Skip to main content
Log in

A geometric approach to nonlinear dissipative balanced reduction I: exogenous signals

  • Original Article
  • Published:
Mathematics of Control, Signals, and Systems Aims and scope Submit manuscript

Abstract

This paper is divided into two parts and provides a unified geometric theory for nonlinear dissipative and Hankel balanced reduction with the help of a framework based on differential geometry, dissipativity theory, Lie-semigroups and sub manifold Hilbert theory. Part I presents a theory for the invariants of the behavior using classical Gauss’ curvature theory. Furthermore, known concepts of the theory of balanced reduction like the Hankel operator, Schmidt decomposition, etc., can be understood in a proper and general perspective.

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.

Similar content being viewed by others

References

  1. Benenti S (1992) Orthogonal separable dynamical systems. In: Proceedings of conference on differential geometry and its applications, Opava, Czechoslovakia, pp 163–184

  2. Benenti S (2004) Separability on Riemannian manifolds. Unpublished. http://www2.dm.unito.it/~benenti/

  3. Burghelea D, Kuiper NH (1969) Hilbert manifolds. Ann Math 90:379–417

    Article  MathSciNet  MATH  Google Scholar 

  4. Clément Ph, Heijmans HJAM (1987) One-parameter semigroups. CWI monographs, vol 5. North Holland, Amsterdam

  5. Delaleau E, Respondek W (1995) Lowering the orders of derivatives of controls in generalized state space systems. J Math Syst Estim Control 5(3):375–378

    MathSciNet  MATH  Google Scholar 

  6. Elkin VI (1999) Reduction of nonlinear control systems: a differential geometric approach. In: Mathematics and its applications, vol 472. Kluwer, Dordrecht

  7. Frankel T (2001) The geometry of physics: an introduction. Cambridge University Press, Cambridge

    Google Scholar 

  8. Fujimoto K, Scherpen JMA (Jan. 2005) Nonlinear input-normal realizations based on the differential eigenstructure of Hankel operators. IEEE Trans Autom Control 50(1):2–18

    Google Scholar 

  9. Gray WS, Scherpen JMA (2005) Hankel singular value functions from Schmidt pairs for nonlinear input-output systems. Syst Control Lett 54:135–144

    Article  MathSciNet  MATH  Google Scholar 

  10. Green M, Limebeer DJN (1995) Linear robust control. Prentice Hall, New Jersey

    MATH  Google Scholar 

  11. Hilgert J, Neeb KH (1993) Basic theory of Lie-semigroups and applications. In: Lecture notes in mathematics, vol 1552. Springer, Berlin

  12. Hill DJ, Moylan PJ (May 1980) Dissipative dynamical systems: basic input-output and state properties. J Franklin Inst 309(5):327–357

    Google Scholar 

  13. Hofmann KH, Lawson JD (1983) Foundations of Lie semigroups. In: Hofmann KH, Jürgensen H, Weinert HJ (eds) Recent developments in the algebraic, analytical and topological theory of semigroups. In: Lecture notes in mathematics, vol 998. Springer, Berlin, pp 128–201

  14. Isidori A (1995) Nonlinear control systems, 3rd edn. Springer, Berlin

  15. Lang S (1999) Fundamentals of differential geometry. Springer, New York

    Book  MATH  Google Scholar 

  16. Lee JM (1997) Riemannian manifolds: an introduction to curvature. In: Graduate texts in mathematics, vol 176. Springer, New York

  17. Levi-Civita T (1904) Sulla integrazione della equazione di Hamilton-Jacobi per separazione di variabili. Math Ann 59:383–397

    Article  MathSciNet  MATH  Google Scholar 

  18. Lopezlena R (2006) Invariants of the behavioral operator using classical curvature theory. In: Mems Cong Nacional de Control Automático 2006, Mexico City, Mexico, pp 85–90

  19. Lopezlena R, Scherpen JMA (2006) Energy functions for dissipativity-based balancing of discrete-time nonlinear systems. Math Control Signs Syst 18:345–368

    Article  MathSciNet  MATH  Google Scholar 

  20. Lopezlena R, Scherpen JMA (2006) Nonlinear behavioral balancing by extension of Lie semigroups. In: Proceedings of 3rd IFAC workshop on Lagr. and Ham. Meth. for Nonl. Cont., Nagoya, Japan

  21. Lopezlena R, Scherpen JMA, Fujimoto K (2003) Energy-storage balanced reduction of port-Hamiltonian systems. In: 2nd IFAC workshop on Lagr. and Ham. Meth. for Nonl. Cont., Seville, Spain, pp 79–84

  22. Moore BC (1981) Principal components analysis in linear systems: controllability, observability, and model reduction. IEEE Trans Autom Control 26(1):17–32

    Google Scholar 

  23. Nijmeijer H, van der Schaft AJ (1990) Nonlinear dynamical control systems. Springer, New York

    Book  MATH  Google Scholar 

  24. Olver PJ (1993) Applications of Lie groups to differential equations. In: Graduate texts in mathematics, 2nd edn, vol 107. Springer, New York

  25. Olver PJ (1995) Equivalence, invariants and symmetry. Cambridge University, New York

  26. Palais RS (1963) Morse theory on Hilbert manifolds. Topology 2:299–340

    Article  MathSciNet  MATH  Google Scholar 

  27. Palais RS (1966) Lusternik-Schnirelman theory on Banach manifolds. Topology 5:115–132

    Article  MathSciNet  MATH  Google Scholar 

  28. Palais RS, Smale S (1964) A generalized Morse theory. Bull Am Math Soc 70:165–172

    Article  MathSciNet  MATH  Google Scholar 

  29. Palais RS, Terng C (1988) Critical point theory and submanifold geometry. In: Lecture notes in mathematics, vol 1353. Springer, New York

  30. Pipkin AC (1991) A course on integral equations. In: Texts on application mathematics, vol 9. Springer, New York

  31. Polderman JW, Willems JC (1998) Introduction to mathematical systems theory: a behavioral approach. In: Texts in application mathematics, vol 26. Springer, New York

  32. Roorda B, Weiland S (2001) Optimal angle reduction: a behavioral approach to linear system approximation. Linear Algebra Appl 337:189–235

    Article  MathSciNet  MATH  Google Scholar 

  33. Rugh WJ (1981) Nonlinear system theory: the Volterra/Wiener approach. The Johns Hopkins University, USA

  34. Sussmann HJ (1975) A generalization of the closed subgroup theorem to quotients of arbitrary manifolds. J Differ Geometry 10:151–166

    MathSciNet  MATH  Google Scholar 

  35. Vidyasagar M (1993) Nonlinear systems analysis, 2nd edn. Prentice Hall, New Jersey

  36. Weiland S (1991) Theory of approximation and disturbance attenuation for linear systems. PhD dissertation, Rijksuniversiteit Groningen

  37. Weiland S (1994) A behavioral approach to balanced representations of dynamical systems. Linear Algebra Appl 205–206:1227–1252

    Article  MathSciNet  Google Scholar 

  38. Willems JC (1972) Dissipative dynamical systems. Part 1: general theory. Arch Ration Mech Anal 45:321–351

    Article  MathSciNet  MATH  Google Scholar 

  39. Willems JC (1991) Paradigms and puzzles in the theory of dynamical systems. IEEE Trans Autom Control 36(3):259–294

    Article  MathSciNet  MATH  Google Scholar 

  40. Yosida K (1980) Functional analysis. Grund. der Math Wiss, 123. Springer, Berlin

  41. Young N (1988) An introduction to Hilbert Space. Cambridge University, London

    Book  MATH  Google Scholar 

Download references

Acknowledgments

The author gratefully acknowledges Prof. Dr. S. Weiland, TU/E, for inspiring conversations on linear behavioral balancing theory during the DISCourses of Winter 2001.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Ricardo Lopezlena.

Additional information

Enrolled as promovendus at the Delft Center for Systems and Control, Delft University of Technology, Mekelweg 2, 2628 CD Delft, The Netherlands.

Appendix

Appendix

Proof of Proposition 2.2

In the differential geometric framework, duality of these spaces follows from the known contraction operation \(i_{\xi } \alpha \) between vector fields and differential 1-forms. Since the inner products in Table 1 are positive definite, symmetric, bilinear forms on (the space of vector fields) \(T_p\mathcal W \) (on the space of 1-forms \(T_p ^*\mathcal W \)) such products are admissible. The remaining details can be seen in [26] and in the more general proof of Proposition 5, pg. 55 assuming \(\mathcal W \) an infinite-dimensional Sobolev space (without differential geometry) in [40]. From Definition 2.6, the duality product \(\langle \xi , \alpha \rangle _{T\mathcal W \times T^{*}\mathcal W }\) is well defined since for a \(\phi \in \mathcal W ^*\) the dual of \(d\phi \in T^*\mathcal W \) is given by (the gradient) \(\nabla \phi \in T\mathcal W \) satisfying \(\phi (w)=\langle \nabla \phi , \dot{w} \rangle _{T\mathcal W }\), \(\dot{w} \in T\mathcal W \). Coordinate invariance of \(\langle \xi , \alpha \rangle _{T\mathcal W \times T^*\mathcal W }\) follows from known facts of differential geometry (see e.g. [6]).

Proof of Proposition 2.3

Since by construction the inner products satisfy Assumption 2.1 on dissipativity, semipositive definiteness, symmetry and boundedness are satisfied and such inner products are admissible. The remaining Hilbert manifold properties are inherited to this structure from \((\mathcal W , \langle \cdot , \cdot \rangle _{T\mathcal W })\) and \((\mathcal W ^*, \langle \cdot , \cdot \rangle _{T^*\mathcal W })\). Duality of \(\alpha _{-}\) with \(\xi ^{+}\) is identified by the abstract duality pairing \(\langle \xi ^{+}, \alpha _{-} \rangle _{T\mathfrak B ^{-} \times T\mathfrak B ^{+}}\) for an isometry \(\tilde{\Gamma }\) s.t. \(\xi ^{+}=\tilde{\Gamma }^* \alpha _{-}\), with \(\xi ^{+}\) dual to \(\alpha _{+}\), which according to Definition 2.6 must satisfy \(\langle \alpha _{-}, \alpha _{-} \rangle _{T\mathfrak B ^-}=\langle \xi ^{+}, \alpha _{-} \rangle _{T\mathfrak B ^{-} \times T\mathfrak B ^{+}}=\langle \xi ^{+}, \tilde{\Gamma }^* \,\alpha _{-} \rangle _{T\mathfrak B ^+}\), which is equivalent to \(\Vert \alpha _{-}\Vert _{T\mathfrak B ^-}=\Vert \xi ^{+}\Vert _{T\mathfrak B ^+}\).

Proof of Proposition 2.4

1) This result is a consequence of the Definition 2.6 for the duality pairing \(\langle \xi ^{+}, \alpha _{-} \rangle _{T\mathfrak B ^{-} \times T\mathfrak B ^{+}}\). By Definition 2.6, duality of \(\alpha _{-}\) with \(\xi ^{+}\) requires the preservation of the relationship

$$\begin{aligned} \langle \alpha _{-}, \alpha _{-} \rangle _{T\mathfrak B ^-}\!=\!\langle \xi ^{+}, \alpha _{-} \rangle _{T\mathfrak B ^{-} \times T\mathfrak B ^{+}}\!=\!\langle \xi ^{+}, \tilde{\Gamma }^* \,\alpha _{-} \rangle _{T\mathfrak B ^+},\quad \forall \alpha _{-} \in \mathfrak B ^-,\, \xi ^{+} \in \mathfrak B ^+.\nonumber \\ \end{aligned}$$
(41)

for inner products defined as in Proposition 2.3, for an isometric isomorphism \(\tilde{\Gamma }:\mathfrak B ^- \rightarrow \mathfrak B ^+\) defined by Eq. (3), which from Definition 2.5 is seen to satisfy commutativity in diagram (15) with the following structure at \(T\mathfrak B \):

$$\begin{aligned} \left[\begin{array}{c} \xi _{u} ^{+}(w)\\ \\ \xi _{y}^{+} (w) \end{array} \right]_{T\mathfrak B ^{+}}= \left[\begin{array}{c c} 0&\Gamma ^{\dagger }(w) |_* \\ \Gamma (w)|_*&0 \end{array} \right] \left[\begin{array}{c} \xi _{u}^{-} (w)\\ \xi _{y}^{-} (w) \end{array} \right], \end{aligned}$$
(42)

where the map \(\Gamma : \mathcal U \rightarrow \mathcal Y \) has an adjoint map \(\Gamma ^{\dagger }: \mathcal Y ^* \rightarrow \mathcal U ^*\).

2) Define by \(\tilde{\Gamma }^\dagger :\mathfrak B ^+ \rightarrow \mathfrak B ^-\) an associated operator to the behavioral operator in Definition 2.3 with the following structure at \(T\mathfrak B \):

$$\begin{aligned} \left[\begin{array}{l} \alpha _{-} ^{\hat{u}}(\hat{w})\\ \\ \alpha _{-} ^{\hat{y}} (\hat{w}) \end{array} \right]_{T\mathfrak B ^{-}}= \left[\begin{array}{ll} 0&\Gamma ^{\dagger }(\hat{w}) |^* \\ \Gamma (\hat{w})|^*&0 \end{array} \right] \left[\begin{array}{l} \alpha _{+} ^{\hat{u}} (\hat{w})\\ \alpha _{+} ^{\hat{y}} (\hat{w}) \end{array} \right], \end{aligned}$$
(43)

such operator is such that \(\langle \xi , \Gamma _*\zeta \rangle _{T\mathfrak B ^-}=\langle \Gamma _*\xi , \zeta \rangle _{T\mathfrak B ^+}\), \( \xi ,\zeta \in T\mathfrak B ^+\) satisfying Definition 2.12 and therefore is selfadjoint. Since \(\tilde{\Gamma }\) is isometric, it satisfies \(\langle \xi ^{-},\zeta ^{-} \rangle _{T\mathfrak B ^-}=\langle \tilde{\Gamma }_* \xi ^{-}, \tilde{\Gamma }_* \zeta ^{-} \rangle _{T\mathfrak B ^+}\equiv \langle \xi ^{+}, \zeta ^{+} \rangle _{T\mathfrak B ^+}\), \(\xi ^{-},\zeta ^{-} \in T\mathfrak B ^-\), \(\xi ^{+},\zeta ^{+} \in T\mathfrak B ^+\). Let the map \(\tilde{\Gamma }^\dagger :\mathfrak B ^{+} \rightarrow \mathfrak B ^{-}\) satisfy \(\langle \xi ^{+},\zeta ^{+} \rangle _{T\mathfrak B ^+}\mathop {=}\limits ^\mathrm{def}\langle \tilde{\Gamma }^\dagger _* \xi ^{+}, \tilde{\Gamma }^\dagger _* \zeta ^{+} \rangle _{T\mathfrak B ^-}\) hence isometric, with \(\xi ^{-} \mathop {=}\limits ^\mathrm{def}\tilde{\Gamma }^\dagger _* \xi ^{+}\). Then Definition 2.11 is verified since \(\langle \tilde{\Gamma }^\dagger _* \xi ^{+}, \zeta ^{-} \rangle _{T\mathfrak B ^-}=\langle \xi ^{+}, \tilde{\Gamma }_* \zeta ^{-} \rangle _{T\mathfrak B ^+}\).

3) Since we may write \(\langle \xi ^{-}, \zeta ^{-} \rangle _{T\mathfrak B ^-}=\langle (\tilde{\Gamma }^\dagger \circ \tilde{\Gamma })_* \xi ^{-}, \zeta ^{-} \rangle _{T\mathfrak B ^-}=\langle \xi ^{-}, (\tilde{\Gamma }^\dagger \circ \tilde{\Gamma })_* \zeta ^{-} \rangle _{T\mathfrak B ^-}\) and \(\langle \xi ^{+}, \zeta ^{+} \rangle _{T\mathfrak B ^+}=\langle \xi ^{+}, (\tilde{\Gamma }\circ \tilde{\Gamma }^\dagger )_* \zeta ^{+} \rangle _{T\mathfrak B ^+}=\langle (\tilde{\Gamma }\circ \tilde{\Gamma }^\dagger )_* \xi ^{+}, \zeta ^{+} \rangle _{T\mathfrak B ^+}\), the operators \(\tilde{\Gamma }^\dagger \circ \tilde{\Gamma }(\cdot )\) and \(\tilde{\Gamma }\circ \tilde{\Gamma }^\dagger (\cdot )\) are self-adjoint on \(\mathfrak B ^-\) and \(\mathfrak B ^+\), respectively, satisfying Definition 2.12.

4) Direct comparison of operators \(\tilde{\Gamma }\) and \(\tilde{\Gamma }^\dagger \) in Eqs. (3) and (16) shows that both are equivalent and their tangent maps Eqs. (42) and (43) satisfy Definition 2.12.

5) Denote by \(\mathcal W |_\mathcal{U },\mathcal W |_\mathcal{Y }\) the restriction of a manifold \(\mathcal W \supset \mathcal U \oplus \mathcal Y \) by their disjoint subsets and consider the restricted maps \(\Gamma : \mathfrak B ^-|_{u} \rightarrow \mathfrak B ^+|_{y}\), \(\Gamma ^\dagger : \mathfrak B ^-|_{y} \rightarrow \mathfrak B ^+|_{u}\). Since we may write \(\langle \xi ^{-}, \zeta ^{-} \rangle _{T\mathfrak B ^-|_{u}}=\langle (\Gamma ^\dagger \circ \Gamma )_* \xi ^{-}, \zeta ^{-} \rangle _{T\mathfrak B ^-|_{u}}=\langle \xi ^{-}, ( \Gamma ^\dagger \circ \Gamma )_* \zeta ^{-} \rangle _{T\mathfrak B ^-|_{u}}\) and \(\langle \xi ^{+}, \zeta ^{+} \rangle _{T\mathfrak B ^+|_{y}}=\langle \xi ^{+}, (\Gamma \circ \Gamma ^\dagger )_* \zeta ^{+} \rangle _{T\mathfrak B ^+|_{y}}=\langle (\Gamma \circ \Gamma ^\dagger )_* \xi ^{+}, \zeta ^{+} \rangle _{T\mathfrak B ^+|_{y}}\), the operators \(\Gamma ^\dagger \circ \Gamma (\cdot )\) and \(\Gamma \circ \Gamma ^\dagger (\cdot )\) are self-adjoint on \(\mathfrak B ^-|_{u}\) and \(\mathfrak B ^+|_{y}\), respectively, satisfying Definition 2.12. \(\square \)

Proof of Proposition 2.6

1) Consider the map \(\Upsilon ^t: \mathbb R ^1 \times \mathcal W \rightarrow \mathcal W \) for the first eigenproblem \(\Upsilon (\varrho (t))=\lambda \varrho (t)\) and some eigenvalue \(\lambda \) along the path \(\varrho (t) \in \mathcal W \). Such path is the integral trajectory of a vector field satisfying \(\frac{d\varrho (t)}{dt}= \xi (\varrho (t))\). On the other side, time derivation at both sides of the equation defining the eigenvalue problem yields \(\frac{\partial \Upsilon }{\partial x}\, \frac{d\varrho (t)}{dt} =\lambda \frac{d\varrho (t)}{dt}\) which is precisely \(\Upsilon _*|_x\,\xi =\lambda \xi \) (a linear eigenvalue problem) and then the problem is well defined. The inverse implication is straightforward. Since a generator \(\xi \) of \(\Upsilon ^t\) acts on differentiable functions as \(\xi f=\frac{\mathrm{d}}{\mathrm{d}t} f(\Upsilon ^t(x))|_{t=0}\) a similar reasoning follows for the proof of Proposition 2.6(2). \(\square \)

Proof of Proposition 2.7

1) Evident from the differentiable structure of \(\tilde{\Gamma }\).

2) Express the map \(\tilde{\Gamma }_*\) as a matrix with components \(G_i ^j=[\partial _{w^j} \tilde{\Gamma }_{i}]\) and rewrite the eigenproblem (18) for each entry as \(\hat{w}_i ^{j}(0)+ \int _{0} ^{\infty } \sum _{j} G_i ^j \beta _{j} ^i(w^{-}(t)) \,\mathrm{d}\mu (\tau )\) which defines a system of \(1\)st-kind Fredholm equations (for generic definitions see [30]).

3) Using Proposition 2.6(1) the eigenvalue problem in terms of \(T\mathfrak B \) with tangent map \(\Gamma _*:T\mathfrak B \rightarrow T\mathfrak B \) yields the eigenproblem presented above with tangent eigenvectorfield \(\xi _{i} \in T\mathfrak B \).

Proof of Theorem 3.1

1) Since by Assumption 2.3, \(\mathfrak g _\mathfrak{B } ^{-}(w^{-},0) \mathop {=}\limits ^\mathrm{def}S_r ^*(\hat{w}^0,r_r)=\langle \alpha _{-} , \alpha _{-} \rangle _{T\mathfrak B ^-}\), \(\mathfrak g _\mathfrak{B } ^{+}(w^{+},0) \mathop {=}\limits ^\mathrm{def}S_a(w^0,r_a) =\langle \xi ^{+}, \xi ^{+} \rangle _{T\mathfrak B ^+ }\) which by the isometric isomorphism \(\xi ^{+}=\tilde{\Gamma }^* \alpha _{-} \) transforms the future metrics in terms of the past metrics by \(\langle \xi ^{+}, \xi ^{+} \rangle _{T\mathfrak B ^+ }=\langle \tilde{\Gamma }^* \alpha _{-} , \tilde{\Gamma }^* \alpha _{-} \rangle _{T\mathfrak B ^+ }=\langle \tilde{\Gamma }_* ^\dagger \tilde{\Gamma }^*\alpha _{-} , \alpha _{-} \rangle _{T\mathfrak B ^-}\), and therefore \(A_\eta ^\mathfrak B =\tilde{\Gamma }_* ^\dagger \tilde{\Gamma }^*\).

2) Since the quotient (23) is such that \(K(\varepsilon \alpha _{-} )=K(\alpha _{-} )\) for any scalar \(0 \ne \varepsilon \in \mathbb R ^1\) and \(\alpha _{-} \in T_p\mathfrak B ^{-}\), this implies that any tangent 1-form candidate solution to the eigenvalue problem satisfies \(\langle \alpha _{-} , \alpha _{-} \rangle _{T\mathfrak B ^{-}}=1\). Therefore, denote by \(\beta _{-}\) the elements of a sphere \(\mathcal S \subset T\mathfrak B ^{-}\) defined by \(\mathcal S =\{\beta _{-} \in T\mathfrak B ^{-} | \langle \beta _{-}, \beta _{-} \rangle _{T\mathfrak B ^{-}}=1\}\). By the Stone-Weierstrass Theorem (e.g. [40]) any continuous function supported on \(\mathcal S \) (being a compact subset) attains in there its maximum and minimum. Express the numerator of (23) by \(\phi :\mathcal S \rightarrow \mathbb R ^1\), i.e., \(\phi (\beta _{-})\mathop {=}\limits ^\mathrm{def}II_\mathfrak{B }(\beta _{-}, \beta _{-})=\langle A_\eta ^\mathfrak B \beta _{-}, \beta _{-} \rangle _{T\mathfrak B ^{-}}\). A point \(\beta _{-} ^0 \in T\mathfrak B ^{-}\) is a critical point in \(\mathcal S \) if for any curve \(\varrho :[-1,1] \rightarrow \mathcal S \), \(\varrho (0)=\beta _{-} ^0\), satisfies \(d \phi (\varrho )/d\tau |_{\tau =0}=0\). Performing this operation yields,

$$\begin{aligned} \left. \frac{\mathrm{d}}{\mathrm{d}\tau } \langle A_\eta ^\mathfrak B \varrho (\tau ), \varrho (\tau ) \rangle _{T\mathfrak B ^{-}} \right|_{\tau =0}&= \langle A_\eta ^\mathfrak B \dot{\varrho }(0), \varrho (0) \rangle _{T\mathfrak B ^{-}}+\langle A_\eta ^\mathfrak B \varrho (0), \dot{\varrho }(0) \rangle _{T\mathfrak B ^{-}} \\&=\langle A_\eta ^\mathfrak B \dot{\varrho }(0), \beta _{-} ^0 \rangle _{T\mathfrak B ^{-}}+\langle A_\eta ^\mathfrak B \beta _{-} ^0, \dot{\varrho }(0) \rangle _{T\mathfrak B ^{-}}\\&= 2 \langle A_\eta ^\mathfrak B \beta _{-} ^0, \dot{\varrho }(0) \rangle _{T\mathfrak B ^{-}} \end{aligned}$$

being the last equality due to selfadjointness of \(A_\eta ^\mathfrak B \). This result implies that \(\beta _{-} ^0\) is a critical point of \(\phi (\beta _{-})\) on \(\mathcal S \) if \(A_\eta ^\mathfrak B \beta _{-} ^0\) is orthogonal to any tangent vector field \(\dot{\varrho }(0) \in T\mathcal S \). Due to the geometry of the sphere \(\mathcal S \), any tangent vector field on \(T\mathcal S \) is orthogonal to a (possibly scaled) normal vector field on \(\mathcal S \). Thus if \(\beta _{-} ^0\) is an eigenform of \(A_\eta ^\mathfrak B \) there is a \(\kappa \) scaling \(\beta _{-} ^0\) normal in \(\mathcal S \), implying \(\beta _{-}\) is a critical point of \(\phi \), concluding the proof.

3) By induction. Let \(\mathrm{dim} \mathcal W =\omega \) for \(\omega =1\) the result is trivial. Assume it is true for \(\omega =k\). Let \(\omega =k+1\); from Theorem 3.1(2), there exists at least a unitary 1-form \(\beta _{-} ^1 \in T_p \mathfrak B ^{-}\) of \(A_\eta ^\mathfrak B \). By Remark 2.5, define locally an orthogonal complementary codistribution by \(W_1 ^*=\{\alpha _{-} \,|\, \langle \beta _{-} ^1, \alpha _{-} \rangle _{T\mathfrak B ^{-}}=0,\, \alpha _{-} \in T_p\mathfrak B ^{-}\}\) then since \(\langle A_\eta ^\mathfrak B \alpha _{-},\beta _{-} ^1 \rangle _{T\mathfrak B ^{-}}=\langle \alpha _{-}, A_\eta ^\mathfrak B \beta _{-} ^1 \rangle _{T\mathfrak B ^{-}}=\langle \alpha _{-},\lambda _1 \beta _{-} ^1 \rangle _{T\mathfrak B ^{-}}=\lambda _1 \langle \alpha _{-},\beta _{-} ^1 \rangle _{T\mathfrak B ^{-}}=0\), locally \(W_1 ^*\) is an \(A_\eta ^\mathfrak B \)-invariant eigencodistribution, i.e., \(A_\eta ^\mathfrak B W_1 ^* \subseteq W_1 ^*\). It is easy to see that the restriction of \(A_\eta ^\mathfrak B \) to \(W_1 ^*\), \(A_\eta ^\mathfrak B |_{W_1 ^*}\) is also self-adjoint. Since \(\mathrm{dim} W_1 ^*= \mathrm{dim} \mathcal W -1=k\) there are \(k\)-basis 1-forms \(\{ \beta _{-} ^2, \ldots , \beta _{-} ^{\omega } \}\) for \(W_1 ^*\) that are eigenforms of \(A_\eta ^\mathfrak B |_{W_1 ^*}\). Though, each eigenform of \(A_\eta ^\mathfrak B |_{W_1 ^*}\) must be an eigenform of \(A_\eta ^\mathfrak B \). Conclude that the eigenforms of \(A_\eta ^\mathfrak B \), \(\{\beta _{-} ^1, \beta _{-} ^2, \ldots , \beta _{-} ^{\omega } \} \in T_p\mathfrak B ^{-}\) must be an orthonormal basis for \(T_p\mathfrak B ^{-}\) at each \(p \in \mathfrak B \) and thereby \(T_p\mathfrak B ^{-}\) can be partitioned locally in eigencodistributions as \(W_1 ^* \oplus \ldots \oplus W_\omega ^*\).

Proof of Proposition 3.1

Each component of \(\mathbb G \) can be calculated from \(g_{ij}=\langle \beta _i,\beta _j\rangle _{T\mathfrak B ^-}\). Since \(\langle \beta _j,\eta \rangle _{T\mathfrak B ^{-}}=0\) for \(\beta _j \in T\mathfrak B ^-\), \(\eta \in (T\mathfrak B ^-)^\perp \), then after derivation \(\partial _{\beta _i}\langle \beta _j,\eta \rangle _{T\mathfrak B ^-}=\langle \partial _{\beta _i}\beta _j,\eta \rangle _{T\mathfrak B ^-}+\langle \beta _j,\partial _{\beta _i}\eta \rangle _{T\mathfrak B ^-}=0\) yields \(\langle \partial _{\beta _i} \beta _j,\eta \rangle _{T\mathfrak B ^-}=\langle \beta _j,-\partial _{\beta _i} \eta \rangle _{T\mathfrak B ^-}=\langle \beta _j,-\bar{\nabla }_\eta \beta _i \rangle _{T\mathfrak B ^-}\). This implies that \(II_\mathfrak{B }(\beta _i,\beta _j):=\langle A_\eta ^\mathfrak B \beta _i, \beta _j \rangle _{T\mathfrak B ^-}=\langle \partial _{\beta _i} \beta _j,\eta \rangle _{T\mathfrak B ^-}\). In consequence the components of \(\mathbb Q \) can be calculated from \(q_{ij}=\langle \partial _{\beta _i} \beta _j,\eta \rangle _{T\mathfrak B ^-}\). Now since

$$\begin{aligned} q_{ij}\!=\!\langle A_\eta ^\mathfrak B \beta _i,\beta _j \rangle _{T \mathfrak B ^-}=\left< \sum _{k=1} ^{\omega } a^{ik}\beta _k , \beta _j \right>_{T\mathfrak B ^-}\!=\! \sum _{k=1} ^{\omega } a^{ik} \langle \beta _k, \beta _j \rangle _{T\mathfrak B ^-}= \sum _{k=1} ^{\omega } a^{ik} g_{kj}\nonumber \\ \end{aligned}$$
(44)

conclude that \(\mathbb Q =A_\eta ^\mathfrak B \mathbb G \). Since \(\mathbb G \) is a positive definite matrix then it is invertible and multiplication by its inverse yields the desired result. By linear algebra arguments \(K(\beta )=\det \mathbb Q /\det \mathbb G \), is another expression for Gauss (total) curvature.

Proof of Proposition 3.2

Any vector field \(\xi ^{+} \in T_w \mathfrak B ^{+}\) (1-form \(\alpha _{-} \in T_w \mathfrak B ^{-}\)) can be expressed in local coordinates by a combination of their (co-) frame elements \(\xi ^{+}=\sum _{i=1} ^{\omega } a_i \zeta _i ^{+}\) ( resp. \(\alpha _{-}=\sum _{i=1} ^{\omega } b_i \beta _{-} ^{i}\)) for scalar functions \(a_i, b_i \in C^{\infty }(\mathcal W )\), see e.g. Chapter 8 in [25]. In particular, let \(\xi ^{+} \in T_w \mathfrak B ^{+}\) have an orthogonal projection \(\xi _{\text{ op}} ^{+} \in \mathfrak B _{\text{ red}} ^{+} \subseteq \mathfrak B ^{+}\) on a submanifold satisfying \(T\mathfrak B _{\text{ red}} ^{+} \mathop {=}\limits ^\mathrm{def}\mathrm{span} \{ \zeta _1 ^{+}, \zeta _2 ^{+}, \ldots \zeta _{\varpi } ^{+} \}\), \(\varpi \le \omega \). Then their difference \(\xi ^{+}- \xi _0 ^{+}\) must be orthogonal to each element of the frame (by duality, see Remark 2.6), i.e., \(\langle \xi ^{+}- \xi _{\text{ op}} ^{+}, \zeta _i ^{+} \rangle _{T\mathfrak B ^{+}}=0\), \(i=1, \ldots ,\varpi \) where orthogonality can be defined in terms of the duality pairing \(\langle \xi ^{+}, \alpha _{-} \rangle _{T\mathfrak B ^{-} \times T\mathfrak B ^{+}}\), see Remark 2.5 and Definition 2.8. Express \(\xi _{\text{ op}} ^{+}\) by the \(\varpi \)-elements of the frame. The \(a_i\) satisfying \(\langle \xi ^{+}- \sum _{i=1} ^{\varpi } a_i \zeta _i ^{+}, \zeta _i ^{+} \rangle _{T\mathfrak B ^{+}}=0\) is equal to the projection of each vector field on the elements of the frame, i.e., \(a_i=\langle \xi ^{+},\zeta _i ^{+} \rangle _{T\mathfrak B ^{+}}\). An equivalent reasoning for \(\alpha _{-} \in T_w \mathfrak B ^{-}\), results in \(b_i=\langle \alpha _{-},\beta _{-} ^{i} \rangle _{T\mathfrak B ^{-}}\), concluding the proof.

Proof of Theorem 3.2

Under the orthonormal frame \(\{ \beta _{-} ^i\}\) of Proposition 3.2 any \(\alpha _{-} \in T_w \mathfrak B ^{-}\), \(\Vert \alpha _{-} \Vert =1\), can be expressed as \(\alpha _{-}=\sum _{i=1} ^{\omega } \langle \alpha _{-},\beta _{-} ^{i} \rangle _{T\mathfrak B ^{-}} \beta _{-} ^i\). Since \(\Vert \alpha _{-} \Vert =1\) then \(K(\alpha _{-})= \langle A_\eta ^\mathfrak B (\alpha _{-}),\alpha _{-} \rangle _{T\mathfrak B ^{-}}\) can be expressed as

$$\begin{aligned} K(\alpha _{-})&= \left< A_\eta ^\mathfrak B \sum _{i=1} ^{\omega } \langle \alpha _{-},\beta _{-} ^{i} \rangle _{T\mathfrak B ^{-}} \beta _{-} ^i,\sum _{i=1} ^{\omega } \langle \alpha _{-},\beta _{-} ^{i} \rangle _{T\mathfrak B ^{-}} \beta _{-} ^i \right>_{T\mathfrak B ^{-}}\\&= \left< \sum _{i=1} ^{\omega } \kappa _i(w) \langle \alpha _{-},\beta _{-} ^{i} \rangle _{T\mathfrak B ^{-}} \beta _{-} ^i,\sum _{i=1} ^{\omega } \langle \alpha _{-},\beta _{-} ^{i} \rangle _{T\mathfrak B ^{-}} \beta _{-} ^i \right>_{T\mathfrak B ^{-}} \end{aligned}$$

notice that \(\left< \beta _{-} ^{i}, \beta _{-} ^{i}\right>_{T\mathfrak B ^{-}}=1\) (by orthogonality) and we may write

$$\begin{aligned} K(\alpha _{-})&= \sum _{i=1} ^{\omega } \kappa _i(w) \langle \alpha _{-},\beta _{-} ^{i} \rangle _{T\mathfrak B ^{-}} ^2 \int _{-\infty } ^{0} \beta _{j} ^i \beta _{j} ^i \;\mathrm{d}\mu (t)= \sum _{i=1} ^{\omega } \kappa _i(w) \langle \alpha _{-},\beta _{-} ^{i} \rangle _{T\mathfrak B ^{-}} ^2 \end{aligned}$$

and since the angle between \(\alpha _{-}\) and \(\beta _{-} ^{i}\) is given by \(\theta _i=\mathrm{angcos}(\langle \alpha _{-},\beta _{-} ^{i} \rangle _{T\mathfrak B ^{-}})\), yields (29). \(\square \)

Proof of Proposition 3.3

1) From Eq. (30) \(w_{-} ^i=\frac{1}{\sigma _i}\Gamma \circ \hat{w}_i(\tau )\) substitution in (31) yields \(\Gamma ^\dagger \circ \Gamma \circ w_i ^{-}=\sigma _i^2 w_i ^{-}\) with \(\lambda _i =\sigma _i ^2\) we obtain Eq. (32). Substitution of Eqs. (31) in (30) and the same steps verifies Eq. (33).

2) At the tangent space, Eqs. (32) and (33) are written as

$$\begin{aligned} (\tilde{\Gamma }^\dagger \circ \tilde{\Gamma })|^* \alpha _{-} ^{i}-\lambda _i \alpha _{-} ^{i}&= 0, \quad i=1, \ldots , \omega , \end{aligned}$$
(45)
$$\begin{aligned} (\tilde{\Gamma }\circ \tilde{\Gamma }^\dagger )|_* \xi _{i} ^{+}-\lambda _i \xi _{i} ^{+}&= 0, \quad i=1, \ldots , \omega . \end{aligned}$$
(46)

which (by commutativity of composition under the differential operation) is precisely the eigenvalue problem for the selfadjoint Shape operator in Theorem 3.1(1). By Theorem 3.1(2)–(3) the elements of the coframe \(\mathrm{span} \{ \beta _{-} ^{1}, \ldots ,\beta _{-} ^{\omega }\}=T_w \mathfrak B ^{-}\) are solution to the eigenproblem (45). Notice that since \(\zeta _{i} ^{+}=\tilde{\Gamma }^* \beta _{-} ^{i}\) the elements of the frame \(\mathrm{span} \{ \zeta _1 ^{+}, \ldots ,\zeta _{\omega } ^{+} \}=T_w \mathfrak B ^{+} \) are solution of the eigenproblem (46).

3) From Proposition 3.2 the orthogonal projection of \(\xi ^{+} \in T_w \mathfrak B ^{+}\) is given by Eq. (27). Since \(\xi ^{+}=\tilde{\Gamma }^* \alpha _{-}\) is on \(T_w \mathfrak B ^{+}\), we obtain \(\tilde{\Gamma }^* \alpha _{-}=\sum _{i=1} ^{\omega } \langle \tilde{\Gamma }^* \alpha _{-},\zeta _i ^{+} \rangle _{T\mathfrak B ^{+}} \zeta _i ^{+}\) where after dualization from Remark 2.6, the inner product can be written as \(\langle \tilde{\Gamma }^* \alpha _{-},\tilde{\Gamma }^* \beta _{-} ^i \rangle _{T\mathfrak B ^{-} \times T\mathfrak B ^{+}}=\langle \tilde{\Gamma }_* ^\dagger \tilde{\Gamma }^* \alpha _{-}, \beta _{-} ^i \rangle _{T\mathfrak B ^{-}}\). Since by Theorem 3.1(3), \(A_{\eta } ^\mathfrak{B } \alpha _{-}=\tilde{\Gamma }_* ^\dagger \tilde{\Gamma }^* \alpha _{-}\), we may write \(\tilde{\Gamma }^* \alpha _{-}=\sum _{i=1} ^{\omega } \langle \lambda _i \alpha _{-}, \beta _{-} ^{i} \rangle _{T\mathfrak B ^{-}} \zeta _i ^{+}=\sum _{i=1} ^{\omega } \lambda _i \langle \alpha _{-}, \beta _{-} ^{i} \rangle _{T\mathfrak B ^{-}} \zeta _i ^{+}\), i.e., Eq. (34). Departing from the orthogonal projection of \(\alpha _{-}=\tilde{\Gamma }_* ^\dagger \xi ^{+}\) from Eq. (28) \(\tilde{\Gamma }_* ^\dagger \xi ^{+}=\sum _{i=1} ^{\omega } \langle \Gamma _* ^\dagger \xi ^{+},\beta _{-} ^{i} \rangle _{T\mathfrak B ^{-}} \beta _{-} ^{i}\), \(\alpha _{-} \in T_w \mathfrak B ^{-}\). Since \(\langle \Gamma _* ^\dagger \xi ^{+},\beta _{-} ^{i} \rangle _{T\mathfrak B ^{-}}=\langle \xi ^{+}, \Gamma ^* \Gamma _* ^\dagger \zeta _{i} ^{+} \rangle _{T\mathfrak B ^{+}}\) implying \(\langle \xi ^{+}, \lambda _i \zeta _{i} ^{+} \rangle _{T\mathfrak B ^{+}}=\lambda _i \langle \xi ^{+}, \zeta _{i} ^{+} \rangle _{T\mathfrak B ^{+}}\) and Eq. (35) is obtained. \(\square \)

Proof of Proposition 3.5

1) By condition (39) \(S_r ^* (\hat{w}^0,r_r)\) and \(S_a (w^0,r_a)\) are orthogonally separated on their respective spaces. Hence, associated with each integral invariant function \(S_r ^i\) is an (exact) differential 1-form \(\zeta _i ^{+}\) defining a distribution \(\Delta _i\) with annihilator \(\Delta _i ^{\perp }\mathop {=}\limits ^\mathrm{def}\{\alpha _{-} \in T_p \mathfrak B ^{-} \,|\, \langle \alpha _{-}, \zeta _i ^{+} \rangle _{T\mathfrak B ^{-} \times T\mathfrak B ^{+}}=0,\, \zeta _i ^{+} \in \Delta _i \}\) (dual to the distribution obtained in Proposition 3.1(3)). From Definition 3.5, the set of annihilators \(\{ \Delta _1 ^{\perp } ,\ldots ,\Delta _\omega ^{\perp } \}\) is an orthogonal web. Similarly, associated with each integral invariant function \(S_a ^{i}\) is an exact differential 1-form \(\beta _{-} ^i\) and the family of annihilators \(\{ \Omega _1 ^{\perp }, \ldots , \Omega _\omega ^{\perp } \}\), defined by \(\Omega _i ^{\perp }\mathop {=}\limits ^\mathrm{def}\{\xi ^{+} \in T_p \mathfrak B ^{+} \,|\, \langle \beta _{-} ^i, \xi ^{+} \rangle _{T\mathfrak B ^{-} \times T\mathfrak B ^{+}}=0,\, \beta _{-} ^i \in \Omega _i \}\) defining the orthogonal web for \(S_a (w^0,r_a)\).

2) Since by virtue of Theorem 3.1 each element in the set of orthogonal differential 1-forms \(\{\beta _{-} ^{i}\} \in T^*\mathcal W \) has an associated integral functional \(\hat{S}_i \in \mathcal W \), from the proof of Theorem 3.1(2), each principal direction \(\beta _{-} ^{i}\) is normal to an orthogonal hypersurface \(\hat{\mathcal{N }}_i\) defined by a (locally) integrable distribution (i.e. its annihilator) whose integral invariant function is precisely \(S_i \in \mathcal W \). By duality, the set of tangent vector fields \(\{\zeta _{i} ^{+}\} \in T\mathcal W \) is normal to an orthogonal hypersurface \(\mathcal N _i\) defined by a (locally) integrable codistribution (i.e. its annihilator) whose integral invariant function is precisely \(S_i \in \mathcal W \). \(\square \)

Proof of Lemma 3.1

Given a completely separated function \(S(w)\) on \(\mathcal W \), a Riemannian metric for \(\mathcal W \) can be defined by \( ds^2=\sum _{i=1} ^{n} S_i(w) \, ds_i ^2,\) and \(\{(\mathcal W _i, S_i)\}\) defines a partition of unity. For a principal frame-balanced realization \(S_r ^*(\hat{w}^0,r)\) and \(S_a(w^0,r)\) can be expressed as in Eq. (36) and metrics defined by \(ds_p^2=\sum _{i=1} ^{\omega } S^r _i(w) \, ds_i ^2\) and \(ds_f ^2=\sum _{i=1} ^{\omega } S^a _i(w) \, ds_i ^2\), with partitions of unity \(\{(\mathcal W _i, S^r _i)\}\) and \(\{(\mathcal W _i, S^a _i)\}\) are well-defined Riemannian metrics since each \(S_i ds_i ^2\) makes sense on all \(\mathcal W \) and \(S_i=0\) outside each \(\mathcal W _i\).

Proof of Proposition 4.1

Steps 1) and 2) are slight modifications of the corresponding arguments presented in Part II of this paper (Propositions 3.1, 3.2 and 3.5 for internal signals) and therefore their proofs are omitted. For step 3), we depart from an extended group \(\{ \Phi ^t(w)\}_{t}\), regular by assuming Eq. (38) satisfied. We are looking for a bijection \(\mathcal R (w) \in C^\infty (\mathcal W )\) defining a regular equivalence relation compatible with the group, i.e., \(w^1\mathcal R w^2, \Phi \in \{ \Phi ^t(w)\}_{t}\), \(\Rightarrow \) \(\Phi (w^1) \mathcal R \Phi (w^2)\), where \(\mathcal R _i(w) \in C^\infty (\mathcal W ), \; i=1 \ldots \omega \) are components of a vector function \(\mathcal R (w): \mathbb R ^\omega \rightarrow \mathbb R ^{\omega -r},\; w \in \mathcal W \) s.t. \(\det [\partial \mathcal R (w)/\partial w] \ne 0, \forall w \in \mathcal W \). Since the orbit of \(\{ \Phi ^t(w)\}_{t \ge 0}\) is given by the set \(\mathcal O \mathop {=}\limits ^\mathrm{def}\{ w \in \mathcal O | \Phi ^t(w) \in \mathcal O , \}\) the natural equivalence relation for all trajectories on the orbit \(\mathcal O \) is precisely given by the natural projection \(\pi : \mathcal W \rightarrow \mathcal W /\mathcal R \) since if two points \(w^1, w^2 \in \mathcal W \) lie on the same orbit of \(\Phi \in \{ \Phi ^t(w)\}_{t}\) such that \(\Phi \cdot w \in \mathcal W \), the induced equivalence relation among them is given by \(\pi (w^1)=\pi (w^2)\) s.t. \(\pi (\Phi \cdot w)=\pi (w)\). In such regular equivalence relation, every equivalence class in the quotient set \(\mathcal W /\Phi \) is an \(r\)-dimensional submanifold \(\mathcal R _i(w)=k_i ,\; i=1, \ldots , r\), for constants \(k_i \in \mathbb R ^1\), [6, 24] (see also [34]). It is a bijection since if \(w \in \mathcal W /\Phi \), then its corresponding orbit in \(\mathcal W \) is given by \(\pi ^{-1}\{w\}\). Finally, since by assumption \(\{ \Phi ^t(w)\}_{t}\) acts regularly on \(\mathcal W \) and \(\pi \) is open in \(\mathcal W /\Phi \), then \(\mathcal W /\Phi \) is endowed with the structure of a smooth manifold [6, 24].

4) Associated with each separated component function \(S_r ^i (\hat{w})\) and \(S_a ^i (w)\) is a subgroup \(\Phi ^t _i (w)=-\nabla _w S_a ^i (w)\) whose orbit is defined on their corresponding orthogonal web as in Definition 3.5. This is natural since each generating function is an integral invariant of the groups it generates. There is thus a partition of the space s.t.

$$\begin{aligned} \begin{array}{c c l} \mathcal W _a&\mathop {=}\limits ^\mathrm{def}&\mathcal W _1\oplus \mathcal W _2 \oplus \cdots \oplus \mathcal W _r, \\ \mathcal W _b&\mathop {=}\limits ^\mathrm{def}&\mathcal W _{r+1}\oplus \mathcal W _{r+2} \oplus \cdots \oplus \mathcal W _{\omega }, \end{array} \;\; r \le \omega , \end{aligned}$$
(47)

Since associated with each subgroup there is a submanifold \(\mathcal W _i \subset \mathcal W \) and a particular principal curvature \(\kappa _i\), a local partial order for the submanifolds \(\mathcal W _i\) can be provided by inequality relations for the principal curvatures (singular values): \(\kappa _1 \ge \kappa _2 \ge \cdots \ge \kappa _\omega \).

5) Consider the submanifold partition in (47). By orthogonality a quotient manifold \(\mathcal W /\mathcal W _b\) can be defined always and the reduced subsystem can be obtained with the use of the natural projection \(\pi _b: \mathcal W \rightarrow \mathcal W /\mathcal W _b\), see [24]. In view of Proposition 3.2, the generator vector field \(\xi _{\text{ op}} ^{+}\) restricted to the quotient space \(\mathcal W /\mathcal W _b\) defines uniquely the reduced-order (quotient) system.

Proof of Proposition 4.2

Consider the set of past trajectories \(\hat{w} \in \mathfrak B _{\text{ red}}^{-} \subset \mathfrak B ^{-}\) adapted to the behavior \(\mathfrak B ^{-}\). From Eq. (28) any such adapted past trajectory has an adapted tangent vector field written as \(\alpha _{-}=\sum _{i=1} ^{r} b_i \beta _{-} ^{i}\), \(b_i \in C^{\infty }(\mathcal W )\), \(\alpha _{-} \in \mathfrak B ^{-} \). Consider set of input vector fields given by \(\bar{\alpha }_{-}=\alpha _{-}+b_{r+1} \beta _{-} ^{r+1}\) for \(\Vert \tilde{\Gamma }-\tilde{\Gamma }_{\text{ red}} \Vert _\mathfrak{B }\). Since by construction \(0 \ne b_{r+1}\in C^{\infty }(\mathcal W )\) can be selected such that \(\bar{\alpha }_{-} \in \mathrm{Ker}\, \tilde{\Gamma }_{\text{ red}}\), \(\Vert (\tilde{\Gamma }-\tilde{\Gamma }_{\text{ red}})\circ \hat{w}_{-} \Vert _\mathfrak{B ^{+}} ^2=\Vert \tilde{\Gamma }\circ \hat{w}_{-} \Vert _\mathfrak{B ^{+}} ^2\) using Eq. (34) and Euler’s formula (29) the squared norm becomes \(\langle \tilde{\Gamma }^* \bar{\alpha }_{-},\tilde{\Gamma }^* \bar{\alpha }_{-} \rangle _{T\mathfrak B ^{+}}= \sum _{i=1} ^{r+1} \kappa _i(w) \langle \alpha _{-},\beta _{-} ^{i} \rangle _{T\mathfrak B ^{-}} ^2= \sum _{i=1} ^{r+1} \kappa _i(w) \langle b_i\beta _{-} ^{i},\beta _{-} ^{i} \rangle _{T\mathfrak B ^{-}} ^2\) which by orthonormality of \(\beta _{-} ^{i}\) yields \(\sum _{i=1} ^{r+1} \kappa _i(w) b_i ^2 \ge \kappa _{r+1}(w) \sum _{i=1} ^{r+1} b_i ^2=\kappa _{r+1}(w) \Vert w\Vert _\mathfrak{B ^{-}} ^2\), concluding the proof.

Rights and permissions

Reprints and permissions

About this article

Cite this article

Lopezlena, R. A geometric approach to nonlinear dissipative balanced reduction I: exogenous signals. Math. Control Signals Syst. 25, 63–96 (2013). https://doi.org/10.1007/s00498-012-0092-0

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s00498-012-0092-0

Keywords

Navigation