Skip to main content
Log in

Sensitivity matrices as keys to local structural system properties of large-scale nonlinear systems

  • Original Paper
  • Published:
Nonlinear Dynamics Aims and scope Submit manuscript

Abstract

Sensitivities are shown to play a key role in a highly efficient algorithm, presented in this paper, to establish three fundamental structural system properties: local structural identifiability, local observability, and local strong accessibility. Sensitivities have the advantageous property to be governed by linear dynamics, also if the system itself is nonlinear. By integrating their linear dynamics over a short time period, and by sampling the result, a sensitivity matrix is obtained. If this sensitivity matrix satisfies a rank condition, then the local structural system property under investigation holds. This rank condition will be referred to in this paper as the sensitivity rank condition (SERC). Applying a singular value decomposition (SVD) to the sensitivity matrix not only determines its rank but also pinpoints exactly the system components causing a possible failure to satisfy the local structural system property. The algorithm is highly efficient because integration of linear systems over short time-periods and computation of an SVD are computationally cheap. Therefore, it allows for the handling of large-scale systems in the order of seconds, as opposed to conventional algorithms that mostly rely on Lie series expansions and a corresponding Lie algebraic rank condition (LARC). The SERC and LARC algorithms are mathematically and computationally compared through a series of 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

Similar content being viewed by others

References

  1. Stigter, J.D., Molenaar, J.: A fast algorithm to assess local structural identifiability. Automatica 58, 118–124 (2015)

    Article  MathSciNet  Google Scholar 

  2. Stigter, J.D., Beck, M.B., Molenaar, J.: Assessing local structural identifiability for environmental models. Environ. Model. Softw. 93, 398–408 (2017)

    Article  Google Scholar 

  3. Stigter, J.D., Joubert, D., Molenaar, J.: Observability of complex systems: finding the gap. Sci. Rep. 7, 16566 (2017). https://doi.org/10.1038/s41598-017-16682-x

    Article  Google Scholar 

  4. Joubert, D., Stigter, J.D., Molenaar, J.: An efficient procedure to assist in the re-parametrization of structurally unidentifiable models. Math. Biosci. 323, 108328 (2020)

    Article  MathSciNet  Google Scholar 

  5. Hermann, R., Krener, A.J.: Nonlinear controllability and observability”. IEEE Trans. Aut. Contr. 22(5), 728–740 (1977)

    Article  MathSciNet  Google Scholar 

  6. Isidori, A.: Nonlinear control systems. Springer, Heidelberg (1989)

    Book  Google Scholar 

  7. Nijmeijer, H., Van der Schaft, A.: Nonlinear dynamical control systems, vol. 175. Springer, New York (1990)

    Book  Google Scholar 

  8. Kwatny, H.G., Blankenship, G.L.: Nonlinear control and analytical mechanics: a computational approach. Birkhäuser, Boston (2000)

    Book  Google Scholar 

  9. Tunali, T., Tarn, T.J.: New results for identifiability of nonlinear systems. IEEE Trans. Autom. Control 32(2), 146–154 (1987)

    Article  MathSciNet  Google Scholar 

  10. Maes, K., Chatzis, M.N., Lombaert, G.: Observability of nonlinear systems with unmeasured inputs. Mech. Syst. Signal Process. 130, 378–394 (2019). https://doi.org/10.1016/j.ymssp.2019.05.010

    Article  Google Scholar 

  11. Yang-Yu Liu, J.J., Slotine, A.L.B.: Controllability of complex networks. Nature 473, 167–173 (2011)

    Article  Google Scholar 

  12. Cowan, N.J., Chastain, E.J., Vilhena, D.A., Freudenberg, J.S., Bergstrom, C.T.: Nodal dynamics, not degree distributions, determine the structural controllability of complex networks. PLoS ONE 7(6), e38398 (2012)

    Article  Google Scholar 

  13. Miao, H., Xia, X., Perelson, A.S., Wu, H.: On identifiability of nonlinear ODE models and applications in viral dynamics. SIAM Rev. 53(1), 3–39 (2011)

    Article  MathSciNet  Google Scholar 

  14. Wieland, F.G., Hauber, A.L., Rosenblatt, M., Tönsing, C., Timmer, J.: On structural and practical identifiability. Curr. Opin. Syst. Biol. 25, 60–69 (2021)

    Article  Google Scholar 

  15. Chappell, M.J., Godfrey, K.R., Vajda, S.: Global identifiability of the parameters of nonlinear systems with specified inputs: a comparison of methods. Math. Biosci. 102(1), 41–73 (1990). https://doi.org/10.1016/0025-5564(90)90055-4

    Article  MATH  Google Scholar 

  16. Chis, O.-T., Banga, J.R., Balsa-Canto, E.: Structural identifiability of systems biology models: a critical comparison of methods. PLoS ONE 6(11), e27755 (2011). https://doi.org/10.1371/journal.pone.0027755

    Article  Google Scholar 

  17. Mir, I., Taha, H., Eisa, S.A., Maqsood, A.: A controllability perspective of dynamic soaring. Nonlinear Dyn. 94(4), 2347–2362 (2018). https://doi.org/10.1007/s11071-018-4493-6

    Article  MATH  Google Scholar 

  18. Van Willigenburg, L.G., Stigter, J.D., Molenaar, J.: Establishing local strong accessibility of large-scale nonlinear systems by replacing the lie algebraic rank condition. In: Proceedings European Control Conference (ECC), Rotterdam, The Netherlands, pp. 2645–2650, June 29 - July 2 (2021)

  19. Kawsky, A.: On the problem whether controllability is finitely determined. In: Proceedings of the 17th International Symposium on Mathematical Theory of Networks and Systems, Kyoto, Japan, July 24–28 (2006)

  20. Stigter, J.D., Van Willigenburg, L.G., Molenaar, J.: An Efficient method to assess local controllability and observability for non-linear systems. In: Preprints of the 9th Vienna International Conference on Mathematical Modelling, Vienna, Austria, February 21–23 (2018)

  21. Hassan, A.M., Taha, H.E.: Geometric control formulation and nonlinear controllability of airplane flight dynamics. Nonlinear Dyn. 88(4), 2651–2669 (2017). https://doi.org/10.1007/s11071-017-3401-9

    Article  MathSciNet  Google Scholar 

  22. Liu, X., Gao, J., Wang, G., Chen, Z.-W.: Controllability analysis of the neural mass model with dynamic parameters. Neural Comput. 29(2), 485–501 (2017). https://doi.org/10.1162/NECO_a_00925

    Article  MathSciNet  MATH  Google Scholar 

  23. Cobelli, C., DiStefano, J.: Parameter and structural identifiability concepts and ambiguities: a critical review and analysis. Am. J. Physiol. Regul. Integr. Comp. Physiol. 239, 7–24 (1980)

    Article  Google Scholar 

  24. Weiss, L., Kalman, R.E.: Contributions to linear system theory. Int. J. Eng. Sci. 3, 141–171 (1965)

    Article  MathSciNet  Google Scholar 

  25. Van Willigenburg, L.G., De Koning, W.L.: Linear systems theory revisited. Automatica 44, 1669–1683 (2008)

    MathSciNet  MATH  Google Scholar 

  26. Kwakernaak, H., Sivan, R.: Linear optimal control systems. Wiley, New York (1972)

    MATH  Google Scholar 

  27. Athans, M.: The role and use of the Linear-Quadratic-Gaussian problem in control system design. IEEE Trans. Aut. Contr. 16(6), 529–552 (1971)

    Article  MathSciNet  Google Scholar 

  28. Neidinger, R.D.: Introduction to automatic differentiation and MATLAB object-oriented programming. SIAM Rev. 52(3), 545–563 (2010)

    Article  MathSciNet  Google Scholar 

  29. Grancharova, A., Johanson, T.A.: Explicit nonlinear model predictive control: theory and applications. Springer, New York (2012)

    Book  Google Scholar 

  30. Martins, J., Sturdza, J. P., Alonso, J.: The connection between the complex-step derivative approximation and algorithmic differentiation. American Institute of Aeronautics and Astronautics (2001)

  31. Saccomani, M.P., Audoly, S., D’Angi, L.: Parameter identifiability of nonlinear systems: the role of initial conditions. Automatica 39, 619–632 (2003)

    Article  MathSciNet  Google Scholar 

  32. McMickell, M.B., Goodwine, B.: Reduction and non-linear controllability of symmetric distributed systems. Int. J. Control 76(18), 1809–1822 (2003)

    Article  MathSciNet  Google Scholar 

  33. Permana, V., Shoureshi, R.: Controllability and observability of a large-scale thermodynamical system via connectability approach. In: ASME 2010 Dynamic Systems and Control Conference, Cambridge, Massachusetts, USA, September 12–15, Vol. 2, 217–224 (2010)

  34. Permana, V.: Controllability and observability of a large-scale thermodynamical system via connectability approach. Electronic Theses and Dissertations. Paper 506, University of Denver (2010)

  35. Stigter, J.D., Joubert, D., van Willigenburg, L.G., Molenaar, J.: A note on the accurate computation of structural properties for dynamic control systems. Submitted to MATHMOD 2022, 16–18 February 2022, Vienna (2021)

  36. Structural vs practical identifiability of nonlinear differential equation models in systems biology. Dynamics of Mathematical Models in Biology. Springer Switzerland (2016) doi: https://doi.org/10.1007/978-3-319-45723-9_3.

  37. Sontag, E.: Universal nonsingular controls. Systems Control Lett. 19, 221–224 (1992)

    Article  MathSciNet  Google Scholar 

  38. Villaverde, A.F., Evans, N.D., Chappell, M.J., Banga, J.R.: Sufficiently exciting inputs for structurally identifiable systems biology models. IFAC-Papers OnLine 19, 16–19 (2018)

    Article  Google Scholar 

  39. Sussmann, H.J.: Single-input observability of continuous time systems. Math. Syst. Theory 12, 371–393 (1979)

    Article  MathSciNet  Google Scholar 

Download references

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to L. G. Van Willigenburg.

Ethics declarations

Conflict of interest

The authors have no relevant financial or non-financial interests to disclose. Data will be made available on reasonable request.

Additional information

Publisher's Note

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

Appendices

Appendix 1: Dynamic model of the combined heat and power system of a hotel building [33, 34] used in Example 5

$$ f\left( {x,u,\theta } \right) = \left[ {\begin{array}{*{20}c} {k_{1} \left( {x_{2} - x_{1} } \right) + \left( {\frac{{b_{1} }}{{V_{1} }} - \frac{{x_{1} }}{{V_{1} }}} \right)u_{1} } \\ {k_{2} \left( {x_{1} - x_{2} } \right) + \left( {\frac{{x_{4} }}{{V_{2} }} - \frac{{x_{2} }}{{V_{2} }}} \right)u_{2} } \\ {\left( {\frac{{x_{2} }}{{V_{3} }} - \frac{{x_{3} }}{{V_{3} }}} \right)u_{2} + \left( {\frac{{x_{4} }}{{V_{3} }} - \frac{{x_{3} }}{{V_{3} }}} \right)u_{3} } \\ {\left( {\frac{{x_{6} }}{{V_{4} }} - \frac{{x_{4} }}{{V_{4} }}} \right)u_{4} } \\ {\left( {\frac{{x_{3} }}{{V_{5} }} - \frac{{x_{5} }}{{V_{5} }}} \right)u_{4} } \\ {\left( {\frac{{x_{7} }}{{V_{6} }} - \frac{{x_{6} }}{{V_{6} }}} \right)u_{7} + \left( {\frac{{x_{9} }}{{V_{6} }} - \frac{{x_{6} }}{{V_{6} }}} \right)u_{9} + \left( {\frac{{x_{12} }}{{V_{6} }} - \frac{{x_{6} }}{{V_{6} }}} \right)u_{5} } \\ {k_{7} \left( {x_{8} - x_{7} } \right) + \left( {\frac{{x_{5} }}{{V_{7} }} - \frac{{x_{7} }}{{V_{7} }}} \right)u_{7} } \\ {k_{8} \left( {x_{7} - x_{8} } \right) + \left( {\frac{{b_{2} }}{{V_{8} }} - \frac{{x_{8} }}{{V_{8} }}} \right)u_{8} } \\ {k_{9} \left( {x_{10} - x_{9} } \right) + \left( {\frac{{x_{5} }}{{V_{9} }} - \frac{{x_{9} }}{{V_{9} }}} \right)u_{9} } \\ {k_{10} \left( {x_{9} - x_{10} } \right) + \left( {\frac{{b_{3} }}{{V_{10} }} - \frac{{x_{10} }}{{V_{10} }}} \right)u_{10} } \\ {\left( {\frac{{x_{5} }}{{V_{11} }} - \frac{{x_{11} }}{{V_{11} }}} \right)u_{5} + \left( {\frac{{x_{12} }}{{V_{11} }} - \frac{{x_{11} }}{{V_{11} }}} \right)u_{12} } \\ {\left( {\frac{{x_{11} }}{{V_{12} }} - \frac{{x_{12} }}{{V_{12} }}} \right)u_{11} } \\ {k_{13} \left( {x_{14} - x_{13} } \right) + \left( {\frac{{x_{12} }}{{V_{13} }} - \frac{{x_{13} }}{{V_{13} }}} \right)u_{11} } \\ {k_{14} \left( {x_{13} - x_{14} } \right) + \left( {\frac{{b_{4} }}{{V_{14} }} - \frac{{x_{14} }}{{V_{14} }}} \right)u_{6} } \\ \end{array} } \right]\,,\, $$
(A1)
$$ \theta = \left[ {V_{1} ,\,V_{2} ,..,V_{14} ,\,\,k_{1} ,k_{2} ,k_{7} ,k_{8} ,k_{9} ,k_{10} ,k_{13} ,k_{14} ,b_{1} ,b_{2} ,b_{3} ,b_{4} } \right]^{T} $$
(A2)

Appendix 2: Proof of generic equivalence between SERC and LARC for local strong accessibility and local observability

2.1 Proof for local strong accessibility

Concerning local strong accessibility, the examples in this paper show that the rank given by SERC equals the one given by LARC, provided the selected trajectory (2.3) of system (2.1) is nonsingular, as shown in Fig. 12. For local strong accessible systems (2.1), it is proved in [37] that trajectories (2.3) of system (2.1) are generically nonsingular. Then, LARC gives full rank for \(x_{L}\) in Fig. 12 while SERC generically does, namely if the selected trajectory (2.3) of system (2.1) containing \(x_{L}\) is nonsingular. In this appendix, we will prove that if the rank obtained from LARC is not full, SERC still generically produces the same rank for \(x_{L}\) in Fig. 12. To that end, we will first summarize the proof given in [37], when the rank obtained from LARC is full.

Theorem 2 in [37] states that for local strong accessible systems, trajectories are generically nonsingular. Its proof relies mostly on proposition 2.1 in [37]. This proposition concerns an observability-related property developed in [39]. Recall that LARC for local strong accessibility is fully determined by the dynamics (2.1). The proof of theorem 2 in [37] requires extension of these dynamics with

$$ \frac{{{\text{d}}Q\left( t \right)}}{{{\text{d}}t}} = A\left( t \right)Q\left( t \right) + Q\left( t \right)A\left( t \right) + B\left( t \right)B^{T} \left( t \right),\,\,\,t \ge t_{0} ,\,\,Q\left( {t_{0} } \right) = 0 $$
(A3)
$$ A\left( t \right) = \frac{{{\text{d}}f}}{{{\text{d}}x}}\left( {x\left( t \right),u\left( t \right)} \right),\,\,B\left( t \right) = \frac{{{\text{d}}f}}{{{\text{d}}u}}\left( {x\left( t \right),u\left( t \right)} \right),\,\,Q\left( t \right) \in R^{n \times n} $$
(A4)

In addition, the following output equation is required:

$$ y_{e} \left( t \right) = h_{e} \left( {x\left( t \right),Q\left( t \right)} \right) = \det \left( {Q\left( t \right)} \right). $$
(A5)

Note that, \(Q\left( t \right)\), \(t > t_{0}\) in (A3) is the controllability matrix of the time-varying linear system (2.18) that describes the propagation of sensitivities. If \(Q\left( t \right)\), \(t > t_{0}\) is not full rank, the linear dynamics (2.18) is not controllable and the trajectory is singular. In that case, the determinant of \(Q\left( t \right)\), \(t > t_{0}\), being the output \(y_{e} \left( t \right)\) in (A5), equals zero. System (2.1), extended with (A3), (A5), when written in the standard state-space format becomes

$$ \begin{aligned} \frac{{{\text{d}}x_{e} \left( t \right)}}{{{\text{d}}t}} = &\, f_{e} \left( {x_{e} \left( t \right),u\left( t \right),\theta } \right) = \left[ {\begin{array}{*{20}c} {f\left( {x\left( t \right),u\left( t \right),\theta } \right)} \\ {st\left( {A\left( t \right)Q\left( t \right) + Q\left( t \right)A\left( t \right) + B\left( t \right)B^{T} \left( t \right)} \right)} \\ \end{array} } \right], \\ x_{e} \left( t \right) = & \left[ {\begin{array}{*{20}c} {x\left( t \right)} \\ {st\left( {Q\left( t \right)} \right)} \\ \end{array} } \right], \\ \end{aligned} $$
(A6)
$$ y_{e} \left( t \right) = h_{e} \left( {x_{e} \left( t \right)} \right) = \det \left( {Q\left( t \right)} \right), $$
(A7)

where \(st\) denotes the stack operator that stacks all columns of the matrix argument, starting with the first one. The proof of theorem 2 in [37] comes down to show, by means of proposition 2.1, that trajectories having initial state \(x_{e} = \left[ {\begin{array}{*{20}c} x \\ 0 \\ \end{array} } \right]\), generically, do not produce a zero output \(y_{e} \left( t \right)\) in (A5), and therefore are nonsingular. To obtain this result, local strong accessibility of system (2.1) is required to guarantee that every state \(x_{e} = \left[ {\begin{array}{*{20}c} x \\ 0 \\ \end{array} } \right]\) of the extended system belongs to the set \(G\), mentioned in proposition 2.1 of [37].

Here, we would like to establish generic equivalence for any rank given by LARC, for any \(x_{L}\) in Fig. 12. This can be achieved since \(f\) in (2.1) is assumed analytic. Then, the Hermann Nagano theorem [5] holds, according to which the state space separates into one or more manifolds having equal or different dimensions. Then by Chow’s theorem [5], a realization of each manifold exists, that is local strong accessible, having a state dimension equal to the rank given by LARC. Given their local strong accessibility, we are allowed again to apply the proof given in [37], to these realizations. This proves the generic equivalence between SERC and LARC for any rank, and for any state \(x_{L}\) in Fig. 12.

2.2 Proof for local observability

For local observability, to show that the rank from LARC generically equals the one from SERC, we can again apply proposition 2.1 from [37]. Now, LARC is obtained from Lie differentiation instead of Lie bracketing [5], [7], [8] and nonsingular trajectories are those leading to an observable linear system (2.14). Therefore, we must extend the dynamics (2.1) with dynamics describing the observability matrix propagation of the time-varying linear system (2.14),

$$ \frac{{{\text{d}}Q\left( t \right)}}{{{\text{d}}t}} = A^{T} \left( t \right)Q\left( t \right) + Q\left( t \right)A^{T} \left( t \right) + C^{T} \left( t \right)C\left( t \right),\,\,\,t \ge t_{0} ,\,\,Q\left( {t_{0} } \right) = 0, $$
(A8)
$$ A\left( t \right) = \frac{{{\text{d}}f}}{{{\text{d}}x}}\left( {x\left( t \right),u\left( t \right)} \right),\,\,C\left( t \right) = \frac{{{\text{d}}h}}{{{\text{d}}x}}\left( {x\left( t \right),u\left( t \right)} \right),\,\,Q\left( t \right) \in R^{n \times n} , $$
(A9)

while taking as the output,

$$ y_{e} \left( t \right) = h\left( {x\left( t \right),Q\left( t \right)} \right) = \det \left( {Q\left( t \right)} \right). $$
(A10)

If \(Q\left( t \right)\), \(t > t_{0}\) is not full rank, the linear dynamics (2.14) are not observable and the trajectory is singular. In that case, the determinant of \(Q\left( t \right)\), \(t > t_{0}\), being the output \(y_{e} \left( t \right)\) in (A10), equals zero. The state-space representation of the extended system now becomes

$$ \begin{aligned} \frac{{{\text{d}}x_{e} \left( t \right)}}{{{\text{d}}t}} = &\, f_{e} \left( {x_{e} \left( t \right),u\left( t \right),\theta } \right) = \left[ {\begin{array}{*{20}c} {f\left( {x\left( t \right),u\left( t \right),\theta } \right)} \\ {st\left( {A^{T} \left( t \right)Q\left( t \right) + Q\left( t \right)A^{T} \left( t \right) + C^{T} \left( t \right)C\left( t \right)} \right)} \\ \end{array} } \right], \\ x_{e} \left( t \right) = &\, \left[ {\begin{array}{*{20}c} {x\left( t \right)} \\ {st\left( {Q\left( t \right)} \right)} \\ \end{array} } \right], \\ \end{aligned} $$
(A11)
$$ y_{e} \left( t \right) = h_{e} \left( {x_{e} \left( t \right)} \right) = \det \left( {Q\left( t \right)} \right). $$
(A12)

Again the proof follows from proposition 2.1 of [37] if all states \(x_{e} = \left[ {\begin{array}{*{20}c} x \\ 0 \\ \end{array} } \right]\) of the extended system belong to the set \(G\), mentioned in proposition 2.1. Now, this requires the original system (2.1), (2.2) to be local observable. That this is so follows from application of the same proposition 2.1, now taking (2.1) as the system and (2.2) as the output. So if LARC gives full rank for any \(x_{L}\) in Fig. 12, SERC generically does, namely if the selected trajectory (2.3) of system (2.1) is nonsingular.

To establish generic equivalence between LARC and SERC for any rank, we rely on our assumption that \(f\) in (2.1) and \(h\) in (2.2) are analytic. Then, the Hermann Nagano and Chow theorems [5] provide us with local accessible realizations of each manifold in which the state space separates. Moreover, these realizations each preserve the observation space, the dimension of which equals the rank obtained from LARC. Next, for each of those realizations, we may obtain a minimal realization that also preserves the observation space [7]. This minimal realization is locally observable and has a state dimension equal to the one of the observation space given by the rank obtained from LARC. Because these minimal realizations are locally observable, we may apply proposition 2.1 of [37] again, which proves the generic equivalence between LARC and SERC for any rank, and any state \(x_{L}\) in Fig. 12.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Van Willigenburg, L.G., Stigter, J.D. & Molenaar, J. Sensitivity matrices as keys to local structural system properties of large-scale nonlinear systems. Nonlinear Dyn 107, 2599–2618 (2022). https://doi.org/10.1007/s11071-021-07125-4

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s11071-021-07125-4

Keywords

Navigation