$2\pi$-periodic self-similar solutions for the anisotropic affine curve shortening problem II

advertisement
2πœ‹-periodic self-similar solutions for the anisotropic
affine curve shortening problem II
Meiyue Jiang
∗
Juncheng Wei
†
July 5, 2012
Abstract
The existence of 2πœ‹-periodic positive solutions of the equation
𝑒′′ + 𝑒 =
π‘Ž(π‘₯)
𝑒3
is studied, where π‘Ž is a positive smooth 2πœ‹-periodic function. Under some
non-degenerate conditions on π‘Ž, the existence of solution to the equation is
established.
1
Introduction and statement of the results
This paper is a continuation of [21] and studies the existence of 2πœ‹-periodic positive
solutions of the equation
π‘Ž(π‘₯)
𝑒′′ + 𝑒 = 3
(1.1)
𝑒
for positive, smooth and 2πœ‹-periodic π‘Ž. Equation (1.1) arises from self-similar solutions of the following generalized curve shortening problem
∂𝛾
= Φ(πœƒ)βˆ£π‘˜βˆ£πœŽ−1 π‘˜π‘,
∂𝑑
𝜎 > 0,
π‘₯ ∈ π•Š1 = ℝ/2πœ‹β„€,
(1.2)
where 𝛾(⋅, 𝑑) is a planar curve, π‘˜(⋅, 𝑑) is its curvature with respect to the unit normal
𝑁 , and Φ is a positive function depending on the normal angle π‘₯ of the curve. This
problem has been extensively studied in the last three decades, see [1-7, 9, 10, 12,
∗
LMAM, School of Mathematical Sciences, Peking University, Beijing, 100871, P. R. China,
E-mail: mjiang@math.pku.edu.cn.
†
Department of Mathematics, The Chinese University of Hong Kong, Shatin, Hong Kong. Email: wei@math.cuhk.edu.hk.
1
13, 15-19, 23]. Assuming that 𝛾(⋅, 𝑑) is convex, and 𝑀(π‘₯, 𝑑) is its support function,
then using the normal angle π‘₯ to parameterize 𝛾, equation (1.2) is equivalent to
∂𝑀
−Φ(π‘₯)
= ′′
,
∂𝑑
(𝑀 + 𝑀)𝜎
π‘₯ ∈ π•Š1 .
(1.3)
Self-similar solutions of (1.3) are solutions having the form 𝑀(π‘₯, 𝑑) = πœ‰(𝑑)𝑒(π‘₯), which
are important in understanding the long time behaviors and the structure of singularities of (1.2). It is rather easy to see that πœ‰(𝑑)𝑒(π‘₯) is a self-similar solution if and
only if 𝑒 satisfies
π‘Ž(π‘₯)
𝑒′′ + 𝑒 = 𝑝+1 , π‘₯ ∈ π•Š1
(1.4)
𝑒
1
with π‘Ž(π‘₯) = Φ πœŽ (π‘₯), 𝑝 + 1 = 𝜎1 and βˆ£πœ‰(𝑑)∣𝜎−1 πœ‰(𝑑)πœ‰ ′ (𝑑) = −𝐢, where 𝐢 is a positive
constant. Equation (1.4) also appears in image processing [23], 2-dimensional 𝐿𝑝 Minkowski problem [8, 20] and other problems [19].
Equation (1.4) with 𝑝 βˆ•= 2 has been studied by many authors. When π‘Ž ≡ 1,
all solutions of (1.4) can be classified, see [1, 5]; and see [12, 13] for some results
when π‘Ž is 2πœ‹-periodic. In particular, Matano-Wei [22] proved that (1.4) is solvable
if 0 ≤ 𝑝 < 7 and π‘Ž is 2πœ‹-periodic and positive. An equation closely related to (1.4)
is
π‘Ž(π‘₯)
𝑒′′ + πœ†π‘’ = 𝜈 .
(1.5)
𝑒
Using the Poincaré-Birkhoff fixed point theorem, del Pino, Manásevich and Montero
2
in [11] proved that that (1.5) possesses a 2πœ‹−periodic solution if πœ† βˆ•= (𝑛+1)
, 𝑛=
4
0, 1, ⋅ ⋅ ⋅ . An important step in establishing the existence of solutions of (1.4) and
(1.5) is to get an a priori estimate for all solutions.
The case 𝜎 = 13 in (1.2) is called the affine curve shortening problem. Thus 𝑝 = 2
and equation (1.4) becomes
𝑒′′ + 𝑒 =
π‘Ž(π‘₯)
𝑒3
π‘₯ ∈ π•Š1 ,
(1.6)
and a solution of equation (1.6) is a self-similar solution of the anisotropic affine
curve shortening problem. All results mentioned above do not cover the affine case.
Indeed, the situation for the affine case is quite different. It is known that there
are some obstructions for the existence and one can’t get a priori estimates for the
solutions of (1.6) without additional assumptions on π‘Ž due to the invariance of the
problem. To see this, let us consider its simplest form
𝑒′′ + 𝑒 =
1
,
𝑒3
π‘₯ ∈ π•Š1 .
(1.7)
Equation (1.7) is invariant under an action of the special linear group 𝑆𝐿(2, ℝ), and
all solutions of (1.7) are given by a 2-parameter family of functions
(
)1
π‘’πœ€,πœƒ (π‘₯) = πœ€2 cos2 (π‘₯ − πœƒ) + πœ€−2 sin2 (π‘₯ − πœƒ) 2 ,
2
(1.8)
for (πœ€, πœƒ) ∈ (0, 1] × [0, πœ‹). Thus the set of 2πœ‹-periodic solutions of (1.7) is not
bounded. For more details on the group invariance of (1.7), see [2, 21].
To state the results for equation (1.6), we need two functions. Let π‘Ž be a positive
2πœ‹-periodic and 𝐢 2 -function, and let
π‘Ž′ (πœƒ)
π‘Ž′ (πœƒ + πœ‹)
𝐴2 (πœƒ) = √
+√
π‘Ž(πœƒ)
π‘Ž(πœƒ + πœ‹)
and
∫
𝐡2 (πœƒ) =
πœ‹
2
− πœ‹2
(
)
π‘Ž′ (πœƒ + 𝑑) + π‘Ž′ (πœƒ + 𝑑 + πœ‹) − π‘Ž′ (πœƒ) − π‘Ž′ (πœƒ + πœ‹) sin 2𝑑
𝑑𝑑.
sin2 𝑑
(1.9)
(1.10)
Note that by definition, 𝐴2 (πœƒ) and 𝐡2 (πœƒ) are πœ‹−periodic, so they are defined on
π•Š11 = ℝ/πœ‹β„€. A function π‘Ž is called 𝐡-nondegenerate if (𝐴2 (πœƒ), 𝐡2 (πœƒ)) βˆ•= (0, 0) for
πœƒ ∈ π•Š11 .
In [2], Ai, Chou and Wei proved that if π‘Ž is a positive, B-nondegenerate and 𝐢 2 function of period πœ‹, then one can get a priori estimates for all πœ‹-periodic solutions
of (1.6). That is, there exists a constant 𝐢 depending on π‘Ž only such that
1
≤𝑒≤𝐢
𝐢
(1.11)
for any πœ‹-periodic solution 𝑒 of (1.6). This result was generalized in [21] to the case
that π‘Ž is 2πœ‹-periodic. In establishing a priori estimates, a major problem is to study
possible blow-ups. The difference between πœ‹-periodic and 2πœ‹-periodic cases is, for a
blow-up sequence π‘’πœ€,𝑑 , only single blow-up can occur when is π‘Ž is πœ‹−periodic, while
π‘Ž is 2πœ‹-periodic, there are 2 possible blow-ups. We have to analyze the interaction
between different blow-ups.
Let π‘Ž is a positive and B-nondegenerate. Then the map 𝐺(πœƒ) = (𝐴2 (πœƒ), 𝐡2 (πœƒ))
satisfies 𝐺(πœƒ) βˆ•= 0 for πœƒ ∈ π•Š11 , and the degree 𝑑𝑒𝑔(𝐺, π•Š11 ) is well-defined, where
𝑑𝑒𝑔(𝐺, π•Š11 ) = 𝑑𝑒𝑔(𝐺, 𝐷; 0), 𝐷 = {(π‘₯, 𝑦)∣π‘₯2 + 𝑦 2 ≤ 1} = {(π‘Ÿ cos 2πœƒ, π‘Ÿ sin 2πœƒ)∣0 ≤ π‘Ÿ ≤
1, πœƒ ∈ [0, πœ‹]}, ∂𝐷 ≈ π•Š11 = {(cos 2πœƒ, sin 2πœƒ)βˆ£πœƒ ∈ [0, πœ‹]}, 𝐺 : 𝐷 → ℝ2 is a continuous
extension of 𝐺 : π•Š11 → ℝ2 , and 𝑑𝑒𝑔(𝐺, 𝐷; 0) is the Brouwer degree. It is well known
that 𝑑𝑒𝑔(𝐺, 𝐷; 0) is determined by the map 𝐺.
The existence results in [2] and [21] now can be stated as follows.
Theorem 1.1 Let π‘Ž be a positive, 𝐢 2 and 2πœ‹-periodic and B-nondegenerate function. Then
(1) equation (1.6) has a πœ‹−periodic solution if π‘Ž is πœ‹-periodic and 𝑑𝑒𝑔(𝐺, π•Š11 ) βˆ•=
−1;
(2) equation (1.6) has a 2πœ‹−periodic solution if π‘Ž is 2πœ‹-periodic, βˆ₯1 − π‘Žβˆ₯𝐢 2 << 1
and 𝑑𝑒𝑔(𝐺, π•Š11 ) βˆ•= −1.
The part (1) in the above theorem was proved in [2] and part (2) was proved in
[21]. The statement in [21] is slightly different, where the degree of 𝐺 is computed
as a 2πœ‹-periodic map that equals to 2𝑑𝑒𝑔(𝐺, π•Š11 ).
3
The aim of this paper is to remove the assumption βˆ₯1 − π‘Žβˆ₯𝐢 2 << 1 in the case
that π‘Ž is 2πœ‹−periodic. Namely, we will prove
Theorem 1.2 Let π‘Ž be a positive, 𝐢 2 and 2πœ‹-periodic and B-nondegenerate function. Then equation (1.6) has a 2πœ‹−periodic solution provided that 𝑑𝑒𝑔(𝐺, π•Š11 ) βˆ•= −1.
Now we explain briefly the reason that we made the assumption βˆ₯1 − π‘Žβˆ₯𝐢 2 << 1
in [21]. For πœ‹−peridic case, we can fix πœ€ << 1 and consider the homotopy of π‘Žπ‘  (π‘₯) =
(1 − 𝑠)(1 + πœ€π‘Ž(π‘₯)) + π‘ π‘Ž(π‘₯). Then for 𝑠 ∈ [0, 1], the function π‘Žπ‘  is 𝐡−nondegenerate
if π‘Ž is, and thus one can get uniform a priori estimate for all πœ‹-periodic solutions of
𝑒′′ + 𝑒 =
π‘Žπ‘  (π‘₯)
.
𝑒3
(1.12)
By Lyapunov-Schmit reduction, one can solve (1.12) for small 𝑠 if 𝑑𝑒𝑔(𝐺, π•Š11 ) βˆ•= −1,
and the solution can be continued to 𝑠 = 1 thanks to the priori estimate, see [2].
Lyapunov-Schmit reduction can be applied to 2πœ‹-periodic case as well. However,
for 2πœ‹-periodic π‘Ž, we do not know how to construct a homotopy like π‘Žπ‘  as in the
πœ‹-periodic case. The 𝐡−nondegenerate condition on π‘Ž is a nonlinear restriction.
Thus we imposed the assumption βˆ₯1 − π‘Žβˆ₯𝐢 2 << 1.
In this paper, we take a different and more direct approach to prove Theorem
1.2. The key point and the main difference between this paper and [21] is that we
do not consider periodic solutions of (1.6), but instead the initial value problem of
(1.6). Let 𝑒(𝑒0 , 𝑒′0 ; π‘₯) be the solution of (1.6) with the initial boundary condition
𝑒(𝑒0 , 𝑒′0 ; 0) = 𝑒0 ,
𝑒′ (𝑒0 , 𝑒′0 ; 0) = 𝑒′0 .
An important observation is that the analysis of periodic solutions in [21] is valid
for solutions of the initial value problem as well, which enables us to use the degree
argument to find fixed point (𝑒0 , 𝑒′0 ) of the Poincaré map:
𝑒(𝑒0 , 𝑒′0 ; 2πœ‹) = 𝑒0 ,
𝑒′ (𝑒0 , 𝑒′0 ; 2πœ‹) = 𝑒′0 .
(1.13)
By the periodicity of π‘Ž, (1.13) implies that 𝑒(𝑒0 , 𝑒′0 ; π‘₯) is a 2πœ‹-periodic solution of
(1.6). But we do not use degree theory to find zeros of the map
(𝑒0 , 𝑒′0 ) → (𝑒(𝑒0 , 𝑒′0 ; 2πœ‹) − 𝑒0 ), 𝑒′ (𝑒0 , 𝑒′0 ; 2πœ‹) − 𝑒′0 ))
directly, instead we introduce another map (𝑒0 , 𝑒′0 ) → β„±(𝑒0 , 𝑒′0 ) defined by
∫ 2πœ‹
∫ 2πœ‹
π‘Ž′ (π‘₯)
π‘Ž′ (π‘₯)
′
β„±(𝑒0 , 𝑒0 ) = (
cos
2π‘₯𝑑π‘₯,
cos 2π‘₯𝑑π‘₯).
𝑒2 (𝑒0 , 𝑒′0 ; π‘₯)
𝑒2 (𝑒0 , 𝑒′0 ; π‘₯)
0
0
A key fact that will be used is that β„±(𝑒0 , 𝑒′0 ) = 0 implies (1.13). The advantage to
consider the map β„±(𝑒0 , 𝑒′0 ) is that one can get the asymptotic expansion as πœ† → 0
via the blow-up analysis. Having the asymptotic expansion, the existence of (𝑒0 , 𝑒′0 )
satisfying β„±(𝑒0 , 𝑒′0 ) = 0 can be accomplished by degree theory.
4
The paper is organized as follows. In Section 2, it is shown that the existence
of 2πœ‹-periodic solutions of (1.6) is equivalent to that β„± has zeros. We analyze the
critical points of the solution 𝑒(πœ†, πœƒ; π‘₯) in detail in Section 3. An asymptotical
expansion of β„± is given in Section 4 following the blow-up analysis of the solutions
𝑒(πœ†, πœƒ; π‘₯), and the main theorem is proved in Section 5 by making use of asymptotic
behavior of the map β„± and the degree theory.
2
A Map
In this section we define a map β„± : (0, 1] × β„/πœ‹β„€ → ℝ2 such that each zero of the
map corresponds to a 2πœ‹-periodic solution of
𝑒′′ + 𝑒 =
π‘Ž(π‘₯)
.
𝑒3
(2.1)
Let 𝑒(πœ†, πœƒ; π‘₯) be the solution of (2.1) satisfying the initial value conditions
𝑒(πœ†, πœƒ; 0) =
√
(πœ†2 − πœ†−2 ) sin πœƒ cos πœƒ
.
𝑒′ (πœ†, πœƒ; 0) = √
πœ†2 cos2 πœƒ + πœ†−2 sin2 πœƒ
πœ†2 cos2 πœƒ + πœ†−2 sin2 πœƒ;
It is easy to see that any initial value 𝑒0 > 0, 𝑒′0 ∈ ℝ can be written as the above
form for (πœ†, πœƒ) ∈ (0, 1] × β„/πœ‹β„€. Hence all solutions of (2.1) are given by 𝑒(πœ†, πœƒ; π‘₯).
By the periodicity of π‘Ž, we know that 𝑒(πœ†, πœƒ; π‘₯) is 2πœ‹-periodic solution of (2.1) if
and only if
𝑒(πœ†, πœƒ; 2πœ‹) = 𝑒(πœ†, πœƒ; 0), 𝑒′ (πœ†, πœƒ; 2πœ‹) = 𝑒′ (πœ†, πœƒ; 0).
(2.2)
We start with the following simple lemma.
Lemma 2.1 For any (πœ†, πœƒ), the solution 𝑒(πœ†, πœƒ; π‘₯) exists for π‘₯ ∈ ℝ and there is a
constant 𝐢 independent of πœ† such that
1 2
π‘Ž(π‘₯)
(πœ† + πœ†−2 ) ≤ 𝑒2 (πœ†, πœƒ; π‘₯) + 𝑒′2 (πœ†, πœƒ; π‘₯) + 2
≤ 𝐢(πœ†2 + πœ†−2 ),
𝐢
𝑒 (πœ†, πœƒ; π‘₯)
π‘₯ ∈ [0, 2πœ‹].
Proof. Let 𝑒 be a solution of equation (2.1). We define 𝐹 (π‘₯) = 𝑒′2 + 𝑒2 +
Then by (2.1), one can easily get that
𝐹 ′ (π‘₯) =
which implies that
π‘Ž(π‘₯)
.
𝑒2 (π‘₯)
π‘Ž′ (π‘₯)
,
𝑒2
∣𝐹 ′ (π‘₯)∣ ≤ 𝐢∣𝐹 (π‘₯)∣
since π‘Ž(π‘₯) is smooth and positive. Thus
∣(log 𝐹 )′ ∣ ≤ 𝐢,
5
(2.3)
(2.4)
which leads to, for all 𝑅 > 0, there exists a constant 𝐢(𝑅) > 0 such that
𝐹 (π‘₯) ≤ 𝐢(𝑅),
π‘₯ ∈ [−𝑅, 𝑅].
Now the global existence of 𝑒 follows immediately.
From (2.4) and Gronwall inequality we have
1
𝐹 (0) ≤ 𝐹 (π‘₯) ≤ 𝐢𝐹 (0),
𝐢
π‘₯ ∈ [0, 2πœ‹].
(2.5)
The inequalities in (2.3) follows from (2.5),
1 ′2
1
1
(𝑒 (π‘₯) + 𝑒2 (π‘₯) + 2 ) ≤ 𝐹 (π‘₯) ≤ 𝐢(𝑒′2 (π‘₯) + 𝑒2 (π‘₯) + 2 )
𝐢
𝑒 (π‘₯)
𝑒 (π‘₯)
and
𝑒′2 (0) + 𝑒2 (0) +
This completes the proof.
Set
= πœ†2 + πœ†−2 .
β–‘
∫
𝐴(πœ†, πœƒ) =
𝐡(πœ†, πœƒ) =
2πœ‹
π‘Ž′ (π‘₯)
cos 2π‘₯𝑑π‘₯
𝑒2 (πœ†, πœƒ; π‘₯)
2πœ‹
π‘Ž′ (π‘₯)
sin 2π‘₯𝑑π‘₯
𝑒2 (πœ†, πœƒ; π‘₯)
0
∫
and
1
𝑒2 (0)
0
β„± : (0, 1] × β„/πœ‹β„€ → ℝ2 : β„±(πœ†, πœƒ) = (𝐴(πœ†, πœƒ), 𝐡(πœ†, πœƒ)).
(2.6)
We know that the map β„± is well defined for all (πœ†, πœƒ) by Lemma 2.1. An important
feature of the map β„± is that when πœ† = 1, the solution 𝑒(πœ†, πœƒ; π‘₯) is independent of
πœƒ, hence so is the map β„±(πœ†, πœƒ). This is important for later purpose and the reason
that we write the initial value in terms of πœ† and πœƒ. But we will see this also causes
some difficulties in choosing the blow-up point when we perform blow-up analysis
at the same time.
The following result is the starting point of our approach later, which relates the
2πœ‹-periodic solutions of (2.1) to the zeros of the map β„±(πœ†, πœƒ).
Proposition 2.2 Let π‘Ž be a positive, 𝐢 2 and 2πœ‹-periodic function and β„± be given
by (2.6). Then 𝑒(πœ†, πœƒ; π‘₯) is a 2πœ‹-periodic solution of (2.1) if and only if β„±(πœ†, πœƒ) = 0.
Proof. For any solution 𝑒 of (2.1), we have
(
𝑒2
π‘Ž′ (π‘₯)
𝑒2 ′′′
) + 4( )′ = 2 .
2
2
𝑒
(2.7)
Multiplying by cos 2π‘₯, sin 2π‘₯ and integration over [0, 2πœ‹], we get β„±(πœ†, πœƒ) = 0 if
𝑒(πœ†, πœƒ; π‘₯) is a 2πœ‹-periodic solution of (2.1).
6
Conversely, let β„±(πœ†, πœƒ) = 0. It follows from (2.7) and integration by parts that
∫ 2πœ‹ 2
𝑒2
𝑒
𝐴(πœ†, πœƒ) =
[( )′′′ + 4( )′ ] cos 2π‘₯𝑑π‘₯
2
2
0
2
𝑒
= ( )′′ cos 2π‘₯∣2πœ‹
0
(2.8)
2
′2
′′
2πœ‹
= (𝑒 + 𝑒 𝑒)∣0
π‘Ž(π‘₯)
= (𝑒′2 + 2 − 𝑒2 )∣2πœ‹
0 .
𝑒
Similarly, we have
∫
2πœ‹
𝑒2 ′′′
𝑒2 ′
𝐡(πœ†, πœƒ) =
[( ) + 4( ) ] sin 2π‘₯𝑑π‘₯
2
2
0
2
𝑒
= −2( )′ cos 2π‘₯∣2πœ‹
0
2
= −2𝑒′ π‘’βˆ£2πœ‹
0 .
(2.9)
Hence β„±(πœ†, πœƒ) = 0 and (2.8), (2.9) imply
𝑒′2 (2πœ‹) − 𝑒′2 (0) = 𝑒2 (2πœ‹) − 𝑒2 (0) +
′
π‘Ž(0)
𝑒2 (0)𝑒2 (2πœ‹)
(𝑒2 (2πœ‹) − 𝑒2 (0)),
′
(2.10)
𝑒 (2πœ‹)𝑒(2πœ‹) = 𝑒 (0)𝑒(0).
Let us assume 𝑒′ (0) βˆ•= 0 and 𝑐 =
𝑒′ (2πœ‹)
.
𝑒′ (0)
Then the second equality in (2.10) gives
𝑒′ (2πœ‹)
𝑒(0)
=
= 𝑐 > 0.
′
𝑒 (0)
𝑒(2πœ‹)
(2.11)
Substituting (2.11) into the first equality of (2.10) we get
(𝑐2 − 1)𝑒′2 (0) = (1 − 𝑐2 )(1 +
Therefore, 𝑐 = 1 and
𝑒′ (2πœ‹) = 𝑒′ (0),
π‘Ž(0)
)𝑒2 (2πœ‹).
𝑐2 𝑒4 (2πœ‹)
𝑒(2πœ‹) = 𝑒(0).
(2.12)
(2.13)
The case 𝑒′ (0) = 0 can be treated similarly. Thus (2.13) always holds true, that is,
𝑒(πœ†, πœƒ; π‘₯) is a 2πœ‹-periodic solution of (2.1).
β–‘
Remark 2.3 Let
∫
𝐢(πœ†, πœƒ) =
2πœ‹
0
π‘Ž′ (π‘₯)
𝑑π‘₯.
𝑒2 (πœ†, πœƒ; π‘₯)
Then it follows from (2.7) that 𝑒(πœ†, πœƒ; π‘₯) is a 2πœ‹-periodic solution of (2.1) also
implies 𝐢(πœ†, πœƒ) = 0. But this equation is not independent of β„±(πœ†, πœƒ) = 0.
7
In view of Proposition 2.2, in order to find 2πœ‹-periodic solutions of (2.1), it
suffices to find a solution of
β„±(πœ†, πœƒ) = 0.
(2.14)
This will be accomplished by a degree argument. To this end we need an asymptotic
expansion of β„±(πœ†, πœƒ) as πœ† → 0, which follows from blow-up analysis of 𝑒(πœ†, πœƒ; π‘₯) as
πœ† → 0.
3
Analysis of Critical Points of 𝑒(πœ†, πœƒ; π‘₯)
The computation of the degree of β„±(πœ†, πœƒ) relies on the asymptotic expansion, which
is based on the blow-up analysis of 𝑒(πœ†, πœƒ; 𝑦) as πœ† → 0. Since the blow-up is made
at a critical point of 𝑒(πœ†, πœƒ; 𝑦), we first give some analysis of the critical points of
𝑒(πœ†, πœƒ; 𝑦) in this section. This part is necessary and important in our later arguments,
which differs from [21], where we only consider periodic solutions, and the existence
of critical points is trivial.
From now on we always assume that π‘Ž is a positive, 𝐢 2 and 2πœ‹-periodic function.
And unless otherwise stated, the letter 𝐢 will always denote various generic constants
which are independent of π‘˜. We denote 𝐴 ∼ 𝐡 or 𝐴 = 𝑂(𝐡) if there exist two
positive uniform constants 𝐢1 and 𝐢2 such that 𝐢1 𝐴 ≤ 𝐡 ≤ 𝐢2 𝐴, and π‘œ(1) means a
quantity that goes to zero as πœ† → 0.
Lemma 3.1 Let 𝑒 be a monotonic solution of
𝑒′′ + 𝑒 =
π‘Ž(π‘₯ + πœƒ)
,
𝑒3
π‘₯ ∈ [0, 𝑇 ].
(3.1)
Then for 𝛿 > 0, there is an πœ€0 (𝛿) independent of πœƒ such that if 𝑒(0) ≤ πœ€0 or 𝑒(𝑇 ) ≤
πœ€0 , we have
𝛿
𝛾 = ∣{π‘₯ ∈ [0, 𝑇 ]βˆ£π‘’(π‘₯) ≤ πœ€0 }∣ ≤ .
4
Proof: Let 𝑒 be monotonic increasing. We fix a constant πœ€0 ≤
8
π‘Ž(π‘₯)
− 𝑒 ≥ 2,
3
𝑒
𝛿
0 < 𝑒 ≤ πœ€0 , π‘₯ ∈ [0, 2πœ‹].
1
8
such that
(3.2)
Then 𝐼 = {π‘₯ ∈ [0, 𝑇 ]βˆ£π‘’(π‘₯) ≤ πœ€0 } = [0, πœ‚] for some πœ‚ ≥ 0. By the mean value
theorem, we can find πœ‰ ∈ ( πœ‚2 , πœ‚) such that
πœ‚
πœ‚
𝑒(πœ‚) − 𝑒( ) = 𝑒′ (πœ‰) .
2
2
Therefore,
1
2
0 ≤ 𝑒′ (πœ‰) ≤ πœ€0 ≤ .
πœ‚
4πœ‚
8
(3.3)
Applying the mean value theorem to 𝑒′ , using (3.1)-(3.3) we get
𝑒′ (πœ‰) − 𝑒′ (0) ≥
8
4
πœ‰ ≥ 2 πœ‚.
2
𝛿
𝛿
(3.4)
Hence it follows from (3.3) and (3.4) that
4
1
πœ‚≤ ,
2
𝛿
4πœ‚
that is, πœ‚ ≤ 4𝛿 , thanks to the fact 𝑒′ (0) ≥ 0. The proof is complete.
With the aid of Lemma 3.1 now we have
β–‘
Proposition 3.2 Let 𝑒 be a monotonic solution of (3.1) such that
𝑒2 (π‘₯) + 𝑒′2 (π‘₯) +
π‘Ž(π‘₯)
∼ πœ†2 + πœ†−2 ,
𝑒2 (π‘₯)
π‘₯ ∈ [0, 𝑇 ].
(3.5)
Then for 𝛿 > 0, there is a πœ†0 independent of πœƒ such that for all πœ† ≤ πœ†0 , we have
𝑇 ≤ πœ‹2 + 𝛿. In particular, in any interval having length more that πœ‹2 + 𝛿, the function
𝑒(π‘₯) has at least one critical point.
Proof: As in Lemma 3.1, let us assume that 𝑒 is monotone increasing. Let 𝛿, πœ€0 be
the constants given by in Lemma 3.1 and 𝑇1 ∈ [0, 𝑇 ] such that 𝑒(𝑇1 ) = πœ€0 . Then by
Lemma 3.1 we have 𝑇1 ≤ 4𝛿 . Now we show 𝑇 − 𝑇1 ≤ πœ‹2 + 2𝛿 if πœ† is small.
We argue by contradiction. Suppose there is a sequence πœ†π‘› → 0, πœƒπ‘› , monotone
increasing functions 𝑒𝑛 defined on a subinterval 𝐼𝑛 of [0, 𝑇 ] having length more than
πœ‹
+ 2𝛿 satisfies (3.5) and
2
π‘Ž(π‘₯ + πœƒπ‘› )
𝑒′′𝑛 + 𝑒𝑛 =
.
(3.6)
𝑒3𝑛
After translation we may assume 𝐼𝑛 = [0, πœ‹2 + 2𝛿 ]. First we note
𝑒𝑛 (π‘₯) ≥ πœ€0 ,
π‘₯ ∈ [0,
πœ‹ 𝛿
+ ]
2 2
due to the fact that 𝑒𝑛 is monotonic increasing. Then
∣
π‘Ž(π‘₯)
∣ ≤ 𝐢,
𝑒3𝑛 (π‘₯)
π‘₯ ∈ [0,
πœ‹ 𝛿
+ ].
2 2
(3.7)
By the standard estimate to equation (3.6) we find
βˆ₯𝑒𝑛 βˆ₯𝐢 1 ≤ 𝐢(𝑀𝑛 + 1),
(3.8)
where 𝑀𝑛 = maxπ‘₯∈[0, πœ‹2 +𝛿] 𝑒𝑛 (π‘₯). It follows from (3.5) and (3.8) that 𝑀𝑛 → ∞ as
𝑒𝑛
→ 𝑣, maxπ‘₯∈[0, πœ‹2 ] 𝑣(π‘₯) = 1
𝑛 → ∞. Divided (3.6) by 𝑀𝑛 and letting 𝑛 → ∞, we get 𝑀
𝑛
and
πœ‹ 𝛿
(3.9)
𝑣 ′′ + 𝑣 = 0, π‘₯ ∈ [0, + ].
2 2
9
Since 𝑒𝑛 is monotone increasing on [0, πœ‹2 + 2𝛿 ], so is 𝑣. This is impossible as 𝑣 satisfies
(3.9) and 𝑣 ≩ 0. The proof is finished.
β–‘
Recall that 𝑒(πœ†, πœƒ; π‘₯) is the solution of
𝑒′′ + 𝑒 =
π‘Ž(π‘₯)
𝑒3
(3.10)
satisfying the initial value conditions
𝑒(πœ†, πœƒ; 0) =
√
πœ†2
cos2
πœƒ+
πœ†−2
(πœ†2 − πœ†−2 ) sin πœƒ cos πœƒ
√
𝑒 (πœ†, πœƒ; 0) =
.
πœ†2 cos2 πœƒ + πœ†−2 sin2 πœƒ
′
2
sin πœƒ;
According to Lemma 2.1, we know that
𝑒2 (π‘₯) + 𝑒′2 (π‘₯) +
π‘Ž(π‘₯)
∼ πœ†2 + πœ†−2 ,
𝑒2 (π‘₯)
π‘₯ ∈ [−2πœ‹, 2πœ‹].
(3.11)
Therefore, it follows from Proposition 3.2 that the function 𝑒(πœ†, πœƒ; π‘₯) has critical
points in [0, πœ‹2 + 𝛿] and [− πœ‹2 − 𝛿, 0], respectively. It is easy to see from (3.10) and
(3.11) that all critical points of 𝑒(πœ†, πœƒ; π‘₯) are nondegenerate for small πœ†, hence they
are isolated. Using these facts now we define two functions Θ(πœ†, πœƒ) and Θ1 (πœ†, πœƒ).
We restrict πœƒ ∈ [0, πœ‹2 ) first. Let Θ(πœ†, πœƒ) and Θ1 (πœ†, πœƒ) be the first critical point of
𝑒(πœ†, πœƒ; π‘₯) in π‘₯ ≥ 0 and π‘₯ ≤ 0, respectively. By definition we have that
Θ(πœ†, 0) = Θ1 (πœ†, 0) = 0,
Θ(πœ†, πœƒ) is a local minimizer and Θ1 (πœ†, πœƒ) is a local maximizer if πœƒ ∈ [0, πœ‹2 ), since πœ† is
small and 𝑒′ (πœ†, πœƒ; 0) ≤ 0. Moreover, as a consequence of (3.11) we see that
π‘š(πœ†, πœƒ) = 𝑒(πœ†, πœƒ; Θ(πœ†, πœƒ)) ∼ πœ†,
and
𝑀 (πœ†, πœƒ) = 𝑒(πœ†, πœƒ; Θ1 (πœ†, πœƒ)) ∼ πœ†−1 ,
πœ‹
πœƒ ∈ [0, )
2
πœ‹
πœƒ ∈ [0, ).
2
The following lemma is proved in [21].
Lemma 3.3 As πœ† → 0, the following estimate holds uniformly in πœƒ:
Θ(πœ†, πœƒ) − Θ1 (πœ†, πœƒ) =
πœ‹
+ π‘œ(1).
2
(3.12)
Using the lemma above we can prove
Proposition 3.4 Let 𝑒(πœ†, πœƒ; π‘₯) and Θ(πœ†, πœƒ) be as above. Then as πœ† → 0, the following estimate holds uniformly in πœƒ ∈ [0, πœ‹2 ):
Θ(πœ†, πœƒ) = πœƒ + π‘œ(1),
10
πœ‹
πœƒ ∈ [0, ).
2
(3.13)
Proof: Since πœƒ ∈ [0, πœ‹2 ), 𝑒(πœ†, πœƒ; π‘₯) is monotone decreasing on [Θ1 (πœ†, πœƒ), Θ(πœ†, πœƒ)]. We
are going to prove: ∀𝛿 > 0, ∃πœ†0 (𝛿) > 0 such that for all πœ† ≤ πœ†0 ,
∣Θ(πœ†, πœƒ) − πœƒβˆ£ ≤ 𝛿,
πœ‹
πœƒ ∈ [0, ).
2
(3.14)
To this end, first applying Lemma 3.1 to 𝑒(πœ†, πœƒ; π‘₯) in the interval [Θ1 , Θ], we
conclude that there is an πœ€0 (𝛿) such that
𝛿
∣{π‘₯ ∈ [Θ1 , Θ]βˆ£π‘’(πœ†, πœƒ; π‘₯) ≤ πœ€0 }∣ ≤ .
4
(3.15)
Let 𝑇 (πœ†, πœƒ) be given by 𝑒(πœ†, πœƒ; 𝑇 (πœ†, πœƒ)) = πœ€0 . Then (3.15) implies
𝛿
∣Θ(πœ†, πœƒ) − 𝑇 (πœ†, πœƒ)∣ ≤ .
4
(3.16)
In order to get (3.14), it suffice to show it holds for πœƒ ≤ 𝑇 (πœ†, πœƒ), since if πœƒ ≥ 𝑇 (πœ†, πœƒ),
(3.16) implies
𝛿
∣Θ(πœ†, πœƒ) − πœƒ)∣ ≤ ∣Θ(πœ†, πœƒ) − 𝑇 (πœ†, πœƒ)∣ ≤ .
4
In the remaining we show ∃πœ†0 (𝛿) > 0 such that for all πœ† ≤ πœ†0 ,
∣
πœ‹
𝛿
− πœƒ + Θ1 (πœ†, πœƒ)∣ ≤ ,
2
2
πœ‹
πœƒ ∈ [0, ).
2
(3.17)
Having this estimate, (3.14) follows from Lemma 3.3 immediately.
To prove (3.17) we consider the function π‘£πœ† (π‘₯) = πœ†π‘’(πœ†, πœƒ; π‘₯ + Θ1 ), which satisfies
π‘£πœ†′′ + π‘£πœ† = πœ†
and
π‘Ž(π‘₯ + Θ1 )
:= πœ†π‘“πœ† (π‘₯),
+ Θ1 )
𝑒3 (πœ†, πœƒ; π‘₯
π‘₯ ∈ [0, 𝑇 (πœ†, πœƒ) − Θ1 ]
π‘£πœ†′ (0) = πœ†π‘’′ (πœ†, πœƒ; Θ1 ) = 0,
√
π‘£πœ† (πœƒ − Θ1 ) = πœ†π‘’(πœ†, πœƒ; 0) = πœ† πœ†2 cos2 πœƒ + πœ†−2 sin2 πœƒ.
Thus
∫
π‘£πœ† (π‘₯) = − cos π‘₯
0
π‘₯
∫
πœ†π‘“πœ† (𝑠) sin 𝑠𝑑𝑠 + sin π‘₯
+ π‘£πœ† (0) cos π‘₯.
0
π‘₯
πœ†π‘“πœ† (𝑠) cos 𝑠𝑑𝑠
By definition we have
𝑒(πœ†, πœƒ; π‘₯) ≥ πœ€0 ,
∣
π‘₯ ∈ [Θ1 , 𝑇 (πœ†, πœƒ)],
π‘Ž(π‘₯)
∣ ≤ 𝐢(𝛿),
𝑒(πœ†, πœƒ; π‘₯)
11
π‘₯ ∈ [Θ1 , 𝑇 (πœ†, πœƒ)].
(3.18)
(3.19)
(3.20)
(3.21)
It follows that
βˆ£π‘“πœ† (π‘₯)∣ ≤ 𝐢(𝛿),
π‘₯ ∈ [0, 𝑇 (πœ†, πœƒ) − Θ1 ].
Hence from (3.21) we deduce that
π‘£πœ† (πœƒ − Θ1 ) = π‘£πœ† (0) cos(πœƒ − Θ1 ) + π‘œ(1),
and
π‘£πœ†′ (πœƒ − Θ1 ) = −π‘£πœ† (0) sin(πœƒ − Θ1 ) + π‘œ(1),
πœ†→0
(3.22)
πœ† → 0.
(3.23)
On the other hand, from (3.20) we infer that
π‘£πœ† (πœƒ − Θ1 ) = sin πœƒ + π‘œ(1) πœ† → 0
(3.24)
holds uniformly in πœƒ. Therefore, inserting (3.24) into (3.22) we have
π‘£πœ† (0) cos(πœƒ − Θ1 ) = sin πœƒ + π‘œ(1),
πœ†→0
(3.25)
uniformly in πœƒ, too.
Now we separate two cases:
(1) πœƒ ≤ 8𝛿 . In this case, using π‘£πœ† (0) ∼ 1 and (3.25), one can find πœ†1 (𝛿) such that
𝛿
∣ cos(πœƒ − Θ1 )∣ = sin ,
4
πœ† ≤ πœ†1 (𝛿).
This shows that
πœ‹
𝛿
− (πœƒ − Θ1 )∣ ≤ , πœ† ≤ πœ†1 (𝛿).
2
4
𝛿
(2) πœƒ ≥ 8 . In this case, we see that, as πœ† → 0,
∣
π‘£πœ†′ (πœƒ − Θ1 ) = πœ†π‘’′ (πœ†, πœƒ; 0)
(πœ†−2 − πœ†2 ) sin πœƒ cos πœƒ
= πœ†√
= − cos πœƒ + π‘œ(1)
πœ†2 cos2 πœƒ + πœ†−2 sin2 πœƒ
(3.26)
(3.27)
uniformly in πœƒ. Combining (3.23) with (3.27) we arrive at
π‘£πœ† (0) sin(πœƒ − Θ1 ) = − cos πœƒ + π‘œ(1).
(3.28)
Thus as a consequence of (3.25) and (3.28), we can find πœ†2 (𝛿) such that
∣
𝛿
πœ‹
− (πœƒ − Θ1 )∣ ≤ ,
2
4
πœ† ≤ πœ†2 (𝛿)
(3.29)
and βˆ£π‘£πœ† (0) − 1∣ ≤ 2𝛿 . Combining (3.26) with (3.29) we get (3.17). This completes the
proof of (3.17) and (3.14).
β–‘
In the next section we will perform the blow-up analysis at a minimum point
of 𝑒(πœ†, πœƒ : π‘₯) as πœ† → 0 for all πœƒ ∈ [0, πœ‹). (The reason for performing blow-up
12
analysis at a minimum point is only for the simplicity of notation since one can do
the same process at a maximum point as well.) As shown before, the first critical
point of 𝑒(πœ†, πœƒ; π‘₯) in π‘₯ ≥ 0 is a minimum point if πœƒ ∈ [0, πœ‹2 ). But this is not the
case if πœƒ ∈ [ πœ‹2 , πœ‹), so we need to define the function Θ(πœ†, πœƒ) slightly different. Let
Θ2 (πœ†, πœƒ) be the first critical point of 𝑒(πœ†, πœƒ; π‘₯) in π‘₯ ≥ 0 which is a maximum point as
πœƒ ∈ [ πœ‹2 , πœ‹), and let Θ(πœ†, πœƒ) be the critical point of 𝑒(πœ†, πœƒ; π‘₯) next to Θ2 (πœ†, πœƒ). Clearly,
Θ(πœ†, πœƒ) is a minimum point and we have
π‘š(πœ†, πœƒ) = 𝑒(πœ†, πœƒ; Θ(πœ†, πœƒ)) ∼ πœ†,
𝑀 (πœ†, πœƒ) = 𝑒(πœ†, πœƒ; Θ2 (πœ†, πœƒ)) ∼ πœ†−1 .
By the same argument, one can show that the estimate (3.13) holds for πœƒ ∈ [ πœ‹2 , πœ‹).
Thus we have
Proposition 3.5 Let 𝑒(πœ†, πœƒ; π‘₯) be as above. Then for πœ† << 1, there exists a function Θ(πœ†, πœƒ) which is a local minimum of 𝑒(πœ†, πœƒ; π‘₯) such that π‘š(πœ†, πœƒ) = 𝑒(πœ†, πœƒ; Θ(πœ†, πœƒ)) ∼
πœ† and
Θ(πœ†, πœƒ) = πœƒ + π‘œ(1)
(3.30)
holds uniformly in πœƒ ∈ [0, πœ‹).
Remark 3.6 One can show that Θ(πœ†, πœƒ) is the first local minimum of 𝑒(πœ†, πœƒ; π‘₯) in
π‘₯ ≥ 0, which is 𝐢 1 in (πœ†, πœƒ) via the implicit function theorem. But one should
note that Θ(πœ†, πœƒ + πœ‹) βˆ•= Θ(πœ†, πœƒ) + πœ‹ in general. However, we will see that (3.30) is
sufficient for the degree argument in Section 5.
4
Asymptotic Expansion of β„±(πœ†, πœƒ)
In this section, following the argument in [21], we give an asymptotic expansion of
the solution β„±(πœ†, πœƒ) given in the last section. This is crucial in the computation of
the degree of β„±(πœ†, πœƒ) in the next section. We note that in [21], the blow-up argument
was used to obtain a priori estimates of 2πœ‹-periodic solutions of (2.1). An important
observation here is that the argument not only works for the periodic solutions of
(2.1), but also for all solutions of (2.1).
Let π‘š(πœ†, πœƒ) = 𝑒(πœ†, πœƒ; Θ(πœ†, πœƒ)) and
(
)1
π‘ˆπœ€ (π‘₯) = πœ€2 cos2 π‘₯ + πœ€−2 sin2 π‘₯ 2 .
We define a transformation
∫
π‘₯ = Θ(πœ†, πœƒ) + πœ“πœ†,πœƒ (𝑦) = Θ(πœ†, πœƒ) +
1
0
𝑦
1
π‘ˆ 2 πœ€−1 (𝜏 )
π‘‘πœ,
where πœ€(πœ†, πœƒ) = (π‘Ž(Θ(πœ†, πœƒ))− 4 π‘š(πœ†, πœƒ) ∼ πœ†. It induces a rule of transformation of the
equation
π‘Ž(π‘₯)
(4.1)
𝑒′′ + 𝑒 = 3 , π‘₯ ∈ [0, 𝑇 ]
𝑒
13
as follows: let
𝑣(πœ†, πœƒ; 𝑦) = π‘ˆπœ€−1 (𝑦)𝑒(Θ(πœ†, πœƒ) + πœ“πœ†,πœƒ (𝑦)) = π‘ˆπœ€−1 (π‘₯ − Θ(πœ†, πœƒ))𝑒(πœ†, πœƒ; π‘₯),
using
𝑑𝑦
= π‘ˆπœ€−2 (π‘₯ − Θ(πœ†, πœƒ)),
and
𝑑π‘₯
)
𝑑2 𝑦 ( 2
= πœ€ − πœ€−2 sin 2(π‘₯ − Θ(πœ†, πœƒ))π‘ˆπœ€−4 (π‘₯ − Θ(πœ†, πœƒ)),
2
𝑑π‘₯
one can verify that
𝑑2 𝑣(πœ†, πœƒ; 𝑦)
π‘Ž(Θ(πœ†, πœƒ) + πœ“πœ†,πœƒ (𝑦))
+
𝑣(πœ†,
πœƒ;
𝑦)
=
.
𝑑𝑦 2
𝑣 3 (πœ†, πœƒ; 𝑦)
(4.2)
Using equation (4.2), the following result on the asymptotical behavior of 𝑒(πœ†, πœƒ; π‘₯)
as πœ† → 0 was proved in [21], where the periodic solutions was dealt with. The argument is valid for the solution 𝑒(πœ†, πœƒ; π‘₯). Hence the detail of the proof is omitted.
Proposition 4.1 Let π‘Ž be a positive, 𝐢 2 and 2πœ‹-periodic function. Then ∃ 𝐢 > 0
such that
1
≤ 𝑣(πœ†, πœƒ; 𝑦) ≤ 𝐢, βˆ£π‘£π‘¦ (πœ†, πœƒ; 𝑦)∣ ≤ 𝐢, 𝑦 ∈ [−2πœ‹, 2πœ‹],
(4.3)
𝐢
and
βˆ₯𝑣(πœ†, πœƒ; 𝑦) − 𝑣(Θ; 𝑦)βˆ₯𝐢 1 ([−2πœ‹,2πœ‹]) ≤ πΆπœ†2 ∣ log πœ†βˆ£, πœ† → 0,
(4.4)
where 𝑣(Θ; 𝑦) is the 2πœ‹-periodic function given by
{ 1
π‘Ž 4 (Θ),
𝑣(Θ; 𝑦) = ( 1
)1
1
π‘Ž 2 (Θ) sin2 𝑦 + π‘Ž(Θ + πœ‹)π‘Ž− 2 (Θ) cos2 𝑦 2 ,
𝑦 ∈ [− πœ‹2 , πœ‹2 ],
𝑦 ∈ [ πœ‹2 , 3πœ‹
].
2
(4.5)
Remark 4.2 Extending 𝑣(Θ; 𝑦) to a 2πœ‹-periodic function, then one can prove that
the estimate
βˆ₯𝑣(πœ†, πœƒ; 𝑦) − 𝑣(Θ; 𝑦)βˆ₯𝐢 1 ([π‘Ž,𝑏]) ≤ πΆπœ†2 ∣ log πœ†βˆ£
(4.6)
holds for any bounded interval [π‘Ž, 𝑏], but the constant 𝐢 may depend on the interval.
This fact will be used later.
Recall that for (πœ†, πœƒ) ∈ (0, 1] × π•Š11 ,
β„±(πœ†, πœƒ) = (𝐴(πœ†, πœƒ), 𝐡(πœ†, πœƒ)),
where
∫
𝐴(πœ†, πœƒ) =
0
2πœ‹
π‘Ž′ (π‘₯)
cos 2π‘₯𝑑π‘₯,
𝑒2 (πœ†, πœƒ; π‘₯)
14
(4.7)
∫
𝐡(πœ†, πœƒ) =
0
2πœ‹
π‘Ž′ (π‘₯)
sin 2π‘₯𝑑π‘₯.
𝑒2 (πœ†, πœƒ; π‘₯)
To compute the degree we need the asymptotic behavior of β„±(πœ†, πœƒ) as πœ† → 0.
Let Θ(πœ†, πœƒ) be the function in Proposition 3.5, π‘š(πœ†, πœƒ) = 𝑒(πœ†, πœƒ; Θ(πœ†, πœƒ)), πœ€(πœ†, πœƒ) =
1
(π‘Ž(Θ(πœ†, πœƒ))− 4 π‘š(πœ†, πœƒ) ∼ πœ†. For simplicity of notations we omit (πœ†, πœƒ) in 𝑒 and Θ.
Then
∫ 2πœ‹ ′
π‘Ž (π‘₯)
𝐴(πœ†, πœƒ) =
cos 2π‘₯𝑑π‘₯
𝑒2 (π‘₯)
0
∫ 2πœ‹−Θ ′
π‘Ž (π‘₯ + Θ)
=
cos 2(π‘₯ + Θ)𝑑π‘₯
𝑒2 (π‘₯ + Θ)
−Θ
∫ 2πœ‹−Θ ′
∫ 2πœ‹−Θ ′
π‘Ž (π‘₯ + Θ)
π‘Ž (π‘₯ + Θ)
= cos 2Θ
cos 2π‘₯𝑑π‘₯ − sin 2Θ
sin 2π‘₯𝑑π‘₯
2
𝑒 (π‘₯ + Θ)
𝑒2 (π‘₯ + Θ)
−Θ
−Θ
and
∫
2πœ‹
π‘Ž′ (π‘₯)
sin 2π‘₯𝑑π‘₯
𝑒2 (π‘₯)
0
∫ 2πœ‹−Θ ′
π‘Ž (π‘₯ + Θ)
sin 2(π‘₯ + Θ)𝑑π‘₯
=
𝑒2 (π‘₯ + Θ)
−Θ
∫ 2πœ‹−Θ ′
∫ 2πœ‹−Θ ′
π‘Ž (π‘₯ + Θ)
π‘Ž (π‘₯ + Θ)
= sin 2Θ
cos 2π‘₯𝑑π‘₯ + cos 2Θ
sin 2π‘₯𝑑π‘₯.
2
𝑒 (π‘₯ + Θ)
𝑒2 (π‘₯ + Θ)
−Θ
−Θ
𝐡(πœ†, πœƒ) =
Using the asymptotic expansion given by Proposition 4.1 now we prove
Proposition 4.3 Let π‘Ž be a positive, 𝐢 2 and 2πœ‹-periodic function. Then as πœ† → 0,
we have
∫ 2πœ‹−Θ ′
π‘Ž (π‘₯ + Θ)
π‘Ž′ (Θ)
π‘Ž′ (πœ‹ + Θ)
√
√
cos
2π‘₯𝑑π‘₯
=
+
+ 𝑂(πœ†2 ∣ log πœ†βˆ£),
(4.8)
2 (π‘₯ + Θ)
𝑒
π‘Ž(Θ)
π‘Ž(πœ‹
+
Θ)
−Θ
and
∫
2πœ‹−Θ
π‘Ž′ (π‘₯ + Θ)
sin 2π‘₯𝑑π‘₯
𝑒2 (π‘₯ + Θ)
−Θ
)
∫ πœ‹ ( ′
′
′
′
2
π‘Ž
(π‘₯
+
Θ)
+
π‘Ž
(π‘₯
+
πœ‹
+
Θ)
−
π‘Ž
(Θ)
−
π‘Ž
(πœ‹
+
Θ)
sin 2π‘₯
=π‘š2 (πœ†, πœƒ)
𝑑π‘₯ + π‘œ(πœ†2 ).
2
πœ‹
sin
π‘₯
−2
(4.9)
The proof is similar to that of Proposition 3.1 in [21].
Proof. Let
∫ 𝑦
1
π‘₯ = πœ“πœ†,πœƒ (𝑦) =
π‘‘πœ,
−2 cos2 𝜏 + πœ€2 sin2 𝜏
0 πœ€
15
1
𝑣(πœ†, πœƒ; 𝑦) = (πœ€−2 cos2 𝑦 + πœ€2 sin2 𝑦) 2 𝑒(Θ(πœ†, πœƒ) + πœ“πœ†,πœƒ (𝑦)).
Then
cos2 𝑦 − πœ€4 sin2 𝑦
,
cos2 𝑦 + πœ€4 sin2 𝑦
2πœ€2 sin 𝑦 cos 𝑦
sin 2π‘₯ =
.
cos2 𝑦 + πœ€4 sin2 𝑦
Set Θ = πœ“πœ†,πœƒ (Θ1 ). Then using the change of variable π‘₯ = πœ“πœ†,πœƒ (𝑦) and Proposition
4.1 we obtain
∫ 2πœ‹−Θ ′
π‘Ž (π‘₯ + Θ)
cos 2π‘₯𝑑π‘₯
𝑒2 (π‘₯ + Θ)
−Θ
∫ 2πœ‹−Θ1 ′
π‘Ž (πœ“πœ†,πœƒ (𝑦) + Θ) cos2 𝑦 − πœ€4 sin2 𝑦
=
𝑑𝑦
𝑣 2 (πœ†, πœƒ; 𝑦) cos2 𝑦 + πœ€4 sin2 𝑦
−Θ1
∫ 2πœ‹−Θ1
cos2 𝑦 − πœ€4 sin2 𝑦
π‘Ž′ (πœ“πœ†,πœƒ (𝑦) + Θ)
𝑑𝑦
=
(4.10)
𝑣 2 (Θ; 𝑦) + 𝑂(πœ€2 ∣ log πœ€βˆ£) cos2 𝑦 + πœ€4 sin2 𝑦
−Θ1
∫ 2πœ‹−Θ1 ′
π‘Ž (πœ“πœ†,πœƒ (𝑦) + Θ) cos2 𝑦 − πœ€4 sin2 𝑦
=
𝑑𝑦 + 𝑂(πœ€2 ∣ log πœ€βˆ£)
2 (Θ; 𝑦)
2 𝑦 + πœ€4 sin2 𝑦
𝑣
cos
−Θ1
∫ 3πœ‹ ′
2
π‘Ž (πœ“πœ†,πœƒ (𝑦) + Θ) cos2 𝑦 − πœ€4 sin2 𝑦
=
𝑑𝑦 + 𝑂(πœ€2 ∣ log πœ€βˆ£).
2 (Θ; 𝑦)
2 𝑦 + πœ€4 sin2 𝑦
πœ‹
𝑣
cos
−2
cos 2π‘₯ =
In the above equality we have used the fact that the function 𝑣(Θ; 𝑦) is 2πœ‹-periodic.
Now we estimate the right-hand side of (4.10). First we have
∫ πœ‹ ′
2
′
2
4
2 π‘Ž (πœ“
πœ†,πœƒ (𝑦) + Θ) − π‘Ž (Θ) cos 𝑦 − πœ€ sin 𝑦
∣
π‘‘π‘¦βˆ£
𝑣 2 (Θ; 𝑦)
cos2 𝑦 + πœ€4 sin2 𝑦
− πœ‹2
∫ πœ‹
2
(4.11)
≤𝐢
βˆ£π‘Ž′ (πœ“πœ†,πœƒ (𝑦) + Θ) − π‘Ž′ (Θ)βˆ£π‘‘π‘¦
− πœ‹2
=𝑂(πœ€2 ∣ log πœ€βˆ£),
∫
πœ‹
2
− πœ‹2
π‘Ž′ (Θ) cos2 𝑦 − πœ€4 sin2 𝑦
𝑑𝑦 =
𝑣 2 (Θ; 𝑦) cos2 𝑦 + πœ€4 sin2 𝑦
=
∫
πœ‹
2
− πœ‹2
∫
πœ‹
2
− πœ‹2
′
2πœ€4 sin2 𝑦
π‘Ž′ (Θ)
(1 −
)𝑑𝑦
𝑣 2 (Θ; 𝑦)
cos2 𝑦 + πœ€4 sin2 𝑦
π‘Ž′ (Θ)
𝑑𝑦 + 𝑂(πœ€2 )
𝑣 2 (Θ; 𝑦)
(4.12)
π‘Ž (Θ)
+ 𝑂(πœ€2 ).
=√
π‘Ž(Θ)
It follows from (4.11) and (4.12) that
∫ πœ‹ ′
2
2
4
2 π‘Ž (πœ“
π‘Ž′ (Θ)
πœ†,πœƒ (𝑦) + Θ) cos 𝑦 − πœ€ sin 𝑦
√
𝑑𝑦
=
+ 𝑂(πœ€2 ∣ log πœ€βˆ£).
2
2
2
4
𝑣 (Θ; 𝑦)
cos 𝑦 + πœ€ sin 𝑦
π‘Ž(Θ)
− πœ‹2
16
(4.13)
Similarly we have
∫
3
πœ‹
2
πœ‹
2
π‘Ž′ (πœ“πœ†,πœƒ (𝑦) + Θ) cos2 𝑦 − πœ€4 sin2 𝑦
π‘Ž′ (Θ + πœ‹)
√
+ 𝑂(πœ€2 ∣ log πœ€βˆ£).
2 𝑑𝑦 =
2
2
4
𝑣 (Θ; 𝑦)
cos 𝑦 + πœ€ sin 𝑦
π‘Ž(Θ + πœ‹)
(4.14)
Inserting (4.13) and (4.14) into (4.10) we get (4.8) as πœ† ∼ πœ€.
Now we prove (4.9). The proof is more involved. First we obtain
∫
2πœ‹−Θ
π‘Ž′ (π‘₯ + Θ)
sin 2π‘₯𝑑π‘₯
𝑒2 (π‘₯ + Θ)
−Θ
∫ 2πœ‹−Θ1 ′
π‘Ž (πœ“πœ†,πœƒ (𝑦) + Θ) 2πœ€2 sin 𝑦 cos 𝑦
=
𝑑𝑦
𝑣 2 (πœ†, πœƒ; 𝑦) cos2 𝑦 + πœ€4 sin2 𝑦
−Θ1
∫ 2πœ‹−Θ1
2πœ€2 sin 𝑦 cos 𝑦
π‘Ž′ (πœ“πœ†,πœƒ (𝑦) + Θ)
=
𝑑𝑦
𝑣 2 (Θ; 𝑦) + 𝑂(πœ€2 ∣ log πœ€βˆ£) cos2 𝑦 + πœ€4 sin2 𝑦
−Θ1
∫ 2πœ‹−Θ1 ′
2πœ€2 sin 𝑦 cos 𝑦
π‘Ž (πœ“πœ†,πœƒ (𝑦) + Θ)
2
𝑑𝑦
=
+ 𝑂(πœ€ ∣ log πœ€βˆ£)) 2
(
𝑣 2 (Θ; 𝑦)
cos 𝑦 + πœ€4 sin2 𝑦
−Θ1
(4.15)
By elementary computations one has
∫
2πœ‹−Θ1
−Θ1
2πœ€2 sin 𝑦 cos 𝑦
βˆ£π‘‘π‘¦ =
∣ 2
cos 𝑦 + πœ€4 sin2 𝑦
∫
3
πœ‹
2
− πœ‹2
∣
2πœ€2 sin 𝑦 cos 𝑦
2
2 βˆ£π‘‘π‘¦ = 𝑂(πœ€ βˆ£π‘™π‘œπ‘”πœ€βˆ£).
2
4
cos 𝑦 + πœ€ sin 𝑦
(4.16)
So from (4.15) and (4.16) we see that
∫
2πœ‹−Θ
π‘Ž′ (π‘₯ + Θ)
sin 2π‘₯𝑑π‘₯
𝑒2 (π‘₯ + Θ)
−Θ
∫ 2πœ‹−Θ1 ′
π‘Ž (πœ“πœ†,πœƒ (𝑦) + Θ) 2πœ€2 sin 𝑦 cos 𝑦
𝑑𝑦 + 𝑂(πœ€4 ∣ log πœ€βˆ£2 )
=
2 (Θ; 𝑦)
2 𝑦 + πœ€4 sin2 𝑦
𝑣
cos
−Θ1
∫ 3πœ‹ ′
2
π‘Ž (πœ“πœ†,πœƒ (𝑦) + Θ) 2πœ€2 sin 𝑦 cos 𝑦
=
𝑑𝑦 + 𝑂(πœ€4 ∣ log πœ€βˆ£2 ).
2 (Θ; 𝑦)
2 𝑦 + πœ€4 sin2 𝑦
πœ‹
𝑣
cos
−2
(4.17)
In order to get (4.9) we express the integral in the right-hand side of (4.17) in
17
−1
terms of π‘₯, that is, we use the change of variable 𝑦 = πœ“πœ†,πœƒ
(π‘₯). Then
∫
∫
πœ‹
2
− πœ‹2
π‘Ž′ (πœ“πœ†,πœƒ (𝑦) + Θ) 2πœ€2 sin 𝑦 cos 𝑦
𝑑𝑦
𝑣 2 (Θ; 𝑦)
cos2 𝑦 + πœ€4 sin2 𝑦
πœ‹
2
πœ€2 sin 2π‘₯
π‘Ž′ (π‘₯ + Θ)
−1
2 𝑑π‘₯
4
2
2
− πœ‹2 𝑣 (Θ; πœ“πœ†,πœƒ (π‘₯)) πœ€ cos π‘₯ + sin π‘₯
∫ πœ‹ ′
2 π‘Ž (π‘₯ + Θ) sin 2π‘₯
πœ€2
= 1
2 𝑑π‘₯
π‘Ž 2 (Θ) − πœ‹2 πœ€4 cos2 π‘₯ + sin π‘₯
∫ πœ‹ ′
2 π‘Ž (π‘₯ + Θ) − π‘Ž′ (Θ)
πœ€2
sin π‘₯ sin 2π‘₯
𝑑π‘₯
= 1
4
πœ‹
sin π‘₯
πœ€ cos2 π‘₯ + sin2 π‘₯
π‘Ž 2 (Θ) − 2
∫ πœ‹ ′
2 π‘Ž (π‘₯ + Θ) − π‘Ž′ (Θ)
πœ€2
= 1
sin 2π‘₯𝑑π‘₯ + π‘œ(πœ€2 )
sin2 π‘₯
π‘Ž 2 (Θ) − πœ‹2
∫ πœ‹ ′
2 π‘Ž (π‘₯ + Θ) − π‘Ž′ (Θ)
2
sin 2π‘₯𝑑π‘₯ + π‘œ(πœ€2 )
=π‘š (πœ†, πœƒ)
2
πœ‹
sin
π‘₯
−2
=
(4.18)
by Lebesgue’s dominated convergent theorem. Similarly, we have
∫
∫
=
=πœ€
3
πœ‹
2
π‘Ž′ (πœ“πœ†,πœƒ (𝑦) + Θ) 2πœ€2 sin 𝑦 cos 𝑦
𝑑𝑦
𝑣 2 (Θ; 𝑦)
cos2 𝑦 + πœ€4 sin2 𝑦
3
πœ‹
2
π‘Ž′ (π‘₯ + Θ)
πœ€2 sin 2π‘₯
𝑑π‘₯
−1
𝑣 2 (Θ; πœ“πœ†,πœƒ
(π‘₯)) πœ€4 cos2 𝑦 + sin2 π‘₯
πœ‹
2
πœ‹
2
2
∫
∫
3
πœ‹
2
π‘Ž′ (π‘₯ + Θ) sin 2π‘₯
− 12
πœ€4 π‘Ž
πœ‹
2
1
(Θ)π‘Ž(Θ + πœ‹) cos2 π‘₯ + π‘Ž 2 (Θ) sin2 π‘₯
3
πœ‹
2
𝑑π‘₯
π‘Ž′ (π‘₯ + Θ) − π‘Ž′ (Θ)
sin π‘₯ sin 2π‘₯
𝑑π‘₯
1
1
4 π‘Ž− 2 (Θ)π‘Ž(Θ + πœ‹) cos2 π‘₯ + π‘Ž 2 (Θ) sin2 π‘₯
πœ‹
sin
π‘₯
πœ€
2
∫ 3πœ‹ ′
2
2
πœ€
π‘Ž (π‘₯ + Θ) − π‘Ž′ (Θ)
= 1
sin 2π‘₯𝑑π‘₯ + π‘œ(πœ€2 )
sin2 π‘₯
π‘Ž 2 (Θ) πœ‹2
∫ πœ‹ ′
2 π‘Ž (π‘₯ + Θ + πœ‹) − π‘Ž′ (Θ + πœ‹)
πœ€2
= 1
sin 2π‘₯𝑑π‘₯ + π‘œ(πœ€2 )
2
πœ‹
sin
π‘₯
2
π‘Ž (Θ) − 2
∫ πœ‹ ′
2 π‘Ž (π‘₯ + Θ + πœ‹) − π‘Ž′ (Θ + πœ‹)
2
=π‘š (πœ†, πœƒ)
sin 2π‘₯𝑑π‘₯ + π‘œ(πœ€2 ).
2
πœ‹
sin
π‘₯
−2
=πœ€
2
(4.19)
Finally, putting (4.17), (4.18) and (4.19) together and taking πœ† ∼ πœ€ into account
we get (4.9). This completes the proof of Proposition 4.3.
β–‘
18
5
Proof of Theorem 1.2
Now we are ready to find solutions of β„±(πœ†, πœƒ) = 0 via the degree theory and prove the
main theorem of this paper, Theorem 1.2. Recall that the map 𝐺(πœƒ) = (𝐴2 (πœƒ), 𝐡2 (πœƒ))
in Theorem 1.2 and β„±(πœ†, πœƒ for fixed πœ† are πœ‹-periodic and defined on π•Š11 = ℝ/πœ‹β„€.
We first need
Lemma 5.1 For 0 < πœ† << 1, we have
β„±(πœ†, πœƒ) = (𝐴(πœ†, πœƒ), 𝐡(πœ†, πœƒ)) βˆ•= 0,
πœƒ ∈ π•Š11 .
(5.1)
Hence 𝑑𝑒𝑔(β„±(πœ†, πœƒ), π•Š11 ) is well-defined.
Proof. This can be easily proved by contradiction. Suppose there is a sequence
πœ†π‘˜ → 0, πœƒπ‘˜ ∈ π•Š11 such that β„±(πœ†π‘˜ , πœƒπ‘˜ ) = 0. Then we have
∫ 2πœ‹−Θπ‘˜ ′
π‘Ž (π‘₯ + Θπ‘˜ )
cos 2π‘₯𝑑π‘₯ = 0
(5.2)
𝑒2 (π‘₯ + Θπ‘˜ )
−Θπ‘˜
and
∫
2πœ‹−Θπ‘˜
−Θπ‘˜
π‘Ž′ (π‘₯ + Θπ‘˜ )
sin 2π‘₯𝑑π‘₯ = 0,
𝑒2 (π‘₯ + Θπ‘˜ )
(5.3)
where Θπ‘˜ = Θ(πœ†π‘˜ , πœƒπ‘˜ ). Then Proposition 4.3, (5.2) and (5.3) yield
π‘Ž′ (πœ‹ + Θπ‘˜ )
π‘Ž′ (Θπ‘˜ )
+√
→0
𝐴2 (Θπ‘˜ ) = √
π‘Ž(Θπ‘˜ )
π‘Ž(πœ‹ + Θπ‘˜ )
and
(5.4)
)
π‘Ž′ (π‘₯ + Θπ‘˜ ) + π‘Ž′ (π‘₯ + πœ‹ + Θπ‘˜ ) − π‘Ž′ (Θπ‘˜ ) − π‘Ž′ (πœ‹ + Θπ‘˜ ) sin 2π‘₯
𝐡2 (Θπ‘˜ ) =
𝑑π‘₯ → 0.
sin2 π‘₯
− πœ‹2
(5.5)
2
2
1
But we know that 𝐴2 (πœƒ) + 𝐡2 (πœƒ) is continuous and does not vanish for πœƒ ∈ π•Š1 , there
is a constant 𝑐0 such that
∫
πœ‹
2
(
𝐴22 (πœƒ) + 𝐡22 (πœƒ) ≥ 𝑐0 > 0,
πœƒ ∈ π•Š11 .
Clearly, (5.6) contradicts to (5.4) and (5.5).
(5.6)
β–‘
Proposition 5.2 For 0 < πœ† << 1, the following equality holds:
𝑑𝑒𝑔(β„±(πœ†, πœƒ), π•Š11 ) = 𝑑𝑒𝑔(𝐺(πœƒ), π•Š11 ) + 1.
Proof of Proposition 5.2: the case 𝐴2 (πœƒ) ≡ 0.
We use homotopy invariance of the degree to prove (5.7). To this end, set
β„±1 (πœ†, πœƒ) = (𝐴2 (πœƒ), πœ†2 𝐡2 (πœƒ)) = (0, πœ†2 𝐡2 (πœƒ))
19
(5.7)
and
𝐻1 (𝑠, πœ†, πœƒ) = 𝑠𝑒−2π‘–πœƒ β„±(πœ†, πœƒ) + (1 − 𝑠)β„±1 (πœ†, πœƒ).
Here we denote π‘’π‘–πœƒ (π‘₯, 𝑦) = (cos πœƒπ‘₯ − sin πœƒπ‘¦, cos πœƒπ‘¦ + sin πœƒπ‘₯). It is obvious that
𝐻1 (𝑠, πœ†, πœƒ) is continuous in (𝑠, πœ†, πœƒ) and πœ‹-periodic in πœƒ.
Claim: For πœ† << 1,
𝐻1 (𝑠, πœ†, πœƒ) βˆ•= 0,
𝑠 ∈ [0, 1], πœƒ ∈ π•Š11 .
Indeed, if 𝐻1 (𝑠, πœ†, πœƒ) = 0, then
𝑠𝑒−2𝑖Θ(πœ†,πœƒ) β„±(πœ†, πœƒ) + (1 − 𝑠)𝑒2𝑖(πœƒ−Θ(πœ†,πœƒ)) (0, πœ†2 𝐡2 (πœƒ)) = 0.
(5.8)
It follows from Proposition 4.3 that the second coordinate equation in (5.8) is
π‘ π‘š2 (πœ†, πœƒ)𝐡2 (Θ(πœ†, πœƒ)) + (1 − 𝑠)πœ†2 𝐡2 (πœƒ) cos 2(πœƒ − Θ(πœ†, πœƒ)) + π‘œ(πœ†2 ) = 0.
Hence
𝐡2 (πœƒ) = 0
2
(5.9)
(5.10)
2
provided by π‘š (πœ†, πœƒ) ∼ πœ† and πœƒ − Θ(πœ†, πœƒ) = π‘œ(1). But we know that 𝐡2 (πœƒ) does not
vanish since π‘Ž is B-nondegenerate and 𝐴2 ≡ 0, a contradiction.
By the definition of the degree, it is easy to see that
𝑑𝑒𝑔(β„±1 (πœ†, πœƒ), π•Š11 ) = 𝑑𝑒𝑔(𝐺(πœƒ), π•Š11 ) = 0.
Then it follows from the claim and the homotopy invariance of the degree that
𝑑𝑒𝑔(𝑒−2π‘–πœƒ β„±(πœ†, πœƒ), π•Š11 ) = 𝑑𝑒𝑔(β„±1 (πœ†, πœƒ), π•Š11 ) = 0.
(5.11)
The left-hand side of (5.11) equals to −1 + 𝑑𝑒𝑔(β„±(πœ†, πœƒ), π•Š11 ). So as a result of (5.11),
𝑑𝑒𝑔(β„±(πœ†, πœƒ), π•Š11 ) = 1 = 𝑑𝑒𝑔(𝐺(πœƒ), π•Š11 ) + 1.
This completes the proof.
β–‘
Next we turn to the case that 𝐴2 (πœƒ) is not identically zero. After shifting, in this
case we may assume 𝐴2 (0) = 𝐴2 (πœ‹) βˆ•= 0. It is more involved since we are not able
to prove 𝐻1 (𝑠, πœ†, πœƒ) βˆ•= 0 anymore as before due to the fact that in the asymptotic
expansion in Proposition 4.3, the orders of πœ† in (4.8) and (4.9) are different. Some
modifications of the homotopy are needed.
Set
β„±2 (πœ†, πœƒ) = (𝐴2 (Θ(πœ†, πœƒ)), πœ†2 𝐡2 (Θ(πœ†, πœƒ)))
and
𝐻2 (𝑠, πœ†, πœƒ) = 𝑠𝑒−2𝑖Θ(πœ†,πœƒ) β„±(πœ†, πœƒ) + (1 − 𝑠)β„±2 (πœ†, πœƒ) := (𝐾1 (𝑠, πœ†, πœƒ), 𝐾2 (𝑠, πœ†, πœƒ)).
It is obvious that 𝐻2 (𝑠, πœ†, πœƒ) is continuous in (𝑠, πœ†, πœƒ), but is not πœ‹-periodic in πœƒ in
general, thus the degree of this map is not well-defined. However, in view of the
asymptotic estimate Θ(πœ†, πœƒ) given by Proposition 3.5, 𝐻2 (𝑠, πœ†, πœƒ) is very close to
a πœ‹-periodic one. This enables us to complete the proof by several steps via the
difference of argument of the map πœƒ → 𝐻2 (𝑠, πœ†, πœƒ) at πœƒ = πœ‹ and πœƒ = 0.
20
Lemma 5.3 For 0 < πœ† << 1,
𝐻2 (𝑠, πœ†, πœƒ) βˆ•= 0,
πœƒ ∈ [0, πœ‹],
𝑠 ∈ [0, 1].
(5.12)
Proof. According to Proposition 4.3,
∫ 2πœ‹−Θ ′
∫ 2πœ‹−Θ ′
π‘Ž (π‘₯ + Θ)
π‘Ž (π‘₯ + Θ)
−2𝑖Θ(πœ†,πœƒ)
𝑒
β„±(πœ†, πœƒ) = (
cos 2π‘₯𝑑π‘₯,
cos 2π‘₯𝑑π‘₯)
2
𝑒 (π‘₯ + Θ)
𝑒2 (π‘₯ + Θ)
−Θ
−Θ
= (𝐴2 (Θ), π‘š2 (πœ†, πœƒ)𝐡2 (Θ)) + (𝑂(πœ†2 ∣ log πœ†βˆ£), π‘œ(πœ†2 )).
Thus we get
𝐻2 (𝑠, πœ†, πœƒ) = (𝐴2 (Θ), (π‘ π‘š2 (πœ†, πœƒ) + (1 − 𝑠)πœ†2 )𝐡2 (Θ)) + (𝑂(πœ†2 ∣ log πœ†βˆ£), π‘œ(πœ†2 )).
The conclusion follows as in Lemma 5.1.
β–‘
For fixed (𝑠, πœ†), we know that the degree of 𝐻2 (𝑠, πœ†, πœƒ) may not be well-defined,
so instead of the degree we consider
˜ 2 (𝑠, πœ†, πœƒ)) = 1
𝑑(𝐻
2πœ‹
∫
πœ‹
0
𝑑
𝑑
−( π‘‘πœƒ
𝐾1 )𝐾2 + ( π‘‘πœƒ
𝐾2 )𝐾1
π‘‘πœƒ.
2
2
𝐾1 + 𝐾2
It is known that the integral on the right-hand side is the difference of argument
˜ 2 (𝑠, πœ†, πœƒ)) is the same
of the map πœƒ → 𝐻2 (𝑠, πœ†, πœƒ) at πœƒ = πœ‹ and πœƒ = 0, and 𝑑(𝐻
as the degree 𝑑𝑒𝑔(𝐻2 (𝑠, πœ†, πœƒ), π•Š11 ) if 𝐻2 (𝑠, πœ†, πœƒ) is πœ‹-periodic in πœƒ. It enjoys similar
properties such as homotopy invariance as the degree.
Lemma 5.4 For 0 < πœ† << 1, we have
˜ 2 (1, πœ†, πœƒ)) = −1 + 𝑑𝑒𝑔(β„±(πœ†, πœƒ), π•Š1 ) + π‘œ(1).
𝑑(𝐻
1
(5.13)
Proof. From the definition we see that
˜ 2 (1, πœ†, πœƒ)) = 𝑑(𝑒
˜ −2𝑖Θ(πœ†,πœƒ) β„±(πœ†, πœƒ))
𝑑(𝐻
Θ(πœ†, πœ‹) − Θ(πœ†, 0) ˜
=−
+ 𝑑(β„±(πœ†, πœƒ))
πœ‹
= −1 + 𝑑𝑒𝑔(β„±(πœ†, πœƒ), π•Š11 ) + π‘œ(1)
thanks to Proposition 3.5 and the fact that β„±(πœ†, πœƒ) is πœ‹−periodic in πœƒ.
(5.14)
β–‘
Lemma 5.5 Let 𝐴2 (0) = 𝐴2 (πœ‹) βˆ•= 0. Then for 0 < πœ† << 1,
˜ 2 (0, πœ†, πœƒ)) = 𝑑(𝐻
˜ 2 (1, πœ†, πœƒ)) + π‘œ(1).
𝑑(𝐻
21
(5.15)
Proof. Let 0 < πœ† << 1 be fixed and 𝑅 = {(𝑠, πœƒ)βˆ£π‘  ∈ [0, 1], πœƒ ∈ [0, πœ‹]}. By
Lemma 5.3 we have
𝐻2 (𝑠, πœ†, πœƒ) = (𝐾1 (𝑠, πœ†, πœƒ), 𝐾2 (𝑠, πœ†, πœƒ)) βˆ•= 0, (𝑠, πœƒ) ∈ 𝑅.
Hence
1
2πœ‹
∫
∂𝑅
−𝑑𝐾1 𝐾2 + 𝑑𝐾2 𝐾1
= 0,
𝐾12 + 𝐾22
(5.16)
the differentials in (5.16) are w.r.t. the variables 𝑠 and πœƒ. It follows from (5.16) that
˜ 2 (0, πœ†, πœƒ)) − 𝑑(𝐻
˜ 2 (1, πœ†, πœƒ))
𝑑(𝐻
∫ 1
1
−𝑑𝐾1 (𝑠, πœ†, 0)𝐾2 (𝑠, πœ†, 0) + 𝑑𝐾2 (𝑠, πœ†, 0)𝐾1 (𝑠, πœ†, 0)
=
2πœ‹ 0
𝐾12 + 𝐾22
∫ 1
1
−𝑑𝐾1 (𝑠, πœ†, πœ‹)𝐾2 (𝑠, πœ†, πœ‹) + 𝑑𝐾2 (𝑠, πœ†, πœ‹)𝐾1 (𝑠, πœ†, πœ‹)
−
2πœ‹ 0
𝐾12 + 𝐾22
=𝐼 + 𝐼𝐼.
(5.17)
By Proposition 4.3 and Θ(πœ†, πœ‹) = πœ‹ + π‘œ(1) we see that for 𝑠 ∈ [0, 1],
𝑑
𝑑
𝐾1 (𝑠, πœ†, 0)∣2 + ∣ 𝐾2 (𝑠, πœ†, 0)∣2
𝑑𝑠
𝑑𝑠
=∣𝐻2 (1, πœ†, 0) − 𝐻2 (0, πœ†, 0)∣2
∣
=βˆ£π‘’−2𝑖Θ(πœ†,πœ‹) β„±(πœ†, πœ‹) − β„±2 (πœ†, πœ‹)∣2 = π‘œ(πœ†4 log2 πœ†)
and
𝐾12 (𝑠, πœ†, 0) + 𝐾22 (𝑠, πœ†, 0) = 𝐴2 (0)2 + π‘œ(1).
Therefore,
∫
∣𝐼∣ ≤
0
Similarly,
1
1
𝑑
𝑑
(∣ 𝑑𝑠
𝐾1 ∣2 + ∣ 𝑑𝑠
𝐾2 ∣2 ) 2
1
(𝐾12 + 𝐾22 ) 2
𝑑𝑠 = π‘œ(1).
∣𝐼𝐼∣ = π‘œ(1).
(5.18)
(5.19)
Now plug (5.18) and (5.19) into (5.17) we obtain (5.15).
β–‘
Lemma 5.6 Let 𝐴2 (0) = 𝐴2 (πœ‹) βˆ•= 0. Then for 0 < πœ† << 1, the following holds:
˜ 2 (0, πœ†, πœƒ)) = 𝑑𝑒𝑔(𝐺(πœƒ), π•Š1 ) + π‘œ(1),
𝑑(𝐻
1
where 𝐺(πœƒ) = (𝐴2 (πœƒ), 𝐡2 (πœƒ)).
22
(5.20)
Proof. From the definition we have
˜ 2 (0, πœ†, πœƒ)) = 𝑑(β„±
˜ 2 (πœ†, πœƒ))
𝑑(𝐻
∫ 𝑑
𝑑
𝐴2 (Θ) ⋅ 𝐡2 (Θ) − π‘‘πœƒ
𝐡2 (Θ) ⋅ 𝐴2 (Θ)
πœ†2 πœ‹ π‘‘πœƒ
=
π‘‘πœƒ
2πœ‹ 0
𝐴2 (Θ)2 + πœ†4 𝐡2 (Θ)2
∫
𝑑
𝑑
𝐴2 (Θ) ⋅ 𝐡2 (Θ) − 𝑑Θ
𝐡2 (Θ) ⋅ 𝐴2 (Θ)
πœ†2 Θ(πœ†,πœ‹) 𝑑Θ
=
𝑑Θ
2
4
(5.21)
2πœ‹ 0
𝐴2 (Θ) + πœ† 𝐡2 (Θ)2
∫
∫
𝑑
𝑑
πœ‹
πœ‹
𝐴2 (Θ) ⋅ 𝐡2 (Θ) − π‘‘Θ π΅2 (Θ) ⋅ 𝐴2 (Θ)
πœ†2
𝑑Θ
=
[
−
] 𝑑Θ
2πœ‹ 0
𝐴2 (Θ)2 + πœ†4 𝐡2 (Θ)2
Θ(πœ†,πœ‹)
= 𝑑𝑒𝑔(𝐺(πœ†, πœƒ), π•Š11 ) + π‘œ(1)
= 𝑑𝑒𝑔(𝐺(πœƒ), π•Š11 ) + π‘œ(1),
where 𝐺(πœ†, πœƒ) = (𝐴2 (πœƒ), πœ†2 𝐡2 (πœƒ)), since 𝐺(πœ†, πœƒ) is πœ‹−periodic in πœƒ. (5.21) is a
consequence of
𝑑𝑒𝑔(𝐺(πœ†, πœƒ), π•Š11 ) = 𝑑𝑒𝑔(𝐺(πœƒ), π•Š11 )
and
2
πœ†βˆ£
∫
πœ‹
Θ(πœ†,πœ‹)
𝑑
𝑑
( 𝑑Θ
𝐴2 (Θ))𝐡2 (Θ) − ( 𝑑Θ
(𝐡2 (Θ))𝐴2 (Θ)
𝑑Θ∣ = π‘œ(1)
𝐴2 (Θ)2 + πœ†4 𝐡2 (Θ)2
due to Θ(πœ†, πœ‹) = πœ‹ + π‘œ(1) and 𝐴2 (πœ‹) βˆ•= 0.
β–‘
Proof of Proposition 5.2: the case 𝐴2 (πœƒ) βˆ•≡ 0.
We assume that 𝐴2 (πœ‹) βˆ•= 0. From Lemma 5.1, for πœ† << 1 we know β„±(πœ†, πœƒ) βˆ•=
(0, 0) and 𝑑𝑒𝑔(β„±(πœ†, πœƒ), π•Š11 ) is well-defined. Combining Lemmas 5.4-5.6 we arrive at
˜
𝑑𝑒𝑔(β„±(πœ†, πœƒ), π•Š11 ) =𝑑(𝐻(1,
πœ†, πœƒ)) + 1 + π‘œ(1)
˜
=𝑑(𝐻(0,
πœ†, πœƒ)) + 1 = π‘œ(1)
=𝑑𝑒𝑔(𝐺(πœƒ), π•Š11 )
(5.22)
+ 1 + π‘œ(1).
But both 𝑑𝑒𝑔(β„±(πœ†, πœƒ), π•Š11 ) and 𝑑𝑒𝑔(𝐺(πœƒ), π•Š11 ) are integers, there must hold
𝑑𝑒𝑔(β„±(πœ†, πœƒ), π•Š11 ) = 𝑑𝑒𝑔(𝐺(πœƒ), π•Š11 ) + 1.
(5.23)
This finishes the proof.
β–‘
Proof of Theorem 1.2: Set Λ = 1 − πœ†. Then as Λ → 1, the degree 𝑑𝑒𝑔(β„±(1 −
Λ, πœƒ); π•Š11 , 0) is well-defined. We define a continuous map from 𝐷(Λ) = {(𝑋, π‘Œ ) ∈
ℝ2 βˆ£π‘‹ 2 + π‘Œ 2 ≤ Λ2 } to ℝ2 as
𝒒(𝑋, π‘Œ ) = (𝐴(1 − Λ, πœƒ), 𝐡(1 − Λ, πœƒ)) = β„±(1 − Λ, πœƒ),
where (𝑋, π‘Œ ) = (Λ cos πœƒ, Λ sin πœƒ). This is a well-defined and continuous map due to
the fact that β„±(πœ†, πœƒ) is independent of πœƒ when πœ† = 1 or Λ = 0. It follows from the
assumption of Theorem 1.2, 𝑑𝑒𝑔(β„±(1 − Λ, πœƒ); π•Š11 , 0) βˆ•= 0, hence by the basic property
of degree, see [14], β„±(1 − Λ, πœƒ) = 0 has a solution (Λ, πœƒ). The proof is finished. β–‘
The following result, which was proved by a different approach in [21], is an
immediate consequence of Theorem 1.2.
23
Corollary 5.7 Let π‘Ž be a positive, 𝐢 2 and 2πœ‹-periodic and B-nondegenerate function. Suppose that
min 𝐡2 (πœƒ) > 0,
𝐴2 (πœƒ)=0
or max 𝐡2 (πœƒ) < 0,
𝐴2 (πœƒ)=0
then equation
𝑒′′ + 𝑒 =
π‘Ž(π‘₯)
𝑒3
(5.24)
(5.25)
has a 2πœ‹-periodic solution.
Proof: Let (5.24) be satisfied. Then it is rather straightforward to see
𝑑𝑒𝑔(𝐺(πœƒ), π•Š11 ) = 0.
(5.26)
Consequently, Theorem 1.2 can be applied, and equation (5.25) has a 2πœ‹-periodic
solution.
β–‘
We conclude with a generalization of Theorem 1.2. Consider π‘›πœ‹-periodic solution
of equation (5.25), where π‘Ž is a positive, 𝐢 2 and π‘›πœ‹-periodic function. The same
argument works for this case. To state the result, as in 2πœ‹-periodic case, we need
two functions 𝐴𝑛 and 𝐡𝑛 given by
′
𝑛
∑
π‘Ž (πœƒ + (𝑗 − 1)πœ‹)
√
𝐴𝑛 (πœƒ) =
π‘Ž(πœƒ + (𝑗 − 1)πœ‹)
𝑗=1
and
∫
𝐡𝑛 (πœƒ) =
πœ‹
2
− πœ‹2
( ∑𝑛
𝑗=1
′
π‘Ž (πœƒ + 𝑑 + (𝑗 − 1)πœ‹) −
∑𝑛
sin2 𝑑
𝑗=1
)
′
π‘Ž (πœƒ + (𝑗 − 1)πœ‹) sin 2𝑑
(5.27)
𝑑𝑑. (5.28)
They are πœ‹−periodic. A function π‘Ž is called 𝐡-nondegenerate if 𝐡𝑛 (πœƒ) does not
vanish whenever 𝐴𝑛 (πœƒ) = 0. Let 𝐺 : π•Š11 → ℝ2 , 𝐺(πœƒ) = (𝐴𝑛 (πœƒ), 𝐡𝑛 (πœƒ)), πœƒ ∈ π•Š11 . We
know that 𝑑𝑒𝑔(𝐺, π•Š11 ) if π‘Ž is 𝐡-nondegenerate. Then we have
Theorem 5.8 Let π‘Ž be a positive, 𝐢 2 and π‘›πœ‹-periodic and B-nondegenerate function. Then equation (5.25) has an π‘›πœ‹−periodic solution provided that 𝑑𝑒𝑔(𝐺, π•Š11 ) βˆ•=
−1.
Acknowledgments: This research was finished when the first author visited the
Department of Mathematics, The Chinese University of Hong Kong. He thanks the
institute for the hospitalities and providing excellent work conditions. The research
of the first author is partially supported by grants from NSFC, that of the second
author is partially supported by an Earmarked Grant from RGC of Hong Kong.
24
References
[1] U. Abresh, J. Langer, The normalized curved shortening flow and homothetic
solutions, J. Diff. Geom., 23(1986), 175-196.
[2] J. Ai, K.S. Chou, J. Wei, Self-similar solutions for the anisotropic affine curve
shortening problem, Calc. Var., 13(2001), 311-337.
[3] S. Altschuler, Singularities of the curve shrinking flow for space curves, J. Diff.
Geom., 34(1991), 491-514.
[4] B. Andrews, Contraction of convex hypersurfaces by their affine normal, J.
Diff. Geom., 43(1996), 207-230.
[5] B. Andrews, Evolving convex curves, Calc. Var., 7(1998), 315-371.
[6] S. Angenent, On the formation of singularities in the curve shortening flow,
J. Diff. Geom., 33(1991), 601-634.
[7] S. Angenent, M. E. Gurtin, Multiphase thermodynamics with interfacial
structure evolution of an isothermal interface, Arch. Rational Mech. Anal.,
108(1989), 323-391.
[8] W.X. Chen, 𝐿𝑝 -Minkowski problem with not necessarily positive data, Adv.
in Math., 201(2006),77-89.
[9] K.S. Chou, L. Zhang, On the uniqueness of stable ultimate shapes for the
anisotropic curve-shorting problem, Manuscripta Math., 102(2000), no.1, 101110.
[10] K.S. Chou, X.P. Zhu, Anisotropic flows for convex plane curves, Duke Math.
J., 97(1999), 579-619.
[11] M. del Pino, R. Manásevich, A. Montero, 𝑇 -periodic solutions for some second
order differential equation with singularities, Proc. Roy. Soc. Edinburgh, Sect.
A, 120(1992), 231-243.
[12] C. Dohmen, Y. Giga, Self-similar shrinking curves for anisotropic curvature
flow equations, Proc. Japan Acad., Ser. A, 70(1994), 252-255.
[13] C. Dohmen, Y. Giga, N. Mizoguchi, Existence of self-similar shrinking curves
for anisotropic curvature flow equations, Calc. Var., 4(1996), 103-119.
[14] I. Fonseca and W. Gangbo, Degree theory in analysis and applications, Oxford
Science Publications, 1995.
[15] M. E. Gage, Evolving plane curve by curvature in relative geometries, Duke
Math. J., 72(1993), 441-466.
25
[16] M. E. Gage, R. Hamilton, The heat equation shrinking convex plane curves, J.
Diff. Geom. 23(1996), 69-96.
[17] M. E. Gage, Y. Li, Evolving plane curve by curvature in relative geometries
II, Duke Math. J. 75(1994),79-98.
[18] M. Grayson, The heat equation shrinking embedded curves to round points, J.
Diff. Geom., 26(1987), 285-314.
[19] M. E. Gurtin, Thermodynamics of evolving phase boundaries in the plane,
Clarendon Press, Oxford 1993.
[20] M.-Y. Jiang, Remarks on the 2-dimensional 𝐿𝑝 -Minkowski problem, Advanced
Nonlinear Studies, 10(2010), 297-313.
[21] M.-Y. Jiang, L. Wang, J. Wei, 2πœ‹-periodic self-similar solutions for the
anisotropic affine curve shortening problem, Calc. Var., 41(2011), 535-565.
[22] H. Matano, J. Wei, On anisotropic curvature flow equations, preprint.
[23] G. Sapiro, A. Tannenbaum, On affine plane curve evolution, J. Funct. Anal.,
119(1994), 79-120.
26
Download