Skip to main content

Advertisement

Log in

Proportional threshold harvesting in discrete-time population models

  • Published:
Journal of Mathematical Biology Aims and scope Submit manuscript

Abstract

Threshold-based harvesting strategies tend to give high yields while protecting the exploited population. A significant drawback, however, is the possibility of harvesting moratoria with their socio-economic consequences, if the population size falls below the threshold and harvesting is not allowed anymore. Proportional threshold harvesting (PTH) is a strategy, where only a fraction of the population surplus above the threshold is harvested. It has been suggested to overcome the drawbacks of threshold-based strategies. Here, we use discrete-time single-species models and rigorously analyze the impact of PTH on population dynamics and stability. We find that the population response to PTH can be markedly different depending on the specific population model. Reducing the threshold and allowing more harvest can be destabilizing (for the Ricker and Hassell map), stabilizing (for the quadratic map), or both (for the generalized Beverton–Holt map). Similarly, management actions in the form of increasing the threshold do not always improve population stability—this can also be due to bistability. Our results therefore emphasize the importance of a rigorous analysis in investigating the impact of PTH on population stability.

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

Similar content being viewed by others

References

  • Barraquand F, Louca S, Abbott KC, Cobbold CA, Cordoleani F, DeAngelis DL, Elderd BD, Fox JW, Greenwood P, Hilker FM, Murray DL, Stieha CR, Taylor RA, Vitense K, Wolkowicz GS, Tyson RC (2017) Moving forward in circles: challenges and opportunities in modelling population cycles. Ecol Lett 20:1074–1092

    Google Scholar 

  • Beddington JR, May RM (1977) Harvesting natural populations in a randomly fluctuating environment. Science 197:463–465

    Google Scholar 

  • Braverman E, Liz E (2012) Global stabilization of periodic orbits using a proportional feedback control with pulses. Nonlinear Dyn 67:2467–2475

    MathSciNet  MATH  Google Scholar 

  • Cid B, Hilker FM, Liz E (2014) Harvest timing and its population dynamic consequences in a discrete single-species model. Math Biosci 248:78–87

    MathSciNet  MATH  Google Scholar 

  • Clark CW, Kirkwood GP (1986) On uncertain renewable resource stocks: optimal harvest policies and the value of stock surveys. J Environ Econ Manag 13:235–244

    MATH  Google Scholar 

  • Conrad JM, Lopez A, Bjørndal T (1998) Fishery management: the consequences of honest mistakes in a stochastic environment. Technical report SNF-working paper 27/98, Foundation for Research in Economics and Business Administration, Bergen, Norway

  • Cull P (2007) Population models: stability in one dimension. Bull Math Biol 69:989–1017

    MathSciNet  MATH  Google Scholar 

  • Dattani J, Blake JCH, Hilker FM (2011) Target-oriented chaos control. Phys Lett A 375:3986–3992

    MATH  Google Scholar 

  • Deroba JJ, Bence JR (2008) A review of harvest policies: understanding relative performance of control rules. Fish Res 94:201–233

    Google Scholar 

  • Doebeli M (1995) Dispersal and dynamics. Theor Popul Biol 47:82–106

    MATH  Google Scholar 

  • El-Morshedy HA, Jiménez López V (2008) Global attractors for difference equations dominated by one-dimensional maps. J Differ Equ Appl 14:391–410

    MathSciNet  MATH  Google Scholar 

  • Enberg K (2005) Benefits of threshold strategies and age-selective harvesting in a fluctuating fish stock of Norwegian spring spawning herring Clupea harengus. Marine Ecol Prog Ser 298:277–286

    Google Scholar 

  • Engen S, Lande R, Sæther BE (1997) Harvesting strategies for fluctuating populations based upon uncertain population estimates. J Theor Biol 186:201–212

    Google Scholar 

  • FAO (1995) Code of conduct for responsible fisheries. Food and Agriculture Organization of the United Nations, Rome

  • Fieberg J (2004) Role of parameter uncertainty in assessing harvest strategies. N Am J Fish Manag 24:459–474

    Google Scholar 

  • Franco D, Liz E (2013) A two-parameter method for chaos control and targeting in one-dimensional maps. Int J Bifurc Chaos 23:1350003

    MathSciNet  MATH  Google Scholar 

  • Franco D, Perán J (2013) Stabilization of population dynamics via threshold harvesting strategies. Ecol Complex 14:85–94

    Google Scholar 

  • Gatto M, Rinaldi S (1976) Mean value and variability of fish catches in fluctuating environments. J Fish Res Board Can 33:189–193

    Google Scholar 

  • Getz WM, Haight RG (1989) Population harvesting: demographic models of fish, forest, and animal resources. Princeton University Press, Princeton

    Google Scholar 

  • Hall DL, Hilborn R, Stocker M, Walters CJ (1988) Alternative harvest strategies for Pacific herring (Clupea harengus pallasi). Can J Fish Aquat Sci 45:888–897

    Google Scholar 

  • Hilborn R, Walters CJ (1992) Quantitative fisheries stock assessment: choice. Dynamics and uncertainty. Chapman & Hall, London

    Google Scholar 

  • Homans FR, Wilen JE (1997) A model of regulated open access resource use. J Environ Econ Manag 32:1–21

    Google Scholar 

  • Hudson PJ, Dobson AP (2001) Harvesting unstable populations: red grouse Lagopus lagopus scoticus (Lath.) in the United Kingdom. Wildl Biol 7:189–195

    Google Scholar 

  • ICES (2011) Report of the ICES advisory committee. International Council for the Exploration of the Sea, Copenhagen

    Google Scholar 

  • Ishimura G, Punt AE, Huppert DD (2005) Management of fluctuating fish stocks: the case of Pacific whiting. Fish Res 73:201–216

    Google Scholar 

  • Jonzén N, Ranta E, Lundberg P, Kaitala V, Lindén H (2003) Harvesting-induced population fluctuations? Wildl Biol 9:59–65

    Google Scholar 

  • Kaitala V, Jonzén N, Enberg K (2003) Harvesting strategies in a fish stock dominated by low-frequency variability: the Norwegian spring-spawning herring (Clupea harengus). Marine Resour Econ 18:263–274

    Google Scholar 

  • Kokko H (2001) Optimal and suboptimal use of compensatory responses to harvesting: timing of hunting as an example. Wildl Biol 7:141–150

    Google Scholar 

  • Lande R, Engen S, Sæther BE (1995) Optimal harvesting of fluctuating populations with a risk of extinction. Am Nat 145:728–745

    Google Scholar 

  • Lande R, Sæther BE, Engen S (1997) Threshold harvesting for sustainability of fluctuating resources. Ecology 78:1341–1350

    Google Scholar 

  • Lande R, Sæther BE, Engen S (2001) Sustainable exploitation of fluctuating populations. In: Reynolds JD, Mace GM, Redford KH, Robinson JG (eds) Conservation of exploited species. Cambridge University Press, Cambridge, pp 67–86

    Google Scholar 

  • Lande R, Engen S, Sæther BE (2003) Stochastic population dynamics in ecology and conservation. Oxford University Press, Oxford

    MATH  Google Scholar 

  • Lillegård M, Engen S, Sæther BE, Toresen R (2005) Harvesting strategies for Norwegian spring-spawning herring. Oikos 110:567–577

    Google Scholar 

  • Liz E (2007) Local stability implies global stability in some one-dimensional discrete single-species models. Discrete Contin Dyn Syst B 7:1991–1999

    MathSciNet  MATH  Google Scholar 

  • Liz E (2010) How to control chaotic behaviour and population size with proportional feedback. Phys Lett A 374:725–728

    MathSciNet  MATH  Google Scholar 

  • Liz E (2017) Effects of strength and timing of harvest on seasonal population models: stability switches and catastrophic shifts. Theor Ecol 10:235–244

    Google Scholar 

  • Liz E, Hilker FM (2014) Harvesting and dynamics in some one-dimensional population models. In: AlSharawi Z, Cushing J, Elaydi S (eds) Theory and applications of difference equations and discrete dynamical systems. Springer, Berlin, pp 61–73

    MATH  Google Scholar 

  • Liz E, Pilarczyk P (2012) Global dynamics in a stage-structured discrete-time population model with harvesting. J Theor Biol 297:148–165

    MathSciNet  MATH  Google Scholar 

  • Liz E, Ruiz-Herrera A (2012) The hydra effect, bubbles, and chaos in a simple discrete population model with constant effort harvesting. J Math Biol 65:997–1016

    MathSciNet  MATH  Google Scholar 

  • Liz E, Ruiz-Herrera A (2016) Potential impact of carry-over effects in the dynamics and management of seasonal populations. PLoS ONE 11(e0155):579

    Google Scholar 

  • Ludwig D (1998) Management of stocks that may collapse. Oikos 83:397–402

    Google Scholar 

  • May RM, Beddington JR, Horwood JW, Shepherd JG (1978) Exploiting natural populations in an uncertain world. Math Biosci 42:219–252

    MathSciNet  Google Scholar 

  • Maynard Smith J, Slatkin M (1973) The stability of predator-prey systems. Ecology 54:384–391

    Google Scholar 

  • Milner-Gulland EJ, Mace R (1998) Conservation of biological resources. Wiley-Blackwell, Malden

    Google Scholar 

  • Milner-Gulland EJ, Shea K, Possingham H, Coulson T, Wilcox C (2001) Competing harvesting strategies in a simulated population under uncertainty. Animal Conserv 45:157–167

    Google Scholar 

  • Pacific Fishery Management Council (1998) Options and analyses for the coastal pelagic species fishery management plan: appendix B to amendment 8. http://www.pcouncil.org/wp-content/uploads/cpsa8_apdx_b.pdf. Accessed 31 Aug 2019

  • Punt AE (2010) Harvest control rules and fisheries management. In: Hilborn R, Squires D, Tait M, Williams MJ, Grafton RQ (eds) Handbook of marine fisheries conservation and management. Oxford University Press, Oxford, pp 582–594

  • Quinn TJ II, Deriso RB (1999) Quantitative fish dynamics. Oxford University Press, New York

    Google Scholar 

  • Quinn TJ II, Fagen R, Zheng J (1990) Threshold management policies for exploited populations. Can J Fish Aquat Sci 47:2016–2029

    Google Scholar 

  • Reed WJ (1979) Optimal escapement levels in stochastic and deterministic harvesting models. J Environ Econ Manag 6:350–363

    MATH  Google Scholar 

  • Ricker WE (1958) Maximum sustained yields from fluctuating environments and mixed stocks. J Fish Res Board Can 15:991–1006

    Google Scholar 

  • Sæther BE, Engen S, Solberg EJ (2001) Optimal harvest of age-structured populations of moose Alces alces in a fluctuating environment. Wildl Biol 7:171–179

    Google Scholar 

  • Sæther BE, Engen S, Persson J, Brøseth H, Landa A, Willebrand T (2005) Management strategies for the wolverine in Scandinavia. J Wildl Manag 69:1001–1014

    Google Scholar 

  • Sæther BE, Engen S, Odden J, Linnell JDC, Grøtan V, Andrén H (2010) Sustainable harvest strategies for age-structured Eurasian lynx populations: the use of reproductive value. Biol Conserv 143:1970–1979

    Google Scholar 

  • Sinclair ARE, Fryxell JM, Caughley G (2006) Wildlife ecology, conservation, and management. Blackwell, Oxford

    Google Scholar 

  • Steele DH, Andersen R, Green JM (1992) The managed commercial annihilation of northern cod. Nfld Stud 8:34–68

    Google Scholar 

  • Stone L, Hart D (1999) Effects of immigration on the dynamics of simple population models. Theor Popul Biol 55:227–234

    MATH  Google Scholar 

  • Stoop R, Wagner C (2003) Scaling properties of simple limiter control. Phys Rev Lett 90(154):101

    Google Scholar 

  • Thieme HR (2003) Mathematics in population biology. Princeton University Press, Princeton

    MATH  Google Scholar 

  • Tufto J, Sæther BE, Engen S, Swenson JE, Sandegren F (1999) Harvesting strategies for conserving minimum viable populations based on World Conservation Union criteria: brown bears in Norway. Proc R Soc Lond B Biol Sci 266:961–967

    Google Scholar 

  • Walters C (1986) Adaptative management of renewable resources. Macmillan, New York

    Google Scholar 

Download references

Acknowledgements

Eduardo Liz acknowledges the support of the research Grant MTM2017–85054–C2–1–P (AEI/FEDER, UE).

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Frank M. Hilker.

Additional information

Publisher's Note

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

A Appendix

A Appendix

1.1 A.1 Existence and uniqueness of a positive equilibrium of (2.3)

We first prove a technical lemma.

Lemma 1

Let \(g:[a,b]\rightarrow {\mathbb {R}}\) be a twice differentiable map such that \(g'(x)>0\) and \(g''(x)<0\) for all \(x\in (a,b)\), and \(g(a)>a\). Then g can have at most one fixed point p in (ab]. Moreover, the fixed point exists if and only if \(g(b)\le b\).

Proof

Assume that the set of fixed points of g is not empty, and denote by \(p_1\) the infimum of this set. Since \(g(a)>a\), it is clear that \(g'(p_1)\le 1\). By the hypotheses, \(g'(x)<1\) for all \(x>p_1\). This excludes the possibility of more fixed points of g, and implies that \(g(b)\le b\). Bolzano’s Theorem ensures the existence of a fixed point of g when \(g(a)>a\) and \(g(b)\le b\). \(\square \)

Proposition A.1

Assume that f satisifies (H) and \(T< K\). Then \(F_q\) has a unique positive fixed point \(K_1=K_1(q,T)\in (T,K)\) for each \(q\in (0,1)\).

Proof

We first prove the uniqueness of the fixed point of \(F_q\), dividing the proof into two steps.

Step 1:\(F_q\)cannot have positive fixed points in\((0,T]\cup [K,\infty )\).

For a fixed \(q\in (0,1)\), \(F_q\) is a convex combination of f(x) and \(F_1(x)\). Actually, \(F_q(x)=(1-q)f(x)+cF_1(x)\), for all \(x\ge 0\). Hence,

$$\begin{aligned} \left. \begin{array}{c}f(x)\le x,\quad \forall \, x\ge K\\ F_1(x)<x,\quad \forall \, x\ge K\end{array}\right\}&\Longrightarrow F_q(x)<x,\quad \forall \, x\ge K,\\ \left. \begin{array}{c}f(x)> x,\quad \forall \, x\le T\\ F_1(x)\ge x,\quad \forall \, x\le T\end{array}\right\}&\Longrightarrow F_q(x)>x,\quad \forall \, x\le T. \end{aligned}$$

Step 2:\(F_q\)has a unique positive fixed point\(K_1\in (T,K)\).

We notice that \(F_q(x)=(1-q)f(x)+qT\) for all \(x\in (T,K)\), and therefore \(F_q'(x)\) and \(F_q''(x)\) have the same sign as \(f'(x)\) and \(f''(x)\), respectively.

If \(d\ge K\), then \(F_q'(x)>0\) and \(F_q''(x)<0\) for all \(x\in (T,K)\). Since \(F_q(T)>T\) and \(F_q(K)<K\), Lemma 1 ensures that \(F_q\) has a unique fixed point \(K_1\in (T,K)\).

In the following, we assume that \(d<K\) and consider two cases.

Case 1: \(F_q(d)<d\).

In this case, \(F_q\) cannot have fixed points in (dK). By Lemma 1, \(F_q(T)>T\) and \(F_q(d)<d\) imply that \(F_q\) has a unique fixed point \(K_1\in (T,d)\subset (T,K)\).

Case 2: \(F_q(d)\ge d\).

In this case, Lemma 1 guarantees that \(F_q\) cannot have any fixed point in (Td). Since \(F_q'(x)<0\) for all \(x\in (d,K)\), \(F_q(d)\ge d\), and \(F_q(K)<K\), it follows from Bolzano’s and Rolle’s theorems that there exists a unique \(K_1\in [d,K)\subset (T,K)\) such that \(F_q(K_1)=K_1\). \(\square \)

1.2 A.2 Global stability for (2.3) when \(F'(K_1)\ge 0\)

If the map f satisfies (H) and F is nondecreasing at the positive equilibrium \(K_1\), then \(K_1\) attracts all solutions of (2.3) starting at a positive initial condition. This result is a direct consequence of Proposition A.1 and the following auxiliary result:

Lemma 2

(Braverman and Liz 2012, Lemma 1) Let \(g:[0,\infty ) \rightarrow [0,\infty )\) be a continuous function such that \(g(0)=0\), and g has a unique fixed point p such that \(x<g(x)\le p\) for all \(x\in (0,p)\), and \(0<g(x)<x\) for all \(x>p\). Then p is globally attracting for all positive solutions of the equation

$$\begin{aligned} x_{n+1}=g(x_n), \end{aligned}$$
(A.1)

that is, every solution \(\{x_n\}\) of (A.1) with \(x_0>0\) converges to p.

1.3 A.3 Local stability for (2.3) when f is concave in (0, K)

Proposition A.2

Assume that f satisifies (H), and \(f''(x)<0\) for all \(x\in (0,K)\). If K is asymptotically stable for (2.1), then the unique positive equilibrium \(K_1=K_1(q,T)\) of (2.3) is asymptotically stable for all \(q\in (0,1)\) and \(T<K\). Moreover, if \(K_1\) is asymptotically stable for given values of \(q\in (0,1)\) and \(T\in (0,K)\), then decreasing T cannot destabilize the equilibrium.

Proof

Assume first that K is asymptotically stable for (2.1). Since \(f''(x)<0\) for all \(x\in (T,K)\), \(K_1<K\), and \(|f'(K)|\le 1\), it follows that

$$\begin{aligned} |F_q'(K_1)|=(1-q)|f'(K_1)|\le (1-q)|f'(K)|<1. \end{aligned}$$

This implies that \(K_1\) is asymptotically stable.

If \(K_1(q,T_1)\) is asymptotically stable for given values of \(q\in (0,1)\) and \(T_1\in (0,K)\), then the same argument used above applies to prove that \(K_1(q,T_2)\) is asymptotically stable if \(T_2<T_1\). We only need to use that \(T_2<T_1\) implies \(T<K_1(q,T_2)<K_1(q,T_1)<K\) (see Franco and Liz 2013). \(\square \)

1.4 A.4 Global stability for (2.3) with a stable Ricker map

In this section we prove that if the positive equilibrium \(K=1\) of the Ricker map (3.2) is globally asymptotically stable, then the same property holds for all \(q\in (0,1]\) and \(T>0\).

When \(T\ge K\), we prove a general result for maps f satisfying assumption (A). To this end, we shall use the following auxiliary result:

Lemma 3

(El-Morshedy and Jiménez López 2008, Theorem B) Assume that \(f:(0,\infty )\rightarrow (0,\infty )\) has a globally attracting equilibrium p. Let \(g:(0,\infty )\rightarrow (0,\infty )\) be a continuous map satisfying that \(x<g(x)\le \max \{f(x),p\}\) for all \(x<p\), and \(x>g(x)\ge \min \{f(x),p\}\) for all \(x>p\). Then p is a globally attracting equilibrium of g.

Proposition A.3

Assume that f satisifies (A). If \(T\ge K\) and K is a global attractor of (2.1), then K is a global attractor of (2.3) for all \(q\in (0,1]\).

Proof

If \(T\ge K\), then Lemma 3 applies to \(g=F_q\) because \(x<F_q(x)\le f(x)\le \max \{f(x),K\}\) for all \(x<K\), and \(x>F_q(x)=f(x)\ge \min \{f(x),K\}\) for all \(x>K\) (see Fig. 2a). \(\square \)

Next we address the case \(T<1\) for the Ricker map.

Proposition A.4

If \(r\le 2\) and \(T<1\), then the unique positive equilibrium \(K_1=K_1(q,T)\) of (2.3) with the Ricker map \(f(x)=x \,\mathrm {e}^{r(1-x)}\) is globally asymptotically stable.

Proof

It is well known that the equilibrium \(K=1\) for the Ricker map \(f(x)=x \,\mathrm {e}^{r(1-x)}\) is a global attractor if and only if \(r\le 2\) (e.g. Thieme 2003). Moreover, since \(f''(x)<0\) for all \(x<2/r\), it follows that \(f''(x)<0\) for all \(x\in (0,1)\) if \(r\le 2\). Thus, Propositions A.1 and A.2 ensure that (2.3) has a unique positive equilibrium \(K_1\), and it is locally asymptotically stable if \(T<1\) and \(q\in (0,1)\).

Next we prove that \(K_1\) is actually a global attractor. To this end, we use an enveloping theorem due to Cull (2007, Theorem 3). We have to show that \(F(x)<2K_1-x\) if \(x<K_1\), and \(F(x)>2K_1-x\) if \(x>K_1\).

Denote by \(p_1, p_2\) the points such that \(f(p_1)=f(p_2)=T\), \(0<p_1<T<K_1<1<p_2\). F is differentiable in \((p_1,p_2)\). We distinguish three cases (see Fig. 11):

  • If \(x\in (0,p_1)\), then \(F(x)=f(x)\le T<2T-x<2K_1-x\).

  • If \(x\in (p_2,\infty )\), then we use that \(f(x)>2-x\) for all \(x>1\) (e.g. Cull 2007) to show that \(F(x)=f(x)>2-x>2K_1-x\).

  • If \(x\in (p_1,p_2)\) we have \(F'(x)=(1-q)f'(x)>-1\), which makes it impossible that there is a point \(q\in (p_1,K_1)\cup (K_1,p_2)\) such that \(F(q)=2K_1-q\) (assume that such q exists, and apply Rolle’s Theorem to the map \(G(x)=2K_1-x-F(x)\) and the points q and \(K_1\) to arrive at a contradiction).

Thus, F must remain below the line \(2K_1-x\) for \(x<K_1\), and above it for \(x>K_1\), and the proof is complete. \(\square \)

Fig. 11
figure 11

A diagram showing how the modified map \(F(x)=\min \{f(x),(1-q)f(x)+qT\}\) (solid blue line) is enveloped by the line \(2K_1-x\). The original Ricker map f(x) is shown in a red dashed line (color figure online)

1.5 A.5 Stability switches for (2.3) with an unstable Ricker map

Here we rigorously prove the results about the Ricker model stated in Sect. 3.2.

Proposition A.5

Assume that \(r>2\), \(0\le T<1\), and \(q\in (0,1)\).

(I) :

The positive equilibrium \(K_1\) of (2.3) with \(f(x)=x \,\mathrm {e}^{r(1-x)}\) is asymptotically stable if and only if the following inequality holds:

$$\begin{aligned} (1-q)(r g(r,q,T)-1)\le \mathrm {e}^{r(g(r,q,T)-1)}, \end{aligned}$$
(A.2)

where

$$\begin{aligned} g(r,q,T):=\frac{2+rqT+\sqrt{(rqT)^2+4}}{2r} \end{aligned}$$
(A.3)

is the larger root of the quadratic equation \(rx^2-(2+rqT)x+qT=0\).

(II) :

Moreover, for each \(r>2\) fixed, the boundary of the stability region defined by the implicit equation

$$\begin{aligned} (1-q)(r g(r,q,T)-1)= \mathrm {e}^{r(g(r,q,T)-1)} \end{aligned}$$

defines a decreasing curve \(q=q(T)\) in the plane (Tq), \(0<T<1\).

Proof

We first prove (I). Since the positive equilibrium \(K_1\) of (2.3) belongs to the interval (T, 1), function F is defined as \(F(x)=(1-q)f(x)+qT\) in a neighborhood of \(K_1\). This map is unimodal, of class \(C^3\), and has a negative Schwarzian derivative everywhere. Thus, the condition for local asymptotic stability is \(F'(K_1)\ge -1\), which is equivalent to \((1-q)f'(K_1)\ge -1\) (see Liz 2010). In order to get the stability boundary, we solve the system of equations

$$\begin{aligned} x&=(1-q)f(x)+qT \end{aligned}$$
(A.4)
$$\begin{aligned} -1&=(1-q)f'(x) \end{aligned}$$
(A.5)

Since \(f(x)=\mathrm {e}^{r(1-x)}\), system (A.4)–(A.5) reads

$$\begin{aligned} (1-q)\mathrm {e}^{r(1-x)}&=\frac{x-qT}{x} \end{aligned}$$
(A.6)
$$\begin{aligned} (1-q)\mathrm {e}^{r(1-x)}&=\frac{-1}{1-rx} \end{aligned}$$
(A.7)

This system leads to

$$\begin{aligned} \frac{x-qT}{x}=\frac{-1}{1-rx}\Longleftrightarrow rx^2-(2+rqT)x+qT=0. \end{aligned}$$

This quadratic equation has two positive roots

$$\begin{aligned} x_{\pm }=\frac{2+rqT\pm \sqrt{(rqT)^2+4}}{2r}. \end{aligned}$$

Since \(0<q<1\), we get

$$\begin{aligned} 2-\sqrt{(rqT)^2+4}<0<2rT(1-q), \end{aligned}$$

which implies that \(x_{-}<T\).

On the other hand, one can check that \(x_{+}=g(r,q,T)\in (T,1)\) for some q such that

$$\begin{aligned} q<\frac{r-2}{(r-1)T}\quad \text{ and }\quad q>\frac{rT-2}{rT-1} \quad \text{ if }~{ rT>2.} \end{aligned}$$

Replacing \(x=g(r,q,T)\) in (A.7), we get the boundary of the stability region defined in (A.2).

Next we prove (II). The stability boundary can be written as \(H(r,q,T)=0\), where

$$\begin{aligned} H(r,q,T)= \mathrm {e}^{r(g(r,q,T)-1)}-(1-q)(rg(r,q,T)-1). \end{aligned}$$

We show that, for a fixed value of \(r>2\), it holds that \(\partial H/\partial T\) and \(\partial H/\partial q\) are positive at the equilibrium. Thus, the Implicit Function Theorem ensures that \(\partial q/\partial T\) is negative, which proves  (II).

It is clear that \(\partial g/\partial T>0\) and \(\partial g/\partial q>0\), where g is defined in (A.3). On the other hand, it is evident that \(rg(r,q,T)>1\). Since

$$\begin{aligned} \frac{\partial H}{\partial T}&=r\frac{\partial g}{\partial T}\left( \mathrm {e}^{r(g(r,q,T)-1)}-(1-q)\right) ,\\ \frac{\partial H}{\partial q}&=r \frac{\partial g}{\partial q}\left( \mathrm {e}^{r(g(r,q,T)-1)}-(1-q)\right) +(rg(r,q,T)-1) \end{aligned}$$

it only remains to prove that \(\mathrm {e}^{r(g(r,q,T)-1)}>1-q\).

Replacing \(x=g(r,q,T)\) in (A.6), we get

$$\begin{aligned} (1-q)g(r,q,T)\mathrm {e}^{r(1-g(r,q,T))}=g(r,q,T)-qT<g(r,q,T), \end{aligned}$$

from where it follows that \((1-q)\mathrm {e}^{r(1-g(r,q,T))}<1\), as we wanted to prove. \(\square \)

Corollary A.6

Assume that \(r>2\), \(0\le T<1\), and \(q\in (0,1)\). Denote by \(K_1\) the positive equilibrium of (2.3) with \(f(x)=x \,\mathrm {e}^{r(1-x)}\).

(i) :

For \(T=0\), \(K_1\) is asymptotically stable if and only if \(q\ge 1-\mathrm {e}^{2-r}\).

(ii) :

As \(T\rightarrow 1^-\), the limit form of the asymptotic stability condition for \(K_1\) is given by the inequality \(q> (r-2)/(r-1)\).

(iii) :

For a given value of \(T\in (0,1)\), increasing q is stabilizing.

(iv) :

For a fixed value of \(q\in (0,1)\), we have three cases:

(a):

If \(q\ge 1-\mathrm {e}^{2-r}\) (\(r\le 2-\ln (1-q)\)), then \(K_1\) is asymptotically stable, independently of T.

(b):

If \((r-2)/(r-1)<q< 1-\mathrm {e}^{2-r}\), then increasing T is stabilizing.

(c):

If \(q\le (r-2)/(r-1)\) (\(r\ge (2-q)/(1-q)\)), then \(K_1\) is unstable, independently of T.

Proof

(i) :

For \(T=0\), we have \(g(r,q,T)=g(r,q,0)=2/r\), and therefore (A.2) reads \(1-q\le \mathrm {e}^{2-r}\).

(ii) :

For \(T=1\), the solution of (A.4) is \(x=1\), and therefore the limit form of the asymptotic stability condition for \(K_1\) as \(T\rightarrow 1^-\) is given by \(-1<(1-q)f'(1)=(1-q)(1-r),\) which is equivalent to \(q> (r-2)/(r-1)\).

(iii) :

and (iv) are a straightforward consequence of Proposition A.5 and the previous items (i) and (ii).

\(\square \)

1.6 A.6 Stability switches for (2.3) with a quadratic map

In this subsection we prove some analytic results for (2.3) with the quadratic map \(f(x)=rx-(r-1)x^2\). By Proposition A.2, \(r>3\) is a necessary condition for the existence of a stability switch at the positive equilibrium \(K_1\).

Proposition A.7

Assume that \(r>3\), and \(q\in (0,1)\). As \(T\rightarrow 1^-\), the limit form of the asymptotic stability condition for the equilibrium \(K_1\) is given by the inequality \(q> (r-3)/(r-2)\).

Proof

It is clear that the unique equilibrium of \(F(x)=(1-q) f(x)+q\) is \(K_1=1\). Thus, the limit form of the asymptotic stability condition is \(-1< (1-q)f'(1)=(1-q)(2-r)\), which is equivalent to \(q>(r-3)/(r-2)\). \(\square \)

A consequence of Propositions A.2 and A.7 is that, for a given value of \(r>3\), a necessary condition for the existence of a stability switch is \(q< (r-3)/(r-2)\).

Proposition A.8

Assume that \(r>3\), \(0\le T<1\), and \(q< (r-3)/(r-2)\).

(I) :

The boundary of the stability region of the positive equilibrium \(K_1\) of (2.3) with \(f(x)=rx-(r-1)x^2\) is implicitly defined by the equation

$$\begin{aligned} ((1-q)r-1)^2+4q(1-q)(r-1)T=4. \end{aligned}$$
(A.8)

Moreover, for each \(r>3\), (A.8) defines an increasing curve \(q=q(T)\) in the plane (Tq), \(0<T<1\).

(II) :

In the limit case \(T=0\), the stability switch is defined by \(q=(r-3)/r\).

(III) :

A stability switch as T is decreased occurs for all q such that

$$\begin{aligned} \frac{r-3}{r}<q<\frac{r-3}{r-2}. \end{aligned}$$

Proof

(I) :

The positive equilibrium \(K_1\) is the unique positive root of the quadratic equation \(x=(1-q)f(x)+qT\), whose expression is

$$\begin{aligned} K_1=\frac{(1-q)r-1+\sqrt{((1-q)r-1)^2+4q(1-q)(r-1)T}}{2(1-q)(r-1)}. \end{aligned}$$

Notice that \(q< (r-3)/(r-2)\) implies that \(q<(r-1)/r\), and therefore \((1-q)r-1>0\). A stability switch occurs when \((1-q)f'(K_1)=-1\), which leads to (A.8). To show that \(q=q(T)\) is increasing, we use implicit differentiation with the map

$$\begin{aligned} H(q,T)=((1-q)r-1)^2+4q(1-q)(r-1)T-4. \end{aligned}$$

It is clear that \(\partial H/\partial T>0\) for all \(r>1\). On the other hand,

$$\begin{aligned} \partial H/\partial q=-2r((1-q)r-1)+4T(r-1)(1-2q)<0 \end{aligned}$$

if \(q\ge 1/2\). If \(q<1/2\), since \(T<1\), we have

$$\begin{aligned} \partial H/\partial q< & {} -2r((1-q)r-1)+4(r-1)(1-2q)\\= & {} -2((1-q)r^2-(3-4q)r+2(1-2q)), \end{aligned}$$

and we can easily check that the last expression is negative for all \(r>3\). Hence, \(\partial q/\partial T=-(\partial H/\partial T)/(\partial H/\partial q)>0\).

(II) :

For \(T=0\), (A.8) gives \((1-q)r-1=2\), which is equivalent to \(q=(r-3)/r\).

(III) :

is a consequence of (II) and Proposition A.7.\(\square \)

1.7 A.7 Condition for the MSY in (2.3) with a Ricker map

Proposition A.9

System (4.1) with \(f(x)=x \,\mathrm {e}^{r(1-x)}\) has a solution (xT) with \(x>0\), \(T\in (0,K)\), if and only if \(r<q-\ln (1-q)\).

Proof

For \(T=0\) and a fixed \(q\in (0,1)\), we can find the unique solution \((r^*,x^*)\) of system (4.1). Indeed, system (4.1) with \(T=0\) becomes

$$\begin{aligned} f'(x)=1;\quad \quad (1-q)f(x)=x. \end{aligned}$$
(A.9)

Using the expression of f in (A.9), we get

$$\begin{aligned} (1-rx)\mathrm {e}^{r(1-x)}=1=(1-q)\mathrm {e}^{r(1-x)}, \end{aligned}$$

which leads to \(rx=q\). Replacing \(x=q/r\) into \((1-q)\mathrm {e}^{r(1-x)}=1\) gives \((1-q)\mathrm {e}^{r-q}=1\), which is equivalent to \(r=q-\ln (1-q)\). Thus, the solution is \(r^*=q-\ln (1-q)\), \(x^*=q/r^*=q/(q-\ln (1-q))\).

Now, if \(T>0\), then system (4.1) leads to

$$\begin{aligned} (1-rx)\mathrm {e}^{r(1-x)}=1;\quad \quad (1-q)x\mathrm {e}^{r(1-x)}+qT=x, \end{aligned}$$

which, by elementary calculus, gives

$$\begin{aligned} rx=q\frac{x-T}{x-qT}:=q_1<q, \end{aligned}$$

and therefore \(r=q_1-\ln (1-q_1)<r^*\), since the map \(v(q)=q-\ln (1-q)\) is increasing for \(q\in (0,1)\). \(\square \)

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Hilker, F.M., Liz, E. Proportional threshold harvesting in discrete-time population models. J. Math. Biol. 79, 1927–1951 (2019). https://doi.org/10.1007/s00285-019-01415-7

Download citation

  • Received:

  • Revised:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s00285-019-01415-7

Keywords

Mathematics Subject Classification

Navigation