Sharp Lp -entropy inequalities on manifolds

Sharp Lp -entropy inequalities on manifolds

JID:YJFAN AID:7289 /FLA [m1L; v1.159; Prn:13/07/2015; 13:18] P.1 (1-29) Journal of Functional Analysis ••• (••••) •••–••• Contents lists available...

1MB Sizes 0 Downloads 76 Views

JID:YJFAN

AID:7289 /FLA

[m1L; v1.159; Prn:13/07/2015; 13:18] P.1 (1-29)

Journal of Functional Analysis ••• (••••) •••–•••

Contents lists available at ScienceDirect

Journal of Functional Analysis www.elsevier.com/locate/jfa

Sharp Lp -entropy inequalities on manifolds Jurandir Ceccon a , Marcos Montenegro b,∗ a

Departamento de Matemática, Universidade Federal do Paraná, Caixa Postal 19081, 81531-980, Curitiba, PR, Brazil b Departamento de Matemática, Universidade Federal de Minas Gerais, Caixa Postal 702, 30123-970, Belo Horizonte, MG, Brazil

a r t i c l e

i n f o

Article history: Received 4 September 2013 Accepted 22 June 2015 Available online xxxx Communicated by L. Gross MSC: 35J92 41A44 58J05

a b s t r a c t In 2003, Del Pino and Dolbeault [14] and Gentil [19] investigated, independently, best constants and extremals associated to Euclidean Lp -entropy inequalities for p > 1. In this work, we present some contributions in the Riemannian context. Namely, let (M, g) be a compact Riemannian manifold of dimension n ≥ 3. For 1 < p ≤ 2, we establish the validity of the sharp Riemannian Lp -entropy inequality 

⎞ ⎛  n p |u| log(|u| )dvg ≤ log ⎝Aopt |∇g u| dvg + B⎠ p p

Keywords: Entropy inequalities Log-Sobolev type inequalities Best constants

M

p

M

on all functions u ∈ H 1,p (M ) such that ||u||Lp (M ) = 1 for some constant B. Moreover, we prove that the first best constant Aopt is equal to the corresponding Euclidean one. Our approach is inspired on the Bakry, Coulhon, Ledoux and Saloff-Coste’s idea [3] of getting Euclidean entropy inequalities as a limit case of suitable subcritical interpolation inequalities. It is conjectured that the inequality sometimes fails for p > 2. © 2015 Elsevier Inc. All rights reserved.

* Corresponding author. E-mail addresses: [email protected] (J. Ceccon), [email protected] (M. Montenegro). http://dx.doi.org/10.1016/j.jfa.2015.06.016 0022-1236/© 2015 Elsevier Inc. All rights reserved.

JID:YJFAN 2

AID:7289 /FLA

[m1L; v1.159; Prn:13/07/2015; 13:18] P.2 (1-29)

J. Ceccon, M. Montenegro / Journal of Functional Analysis ••• (••••) •••–•••

1. Introduction Logarithmic Sobolev inequalities are a powerful tool in Real Analysis, Complex Analysis, Geometric Analysis, Convex Geometry and Probability. The pioneer work by L. Gross [20] puts forward the equivalence between a class of logarithmic Sobolev inequalities and hypercontractivity of the associated heat semigroup. Particularly, his logarithmic Sobolev inequality with respect to the Gaussian measure plays an important role in Ricci flow theory (e.g. [25]), optimal transport theory (e.g. [27]), probability theory (e.g. [23]), among other applications. Later, Weissler [29] introduced a log-Sobolev type inequality (also known as Euclidean L2 -entropy inequality) equivalent to Gross’s inequality with Gaussian measure. The Euclidean L2 -entropy inequality and its variants have been used in the study of optimal estimates for solutions of certain nonlinear diffusion equations, see for instance [2,6,7,9,15,18,19] and the references therein. The Euclidean Lp -entropy inequality for p ≥ 1 states that, for any function u ∈  W 1,p (Rn ) with Rn |u|p dx = 1, ⎛

 |u|p log(|u|p ) dx ≤

Ent dx (|u|p ) := Rn

n log ⎝A0 (p) p



⎞ |∇u|p dx⎠ ,

(1)

Rn

where n ≥ 2, p ≥ 1 and A0 (p) is the best possible constant in this inequality. As mentioned above, the Euclidean L2 -entropy inequality was established by Weissler in [29]. Thereafter, Carlen [10] showed that its extremal functions are precisely dilations and translations of the Gaussian function u0 (x) = π − 2 e−|x| . 2

n

For p = 1, Ledoux [22] proved the inequality (1) and Beckner [5] classified its extremal functions as normalized characteristic functions of balls. In [5], Beckner also proved that (1) is valid for any 1 < p < n and Del Pino and Dolbeault [14] characterized its extremal functions as dilations and translations of the function u0 (x) = π − 2

n

Γ( n2 + 1) Γ( n(p−1) + 1) p

e−|x|

p p−1

.

(2)

Finally, Gentil [19] established the validity of (1) and that u0 is an extremal function for any p > 1. Thanks to a uniqueness argument due to Del Pino and Dolbeault [14], modulo dilations and translations, the classification also extends for any p > 1. As a byproduct, they derived, for any p > 1, p A0 (p) = n



p−1 e



p−1 π

−p 2

Γ( n2 + 1) Γ( n(p−1) + 1) p

np .

JID:YJFAN

AID:7289 /FLA

[m1L; v1.159; Prn:13/07/2015; 13:18] P.3 (1-29)

J. Ceccon, M. Montenegro / Journal of Functional Analysis ••• (••••) •••–•••

3

In order to introduce sharp Lp -entropy inequalities within the Riemannian environment for 1 ≤ p < n, we first deduce an intermediate entropy inequality. Let (M, g) be a smooth compact Riemannian manifold of dimension n ≥ 2 and 1 ≤ np , Hölder’s inequality gives p < n. For any p ≤ s ≤ p∗ := n−p 1−α ||u||Ls (M ) ≤ ||u||α Lp (M ) ||u||Lp∗ (M ) , ∗

for all functions u ∈ Lp (M ), where α = np−ns+ps . Taking logarithm of both sides, one ps gets

 ||u||Lp (M ) ||u||Ls (M ) + (α − 1) log log ≤0. ||u||Lp (M ) ||u||Lp∗ (M ) Since this inequality trivializes to an equality when s = p, we may differentiate it with respect to s at s = p. Then, a simple computation provides ⎛ ⎞ pp∗   ∗ n |u|p log |u|p dvg ≤ log ⎝ |u|p dvg ⎠ p M

M ∗

for all functions u ∈ Lp (M ) with ||u||Lp (M ) = 1. Using the above inequality and the Sobolev embedding theorem for compact manifolds (e.g. [1]), one gets constants A, B ∈ R such that, for any u ∈ H 1,p (M ) with  |u|p dvg = 1, M ⎛ ⎞   n Ent dvg (|u|p ) := |u|p log |u|p dvg ≤ log ⎝A |∇g u|p dvg + B ⎠ , (L(A, B)) p M

M

where dvg denotes the Riemannian volume element, ∇g is the gradient operator of g and H 1,p (M ) is the Sobolev space defined as the completion of C ∞ (M ) under the norm ⎛ ||u||H 1,p (M ) := ⎝



 |∇g u|p dvg +

M

⎞ p1 |u|p dvg ⎠

.

M

The following definitions and notations related to (L(A, B)) are quite natural when one desires to introduce sharp Lp -entropy inequalities in the Riemannian context. The first best Lp -entropy constant is defined by A0 (p, g) := inf{A ∈ R : there exists B ∈ R such that (L(A, B)) holds  |u|p dvg = 1} . for all u ∈ H 1,p (M ) with M

It follows directly that A0 (p, g) is well defined and A0 (p, g) ≥ −∞ for any 1 ≤ p < n.

JID:YJFAN 4

AID:7289 /FLA

[m1L; v1.159; Prn:13/07/2015; 13:18] P.4 (1-29)

J. Ceccon, M. Montenegro / Journal of Functional Analysis ••• (••••) •••–•••

Assuming for a moment that A0 (p, g) is finite, we have: The first sharp Lp -entropy inequality states that there exists a constant B ∈ R such  that, for any u ∈ H 1,p (M ) with M |u|p dvg = 1, ⎛ ⎞   n |u|p log |u|p dvg ≤ log ⎝A0 (p, g) |∇g u|p dvg + B ⎠ . p M

M

If the preceding inequality is true, then we can define the second best Lp -entropy constant as B0 (p, g) := inf{B ∈ R : (L(A0 (p, g), B)) holds  |u|p dvg = 1} for all u ∈ H 1,p (M ) with M

and the second sharp Lp -entropy inequality as the saturated version of (L(A0 (p, g), B))  on the functions u ∈ H 1,p (M ) with M |u|p dvg = 1, that is ⎛ ⎞   n |u|p log |u|p dvg ≤ log ⎝A0 (p, g) |∇g u|p dvg + B0 (p, g)⎠ . p M

M

Note that (L(A0 (p, g), B0 (p, g))) is sharp with respect to both the first and second best constants in the sense that none of them can be lowered. In a natural way, one introduces the notion of extremal functions of (L(A0 (p, g),  B0 (p, g))). A function u0 ∈ H 1,p (M ) satisfying M |u0 |p dvg = 1 is said to be extremal, if ⎛ ⎞   n |u0 |p log |u0 |p dvg = log ⎝A0 (p, g) |∇g u0 |p dvg + B0 (p, g)⎠ . p M

M

We denote by E0 (p, g) the set of all extremal functions of (L(A0 (p, g), B0 (p, g))). In [8], Brouttelande proved for dimensions n ≥ 3 that A0 (2, g) = A0 (2) and (L(A0 (2, g), B)) is valid for some constant B ∈ R. Subsequently, in [9], he obtained the lower bound B0 (2, g) ≥

1 max Rg , 2nπe M

where Rg stands for the scalar curvature of the metric g, and proved that if the inequality is strict, then the set E0 (p, g) is non-empty. A lower bound for B0 (p, g) involving the volume vg (M ) of M also follows by taking a normalized constant function in (L(A0 (p, g), B0 (p, g))), namely B0 (p, g) ≥ vg (M )−p/n , provided that B0 (p, g) is well defined.

JID:YJFAN

AID:7289 /FLA

[m1L; v1.159; Prn:13/07/2015; 13:18] P.5 (1-29)

J. Ceccon, M. Montenegro / Journal of Functional Analysis ••• (••••) •••–•••

5

Our main contributions are gathered in the following result: Theorem 1.1. Let (M, g) be a smooth compact Riemannian manifold of dimension n ≥ 2. For any 1 < p ≤ 2 and p < n, we have: (a) A0 (p, g) = A0 (p); (b) there exists a constant B ∈ R such that (L(A0 (p, g), B)) holds for all functions  u ∈ H 1,p (M ) with M |u|p dvg = 1. Note that the above result extends to 1 < p ≤ 2 the corresponding one due to Brouttelande. However, his arguments do not apply to p = 2, so we here develop an alternative approach in order to prove Theorem 1.1. The idea of using subcritical interpolation inequalities for obtaining entropy inequalities was introduced by Bakry, Coulhon, Ledoux and Saloff-Coste in [3] within the Euclidean environment. Namely, for 1 ≤ p < n, they showed how to produce nonsharp Lp -entropy inequalities as a limit case of a class of non-sharp Gagliardo–Nirenberg inequalities. Later, this view point was explored in the sharp sense by Del Pino and Dolbeault [14] in order to establish the inequality (1) for 1 < p < n. Indeed, they considered a family of sharp Gagliardo–Nirenberg inequalities, interpolating the Lp -Sobolev and Lp -entropy inequalities, whose extremal functions are explicitly known. In trying to adapt the same idea of getting entropy inequalities as a limit case of subcritical interpolation inequalities to the Riemannian context, the situation changes drastically mainly because extremal functions and second best constants are usually unknown. With the aim to make clear to the reader our program of proof and its main points of difficulty, a brief overview on related sharp Riemannian Nash inequalities should be presented. Let 1 < p ≤ 2 and 1 ≤ q < p. In [12], [13] and [21], Ceccon–Montenegro and Humbert established, respectively for 1 < p < 2 and p = 2, the existence of a constant B ∈ R such that the sharp Lp -Nash inequality ⎛ ⎝

 M

⎞ θ1 ⎛ |u|p dvg ⎠ ≤⎝Aopt



 |∇g u|p dvg + B M

⎞⎛ |u|p dvg ⎠⎝

M

holds for all functions u ∈ H 1,p (M ), where θ = Lp -Nash constant. One also knows that



⎞ p(1−θ) qθ |u|q dvg ⎠

(Ip,q (A, B))

M n(p−q) q(p−n)+np

Aopt = N (p, q) ,

and Aopt is the first best

(3)

JID:YJFAN

AID:7289 /FLA

[m1L; v1.159; Prn:13/07/2015; 13:18] P.6 (1-29)

J. Ceccon, M. Montenegro / Journal of Functional Analysis ••• (••••) •••–•••

6

where N (p, q) stands for the best Euclidean Lp -Nash constant, see [17] for p = 2 and [13] for 1 < p < 2. Namely, N (p, q) is the best possible constant in the Lp -Nash inequality ⎛ ⎝



⎞ θ1



|u|p dx⎠ ≤ A ⎝

Rn

⎞⎛



|∇u|p dx⎠ ⎝

Rn

⎞ p(1−θ) qθ



|u|q dx⎠

Rn

which holds for all functions C0∞ (Rn ), provided that n ≥ 2 and 1 ≤ q < p. This last inequality was first proved by Nash in [24]. An alternative proof was given by Beckner in [4]. For p = 2, we refer to the recent work [11] by Ceccon. One then defines the second best Lp -Nash constant as B(p, q, g) := inf{B ∈ R : (Ip,q (N (p, q), B)) holds for all u ∈ H 1,p (M )} . In this case, the second sharp Lp -Nash inequality automatically holds on H 1,p (M ), namely ⎛ ⎝

 M

⎞ θ1



|u|p dvg ⎠ ≤ ⎝N (p, q)



 |∇g u|p dvg + B(p, q, g)

M

⎞⎛ |u|p dvg ⎠⎝

M



⎞ p(1−θ) qθ |u|q dvg ⎠

.

M

Our strategy consists in rearranging the above inequality and applying logarithmic of both sides, so that ||u||Lp (M ) N (p, q) p log ≤ log θ ||u||Lq (M )

 M

|∇g u|p dvg + B(p, q, g) 

pq q dv |u| g M

 M

|u|p dvg

,

(4)

and then letting the limit as q → p− . The success of this plan will be result of the following three statements for 1 < p ≤ 2 and p < n: (A) N (p, q) converges to A0 (p) as q → p− ; (B) B(p, q, g) is bounded for any q < p close to p; (C) A0 (p, g) ≥ A0 (p). In fact, assume for a moment that (A), (B) and (C) are true. Letting q → p− in (4) and using (A) and (B), after straightforward computations, one obtains the inequality (L(A0 (p), B)) for some constant B. In particular, this implies that A0 (p, g) ≤ A0 (p). Thus, thanks to (C), the conclusion of Theorem 1.1 follows. We then describe some ideas involved in the proof of each one of the claims (A)–(C). We begin by addressing (A) and (C) since their proofs are shortest. On the proof of (A). This claim is proved in Section 2 and uses Jensen’s inequality and the fact to be proved that the best Nash constant N (p, q) is increasing on q.

JID:YJFAN

AID:7289 /FLA

[m1L; v1.159; Prn:13/07/2015; 13:18] P.7 (1-29)

J. Ceccon, M. Montenegro / Journal of Functional Analysis ••• (••••) •••–•••

7

On the proof of (C). This claim is proved in Section 3. Its proof is based on estimates of Gaussian bubbles. Precisely, we consider the following test function in (L(A0 (p, g), B0 (p, g))): n x uε (expx0 (x)) = η(x)ε− p u0 ( ) ε

defined locally around a point x0 on M , where η denotes a cutoff function supported in a ball centered at 0. After using Cartan’s expansion of the metric g around x0 and estimating each involved integral for ε > 0 small enough, the desired conclusion follows. On the proof of (B). This claim is proved in Section 4. Since its proof is rather long and technique, for a better understanding two important steps are presented under the form of lemma. It suffices to prove the assertion for each sequence (qk ) ⊂ (1, q) such that qk → p as k → +∞. For such a sequence and each k, we consider the functional ⎛ Jk (u) = ⎝



 |∇g u| dvg + Ck p

M

⎞⎛ |u| dvg ⎠ ⎝

k qk

|u|

p

M

qk ) ⎞ p(1−θ q θ

 qk

dvg ⎠

M

on the set H = {u ∈ H 1,p (M ) : ||u||Lp (M ) = 1}, where Ck is defined as Ck =

B(p, qk , g) − (p − qk ) . N (p, qk )

From its definition, we readily have Ck < B(p, qk , g)N (p, qk )−1 . Therefore, from the definition of B(p, qk , g), there exists a function wk ∈ H satisfying Jk (wk ) < N (p, qk )−1 , so that νk = inf Jk (u) < N (p, qk )−1 . u∈H

As usual, this last inequality leads to the existence of a C 1 minimizer uk ∈ H for the functional Jk on H. The next steps consist in studying some fine properties satisfied by the sequence (uk ) such as concentration and pointwise estimates. The main tools used here are blow-up method and Moser’s iteration on elliptic PDEs. We conclude the introduction by exposing some still open problems on sharp Riemannian entropy inequalities and related best constants. Perhaps, contrary to what one might expect, it is not clear that the first best entropy constant A0 (p, g) is well defined and is equal to A0 (p) for all p ≥ n. The great difficulty in Riemannian Lp -entropy inequalities is that local-to-global type arguments do not usually  work well. For example, the normalization condition M |u|p dvg = 1 and the involved log functions in (L(A, B)) do not allow a direct comparison to the corresponding flat case. In particular, it does not seem immediate that (L(A, B)) is valid for some constants A and B and also that, for each ε > 0, there exists a constant Bε ∈ R such that

JID:YJFAN 8

AID:7289 /FLA

[m1L; v1.159; Prn:13/07/2015; 13:18] P.8 (1-29)

J. Ceccon, M. Montenegro / Journal of Functional Analysis ••• (••••) •••–•••



⎛ ⎞  n |u|p log |u|p dvg ≤ log ⎝(A0 (p) + ε) |∇g u|p dvg + Bε ⎠ p

M

M



for all functions u ∈ H 1,p (M ) with M |u|p dvg = 1. According to our contributions, the first best entropy constant A0 (p, g) is well defined for all 1 ≤ p < n and equal to A0 (p) for all 1 < p ≤ 2 and p < n. In addition, based on developments over thirty years in the field of sharp entropy and Sobolev inequalities (see [1,16] and the references therein), one expects positive answers for the following questions: Open problem 1. Is A0 (p, g) well defined for all p ≥ n? Open problem 2. Does A0 (p, g) = A0 (p) hold in the three cases p = 1, n ≥ 2, p = 2, n = 2 and p > 2, n ≥ 2? Open problem 3. Is (L(A0 (p, g), B)) valid for some constant B in the two cases p = 1, n ≥ 2 and p = 2, n = 2? Open problem 4. Is (L(A0 (p, g), B)) non-valid whenever p > 2, n ≥ 2 and (M, g) has positive scalar curvature somewhere? 2. Proof of the assertion (A) This section is devoted to the proof of the proposition. Proposition 2.1. Let n ≥ 2 and 1 ≤ q < p. We have lim N (p, q) = A0 (p) .

(5)

N (p, q) ≤ A0 (p)

(6)

q→p−

We first prove that

for all 1 < q < p. Let u ∈ C0∞ (Rn ) such that ||u||Lp (Rn ) = 1. Using Jensen’s inequality, we have ⎛ − ln ⎝







|u| dx⎠ = − ln ⎝



 |u|

q

Rn

q−p

|u| dx⎠ ≤ −

Rn

=

p−q p



p

ln(|u|q−p )|u|p dx

Rn



ln(|u|p )|u|p dx . Rn

Joining this inequality with (1), one obtains ⎛ ⎝



p2 ⎞− n(p−q)

|u| dx⎠ q

Rn

 ≤ A0 (p) Rn

|∇u|p dx .

JID:YJFAN

AID:7289 /FLA

[m1L; v1.159; Prn:13/07/2015; 13:18] P.9 (1-29)

J. Ceccon, M. Montenegro / Journal of Functional Analysis ••• (••••) •••–•••

9

Using the fact that ||u||Lp (Rn ) = 1, we then derive the Nash inequality ⎛ ⎝



⎞ θ1



|u|p dx⎠ ≤ A0 (p) ⎝

Rn

⎞⎛



|∇u|p dx⎠ ⎝

Rn

⎞ p(1−θ) θq



|u|q dx⎠

,

Rn

where θ=

n(p − q) . qp − qn + np

By homogeneity, the above inequality is valid for all functions u ∈ C0∞ (Rn ), so that the assertion (6) follows. We now prove that N (p, q) is monotonically increasing on q. Let 1 < q1 < q2 < p be fixed. A usual interpolation inequality yields ⎛ ⎝



⎞ q1



2

|u|q2 dx⎠

≤⎝

Rn

⎞ qμ ⎛



1

|u|q1 dx⎠



Rn

⎞ 1−μ p



|u|p dx⎠

,

Rn

where q12 = qμ1 + 1−μ p . Plugging this inequality in ⎛ ⎝



⎞ θ1



2

|u|p dx⎠

≤ N (p, q2 ) ⎝

Rn

where θ2 = ⎛ ⎝



⎞⎛



|∇u|p dx⎠ ⎝

Rn

n(p−q2 ) q2 (p−n)+np ,



2

|u| dx⎠

2 2

|u|q2 dx⎠

,

Rn

one gets

2 ⎞1+μ 1−θ θ

p

2) ⎞ p(1−θ q θ





≤ N (p, q2 ) ⎝

Rn

⎞⎛ |∇u| dx⎠ ⎝

1

|u|

p

Rn

2) ⎞ qp (μ 1−θ θ

 q1

dx⎠

2

.

Rn

On the other hand, the definition of μ produces the relations 1+μ

1 − θ2 1 = θ2 θ1

and p q1 where θ1 =

n(p−q1 ) q1 (p−n)+np ,



1 − θ2 μ θ2

=

p 1 − θ1 , q1 θ 1

so that N (p, q1 ) ≤ N (p, q2 ) and the monotonicity follows.

JID:YJFAN 10

AID:7289 /FLA

[m1L; v1.159; Prn:13/07/2015; 13:18] P.10 (1-29)

J. Ceccon, M. Montenegro / Journal of Functional Analysis ••• (••••) •••–•••

We then denote A(p) = lim N (p, q) .

(7)

q→p−

It is clear by (6) that A(p) ≤ A0 (p). The remainder of the proof is devoted to show that A(p) ≥ A0 (p). Rearranging the sharp Nash inequality ⎛ ⎝



⎞ θ1



q

|u| dx⎠ p

≤ N (p, q) ⎝

Rn

where θq =



⎞⎛ |∇u| dx⎠ ⎝ p

Rn

n(p−q) np+pq−nq ,



q) ⎞ p(1−θ qθ q

|u| dx⎠ q

Rn

and taking logarithm of both sides, one has 1 log θq



||u||p ||u||q



 1 ||∇u||pp p ≤ log N (p, q) . ||u||pq

Using the definition of θq and taking the limit on q, one gets p2 1 lim log n q→p− p − q



||u||p ||u||q



 1 ||∇u||pp p ≤ log A(p) . ||u||pp

We now compute the left-hand side limit. We first write  log

||u||p ||u||q



1 1 log(||u||pp ) − log(||u||qq ) p q

q−p 1 = log(||u||q ) + log(||u||pp ) − log(||u||qq ) . p p

=

A straightforward computation then gives 1 log lim q→p− p − q



||u||p ||u||q



1 = p

 Rn

|u|p log ||u||pp



|u| ||u||p

dx .

Therefore,  Rn

|u|p log ||u||pp



|u|p ||u||pp



 ||∇u||pp n dx ≤ log A(p) , p ||u||pp

or equivalently,  Rn

⎛ ⎞  n |u|p log(|u|p ) dx ≤ log ⎝A(p) |∇u|p dx⎠ p Rn

,

JID:YJFAN

AID:7289 /FLA

[m1L; v1.159; Prn:13/07/2015; 13:18] P.11 (1-29)

J. Ceccon, M. Montenegro / Journal of Functional Analysis ••• (••••) •••–•••

11

for all functions u ∈ C0∞ (Rn ) with ||u||Lp (Rn ) = 1. So, A(p) ≥ A0 (p) and the proof of Proposition 2.1 follows. 2 3. Proof of the assertion (C) In this section, we prove the following result: Proposition 3.1. For each n ≥ 2 and 1 < p < n, we have A0 (p, g) ≥ A0 (p) . Using the assumption that n ≥ 2 and 1 < p < n, one gets constants A and B such that (L(A, B)) is valid. It suffices to show that A ≥ A0 (p). One knows that (L(A, B)) is equivalent to 1 ||u||pp

 |u|p log(|u|p ) dvg + (

n − 1) log(||u||pp ) p

M

⎛ ⎞   n ≤ log ⎝A |∇u|pg dvg + B |u|p dvg ⎠ p M

(8)

M

for all functions u ∈ C ∞ (M ).

p p−1

We first fix a point x0 ∈ M and an extremal function u0 (x) = ae−b|x| ∈ W 1,p (Rn ) for the sharp entropy inequality (1) (see [14] or [19]), where a and b are positive constants chosen so that ||u0 ||Lp (Rn ) = 1. Consider now a geodesic ball B(x0 , δ) ⊂ M and a radial cutoff function η ∈ ∞ C (B(0, δ)) satisfying η = 1 in B(0, 2δ ), η = 0 outside B(0, δ), 0 ≤ η ≤ 1 in B(0, δ). For ε > 0 and x ∈ B(0, δ), set n x uε (expx0 (x)) = η(x)ε− p u0 ( ) . ε

The asymptotic behavior of each integral of (8) computed at uε with ε > 0 small enough is now presented. Denote   I1 = up0 log(up0 ) dx, I2 = |∇u0 |p dx , 

Rn



up0 |x|2 dx, J2 =

J1 = Rn

Rn

 up0 log(up0 )|x|2 dx .

|∇u0 |p |x|2 dx, J3 = Rn

Rn

Using the expansion of volume element in geodesic coordinates

JID:YJFAN

AID:7289 /FLA

[m1L; v1.159; Prn:13/07/2015; 13:18] P.12 (1-29)

J. Ceccon, M. Montenegro / Journal of Functional Analysis ••• (••••) •••–•••

12

n  1  detg = 1 − Ric ij (x0 )xi xj + O(r3 ) , 6 i,j=1

where Ric ij denotes the components of the Ricci tensor in these coordinates and r = |x|, one easily checks that  upε dvg = 1 −

Rg (x0 ) J1 ε2 + O(ε4 ) , 6n

(9)

M



Rg (x0 ) Rg (x0 ) J3 ε2 + J1 ε2 log ε + o(ε4 log ε) 6n 6

upε log(upε ) dvg = I1 − n log ε −

(10)

M

and 

−p



|∇uε | dvg = ε p



Rg (x0 ) I2 − J2 ε2 + O(ε4 ) 6n

.

(11)

M

Plugging uε in (8), from (9), (10) and (11), one obtains I1 − n log ε −

Rg (x0 ) Rg (x0 ) 2 J1 ε2 log ε 6n J3 ε + 6 R (x ) 1 − g6n 0 J1 ε2 + O(ε4 )

+ o(ε4 log ε)

 n Rg (x0 ) − 1) log 1 − J1 ε2 + O(ε4 ) p 6n  n Rg (x0 ) 2 p q AJ2 ε + Bε + O(ε ) ≤ −n log ε + log AI2 − p 6n +(

for ε > 0 small enough, where q = min{4, p + 2}. Taylor’s expansion then guarantees that Rg (x0 ) Rg (x0 ) Rg (x0 ) Rg (x0 ) J3 ε2 + J1 ε2 log ε + I1 J1 ε2 − J1 ε2 log ε 6n 6 6n 6 Rg (x0 ) n J1 ε2 + O(ε4 log ε) − ( − 1) p 6n

I1 − n log ε −

≤ −n log ε +

n n Rg (x0 ) J2 2 n B p log(A I2 ) − ε + ε + O(εq ) . p p 6n I2 p A I2

After a suitable simplification, one arrives at n Rg (x0 ) I1 − log(A I2 ) ≤ − p 6n

n J2 n I1 J1 + − ( − 1)J1 − J3 ε2 + O(εp ) p I2 p



for ε > 0 small enough, so that I1 ≤

n log(A I2 ) . p

(12)

JID:YJFAN

AID:7289 /FLA

[m1L; v1.159; Prn:13/07/2015; 13:18] P.13 (1-29)

J. Ceccon, M. Montenegro / Journal of Functional Analysis ••• (••••) •••–•••

13

Thus, since u0 is an extremal function of (1), one has I1 =

n log(A0 (p) I2 ) , p

so that A ≥ A0 (p), and the conclusion of Proposition 3.1 readily follows. 2 4. Proof of the assertion (B) This section is devoted to the proof of the following theorem: Theorem 4.1. Let (M, g) be a smooth compact Riemannian manifold of dimension n ≥ 2. For each fixed 1 < p ≤ 2 and p < n, the best constant B(p, q, g) is bounded for any q < p close to p. Clearly, it suffices to prove this result for an arbitrary sequence (qk ) ⊂ (1, p) converging to p as k → +∞. Define Ck =

B(p, qk , g) − (p − qk ) . N (p, qk )

Since Ck < B(p, qk , g)N (p, qk )−1 , according to the definition of the best constant B(p, qk , g), we have νk = inf Jk (u) < N (p, qk )−1 ,

(13)

u∈H

where H = {u ∈ H 1,p (M ) : ||u||Lp (M ) = 1} and ⎛ Jk (u) = ⎝



 |∇g u|p dvg + Ck

M

⎞⎛ |u|p dvg ⎠ ⎝

M



k) ⎞ p(1−θ q θ k k

|u|qk dvg ⎠

.

M

As can easily be checked, the functional Jk is of class C 1 . The condition (13) and a usual argument of direct minimization lead to a minimizer uk ∈ H of Jk , so that νk := Jk (uk ) = inf Jk (u) . u∈H

(14)

Note that we can assume uk ≥ 0, since |∇g |uk || = |∇g uk | almost everywhere. In addition, since Jk is differentiable, uk satisfies the quasilinear elliptic equation Ak Δp,g uk + Ak Ck up−1 + k

1 − θk νk p−1 Bk ukqk −1 = u on M , θk θk k

(15)

JID:YJFAN 14

AID:7289 /FLA

[m1L; v1.159; Prn:13/07/2015; 13:18] P.14 (1-29)

J. Ceccon, M. Montenegro / Journal of Functional Analysis ••• (••••) •••–•••

where Δp,g = −divg (|∇g |p−2 ∇g ) is the p-Laplace operator of g, n(p − qk ) , np + pqk − nqk k) ⎛ ⎞ p(1−θ qk θk  , Ak = ⎝ uqkk dvg ⎠ θk =

M

⎛ Bk = ⎝



⎞⎛



|∇g uk |p dvg + Ck ⎠ ⎝

M

k ) −1 ⎞ p(1−θ q θ k k

uqkk dvg ⎠

.

M

So, by the strong maximum principle, uk is positive on M and, by Tolksdorf’s regularity theory [28], it follows that uk is of class C 1 . A relation that will be useful is  Bk uqkk dvg = νk . (16) M

Modulo a subsequence, we analyze two possible situations for the sequence (Ak ). (i) limk→+∞ Ak > 0; (ii) limk→+∞ Ak = 0. Assume that (i) occurs. Taking uk as a test function in (15) and after using Proposition 2.1, one gets Ak Ck ≤ νk ≤ c

(17)

for k large enough and some constant c independent of k, so that (Ck ) is bounded. Thus, the conclusion of Theorem 4.1 follows directly from the definition of Ck . The remaining of this section is dedicated to prove the boundedness of (Ck ) by assuming that the situation (ii) occurs. In this case, the inequality n 2

||uk ||L∞ (M ) Akp ≥ 1

(18)

follows from  1=

||uk ||pLp (M )

 upk

= M

dvg ≤

k ||uk ||p−q L∞ (M )

uqkk dvg M

and implies that (uk ) blows up in L∞ (M ). Part of the proof consists in performing a fine analysis of concentration of the sequence (uk ).

JID:YJFAN

AID:7289 /FLA

[m1L; v1.159; Prn:13/07/2015; 13:18] P.15 (1-29)

J. Ceccon, M. Montenegro / Journal of Functional Analysis ••• (••••) •••–•••

15

At first, we have lim νk = A0 (p)−1 .

(19)

k→+∞

In fact, a combination between the Nash inequality ⎛ ⎝





⎞ θ1

k

|u|p dvg ⎠



≤ ⎝N (p, qk )

M

 |∇g u|p dvg + B(p, qk , g)

M

⎛ ×⎝



⎞ |u|p dvg ⎠

M k) ⎞ p(1−θ q θ k k

|u|qk dvg ⎠

M

and the definition of Ck yields ⎛ N (p, qk )−1 ≤ ⎝



⎞⎛ |∇g uk |p dvg + Ck ⎠ ⎝

M



k) ⎞ p(1−θ q θ k k

uqkk dvg ⎠

+

(p − qk ) Ak N (p, qk )

M

= νk + Ak

p − qk . N (p, qk )

By Proposition 2.1, N (p, qk ) remains away from zero for k large enough. Therefore, lim inf νk ≥ A0 (p)−1 k→+∞

and so the limit (19) follows from (13). Let xk ∈ M be a maximum point of uk , that is uk (xk ) = ||uk ||L∞ (M ) .

(20)

The following concentration property satisfied by the sequence (uk ) plays an essential role in what follows. The main tools used in its proof are the blow-up method and Moser’s iteration. Lemma 4.1. We have  lim

upk dvg = 1 .

lim

σ→+∞ k→+∞ 1

B(xk ,σAkp )

Proof. Let σ > 0. For each x ∈ B(0, σ) and k large, we define

JID:YJFAN 16

AID:7289 /FLA

[m1L; v1.159; Prn:13/07/2015; 13:18] P.16 (1-29)

J. Ceccon, M. Montenegro / Journal of Functional Analysis ••• (••••) •••–••• 1

hk (x) = g(expxk (Akp x)) , n 2

1

ϕk (x) = Akp uk (expxk (Akp x)) . Note that the above expressions are well defined because we are assuming that Ak → 0 as k → +∞ (situation (ii)). By (15) and (16), one easily deduces that Δp,hk ϕk + Ak Ck ϕp−1 + k

1 − θk νk p−1 νk ϕkqk −1 = ϕ on B(0, σ) . θk θk k

Using the mean value theorem and the value of θk , one gets  pqk − nqk + np ρk p−1 qk −1 on B(0, σ) ϕk log ϕk Δp,hk ϕk + Ak Ck ϕk = νk ϕk + n for some ρk ∈ (qk − 1, p − 1). Consider ε > 0 fixed such that p + ε <

np n−p .

(21)

Since ϕρkk log(ϕεk ) ≤ ϕp−1+ε , we then get k

 pqk − nqk + np ε+p−1 qk −1 ϕ ϕ on B(0, σ) . Δp,hk ϕk + Ak Ck ϕp−1 ≤ ν + k k k k nε Once all coefficients of this equation are bounded, Moser’s iterative scheme (see [26]) produces   n p p p2 p (Ak ||uk ||L∞ (M ) ) = sup ϕk ≤ cσ ϕk dvhk = cσ upk dvg ≤ cσ B(0,σ)

B(0,2σ)

1

B(xk ,2σAkp )

for all k large enough, where cσ is a constant independent of q. So, using (18) in the above inequality, one readily obtains 1 ≤ ||ϕk ||L∞ (B(0,σ)) ≤ c1/p σ

(22)

for all k large enough. By (17), up to a subsequence, we have lim Ak Ck = C ≥ 0 .

k→+∞

Applying Tolksdorf’s elliptic theory to (21), thanks to (22), one easily checks that ϕk → ϕ 1 in Cloc (Rn ) and ϕ ≡ 0. Letting now k → +∞ in (21) and using (19), one gets   p Δp,ξ ϕ + Cϕp−1 = A0 (p)−1 ϕp−1 + ϕp−1 log(ϕp ) on Rn , n where Δp,ξ stands for the Euclidean p-Laplace operator.

(23)

JID:YJFAN

AID:7289 /FLA

[m1L; v1.159; Prn:13/07/2015; 13:18] P.17 (1-29)

J. Ceccon, M. Montenegro / Journal of Functional Analysis ••• (••••) •••–•••

17

For each σ > 0, we have 



 ϕpk dvhk = lim

ϕp dx = lim

k→+∞ B(0,σ)

B(0,σ)

upk dvg ≤ 1

k→+∞

(24)

1

B(xk ,σAkp )

and, by (13) and Proposition 2.1, 

 |∇ϕ| dx = lim p

k→+∞ B(0,σ)

B(0,σ)

|∇hk ϕk |p dvhk



⎞ 

⎜ ⎜ = lim ⎜Ak k→+∞ ⎝

⎟ ⎟ |∇g uk |p dvg ⎟ ≤ A0 (p)−1 . ⎠

1

(25)

B(xk ,σAkp )

In particular, one has ϕ ∈ W 1,p (Rn ). Consider then a sequence of nonnegative functions (ϕk ) ⊂ C0∞ (Rn ) converging to ϕ in W 1,p (Rn ). Taking ϕk as a test function in (23), we can write n A0 (p) p

 |∇ϕ|p−2 ∇ϕ · ∇ϕk dx +

Rn



ϕp−1 ϕk log(ϕp ) dx +

= Rn

n A0 (p)C p



n p

 ϕp−1 ϕk dx Rn

ϕp−1 ϕk dx Rn





ϕp−1 ϕk log(ϕp ) dx +

= [ϕ≤1]

ϕp−1 ϕk log(ϕp ) dx +

[ϕ≥1]

n p

 ϕp−1 ϕk dx. Rn

Letting k → +∞ and applying Fatou’s lemma and the dominated convergence theorem in the right-hand side, one gets n A0 (p) p







|∇ϕ|p dx ≤

Rn

ϕp log(ϕp ) dx + [ϕ≤1]

[ϕ≥1]



ϕp log(ϕp ) dx +

= Rn

Rewriting this inequality in function of ψ(x) = n A0 (p) p



Rn

ϕp log(ϕp ) dx +

n p



Note that this inequality combined with

ϕp dx Rn

Rn

ϕ(x) ||ϕ||p ,

ψ p log(ψ p ) dx + Rn



ϕp dx .

 |∇ψ|p dx ≤

n p

one has

n p

 ψ p dx + log ||ϕ||pp . Rn

JID:YJFAN 18

AID:7289 /FLA

[m1L; v1.159; Prn:13/07/2015; 13:18] P.18 (1-29)

J. Ceccon, M. Montenegro / Journal of Functional Analysis ••• (••••) •••–•••

 Rn

⎛ ⎞  n ψ p log(ψ p ) dx ≤ log ⎝A0 (p) |∇ψ|p dx⎠ p Rn

produces  A0 (p)

⎛ |∇ψ|p dx ≤ log ⎝A0 (p)

Rn



⎞ |∇ψ|p dx⎠ + 1 +

Rn

p log ||ϕ||pp . n

Since log x ≤ x−1 for all x > 0, we find log ||ϕ||pLp (Rn ) ≥ 0 or, equivalently, ||ϕ||Lp (Rn ) ≥ 1. Thus, by (24), we conclude that ||ϕ||Lp (Rn ) = 1, so that 

 lim

upk

lim

σ→∞ k→+∞ 1 B(xk ,σAkp

)

ϕp dx = 1 .

dvg =

2

Rn

By using the concentration property provided in Lemma 4.1, we now establish a pointwise estimate for the sequence (uk ). This result is the key ingredient in the final part of the proof of Theorem 4.1. The necessary tools in its proof are the same ones of the previous proof. Lemma 4.2. For any λ > 0, there exists a constant cλ > 0, independent of q < p, such that λ n p − p2

dg (x, xk )λ uk (x) ≤ cλ Ak

for all x ∈ M and k large enough, where dg stands for the distance with respect to the metric g. Proof. Suppose by contradiction that the above statement fails. Then, there exist λ0 > 0 and yk ∈ M for each k such that fk,λ0 (yk ) → +∞ as k → +∞, where n 2

fk,λ (x) = dg (x, xk )λ uk (x)Akp

−λ p

.

Without loss of generality, we assume that fk,λ0 (yk ) = ||fk,λ0 ||L∞ (M ) . From (22), we have fk,λ0 (yk ) ≤ c

λ λ − 0 − 0 uk (yk ) dg (xk , yk )λ0 Ak p ≤ cdg (xk , yk )λ0 Ak p , ||uk ||∞

so that 1 −p

dg (xk , yk )Ak

→ +∞ as k → +∞ .

(26)

JID:YJFAN

AID:7289 /FLA

[m1L; v1.159; Prn:13/07/2015; 13:18] P.19 (1-29)

J. Ceccon, M. Montenegro / Journal of Functional Analysis ••• (••••) •••–•••

19

For any fixed ε ∈ (0, 1) and σ > 0, we first claim that 1

B(yk , εd(xk , yk )) ∩ B(xk , σAkp ) = ∅

(27)

for k large enough. Clearly, this assertion follows from 1

dg (xk , yk ) ≥ σAkp + εd(xk , yk ) or equivalently, 1 −p

dg (xk , yk )(1 − ε)Ak

≥σ. 1 −p

But the above inequality is automatically satisfied, since dg (xk , yk )Ak k → +∞. We assert that exists a constant c > 0, independent of k, such that

→ +∞ as

uk (x) ≤ cuk (yk )

(28)

for all x ∈ B(yk , εdg (xk , yk )) and k large enough. Indeed, dg (x, xk ) ≥ dg (xk , yk ) − dg (x, yk ) ≥ (1 − ε)dg (xk , yk ) for all x ∈ B(yk , εdg (xk , yk )). Thus, n 2

dg (xk , yk )λ0 uk (yk )Akp



λ0 p

n 2

= fk,λ0 (yk ) ≥ fk,λ0 (x) = dg (x, xk )λ0 uk (x)Akp n 2

≥ (1 − ε)λ0 dg (xk , yk )λ0 uk (x) Akp



λ0 p

,

so that  uk (x) ≤

1 1−ε

λ0 uk (yk )

for all x ∈ B(yk , εdg (xk , yk )) and k large enough, as claimed. We now define 1

˜ k (x) = g(exp (A p x)) , h yk k n 2

1

ϕ˜k (x) = Akp uk (expyk (Akp x)) . By (15) and (16), it readily follows that Δp,hk ϕ˜k + Ak Ck ϕ˜p−1 + k

1 − θk νk p−1 νk ϕ˜kqk −1 = ϕ˜ on B(0, 3) . θk θk k



λ0 p

JID:YJFAN 20

AID:7289 /FLA

[m1L; v1.159; Prn:13/07/2015; 13:18] P.20 (1-29)

J. Ceccon, M. Montenegro / Journal of Functional Analysis ••• (••••) •••–•••

Applying the mean value theorem, one obtains  Δp,hk ϕ˜k +

Ak Ck ϕ˜p−1 k

= νk

ϕ˜kqk −1



pqk − nqk + np ρk ϕ˜k log(ϕ˜k ) + n

on B(0, 3) ,

(29)

where ρk ∈ (qk − 1, p − 1). np For fixed ε > 0 such that p + ε < n−p , one has ϕ˜ρkk log(ϕ˜εk ) ≤ ϕ˜p−1+ε . So, Moser’s k iterative scheme applied to (29) yields p p−qk

μk

n p2



= (Ak uk (yk )) ≤ sup p

B(0,1)

ϕ˜pk

 ϕ˜pk

≤c

upk dvg

dvh˜ k = c 1 B(yk ,Akp

B(0,2)

(30)

)

for k large enough, where  uqkk dvg .

μk = uk (yk )p−qk

(31)

M

We next analyze two independent situations that can occur: (I) μk ≥ 1 − θk for all k, up to a subsequence; (II) μk < 1 − θk for k large enough. In each case, we derive a contradiction. If the assertion (I) is satisfied, on the one hand, one has p p−qk

≥ e− p . n

lim inf μk k→+∞

On the other hand, by (26), one gets 1

B(yqk , Aqpk ) ⊂ B(yqk , εd(xqk , yqk ))

(32)

for k large enough. Thus, joining Lemma 4.1, (27) and (30), one arrives at the contradiction  n 0 < e− p ≤ lim upqk dvg = 0 . k→+∞

1

B(yqk ,Aqpk )

Assume then the assertion (II). In this case, for k large, we set 1

˜ k (x) = g(exp (A p x)) h yk k

1

ψk (x) = uk (yk )−1 uk (expyk (Akp x)) .

JID:YJFAN

AID:7289 /FLA

[m1L; v1.159; Prn:13/07/2015; 13:18] P.21 (1-29)

J. Ceccon, M. Montenegro / Journal of Functional Analysis ••• (••••) •••–•••

21

Thanks to (15) and (16), we have Δp,hk ψk + Ak Ck ψkp−1 +

1 − θk νk qk −1 νk p−1 ψk = ψ on B(0, 3) . θk μk θk k

Rewriting this equation as Δp,hk ψk + Ak Ck ψkp−1 +

νk θk



 1 − θk νk  p−1 − 1 ψkqk −1 = ψk − ψkqk −1 on B(0, 3) , μk θk

from the mean value theorem, one gets Δp,hk ψk + Ak Ck ψkp−1 + =

νk θk



1 − θk − 1 ψkqk −1 μk

νk (pqk − nqk + np) ρk ψk log(ψk ) on B(0, 3) n

(33)

for some ρk ∈ (qk − 1, p − 1). k Using (26), (28) and the fact that 1−θ μk − 1 > 0 for k large, one easily deduces that 1,p ψk → ψ in W (B(0, 2)) and, by Moser’s iteration, that ψ ≡ 0. Let h ∈ C01 (B(0, 2)) be a fixed cutoff function such that h ≡ 1 in B(0, 1) and h ≥ 0. Taking ψhp as a test function in (33), one obtains lim sup k→+∞

1 θk





1 − θk −1 μk


Therefore, up to a subsequence, we can write 1 lim k→+∞ θk





1 − θk −1 μk

=γ≥0,

so that μk → 1. Using the definition of θk , we then derive p p−qk

lim μk

k→+∞

= e−(1+γ) p . n

At last, combining Lemma 4.1, (27), (30) and the above limit, we obtain the contradiction 0 < e−(1+γ) p ≤ lim



n

upk dvg = 0 .

k→+∞ 1 B(yk ,Akp

2

)

Finally, we turn our attention to the final argument of the proof of Theorem 4.1. We recall that our goal is to prove that the sequence (Ck ) is bounded by assuming that Ak → 0 as k → +∞ (situation (ii)). This step consists of several integral estimates

JID:YJFAN

AID:7289 /FLA

[m1L; v1.159; Prn:13/07/2015; 13:18] P.22 (1-29)

J. Ceccon, M. Montenegro / Journal of Functional Analysis ••• (••••) •••–•••

22

around the maximum point xk of uk and Lemma 4.2 plays a central role on some of them. In what follows, several possibly different positive constants independent of k will be denoted by c or c1 . Assume, without loss of generality, that the radius of injectivity of M is greater than 2. Let η ∈ C01 (R) be a cutoff function such that η = 1 on [0, 1), η = 0 on [2, ∞) and 0 ≤ η ≤ 1 and define ηk (x) = η(dg (x, xk )). The sharp Euclidean Lp -Nash inequality asserts that ⎛ ⎜ ⎝



⎞ θ1



k

⎟ (uk ηk ) dx⎠ p

⎜ ≤ N (p, qk ) ⎝

B(0,2)

⎞⎛



⎟⎜ |∇(uk ηk )| dx⎠ ⎝

k) ⎞ p(1−θ q θ



k k

p

B(0,2)

qk

(uk ηk )

⎟ dx⎠

B(0,2)

Expanding the metric g in normal coordinates around xk , one locally gets (1 − cdg (x, xk )2 )dvg ≤ dx ≤ (1 + cdg (x, xk )2 )dvg and |∇(uk ηk )|p ≤ |∇g (uk ηk )|p (1 + c dg (x, xk )2 ) . Thus, ⎛ ⎜ ⎝



⎞ θ1



k

⎟ upk ηkp dx⎠



⎜ ≤ ⎝N (p, qk )Ak

B(0,2)

|∇g (uk ηk )|p dvg B(xk ,2)





⎟ |∇g (uk ηk )|p dg (x, xk )2 dvg ⎠

+ cAk B(xk ,2)

 ×



B(0,2)

uqkk ηkqk dx

uqk M k

k) p(1−θ q θ k k

dvg

.

Using now the inequality |∇g (uk ηk )|p ≤ |∇g uk |p ηkp + c|ηk ∇g uk |p−1 |uk ∇g ηk | + c|uk ∇g ηk |p , and denoting  ηkp |∇g uk |p dg (x, xk )2 dvg

Xk = Ak M

.

JID:YJFAN

AID:7289 /FLA

[m1L; v1.159; Prn:13/07/2015; 13:18] P.23 (1-29)

J. Ceccon, M. Montenegro / Journal of Functional Analysis ••• (••••) •••–•••

23

and  |∇g uk |p−1 |∇g ηk |uk dvg ,

Yk = Ak M

we deduce that ⎛ ⎜ ⎝

⎞ θ1



k

upk ηkp

⎟ dx⎠

⎛ ≤ ⎝N (p, qk )Ak



 |∇g uk |p ηkp

dvg + cXk + cYk + cAk ⎠

M

B(0,2)

 

B(0,2)

×

uqkk ηkqk dx

uqk M k

k) p(1−θ q θ k k

.

dvg

(34)

On the other hand, choosing uk ηkp as a test function in (15) and using (16), one gets  N (p, qk )Ak M

⎛ ⎞   qk p u η dv 1 g ⎝ up η p dvg − M ⎠  k qk k |∇g uk |p ηkp dvg ≤ 1 − N (p, qk )Ak Ck + k k θk u dvg M k M



|∇g uk |p−1 |∇g ηk |uk dvg .

+c M

Using (13), Proposition 2.1 and Lemma 4.2 with a suitable value of λ, the last integral can be estimated as 

⎛ |∇g uk |p−1 |∇g ηk |uk dvg ≤ c ⎝

M



⎛ ⎞ p−1 p ⎜ |∇g uk |p dvg ⎠ ⎝

M



⎞ p1 ⎟ upk dvg ⎠ ≤ cAk ,

B(xk ,2)\B(xk ,1)

so that  N (p, qk )Ak M

⎛ ⎞   qk p u η dv 1 g ⎝ up η p dvg − M ⎠ + cAk  k qk k |∇g uk |p ηkp dvg ≤ 1 − c1 Ak Ck + k k θk u dvg M k M

(35) for k large enough. Let ⎛ ⎞   qk qk u η dv 1 ⎝ g ⎠ .  k qkk Zk = upk ηkp dvg − M θk u dv g M k M

By the mean value theorem and Lemma 4.2, there exists γk ∈ (qk , p) such that

JID:YJFAN

AID:7289 /FLA

[m1L; v1.159; Prn:13/07/2015; 13:18] P.24 (1-29)

J. Ceccon, M. Montenegro / Journal of Functional Analysis ••• (••••) •••–•••

24

 ⎛ ⎞       qk p q q p   u η dv 1  M ukk (ηkk − ηk ) dvg  1 g k k  Zk − ⎝ up η p dvg − M ⎠   k k  ≤ θk    θk uqk dvg uqk dvg   M k M k M  γ q pqk − nqk + np M ηk k | log ηk |ukk dvg  ≤ n uqk dvg M k  uqk dvg B(xk ,2)\B(xk ,1) k  ≤c ≤ cAk (36) uqk dvg M k for k large enough. Plugging (36) into (35) and after (35) into (34), one arrives at ⎛ ⎜ ⎝



⎞ θ1



k

⎟ upk ηkp dx⎠

≤ (1 − cAk Ck + Zk + cXk + cYk + cAk )



B(0,2)

B(0,2)

uqkk ηkqk dx

uqk M k

k) p(1−θ q θ k k

dvg (37)

for k large enough. In order to estimate Xk , we take uk d2g ηkp as a test function in (15). From this choice, we derive ⎛ ⎞   qk qk 2 u η d (x, x ) dv νk ⎝ g k g p p ⎠ Xk ≤ uk ηk dg (x, xk )2 dvg − M k  k qk θk u dvg M k M

 uk ηkp |∇g uk |p−1 dg (x, xk ) dvg

+ cAk M

 +c

M

uqkk ηkqk dg (x, xk )2 dvg  + cYk + cAk . uqk dvg M k

We now estimate the first two terms of the right-hand side above. Namely, after a change of variable, one has    qk qk 2  p p 1 u η dg (x, xk )2 − uk ηk dqg (x, xk )  dvg  k k θk u k dvg  M k M



1 ≤ cAk θk 2 p

|ϕpk η˜kp − ϕqkk η˜kqk | |x|2 dx −1 p

B(0,2Ak 2 p



)

(ϕk η˜k )ρk |log(ϕk η˜k )| |x|2 dx

= cAk

−1 p

B(0,2Ak

) 1

for some ρk ∈ (qk , p), where η˜k (x) = ηk (Akp x).

JID:YJFAN

AID:7289 /FLA

[m1L; v1.159; Prn:13/07/2015; 13:18] P.25 (1-29)

J. Ceccon, M. Montenegro / Journal of Functional Analysis ••• (••••) •••–•••

25

Thus, using Lemma 4.2 and the assumption p ≤ 2, one obtains    qk qk 2  p p u η dg (x, xk )2 − uk ηk dqg (x, xk )  dvg ≤ cAk .  k k u k dvg 

1 θk

M

(38)

k

M

In particular, since 

 upk ηkp dg (x, xk )2 M

dvg −

M

  upk ηkp dg (x, xk )2 −

= M

uqkk ηkqk dg (x, xk )2 dvg  uqk dvg M k uqkk ηkqk dg (x, xk )2  uqk dvg M k

dvg ,

we have 1 θk

     qk qk 2   u η d (x, x ) dv g k g  up η p dg (x, xk )2 dvg − M k  k q  ≤ cAk k k k   u dv g   k M M

for k large enough. Besides, thanks to (13), Proposition 2.1, Lemma 4.2 and the fact that p ≤ 2, we derive 





uk ηkp |∇g uk |p−1 dg (x, xk ) dvg ≤ c ⎝

M

⎛ ⎞ p−1 p ⎜ |∇g uk |p dvg ⎠ ⎝

M

≤ cAk

⎞ p1 ⎟ upk dg (x, xk )p dvg ⎠

B(xk ,2)

⎞ p−1 p

⎛ 2−p p



⎜ ⎜ ⎜ ⎝



−1 B(0,2Ak p

⎟ ⎟ ϕpk |x|p dx⎟ ⎠

≤c

)

for k large enough. So, the above estimates guarantee that  Xk ≤ c

M

uqkk ηkqk dg (x, xk )2 dvg  + cYk + cAk . uqk dvg M k

Evoking again Lemma 4.2 and the condition p ≤ 2, one has  M

2 uqkk ηkqk dg (x, xk )2 dvg  ≤ cAkp qk u dvg M k

 ϕqkk |x|2 dhk ≤ cAk . −1/p B(0,2Ak )

Also, it follows directly from (13) and Proposition 2.1 that

(39)

JID:YJFAN

AID:7289 /FLA

[m1L; v1.159; Prn:13/07/2015; 13:18] P.26 (1-29)

J. Ceccon, M. Montenegro / Journal of Functional Analysis ••• (••••) •••–•••

26



1

upk dvg ≤ cAk ,

Yk ≤ cAkp

(40)

B(xk ,2)\B(xk ,1)

so that Xk ≤ cAk .

(41)

Thus, plugging (39), (40) and (41) into (37), one gets ⎛ ⎜ ⎝

⎞ θ1





k

⎟ upk ηkp dx⎠



B(0,2)

≤ (1 + Zk − c1 Ak Ck + cAk )

uqk M k

B(0,2)

for k large enough. Taking logarithm of both sides and using the fact that ⎛ 1 ⎜ ⎝log θk



 upk ηkp dx − log

B(0,2)

(uk ηk )qk dx

p(1−θk ) qθk

k) p(1−θ q θ k k

dvg

=

1 θk



n−p n ,

one has



uqk η qk dx ⎟ B(0,2) k k  ⎠ uqk dvg M k

n−p log ≤ log(1 + Zk − c1 Ak Ck + cAk ) − n

 

B(0,2) M

uqkk ηkqk dx

uqkk dvg

.

(42)

By the mean value theorem,



 log B(0,2)

upk ηkp dx − log

uqk η qk B(0,2) k k  uqk dvg M k

dx



=

1 ⎜ ⎝ τk



 upk ηkp dx −

B(0,2)

uqk η qk B(0,2) k k  uqk dvg M k

⎞ dx ⎟ ⎠ (43)

for some number τk between the expressions 

 upk ηkp

dx and

B(0,2)



B(0,2) M

uqkk ηkqk dx

uqkk dvg

.

Using Cartan’s expansion of g in normal coordinates around xk and Lemma 4.2, one obtains ⎫ ⎧    ⎪   ⎪ qk qk  qk qk    ⎨  u η dx u η dv ⎬   B(0,2) k k   k qkk g  ≤ cAk (44) max  upk ηkp dx − upk ηkp dvg  ,   − M qk   ⎪ u dvg u dvg ⎪ ⎭ ⎩ M k M k  M B(0,2) for k large enough. Indeed, since p ≤ 2,

JID:YJFAN

AID:7289 /FLA

[m1L; v1.159; Prn:13/07/2015; 13:18] P.27 (1-29)

J. Ceccon, M. Montenegro / Journal of Functional Analysis ••• (••••) •••–•••

         2   p p p p uk ηk dx − uk ηk dvg  ≤ c upk ηkp dg (x, xk )2 dvg ≤ cAkp    B(0,2)  M M

27

 ϕpk |x|2 dvhk −1/p

B(0,2Ak

)

≤ cAk and, by (39),      uqkk ηkqk dx uqkk ηkqk dvg  uqk η qk d (x, xk )2 dvg  B(0,2) M M k k g  ≤ c − ≤ cAk .     uqk dvg uqk dvg  uqk dvg M k M k M k Moreover, we also have ⎫ ⎧     ⎬ qk qk ⎨     u η dvg  k qkk max  (uk ηk )p dvg − 1 ,  M − 1 ≤ cAk ⎭ ⎩ u dvg  M k

(45)

M

for k large enough. In fact, by Lemma 4.2,               up η p dvg − 1 =  up η p dvg − up dvg  ≤ c k k k k k         M

and

M

 upk dvg ≤ cAk

M \B(xk ,1)

M

   uqk dvg   M uqkk ηkqk dvg M \B(xk ,1) k     − 1 ≤ c ≤ cAk .  uqk dvg uqk dvg k

M

M

k

Thanks to (44) and (45), one easily deduces that τk−1 = 1 + O(Ak ). Then, by (43), ⎛ 1 ⎜ ⎝log θk

uqk η qk B(0,2) k k  uqk dvg M k

upk ηkp dx − log

B(0,2)



=











1 ⎜ ⎝ θk

upk ηkp dx −

B(0,2)

dx ⎟ ⎠ (1 + O(Ak )) .

1 ⎜ ⎝ θk

upk ηkp dx −

B(0,2)



=







1 ⎜ ⎝ θk

uqk η qk B(0,2) k k  uqk dvg M k

B(0,2)

dx ⎟ ⎠ ⎞

 upk ηkp −

uqk η qk  k qk k u dvg M k

dx ⎟ ⎠ ⎞

uqk η qk B(0,2) k k  uqk dvg M k

But, by Cartan’s expansion and (38), we have ⎛



⎟ dx⎠

(46)

JID:YJFAN 28

AID:7289 /FLA

[m1L; v1.159; Prn:13/07/2015; 13:18] P.28 (1-29)

J. Ceccon, M. Montenegro / Journal of Functional Analysis ••• (••••) •••–•••

⎛ ⎞  qk qk η u 1 ⎝ = upk ηkp −  k qkk dvg ⎠ θk u dv g k M M

⎛ 1 ⎝ + θk

  upk ηkp M

uqk η qk −  k qk k u dvg M k



⎞ O(dg (x, xk )2 ) dvg ⎠

= Zk + O(Ak ) for k large enough. Replacing this inequality in (46), one obtains ⎛ 1 ⎜ ⎝log θk



 upk ηkp dx − log

B(0,2)

uqk η qk B(0,2) k k  uqk dvg M k

⎞ dx ⎟ ⎠ ≥ Zk − cAk .

In turn, plugging the above inequality in (42), one has    uqk η qk dx  n − p  B(0,2) k k  Zk − cAk ≤ log(1 + Zk − c1 Ak Ck + cAk ) +  . log  n  uqk dvg M k Finally, using Cartan’s expansion of g in normal coordinates, Taylor’s expansion of the function log and Lemma 4.2, one gets       uqk η qk dx  uqk dvg uqk η qk dg (x, xk )2 dvg B(0,2) k k M \B(xk ,1) k    + c M k  k qk ≤ cAk . log ≤c qk qk   u dvg u dvg u dvg M k M k M k In short, for a certain constant c > 0, we deduce that Zk ≤ log(1 + Zk − c1 Ak Ck + cAk ) + cAk for k large enough. Since log x ≤ x − 1 for all x > 0, we find Zk ≤ Zk − c1 Ak Ck + cAk for k large enough, so that the sequence (Ck ) is bounded and the proof of Theorem 4.1 follows. 2 Acknowledgments The authors are indebted to the referee for his valuable suggestions and comments pointed out concerning this work. The first author was partially supported by CAPES (grant BEX 6961/14-2) through INCTmat and the second one was partially supported by CNPq (grant PQ 306406/2013-6) and FAPEMIG (grant PPM 00223-13).

JID:YJFAN

AID:7289 /FLA

[m1L; v1.159; Prn:13/07/2015; 13:18] P.29 (1-29)

J. Ceccon, M. Montenegro / Journal of Functional Analysis ••• (••••) •••–•••

29

References [1] T. Aubin, Some Nonlinear Problems in Riemannian Geometry, Springer Monogr. Math., SpringerVerlag, Berlin, 1998. [2] D. Bakry, L’hypercontractivité et son utilisation en théorie des semigroupes, in: Lectures on Probability Theory: École D’été de Probabilités de St-Flour 1992, in: Lecture Notes in Mathematics, vol. 1581, Springer, Berlin, 1994, pp. 1–114. [3] D. Bakry, T. Coulhon, M. Ledoux, L. Saloff-Coste, Sobolev inequalities in disguise, Indiana Univ. Math. J. 44 (4) (1995) 1033–1074. [4] W. Beckner, Geometric proof of Nash’s inequality, Int. Math. Res. Not. IMRN 2 (1998) 67–71. [5] W. Beckner, Geometric asymptotics and the logarithmic Sobolev inequality, Forum Math. 11 (1) (1999) 105–137. [6] S. Bobkov, I. Gentil, M. Ledoux, Hypercontractivity of Hamilton–Jacobi equations, J. Math. Pures Appl. 80 (2001) 669–696. [7] S. Bobkov, I. Gentil, M. Ledoux, Hypercontractivity of Hamilton–Jacobi equations. Inhomogeneous random systems, Markov Process. Related Fields 8 (2002) 233–235. [8] C. Brouttelande, The best-constant problem for a family of Gagliardo–Nirenberg inequalities on a compact Riemannian manifold, Proc. R. Soc. Edinb. 46 (2003) 147–157. [9] C. Brouttelande, On the second best constant in logarithmic Sobolev inequalities on complete Riemannian manifolds, Bull. Sci. Math. 127 (2003) 292–312. [10] E.A. Carlen, Superadditivity of Fisher’s information and logarithmic Sobolev inequalities, J. Funct. Anal. 101 (1) (1991) 194–211. [11] J. Ceccon, General optimal Lp -Nash inequalities on Riemannian manifolds, Ann. Sc. Norm. Super. Pisa Cl. Sci. (2015), http://dx.doi.org/10.2422/2036-2145.201311_003, in press. [12] J. Ceccon, M. Montenegro, Optimal Lp -Riemannian Gagliardo–Nirenberg inequalities, Math. Z. 258 (2008) 851–873. [13] J. Ceccon, M. Montenegro, Optimal Riemannian Lp -Gagliardo–Nirenberg inequalities revisited, J. Differential Equations 254 (2013) 2532–2555. [14] M. Del Pino, J. Dolbeault, The optimal Euclidean Lp -Sobolev logarithmic inequality, J. Funct. Anal. 197 (2003) 151–161. [15] M. Del Pino, J. Dolbeault, I. Gentil, Nonlinear diffusions, hypercontractivity and the optimal Lp -Euclidean logarithmic Sobolev inequality, J. Math. Anal. Appl. 293 (2004) 375–388. [16] O. Druet, E. Hebey, The AB program in geometric analysis: sharp Sobolev inequalities and related problems, Mem. Amer. Math. Soc. 160 (761) (2002). [17] O. Druet, E. Hebey, M. Vaugon, Optimal Nash’s inequalities on Riemannian manifolds: the influence of geometry, Int. Math. Res. Not. IMRN 14 (1999) 735–779. [18] I. Gentil, Ultracontractive bounds on Hamilton–Jacobi solutions, Bull. Sci. Math. 126 (2002) 507–524. [19] I. Gentil, The general optimal Lp -Euclidean logarithmic Sobolev inequality by Hamilton–Jacobi equations, J. Funct. Anal. 202 (2003) 591–599. [20] L. Gross, Logarithmic Sobolev inequalities, Amer. J. Math. 97 (1975) 1061–1083. [21] E. Humbert, Best constant in the L2 -Nash inequality, Proc. Roy. Soc. Edinburgh Sect. A 131 (2001) 621–646. [22] M. Ledoux, Isoperimetry and Gaussian analysis, in: Lectures on Probability Theory and Statistics: École D’été de Probabilités de St-Flour 1994, in: Lecture Notes in Mathematics, vol. 1648, Springer, Berlin, 1996, pp. 165–294. [23] M. Ledoux, Concentration of measure and logarithmic Sobolev inequalities, in: Séminaire de Probabilités XXXIII, in: Lecture Notes in Mathematics, vol. 1709, Springer, 1999, pp. 120–216. [24] J. Nash, Continuity of solutions of parabolic and elliptic equations, Amer. J. Math. 80 (1958) 931–954. [25] G. Perelman, The entropy formula for the Ricci flow and its geometric applications, arXiv:math/ 0211159v1. [26] J. Serrin, Local behavior of solutions of quasilinear equations, Acta Math. 111 (1964) 247–302. [27] M. Talagrand, Transportation cost for Gaussian and other product measures, Geom. Funct. Anal. 6 (1996) 587–600. [28] P. Tolksdorf, Regularity for a more general class of quasilinear elliptic equations, J. Differential Equations 51 (1) (1984) 126–150. [29] F.B. Weissler, Logarithmic Sobolev inequalities for the heat-diffusion semigroup, Trans. Amer. Math. Soc. 237 (1978) 255–269.