Startseite Lyapunov Stability of a Fractionally Damped Oscillator with Linear (Anti-)Damping
Artikel Open Access

Lyapunov Stability of a Fractionally Damped Oscillator with Linear (Anti-)Damping

  • Matthias Hinze EMAIL logo , André Schmidt und Remco I. Leine
Veröffentlicht/Copyright: 25. Februar 2020
Veröffentlichen auch Sie bei De Gruyter Brill

Abstract

In this paper, we develop a Lyapunov stability framework for fractionally damped mechanical systems. In particular, we study the asymptotic stability of a linear single degree-of-freedom oscillator with viscous and fractional damping. We prove that the total mechanical energy, including the stored energy in the fractional element, is a Lyapunov functional with which one can prove stability of the equilibrium. Furthermore, we develop a strict Lyapunov functional for asymptotic stability, thereby opening the way to a nonlinear stability analysis beyond an eigenvalue analysis. A key result of the paper is a Lyapunov stability condition for systems having negative viscous damping but a sufficient amount of positive fractional damping. This result forms the stepping stone to the study of Hopf bifurcations in fractionally damped mechanical systems. The theory is demonstrated on a stick-slip oscillator with Stribeck friction law leading to an effective negative viscous damping.

MSC 2010: 34K20; 34K37; 37N05

1 Introduction

This paper is concerned with the development of a Lyapunov stability framework, including Lyapunov’s direct method, for the analysis of stability properties of mechanical systems with fractional damping. The scope of the paper is limited to single degree-of-freedom oscillators with both viscous and fractional damping.

The term fractional refers to fractional calculus, which is a mathematical theory dealing with derivatives and integrals of arbitrary (non-integer) order [1, 2] with a variety of applications in science and engineering [3]. Particularly in mechanics, fractional damping may arise through the modeling of mechanical systems with viscoelastic components. Complex rheological models for viscoelastic materials are often described through an array of classical Kelvin or Maxwell elements, inevitably resulting in a model with a large number of parameters. It has been shown that the viscoelastic behavior of complex materials is in many applications well represented by fractional order elements with only a few parameters [4, 5]. Furthermore, a description using fractional order force-displacements relationships may have much better extrapolation properties on long time-scales. The introduction of fractional calculus in mechanics leads to the concept of a springpot element, being a force law reacting linearly on a fractional derivative of its elongation. In a more general setting, springpot elements may also arise through fractional-order control laws [6].

Many problems in industrial applications originate from (dynamic) instability phenomena, e. g. stick-slip vibrations in oilwell drillstrings, flutter of airfoils, shimmy of vehicles and feedback instabilities in control systems. Methods to rigorously prove stability of linear and nonlinear systems are therefore quintessential. The Lyapunov stability framework, which encompasses the method of Lyapunov functions, forms a central element in the research fields Nonlinear Dynamics and Control Theory [7]. The introduction of springpot elements in (controlled) mechanical systems asks for an extension of the Lyapunov stability framework to non-integer order derivatives. A major complication arises through the non-local character of fractional derivatives, i. e. the force in a springpot element depends on the total history of the elongation. A system with springpot elements has therefore an infinite state which asks for the use of Lyapunov functionals instead of Lyapunov functions in Lyapunov’s direct method, which are introduced in the theory of functional differential equations (FDEs) [89101112]. Special Lyapunov functionals for FDEs with fractional derivatives are introduced in [13, 14], which have been shown to represent the potential energy of an infinite arrangement of springs and dashpots [151617]. The energy expressions for springpots are based on the infinite state or diffusive representation of fractional integrators, which were introduced by Montseny [18], Matignon [19] and have been elaborated by Trigeassou et al. [13, 14, 20, 21]. Beyond that, a lot of work has been done on stability conditions [22, 23] and Lyapunov theory [242526] for fractional differential equations, which cannot directly be used for mechanical systems containing springpots, as the differentiation order is in general irrational for such systems. Furthermore, the Laplace transform method has been used to prove stability of equilibria of mechanical systems containing springpots [27, 28].

The aim of this paper is to give a complete stability analysis of a linear single degree-of-freedom mass-spring-dashpot-springpot system. The analysis encompasses the following results/tasks:

  1. The total mechanical energy of the system is derived through the use of the infinite state representation of the springpot element (Section 3).

  2. The system is put in the form of an FDE and, based on this, definitions of Lyapunov stability and attractivity of fractionally damped mechanical systems are given. Furthermore, a Lyapunov-Krasovskii theorem is presented for this class of systems (Section 4.2).

  3. It is shown that the total mechanical energy is a Lyapunov functional for the system with positive viscous and fractional damping with which stability of the equilibrium can be proven (Section 4.3). Furthermore, a strict Lyapunov functional for positive viscous and fractional damping is derived which rigorously proves asymptotic stability (Section 4.4). However, this Lyapunov functional fails to give a stability result in the case of anti-damping, in which the viscous damping is negative.

  4. An extensive eigenvalue analysis is given and, based on that, an expression for the general solution is derived. The eigenvalue analysis reveals that the equilibrium can still be asymptotically stable in the presence of anti-damping (Section 4.5.1).

  5. A strict Lyapunov functional for the case of anti-damping is derived in Section 4.5.2. This leads to a Lyapunov stability condition for systems having negative viscous damping but a sufficient amount of positive fractional damping. Moreover, this result opens the way to study global asymptotic stability of nonlinear systems with fractional damping. The theory is demonstrated on a stick-slip oscillator with Stribeck friction law leading to an effective negative viscous damping (Section 4.5.3).

The scope of the present paper is limited to a single degree-of-freedom oscillator, being of course a first step into the direction of multi degree-of-freedom systems. A brief outlook on how these results can be extended to more degrees of freedom will be given in the conclusion section of the paper.

2 Fractional calculus and infinite state representation

We will consider the fractional derivative of Caputo type that is based on the fractional Riemann–Liouville integral, which is defined for an integrable function x = x(t) with t ≥ t0 and a scalar value α > 0 as

(1) It0+αx(t)=1Γ(α)t0tx(τ)(tτ)α1dτ,

where Γ(α) is the Gamma function. For α = 0 we set It0+0x:=x and it can be seen directly that the choice α = 1 leads to the classical integral. In this paper, we will describe this integral operator by the infinite state representation [21]

(2) { z˙(ω,t)=ωz(ω,t)+x(t),It0+αx(t)=0μα(ω)z(ω,t)dω.

The infinite state z(ω,t) fulfills the above differential equation ω0 and the fractional integral is obtained by integrating all contributions z(ω,t) weighted by the function

(3) μα(ω):=sin(απ)πωα.

The equivalence of eqs. (1) and (2) is derived in [15]. The representation (2) is very useful to give a mechanical interpretation of a fractional derivative (Section 3) and to formulate Lyapunov functionals for fractionally damped mechanical systems (Section 4). Using the variation of constants formula, we can formulate eq. (2) as

(4) It0+αx(t)=0μα(ω)t0teω(tτ)x˙(τ)dτdω.

Finally we introduce the fractional Caputo derivative for an absolutely continuous function x = x(t) and 0 < α < 1 as

(5) CDt0+αx(t)=It0+1αx˙(t).

3 Springpot: Mechanical representation and potential energy

In this section, we will briefly introduce the Caputo springpot (Figure 1) as an abstract mechanical element and discuss its mechanical representation to gain a potential energy expression which may be used for Lyapunov stability considerations of mechanical systems containing springpots. Again, details may be found in [15]. A springpot is defined by its constitutive equation

(6) f(t)=cCDt0+αq(t),

where f is the force acting on the springpot which results in an elongation q depending on the coefficient c > 0, initialization time t0 < 0 and differentiation order α∈(0,1). The time interval [t0,0] represents the entire significant history of the springpot, i. e. for earlier time-instants t ≤ t0 we assume q(t) = 0 and f(t) = 0.

Figure 1: Force acting on a springpot.
Figure 1:

Force acting on a springpot.

Together with eqs. (2) and (5) we derive the infinite state representation of a Caputo springpot as

(7) { y˙(ω,t)=ωy(ω,t)+q˙(t),y(ω,t0)=0,f(t)=c0μ1α(ω)y(ω,t)dω.

The infinite state y in eq. (7) may, similar as in eq. (4), be expressed by the variation-of-constants formula

(8) y(ω,t)=t0teω(tτ)q˙(τ)dτ=t0t0eωsq˙(t+s)ds,

which will be useful for stability considerations later on. The above representation leads to a mechanical analogue model of a springpot, which is a parallel arrangement of an infinite number of Maxwell elements (Figure 2) as in [16], where the forces g(ω,t)dω of the Maxwell elements are integrated to the resulting force

(9) f(t)=0g(ω,t)dω
Figure 2: Schematic mechanical representation of a Caputo springpot.
Figure 2:

Schematic mechanical representation of a Caputo springpot.

on the system. The springs of the Maxwell elements are characterized by their elongation qs(ω,t) and spring constant k(ω)dω and the dashpots by the elongation qd(ω,t) and constant d(ω)dω such that the elongation of the system q(t) appears as

(10) q(t)=qs(ω,t)+qd(ω,t)ω0

with the incremental internal force

(11) g(ω,t)dω=k(ω)dωqs(ω,t)=d(ω)dωq˙d(ω,t).

Differentiation of eq. (10) and substitution of eq. (11) leads to

(12) q˙(t)=g˙(ω,t)k(ω)+g(ω,t)d(ω).

Comparison of eqs. (9) and (12) to (7) results in the identification

(13) g(ω,t)=cμ1α(ω)y(ω,t),k(ω)=cμ1α(ω),d(ω)=cμ1α(ω)ω,ω=k(ω)d(ω).

Furthermore, we obtain an interpretation of the infinite state y of the Caputo springpot as

(14) y(ω,t)=g(ω,t)cμ1α(ω)=g(ω,t)k(ω)=qs(ω,t).

Finally, we consider the energy of the mechanical equivalent system which is the potential energy stored in the springs of the Maxwell elements, i. e.

(15) E(t)=120k(ω)qs2(ω,t)dω,

which can be reformulated with eqs. (13) and (14) as

(16) E(t)=c20μ1α(ω)y2(ω,t)dω.

4 Stability

4.1 Introduction

The mechanical representation and the potential energy expressed in terms of the infinite state y were derived more detailed in [15]. The energy expression in eq. (16) was used to prove stability of the equilibrium of a mass-spring-springpot system. In the following we want to extend this approach to consider stability of the same system when linear (anti-)damping is introduced, i. e. we regard the system (Figure 3)

(17) mq¨(t)=dq˙(t)cCDt0+αq(t)kq(t),t0

with mass m, elongation q(t), spring coefficient k, springpot coefficient c, damping coefficient d and differentiation order α∈(0,1) and given initial functions φ1,φ2CB((,0];R) such that

(18) q(t)=φ1(t),t0,q˙(t)=φ2(t),t0,

Figure 3: Mass-spring-dashpot-springpot system.
Figure 3:

Mass-spring-dashpot-springpot system.

where φi(t) = 0 for t ≤ t0, i = 1,2. Again, with the help of eqs. (7) and (8) we reformulate eq. (17) as an FDE

(19) { q˙(t)=v(t),v˙(t)=kmq(t)dmq˙(t)cmt0t00μ1α(ω)eωsdωvt(s)ds,

where

vt(s)=v(t+s),s(,0]

and use the associated stability theory. For the cases d > 0 (damping) and d < 0 (anti-damping) we use different methods to prove Lyapunov stability of the equilibrium of eq. (17).

4.2 Theoretical background

The system (19) is a representative of an FDE of the form

(21) x˙(t)=f(t,xt),tt10

with initial time t1 and a map

(22) f:SRn

defined on

(23) S=[0,)×QH,QH:={φCB((,0];Rn) | φ<H}, H>0,

i. e. the function f acts in its second argument on the space CB((,0];Rn)k of continuous and bounded functions defined on the negative half space together with the norm

(24) φ=sups(,0]φ(s)2,

where .2 is the Euclidean norm. Furthermore, let

(25) xC((,T);Rn),T>t1,xt(s)=x(t+s),s(,0],

such that

xtCB((,0];Rn).

We assume f locally Lipschitzian in QH which ensures local existence and uniqueness of a solution x(t1,φ) of eq. (21) for a given initial function φ and initial time t1 [9, 11]. Moreover, let f(t,0)=0tt1, such that the trivial solution x(t1,φ)(t) = 0 is an equilibrium of the system. In the following theorem, we formulate sufficient conditions for (asymptotic) stability of the trivial solution of eq. (21) by a Lyapunov theorem adapted to the space CB((,0];Rn) as in [8, 9, 11]. We need the following definitions.

Definition 4.1.

A solution of eq. (21) with initial function φCB((,0];Rn) and initial time t1 ≥ 0 is a function x(t1,φ) defined and continuous on an interval (,T), T > t1, such that xt(t1,φ)S for t∈[t1,T), xt1(t1,φ)=φ and x(t1,φ)(t) satisfies eq. (21) for t∈[t1,T).

Definition 4.2. (Stability)

The trivial solution x(t1,φ)(t) = 0 of eq. (21) together with (23), (25), initial function φ and initial time t1 is called

  1. stable, if for all ε∈(0,H] there exists a δ  =  δ(t1,ε) > 0 such that x(t1,φ)(t)2<ϵ for t ≥ t1 if φ∈Qδ.

  2. asymptotically stable, if it is stable and for all t1 there exists a δ = δ(t1) > 0 such that limtx(t1,φ)(t)2 = 0 if φ∈Qδ.

Theorem 4.3 (Lyapunov-Krasovskii [9, 11])

Let f:SRn such that f(t,0)=0tt1 and denote ui:[0,)R, i = 1,2 some scalar, continuous, non-decreasing functions such that ui(0) = 0 and ui(r) > 0 for r > 0. Let there exist a continuous functional V:[0,)×QHR such that

(26) u1(φ(0)2)V(t,φ),

(27) V(t,0)=0,V˙(t,xt)0,

then the trivial solution of eq. (21) is stable.

Furthermore, if the additional assumptions

(28) L>0:f(t,φ)2<L, tt1, φQH,

(29) V˙(t,xt)u2(x(t1,φ)(t)2)

are satisfied, then the trivial solution is asymptotically stable.

Remark 4.4.

As f is assumed locally Lipschitzian in QH, condition (28) is fulfilled, if the Lipschitz constant L is independent of time t ≥ t1.

4.3 Undamped Case

First, we consider eq. (17) for the case d = 0 and use the total mechanical energy

(30) V1(t,qt,vt)=12mvt2(0)+12kqt2(0)+12c0μ1α(ω)y2(ω,t)dω

as a Lyapunov functional as in [15], where we use a simplified notation in terms of the infinite state y and keep in mind the formulation in eq. (8) or

(31) y(ω,t)=t0t0eωsvt(s)ds,ω0.

We prove stability of the trivial solution with the help of Theorem 4.3. It is obvious, that inequality (26) holds for V1. Furthermore, as

(32) V˙1(t,qt,vt)=kqt(0)q˙t(0)+mvt(0)v˙t(0)+c0μ1α(ω)y(ω,t)y˙(ω,t)dω=vt(0)(mq¨t(0)+kqt(0)+c0μ1α(ω)y(ω,t)dω)c0ωμ1α(ω)y2(ω,t)dω=c0ωμ1α(ω)y2(ω,t)dω0,

inequality (27) is fulfilled, such that the trivial solution is stable.

4.4 Linear damping

Using the energy functional V1 in the case d > 0 again leads to a non-positive rate of V1

(33) V˙1=dv2(t)c0ωμ1α(ω)y2(ω,t)dω0,

which only proves stability of the equilibrium. However, we could find an augmented candidate Lyapunov functional which contains the potential energy term (16) to prove asymptotic stability with the help of Theorem 4.3.

Proposition 4.5

The trivial solution of the damped mass-spring-springpot system (17) (m,d,k,c > 0, α ∈ [0,1]) is asymptotically stable.

Proof

We introduce a Lyapunov functional as in Theorem 4.3 defined on S using the simplified notation in terms of the infinite state y as in eq. (31). Furthermore, we introduce another infinite state Y = Y(ω,t) which satisfies

(34) Y˙(ω,t)=ωY(ω,t)+q(t),Y(ω,t0)=0.

Similar to eq. (31), this may be reformulated as

(35) Y(ω,t)=t0t0eωsqt(s)ds,ω0

and differentiating eq. (8) and comparing to eqs. (31) and (34) we obtain

(36) y(ω,t)=Y˙(ω,t)=ωY(ω,t)+q(t).

Hence, we formulate the Lyapunov functional

(37) V2(t,qt,vt)=12mvt2(0)+12kqt2(0)+d24mqt2(0)+12dqt(0)vt(0)+12c0μ1α(ω)y2(ω,t)dω+cd4m0μ1α(ω)ωY2(ω,t)dω

and check the conditions in Theorem 4.3. For (26) we can estimate

(38) V2(t,qt,vt)12mvt2(0)+12kqt2(0)+d24mqt2(0)+12dqt(0)vt(0)=14mvt2(0)+12kqt2(0)+d2mqt(0)+m2vt(0)214mvt2(0)+12kqt2(0).

Moreover, we compute the rate of V2 along solution curves

V˙2=mvt(0)v˙t(0)+kqt(0)q˙t(0)+d2q˙t(0)vt(0)+d2qt(0)v˙t(0)+d22mqt(0)q˙t(0)+c0μ1α(ω)y(ω,t)y˙(ω,t)dω+cd2m0μ1α(ω)ωY(ω,t)Y˙(ω,t)dω=dvt2(0)cvt(0)0μ1α(ω)y(ω,t)dω+d2vt2(0)+d2qt(0)kmqt(0)dmvt(0)cm0μ1α(ω)y(ω,t)dω+d22mqt(0)vt(0)+c0μ1α(ω)y(ω,t)vt(0)ωy(ω,t)dω+cd2m0μ1α(ω)qt(0)y(ω,t)y(ω,t)dω=d2vt2(0)kd2mqt2(0)c0μ1α(ω)ω+d2my2(ω,t)dωd2v2(t)kd2mq2(t)

which proves inequality (29). Finally, we still have to check condition (28), which is non-trivial only for the last addend of the right-hand side of the second equation in (19). For this term we have to split the interval of integration in two parts which yields for (1,)

| 1μ1α(ω)t0t0eωsvt(s)dsdω |1μ1α(ω)0eωsdsdωvt=sin(απ)π1ωα2dωvt=sin(απ)(1α)πvt.

For the integration over (0,1), we achieve together with eq. (36)

| 01μ1α(ω)t0t0eωsq˙t(s)dsdω |01μ1α(ω)| q(t)ωt0t0eωsqt(s)ds |dω01μ1α(ω)(qt+ω0eωsdsqt)dω=201μ1α(ω)dωqt=2sin(απ)απqt.

This completes the proof.   □

4.5 Linear anti-damping

For the case d < 0, whose physical interpretation is explained and motivated in Section 4.5.3, we expect the equilibrium of eq. (17) to remain stable only for certain values of d and it appears to be much more difficult to find Lyapunov functionals that yield stability criteria. Therefore, we start with the Laplace transform method to find sufficient conditions for stability and try to find similar conditions with the help of a Lyapunov functional. Finally, we compare the results.

4.5.1 Laplace transform method

Before we introduce the Laplace transform of eq. (17), we derive the Laplace transform of the fractional derivative. Therefore, consider the Laplace transform of (7)

(39) sLy(ω,t)(s)y(ω,0)=ωLy(ω,t)(s)+sQ(s)q0,LCDt0+αq(t)(s)=0μ1α(ω)Ly(ω,t)(s)dω.

with Laplace transform Q(s) of q(t) and initial value q(0) = q0. Substitution of the first equation of (39) in the second results in

(40) LCDt0+αq(t)(s)=0μ1α(ω)sQ(s)q0+y(ω,0)ω+sdω.

We reformulate (40) with the help of the following relation.

Proposition 4.6
(41) 0μα(ω)ω+sdω=sα,sCR, α(0,1)
Proof.

Due to the relation for the Laplace transform of eωt

L{eωt}(s)=0eωtestdt=0e(ω+s)tdt=1ω+se(ω+s)t0=1ω+s

we obtain eq. (41) using the formula

Γ(α)Γ(1α)=πsin(απ)

and Fubini’s Theorem as

0μα(ω)ω+sdω=sin(απ)π0ωα0e(ω+s)tdtdω=sin(απ)π0est0ωαeωtdωdt=sin(απ)π0estΓ(1α)tα1dt=sin(απ)πΓ(1α)Γ(α)sα=sα.

   □

This leads to the Laplace transform of the fractional derivative

(42) LCDt0+αq(t)(s)=sαQ(s)sα1q0+0μ1α(ω)y(ω,0)ω+sdω.

Hence, we obtain the Laplace transform of eq. (17) as

(43) ms2Q(s)sq0v0=kQ(s)c(sαQ(s)sα1q0+0μ1α(ω)y(ω,0)ω+sdω)d(sQ(s)q0)

with Laplace transform Q(s) of q(t) and initial values q(0) = q0, q˙(0)=v0. Solving equation (43) for Q(s) leads to

(44) Q(s)=m(sq0+v0)ms2+ds+csα+k+csα1q00μ1α(ω)y(ω,0)ω+sdωms2+ds+csα+k+dq0ms2+ds+csα+k.

The inverse Laplace transform may be obtained integrating along a Hankel contour and using the residue theorem, similar as described in [27, 28]. We will accomplish the entire derivation later and see, that, similar to the classical case, stability of the equilibrium depends on the real part of the poles of the right-hand side of eq. (44), i. e. we consider the equation

(45) ms2+ds+csα+k=0.

Let s=reiθ, we obtain real and imaginary part of eq. (45) as

(46) mr2cos(2θ)+drcos(θ)+crαcos(αθ)+k=0,mr2sin(2θ)+drsin(θ)+crαsin(αθ)=0.

From (46) we want to derive conditions, such that the roots of eq. (45) are located in the left-half complex plane. Therefore, we first consider the critical case for stability θ=±π2, which turns (46) into

(47) mr2+crαcosαπ2+k=0,dr+crαsinαπ2=0.

For fixed m, c, k and α the first equation of (47) has a unique solution r = r* > 0, which may be inserted in the second equation to compute a critical d < 0 for stability

(48) dcrit=csinαπ2rα1.

By numerical solution of eq. (47), we obtain the critical negative damping parameter dcrit depending on the value of α∈(0,1) and the parameters m, c, k, see Figure 4. The value |dcrit| is a measure for the damping capability of the springpot. As expected, it holds that dcrit0 for α0, as in this case the springpot degenerates to a spring, which stores energy and dcritc for α1, as the springpot becomes a dashpot. The dependency of dcrit on α for α∈(0,1) may change drastically for different parameters and it is quite interesting that dcrit<c can be achieved for certain values of α, i. e. a springpot can induce higher damping than a dashpot with the same coefficient. An example for this phenomenon is given in Section 4.5.3, where the knowledge (and tuning) of the system parameters m, k, c and α lead to the characterization of a Stribeck friction law. From the critical case for stability we now derive the following inequality conditions on r, such that a solution s=reiθ of (45) is located in the left-half complex plane. More specifically, we obtain the following proposition.

Figure 4: Critical negative damping parameter depending on α∈(0,1) for different parameters.
Figure 4:

Critical negative damping parameter depending on α∈(0,1) for different parameters.

Proposition 4.7.

Let the inequalities

(49) mr2+crαcosαπ2+k0,

(50) dr+crαsinαπ2>0

have a non-empty solution set for r > 0. Then there exists a pair of complex conjugate roots s=reiθ, sˉ=reiθ of (45) such that π2<θ<π2α. Furthermore, there exists no solution outside the sectors θπ2,π and θπ,π2.

Remark 4.8.

To depict the solution set of inequalities (49) and (50), consider Figure 5 below, where r<R has to hold.

Figure 5: Representation of the solution set of inequalities (49) and (50).
Figure 5:

Representation of the solution set of inequalities (49) and (50).

Proof of Proposition 4.7.

From (46) we see that for each root s=reiθ of equation (45), its complex conjugate sˉ = reiθ is another root. Therefore, we only consider 0 ≤ θ ≤ π. We examine the following cases.

  1. Case 1: θ = 0In this case, eq. (45) degenerates to one equation

    (51) mr2+dr+crα+k=0.

    Using eqs. (49) and (50), we can estimate

    mr2+dr+crα+k>crα1+cosαπ2sinαπ2+2k.

    As the function

    h(α):=1+cosαπ2sinαπ2

    fulfills

    h(0)=2,h(1)=0,h(α)=π2sinαπ2+cosαπ2<0α(0,1),

    we obtain

    mr2+dr+crα+k>0

    and there exists no solution of eq. (51).

  2. Case 2: θ = πThe second equation of (46) in this case reads as

    crαsin(απ)=0,

    which has no solution for α∈(0,1) except r = 0, which does not solve the first equation of (46)

    mr2dr+crαcosαπ+k=0.
  3. Case 3: 0<θ<π2Multiplying the first equation of (46) by cos(θ) and the second by sin(θ) sums up as

    (52) mr2cos(θ)+dr+crαcos((1α)θ)+kcos(θ)=0.

    The left-hand side of eq. (52) may be estimated with (49) and (50) as

    mr2cos(θ)+dr+crαcos((1α)θ)+kcos(θ)>crα(cos(θ)cosαπ2+cos((1α)θ)sinαπ2)+2kcos(θ)>0,

    because

    cos(θ)cosαπ2>0

    and

    cos((1α)θ)sinαπ2=cos((1α)θ)cos(1α)π2>0.

    Hence, there is no solution of eq. (45) for θ0,π2 and α∈(0,1).

  4. Case 4: π2<θ<πMultiplying the second equation of (46) by cos(αθ) and subtracting the first equation multiplied by sin(αθ) leads to

    mr2sin((2α)θ)+drsin((1α)θ)ksin(αθ)=0,

    which may be solved for r > 0 as

    (53) r(θ)=d2msin((1α)θ)sin((2α)θ)+d2msin((1α)θ)sin((2α)θ)2+kmsin(αθ)sin((2α)θ).

    Furthermore, multiplying the first equation of (46) by sin(2θ) and subtracting the second equation multiplied by cos(2θ) leads to

    (54) drsin(θ)+crαsin((2α)θ)+ksin(2θ)=0.

    As the first and the last term on the left-hand side of eq. (54) are negative for θπ2,π, the second term has to be positive to solve the equation, i. e.

    sin((2α)θ)>0π2<θ<π2α.

    Now, we consider the left-hand side of eq. (54) as a function of θ

    g(θ):=dr(θ)sin(θ)+crα(θ)sin((2α)θ)+ksin(2θ),θπ2,π2α

    with r(θ) given by eq. (53). The function g is continuous for θπ2,π2α and it holds that

    gπ2=dr+crαsin(2α)π2=dr+crαsinαπ2>0

    as follows from (50). Furthermore, it can be seen from (53), that there exists a constant C > 0, such that

    limθπ2αr(θ)=limθπ2αCsin((2α)θ)=,

    so that

    limθπ2αg(θ)=limθπ2α[ dCsin((2α)θ)sin(π2α)+c(Csin((2α)θ))αsin((2α)θ)+ksin(2π2α) ]=.

    Therefore, there is at least one root of g, i. e. one pair of complex conjugate solutions of eq. (45) such that θπ2,π2α.

   □

As we have found conditions for solutions of the characteristic eq. (45) to be in the left-half complex plane, we want to prove asymptotic stability of the trivial solution by inverse Laplace transform using fundamental ideas of complex analysis. Therefore, we reformulate eq. (44) as

(55) Q(s)=ms+d+csα1ms2+ds+csα+kq0+mms2+ds+csα+kv0cms2+ds+csα+k0μ1α(ω)y0(ω)ω+sdω.

Similar as in [27] we consider the function

(56) Ξ(s):=ms+d+csα1ms2+ds+csα+k

and we compute the inverse Laplace transform ξ(t) of Ξ(s)=L{ξ(t)}(s). As

ξ(0)=limssΞ(s)=1,

we obtain

L{ξ˙(t)}(s)=sΞ(s)ξ(0)=ms2+ds+csαms2+ds+csα+k1=kms2+ds+csα+k,

which, together with eq. (55) leads to the solution

(57) q(t)=q0ξ(t)mkv0ξ˙(t)+ck0μ1α(ω)y(ω,0)0teω(tτ)ξ˙(τ)dτdω

of eq. (17) and we examine the asymptotic behavior of q from ξ and ξ˙. Therefore, we determine the inverse Laplace transform

ξ(t)=12πiσiσ+iΞ(s)estds,Re(σ)>0

with the help of the residue theorem

(58) 12πiΣΞ(s)estds=jResΞ(s)est,sj,

where sj are the roots of eq. (45) and the closed curve Σ (Figure 6) is split up in six parts, such that

ξ(t)=jResΞ(s)est,sj12πilimRε0IIVIΞ(s)estds.

First, we compute the residues for a pair of complex conjugate roots s1, s2=s1ˉ of eq. (45). As s1/2 are simple poles of Ξ, we obtain the residue by derivation of the denominator as

ResΞ(s)est,s1+ResΞ(s)est,s2=ms1+d+cs1α12ms1+d+cαs1α1es1t+ms2+d+cs2α12ms2+d+cαs2α1es2t.

As the two addends are conjugate, we obtain with s1=a+ib=reiθ

ResΞ(s)est,s1+ResΞ(s)est,s2=2Rems1+d+cs1α12ms1+d+cαs1α1es1t=2eatcos(bt)f1(r,θ)f3(r,θ)+2eatsin(bt)f2(r,θ)f3(r,θ).

with

Figure 6: Curve Σ used for integration to apply the residue theorem.
Figure 6:

Curve Σ used for integration to apply the residue theorem.

(59) f1(r,θ)=2m2r2+d2+3mrdcos(θ)+(1+α)cdrα1cos((1α)θ)+(2+α)mcrαcos((2α)θ)+c2αr2(α1),f2(r,θ)=(2α)mcrαsin((2α)θ)+mdrsin(θ)+(1α)cdrα1sin((1α)θ),f3(r,θ)=4m2r2+d2+4mdrcos(θ)+2cdαrα1cos((1α)θ)+4mcαrαcos((2α)θ)+c2α2r2(α1).

We continue considering the contribution of the integral along the paths IIVI to the value of ξ. There is no contribution of II, because

12πiIIΞ(s)estds=12πiπ2πΞσ+ReiϕeσteRcos(ϕ)teiRsin(ϕ)tiReiϕdϕ12πC1(t)eC2RtRπ2RR0,

with C1,C2>0 as Ξ(s)0 for s and cos(ϕ) < 0 for ϕ(π2,π). The same argumentation holds for path VI. For path IV, we obtain

12πiIVΞ(s)estds=12πiππΞεeiϕeεt(cos(ϕ)+isin(ϕ))iεeiϕdϕε00,

as sΞ(s)0 for s0. Finally, for III and V we obtain a contribution

12πiIII,VΞ(s)estds=12πi0ΞωeiπΞωeiπeωtdω=1π0ImΞωeiπeωtdω

with

1πImΞωeiπ=1πkcωα1sin(απ)(mω2dω+k)2+2cωαcos(απ)(mω2dω+k)+c2ω2α=μ1α(ω)kc(mω2dω+k)2+2cωαcos(απ)(mω2dω+k)+c2ω2α.

This leads us to the inverse Laplace transform of Ξ

(60) ξ(t)=jodd(2eajtcos(bjt)f1(rj,θj)f3(rj,θj)+2eajtsin(bjt)f2(rj,θj)f3(rj,θj))+0μ1α(ω)Z(ω)eωtdω

for roots sj/j+1=aj±ibj=rje±iθj of eq. (45) where

Z(ω)=kc(mω2dω+k)2+2cωαcos(απ)(mω2dω+k)+c2ω2α.

The asymptotic behavior of ξ is determined by the exponential functions in the first addends of eq. (60), which decay, as aj<0j, if the inequalities (49) and (50) have a non-empty solution set. Furthermore, the asymptotic behavior of the last term in eq. (60) may be estimated as follows. It holds that

(mω2dω+k)2+2cωαcos(απ)(mω2dω+k)+c2ω2α>(mω2dω+kcωα)20.

Hence, Z is continuous and bounded in [0,) and by the mean value theorem, there exists C3 > 0 such that

0μ1α(ω)Z(ω)eωtdω=C30μ1α(ω)eωtdω=C3tαΓ(1α),

which leads to algebraic decay of order α for the last term in eq. (60) for t. For ξ˙, we obtain the expression

(62) ξ˙(t)=jodd2eajtf3(rj,θj)((ajf1(rj,θj)+bjf2(rj,θj))cos(bjt)+(ajf2(rj,θj)bjf1(rj,θj))sin(bjt))0μ1α(ω)Z(ω)ωeωtdω,

where the first terms again describe an exponentially decaying oscillation and the last term fulfills

0μ1α(ω)Z(ω)ωeωtdω=αC3tα1Γ(1α),

which again implies algebraic decay, this time of order 1 + α for t. To conclude asymptotic stability of the trivial solution of eq. (17) from the asymptotic behavior of ξ, we still have to consider the last term in eq. (57). Therefore, we recall from eq. (8), that the initial infinite state y(ω,0) has the form

(63) y(ω,0)=t00eωτq˙(τ)dτ,

which may be estimated as

(64) |y(ω,0)|=t00eωτq˙(τ)dτ=eωτq(τ)t00ωt00eωτq(τ)dτ2q+ω0eωτdτq=3q,

which is bounded if qCB((,0];R). Furthermore, consider the reformulation

(65) 0teω(tτ)ξ˙(τ)dτ=ddt0teω(tτ)ξ(τ)dτξ(0)eωt

of the inner integral in the last term of eq. (57). The last term in eq. (65) results in a term

0μ1α(ω)y(ω,0)eωtdω3qtαΓ(1α)

in eq. (57). Substitution of the exponential terms of ξ in the last term of eq. (57) using eq. (65) leads to the estimation

0μ1α(ω)y(ω,0)ddt0teω(tτ)esjτdτdω3q0μ1α(ω)1ω+sjsjesjt+ωeωtdω3q(sj0μ1α(ω)ω+sjdωeRe(sj)t+0μ1α(ω)eωtdω)=3qsjαeRe(sj)t+tαΓ(1α).

with roots sj of eq. (45). For the algebraic decay part in ξ we obtain, again using the mean value theorem a constant C4 > 0 and the term

0μ1α(ω)y(ω,0)×ddt0teω(tτ)0μ1α(η)Z(η)eητdηdτdωC4ddt0μ1α(ω)0teω(tτ)ταΓ(1α)dτdω=C4ddtCD0+αt1αΓ(2α)=C4|12α|t2αΓ(22α).

in the last term of eq. (57). In summary, we obtain sufficient conditions (49) and (50) for global asymptotic stability of the equilibrium of eq. (17), from which we retrieve a Lyapunov functional in Section 4.5.2.

Remark 4.9.

It is even possible to obtain a purely exponential solution of eq. (17) without algebraic decay. Choose the initial function q(τ)=esjτ for τ(,0] (which is not in CB((,0];R)) for a root sj of eq. (45). This leads to initial conditions

q˙(τ)=sjesjt,y(ω,0)=0eωτsjesjτdτ=sjω+sj.

Using this function in the Laplace transform (55) leads to

Q(s)=ms+d+csα1ms2+ds+csα+k+msjms2+ds+csα+kcsjms2+ds+csα+k0μ1α(ω)(ω+s)(ω+sj)dω=ms+d+csα1+msjms2+ds+csα+kcsj0μ1α(ω)ω+sjdω0μ1α(ω)ω+sdω(ssj)(ms2+ds+csα+k)=ms+d+csα1+msj(ssj)(ssj)(ms2+ds+csα+k)csjsjα1sα1(ssj)(ms2+ds+csα+k)=1ssj.

Hence, the solution is

q(t)=esjt,t,

which shows that the integral term in eq. (42) should in general not be omitted.

4.5.2 Lyapunov functional

To formulate a Lyapunov functional, we will use the identities of the next proposition.

Proposition 4.10.

For α∈(0,1) and r > 0, the identities

(66) 0μ1α(ω)ω2+r2dω=cos(απ2)rα2

(67) 0μ1α(ω)ωω2+r2dω =sin(απ2)rα1

hold.

Proof.

Substitute η = ω2 and dη=2ωdω in the integral and obtain

0μ1α(ω)ω2+r2dω=sin(απ)π0ωα1ω2+r2dω=sin(απ)2π0ηα21η+r2dη=sin(απ)2sinαπ20μ1α2(η)η+r2dη.

Using sine-double-angle formula and eq. (41), we directly obtain eq. (66). The proof of eq. (67) is analogous.   □

In the following, let r* be the solution of the first equation of (47). We reformulate eq. (17) using eqs. (7), (36) and Proposition 4.10 as

(68) mq¨(t)=kq(t)dq˙(t)cr20μ1α(ω)ω2+r2y(ω,t)dωc0μ1α(ω)ωω2+r2ωy(ω,t)dω=k+ccosαπ2rαq(t)d+csinαπ2rα1q˙(t)+cr20μ1α(ω)ω2+r2ωY(ω,t)dω+c0μ1α(ω)ω2+r2ωy˙(ω,t)dω.

Introducing the quantities

(69) k˜=k+ccosαπ2rα,d˜=d+csinαπ2rα1,

and new coordinates

(70) q˜(t)=q(t)ck˜r20μ1α(ω)ω2+r2ωY(ω,t)dω=kk˜q(t)+ck˜r20μ1α(ω)ω2+r2y(ω,t)dω

and

(71) v˜(t)=q˙(t)cm0μ1α(ω)ω2+r2ωy(ω,t)dω

we obtain the system

(72) { q ˜ . (t)=   v ˜ . (t), v ˜ . (t),= k ˜ m q ˜ (t) d ~ m v ˜ (t) d ~ c m 2 0 μ 1α (ω) ω 2 + r * 2 ωy(ω,t)dω.

Note that the first equation in (72) holds, as r* is a solution of the first equation of (47). Moreover, we consider the candidate Lyapunov functional

(73) V3(t,qt,vt)=m2v˜t2(0)+k˜2q˜t2(0)+d˜c2m0μ1α(ω)ω2+r2ωy2(ω,t)dω

and prove inequality (26) for V3 w.r.t. the functions qt and vt. Therefore, consider the split of the integral term in eq. (73)

0μ1α(ω)ω2+r2ωy2(ω,t)dω=01μ1α(ω)ω2+r2ωy2(ω,t)dω+1μ1α(ω)ω2+r2ωy2(ω,t)dω

and use the mean value theorem for the first term and the inequality ω ≥ 1 in the second term to find a constant C˜>0, such that

(74) 0μ1α(ω)ω2+r2ωy2(ω,t)dωC˜0μ1α(ω)ω2+r2y2(ω,t)dω.

Moreover, we use Hölder’s inequality to obtain

(75) 0μ1α(ω)ω2+r2y(ω,t)dω20μ1α(ω)ω2+r2dω0μ1α(ω)ω2+r2y2(ω,t)dω

and

(76) 0μ1α(ω)ω2+r2ωy(ω,t)dω20μ1α(ω)ωω2+r2dω0μ1α(ω)ω2+r2ωy2(ω,t)dω.

Using the three inequalities above and Proposition 4.10, we can estimate (73) as

(77) V3(t,qt,vt)m2v˜t2(0)+d˜c4msinαπ2rα1×0μ1α(ω)ω2+r2ωy(ω,t)dω2+k˜2q˜t2(0)+d˜cC˜4mcosαπ2rα2×0μ1α(ω)ω2+r2y(ω,t)dω2

Finally, applying the general relation

(78) (a+b)2+γb2=γ1+γa2+a1+γ+1+γb2

for a,b,γR, γ>0 on the first two and the last two terms of (77) using eqs. (70) and (71), we obtain inequality (26) for V3. Furthermore, we compute the rate of V3 as

V˙3=mv˜˙t(0)v˜t(0)+k˜q˜t(0)q˜˙t(0)+d˜cm0μ1α(ω)ω2+r2ωy(ω,t)y˙(ω,t)dω.

Inserting the dynamics from (72), we obtain

(79) V˙3=d˜v˜2(t)+d˜c2m2(0μ1α(ω)ω2+r*2ωy(ω,t)dω)2

(80) d˜cm0μ1α(ω)ω2+r*2ω2y2(ω,t)dω.

Again, using Hölder’s inequality leads to

(81) 0μ1α(ω)ω2+r2ωy(ω,t)dω20μ1α(ω)ω2+r2dω0μ1α(ω)ω2+r2ω2y2(ω,t)dω

and together with eq. (66), we finally obtain

(82) V˙3d˜v˜2(t)d˜cm1cmcosαπ2rα2×0μ1α(ω)ω2+r2ω2y2(ω,t)dω

where, due to eq. (47)

1cmcosαπ2rα2>0mr2ccosαπ2rα=k>0.

In summary, we found a Lyapunov functional V3, such that V˙30, which has the form of an energy functional w.r.t. the new coordinates q˜t and v˜t but it has the disadvantage that we cannot prove asymptotic stability of the trivial solution of eq. (17) with the help of V3 as in the classical case of the damped linear oscillator using the total mechanical energy as a Lyapunov function. Hence, we introduce another Lyapunov functional V4, which is related to the functional V2 in Section 4.4, to prove the following proposition.

Proposition 4.11.

Let m,k,c > 0, α∈(0,1) and let r = r* > 0 be the solution of

mr2+crαcosαπ2+k=0.

Let dR be such that the inequality

dr+crαsinαπ2>0

holds. Then the trivial solution of eq. (17) is asymptotically stable.

Proof.

We introduce the Lyapunov functional

(83) V4(t,qt,vt)=12mv˜t2(0)+12k˜q˜t2(0)+d˜24mq˜t2(0)+12d˜q˜t(0)v˜t(0)+d˜c2m0μ1α(ω)ω2+r2ωy2(ω,t)dω+d˜2c4m20μ1α(ω)ω2+r2ω2Y2(ω,t)dωd˜2c4m2ck˜r20μ1α(ω)ω2+r2ωY(ω,t)dω2

and check the conditions in Theorem 4.3. Once again, we use Hölder’s inequality to obtain

0μ1α(ω)ω2+r2ωY(ω,t)dω20μ1α(ω)ω2+r2dω0μ1α(ω)ω2+r2ω2Y2(ω,t)dω.

Together with (76), estimation (38) and Proposition 4.10 we obtain

V4(t,qt,vt)14mv˜t2(0)+12k˜q˜t2(0)+d˜c2msin(απ2)rα10μ1α(ω)ω2+r2ωy(ω,t)dω2+d˜2c4m2ck˜r2k˜ccos(απ2)rα1×0μ1α(ω)ω2+r2ωY(ω,t)dω2.

All coefficients in this estimation are positive and, again using eqs. (70), (71) and (78), we obtain relation (26) for V4. Furthermore, we compute the rate of V4 as

V˙4=mv˜t(0)v˜˙t(0)+k˜q˜t(0)q˜˙t(0)+d˜2v˜t(0)q˜˙t(0)+d˜2q˜t(0)v˜˙t(0)+d˜22mq˜t(0)q˜˙t(0)+d˜cm0μ1α(ω)ω2+r2ωy(ω,t)y˙(ω,t)dω+d˜2c2m20μ1α(ω)ω2+r2ωY(ω,t)ωY˙(ω,t)dωd˜2c2m2ck˜r20μ1α(ω)ω2+r2ωY(ω,t)dω×0μ1α(ω)ω2+r2ωY˙(ω,t)dω

Inserting the dynamics and using eqs. (81) and (74), we obtain

V˙4d˜2v˜2(t)d˜cm1cmcosαπ2rα2×0μ1α(ω)ω2+r2ω2y2(ω,t)dωk˜d˜2mq˜2(t)d˜2cC˜2m20μ1α(ω)ω2+r2y2(ω,t)dω.

Again, with the help of the estimations (75), (81) and eqs. (70), (71) and (78), one can prove that V˙4 fulfills inequality (29) w.r.t. q and v. This completes the proof.   □

Remark 4.12.

  1. As a special case, Proposition 4.11 proves asymptotic stability of the trivial solution of eq. (17) for the case d = 0. Previously in Section 4.3, with the help of the energy functional we could only prove stability but not attractivity of the trivial solution.

  2. The conditions for asymptotic stability in Proposition 4.11 are equivalent to the necessary and sufficient conditions obtained by the eigenvalue analysis in Proposition 4.7. In that sense, the choice of the functionals V3 and V4 is optimal.

4.5.3 Example: Stick-slip oscillator [293031]

In this section, we describe a model of a mechanical system, where effective negative linear damping occurs in the linearization of the equation of motion around an equilibrium and we give sufficient conditions for local asymptotic stability of the equilibrium using Proposition 4.11. Consider a mass m suspended by a spring with spring coefficient k, and a springpot with coefficient c and differentiation order α∈(0,1), which is sliding on a conveyor belt as in Figure 7. By q we denote the displacement of the mass. The belt moves with a constant velocity vdr>0 in the direction of q and we assume friction between the mass and the belt, which leads to a friction force FT, such that the equation of motion reads as

(84) mq¨(t)=FTcCDt0+αq(t)kq(t).
Figure 7: Fractionally damped oscillator with dry friction and graph of a set-valued force law describing the Stribeck effect.
Figure 7:

Fractionally damped oscillator with dry friction and graph of a set-valued force law describing the Stribeck effect.

For the friction force FT in the slip phase we consider the force law

(85) FT=μ(vrel)FNsign(vrel),vrel0

where

(86) vrel(t)=q˙(t)vdr

is the relative velocity between mass and belt,

(87) FN=mg

is the normal force acting on the mass and µ = µ(v) is the friction coefficient depending on the relative velocity, where the function µ increases (at least) for small negative values of vrel, i. e.

(88) μ(|v|)>0,|v|1,

which is known as the Stribeck effect. We consider the slip equilibrium q* of eq. (84)

(89) 0=FTkq*,

(90) FT=μ(vdr)FNsign(vdr)=μ(vdr)mg,

which implies

(91) q=μ(vdr)mgk

We introduce a new coordinate

(92) qˉ=qq,

such that we reformulate eq. (84) in terms of qˉ as

(93) mqˉ¨(t)+cCDt0+αqˉ(t)+k(qˉ(t)+q)=μ(qˉ˙(t)vdr)mgsign(qˉ˙(t)vdr).

Linearizing the right-hand side of eq. (93) near the equilibrium leads to

(94) mqˉ¨(t)+cCDt0+αqˉ(t)+k(qˉ(t)+q)=μ(vdr)mg+μ(vdr)mgqˉ˙(t)+Oqˉ˙2,

which together with eq. (91) leads to the linearized equation

(95) mqˉ¨(t)μ(vdr)mgqˉ˙(t)+cCDt0+αqˉ(t)+kqˉ(t)=Oqˉ˙2

Hence, using Proposition 4.11, we obtain the condition

(96) crα1sinαπ2>μ(vdr)mg

for local asymptotic stability of the slip equilibrium q*, where again r* is the solution of

mr2+crαcosαπ2+k=0.

5 Conclusion

The previous sections have given a complete stability analysis of a linear single degree-of-freedom mass-spring-dashpot-springpot system, including an eigenvalue analysis, a derivation of total mechanical energy and Lyapunov functionals for various cases. The major merit of the paper lies in the extension of the Lyapunov stability framework to fractionally damped mechanical systems as this step is essential for rigorous proofs of global stability properties of nonlinear systems and further bifurcation analysis. Specifically, the results in Section 4 show how the terms in the Lyapunov functional related to the springpot element have to be split, i. e. how one has to deal with fractional damping in a Lyapunov setting.

The scope has been limited to a single degree-of-freedom oscillator with viscous and fractional damping. Obviously, these results need to be generalized to multi degree-of-freedom systems. That such a generalization is possible, at least in special cases, can be seen by looking at a linear multi degree-of-freedom system

Mq¨+Dvq˙+DfDt0+αq+Kq=0

with symmetric system matrices. If the viscous damping matrix Dv and fractional damping matrix Df are proportional to the mass and stiffness matrices (or, more generally, Caughey damping), then one can use known results from linear vibration analysis to show that the modal equations are decoupled single degree-of-freedom fractionally damped oscillators as dealt with here. Clearly, further research is needed for more general cases.

A key result of the paper is the Lyapunov stability condition for systems having negative viscous damping but a sufficient amount of positive fractional damping. The example of a stick-slip oscillator demonstrates that negative viscous damping is relevant. More generally, all Hopf bifurcation instabilities in non-conservative mechanical systems are due to negative effective viscous damping. The mechanism leading to effective negative damping may be quite complicated, e. g. be caused by mode coupling, follower forces or aerodynamic forces (flutter), and also be attributed to Ziegler’s paradox. A typical task within control theory is the design of a feedback law which stabilizes the equilibrium and a common tool to achieve this is Lyapunov-based control design. Hence, Lyapunov methods can be used to stabilize systems with negative effective viscous damping, either with integer or fractional PID control. The Lyapunov techniques developed here, specifically the results on anti-damping, may prove to be instrumental for this purpose.

Award Identifier / Grant number: 01IS17096B

Funding statement: This work was supported by the Federal Ministry of Education and Research of Germany (BMBF) (Funder Id: http://doi.org/10.13039/501100002347, Grant Number: 01IS17096B).

References

[1] K. Diethelm, The analysis of fractional differential equations: An application-oriented exposition using differential operators of Caputo type, Lecture notes in mathematics, Springer, Berlin, 2010.10.1007/978-3-642-14574-2_8Suche in Google Scholar

[2] I. Podlubny, Fractional differential equations. An introduction to fractional derivatives, fractional differential equations, to methods of their solution and some of their applications, Mathematics in science and engineering Vol. 198, Academic Press, San Diego, 1999, XXIV.Suche in Google Scholar

[3] H. Sun, Y. Zhang, D. Baleanu, W. Chen and Y. Chen, A new collection of real world applications of fractional calculus in science and engineering, Commun. Nonlinear Sci. Numer. Simul. 64 (2018), 213–231.10.1016/j.cnsns.2018.04.019Suche in Google Scholar

[4] R. L. Bagley and P. J. Torvik, Fractional calculus in the transient analysis of viscoelastically damped structures, AIAA J. 23 (1985), 918–925.10.2514/6.1983-901Suche in Google Scholar

[5] A. Schmidt and L. Gaul, Finite element formulation of viscoelastic constitutive equations using fractional time derivatives, Nonlinear Dyn. 29 (2002), 37–55.10.1023/A:1016552503411Suche in Google Scholar

[6] P. Shah and S. Agashe, Review of fractional PID controller, Mechatronics 38 (2016), 29–41.10.1016/j.mechatronics.2016.06.005Suche in Google Scholar

[7] H. K. Khalil, Nonlinear systems, 3. ed., Prentice Hall, Upper Saddle River, 2002.Suche in Google Scholar

[8] T. A. Burton, Volterra integral and differential equations, Mathematics in science and engineering Vol. 202, Amsterdam, Elsevier, 2005.Suche in Google Scholar

[9] T. A. Burton, Stability and periodic solutions of ordinary and functional differential equations, Mathematics in science and engineering Vol. 178, Academic Press, Orlando, 1985.Suche in Google Scholar

[10] J. K. Hale, Theory of functional differential equations, 2 ed, Applied mathematical sciences Vol. 3, Springer, New York Heidelberg Berlin, 1977.10.1007/978-1-4612-9892-2Suche in Google Scholar

[11] V. B. Kolmanovskii and V. R. Nosov, Stability of functional differential equations, Mathematics in science and engineering Vol. 180, Academic Press, London, 1986.Suche in Google Scholar

[12] J. P. LaSalle and Z. Artstein, The stability of dynamical systems, Regional Conference Series in Applied Mathematics, vol. 25, Society for Industrial and Applied Mathematics, Philadelphia, 1976.Suche in Google Scholar

[13] J.-C. Trigeassou, N. Maamri and A. Oustaloup, Lyapunov stability of commensurate fractional order systems: A physical interpretation, J. Comput. Nonlinear Dyn. 11 (2016), 051007.10.1115/1.4032387Suche in Google Scholar

[14] J.-C. Trigeassou, N. Maamri, J. Sabatier and A. Oustaloup, A Lyapunov approach to the stability of fractional differential equations, Signal Process. 91 (2011), 437–445.10.1016/j.sigpro.2010.04.024Suche in Google Scholar

[15] M. Hinze, A. Schmidt and R. I. Leine, Mechanical representation and stability of dynamical systems containing fractional springpot elements, Proceedings of the IDETC Quebec, Canada, 2018.10.1115/DETC2018-85146Suche in Google Scholar

[16] K. D. Papoulia, V. P. Panoskaltsis, N. V. Kurup and I. Korovajchuk, Rheological representation of fractional order viscoelastic material models, Rheologica Acta 49 (2010), 381–400.10.1007/s00397-010-0436-ySuche in Google Scholar

[17] H. Schiessel and A. Blumen, Hierarchical analogues to fractional relaxation equations, J Phys. A Math. Gen. 26 (1993), 5057–5069.10.1088/0305-4470/26/19/034Suche in Google Scholar

[18] G. Montseny, Diffusive representation of pseudo-differential time-operators, ESAIM: Proc. 5 (1998), 159–175.10.1051/proc:1998005Suche in Google Scholar

[19] D. Matignon, Stability properties for generalized fractional differential systems, ESAIM: Proc. 5 (1998), 145–158.10.1051/proc:1998004Suche in Google Scholar

[20] J. Trigeassou, N. Maamri, J. Sabatier and A. Oustaloup, State variables and transients of fractional order differential systems, Comput. Math. Appl. 64 (2012), 3117–3140. Advances in FDE, III.10.1016/j.camwa.2012.03.099Suche in Google Scholar

[21] J.-C. Trigeassou, N. Maamri, J. Sabatier and A. Oustaloup, Transients of fractional-order integrator and derivatives, Signal Image Video Process. 6(3) (2012), 359–372.10.1007/s11760-012-0332-2Suche in Google Scholar

[22] D. Matignon, Stability results for fractional differential equations with applications to control processing, Computational Engineering in Systems Applications Multiconference, IMACS, IEEE-SMC, Lille, France (1996), 963–968.Suche in Google Scholar

[23] J. Sabatier, M. Moze and C. Farges, LMI stability conditions for fractional order systems, Comput. Math. Appl. 59 (2010), 1594–1609.10.1016/j.camwa.2009.08.003Suche in Google Scholar

[24] R. Agarwal, D. O’Regan and S. Hristova, Stability of Caputo fractional differential equations by Lyapunov functions, Appl. Math. 60 (2015), 653–676.10.1007/s10492-015-0116-4Suche in Google Scholar

[25] M. A. Duarte-Mermoud, N. Aguila-Camacho, J. A. Gallegos and R. Castro-Linares, Using general quadratic Lyapunov functions to prove Lyapunov uniform stability for fractional order systems, Commun. Nonlinear Sci. Numer. Simul. 22 (2015), 650–659.10.1016/j.cnsns.2014.10.008Suche in Google Scholar

[26] Y. Li, Y. Q. Chen and I. Podlubny, Stability of fractional-order nonlinear dynamic systems: Lyapunov direct method and generalized Mittag–Leffler stability, Comput. Math. Appl. 59 (2010), 1810–1821.10.1016/j.camwa.2009.08.019Suche in Google Scholar

[27] L.-L. Liu and J.-S. Duan, A detailed analysis for the fundamental solution of fractional vibration equation, Open Math. 13 (2015), 826–838.10.1515/math-2015-0077Suche in Google Scholar

[28] M. Naber, Linear fractionally damped oscillator, Int. J. Diff. Eq. 2010 (2010), 1–12, Article ID: 197020.10.1155/2010/197020Suche in Google Scholar

[29] U. Galvanetto, S. R. Bishop and L. Briseghella, Mechanical stick–slip vibrations, Int. J. Bifurcation Chaos 5 (1995), 637–651.10.1142/S0218127495000508Suche in Google Scholar

[30] R. A. Ibrahim, Friction-induced vibration, chatter, squeal, and chaos; Part I: Mechanics of contact and friction, ASME Appl. Mech. Rev. 47 (1994), 209–226.10.1115/1.3111079Suche in Google Scholar

[31] R. I. Leine and H. Nijmeijer, Dynamics and bifurcations of non-smooth mechanical systems, Lecture notes in applied and computational mechanics Vol. 18, Springer, Berlin Heidelberg New York, 2004, XII.10.1007/978-3-540-44398-8Suche in Google Scholar

Received: 2018-12-21
Accepted: 2020-02-02
Published Online: 2020-02-25
Published in Print: 2020-07-28

© 2020 Hinze et al., published by De Gruyter

This work is licensed under the Creative Commons Attribution-NonCommercial-NoDerivatives 4.0 International License.

Heruntergeladen am 1.10.2025 von https://www.degruyterbrill.com/document/doi/10.1515/ijnsns-2018-0381/html
Button zum nach oben scrollen