl Math Comput 218(14):7347–7361

    MathSciNet  MATH  Google Scholar 

  • Tumwiine J, Luckhaus S, Mugisha JYT, Luboobi LS (2008) An age-structured mathematical model for the within host dynamics of malaria and the immune system. J Math Model Algor 7(1):79–97

    MathSciNet  MATH  Google Scholar 

  • Tumwiine J, Mugisha JYT, Luboobi LS (2008) On global stability of the intra-host dynamics of malaria and the immune system. J Math Anal Appl 341(2):855–869

    MathSciNet  MATH  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):29–48

    MathSciNet  MATH  Google Scholar 

  • Verhulst PF (1838) Notice sur la loi que la population suit dans son acroissement. Correspondence Mathématiwue et Physique 10:113–121

    Google Scholar 

  • Wahlgren M, Perlmann P (1999) Malaria: molecular and clinical aspects. CRC Press, Boca Raton

    Google Scholar 

  • Weekley C, Smith DS (2013) Malaria: the clinical basics. Global Health Education Consortium (GHEC)

  • WHO (2015) World malaria report 2015. World Health Organisisation Bulletine

  • Willekens FLA, Werre JM, Groenen-Döpp YAM, Roerdinkholder-Stoelwinder B, De Pauw Ben, Bosman Giel JCGM (2008) Erythrocyte vesiculation: a self-protective mechanism? Br J Haematol 141(4):549–556

    Google Scholar 

  • Wongsrichanalai C, Barcus MJ, Muth S, Sutamihardja A, Wernsdorfer Walther H (2007) A review of malaria diagnostic tools: microscopy and rapid diagnostic test (rdt). Am J Trop Med Hyg 77(6 Suppl):119–127

    Google Scholar 

  • World Health Organization and Center for Disease Control (2010) Basic malaria microscopy: tutor’s guide. World Health Organization

Download references

Acknowledgements

The first author, WA, acknowledges support from the Department of Mathematics at Lehigh University, MIT-E’s Lehigh University Development fund, and Lehigh University as a whole, for supporting him and making available to him Lehigh’s resources, and for sponsoring as well as hosting him for more than two months as a visiting pre-doctoral scholar, enabling him to make significant progress on the work related to this manuscript and his thesis under the mentorship of MIT-E in conjunction with GAN via SKYPE. WA also acknowledges support from the African Institute for the Mathematical Sciences (AIMS) Cameroon that paid his flight for him to visit Lehigh University as a visiting pre-doctoral scholar, paving the path towards a successful completion of this manuscript and his doctoral dissertation. GAN acknowledges the grants and support of the Cameroon Ministry of Higher Education through the initiative for the modernization of research in Cameroon’s Higher Education. All three authors, WA, MIT-E and GAN acknowledge the support of the NSF -Directorate for Mathematical and Physical Science grant DMS-1544434 that created the opportunity for all three authors, who were present at the grant related activities (on School on Stochastic Analysis, Financial and Actuarial Mathematics with Applications) to meet as a unit for the first time and commence discussions on the manuscript and related project.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Miranda I. Teboh-Ewungkem.

Additional information

W. A. Woldegerima: Part of this work was done while the author was serving as Pre-doctoral visiting Scholar at Lehigh University.

Appendix

Appendix

1.1 Positivity and Positive Invariance Solution

Theorem 7

(Statement of the positivity and positive invariance of solution theorem) Consider system (8)–(14) with initial conditions in (15) and under the conditions given for \(\psi (R_h)\) and \(H_{i}(E_{i})\) as stated in Sect. 2.2. Then, every solution of the system with initial condition in \(\mathbb {R}_+^7\) remains in \(\mathbb {R}_+^7\). Additionally, if \(\varvec{x}(0)\equiv \varvec{0}\), the solution of system (8)–(14) will remain zero (or positively bounded depending on the form of \(\psi (R_{h})\)), for all time \(t>0\) . That is, \(\mathbb {R}_+^7\) is positively invariant and attracting with respect to the system. Furthermore, the system has a forward positive solution in \(\mathbb {R}_+^7 \) provided that it starts in it.

Proof

We show that the region \(\mathbb {R}^7_+\) is positively invariant, that is, whenever \(\varvec{x}(0)\in \mathbb {R}^{7}_{+}\), \(\varvec{x}\in \mathbb {R}^{7}_{+},\forall t\ge 0\). It suffices to show that there is no solution of the system starting in \(\mathbb {R}^7_+\) which is non-positive. Thus, we are required to show that the rate of change of each state variable, that is each \(\phi _i, ~1\le i\le 7,\) is non-negative at the origin \(\varvec{0}= (0, 0,0,0,0,0,0)\), and on each of the coordinate axis. Notice that at the origin \(\varvec{0}\), if \(\varvec{x}_0 = \varvec{0}\), then \(\varvec{x}'(0) = \Phi (\varvec{0})= \varvec{0}\) if \(\psi (R_{h}) = \Lambda -\tilde{\mu }_{h}R_{h}\), or \(\varvec{x}'(0) = (\Theta ,0,0,0,0,0,0)\) if \(\psi (R_{h})=\frac{\Theta }{R_{h}}\). Thus, if \(\varvec{x}(0)=\varvec{0}\), each component of \(\varvec{x}\) remains stationary at zero or increases from zero depending on the form of \(\psi \). On the other hand, if any one of the components of \(\varvec{x}\) is zero, the rate of change of that component with time is non-negative, showing that no trajectory of the system passes out of \(\mathbb {R}^{7}_{+}\) through that component’s zero axes. For example, when \(R_{p}=0\), \(R_{p}'=\frac{\beta _{1} R_{h}M}{1+\xi _0 E_a}\ge 0\), since \(R_{h}\) and M are non-negative for all time, showing that no solution of the system passes out of \(\mathbb {R}^{7}_{+}\) through the \(R_{p}=0\) axis. This implies the vector field of the system is inward pointing on the boundary of \(\mathbb {R}^{7}_{+}.\) That is, if \( \varvec{x}_0 \in \mathbb {R}^7_+,\) then \( \varvec{x} \in \mathbb {R}^7_+, ~\forall t\ge 0.\) Therefore, the region \( \mathbb {R}^{7}_{+} \) is positively invariant and attracting.

Next to prove the positivity of the solution, we follow the steps in Ngwa et al. (2016), Page 8. Suppose there exists \(t_1 >0\) such that \(R_h (t_1) =0, ~~ R_h'(t_1) < 0\) and \(R_h, R_p, M, G_e, G_l, E_i, E_a > 0\) for all \( 0<t < t_1.\) Then

$$\begin{aligned} R'_h (t_1) =\underbrace{R_{h}(t_{1})\psi (R_h(t_1))}_{:=\psi _{0}} -\underbrace{\mu _h R_h(t_1)}_ { :=0}-\underbrace{\frac{\beta _{1} R_h(t_1)M(t_1)}{1+\xi _0 E_a(t_1)}}_{:=0} = \left\{ \begin{array}{lcl} \Theta &{} \text{ if } &{} \psi (R_{h})=\frac{\Theta }{R_{h}}\\ 0 &{} \text{ if } &{}\psi (R_{h}) = \Lambda -\tilde{\mu }_{h}R_{h}.\end{array}\right. \end{aligned}$$

In either case, \(R'_h (t_1)\ge 0\), leading to a contradiction to the assumption that \(R'_h(t_1) < 0.\) So, no such \(t_1\) exists, and hence \( R_h \ne 0\). Thus, \( R_h (t) > 0, ~~\forall t\ge 0.\) Next suppose that there exists \(t_2\) such that \(R_p(t_2)=0, R'_p (t_2) <0\) and \(R_h, R_p, M, G_e, G_l > 0\) for all \( 0 <t \le t_2.\) Then, \(R'_p (t_2)= \frac{\beta _{1} R_h(t_2)M(t_2)}{1+\xi _0 E_a(t_2)} -\underbrace{(\gamma _{p}+\mu _p)R_p(t_2)}_{:0} = \frac{\beta _{1} R_h(t_2)M(t_2)}{1+\xi _0 E_a(t_2)} >0, \) which is a contradiction to the assumption that \(R'_p (t_2) < 0.\) Hence, \( R_p(t)> 0, ~~\forall t>0.\) Similarly, one can show that \( M(t)> 0,~ G_e (t)> 0, ~G_l (t)>0, E_i(t) >0\) and \( E_a (t) >0\) for all \( t>0.\) Therefore, any solution of the system with an initial condition in \(\mathbb {R}_+^7\) is positive. \(\square \)

1.2 Boundedness of Solution

Theorem 8

(Statement of the boundedness of solution theorem) Consider system (8)–(14) with initial conditions in (15) and under the conditions for \(\psi (R_h)\) and \(H_{i}(E_{i})\) as stated in Sect. 2.2. Then, every forward solution of the system in \(\mathbb {R}_+^7\), with initial condition in \(\mathbb {R}_+^7\), is bounded. Moreover, the system is uniformly dissipative in \(\mathbb {R}_+^7.\)

Proof

To start the proof of boundedness, we first note the following about boundedness of \(f(R_h)\) and \(H_i(E_i).\)

  1. 1.

    For all values of \(R_{h}\), we have

    $$\begin{aligned} R_{h}\psi (R_{h}) \le \mathcal {K_R}, \text{ where } \mathcal {K_R}=\left\{ \begin{array}{lll} \Theta &{}\quad \text{ if }&{}\psi (R_{h})=\frac{\Theta }{R_{h}}\\ \frac{\Lambda ^{2}}{4\tilde{\mu }_{h}}&{}\quad \text{ if }&{}\psi (R_{h})=\Lambda -\tilde{\mu }_{h}R_h.\end{array}\right. \end{aligned}$$
    (61)

    We note that the requirement that \(\psi \) be monotone non-increasing tacitly comes along with the requirement that \(R_{h}\psi (R_{h})\) be continuous from right at the origin. In particular, \(\psi (R_{h})\) satisfies conditions (1)–(3) of Sect. 2.2. Other examples, besides the two studied in this manuscript, of recruitment functions \(\psi (R_{h})\) found in the biological literature that satisfy conditions (1)–(3) of Sect. 2.2 are \(\psi (R_{h})=\Lambda e^{-\tilde{\mu }_{h}R_{h}}\), Ricker recruitment function and \(\psi (R_{h})=\frac{\Lambda }{1+ \left( \frac{R_{h}}{L}\right) ^{n}}\), \(\Lambda ,L,n,\tilde{\mu }_{h}>0\) which is the Maynard–Smith–Slatkin function. Details on these types of recruitment functions can be found in Brännström and Sumpter (2005) and Ngonghala et al. (2016).

  2. 2.

    Similarly, for all values of \(E_{i}\) we have

    $$\begin{aligned}&H_{i}(E_{i})\le \mathcal {K}_{i} \text{ where } \mathcal {K}_{i} \nonumber \\&\quad = \left\{ \begin{array}{lcl} \frac{\delta _{i}K_{i}}{4}&{}\quad \text{ if }&{}\quad H(E_{i})=\delta _{i}E_{i}\left( 1-\frac{E_{i}}{K_{i}}\right) \\ \max (A_{1},A_{2},0)&{}\quad \text{ if }&{}\quad H_{i}(E_{i})=\delta _{i}E_{i}\left( 1-\frac{E_{i}}{K_{i}}\right) \left( \frac{E_{i}}{M_{i}}-1\right) \end{array}\right. \end{aligned}$$
    (62)

    where on setting \(B = -K_i M_i+K_i^2+M_i^2=\left( M_{i}-\frac{1}{2}K_{i}\right) ^2 + \frac{3}{4}K_{i}^2>0\) we can obtain

    $$\begin{aligned} A_{1}= & {} \frac{\delta _i \left( -\sqrt{B}+K_i-2 M_i\right) \left( -\sqrt{B}+K_i+M_i\right) \left( \sqrt{B}+2 K_i-M_i\right) }{27 K_i M_i}\\ A_{2}= & {} -\frac{\delta _i \left( \sqrt{B}+K_i-2 M_i\right) \left( \sqrt{B}-2 K_i+M_i\right) \left( \sqrt{B}+K_i+M_i\right) }{27 K_i M_i}. \end{aligned}$$

    Thus, the functions \(H_{i}\) and \(R_{h}\psi \) defined are bounded.

Now to prove the boundedness of the \(R_h\) and \(R_p,\) let \(R (t)= R_{h}(t)+R_{p}(t)\) be the total size of red blood cells within the human at time t, (healthy plus infected red blood cells) with \(R (0)= R_{h}(0)+R_{p}(0)=R(0)\). Then, we have from the first two equations of system (8)–(14)

$$\begin{aligned} \frac{\mathrm{d} R}{\mathrm{d} t}= & {} R_{h}\psi (R_{h})-\mu _{h}R_{h}-(\gamma _{p}+\mu _p)R_p-(\rho _{p}+\rho _a E_a)R_pE_i\\\le & {} f(R_{h})-\mu R, \text{ where } \mu = \min (\mu _{h},\gamma _{p}+\mu _p), \end{aligned}$$

where \(f(R_{h}) = R_{h}\psi (R_{h})\) with \(\psi :[0,\infty )\rightarrow \mathbb {R}_+\) a monotone decreasing continuously differentiable function. So, the function \(f:[0,\infty )\rightarrow \mathbb {R}_+\) has a maximum value which is either constant when f is the constant function, or that occurs at the point \(R_{h}^{*}\in [0,\infty )\), where \(R_{h}^{*}\) satisfies the equation \(f'(R_{h}^{*})=\psi (R_{h}^{*})+R_{h}^{*}\psi '(R_{h}^{*})=0\). Set \(\mu =\min (\mu _{h},\gamma _{p}+\mu _p)\) and suppose that the maximum value of f is \(\mathcal {K}_{R}\), then we have from above,

$$\begin{aligned} \frac{\mathrm{d} R}{\mathrm{d} t} +\mu R\le & {} \mathcal {K}_{R}\Rightarrow R(t)\le \frac{\mathcal {K}_{R}}{\mu }+ A e^{-\mu t}, \end{aligned}$$

where A is an arbitrary constant that can be determined from initial data. Observe that if the initial condition, R(0), is such that \(R(0)> \frac{\mathcal {K}_{R}}{\mu }\), then A is always positive and the bound for R(t) is decreasing with time. When \(R(0)= \frac{\mathcal {K}_{R}}{\mu }\), then A is non-negative and the bound for R(t) is non-increasing with time. Finally, if \(R(0)< \frac{\mathcal {K}_{R}}{\mu }\), A can be a negative number and the bound for R(t) will be an increasing function of t. If at any of the instances we see that

$$\begin{aligned} \lim _{t\rightarrow \infty }{\sup {R(t)}}\le \frac{\mathcal {K}_{R}}{\mu }. \end{aligned}$$
(63)

Thus, \(0\le R_{h}(t)+R_{p}(t)\le \frac{\mathcal {K}_{R}}{\mu }, ~\forall t\ge 0.\) So there exist \(R_{h}^{\infty }\) and \(R_{p}^{\infty }\) with the property that \(0\le R_{h}(t)\le R_{h}^{\infty }\) and \(0\le R_{p}(t)\le R_{p}^{\infty }\), \(\forall t\ge 0\). Hence, \(R_h\) and \(R_p\) are bounded solutions.

Next we consider the equation for M, namely,

$$\begin{aligned} \frac{\mathrm{d} M}{\mathrm{d}t}= & {} \frac{r \gamma _{p}(1-\sigma ) R_p}{1+\xi _1 E_a (t)} - \mu _m M-\left( \frac{\beta _{2} R_h}{1+\xi _0E_a} + \frac{\beta _{3} R_p}{1+\xi _0E_a} +(\rho _m +\rho _{n} E_{a})E_{i}\right) M, \end{aligned}$$

and observe that when we take into consideration the fact that the quantity \(\frac{1}{1+\xi _1 E_a }\) is largest when \(E_{a}=0\), we have that

$$\begin{aligned} \frac{\mathrm{d} M}{\mathrm{d} t}\le & {} r \gamma _{p}(1-\sigma ) R_p^{\infty } - \mu _{m}M \Rightarrow M(t)\le \frac{ r \gamma _{p}(1-\sigma ) R_p^{\infty }}{\mu _{m}}+ B e^{-\mu _{m} t}, \end{aligned}$$

where B is an arbitrary constant. As above we arrive at the conclusion that there exist \(M^{\infty }\) such that \(0\le \sup {M(t)}\le M^{\infty },~\forall t\ge 0\). So M is bounded.

Next to prove the boundedness of \(G_e\) and \(G_l\), we set \(G(t)=G_{e}(t)+G_{l}(t)\) to be the total size of gametocytes within the human and see that

$$\begin{aligned} \frac{\mathrm{d} G}{\mathrm{d} t}= & {} \frac{s \sigma \gamma _{p} R_p}{1+\xi _1 E_a}-\left( \gamma _{l}+ \mu _e\right) G_e -(\rho _g + \rho _{q}E_a)G_{e}E_{i} + \frac{\gamma _{l} G_e}{1+\xi _1 E_a }- \mu _l G_l -\rho _l E_i G_l\\\le & {} s \sigma \gamma _{p} R_p^{\infty }-\min (\mu _{e},\mu _{l})G. \end{aligned}$$

Thus

$$\begin{aligned} \dfrac{\mathrm{d} G}{\mathrm{d}t}+\min (\mu _{e},\mu _{l})G\le r \sigma \gamma _{p,m} R_p^{\infty }\Rightarrow G(t)\le \frac{r \sigma \gamma _{p,m} R_p^{\infty }}{\min (\mu _{e},\mu _{l})}+ C e^{-\min (\mu _{e},\mu _{l}) t}. \end{aligned}$$

Therefore, as before, there exist \(G_{l}^{\infty }\) and \(G_{e}^{\infty }\) with the property that for \(0\le G_{e}(t)\le G_{e}^{\infty }\) and \(0\le G_{l}(t)\le G_{l}^{\infty }\), \(\forall t\ge 0\). So \(G_{e}\) and \(G_{l}\) are bounded whenever the preceding variables are bounded.

To establish boundedness of the solutions for the equations of the innate and adaptive immune responses, we proceed as follows. From the last two equations of the general model, system (8)–(14), and using the above results, we get

$$\begin{aligned} \dfrac{\mathrm{d} E_i}{\mathrm{d}t}+\left( \lambda _{1} R_p^{\infty }+ \lambda _{2}M^{\infty } \right) E_i\le & {} \mathcal {K}_{i} + \vartheta _1 R_p^{\infty } + \vartheta _{2}M^{\infty }\\ \dfrac{\mathrm{d} E_a}{\mathrm{d}t}+\left( \mu _a+\theta _{1} R_p^{\infty }+\theta _{2} M^{\infty }) \right) E_a\le & {} \varrho _{1}R_p^{\infty } + \varrho _{2}M^{\infty }, \end{aligned}$$

with the right hand side here being only constants and we can again argue as above to come to the conclusion that each \(E_{i}\) will show bounded growth whenever M, \(R_{h}\) and \(R_{P}\) are bounded. This completes the prove for boundedness. So, if we let

$$\begin{aligned} B^{\infty } = \max \{ R_h^{\infty },R_h^{\infty }, M^{\infty }, G_e^{\infty }, G_l^{\infty }, E_i^{\infty } , E_a^{\infty } \}, \end{aligned}$$

then each of \(R_h,~R_p, ~M, ~G_e, ~G_l, ~E_i, ~E_a \le B^{\infty }.\) In the absence of disease, system (8)–(14) reduces to the decoupled equations for the healthy red blood cell population and for the immune cells as follows:

$$\begin{aligned} \frac{\mathrm{d} R_{h}}{\mathrm{d} t}= & {} R_{h}\psi (R_{h})-\mu _{h}R_{h}, ~~\frac{\mathrm{d} E_i}{\mathrm{d}t}= H_i(E_i), ~~ \frac{\mathrm{d} E_a}{\mathrm{d}t}= -\mu _{a}E_{a}. \end{aligned}$$

Note here that in the absence of Allee effect, \(H_i(E_i)\) can have similar forms as \(F(R_h).\) That is, we can write

$$\begin{aligned} H_i(E_i)= E_i\varphi (E_i)- \mu _i E_i, \end{aligned}$$

where \(\varphi :[0, \infty )\rightarrow \mathbb {R}_+\) is a function defined similarly as \(\psi \) and satisfies the conditions stated for \(\psi .\)

So, to prove the boundedness of these functions we observe that the equation for the healthy red blood cell population then satisfies the relation

$$\begin{aligned} \frac{\mathrm{d} R_{h}}{\mathrm{d} t} = R_{h}\psi (R_{h})-\mu _{h}R_{h}\Rightarrow t(R_{h}) = \int \frac{1}{R_{h}(\psi (R_{h})-\mu _{h})}\mathrm{d} R_{h} + C, \end{aligned}$$

where C is an arbitrary constant of integration. For the functional forms of \(\psi \) used here, if at time \(t=0, R_{h}(t)=R_{0h}\), we have

$$\begin{aligned} t(R_{h}) = \left\{ \begin{array}{lll} \frac{1}{\mu _{h}}\ln \left( \frac{\Theta -\mu _{h}R_{0h}}{\Theta -\mu _{h}R_{h}}\right) &{}\quad \text{ if }&{}\psi (R_{h})= \frac{\Theta }{R_{h}}\\ \frac{1}{\omega }\ln \left( \frac{R_{h}(R_{0h}-K)}{R_{0h}(R_{h}-K)}\right) &{}\quad \text{ if }&{}\psi (R_{h})=\Lambda -\tilde{\mu }_{h}R_{h}\end{array}\right) , \end{aligned}$$

where \(\omega = \Lambda -\mu _{h} \text{ and } K=\frac{\omega }{\tilde{\mu }_{h}}\). For both forms of recruitment, it is clear that \(t(R_{h})\rightarrow \infty \) when \(R_{h}\rightarrow \Theta /\mu _{h}\) or \(R_{h}\rightarrow K\), respectively. So the solutions remain bounded. Also, in the absence of disease, the expression for the innate immunity at any time can be written as an exact integral. That is

$$\begin{aligned} \frac{\mathrm{d} E_{i}}{\mathrm{d} t} = H_{i}(E_{i})\Rightarrow t(E_{i}) = \int \frac{1}{H_{i}(E_{i})}\mathrm{d}E_{i} +C, \end{aligned}$$
(64)

where C is a constant whose values can be determined by the initial conditions. So,

$$\begin{aligned} t(E_{i}) = \left\{ \begin{array}{lcl} \frac{K_i \left( \ln \left( E_{i}-M_i\right) -\ln (E_{i})\right) +M_i \left( \ln (E_{i})-\ln \left( E_{i}-K_i\right) \right) }{\delta _i \left( K_i-M_i\right) } +C &{}\quad \text{ if }&{}H_{i}(E_{i})=E_{i} \delta _i \left( 1-\frac{E_{i}}{K_i}\right) \left( \frac{E_{i}}{M_i}-1\right) \\ \frac{\ln (E_{i})-\ln \left( E_{i}-K_i\right) }{\delta _i}+ C&{}\quad \text{ if }&{}H_{i}(E_{i})=E_{i} \delta _i \left( 1-\frac{E_{i}}{K_i}\right) \end{array}\right) , \end{aligned}$$

so that if at time \(t=0\), \(E_{i}(0) =E_{0i}\), we have the implicit solution

$$\begin{aligned} t(E_{i}) = \left\{ \begin{array}{lcl} \frac{1}{\delta _{i}(K_{i}-M_{i})}\ln \left( \left( \frac{E_{0i}(E_{i}-M_{i})}{E_{i}(E_{0i}-M_{i})}\right) ^{K_{i}}\left( \frac{E_{i}(E_{0i}-K_{i})}{(E_{i}-K_{i})E_{0i}}\right) ^{M_{i}}\right) &{}\text{ if }&{}H_{i}(E_{i})=E_{i} \delta _i \left( 1-\frac{E_{i}}{K_i}\right) \left( \frac{E_{i}}{M_i}-1\right) \\ \frac{1}{\delta _{i}}\ln \left( \frac{E_{i}(E_{0i}-K_{i}}{E_{0i}(E_{i}-K_{i})}\right) &{}\text{ if }&{}H_{i}(E_{i})=E_{i} \delta _i \left( 1-\frac{E_{i}}{K_i}\right) \end{array}\right. \end{aligned}$$

We then see clearly that for the logistic case, \(t\rightarrow \infty \) whenever \(E_{i}\rightarrow K_{i}\) for any starting value of \(E_{0i}>0\). In the case with the Allee effect, if \(0<M_{i}<K_{i}\) then \(0<E_{0i}<M_{i}\), \(t(E)\rightarrow \infty \) as \(E_{i}\rightarrow 0+\), while if \(E_{0i}>M_{i}\), then again, \(t(E)\rightarrow \infty \) as \(E_{i}\rightarrow K_{i}\). This shows that in either case, the solutions remain bounded. The inverse function theorem can be applied to obtain the solution \(E_{i}(t)\) in some special cases of values of \(M_{i}\) and \(K_{i}\). We have thus established boundedness of the solutions in all cases in both the presence and absence of the infection. \(\square \)

1.3 Uniqueness of Solution

Theorem 9

(Statement on the Uniqueness of Solution) The positive and bounded solution for the system (8)–(14) whenever it exists, is unique.

Proof

We show that the function \(\Phi \) defined above is globally Lipschitz in \(\mathbb {R}_+^7\) and hence the equation \(\varvec{x}'(t)=\Phi (\varvec{x}(t)), \varvec{x}(0)=\varvec{x}_{0}\) has a unique solution. It is clear that \(\mathbb {R}_+^7\) is a convex set, \(\Phi \) is continuously differentiable, since the partial derivatives \( \frac{\partial \Phi }{\partial x_i}, ~~i=1,2,\ldots ,7\) exist, and are continuous. We show that these partial derivatives are bounded in \(\mathbb {R}_+^7:\) Since \(R_h \psi (R_h)\) is monotone decreasing, continuously differentiable function and each state variable \(R_h, R_p, M, G_e, G_l, E_i, E_a\) are continuously differentiable, then each component \(\Phi _i,~~i=1,2,3\ldots ,7\) of the vector valued function \(\Phi \) on right hand side of system (8)–(14) exists and are continuously differentiable because they are rational functions of the state variables. It suffices to show that \(\Vert \frac{\partial \Phi }{\partial x_{i}}\Vert _{\infty }\)\(i=1,2, \cdots , 7\) are bounded where \((x_1,x_{2},x_{3},x_{4},x_{5},x_{6},x_{7})=(R_h, R_p, M, G_e, G_l, E_i, E_a)\). Observe, for example, that

$$\begin{aligned} \left\| \frac{\partial \Phi }{\partial R_{h}}\right\| _{\infty }= & {} \max \left\{ \left| \frac{\partial \phi _{i}}{\partial R_{h}}\right| , i= 1,2,\ldots ,7\right\} \\= & {} \max \left\{ \left| \psi (R_{h})+R_{h}\psi '(R_{h})-\mu _{h}-\frac{\beta _{1}M}{1+\xi _{0}E_{a}}\right| , ~\left| \frac{\beta _{1}M}{1+\xi _{0}E_{a}}\right| ,0,0,0,0,0\right\} \\\le & {} \left| \psi (R_{h})|+R_{h}|\psi '(R_{h})\right| + \left| \mu _{h}+\frac{\beta _{1}M}{1+\xi _{0}E_{a}}\right| \\\le & {} B_{1} \end{aligned}$$

for some \(B_{1}\), since M and \(E_{a}\) are bounded, and \(\psi \) is monotone decreasing so that \(|\psi '|\) is monotone increasing and bounded by say K where K is the carrying capacity for \(R_{h}\), so \(B_{1}\) exists. Similarly, there exists \(B_i < \infty \), for \(i=2, 3, \cdots , 7\) such that

$$\begin{aligned} \left\| \frac{\partial \Phi }{\partial R_p}\right\| _{\infty }= & {} B_2< \infty , ~\left\| \frac{\partial \Phi }{\partial M}\right\| _{\infty }= B_3< \infty , \left\| \frac{\partial \Phi }{\partial G_e}\right\| _{\infty } = B_4<\infty ,\\ \left\| \frac{\partial \Phi }{\partial G_l}\right\| _{\infty }= & {} B_5<\infty ,~~\left\| \frac{\partial \Phi }{\partial E_i}\right\| _{\infty }= B_6<\infty , ~\left\| \frac{\partial \Phi }{\partial E_{a}}\right\| _{\infty }= B_7 <\infty . \end{aligned}$$

We would have established that the partial derivatives are bounded and hence the function \(\Phi (\varvec{x})\) defined by the right hand side of (8)–(14) is Lipschitzian. Now let \(\varvec{x}_1, ~\varvec{x}_2 \) be two arbitrary points in \(\mathbb {R}^7_+.\) Then define,

$$\begin{aligned} \mathbf z (\varvec{x}_1, \varvec{x}_2; \theta ) = \{ \varvec{x}_1+\theta (\varvec{x}_2-\varvec{x}_1),~~0\le \theta \le 1 \} . \end{aligned}$$

Then, \(\mathbf z (\varvec{x}_1, \varvec{x}_2; \theta )\) is a line segment joining points \(\varvec{x}_1\) and \(\varvec{x}_2\) in \(\mathbb {R}_+^7\) for \(\theta \in [0, 1].\) Furthermore, \(\mathbf z (\varvec{x}_1, \varvec{x}_2; \theta )\) is a convex function and since \(\mathbb {R}_+^7\) is a convex set, then \(\mathbf z (\varvec{x}_1, \varvec{x}_2; \theta ) \in \mathbb {R}_+^7\) for each \(\theta \in [0, 1].\)

Using the mean value theorem for differentiable functions in \(\mathbb {R}_+^n\), one can show that

$$\begin{aligned} \Vert \Phi (\varvec{x}_1)- \Phi (\varvec{x}_2)\Vert _{\infty } =\Vert D_{\Phi }(\varvec{c}; \varvec{x}_1-\varvec{x}_2)\Vert _{\infty }, \end{aligned}$$

where \(\mathbf c \) is the mean value point and \(D_{\Phi }\) is the directional derivative of \(\Phi \) at the point \(\mathbf c \) in the direction of the vector \(\varvec{x}_1 -\varvec{x}_2.\) Using the expression for the directional derivative, as well as applying the triangle inequality and the Cauchy–Schwartz inequality, we see that

$$\begin{aligned} \Vert D_{\Phi }( \mathbf z ; \varvec{x}_1 - \varvec{x}_2) \Vert _{\infty }= & {} \left\| \sum _{k=1}^7 \nabla \Phi _k (\mathbf z ) \cdot (\varvec{x}_1 - \varvec{x}_2)\mathbf e _k \right\| _{\infty } \\ {}\le & {} \left\| \sum _{k=1}^7 \nabla \Phi _k (\mathbf z ) \right\| _{\infty } \Vert (\varvec{x}_1 - \varvec{x}_2)\Vert _{\infty }\\\le & {} \sum _{k=1}^7 \Vert \nabla \Phi _k (\mathbf z ) \Vert _{\infty } \Vert (\varvec{x}_1 - \varvec{x}_2)\Vert _{\infty } \le \mathcal {L} \Vert (\varvec{x}_1 - \varvec{x}_2)\Vert _{\infty }, \end{aligned}$$

for some constant \(\mathcal {L} = 7\max \{B_{1},B_{2},B_{3},B_{4},B_{5},B_{6},B_{7}\}\) where the last inequality comes from the fact that each partial derivative of \(\Phi \) is bounded and \(\mathbf e _k\) is the \(k^{th}\) unit vector in \(\mathbb {R}_+^7.\) Hence, there exists a constant \(\mathcal {L} >0\) such that

$$\begin{aligned} \Vert \Phi (\varvec{x}_1) -\Phi ( \varvec{x}_2) \Vert _{\infty } \le \mathcal {L} \Vert (\varvec{x}_1 - \varvec{x}_2)\Vert _{\infty }. \end{aligned}$$

Hence, \(\Phi \) is Lipschitz continuous and therefore, by the Picard’s existence and uniqueness theorem, the system under study has a unique solution. \(\square \)

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Woldegerima, W.A., Teboh-Ewungkem, M.I. & Ngwa, G.A. The Impact of Recruitment on the Dynamics of an Immune-Suppressed Within-Human–Host Model of the Plasmodium falciparum Parasite. Bull Math Biol 81, 4564–4619 (2019). https://doi.org/10.1007/s11538-018-0436-0

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s11538-018-0436-0

Keywords

Navigation