• Non ci sono risultati.

Andrea Sfecci

N/A
N/A
Protected

Academic year: 2021

Condividi "Andrea Sfecci"

Copied!
18
0
0

Testo completo

(1)

PERIODIC SOLUTIONS OF A SYSTEM OF COUPLED OSCILLATORS WITH ONE-SIDED SUPERLINEAR

RETRACTION FORCES Alessandro Fonda

Dipartimento di Matematica e Geoscienze, Universit` a di Trieste P.le Europa 1, I-34127 Trieste, Italy

Andrea Sfecci

SISSA - International School for Advanced Studies Via Bonomea 265, I-34136 Trieste, Italy

(Submitted by: Reza Aftabizadeh)

Abstract. We generalize the phase-plane method approach introduced in [9] to the case of higher dimensions. To this aim, the phase-space is assumed to be decomposable as a product of planes, where the cor- responding components of the solutions can be controlled by means of suitable plane curves. We then apply our general result to the periodic problem associated to a system of coupled oscillators, with retraction forces having a linear growth, or with one-sided superlinear nonlineari- ties.

1. Introduction

The aim of this paper is to extend to higher dimensions the existence results obtained in [9] for the periodic problem associated to some planar systems of ordinary differential equations.

We want to prove the existence of T -periodic solutions for a system like

u 0 = F(t, u), (1.1)

where F : R × R 2h → R 2h is a continuous function which is T -periodic in its first variable. As a particular case, we have in mind a system of coupled oscillators of the type

 

 

 

 

x

001

+ φ

1

(t, x

1

) = e

1

(t, x

1

, . . . , x

h

) x

002

+ φ

2

(t, x

2

) = e

2

(t, x

1

, . . . , x

h

)

.. .

x

00h

+ φ

h

(t, x

h

) = e

h

(t, x

1

, . . . , x

h

).

(1.2)

Accepted for publication: August 2012.

AMS Subject Classifications: 34C25, 34C15.

993

(2)

Alessandro Fonda and Andrea Sfecci

Some existence theorems for general systems of this type have been provided by the use of functional analytical methods, typically using bifurcation the- ory, or degree theory. We refer to [1, 10] and the references therein, for some classical results in this direction. Moreover, in the variational setting, when (1.1) is a Hamiltonian system, there is a large literature on this type of problem. See, e.g., [3, 4, 13, 15] and the references therein.

In [9], we studied the case h = 1, by the use of phase-plane methods. Such methods are frequently applied to planar systems, but very rarely used in higher dimensions, due to the difficulty in controlling the solutions in the phase-space. The aim of the present paper is to provide a setting where it is possible to have such control, at least when the coupling forces e i (t, ·) have a sublinear growth at infinity. On the other hand, we are able to deal with many different situations involving the growth of the functions φ i (t, ·). As in [9], we can deal with functions having a linear growth, assuming either nonresonance at infinity, or a Landesman–Lazer type of situation, or even with one-sided superlinear nonlinearities. Notice that our result is not of perturbative type, like e.g. the ones in [5, 11, 14], and many others, in the sense that we do not require the functions e i (t, ·) to depend on a small parameter.

We will mainly concentrate on the situation of one-sided superlinear re- traction forces, since, in our opinion, it has not yet been sufficiently studied in the literature, for higher-dimensional systems like (1.2). We recall that, in the case of the periodic problem for a second-order scalar equation, one-sided superlinear growth was first considered in the pioneering papers by Mawhin–

Ward [12] and Fabry–Habets [7], while a particular higher-dimensional sit- uation was studied by Arioli and Ruf in [2], by the use of a variational method.

Let us explain the main result of this paper, recalling first the approach we used in [9]. The idea there was inspired by [7], where a particular curve γ was constructed in order to control the solutions in the phase plane. We then used such a curve to have the necessary estimates in order to apply the Poincar´ e–Bohl fixed-point theorem to the Poincar´ e map associated to (1.1), thus obtaining the existence of a T -periodic solution. Passing to higher di- mensions, one could try to generalize this approach, introducing some kind of manifolds in order to have the same type of control. This seems a very del- icate problem, and it is not clear to us how such manifolds could be defined.

As an alternative approach, we separate the phase space as the product of

h planes, and on each of them we construct a curve γ i , which controls the

solutions in that particular plane. Assuming the coupling forces e i (t, ·) to

(3)

have an appropriate sublinear growth at infinity, the behaviour of a large- amplitude solution x(t) = (x 1 (t), . . . , x h (t)) of (1.2) will be approximately the same as if the oscillators were uncoupled, so that each component x i (t) of the solution will be controlled by the corresponding curve γ i .

Let us briefly describe how the paper is organized. In Section 2, we provide our general setting for a system in R 2h , with h ≥ 1. We state and prove our main existence result, thus generalizing what we have done in [9] in the case h = 1. In Section 3 we particularize our assumptions in view of the applications we have in mind. In Section 4, we deal with a system with nonlinearities having either linear growth, or one-sided superlinear growth.

In Section 5, we show how our existence result applies for a system of coupled oscillators.

A few words about the notation to be used on each phase plane. We denote by h·, ·i the Euclidean scalar product in R 2 , and by | · | the corresponding norm. The open ball, centered at the origin, with radius R > 0, is B 2 R = {v ∈ R 2 : |v| < R}, and by S 1 we denote the set {v ∈ R 2 : |v| = 1}. The cone determined by two angles θ 1 < θ 2 is defined as

Θ(θ 1 , θ 2 ) = {v ∈ R 2 : v = ρe , ρ ≥ 0 , θ ∈ [θ 1 , θ 2 ]}.

(It will be sometimes convenient to use the complex notation for the points in R 2 .) The closed segment joining two points v 1 and v 2 is denoted by [v 1 , v 2 ].

Finally, we use the standard notation J =

 0 −1

1 0

 . 2. The main result

In this section we are going to introduce a generalization of the main result obtained in [9] to a particular system in R 2h with h ≥ 1.

Consider an open set A, containing the origin, with the following shape:

A = U 1 × · · · × U h , where, for i ∈ {1, . . . , h}, the set U i ⊂ R 2 is open and bounded. In the following we will write any point u ∈ R 2h using the coordinates (u 1 , . . . , u h ), where u i ∈ R 2 . In the same way, all the functions ϕ with image in R 2h will be written using the components (ϕ 1 , . . . , ϕ h ).

We start by stating the following result, reminiscent of the Poincar´ e–Bohl theorem, whose proof is a standard application of Brouwer degree theory (see, e.g., [6]).

Theorem 2.1. Let ϕ : A → R 2h be a continuous function such that, for every i ∈ {1, . . . , h}, the following property holds:

ϕ i (u 1 , . . . , u h ) 6= µu i , for every µ > 1 and for every

u ∈ U 1 × · · · × U i−1 × ∂U i × U i+1 × · · · × U h .

(4)

Alessandro Fonda and Andrea Sfecci

Then ϕ has a fixed point in A.

We consider the equation

u 0 = F(t, u), (2.1)

where u = (u 1 , . . . , u h ) ∈ R 2h and F = (f 1 , . . . , f h ), f i : R × R 2h → R 2 being continuous functions which are T -periodic in the first variable.

Let us recall the definition of clockwise rotating regular spiral in a plane, which can be found in [9].

Definition 2.2. A clockwise rotating regular spiral is a continuous and injective curve γ : [0, +∞) → R 2 , satisfying the following properties:

(1) there exists an unlimited strictly increasing sequence 0 = σ 0 < σ 1 < σ 2 < · · · < σ k < σ k+1 < · · ·

such that the restriction of γ to every closed interval [σ k , σ k+1 ] is continuously differentiable, and such that

hJ ˙γ(s), γ(s)i > 0, for every s ∈ [σ k , σ k+1 ] ; (2.2) (2) the curve grows in norm to infinity:

s→+∞ lim |γ(s)| = +∞ ; (2.3)

(3) the curve rotates clockwise infinitely many times:

Z +∞

0

hJ ˙γ(s), γ(s)i

|γ(s)| 2 ds = +∞. (2.4)

A similar definition can be given for a counterclockwise rotating regular spi- ral, by changing the inequality in (2.2), and requiring the integral in (2.4) to be equal to −∞.

Without loss of generality, we will assume that the spiral has the following parametrization: γ(s) = |γ(s)|(cos s, − sin s).

Now we want to introduce an admissibility condition for such spirals, extending the definition in [9] to the case of a higher-dimensional space.

Definition 2.3. A clockwise rotating regular spiral γ is said to be i-adm- issible for system (2.1), with i ∈ {1, . . . , h}, if, when restricted to any subin- terval [σ k , σ k+1 ], it satisfies

hJ ˙γ(s), f i (t, u 1 , . . . , u i−1 , γ(s), u i+1 , . . . , u h )i < 0, (2.5)

for every t ∈ [0, T ], s ∈ [σ k , σ k+1 ], and u j ∈ R 2 with j 6= i. (The sequence

k } k is the one introduced in Definition 2.2.) Moreover, given a subset U

(5)

of R 2 , the spiral is said to be i-admissible in U for system (2.1) if (2.5) is satisfied whenever γ(s) ∈ U .

Starting from a spiral γ i which is i-admissible for (2.1), it is convenient to define, for every n ∈ N, the set Ω i n : it is the open region delimited by the Jordan curve Γ i n obtained by gluing together the piece of curve γ i going from γ i (2πn) to γ i (2π(n + 1)), and the segment joining the two endpoints:

Γ i n = γ i (s) : s ∈ [2πn, 2π(n + 1)]

∪ γ i (2πn), γ i (2π(n + 1)).

Notice moreover that, by the injectivity, one has

i (s)| < |γ i (s + 2π)| for every s > 0. (2.6) Let us now state our main result.

Theorem 2.4. Suppose that the following assumptions hold, for every i ∈ {1, . . . , h}.

(H1 i ) There exists a clockwise rotating regular spiral γ i : [0, +∞[ → R 2 which is i-admissible for (2.1).

(H2 i ) There exists R i > 0 such that, for any solution u : [0, T ] → R 2h of (2.1), satisfying

|u i (t)| ≥ R i , for every t ∈ [0, T ], one has that, either |u i (T )| < |u i (0)|, or

Z T 0

hJ u 0 i (t), u i (t)i

|u i (t)| 2 dt / ∈ 2πN.

(H3 i ) There exist C i > 0 and θ i 1 < θ i 2 such that hJ f i (t, u), u i i ≤ C i (|u i | 2 + 1), for every t ∈ [0, T ] and u ∈ R 2h with u i ∈ Θ(θ i 1 , θ i 2 ).

Then, a T -periodic solution of equation (2.1) exists.

Proof. Take R ≥ max{1, R 1 , . . . , R h } such that Ω 0 i ⊆ B R 2 for every i. Let m 1 be a positive integer such that B 2 R ⊆ Ω i m

1

for every i, and let ¯ n be an integer such that, for every i,

¯

n > (C i + 1)T

θ 2 i − θ 1 i . (2.7)

We can find a R 1 > R such that Ω m i

1

n+1 ⊆ B R 2

1

for every i. In the same way we can find an integer m 2 > m 1 + ¯ n + 1 such that B 2 R

1

⊆ Ω i m

2

for every i, and a constant R 2 > R 1 such that Ω m i

2

n+1 ⊆ B R 2

2

for every i.

(6)

Alessandro Fonda and Andrea Sfecci

Define, for any r > 0, B r = (B 2 r ) h = B r 2 × · · · × B r 2

| {z }

h times

⊂ R 2h . Consider a sequence ( F n ) n = (f 1 n , . . . , f h n ) n of locally Lipschitz-continuous functions converging to F uniformly on [0, T ] × B R

2

. For any i, by (2.5), as long as, for some s, ˜ u = (u 1 , . . . , u i−1 , γ i (s), u i+1 , . . . , u h ) belongs to B R

2

, then, for n large enough,

hJ ˙γ i (s), f i n (t, ˜ u)i < 0, for every t ∈ [0, T ] ; (2.8) moreover, by (H3 i ), for n sufficiently large,

hJ f i n (t, v), v i i

|v i | 2 ≤ C i + 1, (2.9)

for every t ∈ [0, T ] and v ∈ B R

2

whose i-th component is such that v i ∈ Θ(θ i 1 , θ i 2 ) ∩ (B 2 R

2

\ B R 2 ).

The solution to the Cauchy problem associated to the equation

u 0 = F n (t, u) (2.10)

is unique for every n, and, if u n is a solution of (2.10) satisfying |u n i (0)| ≤ R 1

for every i, then, for sufficiently large n,

|u n i (t)| < R 2 , for every t ∈ [0, T ] and for every i (2.11) (i.e., u n (t) ∈ B R

2

for every t ∈ [0, T ]). Indeed, for such n, assuming for the sake of contradiction that max{|u n i (t)| : t ∈ [0, T ], i = 1, . . . , h} ≥ R 2 (i.e., u n (t) / ∈ B R

2

for at least one t ∈ [0, T ]), there exists an index j and t 1 , t 2 ∈ [0, T ] with t 1 < t 2 , such that

|u n j (t 1 )| = R 1 , |u n j (t 2 )| = R 2 , R 1 < |u n j (t)| < R 2 for every t ∈ (t 1 , t 2 ), (2.12)

|u n i (t)| ≤ R 2 for every t ∈ [0, t 2 ] and every i 6= j. (2.13) Then, for t varying from t 1 to t 2 , by (2.8) the component u n j of the solution would be driven by the curve γ j to make at least ¯ n + 1 clockwise revolutions around the origin, thus crossing at least ¯ n times the cone Θ(θ j 1 , θ j 2 ), in the clockwise sense. Writing the solution in polar coordinates

u i (t) = ρ i (t)(cos(ϑ i (t)), sin(ϑ i (t)) ), (2.14) from (2.9) we have that, if θ 1 j ≤ ϑ n j (t) ≤ θ 2 j , then

−(ϑ n j ) 0 (t) = hJ f j n (t, u n (t)) , u n j (t)i

|u n j (t)| 2 ≤ C j + 1.

(7)

So, the time to cross the cone Θ(θ 1 j , θ 2 j ) in the clockwise sense is at least (θ 2 j − θ j 1 )/(C j + 1), and then, by (2.7), the time to cross it ¯ n times should be greater than T . Hence, t 2 − t 1 > T , which is impossible.

The Poincar´ e map associated to (2.10) is then well defined on B R

1

. Let us now see that Theorem 2.1 can be applied for every n large enough, up to a subsequence, taking as A the set B R

1

, in order to find a periodic solution to the equation (2.10).

Assume for the sake of contradiction that, for every n large enough, there exist µ n > 1 and ¯ u n ∈ ∂ B R

1

with

¯

u n = (¯ u n 1 , . . . , ¯ u n i

n

, . . . , ¯ u n h ) ∈ B 2 R

1

× · · · × ∂B R 2

1

× · · · × B 2 R

1

(2.15) for a suitable i n , such that the solution u n of (2.10) with u n (0) = ¯ u n satisfies u n i

n

(T ) = µ n u n i

n

(0). We claim that, for n large enough, it has to be

R < |u n i

n

(t)| < R 2 , for every t ∈ [0, T ]. (2.16) Indeed, we already proved above that max{|u n i (t)| : t ∈ [0, T ], i = 1, . . . , h} <

R 2 . Assume for the sake of contradiction that min{|u n i

n

(t)| : t ∈ [0, T ]} ≤ R.

Then, since |u n i

n

(T )| > R 1 , there would be ˆ t 1 and ˆ t 2 in [0, T ], with ˆ t 1 < ˆ t 2 , such that |u n i

n

(ˆ t 1 )| = R, |u n i

n

(ˆ t 2 )| = R 1 , and R < |u n i

n

(t)| < R 1 , for every t ∈ (ˆ t 1 , ˆ t 2 ). Then, for t varying from ˆ t 1 to ˆ t 2 , by (2.8) the component u n i

n

of the solution would be driven by the curve γ i

n

to make at least ¯ n + 1 clock- wise revolutions around the origin, thus crossing at least ¯ n times the cone Θ(θ i 1

n

, θ i 2

n

), in the clockwise sense. Arguing as above, we see that ˆ t 2 − ˆ t 1 > T , which is impossible. By (2.16), necessarily it has to be 1 < µ n < R R

2

1

, so, up to subsequences, we can assume that

i n ≡ ι, µ n → ¯ µ ∈ [1, R R

2

1

], and u ¯ n → ¯ u ∈ ∂B R

1

.

Moreover, since ( F n ) n converges to F uniformly in [0, T ] × B R

2

, there is a constant M > 0 such that

| F n (t, u)| ≤ M, for every n ∈ N, t ∈ [0, T ], and u ∈ B R

2

.

By (2.11), u n (t) ∈ B R

2

for every t ∈ [0, T ], so (u n ) n is bounded in C 1 ([0, T ]) and, by the Ascoli–Arzel` a theorem, there is a continuous function u : [0, T ] → R 2h such that, up to a subsequence, u n → u uniformly. Passing to the limit in

u n (t) = ¯ u n + Z t

0

F n (τ, u n (τ )) dτ, we obtain

u(t) = ¯ u + Z t

0

F(τ, u(τ)) dτ,

(8)

Alessandro Fonda and Andrea Sfecci

so that u is a solution to the equation (2.1) with initial value u(0) = ¯ u ∈

∂ B R

1

. By (2.16),

R ≤ |u ι (t)| ≤ R 2 , for every t ∈ [0, T ], (2.17) and u ι (T ) = ¯ µu ι (0). Hence, |u ι (T )| ≥ |u ι (0)| and, using polar coordinates as in (2.14), there is an integer k such that

ϑ ι (T ) = ϑ ι (0) − 2πk.

By the angular velocity formula,

−(ϑ ι ) 0 (t) = hJ u 0 ι (t), u ι (t)i

|u ι (t)| 2 = hJ f ι (t, u(t)), u ι (t)i

|u ι (t)| 2 ;

as a consequence of (H2 ι ) we have that k ≤ −1. Taking into account (2.17) and the fact that Ω ι 0 ⊆ B R 2 , let ¯ m ∈ Z be such that

ι (−ϑ ι (0) + 2π( ¯ m − 1))| < |u ι (0)| ≤ |γ ι (−ϑ ι (0) + 2π ¯ m)|.

(Recall that γ ι is parametrized in clockwise polar coordinates.) Then, by the admissibility of the curve γ ι and (2.17), since B 2 R contains Ω 0 ι , it has to be that

|u ι (t)| < |γ ι (−ϑ ι (t) + 2π ¯ m)|, for every t ∈ ]0, T ].

So, using (2.6),

|u ι (T )| < |γ ι (−ϑ ι (T ) + 2π ¯ m)| = |γ ι (−ϑ ι (0) + 2π( ¯ m + k))|

≤ |γ ι (−ϑ ι (0) + 2π( ¯ m − 1))| < |u ι (0)|,

and we get a contradiction with the fact that |u ι (T )| ≥ |u ι (0)|.

So, up to a subsequence, for every ¯ u n ∈ ∂ B R

1

(associated, as in (2.15), with an index i n such that |¯ u n i

n

| = R 1 ), the solution u n of (2.10) with u n (0) = ¯ u n is such that u n i

n

(T ) 6= µ¯ u n i

n

, for every µ > 1. We can then apply Theorem 2.1 to find a T -periodic solution v n (t) of (2.10), for n large enough, up to a subsequence, starting from a point ¯ v n ∈ B R

1

. Using the Ascoli–Arzel` a theorem again, we find that, up to a subsequence, (v n ) n con- verges to a T -periodic solution of equation (2.1). 

3. Some applicative conditions

In this section we introduce three other hypotheses which are useful to

obtain (H1 i ), (H2 i ), and (H3 i ). These conditions are simply the general-

izations to the 2h-dimensional case of the conditions (H4), (H5), and (H6)

introduced in [9].

(9)

(H4 i ) There exist R > 0 and η > 0 such that, for every v ∈ R 2h ,

|v i | ≥ R ⇒ hJ f i (t, v), v i i ≥ η|v i | 2 , for every t ∈ [0, T ].

(H5 i ) There exists a continuous function χ : [0, +∞) → (0, +∞) such that hf i (t, v), v i i ≤ χ(|v i |), for every t ∈ [0, T ] and v ∈ R 2h , (3.1) and

Z +∞

0

r dr

χ(r) = +∞.

(H6 i ) There exist some values w 1 , . . . , w m ∈ S 1 and two positive functions ψ 1 , ψ 2 : S 1 \ {w 1 , . . . , w m } → (0, +∞],

not identically equal to +∞, with the following properties:

(i) in each open arc of the domain these functions are either con- tinuous and bounded with all values in R, or identically equal to +∞;

(ii) one has ψ 1 (w) ≤ lim inf

α→+∞

D J f i (t, u 1 , . . . , u i−1 , αw, u i+1 , . . . , u h )

α , w

E

(3.2)

≤ lim sup

α→+∞

D J f i (t, u 1 , . . . , u i−1 , αw, u i+1 , . . . , u h )

α , w E

≤ ψ 2 (w), uniformly for t ∈ [0, T ], (u 1 , . . . , u i−1 , u i+1 , . . . , u h ) ∈ R 2h−2 , and w in any compact subset of S 1 \ {w 1 , . . . , w n };

(iii) moreover, h Z

0

dθ ψ 2 (e ) ,

Z

0

dθ ψ 1 (e )

i ∩ n T

N : N ∈ N 0

o

= Ø, (3.3) where N 0 denotes the set of positive integers.

Notice that, in (3.3), we use the convention that +∞ 1 = 0, and implic- itly assume that the integrals have finite values. In [9], the following two propositions have been proved.

Proposition 3.1. If (H4 i ) and (H5 i ) hold, then (H1 i ) is satisfied.

Proposition 3.2. If (H4 i ) and (H6 i ) hold, then (H2 i ) and (H3 i ) are satis- fied.

We will see below how condition (H5 i ) can be weakened when dealing

with different regions in the phase plane.

(10)

Alessandro Fonda and Andrea Sfecci

4. Applications

In this section we are going to prove the existence of a T -periodic solution to the following system:

 

 

 

 

J u 0 1 = g 1 (t, u 1 ) + r 1 (t, u 1 , . . . , u h ) J u 0 2 = g 2 (t, u 2 ) + r 2 (t, u 1 , . . . , u h )

.. .

J u 0 h = g h (t, u h ) + r h (t, u 1 , . . . , u h ).

(4.1)

We assume that, for every i ∈ {1, . . . , h}, writing u i = (x i , y i ), the i-th equation of the system has the following form:

( −y 0 i = g i,1 (t, x i ) + r i,1 (t, x 1 , y 1 , . . . , x h , y h )

x 0 i = g i,2 (t, y i ) + r i,2 (t, x 1 , y 1 , . . . , x h , y h ), (4.2) where the functions r i,j : R×R 2h → R are continuous, and T -periodic in their first variable. Moreover, we assume that there exist functions p i,j : R 2 → R such that

|r i,j (t, u 1 , . . . , u h )| ≤ p i,j (u i ), with lim

|u

i

|→+∞

p i,j (u i )

|u i | = 0.

The functions g i,j : R × R → R are continuous and T -periodic in their first variable. We assume that, for j, k ∈ {1, 2} and i ∈ {1, . . . , h}, there are some constants µ i j,k , ν j,k i ∈ (0, +∞] such that

µ i j,1 ≤ lim inf

ξ→+∞

g i,j (t, ξ)

ξ ≤ lim sup

ξ→+∞

g i,j (t, ξ)

ξ ≤ µ i j,2 , (4.3) ν j,1 i ≤ lim inf

ξ→−∞

g i,j (t, ξ)

ξ ≤ lim sup

ξ→−∞

g i,j (t, ξ)

ξ ≤ ν j,2 i . (4.4) With the usual convention that +∞ 1 = 0, let

τ k i = π 2

 1

q

µ i 1,k µ i 2,k

+ 1

q

ν 1,k i µ i 2,k

+ 1

q

ν 1,k i ν 2,k i

+ 1

q

µ i 1,k ν 2,k i



, (4.5)

for i ∈ {1, . . . , h} and k ∈ {1, 2}.

Theorem 4.1. Assume that all the constants in (4.3) and (4.4) are finite, and

2 i , τ 1 i ] ∩ n

T

N : N ∈ N 0

o

= Ø, for every i ∈ {1, . . . , h}. (4.6)

(11)

Then, system (4.1) has a T -periodic solution. The same is true if, for one or more indices i, one of the constants µ i 1,2 , ν 1,2 i , µ i 2,2 , and ν 2,2 i is equal to +∞, the three others being finite.

The procedure to verify that (H1 i ), (H2 i ), and (H3 i ) hold for every i is independent of the index i, so in the following we will consider the case i = 1 and, to simplify the notation, we will often write u 1 = (x, y) instead of (x 1 , y 1 ), and λ = (x 2 , y 2 , . . . , x h , y h ). Hence we have

J f 1 (t, x, y, λ) = g 1 (t, x, y) + r 1 (t, x, y, λ).

In the proof we will need the following refinement of Proposition 3.1, where we focus our attention on the admissibility of a regular spiral on a particular subset U of R 2 .

Lemma 4.2. Given a positive constant η, a point P 0 ∈ R 2 , and a continuous function χ : [0, +∞) → (0, +∞) such that

Z +∞

0

r dr

χ(r) = +∞,

it is possible to build a clockwise rotating regular spiral ˜ γ, passing through P 0 , such that s 7→ |˜ γ(s)| is strictly increasing, which is 1-admissible for system (4.1) in any set U where

(1) hJ f 1 (t, x, y, λ), (x, y)i ≥ η(x 2 + y 2 ), (2) hf 1 (t, x, y, λ), (x, y)i ≤ χ( p

x 2 + y 2 ), for every t ∈ [0, T ], (x, y) ∈ U , and λ ∈ R 2h−2 .

Proof. Let P 0 = r 0 e −is

0

. We define the curve ˜ γ : [0, +∞) → R 2 as

˜

γ(s) = r(s)(cos s, − sin s), where r(s) is the solution of the Cauchy problem

˙r = 2 η

χ(r)

r , r(s 0 ) = r 0 .

Notice that γ(s 0 ) = P 0 . Since this curve is smooth, the sequence (σ k ) k , in this case, is arbitrary. Clearly, (2.2) and (2.4) hold, since s is the angle in clockwise polar coordinates. We see that r(s) is strictly increasing, and remains bounded for s bounded. Moreover, r(s) → +∞ for s → +∞, so that condition (2.3) is satisfied as well. We compute:

hJ ˙˜ γ(s), f 1 (t, ˜ γ(s), λ)i = ˙r(s)

r(s) hJ ˜ γ(s), f 1 (t, ˜ γ(s), λ)i + h˜ γ(s), f 1 (t, ˜ γ(s), λ)i.

(12)

Alessandro Fonda and Andrea Sfecci

Using the assumptions, we have that, for ˜ γ(s) ∈ U ,

hJ ˙˜ γ(s), f 1 (t, ˜ γ(s), λ)i ≤ −η ˙r(s)r(s) + χ(r(s)) < 0,

for every t ∈ [0, T ] and λ ∈ R 2h−2 , thus completing the proof.  Proof of Theorem 4.1. It is easy to see that the functions ψ 1 and ψ 2 which are involved in (H6 1 ) are

ψ 1 (e ) =

 

 

 

 

µ 1 1,1 cos 2 θ + µ 1 2,1 sin 2 θ, if θ ∈ (0, π 2 ), ν 1,1 1 cos 2 θ + µ 1 2,1 sin 2 θ, if θ ∈ ( π 2 , π), ν 1,1 1 cos 2 θ + ν 2,1 1 sin 2 θ, if θ ∈ (π, 2 ), µ 1 1,1 cos 2 θ + ν 2,1 1 sin 2 θ, if θ ∈ ( 2 , 2π), and

ψ 2 (e ) =

 

 

 

 

µ 1 1,2 cos 2 θ + µ 1 2,2 sin 2 θ, if θ ∈ (0, π 2 ), ν 1,2 1 cos 2 θ + µ 1 2,2 sin 2 θ, if θ ∈ ( π 2 , π), ν 1,2 1 cos 2 θ + ν 2,2 1 sin 2 θ, if θ ∈ (π, 2 ), µ 1 1,2 cos 2 θ + ν 2,2 1 sin 2 θ, if θ ∈ ( 2 , 2π).

Since all the constants in (4.3) and (4.4) are strictly positive, (H4 1 ) holds.

Solving the integral, it is easy to see that also (H6 1 ) holds (see [8] for com- putations). So, by Proposition 3.2, conditions (H2 1 ) and (H3 1 ) hold.

If all the constants µ 1 1,2 , ν 1,2 1 , µ 1 2,2 , and ν 2,2 1 are finite, the nonlinearity has an at-most linear growth, so (H5 1 ) holds with χ(r) = ar 2 + b for some suitable constants a and b. By Proposition 3.1, condition (H1 1 ) is satisfied, and the proof is completed in this case, by Theorem 2.4.

We now consider the case in which one of these constants is equal to +∞.

For example, we assume µ 1 1,2 = +∞.

In order to build an admissible spiral γ 1 in this case, we will glue together pieces of curves belonging to some regions of the plane. By construction the curve will pass through some points P α , whose distance from the origin gradually increases, giving to it the shape of a regular spiral. In what follows, we will sometimes use Lemma 4.2, whose condition 1 is satisfied by (H4 1 ), so that we will only need to find a suitable function χ in order to apply it.

Let  > 0 be fixed, in such a way that  < 1 8 min µ 1 1,1 , ν 1,1 1 , µ 1 2,1 , ν 2,1 1 . Then, there exists R > 0 such that, for every (x, y) for which |x| ≥ R and

|y| ≥ R,

|r i,j (t, x, y, λ)| ≤ p 1,j (x, y) ≤ ( |x| + |y| ), j = 1, 2.

We can assume R > 0 large enough to have

x ≥ R ⇒ 0 < µ 1 1,1 x ≤ g 1,1 (t, x),

(13)

x ≤ −R ⇒ ν 1,2 1 x ≤ g 1,1 (t, x) ≤ ν 1,1 1 x < 0, y ≥ R ⇒ 0 < µ 1 2,1 y ≤ g 1,2 (t, y) ≤ µ 1 2,2 y, y ≤ −R ⇒ ν 2,2 1 y ≤ g 1,2 (t, y) ≤ ν 2,1 1 y < 0,

slightly modifying these constants, if necessary, without affecting (4.6).

Moreover, we have the existence of a constant C > 0 such that

|x| ≤ R ⇒ |g 1,1 (t, x)| ≤ C,

|y| ≤ R ⇒ |g 1,2 (t, y)| ≤ C.

We consider five different regions in the phase plane (see Figure 1):

W = (−∞, R] × R, N E = [R, +∞) × [R, +∞), E = [R, +∞) × [−R, R],

ESE = [R, +∞) × (−∞, −R] ∩ {(x, y) : x ≥ −y}, SSE = [R, +∞) × (−∞, −R] ∩ {(x, y) : x ≤ −y}.

Figure 1. The regions in the phase plane.

The regular spiral γ 1 will be constructed by gluing together pieces of curves belonging to each of these regions.

Region W. We note that, in this region,

|g 1,2 (t, y)| ≤ C + max{µ 1 2,1 , µ 1 2,2 , ν 2,1 1 , ν 2,2 1 }|y|,

|g 1,1 (t, x)| ≤ C + max{ν 1,2 1 , ν 1,1 1 }|x|,

(14)

Alessandro Fonda and Andrea Sfecci

giving us, for every (x, y) ∈ W , f 1 (t, x, y, λ)(x, y)

= g 1,2 (t, y)x − g 1,1 (t, x)y + r 1,2 (t, x, y, λ)x − r 1,1 (t, x, y, λ)y

≤ C 1 (x 2 + y 2 ) + C 2 ,

for some suitable constants C 1 and C 2 . Fix a point P 0 = (R, y 0 ) with y 0 < −R. By Lemma 4.2, taking U = W , we can build the spiral ˜ γ which passes through P 0 = ˜ γ(s 0 ). There exists s 1 > s 0 such that ˜ γ([s 0 , s 1 ]) ⊂ W and ˜ γ(s 1 ) = P 1 = (R, y 1 ) with y 1 > R. The spiral γ 1 in W consists of the branch of ˜ γ which goes from P 0 to P 1 , and it is abmissible in W by construction.

Region NE. We have, for every (x, y) ∈ N E, f 1 (t, x, y, λ)(x, y)

= g 1,2 (t, y)x − g 1,1 (t, x)y + r 1,2 (t, x, y, λ)x − r 1,1 (t, x, y, λ)y

≤ µ 1 2,2 xy + 0 + (x + y) 2 ≤ M (x 2 + y 2 )

for a suitable constant M . Similarly to what has been done in the region W , applying Lemma 4.2 with U = N E, we can construct γ 1 going from P 1 to a point P 2 = (x 2 , R) with x 2 > R.

Region E. In this region, we construct the spiral γ 1 as a line y = −mx where 0 < m < 1 is sufficiently small. We recall that, here,

|g 1,2 (t, y)| ≤ C, g 1,1 (t, x) ≥ µ 1 1,1 x, |r 1,i (t, x, y, λ)| ≤ (|x|+|y|) ≤ 2x, i = 1, 2.

Hence, we have

J ˙γ 1 (s)f 1 (t, γ 1 (s), λ) = (m, 1)(x 0 , y 0 )

= m(g 1,2 (t, y) + r 1,2 (t, x, y, λ)) − (g 1,1 (t, x) + r 1,1 (t, x, y, λ))

≤ mC + 2mx − µ 1 1,1 x + 2x

≤ mC − (µ 1 1,1 − 4)x ≤ mC − 1 2 µ 1 1,1 R,

which is negative choosing m < µ 1 1,1 R/2C. In this way we build a branch of the spiral γ 1 which goes from P 2 to a point P 3 = (x 3 , −R), with x 3 > x 2 > R.

Region ESE. In this region the spiral γ 1 simply coincides with the line y = −2(x − x 3 ) − R. Let P 4 = (x 4 , −x 4 ) be the intersection between this line and the line y = −x. We recall that, here,

g 1,2 (t, y) ≤ ν 2,1 1 y < 0, g 1,1 (t, x) ≥ µ 1 1,1 x,

|r 1,i (t, x, y, λ)| ≤ (|x| + |y|) ≤ 2 x, i = 1, 2.

(15)

So we have

J ˙γ 1 (s)f 1 (t, γ 1 (s), λ) = (2, 1)(x 0 , y 0 )

= 2(g 1,2 (t, y) + r 1,2 (t, x, y, λ)) − (g 1,1 (t, x) + r 1,1 (t, x, y, λ))

≤ 2(0 + 2 x) − (µ 1 1,1 x − 2 x) = −(µ 1 1,1 − 6) x < 0.

Region SSE. In this region the following inequalities hold for i = 1, 2:

g 1,2 (t, y) ≤ ν 2,1 1 y < 0, g 1,1 (t, x) ≥ µ 1 1,1 x, |r 1,i (t, x, y, λ)| ≤ (x − y) ≤ −2y.

At first, we note that for a solution of (4.1), with the above notation, as long as (x(t), y(t)) belongs to this region, we have that

x 0 (t) = g 1,2 (t, y(t)) + r 1,2 (t, x(t), y(t), λ) ≤ (ν 2,1 1 − 2) y(t) < 0. (4.7) We have to build the spiral γ 1 starting from the point P 4 = (x 4 , −x 4 ). Call SSE good the region SSE ∩ {x ≤ x 4 } and SSE bad the region SSE ∩ {x > x 4 } (see Figure 2).

Figure 2. The construction of the curve in the region SSE.

Note that, for every (x, y) ∈ SSE good , f 1 (t, x, y, λ)(x, y)

= g 1,2 (t, y)x − g 1,1 (t, x)y + r 1,2 (t, x, y, λ)x − r 1,1 (t, x, y, λ)y

≤ 0 − M 1 y + (x − y) 2 ≤ M 2 (x 2 + y 2 ) + M 1 ,

for some suitable constants M 1 and M 2 . Setting U = SSE good , it is possible

to apply Lemma 4.2 to obtain a spiral ˜ γ which links P 4 to a point P 6 =

(R, y 6 ), with y 6 < −R, passing through SSE. By construction, this spiral

(16)

Alessandro Fonda and Andrea Sfecci

is 1-admissible only in SSE good . Nothing tells us that ˜ γ does not enter in the region SSE bad , but there exists a point P 5 = (x 5 , y 5 ) with x 5 = x 4 and y 5 ≤ y 4 (possibly P 5 = P 4 ) on the curve ˜ γ after which ˜ γ is contained in SSE good . Using (4.7), we choose the spiral γ 1 to be made of the vertical line linking P 4 and P 5 and of the branch of ˜ γ linking P 5 and P 6 .

With such a procedure we have constructed the first lap of the spiral γ 1 . In the same way we can obtain the other ones. Such a spiral is 1-admissible in the whole plane by construction. So, (H1 1 ) holds, and the proof is completed

in this case, too. 

5. An example: coupled oscillators

As a particular case of (4.1), we have the following system of coupled oscillators:

 

 

 

 

x 00 1 + φ 1 (t, x 1 ) = e 1 (t, x 1 , . . . , x h ) x 00 2 + φ 2 (t, x 2 ) = e 2 (t, x 1 , . . . , x h )

.. .

x 00 h + φ h (t, x h ) = e h (t, x 1 , . . . , x h ).

(5.1)

Here we assume that, for every i,

|e i (t, x 1 , . . . , x h )| ≤ p i (x i ), with lim

|x

i

|→+∞

p i (x i )

|x i | = 0, and that the function φ i satisfies

µ i 1 ≤ lim inf

ξ→+∞

φ i (t, ξ)

ξ ≤ lim sup

ξ→+∞

φ i (t, ξ)

ξ ≤ µ i 2 , (5.2) ν 1 i ≤ lim inf

ξ→−∞

φ i (t, ξ)

ξ ≤ lim sup

ξ→−∞

φ i (t, ξ)

ξ ≤ ν 2 i , (5.3) for some suitable constants in (0, +∞]. With the usual convention that

1

+∞ = 0, let

τ 1 i = π q

µ i 1

+ π

q ν 1 i

, τ 2 i = π q

µ i 2

+ π

q ν 2 i

, (5.4)

for i ∈ {1, . . . , h}. As an immediate consequence of Theorem 2.4, we have the following.

Corollary 5.1. Assume that all the constants in (5.2) and (5.3) are finite, and

2 i , τ 1 i ] ∩  T

N : N ∈ N 0 = Ø, for every i ∈ {1, . . . , h}. (5.5)

(17)

Then, system (5.1) has a T -periodic solution. The same is true if, for one or more indices i, one of the constants µ i 2 and ν 2 i is equal to +∞, the other being finite.

Let us show an example of a situation which permits us to apply the pre- vious corollary. We will use the following notation: for every ξ ∈ R we write ξ + = max{ξ, 0} and ξ = max{−ξ, 0}, and for every x = (x 1 , x 2 , . . . , x h ) ∈ R h we write x + = (x + 1 , x + 2 , . . . , x + h ), x = (x 1 , x 2 , . . . , x h ), and exp(x) = (e x

1

, e x

2

, . . . , e x

h

). Consider the equation in R h

x 00 − B(t) arctan(kxk) x + exp(x + ) = a(t), (5.6) where kxk is a norm in R h , a : R → R h is a T -periodic continuous function, and B(t) = diag(b 1 (t), . . . , b h (t)) is a diagonal matrix where each b i : R → R is continuous and T -periodic. We assume that there exist some positive integers N i , and a constant δ > 0, such that

1

2π (λ N

i

+ δ) < b i (t) < 1N

i

+1 − δ),

where λ k = (2πk/T ) 2 is the k-th eigenvalue for the T -periodic problem. We can see that this is a particular case of system (5.1), with

φ i (t, x i ) = − π

2 b i (t) x i + exp(x + i ),

e i (t, x 1 , . . . , x h ) = arctan(kxk) − π 2  b i (t) x i + a i (t), and we find the values

µ i 2 = +∞, ν 1 i = 1 4N

i

+ δ), and ν 2 i = 1 4N

i

+1 − δ), for every i. Thus, choosing µ i 1 large enough, we have

T

N i + 1 < τ 2 i = 2π

N

i

+1 − δ < τ 1 i = 2π

N

i

+ δ + π q

µ i 1

< T N i

,

for every i. Corollary 5.1 can thus be applied, so that equation (5.6) has a T -periodic solution.

Remark 5.2. We have focused our attention on one particular situation

where our conditions (H1 i ), (H2 i ) and (H3 i ) hold, for every i. As we have

shown in [9], many other different cases can be treated with the same ap-

proach, like, e.g., nonlinearities controlled by positively homogeneous Hamil-

tonian functions, Landesman–Lazer situations at resonance, and nonlinear-

ities with a singularity. Our theorem permits us to mix together all these

situations. For example, one could think about a system in R 6 with a one-

sided superlinearity in the first couple of variables, a resonance case with a

(18)

Alessandro Fonda and Andrea Sfecci

Landesman–Lazer condition in the second one, and a singularity in the last one.

Remark 5.3. One can extend our results to the case when the phase space R 2h is replaced by a space of the type R 2h+k , for some k ≥ 1, introducing some hypotheses on the last k coordinates. For example, one could think of some kind of dissipative situation, so that a variant of Theorem 2.1 will be applicable. For briefness, we will not enter into details here.

References

[1] A.A. Andronov, A.A. Vitt, and S.E. Khaikin, “Theory of Oscillators,” reprint of the 1966 translation, Dover Publications, New York, 1987.

[2] G. Arioli and B. Ruf, Periodic solutions for a system of forced and nonlinearly coupled oscillators with applications to electrical circuits, Dynam. Systems Appl., 4 (1995), 87–102.

[3] A. Capietto, J. Mawhin, and F. Zanolin, A continuation approach to superlinear periodic boundary value problems, J. Differential Equation, 88 (1990), 347–395.

[4] A. Castro and A.C. Lazer, On periodic solutions of weakly coupled systems of differ- ential equations, Boll. Un. Mat. I/d., B (5) 18 (1981), 733–742.

[5] C. Chicone, Periodic solutions of coupled oscillators near resonance, SIAM J. Math.

Anal., 26 (1995), 1257–1283.

[6] T.R. Ding, Some fixed point theorems and periodically perturbed nondissipative sys- tems, Chinese Ann. Math., 2 (1981), 281–300.

[7] C. Fabry and P. Habets, Periodic solutions of second order differential equations with superlinear asymmetric nonlinearities, Arch. Math., 60 (1993), 266–276.

[8] A. Fonda and J. Mawhin, Planar differential systems at resonance, Adv. Differential Equations, 11 (2006), 1111–1133.

[9] A. Fonda and A. Sfecci, A general method for the existence of periodic solutions of differential systems in the plane, J. Differential Equations, 252 (2012), 1369-1391.

[10] E. Gaines and J. Mawhin, “Coincidence Degree and Nonlinear Differential Equations,”

Lect. Notes Math. 568, Springer, Berlin, 1977.

[11] E.A. Jackson, Nonlinear coupled oscillators I. Perturbation theory; ergodic problem, J. Math. Phys., 4 (1963), 551–558.

[12] J. Mawhin and J.R. Ward, Periodic solutions of some forced Li´ enard differential equa- tions at resonance, Arch. Math., 41 (1983), 337–351.

[13] J. Mawhin and M. Willem, “Critical Point Theory and Hamiltonian Systems. Applied Mathematical Sciences,” 74, Springer, Berlin, 1989.

[14] R.E. Mickens, A regular perturbation technique for nonlinearly coupled oscillators in resonance, J. Sound Vibration, 81 (1982), 307–310.

[15] M. Struwe, “Variational Methods. Applications to Nonlinear Partial Differential Equa-

tions and Hamiltonian Systems,” fourth edition, Springer, Berlin, 2008.

Riferimenti

Documenti correlati

Definition (the preorder induced by a separable capitalization factor). Let ≥ fe the usual preorder of the financial events. We shall give two proofs: algebraic and geometrical. the

For example, in addition to the influence of 2,3-DPG (2~-diphosphogly- cerate) on oxygen affinity due to its allosteric effect, it can alter red cell pH (pH,) by its effect on

We allow for discontinuous coefficients and random initial condition and, under suitable assumptions, we prove that in the limit as the number of particles grows to infinity

images inaZude the point, i.e.. - Bv Prooosition 2, we have only to Drove that an o-regular function is c. o-regular iff ii) is true.. Then let f be a

into a product of an orthogonal matrix Q 1 and a positive (i.e., with all positive entries along the diagonal) upper triangular matrix R 1 by using the Gram–Schmidt

We hold these truths to be self-evident, that all men are created equal, that they are endowed by their creator with certain inalienable rights, that among these are life, liberty,

In this paper we show that this condition (Assumption 1 below) is necessary and sufficient in order that, for any smooth boundary datum φ and for any bounded Ω with smooth boundary,

As two graded bundle structure on the same manifold are just two homogeneity structures, the obvious concept of compatibility leads to the following:..