1 Introduction

Many species of animals like whales, elephant seals, bisons and rhinoceroses, are at risk of being harvested to extinction (Gulland 1971; Reiter et al. 1981; Ludwig et al. 1993; Primack 2006). Excessive harvesting has already led to both local and gobal extinctions of species (Lande et al. 1995). In fact, a significant percentage of the endangered birds and mammals of the world are threatened by harvesting, hunting or other types of overexploitation (Lande et al. 1995), and there are similar problems for many species of fish (Hutchings and Reynolds 2004). This is why harvesting strategies have to be carefully chosen. After significant harvests, it takes time for the harvested population to get back to the pre-existing level. Moreover, the harvested population fluctuates randomly in time due to environmental stochasticity. As a result, an overestimation of the ability of the population to rebound can lead the harvester to overharvest the population to extinction (Lande et al. 1995). A less common but nevertheless important problem is an insufficient rate of harvesting. Because of instraspecific competition, the population is bounded in a specific environment, so an extraction rate that is too low would lead to a loss of precious resources. For the same reason, choosing an efficient extraction strategy for valuable species is important (Kokko 2001).

We present a stochastic model of population harvesting and find the optimal harvesting strategy that maximizes the asmptotic yield of harvested individuals. We consider a novel framework, the one of optimal ergodic harvesting. This is based on the theory of ergodic control (Arapostathis et al. 2012). In most stochastic models that exist in the literature, for example Lande et al. (1995), Alvarez and Shepp (1998) and Lungu and Øksendal (1997), the population is either assumed to become extinct in finite time, or it can end up being harvested to extinction. In our framework, if the population goes extinct under some harvesting strategy, the asymptotic yield is 0 and therefore this strategy cannot be optimal. If one wants to ensure that harvested species are preserved, this framework is a natural candidate. Our aim is to present a theory of optimal harvesting that includes the risks of extinction from both environmental noise and harvesting. We assume that the population is homogeneous and can be described by a one dimensional diffusion. The harvesting rate is assumed to be bounded, as infinite harvesting rates would imply an unlimited harvesting capacity, something that is clearly not realistic.

In most cases, environmental noise can be introduced in the system by transforming differential equations into stochastic differential equations (SDE). Such techniques require dealing with significant mathematical difficulties, but their use is not just a case of honoring generality. First, there are direct effects of stochasticity on the predictions of the model, and the parameters quantifying it show up in the results. Second, any realistic biological system will depend on environmental variables that are not, or cannot be, accounted for. The role of stochasticity is to ensure that the solutions proposed are robust to such omissions. For example, if avoiding extinction is important, deterministic models can give misleading solutions even when their parameters are corrected for noise (Smith 1978). The transformation to SDE works especially well when the environmental fluctuations are small and there is no chaos (Lande et al. 1995). We focus on models with environmental stochasticity and neglect the demographic stochasticity which arises from the randomness of birth and death rates of each indiviual of a population. Throughout the paper we assume that environmental stochasticity mainly affects the growth rate of the population (see Turelli 1977; Beddington and May 1977; May et al. 1978; Leigh 1981; Braumann 2002; Gard 1988; Evans et al. 2015, 2013; Schreiber et al. 2011; Hening and Nguyen 2018a for more details). For computational tractability and for clarity of exposition, we look at a one-dimensional model. Nevertheless, our framework works for any model that can be written as a system of stochastic differential equations (satisfying some mild assumptions—see Arapostathis et al. 2012).

A major limitation of existing models in the literature is the dependence of the optimal solutions on parameters that are hard to quantify. For example, in Lande et al. (1995) the level at which the population becomes extinct—the minimal viable population—must be assumed; without it the yields become infinite. In Alvarez and Shepp (1998) the yield must be time discounted to avoid maximizing over yield infinities, and this requires providing a time value for resources. The minimal viable population is a difficult scientific question (Shaffer 1981; Traill et al. 2007), and the time value of yields is a difficult economics and policy question, because it implies the comparison of the utility of present and future generations (Drèze and Stern 1987). In contrast, our model sidesteps the issue by assuming no time preference—and therefore no bias towards extracting in the present, and resolves the problem of maximizing over infinite yields naturally by looking at asymptotic behavior.

A particular case of our model was studied in Abakuks and Prajneshu (1981).Footnote 1 The authors limited themselves to the analysis of harvesting strategies that were of bang–bang type. In Abakuks (1979), one of the co-authors in Abakuks and Prajneshu (1981) proved that an optimal gathering strategy was necessarily of a bang–bang type in a continuous time Markov chain model, making use of the simplifying assumption of a finite state space. Here, instead, we look at very general possible harvesting strategies in a continuous state stochastic model, and show that the optimal one is of bang–bang type. Our contribution is therefore twofold. We generalize the setting of Abakuks and Prajneshu (1981) significantly by looking at very general density-dependent growth rates, not just the logistic case. Moreover, we prove what the authors of Abakuks and Prajneshu (1981) intuited, namely that the optimal strategy is of bang–bang type; and furthermore that this is true for the larger class of convex yield functions.

Stochastic optimal control applications are common in the finance literature. Following the seminal contributions of Merton (1969, 1971), objective functions that are integrals of time discounted instantaneous utility flows are now standard. The crucial simplifying assumption is that of time-additive total utility. The utility flow usually depends on consumption flows, and therefore indirectly on other variables and stochastic constraints. With the time-additive utility assumption, our general yield function can also be interpreted as an instantaneous utility function dependent on yield, and our objective function can be the asymptotic expected utility flow dependent on yield. Because a population stock cannot grow indefinitely in our biological model, we diverge from the general finance literature, where financial returns do not usually depend on the size of the holdings of an individual.

Finally, we generalise a result from one of the stochastic models in Smith (1978), where the equivalent to our yield function has a specific simple form. We show that, when the yield function is weakly convex, the optimal control is bang–bang. However, if the yield function is strictly concave, then the optimal harvesting strategy has to be continuous, in contrast to the bang–bang type optimal strategy we find for a linear yield function. This generalization is useful for economic welfare analysis (a more general form of cost-benefit analysis), which typically relies on a concave utility function, equivalent to the concave yield function herein. In economic models, concavity is assumed to model risk aversion [see Mas-Colell et al. (1995, Proposition 6.C.1) for justification], and for the convenience of interior solutions to maximisation problems. Concave utility leads to a trade-off between risk and returns in asset choice (Merton 1971), so the connection between yield concavity and strategy continuity mentioned above is suggestive of risk management. However, risk management interpretations from the finance literature are not directly applicable here. First, financial asset returns are assumed reasonably to not be decreasing in the asset value owned by investors.Footnote 2 Moreover, the risk-return trade-off is captured in models with choice between at least two assets with different risk profiles.Footnote 3 If anything, finding a bang–bang optimal strategy when yield is linear is more related to finding corner solutions in maximisation problems with linear utility. A bang–bang strategy uses one of the two extremes of the harvesting rate, depending on the momentary population stock.

The rest of the paper is organized as follows. In Sect. 2 we introduce our model and results. We prove that, if the population in the absence of harvesting survives, the yield function is the identity and the harvesting rate is bounded above by some number \(M>0\), then the optimal strategy is always a bang–bang type solution: there exists an \(x^*>0\) such that one does not harvest if the current population size lies in the interval \(\left[ 0,x^*\right] \) and harvests at the maximal possible rate, M, if the current population size lies in the interval \(\left( x^*,\infty \right) \). The proofs of the above results are collected in Appendix A. In Sect. 4 we apply our results to the special setting of the logistic Verhulst–Pearl model. In Sect. 3 (proofs in Appendix B) we show that if the yield function is strictly concave, the optimal harvesting strategy is continuous, and when the yield function is more generally weakly convex, the optimal strategy is bang–bang.

Finally, in Sect. 5 we offer some numerical simulations that show how the optimal harvesting strategies and optimal asymptotic change with respect to the parameters of the model. We also provide a discussion of our results.

2 Optimal ergodic harvesting

We consider a population whose density \({\tilde{X}}(t)\) at time \(t\ge 0\), in the absence of harvesting, follows the stochastic differential equation (SDE)

$$\begin{aligned} d{\tilde{X}}(t) = {\tilde{X}}(t)\mu ({\tilde{X}}(t))\,dt + \sigma \tilde{X}(t)\,dB(t), \quad {\tilde{X}}(0)=x>0, \end{aligned}$$
(2.1)

where \((B(t))_{t\ge 0}\) is a standard one dimensional Brownian motion. This describes a population \({\tilde{X}}\) with per-capita growth rate given by \(\mu (x)>0\) when the density is \(\tilde{X}=x\). The infinitesimal variance of fluctuations in the per-capita growth rate is given by \(\sigma ^2\).

The following is a standing assumption throughout the paper.

Assumption 2.1

The function \(\mu :[0,\infty )\rightarrow \mathbb {R}\) satisfies:

  • \(\mu \) is locally Lipschitz.

  • \(\mu \) is decreasing.

  • As \(x\rightarrow \infty \) we have \(\mu (x)\rightarrow -\infty \).

  • The function \(p(x):=x\mu (x)\) has a unique maximum.

  • There is no interval \((u,v)\subset \mathbb {R}_+\) such that \(p(\cdot )\) is constant on (uv).

The behavior of (2.1) is not hard to study. In the particular case when \(\mu (x)=\overline{\mu }- \kappa x\) see Evans et al. (2015) and Dennis and Patil (1984). The methods there can be easily adapted to our setting. Alternatively, one could use the general results from Hening and Nguyen (2018a). The process \(\tilde{X}\) does not reach 0 or \(\infty \) in finite time and the stochastic growth rate \(\mu (0)-\frac{\sigma ^2}{2}\) determines the long-term behavior in the following way:

  • If \(\mu (0)-\frac{\sigma ^2}{2}>0\) and \( \tilde{X}(0)=x>0\), then \((\tilde{X}(t))_{t\ge 0}\) converges weakly to its unique invariant probability measure \(\nu \) on \((0,\infty )\).

  • If \(\mu (0)-\frac{\sigma ^2}{2}<0\) and \( \tilde{X}(0)=x>0\), then \(\lim _{t\rightarrow \infty } \tilde{X}(t)=0\) almost surely.

We let \(\mathbb {R}_+:=[0,\infty )\) and \(\mathbb {R}_{++}:=(0,\infty )\) throughout the paper.

Assume that the population is harvested at time \(t\ge 0\) at the stochastic rate\(h(t)\in U:=[0,M]\) for some fixed \(M>0\). Adding the harvesting to (2.1) yields the SDE

$$\begin{aligned} dX(t) = X(t)(\mu (X(t)) -h(t) )\,dt + \sigma X(t)\,dB(t),\quad X(0)=x>0. \end{aligned}$$
(2.2)

A stochastic process \((h(t))_{t\ge 0}\) taking values in U is said to be an admissible strategy if \((h(t))_{t\ge 0}\) is adapted to the filtration \((\mathcal {F}_t)_{t\ge 0}\) generated by the Brownian motion \((B(t))_{t\ge 0}\). Let \(\mathfrak {U}\) be the class of admissible strategies. An important subset of \(\mathfrak {U}\) is the class \(\mathfrak U_{sm}\) of stationary Markov strategies, that is, admissible strategies of the form \(h(t)=v(X(t))\) where \(v:\mathbb {R}_{++}\mapsto U\) is a measurable function. By abuse of terminology, we often refer to the map \(v(\cdot )\) as the stationary Markov strategy. Using a stationary Markov strategy \(v(\cdot )\), (2.2) becomes

$$\begin{aligned} dX(t) = X(t)(\mu (X(t)) -v(X(t)) )\,dt + \sigma X(t)\,dB(t), \quad X(0)=x>0. \end{aligned}$$
(2.3)

Remark 2.1

The sigma algebra \(\mathcal {F}_t\) gives one the information available from time 0 to time t. An admissible harvesting strategy is therefore a strategy which can take into account all the information from the start of the harvesting to the present. These strategies are much more general than constant strategies. Stationary Markov strategies are the harvesting strategies which only depend on the present state of the population density.

We associate with X(t) the family of generators \(({\mathcal L}_u)_{u\in [0,M]}\) defined by their action on \(C^2\) functions with compact support in \(\mathbb {R}_{++}\) as

$$\begin{aligned} {\mathcal L}_u f(x):=x[\mu (x)-u]f_x+\dfrac{1}{2}\sigma ^2 x^2f_{xx}. \end{aligned}$$
(2.4)

We will call \(\Phi :\mathbb {R}_+\rightarrow \mathbb {R}_+\) a yield function if the following assumption holds.

Assumption 2.2

The function \(\Phi :\mathbb {R}_+\rightarrow \mathbb {R}_+\) satisfies:

  • \(\Phi \) is continuous.

  • \(\Phi (0)=0\).

  • \(\Phi \) has subpolynomial growth that is, there is \(n \in \mathbb {N}\) such that \(\frac{\Phi (x)}{x^n} \rightarrow 0\) for \(x\rightarrow \infty \).

Our aim is to find the optimal strategy h(t) that almost surely maximizes the asymptotic yield

$$\begin{aligned} \liminf _{T\rightarrow \infty }\dfrac{1}{T}\int _0^T \Phi \Big (X(t)h(t)\Big )\,dt. \end{aligned}$$
(2.5)

In other words we want to find v such that, for any initial population size \(X(0)=x>0\), we have with probability 1 that

$$\begin{aligned} \liminf _{T\rightarrow \infty }\dfrac{1}{T}\int _0^T \Phi \Big (X(t)v(X(t))\Big )\,dt = \sup _{h\in \mathfrak {U}} \liminf _{T\rightarrow \infty }\dfrac{1}{T}\int _0^T \Phi \Big (X(t)h(t)\Big )\,dt=:\rho ^*. \end{aligned}$$

We note that many of the existing models that look at the optimal harvesting of a population in a stochastic environment (Lungu and Øksendal 1997; Alvarez and Shepp 1998; Lande et al. 1995) assume that the yield function \(\Phi \) is the identity i.e. \(\Phi (x)=x, x\ge 0\). This assumption is not always justifiable (see Alvarez 2000) and as such we present in Sect. 3 results for more general functions \(\Phi \).

Remark 2.2

We note that if X has an invariant probability measure \(\pi \) on \(\mathbb {R}_{++}\), then for any \(X(0)=x>0\) almost surely

$$\begin{aligned} \lim _{T\rightarrow \infty }\dfrac{1}{T}\int _0^T \Phi \Big (X(t)v(X(t))\Big )\,dt = \int _{\mathbb {R}_{++}} \Phi (xv(x))\pi (dx). \end{aligned}$$

In particular, if X goes extinct, that is, for any \(X(0)=x>0\) we have with probability 1

$$\begin{aligned} \lim _{t\rightarrow \infty }X(t)=0, \end{aligned}$$

then the only invariant ergodic measure of X on \(\mathbb {R}_+\) is \(\delta _0\) the point mass at 0, and hence, we get that with probability 1

$$\begin{aligned} \lim _{T\rightarrow \infty }\dfrac{1}{T}\int _0^T \Phi \Big (X(t)v(X(t))\Big )\,dt =0. \end{aligned}$$

Our method for maximizing the asymptotic yield forces the optimal harvesting to be such that the population persists.

Remark 2.3

By Arapostathis et al. (2012, Theorems 2.2.2 and 2.2.12), the controlled systems (2.2) and (2.3) have unique local solutions on \(\mathbb {R}_{++}\) for any admissible control h(t) and stationary Markov control v respectively. Note that one can find \(N>0\) large enough such that

$$\begin{aligned} \begin{aligned} {\mathcal L}_u\left( x+\frac{1}{x}\right)&=x(\mu (x)-u)\left( \frac{x^2-1}{x^2}\right) + \sigma ^2x^2\frac{1}{x^3}\\&\le N(\sigma ^2+M)\left( x+\frac{1}{x}\right) ,\quad x\in \mathbb {R}_{++}, u\in U. \end{aligned} \end{aligned}$$

With this fact in hand, we can use the arguments from Khasminskii (2012, Theorem 3.5), to obtain the existence of global solutions on \(\mathbb {R}_{++}\) of (2.2) and (2.3). In particular we get that

$$\begin{aligned} \mathbb {P}_x\left( X(t)\in \mathbb {R}_{++}, t\ge 0\right) =1, \quad x\in \mathbb {R}_{++}. \end{aligned}$$

The main result of the paper is the following.

Theorem 2.1

Assume that \(\Phi (x)=x, x\in (0,\infty )\) and that the population survives in the absence of harvesting, that is \(\mu (0)-\frac{\sigma ^2}{2}>0\). Furthermore assume that the drift function \(\mu (\cdot )\) satisfies Assumption 2.1. The optimal control (the optimal harvesting strategy) v has the bang–bang form

$$\begin{aligned} \begin{aligned} v(x)&= {\left\{ \begin{array}{ll} 0 &{}\quad \text{ if } \; 0< x\le x^* \\ M &{}\quad \text{ if } \; x>x^* \end{array}\right. } \end{aligned} \end{aligned}$$
(2.6)

for a unique \(x^*\in (0,\infty )\). Furthermore, we have the following upper bound for the optimal asymptotic yield

$$\begin{aligned} \rho ^* \le \sup _{x\in \mathbb {R}_+} x\mu (x). \end{aligned}$$
(2.7)

3 Continuous versus bang–bang optimal harvesting strategies

As showcased in Theorem 2.1, when \(\Phi \) is the identity function the optimal harvesting strategy is of bang–bang type. In Appendix B we prove the following result.

Theorem 3.1

Suppose Assumption 2.1 holds and the yield function satisfies

  1. (1)

    \(\Phi \in C^2(\mathbb {R}_+)\),

  2. (2)

    \(\Phi \) is strictly concave.

Then the optimal harvesting strategy is continuous and given by

$$\begin{aligned}\begin{aligned} v&= {\left\{ \begin{array}{ll} 0 &{}\quad \text{ if } \; [\Phi ']^{-1}(V_x^*(x))\le 0,\\ \displaystyle \frac{[\Phi ']^{-1}(V_x^*(x))}{x} &{}\quad \text{ if } \; 0<[\Phi ']^{-1}(V_x^*(x))<xM,\\ M &{}\quad \text{ if } \; [\Phi ']^{-1}(V_x^*(x))\ge xM. \end{array}\right. } \\ \end{aligned} \end{aligned}$$

Furthermore, the HJB equation for the system becomes

$$\begin{aligned} \begin{aligned} \rho&= {\left\{ \begin{array}{ll} x\mu (x)f_x+\dfrac{1}{2}\sigma ^2 x^2f_{xx} &{} \text{ if } [\Phi ']^{-1}(f_x(x))\le 0,\\ x\mu (x)f_x+\dfrac{1}{2}\sigma ^2 x^2f_{xx}\\ -f_x[\Phi ']^{-1}(f_x) + \Phi ([\Phi ']^{-1}(f_x)) &{}\text{ if } 0<[\Phi ']^{-1}(f_x(x))<xM,\\ x(\mu (x)-M)f_x+\dfrac{1}{2}\sigma ^2 x^2f_{xx}+\Phi (xM) &{}\text{ if } [\Phi ']^{-1}(f_x(x))\ge xM. \end{array}\right. } \end{aligned} \end{aligned}$$
(3.1)

Remark 3.1

We cannot find the exact form of the optimal harvesting strategies in this case. Note that in Theorem 2.1 we have \(\Phi (x)=x\) which is not strictly concave nor strictly convex.

Intuitively, this is not unlike maximising a strictly concave objective function under a linear constraint. The optimal choice usually moves smoothly over the domain as the direction of the constraint changes. However, when the objective function is weakly convex, e.g. linear, the optimum will jump on the allowed interval.

Here, we show that if the yield function is weakly convex, the optimal control is bang–bang. The optimal strategy has a similar form to the one for linear yield, if a further assumption on the joint rates of change of the population growth rate and the yield function is made.

Theorem 3.2

Assume that \(\Phi :\mathbb {R}_+\rightarrow \mathbb {R}_+\) is weakly convex, \(\Phi \) grows at most polynomially, \(\Phi \in C^1(\mathbb {R_+})\) and the population survives in the absence of harvesting, that is \(\mu (0)-\frac{\sigma ^2}{2}>0\). Furthermore assume that the drift function \(\mu (\cdot )\) satisfies the following modification of Assumption 2.1:

  1. (i)

    \(\mu \) is locally Lipschitz.

  2. (ii)

    \(\mu \) is decreasing.

  3. (iii)

    As \(x\rightarrow \infty \) we have \(\mu (x)\rightarrow -\infty \).

  4. (iv)

    The function

    $$\begin{aligned} G(x)= \Phi (xM)\left( 1-\frac{2}{\sigma ^2}\mu (x)\right) - xM\Phi '(xM) \end{aligned}$$
    (3.2)

    has a unique extreme point in \((0,\infty )\) which is a minimum, and is not constant on any interval \((u,v)\subset \mathbb {R}_+\).

If the assumptions (i)–(iii) hold, the optimal control has a bang–bang form (i.e., the harvesting rate is either 0 or the maximal M). If assumptions (i)–(iv) hold, the optimal harvesting strategy v has a bang–bang form with one threshold

$$\begin{aligned} \begin{aligned} v(x)&= {\left\{ \begin{array}{ll} 0 &{} \quad \text{ if } \; 0< x\le x^* \\ M &{}\quad \text{ if } \; x>x^* \end{array}\right. } \end{aligned} \end{aligned}$$

for some \(x^*\in (0,\infty )\).

4 The logistic case: \(\mu (x)=\overline{\mu }-\kappa x\)

Throughout this section we provide a thorough analysis of the logistic Verhulst–Pearl model. As such, we will assume that the growth rate is \(\mu (x)=\overline{\mu }-\kappa x\) for positive constants \(\overline{\mu }, \kappa >0\). It is clear that this \(\mu (\cdot )\) satisfies Assumption 2.1. If we harvest according to a constant strategy\(\ell >0\) then the SDE (2.3) becomes

$$\begin{aligned} dX(t) = X(t)(\overline{\mu }- \kappa X(t) -\ell )\,dt + \sigma X(t)\,dB(t). \end{aligned}$$

It is then easy to see that, as long as \(\overline{\mu }-\ell -\frac{\sigma ^2}{2}>0\), the asymptotic yield is

$$\begin{aligned} L(\ell ):=\lim _{T\rightarrow \infty }\dfrac{1}{T}\int _0^T \ell X(t)\,dt = \ell \frac{\overline{\mu }-\ell -\frac{\sigma ^2}{2}}{\kappa }. \end{aligned}$$

We can maximize this yield \(L(\ell )\), which is quadratic in \(\ell \). The maximum will be at

$$\begin{aligned} \ell ^* = \frac{1}{2}\left( \overline{\mu }-\frac{\sigma ^2}{2}\right) \end{aligned}$$

and the maximal asymptotic yield (among constant harvesting strategies) is

$$\begin{aligned} L(\ell ^*)= \frac{\left( \overline{\mu }-\frac{\sigma ^2}{2}\right) ^2}{4\kappa }. \end{aligned}$$

Note that \(L(\ell ^*)\) is also called maximum sustainable yield (MSY) in the literature. Since \(x\mu (x)=\overline{\mu }x-\kappa x^2\) we note that

$$\begin{aligned} \sup _{x\in \mathbb {R}_+} x\mu (x)=\frac{\overline{\mu }^2}{4\kappa }. \end{aligned}$$

Combining this with (2.7) one sees that the optimal asymptotic yield \(\rho ^*\) satisfies

$$\begin{aligned} \frac{\left( \overline{\mu }-\frac{\sigma ^2}{2}\right) ^2}{4\kappa }\le \rho ^*\le \frac{\overline{\mu }^2}{4\kappa }. \end{aligned}$$

Note that Theorem 2.1 does not give us information about \(x^*\), the point at which one starts harvesting.

One possible strategy to find out more information about \(x^*\) is the following: Look at controls of bang–bang type that have a threshold at \(\eta \) and then maximize over all possible \(\eta \). This will then give us a way of finding \(x^*\). Let \(w(x;\eta )\) be the harvesting strategy

$$\begin{aligned} \begin{aligned} w(x;\eta )&= {\left\{ \begin{array}{ll} 0 &{}\quad \text{ if } \; 0< x\le \eta \\ M &{}\quad \text{ if } \; x>\eta . \end{array}\right. } \end{aligned} \end{aligned}$$
(4.1)

For this control w our diffusion (2.3) (with \(h\equiv w\)) is of the form

$$\begin{aligned} dX(t) = a(X(t))\,dt + b(X(t))\,dB(t) \end{aligned}$$
(4.2)

for

$$\begin{aligned} a(x)= x(\overline{\mu }-w(x,\eta )-\kappa x) \end{aligned}$$

and

$$\begin{aligned} b(x)=\sigma x. \end{aligned}$$

Standard diffusion theory shows (see Hening and Kolb 2018; Borodin and Salminen 2012) that the boundary 0 is natural and the boundary \(\infty \) is entrance for the process X from (4.2). As a result, when \(\mu -\frac{\sigma ^2}{2}>0\), one can show using Borodin and Salminen (2012) that the density \(\rho :(0,\infty )\rightarrow (0,\infty )\) of the invariant measure \(\pi \) is of the form

$$\begin{aligned} \begin{aligned} \rho (y)&= \frac{C_1}{b^2(y)}\exp \left( 2\int _{\eta }^y \frac{a(z)}{b^2(z)}\,dz\right) \\&=\frac{C_1}{\sigma ^2 y^2} \exp \left( 2\int _{\eta }^y \frac{z(\overline{\mu }-w(z,\eta )-\kappa z)}{\sigma ^2 z^2}\,dz\right) \\&= {\left\{ \begin{array}{ll} \frac{C_1}{\sigma ^2 y^2}\left( \frac{y}{\eta }\right) ^{\frac{2\overline{\mu }}{\sigma ^2}}e^{-\frac{2\kappa }{\sigma ^2}(y-\eta )} &{} \text{ if } 0< y\le \eta \\ \frac{C_1}{\sigma ^2 y^2}\left( \frac{y}{\eta }\right) ^{\frac{2(\overline{\mu }-M)}{\sigma ^2}}e^{-\frac{2\kappa }{\sigma ^2}(y-\eta )} &{} \text{ if } y>\eta , \end{array}\right. } \end{aligned} \end{aligned}$$
(4.3)

where \(C_1\) is a normalizing constant given by

$$\begin{aligned} \frac{1}{C_1} = \int _0^{\eta }\frac{1}{\sigma ^2 y^2}\left( \frac{y}{\eta }\right) ^{\frac{2\overline{\mu }}{\sigma ^2}}e^{-\frac{2\kappa }{\sigma ^2}(y-\eta )}\,dy + \int _{\eta }^\infty \frac{1}{\sigma ^2 y^2}\left( \frac{y}{\eta }\right) ^{\frac{2(\overline{\mu }-M)}{\sigma ^2}}e^{-\frac{2\kappa }{\sigma ^2}(y-\eta )}\,dy. \end{aligned}$$

In this case the harvesting yield is

$$\begin{aligned} \begin{aligned} H(\eta )&:=\lim _{T\rightarrow \infty }\dfrac{1}{T}\int _0^T \Phi \Big (X(t)w(X(t),\eta )\Big )\,dt \\&= \int _{\mathbb {R}_{++}} y w(y,\eta )\pi (dy) \\&= \int _0^\infty y w(y,\eta )\rho (y)dy\\&= \int _{\eta }^\infty yM\frac{C_1}{\sigma ^2 y^2}\left( \frac{y}{\eta }\right) ^{\frac{2(\overline{\mu }-M)}{\sigma ^2}}e^{-\frac{2\kappa }{\sigma ^2}(y-\eta )}\,dy\\&= \frac{\displaystyle \int _{\eta }^\infty yM\frac{1}{\sigma ^2 y^2}\left( \frac{y}{\eta }\right) ^{\frac{2(\overline{\mu }-M)}{\sigma ^2}}e^{-\frac{2\kappa }{\sigma ^2}(y-\eta )}\,dy}{\displaystyle \int _0^{\eta }\frac{1}{\sigma ^2 y^2}\left( \frac{y}{\eta }\right) ^{\frac{2\overline{\mu }}{\sigma ^2}}e^{-\frac{2\kappa }{\sigma ^2}(y-\eta )}\,dy + \displaystyle \int _{\eta }^\infty \frac{1}{\sigma ^2 y^2}\left( \frac{y}{\eta }\right) ^{\frac{2(\overline{\mu }-M)}{\sigma ^2}}e^{-\frac{2\kappa }{\sigma ^2}(y-\eta )}\,dy} \end{aligned} \end{aligned}$$
(4.4)

By Theorem 2.1 the point \(x^*\) has to satisfy:

$$\begin{aligned} H(x^*)= \max _{\eta \in (0,\infty )}H(\eta ). \end{aligned}$$

It is clear that H is differentiable, that \(x^*\) exists and satisfies \(x^*\in (0,\infty )\). Therefore, \(x^*\) is a solution of

$$\begin{aligned} H'(\eta )=0. \end{aligned}$$
(4.5)

The condition above can be restated as an equation involving incomplete gamma functions. We were not able to prove analytically that (4.5) has a unique solution. Berg and Pedersen (2006, 2008) show possible analytical methods that can be applied to such equations in a simple case. However, numerical experiments that we have done support this conjecture (see Fig. 1).

Fig. 1
figure 1

Typical shape of the asymptotic yield function H(x) as a function of the harvesting threshold x, where one begins to harvest. Here for \(\sigma ^2=1\) and \(M=\overline{\mu }=\kappa =1\)

Conjecture 4.1

There exists a unique \(x^*\in (0,\infty )\) such that \(H'(x^*)=0\). Furthermore, the optimal harvesting strategy is given by

$$\begin{aligned} \begin{aligned} v(x)&= {\left\{ \begin{array}{ll} 0 &{}\quad \text{ if } \; 0< x\le x^* \\ M &{}\quad \text{ if } \; x>x^*. \end{array}\right. } \end{aligned} \end{aligned}$$

5 Discussion and future research

We have analysed a population whose dynamics evolves according to generalization of the logistic Verhulst–Pearl model in a stochastic environment, but subjected to strategic harvesting. The rate at which the population gets harvested is bounded above by a constant \(M>0\), and the harvested infinitesimal amount is proportional to the current size of the population. We show that the harvesting strategy v, which describes the harvesting rate and is chosen to maximize the asymptotic harvesting yield

$$\begin{aligned} \liminf _{T\rightarrow \infty }\dfrac{1}{T}\int _0^T X(t)h(t)\,dt, \end{aligned}$$

is of bang–bang type, i.e. there exists \(x^*>0\) such that

$$\begin{aligned} \begin{aligned} v(x)&= {\left\{ \begin{array}{ll} 0 &{} \text{ if } 0< x\le x^* \\ M &{} \text{ if } x>x^*. \end{array}\right. } \end{aligned} \end{aligned}$$
Fig. 2
figure 2

Graph of the maximal asymptotic yield H(x) as a function of the harvesting threshold x for different values of the growth rate \(\overline{\mu }\). We take \(\sigma ^2=1\), \(M=\kappa =1\), and \(\overline{\mu }=1\) (blue), \(\overline{\mu }=1.5\) (orange), \(\overline{\mu }=2\) (green), \(\overline{\mu }=2.5\) (red), and \(\overline{\mu }=3\) (purple) (color figure online)

5.1 Logistic Verhulst–Pearl

In the particular case when \(\mu (x)=\overline{\mu }- \kappa x\), we can give more information about \(x^*\) as follows: The harvesting yield function \(H(\eta )\) is determined, by letting the jump in the bang–bang control be at \(\eta \). That is, we look at the yield when the control is

$$\begin{aligned} \begin{aligned} v(x)&= {\left\{ \begin{array}{ll} 0 &{} \text{ if } 0< x\le \eta \\ M &{} \text{ if } x>\eta . \end{array}\right. } \end{aligned} \end{aligned}$$

The typical behavior of the point \(x^*\) where H is maximized and of \(H(x^*)\) as the parameters \(\overline{\mu }, \kappa \) and M change was analyzed numerically and is presented in Figs. 2, 3 and 4, with the normalization \(\sigma ^2=1\).

We note from numerical experiments that increasing the growth rate \(\overline{\mu }\) increases the threshold \(x^*\) at which one should start optimally harvesting (Fig. 2). This is an intuitive result, since an increased growth rate increases the maximal equilibrium value of the population in the equivalent deterministic growth model with competition (Smith 1978). Therefore, it should also increase asymptotic harvesting yield, as well as the point at which harvesting should start. Moreover, higher growth rates make the population get faster to the point \(x^*\) where one starts harvesting, reducing the cost of a delay.

If one increases the maximal harvesting rate M then the harvesting threshold \(x^*\) is also increased (Fig. 3). This also makes sense because if \(\overline{\mu }-\frac{\sigma ^2}{2}-M<0\), then a population with constant harvesting rate M will go extinct almost surely. An increase in the harvesting threshold \(x^*\) is necessary to make sure that there is no extinction. Moreover, as M gets larger one can wait longer to start harvesting. With a larger maximal rate available, there is less chance that there will be losses because the population overshoots the optimal extraction point. Similarly, increasing the harvesting rate M also increases the maximal asymptotic harvesting yield, for the obvious reason that there is better control on the population level and therefore extraction can happen closer to the optimal level.

In contrast, if one increases the intraspecific competition rate \(\kappa \), then the harvesting threshold decreases (Fig. 4). The equilibrium value of the population in the equivalent deterministic model (Smith 1978) decreases with \(\kappa \), and as a result so does the extraction rate. Evidently, even in the stochastic model, if competition is very strong the population cannot spend much time at high densities, and therefore one has to start harvesting early. An increase in \(\kappa \) will also decrease the maximal asymptotic harvesting yield.

Fig. 3
figure 3

Graph of the maximal asymptotic yield H(x) as a function of the harvesting threshold x for different values of the maximal harvesting rate M. We take \(\sigma ^2=1\), \(\overline{\mu }=\kappa =1\), and \(M=0.1\) (blue), \(M=0.2\) (orange), \(M=0.5\) (green), \(M=1\) (red), \(M=2\) (purple), and \(M=5\) (brown) (color figure online)

Fig. 4
figure 4

Graph of the maximal asymptotic yield H(x) as a function of the harvesting threshold x for different values of the intra-competition rate. We take \(\sigma ^2=1\), \(M=\overline{\mu }=1\), and \(\kappa =1\) (blue), \(\kappa =2\) (orange), \(\kappa =3\) (green), \(\kappa =4\) (red), and \(\kappa =5\) (purple) (color figure online)

We are able to prove that the maximal asymptotic yield \(\rho ^*\) satisfies the inequality

$$\begin{aligned} \frac{\left( \overline{\mu }-\frac{\sigma ^2}{2}\right) ^2}{4\kappa }\le \rho ^*\le \frac{\overline{\mu }^2}{4\kappa }. \end{aligned}$$

In particular, the bang–bang optimal strategy has a higher asymptotic yield than the optimal constant harvesting strategy. Moreover, the bang–bang optimal strategy gives a lower asymptotic yield than the optimal constant harvesting strategy in the absence of noise. This means that the analysis of the more complex stochastic model was fruitful, recommending a qualitatively different strategy. Moreover, environmental fluctuations decrease the maximal asymptotic yield and, because the correction is negative, protecting a population from extinction requires a careful measurement of natural fluctuations when designing optimal harvesting. When environmental stochasticity was not taken into account, harvesting often lead populations to extinction (Lande et al. 1995).

Real populations do not evolve in isolation. As a result, ecology is concerned with understanding the characteristics that allow species to coexist. Harvesting can disturb the coexistence of species. In future research we intend to tackle multi-dimensional analogues of the setting treated in the current article. Natural models for which one can add harvesting would be predator-prey food chains (Gard and Hallam 1979; Gard 1984; Hening and Nguyen 2018b, c; Tyson and Lutscher 2016), more general Kolmogorov systems (Schreiber et al. 2011; Hening and Nguyen 2018a) and structured populations where there can be asymmetric harvesting (Evans et al. 2013, 2015; Hening et al. 2018; Roth and Schreiber 2014; Benaïm and Schreiber 2009; Schreiber and Ryan 2011). In the multi-dimensional setting the Hamilton–Jacobi–Bellman (HJB) equation becomes a PDE and the analysis becomes significantly more complex. New tools will have to be developed to tackle these problems.

Above we have imposed a bound on the extraction rate, M. This was because it is a realistic feature, but it was also practical for the analysis. Nevertheless, it is interesting to consider the case when the extraction rate is unbounded. A practical model with no extraction limit corresponds to having unlimited control over a target population, which is sometimes the case. Such a model would have the benefit of not requiring a nuisance parameter that may be hard to determine.

5.2 Concave and convex yield functions

We have also studied the more general case involving concave and convex yield functions. When the yield function is strictly concave, it was shown that the optimal control is not bang–bang, but continuous in the population parameter. Vice-versa, when the yield function is weakly convex, we have shown that the optimal control is necessarily bang–bang. Moreover, if a certain further assumption on the relative rate of growth of \(\mu \) and \(\Phi \) holds, we can also show that the bang–bang optimal control has a single threshold \(x^*\) where the extraction rate goes from 0 to M—as in the linear special case.

This generalization allows us to think of applications of population harvesting where the yield function is in fact a utility function, or some other more general social welfare measure.

5.3 Unbounded harvesting

If we allow for general, possibly unbounded, harvesting we would have to study the Skorokhod SDE

$$\begin{aligned} d\tilde{X}(t) = \tilde{X}(t)(\mu - \kappa \tilde{X}(t))\,dt + \sigma \tilde{X}(t)\,dB(t) - dZ_t, \quad \tilde{X}(0)=x>0. \end{aligned}$$
(5.1)

where \((Z_t)_{t\ge 0}\) is supposed to be non-negative, increasing, right-continuous and adapted to \((\mathcal {F}_t)_{t\ge 0}\)—we denote the set of all such strategies by A. Then the problem is to maximize the asymptotic yield, i.e. find

$$\begin{aligned} V(x) = \sup _{(Z_t)_{t\ge 0}\in A}\liminf _{T\rightarrow \infty }{\mathbb {E}}_x \frac{1}{T}\int _0^T dZ_t = \sup _{(Z_t)_{t\ge 0}\in A}\liminf _{T\rightarrow \infty } \frac{{\mathbb {E}}_x Z_T}{T} \end{aligned}$$

We want to find the harvesting strategy \((Z_t^*)_{t\ge 0}\in A\), which we call the optimal harvesting strategy, such that

$$\begin{aligned} V(x) = \liminf _{T\rightarrow \infty }\frac{{\mathbb {E}}_x Z^*_T}{T}. \end{aligned}$$

The analysis above, for the bounded harvesting rate, determined that the optimal strategy has a bang–bang property, where extraction is maximal after some cut-off. This suggests that raising the maximum would not change the bang–bang property, but determining that result required a bounded extraction rate. Thinking of the limiting behavior of the yield function above shows the difficulty: as \(M \rightarrow \infty \), the density of the distribution above the cut-off \(x^*\) goes to 0 (see (4.3)). The conjectured optimal solution is akin to having a reflective boundary at \(x^*\), and the yield is determined by the time spent close to the boundary.

Conjecture 5.1

Assume that the population survives in the absence of harvesting i.e. \(\mu -\frac{\sigma ^2}{2}>0\). The optimal extraction strategy \((Z^*_t)_{t\ge 0}\) has the form

$$\begin{aligned} \begin{aligned} Z_t^*(x)&= {\left\{ \begin{array}{ll} (x-x^*)^+ &{} \text{ if } t=0 \\ L(t,x^*)&{} \text{ if } t>0. \end{array}\right. } \end{aligned} \end{aligned}$$
(5.2)

for some \(x^*\in (0,\infty )\), where \(L(t,x^*)\) is the local time at \(x^*\) of the process \(\tilde{X}\) from (5.1).

This conjecture is supported by the results from Alvarez and Shepp (1998) where the authors study the maximization of the discounted yield

$$\begin{aligned} V(x):=\sup _{(Z_t)_{t\ge 0}\in A}{\mathbb {E}}_x \int _0^{\tau }e^{-rt} dZ_t \end{aligned}$$

and \(\tau :=\inf \{t\ge 0: \tilde{X}_t=0\}\) is the extinction time. It is shown in Alvarez and Shepp (1998) that the optimal harvesting strategy is of the form (5.2). One possible approach to prove Conjecture 5.1 would be to use the results from Alvarez and Shepp (1998) and then let the discount factor r go to 0.