Skip to main content
Log in

Optimal Rotation Age in Fast Growing Plantations: A Dynamical Optimization Problem

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

Abstract

Forest plantations are economically and environmentally relevant, as they play a key role in timber production and carbon capture. It is expected that the future climate change scenario affects forest growth and modify the rotation age for timber production. However, mathematical models on the effect of climate change on the rotation age for timber production remain still limited. We aim to determine the optimal rotation age that maximizes the net economic benefit of timber volume in a negative scenario from the climatic point of view. For this purpose, a bioeconomic optimal control problem was formulated from a system of Ordinary Differential Equations (ODEs) governed by the state variables live biomass volume, intrinsic growth rate, and area affected by fire. Then, four control variables were associated to the system, representing forest management activities, which are felling, thinning, reforestation, and fire prevention. The existence of optimal control solutions was demonstrated, and the solutions of the optimal control problem were also characterized using Pontryagin’s Maximum Principle. The solutions of the model were approximated numerically by the Forward–Backward Sweep method. To validate the model, two scenarios were considered: a realistic scenario that represents current forestry activities for the exotic species Pinus radiata D. Don, and a pessimistic scenario, which considers environmental conditions conducive to a higher occurrence of forest fires. The optimal solution that maximizes the net benefit of timber volume consists of a strategy that considers all four control variables simultaneously. For felling and thinning, regardless of the scenario considered, the optimal strategy is to spend on both activities depending on the amount of biomass in the field. Similarly, for reforestation, the optimal strategy is to spend as the forest is harvested. In the case of fire prevention, in the realistic scenario, the optimal strategy consists of reducing the expenses in fire prevention because the incidence of fires is lower, whereas in the pessimistic scenario, the opposite is true. It is concluded that the optimal rotation age that maximizes the net economic benefit of timber volume in P. radiata plantations is 24 and 19 years for the realistic and pessimistic scenarios, respectively. This corroborates that the presence of fires influences the determination of the optimal rotation age, and as a consequence, the net economic benefit.

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

Similar content being viewed by others

References

  • Alon U (2007) Chapman & Hall/CRC mathematical and computational biology series. Chapman & Hall/CRC, Boca Raton, FL

    Google Scholar 

  • Altamirano-Fernández A, Rojas-Palma A, Espinoza-Meza S (2022) A mathematical model to study the dynamics of carbon capture in forest plantations. J Phys Conf Series 2159:012001

    Google Scholar 

  • Altamirano-Fernández A, Rojas-Palma A, Espinoza-Meza S (2023) Optimal management strategies to maximize carbon capture in forest plantations: a case study with Pinus radiata D. Don. Forests 14(1):82

    Google Scholar 

  • Brazee RJ (2001) The faustmann formula: fundamental to forest economics 150 years after publication. Forest Sci 47(4):441

    Google Scholar 

  • Brown TJ, Hall BL, Westerling AL (2004) The impact of twenty-first century climate change on wildland fire danger in the western United States: an applications perspective. Clim Change 62(1):365–388

    Google Scholar 

  • Bulirsch R, Nerz E, Pesch HJ, Stryk O (1994) Combining direct and indirect methods in optimal control: range maximization of a hang glider. Springer, Birkhäuser Basel

    Google Scholar 

  • Caetano MAL, Gherardi DFM, Paula Ribeiro G, Yoneyama T (2009) Reduction of CO\(_2\) emission by optimally tracking a pre-defined target. Ecol Model 220(19):2536–2542

    Google Scholar 

  • Campanharo WA, Lopes AP, Anderson LO, Silva TF, Aragão LE (2019) Translating fire impacts in southwestern amazonia into economic costs. Remote Sens 11(7):764

    Google Scholar 

  • Chang W-Y, Wang S, Gaston C, Cool J, An H, Thomas BR (2019) Economic evaluations of tree improvement for planted forests: a systematic review. BioProducts Bus 4(1):1–14

    Google Scholar 

  • Citaristi I (2022) Organization of American States-OAS:(Organización de los Estados Americanos-OEA). In: Publications E (ed) The Europa Directory of International Organizations, vol 2022. Routledge, London, p 942

    Google Scholar 

  • Clark JS (1988) Effect of climate change on fire regimes in northwestern Minnesota. Nature 334(6179):233–235

    Google Scholar 

  • Couture S, Reynaud A (2011) Forest management under fire risk when forest carbon sequestration has value. Ecol Econ 70(11):2002–2011

    Google Scholar 

  • Cubbage F, Mac Donagh P, Sawinski Júnior J, Rubilar R, Donoso P, Ferreira A, Hoeflich V, Olmos VM, Ferreira G, Balmelli G et al (2007) Timber investment returns for selected plantations and native forests in South America and the Southern United States. New Forest 33(3):237–255

    Google Scholar 

  • Dubey B, Patra A (2013) A mathematical model for optimal management and utilization of a renewable resource by population. J Math. https://doi.org/10.1155/2013/613706

    Article  Google Scholar 

  • Englin J, Boxall P, Hauer G (2000) An empirical examination of optimal rotations in a multiple-use forest in the presence of fire risk. J Agric Resour Econ 15(1):14–27

    Google Scholar 

  • Eshetu G, Johansson T, Garedew W, Yisahak T (2021) Determinants of smallholder farmers’ adaptation options to climate change in a coffee-based farming system of Southwest Ethiopia. Climate Dev 13(4):318–325

    Google Scholar 

  • Faustmann M (1849) Berechnung des Werthes, welchen Waldboden, sowie noch nicht haubare Holzbestande fur die Waldwirthschaft besitzen [Calculation of the value which forest land and immature stands possess for forestry]. Allgemeine Fotst-und Jagd-Zeitung 25:441–455

    Google Scholar 

  • Flannigan M, Wagner CV (1991) Climate change and wildfire in Canada. Can J For Res 21(1):66–72

    Google Scholar 

  • Flannigan MD, Krawchuk MA, Groot WJ, Wotton BM, Gowman LM (2009) Implications of changing climate for global wildland fire. Int J Wildland Fire 18(5):483–507

    Google Scholar 

  • Fleming W, Rishel R, Marchuk G, Balakrishnan A, Borovkov A, Makarov V, Rubinov A, Liptser R, Shiryayev A, Krassovsky N et al (1975) Applications of mathematics. Deterministic and stochastic optimal control. Springer, Berlin

    Google Scholar 

  • Freedman HI, So J-H (1985) Global stability and persistence of simple food chains. Math Biosci 76(1):69–86

    MathSciNet  Google Scholar 

  • Gaoue OG, Ngonghala CN, Jiang J, Lelu M (2016) Towards a mechanistic understanding of the synergistic effects of harvesting timber and non-timber forest products. Methods Ecol Evol 7(4):398–406

    Google Scholar 

  • Gaoue OG, Jiang J, Ding W, Agusto FB, Lenhart S (2016) Optimal harvesting strategies for timber and non-timber forest products in tropical ecosystems. Thyroid Res 9:287–297

    Google Scholar 

  • González JR, Pukkala T, Palahí M (2005) Optimising the management of Pinus sylvestris L. stand under risk of fire in Catalonia (north-east of Spain). Ann For Sci 62(6):493–501

    Google Scholar 

  • González-González JM, Vázquez-Méndez ME, Diéguez-Aranda U (2022) Multi-objective models for the forest harvest scheduling problem in a continuous-time framework. Forest Policy Econ 136:102687

    Google Scholar 

  • Goodrick SL (2002) Modification of the fosberg fire weather index to include drought. Int J Wildland Fire 11(4):205–211

    Google Scholar 

  • Hackbusch W (1978) A numerical method for solving parabolic equations with opposite orientations. Computing 20(3):229–240

    MathSciNet  Google Scholar 

  • Hale JK (2009) Ordinary differential equations. Courier Corporation, Malabar, FL

    Google Scholar 

  • Hassan KK (2014) Nonlinear systems. Pearson New International Edition, New Jersey

    Google Scholar 

  • Heidari H, Arabi M, Warziniack T (2021) Effects of climate change on natural-caused fire activity in western US national forests. Atmosphere 12(8):981

    Google Scholar 

  • Jacobsen JB, Jensen F, Thorsen BJ (2018) Forest value and optimal rotations in continuous cover forestry. Environ Resource Econ 69(4):713–732

    Google Scholar 

  • Lampin-Maillet C, Jappiot M, Long M, Morge D, Ferrier J-P (2009) Characterization and mapping of dwelling types for forest fire prevention. Comput Environ Urban Syst 33(3):224–232

    Google Scholar 

  • Le HD, Smith C, Herbohn J (2014) What drives the success of reforestation projects in tropical developing countries? the case of the Philippines. Glob Environ Chang 24:334–348

    Google Scholar 

  • Lenhart S, Workman JT (2007) Mathematical and computational biology series. CRC Press, London

    Google Scholar 

  • Littell JS, Peterson DL, Riley KL, Liu Y, Luce CH (2016) A review of the relationships between drought and forest fire in the United States. Glob Change Biol 22(7):2353–2369

    Google Scholar 

  • Lizundia-Loiola J, Otón G, Ramo R, Chuvieco E (2020) A spatio-temporal active-fire clustering approach for global burned area mapping at 250 m from MODIS data. Remote Sens Environ 236:111493

    Google Scholar 

  • Log T (2019) Modeling indoor relative humidity and wood moisture content as a proxy for wooden home fire risk. Sensors 19(22):5050

    Google Scholar 

  • Lukes DL (1982) Differential equations: classical to controlled. Academic Press, New York

    Google Scholar 

  • Macdonald E, Hubert J (2002) A review of the effects of silviculture on timber quality of sitka spruce. Forestry 75(2):107–138

    Google Scholar 

  • Martínez-Ramos M, Balvanera P, Villa FA, Mora F, Maass JM, Méndez SM-V (2018) Effects of long-term inter-annual rainfall variation on the dynamics of regenerative communities during the old-field succession of a neotropical dry forest. For Ecol Manage 426:91–100

    Google Scholar 

  • Mataix-Solera J, Guerrero C (2007) Efectos de los incendios forestales en las propiedades edáficas. Caja Mediterráneo, CEMACAM Font Roja-Alcoi. Alicante, 5–40

  • Mattioli W, Ferrara C, Lombardo E, Barbati A, Salvati L, Tomao A (2022) Estimating wildfire suppression costs: a systematic review. Int For Rev 24(1):15–29

    Google Scholar 

  • Mauro F, Hardison PD (2000) Traditional knowledge of indigenous and local communities: international debate and policy initiatives. Ecol Appl 10(5):1263–1269

    Google Scholar 

  • McDowell N, Allen CD, Anderson-Teixeira K, Brando P, Brienen R, Chambers J, Christoffersen B, Davies S, Doughty C, Duque A et al (2018) Drivers and mechanisms of tree mortality in moist tropical forests. New Phytol 219(3):851–869

    Google Scholar 

  • Nilsson U, Berglund M, Bergquist J, Holmström H, Wallgren M (2016) Simulated effects of browsing on the production and economic values of Scots pine (Pinus sylvestris) stands. Scand J For Res 31(3):279–285

    Google Scholar 

  • Noss RF (2001) Beyond kyoto: forest management in a time of rapid climate change. Conserv Biol 15(3):578–590

    Google Scholar 

  • Ochuodho TO, Lantz VA (2014) Economic impacts of climate change in the forest sector: a comparison of single-region and multiregional CGE modeling frameworks. Can J For Res 44(5):449–464

    Google Scholar 

  • Patto JV, Rosa R (2022) Adapting to frequent fires: optimal forest management revisited. J Environ Econ Manag 111:102570

    Google Scholar 

  • Peng C (2000) Growth and yield models for uneven-aged stands: past, present and future. For Ecol Manage 132(2–3):259–279

    Google Scholar 

  • Peñuelas J, Sardans J (2021) Global change and forest disturbances in the mediterranean basin: breakthroughs, knowledge gaps, and recommendations. Forests 12(5):603

    Google Scholar 

  • Pontryagin LS (1987) Mathematical Theory of Optimal Processes. CRC Press, London, New York

    Google Scholar 

  • Routledge R (1980) The effect of potential catastrophic mortality and other unpredictable events on optimal forest rotation policy. Forest Sci 26(3):389–399

    Google Scholar 

  • Shukla J, Lata K, Misra A (2011) Modeling the depletion of a renewable resource by population and industrialization: Effect of technology on its conservation. Nat Resour Model 24(2):242–267

    MathSciNet  Google Scholar 

  • Sohngen B, Sedjo R (2005) Impacts of climate change on forest product markets: implications for North American producers. For Chron 81(5):669–674

    Google Scholar 

  • Starbuck CM, Berrens RP, McKee M (2006) Simulating changes in forest recreation demand and associated economic impacts due to fire and fuels management activities. Forest Policy Econ 8(1):52–66

    Google Scholar 

  • Susaeta A (2018) On Pressler’s indicator rate formula under the generalized Reed model. J For Econ 30:32–37

    Google Scholar 

  • Tessler N, Wittenberg L, Greenbaum N (2016) Vegetation cover and species richness after recurrent forest fires in the Eastern Mediterranean ecosystem of Mount Carmel, Israel. Sci Total Environ 572:1395–1402

    Google Scholar 

  • Úbeda X, Sarricolea P (2016) Wildfires in Chile: a review. Global Planet Change 146:152–161

    Google Scholar 

  • Urrutia-Jalabert R, González ME, González-Reyes Á, Lara A, Garreaud R (2018) Climate variability and forest fires in central and south-central Chile. Ecosphere 9(4):02171

    Google Scholar 

  • Vanclay JK (1994) Modelling forest growth and yield: applications to mixed tropical forests. CRC Press, Lismore

    Google Scholar 

  • Wang D, Guan D, Zhu S, Kinnon MM, Geng G, Zhang Q, Zheng H, Lei T, Shao S, Gong P et al (2021) Economic footprint of California wildfires in 2018. Nat Sustainabil 4(3):252–260

    Google Scholar 

  • Zamir M, Abdeljawad T, Nadeem F, Wahid A, Yousef A (2021) An optimal control analysis of a COVID-19 model. Alex Eng J 60(3):2875–2884

    Google Scholar 

Download references

Acknowledgements

A. Altamirano would like to thank the Vicerrectoría de Investigación y Postgrado at Universidad Católica del Maule, Chile. This work is part of A. Altamirano Ph.D. thesis in the program of Doctorado en Modelamiento Matemático Aplicado.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Alex Altamirano-Fernández.

Additional information

Publisher's Note

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

Appendices

Appendix A. Computational Implementation

To continue with the numerical solution of the optimal control problem. First, we rewrite system (1) as follows.

$$\begin{aligned} \begin{array}{rlr} D'(t)&=g(t, D(t), U(t)) \end{array} \end{aligned}$$
(19)

where \(D(t)=[B(t), r(t), I(t)]^\textrm{T}\) is the vector of solutions of the system for equations of state (1); and \(U(t)=[F(t), T(t), R(t), S(t)]^\texttt {T}\) represents the control variables, g is the right-hand side vector of the equations of state. In addition, we rewrite the objective functional of Eq. (13), as

$$\begin{aligned} \begin{array}{rlr} J(F,T,R,S)=\int _{0}^{T_f}f[t, D(t), U(t)]dt. \end{array} \end{aligned}$$

Where f is the integrand of the objective functional (13). Following Eq. (16), we denote the adjoint variables \({\lambda }=(\lambda _1, \lambda _2, \lambda _3)^\textrm{T}\) that were defined in Eq. (14), We now rewrite the Hamiltonian function, as follows:

$$\begin{aligned} H= f[t, D(t), U(t)] + {\lambda }[g(t, D(t), U(t))] \end{aligned}$$

then the optimality condition is expressed

$$\begin{aligned} \frac{\partial H}{\partial U}=0, \quad \text {at} \quad U^*\Rightarrow f_U + {\lambda } g_U=0, \end{aligned}$$

the attached variables and cross-cutting conditions are also expressed, as follows

$$\begin{aligned} {\lambda }'=-\frac{\partial H}{\partial D}=-(f_D + {\lambda } g_U), \quad {\lambda }(T_f)=0. \end{aligned}$$

The Forward–Backward Sweep Method (FBSM) is a numerical technique for solving optimal control problems. FBSM is one of the indirect methods valid for several areas that is designed to numerically solve ODEs generated by Pontryagin’s Maximum Principle (Lenhart and Workman 2007). It is known that indirect methods present difficulties for complex and difficult to-analyze optimal control problems. For this, there are direct methods that allow obtaining an approximation of the optimal solution for complex problems (Bulirsch et al. 1994). However, it is interesting to know the convergence of the FBSM method. In Mattioli et al. (2022), they present a convergence result for an optimal control problem. The authors do not claim that the FBSM overcomes some difficulties of a more complex optimal control problem, but that the method is easy to apply to small problems and can provide a quick check of previous results of more efficient methods. For information on the convergence and stability of this method, see (Hackbusch 1978). The steps of the FBSM algorithm are then presented.

Step 1.:

Make an initial guess for U on the interval.

Step 2.:

Using the initial condition \(D1 = D(0) = [B(0), r(0), I(0)]^{\textrm{T}}=[B_0,r_0,I_0]^{\textrm{T}}\) and the values of U, solve D forward in time according to its differential equation in the optimality system (19).

Step 3.:

Using the transversality condition \({\lambda }_{N+1} = (\lambda _1(T_f),\lambda _2(T_f),\lambda _3(T_f))= (0,0,0)\), where \(N+1\) is the step number and the values of U and D, solve \({\lambda }\) backward in time according to its differential equation in the optimality system.

Step 4.:

Update U by introducing the new values of D and \({\lambda }\) in the optimal control characterization.

Step 5.:

Check convergence. If the values of the variables in this iteration and in the last iteration are very close, the current values are output as solutions. If the values are not close, go back to Step 2.

Appendix B. Proof of Condition \((A_4)\)

To show that the objective functional is concave in U, we define

$$\begin{aligned} \mathcal {Z}(t,X,U)=[ p_1FB+p_2TB-c_1F^2 - c_2T^2 - c_3R^2 - c_4S^2]. \end{aligned}$$

The definition of the convex set is as follows

$$\begin{aligned} (1-q)\mathcal {Z}(t,X,u) + q\mathcal {Z}(t,X,s)\le \mathcal {Z}(t,X,(1-q)u + qs); \end{aligned}$$
(20)

where us are two vectors and \(q\in (0,1)\). First, we will develop the right-hand side of Equation (20), such that

$$\begin{aligned} \begin{aligned} (1-q)\mathcal {Z}(t,X,u) + q\mathcal {Z}(t,X,s)&=p_1Bu_1+p_2Bu_2 - c_1u_1^2-c_2u_2^2-c_3u_3^2-c_4u_4^2\\&\quad - qp_1Bu_1- qp_2Bu_2 + qc_1u_1^2+qc_2u_2^2\\&\quad +qc_3u_3^2+qc_4u_4^2 +qp_1Bs_1+qp_2Bs_2\\&\quad - qc_1s_1^2-qc_2s_2^2-qc_3s_3^2-qc_4s_4^2, \end{aligned} \end{aligned}$$
(21)

now we will develop the left-hand side of Eq. (20), we obtain that

$$\begin{aligned} \begin{aligned} \mathcal {Z}(t,X,(1-q)u + qs)&=p_1Bu_1 -p_1qBu_1 + p_1qBs_1 + p_2Bu_2 - p_2qBu_2 +p_2qBs_2 \\&\quad -[c_1u_1^2-c_1qu_1^2 + c_1[(q^2-q)u_1^2 + 2q(1-q)u_1s_1 + q^2s_1^2]]\\&\quad -[c_2u_2^2-c_2qu_2^2 + c_2[(q^2-q)u_2^2 + 2q(1-q)u_2s_2 + q^2s_2^2]]\\&\quad -[c_3u_3^2-c_3qu_3^2 + c_3[(q^2-q)u_3^2 + 2q(1-q)u_3s_3 + q^2s_3^2]]\\&\quad -[c_4u_4^2-c_4qu_4^2 + c_4[(q^2-q)u_4^2 + 2q(1-q)u_4s_4 + q^2s_4^2]]\\ \end{aligned} \end{aligned}$$
(22)

replacing Eqs. (22) and (21) in Eq. (20), it is obtained that

$$\begin{aligned} \begin{aligned}&(1-q)\mathcal {Z}(t,X,u) + q\mathcal {Z}(t,X,s)-\mathcal {Z}(t,X,(1-q)u + qs)\\&\quad =- \sum _{i=1}^4c_i\left[ \sqrt{q(1-q)}u_i-\sqrt{q(1-q)}s_i\right] ^2< 0, \end{aligned} \end{aligned}$$
(23)

furthermore, as the objective functional (13) there is a discount factor \(e^{-\delta t}\) and it is always positive. Therefore, from Eq. (23) it is justified that the objective functional is is concave in U, which determines the demonstration of \((A_4)\).

Rights and permissions

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

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Altamirano-Fernández, A., Rojas-Palma, A. & Espinoza-Meza, S. Optimal Rotation Age in Fast Growing Plantations: A Dynamical Optimization Problem. Bull Math Biol 86, 51 (2024). https://doi.org/10.1007/s11538-024-01262-8

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1007/s11538-024-01262-8

Keywords

Navigation