Skip to main content
Log in

A Survey of the Hysteretic Duhem Model

  • Original Paper
  • Published:
Archives of Computational Methods in Engineering Aims and scope Submit manuscript

An Erratum to this article was published on 08 August 2017

This article has been updated

Abstract

The Duhem model is a simulacrum of a complex and hazy reality: hysteresis. Introduced by Pierre Duhem to provide a mathematical representation of thermodynamical irreversibility, it is used to describe hysteresis in other areas of science and engineering. Our aim is to survey the relationship between the Duhem model as a mathematical representation, and hysteresis as the object of that representation.

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
Fig. 15
Fig. 16

Similar content being viewed by others

Change history

  • 08 August 2017

    An erratum to this article has been published.

Notes

  1. There are no specified authors. The one-page preface written by M. V. cites the following people as co–authors or co–authors to be: M. Edouard Jordan, M. J. Hadamard, M. L. Marchis, M. H. Pélabon, M. Ed. Le Roy, and M. Darbon. However, the chapters bear the following names. The biography of P. Duhem is written by E. Jordan. The “Notice sur les titres et travaux scientifiques de Pierre Duhem” is the note Duhem wrote himself when he applied to the Académie des Sciences. The following chapter “La physique de P. Duhem” is written by Octave Manville. The chapter “L’œuvre de Pierre Duhem dans son aspect mathématique” is authored by J. Hadamard. Finally “L’histoire des sciences dans l’œuvre de P. Duhem” is written by A. Darbon.

  2. For a detailed study of the life and work of Pierre-Maurice-Marie Duhem (9 June 1861–14 September 1916) see Refs. [39] or [67].

  3. We are indebted to Jean François Stoffel for this information.

  4. Ref. [48] cites a translation into German of the original memoir Ref. [16] which is written in French.

  5. Quoting from Ref. [48, p. 96]: “the Madelung paper does not use a differential equation or integral operator. In fact, Madelung allows nonuniqueness of trajectories through a point \(\ldots \) which would make a differential equation model difficult.”

  6. The term “rate independence” is attributed to Truesdell and Noll (Section 99, Encyclopedia of Phyics, volume III/3, 1965) by Visintin [64, p. 13]. We read Section 99 of the 2004 edition [62] of the original treatise by Truesdell and Noll but found no clear evidence of the correctness of the attribution.

  7. Called the Madelung model in Ref. [43].

  8. In this paper we avoid the words “positive”, “negative”, “increasing”, “decreasing” as they mean different things in different books.

  9. If the functions \(h_{\ell }\) and \(h_r\) are continuous then they are Borel and locally bounded. Continuity is the condition that appears in Ref. [48].

  10. Ref. [54] considers that u is continuous and piecewise \(C^1\). However, the results that we present here are also valid for inputs belonging to \(W^{1,\infty }(\mathbb {R}_+,\mathbb {R}).\)

  11. The uniqueness of \(x_{0,u}\) is not asked in [54, Definition 2.2]. However without uniqueness the equality in Condition (i) of Definition 4 would have no meaning since \(\mathcal {C}_{u,\gamma }\) would not correspond to a single mathematical object.

  12. u is non constant if \(\exists t_1,t_2 \in \mathbb {R}_+\) such that \(u(t_1)\ne u(t_2)\).

  13. In the proof of [54, Proposition 5.1] Oh and Bernstein use as input \(u \circ s_\gamma \) where \(u \in \Lambda \), and obtain by a limiting process a rate-independent semilinear Duhem model. In Ref. [35], Ikhouane extends this idea to causal operators \(\mathcal {H}:W^{1,\infty }(\mathbb {R}_+,\mathbb {R}^p) \times \Xi \rightarrow L^{\infty }(\mathbb {R}_+,\mathbb {R}^m)\) that satisfy Assumption 3, and to inputs that belong to \(W^{1,\infty }(\mathbb {R}_+,\mathbb {R}^p) \).

  14. Definition 6, Assumption 4, and Proposition 1 do not appear in Ref. [35].

  15. To the best of our knowledge, proposing a formal definition of hysteresis based on the existence of a hysteresis loop was first done by Oh and Bernstein in Ref. [54] for the generalized Duhem model, and for inputs belonging to \(\Lambda \). Ikhouane used a different perspective to generalize this idea to causal operators \(\mathcal {H}:W^{1,\infty }(\mathbb {R}_+,\mathbb {R}^p) \times \Xi \rightarrow L^{\infty }(\mathbb {R}_+,\mathbb {R}^m)\) that satisfy Assumption 3, and to periodic inputs that belong to \(W^{1,\infty }(\mathbb {R}_+,\mathbb {R}^p) \) [35].

  16. Definition 9 does not appear in Ref. [35]. Compare with Condition (ii) of Definition 4.

  17. Indeed, if \(\lambda \in \;]0,1[\), Eqs. (17)–(18) lead to \(x(t)=x_0,\forall t \in \mathbb {R}_+\). If \(\lambda \in \;]1,\infty [\), \(\varphi _u^{\star }\) is identically \(x_0\) which implies that \(\varphi _u^{\circ }\) is identically \(x_0\). In both cases the operator \(\mathcal {H}_s\) has a trivial hysteresis loop with respect to all inputs and initial states (see Definition 9).

  18. The condition that functions \(\lambda _1,\lambda _2\) are bounded on any bounded interval does not appear in Ref. [40]. However, without this condition there is no guarantee that the maximal interval of existence of the solutions of (34)–(36) is \([0,\infty [\), see Sect. 4.2. In [43, p. 278] it is considered that \(\lambda _1=\lambda _2\) is continuous so that the local boundedness condition holds.

  19. Since all the results of this section are proved for a finite time interval, Ref. [64] considers that the differential equation (43)–(44) holds almost everywhere on that finite time interval. We consider that the differential equation (43)–(44) holds almost everywhere on \(\mathbb {R}_+\) to simplify the discussion of Sect. 12.2 without loss of generality.

  20. If \(\displaystyle {\lim _{w \downarrow 0}\bar{g}_1(w)=a_1\ne 0}\) and \(\displaystyle {\lim _{w \uparrow 0}\bar{g}_2(w)=-a_2 \ne 0}\), the constants \(a_1\) and \(a_2\) are incorporated into the matrices \(A_1\) and \(A_2\) respectively.

  21. A matrix is stable if all its eigenvalues have strictly negative real parts.

  22. These special cases of are not studied in Ref. [35].

References

  1. Angeli D (2006) Systems with counterclockwise input-ouput dynamics. IEEE Trans Autom Control 51:1130–1143

    Article  MathSciNet  Google Scholar 

  2. Åström KJ, Canudas de Wit C (2008) Revisiting the LuGre friction model. IEEE Control Syst 28:101–114

    MATH  Google Scholar 

  3. Babuška I (1952) Die nichtlineare theorie der inneren reibung. Apl Mat 4:303–321

    MathSciNet  Google Scholar 

  4. Barker JA, Schreiber DE, Huth BG, Everett DH (1983) Magnetic hysteresis and minor loops: models and experiments. Proc R Soc Lond A 386:251–261

    Article  Google Scholar 

  5. Bernardini D, Vestroni F (2000) Hysteretic modeling of shape memory alloy vibration reduction devices. J Mater Process Manuf Sci 9:101–112

    Article  Google Scholar 

  6. Bernstein DS (2007) Ivory ghost [ask the experts]. IEEE Control Syst 27:16–17

    Google Scholar 

  7. Bertotti G, Mayergoyz I (eds) (2006) The science of hysteresis, vol 3. Elsevier, Amsterdam

    MATH  Google Scholar 

  8. Bouc R (1971) Modèle mathématique d’hystérésis. Acustica 21:16–25

    MATH  Google Scholar 

  9. Branciforte M, Meli A, Muscato G, Porto D (2011) ANN and non-integer order modeling of ABS solenoid valves. IEEE Trans Control Syst Technol 19:628–635

    Article  Google Scholar 

  10. Brokate M, Sprekels J (1996) Hysteresis and phase transitions. Springer, New York

    Book  Google Scholar 

  11. Canudas de Wit C, Olsson H, Åström KJ, Lischinsky P (1995) A new model for control of systems with friction. IEEE Trans Autom Control 40:419–425

    Article  MathSciNet  Google Scholar 

  12. Coleman BD, Hodgdon ML (1986) A constituve relation for rate-independent hysteresis in ferromagnetically soft materials. Int J Eng Sci 24:897–919

    Article  Google Scholar 

  13. Dahl P (1976) Solid friction damping of mechanical vibration. Am Inst Aeronaut Astronaut (AIAA) 14:1675–1682

    Article  Google Scholar 

  14. Della Torre E (1999) Magnetic hysteresis. IEEE Press, New York

    Google Scholar 

  15. Drinčić B (2012) Mechanical models of friction that exhibit hysteresis, stick–slip, and the Stribeck effect, PhD thesis, University of Michigan

  16. Duhem P (1896) Sur les déformations permanentes et l’hystérésis. Premier Mémoire, tome LIV, Mémoires couronnés et Mémoires des savants étrangers, l’Académie royale des sciences, des lettres et des beaux–arts de Belgique [The Cornell University library digital collections]

  17. Duhem P (1896) Les modifications permanentes du soufre. Deuxième Mémoire, tome LIV, Mémoires couronnés et Mémoires des savants étrangers, l’Académie royale des sciences, des lettres et des beaux–arts de Belgique [The Cornell University library digital collections]

  18. Duhem P (1896) Théorie générale des déformations permanentes. Troisième Mémoire, tome LIV, Mémoires couronnés et Mémoires des savants étrangers, l’Académie royale des sciences, des lettres et des beaux–arts de Belgique [The Cornell University library digital collections]

  19. Duhem P (1898) Étude de divers systèmes dépendant d’une seule variable. Quatrième Mémoire, tome LVI, Mémoires couronnés et Mémoires des savants étrangers, l’Académie royale des sciences, des lettres et des beaux–arts de Belgique [The Cornell University library digital collections]

  20. Duhem P (1898) Étude de divers systèmes dépendant de deux variables. Cinquième Mémoire, tome LVI, Mémoires couronnés et Mémoires des savants étrangers, l’Académie royale des sciences, des lettres et des beaux–arts de Belgique [The Cornell University library digital collections]

  21. Duhem P (1902) L’inégalité de Clausius et l’hystérésis. Sixième Mémoire, tome LXII, Mémoires couronnés et autres Mémoires, l’Académie royale de Belgique [The Cornell University library digital collections]

  22. Duhem P (1902) Hystérésis et viscosité. Septième Mémoire, tome LXII, Mémoires couronnés et autres Mémoires, l’Académie royale de Belgique [The Cornell University library digital collections]

  23. Edgar GA (1990) Measure, topology and fractal geometry. Springer-Verlag, New York

    Book  Google Scholar 

  24. Everett DH, Whitton WI (1952) A general approach to hysteresis. Trans Faraday Soc 48:749–757

    Article  Google Scholar 

  25. Everett DH, Smith FW (1954) A general approach to hysteresis. Part 2: development of the domain theory. Trans Faraday Soc 50:187–197

    Article  Google Scholar 

  26. Everett DH (1954) A general approach to hysteresis. Part 3: a formal treatment of the independent domain model of hysteresis. Trans Faraday Soc 50:1077–1096

    Article  Google Scholar 

  27. Everett DH (1955) A general approach to hysteresis. Part 4: an alternative formulation of the domain model. Trans Faraday Soc 51:1551–1557

    Article  Google Scholar 

  28. Ewing JA (1881) On the production of transient electric currents in iron and steel conductors by twisting them when magnetised or by magnetising them when twisted. Proc R Soc Lond 33:21–23

    Article  Google Scholar 

  29. Filippov AF (1988) Differential equations with discontinuous righthand sides. Kluwer Academic Publishers, Berlin

    Book  Google Scholar 

  30. Celanovic MA (1997) A lumped parameter electromechanical model for describing the nonlinear behavior of piezoelectric actuators. J Dyn Syst Meas Control 119:479

    MATH  Google Scholar 

  31. Hamimid M, Mimoune SM, Feliachi M, Atallah K (2014) Non centered minor hysteresis loops evaluation based on exponential parameters transforms of the modified inverse Jiles-Atherton model. Phys B 451:16–19

    Article  Google Scholar 

  32. Hartman P (2002) Ordinary differential equations. SIAM classics in applied mathematics. SIAM, Philadelphia

    Google Scholar 

  33. Hassani V, Tjahjowidodo T, Do TN (2014) A survey on hysteresis modeling, identification and control. Mech Syst Signal Proces 49:209–233

    Article  Google Scholar 

  34. Hodgdon ML (1988) Applications of a theory of ferromagnetic hysteresis. IEEE Trans Magn 24:218–221

    Article  Google Scholar 

  35. Ikhouane F (2013) Characterization of hysteresis processes. Math Control Signal Syst 25:294–310

    Article  MathSciNet  Google Scholar 

  36. Ikhouane F, Rodellar J (2006) A linear controller for hysteretic systems. IEEE Trans Autom Control 51(2):340–344

    Article  MathSciNet  Google Scholar 

  37. Ikhouane F, Rodellar J (2007) Systems with hysteresis: analysis, identification and control using the Bouc–Wen model. Wiley, New york

    Book  Google Scholar 

  38. Ismail M, Ikhouane F, Rodellar J (2009) The hysteresis Bouc–Wen model, a survey. Arch Comput Meth Eng 16:161–188

    Article  Google Scholar 

  39. Jaki SL (1987) Uneasy genius: the life and work of Pierre Duhem. Martinus Nijhoff Publishers, Leiden

    Book  Google Scholar 

  40. Jayawardhana B, Ouyang R, Andrieu V (2012) Stability of systems with the Duhem hysteresis operator: the dissipativity approach. Automatica 48:2657–2662

    Article  MathSciNet  Google Scholar 

  41. Jayawardhana B, Ouyang R, Andrieu V (2011) Dissipativity of general Duhem hysteresis models. In: 50th IEEE Conference on Decision and control and European control conference, Orlando, Florida, USA, December 12–15, pp. 3234–3239

  42. Khalil H (2002) Nonlinear systems, 3rd edn. Prentice Hall, New Jersey

    MATH  Google Scholar 

  43. Krasnosel’skiǐ MA, Pokrovskiǐ AV (1989) Systems with hysteresis. Springer, New York

    Book  Google Scholar 

  44. Lagrange C (1895) Rapport de M. Ch. Lagrange, premier commisaire. Bulletins de l’Académie royale des sciences, des lettres et des beaux-arts de Belgique, 65e année, 3e série, t. XXIX, no 6, 805–820

  45. Leoni G (2009) A first course in Sobolev spaces. The American Mathematical Society, Providence

    Book  Google Scholar 

  46. Loloee R, Pence TJ, Grummon DS (1995) Phase–fraction evolution during incomplete cyclic transformation in TiNi: correlation and analytical models with magnetic susceptibility measurements. J Phys IV 5(C8):C8–545

    Google Scholar 

  47. Logemann H, Ryan EP, Shvartsmann I (2008) A class of differential-delay systems with hysteresis: asymptotic behaviour of solutions. Nonlinear Anal 69:363–391

    Article  MathSciNet  Google Scholar 

  48. Macki JW, Nistri P, Zecca P (1993) Mathematical models for hysteresis. SIAM Rev 35:94–123

    Article  MathSciNet  Google Scholar 

  49. Mayergoyz I (2003) Mathematical models of hysteresis., Elsevier series in electromagnetism. Elsevier, Amsterdam

    MATH  Google Scholar 

  50. Morris KA (2011) What is hysteresis? Appl Mech Rev 64:14

    Google Scholar 

  51. Mohammad Naser MF, Ikhouane F (2013) Consistency of the Duhem model with hysteresis. Math Prob Eng

  52. Ni YQ, Ying ZG, Ko JM, Zhu WQ (2002) Random response of integrable Duhem hysteretic systems under non-white excitation. Non-Linear Mech 37:1407–1419

    Article  Google Scholar 

  53. Ouyang R, Andrieu V, Jayawardhana B (2013) On the characterization of the Duhem hysteresis operator with clockwise input-output dynamics. Syst Control Lett 62:286–293

    Article  MathSciNet  Google Scholar 

  54. Oh J, Bernstein DS (2005) Semilinear Duhem model for rate-independent and rate-dependent hysteresis. IEEE Trans Autom Control 50:631–645

    Article  MathSciNet  Google Scholar 

  55. Oh J, Drinčić B, Bernstein DS (2009) Nonlinear feedback models of hysteresis. IEEE Control Syst 29:100–119

    Article  MathSciNet  Google Scholar 

  56. Preisach F (1935) Über die magnetische Nachwirkung. Zeitschrift für Physik 94:277–302

    Article  Google Scholar 

  57. Padthe AK, Drincic B, Oh J, Rizos DD, Fassois SD, Bernstein DS (2008) Duhem modeling of friction-induced hysteresis. IEEE Control Syst 28:90–107

    MATH  Google Scholar 

  58. Padthe AK, Oh J, Bernstein DS (2005) Counterclockwise dynamics of a rate-independent semilinear Duhem model. In: 44th IEEE Conference on Decision and Control, 2005 and 2005 European Control Conference, Seville, Spain, December 12–15, pp. 8000–8005

  59. Rayleigh L (1887) On the behavior of iron and steel under the operations of feeble magnetic forces. Lond Edinb Dublin Philos Mag J Sci 23:225–245

    Article  Google Scholar 

  60. Rudin W (1987) Real and complex analysis, 3rd edn., McGraw-Hill series in higher mathematics. McGraw-Hill, New york

    MATH  Google Scholar 

  61. Schmitt K, Thompson RC (2000) Nonlinear analysis and differential equations. An introduction, Free E-Book

  62. Truesdell C, Noll W(2004) In: Antman SS (ed) The non-linear field theories of mechanics, 3rd edn. Springer, New York

  63. Visintin A (2015) PDES with hysteresis 30 years later. Discrete Contin Dyn Syst Ser S 8:793–816

    Article  MathSciNet  Google Scholar 

  64. Visintin A (1994) Differential models of hysteresis. Springer, Berlin

    Book  Google Scholar 

  65. Wiedemann G (1886) Magnetic researches. Lond Edinb Dublin Philos Mag J Sci 22:50–70

    Article  Google Scholar 

  66. Yen EH, Van Der Vaart HR (1966) On measurable functions, continuous functions and some related concepts. Am Math Mon 73:991–993

    MathSciNet  MATH  Google Scholar 

  67. Duhem P (1917) Sa vie–Ses œuvres. Gauthier–Villars, Mémoires de la Sociétédes sciences physiques et naturelles de Bordeaux, 7éme série, Paris, France

Download references

Funding

This study was funded by the Spanish Ministry of Economy, Industry and Competitiveness (Grant Number DPI2016-77407-P (AEI/FEDER, UE).

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Fayçal Ikhouane.

Ethics declarations

Conflict of interest

The author declares that they have no conflict of interest.

Additional information

The original version of this article was revised.

Appendices

Appendix

1.1 On the Existence and Uniqueness of Solutions of Differential Equations

In this section we present some existence and uniqueness theorems for the solutions of ordinary differential equations. To this end, let \(\mathcal {D}\) be a domain, that is an open connected subset of \(\mathbb {R} \times \mathbb {R}^{n}\) where \(n>0\) is an integer. Let \((t_0,x_0) \in \mathcal {D}\) and let \(a,b \in \; ]0,\infty [.\) Define the parallelepiped \(Q_{a,b}\) by

$$\begin{aligned} Q_{a,b}=\left\{ (t,w) \in \mathbb {R} \times \mathbb {R}^{n} \mid |t-t_0| \le a,|w-x_0| \le b\right\} . \end{aligned}$$
(127)

We say that the map \(F:\mathcal {D} \rightarrow \mathbb {R}^{n}\) satisfies the Carathéodory conditions on the domain \(\mathcal {D}\) if Conditions (i)–(iii) hold on any parallelepiped \(Q_{a,b} \subset \mathcal {D}\) [61, p. 68].

  1. (i)

    The function F is defined and continuous in w for almost all t;

  2. (ii)

    the function F is measurable in t for each fixed w;

  3. (iii)

    for each \(Q_{a,b} \subset \mathcal {D}\) there exists a measurable function \(m_{Q_{a,b}} \in L^1\big ([t_0-a,t_0+a],\mathbb {R}\big )\) such that

    $$\begin{aligned}&|F(t,w)| \le m_{Q_{a,b}}(t), \; \forall w \in \mathbb {R}^{n} \text { and for almost all } \nonumber \\&t \in [t_0-a,t_0+a] \text { satisfying } (t,w) \in Q_{a,b}. \end{aligned}$$
    (128)

Now, consider the differential equation

$$\begin{aligned} \dot{x}(t)&= F\big (t,x(t)\big ), \end{aligned}$$
(129)
$$\begin{aligned} x(t_0)&= x_0, \end{aligned}$$
(130)

where \(F:\mathcal {D} \rightarrow \mathbb {R}^{n}\) satisfies the Carathéodory conditions on the domain \(\mathcal {D} \subset \mathbb {R} \times \mathbb {R}^{n}\) and \((t_0,x_0) \in \mathcal {D}\).

Theorem 10

[61, p. 68] The differential equation (129)–(130) has a solution on some nonempty open interval \(I \ni t_0,\) in the sense that there exists an absolutely continuous function \(x:I\rightarrow \mathbb {R}^{n}\) such that the following properties (i)–(iii) are satisfied.

  1. (i)

    The initial condition (130) holds;

  2. (ii)

    \(\forall t \in I\) we have \(\big (t,x(t)\big ) \in \mathcal {D}\);

  3. (iii)

    and the differential equation (129) is satisfied almost everywhere in I.

A lower bound on the size of the interval I is obtained by solving the inequality

$$\begin{aligned} \int _{t_0-c}^{t_0+c} m_{Q_{a,b}}(t) \, \text {d}t \le b, \end{aligned}$$
(131)

where \(a,b \in \; ]0,\infty [\) are chosen so that \((t_0,x_0) \in Q_{a,b} \subset \mathcal {D}\). Observe that the function \(c \rightarrow \int _{t_0-c}^{t_0+c} m_{Q_{a,b}}(t) \, \text {d}t\) is continuous and is zero at \(c=0\). This implies that there exists at least a \(0<c\le a\) such that (131) holds. Then we have \(]t_0-c,t_0+c[ \,\subset I\) [61, p. 69].

Theorem 11

[61, p. 70 and p. 80] Assume that \(F:\mathcal {D}\rightarrow \mathbb {R}^{n}\) satisfies the Carathéodory conditions on the domain \(\mathcal {D}\). Let x be a solution of the differential equation (129)–(130) defined on some interval I. Then x may be extended as a solution of (129)–(130) to a maximal interval of existence \(]\omega _-,\omega _+[\) and \(\big (t,x(t)\big ) \rightarrow \partial \mathcal {D}\) as \(t \rightarrow \omega _{\pm }\), where \(\partial \mathcal {D}\) is the boundary of \(\mathcal {D}\).

Theorem 12

[29, p. 5] Assume that \(F:\mathcal {D}\rightarrow \mathbb {R}^{n}\) satisfies the Carathéodory conditions on the domain \(\mathcal {D}\). Assume that there exists a function \(l \in L^1\big (J , \mathbb {R}_+ \big )\) for every finite interval \(J \subset \mathbb {R}\) which satisfies the following. For almost all \(t \in \mathbb {R}\) and \(\forall w_1,w_2 \in \mathbb {R}^{n}\) such that \( (t,w_1),(t,w_2) \in \mathcal {D}\) we have

$$\begin{aligned} \big |F(t,w_1)-F(t,w_2)\big | \le l(t)|w_1-w_2|. \end{aligned}$$
(132)

Then in the domain \(\mathcal {D}\) there exists at most one solution to the differential equation (129)–(130).

The local Lipschitz condition (132) can be relaxed as follows [29, p. 5].

$$\begin{aligned} \big (F(t,w_1)-F(t,w_2)\big )\cdot&\, (w_1-w_2) \le l_{1}(t)|w_1-w_2|^2, \nonumber \\&\text { for almost all } t \ge t_0, \end{aligned}$$
(133)
$$\begin{aligned} \big (F(t,w_1)-F(t,w_2)\big )\cdot&\,(w_1-w_2) \ge -l_{2}(t)|w_1-w_2|^2, \nonumber \\&\text { for almost all } t \le t_0, \end{aligned}$$
(134)

where the product is understood as the scalar product if \(F(t,w_1),F(t,w_2),w_1,w_2\) are vectors; the functions \(l_{1},l_{2} \in L^1\big (J , \mathbb {R}_+ \big )\) for every finite interval \(J \subset \mathbb {R}\), and \(w_1,w_2 \in \mathbb {R}^{n}\) are such that \( (t,w_1),(t,w_2) \in \mathcal {D}\).

Finally we provide a result we could not find in the literature, and which is useful to the present paper.

Lemma 12

Suppose that the application \(F:\mathbb {R} \times \mathbb {R} \rightarrow \mathbb {R}\) satisfies the Carathéodory conditions on the domain \(\mathbb {R}^2\). Assume that there exists \(k \in [0,\infty [\) such that

$$\begin{aligned} \big (F(t,w_1)-F(t,w_2)\big )\cdot (w_1-w_2) \le k|w_1-w_2|^2, \nonumber \\ \text {for almost all } t \ge t_0, \forall w_1,w_2 \in \mathbb {R}. \end{aligned}$$
(135)

Then the differential equation (129)–(130) has exactly one solution defined on \([t_0,\infty [\).

Proof

From Theorems 10, 11, and 12 it follows that there exists a unique solution x to the differential equation (129)–(130) defined on a maximal interval of existence \([t_0,\omega _+[\) where \(\omega _+ \in \; ]t_0,\infty ]\). Assume that \(\omega _+ < \infty \), and let \(w \in \mathbb {R}\) be fixed. It comes from Theorem 11 that \(\exists \, t_w \in \;]t_0,\omega _+[\) such that \(\forall t \in [t_w,\omega _+[\) we have \(|x(t)|> |w|\). Consider the case \(\forall t \in [t_w,\omega _+[,\,x(t)> |w| \ge w\) (a similar proof holds for the case \(\forall t \in [t_w,\omega _+[,\,x(t) <- |w|\)). Then Inequality (135) leads to

$$\begin{aligned} F(t,x(t)) \le F(t,w)+k\big (x(t)-w\big ), \text { for almost all } t \in [t_w,\omega _+[. \end{aligned}$$
(136)

Integrating both sides of (136) on the time interval \([t_w,t]\) it follows that

$$\begin{aligned} |x(t)|&=x(t)=x(t_w)+\int _{t_w}^t F(s,x(s))\,\text {d}s \nonumber \\&\le C+k \int _{t_w}^t |x(s)|\,\text {d}s, \forall t \in [t_w,\omega _+[, \\C&= x(t_w)+\int _{t_w}^{\omega _+}|F(s,w)|\,\text {d}s +k|w|(\omega _+-t_w) <\infty . \nonumber \end{aligned}$$
(137)

Using Gronwall’s lemma [32, p. 24] it comes from Inequality (137) that

$$\begin{aligned} |x(t)| \le Ce^{t-t_w} \le Ce^{\omega _+ -t_w}, \forall t\in [t_w,\omega _+[. \end{aligned}$$
(138)

Inequality (138) contradicts the fact that \(|x(t)| \rightarrow \infty \) as \(t \rightarrow \omega _+\).

Proof of Lemma 13

Lemma 13

Let \(u \in W^{1,\infty }(\mathbb {R}_+,\mathbb {R})\) be non constant. There exists a unique function \(v_{u}\in L^{\infty }\left( I_{u},\mathbb {R}\right) \) that is defined by \(v_{u}\circ \rho _{u}=\dot{u}\). Moreover, \(\Vert v_u\Vert _{I_u} \le \Vert \dot{u}\Vert \) and \(v_{u}\) is nonzero almost everywhere on \(I_u\).

Proof

The operator \(\Delta _-\) defined in Section 11.2 is causal and satisfies Assumption 3. Using Lemma 3 it follows that \(v_{u}\in L^{\infty }\left( I_{u},\mathbb {R}\right) \) and \(\Vert v_u\Vert _{I_u} \le \Vert \dot{u}\Vert \). Now, define the following sets:

$$\begin{aligned} A&=\{\varrho \in I_u \mid v_u(\varrho )=0\}, \\ B&=\{t \in \mathbb {R}_+ \mid \dot{u}(t) =0\}, \\ B_1&=\{t \in \mathbb {R}_+ \mid \dot{\rho }_u(t) \text{ is } \text{ not } \text{ defined } \text{ at } t\}, \\ B_2&=\{t \in \mathbb {R}_+ \mid \dot{u}(t) \text{ is } \text{ defined, } \dot{\rho }_u(t) \text { is defined, and} \\&\quad \quad |\dot{u}(t)| \ne \dot{\rho }_u(t)\}, \\ C&=\{t \in \mathbb {R}_+ \mid \dot{\rho }_u(t) =0\}. \end{aligned}$$

Since \(\rho _u\) is absolutely continuous on \(\mathbb {R}_+\), we get from [45, Corollary 3.41] that \(\mu (B_1)=0\). Since \(\dot{u}\in L^{\infty }\left( \mathbb {R}_+,\mathbb {R}\right) \) we get from [45, Lemma 3.31] that \(\dot{\rho }_u = |\dot{u}|\) almost everywhere on \(\mathbb {R}_+\), which implies that \(\mu (B_2)=0\). Also, from [45, Corollary 3.14] it follows that \(\mu \big (\rho _u(C)\big )=0.\) Since \(\rho _u\) is absolutely continuous on \(\mathbb {R}_+\), and since \(\mu (B_1)=\mu (B_2)=0\) it follows from [45, Corollary 3.41] that \(\mu \big (\rho _u(B_1)\big )=\mu \big (\rho _u(B_2)\big )=0\). Now, observe that \(B \subset C \cup B_1 \cup B_2\), thus \(\rho _u(B) \subset \rho _u(C) \cup \rho _u(B_1) \cup \rho _u(B_2)\) which implies that \(\mu \big (\rho _u(B)\big )=0\). Since \(A=\rho _u(B)\) it follows that \(\mu (A)=0\).

Proof of Theorem 8

We get from Equation (68) that \(\exists \delta _1>0\) such that \(\forall w \in (0,\delta _1)\) we have \(\left| \bar{g}_1(w)-1\right| <\frac{1}{2}\), and \(\exists \delta _2>0\) such that \(\forall w \in ( -\delta _2,0)\) we have \(\left| \bar{g}_2(w)+1\right| <\frac{1}{2}\). Define

$$\begin{aligned} \gamma _0 = \frac{\Vert \dot{u}\Vert }{\min (\delta _1,\delta _2)}. \end{aligned}$$
(139)

Observe that \(0<\gamma _0<\infty \) since \(u \in \Lambda _{u_{\min },u_{\max },\alpha _1,T}\). Let \(\gamma \in \; ]\gamma _0,\infty [\) be fixed, and define \(x_\gamma = \mathcal {H}_s(u \circ s_\gamma ,x_{0})\). From Equations (63) and (64) we get

$$\begin{aligned} {\begin{matrix} x_\gamma (t)={}&{}x_0+\int _0^t g_1\big (\dot{u}_\gamma (\tau )\big ) \big ( A_1 x_\gamma (\tau )+ B_1u_\gamma (\tau )+ E_1 \big ) \\ &{} +g_2\big (\dot{u}_\gamma (\tau )\big ) \big ( A_2 x_\gamma (\tau )+ B_2 u_\gamma (\tau )+ E_2 \big ) \,\text {d}\tau , \forall t \in \mathbb {R}_+ \end{matrix}} \end{aligned}$$
(140)

where \(u_\gamma = u \circ s_\gamma \). Consider the change of variable \(\tau ^\prime = \frac{\tau }{\gamma }\), then

$$\begin{aligned} {\begin{matrix} x_\gamma (t)={}&{}x_0+\gamma \int _0^{\frac{t}{\gamma }} g_1\left( \frac{\dot{u}(\tau ^\prime )}{\gamma }\right) \big [ A_1 x_\gamma (\gamma \tau ^\prime )+ B_1 u(\tau ^\prime ) + E_1 \big ] \\ &{} +g_2\left( \frac{\dot{u}(\tau ^\prime )}{\gamma }\right) \times \big [ A_2 x_\gamma (\gamma \tau ^\prime )+ B_2 u(\tau ^\prime )+ E_2 \big ] \, \text {d}\tau ^\prime , \\ {} &{} \forall t \in \mathbb {R}_+. \end{matrix}} \end{aligned}$$
(141)

Define \(\sigma =\frac{t}{\gamma }\) and \(z:\mathbb {R}_+ \rightarrow \mathbb {R}\) by \(z(\sigma )=x_\gamma (\gamma \sigma ), \forall \sigma \in \mathbb {R}_+\); then

$$\begin{aligned} {\begin{matrix} z(\sigma )={}&{}x_0+\gamma \int _0^{\sigma } g_1\left( \frac{\dot{u}(\tau ^\prime )}{\gamma }\right) \big (A_1 z(\tau ^\prime )+ B_1 u(\tau ^\prime )+ E_1 \big ) \\ &{} +g_2\left( \frac{\dot{u}(\tau ^\prime )}{\gamma }\right) \big (A_2 z(\tau ^\prime )+ B_2 u(\tau ^\prime )+ E_2 \big ) \, \text {d}\tau ^\prime , \\ &{} \forall \sigma \in \mathbb {R}_+. \end{matrix}} \end{aligned}$$
(142)

For any \(m \in \mathbb {N}\) define \(z_{m}:[0,T] \rightarrow \mathbb {R}\) by

$$\begin{aligned} z_{m}(\sigma )=z(\sigma +mT),\forall \sigma \in [0,T]. \end{aligned}$$
(143)

The objective of the following analysis is to show that the sequence \(\{z_m\}_{m \in \mathbb {N}}\) converges in the Banach space \(C^0\left( [0,T],\mathbb {R}\right) \) endowed with the norm \(\Vert \cdot \Vert _{[0,T]}\). To this end, we prove that \(\{z_m\}_{m \in \mathbb {N}}\) is a Cauchy sequence. For any \(m_1,m_2 \in \mathbb {N}\) define

$$\begin{aligned} z_{m_1,m_2}=z_{m_1}-z_{m_2}. \end{aligned}$$
(144)

Owing to the T–periodicity of both u and \(\dot{u}\) it follows from Equations (142)–(144) that

$$\begin{aligned} \dot{z}_{m_1,m_2}(\sigma )=&\;\gamma \left( A_1 g_1\left( \frac{\dot{u}(\sigma )}{\gamma }\right) +A_2 g_2\left( \frac{\dot{u}(\sigma )}{\gamma }\right) \right) z_{m_1,m_2}(\sigma ),\nonumber \\&\forall \sigma \in \;]0,\alpha _1[\;\cup \;]\alpha _1,T[. \end{aligned}$$
(145)

Let \(\sigma \in (0,\alpha _1)\) then \(\dot{u}(\sigma ) \ge 0\) since \(u \in \Lambda _{u_{\min },u_{\max },\alpha _1,T}\). We study two cases: \(\dot{u}(\sigma )> 0\) and \(\dot{u}(\sigma ) = 0\).

Case \(\dot{u}(\sigma )> 0\). Since \(0<\frac{\dot{u}(\sigma )}{\gamma }< \frac{\Vert \dot{u}\Vert }{\gamma _0}\le \delta _1\) it follows that \(\left| \bar{g}_1\left( \frac{\dot{u}(\sigma )}{\gamma }\right) -1\right| <\frac{1}{2}\) which, using Equation (66), leads to

$$\begin{aligned} \frac{3A_1}{2}\dot{u}(\sigma ) \le \gamma A_1 g_1\left( \frac{\dot{u}(\sigma )}{\gamma }\right) \le \frac{A_1}{2}\dot{u}(\sigma ). \end{aligned}$$
(146)

Case \(\dot{u}(\sigma ) = 0\). In this case, Inequality (146) holds by definition of the function \(g_1\). That is we have

$$\begin{aligned} \frac{3A_1}{2}\dot{u}(\sigma ) \le \gamma A_1 g_1\left( \frac{\dot{u}(\sigma )}{\gamma }\right) \le \frac{A_1}{2}\dot{u}(\sigma ), \forall \sigma \in \; ]0,\alpha _1[. \end{aligned}$$
(147)

Similarly, it can be shown that

$$\begin{aligned} \frac{3A_2}{2}\dot{u}(\sigma ) \le \gamma A_2 g_2\left( \frac{\dot{u}(\sigma )}{\gamma }\right) \le \frac{A_2}{2}\dot{u}(\sigma ), \forall \sigma \in \; ]\alpha _1,T[. \end{aligned}$$
(148)

Now, define the function \(V:[0,T] \rightarrow \mathbb {R}\) by

$$\begin{aligned} V(\sigma )=\frac{1}{2}z^2_{m_1,m_2}(\sigma ), \forall \sigma \in [0,T]. \end{aligned}$$
(149)

Then, V is continuous on [0, T] and is \(C^1\) on \(]0,\alpha _1[\;\cup \; ]\alpha _1,T[\). From Eq. (145) we obtain

$$\begin{aligned} \dot{V}(\sigma )=&\;2\gamma \left( A_1 g_1\left( \frac{\dot{u}(\sigma )}{\gamma }\right) +A_2 g_2\left( \frac{\dot{u}(\sigma )}{\gamma }\right) \right) V(\sigma ),\nonumber \\&\forall \sigma \in \;]0,\alpha _1[ \;\cup \;]\alpha _1,T[. \end{aligned}$$
(150)

Combining Eqs. (150), (147) and (148) it follows that

$$\begin{aligned} \dot{V}(\sigma )&\le A_1\dot{u}(\sigma )V(\sigma ), \forall \sigma \in \;]0,\alpha _1[,\end{aligned}$$
(151)
$$\begin{aligned} \dot{V}(\sigma )&\le A_2\dot{u}(\sigma )V(\sigma ), \forall \sigma \in \;]\alpha _1,T[. \end{aligned}$$
(152)

Define the continuous function \(W:[0,\alpha _1] \rightarrow \mathbb {R}\) as being the solution of the following differential equation

$$\begin{aligned} \dot{W}(\sigma )&= A_1\dot{u}(\sigma )W(\sigma ), \forall \sigma \in \; ]0,\alpha _1[,\end{aligned}$$
(153)
$$\begin{aligned} W(0)&=V(0). \end{aligned}$$
(154)

Integrating (153)–(154) gives

$$\begin{aligned} W(\sigma )=V(0)\exp \left( \frac{A_1}{\gamma }\big (u(\sigma )-u_{\min } \big )\right) , \forall \sigma \in [0,\alpha _1]. \end{aligned}$$
(155)

Using the Comparison Lemma [42, p. 102] it comes from Eqs. (151), (153), (154), and (155) that

$$\begin{aligned} V(\alpha _1) \le W(\alpha _1) = V(0)\exp \big (A_1\left( u_{\max }-u_{\min } \right) \big ). \end{aligned}$$
(156)

Using a similar argument on the interval \([\alpha _1,T]\) it follows that

$$\begin{aligned} V(T) \le W(\alpha _1)\exp \big (A_2\left( u_{\min }-u_{\max } \right) \big ). \end{aligned}$$
(157)

As a conclusion, we have proved that

$$\begin{aligned} V(T)&\le r V(0), \end{aligned}$$
(158)
$$\begin{aligned} 0<r&=\exp \big ((A_1-A_2)(u_{\max }-u_{\min })\big ) <1, \end{aligned}$$
(159)
$$\begin{aligned} \Vert V\Vert _{[0,T]}&\le V(0). \end{aligned}$$
(160)

Note that (160) is due to the inequality \(\dot{V}(\sigma ) \le 0\), \(\forall \sigma \in \;]0,\alpha _1[\;\cup \;]\alpha _1,T[\) b‘ecause of Inequalities (151)–(152).

Combining Eqs. (158), (149), (144), and (143) we get

$$\begin{aligned} \Big [ z\big ( (m_1+1)T\big )-z\big ( (m_2+1)T\big )\Big ]^2&\;\le r \Big [ z\big ( m_1T\big ) -z\big ( m_2T\big )\Big ]^2,\nonumber \\&\forall m_1,m_2 \in \mathbb {N}. \end{aligned}$$
(161)

An argument by induction shows that from (161) we get

$$\begin{aligned} {\begin{matrix} V(0)=&{}\;\frac{1}{2}\Big [ z\big ( m_1T\big )-z\big ( m_2T\big )\Big ]^2 \\ &{}\le \frac{1}{2}r^{\min (m_1,m_2)} \Big [ z\left( 0\right) -z\big (| m_2-m_1|T\big )\Big ]^2 \\ &{} \le 2 r^{\min (m_1,m_2)} \Vert z\Vert ^2, \forall m_1,m_2 \in \mathbb {N}. \end{matrix}} \end{aligned}$$
(162)

Observe that, owing to Theorem 5, we have \(\Vert z\Vert <\infty \). Hence, from Eqs. (162), (160), (149), (144), and (159) it comes that \(\{z_m\}_{m \in \mathbb {N}}\) is a Cauchy sequence. Therefore there exists \(z_{\infty } \in C^0\left( [0,T],\mathbb {R}\right) \) such that \(\lim _{m \rightarrow \infty } \Vert z_m - z_{\infty }\Vert _{[0,T]}=0\). Thus we get \(\lim _{m \rightarrow \infty }\left| z_m(0)-z_{\infty }(0)\right| =0\) and

\(\lim _{m \rightarrow \infty } \left| z_m(T)-z_{\infty }(T)\right| =0\). Note that \(z_m(0)=z(mT)\) and \(z_m(T)=z\big ( (m+1)T \big )\) by (143). Take \(m_1=m\) and \(m_2=m+1\) in Inequality (162). Then we get \(\lim _{m \rightarrow \infty } \left| z(mT)-z\big ( (m+1)T \big )\right| =0\). All these facts show that we have

$$\begin{aligned} z_{\infty }(0)=z_{\infty }(T). \end{aligned}$$
(163)

Combining Eqs. (142) and (143) it comes that

$$\begin{aligned} z_m(\sigma )&\;=z_m(0)+\gamma \int _0^{\sigma } g_1\left( \frac{\dot{u}(\tau )}{\gamma }\right) \big [A_1 z_m(\tau )+B_1 u(\tau ) + E_1 \big ] \nonumber \\&\quad +g_2\left( \frac{\dot{u}(\tau )}{\gamma }\right) \big (A_2 z_m(\tau )+B_2 u(\tau )+ E_2 \big ) \, \text {d}\tau ,\nonumber \\&\forall \sigma \in [0,T], \forall m \in \mathbb {N}. \end{aligned}$$
(164)

Note that \(\Vert z_m\Vert \le \Vert z\Vert < \infty \). Also, \(\left| \frac{\dot{u}(\tau )}{\gamma } \right| \le \frac{\Vert \dot{u}\Vert }{\gamma _0}\) so that, by the continuity of the functions \(g_1\) and \(g_2\) we have \(\left| g_1\left( \frac{\dot{u}(\tau )}{\gamma }\right) \right| \le k_1\) and \(\left| g_2\left( \frac{\dot{u}(\tau )}{\gamma }\right) \right| \le k_2\), where \(k_1, k_2 \in \mathbb {R}_+\) are independent of \(\tau \) and m. This means that the term under the integral in Eq. (164) is bounded by a constant independent of \(\tau \) and m. Using the Lebesgue Dominated Convergence Theorem it follows from (164) that

$$\begin{aligned} z_\infty (\sigma )&=z_\infty (0)+\gamma \!\int _0^{\sigma }\!\! g_1\left( \frac{\dot{u}(\tau )}{\gamma }\right) \big [ A_1 z_\infty (\tau )+B_1 u(\tau )+E_1 \big ] \nonumber \\ {}&\quad +g_2\left( \frac{\dot{u}(\tau )}{\gamma }\right) \big [ A_2 z_\infty (\tau )+B_2 u(\tau )+E_2 \big ] \,\text {d}\tau ,\nonumber \\&\forall \sigma \in [0,T]. \end{aligned}$$
(165)

Define \(\bar{z}_\gamma :\mathbb {R}_+ \rightarrow \mathbb {R}\) by

$$\begin{aligned} \bar{z}_\gamma (\sigma +mT)=z_\infty (\sigma ),\forall \sigma \in [0,T],\forall m \in \mathbb {N}. \end{aligned}$$
(166)

Then it comes from Eqs. (166), (165) and (163) that \(\bar{z}_\gamma \) is T–periodic and

$$\begin{aligned} \bar{z}_\gamma (\sigma )&=\bar{z}_\gamma (0)+\gamma \int _0^{\sigma } g_1\left( \frac{\dot{u}(\tau )}{\gamma }\right) \big [ A_1 \bar{z}_\gamma (\tau )+ B_1 u(\tau )+E_1 \big ] \nonumber \\&\quad +g_2\left( \frac{\dot{u}(\tau )}{\gamma }\right) \big [ A_2 \bar{z}_\gamma (\tau )+ B_2 u(\tau )+E_2 \big ]\, \text {d}\tau ,\nonumber \\&\forall \sigma \in \mathbb {R}_+. \end{aligned}$$
(167)

As a conclusion, we have proved that there exists

$$\begin{aligned} x_{0,\gamma }=\bar{z}_\gamma (0) \end{aligned}$$
(168)

such that

$$\begin{aligned} \mathcal {H}_s(u \circ s_\gamma ,x_{0,\gamma })=\bar{z}_\gamma \circ s_\gamma \end{aligned}$$
(169)

is \(T \gamma \)–periodic.

To prove the uniqueness of \(x_{0,\gamma }\) we use an argument similar to the one used for the proof of the existence. Take \(\gamma> \gamma _0\) and suppose that there exists \(x^\prime _{0,\gamma }\) such that \( \mathcal {H}_s(u \circ s_\gamma ,x^\prime _{0,\gamma })\) is \(T \gamma \)–periodic. Define \(\bar{z}_\gamma ^\prime : \mathbb {R}_+ \rightarrow \mathbb {R}\) by \(\bar{z}_\gamma ^\prime =\mathcal {H}_s(u \circ s_\gamma ,x^\prime _{0,\gamma }) \circ s_{\frac{1}{\gamma }}\). Then, \(\bar{z}_\gamma ^\prime (0)=x^\prime _{0,\gamma }\) and \(\bar{z}_\gamma ^\prime \) satisfies Eq. (167) with \(\bar{z}_\gamma \) replaced by \(\bar{z}_\gamma ^\prime \). Considering the difference \(\varepsilon =\bar{z}_\gamma -\bar{z}_\gamma ^\prime \) it follows that \(\varepsilon \) satisfies Eq. (145) with \(z_{m_1,m_2}\) replaced by \(\varepsilon \). A function V can be defined as in Eq. (149) with \(z_{m_1,m_2}\) replaced by \(\varepsilon \) which leads to Inequality (158). Since \(V(0)=V(T)\) owing to the T–periodicity of V, it follows that \(V(0)=0\) as V is nonnegative. Thus \(x^\prime _{0,\gamma }=x_{0,\gamma }\).

Proof of Theorem 9

Let \(\gamma \in \;]\gamma _0,\infty [\) where \(\gamma _0\) is given by Eq. (139). From Eq. (147) it follows that

$$\begin{aligned} \frac{3A_1}{2}\big (u_{\max }-u_{\min } \big ) \le \int _0^{\tau }\frac{3A_1}{2}\dot{u}(t)\text {d}t \le \int _0^{\tau }\gamma A_1 g_1\left( \frac{\dot{u}(t)}{\gamma }\right) \text {d}t, \end{aligned}$$
(170)

and

$$\begin{aligned} \left| \gamma A_1 g_1\left( \frac{\dot{u}(\tau )}{\gamma }\right) \right| \le \frac{3|A_1|}{2}\Vert \dot{u}\Vert , \forall \tau \in \;]0,\alpha _1[. \end{aligned}$$
(171)

Also, From Eq. (148) it follows that

$$\begin{aligned} \frac{3A_2}{2}\big (u_{\min }-u_{\max } \big ) \le \int _{\alpha _1}^{\tau }\frac{3A_2}{2}\dot{u}(t)\text {d}t \le \int _{\alpha _1}^{\tau }\gamma A_2 g_2\left( \frac{\dot{u}(t)}{\gamma }\right) \text {d}t, \end{aligned}$$
(172)

and

$$\begin{aligned} \left| \gamma A_2 g_2\left( \frac{\dot{u}(\tau )}{\gamma }\right) \right| \le \frac{3A_2}{2}\Vert \dot{u}\Vert , \forall \tau \in \; ]\alpha _1,T[. \end{aligned}$$
(173)

Equations (170)–(173) show that we can apply the Lebesgue Dominated Convergence Theorem in (103) so that we get

$$\begin{aligned} \lim _{\gamma \rightarrow \infty } \bar{z}_\gamma (0)= \bar{z}(0)=\theta . \end{aligned}$$
(174)

Observe that using the same theorem we can show that \(\forall \sigma \in [0,T]\) we have \(\lim _{\gamma \rightarrow \infty } |\bar{z}_\gamma (\sigma )-\bar{z}(\sigma )|=0\). However, this simple convergence does not imply Theorem 9; we need to prove the uniform convergence of \(\bar{z}_\gamma \) to \(\bar{z}\) on the interval [0, T]. This is the aim of the following analysis.

Inequalities (170)–(173) along with Eqs. (99), (101) and (102) lead to

$$\begin{aligned} \Vert \bar{z}_\gamma \Vert _{[0,T]} \le c_1, \forall \gamma \in \; ]\gamma _0,\infty [ \end{aligned}$$
(175)

where \(c_1 \in \mathbb {R}_+\) is independent of \(\gamma \).

On the other hand, it can be checked that Eqs. (93), (94), (90), (104), (105) lead to

$$\begin{aligned} \dot{\bar{z}}(\sigma )&= \dot{u}(\sigma ) \left( A_1 \bar{z}(\sigma )+ B_1 u(\sigma )+ E_1 \right) , \; \forall \sigma \in \; ]0,\alpha _1[, \end{aligned}$$
(176)
$$\begin{aligned} \dot{\bar{z}}(\sigma )&= \dot{u}(\sigma ) \left( A_2 \bar{z}(\sigma )+ B_2 u(\sigma )+ E_2 \right) , \; \forall \sigma \in \;]\alpha _1,T[. \end{aligned}$$
(177)

Define the function \(V:[0,T] \rightarrow \mathbb {R}\) by the relation

$$\begin{aligned} V_\gamma (\sigma )=\frac{1}{2}\big (\bar{z}(\sigma )- \bar{z}_\gamma (\sigma )\big )^2, \forall \sigma \in [0,T]. \end{aligned}$$
(178)

Take \(\sigma \in \;]0,\alpha _1[\), then it comes from Eqs. (167) and (176) that

$$\begin{aligned} \dot{V}_\gamma (\sigma )=&\,\big (\bar{z}(\sigma )- \bar{z}_\gamma (\sigma )\big )\Big [\dot{u}(\sigma )\big (A_1 \bar{z}(\sigma )+ B_1 u(\sigma )+ E_1 \big ) \nonumber \\&-\gamma g_1\left( \frac{\dot{u}(\sigma )}{\gamma }\right) \big (A_1 \bar{z}_\gamma (\sigma ) + B_1 u(\sigma )+ E_1 \big )\Big ]\nonumber \\ =&\,\big (\bar{z}(\sigma )- \bar{z}_\gamma (\sigma )\big )\big ( B_1u(\sigma )+E_1\big ) \left( \dot{u}(\sigma )-\gamma g_1\left( \frac{\dot{u}(\sigma )}{\gamma }\right) \right) \nonumber \\ {}&+A_1\big (\bar{z}(\sigma )- \bar{z}_\gamma (\sigma )\big ) \left( \dot{u}(\sigma )-\gamma g_1\left( \frac{\dot{u}(\sigma )}{\gamma }\right) \right) \bar{z}(\sigma ) \\&+2A_1\gamma g_1\left( \frac{\dot{u}(\sigma )}{\gamma }\right) V_\gamma (\sigma ),\forall \sigma \in \;]0,\alpha _1[, \forall \gamma> \gamma _0. \nonumber \end{aligned}$$
(179)

Let \(\varepsilon>0\). From Eqs. (66) and (68) it follows that \(\exists \delta _\varepsilon>0\) such that \(\forall w \in \;]0,\delta _\varepsilon [\) we have \(\left| \bar{g}_1(w)-1\right| <\frac{\varepsilon }{\Vert \dot{u}\Vert }\). Thus, \(\exists \gamma _\varepsilon =\min \left( \gamma _0,\frac{\Vert \dot{u}\Vert }{\delta _\varepsilon } \right) \) such that \(\forall \gamma> \gamma _\varepsilon \) we have

$$\begin{aligned} \left| \gamma g_1\left( \frac{\dot{u}(\sigma )}{\gamma }\right) -\dot{u}(\sigma )\right| \le \varepsilon , \forall \sigma \in \;]0,\alpha _1[. \end{aligned}$$
(180)

Combining Eqs. (178)–(180) along with Inequalities (175) and (147) it comes that

$$\begin{aligned} \dot{V}_\gamma (\sigma ) \le A_1 \dot{u}(\sigma ) V_\gamma (\sigma )+c_2 \varepsilon \sqrt{V_\gamma (\sigma )}, \forall \sigma \in \;]0,\alpha _1[, \forall \gamma> \gamma _\varepsilon . \end{aligned}$$
(181)

where \(c_2 \in \mathbb {R}_+\) is independent of \(\gamma \). Define the continuous function \(W:[0,\alpha _1] \rightarrow \mathbb {R}_+\) as the solution of the following differential equation

$$\begin{aligned} \dot{W}(\sigma )&= A_1\dot{u}(\sigma )W(\sigma )+c_2\varepsilon \sqrt{W(\sigma )}, \forall \sigma \in \;]0,\alpha _1[, \end{aligned}$$
(182)
$$\begin{aligned} W(0)&=V(0). \end{aligned}$$
(183)

Integrating (182)–(183) gives

$$\begin{aligned} W(\sigma )&=e^{A_1u(\sigma )}\Big ( \sqrt{V(0)}e^{-\frac{A_1}{2}u_{\min }} +\frac{c_2}{2} \varepsilon \int _0^{\sigma } e^{-\frac{A_1}{2}u(\tau )} \text {d}\tau \Big )^2,\nonumber \\&\forall \sigma \in [0,\alpha _1],\nonumber \\&\le \; e^{A_1u_{\min }}\Big ( \sqrt{V(0)}e^{-\frac{A_1}{2}u_{\min }} +\frac{c_2}{2} \varepsilon \int _0^{\alpha _1} e^{-\frac{A_1}{2}u(\tau )} \text {d}\tau \Big )^2,\nonumber \\&\forall \sigma \in [0,\alpha _1]. \end{aligned}$$
(184)

Using the Comparison Lemma [42, p. 102] it follows from (181)–(184) that

$$\begin{aligned} V_\gamma (\sigma ) \le&\;e^{A_1u_{\min }}\Big ( \sqrt{V(0)}e^{-\frac{A_1}{2}u_{\min }} +\frac{c_2}{2} \varepsilon \int _0^{\alpha _1} e^{-\frac{A_1}{2}u(\tau )} \text {d}\tau \Big )^2,\nonumber \\&\forall \sigma \in [0,\alpha _1],\forall \gamma> \gamma _\varepsilon . \end{aligned}$$
(185)

Equations (185), (174) and (178) show that

\(\lim _{\gamma \rightarrow \infty }\Vert V_\gamma \Vert _{[0,\alpha _1]}=0\). A similar argument on the interval \([\alpha _1,T]\) shows that \(\lim _{\gamma \rightarrow \infty }\Vert V_\gamma \Vert _{[0,T]}=0\). The uniform convergence of \(\bar{z}_\gamma \) (restricted to the interval [0, T]) to \(\bar{z}\) has thus been demonstrated, which completes the proof.

Proof of Lemma 8

(i) \(\Rightarrow \) (ii). From Eq. (80) and \(C\ne 0\) it comes that \(\forall \varrho _1,\varrho _2 \in [0,\rho _u(T)]\) we have \(\varphi _u^\circ (\varrho _1)=\varphi _u^\circ (\varrho _2) \Leftrightarrow x_u^\circ (\varrho _1) = x_u^\circ (\varrho _2)\). Condition (i) implies that \(\forall \nu \in [u_{\min },u_{\max }]\) we have \(\xi _1(\nu )=\xi _2(\nu )\). Therefore \(\forall \nu \in \;]u_{\min },u_{\max }[\) we have \(\dot{\xi }_1(\nu )=\dot{\xi }_2(\nu ).\) Thus we get from (91)–(92) that

$$\begin{aligned} \xi _1(\nu )=\xi _2(\nu )=\frac{B_1-B_2}{A_2-A_1}\nu +\frac{E_1-E_2}{A_2-A_1}, \forall \nu \in \;]u_{\min },u_{\max }[. \end{aligned}$$
(186)

Consider the functions \(f_1,f_2,f_3,\mathbf {0}:\;]u_{\min },u_{\max }[\; \rightarrow \mathbb {R}\) defined by \(\forall \nu \in \;]u_{\min },u_{\max }[, f_1(\nu )=1,\) \(f_2(\nu )=\nu, \) \(f_3(\nu )= e^{A_1(\nu -u_{\min })},\) and \(\mathbf {0}(\nu )=0.\) Then Eq. (186) along with (93)–(94) lead to

$$\begin{aligned} {\begin{matrix} &{}\left( \frac{E_1-E_2}{A_2-A_1}+\frac{E_1}{A_1}+\frac{B_1}{A_1^2}\right) f_1+\left( \frac{B_1-B_2}{A_2-A_1}+\frac{B_1}{A_1}\right) f_2\\ &{} -\left( \frac{B_1}{A_1}u_{\min } +\frac{E_1}{A_1}+\frac{B_1}{A_1^2}+\theta \right) f_3=\mathbf {0}, \end{matrix}} \end{aligned}$$
(187)
$$\begin{aligned} {\begin{matrix} &{} \left( \frac{E_1-E_2}{A_2-A_1}+\frac{E_2}{A_2}+\frac{B_2}{A_2^2}\right) f_1+\left( \frac{B_1-B_2}{A_2-A_1}+\frac{B_2}{A_2}\right) f_2 \\ &{}- \left( \frac{B_2}{A_2}u_{\min } +\frac{E_2}{A_2}+\frac{B_2}{A_2^2}+\theta \right) f_3=\mathbf {0}. \end{matrix}} \end{aligned}$$
(188)

Consider the vector space of functions \(\{p:\;]u_{\min },u_{\max }[\; \rightarrow \mathbb {R}\}\) with its usual binary operations of vector addition and scalar multiplication. Then the functions \(f_1,f_2,f_3\) are linearly independent vectors so that, owing to Eqs. (187)–(188), we must have

$$\begin{aligned} \frac{E_1-E_2}{A_2-A_1}+\frac{E_1}{A_1}+\frac{B_1}{A_1^2}&=0, \end{aligned}$$
(189)
$$\begin{aligned} \frac{B_1-B_2}{A_2-A_1}+\frac{B_1}{A_1}&=0,\end{aligned}$$
(190)
$$\begin{aligned} \frac{B_1}{A_1}u_{\min } +\frac{E_1}{A_1}+\frac{B_1}{A_1^2}+\theta&=0,\end{aligned}$$
(191)
$$\begin{aligned} \frac{E_1-E_2}{A_2-A_1}+\frac{E_2}{A_2}+\frac{B_2}{A_2^2}&=0,\end{aligned}$$
(192)
$$\begin{aligned} \frac{B_1-B_2}{A_2-A_1}+\frac{B_2}{A_2}&=0, \end{aligned}$$
(193)
$$\begin{aligned} \frac{B_2}{A_2}u_{\min } +\frac{E_2}{A_2}+\frac{B_2}{A_2^2}+\theta&=0. \end{aligned}$$
(194)

Simple calculations show that Eqs. (189)–(194) lead to (96)–(97).

(ii) \(\Rightarrow \) (i). It can be checked that Eqs. (96)–(97) lead to (189)–(194) so that the opertor \(\mathcal {H}_o\) has a trivial hysteresis loop with respect to all \((u,x_0) \in \Lambda _{u_{\min },u_{\max },\alpha _1,T} \times \mathbb {R}\).

Proof of Lemma 10

Using Eq. (40) the functions \(F_1,F_2: \mathbb {R}^2 \rightarrow \mathbb {R}\) are given by

$$\begin{aligned} F_1(x_1,v) =&\frac{A_1-A_2}{2}x_1 + \frac{B_1-B_2}{2}v+\frac{E_1-E_2}{2}, \end{aligned}$$
(195)
$$\begin{aligned} F_2(x_1,v) =&\frac{A_1+A_2}{2}x_1 + \frac{B_1+B_2}{2}v+\frac{E_1+E_2}{2}. \end{aligned}$$
(196)

Then Assumption 7 holds since \(A_1 \ne A_2\). The anhysteresis function is

$$\begin{aligned} f_{\text {an}}(v)=-\frac{B_1}{A_1}v+\frac{E_2-E_1}{A_1-A_2}, \forall v \in \mathbb {R} \end{aligned}$$
(197)

where (114) has been used. For every pair \((x_0,u_0) \in \mathbb {R}^2\), let \(\omega _{\Phi ,1}(\cdot ,x_0,u_0):[u_0,\infty ) \rightarrow \mathbb {R}\) be the solution z of \(z(\sigma )-x_0=\int _{u_0}^{\sigma } A_1 z(\tau ) + B_1 \tau + E_1\,\text {d}\tau \), for all \(\sigma \in [u_0,\infty [\) and let \(\omega _{\Phi ,2}(\cdot ,x_0,u_0):\;]-\infty ,u_0] \rightarrow \mathbb {R}\) be the solution z of \(z(\sigma )-x_0=\int _{u_0}^{\sigma } A_2 z(\tau )+B_2\tau +E_2\,\text {d}\tau \), for all \(\sigma \in \; ]-\infty ,u_0]\). Then

$$\begin{aligned} \omega _{\Phi ,1}(\sigma ,x_0,u_0) =&\, \frac{A_1B_1u_0+A_1E_1+B_1}{A_1^2}e^{(\sigma -u_0)A_1}\nonumber \\&-\frac{A_1B_1\sigma +A_1E_1+B_1}{A_1^2}\nonumber \\&\,+e^{(\sigma -u_0)A_1}x_0,\forall \sigma \in [u_0,\infty [, \end{aligned}$$
(198)
$$\begin{aligned} \omega _{\Phi ,2}(\sigma ,x_0,u_0) =&\, \frac{A_2B_2u_0+A_2E_2+B_2}{A_2^2}e^{(\sigma -u_0)A_2}\nonumber \\&-\frac{A_2B_2\sigma +A_2E_2+B_2}{A_2^2} \nonumber \\&\,+e^{(\sigma -u_0)A_2}x_0,\forall \sigma \in \;]-\infty ,u_0]. \end{aligned}$$
(199)

Equations (198)–(199) are valid since \(A_1\ne 0\) and \(A_2 \ne 0\). Define the function \(\omega _{\Phi }(\cdot ,x_0,u_0)\) by Eq. (41). Then, the intersecting function \(\Omega \) should satisfy

$$\begin{aligned} \omega _{\Phi }\big (\Omega (x_1,v) ,x_1,v\big )=f_{\text {an}}\big (\Omega (x_1,v)\big ),\forall (x_1,v) \in \mathbb {R}^2. \end{aligned}$$
(200)

Define

$$\begin{aligned} M_1=&\;\Big (B_1 A_1^{-1}\big (A_2^{-1}-A_1^{-1} \big )-E_1 A_1^{-1}+E_2A_2^{-1}\Big ) \frac{A_2}{A_1-A_2}, \end{aligned}$$
(201)
$$\begin{aligned} M_2=&\;\Big (B_1 A_1^{-1}\big (A_2^{-1}-A_1^{-1} \big )-E_1 A_1^{-1}+E_2A_2^{-1}\Big ) \frac{A_1}{A_1-A_2}. \end{aligned}$$
(202)

Note that \(M_1>0\) and \(M_2<0\) owing to (114)–(116). Combining (197)–(200) and (114)–(116) it follows from the definition of function \(\Omega \) (in Sect. 8.3) that

$$\begin{aligned} \Omega (x_1,v) = {\left\{ \begin{array}{ll}&{}v-\frac{1}{A_1} \log \left( 1+\frac{x_1 - f_{\text {an}}(v)}{M_1} \right) \text { if } \quad x_1 \ge f_{\text {an}}(v),\\ &{}v-\frac{1}{A_2} \log \left( 1-\frac{f_{\text {an}}(v)-x_1}{M_2} \right) \text { if } \quad x_1 \le f_{\text {an}}(v), \end{array}\right. } \end{aligned}$$
(203)

where \(\log \) sets for the natural logarithm. The function \(\varsigma \) in Eq. (42) can be determined explicitly as

$$\begin{aligned} \varsigma (x_1,v) =&\; x_1v-\frac{1}{A_1}x_1+\frac{E_1}{A_1}v+\frac{B_1}{2A_1}v^2 -M_1 \Omega (x_1,v) \nonumber \\&+\frac{E_2-E_1}{A_1(A_1-A_2)} \quad \text { if } \quad x_1 \ge f_{\text {an}}(v), \end{aligned}$$
(204)
$$\begin{aligned} \varsigma (x_1,v) =&\; x_1v-\frac{1}{A_2}x_1+\frac{E_2}{A_2}v+\frac{B_1}{2A_1}v^2 -M_2 \Omega (x_1,v) \nonumber \\&+\frac{E_2-E_1}{A_2(A_1-A_2)} \quad \text { if } \quad x_1 \le f_{\text {an}}(v). \end{aligned}$$
(205)

It can be checked that

$$\begin{aligned} \varsigma (x_1,v) =-\frac{B_1}{2A_1}v^2 \quad \text {if} \quad x_1=f_{\text {an}}(v). \end{aligned}$$
(206)

The fact that Inequality (39) holds for any input \( u \in AC(\mathbb {R}_+,\mathbb {R})\) and any initial condition \(x_0 \in \mathbb {R}\) follows from Theorem 3. However, \(\varsigma \) is not nonnegative: it can be checked that for any fixed \(x_1\) we have \(\lim _{v \rightarrow \pm \infty } \varsigma (x_1,v)= -\infty \).

The aim of the following analysis is to show that \(\forall (x_1,v) \in \mathbb {R} \times \left[ \frac{1}{A_1},\frac{1}{A_2} \right] \) we have \(\varsigma (x_1,v) \ge 0\). To this end, observe that, from (114) and (206), we have

$$\begin{aligned} \varsigma (x_1,v) \ge 0 \quad \text {whenever} \quad x_1=f_{\text {an}}(v). \end{aligned}$$
(207)

Now, fix \(v \in \left[ \frac{1}{A_1},\frac{1}{A_2} \right] \). From (203)–(204) and (114)–(116) it follows that

$$\begin{aligned} \lim _{x_1 \rightarrow \infty } \varsigma (x_1,v) = \infty .\end{aligned}$$
(208)

Suppose that there exists \(x_2 \in \;]f_{\text {an}}(v),\infty [\) such that \(\varsigma (x_2,v)<0\). Then, from (207)–(208) it follows that \(\varsigma (\cdot ,v)\) should have a minimum at \(x_3 \in \; ]f_{\text {an}}(v),\infty [\) such that \(\varsigma (x_3,v)<0\). A necessary condition for this to happen is \(\frac{\partial \varsigma }{\partial x_1}(x_3,v)=0\). It can be checked from Eq. (204) that this last equality cannot hold. A similar argument can be used for Equation (205).

Proof of Lemma 11

Observe that, for Theorem 5 to hold, it is needed that \(A_1\) and \(-A_2\) are both stable. Since \(n=1\), this condition translates into \(A_1<0\) and \(A_2>0\) so that the results of Theorems 5, 6, and 7 apply.

The proof is done in two steps. In Step 1 we consider a specific T–periodic input \(u \in W^{1,\infty }(\mathbb {R}_+,\mathbb {R})\) and an arbitrary initial condition \(x_0\). Using Theorem 7 it follows that the function \(\varphi _u^\circ \) that characterizes the hysteresis loop satisfies the differential state equation (79) and the output equation (80). The aim of Step 1 is to find the initial state \(x_u^\circ (0)\) since the latter may be different from \(x_0\). In Step 2 we use the knowledge of \(x_u^\circ (0)\) to prove that, if Assumption 8 holds, then the relations (251)–(252) hold.

Step 1. Let \(\alpha \in \;]0,1[;\) define \(\varrho _1=1\), \(\varrho _2=2-\alpha \), \(\varrho _3=3-2\alpha \), \(\varrho _4=4-2\alpha \). Note that \(0<\varrho _1<\varrho _2<\varrho _3<\varrho _4\). We consider the \(\varrho _4\)-periodic input \(u : \mathbb {R}_+ \rightarrow \mathbb {R}\) defined on the interval \([0,\varrho _4]\) by

$$\begin{aligned} u(\varrho )&=\varrho , \forall \varrho \in [0,\varrho _1], \end{aligned}$$
(209)
$$\begin{aligned} u(\varrho )&=2-\varrho , \forall \varrho \in [\varrho _1,\varrho _2], \end{aligned}$$
(210)
$$\begin{aligned} u(\varrho )&=2\alpha -2+\varrho , \forall \varrho \in [\varrho _2,\varrho _3], \end{aligned}$$
(211)
$$\begin{aligned} u(\varrho )&=4-2\alpha -\varrho , \forall \varrho \in [\varrho _3,\varrho _4]. \end{aligned}$$
(212)

Observe that \(u(0)=0,\) \(u(\varrho _1)=1,\) \(u(\varrho _2)=\alpha, \) \(u(\varrho _3)=1,\) \(u(\varrho _4)=0,\) and that \(u \in W^{1,\infty }(\mathbb {R}_+,\mathbb {R}). \) Observe also that \(|\dot{u}(\varrho )|=1\) for almost all \(\varrho \in \mathbb {R}_+\) so that \(\rho _u\) is the identity function which gives \(\psi _u=u.\) Let \(x_0 \in \mathbb {R}\) and consider the scalar semilinear Duhem model with input u and initial condition \(x_0\) (Eqs. (63)–(65)). Since all conditions of Theorem 7 hold, we get from Equality (80) that

$$\begin{aligned} \varphi _u^\circ (\varrho )=C x_u^\circ (\varrho ) + D u(\varrho ), \forall \varrho \in [0,\varrho _4], \end{aligned}$$
(213)

where \(x_u^\circ \) satisfies the differential equation (79). To find the initial condition \(x_u^\circ (0)\) we compute \(x_u^\circ (\varrho _k), k=1,\ldots ,4\) as a function of \(x_u^\circ (0)\) and we use the fact that, by Theorem 7, we have \(x_u^\circ (0)=x_u^\circ (\varrho _4).\) We start by computing \(x_u^\circ (\varrho _1)\) as a function of \(x_u^\circ (0).\) In the interval \([0,\varrho _1],\) the differential equation (79) becomes

$$\begin{aligned} \frac{\text{ d }x_u^\circ }{\text{ d }\varrho }(\varrho ) = A_1x_u^\circ (\varrho )+B_1u(\varrho )+E_1, \forall \varrho \in \;]0,\varrho _1[. \end{aligned}$$
(214)

Equation (214) can be solved explicitly and it gives

$$\begin{aligned} x_u^\circ (\varrho _1) =e^{\varrho _1 A_1 }x_u^\circ (0)+e^{\varrho _1 A_1 } \int _{0}^{\varrho _1} e^{-\tau A_1} \left( B_1u(\tau )+E_1\right) \, \text{ d }\tau . \end{aligned}$$
(215)

Taking into account Eq. (209) it follows that

$$\begin{aligned} x_u^\circ (1)&=e^{ A_1 }x_u^\circ (0)+ \beta _{11}, \end{aligned}$$
(216)
$$\begin{aligned} \beta _{11}&=A_1 ^{-1} \Big [\left( -1 -A_1^{-1} +A_1^{-1}e^{ A_1} \right) B_1 + \left( -1 + e^{ A_1} \right) E_1\Big ]. \end{aligned}$$
(217)

In the interval \([\varrho _1,\varrho _2]\), the differential equation (79) becomes

$$\begin{aligned} \frac{\text{ d }x_u^\circ }{\text{ d }\varrho }(\varrho ) = -A_2x_u^\circ (\varrho )-u(\varrho )B_2-E_2, \forall \varrho \in \;]\varrho _1,\varrho _2[. \end{aligned}$$
(218)

Equation (218) can be solved explicitly and it gives

$$\begin{aligned} x_u^\circ (\varrho _2) =&\;e^{-(\varrho _2-\varrho _1) A_2 }x_u^\circ (\varrho _1)\nonumber \\&-e^{-\varrho _2 A_2 } \int _{\varrho _1}^{\varrho _2} e^{\tau A_2} \left( u(\tau )B_2+E_2\right) \, \text{ d }\tau . \end{aligned}$$
(219)

Taking into account Eq. (210) it follows that

$$\begin{aligned} x_u^\circ (\varrho _2)&=e^{(\alpha -1) A_2 }x_u^\circ (1) + \beta _{21}e^{A_2 \alpha }+\beta _{22} \alpha +\beta _{23}, \end{aligned}$$
(220)
$$\begin{aligned} \beta _{21}&=A_2 ^{-1} e^{-A_2}\left( B_2(1+A_2^{-1})+E_2\right) , \end{aligned}$$
(221)
$$\begin{aligned} \beta _{22}&=-A_2^{-1}B_2, \end{aligned}$$
(222)
$$\begin{aligned} \beta _{23}&=-A_2^{-1}\left( A_2^{-1}B_2+E_2 \right) . \end{aligned}$$
(223)

In the interval \([\varrho _2,\varrho _3]\), the differential equation (79) becomes

$$\begin{aligned} \frac{\text{ d }x_u^\circ }{\text{ d }\varrho }(\varrho ) = A_1x_u^\circ (\varrho )+B_1u(\varrho )+E_1, \forall \varrho \in \;]\varrho _2,\varrho _3[. \end{aligned}$$
(224)

Equation (224) can be solved explicitly and it gives

$$\begin{aligned} {\begin{matrix} x_u^\circ (\varrho _3) = &{}\;e^{(\varrho _3-\varrho _2) A_1 }x_u^\circ (\varrho _2)\\ &{}+e^{\varrho _3 A_1 } \int _{\varrho _2}^{\varrho _3} e^{-\tau A_1} \left( B_1u(\tau )+E_1\right) \, \text{ d }\tau . \end{matrix}} \end{aligned}$$
(225)

Taking into account Eq. (211) it follows that

$$\begin{aligned} x_u^\circ (\varrho _3)&=e^{(1-\alpha ) A_1 }x_u^\circ (\varrho _2)+\beta _{31}\alpha e^{-A_1\alpha } +\beta _{32} e^{-A_1\alpha }+ \beta _{33}, \end{aligned}$$
(226)
$$\begin{aligned} \beta _{31}&=A_1 ^{-1} B_1e^{A_1}, \end{aligned}$$
(227)
$$\begin{aligned} \beta _{32}&= A_1 ^{-1} e^{A_1} \left( A_1^{-1}B_1+E_1\right) , \end{aligned}$$
(228)
$$\begin{aligned} \beta _{33}&=A_1^{-1}\left( -(1+A_1^{-1})B_1-E_1 \right) . \end{aligned}$$
(229)

In the interval \([\varrho _3,\varrho _4]\), the differential equation (79) becomes

$$\begin{aligned} \frac{\text{ d }x_u^\circ }{\text{ d }\varrho }(\varrho ) = -A_2x_u^\circ (\varrho )-u(\varrho )B_2-E_2, \forall \varrho \in \;]\varrho _3,\varrho _4[. \end{aligned}$$
(230)

Eq. (230) can be solved explicitly and it gives

$$\begin{aligned} x_u^\circ (\varrho _4) =&\;e^{-(\varrho _4-\varrho _3) A_2 }x_u^\circ (\varrho _3)\nonumber \\ {}&-e^{-\varrho _4 A_2 } \int _{\varrho _3}^{\varrho _4} e^{\tau A_2} \left( u(\tau )B_2+E_2\right) \, \text{ d }\tau . \end{aligned}$$
(231)

Taking into account Eq. (212) it follows that

$$\begin{aligned} x_u^\circ (\varrho _4)&=e^{- A_2 }x_u^\circ (\varrho _3)+\beta _{41}, \end{aligned}$$
(232)
$$\begin{aligned} \beta _{41}&= -A_2^{-1}\Big [B_2\left( -e^{-A_2}-A_2^{-1}e^{-A_2}+A_2^{-1} \right) \nonumber \\&\quad +E_2\left( 1-e^{-A_2} \right) \Big ]. \end{aligned}$$
(233)

Now we use the relation \(x_u^\circ (0)=x_u^\circ (\varrho _4)\) to find \(x_u^\circ (0)\) using Eqs. (216)–(217), (220)–(223), (226)–(229) and (232)–(233). We get

$$\begin{aligned} x_u^\circ (0)&=\frac{\beta _{51} + \beta _{52}e^{(A_2-A_1)\alpha }+ \beta _{53} \alpha e^{-A_1\alpha }+\beta _{54}e^{-A_1\alpha } }{1+\beta _{55}e^{(A_2-A_1)\alpha }}, \end{aligned}$$
(234)
$$\begin{aligned} \beta _{51}&=\beta _{33}e^{-A_2}+\beta _{41}, \end{aligned}$$
(235)
$$\begin{aligned} \beta _{52}&=\beta _{21}e^{A_1-A_2}+\beta _{11}e^{A_1-2A_2}, \end{aligned}$$
(236)
$$\begin{aligned} \beta _{53}&=\beta _{22}e^{A_1-A_2} + e^{-A_2}\beta _{31}, \end{aligned}$$
(237)
$$\begin{aligned} \beta _{54}&=\beta _{23}e^{A_1-A_2}+e^{-A_2}\beta _{32},\end{aligned}$$
(238)
$$\begin{aligned} \beta _{55}&=-e^{2(A_1-A_2)}. \end{aligned}$$
(239)

Note that, since \(0<\alpha <1\), \(A_1<0\) and \(A_2>0\) it follows that \(0<e^{(2-\alpha )(-A_2+A_1)}<1\) so that the denominator in Equ. (234) is nonzero.

Step 2. By Assumption 8 it follows that \(\varphi _u^\circ (\varrho _1)=\varphi _u^\circ (\varrho _3)\). This means that \(x_u^\circ (1)=x_u^\circ (\varrho _3)\) because \(C \ne 0\). Since \(x_u^\circ (0)\) has been computed explicitly, \(x_u^\circ (1)\) and \(x_u^\circ (\varrho _3)\) are available explicitly using Eqs. (216)–(217) and (226)–(229) respectively. We get

$$\begin{aligned} {\begin{matrix} x_u^\circ (1)=&{}\;\frac{e^{A_1}}{1+\beta _{55}e^{(A_2-A_1)\alpha }} \cdot \Big (\beta _{51} + \beta _{52}e^{(A_2-A_1)\alpha } \\ &{}+ \beta _{53} \alpha e^{-A_1\alpha }+\beta _{54}e^{-A_1\alpha } \Big )+\beta _{11}, \end{matrix}} \end{aligned}$$
(240)
$$\begin{aligned} {\begin{matrix} x_u^\circ (\varrho _3)={}&{}\;\frac{e^{2A_1-A_2}}{1+\beta _{55}e^{(A_2-A_1)\alpha }}\cdot \Big (\beta _{51} e^{(A_2-A_1) \alpha } \\ &{}+\beta _{52}e^{2(A_2-A_1)\alpha }+ \beta _{53} \alpha e^{(A_2-2A_1)\alpha }\\ {} &{}+\beta _{54}e^{(A_2-2A_1)\alpha } \Big ) + \beta _{61} e^{(A_2-A_1) \alpha }\\ {} &{}+ \beta _{62}\alpha e^{-A_1 \alpha } + \beta _{63}e^{-A_1 \alpha } + \beta _{33}, \end{matrix}} \end{aligned}$$
(241)

where

$$\begin{aligned} \beta _{61}&= \left( \beta _{21} +\beta _{11}e^{-A_2}\right) e^{A_1},\end{aligned}$$
(242)
$$\begin{aligned} \beta _{62}&=\beta _{22}e^{A_1}+\beta _{31},\end{aligned}$$
(243)
$$\begin{aligned} \beta _{63}&=\beta _{23}e^{A_1}+\beta _{32}. \end{aligned}$$
(244)

Our aim in the following analysis is to find the conditions under which we have \(x_u^\circ (1)=x_u^\circ (\varrho _3)\) for all inputs u that satisfy the relations (209)–(212). This means finding the conditions under which we have \(x_u^\circ (1)=x_u^\circ (\varrho _3)\) for all \(\alpha \in \;]0,1[\). In the equality \(x_u^\circ (1)=x_u^\circ (\varrho _3)\) we multiply both terms with \(1+\beta _{55}e^{(A_2-A_1)\alpha }\) so that we get from Equalities (240)–(244) that

$$\begin{aligned} \beta _{71}+\beta _{72}e^{(A_2-A_1)\alpha }+&\;\beta _{73}\alpha e^{-A_1\alpha }+\beta _{74}e^{-A_1\alpha }=0, \nonumber \\&\forall \alpha \in \; ]0,1[, \end{aligned}$$
(245)

where

$$\begin{aligned} \beta _{71}&=e^{A_1}\beta _{51}+\beta _{11}-\beta _{33}, \end{aligned}$$
(246)
$$\begin{aligned} \beta _{72}&=e^{A_1}\beta _{52}+\beta _{11}\beta _{55}-e^{2A_1-A_2}\beta _{51}-\beta _{61}-\beta _{55}\beta _{33}, \end{aligned}$$
(247)
$$\begin{aligned} \beta _{73}&=e^{A_1}\beta _{53}-\beta _{62},\end{aligned}$$
(248)
$$\begin{aligned} \beta _{74}&=e^{A_1}\beta _{54}-\beta _{63}. \end{aligned}$$
(249)

Consider the functions \(f_1,f_2,f_3,f_4,\mathbf {0}:\;]0,1[\; \rightarrow \mathbb {R}\) defined by \(\forall \alpha \in \;]0,1[, f_1(\alpha )=1\), \(f_2(\alpha )=e^{(A_2-A_1)\alpha }\), \(f_3(\alpha )=\alpha e^{-A_1\alpha }\), \(f_4(\alpha )=e^{-A_1\alpha }\), and \(\mathbf {0}(\alpha )=0\). Then Eq. (245) can be written as

$$\begin{aligned} \beta _{71}\cdot f_1+\beta _{72}\cdot f_2+\beta _{73}\cdot f_3+\beta _{74}\cdot f_4=\mathbf {0}. \end{aligned}$$
(250)

Consider the vector space of functions \(\{p:\;]0,1[\; \rightarrow \mathbb {R}\}\) with its usual binary operations of vector addition and scalar multiplication. Then the functions \(f_1,f_2,f_3,f_4\) are linearly independent vectors so that, owing to Eq. (250), we have \(\beta _{71}=\beta _{72}=\beta _{73}=\beta _{74}=0\) since \(\beta _{ij}\) is independent of \(\alpha \) for all possible i and j.

We start by solving Equation \(\beta _{73}=0\). Combining Eqs. (248), (237), (243), (222), and (227) it comes that

$$\begin{aligned} A_2^{-1} B_2 = A_1^{-1}B_1. \end{aligned}$$
(251)

Now we solve Equation \(\beta _{71}=0\). Combining Eqs. (251), (246), (235), (229), (233), and (217) it follows that

$$\begin{aligned} B_1 A_1^{-1}\left( A_2^{-1}-A_1^{-1} \right) -E_1 A_1^{-1}+E_2A_2^{-1}=0. \end{aligned}$$
(252)

It can be checked that Equalities (251)–(252) imply that \(\beta _{71}=\beta _{72}=\beta _{73}=\beta _{74}=0\).

Lemma 11 follows from Lemma 8.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Ikhouane, F. A Survey of the Hysteretic Duhem Model. Arch Computat Methods Eng 25, 965–1002 (2018). https://doi.org/10.1007/s11831-017-9218-3

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s11831-017-9218-3

Navigation