Skip to main content
Log in

Integral control for population management

  • Published:
Journal of Mathematical Biology Aims and scope Submit manuscript

Abstract

We present a novel management methodology for restocking a declining population. The strategy uses integral control, a concept ubiquitous in control theory which has not been applied to population dynamics. Integral control is based on dynamic feedback—using measurements of the population to inform management strategies and is robust to model uncertainty, an important consideration for ecological models. We demonstrate from first principles why such an approach to population management is suitable via theory and examples.

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

Similar content being viewed by others

References

  • Astolfi A, Karagiannis D, Ortega R (2008) Nonlinear and adaptive control with applications. Communications and Control Engineering Series. Springer, London Ltd., London (ISBN 978-1-84800-065-0)

  • Åström KJ (1980) Why use adaptive techniques for steering large tankers? Int J Control 32:689–708

    Article  Google Scholar 

  • Åström KJ, Hägglund T (1995) PID controllers: theory, design, and tuning. Instrument Society of America, Research Triangle Park, North Carolina

  • Berman A, Plemmons RJ (1994) Nonnegative matrices in the mathematical sciences, volume 9 of Classics in applied mathematics. Society for Industrial and Applied Mathematics (SIAM), Philadelphia. ISBN 0-89871-321-8. doi:10.1137/1.9781611971262. URL http://dx.doi.org/10.1137/1.9781611971262

  • Bieber C, Ruf T (2005) Population dynamics in wild boar sus scrofa: ecology, elasticity of growth rate and implications for the management of pulsed resource consumers. J Appl Ecol 42(6):1203–1213

    Article  Google Scholar 

  • Blackwood J, Hastings A, Costello C (2010) Cost-effective management of invasive species using linear-quadratic control. Ecol Econ 69(3):519–527

    Article  Google Scholar 

  • Brenner SC, Scott LR (1994) The mathematical theory of finite element methods, volume 15 of Texts in applied mathematics. Springer, New York (ISBN 0-387-94193-2)

  • Briggs J, Dabbs K, Holm M, Lubben J, Rebarber R, Tenhumberg B, Riser-Espinoza D (2010) Structured population dynamics. Math Mag 83(4):243–257

    Article  MATH  MathSciNet  Google Scholar 

  • Cassandra AR (1998) Exact and approximate algorithms for partially observable Markov decision processes. PhD thesis, Brown University, Department of Computer Science, Providence. Available from: http://cs.brown.edu/research/ai/pomdp/papers/

  • Caswell H (2001) Matrix population models : construction, analysis, and interpretation. Sinauer Associates, Sunderland

  • Chadès I, McDonald-Madden E, McCarthy MA, Wintle B, Linkie M, Possingham HP (2008) When to stop managing or surveying cryptic threatened species. Proc Natl Acad Sci 105(37):13936–13940

    Article  Google Scholar 

  • Chadès I, Martin TG, Nicol S, Burgman MA, Possingham HP, Buckley YM (2011) General rules for managing and surveying networks of pests, diseases, and endangered species. Proc Natl Acad Sci 108(20):8323–8328

    Article  Google Scholar 

  • Childs D, Rees M, Rose K, Grubb P, Ellner S (2003) Evolution of complex flowering strategies: an age-and size-structured integral projection model. Proc R Soc Lond Ser B Biol Sci 270(1526):1829–1838

    Article  Google Scholar 

  • Coughlan J (2007) Absolute stability results for infinite-dimensional discrete-time systems with applications to sampled-data integral control. PhD thesis, University of Bath. Available from: http://people.bath.ac.uk/mashl/research.html

  • Coughlan JJ, Logemann H (2009) Absolute stability and integral control for infinite-dimensional discrete-time systems. Nonlinear Anal 71(10):4769–4789, 2009. ISSN 0362–546X. doi:10.1016/j.na.2009.03.072. URL http://dx.doi.org/10.1016/j.na.2009.03.072

  • Crooks KR, Sanjayan M, Doak DF (1998) New insights on cheetah conservation through demographic modeling. Conserv Biol 12(4):889–895

    Article  Google Scholar 

  • Cushing JM (1998) An introduction to structured population dynamics. SIAM, Philadelphia

  • Davison EJ (1975) Multivariable tuning regulators: the feedforward and robust control of a general servomechanism problem. In Proceedings of the IEEE Conference on Decision and Control including the 14th Symposium on Adaptive Processes. Institute of Electrical and Electronics Engineers, Houston, pp 180–187

  • Davison EJ (1976) Multivariable tuning regulators: the feedforward and robust control of a general servomechanism problem. IEEE Trans Autom Control AC 21(1):35–47 (ISSN 0018–9286)

  • de Kroon H, Plaisier A, van Groenendael J, Caswell H (1986) Elasticity: the relative contribution of demographic parameters to population growth rate. Ecology 67(5):1427–1431

    Article  Google Scholar 

  • Degla G (2008) An overview of semi-continuity results on the spectral radius and positivity. J Math Anal Appl 338(1):101–110, 2008 (ISSN 0022–247X). doi:10.1016/j.jmaa.2007.05.011. URL http://dx.doi.org/10.1016/j.jmaa.2007.05.011

  • Deines A, Peterson E, Boeckner D, Boyle J, Keighley A, Kogut J, Lubben J, Rebarber R, Ryan R, Tenhumberg B et al (2007) Robust population management under uncertainty for structured population models. Ecol Appl 17(8):2175–2183

    Article  Google Scholar 

  • Demetrius L (1969) The sensitivity of population growth rate to perturbations in the life cycle components. Math Biosci 4(1–2):129–136

    Article  MATH  Google Scholar 

  • Doyle J (1978) Guaranteed margins for IQG regulators. IEEE Trans Autom Control 23(4):756–757 (ISSN 0018–9286). doi:10.1109/TAC.1978.1101812.

  • Doyle JC, Francis BA, Tannenbaum AR (1992) Feedback control theory. Macmillan Publishing Company, New York (ISBN 0-02-330011-6)

  • Dutech A, Scherrer B (2013) Partially observable Markov decision processes. Wiley, pp 185–228 (ISBN 9781118557426). doi:10.1002/9781118557426.ch7. URL http://dx.doi.org/10.1002/9781118557426.ch7

  • Easterling M, Ellner S, Dixon P (2000) Size-specific sensitivity: applying a new structured population model. Ecology 81(3):694–708

    Article  Google Scholar 

  • El-Samad H, Goff J, Khammash M (2002) Calcium homeostasis and parturient hypocalcemia: an integral feedback perspective. J Theor Biol 214(1):17–29

    Article  MathSciNet  Google Scholar 

  • Ellner S, Rees M (2006) Integral projection models for species with complex demography. Am Nat 167(3):410–428

    Article  Google Scholar 

  • Evans LC (2010) Partial differential equations, volume 19 of Graduate Studies in Mathematics. American Mathematical Society, Providence, 2nd edn (ISBN 978-0-8218-4974-3)

  • Evans WR (1948) Graphical analysis of control systems. Trans Am Inst Electr Eng 67(1):547–551 (ISSN 0096–3860). doi:10.1109/T-AIEE.1948.5059708

  • Evans WR (1950) Control system synthesis by root locus method. Trans Am Inst Electr Eng 69(1):66–69 (ISSN 0096–3860). doi:10.1109/T-AIEE.1950.5060121

  • Fackler P, Pacifici K (2014) Addressing structural and observational uncertainty in resource management. J Environ Manag 133:27–36

    Article  Google Scholar 

  • Francis B, Wonham W (1976) The internal model principle of control theory. Automatica 12(5):457–465. (ISSN 0005–1098). doi:10.1016/0005-1098(76)90006-6. URL http://www.sciencedirect.com/science/article/pii/0005109876900066

  • Franklin G, Powell J, Emami-Naeini A, Powell J (1994) Feedback control of dynamic systems, vol 3. Addison-Wesley Reading, MA

    Google Scholar 

  • Gouzé J, Rapaport A, Hadj-Sadok M (2000) Interval observers for uncertain biological systems. Ecol Model 133(1):45–56

    Article  Google Scholar 

  • Green M, Limebeer DJN (1995) Linear robust control. Prentice-Hall Inc., Upper Saddle River (ISBN 0-13-102278-4)

  • Grosdidier P, Morari M, Holt BR (1985) Closed-loop properties from steady-state gain information. Ind Eng Chem Fundam 24(2):221–235

    Article  Google Scholar 

  • Haight RG, Polasky S (2010) Optimal control of an invasive species with imperfect information about the level of infestation. Resour Energy Econ 32(4):519–533

    Article  Google Scholar 

  • Haines R, Hittle D (2003) Control systems for heating, ventilating, and air conditioning. Springer, New York

    Google Scholar 

  • Hinrichsen D, Pritchard AJ (1986a) Stability radii of linear systems. Syst Control Lett 7(1):1–10 (ISSN 0167–6911). doi:10.1016/0167-6911(86)90094-0. URL http://dx.doi.org/10.1016/0167-6911(86)90094-0

  • Hinrichsen D, Pritchard AJ (1986b) Stability radius for structured perturbations and the algebraic Riccati equation. Syst Control Lett 8(2):105–113 (ISSN 0167–6911). doi:10.1016/0167-6911(86)90068-X. URL http://dx.doi.org/10.1016/0167-6911(86)90068-X

  • Hodgson D, Townley S (2004) Methodological insight: linking management changes to population dynamic responses: the transfer function of a projection matrix perturbation. J Appl Ecol 41(6):1155–1161

    Article  Google Scholar 

  • Hodgson D, Townley S, McCarthy D (2006) Robustness: predicting the effects of life history perturbations on stage-structured population dynamics. Theor Popul Biol 70(2):214–224

    Article  MATH  Google Scholar 

  • Hunt L (2001) Heterogeneous grazing causes local extinction of edible perennial shrubs: a matrix analysis. J Appl Ecol 38(2):238–252 (ISSN 1365–2664). doi:10.1046/j.1365-2664.2001.00586.x. URL http://dx.doi.org/10.1046/j.1365-2664.2001.00586.x

  • Ilchmann A (1991) Non-identifier-based adaptive control of dynamical systems: a survey, Universal stabilizers and high-gain adaptive control. IMA J Math Control Inf 8(4):321–366 (ISSN 0265–0754). doi:10.1093/imamci/8.4.321. URL http://dx.doi.org/10.1093/imamci/8.4.321

  • Ilchmann A, Ryan EP (1994) Universal \(\lambda \)-tracking for non-linearly-perturbed systems in the presence of noise. Automatica J. IFAC 30(2):337–346 (ISSN 0005–1098). doi:10.1016/0005-1098(94)90035-3. URL http://dx.doi.org/10.1016/0005-1098(94)90035-3

  • Ilchmann A, Ryan EP, Sangwin CJ (2002). Tracking with prescribed transient behaviour. ESAIM Control Optim Calc Var 7:471–493. (ISSN 1292–8119). doi:10.1051/cocv:2002064. URL http://dx.doi.org/10.1051/cocv:2002064

  • Jedrzejewska B, Jedrzejewsk W, Bunevich AN, Milowski L, Krasinski ZA (1997) Factors shaping population densities and increase rates of ungulates in bialowieia primeval forest. Acta Theriologica 42(4):399–451

    Article  Google Scholar 

  • Johnson C (1987) Numerical solution of partial differential equations by the finite element method. Cambridge University Press, Cambridge (ISBN 0-521-34514-6; 0-521-34758-0)

  • Källström C, Åström KJ, Thorell NE, Eriksson J, Sten L (1979) Adaptive autopilots for tankers. Automatica 15:241–254

    Article  Google Scholar 

  • Ke Z, Logemann H, Townley S (2009) Adaptive sampled-data integral control of stable infinite-dimensional linear systems. Syst Control Lett 58(4):233–240 (ISSN 0167–6911). doi:10.1016/j.sysconle.2008.10.015. URL http://dx.doi.org/10.1016/j.sysconle.2008.10.015

  • Kot M (1992) Discrete-time travelling waves: ecological examples. J Math Biol 30(4):413–436

    Article  MATH  MathSciNet  Google Scholar 

  • Kot M, Lewis MA, van den Driessche P (1996) Dispersal data and the spread of invading organisms. Ecology 77(7):2027–2042

    Article  Google Scholar 

  • Landau YD (1979) Adaptive control: the model reference approach, volume 8 of Control and Systems Theory. Marcel Dekker Inc., New York (ISBN 0-8247-6548-6)

  • Lenhart S, Workman JT (2007) Optimal control applied to biological models. Chapman & Hall/CRC Mathematical and Computational Biology Series. Chapman & Hall/CRC, Boca Raton (ISBN 978-1-58488-640-2; 1-58488-640-4)

  • Littman ML (2009) A tutorial on partially observable Markov decision processes. J Math Psychol 53(3):119–125

    Article  MATH  MathSciNet  Google Scholar 

  • Logemann H, Ryan EP (2000) Time-varying and adaptive discrete-time low-gain control of infinite-dimensional linear systems with input nonlinearities. Math. Control Signals Syst 13(4):293–317 (ISSN 0932–4194). doi:10.1007/PL00009871. URL http://dx.doi.org/10.1007/PL00009871

  • Logemann H, Townley S (1997) Discrete-time low-gain control of uncertain infinite-dimensional systems. IEEE Trans Automat Control 42(1):22–37 (ISSN 0018–9286). doi:10.1109/9.553685. URL http://dx.doi.org/10.1109/9.553685

  • Lubben J, Tenhumberg B, Tyre A, Rebarber R (2008) Management recommendations based on matrix projection models: the importance of considering biological limits. Biol Conserv 141(2):517–523

    Article  Google Scholar 

  • Lubben J, Boeckner D, Rebarber R, Townley S, Tenhumberg B (2009) Parameterizing the growth-decline boundary for uncertain population projection models. Theor Popul Biol 75(2–3):85–97

    Article  MATH  Google Scholar 

  • Lunze J (1985) Determination of robust multivariable I-controllers by means of experiments and simulation. Syst Anal Model Simul 2(3):227–249 (ISSN 0232–9298)

  • McCarthy MA, Possingham HP, Gill A (2001) Using stochastic dynamic programming to determine optimal fire management for Banksia ornata. J Appl Ecol 38(3):585–592

    Article  Google Scholar 

  • Meir E, Andelman S, Possingham HP (2004) Does conservation planning matter in a dynamic and uncertain world? Ecol Lett 7(8):615–622

    Article  Google Scholar 

  • Monahan GE (1982) State of the art survey of partially observable markov decision processes: theory, models, and algorithms. Manag Sci 28(1):1–16

    Article  MATH  MathSciNet  Google Scholar 

  • Morari M (1985) Robust stability of systems with integral control. IEEE Trans Automat Control 30(6):574–577 (ISSN 0018–9286). doi:10.1109/TAC.1985.1104012. URL http://dx.doi.org/10.1109/TAC.1985.1104012

  • Nichols JD, Sauer JR, Pollock KH, Hestbeck JB (1992) Estimating transition probabilities for stage-based population projection matrices using capture-recapture data. Ecology 73(1):306–312 (ISSN 00129658). URL http://www.jstor.org/stable/1938741

  • O’Connor T (1993) The influence of rainfall and grazing on the demography of some African Savanna grasses: a matrix modelling approach. J Appl Ecol, pp 119–132

  • Ozgul A, Childs DZ, Oli MK, Armitage KB, Blumstein DT, Olson LE, Tuljapurkar S, Coulson T (2010) Coupled dynamics of body mass and population growth in response to environmental change. Nature 466(7305):482–485

    Article  Google Scholar 

  • Pulliam HR (1988) Sources, sinks, and population regulation. Am Nat 652–661

  • Puterman ML (1994) Markov decision processes: discrete stochastic dynamic programming. Applied probability and statistics, Wiley Series in Probability and Mathematical Statistics. Wiley , New York (ISBN 0-471-61977-9)

  • Ran ACM, Reurings MCB (2002) The symmetric linear matrix equation. Electron J Linear Algebra 9:93–107 (ISSN 1081–3810)

  • Rees M, Ellner S (2009) Integral projection models for populations in temporally varying environments. Ecol Monogr 79(4): 575–594 (ISSN 00129615). URL http://www.jstor.org/stable/40385228

  • Regan HM, Colyvan M, Burgman MA (2002) A taxonomy and treatment of uncertainty for ecology and conservation biology. Ecol Appl 12(2):618–628

    Article  Google Scholar 

  • Regan TJ, Chades I, Possingham HP (2011) Optimally managing under imperfect detection: a method for plant invasions. J Appl Ecol 48(1):76–85

    Article  Google Scholar 

  • Rose KE, Louda SM, Rees M (2005) Demographic and evolutionary impacts of native and invasive insect herbivores on \(Cirsium canescens\). Ecology 86(2):453–465

    Article  Google Scholar 

  • Roy M, Holt RD, Barfield M (2005) Temporal autocorrelation can enhance the persistence and abundance of metapopulations comprised of coupled sinks. Am Nat 166(2):246–261

    Article  Google Scholar 

  • Safonov M, Fan K (1997) Special issue: multivariable stability margin. Int J Robust Nonlinear Control 7(2):97–226

    Article  Google Scholar 

  • Samad T, Annaswamy A (eds) (2011) The impact of control technology. IEEE Control Systems Society. Available at http://ieeecss.org/general/impact-control-technology

  • Sarrazin F, Barbault R (1996) Reintroduction: challenges and lessons for basic ecology. Trends Ecol Evol 11(11):474–478

    Article  Google Scholar 

  • Saunders PT, Koeslag JH, Wessels JA (1998) Integral rein control in physiology. J Theor Biol 194(2):163–173

    Article  Google Scholar 

  • Shea K, Possingham HP (2000) Optimal release strategies for biological control agents: an application of stochastic dynamic programming to population management. J Appl Ecol 37(1):77–86

    Article  Google Scholar 

  • Shea K, Possingham HP, Murdoch WW, Roush R (2002) Active adaptive management in insect pest and weed control: intervention with a plan for learning. Ecol Appl 12(3):927–936

    Article  Google Scholar 

  • Sragovich VG (2006) Mathematical theory of adaptive control, volume 4 of Interdisciplinary mathematical sciences. In: Spaliński J, Stettner L, Zabczyk J (eds) World Scientific Publishing Co., Pt. Ltd., Hackensack (ISBN 981-256-371-7). Translated from the Russian by Sinitzin IA

  • Stott I, Townley S, Carslake D, Hodgson D (2010) On reducibility and ergodicity of population projection matrix models. Methods Ecol Evol 1(3):242–252

    Google Scholar 

  • Stott I, Hodgson DJ, Townley S (2012) Beyond sensitivity: nonlinear perturbation analysis of transient dynamics. Methods Ecol Evol 3(4):673–684

    Article  Google Scholar 

  • Tenhumberg B, Tyre AJ, Shea K, Possingham HP (2004) Linking wild and captive populations to maximize species persistence: optimal translocation strategies. Conserv Biol 18(5):1304–1314

    Article  Google Scholar 

  • Walker B (1998) The art and science of wildlife management. Wildl Res 25(1):1–9

    Article  Google Scholar 

  • Westphal MI, Pickett M, Getz WM, Possingham HP (2003) The use of stochastic dynamic programming in optimal landscape reconstruction for metapopulations. Ecol Appl 13(2):543–555

    Article  Google Scholar 

  • Williams BK (2001) Uncertainty, learning, and the optimal management of wildlife. Environ Ecol Stat 8(3):269–288

    Article  MathSciNet  Google Scholar 

  • Williams BK (2011a) Passive and active adaptive management: approaches and an example. J Environ Manag 92(5):1371–1378

    Article  Google Scholar 

  • Williams BK (2011b) Adaptive management of natural resources—framework and issues. J Environ Manag 92(5):1346–1353

    Article  Google Scholar 

  • Wilson PH (2003) Using population projection matrices to evaluate recovery strategies for snake river spring and summer chinook salmon. Conserv Biol 17(3):782–794 (ISSN 1523–1739). doi:10.1046/j.1523-1739.2003.01535.x. URL http://dx.doi.org/10.1046/j.1523-1739.2003.01535.x

  • Yi T-M, Huang Y, Simon MI, Doyle J (2000) Robust perfect adaptation in bacterial chemotaxis through integral feedback control. Proc Natl Acad Sci 97(9):4649–4653

    Article  Google Scholar 

  • Zhou K, Doyle JC (1998) Essentials of robust control, vol 104. Prentice Hall, Upper Saddle River

    Google Scholar 

  • Zhou K, Doyle J, Glover K (1996) Robust and optimal control. Prentice Hall, Englewood Cliffs

    MATH  Google Scholar 

Download references

Acknowledgments

Chris Guiver and Adam Bill are fully supported and Dave Hodgson, Stuart Townley, Richard Rebarber and Hartmut Logemann are partially supported by the UK Engineering and Physical Sciences Research Council (EPSRC) grant EP/I019456/1.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Chris Guiver.

Appendix

Appendix

1.1 The constants \(\upgamma ,\kappa ,\mu _g\) and \(\nu _g\)

\(\bullet \) We first prove the inequality (3.24). In the proceeding arguments, for a sequence \(v\) we use the notation \(\hat{v}\) to denote the \(Z\)-transform of \(v\) given by

$$\begin{aligned} \hat{v}(z) = \sum _{j=0}^\infty v(j) z^{-j}, \end{aligned}$$

defined for all complex \(z\) where the summation converges absolutely. The step response of the linear system (3.5) is the output of (3.5) subject to zero initial state (\(x^0 = 0\)) and constant input \(\tilde{u} =1\) and is given by

$$\begin{aligned} s(0) = 0, \quad s(t) = \sum _{j=0}^{t-1} c^TA^j b, \quad t= 1,2,\dots . \end{aligned}$$

Assumption (A1) ensures that \(s(t) \rightarrow G(1)\) as \(t \rightarrow \infty \). Furthermore, a calculation shows that \(s\) has \(Z\)-transform

$$\begin{aligned} \hat{s}(z) = \frac{zG(z)}{z-1}, \quad z \in \mathbb {C}, \quad \vert z \vert >1\,. \end{aligned}$$

Under the assumptions that \(A,b,c^T \ge 0\) and (A2) it follows that \(s(t) \ge 0\) and is non-decreasing. We define the step response error

$$\begin{aligned} e(t) = s(t) - G(1), \quad t =0,1,2,\dots , \end{aligned}$$

which is consequently non-positive, non-decreasing and converges to 0. Furthermore, the \(Z\)-transform of \(e\) satisfies

$$\begin{aligned} \frac{\hat{e}(z)}{z} = \frac{G(z) - G(1)}{z-1}, \end{aligned}$$
(5.1)

for every complex \(z\) with modulus greater than one. Since \(G\) is differentiable at \(z=1\) we note that

$$\begin{aligned} \lim _{z \rightarrow 1} \hat{e}(z) = \lim _{z \rightarrow 1} \frac{\hat{e}(z)}{z} = \lim _{z \rightarrow 1} \frac{G(z) - G(1)}{z-1} = G^{\prime }(1)\,. \end{aligned}$$
(5.2)

As \(z \mapsto \tfrac{\hat{e}(z)}{z}\) is continuous outside of the unit disc the above shows that we can extend \(z \mapsto \tfrac{\hat{e}(z)}{z}\) continuously to \(z=1\) with

$$\begin{aligned} \hat{e}(1) =\frac{\hat{e}(1)}{1} = G^{\prime }(1)\,. \end{aligned}$$
(5.3)

We now use (5.3) and the property that \(e(t) \le 0\) for every \(t\) to show that for any complex \(z\) with modulus one,

$$\begin{aligned} -G^{\prime }(1)&= -\hat{e}(1) = -\sum _{k=0}^\infty e(k) = \sum _{k=0}^\infty \vert e(k) \vert \cdot \vert z^{-(k+1)} \vert \ge \left| \frac{\hat{e} (z)}{z} \right| \ge \mathrm{Re } \,\left( \frac{-\hat{e} (z)}{z} \right) \nonumber \\&\quad = \mathrm{Re} \, \left[ \frac{G(1) - G(z)}{z-1}\right] \nonumber \\&\quad = -\frac{G(1)}{2} - \mathrm{Re} \, \left[ \frac{ G(z)}{z-1}\right] . \end{aligned}$$
(5.4)

Rearranging (5.4) gives

$$\begin{aligned} \mathrm{Re}\, \left[ \frac{G(z)}{z-1} \right] \ge G^{\prime }(1) - \frac{G(1)}{2}, \quad \text {for all complex } z \text { with modulus one,} \end{aligned}$$

which implies that

$$\begin{aligned} \inf _{\vert z \vert =1 } \mathrm{Re}\, \left[ \frac{G(z)}{z-1}\right] =\inf _{\theta \in [0,2\pi )} \mathrm{Re}\, \left[ \frac{G(\mathrm {e}^{\mathrm {i}\theta })}{\mathrm {e}^{\mathrm {i}\theta }-1}\right] \ge G^{\prime }(1) - \frac{G(1)}{2}. \end{aligned}$$
(5.5)

From Coughlan and Logemann (2009) we have that \(\upgamma \) satisfies \(-\infty <\upgamma \le \tfrac{-G(1)}{2} <0\) as \(G(1) >0\), and so

$$\begin{aligned} 0 > \upgamma \ge \inf _{\theta \in [0,2\pi )} \mathrm{Re} \left[ \frac{ G(\mathrm {e}^{\mathrm {i}\theta })}{\mathrm {e}^{\mathrm {i}\theta } -1} \right] \ge G^{\prime }(1) - \frac{G(1)}{2}\,, \end{aligned}$$
(5.6)

where we have used the estimate (5.5). It is clear from \(\upgamma \le \tfrac{-G(1)}{2}\) and (5.6) that \(G^{\prime }(1) <0\) and consequently inequality (5.6) is equivalent to

$$\begin{aligned} 0 < -\upgamma = \vert \upgamma \vert \le -G^{\prime }(1) + \frac{G(1)}{2} = \vert G^{\prime }(1) \vert + \frac{G(1)}{2} {=:} \frac{1}{\kappa }, \end{aligned}$$

which implies (3.24).

\(\bullet \) The constants \(\mu _g\) and \(\nu _g\) in (3.16) and (3.20) are given by

$$\begin{aligned} \mu _g {:=} \sum _{j=0}^\infty \left| \begin{bmatrix}c^T&0 \end{bmatrix}\begin{bmatrix}A&\quad b \\ -gc^T&\quad 1 \end{bmatrix}^j\begin{bmatrix}0 \\ g \end{bmatrix}\right| \quad \text {and} \quad \nu _g {:=} \sum _{j=0}^\infty \left| \begin{bmatrix}c^T&0 \end{bmatrix}\begin{bmatrix}A&\quad b \\ -gc^T&\quad 1 \end{bmatrix}^j\begin{bmatrix}b \\ 0 \end{bmatrix}\right| \,, \end{aligned}$$

respectively, which are both finite since by assumption \(g >0\) is such that has \(\rho (A_g) <1\).

\(\bullet \) We now derive the inequality (3.29). For complex \(z\) with modulus greater than or equal to one the transfer function \(G\) given by (3.11) of the linear system (3.5) can be written as

$$\begin{aligned} G(z) = \sum _{j =0}^\infty g_j z^{-j}, \quad \text {where} \quad g_j = \left\{ \begin{aligned}&0,&j=0, \\&c^T A^{j-1}b,&j \ge 1\,. \end{aligned} \right. \end{aligned}$$

We define \(z \mapsto \tilde{G}(z) {:=} zG(z)\) and introduce the constant

$$\begin{aligned} \tilde{\upgamma }{:=}\sup _{q\ge 0}\left\{ \inf _{\theta \in [0,2\pi )}\mathrm{Re} \left[ \left( \frac{q}{\mathrm {e}^{i\theta }} +\frac{1}{\mathrm {e}^{i\theta }-1}\right) \tilde{G}(\mathrm {e}^{i\theta })\right] \right\} \,. \end{aligned}$$

We know that \(-\infty < \tilde{\upgamma } \le -\tfrac{\tilde{G}(1)}{2} = -\tfrac{G(1)}{2}\). By inspection of the definition of \(\tilde{G}\), the constant \(\tilde{\upgamma }\), and \({\upgamma }_0\) in (3.29) we see that

$$\begin{aligned} \tilde{\upgamma }&= {\upgamma }_0, \end{aligned}$$
(5.7)
$$\begin{aligned} \tilde{G}^{\prime }(z)&= G(z) + zG^{\prime }(z) \quad \text {and so} \quad \tilde{G}^{\prime }(1) = G(1) + G^{\prime }(1)\,. \end{aligned}$$
(5.8)

We note from (5.8) that

$$\begin{aligned} \tilde{G}^{\prime }(1) = G(1) + G^{\prime }(1) = \sum _{j =1}^\infty g_j z^{-j} - \sum _{j =1}^\infty j g_j z^{-j} = \sum _{j =1}^\infty (1 -j)g_j z^{-j} \le 0\,, \end{aligned}$$
(5.9)

and consequently we can apply the estimate (3.24) to \(\tilde{G}\) to yield that

$$\begin{aligned} \frac{2}{\tilde{G}(1) + 2 \vert \tilde{G}^{\prime }(1)\vert } \le \frac{1}{\vert \tilde{\upgamma } \vert }\,. \end{aligned}$$
(5.10)

In light of (5.7), (5.8) and the following definition of \(\kappa _0\), (5.10) implies that

$$\begin{aligned} \kappa _0 : = \frac{2}{G(1) + 2 \vert G(1) + G^{\prime }(1) \vert } = \frac{2}{\tilde{G}(1) + 2 \vert \tilde{G}^{\prime }(1)\vert } \le \frac{1}{\vert \tilde{\upgamma } \vert } = \frac{1}{\vert {\upgamma }_0 \vert }\,, \end{aligned}$$

as required. Finally, as \(G(1) >0\), it is clear from (5.9) that \(G^{\prime }(1) \le \tilde{G}^{\prime }(1) \le 0\) and thus

$$\begin{aligned} \vert G(1) + G^{\prime }(1) \vert = \vert \tilde{G}^{\prime }(1) \vert < \vert G^{\prime }(1)\vert \,. \end{aligned}$$
(5.11)

From inequality (5.11) we deduce that

$$\begin{aligned} \kappa = \frac{2}{G(1) + 2 \vert G^{\prime }(1) \vert } < \frac{2}{G(1) + 2 \vert G(1) + G^{\prime }(1) \vert } = \kappa _0\,. \end{aligned}$$

1.2 Proof of Theorem 2

Let \((x,u)\) denote the solution of

$$\begin{aligned} \left. \begin{aligned} x(t+1)&= Ax(t)+bu(t),\,\,&x(0)=x^0, \\ u(t+1)&= u(t)+g(r-(1+\varepsilon (t))c^Tx(t)),&u(0)=u^0, \end{aligned}\right\} \quad t = 0,1,2,\dots , \end{aligned}$$
(5.12)

the integral control system (3.10) with proportional observation errors \(\varepsilon (t)\). When \(\varepsilon \) is a sequence of random variable then so are \(x\) and \(u\). We let

$$\begin{aligned} x_*= (I-A)^{-1}b\frac{r}{G(1)}, \quad u_*= \frac{r}{G(1)}, \end{aligned}$$

which are equilibria of (3.10) as in (3.12). For notational convenience we define the random variable

$$\begin{aligned} z(t) {:=} \begin{bmatrix}x(t) - x_*\\ u(t) - u_* \end{bmatrix}, \quad t =0,1,2,\dots , \end{aligned}$$
(5.13)

a vector with \(n+1\) components. A short calculation using (3.12) and (5.12) demonstrates that \(z(t)\) has dynamics given by

$$\begin{aligned} z(t+1) = \left[ \begin{bmatrix}A&\quad b \\ -gc^T&\quad 1 \end{bmatrix} - \begin{bmatrix}0 \\ 1 \end{bmatrix}g\varepsilon (t) \begin{bmatrix}c^T&0 \end{bmatrix}\right] z(t) - \begin{bmatrix}0 \\ 1 \end{bmatrix}gr \varepsilon (t), \quad t = 0,1,2,\dots . \end{aligned}$$
(5.14)

We introduce the notation

$$\begin{aligned} A_g {:=} \begin{bmatrix}A&\quad b \\ -gc^T&\quad 1 \end{bmatrix}, \quad D {:=} \begin{bmatrix}0_{n \times 1} \\ 1 \end{bmatrix}, \quad E {:=} \begin{bmatrix}c^T&0 \end{bmatrix}, \end{aligned}$$

where recall that \(0_{n \times 1}\) is a column vector of \(n\) zeros. With this notation (5.14) can be more concisely expressed as

$$\begin{aligned} z(t+1) = \left[ A_g - g\varepsilon (t) DE\right] z(t) - Dgr \varepsilon (t), \quad t = 0,1,2,\dots . \end{aligned}$$
(5.15)

Letting \(\overline{z(t)} = \mathbb {E}(z(t))\) denote the expectation of \(z(t)\), we take expectations in (5.15) to yield that

$$\begin{aligned} \overline{z(t+1)} = A_g \overline{z(t)}, \quad t = 0,1,2,\dots , \end{aligned}$$
(5.16)

where we have used the facts that expectation is linear, \(\overline{\varepsilon (t)} =0\) and that \(\varepsilon (t)\) and \(z(t)\) are independent. We are assuming that the gain parameter \(g>0\) is such that \(\rho (A_g) <1\), and hence from (5.16) we conclude that

$$\begin{aligned} \lim _{t \rightarrow \infty } \overline{z(t)} = 0, \quad \text {and thus} \quad \lim _{t \rightarrow \infty } \overline{y(t)} = \lim _{t \rightarrow \infty } \begin{bmatrix}c^T&0 \end{bmatrix}\overline{z(t)} +r =r, \end{aligned}$$

establishing claim (1). We now consider the covariance

$$\begin{aligned} \mathrm{cov} \, (z(t), z(t))&= \mathbb {E}\left( (z(t) - \overline{z(t)})(z(t) - \overline{z(t)})^T \right) = \mathbb {E}(z(t)z^T(t)) - \overline{z(t)} \cdot \overline{z^T(t)} \nonumber \\&{=:} C(t) - \overline{z(t)} \cdot \overline{z^T(t)}, \quad t = 0,1,2,\dots , \end{aligned}$$
(5.17)

where \(z^T(t) = (z(t))^T\). We focus on the quantity \(C(t)\), which (appealing to (5.15)) has dynamics

$$\begin{aligned} C(t+1)&= \mathbb {E}(z(t+1)z^T(t+1)) \nonumber \\&= \mathbb {E}\left( \left[ [A_g - g\varepsilon (t) DE] z(t) - Dg r\varepsilon (t)\right] \left[ [A_g - g\varepsilon (t) DE] z(t) - Dgr \varepsilon (t)\right] ^T \right) \nonumber \\&= \mathbb {E}(A_g z(t) (A_g z(t))^T) \nonumber \\&\quad +\, \underbrace{\mathbb {E}(A_g z(t) z^T(t)(-g\varepsilon (t))(DE)^T) + \mathbb {E}((A_g z(t) z^T(t)(-g\varepsilon (t))(DE)^T)^T)}_{=0} \nonumber \\&\quad +\, \underbrace{\mathbb {E}(A_g z(t)\varepsilon (t) D^Tgr) + \mathbb {E}((A_g z(t)\varepsilon (t) D^Tgr)^T)}_{=0} \nonumber \\&\quad +\, g^2 {\sigma }^2 \mathbb {E}(DE z(t) z^T(t) (DE)^T ) \nonumber \\&\quad +\, g^2r^2{\sigma }^2DD^T + \mathbb {E}(D(-g\varepsilon (t))Ez(t)\varepsilon (t)(-D^Tgr)) \nonumber \\&\quad +\, \mathbb {E}((D(-g\varepsilon (t))Ez(t)\varepsilon (t)(-D^Tgr))^T), \end{aligned}$$
(5.18)

for \(t = 0,1,2,\dots \). Equation (5.18) simplifies to

$$\begin{aligned} C(t+1)&= A_g C(t) A_g^T + g^2 {\sigma }^2 (DE)C(t) (DE)^T + g^2 {\sigma }^2 r^2 DD^T\nonumber \\&\quad + rg^2{\sigma }^2 DE \overline{z(t)} D^T + rg^2{\sigma }^2 (DE)^T D\overline{z^T(t)}, \quad t =0,1,2,\dots . \end{aligned}$$
(5.19)

Define \(A_1 {:=} A_g\), \(A_2 {:=} g \sigma DE\) and write

$$\begin{aligned}&c(t)\, {:=}\, \mathrm{vec}\,C(t),\quad p(t) {:=} \mathrm{vec}\,\left( g^2 {\sigma }^2 r^2 DD^T + rg^2{\sigma }^2 DE \overline{z(t)} D^T \nonumber \right. \\&\quad \left. +\, rg^2{\sigma }^2 (DE)^T D\overline{z^T(t)}\right) , \end{aligned}$$
(5.20)

where if \(x_i\) are the columns of the \(n \times n\) matrix \(X = [x_1, x_2, \dots , x_n]\) then

$$\begin{aligned} \mathrm{vec}\,X {:=} \begin{bmatrix}x_1^T&x_2^T&\dots&x_n^T \end{bmatrix}^T \in {\mathbb {R}}^{n^2}. \end{aligned}$$

Arguing now as in Ran and Reurings (2002), the matrix difference equation (5.19) can be written as the \((n+1)^2 \times (n+1)^2 \) linear system

$$\begin{aligned} c(t+1) = \left( \sum _{i=1}^2 A_i \otimes A_i \right) c(t) + p(t), \quad t =0,1,2,\dots , \end{aligned}$$
(5.21)

where \(\otimes \) denotes the Kronecker product. Using the fact that \(\overline{z(t)} \rightarrow 0\) as \(t \rightarrow \infty \) it follows from (5.20) that

$$\begin{aligned} \lim _{t\rightarrow \infty } p(t) = \mathrm{vec}\,(g^2 {\sigma }^2 r^2 DD^T) {=:} p_{\infty }. \end{aligned}$$

Consequently, if

$$\begin{aligned} \rho \left( \sum _{i=1}^2 A_i \otimes A_i \right) = \rho ( A_1 \otimes A_1 + A_2 \otimes A_2)<1, \end{aligned}$$
(5.22)

then for any initial condition \(c(0)\) the solution \(c\) of (5.21) converges to a finite limit \(c_{\infty }\) satisfying

$$\begin{aligned} c_{\infty } = \left( \sum _{i=1}^2 A_i \otimes A_i \right) c_{\infty } + p_{\infty }. \end{aligned}$$
(5.23)

Assuming that (5.22) holds, defining \(C_{\infty }\) as the matrix such that

$$\begin{aligned} c_{\infty } = \mathrm{vec}\, C_{\infty }, \end{aligned}$$

we have from (5.23) that \(C_{\infty }\) must satisfy

$$\begin{aligned} C_{\infty } = A_g C_{\infty } A_g^T + g^2 {\sigma }^2 (DE)C_{\infty } (DE)^T + g^2 {\sigma }^2 r^2 DD^T. \end{aligned}$$

Furthermore, as \(C(t)\) converges to \(C_{\infty }\) the iterative scheme (5.19) provides a method for approximating \(C_{\infty }\).

It remains to find a characterisation of the stability condition (5.22). Recalling that for square matrices \(X,Y\)

$$\begin{aligned} \sigma (X\otimes Y) = \left\{ \lambda \mu \, : \, \lambda \in \sigma (X), \mu \in \sigma (Y) \right\} , \end{aligned}$$

we have that

$$\begin{aligned} \rho (A_1 \otimes A_1) = \rho (A_g \otimes A_g) = \rho (A_g)^2 <1, \end{aligned}$$

and thus we can view \(A_1 \otimes A_1 + A_2 \otimes A_2\) as a structured perturbation of \(A_1 \otimes A_1\). Therefore we can characterise the condition (5.22) by appealing to stability radius arguments (Hinrichsen and Pritchard 1986a, b). A calculation shows that \(A_2 \otimes A_2\) is a rank one perturbation, namely

$$\begin{aligned}&A_2 \otimes A_2 = g^2 {\sigma }^2 \begin{bmatrix}0&\quad 0 \\ c^T&\quad 0 \end{bmatrix}\otimes \begin{bmatrix}0&\quad 0 \\ c^T&\quad 0 \end{bmatrix} = g^2 {\sigma }^2 \begin{bmatrix}0_{(n^2 + 2n) \times 1} \\ 1 \end{bmatrix}\begin{bmatrix} (c^T&\quad 0) \otimes (c^T&\quad 0) \end{bmatrix}\\&\quad {=:} g^2 {\sigma }^2 \tilde{D} \tilde{E}. \end{aligned}$$

Hence, condition (5.22) is satisfied if, and only if,

$$\begin{aligned} {\sigma }^2 g^2 < \frac{1}{ {\displaystyle \mathop {\max }_{ \vert z \vert =1 }}\, \vert \tilde{E}(zI- A_g \otimes A_g)^{-1}\tilde{D} \vert }, \end{aligned}$$

which is equivalent to the condition (3.18). We can now take the limit as \(t \rightarrow \infty \) in (5.17) and use that \(\overline{z(t)}\) converges to zero to deduce that

$$\begin{aligned} \lim _{t \rightarrow \infty } \mathrm{cov} \, (z(t), z(t)) = \lim _{t \rightarrow \infty } C(t) = C_{\infty }. \end{aligned}$$
(5.24)

The variance of the output satisfies

$$\begin{aligned} \mathrm{var}\,y(t)&= \mathrm{var}\,(y(t) -r) = \mathrm{var}\left( \begin{bmatrix}c^T&\quad 0 \end{bmatrix}z(t)\right) = \mathrm{cov} \left( \begin{bmatrix}c^T&\quad 0 \end{bmatrix}z(t), \begin{bmatrix}c^T&\quad 0 \end{bmatrix}z(t) \right) \nonumber \\&= \begin{bmatrix}c^T&0 \end{bmatrix} \mathrm{cov} \left( z(t), z(t) \right) \begin{bmatrix}c \\ 0 \end{bmatrix}, \quad t =0,1,2, \dots . \end{aligned}$$
(5.25)

Therefore taking limits in (5.25) and invoking (5.24) we have that

$$\begin{aligned} \lim _{t \rightarrow \infty } \mathrm{var}\,y(t) = \begin{bmatrix}c^T&\quad 0 \end{bmatrix} C_{\infty } \begin{bmatrix}c \\ 0 \end{bmatrix} <\infty , \end{aligned}$$

proving claim (2).

1.3 More general input nonlinearities

We comment further on Remarks 2 and 3. Theorem 3 applies when \(\phi \) in (3.21) is replaced by any function \(\phi : \mathbb {R}\rightarrow \mathbb {R}\) that satisfies a so-called Lipschitz condition, namely

(A3) :

there exists \(l>0\) such that \(0\le \phi (v)-\phi (w)\le l(v-w)\) for all \(v\ge w\).

The constant \(l\) in assumption (A3) is called the Lipschitz constant of \(\phi \) and, for example, the function \(\phi \) in (3.21) satisfies (A3) with \(l =1\).

For a function \(\phi : \mathbb {R}\rightarrow \mathbb {R}\) and a set \(X \subseteq \mathbb {R}\) we let \(\mathrm{im }\, \phi \) and \(\phi ^{-1}(X)\) denote the image of \(\phi \) and preimage of \(X\) of under the function \(\phi \) respectively.

In this more general setting, Theorem 3 can be restated as: Assume that (3.22) satisfies (A1)(A3). Then, for every \(r \in \mathbb {R}\) such that \(r/G(1)\in \mathrm{im}\,\phi \), every \(g\in (0,1/|\upgamma l|)\) and all initial conditions \((x^0,u^0)\in {\mathbb {R}}^n\times \mathbb {R}\), statements (1), (2) and (3) hold. Moreover, if additionally \(\phi ^{-1}(r/G(1))\) is a singleton then \((x^r,u^r)\) is a globally asymptotically stable equilibrium of (3.22).

The adaptive integral control result, Theorem 4, can be restated as: Assume that (3.26) satisfies assumptions (A1)-(A3). Then, for every \(r\in \mathbb {R}\) such that \(r/G(1)\in \mathrm{im}\,\phi \), and all initial conditions \((x^0,u^0,h^0)\in {\mathbb {R}}^n\times \mathbb {R}\times (0,\infty )\),

  1. (1)

    \(\displaystyle {\lim _{t\rightarrow \infty }}u(t)=\displaystyle {\frac{r}{G(1)}}\),

  2. (2)

    \(\displaystyle {\lim _{t\rightarrow \infty }}x(t)=x^r{:=}(I-A)^{-1}b\displaystyle {\frac{r}{G(1)}}\),

  3. (3)

    \(\displaystyle {\lim _{t\rightarrow \infty }}y(t)=\lim _{t\rightarrow \infty } c^Tx(t)=r\).

Moreover, if \(\phi ^{-1}(r/G(1))\) is a singleton, then

  1. (4)

    the non-increasing gain \(k(t)=1/h(t)\) converges to a positive limit as \(t\rightarrow \infty \),

  2. (5)

    \(\displaystyle {\lim _{t\rightarrow \infty }}w(t)=w^r\), where \(\phi (w^r)=\displaystyle {\frac{r}{G(1)}}\).

1.4 Proof of Theorem 5

By assumption \(k>0\) is chosen so that \(A - kbc^T\) is componentwise nonnegative. Since \(A, b\) and \(c^T\) are also componentwise nonnegative we clearly have that \(A \ge A- kbc^T\) (the inequality is understood componentwise) and so Berman and Plemmons (1994, p. 27) implies that

$$\begin{aligned} 0 \le \rho (A-kbc^T) \le \rho (A) <1. \end{aligned}$$

We deduce that assumption (A1) holds for \(A - kbc^T\). Moreover, one can show that the transfer function of \((A-kbc^T, b,c^T)\) is

$$\begin{aligned} z \mapsto G_k(z) = \frac{G(z)}{1+kG(z)}, \quad \text {so that} \quad G_k(1) = \frac{G(1)}{1+kG(1)} >0, \end{aligned}$$

implying that assumption (A2) applies to \((A-kbc^T, b,c^T)\). Therefore, Theorem 1 now applies to the feedback system (3.30), that is the original integral control system (3.10) with \(A\) replaced by \(A-kbc^T\). It is straightforward to demonstrate that the equilibria \((x^{*}, u^{*})\) of (3.30) are the same as those of (3.10).

1.5 IPM example

Following Briggs et al. (2010) we take \(\varOmega = [\mathrm {e}^{-0.5}, \mathrm {e}^{3.5}]\), so that \(\alpha =\mathrm {e}^{-0.5} \sim 0.6 \) and \(\beta =\mathrm {e}^{3.5} \sim 33\). The kernel \(k\) is divided into

$$\begin{aligned} k(y,x) = p(y,x) + f(y,x), \end{aligned}$$

where \(p\) denotes the survival component and \(f\) denotes the reproductive component. These have respective decompositions

$$\begin{aligned} p(y,x) = s(x)(1-f_p(x))g(y,x), \quad \text {and} \quad \text{ f }(y,x) = P_eJ(y)s(x)f_p(x)S(x). \end{aligned}$$

The functions \(s,f_p,g,J, S\) and constant \(P_e\) are as in (Briggs et al. (2010), Table 1), where a biological interpretation is also provided. For our simulations we have altered \(s\), \(S\) and \(P_e\) to

$$\begin{aligned} s(x) = 0.7\frac{\mathrm {e}^{0.85x -0.62}}{1+ \mathrm {e}^{0.85x -0.62}}, \quad S(x) = \mathrm {e}^{1.85x +0.37}, \quad P_e = 0.05. \end{aligned}$$

We have made these alterations so that the population is declining, and we can apply our results.

Finite element approximations are one method of reducing the infinite-dimensional IPM to a finite-dimensional difference equation by discretising the spatial domain. That is, the function space \(L^1(\varOmega )\) is approximated by an indexed sequence of finite-dimensional subspaces which get ‘closer’ to \(L^1(\varOmega )\) as the index \(N\) increases. In what follows we give a very brief description of how finite elements is used to derive an approximation of the IPM and refer the reader to the texts by Johnson (1987) or Brenner and Scott (1994) for a thorough treatment.

For an integer \(N\), the interval \([\alpha , \beta ]\) is partitioned into \(N\) subintervals with \(N+1\) equally spaced endpoints \(s_i\) defined by

$$\begin{aligned} s_i = \alpha + \frac{(i-1)(\beta -\alpha )}{N}, \quad 1 \le i \le N+1. \end{aligned}$$

In particular \(s_1 = \alpha \) and \(s_{N+1} =\beta \). The \(N+1\) ‘hat’ or ‘tent’ functions \(\delta _i\) are defined by

$$\begin{aligned} \delta _i(s) = \left\{ \begin{aligned}&\frac{s-s_{i-1}}{s_i - s_{i-1}}&\quad s \in [s_{i-1}, s_i], \\&\frac{s_{i+1}-s}{s_{i+1} - s_i}&\quad s \in [s_i, s_{i+1}], \\&0&\quad \text {otherwise,} \end{aligned} \right. \quad 1\le i \le N+1, \end{aligned}$$
(5.26)

where \(s_0 = s_1 = \alpha \) and \(s_{N+2} = s_{N+1} = \beta \). The hat functions are more readily understood visually, and some examples are plotted in Fig. 15.

Fig. 15
figure 15

Three sample hat functions defined by (5.26) with \(\alpha = 0\), \(\beta = 1\) and \(N=10\). The functions \(\delta _1\), \(\delta _5\) and \(\delta _{11}\) are plotted in solid, dashed and dashed–dotted lines respectively

Loosely speaking, the finite element method assumes that functions in \(L^1(\varOmega )\) are well approximated by a linear combination of finitely many of the \(\delta _i\) functions. And so, supposing that \(n\) is a solution of the IPM (3.5), using (3.32), (3.34) and (3.35), with input \(u\) and output \(y\) then for any continuous function \(v\) the following equation is satisfied

$$\begin{aligned} \int _{\xi \in \varOmega } v(\xi ) \left[ n(\xi ,t+1) - (An)(\xi ,t) - b(\xi )u(t)\right] \,d\xi =0, \quad t =0,1,2,\dots . \end{aligned}$$
(5.27)

We assume that \(v\) and \(n\) can be written as a linear combination of the \(\delta _i\), that is, as

$$\begin{aligned} v(t,\xi ) =\sum _{i=1}^{N+1} v_i(t) \delta _i(\xi ), \quad n(t,\xi ) =\sum _{j=1}^{N+1} n_j(t) \delta _j(\xi ), \end{aligned}$$
(5.28)

for some coefficients \(v_i\) and \(n_j\). Substituting (5.28) into (5.27) and simplifying gives the following matrix equation

$$\begin{aligned} M\mathtt{n }(t+1) = D \mathtt{n }(t) + J u(t), \quad t = 0,1,2,\dots , \end{aligned}$$
(5.29)

where \(\mathtt{n }(t) = \begin{bmatrix}n_1(t)&\dots&n_{N+1}(t) \end{bmatrix}^T\) and the matrices \(M,D\) and vector \(J\) have components given by

$$\begin{aligned} \left. \begin{aligned} M_{ij}&= \int _{\xi \in \varOmega } \delta _i(\xi ) \delta _j(\xi ) \, d\xi ,&D_{ij}&= \int _{\xi \in \varOmega } \delta _i(\xi ) \int _{s \in \varOmega } k(\xi ,s) \delta _j(s) \,ds d\xi , \\ J_i&= \begin{bmatrix}J_1&\dots&J_{N+1} \end{bmatrix}^T,&J_i&=\int _{\xi \in \varOmega } \delta _i(\xi ) b(\xi ) \, d\xi , \\ \end{aligned} \right\} \quad 1 \le i,j \le N+1. \end{aligned}$$

It is straightforward to see that the matrix \(M\) is invertible; if \(q \in \mathbb {C}^{N+1}\) has \(i^\mathrm{th}\) component \(q_i\) then we see that

$$\begin{aligned} \overline{q}^T Mq = \sum _{i,j=1}^{N+1} \overline{q_i} M_{ij} q_j = \int _{\xi \in \varOmega } \left\| \sum _{i=1}^{N+1} q_i \delta _i(\xi ) \right\| ^2 \, d\xi \ge 0\,. \end{aligned}$$
(5.30)

Furthermore, if \(\overline{q}^T Mq =0\) then as \( \xi \mapsto \sum _{i=1}^{N+1} q_i \delta _i(\xi ) \) is continuous it follows from (5.30) that

$$\begin{aligned} \sum _{i=1}^{N+1} q_i \delta _i(\xi ) = 0, \quad \forall \, \xi \in \varOmega \quad \Rightarrow \quad q_i =0, \quad \forall \, i \in \{1,2,\dots , N+1\}\,, \end{aligned}$$

and thus \(q=0\), proving that \(M\) is invertible.

When the output is of the form

$$\begin{aligned} y(t) = \int _{\xi _1}^{\xi _2} n(s,t) \,ds, \quad t = 0,1,2,\dots \,, \end{aligned}$$
(5.31)

where \(\xi _1 < \xi _2\) denote the range of stage-classes observed, then substituting (5.28) into Eq.  (5.31) gives \(y(t) = F\mathtt{n }(t)\), where the row vector \(F = \begin{bmatrix}F_1&\dots&F_{N+1} \end{bmatrix}\) has components

$$\begin{aligned} F_i = \int _{\xi _1}^{\xi _2} \delta _i(s) \,ds, \quad 1 \le i \le N+1. \end{aligned}$$

Therefore, we have the following system with \(N+1\) states

$$\begin{aligned} \left. \begin{aligned} \mathtt{n }(t+1)&= M^{-1}D \mathtt{n }(t) + M^{-1}J u(t), \\ y(t)&= F\mathtt{n }(t), \end{aligned} \right\} \quad t = 0,1,2,\dots , \end{aligned}$$
(5.32)

which is an approximation of the IPM (3.5) and can be readily implemented. The matrix \(M\) and vector \(F\) can be found analytically, whilst \(D\) and \(J\) generally need to be computed numerically. This can be achieved using quadrature, or for example the Matlab functions integral and integral2. In principle, larger \(N\) gives rise to a closer approximation, but clearly adds complexity to simulations. We denote by \(G_N\) the transfer function of (5.32) so that the steady-state gain of (5.32) is

$$\begin{aligned} G_N(1) = F(I - M^{-1}D)^{-1}M^{-1}J, \end{aligned}$$

(whenever \(\rho (M^{-1}D) <1\)). For our example we worked on the log of the interval \([\alpha , \beta ]\), as this gave better results. As such the above goes through with \(s_1 = -0.5\), \(s_{N+1} = 3.5\). Figure 16 plots both the spectral radius of \(M^{-1}D\) and the steady state gain \(G_N(1)\) for increasing \(N\). The figure suggests that both converge for \(N \ge 10\) and thus we choose \(N=12\) for the simulations in Fig. 14. Furthermore, this suggests that the model in Example 9 satisfies both assumptions (A1 \(^{\prime }\)) and (A2 \(^{\prime }\)).

Fig. 16
figure 16

Spectral radius in solid-crossed and steady-state gain in solid-circled of the finite element approximations (5.32) of the IPM model of platte thistle of Example 9

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Guiver, C., Logemann, H., Rebarber, R. et al. Integral control for population management. J. Math. Biol. 70, 1015–1063 (2015). https://doi.org/10.1007/s00285-014-0789-4

Download citation

  • Received:

  • Revised:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s00285-014-0789-4

Keywords

Mathematics Subject Classification (2010)

Navigation