DISCRETE AND CONTINUOUS Website: http://AIMsciences.org DYNAMICAL SYSTEMS Volume 9,Number2,March2003 pp. 339–358

OSCILLATIONS IN A SECOND-ORDER DISCONTINUOUS SYSTEM WITH DELAY

Eugenii Shustin

School of Mathematical Sciences University Ramat Aviv, 69978 Tel Aviv, Emilia Fridman

Department of & Systems Tel Aviv University Ramat Aviv, 69978 Tel Aviv, Israel Leonid Fridman

Division of Postgraduate and Investigation Chihuahua Institute of Technology Chihuahua, Chi, C.P. 31160,

(Communicated by Hans-Otto Walther)

Abstract. We consider the equation

αx(t)=−x(t)+F (x(t),t) − sign x(t − h),α=const> 0,h=const> 0,

which is a model for a scalar system with a discontinuous negative delayed feedback, and study the dynamics of oscillations with emphasis on the existence, frequency and stability of periodic oscillations. Our main conclusion is that, in the autonomous case F (x, t) ≡ F (x), for |F (x)| < 1, there are periodic solutions with different frequencies of oscillations, though only slowly-oscillating solutions are (orbitally) stable. Under extra conditions we show the uniqueness of a periodic slowly-oscillating solution. We also give a criterion for the existence of bounded oscillations in the case of unbounded function F (x, t). Our approach consists basically in reducing the problem to the study of dynamics of some discrete scalar system.

Introduction. Equations with discontinuity often appear in various models [2, 4, 5, 10, 11, 15, 18]. Time delay, always existing in real systems, usually results in oscillations around the discontinuity surface. In the present paper we study various aspects of such oscillations for the system αx(t)=−x(t)+F (x(t),t) − sign x(t − h) , (0.1) where F is a smooth function and α, h are positive constants (the positivity of α means that − sign x(t − h) provides a negative feedback). An important point is a

1991 Mathematics Subject Classification. 34K11, 34K13, 34K35. Key words and phrases. delay-differential equation, periodic solutions, orbital stability.

339 340 E. SHUSTIN, E. FRIDMAN, L. FRIDMAN choice of sign 0. From the control theory point of view signum should be a binary sensor or actuator, i.e., taking only values ±1. Following [1], we choose  1 , as x(t) > 0, sign x(t)= −1 , as x(t) < 0,  z(t) , as x(t)=0, where z(t) is any measurable function with |z(t)| = 1. The convenience of this choice is explained by Lemma 1.1 below, which claims that, for any solution of (0.1), mes{t ≥ 0:x(t)=0} =0. For small α, equation (0.1) can be viewed as a singular perturbation of the equation x(t)=F (x(t),t) − sign x(t − h) , (0.2) which has thoroughly been studied in [7, 16] (see also [20], where signum is replaced by a close continuous odd function), and we show that equation (0.1) inherits the basic properties of (0.2), especially, in the case

|F (x, t)|≤p =const< 1 . (0.3)

First of all, for oscillating solutions of (0.1) we define a frequency function which is non-increasing as a similar function for equation (0.2). Moreover, stability is observed for the only oscillating solutions with a minimal oscillation frequency, which, in fact, is the case for similar systems of the first order [6, 14, 16, 19]. The autonomous case F (x, t) ≡ F (x) is of special interest. In contrast to the first-order autonomous system (0.2), whose slowly oscillating solutions coincide up to shifts in t, the variety of slowly oscillating solutions of (0.1) reveals a more complicated structure. We prove the existence of periodic slowly oscillating solutions and show, under some conditions, the uniqueness of such a solution, and in the last case, the orbital asymptotic stability of the periodic solution. We show that there always exist periodic solutions with oscillation frequency greater than the minimal one. These fast oscillations seem to be unstable (as occurs in (0.2)), but it is still a question. We point out the difference between the following two situations. Under assump- tion (0.3), all solutions oscillate around zero, and the main problem is to describe the periodic and stable solutions. If |F (x)| can be greater than 1, one observes both bounded oscillatory solutions and unbounded solutions, in which case we give a criterion for the existence of bounded solutions. The latter problem reflects the following phenomenon: the associated equation αx(t)=−x(t)+F (x(t)) does not have nontrivial bounded solutions, whereas equation (0.1) does. This has been studied in detail for the first-order system (0.2) in [6, 16] and for system (0.1) with F (x)=kx, k>0, in [17]. We should mention that a similar equation x(t)=−αx(t)+f(x(t − h)) with a two-level piecewise constant function f was considered in [3, 8, 9] with emphasis on the existence and stability of periodic solutions. The material is organized as follows. Section 1 to 3 contain statements of results and discussion, section 4 contains the proofs. Acknowledgments. We would like to thank the anonymous referee for his valuable remarks which allowed us to correct mistakes and improve the readability of the paper. OSCILLATIONS IN A SECOND-ORDER SYSTEM 341

1. Oscillatory solutions. For equation (0.1) we state the initial value problem by defining the initial data space to be the space C0[−h, 0] of continuous functions ϕ :[−h, 0] → R differentiable at the origin. We intend to characterize oscillations in (0.1) via the distribution of zeroes of solutions, and the following Lemma provides basics for the further study. Lemma 1.1. Equation (0.1), satisfying (0.3), with initial condition x(t)=ϕ(t),t∈ [−h, 0],ϕ∈ C0[−h, 0],x(0) = ϕ (0), (1.4) has a unique continuous solution xϕ(t), t ∈ [−h, ∞), and this solution satisfies

mes{t ≥ 0:xϕ(t)=0} =0. (1.5)

Moreover, in the interval (0, ∞), xϕ is differentiable, its derivative is absolutely continuous and differentiable almost everywhere.

1.1. Frequency of oscillations. Denote by Zϕ the set of zeroes of xϕ(t)in [−h, ∞). We say that xϕ changes sign at an isolated zero t ∈ Zϕ if xϕ(t − ε)xϕ(t + 0 ε) < 0 for all sufficiently small ε>0. Denote by Zϕ the set of zeroes t ∈ Zϕ where xϕ changes sign. A solution xϕ(t) is said to be oscillatory if Zϕ is unbounded.

Lemma 1.2. For any ϕ ∈ C0[−h, 0], a solution of the initial value problem (0.1), (0.3), (1.4) is oscillatory.

For any (oscillatory) solution xϕ put ∗ ∗ ∞, if card(Zϕ ∩ (t − h, t )) = ∞, νϕ t ( )= 0 ∗ ∗ ∗ ∗ card(Zϕ ∩ (t − h, t )), if card(Zϕ ∩ (t − h, t )) < ∞, ∗ where t ≥ 0, t = max{τ ≤ t : xϕ(τ)=0}. Define the frequency function by νϕ(t)+1 νϕ(t)=∞ if νϕ(t)=∞, νϕ(t)= if νϕ(t) < ∞, 2 where the brackets mean the entire part. The following property of the frequency function, analogous to one in the case of the first-order equation [7, 12, 13, 16, 19] determines the leading role of this parameter in the theory.

Theorem 1.3. If xϕ is a solution of the initial value problem (0.1), ((0.3), 1.4), then the function νϕ(t) is non-increasing.

Notice that in the first-order case the function νϕ itself is non-increasing. Theorem 1.3 immediately yields

Corollary 1.4. Under the conditions of Theorem 1.3, for any solution xϕ, there exists the limit frequency (shortly, l-frequency)

Nϕ = lim νϕ(t) ∈ N ∪{0}∪{∞}. t→∞

In particular, the set of initial functions C0[−h, 0] splits into the disjoint union ∞ C0[−h, 0] = Un ∪ U∞,Un = {ϕ ∈ C0[−h, 0] : Nϕ = n}. n=0 We shall see later that the sets Un, n>0, can be nonempty. However, Theorem 1.3 indicates that the set U0 constitutes the main part of C0[−h, 0]. Namely, the next claim shows that the property of having zero l-frequency is stable in the following sense. 342 E. SHUSTIN, E. FRIDMAN, L. FRIDMAN

T’a Ta

Figure 1. Solution xa

Theorem 1.5. Under the condition of Theorem 1.3, the (nonempty) set U 0 {ϕ ∈ C −h, N , ϕ−1 } 0 = 0[ 0] : ϕ =0 mes( (0)) = 0 is contained in the interior of U0 in C0[−h, 0] with respect to the norm 0 ||ϕ||∞ = max |ϕ| + |ϕ (0)| . [−h,0] U For the first-order system (0.2), satisfying (0.3), the set n>0 n (under some mild conditions [7, 16]) is nowhere dense in C[−h, 0]. We conjecture that the similar result holds for the second order system (0.1), satisfying (0.3).

2. Periodic and stable solutions of the autonomous equation. We intend to study in more detail solutions of the autonomous equation αx(t)=−x(t)+F (x(t)) − sign x(t − h) . (2.6) Any solution of the first-order autonomous system (0.2) satisfying (0.3) becomes periodic after a period of time [7, 16]. This is not the case in our situation, however, periodic solutions play the central role in the study. Theorem 2.1. For any integer n ≥ 0, equation (2.6), satisfying (0.3), has a periodic solution with a constant frequency function ν(t) ≡ n.

Theorem 1.5 states, in particular, that the frequency function νϕ(t) with Nϕ =0 is stable if ϕ has finitely many zeroes in [−h, 0], but this does not imply that the corresponding solution xϕ(t) is stable in any sense. The zero l-frequency periodic solutions are natural candidates to be stable, and we show that this is, in fact, the case. Moreover, the stability problem is related to the question on the uniqueness of the zero l-frequency periodic solution. Since the solutions of (2.6) are invariant with respect to shifts in t, we consider only solutions with zero l-frequency such that x(0) = 0,x(t) < 0,t∈ (−h, 0) , (2.7) and ask the question: How many periodic solutions of (2.6) with zero l-frequency, satisfying (2.7), are there ? OSCILLATIONS IN A SECOND-ORDER SYSTEM 343

We reformulate the question as follows. As a consequence of Lemma 1.2 and Theorem 1.3, we obtain that for any a ≥ 0, there is a zero l-frequency solution xa(t) to (2.6), (0.3), satisfying (2.7) and the condition x(0) = a. Moreover, this solution is uniquely defined in the interval [0, ∞). Define the map Φ : [0, ∞) → [0, ∞)by Φ(a)=xa(Ta), where Ta is the second positive zero of xa(t) (see Figure 1). Clearly, if Φ(a)=a then xa is periodic. In the proof of Theorem 2.4, section 4.7 below, we shall show that all periodic zero l-frequency solutions correspond to the roots of the equation Φ(a)=a. So, the above question reduces to counting roots of the latter equation. Since the function F (x) in (2.6) is smooth, so is Φ. A periodic zero l-frequency solution xa(t) is said to be simple if the root a of the function Φ(a) − a has multiplicity one. Conjecture 2.2. Equation (2.6), satisfying (0.3), has only one zero l-frequency periodic solution with property (2.7), and this solution is simple. According to [17], Conjecture 2.2 holds in the case F (x)=kx, provided the constant k satisfies h2 +4α2π2 0 ≥ k ≥− , 4αh or √ √ 1+4αk − 1 1+4αk k>0, exp −h > √ . 2α 1+4αk +1 Here we supplement this result with Theorem 2.3. (1) Conjecture 2.2 holds if F (x) ≡ const. (2) There exists a positive function A(M,p,h), p ∈ [0, 1), M ≥ 0, such that Conjecture 2.2 holds for any α ≤ A(M,p,h), provided, |F (x)|≤p, |dF/dx|≤M.

A solution xϕ(t) of equation (2.6) is called orbitally asymptotically stable if, for any ψ ∈ C0[−h, 0] in a small neighborhood of ϕ there is t0 with

lim |xψ(t) − xϕ(t + t )| =0. (2.8) t→∞ 0

A periodic zero l-frequency solution xa(t) is called quasi-stable if a is an attractor of the dynamical system defined by Φ. Clearly, an orbitally-stable periodic zero l-frequency solution is quasi-stable. Theorem 2.4. (1) If the function F (x) is analytic and satisfies (0.3), then equation (2.6) has a quasi-stable zero l-frequency periodic solution. (2) If equation (2.6), satisfying (0.3), has only one zero l-frequency periodic so- lution with property (2.7), and this solution is simple, then it is orbitally asymptot- ically stable. Moreover, any other zero l-frequency solution approaches the periodic solution in the sense of (2.8). We should like to comment the difference between quasi-stable and orbitally- stable solutions. Assume, for example, that F is twice differentiable, and so is Φ, and that a non-simple stable root a0 of the function Φ(a) − a has multiplicity ≥ 2, in particular, that, for ε>0 small, 2 a0 + ε>Φ(a0 + ε) ≥ a0 + ε − αε ,α>0 .

Let a1 = a0 + ε0 be close to a0. Then Φ generates the sequence an+1 =Φ(an), n ≥ 1, such that εn = an+1 − a0, n ≥ 1, satisfy 2 εn+1 ≥ εn − αεn,n≥ 1 . 344 E. SHUSTIN, E. FRIDMAN, L. FRIDMAN

This in turn implies that C ∞ ε ≥ ,n≥ ,C > , ⇒ ε ∞ . n n 1 =const 0 = n = (2.9) n=0

Suppose, in addition, that dTa/da > 0 at the point a = a0, where Ta is the second positive zero of xa. Then, for any a1 >a0, the 2n-th positive zero T of the solution x t a1 ( ) satisfies n−1 T ≥ nT C · ε ,C > , a0 + 1 i 1 =const 0 i=0 x which in view of (2.9) says that the solution a0 is not orbitally stable. Remark 2.5. One may ask what is common in the three proven cases of Conjecture 2.2, when F (x)=kx,orF (x) = const,orα is sufficiently small, and what is the difficulty in handling the general situation. In principle, one could write certain implicit equations for the function Φ introduced above, but such general expressions seem to be hardly treatable. It is easy to show (see section 5) that Φ takes the interval [0, 2] into itself, so one can expect the uniqueness and simplicity of the root of the equation Φ(a)=a when −1 < Φ(a) < 1. This, for instance, holds for small values of α (the case (2) of Theorem 2.3). Indeed, then equation (0.1) is close to (0.2), which corresponds to the constant function Φ. We omit the proof of Theorem 2.3(2), since it is just a routine computation. For (relatively) large values of α, Φ(a) can be greater than 1, the second de- rivative Φ(a) can change sign etc. One can produce such examples playing with formulas of section 4.8 below. If F (x)=kx,orF (x) = const, we integrate equation (0.1) and find implicit formulas for Φ via elementary functions, which finally reduce (in different ways) to one equation in one unknown with two parameters. To prove the uniqueness of the root of Φ(a)=a, we then perform rather delicate (and rather different) computations, which do not indicate if this argument can be generalized in any way.

3. Existence of bounded solutions in the presence of an unbounded per- turbation. If the function F (x, t) (which we call perturbation) in equation (0.1) does not satisfy |F |≤1, one can observe unbounded solution. In this situation, we intend to provide a sufficient condition for the existence of bounded oscillatory solutions, which, furthermore, form an open set in the space of all solutions. Theorem 3.1. Let a smooth function F (x, t) satisfy for all x, t, F x, t − p F ,t p ,p ∈ − , , ≤ ( ) 0 ≤ k, (0 )= 0 0 =const ( 1 1) 0 x (3.10) where the positive constant k satisfies √ √ 1+4αk − 1 (1 + |p |)(2 + ( 1+4αk − 1)(1 − e−τ/α)) exp −h ≥ 0 , (3.11) 2α 4 and τ is the positive root of the equation |p | α − e−τ/α τ − 1+ 0 . (1 )= k (3.12) Then any solution x(t) of equation (0.1) such that x(t) < 0,t∈ [−h, 0),x(0) = 0, 0 ≤ x (0) <ξ(h, α, k, p0) (3.13) OSCILLATIONS IN A SECOND-ORDER SYSTEM 345 or x(t) > 0,t∈ [−h, 0),x(0) = 0, 0 ≥ x (0) > −ξ(h, α, k, −p0) , (3.14) is bounded, oscillatory, and has zero l-frequency, where √ 1+4αk − 1 2 ξ(h, α, k, p)= 2exp −h − 1 − p · √ . 2α 1+4αk − 1 In the following special case we can weaken condition (3.11)

Theorem 3.2. The statement of Theorem 3.1 holds if F (x, t)=kx + p0,where p ∈ (−1, 1) and k is positive, satisfying 0 √ √ 1+4αk − 1 1+4αk + |p | exp −h > √ 0 . (3.15) 2α 1+4αk +1 Remark 3.3. (1) The left-hand side of (3.15) is less than that of (3.11), i.e., Theorem 3.2 strengthens Theorem 3.1 for a particular case. (2) Restriction (3.11) may, in principle, be weakened, but cannot be replaced by (3.15) even in the case F (x, t) ≡ F (x), F (0) = 0. Indeed,inanexamplewith F (x)=k2x, x ≥ 0,andF (x)=k1x, x<0,wherek0 >k1 >k2 < 0, substitution of k = k0 > 0 in (3.15) turns it into an equality, k1 is close to k0, k1 −k2 k0 −k1,a direct integration shows that there are no oscillating solutions with zero l-frequency. (3) In fact, under condition (3.13) or (3.14), we have |F (x(t),t)|≤p1 < 1 with some constant p1, i.e., such solutions x(t) possess the properties asserted in sections 1, 2. In particular, by Theorem 1.5 all solutions of (0.1) with an initial ϕ ∈ C −h, x t function 0[ 0],closeto ( ) [−h,0] satisfying (3.13) or (3.14), are bounded, oscillatory, and have zero l-frequency. (4) If α tends to zero, both conditions (3.10) and (3.15) converge to 2 kh < log , 1+|p0| known to be sufficient for the existence of bounded zero l-frequency oscillations in the first-order system (0.2) [7, 16, 6]. (5) It is easy to verify that (3.15) holds for all α>0,provided2hk ≤ 1 −|p0|. 4. Proofs.

4.1. Proof of Lemma 1.1. Assume that a continuous solution xϕ of (0.1), (1.4) is uniquely defined on [−h, T ] and is differentiable on the interval [0,T]. Show that xϕ can be extended to [T,T + h] as a continuous solution of (0.1), and that such an extension is unique and differentiable. The latter requirement is reduced to solving the initial value problem αx = −x + F (x, t)+ψ(t),x(0) = a0,x(0) = a1,t∈ [0,h], (4.16) where ψ(t) is a measurable bounded function. Integrating (4.16) twice, one succes- sively obtains t x t a e−t/α 1 F x τ ,τ ψ τ e(τ−t)/αdτ t ≥ , ( )= 1 + α ( ( ( ) )+ ( )) 0 (4.17) 0 t −t/α (τ−t)/α x(t)=a0 +αa1(1−e )+ (1−e )(F (x(τ),τ)+ψ(τ))dτ, t ≥ 0 . (4.18) 0 Solutions to (4.16) are, clearly, bounded on the segment [0,h] by some constant. Hence we can assume |Fx(x, t)| 0 in our situation. Denote by 346 E. SHUSTIN, E. FRIDMAN, L. FRIDMAN

C[0, 1/M ] the space of continuous functions on [0, 1/(2M)], equipped with the sup- norm ||x||∞ =sup|x(τ)|. Then the right hand side of (4.18) defines an operator Q : C[0, 1/(2M)] → C[0, 1/(2M)], which is contraction due to t (τ−t)/α ||Q(x1) − Q(x2)||∞ =sup (1 − e )(F (x1(τ)) − F (x2(τ))) dτ 0 1

4.2. Proof of Lemma 1.2. Assume on the contrary that Zϕ is bounded, and that T = max Zϕ. Without loss of generality suppose that xϕ(t) > 0ast>T. Then one derives from (0.1), (0.3) that, for t ≥ T + h, − p αx x − F x, t ≤− p ⇒ xet/α ≤−1 et/α + = 1+ ( ) 1+ = α −t/α =⇒ x(t) ≤ C0 + C1e − (1 − p)t, with some constants C0,C1, which implies xϕ(t) < 0ast>(|C0| + |C1|)/(1 − p), contradicting the above assumption and proving Lemma.

4.3. Proof of Theorem 1.3. (1) Assume that xϕ(T )=0,νϕ(T ) = 0, and xϕ(t) ≤ −1 0, t ∈ [T − h, T ], whereas xϕ (0) ∩ [T − h, T ] is finite. We shall show that xϕ(t) > 0 as t ∈ (T,T + h), which will imply thatνϕ(t)=0fort>T. Observe that the replacement of xϕ by any smooth function, which is [T −h,T ] T − h, T T xϕ negative in [ ) and zero at , does not influence on [T,∞) in view of (4.18). Thus, x = xϕ satisfies αx = −x + F (x, t)+1,t∈ [T,T + h],x(T )=0,x(T ) ≥ 0.

Hence t x t x T e(T −t)/α 1 F x τ ,τ e(τ−t)/αdτ ( )= ( ) + α (1 + ( ( ) )) T 1 − p t ≥ e(τ−t)/αdτ =(1− p)(1 − e(T −t)/α),t∈ [T,T + h]. (4.19) α T In particular, x(t) strictly increases in (T,T + h], and so is positive there. (2) Assume that νϕ(T )=n>0, n is even, xϕ(T )=0,xϕ(t) < 0asT −εT +h, where ∗∗ T = min{t>T : xϕ(t)=0}, (4.21) ∗∗ implying νϕ(T ) ≤ νϕ(T ), and we are done in this case. (3) Assume that νϕ(T )=n>0, n is odd, xϕ(T )=0,xϕ(t) < 0asT −ε

ϕ ∈ U 0 T>h 4.4. Proof of Theorem 1.5. Let 0 . Then there exists such that xϕ(T )=0,νϕ(T ) = 0 and, for instance, xϕ(t) ≤ 0ast ∈ [T − h, T ]. Then (4.19) tells us that xϕ(t) > 0ast ∈ (T,T + h +2ε) for some ε>0. If ψ ∈ C0[−h;0]is close to ϕ, then ψ−1(0) is contained in a sufficiently small neighborhood of ϕ−1(0), and mes({ϕ>0}◦{ψ>0}), mes({ϕ<0}◦{ψ<0}) are small enough, where A ◦ B denotes (A\B) ∪ (B\A). Hence Zψ ∩ [0; T +2h] is contained in a sufficiently small neighborhood of Zϕ ∩ [0; T +2h]. Therefore xψ(t) > 0ast ∈ (T + ε/2,T + h +3ε/2), which, as in the proof of Theorem 1.3, implies Nψ =0. 4.5. Proof of Theorem 2.1 in the zero l-frequency case. Since by Lemma 1.2 any solution xϕ of (2.6), satisfying (0.3), oscillates around zero, so does xϕ. Hence ∗ sup xϕ(t)=supxϕ(t ), t≥T t∗ where t∗ runs through the local maxima t∗ ≥ T ,andT is the first local extremum ∗ of xϕ. In particular, in a neighborhood of a local maximum t of xϕ we have ∗ ∗ ∗ ∗ xϕ(tk) ≥ 0, tk → t , tk 0 and consider the family of equations αx(t)=−x(t)+F (x(t)) − sign x(t − hκ),κ∈ (0, ∞) , (4.23) with the initial conditions

x(t)=ϕ(t),ϕ∈ C0[−κh],ϕ(0) = 0,ϕ(t) < 0ast<0 . (4.24) Let Ψ : (0, ∞) × [0, 2] → [0, 2] be a map such, for any κ ∈ (0, ∞), the restriction of Ψonthesegment{κ}×[0, 2] is the map Φ for system (4.23) as defined in section 2. Since it is continuous, the set

A = {(κ, a) ∈ (0, ∞) × [0, 2] : Φκ(a)=a} is closed in (0, ∞) × [0, 2]. Introduce one more map, Π : (0, ∞) × [0, 2] → R,which sends a point (κ, a) to the second positive zero of the solution to (4.23), (4.24). It is continuous as well, and sends any point (κ, a) ∈ A to the minimal positive period of the corresponding periodic solution. We intend to show that the image of the map Θ:A → R, Θ(κ, a)=hκ + nΠ(κ, a) is the whole interval (0, ∞). In particular, there exists (κ, a) ∈ A such that hκ + nΠ(κ, a)=h, and thus, for the corresponding Π(κ, a)-periodic solution to (4.23), we have αx(t)=−x(t)+F (x(t)) − sign x(t − hκ) = −x(t)+F (x(t)) − sign x(t − hκ − nΠ(κ, a)) 348 E. SHUSTIN, E. FRIDMAN, L. FRIDMAN

= −x(t)+F (x(t)) − sign x(t − h) , i.e., it is a periodic solution to (2.6) with ν ≡ n. Step 2. Show that lim Π(κ, a)=∞, lim Π(κ, a)=0 κ→∞ κ→0 (κ,a)∈A uniformly on a ∈ [0, 2]. The first relation immediately follows from Π(κ, a) >hκ. Assume that the second relation does not hold. Then there exists a sequence ∗ ∗ (κk,ak) ∈ A such that κk → 0, ak → a ≥ 0, Π(κk,ak) → T > 0ask →∞. Observe that T ∗ < ∞. Indeed, equation (4.23) on the segment [0,hκ] turns into αx = −x + F (x)+1,x(0) = 0,x(0) = a ∈ [0, 2] . Hence by formulas (4.17), (4.18), p hκ p hκ x hκ ≤ a (1 + ) ≤ (1 + ) , ( ) + α 3+ α hκ x(hκ) ≤ αa(1−e−hκ/α)+(1+p) (1−e(τ−hκ)/α)dτ ≤ (a+1+p)hκ ≤ (3+p)hκ . 0 On the segment defined by t ≥ hκ, x(t) ≥ 0, we have p hκ αx −x F x − ,xhκ ≤ p hκ, x ≤ (1 + ) = + ( ) 1 ( ) (3 + ) (0) 3+ α ; hence again by (4.18) t x(t) ≤ (3 + p)hκ +(3α +(1+p)hκ)(1 − e(t−hκ)/α) − (1 − p) (1 − e(τ−t)/α)dτ hκ =(3+p)hκ +(3α +(1+p)hκ)(1 − e(hκ−t)/α) − (1 − p)(t − hκ − α(1 − e(hκ−t)/α)) ≤ 6hκ +4α − (1 − p)t.

Hence x(t) has its first positive zero at some point T1 ≤ (6hκ +4α)/(1 − p), and consequently its second positive zero at some point T2 ≤ (12hκ +8α)/(1 − p). The limit of the zero l-frequency solutions (the convergence is evident in view of the continuous dependence on κ and a, including the zero values) to the corre- sponding equations is a T ∗-periodic function x(t) such that, for some T ∈ [0,T∗], x(0) = x(T )=x(T ∗)=0,x(0) = x(T ∗)=a∗, (4.25) αx(t)=−x(t)+F (x(t)) − 1,x(t) ≥ 0,t∈ [0,T], (4.26) αx(t)=−x(t)+F (x(t)) + 1,x(t) ≤ 0,t∈ [T,T∗] . Without loss of generality assume that T>0. Then a∗ = x(0) > 0, since otherwise (4.26) would imply x(t) < 0 for small positive t.Ifx (T0) = 0 for some T0 ∈ [0,T] then x (T0)=F (x(T0)) − 1 ≤ p − 1 < 0, i.e., T0 is a strong local maximum of x(t). Hence such a point T0 is unique, belongs to (0,T) and is the global maximum of x(t)on[0,T]. Moreover, x (t) > 0, t ∈ [0,T0), and x (t) < 0, t ∈ (T0,T]. We shall show that a∗ = x(0) > |x(T )| . (4.27)

Since x(t) is strictly monotone in [0,T0] and in [T0,T], we can replace (4.26) by the equation dy − F x − 1 − 1 ( ), ≤ x ≤ x T ,y a∗,yx T , dx = α αy 0 ( 0) (0) = ( ( 0)) = 0 OSCILLATIONS IN A SECOND-ORDER SYSTEM 349

corresponding to the interval [0,T0] with y = x , and the equation dz − F x 1 − 1 ( ), ≤ x ≤ x T ,z −x T ,zx T , dx = α αz 0 ( 0) (0) = ( ) ( ( 0)) = 0 corresponding to the interval [0,T0] with z = −x . This immediately implies y(x) > z(x) for any 0 ≤ xx(T ∗), which together with (4.27) gives a∗ >x(T ∗), contradicting (4.25). Step 3. Since A ∩ ({κ}×[0, 2]) = ∅ for any κ ∈ (0, ∞), lim Θ(κ, a)=0, lim Θ(κ, a)=∞ (4.28) κ→0 κ→∞ (κ,a)∈A (κ,a)∈A uniformly on a ∈ [0, 2]. Thus, to complete the proof, we have to establish the connectedness of Θ(A) ⊂ R. We argue on the contrary. Assume that Θ(A) ⊂ (0,θ) ∪ (θ, ∞), θ>0. Then the sets −1 −1 A− = A ∩ Θ (0,θ),A+ = A ∩ Θ (θ, ∞) are nonempty and closed in (0, ∞) × [0, 2]. Notice that A ∩{κ ∈ (0, ∞),a=2} = A ∩{κ ∈ (0, ∞),a=0} = ∅ . Indeed, Ψ(κ, 2) ≤ 1+p<2 and Ψ(κ, 0) > 0. Here the equality Ψ(κ, 0) = 0 is impossible, because otherwise the corresponding periodic zero l-frequency solution x with x(0) = x(0) = 0 would have a local minimum at 0 in view of x(0) = (1 + F (0))/α > 0, but x changes its sign at 0 from minus to plus by construction. The uniform convergence in (4.28) implies also that A+ does not intersect some ∗ ∗ strip (0,ε) × [0, 2]. Take a sufficiently large κ > 0 such that A− ⊂ (0,κ ) × [0, 2] and A∗ A ∩ ,κ∗ × , ∅ . + = + (0 ] [0 2] = A∗ { }× , ,κ∗ ×{ } Furthermore, + is compact and disjoint with the segments 0 [0 2], [0 ] 0 , ,κ∗ ×{ } A A∗ [0 ] 2 , and with the compact set −. In particular, we can cover + by the union U of finitely many open discs such that the closure U is disjoint with the above ∗ three segments as well as with A−. take the segment S = {κ = κ , 0 ≤ a ≤ 2}. By construction, it does not meet A−. If it crosses U, then we construct a path in (0,κ∗]×[0, 2], starting from the point (κ∗, 0), then along S to the intersection point with U, then along finitely many circle arcs of ∂U until we come to S again. After finitely many such moves along S and ∂U, we end up at (κ∗, 2). The continuous path constructed lies inside (0,κ∗] × [0, 2]\A, but this is impossible, since the difference Ψ(κ, a) − a changes its sign along this path from plus at (κ∗, 0) to minus at (κ∗, 2), so must be intersection point with A. 4.7. Proof of Theorem 2.4. The problem reduces to the study of the function Φ defined in section 2. We intend to show that Φ(a) ≥ 0,a∈ [0, 2] . (4.29) If F (x) is analytic then Φ(a) is analytic as well. Hence there are only finitely many roots of the equation Φ(a)=a in the segment [0, 2], because Φ(a) ≡ a in view of Φ(2) ≤ 1+p<2. The latter inequality together with Φ(0) > 0 (see section 4.6) says that there is a0 ∈ (0, 2) such that Φ(a0)=a0 and Φ(a) − a changes its sign from plus to minus at a0. Assuming (4.29) to be true, this means that a0 is an attractor of the dynamical system a → Φ(a)on[0, 2]; hence the corresponding zero l-frequency periodic solution is quasi-stable. 350 E. SHUSTIN, E. FRIDMAN, L. FRIDMAN

Inequality (4.29) implies statement (2) of Theorem 2.4 as well. Indeed, the simplicity of the (unique) root a0 of the function Φ(a) − a means that Φ (a0) < 1. Together with (4.29) this implies that any sequence a1 ∈ [0, 2], ai+1 =Φ(ai), i ≥ 1, satisfies εi+1 ≤ qεi,q=const∈ (0, 1),εi = ai − a0,i≥ 1; hence ∞ εi < ∞ , i=1 which in view of the boundedness of |dTa/da| in a neighborhood of a0, gives the existence of the limit (n) lim (Ta − nTa ) , n→∞ 1 0 T (n) n x where a1 is the 2 -th positive zero of a1 . Now we prove (4.29). Let a>0. Denote by Ta and Ta the first and the second positive zero of xa and introduce the functions Ψ+, Ψ− :[0, 2] → [0, 2] by Ψ+(a)=−xa(Ta), Ψ−(−xa(Ta)) = xa(Ta). To prove (4.29) it is sufficient to show s ≥ s ≥ s ∈ , that Ψ+( ) 0, Ψ−( ) 0, [0 2], or, equivalently, that Ψ+ and Ψ− are increasing. Let us consider Ψ+.Putb = a + δ, where δ>0 is a small parameter, and show that Ψ+(b) > Ψ+(a). As seen in the proof of Theorem 1.3, section 4.3, any solution x(t) of (2.6) with x(0) = 0, x(0) > 0, is strictly increasing in [0,h]. Hence (2.6) reduces to the equation dy F x − 1 1+ ( ),y x , ≤ x ≤ x h , dx = α + αy (0) = (0) 0 ( ) (4.30) where y(x):=x (t). If ya(x), yb(x) are solutions to (4.30) with ya(0) = aya(x) on the whole interval 0 ≤ x ≤ xa(h). In other words, xb(τ(t)) >xa(t) for any t ∈ [0,h]andτ(t) ∈ [0,h] such that xb(τ(t)) = xa(t). It follows then that (see Figure 2)

xb(t) >xa(t)andτ(t) hbe the minimal positive such that xa(t1)=xb(h) (see Figure 2). Relations (4.32) imply that 2 t1 − h = h − τ(h)+O(δ ) . Then x h x τ h 1 −x τ h F x h h − τ h O δ2 b( )= b( ( )) + α( b( ( )) + ( a( )) + 1)( ( )) + ( ) M x h M δ O δ2 ,M M 1 −x h F x h , = a( )+ 3 + ( ) 3 = 2 + α ( a( )+ ( a( )) + 1) x t x h 1 −x h F x h − t − h O δ2 a( 1)= a( )+α( a( )+ ( a( )) 1)( 1 )+ ( ) x h 1 −x h F x h − h − τ h O δ2 = a( )+α( a( )+ ( a( )) 1)( ( )) + ( ) M x h M δ O δ2 ,M 1 −x h F x h − . = a( )+ 4 + ( ) 4 = α ( a( )+ ( a( )) 1) OSCILLATIONS IN A SECOND-ORDER SYSTEM 351

x(a T)0

xb (h)

xa (h)

h t 1 T0 T1 T’abT’

Figure 2. Comparison of xa and xb

Clearly, M3 >M4; hence, for a sufficiently small δ>0, xb(h) >xa(t1) . (4.33)

As seen in the proof of Theorem 2.1, section 4.6, the solution xa(t) is positive in [h, Ta) and has in [h, Ta) a unique extremum, a global maximum, at some point T0 >h.So,xa satisfies the equation dy − F x − 1 1+ ( ), dx = α + αy (4.34) y(xb(h)) = xa(t1),xb(h) ≤ x ≤ xa(T0), where x = xa, y = xa. The function xb satisfies equation (4.34) with the initial data y(xb(h)) = xb(h) (4.35) in a neighborhood of the point x = xb(h). Since xa(t) > 0ast ∈ [t1,T0), and xb(τ+(t)) >xa(t)fort in a neighborhood of t1 and τ+(t) such that xb(τ+(t)) = xa(t), according to (4.33), the initial value problem (4.34), (4.35) has a solution on the whole segment [xb(h),xa(T0)], and, moreover, xb(τ+(T0)) >xa(T0)=0. This implies that def xb(T ) = max xb(t) >xa(T ) = max xa(t), 1 0 t∈[0,Tb] t∈[0,Ta] and xb(τ−(T0)) −xa(Ta)=Ψ+(a), since xa and xb induce solutions to (4.34) in 0 ≤ x ≤ xa(T0) with the initial data y(xa(T0)) = 0 and y(xa(T0)) = xb(τ−(T0)), respectively. 352 E. SHUSTIN, E. FRIDMAN, L. FRIDMAN

In the same way one can show that Ψ− is increasing.

4.8. Proof of Theorem 2.3(1). We assume that F (x) ≡ q =const,0≤ q<1. We have to show that the equation Φ(a)=a has a unique solution in the segment [0, 1+q]. Step1. The case q = 0 has been treated in [17], section 3.6. Assume that 0 < q<1. Let x(t) be a zero l-frequency solution of (2.6) with x(0) = 0, x(0) = a ≥ 0. It satisfies αx = −x + q +1, 0 ≤ t ≤ h, u + h ≤ t ≤ u + v, αx = −x = q − 1,h≤ t ≤ u + h, where u and u + v are the first and the second positive zeroes of x(t). One can easily get that x(t)=(1+q − a)α(e−t/α − 1) + (1 + q)t, 0 ≤ t ≤ h, x(t)=α(1 + q − a − 2eh/α)e−t/α +2h + α(1 + a − q) − (1 − q)t, x(t)=(2eh/α − 1 − q + a)e−t/α − 1+q, h ≤ t ≤ u. Denoting b = −x(u), c = x(u + v), we obtain the equations α(1 + q − a − 2eh/α)e−u/α +2h + α(1 + a − q) − (1 − q)u =0, (4.36)

def h/α −u/α b =Ψ+(a)=1− q − (2e − 1 − q + a)e , (4.37) α(1 − q − b − 2eh/α)e−v/α +2h + α(1 + b + q) − (1 + q)u =0, (4.38)

def h/α −v/α c =Ψ−(b)=1+q − (2e − 1+q + b)e . (4.39) For any α, q, we count solutions to (4.36)-(4.39) with c = a. Assuming the latter equality and introducing parameters a =1+q − a, b =1− q − b, w =2− a − b, we rewrite (4.36)-(4.39) as b =(2eh/α − a)e−u/α,a =(2eh/α − b)e−v/α, (4.40) 2h + αw =(1− q)u =(1+q)v. (4.41) Note that a ≤ 1+q, b ≤ 1 − q by (4.37), (4.39). Hence 0 ≤ w ≤ 2. Also, knowing the value of w, one can uniquely restore u, v, a,b, and hence a, b, from (4.40), (4.41). Solving (4.40), (4.41) with respect to a,b,u,v and substituting this into w =2− a − b, we obtain def 4h 2w G(w, α, q) =(2− w) exp α −q2 + −q2 − 1 (1 ) 1 − eh/α 2h w 2h w − . 2 exp α(1−q) + 1−q +exp α(1+q) + 1+q 2 =0 (4.42) Hence our problem reduces to showing that, given α ∈ (0, ∞), q ∈ (0, 1), there exists at most one (counting multiplicities) solution w ∈ [0, 2] to (4.42). In [17] it was shown for q = 0. If, for some α, q there exist two solutions (counting multiplicities) to (4.42), then there exists a point (w, α, q) ∈ Π=[0, 2] × (0, ∞) × [0, 1) such that G(w, α, q)=Gw(w, α, q) = 0. We shall show that the set {G = Gw =0}∩Πis empty. Step 2. We shall now prove that there exists a function α∞ :[0, 1) → (0, ∞) such that |Gw(w, α, q)| + |G(w, α, q)| > 0asα>α∞(q). OSCILLATIONS IN A SECOND-ORDER SYSTEM 353

Let us introduce the function H(w, β, q)=G(w, h/α, q). Then 2w w w H(w, 0,q)=(2−w) exp − 1 −2 exp +exp − 2 . 1 − q2 1 − q 1+q For any q ∈ [0, 1), the equation H(w, 0,q) = 0 has the solution w = 0. We claim that this is the only solution. Indeed, 2(2 − w) 2w Hw(w, 0,q)=1+ − 1 exp 1 − q2 1 − q2 1 w 1 w −2 exp + exp , 1 − q 1 − q 1+q 1+q which is zero at w = 0 and negative for w>0. The latter statement follows from ∂ 2w Hw(w, 0,q)exp − ∂w 1 − q2 2 w w = − 1 − exp − 1 − exp − , 1 − q2 1 − q 1+q which is zero at w = 0 and negative for w>0. One also easily verifies that H(0, 0,q)=0,q ∈ [0, 1). To complete the proof of the required statement, we have to show that no se- quence (wk,βk,qk) with βk > 0, H(wk,βk,qk)=Hw(wk,βk,qk) = 0 can approach a point (0, 0,q), q ∈ [0, 1). Indeed, in a small neighborhood U of any point (0, 0,q), one has 2 1 H(w, β, q)= 2β3 +3β2w + βw2 − w4 + h.o.t., 1 − q2 12(1 − q2)2 2 2 1 3 Hw(w, β, q)= 3β +2βw − w + h.o.t.. 1 − q2 3(1 − q2)2 Thus, the system H = Hw = 0 reduces in U ∩ Πto 1 2 k 1 2 k β = w + Ak(q)w ,β= w + Bk(q)w , 12(1 − q2)2 6(1 − q2)2 k≥3 k≥3 which has no solutions in U ∩ Π. Step 3. We next claim that there exists a function α0 :[0, 1) → (0, ∞) such that |Gw(w, α, q)| + |G(w, α, q)| > 0as0<α<α0(q). Indeed, if not, one would have a sequence (wk,αk,qk) approaching a point (w, 0,q), q ∈ [0, 1), such that G(wk,αk,qk)=Gw(wk,αk,qk) = 0. However, these equations can be written as 4h 2w 2 − w =(2− w)exp − − α(1 − q2) 1 − q2 h(1 − q) w h(1 + q) w +2 exp − − +exp − − α(1 + q) 1+q α(1 − q) 1 − q h(3 + q2) 2w −2exp − − , α(1 − q2) 1 − q2 3+q2 − 2w 4h 2w = − exp − − − q2 α − q2 − q2 1 (1 ) 1 1 h(1 − q) w 1 h(1 + q) w +2 exp − − + exp − − , 1 − q α(1 + q) 1+q 1+q α(1 − q) 1 − q 354 E. SHUSTIN, E. FRIDMAN, L. FRIDMAN which imply the inconsistent relations 3+q2 lim w = 2 and lim w = , α→0 α→0 2 respectively. Step 4. Assume that {G = Gw =0}∩Π = ∅. Since this set is closed in Π, and in view of the results of Steps 1-4, there exists

q0 = min{q :(w, α, q) ∈ Π,G(w, α, q)=Gw(w, α, q)=0} > 0 . At the corresponding point (w, α, q )wehave 0 Gw Gα G = Gw = =0, Gww Gwα which is equivalent to G = Gw = GαGww =0. Assuming that G = Gw = Gα = 0, we obtain α2 2(2 − w) 4h 2w − Gα = exp + 2h 1 − q2 α(1 − q2) 1 − q2 − q h w q h w −eh/α 3 2 3+ 2 − − q exp α − q + − q + q exp α q + q 2 1 (1 ) 1 1+ (1 + ) 1+ 1 w 4h 2w = Gw + G + exp + − 1 =0, 2 2 α(1 − q2) 1 − q2 which together with G = Gw =0andw = 0 gives a contradiction. The remaining case to consider is G = Gw = Gww = 0. In the notation 4h 2w E = eh/α − 1,X=exp + , α(1 − q2) 1 − q2 2h w 2h w Y =exp + ,Z=exp + , α(1 − q) 1 − q α(1 + q) 1+q the equations G = Gw = Gww = 0 read as (2 − w)X − 2(1 + E)Y − 2(1 + E)Z = −4E − w − 2 , (3 + q2 − 2w)X − 2(1 + E)(1 + q)Y + 2(1 + E)(1 − q)Z = −1+q2 , 2(1 + q2 − w)X − (1 + E)(1 + q)2Y +(1+E)(1 − q)2Z =0, which we solve with respect to X, Y, Z: 4E +2w X = , w(1 + E) 4Eq +2wq − w2 − 4Ew +2E +2Eq2 Y = , 2wq(1 + E) 4Eq +2wq + w2 +4Ew − 2E − 2Eq2 Z = . 2wq(1 + E) We plug these formulae in the identity X − YZ= 0 and obtain w4 +4q2w2E − 4w2E +4E2 − 8q2E2 +4q4E2 +16q2wE2 +16q2wE3 +4q2w2E2 +16w2E2 − 16wE2 +8w3E =0, which transforms into an impossible equality (w2 +2E(2w − 1+q2))2 +4q2w2E2 +16q2wE3 =0, OSCILLATIONS IN A SECOND-ORDER SYSTEM 355 and we are done.

4.9. Proof of Theorem 3.2. We consider the equation αy (t)=−y (t)+ky(t)+p0 − sign y(t − h) , (4.43) corresponding to F (x, t) ≡ F (x)=kx + p0. Introduce the roots λ1 > 0 >λ2 of the characteristic equation αλ2 = −λ + k. Assuming y(t) < 0,t∈ [−h, 0),y(0) = 0,y(0) = a ≥ 0 , we obtain a solution to (4.43) λ (1 + p ) − ak λ (1 + p ) − ak 1+p y(t)= 2 0 eλ1t + 1 0 eλ2t − 0 k(λ2 − λ1) k(λ1 − λ2) k as t ∈ [0,h], and, respectively, λ (1 + p − 2e−λ1h) − ak y(t)= 2 0 eλ1t k(λ2 − λ1) λ (1 + p − 2e−λ2h) − ak 1 − p + 1 0 eλ2t + 0 k(λ1 − λ2) k as t ≥ h. Since λ1 > 0 >λ2, the solution y(t) necessarily has a positive zero λ1t t = T1 >hif the coefficient of e is negative, i.e., eλ1h − − p a<2 1 0 , αλ1 which is equivalent to (3.13). Moreover, max{y(t),t∈ [0,T1]} < (1 − p0)/k, since, otherwise, attaining the value (1 − p0)/k, the solution y(t) would further be greater than (1 − p0)/k.So, F (y(t)) ≤ p1(a) < 1ast ∈ [0,T1]. Denote b = −y (T1). It is, evidently, a nonnegative number. From the explicit formula for y(t) it is not difficult to derive that

λ2T1 −λ2h e (2e − 1 − p0 − αλ2a)=1− p0 + αλ2b. Since −λ2h 2e − 1 − p0 − αλ2a ≥ 2 − 1 − p0 =1− p0 > 0 , we obtain 1 − p0 b<− = η(h, α, k, −p0) <ξ(h, α, k, −p0) , (4.44) αλ2 where the latter inequality can be rewritten as √ √ 1+4αk − 1 1+4αk + p exp −h > √ 0 , 2α 1+4αk +1 which follows from (3.15). In view of inequality (4.44), equivalent, in fact, to (3.14), the above argument (after switching the sign of y(t)) implies that y(t) has the second positive zero T2 >T1 + h and is negative for t ∈ (T1,T2). Furthermore, min{y(t),t∈ [T1,T2]} > −(1 + p0)/k; hence F (y(t)) ≥−p2(a) > −1ast ∈ [T1,T2]. Denoting c = y (T2), we derive as above

λ2(T2−T1) −λ2h e (2e − 1+p0 − αλ2b)=1+p0 + αλ2c. 356 E. SHUSTIN, E. FRIDMAN, L. FRIDMAN

Hence 1+p0 c<− <ξ(h, α, k, p0) , αλ2 the latter inequality being √ √ 1+4αk − 1 1+4αk − p exp −h > √ 0 , 2α 1+4αk +1 and hence coming from (3.11). Extending y(t) further we obtain in the same manner a bounded, oscillatory, zero l-frequency solution. In addition, we have shown that |F (y(t)|≤max{p1(a),p2(a)} < 1 confirming Remark 3.3(3). 4.10. Proof of Theorem 3.1. In the notation of the previous section, inequality (3.11) can be rewritten as

−λ1h −λ1h 2e − 1 − p0 2e − 1+p0 ≥ 1+p0, ≥ 1 − p0 . αλ1 αλ1 Starting with the initial data (3.13), we obtain that the corresponding solution x(t) of (0.1) satisfies αx (t)=−x (t)+F (x(t),t) − sign x(t − h) ≤−x (t)+kx(t)+p0 − sign x(t − h) , as far as x(t) ≥ 0. Hence x(t) ≤ y(t) as far as x(t) ≥ 0, where y(t) is the solution to equation (4.43) with the same initial data (3.13). This immediately implies that x(t) has a positive zero Ta >hin the interval [0,T1], and max |F (x(t),t)|≤ max (ky(t)+p ) ≤ p (a) < 1 . 0 1 t∈[0,Ta] t∈[0,T1] Now we estimate x (Ta). Note that x(t) has a unique maximum point T0 ∈ h, T T 0ast

Since G(x) ≥ p0, x ∈ [0,x(T0)], we have x (Ta)=z(0) ≥ w(0) , where dw − p − 1 − 1 0 , ≤ x ≤ x T ,wx T . dx = α αw 0 ( 0) ( ( 0)) = 0 A direct computation gives −τ /α x (Ta) ≥ w(0) = −(1 − p0)(1 − e ) , OSCILLATIONS IN A SECOND-ORDER SYSTEM 357 where τ is the positive root of the equation −τ /α α(1 − e )=τ − x(T0) . Since x(T0) < (1 − p0)/k, we obtain that τ is less that τ defined in Theorem 3.1, and hence

−λ1h −τ/α 2e − 1+p0 x (Ta) > −(1 − p0)(1 − e ) ≥− ≥−ξ(h, α, k, |p0|) , (4.45) αλ1 the middle inequality being (3.11) with |p0| replaced by −p0. In fact, we get in the situation of initial value problem (3.14), and hence by the same reason, we can extend x(t) to infinity as bounded, oscillatory, zero l-frequency solution.

REFERENCES

[1] Akian, M., and Bliman, P.-A. (2000): On super-high-frequencies in discontinuous 1st-order delay-differential equations. J. Differential Equations 162, no. 2, 326–358 [2] Andre, J., and Seibert, P. (1956): Uber¨ die st¨uckweise linearen Differentialgleichungen, die bei Regelungsproblemen auftreten. Arch. Math. 7, no. 3, 148–164 [3] Bayer, W., and an der Heiden, U. (1998): Oscillation types and bifurcations of a nonlinear second-order differential-difference equation. J. Dynam. Diff. Equations 10, no. 2, 303–326 [4] Choi, S.-B., and Hedrick, J. K. (1996): Robust Throttle Control of Automotive Engines: Theory and Experiment. ASME J. of Dynamic Systems, Measurement, and Control 118, no. 1, 92–98 [5] Drakunov, S. V., and Utkin, V. I. (1992): Sliding mode control in dynamic systems. Int. J. Control 55, no. 4, 1029–1037 [6] Fridman, E., Fridman, L., and Shustin, E. (2000): Steady modes and stability in discontinuous delay systems with periodic disturbances. ASME J. of Dynamic Systems, Measurement, and Control 122, no. 4, 732–737 [7] Fridman, L., Fridman, E., and Shustin, E. (1993): Steady modes in an autonomous system with break and delay. Differential Equations 29, no. 8, 1161–1166 [8] an der Heiden, U., Longtin, A., Mackey, M. C., Milton, J. G., and Scholl, R. (1990): Oscil- latory modes in a nonlinear second-order differential equation with delay. J. Dynam. Diff. Equations 2, no. 4, 423–449 [9] an der Heiden, U., and Reichard, K. (1990): Multitude of oscillatory behavior in a nonlinear second order differential-difference equation. Z. Angew. Math. Mech. 70, no. 6, PT621–T624 [10] Longtin, A., and Milton, J. G. (1989): Modelling autonomous oscillations in the human pupil light reflex using nonlinear delay-differential equation. Bull. Math. Biol. 51, no. 5, 605–624 [11] Longtin, A., and Milton, J. G. (1989): Insight into the transfer function, gain, and oscillation onset for the pupil light reflex using nonlinear delay-differential equations. Biol. Cybern. 61, no. 1, 51–58 [12] Mallet-Paret, J. (1988): Morse decomposition for delay-differential equations. J. Differential Equations 72, 270–315 [13] Mallet-Paret, J., and Nussbaum, R. D. (1992): Boundary layer phenomena for differential- delay equations with state-dependent time lags, I. Arch. Rational Mechanics and Analysis 120, 99–146 [14] Mallet-Paret, J., and Walther, H.-O. (1994): Rapid oscillations are rare in scalar systems governed by monotone negative feedback with a time lag,Preprint [15] Moskwa, J. J., and Hedrick, K. (1989): Sliding mode control of automotive engines. In: Proc. of the American Contr. Conf., vol. 2, pp. 1040–1045 [16] Shustin, E., Fridman, E., and Fridman, L. (1993): Steady modes in a discontinuous control system with time delay. Pure Math. Appl. 4, no. 1, 59–73 [17] Shustin, E., Fridman, E., and Fridman, L. (1998): Stable oscillations in a discontinuous delay system of the second order. In: Dynamical Systems and Differential equations (W. Chen and S. Hu, eds.), Proc. International Conf. on Dynamical Systems and Differential equations, Springfield, Missouri 1996, vol. 2/ An added volume to Discrete and Continuous Dynamical Systems, pp. 209-233 [18] Utkin, V. I. (1991): Sliding modes in control optimization, Springer 358 E. SHUSTIN, E. FRIDMAN, L. FRIDMAN

[19] Walther, H.-O. (1981): Density of slowly oscillating solutions ofx ˙ (t)=−f(x(t−1)). J. Math. Anal. Appl. 79, 127–140 [20] Walther, H.-O. (2001): Contracting return maps for monotone delayed feedback. Discrete and Continuous Dynamical Systems 7, no. 2, 259–274 Received March 2001; revised August 2001; second revision February 2002. E-mail address: [email protected]