Home On the Blow-Up of Solutions to Liouville-Type Equations
Article Publicly Available

On the Blow-Up of Solutions to Liouville-Type Equations

  • Tonia Ricciardi EMAIL logo and Gabriella Zecca
Published/Copyright: December 8, 2015

Abstract

We estimate some complex structures related to perturbed Liouville equations defined on a compact Riemannian 2-manifold. As a byproduct, we obtain a quick proof of the mass quantization and we locate the blow-up points.

MSC 2010: 35J20; 35J60

1 Introduction

In the article [6], Nagasaki and Suzuki considered the Liouville-type problem

(1.1) { - Δ u = ρ f ( u ) in Ω , u = 0 on Ω ,

where Ω2 is a smooth bounded domain, ρ>0, and f: is a smooth function such that

(1.2) f ( t ) = e t + φ ( t ) with φ ( t ) = o ( e t ) as t + .

Equations of the form (1.1) are of actual interest in several contexts, including turbulent Euler flows, chemotaxis, and the Nirenberg problem in geometry; see, e.g., [5] and the references therein. A recent example is given by the mean field equation

(1.3) { - Δ u = λ [ - 1 , 1 ] α e α u 𝒫 ( d α ) [ - 1 , 1 ] × Ω e α u 𝒫 ( d α ) 𝑑 x in Ω , u = 0 on Ω ,

which was derived in [7] for turbulent flows with variable intensities, where 𝒫([-1,1]) is a probability measure related to the vortex intensity distribution. In this case, setting

f ( t ) = [ - 1 , 1 ] α e α t 𝒫 ( d α ) , ρ = λ ( [ - 1 , 1 ] × Ω e α u 𝒫 ( d α ) 𝑑 x ) - 1 ,

it is readily seen that if 𝒫({1})>0, then along a blow-up sequence, (1.3) is of the form (1.1). See [10, 11, 13, 12] for more details, where the existence of solutions by variational arguments and blow-up analysis are also considered. Blow-up solution sequences for (1.3) have also been recently constructed in [9] following the approach introduced in [4].

In [6], Nagasaki and Suzuki derived a concentration-compactness principle for (1.1), mass quantization, and the location of blow-up points, under some additional technical assumptions for f. More precisely, they assumed that

(1.4) | φ ( t ) - φ ( t ) | 𝒢 ( t ) for some 𝒢 C 1 ( , ) satisfying 𝒢 ( t ) + | 𝒢 ( t ) | C e γ t with γ < 1 4

and

(1.5) f ( t ) 0 for all t 0 .

By a complex analysis approach, they established the following result.

Theorem 1.1

Theorem 1.1 ([6])

Let f satisfy assumptions (1.2), (1.4), and (1.5). Let un be a solution sequence to (1.1) with ρ=ρn0. Suppose un converges to some nontrivial function u0. Then,

u 0 ( x ) = 8 π j = 1 m G Ω ( x , p j )

for some p1,,pmΩ, m, where GΩ denotes the Green’s function for the Dirichlet problem on Ω. Furthermore, at each blow-up point pj, j=1,,m, there holds that

[ G Ω ( x , p j ) + 1 2 π log | x - p j | ] | x = p j + [ i j G Ω ( p j , p i ) ] = 0 .

The original estimates in [6] are involved and require the technical assumption γ(0,14). It should be mentioned that this assumption was later weakened to the natural assumption γ(0,1) in [14], by taking a different viewpoint on the line of [1].

Here, we are interested in revisiting the complex analysis framework introduced in [6]. In particular, we study the effect of the lower-order terms which naturally appear when the equation is considered on a compact Riemannian 2-manifold. We observe that, although the very elaborate key L-estimate obtained in [6], namely, Proposition 1.2 below, may be extended in a straightforward manner to the case of manifolds (see Appendix A for the details), the lower-order terms are naturally estimated only in L1. Therefore, we are led to consider an L1-framework, which turns out to be significantly simpler and which holds under the weaker assumption γ(0,12). As a byproduct, we obtain a quick proof of mass quantization and blow-up point location for the case γ(0,12).

In order to state our results, for a function uC2(Ω), we define the quantity

(1.6) S ( u ) = u z 2 2 - u z z ,

where

z = x - i y 2 , z ¯ = x + i y 2 .

Then, if u is a solution to (1.1), we have

z ¯ [ S ( u ) ] = - ρ 4 u z [ f ( u ) - f ( u ) ] = ρ 4 u z [ φ ( u ) - φ ( u ) ] .

In particular, in the Liouville case f(u)=eu, the function S(u) is holomorphic. Therefore, the complex derivative z¯[S(u)] may be viewed as an estimate of the “distance” between the equation in (1.1) and the standard Liouville equation.

We recall that the main technical estimate in [6] is given by the following proposition.

Proposition 1.2

Proposition 1.2 ([6])

Let uρ be a blow-up sequence for (1.1). Assume (1.2), (1.4), and (1.5). Then,

z ¯ S ( u ) L ( Ω ) = ρ 4 u ρ ( f ( u ρ ) - f ( u ρ ) ) L ( Ω ) 0 .

It is natural to expect that corresponding results should hold on a compact Riemannian 2-manifold (M,g) without boundary. We show that, in fact, the L-convergence as stated in Proposition 1.2 still holds true on M (see Proposition A.1 in Appendix A). However, a modified point of view is needed in order to suitably locally define a function S corresponding to (1.6), such that the lower-order terms may be controlled, as well as to prove its convergence to a holomorphic function in some suitable norm, so that the mass quantization and the location of the blow-up points may be derived. As we shall see, our point of view holds under the weaker assumption γ(0,12) and is significantly simpler than the original L-framework.

More precisely, on a compact Riemannian 2-manifold without boundary (M,g), we consider the problem

(1.7) { - Δ g u = ρ f ( u ) - c ρ in M , M u 𝑑 x = 0 ,

where cρ=ρ|M|-1Mf(u)𝑑x, dx denotes the volume element on M, and Δg denotes the Laplace–Beltrami operator. We assume that f(t)=et+φ(t) satisfies (1.2) and, moreover, that

(1.8) | φ ( t ) - φ ( t ) | 𝒢 ( t ) for some 𝒢 C 1 ( , ) satisfying 𝒢 ( t ) + | 𝒢 ( t ) | C e γ t with γ < 1 2

and

(1.9) f ( t ) - C for all t 0 .

In the spirit of [3], we assume that along a blow-up sequence we have

(1.10) ρ M f ( u ) 𝑑 x C .

In particular, without loss of generality, we may assume that

(1.11) c ρ c 0 as ρ 0 + .

We note that (1.9) implies that u-C. We now define the modified quantity corresponding to S(u). Let 𝒮={p1,,pm} denote the blow-up set. Let p𝒮 and denote X=(x1,x2). We consider a local isothermal chart (Ψ,𝒰) such that ε(p)𝒰, Ψ(p)=0, ε(p)𝒮=, g(X)=eξ(X)(dx12+dx22), and ξ(0)=0. For the sake of simplicity, we identify here functions on M with their pullback functions to B=B(0,r)=Ψ(ε(p)). We denote by GB(X,Y) the Green’s function of ΔX=x12+x22 on B. We set

(1.12) K ( X ) = - B G B ( X , Y ) e ξ ( Y ) 𝑑 Y + c 1 z

with c1 defined by

(1.13) z [ ξ ( z , z ¯ ) + c 0 K ( z , z ¯ ) ] | z = 0 = 0 ,

where c0 is defined in (1.11). Let u denote a solution sequence to (1.7). We define w(z)=u-cρK, so that -Δw=eξρf(u) in B. Finally, consider S(w), where S is defined in (1.6). Our main estimate is given in the following theorem.

Theorem 1.3

Assume that f(t)=et+φ(t) satisfies (1.2), (1.8), and (1.9). Let uρ be a blow-up solution sequence for (1.7). Then,

  1. for every 1s<(γ+12)-1,

    ρ u ρ ( f ( u ρ ) - f ( u ρ ) ) L s ( M ) 0 as ρ 0 + ;

  2. for every blow-up point p𝒮, the function S(w)S0 in L1(B) as ρ0+, where S0 is holomorphic in B.

Consequently, we derive the following corollary.

Corollary 1.4

Assume that f(t)=et+φ(t) satisfies (1.2), (1.8), and (1.9). Suppose un converges to some nontrivial function u0. Then,

(1.14) u 0 ( x ) = 8 π j = 1 m G M ( x , p j ) .

Moreover, for all p𝒮, we have the relation

(1.15) [ X ( q 𝒮 { p } G M ( Ψ - 1 ( X ) , q ) + G M ( Ψ - 1 ( X ) , p ) + 1 2 π log | X | + 1 8 π ξ ( X ) ) ] | X = 0 = 0 .

We provide the proofs of Theorem 1.3 and Corollary 1.4 in Section 2. For the sake of completeness and in order to readily allow a comparison with the L-framework employed in [6], in Appendix A we extend Proposition 1.2 to the case of Riemannian 2-manifolds without boundary.

Throughout this note, we denote by C>0 a constant whose actual value may vary from line to line.

2 Proof of Theorem 1.3

We begin by establishing the following result.

Lemma 2.1

Let u be a solution to (1.7). For every r>0, we have

(2.1) r M e - r u | u | 2 𝑑 x C ,

where C=C(r,M,φ,c0).

Proof.

We multiply the equation -Δgu=ρf(u)-cρ by e-ru. Integrating, we have

r M e - r u | u | 2 𝑑 x = M e - r u Δ g u 𝑑 x
= - ρ M e - r u f ( u ) 𝑑 x + c ρ M e - r u 𝑑 x
ρ M e - r u | φ ( u ) | 𝑑 x + c ρ M e r C 𝑑 x
ρ M e - r u | φ ( u ) | 𝑑 x + c ρ e r C | M | ,

since u-C. Using the assumptions on φ, there exists t0>0 such that |g(u)|<eu for u>t0, so that

r M e - r u | u | 2 𝑑 x C + ρ ( { u > t 0 } e ( 1 - r ) u 𝑑 x + { u t 0 } e - r u | φ ( u ) | 𝑑 x ) C + ρ ( M e u 𝑑 x + { u t 0 } e - r u | φ ( u ) | 𝑑 x ) ,

and the claim follows using again the fact that u-C. ∎

The following proposition proves Theorem 1.3(i).

Proposition 2.2

Let u be a solution to (1.7). Then, for every 1s<(γ+12)-1 and for every ε>0, we have

u ( f ( u ) - f ( u ) ) L s ( M ) C ρ - γ - ε

for 0<ρ<1.

Proof.

In view of (1.8), we have

0 | f ( u ) - f ( u ) | C e γ u .

Hence,

(2.2) ( f ( u ) - f ( u ) ) u L s C e γ u u L s .

Moreover, (1.10) implies that

M e u 𝑑 x c ρ - 1 .

Then, for every 1q<γ-1, using Hölder’s inequality we have

(2.3) e γ u L q ( M ) C | M | 1 q - γ ρ - γ .

Let 0<r<1-s(γ+12). By Lemma 2.1, for

q = s + r γ 1 - s 2 < 1 γ ,

using Hölder’s inequality again, we have

(2.4) e γ u u L s ( M ) s = M e ( s γ + r ) u ( e - r u | u | s ) 𝑑 x ( M e γ u q 𝑑 x ) 1 - s 2 ( M e - 2 r u | u | 2 𝑑 x ) s 2 C e γ u L q ( M ) s + r γ .

Then, by (2.3) and (2.4) we have

(2.5) e γ u u L s ( M ) C ρ - γ - r s .

Combining (2.2) and (2.5), the claim is proved. ∎

Let p𝒮. We denote by (Ψ,𝒰) an isothermal chart satisfying

𝒰 ¯ 𝒮 = { p } , Ψ ( 𝒰 ) = 𝒪 2 , Ψ ( p ) = 0 , g ( X ) = e ξ ( X ) ( d x 1 2 + d x 2 2 ) , ξ ( 0 ) = 0 ,

where X=(x1,x2) denotes a coordinate system on 𝒪. We consider ε>0 sufficiently small so that (p,ε)𝒰 and let B=B(0,r)=Ψ((p,ε)). The Laplace–Beltrami operator Δg is then mapped to the operator e-ξ(X)ΔX on 𝒪, where ΔX=x122+x222. By GB(X,Y) we denote the Green’s function of ΔX on B, namely,

{ - Δ X G B ( X , Y ) = δ Y in B , G B ( X , Y ) = 0 on B .

We recall from (1.12) that

K ( X ) = - B G B ( X , Y ) e ξ ( Y ) 𝑑 Y + c 1 z

with c1 the constant defined by (1.13), namely,

z [ ξ ( z , z ¯ ) + c 0 K ( z , z ¯ ) ] | z = 0 = 0 ,

where c0=limρ0cρ. Then, KC(B) and

Δ X K = e ξ in B ¯ .

Let uρ be a blow-up solution sequence for (1.7). As ρ0, uu0 in Cloc(M𝒮), u-u0W1,q(M) for 1q<2, and f(u)f(u0) in Cloc(M𝒮), we have ΔguΔgu0 in Cloc(M𝒮), so that

Δ g u 0 = c 0 in M 𝒮 .

We consider the following functions defined in B:

(2.6)

u ~ = u Ψ - 1 , u ~ 0 = u 0 Ψ - 1 ,
w ( z ) = u ~ - c ρ K , w 0 ( z ) = u ~ 0 - c 0 K ,
S ( w ) = w z z - 1 2 w z 2 , S 0 = w 0 z z - 1 2 w 0 z 2 .

The following proposition proves Theorem 1.3(ii).

Proposition 2.3

The complex function S0 defined in (2.6) is holomorphic in B and SS0 in L1(B).

Proof.

By (2.6) we have

- Δ X w = ρ f ( u ~ ) e ξ and w z = u ~ z - c ρ K z .

Then, using ΔX=4zz¯ we compute

z ¯ [ S ( w ) ] = 1 4 ( z Δ X w - w z Δ X w )
= - ρ 4 e ξ ( f ( u ~ ) ξ z + u ~ z f ( u ~ ) ) + ρ 4 e ξ f ( u ~ ) ( u ~ z - c ρ K z )
(2.7) = ρ 4 e ξ ( f ( u ~ ) - f ( u ~ ) ) u ~ z - ρ 4 e ξ f ( u ~ ) ( ξ z + c 0 K z ) + ( c 0 - c ρ ) ρ 4 e ξ f ( u ~ ) K z .

Using (2) we derive that

(2.8) z ¯ S 0 in L 1 ( B ) .

Indeed, this follows by Proposition 2.2, (1.13), and by the fact that |ρf(u~)|*aδ0(dx) for some a>0. On the other hand, by (2.6), since uu0 in Cloc(M𝒮), we have

w w 0 in C loc ( B ¯ { 0 } )

and then

(2.9) S S 0 in C loc ( B ¯ { 0 } ) .

At this point, we set Ξ=(ξ1,ξ2) and ζ=ξ1+iξ2 and we observe that by the Cauchy integral formula we may write

(2.10) [ S ( w ) ] ( ζ ) = 1 π B z ¯ S ( z ) ζ - z 𝑑 X + i 2 π + B [ S ( w ) ] ( z ) ζ - z 𝑑 z = g ( ζ ) + h ( ζ ) .

We have

(2.11) h ( ζ ) h 0 ( ζ ) = i 2 π + B S 0 ( z ) ζ - z 𝑑 z in C loc 0 ( B )

and h0 is holomorphic in B. On the other hand, we have

(2.12) g 0 in L 1 ( B ) .

To prove (2.12), it is sufficient to observe that for every zB=B(0,r), we have BB(z,2r) and then

g L 1 ( B ) B × B | z ¯ S ( z ) | 1 | ζ - z | 𝑑 X 𝑑 Ξ B | z ¯ S ( z ) | ( B ( z , 2 r ) 1 | ζ - z | 𝑑 Ξ ) 𝑑 X = 4 π r B | z ¯ S ( z ) | 𝑑 X ,

which tends to zero by (2.8). Combining (2.10), (2.11), and (2.12), we have

S h 0 in L 1 ( B ) as ρ 0 ,

and hence, up to subsequences,

S h 0 a.e. in B as ρ 0 ,

so that by (2.9),

S 0 ( ζ ) = h 0 ( ζ ) for all ζ B { 0 } .

This completes our proof. ∎

Finally, we use the following result from [2].

Proposition 2.4

Proposition 2.4 ([2])

For B=B(0,1)n, n2, the conditions vW1,p(B), 1<p<, and Δv=0 in B{0} imply that H=v-E is harmonic in B, where is some constant and

E ( x ) = { | x | 2 - n if n > 2 , log | x | if n = 2 .

Now, we are ready to prove Corollary 1.4. By GM we denote the Green’s function on the manifold M, defined by

{ - Δ g G M ( x , y ) = δ y - 1 | M | M G M ( x , y ) 𝑑 x = 0 .

Proof of Corollary 1.4.

Assume that p𝒮. Let us start by observing that w0 in (2.6) is harmonic in B{0} by definition and that w0W1,q(B) for all 1<q<2. Hence, also by using Proposition 2.4, we have

w 0 ( z ) = log 1 | z | + H ( z ) ,

where H is harmonic in B and 0. Then, using the fact that

z log | z | = 1 2 z log ( z z ¯ ) = ( 2 z ) - 1 ,

we compute

w 0 z = - 2 z + H z , w 0 z z = 2 z 2 + H z z .

Therefore,

S 0 = w 0 z z - 1 2 w 0 z 2 = 2 z 2 + H z z - 1 2 ( 2 z - H z ) 2 = ( 4 - ) 8 z 2 + 2 z H z + H z z - 1 2 H z 2 .

By Proposition 2.3, we know that S0 is holomorphic. Hence, we can conclude that =4 and Hz(0)=0. Since

(2.13) H = w 0 - 4 log 1 | z |

is harmonic in B, we have

Δ X ( u ~ 0 - 4 log 1 | z | ) = c 0 e ξ in B ( 0 , r )

and, therefore,

Δ g ( u 0 ( x ) - 8 π G M ( x , p ) ) = c 0 - 8 π | M | + h p in ( p , ε )

for some harmonic function hp. Arguing similarly for each p𝒮={p1,p2,,pm}, we conclude that

Δ g ( u 0 ( x ) - 8 π j = 1 m G M ( x , p j ) ) = c 0 - 8 π m | M | in M .

In particular, we obtain

u 0 ( x ) - 8 π j = 1 m G M ( x , p j ) = constant in M .

Observing that Mu0=0, this completes the proof of (1.14). To obtain (1.15) it is sufficient to observe that, in view of (2.13) and (1.13), we have

0 = 1 8 π z H ( X ) | X = 0
= z [ q 𝒮 G M ( Ψ - 1 ( X ) , q ) + 1 2 π log | X | ] | X = 0 - [ m | M | z K ( X ) ] | X = 0
= z [ q 𝒮 G M ( Ψ - 1 ( X ) , q ) + 1 2 π log | X | - 1 8 π ξ ( X ) ] | X = 0 .

Now, Corollary 1.4 is completely established. ∎

Funding statement: The first author acknowledges the support of FP7-MC-2009-IRSES-247486 “MaNEqui”.

A The L-estimate on M

In this appendix, for the sake of completeness and in order to outline the original arguments in [6], so that the simplification of our L1-approach may be seen, we check that Proposition 1.2 may be actually extended to (1.7) on a compact Riemannian 2-manifold (M,g) without boundary with minor modifications. We consider a solution sequence for (1.7). We assume that f satisfies (1.2), (1.4), and (1.5). Moreover, we assume (1.10), so that cρc0 as ρ0+. We show the following proposition.

Proposition A.1

Let u be a solution to (1.7). Then,

ρ u ( f ( u ) - f ( u ) ) L ( M ) 0 as ρ 0 .

The proof relies on the following relation, due to Obata.

Lemma A.2

Lemma A.2 ([8])

Let w=w(x)>0 be a solution to

(A.1) Δ w = | w | 2 w + F ( w ) on M ,

where F is a C1-function. Then, there holds the identity

(A.2) div V = J + 1 2 | w | 2 w - 2 ( F ( w ) + w F ( w ) ) ,

where, in local coordinates,

V j = w - 1 { ( w x i ) w - 1 2 w x i Δ w } , j = 1 , 2 ,

and

J = w - 1 { i , j = 1 2 ( 2 w x i x j ) 2 - 1 2 ( Δ w ) 2 } 0 .

Lemma A.3

Let u be a solution to (1.7). Then, for every r>0, there holds

(A.3) ρ M e - r u | u | 2 ( 2 r f ( u ) - f ( u ) ) 2 r c ρ M e - r u | u | 2 .

Proof.

Let u be a solution to (1.7). Denoting w=e-ru, it is easy to see that Obata’s assumption (A.1) is satisfied by the function w with

F ( w ) = r e - r u ( ρ f ( u ) - c ρ ) .

On the other hand, we have

F ( w ) + w F ( w ) = ρ e - r u ( 2 r f ( u ) - f ( u ) ) - 2 r e - r u c ρ .

In view of Obata’s identity (A.2), we conclude that

M | w | 2 w 2 ( F ( w ) + w F ( w ) ) 2 M div V = 0 .

In particular, since

w w = - r u ,

by the last inequality we obtain

M r 2 | u | 2 ( F ( w ) + w F ( w ) ) = r 2 ρ M e - r u | u | 2 ( 2 r f ( u ) - f ( u ) ) - 2 r 3 c ρ M e - r u | u | 2 0 .

We note that combining (A.3) and (2.1), for 12<r<1, we obtain

(A.4) ρ M e - r u | u | 2 f ( u ) C ( 1 + ρ M e - ( r - γ ) u | u | 2 ) .

Since γ<14, combining (2.1) and (A.4) we obtain

ρ M e - r u | u | 2 f ( u ) 𝑑 x C

and then, since u-C, using (2.1) again we have

(A.5) ρ M e - r u | u | 2 | f ( u ) | 𝑑 x C if 1 2 < r < 1 .

For r>0, we define

G r ( t ) = 0 t e - r 2 s | f ( s ) | 𝑑 s .

Then, (A.5) may be written in the form

(A.6) G r ( u ) L 2 ( M ) C ρ .

Lemma A.4

There holds

(A.7) G r ( u ) L 1 ( M ) C ρ .

Proof.

The proof can be easily obtained as in Lemma 2.1. Let us observe that in our assumption, for every 12<r<1, we have

(A.8) { x M : u ( x ) 0 } G r ( u ) 𝑑 x 2 r { u 0 } | f ( u ) | 𝑑 x C ( M | f ( u ) | 𝑑 x ) 1 2 C ρ .

On the other hand, since -uC, we have

(A.9) { x M : u ( x ) 0 } | G r ( u ) | 𝑑 x C { u 0 } 𝑑 x u 0 e C r 2 C e C r 2 | M | C .

Combining (A.8) and (A.9), we conclude the proof of (A.7). ∎

Reducing (A.6) to

G r ( u ) L p ( M ) C ρ for 1 < p < 2 ,

and using (A.7) and the Sobolev embedding, we obtain

G r ( u ) L p * ( M ) C ρ , 1 p * = 1 p - 1 2 .

Moreover, we have

| f ( t ) | 1 2 σ C ( | G r ( t ) | + 1 )

for σ=11-r (>2). We choose 12<r<1 such that

(A.10) ( γ + 1 2 ) σ < 3 2 .

Arguing as in [6], for every ε>0, we obtain

(A.11) f ( u ) L p ( M ) C ρ - σ + σ - 1 p - ε , 1 < p < ,

and, for q>2,

(A.12) u L q ( M ) C ρ ( - 1 2 + 1 q ) ( σ - 1 ) - ε .

Now, we conclude the proof of Proposition A.1.

Proof of Proposition A.1.

There holds

(A.13) ( f ( u ) - f ( u ) ) u L ( M ) C e γ u u L ( M ) = C γ e γ u L ( M ) .

Moreover, by (1.7) we have

- Δ g e γ u = - γ 2 e γ u | u | 2 + ρ γ e γ u f ( u ) - c ρ γ e γ u in M .

Hence, for p>2, we have

e γ u L ( M ) C ( Δ g e γ u L p ( M ) + e γ u L 1 ( M ) ) C ( e γ u | u | 2 L p ( M ) + ρ e γ u f ( u ) L p ( M ) + c ρ e γ u L p ( M ) ) .

Now, observing that euC(f(u)+1), by (A.11) we obtain

(A.14) ρ e γ u f ( u ) L p ( M ) C ρ e ( γ + 1 ) u L p ( M ) = C ρ e u L p ( γ + 1 ) ( M ) γ + 1 C ρ τ - ε

for every ε>0 with

(A.15) τ = 1 + ( γ + 1 ) ( σ - 1 p ( γ + 1 ) - σ ) = 1 + σ - 1 p - σ ( γ + 1 ) .

Hence, as p2, we have

(A.16) τ 1 + 1 2 ( σ - 1 ) - σ ( γ + 1 ) > - 1

by (A.10). On the other hand, by (2.3), for 1p<1γ, we have

c ρ e γ u L p ( M ) C ρ - γ .

Moreover, if q>12γ(>2), then

e γ u | u | 2 L p ( M ) e γ u L p q ( M ) u | L 2 p q ( M ) 2 ,

where qq=q+q. By (A.12), for every ε>0 and since 2pq>2, we have

u L 2 p q ( M ) 2 C ρ ( - 1 + 1 p q ) ( σ - 1 ) - ε .

Using again (A.11), for every ε>0, we have

e γ u L p q ( M ) C e u L p q γ ( M ) γ C ρ - γ σ + σ - 1 p q - ε .

Then, for every ε>0, we have

(A.17) e γ u | u | 2 L p ( M ) C ρ τ - ε

with τ defined by (A.15). Combining (A.13)–(A.14) and (A.16)–(A.17), we complete the proof. ∎

References

[1] Bethuel F., Brezis H. and Hélein F., Ginzburg–Landau Vortices, Progr. Nonlinear Differential Equations Appl. 13, Birkhäuser, Boston, 1994. 10.1007/978-1-4612-0287-5Search in Google Scholar

[2] Brézis H. and Lions P. L., A note on isolated singularities for linear elliptic equations, Mathematical Analysis and Applications, Part A, Adv. in Math. Suppl. Stud. 7A, Academic Press, New York (1981), 263–266. Search in Google Scholar

[3] Brezis H. and Merle F., Uniform estimates and blow-up behavior for solutions of -Δu=V(x)eu in two dimensions, Comm. Partial Differential Equations 16 (1991), no. 8–9, 1223–1253. 10.1080/03605309108820797Search in Google Scholar

[4] Esposito P., Grossi M. and Pistoia A., On the existence of blowing-up solutions for a mean field equation, Ann. Inst. H. Poincaré Anal. Non Linéaire 22 (2005), no. 2, 227–257. 10.1016/j.anihpc.2004.12.001Search in Google Scholar

[5] Lin C.-S., An expository survey on the recent development of mean field equations, Discrete Contin. Dyn. Syst. 19 (2007), no. 2, 387–410. 10.3934/dcds.2007.19.387Search in Google Scholar

[6] Nagasaki K. and Suzuki T., Asymptotic analysis for two-dimensional elliptic eigenvalue problems with exponentially-dominated nonlinearities, Asymptot. Anal. 3 (1990), no. 2, 173–188. 10.3233/ASY-1990-3205Search in Google Scholar

[7] Neri C., Statistical mechanics of the N-point vortex system with random intensities on a bounded domain, Ann. Inst. H. Poincaré Anal. Non Linéaire 21 (2004), no. 3, 381–399. 10.1016/j.anihpc.2003.05.002Search in Google Scholar

[8] Obata M., The conjectures on conformal transformations of Riemannian manifolds, J. Differential Geom. 6 (1971), no. 2, 247–258. 10.4310/jdg/1214430407Search in Google Scholar

[9] Pistoia A. and Ricciardi T., Concentrating solutions for a Liouville type equation with variable intensities in 2D-turbulence, preprint 2015, http://arxiv.org/abs/1505.05304. 10.1088/0951-7715/29/2/271Search in Google Scholar

[10] Ricciardi T. and Suzuki T., Duality and best constant for a Trudinger–Moser inequality involving probability measures, J. Eur. Math. Soc. (JEMS) 16 (2014), no. 7, 1327–1348. 10.4171/JEMS/462Search in Google Scholar

[11] Ricciardi T. and Zecca G., Blow-up analysis for some mean field equations involving probability measures from statistical hydrodynamics, Differential Integral Equations 25 (2012), no. 3–4, 201–222. 10.57262/die/1356012734Search in Google Scholar

[12] Ricciardi T. and Zecca G., Mean field equations with probability measure in 2D-turbulence, Ric. Mat. 63 (2014), no. 1, suppl., S255–S264. 10.1007/s11587-014-0208-6Search in Google Scholar

[13] Ricciardi T. and Zecca G., Mass quantization and minimax solutions for Neri’s mean field equation in 2D-turbulence, J. Differential Equations 260 (2016), no. 1, 339–369. 10.1016/j.jde.2015.08.045Search in Google Scholar

[14] Ye D., Une remarque sur le comportement asymptotique des solutions de -Δu=λf(u), C. R. Acad. Sci. Paris Sér I Math. 325 (1997), no. 12, 1279–1282. 10.1016/S0764-4442(97)82353-1Search in Google Scholar

Received: 2015-01-28
Accepted: 2015-05-19
Published Online: 2015-12-08
Published in Print: 2016-02-01

© 2016 by De Gruyter

Downloaded on 23.9.2025 from https://www.degruyterbrill.com/document/doi/10.1515/ans-2015-5015/html
Scroll to top button