Skip to main content
Log in

Mathematical Study of a Resource-Based Diffusion Model with Gilpin–Ayala Growth and Harvesting

  • Original Article
  • Published:
Bulletin of Mathematical Biology Aims and scope Submit manuscript

Abstract

This paper focuses on a Gilpin–Ayala growth model with spatial diffusion and Neumann boundary condition to study single species population distribution. In our heterogeneous model, we assume that the diffusive spread of population is proportional to the gradient of population per unit resource, rather than the population density itself. We investigate global well-posedness of the mathematical model, determine conditions on harvesting rate for which non-trivial equilibrium states exist and examine their global stability. We also determine conditions on harvesting that leads to species extinction through global stability of the trivial solution. Additionally, for time periodic growth, resource, capacity and harvesting functions, we prove existence of time-periodic states with the same period. We also present numerical results on the nature of nonzero equilibrium states and their dependence on resource and capacity functions as well as on Gilpin–Ayala parameter \(\theta \). We conclude enhanced effects of diffusion for small \(\theta \) which in particular disallows existence of nontrivial states even in some cases when intrinsic growth rate exceeds harvesting at some locations in space for which a logistic model allows for a nonzero equilibrium density.

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
Fig. 12
Fig. 13
Fig. 14
Fig. 15
Fig. 16
Fig. 17
Fig. 18

Similar content being viewed by others

References

  • Bai L, Wang K (2005) Gilpin–Ayala model with spatial diffusion and its optimal harvesting policy. Appl Math Comput 171:531–546

    MathSciNet  MATH  Google Scholar 

  • Bai L, Wang K (2007) A diffusive single-species model with periodic coefficients and its optimal harvesting policy. Appl Math Comput 187:873–882

    MathSciNet  MATH  Google Scholar 

  • Belgacem F, Conser C (1995) The effects of dispersal along environmental gradients on the dynamics of population in heterogeneous environment. Can Appl Math Q 3:379–397

    MathSciNet  Google Scholar 

  • Braverman E, Braverman L (2009) Optimal harvesting of diffusive models in a non-homogeneous environment. Nonlinear Anal Theory Methods Appl 71(2009):e2173–e2181

    Article  Google Scholar 

  • Braverman E, Kamrujjaman M (2016a) Lotka systems with directed dispersal dynamics: competition and influence of diffusion strategies. Math Biosci 279:1–12

    Article  MathSciNet  Google Scholar 

  • Braverman E, Kamrujjaman M (2016b) Competitive-cooperative models with various diffusion strategies. Comput Math Appl 72(6):653–662

    Article  MathSciNet  Google Scholar 

  • Burden RL, Faires JD, Burden AM (2015) Numerical analysis, 10th edn. Cengage Learning, Boston

    MATH  Google Scholar 

  • Cantrell RS, Conser CC (1991) Diffusive logistic equation with indefinite weights: population model in disrupted environments II. SIAM J Math Anal 22:1043–1069

    Article  MathSciNet  Google Scholar 

  • Cantrell RS, Cosner C (2003) Spatial ecology via reaction-diffusion equations. Wiley series in mathematical and computational biology. Wiley, Chichester

    MATH  Google Scholar 

  • Cantrell RS, Cosner C, Martinez S (2020) Persistence of a two-stage reaction–diffusion system. Mathematics 8:396

    Article  Google Scholar 

  • Clark F, Brook BW, Delean S, Akcakaya HR, Bradshaw CJA (2010) The theta-logistic is unreliable for modelling most census data. Meth Ecol Evol 1:253–262

    Article  Google Scholar 

  • Conser C (2008) Reaction–diffusion equations and ecological modelling. Book chapter of the book: (Evolution and Ecology) A Friedman (ed). Springer

  • Dockery J, Hutson V, Mishainkow K, Pernarow M (1998) The evolution of slow dispersal rates: a reaction diffusion model. J Math Biol 37(1):61–83

    Article  MathSciNet  Google Scholar 

  • Follak S, Strauss G (2010) Potential distribution and management of the invasive weed Solanum carolinense in Central Europe. Weed Res 50(6):544–552

    Article  Google Scholar 

  • Gilpin ME, Ayala FJ (1973) Global models of growth and competition. Proc Natl Acad Sci USA 70(12):3590–3593

    Article  Google Scholar 

  • Jiang Y, Liu Z, Yang J, Tan Y (2020) Dynamics of a stochastic Gilpin–Ayala population model with Markovian switching and impulsive perturbations. Adv Differ Equ 530:1–17

    MathSciNet  MATH  Google Scholar 

  • Kamrujjaman M (2019) Directed vs regular diffusion strategy: evolutionary stability analysis of a competition model and an ideal free pair. Differ Equ Appl 11(2):267–290

    MathSciNet  MATH  Google Scholar 

  • Korobenko L, Braverman E (2009) A logistic model with a carrying capacity driven diffusion. Can Appl Math Q 17:85–104

    MathSciNet  MATH  Google Scholar 

  • Korobenko L, Kamrujjaman M, Braverman E (2013) Persistence and extinction in Spatial model with a carrying capacity driven diffusion and harvesting. J Math Anal Appl 399:352–368

    Article  MathSciNet  Google Scholar 

  • Man-Jun MA, Hu J-J, Zhang J, Tao J-C (2013) A reaction-diffusion model with nonlinearity driven diffusion. Appl Math J Chin Univ 28(3):290–302

    Article  MathSciNet  Google Scholar 

  • Morton KW, Mayers DF (2005) Numerical solution of partial differential equations, 2nd edn. Cambridge University Press, Cambridge

    Book  Google Scholar 

  • Oesterheld M, Semmartin M (2011) Impact of grazing on species composition: adding complexity to a generalized model. Austral Ecol 36(8):881–890

    Article  Google Scholar 

  • Padron V, Trevisan MC (2006) Environmentally induced dispersal under heterogeneous logistic growth. Math Biosci 199(2):160–74

    Article  MathSciNet  Google Scholar 

  • Pao CV (1992) Nonlinear parabolic and elliptic equations. Plenum, New York

    MATH  Google Scholar 

  • Saether BE, Lillegard M, Groton V, Derever MC, Engen S, Nudds TD, Podruzny KM (2008) Geographical gradients in the population dynamics of North-America Prairie ducks. J Anim Ecol 77(5):869–882

    Article  Google Scholar 

  • Smith GD (1985) Numerical solution of partial differential equations: finite difference methods, 3rd edn. Oxford University Press, New York

    Google Scholar 

  • Williams S, Chow P (1978) Nonliner reaction–diffusion models for interacting population. J Math Anal Appl 62:157–159

  • Zhao XQ (2003) Dynamical system in population biology. Springer, New York

    Book  Google Scholar 

  • Zhao K (2018) Global exponential stability of positive periodic solution of the n-species impulsive Gilpin–Ayala competition model with discrete and distributed time delays. J Biol Dyn 12(1):433–454

    Article  MathSciNet  Google Scholar 

Download references

Acknowledgements

The authors are grateful to anonymous reviewers for their valuable comments that significantly contributed to the presentation of the paper. The author M. Kamrujjaman research was partially supported by the University Grants Commission (UGC) and the Ministry of Science and Technology (MOST), Grant EAS 488, Bangladesh. The research of ST was supported in part by the US National Science Foundation (NSF-DMS-1515755) and the Simons Foundation Grant 712318.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Md. Kamrujjaman.

Ethics declarations

Conflict of interest

The authors declare no conflict of interest.

Additional information

Publisher's Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

A Appendix

A Appendix

In this section, we present some lemmas that are already known (see, for example Pao 1992) and presented here for the sake of completeness. Consider

$$\begin{aligned} \begin{aligned} \frac{\partial v}{\partial t} +{\mathcal {L}} v&= \mu (t, x, v) \ , \quad (t, x)\in {\mathbb {A}}_T \ , \\ \frac{\partial v}{\partial n}&= 0 \ , \quad (t, x) \in S_T \ , \\ v (0, x)&= v_0 (x) \ , \quad x \in \Omega . \end{aligned} \end{aligned}$$
(68)

where \({\mathcal {L}} v\) is a uniformly elliptic operator. It will be assumed that \(\mu \) is a \(C^1\) function of its arguments for \((t,x, v) \in \overline{{\mathbb {Z}}} \times I\) where \(I \subset {\mathbb {R}}\) is some suitable finite sub-interval that contains some upper and lower solutions as defined shortly.

Definition 1

A function \(v^* \in {\mathcal {S}}_T \) is an upper solution of (68) if it satisfies

$$\begin{aligned} \begin{aligned}&\frac{\partial v^*}{\partial t} + {\mathcal {L}} v^* - \mu (t, x, v^*) \ge 0 \ , \quad (t, x)\in {\mathbb {A}}_T \ , \\&\frac{\partial v^*}{\partial n} \ge 0 \ , \quad (t, x) \in S_T \ , \\&v^* (0, x) - v_0 (x) \ge 0 \ , \quad x \in \Omega . \end{aligned} \end{aligned}$$
(69)

A function \(v_* \in {\mathcal {S}}_T\) is a lower solution of (68) if \(v_*\) satisfies the same conditions as in (69), except with inequalities \( \ge 0\) replaced by \(\le 0\). The upper-lower solution pair \((v_*, v^*)\) is said to be ordered if \(v^* \ge v_* \) in \({\mathbb {A}}_T\). We denote \(\langle v_*, v^* \rangle \) to be the all continuous functions \(v \in C \left( \overline{{\mathbb {A}}_T} \right) \) with the property that \( v_* \le v \le v^*\).

In what follows, we will assume that for an ordered set of lower and upper solutions \((v_*, v^*)\), there exists \(m_* (t, x), m^* (t, x) \in C^1 \left( \overline{{\mathbb {A}}_T} \right) \) such that for \(v_* \le v_2 \le v_1 \le v^*\),

$$\begin{aligned} -m_* (v_1-v_2) \le \mu \left( t, x, v_1 \right) - \mu \left( t, x, v_2 \right) \le m^* (v_1 - v_2) \end{aligned}$$
(70)

We define

$$\begin{aligned} {\mathbb {G}} (t, x, v) = m_* v + \mu \left( t, x, v \right) \ . \end{aligned}$$
(71)

The left inequality in (70) implies that \({\mathbb {G}} (t, x, v) \) is a non-decreasing function of \(v \in (v_*, v^*)\).

Definition 2

Define lower \(\left\{ {\underline{v}}^{(k)} \right\} _{k=1}^\infty \) and upper sequences \(\left\{ {\overline{v}}^{(k)} \right\} _{k=1}^\infty \) and each of which satisfy the recursion relation

$$\begin{aligned} \begin{aligned} \frac{\partial {v}^{(k)}}{\partial t} + {\mathcal {L}} {v}^{(k)} + m_* {v}^{(k)}&= {\mathbb {G}} \left( t, x, {v}^{(k-1)} \right) \ , \quad (t, x) \in {\mathbb {A}}_T \\ \frac{\partial v^{(k)}}{\partial n}&= 0 \ , \quad (t, x) \in S_T\ , \\ {v}^{(k)} (0, x)&= v_0 (x), \quad x \in \Omega . \end{aligned} \end{aligned}$$
(72)

where \(v^{(0)} = v_*\) and \(v^{(0)} = v^*\) in cases of lower and upper-sequences, respectively.

Remark 5

If \(v^{(k-1)} \in C^{1,2} \left( {\mathbb {A}}_T \right) \cup C \left( \overline{{\mathbb {A}}_T} \right) \), from given properties of \(\mu \) and \(m_*\), \({\mathbb {G}} \left( t, x, {v}^{(k-1)} (t, x) \right) \in C^1 \left( {\mathbb {A}}_T \right) \cup C \left( \overline{{\mathbb {A}}_T} \right) \), and therefore, using properties of Green’s function of the operator \(\partial _t +{\mathcal {L}} + m_*\) with Neumann conditions, it follows that \(v^{(k)} \in C^{1,2} \left( {\mathbb {A}}_T \right) \cup C \left( \overline{{\mathbb {A}}_T} \right) \). Therefore, since \(v^0 \in C^{1,2} \left( {\mathbb {A}}_T \right) \cup C \left( \overline{{\mathbb {A}}_T} \right) \), it follows by induction that \(v^{(k)} \in C^{1,2} \left( {\mathbb {A}}_T \right) \cup C \left( \overline{{\mathbb {A}}_T} \right) \) for all \(k \in {\mathbb {Z}}^+\).

Theorem 7

For any \((t, x) \in {\mathbb {A}}_T\), \( \left\{ {\underline{v}}_k (t, x) \right\} _{k=1}^\infty \) and \( \left\{ {\overline{v}}_k (t, x) \right\} _{k=1}^\infty \) are non-decreasing and non-increasing sequences respectively satisfying following inequalities for each k

$$\begin{aligned} v_* (t, x) \le {\underline{v}}_k (t, x) \le {\overline{v}}_k (t, x) \le v^* (t, x) \end{aligned}$$
(73)

and each of the sequences \({\underline{v}}_k, {\overline{v}}_k\) converge in \(C^{1, 2} \left( {\mathbb {A}}_T \right) \cup C \left( \overline{{\mathbb {A}}} \right) \) to the same solution v(tx) of the initial value problem.

Proof

Assume the statement is true for a certain k. Now, consider

$$\begin{aligned} w = {\underline{v}}^{(k+1)}-{\underline{v}}^{(k)} \end{aligned}$$
(74)

Clearly since \({\mathbb {G}} (t, x, v)\) is a non-decreasing function of v, w satisfies

$$\begin{aligned} \frac{\partial w}{\partial t} + {\mathcal {L}} w + m_* w = {\mathbb {G}} \left( t, x, {\underline{v}}^{(k)} \right) -{\mathbb {G}} \left( t, x, {\underline{v}}^{(k-1)} \right) \end{aligned}$$
(75)

while satisfying homogeneous initial and boundary conditions. Therefore \(w \ge 0\). Furthermore, from maximum principle if \(w =0\) at some interior point, it must be identically zero. Therefore \({\underline{v}}^{(k+1)} (t, x) \ge {\underline{v}}^{(k)} (t, x)\), the equality holding only if they are identical. Similarly, if we define \(w= {\overline{v}}^{(k)} - {\overline{v}}^{(k+1)} \), the induction statement applied to k and the fact that \({\mathbb {G}}\) is non-decreasing in v implies \( {\mathbb {G}} \left( t, x, {\overline{v}}^{(k-1)} \right) \ge {\mathbb {G}} \left( t, x, {\overline{v}}^{(k)} \right) \ge 0 \), it follows in this case as well, \(\partial _t w + {\mathcal {L}} w + m_* w \ge 0\) while satisfying homogeneous initial and boundary condition. It follows that \(w \ge 0\), with \(w=0\) at some interior point only if \({\overline{v}}^{(k+1)} (t, x) = {\overline{v}}^{(k)} (t,x)\) for any \((t, x) \in {\mathbb {A}}_T\). Again taking \(w = {\overline{v}}^{(k+1)} - {\underline{v}}^{(k+1)} \), it follows that

$$\begin{aligned} \partial _t w + {\mathcal {L}} w + m_* w = {\mathbb {G}} \left( t, x, {\overline{v}}^{(k)} \right) -{\mathbb {G}} \left( t, x, {\underline{v}}^{(k)} \right) \ge 0 \end{aligned}$$
(76)

and the same argument holds, implying that \(w \ge 0\), and thus \({\overline{v}}^{(k+1)} (t, x) \ge {\underline{v}}^{(k+1)} (t, x) \) with equality holding at some point in \({\mathbb {A}}_T\) only if the two functions are identical. Therefore, we have proved (77). Therefore, from monotonicity of each of the sequences \(\left\{ {\underline{v}}^{(k)} (t, x) \right\} _{k=1}^\infty \) and \(\left\{ {\overline{v}}^{(k)} (t, x) \right\} _{k=1}^\infty \) for any \((t, x) \in {\mathbb {A}}_T\), one increasing with an upper bound, the other decreasing with a lower bound, we have pointwise convergence to a function \({\underline{v}} (t, x)\) and \({\overline{v}} (t, x)\), respectively. Now, since using Green’s function we may write the solution \(v^{(k)} (t, x) \in C^{1,2} ({\mathbb {A}}_T ) \cup C \left( \overline{{\mathbb {A}}}_T \right) \) satisfying (72) may be written as an integral equation

$$\begin{aligned} v^{(k)} (t, x)&= \int _0^t \int _{\Omega } {\mathbb {G}} \left( t, x, \tau , \xi \right) {\mathbb {G}} \left( \tau , \xi , v^{(k-1)} \left( \tau , \xi \right) \right) \mathrm{d}\xi \mathrm{d}\tau \nonumber \\&\quad + \int _{\Omega } {\mathbb {G}}_0 \left( t, x, \xi \right) v_0 (\xi ) \mathrm{d}\xi \end{aligned}$$
(77)

applying dominating convergence theorem, it follows that the limiting functions, either \({\underline{v}}\) or \({\overline{v}}\) solves the integral equation

$$\begin{aligned} v (t, x) = \int _0^t \int _{\Omega } {\mathbb {G}} \left( t, x, \tau , \xi \right) {\mathbb {G}} \left( \tau , \xi , v \left( \tau , \xi \right) \right) \mathrm{d}\xi \mathrm{d}\tau + \int _{\Omega } {\mathbb {G}}_0 \left( t, x, \xi \right) v_0 (\xi ) \mathrm{d}\xi \end{aligned}$$
(78)

and therefore is in \(C^{(1,2)} \left( {\mathbb {A}}_T \right) \cup C \left( \overline{{\mathbb {A}}_T } \right) \) and a solution to the nonlinear parabolic initial value problem (68). We now argue that \({\underline{v}} = {\overline{v}}\). If we define \(w = {\underline{v}} - {\overline{v}} \), using (70) we note it satisfies

$$\begin{aligned} \frac{\partial w}{\partial t} + {\mathcal {L}} w -m^* w \ge 0 , \quad \frac{\partial w}{\partial n} = 0 \;\mathrm{on}\quad \partial \Omega \ , \quad \quad w (0, x) = 0 \end{aligned}$$
(79)

implying that \(w \ge 0\). Since \(w \le 0\) as well, it follows \(w=0\). \(\square \)

Consider now the steady state problem

$$\begin{aligned} {\mathcal {L}} v = \mu (x, v) \ , \quad \frac{\partial v}{\partial n} = 0 ~\mathrm{on}~ \partial \Omega \end{aligned}$$
(80)

We will assume again that for \(v \in J \), there exists \(m_*, m^* \in C^1 ({\overline{\Omega }} ) \) with the property that for \(v_1 \ge v_2\)

$$\begin{aligned} \mu (x, v_1) - \mu (x, v_2) \ge - m_* (v_1-v_2) \end{aligned}$$
(81)

where \(m_* (x) \ge 0\) and not identically zero. From (81),

$$\begin{aligned} {\mathbb {G}} (x, v) = \mu (x, v) + m_* v \end{aligned}$$
(82)

is a non-decreasing function of v.

Definition 3

We define \(v_* , v^* \in C^{2} \left( \Omega \right) \cup C^{\alpha } \left( {\overline{\Omega }} \right) \) to be positive lower and upper solutions of (80) satisfying \(v^* (x) \ge v_* (x) \) for each \(x \in \Omega \). Furthermore, we define lower and upper sequences \(\left\{ {\underline{v}}^{(k)} \right\} _{k=1}^\infty \) and \(\left\{ {\overline{v}}^{(k)} \right\} _{k=1}^\infty \) respectively so that each satisfy the recurrence relation

$$\begin{aligned} \begin{aligned}&{\mathcal {L}} {v}^{(k)} + m_* {v}^{(k)} = {\mathbb {G}} \left( x, {v}^{(k-1)} \right) ,\quad x \in \Omega \\&\frac{\partial v^{(k)}}{\partial n} = 0 \ , \quad x \in \partial \Omega . \\ \end{aligned} \end{aligned}$$
(83)

where \(v^{0} = v_*\) and \({v}^*\) for lower and upper sequences, respectively.

Remark 6

Since from assumptions \({\mathbb {G}} \) is \(C^1\) in v and x, it follows that that if \(v^{(k-1)} \in C^2 \left( \Omega \right) \cup C^{\alpha } \left( {\overline{\Omega }} \right) \), \({v}^{(k)} \in C^{2} \left( \Omega \right) \cup C^{\alpha } \left( {\overline{\Omega }} \right) \) from theory of linear elliptic equations.

Theorem 8

For any \(x \in \Omega \), \( \left\{ {\underline{v}}_k (x) \right\} _{k=1}^\infty \) and \( \left\{ {\overline{v}}_k (x) \right\} _{k=1}^\infty \) are non-decreasing and non-increasing sequences respectively satisfying following inequalities for each k

$$\begin{aligned} v_* (x) \le {\underline{v}}_k (x) \le {\overline{v}}_k (x) \le v^* (x) \end{aligned}$$
(84)

and each of the sequences \({\underline{v}}_k, {\overline{v}}_k\) converge in \(C^{2} \left( \Omega \right) \cup C^\alpha \left( {\overline{\Omega }} \right) \) to \({\underline{v}}\) and \({\overline{v}} \), respectively, each of is a nonzero solution of (80).

Proof

Assume the statement is true for a certain k. Now, consider

$$\begin{aligned} w = {\underline{v}}^{(k+1)}-{\underline{v}}^{(k)} \end{aligned}$$
(85)

Clearly since \({\mathbb {G}} (x, v)\) is a non-decreasing function of v, w satisfies

$$\begin{aligned} {\mathcal {L}} w + m_* w = {\mathbb {G}} \left( x, {\underline{v}}^{(k)} \right) -{\mathbb {G}} \left( x, {\underline{v}}^{(k-1)} \right) \ge 0 \end{aligned}$$
(86)

while satisfying homogeneous Neumann boundary condition. Therefore \(w \ge 0\) since \(m_* \ge 0\) and not indentically zero. Furthermore, if \(w =0\) at some interior point, it must be identically zero from maximum principle. Therefore, \({\underline{v}}^{(k+1)} (x) \ge {\underline{v}}^{(k)} (x)\), the equality holding only if they are identical. Similarly, if we define \(w= {\overline{v}}^{(k)} - {\overline{v}}^{(k+1)} \), the induction statement applied to k and the fact that \({\mathbb {G}}\) is non-decreasing in v implies \( {\mathbb {G}} \left( x, {\overline{v}}^{(k-1)} \right) \ge {\mathbb {G}} \left( x, {\overline{v}}^{(k)} \right) \ge 0 \), it follows in this case as well, \({\mathcal {L}} w + m_* w \ge 0\) while satisfying homogeneous Neumann boundary condition. It follows that \(w \ge 0\), with \(w=0\) at some interior point only if \({\overline{v}}^{(k+1)} (t, x) = {\overline{v}}^{(k)} (t,x)\) for any \((t, x) \in {\mathbb {A}}_T\). Again taking \(w = {\overline{v}}^{(k+1)} - {\underline{v}}^{(k+1)} \), it follows that

$$\begin{aligned} {\mathcal {L}} w + m_* w = {\mathbb {G}} \left( x, {\overline{v}}^{(k)} \right) -{\mathbb {G}} \left( x, {\underline{v}}^{(k)} \right) \ge 0 \end{aligned}$$
(87)

and the same argument holds, implying that \(w \ge 0\), and thus \({\overline{v}}^{(k+1)} (x) \ge {\underline{v}}^{(k+1)} (x) \) with equality holding at some point in \(\Omega \) only if the two functions are identical. Therefore, we have proved (84). Therefore, from monotonicity of each of the sequences \(\left\{ {\underline{v}}^{(k)} (x) \right\} _{k=1}^\infty \) and \(\left\{ {\overline{v}}^{(k)} (x) \right\} _{k=1}^\infty \) for any \(x \in \Omega \), one increasing with an upper bound, the other decreasing with a lower bound, we have pointwise convergence to a function \({\underline{v}} (x)\) and \({\overline{v}} (x)\), respectively. Now, since using Green’s function we may write the solution \(v^{(k)} (t, x) \in C^{2} (\Omega ) \cup C^\alpha \left( {\overline{\Omega }} \right) \) satisfying (83), may be written as an integral equation

$$\begin{aligned} v^{(k))} (x) = \int _{\Omega } {\mathbb {G}} \left( x, \xi \right) {\mathbb {G}} \left( \xi , v^{(k-1)} \left( \tau , \xi \right) \right) \mathrm{d}\xi \end{aligned}$$
(88)

applying dominating convergence theorem, it follows that the limiting functions, either \({\underline{v}}\) or \({\overline{v}}\) solves the integral equation

$$\begin{aligned} v (x) = \int _{\Omega } {\mathbb {G}} \left( x, \xi \right) {\mathbb {G}} \left( \xi , v \left( \xi \right) \right) \mathrm{d}\xi \end{aligned}$$
(89)

and therefore is in \(C^{2} \left( \Omega \right) \cup C^\alpha \left( {\overline{\Omega }} \right) \) and a solution of (80). \(\square \)

Definition 4

Grönwall’s inequality: Let \(\rho :[0,T]\rightarrow \mathbb {R+}\) be a positive differential function for which there exists a constant C so that

$$\begin{aligned} \frac{\mathrm{d}}{\mathrm{d}t}(\rho (t))\le C\rho (t)\ , \quad \mathrm{for all} \quad t\in [0,T] \end{aligned}$$

Then

$$\begin{aligned} \rho (t)\le e^{Ct}\rho (0) \end{aligned}$$

Rights and permissions

Springer Nature or its licensor holds exclusive rights to this article under a publishing agreement with the author(s) or other rightsholder(s); author self-archiving of the accepted manuscript version of this article is solely governed by the terms of such publishing agreement and applicable law.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Zahan, I., Kamrujjaman, M. & Tanveer, S. Mathematical Study of a Resource-Based Diffusion Model with Gilpin–Ayala Growth and Harvesting. Bull Math Biol 84, 120 (2022). https://doi.org/10.1007/s11538-022-01074-8

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1007/s11538-022-01074-8

Keywords

Mathematics Subject Classification

Navigation