Loading [MathJax]/jax/output/SVG/jax.js
Research article Special Issues

Averaging methods for piecewise-smooth ordinary differential equations

  • The averaging method is developed for periodic piecewise-smooth systems. We discuss the behavior of solutions intersecting the discontinuity boundary and the problems it introduces. We illustrate these difficulties on specific examples. In the case of transversal and sliding solutions, we introduce conditions that allow us to prove averaging theorems for piecewise-smooth periodic differential equations.

    Citation: Michal Fečkan, Július Pačuta, Michal Pospíśil, Pavol Vidlička. Averaging methods for piecewise-smooth ordinary differential equations[J]. AIMS Mathematics, 2019, 4(5): 1466-1487. doi: 10.3934/math.2019.5.1466

    Related Papers:

    [1] Abdon Atangana, Seda İğret Araz . Piecewise differential equations: theory, methods and applications. AIMS Mathematics, 2023, 8(7): 15352-15382. doi: 10.3934/math.2023785
    [2] Woocheol Choi, Young-Pil Choi . A sharp error analysis for the DG method of optimal control problems. AIMS Mathematics, 2022, 7(5): 9117-9155. doi: 10.3934/math.2022506
    [3] Xumei Zhang, Junying Cao . A high order numerical method for solving Caputo nonlinear fractional ordinary differential equations. AIMS Mathematics, 2021, 6(12): 13187-13209. doi: 10.3934/math.2021762
    [4] Xin Liu, Yan Wang . Averaging principle on infinite intervals for stochastic ordinary differential equations with Lévy noise. AIMS Mathematics, 2021, 6(5): 5316-5350. doi: 10.3934/math.2021314
    [5] Shifan Luo, Dongshu Wang, Wenxiu Li . Dynamic analysis of a SIV Filippov system with media coverage and protective measures. AIMS Mathematics, 2022, 7(7): 13469-13492. doi: 10.3934/math.2022745
    [6] Essam R. El-Zahar, Ghaliah F. Al-Boqami, Haifa S. Al-Juaydi . Piecewise approximate analytical solutions of high-order reaction-diffusion singular perturbation problems with boundary and interior layers. AIMS Mathematics, 2024, 9(6): 15671-15698. doi: 10.3934/math.2024756
    [7] Saima Rashid, Fahd Jarad, Sobhy A. A. El-Marouf, Sayed K. Elagan . Global dynamics of deterministic-stochastic dengue infection model including multi specific receptors via crossover effects. AIMS Mathematics, 2023, 8(3): 6466-6503. doi: 10.3934/math.2023327
    [8] Wedad Albalawi, Muhammad Imran Liaqat, Kottakkaran Sooppy Nisar, Abdel-Haleem Abdel-Aty . Qualitative study of Caputo Erdélyi-Kober stochastic fractional delay differential equations. AIMS Mathematics, 2025, 10(4): 8277-8305. doi: 10.3934/math.2025381
    [9] Yanshou Dong, Junfang Zhao, Xu Miao, Ming Kang . Piecewise pseudo almost periodic solutions of interval general BAM neural networks with mixed time-varying delays and impulsive perturbations. AIMS Mathematics, 2023, 8(9): 21828-21855. doi: 10.3934/math.20231113
    [10] Dumitru Baleanu, Babak Shiri . Nonlinear higher order fractional terminal value problems. AIMS Mathematics, 2022, 7(5): 7489-7506. doi: 10.3934/math.2022420
  • The averaging method is developed for periodic piecewise-smooth systems. We discuss the behavior of solutions intersecting the discontinuity boundary and the problems it introduces. We illustrate these difficulties on specific examples. In the case of transversal and sliding solutions, we introduce conditions that allow us to prove averaging theorems for piecewise-smooth periodic differential equations.


    Averaging method has many applications and it is well-developed for ordinary and partial differential equations [4,8], for impulsive and differential inclusions [7], and for more general types of differential equations as well [6,5]. Non-smooth dynamical systems also possess many applications [2]. The goal of this paper is to extend the tool set available for analyzing periodic piecewise-smooth differential equations by generalizing the averaging method. Several intrinsic problems surfaced while trying to generalize the results to the piecewise-smooth setting. We focus on the cases of transversal and sliding solutions of the piecewise-smooth systems. In these cases, we discuss the problems and try to shed some light on their nature and causes by considering concrete examples. Several conditions are introduced which allow us to prove results analogous to the averaging theorem for smooth systems.

    Section 2 is a brief introduction to the theory of averaging for smooth systems and contains an averaging theorem whose statement serves as a prototype for statements of our generalizations for piecewise-smooth systems. Section 3 introduces piecewise-smooth systems and our main definitions. Section 4 provides examples that motivated us for these definitions. Section 5 presents some auxiliary results concerning the discontinuity boundary and its intersection with a solution that are used in the next sections. Sections 6 and 7 deal with the proofs of our main results. The final Section 8 summarizes our results with a possible outline for further study.

    Throughout the paper, we shall denote uv the inner product of vectors u,vRn and |u| the Euclidean norm of uRn. Furthermore, the distance of sets A,BRn is defined as dist(A,B)=inf{|ab|aA,bB}.

    Let ΩRn be a domain and aΩ. We consider the initial value problem

    ˙x=ϵf(x,t,ϵ),x(0)=a (2.1)

    defined for xΩ, ϵ[0,1] and tR where f(x,t,ϵ) is a continuously differentiable periodic function in the time variable t with the period T.

    Definition 2.1. Let f be the right-hand side of initial value problem (2.1). Denote

    ˉf(z)=1TT0f(z,s,0)ds.

    The averaged system associated with (2.1) is the system

    ˙z=ϵˉf(z), (2.2)

    and the guiding system associated with (2.1) is the system

    ˙w=ˉf(w). (2.3)

    Initial value problem (2.1) gives naturally a rise to the initial value problem for the averaged system (2.2) with z(0)=a, and for the guiding system (2.3) with w(0)=a. The guiding system is only a rescaled averaged system.

    The following theorem from [8,Theorem 2.8.1] is the key result we need and it will be used throughout the whole paper.

    Theorem 2.2. Consider initial value problem (2.1). Suppose that f is Lipschitz continuous and w:[0,L]Ω is solution of the associated guiding system. Then there exist positive constants ϵ0, C such that for all ϵ(0,ϵ0] there exists a unique solution xϵ:[0,Lϵ]Ω such that

    |xϵ(t)w(ϵt)|Cϵ

    for 0tLϵ.

    To simplify some of the statements in the rest of the paper, we collect some of the common assumptions on the right-hand side f in the following definition.

    Definition 2.3. We say, that a function f:Ω×R×[0,1]Rn is T-PCDLB if it is T-periodic in the second argument, continuously differentiable, Lipschitz continuous and bounded.

    The following lemma is an extension of the averaging theorem for smooth systems which will be useful in the next sections.

    Lemma 2.4. Let f be a T-PCDLB function, w:[L1,L2]Ω be a solution of the guiding system ˙w=ˉf(w) with w(L1)=a and A0 be a constant. Let ϵ(0,1]. Consider the local solution of initial value problem ˙x=ϵf(x,t,ϵ), x(L1ϵ+A)=bϵ for bϵΩ such that |abϵ|Dϵ for some constant D>0.

    Then there exist ϵ0,C>0 and a unique solution xϵ:[L1ϵ+A,L2ϵ]Ω of ˙x=ϵf(x,t,ϵ), x(L1ϵ+A)=bϵ for all ϵ(0,ϵ0] such that for all t[L1ϵ+A,L2ϵ] it holds

    |xϵ(t)w(ϵt)|Cϵ. (2.4)

    Proof. Let us at first assume the existence of xϵ for sufficiently small ϵ>0 and prove estimate (2.4). Let Aϵ be such that AϵA, |AϵA|T, L1ϵ+Aϵ=kϵT for some kϵZ. The existence of such Aϵ for all ϵ is obvious. Let ϵ0>0 be so small that L1ϵ+AϵL2ϵ for all ϵ(0,ϵ0]. Let M be a bound of |f|.

    First, we prove inequality (2.4) for t[L1ϵ+A,L1ϵ+Aϵ]. Using the assumption |abϵ|Dϵ and the Lipschitz continuity of solutions xϵ and w, we obtain

    |xϵ(t)w(ϵt)||xϵ(t)xϵ(L1ϵ+A)|+|xϵ(L1ϵ+A)w(ϵL1ϵ)|+|w(ϵL1ϵ)w(ϵt)|ϵM|AϵA|+Dϵ+ϵMAϵϵ(D+MT+M(A+T)).

    Next, we prove inequality (2.4) for t[L1ϵ+Aϵ,L2ϵ]. Since the solution w is defined on interval [L1,L2], we will use the averaging theorem for smooth systems for function ˜w(τ)=w(τ+L1). Hence there exists some C>0 such that for all ϵ>0 sufficiently small, there is a solution yϵ:[0,L2L1ϵ]Ω of the initial value problem ˙y=ϵf(y,t,ϵ), y(0)=a, and for all t[0,L2L1ϵ] it holds

    |yϵ(t)˜w(ϵt)|Cϵ. (2.5)

    Using yϵ we can define a function uϵ:[L1ϵ+Aϵ,L2ϵ+Aϵ]Ω by uϵ(t)=yϵ(t(L1ϵ+Aϵ)). The function uϵ satisfies the equation

    ˙uϵ(t)=˙yϵ(t(L1ϵ+Aϵ))=ϵf(uϵ(t),t(L1ϵ+Aϵ),ϵ)=ϵf(uϵ(t),t,ϵ),

    since L1ϵ+Aϵ=kϵT and f is T-periodic in the second argument.

    Now, let t[L1ϵ+Aϵ,L2ϵ]. Using (2.5) we estimate

    |uϵ(t)w(ϵt)|=|yϵ(tAϵL1ϵ)w(ϵt)||yϵ(tAϵL1ϵ)w(ϵ(tAϵ))|+|w(ϵ(tAϵ))w(ϵt)|ϵ(C+M(A+T)). (2.6)

    The functions uϵ, xϵ satisfy the same equation (2.1) (however, with possibly different initial values), hence it holds

    |uϵ(t)xϵ(t)||uϵ(L1ϵ+Aϵ)xϵ(L1ϵ+Aϵ)|+ϵtL1ϵ+Aϵ|f(uϵ(s),s,ϵ)f(xϵ(s),s,ϵ)|ds.

    Using the Lipschitz continuity of f and the Gronwall lemma, we obtain

    |uϵ(t)xϵ(t)|C|uϵ(L1ϵ+Aϵ)xϵ(L1ϵ+Aϵ)|

    for some constant C>0. The estimates (2.4), (2.6) for t=L1ϵ+Aϵ imply

    |uϵ(L1ϵ+Aϵ)xϵ(L1ϵ+Aϵ)|Cϵ.

    Thus the proof of the estimate (2.4) for t[L1ϵ+Aϵ,L2ϵ] is finished.

    Now we prove the existence and uniqueness. The uniqueness and the local existence of xϵ are consequences of the Lipschitz continuity. The only issue for global existence of xϵ would be if xϵ approaches the boundary Ω.

    Denote d=dist(Ω,w([L1,L2])). Let ϵ0>0 be such small that Cϵ0d2. Then we have

    |xϵ(t)w(ϵt)|Cϵd2

    for any ϵ(0,ϵ0] and t[L1ϵ+A,L2ϵ]. Consequently,

    ddist(Ω,w([L1+Aϵ,L2]))dist(Ω,xϵ([L1ϵ+A,L2ϵ]))+dist(xϵ([L1ϵ+A,L2ϵ]),w([L1+Aϵ,L2]))dist(Ω,xϵ([L1ϵ+A,L2ϵ]))+d2,

    which means that dist(Ω,xϵ(t))d2 for all t[L1ϵ+A,L2ϵ]. Hence, xϵ exists globally on interval [L1ϵ+A,L2ϵ].

    We start with a definition of a discontinuity boundary.

    Definition 3.1. We say that a function G:ΩR is a boundary function if G is continuously differentiable and G(x)0 for every xΩ such that G(x)=0.

    For a boundary function G, let us define

    Ω0={xΩG(x)=0},Ω+={xΩG(x)>0},Ω={xΩG(x)<0}.

    The set Ω0 is a C1-manifold and is called a discontinuity boundary or a discontinuity surface. The notation introduced above will be used throughout our paper.

    Definition 3.2. Let ΩRn be an open set, let f,f+:Ω×R×[0,1]Rn and G:ΩR be functions. We say that the tuple (Ω,G,f,f+) is a piecewise-smooth system if f and f+ are continuously differentiable and G is a boundary function. We say that the tuple (Ω,G,f,f+) is a well-behaved piecewise-smooth system if f and f+ are T-PCDLB and G is a boundary function.

    A shorthand ˙x=f±(x,t,ϵ) will be often used instead of the tuple (Ω,G,f,f+) while talking about a piecewise-smooth system, implicitly including a set Ω and a boundary function G.

    Definition 3.3. An absolutely continuous function x:[a,b]Ω is called a solution of the piecewise-smooth system (Ω,G,f,f+) if for almost all t[a,b],

    ˙x(t){{f(x(t),t,ϵ)}, if x(t)Ω,{f+(x(t),t,ϵ)}, if x(t)Ω+,conv({f(x(t),t,ϵ),f+(x(t),t,ϵ)}), if x(t)Ω0.

    Furthermore, we say that a solution of the piecewise-smooth system is a solution of an initial value problem ˙x=ϵf±(x,t,ϵ), x(a)=xaΩ if x(a)=xa.

    The formalism that we just introduced is called the Filippov regularization of a piecewise-smooth system (see [1,2]).

    Next we define two properties for which we will prove the averaging theorem.

    Definition 3.4. We say that a well-behaved piecewise-smooth system (Ω,G,f,f+) is uniformly transversal at the point x0Ω0 if there exists a positive constant m such that

    f±(x0,t,ϵ)G(x0)m

    for all tR and ϵ[0,1].

    The uniform transversality at x0 also implies that ¯f±(x0)G(x0)m.

    Definition 3.5. We say that a well-behaved piecewise-smooth system (Ω,G,f,f+) is uniformly sliding, if there exists a positive constant m such that

    f(x,t,ϵ)G(x)mandf+(x,t,ϵ)G(x)m

    for all xΩ0, tR and ϵ[0,1].

    The condition of uniform sliding itself is not sufficient to prove an averaging theorem. We need to reformulate the guiding system in a following way. Suppose that the condition of uniform sliding is true for every xΩ0. This implies, that any solution sliding at t0 has to stay on Ω0 up to some t1>t0. Hence the derivative of the solution is tangent to the manifold Ω0 or, equivalently, it is orthogonal to G(x). A short calculation shows that for a given (x,t,ϵ) such that the piecewise smooth system is sliding at (x,t,ϵ), there is a unique element of conv({f(x,t,ϵ),f+(x,t,ϵ)}), which is orthogonal to G(x). We define function g:Ω0×R×[0,1]Rn,

    g=αf+(1α)f+ (3.1)

    where

    α=f+Gf+GfG. (3.2)

    The function g can be used to define an induced differential equation on some neighborhood of (x,t,ϵ) in the set Ω0×R×[0,1]. The smoothness of g depends on the smoothness of f, f+ and G; see [1] for details. We can apply averaging to this function to get

    ˉg(x)=1TT0g(x,s,0)ds (3.3)

    which serves as a right-hand side of the guiding system ˙u=ˉg(u) for the equation induced on the discontinuity surface.

    Definition 3.6. Let (Ω,G,f,f+) be a well-behaved piecewise-smooth system that is uniformly sliding. Let ˉg be defined as in (3.3). The sliding adjusted guiding system associated with the initial value problem ˙x=ϵf±(x,t,ϵ), x(0)=aΩ is given by

    ˙w={ˉf(w), xΩ,ˉg(w), xΩ0.

    An absolutely continuous function w:[0,L]Ω is a solution of the sliding adjusted guiding system, if it satisfies the above equality almost everywhere.

    Suppose that we are given an initial value problem for a piecewise-smooth system (Ω,G,ϵf,ϵf+),

    ˙x=ϵf±(x,t,ϵ),x(0)=a (3.4)

    such that both f and f+ are T-periodic. Our goal is to generalize the averaging theorem, Theorem 2.2, to this setting. Akin to Definition 2.1, we can define

    ¯f±(w)=1TT0f±(w,s,0)ds

    and also define a guiding system associated with (3.4) as

    ˙w=ˉf±(w),w(0)=a. (3.5)

    Given a solution w:[0,T]Ω of the guiding system (3.5), it is clear that if (Im w)Ω0=, we can use the result for smooth systems by restricting ourselves to a neighborhood of Im w that entirely lies inside Ω or inside Ω+.

    That is the reason why we will be concerned only with solutions of the guiding system (3.5), which intersect Ω0. Furthermore, we will assume that w(0)=aΩ. It is easy to see that if aΩ+, it suffices to exchange G for G and all the results hold. We are not going to explicitly consider aΩ0, but such an initial condition is a special case that can usually be easily inferred from presented results.

    Let us start with an example, which shows that we cannot simply adapt the averaging theorem by adding the word `piecewise' to the statement. More specifically, we construct a well-behaved piecewise-smooth system and a transversal solution of the associated initial value problem and we show that the result of the averaging theorem does not hold.

    Example 4.1. Let us define g as a periodic function with period 18, which is defined on the interval [0,18] by the following formula (see Figure 1):

    g(t)={1, t[0,2],20(t2)1, t[2,3],20(t3)+19, t[3,4],1, t[4,18].
    Figure 1.  Graph of function g.

    We can calculate that

    180g(t)dt=18+20212=2.

    We will use the standard mollifier as defined in [3,p.629],

    μ(t)={(11e1y21dy)1e1t21,t(1,1),0,t(1,1).

    For each δ>0, we set

    μδ(t)=1δμ(tδ)

    and we define

    gδ(t)=(μδg)(t)=δδμδ(y)g(ty)dy.

    By [3,p.630,Theorem 6(ⅰ)], gδC(R). Let 0<δ<2 be fixed such that

    suptR|g(t)gδ(t)|=supt[0,18]|g(t)gδ(t)|<12

    and 180gδ(t)dt>0, which can be done due to [3,p.630,Theorem 6(ⅲ)]. We can immediately see that

    gδ(t){12,tkZ([0+18k,2+18k][4+18k,20+18k]),12,tkZ[2.1+18k,3.9+18k]. (4.1)

    We will show that gδ has exactly two zeros in [0,18]. The derivative of gδ is

    (gδ)(t)=(μδh)(t)=δδμδ(y)h(ty)dy

    where h is the periodic continuation of

    h(t)={0, t[0,2],20, t[2,3],20, t[3,4],0, t[4,18].

    This implies that (gδ)(t)>0 for t(2δ,3) and (gδ)(t)<0 for t(3,4+δ). Hence, by (4.1), gδ has on [0,18] just two zeros: r1(2,3), r2(3,4).

    Consider ˙x=ϵf±(x,t), xR, tR with a boundary function G(x)=x and

    f(x,t)=gδ(t),f+(x,t)=gδ(t).

    Such a system is well-behaved. The corresponding averaged version of these functions is

    ˉf+=¯f+(x)=118180gδ(t)dt=I>0

    and the guiding system is ˙w=I. Consider the initial value problem w(0)=C whose solution is w:[0,L]R, w(t)=C+It where CR+ and L>CI.

    We will show that although the averaged system has a transversal solution for all choices of C and L satisfying the conditions above, the original system has no solution that begins in R and then crosses over to R+.

    Suppose that xϵ(t) is such a solution, meaning that xϵ(0)<0 and there is a t such that xϵ(t)=0 (that is, it intersects the boundary Ω0={0}).

    Let t0=sup{ts[0,t):xϵ(s)<0}. The definition of t0 and existence of t imply that xϵ(t0)=0. Since for all s<t0 we have xϵ(s)<0, for each mN there is tm(t01m,t0) such that ϵgδ(tm)=˙xϵ(tm)>0. This implies that t0[r1+18l,r2+18l] for some lZ. Consequently, for all t[t0,r2+18l] it holds that ϵf(t)0, ϵf+(t)<0 implying that xϵ(t)=0.

    For all t(r2+18l,(18r2)+18l) it holds that ϵf(t)<0, ϵf+(t)<0 which means that ˙xϵ(t)<0. Thus, xϵ(t)<0 and also xϵ((18r2)+18l)<0.

    For all t[(18r2)+18l,r1+18(l+1)], ϵf(t)<0 which implies

    xϵ(r1+18(l+1))<x((18r2)+18l)<0.

    Since

    r2+18(l+1)r2+18lϵf(t)dt>0,

    there is t1(r1+18(l+1),r2+18(l+1)) such that xϵ(t1)=0 and t1=sup{ts(r2+18l,t):xϵ(s)<0}.

    The same analysis as for t0 can be repeated for t1, etc. The solution can be extended up to infinity, but does not cross the boundary. However, the solution of the guiding system crosses the boundary without a problem. The problem is that the guiding system does not reflect the behavior of the original equation at the discontinuity boundary.

    Now, we provide an example of an initial value problem for a well-behaved piecewise-smooth system. The solution of the associated guiding system is transversal, while there is a solution of the original system for which the estimate from the averaging theorem holds and this solution is not transversal.

    Example 4.2. Let G(x)=x be the boundary function and consider a system ˙x=ϵf±(x,t), xR, tR with

    f(x,t)=1,f+(x,t)=12cost

    and an initial condition x(0)=2π.

    The guiding system is ˙w=1, which is transversal at x=0 and we can immediately write the solution of the guiding system as

    w(τ)=2π+τ

    which means that

    wϵ(t)=2π+ϵt.

    The solution of the original system up to the intersection with the discontinuity, that is for t[0,2πϵ], is given by

    xϵ(t)=2π+ϵt=w(ϵt).

    The solution xϵ can cross Ω0 transversally at t=2πϵ only if f+(0,2πϵ)>0. That implies

    tkZ[π3+2kπ,π3+2kπ].

    Let ϵk>0 be such that 1ϵk(16+k,16+k) for some kZ (see Figure 2). For such ϵk, the solution xϵk makes the first contact with the boundary at t=2πϵk(π3+2kπ,π3+2kπ). However, since f+(t)0 up to π3+2kπ, we can say that the solution "sticks" to x=0 up to t=π3+2kπ.

    Figure 2.  Comparison of w and a "sticking" solution xϵ. Graph of the solution w of the guiding system and xϵ for t1ϵ and 1ϵ(16+k,16+k) for some kZ.

    By solving the system ˙x=ϵkf+(x,t), x(π3+2kπ)=0 we obtain a continuation of the solution xϵk, namely

    xϵk(t)=ϵkt2ϵksint+Ck

    where Ck=2ϵksin(π3+2kπ)ϵk(π3+2kπ).

    Such a solution does not make a contact with the boundary again. To see that, it only suffices to look at the values of t2sintπ3+2sin(π3) for t>π3.

    Now, we show that condition (2.4) is satisfied. For t[2πϵk,2kπ+π3],

    |wϵk(t)xϵk(t)|=|w(ϵkt)0||ϵk||t2πϵk|2π3|ϵk|.

    For t>2kπ+π3,

    |wϵk(t)xϵk(t)|=|2π+ϵktϵkt+2ϵksintCk|=|ϵk||2sint2sin(π3+2kπ)+π3+2kπ2πϵk||ϵk|(4+2π3)

    because

    |π3+2kπ2πϵk|2π3.

    The analysis for other values of ϵ is a bit more tricky. The solution crosses Ω0 transversally, but it can make a contact with the boundary again. If it makes the second contact with the boundary, it will continue with the same motion as we just analyzed (see Figure 3).

    Figure 3.  Comparison of w and a "non-sticking" solution xϵ. Graph of the solution w of the guiding system and xϵ for t1ϵ and 1ϵ[16+k,16+k] for each kZ. The solution for ϵ=0.27 intersects the discontinuity boundary again and sticks to it for a bit before continuing to Ω+.

    However, the difference in values of wϵ and xϵ will be less than C|ϵ| for some C>0. That means that the averaging estimate still holds on an interval [0,T] for any T.

    Note that the condition that the solution of the guiding system w is transversal translates to a condition that the system ˙w=ˉf±(w) is transversal at x0. However, that does not imply that the original system ˙x=ϵf(x,t,ϵ) is transversal at x0 for all tR,ϵ[0,1]. This means that the behavior of the original system near x0Ω0 is not reflected by the guiding system and can be more complex. For this reason, we introduce the notion of uniform transversality of the system ˙x=ϵf(x,t,ϵ) at x0 (uniform in t and ϵ).

    Now consider the guiding system ˙w=ˉf±(w) associated with ˙x=ϵf±(x,t,ϵ). The right-hand side of the differential equation induced by this guiding system on the discontinuity boundary is given by

    h(w)=β(x)ˉf(w)+(1β(x))ˉf+(w)

    where β is given by

    β(x)=ˉf+(x)G(x)ˉf+(x)G(x)ˉf(x)G(x).

    The sliding solution of the guiding system ˙w=ˉf±(w) is governed by h, but ˉg of (3.1) is the true right-hand side of the guiding system for the sliding motion on Ω0.

    A question that naturally arise is: What is the relationship between ˉg and h? We could hope that these two functions are equal. However, this is not the case in general. The function ˉg is a result of first inducing an equation on Ω0 and then averaging. For h, those operations are reversed. That is, we first average and induce an equation on Ω0 afterwards. Their relationship is not that simple.

    The following example shows a specific system, where ˉg and h are not equal.

    Example 4.3. Let the discontinuity surface be defined by G:R2R, G(x)=x1 and consider the system ˙x=ϵf±(x,t), xR2, tR with

    f(x,t)=(1,1),f+(x,t)=(1+12sint,3)

    along with the initial condition x(0)=(1,1).

    The differential equation induced on Ω0={0}×R is ˙x=ϵg(x,t), where g depends only on t, and is given by

    g(x,t)=(0,α(t)1+(1α(t))3).

    The function α is defined by the equality

    0=α(t)1+(1α(t))(1+12sint).

    By solving for α and substituting into the previous equality, we get

    g(x,t)=(0,412sint212sint).

    By averaging over the period, we obtain

    ˉg(x)=(0,1+415)(0,2.0328).

    The guiding system for this example is given by the right-hand side

    ˉf(x)=(1,1),ˉf+(x)=(1,3)

    which induces a differential equation on Ω0, ˙x=ϵh(x), where

    h(x)=(0,121+123)=(0,2)ˉg(x).

    Let us compare the solution of the original initial value problem xϵ and wϵ obtained from the guiding system for τ[0,2] (that is, t[0,2ϵ]).

    For t[0,1ϵ] we have

    xϵ(t)=wϵ(t)=(1+ϵt,1+ϵt)

    and the solutions make contact with the boundary at t=1ϵ in x=(0,0). After that time, both solutions slide along the discontinuity boundary Ω0. The motion of xϵ is given by the differential equation ˙xϵ=ϵg(xϵ,t), while the motion of wϵ is given by ˙wϵ=ϵh(wϵ)=ϵ(0,2) (see Figure 4).

    Figure 4.  Difference between xϵ and wϵ on the discontinuity boundary.

    To analyze the behavior of xϵ on the boundary for sufficiently small ϵ, we can employ the averaging theorem for smooth systems. In the following, let us drop the first coordinate (which is constant and equal to 0).

    The guiding system for the equation on the boundary has the right-hand side equal to ˉg0. Using Lemma 2.4 (an extension of the averaging theorem), there is ϵ0>0, C>0 such that for all ϵ(0,ϵ0] and t[1ϵ,2ϵ],

    |xϵ(t)ϵ(1+415)(t1ϵ)|Cϵ.

    In particular, at t=2ϵ we have

    |xϵ(2ϵ)(1+415)|Cϵ.

    On the other side,

    wϵ(2ϵ)=2ϵ(t1ϵ)|t=2ϵ=21+415,

    which means that it cannot be true that there exists a constant ˜C>0 such that for all t[1ϵ,2ϵ] and all ϵ>0 sufficiently small,

    |xϵ(t)wϵ(t)|˜Cϵ.

    Indeed, otherwise one would get a contradiction from

    |2(1+415)||xϵ(2ϵ)2|+|xϵ(2ϵ)(1+415)|ϵ(˜C+C)

    (see Figure 5).

    Figure 5.  Comparison of g, ˉg and h.

    For the purpose of this section, we denote Br(a)={xRn|xa|<r}. The next theorem shows that the function G can be thought of as a distance function on some neighborhood of a point x0Ω0.

    Theorem 5.1. Let ΩRn be an open set and G:ΩR a continuously differentiable function. Suppose that x0Ω0={xΩG(x)=0} and that G(x0)0. Then there exist positive numbers d,α,βR, such that ¯Bd(x0)Ω, and for each x¯Bd(x0),

    α|G(x)|dist(x,Ω0)β|G(x)|.

    Proof. Since the functions (x,y)G(x)G(y) defined on Ω×Ω and x|G(x)| defined on Ω are continuous and G(x0)G(x0)>0, |G(x0)|>0, there are positive constants d, a, A, b, B such that ¯B2d(x0)Ω and for all x,y¯B2d(x0),

    aG(x)G(y)A,b|G(x)|B. (5.1)

    Suppose that x¯Bd(x0) and zΩ0 is such that |xz|=dist(x,Ω0). Since x0Ω0 and |xx0|d, it follows that |xz|d. Hence |x0z|2d, i.e., z¯B2d(x0). Consequently, G(z)0. It follows from the geometry that xz=kG(z) where k=sgn G(x)|xz||G(z)|=sgn G(x)dist(x,Ω0)|G(z)|. Hence, it holds

    G(x)=G(z)+10G(z+t(xz))(xz)dt=sgn G(x)dist(x,Ω0)|G(z)|10G(z+t(xz))G(z)dt.

    Using the inequalities (5.1), we arrive at

    aBdist(x,Ω0)|G(x)|Abdist(x,Ω0)

    and the statement easily follows.

    In the next lemma, we consider a system with C1-smooth right-hand side but included in Ω=Ω+Ω0Ω. So, we can estimate the time and position of the contact with Ω0.

    Lemma 5.2. Let f be T-PCDLB and G be a boundary function. Let w:[0,L]Ω be a solution of the initial value problem ˙w=ˉf(w), w(0)=aΩ and xϵ:[0,Lϵ]Ω be a solution to ˙x=ϵf(x,t,ϵ), x(0)=a. Assume that w intersects Ω0 in exactly one point x0 at τ0(0,L) and ˉf(x0)G(x0)>0.

    Then there exist positive constants ϵ0,A,B such that for all ϵ(0,ϵ0], the solution xϵ intersects Ω0 in some point and if we define tϵ to be the smallest t such that xϵ(t)Ω0, then

    |tϵτ0ϵ|Aand|xϵ(tϵ)x0|Bϵ.

    Proof. Let M be a bound of |f|. Due to the smooth averaging, Lemma 2.4 can be used to show that there exist C>0 and ϵ0>0 sufficiently small such that for all ϵ(0,ϵ0] and t[0,Lϵ], it holds

    |xϵ(t)w(ϵt)|Cϵ. (5.2)

    In the rest of the proof, ϵ0 may decrease.

    Since w(L)Ω+, we have G(w(L))>0. Then the estimate (5.2) and the continuity of G imply that G(xϵ(Lϵ))>0 for all ϵ(0,ϵ0]. Since G(xϵ(0))=G(a)<0, the solution xϵ crosses Ω0 in some point.

    Due to the continuity of mapping xˉf(x)G(x) and Theorem 5.1, there exist positive constants d, m, α, β such that ¯Bd(x0)Ω, and for all x¯Bd(x0), it holds

    ˉf(x)G(x)mandα|G(x)|dist(x,Ω0)β|G(x)|. (5.3)

    Let A be a constant such that A>Cmα where C is from (5.2). Let ϵ0 be such that τ0ϵ0A>0, τ0+ϵ0A<L and MAϵ0d. If we choose ϵ(0,ϵ0], then for all τ[τ0Aϵ,τ0+Aϵ][0,L], it holds

    |w(τ)x0|ττ0|ˉf(w(s))|dsM|ττ0|MAϵd,

    i.e., w(τ)¯Bd(x0). Since w(τ0+Aϵ)Ω+, we have G(w(τ0+Aϵ))>0 and we obtain

    G(w(τ0+Aϵ))=G(w(τ0))+τ0+Aϵτ0ˉf(w(s))G(w(s))dsAϵm

    due to the first inequality of (5.3). Now, we can use the other inequality of (5.3) to get

    dist(w(τ0+Aϵ),Ω0)α|G(w(τ0+Aϵ))|αAϵm.

    Observe that the set w([τ0+Aϵ0,L]) is compact and lies in Ω+. Therefore, dist(w([τ0+Aϵ0,L],Ω0)>0. Consequently, for some ϵ0>0 and all ϵ(0,ϵ0], we obtain

    dist(w([τ0+Aϵ,L]),Ω0)min{dist(w([τ0+Aϵ,τ0+Aϵ0]),Ω0),dist(w([τ0+Aϵ0,L]),Ω0)}αAϵm.

    The estimate (5.2) implies that for any ϵ(0,ϵ0],

    dist(xϵ([τ0+Aϵϵ,Lϵ]),Ω0)dist(w([τ0+Aϵ,L]),Ω0)supt[τ0+Aϵϵ,Lϵ]|xϵ(t)w(ϵt)|(αAmC)ϵ.

    Due to our choice of A, we deduce that xϵ(t)Ω+ for all t[τ0+Aϵϵ,Lϵ], ϵ(0,ϵ0].

    Using similar arguments, we can show that

    dist(w(τ0Aϵ),Ω0)αAϵmandxϵ([0,τ0Aϵϵ])Ω

    for all ϵ(0,ϵ0] for some (possibly smaller than before) ϵ0>0. Hence, there exists tϵ(τ0Aϵϵ,τ0+Aϵϵ) such that xϵ(tϵ)Ω0. Furthermore, the following estimate is valid:

    |xϵ(tϵ)w(τ0)||xϵ(tϵ)w(ϵtϵ)|+|w(ϵtϵ)w(τ0)|Cϵ+M|ϵtϵτ0|ϵ(C+MA).

    In this section, we prove an averaging theorem for uniformly transversal systems. First we prove an auxiliary result.

    Lemma 6.1. Let (Ω,G,f,f+) be a well-behaved piecewise-smooth system that is uniformly transversal at x0Ω0. Let w:[0,L]Ω be a solution of the guiding system ˙w=¯f±(w),w(0)=aΩ which intersects the boundary Ω0 in exactly one point w(τ0)=x0 at τ0(0,L).

    Suppose that there is ϵ1(0,1] such that for all ϵ(0,ϵ1] there is a unique solution xϵ:[0,Lϵ]Ω of the problem ˙xϵ=ϵf±(xϵ,t,ϵ), xϵ(0)=a intersecting the boundary Ω0 at exactly one point in time.

    Then there are ϵ0(0,ϵ1] and C>0 such that for all ϵ(0,ϵ0] and all t[0,Lϵ],

    |xϵ(t)w(ϵt)|Cϵ. (6.1)

    Proof. We know that w is a solution of ˙w(τ)=ˉf(w(τ)) for τ[0,τ0], satisfying w(0)=a. Since the right-hand side ˉf is defined on the whole Ω, there is a solution w:[0,L]Ω of the same problem for some L>τ0. This solution is equal to w on [0,τ0]. The uniform transversality at x0 implies that ˙w(τ0)0. Hence the solution w intersects Ω0 exactly once if L is sufficiently close to τ0.

    In what follows, the constants ϵ0 and C may vary from step to step. By Theorem 2.2, for every ϵ(0,ϵ0], there is a unique solution xϵ:[0,Lϵ]Ω of the initial value problem ˙x=ϵf(x,t,ϵ), x(0)=a such that

    |xϵ(t)w(ϵt)|Cϵ (6.2)

    for all t[0,Lϵ]. The assumption of uniform transversality at x0 and Lemma 5.2 imply that there are positive constants A, B such that for all ϵ(0,ϵ1], the solution xϵ intersects Ω0 and it holds

    |tϵτ0ϵ|Aand|xϵ(tϵ)x0|Bϵ (6.3)

    where tϵ is the earliest time of intersection. Our assumption on f implies that tϵ is the time of the only intersection. Clearly, the solution xϵ is equal to xϵ on [0,tϵ] and xϵ(t)Ω+ for t(tϵ,L].

    Since the functions f± are bounded, the solutions xϵ, w are Lipschitz continuous. For t[τ0ϵA,τ0ϵ+A], we estimate

    |xϵ(t)w(ϵt)||xϵ(t)xϵ(τ0ϵA)|+|xϵ(τ0ϵA)w(τ0ϵA)|+|w(τ0ϵA)w(ϵt)|.

    Since xϵ(τ0ϵA)=xϵ(τ0ϵA) and w(τ0ϵA)=w(τ0ϵA), using Lipschitz continuity of xϵ, w and inequality (6.2), we arrive at (6.1) for t[τ0ϵA,τ0ϵ+A].

    Now, let t[τ0ϵ+A,Lϵ]. For all such t we have that w(t),xϵ(t)Ω+. Hence xϵ is a solution of equation ˙x=ϵf+(x,t,ϵ) and w is a solution of the corresponding guiding system. Lemma 2.4 with L1=τ0, a=w(τ0) and bϵ=xϵ(τ0ϵ+A) implies the existence of ϵ0,C>0 such that for all ϵ(0,ϵ0], it holds

    |xϵ(t)w(ϵt)|Cϵ

    for all t[τ0ϵ+A,Lϵ]. This estimate concludes the proof.

    Theorem 6.2. Let (Ω,G,f,f+) be a well-behaved piecewise-smooth system that is uniformly transversal at x0Ω0. Let w:[0,L]Ω be a solution of the guiding system ˙w=¯f±(w), w(0)=aΩ which intersects the boundary Ω0 in exactly one point w(τ0)=x0 at time τ0(0,L).

    Then there are ϵ0>0 and C>0 such that for all ϵ(0,ϵ0] there exists a unique solution xϵ of the problem ˙xϵ=ϵf±(x,t,ϵ), xϵ(0)=a, intersecting Ω0 exactly once, the intersection is transversal, and for all t[0,Lϵ], it holds

    |xϵ(t)w(ϵt)|Cϵ. (6.4)

    Proof. The main idea of the proof is to use the averaging theorem, Theorem 2.2, for smooth systems in Ω and Ω+ separately and to construct a solution of the piecewise-smooth system ˙x=ϵf±(x,t,ϵ) that crosses Ω0 only once.

    The proof is divided into two parts. Lemma 6.1 shows that, given a transversal solution of ˙x=ϵf±(x,t,ϵ), the averaging estimate (6.4) holds for all sufficiently small ϵ>0.

    Now, it only remains to show that there exists a unique transversal solution for all sufficiently small ϵ. The uniform transversality at x0 implies the existence of d,m>0 such that ¯B2d(x0)Ω and

    f±(x,t,ϵ)G(x)m2

    for all x¯B2d(x0), tR, ϵ[0,1].

    Let M be a bound of both |f+| and |f|. Analogously to the proof of Lemma 6.1, there is L>τ0 and a solution w:[0,L]Ω of ˙w=ˉf(w), w(0)=a that intersects Ω0 exactly once and w(τ)=w(τ) for τ[0,τ0]. Due to the averaging theorem for smooth systems, there is ϵ1(0,1] and C>0 such that for any ϵ(0,ϵ0], there is a unique solution xϵ:[0,Lϵ]Ω of the initial value problem ˙x=ϵf(x,t,ϵ), x(0)=a such that it holds

    |xϵ(t)w(ϵt)|Cϵ

    for all t[0,Lϵ]. In the following again, the constants ϵ0 and C may vary from step to step. By Lemma 5.2, there are constants A, B such that for all ϵ(0,ϵ0] the solution xϵ intersects Ω0 and it holds

    |tϵτ0ϵ|Aand|xϵ(tϵ)x0|Bϵ

    where tϵ is the time of the first contact. Let ϵ0 be so small that the following inequalities are true

    Bϵ0d,ϵ0AMd2,ϵ0Amin{d2M,Lτ0}. (6.5)

    The first inequality of (6.5) implies

    |xϵ(tϵ)x0|d

    for all ϵ(0,ϵ0].

    Now, let ϵ(0,ϵ0] be fixed. Denote x+ϵ(t):[tϵ,s]Ω the unique local solution of the initial value problem ˙x=ϵf+(x,t,ϵ), x(tϵ)=x(tϵ) for some s>tϵ. This solution can be extended until it approaches Ω. For t[tϵ,s], we estimate

    |x+ϵ(t)x0||x+ϵ(t)x+ϵ(tϵ)|+|x+ϵ(tϵ)x0|ϵM|stϵ|+d. (6.6)

    Denote Δ=min{d2M,Lτ0}. From the third inequality of (6.5) we deduce that

    τ0+Δϵτ0+ϵAϵtϵ.

    Assume that sτ0+Δϵ. The inequality tϵτ0ϵAϵ together with (6.5) imply

    ϵM|stϵ|ϵM|τ0+Δϵτ0ϵAϵ|=M(Δ+ϵA)d2+d2=d.

    Combining this estimate with (6.6), we get

    x+ϵ(t)¯B2d(x0)

    for any t[tϵ,s]. This means that the solution x+ϵ is separated from the boundary Ω on [tϵ,s] for arbitrary sτ0+Δϵ. Hence, it exists on [tϵ,τ0+Δϵ]. Moreover, since x+ϵ(t)¯B2d(x0) for all t[tϵ,τ0+Δϵ], it holds

    G(x+ϵ(t))=G(x+ϵ(tϵ))+ϵttϵf+(x+ϵ(s),s,ϵ)G(x+ϵ(s))dsϵm(ttϵ)2>0

    for all t[tϵ,τ0+Δϵ], and so x+ϵ(t)Ω+.

    For ϵ[0,ϵ0], define xϵ:[0,τ0+Δϵ]Ω as follows:

    xϵ(t)={xϵ(t), t[0,tϵ],x+ϵ(t), t(tϵ,τ0+Δϵ].

    This is a solution of the initial value problem ˙x=f±(x,t,ϵ), x(0)=a, which intersects Ω0 exactly once and the intersection is transversal. Lemma 6.1 implies validity of estimate (6.4) for ϵ(0,ϵ0] and t[0,τ0+Δϵ].

    If τ0+Δ=L the proof is done. Thus assume that τ0+Δ<L. Let

    h=dist(w([τ0+Δ,L]),Ω0)>0.

    Since ˙w(τ)=ˉf+(w(τ)) for τ(τ0,L), we can use Lemma 2.4 with L1=τ0+Δ, a=w(L1), bϵ=xϵ(τ0+Δϵ) and A=0. Hence, there is a unique solution x++ϵ:[τ0+Δϵ,Lϵ]Ω such that

    |x++ϵ(t)w(ϵt)|Cϵ

    for all t[τ0+Δϵ,Lϵ] and ϵ(0,ϵ0]. Observe that x++ϵ(t)Ω+ for all t[τ0+Δϵ,Lϵ] and ϵ(0,hC].

    Finally for ϵ[0,ϵ0], we define the desired solution xϵ:[0,Lϵ]Ω as

    xϵ(t)={xϵ(t),t[0,tϵ],x+ϵ(t),t(tϵ,τ0+Δϵ],x++ϵ(t),t(τ0+Δϵ,Lϵ]

    and estimate (6.4) is valid.

    The uniqueness of the solution xϵ follows from the smoothness of f, f+ and the transversality.

    Notice that G is used in the definition of the function g. To ensure that the function g is Lipschitz continuous, which is needed to use the averaging theorem for smooth systems, we will need to impose some further conditions on G. One can see that we deal with a differential equation defined on (n1)-dimensional surface. We will show that it can be analyzed using the techniques available for differential equations defined on open subsets of Rn by extending the equation to an open set in Ω.

    Lemma 7.1. Let (Ω,G,f,f+) be a well-behaved piecewise-smooth system that is uniformly sliding and let g be defined by (3.1). Let M be a compact subset of Ω0 and let G be continuously differentiable and Lipschitz continuous on some open neighborhood V of M. Then there exists an open set UV such that MU, g can be defined by the formula (3.1) on U×R×[0,1], and the function g:U×R×[0,1]Rn is T-PCDLB.

    Proof. We define a function ϕ:ΩR as

    ϕ(x)=mint[0,T],ϵ[0,1]{f(x,t,ϵ)G(x)f+(x,t,ϵ)G(x)}.

    The uniform sliding implies that ϕ(x)2m for all xΩ0, where m is a positive constant from the definition of uniformly sliding systems. For every xM, there is a radius d>0 such that ¯Bd(x)V and ϕ(y)m for all yBd(x). Since the compact set M is covered by these balls, there exists a finite open subcover of M. Let U be the union of the balls from the finite open subcover of M. It is clear that ¯UV and ϕ(x)m for all xU.

    This implies that for all xU, tR and ϵ[0,1],

    f(x,t,ϵ)G(x)f+(x,t,ϵ)G(x)m.

    Hence, the function g can be extended to ¯U and one can prove that it has the desired properties.

    We are now ready to formulate and prove an averaging theorem for sliding solutions of piecewise-smooth systems.

    Theorem 7.2. Let (Ω,G,f,f+) be a well-behaved piecewise-smooth system that is uniformly sliding. Let w:[0,L]Ω be a solution of the sliding adjusted guiding system

    ˙w={ˉf(w), xΩ,ˉg(w), xΩ0

    associated with the initial value problem ˙x=ϵf±(x,t,ϵ), x(0)=aΩ, where g is given by (3.1). Let τ0 be such that 0<τ0<L, w(τ)Ω for τ[0,τ0) and w(τ)Ω0 for τ[τ0,L]. Furthermore, let there be an open set V such that w([τ0,L])VΩ and G is continuously differentiable and Lipschitz continuous on V.

    Then there is a constant ϵ0>0 such that for all ϵ(0,ϵ0], there is a unique solution xϵ of the problem ˙x=ϵf±(x,t,ϵ), x(0)=aΩ defined on [0,Lϵ] and sliding from time tϵ. In addition, there are constants C,A>0 such that |tϵτ0ϵ|A and

    |xϵ(t)w(ϵt)|Cϵ

    for all t[0,Lϵ].

    Proof. The uniform sliding condition implies that

    ˉf(w)G(w)>0.

    Hence there is a solution w of ˙w=ˉf(w), w(0)=a defined on [0,L] for some L>τ0 such that w(τ)Ω+ for all τ(τ0,L].

    The averaging theorem for smooth systems implies that there exist ϵ0,C>0 such that for all ϵ(0,ϵ0], there is a unique solution xϵ of ˙x=ϵf(x,t,ϵ), x(0)=a defined on [0,Lϵ] and it holds

    |xϵ(t)w(ϵt)|Cϵ (7.1)

    for every t[0,Lϵ]. Lemma 5.2 implies that there are constants A, B such that for all ϵ(0,ϵ0], the solution xϵ crosses Ω0 and we have

    |tϵτ0ϵ|A,|xϵ(tϵ)x0|Bϵ,

    where tϵ is the time of the first contact of xϵ with Ω0.

    For τ[τ0,L], we define w0(τ)=w(τ). Thus the function w0 is the solution of the initial value problem ˙w=ˉg(w), w(τ0)=x0.

    Due to Lemma 7.1, there is an open set UV such that g is T-PCDLB on U. The uniform sliding condition implies that the solution xϵ can be uniquely extended as a solution x0ϵ of the initial value problem ˙x=ϵg(x,t,ϵ), x(tϵ)=xϵ(tϵ). Trajectories of solutions of this system lie in Ω0 and the local solution x0ϵ can be extended at least as long as the solution does not approach the boundary of U. Suppose that x0ϵ can be defined at least up to some s>tϵ.

    Assume that sτ0ϵ+A. For every t[tϵ,s], we estimate

    |x0ϵ(t)w0(ϵt)||x0ϵ(t)x0ϵ(tϵ)|+|xϵ(tϵ)xϵ(τ0ϵA)|+|xϵ(τ0ϵA)w(τ0ϵA)|+|w(τ0ϵA)w(τ0)|+|w0(τ0)w0(ϵt)|.

    Using estimate (7.1) and the Lipschitz continuity of solutions x0ϵ, xϵ and w, w0, we arrive at

    |x0ϵ(t)w0(ϵt)|Cϵ

    for tϵtsτ0ϵ+A. Observe that C does not depend on s, if sτ0ϵ+A and x0ϵ is separated from U for ϵ0 sufficiently small. Since s was arbitrary, the solution x0ϵ can be defined up to τ0ϵ+A.

    At this point, we can use Lemma 2.4 on the initial value problem ˙x=ϵg(x,t,ϵ), x(τ0ϵ+A)=x0ϵ(τ0ϵ+A) with w0:[τ0,L]Ω and L1=τ0, a=w0(τ0), L2=L. The assumptions of Lemma 2.4 are satisfied since

    |x0ϵ(τ0ϵ+A)w0(τ0)||x0ϵ(τ0ϵ+A)w0(τ0+ϵA)|+|w0(τ0+ϵA)w0(τ0)|Cϵ.

    Hence for all ϵ(0,ϵ0], the solution x0ϵ can be extended to the interval [τ0ϵ,Lϵ], and it holds

    |x0ϵ(t)w0(ϵt)|Cϵ

    for all t[τ0ϵ+A,Lϵ]. It is clear that the function

    xϵ(t)={xϵ(t), t[0,tϵ],x0ϵ(t), t(tϵ,Lϵ]

    is the desired unique solution.

    We have contributed to the study of piecewise-smooth differential equations by extending the results of averaging methods for periodic systems to the piecewise-smooth setting. Our main contribution is the analysis of how the boundary behavior influences averaging, especially in case of sliding. We have also proposed some sufficient conditions that allowed us to prove results on averaging in case of transversal crossing and sliding. Moreover, we demonstrated how to achieve results for solutions of piecewise-smooth systems by dividing them into solutions of smooth systems and patching together results from the smooth setting. Future opportunities for study include weakening of the assumptions in the transversal and sliding case and extension of the results to situations, in which f+G or fG vanish, which includes grazing solutions. Moreover, it would be interesting to develop an averaging method for differential equations with impacts [2].

    We thank the referee for valuable comments. The paper was supported in part by the Slovak Research and Development Agency under the contract No. APVV-18-0308 and by the Slovak Grant Agency VEGA No. 2/0153/16 and No. 1/0078/17.

    All authors declare no conflicts of interest in this paper.



    [1] F. M. Arscott, A. F. Filippov, Differential Equations with Discontinuous Righthand Sides: Control Systems, Netherlands: Springer, 2013.
    [2] M. Bernardo, C. Budd, A. R. Champneys, et al. Piecewise-Smooth Dynamical Systems: Theory and Applications, London: Springer Science & Business Media, 2008.
    [3] L. C. Evans, Partial Differential Equations, American Mathematical Society, 2010.
    [4] M. Fečkan, A Galerkin-averaging method for weakly nonlinear equations, Nonlinear Anal., 41 (2000), 345-369. doi: 10.1016/S0362-546X(98)00281-8
    [5] S. Ma, Y. Kang, Periodic averaging method for impulsive stochastic differential equations with Lévy noise, Appl. Math. Lett., 93 (2019), 91-97. doi: 10.1016/j.aml.2019.01.040
    [6] J. G. Mesquita, A. Slavík, Periodic averaging theorems for various types of equations, J. Math. Anal. Appl., 387 (2012), 862-877. doi: 10.1016/j.jmaa.2011.09.038
    [7] N. A. Perestyuk, V. A. Plotnikov, A. M. Samoilenko, et al. Differential Equations with Impulse Effects, Multivalued Right-Hand Sides with Discontinuities, Berlin: Walter de Gruyter GmbH & Co. KG, 2011.
    [8] J. A. Sanders, F. Verhulst, J. Murdock, Averaging Methods in Nonlinear Dynamical Systems, 2 Eds, New York: Springer, 2007.
  • This article has been cited by:

    1. Michal Fečkan, Július Pačuta, Averaging Methods for Second-Order Differential Equations and Their Application for Impact Systems, 2020, 8, 2227-7390, 916, 10.3390/math8060916
  • Reader Comments
  • © 2019 the Author(s), licensee AIMS Press. This is an open access article distributed under the terms of the Creative Commons Attribution License (http://creativecommons.org/licenses/by/4.0)
通讯作者: 陈斌, bchen63@163.com
  • 1. 

    沈阳化工大学材料科学与工程学院 沈阳 110142

  1. 本站搜索
  2. 百度学术搜索
  3. 万方数据库搜索
  4. CNKI搜索

Metrics

Article views(4133) PDF downloads(440) Cited by(1)

Figures and Tables

Figures(5)

/

DownLoad:  Full-Size Img  PowerPoint
Return
Return

Catalog