Skip to main content

Advertisement

Log in

Production-based pollution versus deforestation: optimal policy with state-independent and-dependent environmental absorption efficiency restoration process

  • Original Research
  • Published:
Annals of Operations Research Aims and scope Submit manuscript

Abstract

An important yet largely unexamined issue is how the interaction between deforestation and pollution affects economic and environmental sustainability. This article seeks to bridge the gap by introducing a dynamic model of pollution accumulation where polluting emissions can be mitigated and the absorption efficiency of pollution sinks can be restored. We assume that emissions are due to a production activity, and we include deforestation both as an additional source of emissions and as a cause of the exhaustion of environmental absorption efficiency. To account for the fact that the switching of natural sinks to a pollution source can be either possible, and in such a case even reversible, or impossible, we consider that restoration efforts can be either independent from or dependent on environmental absorption efficiency, i.e., state-independent versus state-dependent restoration efforts. We determine (i) whether production or deforestation is the most detrimental from environmental and social welfare perspectives, and (ii) how state-dependent restoration process affects pollution accumulation and deforestation policies and the related environmental and social welfare consequences.

This is a preview of subscription content, log in via an institution to check access.

Access this article

Price excludes VAT (USA)
Tax calculation will be finalised during checkout.

Instant access to the full article PDF.

Fig. 1
Fig. 2
Fig. 3
Fig. 4
Fig. 5
Fig. 6
Fig. 7
Fig. 8
Fig. 9
Fig. 10
Fig. 11
Fig. 12
Fig. 13
Fig. 14
Fig. 15

Similar content being viewed by others

Notes

  1. A similar concept of sink efficiency was introduced by Gloor et al. (2010) and operationalized by Raupach et al. (2014) as the carbon uptake rate by land and ocean sinks.

References

  • Amesbury, M. J., Gallego-Sala, A., & Loisel, J. (2019). Peatlands as prolific carbon sinks. Nature Geoscience,12, 880–881.

    Google Scholar 

  • Andrés-Domenech, P., Martín-Herrán, G., & Zaccour, G. (2015). Cooperation for sustainable forest management: An empirical differential game approach. Ecological Economics,117, 118–128.

    Google Scholar 

  • Angelsen, A., & Rudel, T. K. (2013). Designing and implementing effective REDD + policies: A forest transition approach. Review of Environmental Economics and Policy,7(1), 91–113.

    Google Scholar 

  • Baccini, A., Goetz, S. J., Walker, W. S., Laporte, N. T., Sun, M., Sulla-Menashe, D., et al. (2012). Estimated carbon dioxide emissions from tropical deforestation improved by carbon-density maps. Nature Climate Change,2(3), 182–185.

    Google Scholar 

  • Baccini, A., Walker, W., Carvalho, L., Farina, M., Sulla-Menashe, D., & Houghton, R. A. (2017). Tropical forests are a net carbon source based on aboveground measurements of gain and loss. Science,358(6360), 230–234.

    Google Scholar 

  • Barbier, E. B., & Burgess, J. C. (1997). The economics of forest land use options. Land Economics,73(2), 174–195.

    Google Scholar 

  • Bastin, J. F., Finegold, Y., Garcia, C., Mollicone, D., Rezende, M., Routh, D., et al. (2019). The global tree restoration potential. Science,365(6448), 76–79.

    Google Scholar 

  • Boucekkine, R., Pommeret, A., & Prieur, F. (2013). Optimal regime switching and threshold effects. Journal of Economic Dynamics and Control,37(2), 2979–2997.

    Google Scholar 

  • Canadell, J. G., Le Quéré, C., Raupach, M. R., Field, C. B., Buitenhuis, E. T., Ciais, P., et al. (2007). Contributions to accelerating atmospheric CO2 growth from economic activity, carbon intensity, and efficiency of natural sinks. Proceedings of the National Academy of Sciences,104(47), 18866–18870.

    Google Scholar 

  • Canadell, J. G., & Raupach, M. R. (2008). Managing forests for climate change mitigation. Science,320(5882), 1456–1457.

    Google Scholar 

  • Cox, P. M., Betts, R. A., Jones, C., Spall, S. A., & Totterdell, I. (2000). Acceleration of global warming due to carbon-cycle feedbacks in a coupled climate model. Nature, 408, 184–187.

    Google Scholar 

  • Cramer, W., Bondeau, A., Woodward, F. I., Prentice, I. C., Betts, R. A., Brovkin, V., et al. (2001). Global response of terrestrial ecosystem structure and function to CO2 and climate change: results from six dynamic global vegetation models. Global Change Biology, 7(4), 357–373.

    Google Scholar 

  • Dockner, E., & Feichtinger, G. (1991). On the optimality of limit cycles in dynamic economic systems. Journal of Economics,53(1), 31–50.

    Google Scholar 

  • El Ouardighi, F., Benchekroun, H., & Grass, D. (2014). Controlling pollution and environmental absorption capacity. Annals of Operations Research,220(1), 1–23.

    Google Scholar 

  • El Ouardighi, F., Benchekroun, H., & Grass, D. (2016). Self-regenerating environmental absorption efficiency and the Soylent Green scenario. Annals of Operations Research,238(1), 179–198.

    Google Scholar 

  • El Ouardighi, F., Kogan, K., Gnecco, G., & Sanguineti, M. (2018a). Commitment-based equilibrium environmental strategies under time-dependent absorption efficiency. Group Decision and Negotiation,27(2), 235–249.

    Google Scholar 

  • El Ouardighi, F., Kogan, K., Gnecco, G., & Sanguineti, M. (2018b). Transboundary pollution control and environmental absorption efficiency management. Annals of Operations Research. https://doi.org/10.1007/s10479-018-2927-7,1-29.

    Article  Google Scholar 

  • Forster, B. (1973). Optimal consumption planning in a polluted environment. Economic Record,49(4), 534–545.

    Google Scholar 

  • Fredj, K., Martín-Herrán, G., & Zaccour, G. (2006). Incentive mechanisms to enforce sustainable forest exploitation. Environmental Modeling Assessment,11(2), 145–156.

    Google Scholar 

  • Gattuso, J.-P., Magnan, A. K., Bopp, L., Cheung, W. W. L., Duarte, C. M., Hinke, J., et al. (2018). Ocean solutions to address climate change and its effects on marine ecosystems. Frontiers in Marine Science,5(337), 1–18.

    Google Scholar 

  • Gloor, M., Sarmiento, J. L., & Gruber, N. (2010). What can be learned about carbon cycle climate feedbacks from the CO2 airborne fraction? Atmospheric Chemistry and Physics,10(16), 7739–7751.

    Google Scholar 

  • Gramling, C. (2017). Tropical forests have flipped from sponges to sources of carbon dioxide. Science News, September 28.

  • Grass, D., Caulkins, J. P., Feichtinger, G., Tragler, G., & Behrens, D. A. (2008). Optimal control of nonlinear processes with applications in drugs, corruption, and terror. Heidelberg: Springer.

    Google Scholar 

  • Harris, N. L., Brown, S., Hagen, S. C., Saatchi, S. S., Petrova, S., Salas, W., et al. (2012). Baseline map of carbon emissions from deforestation in tropical regions. Science,336(6088), 1573–1576.

    Google Scholar 

  • Hediger, W. (2009). Sustainable development with stock pollution. Environment and Development Economics,14(6), 759–780.

    Google Scholar 

  • Houghton, R. A., House, J. I., Pongratz, J., van der Werf, G. R., DeFries, R. S., Hansen, M. C., et al. (2012). Carbon emissions from land use and land-cover change. Biogeosciences,9, 5125–5142.

    Google Scholar 

  • Joos, F., Prentice, I. C., Sitch, S., Meyer, R., Hooss, G., Plattner, G.-K., et al. (2001). Global warming feedbacks on terrestrial carbon uptake under the Intergovernmental Panel on Climate Change (IPCC) emission scenarios. Global Biogeochemical Cycles, 15(4), 891–907.

    Google Scholar 

  • Keeler, E., Spence, M., & Zeckhauser, R. (1972). The optimal control of pollution. Journal of Economic Theory,4, 19–34.

    Google Scholar 

  • Leandri, M. (2009). The shadow price of assimilative capacity in optimal flow pollution control. Ecological Economics,68(4), 1020–1031.

    Google Scholar 

  • Leandri, M., & Tidball, M. (2019). Assessing the sustainability of optimal pollution paths in a world with inertia. Environmental Modeling and Assessment,24(2), 249–263.

    Google Scholar 

  • Lenton, T. M., Williamson, M. S., Edwards, N. R., Marsh, R., Price, A. R., Ridgwell, A. J., et al. (2006). Millennial timescale carbon cycle and climate change in an efficient Earth system model. Climate Dynamics, 26(7/8), 687–711.

    Google Scholar 

  • Liebsch, D., Marques, M. C., & Goldenberg, R. (2008). How long does the Atlantic Rain Forest take to recover after a disturbance? Changes in species composition and ecological features during secondary succession. Biological Conservation,141(6), 1717–1725.

    Google Scholar 

  • Maimon, O., Khmelnitsky, E., & Kogan, K. (1998). Optimal flow control in manufacturing systems: Production planning and scheduling. Dordrecht: Kluwer Academic Publishers.

    Google Scholar 

  • Makarieva, A. M., Gorshkov, V. G., Sheil, D., Nobre, A. D., Bunyard, P., & Li, B.-L. (2014). Why does air passage over forest yield more rain? Examining the coupling between rainfall, pressure, and atmospheric moisture content. Journal of Hydrometeorology,15(1), 411–426.

    Google Scholar 

  • Martin, P. A., Newton, A. C., & Bullock, J. M. (2013). Carbon pools recover more quickly than plant biodiversity in tropical secondary forests. Proceedings of the Royal Society B: Biological Sciences,280(1773), 22–36.

    Google Scholar 

  • Michel, P., & Rotillon, G. (1995). Disutility of pollution and endogenous growth. Environmental and Resource Economics,6(3), 279–300.

    Google Scholar 

  • Moser, E., Seidl, A., & Feichtinger, G. (2014). History-dependence in production-pollution-trade-off models: A multi-stage approach. Annals of Operation Research,222, 457–481.

    Google Scholar 

  • Piao, S. L., Ciais, P., Friedlingstein, P., Peylin, P., Reichstein, M., Luyssaert, S., et al. (2008). Net carbon dioxide losses of northern ecosystems in response to autumn warming. Nature, 451, 49–52.

    Google Scholar 

  • Prieur, F. (2009). The environmental Kuznets curve in a world of irreversibility. Economic Theory,40(1), 57–90.

    Google Scholar 

  • Raupach, M. R., Gloor, M., Sarmiento, J. L., Canadell, J. G., Frölicher, T. L., Gasser, T., et al. (2014). The declining uptake rate of atmospheric CO2 by land and ocean sinks. Biogeosciences,11(13), 3453–3475.

    Google Scholar 

  • Rodrigues, A. S. L., Robert, M. E. R. M., Parry, L., Souza, C., Jr., Veríssimo, A., & Balmford, A. (2009). Boom-and-bust development patterns across the Amazon deforestation frontier. Science,324(5933), 1435–1437.

    Google Scholar 

  • Schuur, E. A., McGuire, A. D., Schädel, C., Grosse, G., Harden, J. W., Hayes, D. J., et al. (2015). Climate change and the permafrost carbon feedback. Nature,520(7546), 171–179.

    Google Scholar 

  • Sheil, D., & Murdiyarso, D. (2009). How forests attract rain: An examination of a new hypothesis. BioScience,59(4), 341–347.

    Google Scholar 

  • Sohngen, B., & Mendelsohn, R. (2003). An optimal control model of forest carbon sequestration. American Journal of Agricultural Economics,85(2), 448–457.

    Google Scholar 

  • Southgate, D. (1990). The causes of land degradation along ‘spontaneous’ expanding agricultural frontiers. Land Economics,66(1), 93–101.

    Google Scholar 

  • Stähler, F. (1996). On international compensations for environmental stocks. Environmental Resource Economics,8(1), 1–13.

    Google Scholar 

  • Stern, N. (2006). Stern review report on the economics of climate change. London: HM Treasury.

    Google Scholar 

  • Stern, N. (2015). Why are we waiting? The logic, urgency, and promise of tackling climate change. Cambridge: MIT Press.

    Google Scholar 

  • Tahvonen, O., & Salo, S. (1996). Nonconvexities in optimal pollution accumulation. Journal of Environmental Economics and Management,31(2), 160–177.

    Google Scholar 

  • Tahvonen, O., & Withagen, C. (1996). Optimality of irreversible pollution accumulation. Journal of Economic Dynamics and Control,20(9), 1775–1795.

    Google Scholar 

  • Van Soest, D. (1998). Tropical deforestation: An economic perspective. The Netherlands: Labyrinth Publications.

    Google Scholar 

  • Van Soest, D., & Lensink, R. (2000). Foreign transfers and tropical deforestation: What terms of conditionality? American Journal of Agricultural Economics,82(2), 389–399.

    Google Scholar 

  • Wirl, F. (2007). Do multiple Nash equilibria in Markov strategies mitigate the tragedy of the commons? Journal of Economic Dynamics and Control,31(10), 3723–3740.

    Google Scholar 

Download references

Acknowledgements

The authors acknowledge helpful comments from one anonymous referee. This research was supported by ESSEC Business School (France) and Tel Aviv University (Israel). The first author dedicates this paper to the memory of Mohamed El Houari, a wonderful mentor and friend.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Fouad El Ouardighi.

Additional information

Publisher's Note

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

Appendix

Appendix

A1. The Legendre-Clebsch condition holds for (4) since the Hessian is negative definite, that is:

$$ \left[ {\begin{array}{*{20}c} {H_{uu} } \\ {H_{vu} } \\ {H_{wu} } \\ \end{array} \begin{array}{*{20}c} {H_{uv} } \\ {H_{vv} } \\ {H_{wv} } \\ \end{array} \begin{array}{*{20}c} {H_{uw} } \\ {H_{vw} } \\ {H_{ww} } \\ \end{array} } \right] = \left[ {\begin{array}{*{20}c} { - 1} \\ 0 \\ 0 \\ \end{array} \begin{array}{*{20}c} 0 \\ { - 1} \\ 0 \\ \end{array} \begin{array}{*{20}c} 0 \\ 0 \\ { - 1} \\ \end{array} } \right] $$

The Hamiltonian is therefore concave with respect to \( \left( {u,v,w} \right) \), which guarantees a (local) maximum.

A2. The candidate arc segments of the optimal path are described in the following way.

1.1 Production-based emissions and deforestation arc

In this case, Eq. (7) yields \( \left( {u^{*} , v^{*} , w^{*} } \right) = \left( {u^\circ \left( \lambda \right), v^\circ \left( {\lambda , \varphi } \right), w^\circ \left( {A^{*} , \varphi } \right)} \right) \) and \( \rho_{1} = 0 \), \( \rho_{2} = 0 \), \( \rho_{3} = 0 \), with:

$$ \left( {u^\circ v^\circ w^\circ } \right)^{t} \text{ := }\left( {a + \lambda b + \lambda \alpha - \varphi \varphi A^{\beta } } \right)^{t} $$
(A2.1)

being the solution of \( H_{u} \left( {P, A, u^\circ , v^\circ , w^\circ , \lambda , \varphi } \right) = 0 \), \( H_{v} \left( {P, A, u^\circ , v^\circ , w^\circ , \lambda , \varphi } \right) = 0 \) and \( H_{w} \left( {P, A, u^\circ , v^\circ , w^\circ , \lambda , \varphi } \right) = 0 \), respectively.

Plugging the expressions of \( u^{*} , v^{*} \) and \( w^{*} \) in (1)–(2) and (9)–(10), respectively, gives:

$$ \dot{P} = a + \left( {1 + \alpha^{2} } \right)\lambda + \alpha \left( {b - \varphi } \right) - AP $$
(A2.2)
$$ \dot{A} = \left( {1 + A^{2\beta } } \right)\varphi - b - \lambda \alpha - \gamma P $$
(A2.3)
$$ \dot{\lambda } = \left( {r + A} \right)\lambda + \gamma \varphi + cP $$
(A2.4)
$$ \dot{\varphi } = \left( {r - \beta \varphi A^{2\beta - 1} } \right)\varphi + \lambda P $$
(A2.5)

To analyze the behavior of the canonical system (A2.2)–(A2.5) in the neighborhood of the steady state, if it exists, we compute the Jacobian matrix:

$$ J = \left[ {\begin{array}{*{20}c} { - A} \\ { - \gamma } \\ c \\ \lambda \\ \end{array} \begin{array}{*{20}c} { - P} \\ {2\beta \varphi A^{2\beta - 1} } \\ \lambda \\ { - \beta \left( {2\beta - 1} \right)\varphi^{2} A^{2\beta - 2} } \\ \end{array} \begin{array}{*{20}c} {1 + \alpha^{2} } \\ { - \alpha } \\ {r + A} \\ P \\ \end{array} \begin{array}{*{20}c} { - \alpha } \\ {1 + A^{2\beta } } \\ \gamma \\ {r - \beta \varphi A^{2\beta - 1} } \\ \end{array} } \right] $$
(A2.6)

where \( \left( {P, A,\lambda ,\varphi } \right) \) are evaluated at their steady-state values, and whose determinant is given by:

$$ \left| J \right| = \beta \varphi^{2} A^{2\beta - 2} \left\{ {\left( {1 - 2\beta } \right)\left[ {\left( {\alpha \gamma + A} \right)\left( {r + \alpha \gamma + A} \right) + \gamma^{2} + c} \right] + \left( {1 + 2\beta } \right)A^{2\beta } \left[ {A\left( {r + A} \right) + c\left( {1 + \alpha^{2} } \right)} \right]} \right\} - 2\beta \varphi A^{2\beta - 1} \left\{ {r\left[ {A\left( {r + A} \right) + c\left( {1 + \alpha^{2} } \right)} \right] - \left( {r + 2A} \right)\left( {\alpha \lambda + \gamma P} \right) - 2\left[ {\gamma \lambda \left( {1 + \alpha^{2} } \right) - \alpha cP} \right]} \right\} - \alpha^{2} \lambda^{2} A^{2\beta } + \left( {1 + A^{2\beta } } \right)\left[ {cP^{2} - \lambda \left( {\lambda + rP + 2AP} \right)} \right] - r\lambda \left[ {\gamma \left( {1 + \alpha^{2} } \right) + \alpha A} \right] + P\left\{ {\alpha \left( {rc - 2\gamma \lambda } \right) + \gamma \left[ {\gamma P - r\left( {2r + A} \right)} \right]} \right\} $$

and the sum of the principal minors of \( J \) of order 2 minus the squared discounting rate is:

$$ K = - A\left( {r + A} \right) - c\left( {1 + \alpha^{2} } \right) + 2\left( {\alpha \lambda - \gamma P} \right) + \beta \varphi A^{2\beta - 1} \left\{ {2r + \varphi A^{ - 1} \left[ {2\beta - 1 - \left( {2\beta + 1} \right)A^{2\beta } } \right]} \right\} $$

In the case of state-independent restoration efforts, \( K_{{\left| {\beta = 0} \right.}} < 0 \), which rules out the possibility of limit cycles (Dockner and Feichtinger 1991). For linearly state-dependent restoration efforts, the sign of \( K_{{\left| {\beta = 1} \right.}} = - A\left( {r + A} \right) - c\left( {1 + \alpha^{2} } \right) + 2\left( {\alpha \lambda - \gamma P} \right) + \varphi \left[ {2rA + \varphi \left( {1 - A^{2} } \right)} \right] \) is not clear.

1.2 Production-based emissions-only arc

In this case, the control constraint \( v\left( t \right) \ge 0 \) is active. Using (5), the maximizing condition (6) yields \( \left( {u^{*} , v^{*} , w^{*} } \right) = \left( {u^\circ \left( \lambda \right), 0, w^\circ \left( {A^{*} , \varphi } \right)} \right) \) and \( \rho_{1} = 0 \), \( \rho_{2} \le 0 \), \( \rho_{3} = 0 \), with:

$$ \left( {u^\circ w^\circ } \right)^{t} \text{ := }\left( {a + \lambda \varphi A^{\beta } } \right)^{t} \;{\text{and}}\;\rho_{2} = \varphi - b - \lambda \alpha $$
(A2.7)

being the solution of \( L_{u} \left( {P, A, u^\circ , 0, w^\circ , \lambda , \varphi , 0, \rho_{2} , 0} \right) = 0 \), \( L_{v} \left( {P, A, u^\circ , 0, w^\circ , \lambda , \varphi , 0, \rho_{2} , 0} \right) = 0 \) and \( L_{w} \left( {P, A, u^\circ , 0, w^\circ , \lambda , \varphi , 0, \rho_{2} , 0} \right) = 0 \), respectively. Plugging the expressions of \( u^{*} \) and \( w^{*} \) in (1)–(2) and (9)–(10) respectively gives the canonical system:

$$ \dot{P} = a + \lambda - AP $$
(A2.8)
$$ \dot{A} = \varphi A^{2\beta } - \gamma P $$
(A2.9)
$$ \dot{\lambda } = \left( {r + A} \right)\lambda + \gamma \varphi + cP $$
(A2.10)
$$ \dot{\varphi } = \left( {r - \beta \varphi A^{2\beta - 1} } \right)\varphi + \lambda P $$
(A2.11)

The steady state, if it exists, is obtained by solving (11) where:

$$ \left( {\varphi_{\infty } \lambda_{\infty } A_{\infty } } \right)^{t} = \left( {\frac{{\gamma a\left( {r + A_{\infty } } \right)}}{{{\varPhi }}} - \frac{{a\left( {\gamma^{2} + cA_{\infty }^{2\beta } } \right)}}{{{\varPhi }}} \frac{{a\left( {r + A_{\infty } } \right)A_{\infty }^{2\beta } }}{{{\varPhi }}}} \right)^{t} $$

with \( {{\varPhi }} = A_{\infty }^{2\beta } \left[ {c + A_{\infty } \left( {r + A_{\infty } } \right)} \right] + \gamma^{2} \). It can be shown that the resolution of the system (A2.8)–(A2.11) for the case of state-independent restoration efforts (\( \beta = 0 \)), leads to a steady state that is a saddle-point with either monotonic or spiraling convergence. On the other hand, if \( \beta = 1 \), we obtain \( K_{{\left| {\beta = 1} \right.}} = - A\left[ {r + A\left( {1 + 3\varphi^{2} } \right)} \right] - 2\gamma P - c + 2r\varphi A \). Therefore, the possibility of limit cycles cannot be ruled out for linearly state-dependent restoration efforts.

1.3 Deforestation-only arc

In this case, the control constraint \( u\left( t \right) \ge 0 \) is active. Using (5), the maximizing condition (6) yields \( \left( {u^{*} , v^{*} , w^{*} } \right) = \left( {0, v^\circ \left( {\lambda ,\varphi } \right), w^\circ \left( {A^{*} , \varphi } \right)} \right) \) and \( \rho_{1} \le 0 \), \( \rho_{2} = 0 \), \( \rho_{3} = 0 \), with:

$$ \left( {v^\circ w^\circ } \right)^{t} \text{ := }\left( {b + \lambda \alpha - \varphi \varphi A^{\beta } } \right)^{t} \;{\text{and}}\;\rho_{1} = - \lambda - a $$
(A2.12)

being the solution of \( L_{u} \left( {P, A, 0,v^\circ , w^\circ , \lambda , \varphi , \rho_{1} , 0, 0} \right) = 0 \), \( L_{v} \left( {P, A, 0,v^\circ , w^\circ , \lambda , \varphi , \rho_{1} , 0, 0} \right) = 0 \) and \( L_{w} \left( {P, A, 0,v^\circ , w^\circ , \lambda , \varphi , \rho_{1} , 0, 0} \right) = 0 \), respectively. Plugging the expressions of \( v^{*} \) and \( w^{*} \) in (1)–(2) and (9)–(10) respectively gives the canonical system:

$$ \dot{P} = \alpha \left( {b + \lambda \alpha - \varphi } \right) - AP $$
(A2.13)
$$ \dot{A} = \varphi \left( {1 + A^{2\beta } } \right) - b - \lambda \alpha - \gamma P $$
(A2.14)
$$ \dot{\lambda } = \left( {r + A} \right)\lambda + \gamma \varphi + cP $$
(A2.15)
$$ \dot{\varphi } = \left( {r - \beta \varphi A^{2\beta - 1} } \right)\varphi + \lambda P $$
(A2.16)

The steady state, if it exists, is obtained by solving (11) where:

$$ \left( {\varphi_{\infty } \lambda_{\infty } A_{\infty } } \right)^{t} = \left( {\frac{{b\left( {r + A_{\infty } } \right)\left( {\alpha \gamma + A_{\infty } } \right)}}{{{\varPsi }}} - \frac{{b\left[ {\gamma \left( {\alpha \gamma + A_{\infty } } \right) + \alpha cA_{\infty }^{2\beta } } \right]}}{{{\varPsi }}} \frac{{\alpha b\left( {r + A_{\infty } } \right)A_{\infty }^{2\beta } }}{{{\varPsi }}}} \right)^{t} $$

with \( {{\varPsi }} = \left( {A_{\infty } + \alpha \gamma } \right)\left( {r + A_{\infty } + \alpha \gamma } \right) + A_{\infty }^{2\beta } \left[ {c\alpha^{2} + A_{\infty } \left( {r + A_{\infty } } \right)} \right] \). If \( \beta = 0 \), there can be no limit cycle. If \( \beta = 1 \), \( K_{{\left| {\beta = 1} \right.}} = - A\left[ {r + A\left( {1 + 3\varphi^{2} } \right)} \right] - \alpha^{2} c + 2\left( {\alpha \lambda - \gamma P} \right) + \varphi \left( {\varphi + 2rA} \right) \), which implies that limit cycles are possible for linearly state-dependent restoration efforts.

A3. Using (A2.1) gives:

$$ w^\circ = A^{\beta } \left( {\alpha u^\circ - v^\circ + b - \alpha a} \right) $$

A4. The steady state, if it exists, is obtained by solving (A2.8)–(A2.11) to zero. Equations (A2.8)–(A2.10) are linear in \( \varphi \), \( \lambda \) and \( P \), and after substituting their solutions into (A2.11), we get (11).

A5. The time decomposition method used for the numerical resolution follows the main steps below:

  • Step 1 Set a large enough time horizon \( t \in \left[ {0,T} \right] \). Set the initial states \( P\left( 0 \right) \) and \( A\left( 0 \right) \) at the given values. Set the terminal values of the costate variables at the steady-state values, \( \lambda \left( T \right) = \lambda_{\infty } \) and \( \varphi \left( T \right) = \varphi_{\infty } \).

  • Step 2 Guess feasible control functions, \( u\left( t \right) \), \( v\left( t \right) \) and \( w\left( t \right) \).

  • Step 3 Integrate the state system from left to right.

  • Step 4 Integrate the costate system from right to left.

  • Step 5 If the optimality conditions are satisfied (the Hamiltonian is maximized at each \( t \)) with a required tolerance, stop. Otherwise, go to step 6.

  • Step 6 At each \( t \), where the Hamiltonian is not maximized, change \( u\left( t \right) \), \( v\left( t \right) \) and \( w\left( t \right) \) to \( u\left( t \right) + \delta u\left( t \right) \), \( v\left( t \right) + \delta v\left( t \right) \) and \( w\left( t \right) + \delta w\left( t \right) \) where \( \delta u\left( t \right) \), \( \delta v\left( t \right) \) and \( \delta w\left( t \right) \) are small enough positive/negative increments to make the Hamiltonian rise.

  • Step 7 Go to Step 3.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

El Ouardighi, F., Khmelnitsky, E. & Leandri, M. Production-based pollution versus deforestation: optimal policy with state-independent and-dependent environmental absorption efficiency restoration process. Ann Oper Res 292, 1–26 (2020). https://doi.org/10.1007/s10479-020-03638-0

Download citation

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s10479-020-03638-0

Keywords

Navigation