Skip to main content

Analysis Under Nested Criteria

  • Chapter
  • First Online:
  • 795 Accesses

Part of the book series: SpringerBriefs in Applied Sciences and Technology ((BRIEFSAPPLSCIENCES))

Abstract

The small-gain theorem provides a sufficient condition for the stability of two feedback interconnected dynamical systems for which input-to-state (or input–output) gains can be defined. Roughly speaking, to apply this theorem, the resulting gains’ composition is required to be continuous, increasing, and strictly smaller than the identity function.

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

Buying options

Chapter
USD   29.95
Price excludes VAT (USA)
  • Available as PDF
  • Read on any device
  • Instant download
  • Own it forever
eBook
USD   39.99
Price excludes VAT (USA)
  • Available as EPUB and PDF
  • Read on any device
  • Instant download
  • Own it forever
Softcover Book
USD   54.99
Price excludes VAT (USA)
  • Compact, lightweight edition
  • Dispatched in 3 to 5 business days
  • Free shipping worldwide - see info

Tax calculation will be finalised at checkout

Purchases are for personal use only

Learn about institutional subscriptions

Notes

  1. 1.

    To conclude about the asymptotic stability of this example, one may infer from the LaSalle invariance principle together with the Lyapunov function \(V+W\). Other techniques also apply, see [4], for example.

  2. 2.

    Note that \(0.95\rho _x(2)=0.6\).

  3. 3.

    Note that \(b=\infty \).

  4. 4.

    Note also that, when \(Y(t,\mathbf {Z})\) does not exist, \(Y(t,\mathbf {Z})=\emptyset \).

  5. 5.

    Because (3.8) is of class \(\mathscr {C}^1\) and solutions are unique (see also [13, Corollary 3.1]).

  6. 6.

    Note that, from the previous paragraph, for almost every \(y\in \varOmega _{=M_g}(U_\infty )\), \(\tfrac{\partial U_g}{\partial y}(y)\) exists.

  7. 7.

    From the uniqueness of solutions with respect to initial conditions, the closed orbit C is a simple closed curve.

  8. 8.

    Pfeffer [21, p. 49].

  9. 9.

    In other words, the set \(\varOmega _{=c}(V)\) can be \(\mathscr {H}^1\)-almost everywhere covered by countably many 1-dimensional curves of class \(\mathscr {C}^1\).

  10. 10.

    Adapted from [20, Theorem 1.7] or [21, Theorem 6.5.4]. In the latter, the set where the integral is computed is assumed to have Bounded Variation, in [21, Theorem 6.5.5] it is shown that a set has bounded variation if and only if it has finite perimeter.

  11. 11.

    More specifically from (3.38).

References

  1. Alberti G, Bianchini S, Crippa G (2013) Structure of level sets and Sard-type properties of Lipschitz maps. Ann Scuola Norm Sup Pisa Cl Sci 12(4):863–902

    MathSciNet  MATH  Google Scholar 

  2. Andrieu V, Prieur C (2010) Uniting two control Lyapunov functions for affine systems. IEEE Trans Autom Control 55(8):1923–1927

    Article  MathSciNet  Google Scholar 

  3. Angeli D (2004) An almost global notion of input-to-state stability. IEEE Trans Autom Control 49(6):866–874

    Article  MathSciNet  Google Scholar 

  4. Angeli D, Astolfi A (2007) A tight small-gain theorem for not necessarily ISS systems. Syst Control Lett 56(1):87–91

    Article  MathSciNet  MATH  Google Scholar 

  5. Astolfi A, Praly L (2012) A weak version of the small-gain theorem. In: IEEE 51st conference on decision and control (CDC), 2012, pp 4586–4590

    Google Scholar 

  6. Chaillet A, Angeli D, Ito H (2012) Strong iISS: combination of iISS and ISS with respect to small inputs. In: Proceedings of the 51st IEEE conference on decision and control, pp 2256–2261

    Google Scholar 

  7. Dashkovskiy S, Ruffer BS (2010) Local ISS of large-scale interconnections and estimates for stability regions. Syst Control Lett 59:241–247

    Article  MathSciNet  MATH  Google Scholar 

  8. Dashkovskiy S, Ruffer BS, Wirth F (2010) Small gain theorems for large scale systems and construction of ISS Lyapunov functions. SIAM J Control Optim 48(6):4089–4118

    Article  MathSciNet  MATH  Google Scholar 

  9. DiBenedetto E (2002) Real analysis, series. Advanced texts: Basler Lehrbücher. Birkhäuser, Boston

    Google Scholar 

  10. Evans LC, Gariepy RF (1992) Measure theory and fine properties of functions. CRC Press

    Google Scholar 

  11. Federer H (1945) The Gauss-green theorem. Trans Am Math Soc 58:44–76

    Article  MathSciNet  MATH  Google Scholar 

  12. Freeman RA, Kokotović PV (2008) Robust nonlinear control design: state space and Lyapunov techniques. Birkhäuser

    Google Scholar 

  13. Hartman P (1982) Ordinary differential equations, 2nd edn. SIAM

    Google Scholar 

  14. Isidori A (1995) Nonlinear control systems, series. Communications and control engineering. Springer

    Google Scholar 

  15. Ito H (2006) State-dependent scaling problems and stability of interconnected iISS and ISS systems. IEEE Trans Autom Control 51(10):1626–1643

    Article  MathSciNet  Google Scholar 

  16. Ito H, Jiang Z-P (2009) Necessary and sufficient small gain conditions for integral input-to-state stable systems: a Lyapunov perspective. IEEE Trans Autom Control 54(10):2389–2404

    Article  MathSciNet  Google Scholar 

  17. Jiang Z-P, Mareels IMY, Wang Y (1996) A Lyapunov formulation of the nonlinear small-gain theorem for interconnected ISS systems. Automatica 32(8):1211–1215

    Article  MathSciNet  MATH  Google Scholar 

  18. Jiang Z-P, Teel AR, Praly L (1994) Small-gain theorem for ISS systems and applications. Math Control Signals Syst 7:95–120

    Article  MathSciNet  MATH  Google Scholar 

  19. Liberzon D, Nesić D, Teel AR (2013) Lyapunov-based small-gain theorems for hybrid systems. IEEE Trans Autom Control 59(6):1395–1410

    Google Scholar 

  20. Marzocchi A (2005) Singular stresses and nonsmooth boundaries in continuum mechanics. In: XXX Ravello summer school

    Google Scholar 

  21. Pfeffer WF (2012) The divergence theorem and sets of finite perimeter. CRC Press

    Google Scholar 

  22. Rantzer A (2001) A dual to Lyapunov’s stability theorem. Syst Control Lett 42:161–168

    Article  MathSciNet  MATH  Google Scholar 

  23. Rudin W (1976) Principles of mathematical analysis. McGraw-Hill

    Google Scholar 

  24. Salsa S (2008) Partial differential equations in action: from modelling to theory. Springer

    Google Scholar 

  25. Sanfelice RG (2014) Input-output-to-state stability tools for hybrid systems and their interconnections. IEEE Trans Autom Control 59(5):1360–1366

    Article  MathSciNet  Google Scholar 

  26. Sastry S (1999) Nonlinear systems, series. Interdisciplinary applied mathematics, vol 10. Springer

    Google Scholar 

  27. Sontag ED (1989) Smooth stabilization implies coprime factorization. IEEE Trans Autom Control 34(4):435–443

    Article  MathSciNet  MATH  Google Scholar 

  28. Sontag ED (2001) The ISS philosophy as a unifying framework for stability-like behavior. In: Nonlinear control in the year, vol 259. Springer, pp 443–467

    Google Scholar 

  29. Sontag ED (2008) Input to state stability: basic concepts and results. In: Nonlinear and optimal control theory, series. Lecture notes in mathematics, vol 1932. Springer, pp 163–220

    Google Scholar 

  30. Sontag ED, Wang Y (1995) On characterizations of the input-to-state stability property. Syst Control Lett 24(5):351–359

    Article  MathSciNet  MATH  Google Scholar 

  31. Spiegel MR (1959) Vector analysis and an introduction to tensor analysis. McGraw-Hill

    Google Scholar 

  32. Stein Shiromoto H (2014) Stabilisation sous contraintes locales et globales. PhD thesis, Université Grenoble Alpes

    Google Scholar 

  33. Stein Shiromoto H, Andrieu V, Prieur C (2013) Interconnecting a system having a single input-to-state gain with a system having a region-dependent input-to-state gain. In: Proceedings of the 52nd IEEE conference on decision and control (CDC), Florence, Italy, 2013, pp 624–629

    Google Scholar 

  34. Stein Shiromoto H, Andrieu V, Prieur C (2015) A region-dependent gain condition for asymptotic stability. Automatica 52:309–316

    Article  MathSciNet  MATH  Google Scholar 

  35. Zames G (1966) On the input-output stability of time-varying nonlinear feedback systems part I: conditions derived using concepts of loop gain, conicity, and positivity. IEEE Trans Autom Control 11(2):228–238

    Article  Google Scholar 

Download references

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Humberto Stein Shiromoto .

Appendix of Chap. 3

Appendix of Chap. 3

3.1.1 Technical Lemma

Lemma 3.19

Let \(k\ge 0\) and \(p>0\) be constant integers. Given a function \(h\in \mathscr {C}^k(\mathbb R^n,\mathbb R^p)\), and compact set \(\mathbf {K}\subset \mathbb R^n\) such that, for every \(y\in \mathbf {K}\), \(h(y)\ne 0\). Then, there exists \(\tilde{h}\in \mathscr {C}^k(\mathbb R^n,\mathbb R^p)\) such that \(\mathtt {supp}(\tilde{h})\supset \mathbf {K}\) and, for every \(y\in \mathbf {K}\), \(\tilde{h}(y)=h(y)\).

The proof of Lemma 3.19 is based on [24, p. 370] and can be found in [32].

3.1.2 The Divergence Theorem for Level Sets of a Lyapunov Function

The following definition is recalled from [23]:

Definition 3.20

(Gamma function) The function

$$\begin{aligned} \begin{array}{rrcl} \varGamma :&{}\mathbb {R}_{\ge 0}&{}\rightarrow &{}\mathbb R\\ &{}s&{}\mapsto &{}\displaystyle \int \limits _0^\infty t^{s-1}e^{-t}\,dt \end{array} \end{aligned}$$

is called gamma function.

The next definition is recalled from [21].

Definition 3.21

(Hausdorff measure) Let \(\mathbf {E}\subset \mathbb R^n\), the diameter of the set \(\mathbf {E}\) is the function

$$\begin{aligned} \begin{array}{rrcl} {{\mathrm{\mathtt {diam}}}}:&{}\mathbf {E}\times \mathbf {E}&{}\rightarrow &{}\mathbb {R}_{\ge 0}\\ &{}(x,y)&{}\mapsto &{}\sup \{|x-y|\}. \end{array} \end{aligned}$$

Let \(0\le n<\infty \) and define, for \(0<\delta \le \infty \), the value

$$\begin{aligned} \mathscr {H}^n_\delta (\mathbf {E})=\inf \left\{ \sum _{j\in \mathbb {N}}{{\mathrm{\mathtt {diam}}}}(\mathbf {E}_j)^n:\mathbf {E}\subset \bigcup _{j\in \mathbb {N}}\mathbf {E}_j,{{\mathrm{\mathtt {diam}}}}(\mathbf {E}_j)<\delta ,\mathbf {E}_j\subset \mathbb R^n\right\} . \end{aligned}$$

The n-dimensional unnormalized Hausdorff measure of \(\mathbf {E}\) is the limit

$$\begin{aligned} \widetilde{\mathscr {H}}^n(\mathbf {E})=\lim _{\delta \rightarrow 0}\mathscr {H}^n_\delta (\mathbf {E})=\sup _{\delta >0}\mathscr {H}^n_\delta (\mathbf {E}). \end{aligned}$$

The n-dimensional Hausdorff measure of \(\mathbf {E}\) is given by

$$\begin{aligned} \mathscr {H}^n(\mathbf {E})=\dfrac{\alpha (s)}{2^n}\widetilde{\mathscr {H}}^n(\mathbf {E}), \end{aligned}$$

where

$$\begin{aligned} \alpha (n)=\dfrac{\varGamma \left( \tfrac{n}{2}\right) }{\varGamma \left( \tfrac{n}{2}+1\right) }. \end{aligned}$$

The relation between Hausdorff and Lebesgue measures is explained in the following remark which is based on [10, Sect. 2.2] and [20]:

Remark 3.22

Note that the n-dimensional Lebesgue measure of a set \(\mathbf {E}\subset \mathbb R^n\) is the n-fold product of unidimensional Lebesgue measures (cf. Definitions A.5 and A.7) while the Hausdorff measure is computed in terms of arbitrarily coverings of \(\mathbf {E}\) with small diameters. Moreover, the Lebesgue measure in \(\mathbb R^n\) is equivalent to the n-dimensional Hausdorff measure, i.e., \(\mu =\mathscr {H}^n\). Also, if \(\mathscr {H}^n(\mathbf {E})<\infty \), then \(\mathscr {H}^{n-1}(\mathbf {E})=\infty \) and \(\mathscr {H}^{n+1}(\mathbf {E})=0\).

The next concept, recalled from [21, p. 50], is a measure-theoretical notion of boundaries of a set.

Definition 3.23

(Essential boundaries) For a set \(\mathbf {E}\subset \mathbb R^n\),

  • The essential exterior is the set

    $$\begin{aligned} {{\mathrm{\mathtt {ext}}}}_*(\mathbf {E})= \left\{ x\in \mathbb R^n:\lim _{r\rightarrow 0}\dfrac{\mu \left( \mathbf {E}\cap \mathbf {B}_{\le r}(x)\right) }{\mu \left( \mathbf {B}_{\le r}(x)\right) }=0\right\} ; \end{aligned}$$
  • The essential interior is the set \({{\mathrm{\mathtt {int}}}}_*(\mathbf {E})={{\mathrm{\mathtt {ext}}}}_*(\mathbb R^n\setminus \mathbf {E})\);

  • The essential closure is the set \(\mathtt {cl}_*\{\mathbf {E}\}=\mathbb R^n\setminus {{\mathrm{\mathtt {ext}}}}_*(\mathbf {E})\);

  • The essential boundary is the set \(\partial _*\mathbf {E}=\mathtt {cl}_*\{\mathbf {E}\}\setminus {{\mathrm{\mathtt {int}}}}_*(\mathbf {E})\).

The following properties holdFootnote 8:

$$\begin{aligned} \begin{array}{l} {{\mathrm{\mathtt {int}}}}_*(\mathbf {E})\subset \mathtt {cl}_*\{\mathbf {E}\},{{\mathrm{\mathtt {int}}}}_*(\mathbb R^n\setminus \mathbf {E})={{\mathrm{\mathtt {ext}}}}_*(\mathbf {E}),\\ {\partial _*\mathbf {E}=\mathtt {cl}_*(\mathbf {E})\cap \mathtt {cl}_*\{\mathbb R^n\setminus \mathbf {E}\}=\partial _*(\mathbb R^n\setminus \mathbf {E})=\mathbb R^n\setminus ({{\mathrm{\mathtt {int}}}}_*(\mathbf {E})\cup {{\mathrm{\mathtt {ext}}}}_*(\mathbf {E})).} \end{array} \end{aligned}$$

Definition 3.23 is related to the usual topological concepts as follows:

$$\begin{aligned} {{\mathrm{\mathtt {int}}}}(\mathbf {E})\subset {{\mathrm{\mathtt {int}}}}_*(\mathbf {E}),\quad \mathtt {cl}_*\{\mathbf {E}\}\subset \mathtt {cl}\{\mathbf {E}\},\quad \partial _*\mathbf {E}\subset \partial \mathbf {E}. \end{aligned}$$

Moreover,

$$\begin{aligned} \partial _*\mathbf {E}=\partial \mathbf {E}\Leftrightarrow {{\mathrm{\mathtt {int}}}}(\mathbf {E})={{\mathrm{\mathtt {int}}}}_*(\mathbf {E})\quad \text {and}\quad \mathtt {cl}_*\{\mathbf {E}\}=\mathtt {cl}\{\mathbf {E}\}, \end{aligned}$$

and the inclusion

$$\begin{aligned} {{\mathrm{\mathtt {int}}}}_*(\mathbf {E})\subset \left\{ x\in \mathbb R^n:\lim _{r\rightarrow 0}\dfrac{\mu \left( \mathbf {E}\cap \mathbf {B}_{\le r}(x)\right) }{\mu \left( \mathbf {B}_{\le r}(x)\right) }=1\right\} \end{aligned}$$

becomes an inequality, when \(\mathbf {E}\) is measurable.

The following measure-theoretical notion of perimeter of a set is recalled from [21, Definition 4.5.1]:

Definition 3.24

(Perimeter of a set) The perimeter of a set \(\mathbf {E}\subset \mathbb R^n\) is the measure

$$\begin{aligned} P(\mathbf {E})=\mathscr {H}^{n-1}(\partial _*\mathbf {E}). \end{aligned}$$

The perimeter is finite if \(\mu (\mathbf {E})+P(\mathbf {E})<\infty \).

The notion of a perimeter of a set is an important concept for the next theorem, adapted from [1, Theorem 2.5].

Theorem 3.25

Let \(k\ge 0\) be a constant integer, and consider the Lipschitz map \(V\in \mathscr {C}^k(\mathbb R^2,\mathbb {R}_{\ge 0})\) with \(\mathtt {supp}(V)\) compact. The following statements hold, for almost every \(c\in \mathbb {R}_{\ge 0}\):

  1. 1.

    \(\varOmega _{=c}(V)\) is 1-rectifiableFootnote 9 and \(\mathscr {H}^1\left( \varOmega _{=c}(V)\right) <\infty \);

  2. 2.

    For \(\mathscr {H}^1\)-almost every \(x\in \varOmega _{=c}(V)\), the map V is differentiable at x;

  3. 3.

    Every connected component C of \(\varOmega _{=c}(V)\) is either a point or a closed simple curve with a Lipschitz parametrization \(p:[a,b]\rightarrow C\) which is injective and satisfies, for almost every \(t\in [a,b]\),

    $$\dfrac{dp}{dt}(t)=\tau (p(t)),$$

    where, for every \(x\in C\), \(\tau (x)\) is the vector tangent to C.

From item 2 and since the level set \(\varOmega _{=c}(V)\) is either a point or a simple closed curve of \(\mathbb R^2\), \(\partial \varOmega _{=c}(V)=\mathtt {cl}\{\varOmega _{=c}(V)\}=\varOmega _{=c}(V)\). Moreover, \(\partial _*\varOmega _{=c}(V)\subset \mathtt {cl}\{\varOmega _{=c}(V)\}\). From item 1, the sublevel set \(\varOmega _{\le c}(V)\) has finite perimeter. Thus the inequality

$$\begin{aligned} \int \limits _{\varOmega _{= c}(V)}\,d\mathscr {H}^1<\infty \end{aligned}$$

holds. Note that, from Remark 3.22, this integral is defined in the Lebesgue sense in \(\mathbb R^1\).

The next definition, based on [20, Definition 1.6] and [21, pp. 127–128], recalls the concept of vector being an outward normal to a set.

Definition 3.26

(Outward normal) For every \(x\in \partial _*\mathbf {E}\), denote by \(n_\mathbf {E}(x)\) the unit vector of \(\mathbb R^n\) such that

$$\begin{aligned} \mathbf {H}_\pm (\mathbf {E},x)=\{y\in \mathbb R^n:\pm n_\mathbf {E}(x)\cdot (y-x)\ge 0\}. \end{aligned}$$

The function \(n_\mathbf {E}\) is called outward unit normal of \(\mathbf {E}\subset \mathbb R^n\) if, for every \(x\in \partial _*\mathbf {E}\),

$$\begin{aligned} \begin{array}{rcl} \displaystyle \lim _{r\rightarrow 0}\dfrac{\mu \left( \mathbf {B}_{\le r}(x)\cap \mathbf {H}_+(\mathbf {E},x)\cap \mathbf {E}\right) }{\mu \left( \mathbf {B}_{\le r}(x)\right) }&{}=&{}0,\\ \\ \displaystyle \lim _{r\rightarrow 0}\dfrac{\mu \left( \mathbf {B}_{\le r}(x)\cap \left( \mathbf {H}_-(\mathbf {E},x)\setminus \mathbf {E}\right) \right) }{\mu \left( \mathbf {B}_{\le r}(x)\right) }&{}=&{}0 \end{array} \end{aligned}$$
(3.36)

hold.

From item 2 of Theorem 3.25, for \(\mathscr {H}^1\)-almost every \(x\in \varOmega _{=c}(V)\), \({{\mathrm{\mathtt {grad}}}}\,V(x)\) exists. Thus, the vector field

$$\begin{aligned} \begin{array}{rrcl} n:&{}\varOmega _{= c}(V)&{}\rightarrow &{}\mathbb R^2\\ &{}x&{}\mapsto &{}\left\{ \begin{array}{rcl} \dfrac{{{\mathrm{\mathtt {grad}}}}\,V(x)}{|{{\mathrm{\mathtt {grad}}}}\,V(x)|},&{}\text {if}&{}{{\mathrm{\mathtt {grad}}}}\,V(x)\ \text {exists,}\\ 0,&{}\text {if}&{}\text {otherwise} \end{array} \right. \end{array} \end{aligned}$$
(3.37)

is \(\mathscr {H}^1\)-almost everywhere an outward normal to the set \(\varOmega _{\le c}(V)\). Since the outward normal to sets of finite perimeter is unique (cf. [11, Theorem 3.4]), the vector n satisfies the limits (3.36).

For a further reading on the sets of finite perimeters and on the construction of outward normals for them, the interested reader is invited to see [21, Chaps. 5 and 6].

The next result shows the relationship between the line integral along a closed curve and the integral on the domain bounded by this curve.

Theorem 3.27

(Generalized divergence theorem)Footnote 10 Under the assumptions of Theorem 3.25. Let \(k\ge 0\) be a constant integer, and consider the map \(f\in \mathscr {C}^k(\mathbb R^2,\mathbb R^2)\). Then, the formula

$$\begin{aligned} \iint \limits _{\varOmega _{\le c}(V)}{{\mathrm{\mathtt {div}}}}\,f(x)\,dx=\oint \limits _{[a,b]}f(p(s))\cdot n(p(s))\,ds \end{aligned}$$

holds, where the integral of the left-hand side (resp. right-hand side) is taken in the Lebesgue (resp. 1-dimensional Hasudorff) measure on \(\mathbb R^2\) (resp. \(\mathbb R\)), and \(p:[a,b]\rightarrow \varOmega _{=c}(V)\) is a parametrization of \(\varOmega _{=c}(V)\).

Before showing a sketch of the proof of Theorem 3.27, the following lemma, based on [31, p. 106], is needed. For a detailed proof in \(\mathbb R^n\), the interested reader may consult [21, Chaps. 1–6].

Lemma 3.28

(Green’s Theorem) Let \(C\subset \mathbb R^2\) be a positively oriented, piecewise smooth, simple closed curve with finite length, let \(\mathbf {D}_C\) be the region bounded by C, and let \(f=(f_1,f_2):\mathbb R^2\rightarrow \mathbb R^2\). If \(f_1:\mathbb R^2\rightarrow \mathbb R\) and \(f_2:\mathbb R^2\rightarrow \mathbb R\) are defined on an open region containing \(\mathbf {D}_C\), and f is differentiable in such a region, then

$$\begin{aligned} \oint \limits _C(f_1(x_1,x_2)\,dx_1 + f_2(x_1,x_2)\,dx_2)=\iint \limits _{\mathbf {D}_C}\left( \dfrac{\partial f_2}{\partial x_1}-\dfrac{\partial f_1}{\partial x_2} \right) \,dx_1dx_2, \end{aligned}$$
(3.38)

where the path of integration along C is counterclockwise.

Sketch of the (of Lemma 3.28 )

This proof is based on [31, p. 108]. Since C is a simple closed curve in the plane, the region \(\mathbf {D}_C\) is bounded. The projection of the curve in the x-axis (resp. y-axis) yields an interval [ab] (resp. [ef]). Consider the points of \(A,B\in C\) (resp. \(E,F\in C\)) corresponding to the points a and b (resp. e and f) on the x-axis, the curve C can be seen as the union of the curves AEB and AFB. Figure 3.3 illustrates the curve C, and the intervals [ab], and [ef].

Let the equation of the curve containing the points AEB (resp. AFB) be given by a piecewise continuous function \(\eta _1:[a,b]\rightarrow \mathbb R^2\) (resp. \(\eta _2:[a,b]\rightarrow \mathbb R^2\)).

Integrating the partial derivative of \(f_1\) with respect to \(x_2\) in \(\mathbf {D}_C\) yields

$$\begin{aligned} \displaystyle \iint \limits _{\mathbf {D}_C}\dfrac{\partial f_1}{\partial x_2}(x_1,x_2)\,dx_1dx_2=\displaystyle \int \limits _a^b\int \limits _{\eta _1(x_1)}^{\eta _2(x_1)}\dfrac{\partial f_1}{\partial x_2}(x_1,x_2)\,dx_2dx_1, \end{aligned}$$

where the equality is due to Fubini’s theorem (cf. [9, Theorem 14.1]). Moreover, since C has finite length, the equality

$$\begin{aligned} \begin{array}{rcl} \displaystyle \iint \limits _{\mathbf {D}_C}\dfrac{\partial f_1}{\partial x_2}(x_1,x_2)\,dx_1dx_2&{}=&{}\displaystyle \int \limits _a^b (f_1(x_1,\eta _2(x_1))-f_1(x_1,\eta _1(x_1)))\,dx_1\\ &{}=&{}-\displaystyle \int \limits _a^b f_1(x_1,\eta _1(x_1))\,dx_1-\displaystyle \int \limits _b^a f_1(x_1,\eta _2(x_1))\,dx_1\\ &{}=&{}-\displaystyle \oint \limits _C f_1(x_1,x_2)\,dx_1. \end{array} \end{aligned}$$

holds.

Analogously, integrating the partial derivative of \(f_2\) with respect to \(x_1\) in \(\mathbf {D}_C\) yields

$$\begin{aligned} \displaystyle \iint \limits _{\mathbf {D}_C}\dfrac{\partial f_2}{\partial x_1}(x_1,x_2)\,dx_1dx_2=\displaystyle \oint \limits _C f_2(x_1,x_2)\,dx_2. \end{aligned}$$

From where the conclusion follows.

Fig. 3.3
figure 3

Illustration of the curve C

Now, it is possible to present an idea of the proof of Theorem 3.27.

From Theorem 3.25,

  • The curve \(\varOmega _{=c}(V)\) is piecewise \(\mathscr {C}^1\), because it is rectifiable. Moreover, it is also simple and closed;

  • Since V is \(\mathscr {H}^1\)-a.e. differentiable in \(\varOmega _{=c}(V)\), the outward normal vector defined by (3.37) is \(\mathscr {H}^1\)-a.e. non-nil;

  • The curve \(\varOmega _{=c}(V)\) has finite length, because \(\mathscr {H}^1(\varOmega _{=c}(V))<\infty \);

  • There exists a injective and Lipschitz continuous parametrization \(p:[a,b]\rightarrow \varOmega _{=c}(V)\) that is a.e. differentiable.

Consider the vector field \(\tilde{f}=(-f_2,f_1)\) that is perpendicular to \(f=(f_1,f_2)\). Since \(f=(f_1,f_2)\in \mathscr {C}^1(\mathbb R^2,\mathbb R^2)\), \(\tilde{f}\in \mathscr {C}^1(\mathbb R^2,\mathbb R^2)\). Together with the above, fromFootnote 11 Lemma 3.28,

$$\begin{aligned} \displaystyle \oint \limits _{\varOmega _{=c}(V)}(-f_2(x_1,x_2)\,dx_1+f_1(x_1,x_2)\,dx_2)=\displaystyle \oint \limits _{\varOmega _{=c}(V)}(-f_2(x_1,x_2),f_1(x_1,x_2))\cdot (dx_1,dx_2) \end{aligned}$$

Consider a point \(\bar{x}=(\bar{x}_1,\bar{x}_2)\in C\) for which there exists \(\bar{s}\in [a,b]\) such that \(p(\bar{s})=(\bar{x}_1,\bar{x}_2)\) and \(p'(\bar{s})\) is defined. The unit tangent vector to C at \(\bar{x}\) is given by . The unit normal vector at \(\bar{x}\) is given by \(N(\bar{s})=n(p(s))=(\sigma (\bar{s}),-\tau (\bar{s}))\). For almost every \(s\in [a,b]\),

$$\begin{aligned} \begin{pmatrix}dx_1\\ dx_2 \end{pmatrix}=T(s)\,ds=\begin{pmatrix}\tau (s)\\ \sigma (s)\end{pmatrix}\,ds. \end{aligned}$$

Thus,

$$\begin{aligned} \begin{array}{rcl} \displaystyle \oint \limits _{\varOmega _{=c}(V)}(-f_2(x_1,x_2)\,dx_1+f_1(x_1,x_2)\,dx_2)&{}=&{}\displaystyle \oint \limits _{[a,b]}(-f_2(p(s)),f_1(p(s)))\cdot (\tau (s),\sigma (s))\,ds\\ &{}=&{}\displaystyle \oint \limits _{[a,b]}(f_1(p(s)),f_2(p(s)))\cdot (\sigma (s),-\tau (s))\,ds\\ &{}=&{}\displaystyle \oint \limits _{[a,b]}(f_1(p(s)),f_2(p(s)))\cdot n(p(s))\,ds \end{array} \end{aligned}$$

From (3.38),

$$\begin{aligned} \displaystyle \iint \limits _{\varOmega _{\le c}(V)}\left( \dfrac{\partial f_1}{\partial x_1}(x_1,x_2)+\dfrac{\partial f_2}{\partial x_2}(x_1,x_2)\right) \,dx_1dx_2=\displaystyle \oint \limits _{[a,b]}(f_1(p(s)),f_2(p(s)))\cdot n(p(s))\,ds. \end{aligned}$$

This concludes the sketch of the proof of Theorem 3.27.

3.1.3 Integration Along Solutions of an ODE

Before recalling the main result, the following lemma which is recalled from [22, Lemma A.1] is needed.

Lemma 3.29

(Liouville’s Theorem) Let \(k\ge 1\) and \(p\ge 1\) be constant integers, the function \(\rho \in (\mathscr {C}^k\cap \mathscr {L}^p)(\mathbb R^{n+m},\mathbb {R}_{\ge 0})\). Let also Y(ty) be a solution of (3.8) starting in \(y\in \mathbb R^{n+m}\) and computed at time \(t\in \mathbb {R}_{\ge 0}\). For a measurable set \(\mathbf {Z}\), let \(Y(\cdot ,\mathbf {Z})=\{Y(\cdot ,z):z\in \mathbf {Z}\}\). Then,

$$\begin{aligned} \int \limits _{Y(t,\mathbf {Z})}\rho (y)\,dy-\int \limits _\mathbf {Z}\rho (y)\,dy=\int \limits _0^t\int \limits _{Y(\tau ,\mathbf {Z})}{{\mathrm{\mathtt {div}}}}\,(\rho h)(y)\,dyd\tau . \end{aligned}$$

The main result in this section is recalled from [22, Theorem 1].

Theorem 3.30

(Almost attractivity) Let \(k\ge 1\) and \(p\ge 1\) be constant integers. Suppose that there exists \(\rho \in (\mathscr {C}^k\cap \mathscr {L}^p)(\mathbb R^n,\mathbb {R}_{\ge 0})\) such that,

$$\begin{aligned} \int _{\mathbf {B}_{\ge 1}(0)}\dfrac{(h\rho )(y)}{|y|}\,dy<\infty \end{aligned}$$

and

$$\begin{aligned} y\in \mathbb R^{n+m},\quad {{\mathrm{\mathtt {div}}}}\,(h\rho )(y)>0. \end{aligned}$$

Then, for almost every initial condition \(y\in \mathbb R^{n+m}\),

$$\begin{aligned} \limsup _{t\rightarrow \infty }|Y(t,y)|=0. \end{aligned}$$

Moreover, if the origin is stable, then the conclusion remains valid when \(\rho \) takes negative values.

Rights and permissions

Reprints and permissions

Copyright information

© 2017 The Author(s)

About this chapter

Cite this chapter

Stein Shiromoto, H. (2017). Analysis Under Nested Criteria. In: Design and Analysis of Control Systems. SpringerBriefs in Applied Sciences and Technology. Springer, Cham. https://doi.org/10.1007/978-3-319-52012-4_3

Download citation

  • DOI: https://doi.org/10.1007/978-3-319-52012-4_3

  • Published:

  • Publisher Name: Springer, Cham

  • Print ISBN: 978-3-319-52011-7

  • Online ISBN: 978-3-319-52012-4

  • eBook Packages: EngineeringEngineering (R0)

Publish with us

Policies and ethics