• Aucun résultat trouvé

centroaffine Minkowski problem A remark on rotationally symmetric solutions to the ScienceDirect

N/A
N/A
Protected

Academic year: 2022

Partager "centroaffine Minkowski problem A remark on rotationally symmetric solutions to the ScienceDirect"

Copied!
38
0
0

Texte intégral

(1)

ScienceDirect

J. Differential Equations 266 (2019) 4394–4431

www.elsevier.com/locate/jde

A remark on rotationally symmetric solutions to the centroaffine Minkowski problem

Jian Lu

DepartmentofAppliedMathematics,ZhejiangUniversityofTechnology,Hangzhou310023,China Received 1April2018

Availableonline 6November2018

Abstract

Inthispaperwestudythe solvabilityoftherotationallysymmetriccentroaffineMinkowskiproblem.

Bydelicateblow-upanalyses,weremoveatechnicalconditionintheexistenceresultobtainedbyLuand Wang [30].

©2018ElsevierInc.Allrightsreserved.

MSC:35J96;35J75;53A15;53A07

Keywords:Monge–Ampèreequation;CentroaffineMinkowskiproblem;Existenceofsolutions;Blow-upanalyses

1. Introduction

Given a convex body Xin the Euclidean space Rn+1containing the origin, the centroaffine curvatureof ∂Xat point pis by definition equal to K/dn+2, where Kis the Gauss curvature and d is the distance from the origin to the tangent hyperplane of ∂Xat p. The centroaffine curvature is invariant under unimodular linear transforms in Rn+1and has received much attention in ge- ometry [36,37]. The centroaffine Minkowski problem[11] is a prescribed centroaffine curvature problem, which in the smooth case is equivalent to solving the following Monge–Ampère type equation

ThisworkwassupportedbyNaturalScienceFoundationofChina(11871432,11401527).

E-mailaddress:lj-tshu04@163.com.

https://doi.org/10.1016/j.jde.2018.09.034

0022-0396/©2018ElsevierInc.Allrightsreserved.

(2)

det(∇2H+H I )= f

Hn+2 onSn, (1)

where f is a given positive function, His the support function of a bounded convex body Xin Rn+1, I is the unit matrix, ∇2H =(ijH )is the Hessian matrix of covariant derivatives of H with respect to an orthonormal frame on Sn. When f is a constant, this equation describes affine hyperspheres of elliptic type, and all its solutions are ellipsoids centered at the origin [8].

Equation (1) is also the special case of the Lp-Minkowski problem with p= −n −1. The Lp-Minkowski problem, introduced by Lutwak [31], is an important generalization of the clas- sical Minkowski problem, and is a basic problem in the Lp-Brunn–Minkowski theory in modern convex geometry. It has attracted great attention over the last two decades, see e.g. [5,6,10,11,14, 16,18–21,26,32–34,39,40,42,44,46] and references therein.

Equation (1) naturally arises in anisotropic Gauss curvature flows and describes their self- similar solutions [4,7,12,17,41]. Besides, its parabolic form can be used for image processing [2].

Eq. (1) can be reduced to a singular Monge–Ampère equation in the half Euclidean space Rn++1, the regularity of which was strongly studied in [22,23].

Equation (1) corresponds to the critical case of the famous Blaschke–Santaló inequality in convex geometry [35]:

vol(X)inf

ξX

1 n+1

Sn

dS(x)

(H (x)ξ·x)n+1κn2+1, (2) where X is any convex body in Rn+1, vol(X)is the volume of the convex body X, H is the support function of X, and κn+1is the volume of the unit ball in Rn+1. Also Eq. (1) remains invariant under projective transforms on Sn[11,30]. When f is a constant function, it only has constant solutions up to projective transformations. This result has been known for a long time, see e.g. [8], which implies that there is no a priori estimates on solutions for general f with- out additional assumptions. Besides, Chou and Wang [11] found an obstruction for solutions to Eq. (1), which means it may have no solution for some f. On the other hand, it may also have many solutions for some f [15]. This situation is similar, in some aspects, to the prescribed scalar curvature problem on Sn, which involves critical exponents of Sobolev inequalities and the Kazdan–Warner obstruction [9,38]. So the solvability of Eq. (1) is a rather complicated problem due to these features.

For n =1, the existence of solutions to Eq. (1) was investigated in [1,3,10,12,13,24,25,40,43].

In general, one needs to impose some non-degenerate and topological degree conditions on f to obtain an existence result.

For higher n-dimension, only several special cases were studied, see [29,30] for the rotation- ally symmetric case, [27] for a generalized rotationally symmetric case, [21] for the mirror- symmetric case, and [45] for the discrete case. In these papers, sufficient conditions for the existence of solutions can be found. However, the solvability of Eq. (1) for a general f is still open.

In this paper, we are only concerned about the rotationally symmetric case of Eq. (1). That is, the given function f and solutions H are assumed to be rotationally symmetric with respect to the xn+1-axis in Rn+1with n ≥1. In the spherical coordinates, a rotationally symmetric function f on Sncan be regarded as a function on [0, π], such that

f (θ ):=f (x1,· · ·, xn+1)withxn+1=cosθ.

(3)

In particular, f (0)and f (π )are values of f at the north and south poles respectively. By the correspondence xn+1=cosθ, one can naturally extend f (θ )on [0, π]to be a 2π-periodic and even function on R. Observe that if f ∈Cm(Sn)for some integer m, then f ∈Cm(R). Using the superscript denotes d, we have f(0) =f(π ) =0 if it is differentiable. Throughout this paper, we will always use these conventions.

A typical existence result about the rotationally symmetric case of Eq. (1) was first established in [30] and then supplemented in [29]. To state this result, we introduce two quantities:

ni(f )=

⎧⎪

⎪⎪

⎪⎪

⎪⎩

f(π2), n≥2, π

0

[f(θ )f(π2)]tanθdθ, n=1,

and

pi(f )= π

0

f(θ )cotθdθ.

Theorem A ([29,30]). Assume that fC2(Sn) (requiringC6forn =2), and that f is positive and rotationally symmetric. If f(π2) =0and ni(f ) ·pi(f ) <0, then Eq.(1)admits a rotation- ally symmetric solution.

The assumption f(π2) =0 in the above theorem is not essential, but used to reduce some difficulties in blow-up analyses. It was showed in [29] that this assumption can be removed when fis very close to a positive constant. The aim of this paper is to remove this technical assumption in a general case.

For n =1, 2, we follow the arguments in [29,30], carry out more delicate analyses, and then remove the condition f(π2) =0 completely.

Theorem 1. Assume that fC2(S1)or fC2,α(S2)for some α(0, 1), and that f is positive and rotationally symmetric. If ni(f ) ·pi(f ) <0, then Eq.(1)admits a rotationally symmetric solution.

For n ≥3, the above method is no longer applicable. Inspired by [27], we carry out blow-up analyses for a variational method to obtain the following

Theorem 2. Assume that fC2(Sn)with n ≥3, and that f is positive and rotationally symmet- ric. If ni(f ) <nn++12f(π2)2/f (π2)and pi(f ) >0, then Eq.(1)admits a rotationally symmetric solution.

We see in the case n ≥3, a little more restriction on ni(f )will be needed when the assump- tion f(π2) =0 is removed. However if f(π2) =0, Theorem2just becomes into the existence theorem [30, Theorem 1.3].

The paper is organized as follows. In section2, we provide some basic facts about Eq. (1) and convex bodies. Then we prove Theorem1and 2in section3and section4respectively.

(4)

2. Preliminaries

In this section we state some properties about Eq. (1) and a few facts in convex geometry, which will be used throughout this paper. One can consult [37] for more knowledge about convex geometry.

An obstruction for solutions to Eq. (1) was found by Chou and Wang [11].

Lemma 1 ([11]). Let Hbe a C3-solution to equation (1). Then we have

Sn

ξf

Hn+1=0 (3)

for any projective vector field ξ, given by

ξ(x)=Bx(xTBx)x, xSn, where Bis an arbitrary matrix of order n +1.

In the rotationally symmetric case, (3) is reduced to π

0

f(θ )sinnθcosθ

Hn+1(θ ) =0. (4)

See [30, Proposition 3.1].

We have a volume estimate for any solution to Eq. (1).

Lemma 2 ([30]). There exist positive constants Cn, C˜n, depending only on n, such that for any solution Hto Eq.(1), we have

Cn

fmin≤vol(H )≤ ˜Cn fmax, where fmin=inf

Snf, fmax=sup

Sn

f, and vol(H ) is the volume of the convex body determined byH.

Let X be any convex body in Rn+1, and H be its support function. Under the action of a unimodular linear transform AT ∈SL(n +1), Xbecomes into another convex body XA:=ATX.

Denote the support function of XAby HA. Then HA(x)= |Ax| ·H

Ax

|Ax| , xSn. (5) See e.g. [30, (2.11)].

We remark that if H is a solution to Eq. (1), then HAis a solution to the following equation det(∇2HA+HAI )= fA

HAn+2, fA(x)=f Ax

|Ax| . (6)

See [11] for more details.

(5)

Related to the linear transform, there is an integral variable substitution formula.

Lemma 3 ([28]). For any integral function gon Sn, and any matrix A ∈GL(n +1), we have the following variable substitution for integration:

Sn

g(y)dS(y)=

Sn

g Ax

|Ax|

· |detA|

|Ax|n+1dS(x).

By this lemma and (5), we see for any unimodular linear transform A ∈SL(n +1), there is

Sn

f Hn+1=

Sn

fA

HAn+1, (7)

where fAis the same as in (6).

John’s Lemma in convex geometry says that for any non-degenerate convex body Xin Rn+1, there is a unique ellipsoid Ewhich attains the minimum volume among all ellipsoids contain- ingX. This ellipsoid Eis called the minimum ellipsoidof X. It satisfies

1

n+1EXE,

where λE= {x0+λ(xx0) :xE}with x0the center of E. We say Xis normalizedif the E is a ball.

We denote the area of Snby σn, and the unit vector along xi-axis by eifor i=1, 2, · · ·, n +1.

3. Proof of Theorem1

In this section, we prove Theorem 1. To achieve this, one needs an improvement of [30, Theorem 1.2].

Theorem 3. Assume that fC2(S1)or fC2,α(S2)for some α(0, 1), and that f is positive and rotationally symmetric. If ni(f ) ·pi(f ) =0, then there exist positive constants C,C˜ de- pending only on nand f, such that for any rotationally symmetric solution H to Eq.(1), we have

CH≤ ˜C.

Once we have Theorem3, we can repeat the arguments of [29] to prove that Eq. (1) admits a rotationally symmetric solution provided ni(f ) ·pi(f ) <0, yielding Theorem 1. One can consult [29] for details. So in the rest of this section, we are only concerned about Theorem3, and give its proof.

Let {Hk}be any sequence of rotationally symmetric solutions to Eq. (1). For each Hk, define ak∈Rand Ak∈SL(n +1)as

ak=f (π2)12/Hk(π2)n+1, Ak=diag

a

n+11

k ,· · ·, a

n+11

k , a

n+1n

k

.

(8)

(6)

Let

HAk(x)= |Akx| ·Hk Akx

|Akx|

, xSn. (9)

Then HAk is a rotationally symmetric solution to Eq. (6) with Areplaced by Ak. Note that HAk(π2)=a

1 n+1

k Hk(π2)=f (π2)2n+21 . (10) Lemma 4. There exist positive constants C, C˜ depending only on n, fmaxand fmin, such that

CHAk≤ ˜C. (11)

Proof. By the rotational symmetry of HAk, one can easily see that vol(HAk)κn

n+1HAk(π2)n[HAk(0)+HAk(π )], (12) and

maxHAk

HAk(π2)2+ [HAk(0)+HAk(π )]2. (13) Recalling HAk satisfies equation (6), by the volume estimate given in Lemma2, we have

vol(HAk)Cn

maxfAk=Cn

fmax, (14)

which together with (12) yields

HAk(0)+HAk(π )≤ ˜Cn

fmax·HAk(π2)n

= ˜Cn

fmax·f (π2)2n+n2

≤ ˜Cnf

1

max2 f

n 2n+2

min , where we have used (10) for the equality. Now from (13) we obtain

maxHAkf

1 2n+2

max + ˜Cnf

1

max2 f

n 2n+2

min , (15)

which means the second inequality in (11) is true.

On the other hand, by virtue of [30, Lemma 2.3], there is

minHAk·(maxHAk)n·vol(HAk)Cnfmin.

Combining it with (14) and (15), we easily obtain the first inequality in (11). 2

(7)

To obtain uniform upper and lower bounds for {Hk}, by (9) and Lemma4, we should exclude two cases, namely ak→ +∞or ak→0+when k→ +∞. The second case can be still solved by the method developed in [30]. But for the first case where ak→ +∞, one needs more delicate analyses to deal with. The following are details.

First note that in the rotationally symmetric case, fAk defined in (6) can be written as

fAk(θ )=f (γak(θ )), (16)

where

γak(θ )=arccos cosθ

iak(θ ) , iak(θ )=

ak2sin2θ+cos2θ , (17) see [30, (3.3)–(3.4)].

Lemma 5. Assume ak → +∞ when k→ +∞. Then HAk converges to the constant function f (π2)2n1+2 uniformly on [0, π].

Proof. From Lemma4, we see HAk

is uniformly bounded. By the Blaschke selection theorem, one may assume that

HAk

converges uniformly to some support function Hon Sn, which is also rotationally symmetric. It remains to prove that

Hf (π2)2n1+2 onSn. (18) Recall Eq. (6), namely

det(∇2HAk+HAkI )= fAk

HAn+2

k

onSn. (19)

Note when ak→ +∞, fAkconverges to f (π2)almost everywhere on [0, π], see (16). Passing to the limit in Eq. (19), we see His a generalized solution to

det(∇2H+HI )=f (π2) Hn+2 onSn.

So His an elliptic affine sphere, which must be an ellipsoid [8]. By the rotational symmetry of H, it should be expressed as

H(x)=f (π2)2n1+2|x|, xSn (20) for some ∈SL(n +1)of form

=diag

λn+11,· · ·, λn+11, λn+n1

withλ >0.

Then

H(π2)=f (π2)2n+21 λn+11 .

(8)

On the other hand, recalling (10), we have

H(π2)=f (π2)2n+21 .

Hence λ =1, namely is the identity matrix of order n +1. Now (20) is simplified into (18).

The proof of this lemma is completed. 2

Recall HAksatisfies equation (19), which in the rotationally symmetric case can be simplified into the following form:

(HA

k+HAk)(HA

kcotθ+HAk)n1= fAk HAn+2

k

on[0, π], (21) see [29, (2)].

Lemma 6.

(a) There exist positive constants C, C˜ depending only on n, fmaxand fmin, such that CHA

kcotθ+HAk≤ ˜C, (22)

CHA

k+HAk≤ ˜C. (23)

(b) If ak→ +∞when k→ +∞, then HA

ksin14θ

converges to 0uniformly on [0, π].

Proof. (a) Recalling Lemma4, we obtain from (21) that C1(HA

k+HAk)(HA

kcotθ+HAk)n1C2 (24)

for some positive constants C1, C2depending only on n, fmaxand fmin. Note (HA

kcosθ+HAksinθ )=(HA

k+HAk)cosθ, the above inequality can be written as

C1≤ 1

nsinn1θcosθ · d dθ(HA

kcosθ+HAksinθ )nC2. (25)

When θ∈ [0, π/2], we have by (25) that d

C1sinnθ≤ d dθ(HA

kcosθ+HAksinθ )n≤ d

C2sinnθ, which together with HA

k(0) =0 implies C

1 n

1 sinθHA

kcosθ+HAksinθC

1 n

2 sinθ,θ∈ [0, π/2].

(9)

Similarly, by (25) and HA

k(π ) =0, we also have C

1 n

1 sinθHA

kcosθ+HAksinθC

1 n

2 sinθ,θ∈ [π/2, π]. Therefore

C

1 n

1HA

kcotθ+HAkC

1 n

2,θ∈ [0, π], which is just (22). Now recalling (24), one can obtain (23).

(b) We first note that by (11) and (23), there is

|HA

k| ≤C3 (26)

for some positive constant C3depending only on n, fmaxand fmin.

Now assume ak→ +∞as k→ +∞. We claim that for any δ∈(0, π/2), HA

k⇒0 uniformly on[δ, πδ]. (27) In fact, by (16), fAkf (π2)uniformly on [δ, πδ]. By Lemma5, HAkf (π2)2n+21 uniformly on [0, π], which implies that HAk⇒0 uniformly on [0, π]. Then by (21), when θ∈ [δ, πδ], we have

HA

k =fAkHAn2

k (HA

kcotθ+HAk)1nHAk

f (π2)·f (π2)2nn+2+2 ·f (π2)2n1−n+2f (π2)2n1+2

=0.

Thus (27) is true.

We now prove

HA

ksin14θ⇒0 uniformly on[0, π]. (28) Given any >0. By (26), there exists some δ∈(0, π/2), such that

[0,δ]∪[supπδ,π]|HA

ksin14θ|< ,k. (29) Then by virtue of (27), there exists a k0, such that

[δ,πsupδ]|HA

k|< ,kk0. (30)

Combining (29) and (30), we have

[sup0,π]|HA

ksin14θ|< ,kk0. Thus (28) is true. 2

(10)

With a more detailed analysis, we can strengthen Lemma5for n =1, 2.

Lemma 7. Assume ak→ +∞as k→ +∞. For sufficiently large k, we have

max[0,π]HA

kf (π2)2n1+2C

⎧⎪

⎪⎪

⎪⎪

⎪⎨

⎪⎪

⎪⎪

⎪⎪

π

0

|fAkf (π2)|dθ, ifn=1, π

0

|fAkf (π2)|sin12θdθ, ifn=2,

(31)

where Cis a positive constant depending only on f (π2).

Proof. For simplicity, let

β:=f (π2)2n1+2 and hk(θ ):=HAk(θ )β.

Also we will drop the subscript k in the following proof if no confusion arises. Recall by Lemma5, hconverges uniformly to 0 on [0, π]as k→ +∞.

(a) When n =1. Now Eq. (21) is simplified as h+h+β= fA

HA3. (32)

Observing that

HA3=+h)3

=β3−3β4h+6τ5h2, where τ is between βand HA(θ ), and that β=f (π2)1/4, we have

f (π2)

HA3 =β−3h+6β4τ5h2. Then (32) can be written as

h+h+3h−6β4τ5h2=fAf (π2) HA3 , namely

h+4h=fAf (π2)

HA3 +6β4τ5h2. (33)

(11)

Recalling h(π2) =0 by (10), we can apply Lemma8to equation (33) and then obtain

max|h| ≤

fAf (π2) HA3

L1[0,π]

+6β4τ5h2

L1[0,π]. (34) Since HAβ >0 uniformly on [0, π]as k→ +∞, there exists a large integer k0, such that

max|HAβ| ≤β

2,kk0. Then when k≥k0we have

max|h| ≤β

2 and HA, τβ 2,

2

. (35)

Thus (34) is simplified into

max|h| ≤8β3fAf (π2)

L1[0,π]+192β1π(max|h|)2. By virtue of max|h| →0 as k→ +∞, we also can assume

192β1π·max|h|<1

2 whenkk0. Hence

max|h| ≤16β3fAf (π2)

L1[0,π],kk0, which is just (31) for n =1.

(b) When n =2. Now Eq. (21) is written as

(h+h+β)(hcotθ+h+β)= fA

HA4, (36)

namely

β(h+hcotθ+2h)+β2+(h+h)(hcotθ+h)= fA

HA4. (37)

Observing that

HA4=+h)4

=β4−4β5h+10τ6h2, where τ is between β and HA(θ ), and that β=f (π2)1/6, we have

(12)

f (π2)

HA4 =β2−4βh+10β6τ6h2. (38) Then (37) can be written as

β(h+hcotθ+6h)+(h+h)(hcotθ+h)−10β6τ6h2=fAf (π2) HA4 , namely

h+hcotθ+6h=fAf (π2)

βHA4 +Ra(θ ), (39)

where

Ra(θ )=10β5τ6h2β1(h+h)(hcotθ+h). (40) Applying Lemma10to equation (39), we have

max|h| ≤2 π

0

|fAf (π2)|

βHA4 (2−log sinθ )sinθdθ +2

π

0

|Ra(θ )|(2−log sinθ )sinθ

≤4 π

0

|fAf (π2)|

βHA4 sin12 θdθ+6 π

0

|Ra(θ )|sin34θdθ.

Recalling (35), we obtain

max|h| ≤64β5 π

0

|fAf (π2)|sin12θdθ+6 π

0

|Ra(θ )|sin34θdθ. (41)

We see Rainvolves derivatives of h. To deal with them, we need to explore (36) more care- fully. Note that

(hcosθ+hsinθ+βsinθ )=(h+h+β)cosθ, then Eq. (36) is equivalent to

d

(hcosθ+hsinθ+βsinθ )2= fA

HA4 ·2 sinθcosθ.

Therefore we have

(13)

(hcosθ+hsinθ+βsinθ )2= θ

0

fA

HA4 ·2 sintcostdt, ∀θ∈ [0, π/2]. (42) Since hcotθ+h +β >0, there is

hcosθ+hsinθ+βsinθ=

θ

0

fA

HA4 ·2 sintcostdt

1/2

.

Thus we have

hcosθ+hsinθ=θ

0

fA

HA4 ·2 sintcostdt 1/2βsinθ

= θ

0 fA

HA4 ·2 sintcostdt−β2sin2θ θ

0 fA

HA4 ·2 sintcostdt1/2

+βsinθ

≤ 1 βsinθ

θ

0

fA

HA4·2 sintcostdt−β2sin2θ.

(43)

Recalling (38), there is fA

HA4 =fAf (π2)

HA4 +β2−4βh+10β6τ6h2, which implies that

θ

0

fA

HA4 ·2 sintcostdt= θ

0

fAf (π2)

HA4 ·2 sintcostdt+β2 θ

0

2 sintcostdt

+ θ

0

(−4βh+10β6τ6h2)·2 sintcostdt

= θ

0

fAf (π2)

HA4 ·2 sintcostdt+β2sin2θ

+ θ

0

(−4β+10β6τ6h)h·2 sintcostdt.

Recalling (35), we obtain from the above equality that

(14)

θ

0

fA

HA4 ·2 sintcostdt−β2sin2θ≤32β4 θ

0

|fAf (π2)|sintdt+324β(max|h|)sin2θ.

Then (43) is simplified into

hcosθ+hsinθ≤ 32β5 sinθ

θ

0

|fAf (π2)|sintdt+324(max|h|)sinθ,

namely

hcotθ+hsin12θ≤32β5 sin32θ

θ

0

|fAf (π2)|sintdt+324(max|h|)sin12θ.

Integrating both sides over [0, π/2], we have

π

2

0

|hcotθ+h|sin12θ

≤32β5

π

2

0

dθ sin32θ

θ

0

|fAf (π2)|sintdt+162π(max|h|)

=32β5

π

2

0

|fAf (π2)|sintdt

π

2

t

sin32θ+162π(max|h|).

(44)

Note that

π

2

t

dθ sin32θπ

2

3 2

π

2

t

θ32

=π 2

3

2 ·2[t12(π/2)12]

<4 sin12t, then (44) is reduced into

π

2

0

|hcotθ+h|sin12θdθ≤128β5

π

2

0

|fAf (π2)|sin12tdt+162π(max|h|). (45)

Now similar to (42), we have

(15)

(hcosθ+hsinθ+βsinθ )2= π

θ

fA

HA4 ·2 sint|cost|dt, ∀θ∈ [π/2, π]. Then following almost the same arguments used to obtain (45), one can get

π

π 2

|hcotθ+h|sin12θdθ≤128β5 π

π 2

|fAf (π2)|sin12tdt+162π(max|h|). (46)

Adding (45) and (46) together, we have π

0

|hcotθ+h|sin12θdθ≤128β5 π

0

|fAf (π2)|sin12tdt+324π·max|h|. (47)

Now we can estimate the integral about Rain (41). By the definition of Rain (40), there is π

0

|Ra(θ )|sin34θdθ≤ π

0

10β5τ6h2sin34θ

+ π

0

β1|h+h| |hcotθ+h|sin34θ

≤640πβ1(max|h|)2+mk π

0

|hcotθ+h|sin12θdθ,

where (35) is used, and mkis defined as mk:=β1 max

θ∈[0,π]|h(θ )+h(θ )|sin14θ.

By estimate (47), the above inequality becomes into π

0

|Ra(θ )|sin34θdθ≤640πβ1(max|h|)2+324π mk·max|h|

+128β5mk

π

0

|fAf (π2)|sin12 tdt.

(48)

Recall Lemma6(b), |h(θ ) +h(θ )| sin14θ converges uniformly to 0 on [0, π]when k→ +∞, which implies

mk→0 ask→ +∞.

(16)

Also recall max|h| →0. We can assume when k≥k0that 640πβ1max|h| +324π mk< 1

12. Then (48) is simplified into

π

0

|Ra(θ )|sin34θdθ≤ 1

12max|h| + 1 12β5

π

0

|fAf (π2)|sin12tdt. (49)

Now combining (41) and (49), we obtain

max|h| ≤129β5 π

0

|fAf (π2)|sin12θdθ,

which is just (31) for n =2. 2

The following Lemmas8and 10have been used in the proof of the above Lemma7.

Lemma 8. Assume h C2(R)is 2π-periodic and even. If it satisfies the following differential equation

h+4h=g, (50)

and h(π2) =0, then there is

maxR |h| ≤ gL1[0,π]. Proof. One can easily solve equation (50) to obtain

h(θ )=c1cos 2θ+c2sin 2θ−1 2cos 2θ

θ

0

g(t)sin 2tdt+1 2sin 2θ

θ

0

g(t)cos 2tdt,

where c1and c2are constants to be determined. Then we have

h(θ )= −2c1sin 2θ+2c2cos 2θ+sin 2θ θ

0

g(t)sin 2tdt+cos 2θ θ

0

g(t)cos 2tdt.

From h(0) =0, we get c2=0. And h(π2) =0 implies

c1=1 2

π

2

0

g(t)sin 2tdt.

(17)

Therefore his given by

h(θ )=1 2cos 2θ

π

2

θ

g(t )sin 2tdt+1 2sin 2θ

θ

0

g(t )cos 2tdt.

Hence when θ∈ [0, π],

|h(θ )| ≤1 2

π

2

θ

g(t )sin 2tdt +1

2 θ

0

g(t )cos 2tdt

≤1 2 π

0

|g(t )|dt+1 2 π

0

|g(t )|dt

= π

0

|g(t )|dt,

which leads to the conclusion of this lemma. 2 Lemma 9. The homogeneous differential equation

h+hcotθ+6h=0 in(0, π ) has the following two fundamental solutions:

h1(θ )=1−3 cos2θ, h2(θ )= −3

4cosθ+1

8(1−3 cos2θ )log1−cosθ 1+cosθ. These two solutions have the following properties:

(a) h1(π2) =1, h1(π2) =0and h2(π2) =0, h2(π2) =1.

(b) Abel’s identity: h1h2h1h2=cscθ, θ(0, π ).

(c) h1(θ ) =6 sinθcosθ.

(d) |h2(θ )| ≤2 −log sinθ, ∀ θ(0, π ).

(e) |h2(θ ) sinθ| ≤5/2, ∀ θ(0, π ).

(f) As θ→0+or θπ, there is

h2(θ )=−1/2+o(1) sinθ .

Proof. Direct computations show that h1and h2are solutions to the differential equation in the lemma. And one can easily check (a), (b) and (c).

(18)

We note that 1 2

log1−cosθ 1+cosθ

≤ −log sinθ+log 2, ∀θ(0, π ).

Since both sides are symmetric with respect to θ=π/2, we only need to verify it for θ∈(0, π/2], which is a direct corollary of the following equality:

1 2

log1−cosθ 1+cosθ

=1 2

log 1−cos2θ (1+cosθ )2

=

log sinθ 1+cosθ

.

Now by the expression of h2, there is

|h2(θ )| ≤3 4+1

2

log1−cosθ 1+cosθ

≤3

4−log sinθ+log 2

≤2−log sinθ, which is just (d).

Computing h2, we have h2(θ )=3

4sinθ+3

4sinθcosθlog1−cosθ 1+cosθ +1

4(1−3 cos2θ )cscθ.

Then

|h2(θ )| ≤3 4+3

4sinθ

log1−cosθ 1+cosθ

+1 2cscθ

≤3 4+3

2sinθ·(−log sinθ+log 2)+1 2cscθ

≤2+1 2cscθ, which implies (e).

By the expression of h2, we see as θ→0+or θ→πthat h2(θ )sinθ→ −1

2, yielding (f). 2

Lemma 10. Assume h ∈C2(R)is 2π-periodic and even. If it satisfies the following differential equation

h+hcotθ+6h=g, (51)

(19)

and h(π2) =0, then there is

maxR |h| ≤2 π

0

|g(θ )|(2−log sinθ )sinθdθ. (52)

Proof. Recalling Lemma9, h1and h2are two fundamental solutions to the homogeneous dif- ferential equation:

h+hcotθ+6h=0 in(0, π ).

By method of variation of parameters and Lemma9(b), we solve (51) in (0, π )and obtain

h(θ )=c1h1+c2h2h1 θ

π/2

h2(t)g(t)sintdt+h2 θ

π/2

h1(t)g(t)sintdt, (53)

where c1and c2are constants to be determined. Note the assumption h(π2) =0, and by (53) h(π2)=c1h1(π2)+c2h2(π2)=c1,

there is c1=0. Then

h(θ )=c2h2h1 θ

π/2

h2(t)g(t)sintdt+h2 θ

π/2

h1(t)g(t)sintdt. (54)

To determine c2, we need to compute h(0).

By Lemma9(d), |h2|is an integrable function in (0, π/2]. Then 0

π/2

h2(t)g(t)sintdt

is a finite number. For small θ >0, one can rewrite (54) as 1

h2(θ )

h(θ )+h1(θ ) θ

π/2

h2(t)g(t)sintdt

=c2+ θ

π/2

h1(t)g(t)sintdt. (55)

Letting θ→0+, and recalling h(0) =0, h1(0) =0 and Lemma9(f), we obtain

0=c2+ 0

π/2

h1(t)g(t)sintdt,

Références

Documents relatifs

The Brunn–Minkowski theory (or the theory of mixed volumes) of convex bodies, devel- oped by Minkowski, Aleksandrov, Fenchel, et al., centers around the study of geometric

In w we prove a delicate continuity property for the first variation that is used later in the existence part of Theorem 0.8 and also to extend the validity of the

Aleksandrov [2], [3] and Fenchel-Jessen [22] independently solved the problem in 1938 for arbitrary measures: If µ is not concentrated on any closed hemisphere of S n−1 , then µ is

A non-zero finite even Borel measure on the unit sphere S n−1 is the cone- volume measure of an origin-symmetric convex body in R n if and only if it satisfies the

As mentioned previously, the family of L p dual curvature measures, or sometimes simply referred to as the (p, q)-th dual curvature measure and denoted by C p,q (K, · ) for q ∈ R

Like many other central objects of interest in convex geometry, such as the L p affine and geominimal surface areas and the L p John ellipsoids (see e.g., [27, 34, 37, 42, 43, 52,

Finally, a very general Minkowski-type inequality, involving two Orlicz functions, two convex bodies, and a star body, is proved, that includes as special cases several others in

The logarithmic Minkowski problem asks for necessary and sufficient conditions for a finite Borel measure on the unit sphere so that it is the cone-volume measure of a convex