Skip to main content

Bootstrap Simulation, Markov Decision Process Models, and Role of Discounting in the Valuation of Ecological Criteria in Uneven-Aged Forest Management

  • Chapter
  • First Online:
Continuous Cover Forestry

Part of the book series: Managing Forest Ecosystems ((MAFE,volume 23))

  • 1626 Accesses

Abstract

This chapter presents general methods combining stochastic simulation and Markov decision process models, with a specific application to the issue of discounting ecological criteria. The literature argues for and against discounting ecological costs and benefits like standard financial investments. As part of this debate, this chapter investigates some purely ecological consequences of discounting ecological criteria. The methods are Markov decision process models with infinite time horizon and discounted or undiscounted ecological objectives, based on bootstrap simulations of stand growth. The data were from Douglas-fir/western hemlock forests in the U.S. Pacific Northwest, with the assumption of continuous-cover forestry. The ecological criteria examined here included the stand basal area per hectare, the tree species and size diversity measured with Shannon’s index, and the percentage of the forest in late seral stage. In maximizing expected tree species diversity, 18 out of 64 possible stand states would call for different decisions with discounting. For tree size diversity, the decisions differed for five stand states. For basal area all but two states called for the same decision. For late seral forest frequency, discounting led to different decisions for 13 stand states. However, for only a few stand states were the ecological criteria substantially different immediately after harvest. Thus, discounting would matter only for a few stand states, which in the study area had a low frequency. Given this initial condition, only the expected late seral forest frequency differed substantially after a decade with discounting. In the long run, the initial condition does not matter and discounting the criteria gave very similar values as not discounting.

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

Access this chapter

Chapter
USD 29.95
Price excludes VAT (USA)
  • Available as PDF
  • Read on any device
  • Instant download
  • Own it forever
eBook
USD 189.00
Price excludes VAT (USA)
  • Available as EPUB and PDF
  • Read on any device
  • Instant download
  • Own it forever
Softcover Book
USD 279.99
Price excludes VAT (USA)
  • Compact, lightweight edition
  • Dispatched in 3 to 5 business days
  • Free shipping worldwide - see info
Hardcover Book
USD 279.99
Price excludes VAT (USA)
  • Durable hardcover edition
  • Dispatched in 3 to 5 business days
  • Free shipping worldwide - see info

Tax calculation will be finalised at checkout

Purchases are for personal use only

Institutional subscriptions

References

  • Ambuel B, Temple SA (1983) Area-dependent changes in the bird communities and vegetation of S. Wisconsin forests. Ecology 64:1057–1068

    Article  Google Scholar 

  • Austin MP, Meyers JA (1999) Current approaches to modelling the environmental niche of eucalypts: implication for management of forest biodiversity. For Ecol Manag 85:95–106

    Article  Google Scholar 

  • Barbour RJ, Kellogg RM (1990) Forest management and end-product quality: a Canadian perspective. Can J For Res 20:405–414

    Google Scholar 

  • Bentley WR, Roger FR (1966) A zero interest comparison of forest rent and soil rent. For Sci 12(4):460–460

    Google Scholar 

  • Bevers M, Hof J (1999) Spatially optimizing wildlife habitat edge effects in forest management linear and mixed-integer programs. For Sci 45(2):249–258

    Google Scholar 

  • Bierman HJ (1968) The growth period decision. Manag Sci 14(6):302–309

    Article  Google Scholar 

  • Boscolo M, Buongiorno J, Panayotou T (1997) Simulating options for carbon sequestration through improved management of a lowland tropical rainforest. Environ Dev Econ 2:241–263

    Article  Google Scholar 

  • Buongiorno J (2001) Generalization of Faustmann’s formula for stochastic forest growth and prices with Markov decision process models. For Sci 47(4):466–474

    Google Scholar 

  • Buongiorno J, Gilless JK (2003) Decision Methods for Forest Resource Management. Academic Press, San Diego

    Google Scholar 

  • Buongiorno J, Zhou M (2009) Further generalization of Faustmann’s formula for stochastic interest rate. Paper presented at the third international Faustmann symposium, 28–31 Oct 2009, Darmstadt

    Google Scholar 

  • Burns RM and Honkala BH (1990) Silvics of North America, Vol. 1, Conifers. Washington DC: U.S.D.A. Forest Service Agriculture Handbook 654

    Google Scholar 

  • Chang SJ (1981) Determination of the optimal growing stock and cutting cycle for an uneven-aged stand. For Sci 27(4):739–744

    Google Scholar 

  • Clark CW (1988) Clear-cut economies (should we harvest everything now?). The Sciences, pp 16–20

    Google Scholar 

  • D’Epenoux F (1963) A probabilitstic production and inventory problem. Manag Sci 10(1):98–108

    Article  Google Scholar 

  • Dewar RC, Cannell MR (1992) Carbon sequestration in the trees, products and soils of forest plantations: an analysis using UK examples. Tree Physiol 11:49–71

    PubMed  CAS  Google Scholar 

  • Faustmann M (1849) Calculation of the value which forest land and immature stands possess for forestry. Allgemaine forst- und jagdzeitung 15:441–55

    Google Scholar 

  • Feinberg F, Schwartz A (2001) Handbook of Markov decision processes: methods and applications. Springer, New York, 576 p

    Google Scholar 

  • Franklin JF (1988) Pacific northwest forests. In: Barbour MG, Billings WD (Eds), North American Terrestrial Vegetation. Cambridge University Press, New York, p 434

    Google Scholar 

  • Gobster PH (1999) An ecological aesthetic for forest landscape management. Landsc J 18(1): 54–64

    Google Scholar 

  • Gowdy JM (1996) Discounting, hierarchies, and the social aspects of biodiversity protection. Int J Soc Econ 23(4/4/6):49–63

    Article  Google Scholar 

  • Guldin JM (1996) The role of uneven-aged silviculture in the context of ecosystem management. West J Appl For 11(1):4–12

    Google Scholar 

  • Hansen AJ, Spies TA, Swanson FJ, Ohmann JL (1991) Conserving biodiversity in managed forests: lessons from natural forests. BioScience 41(6):382–292

    Article  Google Scholar 

  • Hartman R (1976) The harvesting decision when the standing tree has value. Econ Inq 14(1):52–58

    Article  Google Scholar 

  • Hillier FS, Lieberman GJ (2005) Introduction to operations research, 8th edn. McGraw-Hill, Boston, 1059 p

    Google Scholar 

  • Holling CS, Dantzig GB, Winkler C (1986) Determining optimal policies for ecosystems, pp 463–467. In: Kallio M et al (eds) Systems analysis in forestry and forest industries. TIMS studies in Management Science vol 21, North Holland, Amsterdam, 487 p

    Google Scholar 

  • Howarth RB (2009) Discounting, uncertainty, and revealed time preference. Land Econ 85(1): 24–40

    Google Scholar 

  • Hummel S, Calkin DE (2005) Costs of landscape silviculture for fire and habitat management. For Ecol Manag 207(3):385–404

    Article  Google Scholar 

  • Insley M (2002) A real options approach to the valuation of a forestry investment. J Environ Econ Manag 44(3):471–492

    Article  Google Scholar 

  • Insley M, Rollins K (2005) On Solving the Multirotational Timber Harvesting Problem with Stochastic Prices: A Linear Complementarity Formulation. Am J Agri Econ 87(3):735–755

    Google Scholar 

  • Jiang H, Strittholt J, Frost P, Slosser N (2004) The classification of late seral forests in the Pacific Northwest, USA using Landsat ETM + imagery. Remote Sens Environ 91:320–331

    Article  Google Scholar 

  • Kangas J, Kuusipalo J (1993) Integrating biodiversity into forest management planning and decision making. For Ecol Manag 61:1–15

    Article  Google Scholar 

  • Kao C (1984) Notes: optimal stocking levels and rotation under uncertainty. For Sci 30(4):921–927

    Google Scholar 

  • Kaya I, Buongiorno J (1987) Economic management of uneven-aged northern hardwood stands. North J Appl For 2:28–31

    Google Scholar 

  • Kaya I, Buongiorno J (1989) A harvesting guide for uneven-aged northern hardwood stands. North J Appl For 2:28–31

    Google Scholar 

  • Lembersky MR, Johnson KN (1975) Optimal policies for managed stands: an infinite horizon Markov decision process approach. For Sci 21(2):109–122

    Google Scholar 

  • Liang J, Buongiorno J, Monserud RA (2005) Growth and yield of all-aged Douglas-fir/western hemlock forest stands: a matrix model with stand diversity effects. Can J For Res 35:2369–2382

    Article  Google Scholar 

  • Liang J, Buongiorno J, Monserud RA (2006) Bootstrap simulation and response surface optimization of management regimes for Douglas-fir/western hemlock stands. For Sci 52(5):579–594

    Google Scholar 

  • Lin C-R, Buongiorno J (1998) Tree diversity, landscape diversity, and economics of maple-birch forests: implications of Markov models. Manag Sci 44(10):1351–1366

    Article  Google Scholar 

  • Miller M, Emmingham B (2001) Can Selection Thinning Convert Even-Age Douglas-Fir Stands to Uneven-Age Structures? Western J Appl For 16(1):35–43

    Google Scholar 

  • Newman D (1988) The optimal forest rotation: a discussion and annotated bibliography. General Technical Report SE-48. U.S. Department of Agriculture, Forest Service, Southeastern Forest Experiment Station, Asheville, 47 p

    Google Scholar 

  • Newman DH, Yin R (1995) A note on the tree-cutting problem in a stochastic environment. J For Econ 1:181–190

    Google Scholar 

  • Plantinga AJ (1998) Optimal harvesting strategies with stationary and non-stationary prices: an option value analysis. For Sci 44:192–202

    Google Scholar 

  • NWP Regional Ecosystem Office (2005) REO GIS data. Regional Ecosystem Office, Portland, Oregon. Available from http://www.reo.gov/gis/data/gisdata/index.htm. Last accessed July 2006

  • O'HARA KL (1998) Silviculture for structural diversity: A new look at multiaged systems. J Forestry 96:4–10

    Google Scholar 

  • Runkle DE (1987) Vector autoregression and reality. J Bus Econ Stat 5:437–442

    Article  Google Scholar 

  • Seely B, Nelson J, Wells R, Peter B, Meitner M, Anderson A, Harshaw H, Sheppard S, Bunnell FL, Kimmins H, Harrison D (2004) The application of a hierarchical, decision-support system to evaluate multi-objective forest management strategies: a case study in northeastern British Columbia, Canada. For Ecol Manag 199:284–305

    Google Scholar 

  • Shannon CE (1948) A mathematical theory of communication. Bell Syst Tech J 27:379–423, 623–656

    Google Scholar 

  • Winston WL (1991) Operations research: applications and algorithms, 2nd edn. PWS-Kent Publishing Co, Boston, 1262 pp

    Google Scholar 

  • Zhou M, Buongiorno J (2006) Forest landscape management in a stochastic environment, with an application to mixed loblolly pine-hardwood forests. For Ecol Manag 223:170–182

    Article  Google Scholar 

  • Zhou M, Liang J, Buongiorno J (2008a) Adaptive versus fixed policies for economic or ecological objectives in forest management. For Ecol Manag 254:178–187

    Article  Google Scholar 

  • Zhou M, Buongiorno J, Liang J (2008b) Economic and ecological effects of diameter caps: a Markov decision model for Douglas-fir/western hemlock forests. For Sci 54(4):397–407

    Google Scholar 

Download references

Acknowledgements

This work has been supported in part by financial support to Joseph Buongiorno by the USDA Forest Service Southern Forest Experiment Station.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Mo Zhou .

Editor information

Editors and Affiliations

Appendix

Appendix

The matrix G and the vector r in the deterministic part of the growth model (8.2) have the following form (Liang et~al. 2005):

$$ \begin{aligned} & {\mathbf{G}} = \left[ {\begin{array}{*{20}{c}} {{{\mathbf{G}}_1}} & {} & {} & {} \\ {} & {{{\mathbf{G}}_2}} & {} & {} \\ {} & {} & {{{\mathbf{G}}_3}} & {} \\ {} & {} & {} & {{{\mathbf{G}}_4}} \\ \end{array} } \right],\ {{\mathbf{G}}_i} = \left[\arraycolsep4pt {\begin{array}{*{20}{c}} {{a_{i1}}} & {} & {} & {} & {} \\ {{b_{i1}}} & {{a_{i2}}} & {} & {} & {} \\ {} & \ddots & \ddots & {} & {} \\ {} & {} & {{b_{i,17}}} & {{a_{i,18}}} & {} \\ {} & {} & {} & {{b_{i,18}}} & {{a_{i,19}}} \\ \end{array} } \right] \\ & {\mathbf{r}} = \left[ {\begin{array}{*{20}{c}} {{{\mathbf{r}}_1}} \\ {{{\mathbf{r}}_2}} \\ {{{\mathbf{r}}_3}} \\ {{{\mathbf{r}}_4}} \\ \end{array} } \right],\quad {{\mathbf{r}}_i} = \left[ {\begin{array}{*{20}{c}} {{r_i}} \\ {0} \\ \vdots \\ {0} \\ \end{array} } \right] \\ \end{aligned} $$

where a ij is the probability that a tree of species i and diameter class j stays alive and in the same diameter class between t and t + 1. i = 1 for Douglas fir, 2 for other shade tolerant species, 3 for western hemlock, and 4 for other shade-tolerant species. b ij is the probability that a tree of species i and diameter class j stays alive and grows into diameter class j + 1, and r i is the number of trees of species group i recruited in the smallest diameter class between t and t + 1, with a time period of 1 year. Recruitment is zero in the higher diameter classes. The b ij probability is equal to the annual tree diameter growth, g ij , divided by the width of the diameter class. Diameter growth is a function of tree diameter D j (cm), stand basal area B (m2/ha), site productivity Q (m3/ha/year), tree species diversity H s , and tree size diversity H d .

$${\fontsize{9}{11}\selectfont{ \begin{aligned} & {g_{1j}} = 0.7860 + 0.0124{D_j} - 0.0001D_j^2 - 0.0107{B} + 0.0267Q + 0.0658{H_s} - 0.2426{H_d}\quad \\ & {g_{2j}} = 0.6104 - 0.0038{D_j} + 0.0001D_j^2 - 0.0080{B} + 0.0170Q + 0.0707{H_s} - 0.0693{H_d}\quad \\ & {g_{3j}} = 0.9026 + 0.0148{D_j} - 0.0001D_j^2 - 0.0107{B} + 0.0061Q - 0.0250{H_s} - 0.1750{H_d}\quad \\ & {g_{4j}} = 0.5851 + 0.0081{D_j} - 0.00003D_j^2 - 0.008{B} + 0.0178Q + 0.1285{H_s} - 0.1441{H_d}\quad \\ \end{aligned} }}$$

The expected recruitment of species i is represented by a Tobit model:

$$ {r_i} = \Phi ({{\mathbf{\beta }}_i}{{\mathbf{x}}_i}/{\sigma_i}){{\mathbf{\beta }}_i}{{\mathbf{x}}_i} + {\sigma_i}\phi ({{\mathbf{\beta }}_i}{{\mathbf{x}}_i}/{\sigma_i}) $$

with:

$$ \begin{array}{llll} & {{\mathbf{\beta }}_1}{{\mathbf{x}}_1} = - 21.9317 - 1.2996B + 0.0971{N_1} + 0.8007Q + 11.119{H_s} - 6.8020{H_d} \\ & {{\mathbf{\beta }}_2}{{\mathbf{x}}_2} = - 23.6333 - 0.8293B + 0.0975{N_2} + 0.2032Q + 8.4122{H_s} - 5.9733{H_d} \\ & {{\mathbf{\beta }}_3}{{\mathbf{x}}_3} = - 30.8842 - 0.9359B + 0.0926{N_3} + 0.6699Q + 14.693{H_s} - 9.8919{H_d} \\ & {{\mathbf{\beta }}_4}{{\mathbf{x}}_4} = - 34.5350 - 0.7512B + 0.0924{N_4} + 0.7375Q + 8.0361{H_s} - 2.2701{H_d} \end{array} $$

where N i is the number of trees per hectare in species group i; Φ and ϕ are respectively the standard normal cumulative and density functions, and the standard deviations of the residuals are, σ 1 = 23.5417, σ 2 = 22.4354, σ 3 = 27.3244, σ 4 = 23.0297.

The probability of tree mortality per year, m ij , is a species-dependent probit function of tree size and stand state:

$${\fontsize{8}{10}\selectfont{ \begin{aligned} & {m_1} = \frac{1}{{10.5}}\Phi ( - 2.1103 - 0.0356{D_j} + 0.0002D_j^2 + 0.0081B - 0.0200C + 0.0059{H_s} + 0.5110{H_d}){ } \\ & {m_2} = \frac{1}{{10.5}}\Phi ( - 1.4063 - 0.0204{D_j} + 0.0002D_j^2 + 0.0053B - 0.0147C + 0.0022{H_s} + 0.1411{H_d}){ } \\ & {m_3} = \frac{1}{{10.5}}\Phi ( - 3.1746 - 0.0416{D_j} + 0.0003D_j^2 + 0.0156B - 0.0230C + 0.3252{H_s} + 0.4192{H_d}){ } \\ & {m_4} = \frac{1}{{10.5}}\Phi ( - 1.5188 - 0.0093{D_j} - 0.0000D_j^2 - 0.0042B - 0.0303C - 0.2721{H_s} + 0.3528{H_d}){ } \end{aligned}}} $$

The probability that a tree stays alive and in the same size class from t to t + 1 is, then:

$$ {a_{ij}} = {1} - {b_{ij}} - {m_{ij}} $$

The expected single tree volume v ij (m3) is a species-dependent function of tree and stand characteristics:

$$ {\fontsize{8.8}{10.8}\selectfont{\hspace*{-1pt}\begin{aligned} & {v_{1j}} = - 0.6116 - 0.0119{D_j} + 0.0012D_j^2 + 0.0132B + 0.0293Q - 0.1249{H_s} - 0.0330{H_d} \\ & {v_{2j}} = - 0.5487 + 0.0038{D_j} + 0.0009D_j^2 + 0.0044B + 0.0173Q + 0.0136{H_s} + 0.0358{H_d} \\ & {v_{3j}} = - 0.4621 - 0.0141{D_j} + 0.0013D_j^2 + 0.0072B{ + 0}{.0104}Q - 0.0659{H_s} + 0.0516{H_d} \\ & {v_{{4 }j}} = - 0.1401 - 0.0170{D_j} + 0.0011D_j^2{ + 0}{.0087}B{ + 0}{.0146}Q - {0}{.2471}{H_s} + 0.0125{H_d} \end{aligned}}} $$

Rights and permissions

Reprints and permissions

Copyright information

© 2012 Springer Science+Business Media B.V.

About this chapter

Cite this chapter

Zhou, M., Buongiorno, J., Liang, J. (2012). Bootstrap Simulation, Markov Decision Process Models, and Role of Discounting in the Valuation of Ecological Criteria in Uneven-Aged Forest Management. In: Pukkala, T., von Gadow, K. (eds) Continuous Cover Forestry. Managing Forest Ecosystems, vol 23. Springer, Dordrecht. https://doi.org/10.1007/978-94-007-2202-6_8

Download citation

Publish with us

Policies and ethics