Skip to main content
Log in

Poverty, social preference for employment, and natural resource depletion

  • Research Article
  • Published:
Environmental Economics and Policy Studies Aims and scope Submit manuscript

Abstract

We show that in poor resource-based communities, the socio-psychological preference for employment, which arises from a strong desire to follow the communal norm of sharing in harvesting efforts, can lead to the optimality of full-employment harvesting until resource extinction. We show that such communities may be able to sustain both their natural resources and full employment by using outside-the-community employment opportunities or by economic diversification. However, to be effective, such policies must ensure that the outside wage rate and the initial capital stock are above certain minimum levels which depend on the existing size of the resource stock, the characteristics of the community’s harvesting technology, and the biological growth characteristics of the resource in question, and which will be higher the longer the remedial policies are delayed.

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.

Similar content being viewed by others

Notes

  1. Diamond’s accounts of these collapses are, however, mainly descriptive histories and fall short of providing analytical, let alone economic, explanations. For an insightful review and assessment of the relevance of diamond’s work to modern civilizations, see Page (2005).

  2. Note that the well-known case of extinction of a population under the open-access regime (sometimes known, incorrectly, also as the “tragedy of the commons”) may be viewed as a special case of this result when the discount rate approaches infinity. However, Cropper and Lee (1979) and Cropper (1988) show that this result need not hold when the harvest price is allowed to be inversely related to the harvest rate rather than assumed to be constant, as in Clark (1973). In that case, the result holds only for sufficiently small initial stocks, but not for sufficiently large initial stocks even if the discount rate is infinite. In a related paper, Smith (1975) explains the mass extinctions of mega fauna in America during the late Pleistocene by using an open-access resource (free-access hunting) model. He shows that if the average biological growth rate at zero stock level is low and less than the equilibrium harvest per unit biomass, the extinction of a species will occur due to hunting pressure.

  3. It has been well known that in the game theoretic framework, the resource extraction towards finite time extinction can occur as a result of “individual” rationality under open-access regime. Gordon (1954) and Hardin’s (1968) influential paper, “the tragedy of the commons”, are among the early works alluding to this, followed by a number of rigorous analyses including Dasgupta (1982); Cornes and Sandler (1983, 1996); and Sorger (1998), among others. It should be emphasized, however, that in our model there is no open access to the resource and no strategic interactions among the community members who are assumed to fully and collectively own the resource.

  4. The assumption of constant population is more of an analytical convenience. An exogenously growing population would only reinforce our result by raising the full employment harvest rate at any time and hence shortening the time period to extinction. We shall have more to say on this in Sect. 5.

  5. For an analysis of the effect of nonpecuniary work motivation on labor supply, see Farzin (2009).

  6. The time arguments of the functions are omitted whenever no confusion arises.

  7. Technically, the condition requires that at the stationary point \(\tilde{E}\), the harvest function H be more concave in E than the welfare function w is, implying that \(\frac{d}{dE}\left( {{{w_{E} } \mathord{\left/ {\vphantom {{w_{E} } {H_{E} }}} \right. \kern-0pt} {H_{E} }}} \right) > 0\) at that point.

  8. To see this formally, recall that \(w(E,S) \equiv W[H(E,S),E] = W(C,E)\). Differentiate 3 with respect to E to get \(W_{C} H_{E} = - W_{E} + \lambda H_{E}\), and use this to obtain \(\begin{gathered} \frac{{ - EH_{EE} }}{{H_{E} }} - \frac{{ - Ew_{EE} }}{{w_{E} }} = \frac{{ - EH_{EE} }}{{H_{E} }} - \frac{{ - E\left( {W_{CC} \left( {H_{E} } \right)^{2} + 2W_{CE} H_{E} + W_{C} H_{EE} + W_{EE} } \right)}}{{\lambda H_{E} }} \hfill \\ = \frac{{ - EH_{EE} }}{{H_{E} }}\left( {1 - \frac{{W_{C} }}{\lambda }} \right)\, - \,\frac{{ - E\left( {W_{CC} \left( {H_{E} } \right)^{2} + 2W_{CE} H_{E} + W_{EE} } \right)}}{{\lambda H_{E} }} = \frac{{ - EH_{EE} }}{{H_{E} }}\left( {\frac{{W_{E} }}{{\lambda H_{E} }}} \right)\, - \,\frac{{ - E\left( {W_{CC} \left( {H_{E} } \right)^{2} + 2W_{CE} H_{E} + W_{EE} } \right)}}{{\lambda H_{E} }} < 0, \hfill \\ \end{gathered}\) where all functions are evaluated at \((\tilde{E}(S,\lambda ),S,\lambda )\).

  9. It is also worth noting that the condition 6 does not hold if the harvest function is convex or linear in E (i.e., if \(H_{EE} \ge 0\)), although the latter has been commonly assumed in the literature.

  10. To see this, suppose \(\lim_{S \to 0} H_{S} (\bar{E},S) = \xi < \infty\). Then, we can approximate the harvest function with \(\xi {\kern 1pt} S\) and the natural growth function with \(rS\) in a neighborhood of the origin. Thus, \(\dot{S} = G(S) - H(E,S) \approx S\left( {r - \xi } \right)\). With a linear differential equation, it takes infinite time to extinction.

  11. Notice that the utility function being bounded away from below is necessary to have an optimal path with finite time extinction. Also, notice that without direct welfare effect of employment, one would conventionally have the condition \(\mathop {\lim }\nolimits_{C \to 0} W_{C} (C) = \infty\), which would defy the optimality of resource extinction in finite time.

  12. For the proof of the existence of an optimal path, see “Appendix A3”.

  13. Notice, again, from 11 and \(\eta_{c} < 1\) that the full employment obsession, and hence the finite time extinction, can occur only if employment also directly enhances the individual’s utility and therefore the community’s welfare, i.e., only if \(\eta_{E} > 0\).

  14. It can also be shown that when economic assistance to the community is in the form of a real flow of lump-sum transfers or foreign grants (denoted by \(X(t) \ge 0,\;\;\forall t \ge 0\)), one has \(\frac{{ - EH_{EE} }}{{H_{E} }} - \frac{{ - Ew_{EE}^{X} }}{{w_{E} }} = \eta_{E} - \alpha + \alpha \eta_{c} {\kern 1pt} \varphi \left( {1 + \frac{X}{H}} \right),\) where \(w^{X} (E,S,t) = W[H(E,S) + X,E] = [H(E,S) + X]^{{\eta_{c} }} E^{{\eta_{E} }}\) and the function \(\varphi\) is defined by \(\varphi (x) \triangleq \frac{{(\alpha + \eta_{E} )x - \alpha (1 - \eta_{c} )}}{{x(\alpha \eta_{c} + \eta_{E} \,x)}}, \, x = \left( {1 + \frac{X}{H}} \right) \ge 1.\) Since \(\lim_{x \to \infty } \varphi (x) = 0\), it follows that such policies can be effective (i.e., \(\eta_{E} - \alpha + \alpha \eta_{c} {\kern 1pt} \varphi \left( {1 + \frac{X}{H}} \right) \le 0\)) if the rate of transfer relative to harvest rate is at any time sufficiently large and if \(\eta_{E} < \alpha\). However, if \(\eta_{E} \ge \alpha ,\), the full employment obsession path is optimal and resource extinction occurs in finite time, no matter how large the flow of transfers.

  15. The effect of this latter type of technological change on resource sustainability depends crucially on whether the marginal stock effect on biological growth rate, \(G_{S}\), is larger or smaller than its effect on the harvest rate, \(H_{S}\). If \(G_{S} > H_{S}\), it may contribute to resource sustainability; otherwise it can accelerate resource extinction.

  16. The assumption of constant real wage rate is made for expositional simplicity. The results that follow also hold for time dependent wage rates \(\omega (t)\).

  17. This assumption is used only to derive the monotonicity of the steady state in the wage rate.

  18. It is not surprising that the potential optimal steady state is greater than the modified golden rule, because the harvest function \(H(E,S)\) implies the presence of the so-called wealth effect.

  19. Also, notice that a concave harvest (production) function seems a quite standard assumption in economics and therefore the emergence of the nonconvexity problem does not seem unusual in resource economics. In our knowledge, however, this has not been addressed so far.

  20. Notice that because of the marginal social cost of employment due to the negative stock effect of increased harvest rate (i.e., \(- \lambda H_{E}\)), the optimal employment allocation requires that the marginal product of employment in the resource sector to exceed the outside wage rate.

  21. This follows from \(\mathop {\lim }\limits_{S \to 0} \left[ {\omega (S)^{\alpha /(1 - \alpha )} /(\alpha \gamma^{1/\alpha } )} \right] = \mathop {\lim }\limits_{S \to 0} \left[ {S^{\beta /(1 - \alpha )} /G(S)} \right] = {{\mathop {\lim }\limits_{S \to 0} [\beta /(1 - \alpha )]S^{(\alpha + \beta - 1)/(1 - \alpha )} } \mathord{\left/ {\vphantom {{\mathop {\lim }\limits_{S \to 0} [\beta /(1 - \alpha )]S^{(\alpha + \beta - 1)/(1 - \alpha )} } r}} \right. \kern-0pt} r} = \infty .\)

  22. Note that we are implicitly assuming that investment in the industry sector is irreversible. We are also assuming that the initial capital stock, \(K_{0}\), is given to the community either as a capital grant from the government or as foreign aid.

  23. Alternatively, we can assume that the manufactured good and the resource harvest are perfect substitutes in consumption. Farm-raised fish and harvested wild fish present an example.

  24. Of course, if the current resource users happen to be historically vigilant of resource stock depletion and if they sufficiently care about the resource stock size they leave for future generations, then the socially optimal outcome may not be resource extinction. Analytically, an easy way to allow for this is to postulate a social welfare function where well-being increases not only with the levels of consumption and employment, but also sufficiently strongly with the remaining stock size at each time.

References

  • Argyle M (2001) The psychology of happiness. Taylor & Francis, New York

  • Brander JA, Taylor MS (1998) The simple economics of Easter Island: a Ricardo-malthus model of renewable resource use. Am Econ Rev 88(1):119–138

    Google Scholar 

  • Clark CW (1973) Profit maximization and the extinction of animal species. J Polit Econ 81:950–961

    Article  Google Scholar 

  • Clark AE (2003) Unemployment as a social norm: psychological evidence from panel data. J Labor Econ 21(2):323–351

    Article  Google Scholar 

  • Clark AE, Oswald AJ (1994) Unhappiness and unemployment. Econ J 104(424):648–659

    Article  Google Scholar 

  • Cornes R, Sandler T (1983) On commons and tragedies. Am Econ Rev 73:787–792

    Google Scholar 

  • Cornes R, Sandler T (1996) The theory of externalities, public goods and club goods, 2nd edn. Cambridge University Press, Cambridge

    Book  Google Scholar 

  • Cropper ML (1988) A note on the extinction of renewable resources. J Environ Econ Manag 15(1):64–70

    Article  Google Scholar 

  • Cropper ML, Lee DR (1979) The optimal extinction of a renewable natural resource. J Environ Econ Manag 6(4):341–349

    Article  Google Scholar 

  • Dasgupta P (1982) The control of resources. Basil Blackwell, Oxford

    Google Scholar 

  • Dasgupta P, Maler KG (1995) Poverty, Institutions, and the Environmental Resource-Base, chapter 39 In: Behrman J, Srinivasan TN (eds) Handbook of Development Economics. vol. IIIA, Elsevier Science, The Netherlands

  • Di Tella R, MacCulloch RJ, Oswald AJ (2001) Preferences over inflation and unemployment: evidence from surveys of happiness. Am Econ Rev 91(1):335–341

    Article  Google Scholar 

  • Di Tella R, MacCulloch RJ, Oswald AJ (2003) The macroeconomics of happiness. Rev Econ Stat 85(4):809–827

    Article  Google Scholar 

  • Diamond J (2005) Collapse: how societies choose to fail or succeed, Viking Penguin

  • Dockner EJ, Nishimura K (2005) Capital accumulation games with a non-concave production function. J Econ Behav Organ 57:408–420

    Article  Google Scholar 

  • Farzin YH (2009) The effect of non-pecuniary motivations on labor supply. Q Rev Econ Finance 49:1236–1259

    Article  Google Scholar 

  • Feather NT (1990) The psychological impact of unemployment. Springer, New York

    Book  Google Scholar 

  • Frey BS, Stutzer A (2002) What can economists learn from happiness research? J Econ Lit XL:402–435

    Google Scholar 

  • Gordon JR (1954) The economic theory of a common property resource: the fishery. J Political Econ 62:1031–1039

    Article  Google Scholar 

  • Hardin G (1968) The tragedy of the commons. Science 162:1243–1247

    Article  Google Scholar 

  • Jahoda M (1981) Work, employment and unemployment: values, theories, and approaches in social research. Am Psychol 36:184–191

    Article  Google Scholar 

  • Karpoff JM (1985) Nonpecuniary benefits in commercial fishing: empirical findings from the Alaska salmon fisheries. Econ Inq 23(1):159–174

    Article  Google Scholar 

  • Lucas RE, Clark AE, Georgellis Y, Diener E (2004) Unemployment alters the set point for life satisfaction. Psychol Sci 15(1):8–13

    Article  Google Scholar 

  • Page SE (2005) Are we collapsing? A review of Jared diamond’s collapse: how societies choose to fail or succeed? J Econ Lit XLIII:1049–1062

    Google Scholar 

  • Ramsey FP (1928) A mathematical theory of saving. Econ J 38:543–559

    Google Scholar 

  • Romer P (1986) Cake eating, chattering, and jumps: existence results for variational problems. Econometrica 54:897–908

    Article  Google Scholar 

  • Smith VL (1975) The primitive hunter culture, pleistocene extinction, and the rise of agriculture. Am Econ Rev 83(4):727–756

    Google Scholar 

  • Sorger G (1998) Markov-perfect Nash equilibria in a class of resource games. J Econ Theory 11:78–100

    Google Scholar 

  • Spence AM (1973) Blue whales and applied control theory. In: Systems approaches to environmental problems, Bavarian Academy of Sciences, conference values, June 1973

  • Veblen T (1899) The theory of the leisure class, A.M. Kelly, New York, 1975

  • Whelan C (1994) Social class, unemployment and psychological distress. Eur Sociol Rev 10(1):49–61

    Google Scholar 

  • Winelmann L, Winkelmann R (1998) Why are the unemployed so unhappy? Evidence from panel data. Economica 257(65):1–15

    Article  Google Scholar 

Download references

Acknowledgments

For helpful comments and suggestions, we thank several annonymous referees, Ed Barbier, Antonio Bento, Carol McAusland, Chris Costello, Bob Deacon, Phil Martin, Chuck Mason, Kazuo Nishimura, Prasanta Pattanaik, Michael Rauscher, Kazuhiro Ueta, participants at The 3rd World Congress of Environmental and Resource Economists, Kyoto, and at seminars at UC-Berkeley, UC-Davis, UC-Riverside, UC-Santa Barbara, Kobe U., Kyoto U., Maryland U., Osaka U., UCF, Waseda U., and Wyoming U. Any remaining shortcoming is our responsibility.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Y. H. Farzin.

Appendices

Appendix

A1. Proof of Proposition 1

We first prove the following lemmas:

Lemma 1

If an interior effort level \(E = \tilde{E} \in (0,\bar{E})\) is optimal for a stock level S > 0, then

$$\varPsi_{EE} \left( {\tilde{E},S,\frac{{w_{E} (\tilde{E},S)}}{{H_{E} (\tilde{E},S)}}} \right) \le 0 .$$
(23)

Proof

By Pontryagin’s maximum principle, the following two conditions hold at \(\tilde{E}\):

$$\varPsi_{E} (\tilde{E},S,\lambda ) = w_{E} (\tilde{E},S) - \lambda H_{E} (\tilde{E},S) = 0,\;\varPsi_{EE} (\tilde{E},S,\lambda ) \le 0.$$
(24)

Combine these to get the inequality in the claim. ■

Lemma 2

Let \(\lambda (E,S) = w_{E} (E,S)/H_{E} (E,S)\) . Assume that problem 2 has a solution for each \(S \in [0,k]\) . If \(\varPsi_{EE} (E,S,\lambda (E,S)) > 0\) for all S > 0 and all \(E \in (0,\bar{E})\) , the optimal path is full employment obsession path.

Proof

The contraposition of Lemma 1 is that if \(\varPsi_{EE} \left( {\tilde{E},S,\lambda (\tilde{E},S)} \right) > 0\) for all S > 0 and all \(\tilde{E} \in (0,\bar{E})\), then optimal effort cannot be interior for any stock level S > 0. Therefore, with the condition \(\varPsi_{EE} \left( {\tilde{E},S,\lambda (\tilde{E},S)} \right) > 0\), the optimal policy \(E^{*} (S)\) for the problem 2 satisfies \(E^{*} (S) \in \{ 0,\bar{E}\}\) for all \(S \in (0,k)\). Note that an interior stock level \(S \in (0,k)\) is sustained by interior effort level \(E \in (0,\bar{E})\)such that \(G(S) - H(E,S) = 0\). Therefore, there is no interior optimal steady state. Dockner and Nishimura (2005, Lemma 2) prove that an optimal state path for a continuous time one-state variable optimal control problem is monotone. Since the state space \([0,k]\) is bounded, every monotone optimal state path converges to an optimal corner steady state \(S_{ss}^{*} = 0\) or \(k\). We eliminate \(S_{ss}^{*} = k\), since it implies \(E^{*} (S_{ss}^{*} ) = 0\) in all periods, which is obviously suboptimal. Therefore, every optimal state path monotonically decreases to the zero stock level. The associated optimal policy is \(E^{*} (S) = \bar{E}\) for all \(S \in (0,k]\).■

Proof of Proposition 1

Since \(\varPsi_{EE} \left( {\tilde{E},S,\lambda (\tilde{E},S)} \right) > 0\) is equivalent to 6, the statement is true by Lemma 2.■

A2. Proof of finite time exhaustion

Proposition A1

Let \(\overset{\lower0.5em\hbox{$\smash{\scriptscriptstyle\frown}$}}{S}\) be the supremum such that if \(0 < S < \overset{\lower0.5em\hbox{$\smash{\scriptscriptstyle\frown}$}}{S}\), \(\dot{S} = G(S) - H(\bar{E},S) < 0\) . When initial stock S 0 is in \((0,\overset{\lower0.5em\hbox{$\smash{\scriptscriptstyle\frown}$}}{S} )\) , the path \(S(t;S_{0} )\) described by

$$\dot{S} = G(S) - H(\bar{E},S) = rS\left( {1 - \frac{S}{k}} \right) - \gamma \bar{E}^{\alpha } S^{\beta } , \, S(0) = S_{0} ,$$
(25)

goes to zero within a finite time.

Proof

Notice that \(\dot{S} = G(S) - H(\bar{E},S) < rS - \gamma \bar{E}^{\alpha } S^{\beta }\). Let \(\sigma (t,S_{0} )\) be the solution of \(\dot{\sigma } = r\sigma - \gamma \bar{E}^{\alpha } \sigma^{\beta } , \, \sigma (0) = S_{0}\). This Bernoulli’s differential equation is solved as \(\sigma (t;S_{0} ) = \left[ {\left( {S_{0}^{1 - \beta } - \frac{{\gamma \bar{E}^{\alpha } }}{r}} \right)e^{r(1 - \beta )t} + \frac{{\gamma \bar{E}^{\alpha } }}{r}} \right]^{{^{{\frac{1}{1 - \beta }}} }}\). From this, \(\sigma (t;S_{0} )\) goes to zero within a finite time if the initial stock satisfies \(S_{0}^{1 - \beta } < \gamma \bar{E}^{\alpha } /r\). By construction, \(S(t;S_{0} ) < \sigma (t;S_{0} )\) and thus \(S(t;S_{0} )\) also go to zero within a finite time. Note that since \(\dot{S}(t;S_{0} ) < 0\), \(\lim_{t \to \infty } S(t;S_{0} ) = 0\) is obvious, and we can arbitrarily choose a sufficiently small initial value.■

Proposition A2

If \(\gamma > \frac{{r[(1 - \beta )k]^{1 - \beta } }}{{(\bar{E})^{\alpha } (2 - \beta )^{2 - \beta } }},\) then \(\dot{S} = G(S) - H(\bar{E},S) < 0\) for all \(S \ge 0\).

Proof

Let \(g(S;\gamma ) = rS\left( {1 - S/k} \right) - \gamma \bar{E}^{\alpha } S^{\beta }\). If there is a unique pair (\(\tilde{\gamma }\), \(\tilde{S}\)) such that \(g(\tilde{S};\tilde{\gamma }) = 0, \, g_{S} (\tilde{S};\tilde{\gamma }) = 0,{\text{ and }}g_{SS} (\tilde{S};\tilde{\gamma }) < 0\), then \(g(S;\gamma ) < 0\) for all S > 0 and \(\gamma > \tilde{\gamma }\). Solve the system of equations \(g(\tilde{S};\tilde{\gamma }) = 0{\text{ and }}g_{S} (\tilde{S};\tilde{\gamma }) = 0\) and we have the unique solution

$$\tilde{S} = \frac{1 - \beta }{2 - \beta }k, \, \tilde{\gamma } = \frac{{r[(1 - \beta )k]^{1 - \beta } }}{{(\bar{E})^{\alpha } (2 - \beta )^{2 - \beta } }}$$
(26)

It is easily seen that \(g_{SS} (\tilde{S};\tilde{\gamma }) = - r(2 - \beta )/k < 0\).■

A3. Proof of the existence of optimal paths

We apply the Romer’s Theorem (Romer 1986) to prove the existence, because the theorem is applicable to a non-convex problem as in our model.

The reduced form welfare function for our model is written as:

$$\omega (S,\dot{S}) = \left\{ {\begin{array}{*{20}c} {\gamma^{{ - \eta_{E} /\alpha }} [G(S) - \dot{S}]^{{\eta_{C} + \eta_{E} /\alpha }} S^{{ - \eta_{E} \beta /\alpha }} } & {{\text{if }}\dot{S} \in [G(S) - H(\bar{E},S),G(S)],} \\ { - \infty } & {{\text{otherwise}}.} \\ \end{array} } \right.$$

Since we have assumed \(\eta_{E} - \alpha (1 - \eta_{C} ) > 0\), \(\omega\) is convex in \(\dot{S}\). Hence, to apply the theorem, we introduce an “artificial” argument,

$$\mathop S\limits^{..} (t){\text{ such that}}\text{ }\dot{S}(t) - \dot{S}(s) \equiv \int\limits_{s}^{t} {\mathop S\limits^{..} (\tau )d\tau } ,$$

and rewrite \(\omega (S,\dot{S})\) as \(\omega (S,\dot{S},\mathop S\limits^{..} )\). Notice that for \(\mathop S\limits^{..}\) to exist almost everywhere, \(\dot{S}\) must be absolutely continuous. This restriction is a cost of applying the Romer’s Theorem. Another restriction is that we have to assume that \(\mathop S\limits^{..}\) is essentially bounded, i.e., there exists a real number \(M\) such that \(\mathop {{\text{ess}} \cdot \sup }\limits_{{t^{3} 0}} |\mathop S\limits^{..} (t)| < M\). The Romer’s Theorem ensures the existence of optimal paths if the following two conditions are met:

  1. (1)

    \(\omega (S,\dot{S},\mathop S\limits^{..} )\) is upper-semi continuous and \(\omega (S,\dot{S}, \cdot )\) is concave for all \((S,\dot{S}) \in {\mathbb{R}}^{2} .\)

  2. (2)

    There is a real number \(m\) such that \(\omega (S(t),\dot{S}(t),\mathop S\limits^{..} (t)) \le m - |\mathop S\limits^{..} (t)|\) holds almost everywhere on \([0,\infty )\) and for all \((S(t),\dot{S}(t),\mathop S\limits^{..} (t)).\)

Our model obviously satisfies the first condition. For the second condition, notice that there is an upper bound of \(\omega (S,\dot{S},\mathop S\limits^{..} )\), say \(\bar{\omega }\) (\(< \infty\)). Thus, choose \(m = M + \bar{\omega }\) and then condition 2 is satisfied.■

A4. Proof of \(d\tilde{S}/d\omega > 0\) for the problem 12

By Pontryagin’s maximum principle, an interior optimal path satisfies, with costate \(\lambda \ge 0\):

$$\begin{aligned} & \varPsi_{E} (E,S,\lambda ) = W_{C} (H_{E} - \omega ) - \lambda H_{E} = 0, \\ & \dot{\lambda } = \rho \lambda - W_{C} H_{S} - \lambda (G^{\prime} - H_{S} ). \\ \end{aligned}$$
(14)

The first equation implies that \(\left( {W_{C} - \lambda } \right) = \omega /H_{E} > 0\). At a steady state (\(\dot{\lambda } = 0\)), the second equation becomes \(\lambda (\rho - G^{\prime}) - (W_{C} - \lambda )H_{S} = 0\). Therefore, we have \(\rho - G^{\prime} > 0\) and \(\lambda > 0\). The latter and the first equation in 14 imply \(H_{E} - \omega > 0\). We use these inequalities below.

Let \((\tilde{E},\tilde{S})\) be the effort and the resource stock level at a steady state. As in the main text, \((\tilde{E},\tilde{S})\) satisfies

$$(H_{E} - \omega )(\rho - G^{\prime} + H_{S} ) - H_{E} H_{S} = 0{\text{ and }}G - H = 0.$$
(15)

Totally differentiate the second equation in 15 to have \(\left( {G^{\prime} - H_{S} } \right)d\tilde{S} - H_{E} d\tilde{E} = 0\). Thus,

$$\frac{{\partial \tilde{E}}}{{\partial \tilde{S}}} = \frac{{G^{\prime} - H_{S} }}{{H_{E} }}.$$

Using this, totally differentiate the first equation in 15

$$\begin{aligned} & - (\rho - G^{\prime} + H_{S} )d\omega + \{ (H_{ES} + H_{EE} \partial \tilde{E}/\partial \tilde{S})(\rho - G^{\prime} + H_{S} ) \\ & \, + (H_{E} - \omega )( - G^{\prime\prime} + H_{SS} + H_{SE} \partial \tilde{E}/\partial \tilde{S}) \\ & \, - \left( {H_{ES} + H_{EE} \partial \tilde{E}/\partial \tilde{S}} \right)H_{S} \\ & \, - H_{E} \left( {H_{SS} + H_{SE} \partial \tilde{E}/\partial \tilde{S}} \right)\} = 0. \\ \end{aligned}$$
(27)

Observe the following calculations:

$$\begin{gathered} H_{ES} + H_{EE} \frac{{\partial \tilde{E}}}{{\partial \tilde{S}}} = \frac{{\alpha H_{S} }}{{\tilde{E}}} + \frac{{\tilde{E}H_{EE} }}{{H_{E} }}\frac{{G^{\prime} - H_{S} }}{{\tilde{E}}} = \frac{{H_{S} - (1 - \alpha )G^{\prime}}}{{\tilde{E}}},{\text{ and}} \hfill \\ H_{SS} + H_{SE} \frac{{\partial \tilde{E}}}{{\partial \tilde{S}}} = \frac{{(\beta - 1)H_{S} }}{{\tilde{S}}} + \frac{{\beta H_{E} }}{{\tilde{S}}}\frac{{G^{\prime} - H_{S} }}{{H_{E} }} = \frac{{\beta G^{\prime} - H_{S} }}{{\tilde{S}}}. \hfill \\ \end{gathered}$$
(28)

Substitute 28 into 27 and we have:

$$\begin{gathered} (\rho - G^{\prime} + H_{S} )d\omega \hfill \\ = \{ \left( {H_{S} - (1 - \alpha )G^{\prime}} \right)(\rho - G^{\prime})/\tilde{E} - \omega (\beta G^{\prime} - H_{S} )/\tilde{S} - G^{\prime\prime}(H_{E} - \omega )\} d\tilde{S}. \hfill \\ \end{gathered}$$
(29)

As shown above, \(\rho - G^{\prime} > 0\) and \(H_{E} - \omega > 0\), and thus \(\rho - G^{\prime} + H_{S} > 0\). We have \(\beta G^{\prime} - H_{S} < 0\) because

$$\beta G^{\prime} - H_{S} = \frac{\beta }{S}\left[ {rS\left( {1 - \frac{2S}{k}} \right) - H} \right] = \frac{\beta }{S}\left[ { - rS\frac{S}{k} + G - H} \right] = \frac{ - rS\beta }{k} < 0.$$
(30)

Finally, by Assumption A5, \(H_{S} - (1 - \alpha )G^{\prime} = H_{S} - \beta G^{\prime} > 0\). Therefore, from 29, we have \(d\tilde{S}/d\omega > 0\).■

About this article

Cite this article

Farzin, Y.H., Akao, K.I. Poverty, social preference for employment, and natural resource depletion. Environ Econ Policy Stud 17, 1–26 (2015). https://doi.org/10.1007/s10018-013-0074-6

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s10018-013-0074-6

Keywords

JEL Classification

Navigation