ArticlePDF Available

Characterization of isochronous foci for planar analytic differential systems

Authors:

Abstract

We consider the two-dimensional autonomous systems of differential equations of the form x = lambda x - y + P(x, y), y = x + lambda y + Q(x, y), where P(x, y) and Q(x, y) are analytic functions of order greater than or equal to 2. These systems have a focus at the origin if lambda not equal 0, and have either a centre or a weal: focus if lambda = 0. In this work we study the necessary and sufficient conditions for the existence of an isochronous critical point at the origin. Our result is, to the best of our knowledge, original when applied to weak foci and gives known results when applied to strong foci or to centres.
Characterization of isochronous foci for
planar analytic differential systems
Jaume Gin´
e & Maite Grau
Departament de Matem`atica. Universitat de Lleida.
Avda. Jaume II, 69. 25001 Lleida, SPAIN.
E–mails: gine@eps.udl.es and mtgrau@matematica.udl.es
Abstract
We consider the two-dimensional autonomous systems of differential equations
of the form:
˙x=λx y+P(x, y),˙y=x+λy +Q(x, y),
where P(x, y) and Q(x, y) are analytic functions of order 2. These systems
have a focus at the origin if λ6= 0, and have either a center or a weak focus if
λ= 0. In this work we study necessary and sufficient conditions for the existence
of an isochronous critical point at the origin. Our result is original when applied
to weak foci and gives known results when applied to strong foci or to centers.
Keywords: nonlinear differential equations, isochronicity.
AMS classification: Primary 34C05; Secondary 34C23, 37G15.
1 Introduction
Let us consider an autonomous differential system:
˙x=λx y+P(x, y),˙y=x+λy +Q(x, y),(1.1)
where P(x, y) and Q(x, y) are analytic functions in a neighborhood Uof the origin O
and of order greater or equal than two. We assume that Ois an isolated singular point
of (1.1). We denote by Xthe equivalent vector field:
X= (λx y+P(x, y))
∂x + (x+λy +Q(x, y ))
∂y .
An isolated singular point of (1.1) is said to be a focus if it has a neighborhood where
all the orbits spiral in forward or backward time. An isolated singular point of (1.1) is
said to be a center if it has a punctured neighborhood filled of periodic orbits. System
(1.1) has a strong focus at the origin if λ6= 0 and it has a weak focus or a center at
The authors are partially supported by a MCYT grant number BFM 2002-04236-C02-01. The first
author is partially supported by DURSI of Government of Catalonia “Distinci´o de la Generalitat de
Catalunya per a la promoci´o de la recerca universit`aria. The second author is also partially supported
by a FPU grant with reference AP2000-3585.
1
the origin if λ= 0.
A main problem is that of studying the existence and properties of periodic solu-
tions in a neighborhood of the origin of (1.1). In this field, different methods have
been used to study isolated periodic solutions, i.e. limit cycles, or non-isolated ones,
i.e. period annulus. The stability of the singular point Odoes not imply the stability
of the cycles close to the singular point. In fact, a non-isolated cycle is Liapunov stable
if and only if every neighboring cycle has the same period. This fact motivates the def-
inition of isochronicity. We give a precise definition of isochronicity in the forthcoming
paragraph. Isochronicity has been widely studied not only for its physical meaning and
for its role in stability theory, but also for its relationship with bifurcation problems
and to boundary value problems.
An essential tool to study the stability of the origin of system (1.1) is the Poincar´e
map, see [14, 15]. Let us consider a neighborhood Uof the origin and let Σ be a section
of system (1.1) through the origin, that is, a transversal curve through the origin for
the flow of system (1.1).
More precisely, we define a section through the origin as a simple arc without contact
with the origin Oas an endpoint. See the book of Andronov et al. [1], page 55, for a
precise definition of simple arc without contact. We also need some assumptions on its
regularity for technical reasons. Given a section through the origin Σ R2, we consider
a parameterization c:RR2such that Σ = {c(σ)|σR}and limσ→−∞ c(σ) = O.
We assume that c(σ) is analytic for all σR.
For each point pΣ, the flow of system (1.1) through pwill cross Σ again at a
point P(p)Σ near p. The map p7→ P (p) is called the Poincar´e map. If we denote by
Φt(p) the flow of system (1.1) with the initial condition Φ0(p) = p, we can define the
Poincar´e map in the following way. Given pΣ, there is a unique analytic function
τ(p) such that Φτ(p)(p)Σ and Φt(p)6∈ Σ for any 0 < t < τ(p), see [14]. In these
terms, we have P(p)=Φτ(p)(p). We remark that both functions Pand τdepend on
the chosen section Σ. The function τ: Σ R+is called the period function. As usual
R+denotes the set of positive real numbers. In this paper we study the existence of
a section Σ such that τ: Σ R+is constant. When such a Σ exists, we say that the
origin Oof (1.1) is isochronous and that Σ is an isochronous section.
We will call a center an analytic system of the form (1.1) with λ= 0 and where
the origin Ois a center. Isochronicity has been widely studied for centers, see for
instance [8] and the references therein. We remark that the period function of a center
does not depend on the chosen section Σ. The main methods used in order to study
isochronicity of centers can be roughly classified in two categories, linearization and
commutation.
Finding a linearization for a center Xmeans finding a transformation φ:U → U
analytic in a neighborhood of the origin such that (O) = I, where I denotes the
2×2 identity matrix, such that the transformed system is a linear center, that is
φ(X) = y ∂/∂x +x ∂/∂ y. If such a transformation exists, then all the orbits have the
same period, coinciding with the period of the linear center. So, a center is isochronous
if and only if a linearization can be found.
Finding a commutator for a center Xmeans finding a second vector field Yanalytic
in a neighborhood of the origin and of the form
Y= (x+A(x, y))
∂x + (y+B(x, y))
∂y ,(1.2)
2
with Aand Banalytic functions of order 2, such that the Lie bracket [X,Y] of the
center Xand Yidentically vanishes.
An isolated singular point of a real planar analytic autonomous system is called a
star node if the linear part of the vector field at the singular point has equal non-zero
eigenvalues and it is diagonalizable. Clearly, the origin is a star node for (1.2). By an
affine change of coordinates any vector field with a star node can be brought to the
form (1.2).
Given two analytic vector fields defined in an open set U,Xand Y, we say that
they are transversal at noncritical points when Xand Yhave isolated singular points,
they both have the same critical points in U, and if p∈ U is such that X(p)6= 0 then
the function given by the wedge product of Xand Yis not zero at p. From now on,
we always assume that Xand Yare analytic vector fields defined in a neighborhood
Uof the origin and transversal at non critical points.
We will always consider analytic vector fields although many of the stated results
apply also for vector fields with weaker differentiability restrictions. The results of
Sabatini [12] go on this direction. We define a smooth function as a function of class
Cin a neighborhood Uof the origin O. Analogously, a smooth vector field is defined
by smooth functions.
The following result, proved in [2] (Theorem 2.4, page 140), characterizes centers
in terms of Lie brackets.
Theorem 1.1 [2] System (1.1) with λ= 0 has a center at the origin if, and only
if, there exists a smooth vector field Uof the form (1.2) and a smooth scalar function
ν(x, y)with ν(0,0) = 0 such that [X, U ] = νX.
The most important result on characterization of isochronous centers appears in [16,
10]. A further study can be found in, for instance, [2, 7] and the references therein. See
[3] for a constructive method of Uand νin special cases for polynomial vector fields.
The following theorem, which is stated and proved in [10] (Theorem 2, page 98),
gives the equivalence between commutation and isochronicity for centers.
Theorem 1.2 [10] Let Obe a center of system (1.1), with λ= 0. Then Ois isochro-
nous for system (1.1) if, and only if, there exists an analytic vector field Yof the form
(1.2), transversal to Xand such that [X,Y]0.
Another work on commuting systems is [11], where M. Sabatini discusses the local and
global behavior of the orbits of a pair of commuting systems and gives several illustra-
tive examples. A wide collection of commutators and linearizations can be found in [4].
When a center is isochronous, it is possible to construct an isochronous section Σ,
see for instance [12]. However, the existence of an isochronous section is not strictly
dependent on the existence of a center. A system can have a singular point of focus type
with an isochronous section. This implies the existence of a neighborhood covered with
solutions spiralling towards the singular point, all meeting Σ at equal time intervals.
Such a behavior may occur, for instance, in a pendulum with friction, or in an electric
circuit with dissipation, see also [12]. Our main result, Theorem 3.1, characterizes
when the origin of system (1.1) is isochronous, even when the origin is a center, a weak
focus or a strong focus. In this paper, we adapt the two different techniques usually
used for isochronous centers, in order to study isochronous foci.
In Section 2 we summarize the known results on isochronicity for foci. It is shown
that a strong focus of an analytic system is always isochronous. All the results described
in Section 2 only apply for systems of the form (1.1) with λ6= 0 or for centers.
3
Section 3 contains the main theorem of this work which characterizes isochronicity
for the origin of a system (1.1). Our result is original when applied to weak foci
and gives known results when applied to strong foci or to centers. We modify the
commutators’ method to study isochronous critical points. We prove that system (1.1)
has a transversal vector field Ysuch that the vector field [X,Y] is proportional to Y
if, and only if, system (1.1) has an isochronous critical point at the origin.
We give two examples of weak isochronous foci and we give an example of a family
of quadratic systems depending on a parameter wRwhich never has an isochronous
critical point at the origin. When w= 0 the system is a center and when w6= 0 the
system is a weak focus (stable if w < 0 and unstable if w > 0). Hence, we show that
there is no isochronous section for any system of this family.
2 Summary of known results
We denote by Uany open neighborhood of the origin and by ρ:U R+×Rthe change
to polar coordinates, that is, ρ(x, y )=(r, θ) with r=px2+y2and θ= arctan(y/x).
As usual, ρis the push-forward defined by ρand ρis the corresponding pull-back.
In order to give the definition of isochronous critical point, we consider the form of
(1.1) in polar coordinates, that is, ρ(X) = rf(r, θ)
∂r +g(r, θ)
∂θ , where fand gare
analytic functions in a neighborhood of ρ(O).
Definition 2.1 The critical point Oof (1.1) is said to be isochronous if there ex-
ists a local analytic change of variables φwith (O)=Iand such that ρφ(X) =
rf (r, θ)
∂r +g(θ)
∂θ .
A system (1.1) with an isochronous critical point at the origin is more easily written
using the arc–length ϕ, defined by ϕ=Rθ
0dθ/g(θ), as new angular variable. In this
formulation we end up to the following definition.
Definition 2.2 The critical point Oof (1.1) is said to be isochronous if there ex-
ists a local analytic change of variables φwith (O)=Iand such that ρφ(X) =
rf (r, θ)
∂r +k
∂θ ,kR,k6= 0.
The existence of an isochronous section is equivalent to the existence of the local ana-
lytic change of variables φ, as we will show in Theorem 3.1. We state the definition of
isochronous critical point by means of φsince this is its classical definition which let
us give the summary of known results.
Linear foci, (λx y)/∂x + (x+λy)∂/∂y, are isochronous since their angular
speed is constant along rays through the origin. For a linear focus, every ray through
the origin is an isochronous section. We say that a vector field Xof the form (1.1) is
linearizable when there exists a local change of variables φwith (O) = Isuch that
φ(X) is a linear focus.
By the above definition, every analytic linearizable focus is isochronous. If φis
the linearizing transformation and Σ is a ray, then φ1(Σ) is an isochronous section
of the analytic linearizable focus. Next theorem, which is a special case of classical
Poincar´e’s Theorem, shows that every strong focus of an analytic system is linearizable
and therefore isochronous. For a proof, see [5, 15].
Theorem 2.3 [15] Let us consider the planar real analytic system
˙x=αx βy +g1(x, y),˙y=βx +αy +g2(x, y),(2.1)
4
with αβ 6= 0, and g1and g2are of second order in xand y. Then there exists a real
local analytic change of variables φ(x, y) = (u, v)with (O)=Iwhich transforms
system (2.1) into ˙u=αu βv,˙v=βu +αv.
This result can also be stated for a system of the form (2.1) satisfying weaker differen-
tiability restrictions. Since we are only concerned with analytic vector fields, we state
the result only for the analytic case.
We have seen that every analytic linearizable focus is isochronous, but finding the
linearization, and hence the isochronous sections, is usually too difficult. Next theorem
proved in [12] shows that it is not necessary to find the explicit form of the linearization,
since the orbits of a suitable commutator are isochronous sections of X.
Theorem 2.4 [12] If the vector field Xgiven by (1.1) has a focus Oand a nontrivial
commutator Ywith a star node at O, then every orbit of Ycontained in a neighborhood
of Ois an isochronous section of X.
This result only applies when the vector field Xhas a strong focus at the origin or has
a center because if the vector field Xhas a weak focus at the origin with a nontrivial
commutator Ywith a star node at O, then by Theorem 1.1 the vector field has a center
at the origin. Next corollary proved in [12] shows that every system with a strong focus
and a nontrivial commutator has a commutator with a star node.
Corollary 2.5 [12] If the vector field Xhas eigenvalues with non-zero real part at
a focus Oand a nontrivial commutator Y, then it has infinitely many isochronous
sections.
In [9] and [12], different sufficient conditions for an analytic vector field to have an
isochronous weak focus at Oare given. In [12], the particular case of a differential
system equivalent to a Li´enard equation is taken into account.
3 Characterization of isochronous critical points
The following theorem characterizes when the origin Oof a system (1.1) has a section
Σ such that the period function τ: Σ R+is constant, that is, it does not depend
on the point pΣ considered. We will see that if such section exists, then there are
an infinite number of them. In particular next theorem characterizes the existence of
isochronous critical points.
Theorem 3.1 Let us consider an analytic system (1.1). The following statements are
equivalent:
(i) There exists an analytic change of variables φ:U → U , where Uis a neighbor-
hood of the origin, with (O)=I, such that the transformed system reads for
ρφ(X) = rf (r, θ)
∂r +g(θ)
∂θ .
(ii) There exists an analytic vector field Ydefined in a neighborhood of the origin of
the form
Y= (x+A(x, y))
∂x + (y+B(x, y))
∂y ,(3.1)
with Aand Banalytic functions of order 2, such that [X,Y] = µ(x, y)Y,
where µ(x, y)is a scalar function with µ(0,0) = 0.
(iii) There exists a section Σsuch that the period function τ: Σ R+is constant.
5
Proof of Theorem 3.1.In order to prove the equivalence of the three statements,
it suffices to show (i)(ii), (ii)(iii) and (iii)(i). We also include the proof of
(ii)(i) and (iii)(ii), for completeness.
In the subsequent, we will denote by a subindex a partial derivative, for instance,
if f(r, θ) is a function of (r, θ), ∂f
∂r is replaced by fr.
(i)(ii) We define
Y=φµx
∂x +y
∂y ,
and we have that Yhas the form described since φis an analytic change such that
(O) = I. Moreover,
[X,Y] = ·φφX, φµx
∂x +y
∂y ¶¸ =φρµ·rf (r, θ)
∂r +g(θ)
∂θ , r
∂r ¸¶
=φρµr2fr(r, θ)
∂r .
We denote by µ(x, y) = φρ(rfr(r, θ)). It is obvious that is an analytic scalar
function with µ(0,0) = 0. We have
[X,Y] = µ(x, y)φρµr
∂r =µ(x, y )φµx
∂x +y
∂y =µ(x, y )Y.
(ii)(i) From normal form theory, see [5, 15], we have that there exists an analytic
change of variables φ, defined in a neighborhood Uof the origin and with Dφ(O) = I,
such that φ(Y) = x
∂x +y
∂y .
Since [X,Y] = µY, we have [φ(X), φ(Y)] = φ(µ)φ(Y). We introduce the fol-
lowing notation ˜µ(r, θ) := ρφ(µ(x, y)) and ρφ(X) := rf (r, θ)
∂r +g(r, θ)
∂θ . Hence,
·rf (r, θ)
∂r +g(r, θ)
∂θ , r
∂r ¸= ˜µ(r, θ)r
∂r .
We compute the Lie bracket and we have the following equality
r2fr(r, θ)
∂r rgr(r, θ)
∂θ = ˜µ(r, θ)r
∂r ,
which implies gr(r, θ)0 and, therefore, g(r, θ) = g(θ). We remark that since the ori-
gin of the system defined by Xis a monodromic critical point, we have that g(θ)>0
or g(θ)<0 for all θR. Moreover, as before, we may consider the arc–length
ϕ=Rθ
0dθ/g(θ). This integral is well defined and it gives a change of variable since
g(θ) has a definite sign for all θR. Then, after this change, the angular speed of the
corresponding system is constant.
(ii)(iii) This statement is a clear corollary of Theorem 2.4. However, a geo-
metric outline of its proof is easy enough to be given here.
Let, for any p∈ U, be Φt(p) the flow of Xand Ψs(p) that of Y, with the initial
condition Φ0(p) = Ψ0(p) = p. Without lack of generality, we can assume that Ois
the unique singular point for Xand Yin U. Let p, q ∈ U ,p, q 6=O. By classical Lie
6
theory, we have that the relation [X,Y] = µ(x, y)Yimplies that if Σ = {Ψs(p)|sR}
is a solution of Y, then for any tR, Φt(Σ) is another solution for Y.
It is clear that Σ is a transversal section for X. Let τ, Pbe the corresponding period
function and Poincar´e map defined on it. We will show that any two points p, q Σ
have the same period function. We have that P(p) = Φτ(p)(p). The time τ(p) leaves Σ
invariant: Φτ(p)(Σ) Σ. Let qΣ, then there exists sRsuch that q= Ψs(p). The
minimal time to meet Σ again, that is τ(q), must coincide with τ(p) since the time
τ(p) brings the solution Σ into itself. Then τ(p) = τ(q).
(iii)(ii) We consider Φt(p) the flow of system Xdefined in the neighborhood
Uof the origin and with the initial condition Φ0(p) = p.
Given a section through the origin, Σ R2, we consider its parameterization by
its arc-parameter σ, that is, there exists a map c:RΣ such that Σ = {c(σ)|σ
R}. We can assume without loss of generality that limσ→−∞ c(σ) = Oand that
limσ→−∞ c0(σ)6= (0,0). As usual, c0(σ) denotes the derivative of the parameterization
of the curve c:σ7→ c(σ) at the value σ. We define the following set of transformations
Ψ : R× U U in the following way, see Figure 1.
If pΣ, that is p=c(σ0) for a certain σ0R, and sRthen Ψs(p) := c(σ0+s).
If p6∈ Σ, there exists t0(p)Rsuch that Φt0(p)(p)Σ, that is, there exists σ0R
such that c(σ0) = Φt0(p)(p). Assume that t0(p)>0 is the lowest positive real with this
property. For any sRwe define Ψs(p) = Φt0(p)(c(σ0+s)).
In the subsequent, for any p∈ U we denote by t0(p) as the lowest positive real
such that Φt0(p)(p)Σ. It is clear that t0:U [0, T ) where T > 0 is the period
defined by the section Σ. We denote by σ0(p)Rthe value of the parameter such
that Φt0(p)(p) = c(σ0(p)).
Figure 1: Definition of Ψs(p).
We are going to prove that the set of transformations defined by Ψsis a one–
parameter Lie group of point transformations. We need to show the following state-
ments:
(a) For all sR, Ψs:U → U is bijective.
(b) Ψ0is the identity map.
(c) For any s1, s2R, Ψs1Ψs2= Ψs1+s2.
(d) Ψ ∈ Cω(R)× Cω(U).
7
(a) Fixed sR, let us consider any p∈ U and we have Ψs(p)=Φt0(p)(c(σ0(p) +
s)). Let p1, p2∈ U . If Ψs(p1) = Ψs(p2), let qbe this point q= Ψs(pi). Then, the points
Φt0(p1)(q) = c(σ0(p1) + s) and Φt0(p2)(q) = c(σ(p2) + s) belong to Σ. Both, t0(p1) and
t0(p2) are defined as the minimum positive time with this property so, t0(p1) = t0(p2).
Therefore, c(σ0(p2) + s) = c(σ0(p1) + s) and this gives σ0(p1) = σ0(p2) which implies
p1=p2. Then, Ψsis injective.
Let us see that it is exhaustive. Given q∈ U let p= Φt0(q)(c(σ0(q)s)). Then,
t0(p) = t0(q), σ0(p) = σ0(q)sand Ψs(p)=Φt0(q)(c(σ0(q))) = q. The fact that the
section Σ is isochronous ensures the well-definition of this p.
(b) Given p∈ U we have that Ψ0(p) = Φt0(p)(c(σ0(p))) where c(σ0(p)) = Φt0(p)(p).
Then, clearly, Ψ0(p) = p.
(c) Given p∈ U, it is clear that t0t0(p)(c(σ0+s1))) = t0t0(p)(c(σ0+s1+
s2))) = t0(p). We have Ψs1Ψs2(p)=Ψs1t0(p)(c(σ0+s1))) = Φt0(p)(c(σ0+s1+
s2)) = Ψs1+s2(p).
(d) The regularity of Ψ is clear due to the regularity of Φ and c.
Once we have that Ψ is a one–parameter Lie group of point transformations, we
apply the first fundamental theorem of Lie, see [13], and we have that there exists
an analytic vector field Ywhose flow coincides with Ψs(p). Moreover, Yis given
by Ψs
∂s (p)|s=0 . By the definition Ψs(p) = Φt0(p)(c(σ0(p) + s)) we have that Y(p) =
t0(p)(c(σ0(p) + s)) ·c0(σ0(p) + s)|s=0 = TΦt0(p)(c(σ0(p))) ·c0(σ0(p)), where TΦt(q)
denotes the jacobian matrix of the analytic change of variables Φtat the point qand,
as before, c0(σ) denotes the derivative of the parameterization of the curve c:σ7→ c(σ)
at the value σ.
Moreover, by construction, Yhas a star node at the origin. This is clear by the
fact that each of its orbits Φt(Σ), t[0, T ), has a different tangent at the origin.
Let Y=ξ(x, y)
∂x +η(x, y)
∂y . Since Yhas a star node at the origin, by a classical
result stated in [17], page 63, we have that ξ(x, y) = xh(x, y) + h.o.t. and η(x, y) =
yh(x, y) + h.o.t., where h(x, y ) is a homogeneous polynomial and h.o.t. denotes higher
order terms. Therefore, in order to see that Yis of the form (1.2), we only need to
show that the divergence of the vector field Y, that is divY, is different from zero at
the origin, where divY(x, y) = ∂ξ
∂x (x, y) + η
∂y (x, y). The divergence of the vector field
Yis related to the inverse integrating factor of Y. The inverse integrating factor of
Yis given by V(x, y)=(λx y+P(x, y))η(x, y)(x+λy +Q(x, y))ξ(x, y) which is
defined in the neighborhood Uof the origin. An easy computation shows that
Vs(x0, y0)) = V(x0, y0) exp ½Zs
0
divYu(x0, y0))du¾(3.2)
for any (x0, y0)∈ U. It is clear that V(0,0) = 0. Let p0:= (x0, y0) U − {(0,0)}and
assume that V(p0) = 0. This implies that the vectors Y(p0) and X(p0) are parallel.
By the definition of Y, we have that
Y(p0) = TΦt0(p0)(c(σ0(p0))) ·c0(σ0(p0)) = TΦt0(p0)t0(p0)(p0)) · Yt0(p0)(p0)).
We denote by q0= Φt0(p0)(p0) and we have TΦt0(p0)(q0)· Y(p0) = Y(q0). Since Φ is
the flow of X, we have TΦt0(p0)(q0)· X (p0) = X(q0).Therefore, if Y(p0) and X(p0) are
parallel, then Y(q0) and X(q0) are parallel. However, q0Σ and the vector Y(q0) is
tangent to Σ at q0, so the parallelism between Y(q0) and X(q0) is a contradiction with
Σ being a transversal section for X. Therefore, we conclude that V(x0, y0)6= 0 for any
(x0, y0)∈ U − {(0,0)}.
8
By using this fact, we prove that divY(0,0) 6= 0. Let us consider (x0, y0)
U − {(0,0)}and we have that lims→−∞ Ψs(x0, y0) = (0,0). By continuity and the
identity (3.2), we have that the integral I(x0, y0) := R0
−∞ divYu(x0, y0))du di-
verges. I(x0, y0) is continuous, so I(0,0) also diverges. Hence, if divY(0,0) = 0,
then I(0,0) = R0
−∞ divYu(0,0))du =R0
−∞ divY(0,0)du = 0, in contradiction with
being divergent. Therefore, divY(0,0) 6= 0.
Moreover, by definition it is clear that the flow of Xtakes solutions of Yto solu-
tions of Y. Another classical result on Lie symmetries gives that Xis a Lie symmetry
for Yand therefore, there exists an analytic scalar function µ:U Rsuch that
[X,Y] = µY. Moreover, µ(0,0) = 0 since both functions defining the vector field
[X,Y] have order two at the origin and the vector field Yhas order one at the origin.
(i)(iii) The ray ˜
Σ = {(x, 0) |x > 0}is an isochronous section for the sys-
tem φ(X), where ρφ(X) = rf (r, θ)
∂r +g(θ)
∂θ , since ˜τ:˜
ΣR+is given by
˜τ(x) = R2π
0dθ/g(θ), which is constant for every x˜
Σ. Then, Σ := φ1(˜
Σ) is an
isochronous section for system (1.1) and the period function is given by τ:= φτ). ¥
Using Theorem 1.1 and Theorem 3.1, we reencounter the following result which
characterizes isochronous centers and which is stated and proved in [2] (Theorem 2.3,
page 140).
Theorem 3.2 System (1.1) with λ= 0 has an isochronous center at the origin if, and
only if, there exists a smooth vector field Zof the form (1.2) such that [X,Z] = 0.
Proof. Assume that system (1.1), with λ= 0, has an isochronous center at the
origin. Then by Theorem 1.1 there exists a smooth vector field Uof the form (1.2)
and a smooth function ν, with ν(0,0) = 0 such that [X,U] = νX. Moreover, by
Theorem 3.1 there exists an analytic vector field Yof the form (1.2) and an analytic
functions µ, with µ(0,0) = 0, satisfying [X,Y] = µY. Since Xand Yare transversal in
a neighborhood of the origin, they define a basis in this neighborhood and therefore,
there exist two smooth functions α, β such that U=αX+βY. Since both Uand Y
have the form (1.2), we have that β= 1 + β1where β1is a smooth function of order
1. We compute
[X,U]=[X, α X+βY] = X(α)X+α[X,X] + X(β)Y+β[X,Y]
=X(α)X+ (X(β) + βµ)Y.
Since [X,Y] = νX, we deduce X(β) = µβ.
We define Z=βYwhich is a smooth vector field with the form (1.2) since β= 1+β1
where β1is a smooth function of order 1. Then, [X,Z] = [X, βY] = β[X,Y] +
X(β)Y=βµY µβY= 0. ¥
We want to remark a difference between this Theorem 3.2 and Theorem 1.2. In The-
orem 1.2 the origin is required to be a center and the result characterizes its isochronic-
ity. In Theorem 3.2, the origin is only a linear center (that the equilibrium is a center
is not required), thanks to Theorem 1.1.
The methods developed in this work can be used to classify isochronous critical
points for polynomial systems. We give some examples of systems of the form (1.1)
with an isochronous critical point at the origin. The determination of the origin being
a focus is straightforward by computing Liapunov constants, see for instance [9]. When
9
the origin is a center, a first integral defined on a neighborhood of it is provided. We
also give an example of a family of quadratic systems depending on a real parameter
w6= 0 which never has an isochronous critical point at the origin. When w= 0, the
system has a center, and when w6= 0 the system has a weak focus at the origin.
Example 1. The following system has an isochronous critical point at the origin.
˙x=y+λ2x3+λ3x2y+λ4xy2,
˙y=x+λ2x2y+λ3xy2+λ4y3,(3.3)
with λ2, λ3, λ4R. In polar coordinates, this system reads for
˙r=r3
2(λ2+λ4+ (λ2λ4) cos(2θ) + λ3sin(2θ)),˙
θ= 1.
Then, by definition, the origin is an isochronous critical point. A first integral for
system (3.3) is given by
H(x, y) = x2+y2
1λ3x2+ (λ2λ4)xy + (λ2+λ4)(x2+y2) arctan( y
x).
When λ2+λ46= 0, the origin is a focus and when λ2+λ4= 0, the origin is a cen-
ter. Let us consider Xthe corresponding vector field and Y=x
∂x +y
∂y . We have
[X,Y] = 2(λ2x2+λ3xy +λ4y2)Y.
Example 2. The following system has an isochronous focus at the origin.
˙x=y2xy +xy22y3+µ2(x3xy2) + µ3x2yy4+
µ2(x2y2+y4)µ2xy4µ3y5µ2y6,
˙y=x+y2+y3+µ2(x2yy3) + µ3xy2+ 2µ2xy3+µ3y4+µ2y5,
(3.4)
where µiare arbitrary real constants for i= 2,3. This system has no constant angular
speed. An easy computation shows that the first Liapunov constant equals 1/2, so
the origin of (3.4) is a stable weak focus. We use Theorem 3.1 to ensure the property
of isochronicity.
Let us consider Xthe corresponding vector field and Ythe following analytic vector
field
Y= (xy2)
∂x +y
∂y .
The Lie bracket [X,Y] gives [X,Y] = 2(y2+µ2(x2y2+2xy2+y4) + µ3(xy +y3)) Y.
Therefore, the hypothesis of Theorem 3.1 are satisfied and the origin of system (3.4)
is an isochronous focus.
Example 3. The following family of quadratic systems depending on the parameter
wR
˙x=y, ˙y=x4wxy + 2y2,(3.5)
never has an isochronous critical point at the origin.
It can be shown that wis the first Liapunov constant for this family of quadratic
systems. Hence, when w > 0 the origin is an unstable weak focus and when w < 0 the
origin is a stable weak focus. When w= 0, we have that H(x, y) = (4x+ 8y21)e4x
defines a first integral which is analytic in a neighborhood of the origin. So, the origin
is a center for w= 0.
10
We will try to construct a vector field Yand a function µsatisfying Theorem 3.1 and
we will get a contradiction. Assume that there exists a vector field Ywith a star node at
the origin such that the Lie bracket between the vector field Xwdefined by (3.5) and Yis
equal to µ(x, y)Yfor a certain scalar analytic function µ(x, y) with µ(0,0) = 0. We can
write Y= (x+Pi>1Ai(x, y)) /∂x + (y+Pi>1Bi(x, y)) ∂/∂y, where Ai(x, y ), Bi(x, y)
are homogeneous polynomials of degree iand µ(x, y) = Pi>0mi(x, y ) where mi(x, y)
is a homogeneous polynomial of degree i.
Equating the terms of order 2 in the equation [Xw,Y] = µYwe get the two following
equations:
y∂A2
∂x +xA2
∂y +B2=m1x,
y∂B2
∂x +xB2
∂y + 4xwy 2y2A2=m1y.
The solution of these two equations is A2(x, y) = ax2+bxy (2/3)y2,B2(x, y) =
(4w/3)x2+axy +by2and m1(x, y ) = (b+ (4w/3))x((4/3) + a)y, where a, b are any
two real numbers.
Equating the terms of order 3 in the equation [Xw,Y] = µY, we get the two
following equations:
y∂A3
∂x +xA3
∂y + (2y24wxy)A2
∂y +B3=m2x+m1A2,
y∂B3
∂x +xB3
∂y + (2y24wxy)B2
∂y + 4wyA2A34(ywx)B2=m2y+m1B2.
Let us write m2(x, y) = Pi+j=2 mij xiyj,A3(x, y) = Pi+j=3 aijxiyjand B3(x, y ) =
Pi+j=3 bij xiyj. We consider the vector of unknowns
v={m20, m11 , m02, a30 , a21, a12 , a03, b30 , b21, b12 , b03}
and we can write the previous two equations as a linear system of eight equations in
these eleven unknowns : M v =k. The matrix Mis
M=
10001001000
01 0 3 0 2 0 0 1 0 0
0 0 1 0 2 0 3 0 0 1 0
0 0 0 0 0 1 0 0 0 0 1
0 0 0 1 0 0 0 0 1 0 0
1 0 0 0 1 0 0 3 0 2 0
01 0 0 0 1 0 0 2 0 3
0 0 1 0 0 0 1 0 0 1 0
which can be seen that it is of rank 7. The vector kis
k=½a
3(3b+ 4w),1
3(4a3a2+ 3b2+ 16bw),1
9(36b9ab 56w),
2
9(16 + 3a),4
9(3b8w)w, 1
9(9ab + 32w36aw),
1
3(2a3a2+ 3b2+ 4bw),1
3(4b3ab + 8w)¾.
11
The matrix (M|k) has rank 8 as the determinant of one of its 8 ×8 minors equals
1 + w2. So, the linear system does not satisfy the compatibility condition and, hence,
no such Ynor µcan exist.
Acknowledgements
The authors would like to thank Prof. I.A. Garc´ıa from Univ. de Lleida and Prof. M.
Sabatini from Univ. degli Studi di Trento for several useful conversations and remarks
about this work.
References
[1] A.A. Andronov, E.A. Leontovich, I.I. Gordon, and A.G. Ma˘ıer. Qualitative theory
of second-order dynamic systems. Translated from the Russian by D. Louvish.
Halsted Press (A division of John Wiley & Sons), New York-Toronto, Ont.; Israel
Program for Scientific Translations, Jerusalem-London, 1973.
[2] A. Algaba, E. Freire and E. Gamero. Isochronicity via normal form. Qual. Theory
Dyn. Syst. 1(2000), no. 2, 133–156.
[3] J. Chavarriga, I.A. Garc´ıa and J. Gin´e. On Lie’s symmetries for planar polynomial
differential systems. Nonlinearity 14 (2001), no. 4, 863–880.
[4] J. Chavarriga and M. Sabatini. A survey of isochronous centers. Qual. Theory
Dyn. Syst. 1(1999), no. 1, 1–70.
[5] S.N. Chow, C. Li and D. Wang. Normal forms and bifurcation of planar vector
fields. Cambrige Univ. Press, 1994.
[6] R. Conti. Uniformly isochronous centers of polynomial systems in R2.Lecture
Notes in Pure Appl. Math., vol. 152, pp. 21–31, M. Dekker, New York, (1994).
[7] E. Freire, A. Gasull and A. Guillamon. A characterization of isochronous centres
in terms of symmetries. Rev. Mat. Iberoamericana,20 (2004), no. 1, 205-222.
[8] A. Gasull, A. Guillamon and J. Villadelprat. The period function for second-order
quadratic ODEs is monotone.Qual. Theory Dyn. Syst.,4(2004), no. 2, 329–352.
[9] J. Gin´e. Isochronous foci for analytic differential systems. Internat. J. Bifur. Chaos
Appl. Sci. Engrg.,13 (2003), no. 6, 1–7.
[10] M. Sabatini. Characterizing isochronous centres by Lie brackets. Differential Equa-
tions Dynam. Systems 5(1997), no. 1, 91–99.
[11] M. Sabatini. Dynamics of commuting systems on two-dimensional manifolds. Ann.
Mat. Pura Appl. (4) 173 (1997), 213–232.
[12] M. Sabatini. Non periodic isochronous oscillations in plane differential systems.
Ann. Mat. Pura Appl. (4) 182 (2003), 487–501.
[13] P.J. Olver. Applications of Lie groups to differential equations. Second edition.
Graduate Texts in Mathematics,107. Springer-Verlag, New York, 1993.
[14] L. Perko. Differential equations and dynamical systems. Third edition. Texts in
Applied Mathematics,7. Springer-Verlag, New York, 2001.
12
[15] H. Poincar´e, “Thesis”, 1879; also “Oeuvres I”, 59-129, Gauthier Villars, Paris,
1928.
[16] M. Villarini. Regularity properties of the period function near a center of a planar
vector field, Nonlinear Analysis T.M.A. 19 (1992), 787–803.
[17] Zhang Zhi Fen et al. Qualitative theory of differential equations. Translations of
Mathematical Monographs, 101. American Mathematical Society, Providence, RI,
1992.
13
... However, in this case, a complete classification is yet to have arrived. Among these solvable systems a particular family of systems has been classified as isochronous systems (see Refs. [1,2,[14][15][16]). These isochronous systems exhibit amplitude-independent frequency of oscillations [2]. ...
... The above expression can be derived from Eq. (14). In general, one can find two solutions for x 1 from Eq. (20). ...
... Through this way, we fix the exact expression for x 1 (t). • Step 4: From the known solution x 1 (t), we can derive an expression for x 2 (t) using the first expression of Eq. (14). ...
Article
Full-text available
In a recent work, Calogero and Payandeh have identified a class of solvable two coupled first order nonlinear ordinary differential equations by connecting the roots of a monic polynomial of degree four with the coefficients of the polynomial. In this paper, we apply one of their earlier works to these first order nonlinear differential equations and enumerate the evolution equations in Newtonian form. Further, we demonstrate that second order evolution equations of the roots also follow the dynamics of the coefficients of the polynomial.
... Similarly to the above problem, we can restrict our analysis to the class of vector fields that remain in type W. This is the case associated to the problem of studying isochronous foci, a problem that was addressed for example by Giné in [19][20][21]. In this special class, the cyclicity problem is also an interesting problem to be approached, which up to our knowledge is not completely solved even for low degree vector fields. ...
... where the series converges for 0 ≤ ϕ ≤ 2π and sufficiently small ρ ≥ 0. In the center case, (20) can be seen as the lowest period of the trajectory of (1) passing through (x, y) = (ρ, 0) for ρ = 0. The coefficients T i of the period function are then given by the expression ...
... Moreover, if α V 3 < 0 and α is small enough, then a unique limit cycle bifurcates from the origin due to a Hopf bifurcation. (ii) The period function (20) when α = 0 has the form ...
Article
Full-text available
The present work introduces the problem of simultaneous bifurcation of limit cycles and critical periods for a system of polynomial differential equations in the plane. The simultaneity concept is defined, as well as the idea of bi-weakness in the return map and the period function. Together with the classical methods, we present an approach which uses the Lie bracket to address the simultaneity in some cases. This approach is used to find the bi-weakness of cubic and quartic Liénard systems, the general quadratic family, and the linear plus cubic homogeneous family. We finish with an illustrative example by solving the problem of simultaneous bifurcation of limit cycles and critical periods for the cubic Liénard family.
... This problem plays an important role in the study of dynamical systems, and has also been intensively investigated over the past three decades. The study on the isochronous center problem is also interesting, and in fact many results have been rediscovered several times, see for instance [15,16,55] and references therein. Several methods have been developed for deriving the necessary conditions under which a center becomes an isochronous center, see [25,27,28,5,7] and references therein. ...
... Proposition 5.3. Under each of the conditions L 12 -L 16 , system (4.6) respectively has a linearizability transformation T k , k = 12, 13, . . . ...
Article
Full-text available
In this paper, we study complex isochronous center problem for cubic complex planar vector fields, which are assumed to be Z2-equivariant with two symmetric centers. Such integrable systems can be classified as 11 cases. A complete classification is given on the complex isochronous centers and proven to have a total of 54 cases. All the algebraic conditions for the 54 cases are derived and, moreover, all the corresponding linearization transformations are obtained. This problem for the Z2-equivariant with two symmetric centers has been completely solved.
... For two-dimensional systems, the so-called isochronous centers have been studied in [4,5,10,14,13,17]. These are centers for which all orbits have the same frequency. ...
... These are centers for which all orbits have the same frequency. Related to isochronous centers are isochronous foci [9,10,16]. Here there are no periodic orbits and the time of return to a transversal cross-section to the orbit in the focus is studied. ...
Preprint
Full-text available
In this article we present a method for determining if the frequency of a family of periodic orbits remains constant when a parameter changes. Two-dimensional systems of ordinary and delayed differential equations are considered. Several examples are given.
... However, as we have shown, unless k 3 = η 2 = 0, the origin of the system is an isochronous focus (see e.g. [1,17] for definitions) but not a center. We note that the conditions for isochronicity of system (1.2) are also given in the survey paper [8]. ...
... However, as it is shown above, generally speaking system (e) has not a center, but an isochronous focus at the origin (see e.g. [1,17] for definitions). So system (e) should not be presented in the statements of Theorem 4.1. ...
Article
Full-text available
In this paper we investigate the problem of linearizability for a family of cubic complex planar systems of ordinary differential equations. We give a classification of linearizable systems in the family obtaining conditions for linearizability in terms of parameters. We also discuss coexistence of isochronous centers in the systems.
... This solvability breaks down when entering the realm of non-linear ordinary differential equation systems where no general solution method exists. Integrable non-linear systems are found only in the minority of cases including for example linearizable planar systems [1,2] or polynomial systems with isochronous centers [3,4]. In most cases, however, non-linear systems are handled with sophisticated numerical methods such as Runge-Kutta schemes [5] or other advanced techniques [6]. ...
Article
Full-text available
We study the autonomous systems of quadratic differential equations of the form $${\dot{x}}_i(t)={\textbf{x}}(t)^T {\textbf{A}}_i {\textbf{x}}(t) + {\textbf{v}}_i^T {\textbf{x}}(t)$$ x ˙ i ( t ) = x ( t ) T A i x ( t ) + v i T x ( t ) with $${\textbf{x}}(t) = (x_1(t),$$ x ( t ) = ( x 1 ( t ) , $$x_2(t), \ldots ,x_i(t),\dots )$$ x 2 ( t ) , … , x i ( t ) , ⋯ ) which, in general, cannot be solved exactly. In the present paper, we introduce a subclass of analytically solvable quadratic systems, whose solution is realized through a multi-dimensional generalization of the inversion which transforms a quadratic system into a linear one. We provide a constructive algorithm which, on one hand, decides whether the system of differential equations is analytically solvable with the inversion transformation and, on the other hand, provides the solution. The presented results apply for arbitrary, finite number of variables.
... Near singular points or limit cycles, many authors have studied the problem of the existence of a transversal section for which the flight return time function is constant. These transversal sections are called isochrones and their existence and computation are useful in biological models, see [1,7,13,15]. Our interest is to show which features of the period function are maintained for the flight return time function, independently of the transversal section, for a perturbation of a differential system with a period annulus. ...
Article
We deal here with planar analytic systems x˙=X(x,ε) which are small perturbations of a period annulus. For each transversal section Σ to the unperturbed orbits we denote by TΣ(q,ε) the time needed by a perturbed orbit that starts from q∈Σ to return to Σ. We call this the flight return time function. We say that the closed orbit Γ of x˙=X(x,0) is a continuable critical orbit in a family of the form x˙=X(x,ε) if, for any q∈Γ and any Σ that passes through q, there exists qε∈Σ a critical point of TΣ(⋅,ε) such that qε→q as ε→0. In this work we study this new problem of continuability. In particular we prove that a simple critical periodic orbit of x˙=X(x,0) is a continuable critical orbit in any family of the form x˙=X(x,ε). We also give sufficient conditions for the existence of a continuable critical orbit of an isochronous center x˙=X(x,0).
Article
In this paper, we describe the obstructions preventing the germ of an order-1 elliptic family from being temporally normalizable in the analytic case. We describe two classes of symmetries on the temporal part of the modulus. They arise out of an anti-holomorphic involution of the complex foliation. We identify the temporal modulus of analytic classification in the Poincaré domain in the parameter space.
Article
Full-text available
In this paper we show that any polynomial planar vector field which is either polynomial or rational integrable possesses a polynomial or rational infinitesimal generator of a Lie symmetry, respectively. Moreover, if all the critical points of the vector field are strong and there exists a polynomial inverse integrating factor that vanishes at all the critical points we show that, independently of the class of their first integral, there exists a polynomial infinitesimal generator of a Lie symmetry.
Article
Full-text available
Very little is known about the period function for large families of centers. In one of the pioneering works on this problem, Chicone [3] conjectured that all the centers encountered in the family of second-order differential equations [(x)\ddot] = V(x,[(x)\dot])\ddot x = V(x,\dot x) , beingV a quadratic polynomial, should have a monotone period function. Chicone solved some of the cases but some others remain still unsolved. In this paper we fill up these gaps by using a new technique based on the existence of Lie symmetries and presented in [8]. This technique can be used as well to reprove all the cases that were already solved, providing in this way a compact proof for all the quadratic second-order differential equations. We also prove that this property on the period function is no longer true whenV is a polynomial which nonlinear part is homogeneous of degreen>2.
Article
Full-text available
We analyze the isochronicity of centres, by means of a procedure to obtain hypernormal forms (simplest normal forms) for the Hopf bifurcation, that uses the theory of transformations based on the Lie transforms. We establish the relation between the period constants and the normal form coefficients, and prove that an equilibrium point is an isochronous centre if and only if a property of commutation holds. Also, we give necessary and sufficient conditions, expressed in terms of the Lie product, to determine if an equilibrium point is a centre. Several examples are also included in order to show the usefulness of the method. In particular, the isochronicity of the origin for the Lienard equation is analyzed in some cases.
Article
Scitation is the online home of leading journals and conference proceedings from AIP Publishing and AIP Member Societies
Article
We give a description of the local and global behaviour of orbits of commuting systems in the plane and on two-dimensional compact, connected, oriented manifolds.