• Non ci sono risultati.

Rigidity of critical points for a nonlocal Ohta-Kawasaki energy

N/A
N/A
Protected

Academic year: 2021

Condividi "Rigidity of critical points for a nonlocal Ohta-Kawasaki energy"

Copied!
12
0
0

Testo completo

(1)

Rigidity of critical points

for a nonlocal Ohta-Kawasaki energy

Serena Dipierro† Matteo Novaga‡ Enrico Valdinoci§

Abstract

We investigate the shape of critical points for a free energy consisting of a nonlocal perimeter plus a nonlocal repulsive term. In particular, we prove that a volume-constrained critical point is necessarily a ball if its volume is sufficiently small with respect to its isodiametric ratio, thus extending a result previously known only for global minimizers.

We also show that, at least in one-dimension, there exist critical points with arbitrarily small volume and large isodiametric ratio. This example shows that a constraint on the diameter is, in general, necessary to establish the radial symmetry of the critical points.

Contents

1 Introduction 1 2 Potential estimates 3 3 Proof of Theorem 1 7 4 Proof of Theorem 2 9

1

Introduction

In recent years the following variational problem has been widely studied: min |E|=mP (E) + Z Z E×E dx dy |x − y|, (1)

where P (E) denotes the perimeter of a set E ⊂ R3. Such problem first appeared in the liquid drop model of the atomic nuclei proposed by Gamow in 1928 [13] and then developed by other researchers [3, 9], and it is also relevant in some models of diblock copolymer melts [5, 17].

In [15, 16] (see also [6]) the authors showed that global minimizers of (1) exist if the volume m is small, and do not exist if the volume is large enough. They also showed that minimizers are necessarily balls if

2010 Mathematics Subject Classification: 49Q10, 49Q20, 35B38, 58J70.

School of Mathematics and Statistics, University of Melbourne, 813 Swanston St, Parkville VIC 3010, Australia, and Weierstraß-Institut f¨ur Angewandte Analysis und Stochastik, Hausvogteiplatz 5/7, 10117 Berlin, Germany, serydipierro@yahoo.it

Department of Mathematics, University of Pisa, novaga@dm.unipi.it

§

School of Mathematics and Statistics, University of Melbourne, 813 Swanston St, Parkville VIC 3010, Australia, Weierstraß-Institut f¨ur Angewandte Analysis und Stochastik, Hausvogteiplatz 5/7, 10117 Berlin, Germany, and Dipartimento di Matematica, Universit`a degli studi di Milano, Via Saldini 50, 20133 Milan, Italy, enrico@math.utexas.edu

(2)

the volume is small enough. This result has been later extended in [14] to connected critical points in two-dimensions. However it is still not known if there exist non-spherical critical points of (1) with arbitrarily small volume.

In this paper we consider the modified energy

F (E) := Ps(E) +

Z Z

E×E

dx dy

|x − y|α, (2)

with s ∈ (0, 1) and α ∈ (0, n), which is a nonlocal extension of the energy in (1), which takes into account long-range interactions in the attractive term of surface tension type.

Here Ps(E) := Z Z E×Ec dx dy |x − y|n+s

denotes the so-called fractional perimeter of E, recently introduced by Caffarelli, Roquejoffre and Savin in [4], where they initiated the analysis of the corresponding Plateau Problem.

The functional (2) has been studied in [11], where the authors showed that volume-constrained minimizers are balls, if the volume is small enough, thus extending the results in [15, 16].

We observe that, for any λ > 0, it holds F (λE) = λn−s  Ps(E) + λn−α+s Z Z E×E dx dy |x − y|α  .

Therefore, minimizing F under the volume constraint |E| = m is equivalent to minimize the functional Fε(E) := Ps(E) + ε

Z Z

E×E

dx dy |x − y|α,

under the constraint |E| = 1, with ε := m1−αn+ s n.

If E is a critical point of Fεwith boundary of class C2, then

κE(x) + cεVE(x) = λε for any x ∈ ∂E, (3)

for a suitable λε∈ R. Here above, c > 0 is a normalizing constant,

VE(x) :=

Z

E

dy

|x − y|α (4)

is the potential, and κE is the fractional mean curvature of E, i.e.,

κE(x) :=

Z

Rn

χRn\E(y) − χE(y)

|x − y|n+s dy, (5)

where the integral is meant in the principal value sense, see e.g. [4, 1, 11] for further details on this notion of mean curvature.

Notice that, if E is a critical point of Fε with boundary of class C2, by elliptic regularity (see [2]) the

boundary of E is in fact of class C∞.

The main result of this paper is that critical points of F are necessarily spherically symmetric, if their volume is small enough with respect to their isodiametric ratio, which is defined as

I(E) := m

1 n

diam(E). (6)

To state this result precisely, we introduce a scale parameter that links volume and diameter of a set, given by

β = β(n, α, s) := n + s − α

(2n + s + 1)n. (7)

(3)

Theorem 1. Let n> 2, s ∈ (0, 1) and α ∈ (0, n − 1).

There existsco = co(n, α, s) > 0 such that, if E?is a bounded, volume-constrained, critical point ofF with

smooth boundary, such that

|E?|β 6 coI(E?), (8)

thenE?is a ball.

Equivalently, ifE is a bounded, volume-constrained, critical point of Fεwith smooth boundary, such that

|E| = 1 and diam(E) 6 coε−

1

2n+s+1, (9)

thenE is a ball.

We observe that condition (8) does not violate the isodiametric inequality, since the right hand side is scale invariant, and so it is not void for small sets. Moreover, we remark that in problems like the ones considered in this paper, in which an aggregating energy is competing with a repulsive one, universal bounds on the diameter of the solutions do not hold in general. Indeed, roughly speaking, the repulsive term of the energy functional may produce critical points with disconnected components which lie “far away” one from the other and which therefore violate uniform diameter bounds. To give a concrete example of this phenomenon we provide the following result:

Theorem 2. Let n = 1, s ∈ (0, 1) and α ∈ (0, 1).

There existsm = ¯¯ m(α, s) > 0 such that, if m ∈ (0, ¯m), there exists a bounded critical point E?ofF with

volumem, made by two disconnected components, whose diameter satisfies diam(E?) > Co,

for someCo> 0.

Equivalently, there existsε = ¯¯ ε(α, s) > 0 such that, if ε ∈ (0, ¯ε), there exists a bounded set E ⊂ R which is a solution of (3), which is made by two disconnected components, whose diameter satisfies

diam(E) > Coε−

1 1+s−α.

In particular, the diameter ofE is not bounded uniformly in ε.

One of the main tools in the proof of Theorem 1 is a nonlocal version of Alexandrov’s Theorem, recently proved in [8], which states that regular sets with almost constant fractional mean curvature are necessarily balls. This result can be applied to solutions to (3), after deriving suitable estimates on the potential VE.

We point out that such quantitative Alexandrov’s Theorem does not hold in the local setting, where it is known that a set with almost constant mean curvature is close to a family of tangent balls (see [7]), and this is the main obstruction for extending our result to the case s = 1, that is, to the functional in (1).

The plan of the paper is the following: in Section 2 we provide some preliminary estimates on the potential VE and on its gradient, which will be useful in the proof of Theorem 1; in Section 3 we prove Theorem 1;

finally in Section 4 we prove Theorem 2, giving an explicit example of a one-dimensional set, composed by two disjoint segments, which is a critical point of (2) and has arbitrarily small volume.

2

Potential estimates

In this section we provide some bounds on the potential V and on its derivatives that will be used in the proof of the main results.

(4)

Lemma 3. Let B be a ball centered at the origin with |B| = |E|. Then, for any x ∈ Rn, VE(x) 6 VB(0) 6 C,

for someC > 0.

Proof. Let ρ > 0 be the radius of B. In this way, we have that x + B = Bρ(x). Also, if y ∈ Bρ(x) \ E,

then |x − y|6 ρ and so Z Bρ(x)\E dy |x − y|α > Z Bρ(x)\E dy ρα = |Bρ(x) \ E| ρα .

Similarly, if y ∈ E \ Bρ(x), then |x − y| > ρ and so

Z E\Bρ(x) dy |x − y|α 6 |E \ Bρ(x)| ρα . Moreover |E \ Bρ(x)| = |E| − |E ∩ Bρ(x)| = |Bρ(x)| − |Bρ(x) ∩ E| = |Bρ(x) \ E|. Consequently, Z E\Bρ(x) dy |x − y|α 6 Z Bρ(x)\E dy |x − y|α.

Therefore, summing the contributions in E ∩ Bρ(x) to both sides of this inequality, we obtain that

Z E dy |x − y|α 6 Z Bρ(x) dy |x − y|α.

This and the integrability of the kernel imply the desired result. Following are additional potential estimates:

Lemma 4. We have that Z E ∇VE(x) · x dx = −α 2 Z E VE(x) dx (10) and Z ∂E VE(x) x · ν(x) dHn−1(x) =  n − α 2 Z E VE(x) dx. (11) Moreover, |∇VE(x)| 6 C, (12)

for anyx ∈ Rn, for a suitableC > 0. Proof. We observe that

∇VE(x) = −α Z E (x − y) dy |x − y|α+2, (13) so we can write β := Z E ∇VE(x) · x dx = −α Z Z E×E (x − y) · x |x − y|α+2dx dy.

Since the role played by the variable in the latter integral is symmetric, we can also write β = −α

Z Z

E×E

(y − x) · y |x − y|α+2 dx dy.

(5)

As a consequence 2β = −α Z Z E×E (x − y) · x |x − y|α+2 dx dy − α Z Z E×E (y − x) · y |x − y|α+2 dx dy = −α Z Z E×E (x − y) · (x − y) |x − y|α+2 dx dy = −α Z Z E×E dx dy |x − y|α. This establishes (10).

Now we use the Divergence Theorem and (10) to see that Z ∂E VE(x) x · ν(x) dHn−1(x) = Z E div VE(x) x dx = n Z E VE(x) dx + Z E ∇VE(x) · x dx =  n − α 2 Z E VE(x) dx,

and so we have proved (11). Finally, by (13), |∇VE(x)| 6 α Z E dy |x − y|α+1,

hence (12) follows as in Lemma 3 (replacing α there with α + 1 ∈ (0, n)).

With this, we can show that smooth critical points possess a uniform bound on λε in the Euler-Lagrange

equation in (3):

Corollary 5. There exist C1,C2 > 0 such that

λε− C1Ps(E)

6 C2ε.

Proof. Given a vector field X ∈ C0∞(Rn), we denote, for any small t > 0, by Φtthe flow induced by X, as

defined by the Cauchy problem

(

∂tΦt(x) = X(Φt(x)),

Φ0(x) = x.

Then (up to normalizing constants that we neglect, see e.g. Formula (6.12) in [11]), we have that ∂ ∂tPs(Φ t(E)) t=0 = Z ∂E κE(x) X(x) · ν(x) dHn−1(x), (14)

being ν the external normal of E.

Now, by taking X(x) := x in the vicinity of E, we have that

Φt(x) = x + tX(x) + O(t2) = (1 + t)x + O(t2) and so |Φt(E)| = (1 + t + O(t2)) |E|. Therefore, by scaling, we have that

Ps(Φt(E)) = (1 + t + O(t2))n−sPs(E)

and thus ∂ ∂tPs(Φ t(E)) t=0= (n − s) Ps(E).

(6)

Comparing this and (14), we obtain Ps(E) =

Z

∂E

κE(x) x · ν(x) dHn−1(x), (15)

up to multiplicative normalization constants that we neglect. Also, by the Divergence Theorem,

Z ∂E x · ν(x) dHn−1(x) = Z E div x dx = n |E|. Hence, by (3), (15) and Lemma 4, up to multiplicative constants,

λε|E| = Z ∂E λεx · ν(x) dHn−1(x) = Z ∂E κE(x) x · ν(x) dHn−1(x) + ε Z ∂E VE(x) x · ν(x) dHn−1(x) = Ps(E) + ε Z E VE(x) dx. (16) Also, by Lemma 3, Z E VE(x) dx 6 C |E|.

From this, (16) and the volume constraint |E| = 1, the desired result plainly follows. Now we control the tangential gradient of VE:

Lemma 6. Let x ∈ ∂E and τE(x) be a tangential vector to ∂E at x. Assume that E = T (B), for a suitable

ballB, with |B| = |E| and let T be a diffeomorphism such that kT − IdkC1(B) 6 µ,

for someµ > 0. Then, there exists µ0 > 0 such that for any µ ∈ [0, µ0]

∇VE(x) · τE(x)

6 Cµ, (17)

for someC > 0.

Proof. We define ¯x := T−1(x) and

τB(¯x) := DT−1(x) τE(x) = DT (¯x)

−1 τE(x).

Notice that τB(¯x) is a tangent vector to ∂B at ¯x ∈ ∂B. Also, by rotational symmetry, the function VB is

constant along ∂B, and therefore

∇VB(¯x) · τB(¯x) = 0. (18)

Moreover, by (13) and using the substitution ¯y := T−1(y), ∇VE(x) · τE(x) = −α Z E (x − y) · τE(x) |x − y|α+2 dy = −α Z B (T (¯x) − T (¯y)) · DT (¯x)τB(¯x) 

|T (¯x) − T (¯y)|α+2 | det DT (¯y)| d¯y.

(7)

Now, by (17), DT (¯x)τB(¯x) − τB(¯x) 6Cµ, det DT (¯y) − 1 6Cµ, and T (¯x) − T (¯y) − (¯x − ¯y) 6 Z 1 0 DT (t¯y + (1 − t)¯x) dt |x − y| 6 Cµ |¯x − ¯y| (20)

for some C > 0, and so

(1 − Cµ) |¯x − ¯y| 6 T (¯x) − T (¯y) 6 (1 + Cµ) |¯x − ¯y|, which in turn implies that

1 |T (¯x) − T (¯y)|α+2 − 1 |¯x − ¯y|α+2 = 1 |¯x − ¯y|α+2 1 − |¯x − ¯y| α+2 |T (¯x) − T (¯y)|α+2 6 Cµ |¯x − ¯y|α+2,

up to renaming C > 0. Using this and (20) inside (19), we obtain that 1 α ∇VE(x) · τE(x) − ∇VB(¯x) · τB(¯x) = Z B (T (¯x) − T (¯y)) · DT (¯x)τB(¯x) 

|T (¯x) − T (¯y)|α+2 | det DT (¯y)| d¯y −

Z B (¯x − ¯y) · τB(¯x) |¯x − ¯y|α+2 d¯y 6 Z B (T (¯x) − T (¯y)) · DT (¯x)τB(¯x) 

|T (¯x) − T (¯y)|α+2 | det DT (¯y)| d¯y

− Z

B

(¯x − ¯y) · DT (¯x)τB(¯x)



|T (¯x) − T (¯y)|α+2 | det DT (¯y)| d¯y

+ Z B (¯x − ¯y) · DT (¯x)τB(¯x) 

|T (¯x) − T (¯y)|α+2 | det DT (¯y)| d¯y

− Z

B

(¯x − ¯y) · τB(¯x)

|T (¯x) − T (¯y)|α+2| det DT (¯y)| d¯y

+ Z B (¯x − ¯y) · τB(¯x)

|T (¯x) − T (¯y)|α+2 | det DT (¯y)| d¯y −

Z B (¯x − ¯y) · τB(¯x) |¯x − ¯y|α+2 | det DT (¯y)| d¯y + Z B (¯x − ¯y) · τB(¯x) |¯x − ¯y|α+2 | det DT (¯y)| d¯y − Z B (¯x − ¯y) · τB(¯x) |¯x − ¯y|α+2 d¯y 6 Cµ Z B d¯y |¯x − ¯y|α+1 6 Cµ,

possibly changing C > 0 from one line to another. From this estimate and (18), the desired result plainly follows.

3

Proof of Theorem 1

By a spatial dilation, we replace E? with a critical point E of the functional Fε, under the volume

con-straint |E| = 1. To this aim, we define

E := m−n1 E? and ε := m1− α n+

s n.

(8)

In particular diam(E) = m−n1 diam(E?) and thus, by (6) and (7), condition (8) becomes ε diam(E)2n+s+1 =ε2n+s+11 · m−n1 diam(E?) 2n+s+1 =m n−α+s (2n+s+1)n · m−n1 diam(E?) 2n+s+1 =  mβ I(E?) 2n+s+1 6 c2n+s+1o , (21)

which is supposed to be a small quantity.

Also, the Euler-Lagrange equation in (3) holds true. Therefore we have that δs(E) := sup x,y∈∂E x6=y |κE(x) − κE(y)| |x − y| = sup x,y∈∂E x6=y c ε VE(x) − λε − c ε VE(y) − λε  |x − y| = cε sup x,y∈∂E x6=y |VE(x) − VE(y)| |x − y| . (22) Thus, by (12), δs(E) 6 cε sup x,y∈Rn x6=y |VE(x) − VE(y)| |x − y| 6 Cε, (23)

for some C > 0. Now, letting

ηs(E) := diam(E)

2n+s+1

δs(E), (24)

by Formula (1.4) in [8] we have that ρ(E) := inf  R − r diam(E), with p ∈ E, Br(p) ⊆ E ⊆ BR(p)  6 C ηs(E), (25)

up to renaming C > 0. Also, by (23) and (24) we get ηs(E) 6 C diam(E)

2n+s+1

ε. (26)

This and (21) give that ηs(E) is small if so is co, hence we are in the position to apply Theorem 1.5 in [8].

In particular, we obtain that ∂E = F (∂B), where B is a ball with volume m, with kF − IdkC2(∂B) 6 C ηs(E),

up to renaming C > 0. That is, if we set T (x) := |x| F 

x |x|



, we have that E = T (B) and kT − IdkC1(B) 6 C ηs(E).

This says that we are in the setting of Lemma 6, with µ := C ηs(E). As a consequence, we obtain that

τVE(x)

6 Cηs(E), for some C > 0. Hence, we insert this into (22) and we find

δs(E) 6 C ε ηs(E),

up to renaming C > 0. Accordingly, in view of (24) and (21), δs(E) 6 C ε diam(E)

2n+s+1

δs(E) 6 C c2n+s+1o δs(E) 6

δs(E)

2 , if cois small enough. This implies that δs(E) = 0.

(9)

4

Proof of Theorem 2

We let E :=  0, 1 2  ∪  d, d + 1 2  , for some d > 1/2, that we choose appropriately in order to fulfill (3). By (5) and (4), for any x ∈ {0, 1/2, d, d + 1/2},

ζE(x) := κE(x) + cε VE(x) = Z 0 −∞ dy |x − y|1+s − Z 1/2 0 dy |x − y|1+s + Z d 1/2 dy |x − y|1+s − Z d+1/2 d dy |x − y|1+s + Z +∞ d+1/2 dy |x − y|1+s +cε Z 1/2 0 dy |x − y|α + cε Z d+1/2 d dy |x − y|α.

By symmetry (i.e. using the substitution ˜y := (d + 1/2) − y), we see that ζE(0) = ζE  d +1 2  and ζE  1 2  = ζE(d). (27)

Now we prove that we can choose d > 1/2 in such a way that ζE(0) = ζE

 1 2



(10)

To this aim, we compute f (d) := ζE  1 2  − ζE(0) = Z 0 −∞ dy |y − 1/2|1+s − Z 1/2 0 dy |y − 1/2|1+s + Z d 1/2 dy |y − 1/2|1+s − Z d+1/2 d dy |y − 1/2|1+s + Z +∞ d+1/2 dy |y − 1/2|1+s − Z 0 −∞ dy |y|1+s + Z 1/2 0 dy |y|1+s − Z d 1/2 dy |y|1+s + Z d+1/2 d dy |y|1+s − Z +∞ d+1/2 dy |y|1+s +cε Z 1/2 0 dy |y − 1/2|α + cε Z d+1/2 d dy |y − 1/2|α −cε Z 1/2 0 dy |y|α − cε Z d+1/2 d dy |y|α = Z −1/2 −∞ dy |y|1+s − Z 0 −1/2 dy |y|1+s + Z d−1/2 0 dy |y|1+s − Z d d−1/2 dy |y|1+s + Z +∞ d dy |y|1+s − Z 0 −∞ dy |y|1+s + Z 1/2 0 dy |y|1+s − Z d 1/2 dy |y|1+s + Z d+1/2 d dy |y|1+s − Z +∞ d+1/2 dy |y|1+s +cε Z 0 −1/2 dy |y|α + cε Z d d−1/2 dy |y|α − cε Z 1/2 0 dy |y|α − cε Z d+1/2 d dy |y|α = −2 Z d d−1/2 dy |y|1+s + 2 Z d+1/2 d dy |y|1+s + cε Z d d−1/2 dy |y|α − cε Z d+1/2 d dy |y|α = 2 s2d −s− (d − 1/2)−s− (d + 1/2)−s + cε 1 − α2d 1−α− (d − 1/2)1−α− (d + 1/2)1−α .

Now we use that, for any β ∈ R and δ > 0 small,

2 − (1 − δ)β− (1 + δ)β = −β (β − 1) δ2+ O(δ3),

therefore, for large d,

f (d) = 2d −s s " 2 −  1 − 1 2d −s −  1 + 1 2d −s# +cε d 1−α 1 − α " 2 −  1 − 1 2d 1−α −  1 + 1 2d 1−α# = 2d −s s ·  −s(1 + s) 4d2 + O(d −3)  + cε d 1−α 1 − α ·  α(1 − α) 4d2 + O(d −3)  = d−1−α cαε 4 − (1 + s)d−1−s+α 2 + O(d −2−s+α ) + O(εd−3)  = d−1−αg(d), (29)

(11)

where g(d) := cαε 4 − (1 + s)d−1−s+α 2 + O(d −2−s+α) + O(εd−3). Notice that g(+∞) = cαε 4 > 0. (30) Also, if dε:=  1 + s cαε 1+s−α1 , we have that g(dε) = − cαε 4 + O(ε 1+1+s−α1 ) + O(ε1+1+s−α3 ) 6 −cαε 8 < 0, if ε is sufficiently small.

From this and (30) we obtain that there exists an appropriate d > dε for which g(d) = 0, and thus,

recall-ing (29), we obtain that f (d) = 0, which proves (28). Now, from (27) and (28), we deduce that

ζE(0) = ζE  d +1 2  = ζE  1 2  = ζE(d) =: λ,

for some λ ∈ R, that is

κE(x) + cε VE(x) = λ

for any x ∈ ∂E, as desired.

Acknowledgements

Part of this work was carried out while Serena Dipierro and Enrico Valdinoci were visiting the University of Pisa, which they wish to thank for the hospitality.

This work has been supported by the Alexander von Humboldt Foundation, the ERC grant 277749 E.P.S.I.L.O.N. “Elliptic Pde’s and Symmetry of Interfaces and Layers for Odd Nonlinearities”, the PRIN grant 201274FYK7 “Aspetti variazionali e perturbativi nei problemi differenziali nonlineari” and the University of Pisa grant PRA- 2015-0017.

Finally, the authors wish to thank Giulio Ciraolo and Cyrill Muratov for useful conversations on the subject of this paper.

References

[1] N. ABATANGELO, E. VALDINOCI: A notion of nonlocal curvature, Numer. Funct. Anal. Optim. 35 (2014), no. 7-9, 793–815.

[2] B. BARRIOS, A. FIGALLI, E. VALDINOCI: Bootstrap regularity for integro-differential operators and its application to nonlocal minimal surfaces, Ann. Sc. Norm. Super. Pisa Cl. Sci. (5) 13 (2014), no. 3, 609–639.

[3] N. BOHR: Neutron capture and nuclear constitution, Nature 137 (1936), 344–348.

[4] L. CAFFARELLI, J.-M. ROQUEJOFFRE, O. SAVIN: Nonlocal minimal surfaces, Commun. Pure Appl.

Math. 63 (2010), no. 9, 1111–1144.

[5] R. CHOKSI, X. REN: On the derivation of a density functional theory for microphase separation of diblock copolymers, J. Statist. Phys. 113 (2003), 151–176.

(12)

[6] M. CICALESE, E. SPADARO: Droplet minimizers of an isoperimetric problem with long-range inter-actions, Comm. Pure Appl. Math. 66 (2013), 1298–1333.

[7] G. CIRAOLO, F. MAGGI: On the shape of compact hypersurfaces with almost constant mean curvature, Preprint, arXiv:1503.06674.

[8] G. CIRAOLO, A. FIGALLI, F. MAGGI, M. NOVAGA: Rigidity and sharp stability estimates for

hy-persurfaces with constant and almost-constant nonlocal mean curvature, to appear on J. Reine Angew. Math. arXiv:1503.00653.

[9] S. COHEN, F. PLASIL, W. J. SWIATECKI: Equilibrium configurations of rotating charged or gravitat-ing liquid masses with surface tension. II, Ann. Physics 82 (1974), 557–596.

[10] A. DI CASTRO, M. NOVAGA, B. RUFFINI, E. VALDINOCI: Nonlocal quantitative isoperimetric in-equalities, Calc. Var. Partial Differential Equations 54 (2015), no. 3, 2421–2464.

[11] A. FIGALLI, N. FUSCO, F. MAGGI, V. MILLOT, M. MORINI: Isoperimetry and stability properties of balls with respect to nonlocal energies, Commun. Math. Phys. 336 (2015), no. 1, 441–507.

[12] N. FUSCO, V. MILLOT, M. MORINI: A quantitative isoperimetric inequality for fractional perimeters, J. Funct. Anal. 261 (2011), no. 3, 697–715.

[13] G. GAMOW: Mass defect curve and nuclear constitution, Proc. R. Soc. Lond. A 126 (1930), 632-644. [14] D. GOLDMAN: Uniqueness results for critical points of a non-local isoperimetric problem via curve

shortening, Preprint, arXiv:1206.5984.

[15] H. KNUPFER¨ , C. MURATOV: On an isoperimetric problem with a competing non- local term. I. The planar case, Comm. Pure Appl. Math. 66 (2013), 1129–1162.

[16] H. KNUPFER¨ , C. MURATOV: On an isoperimetric problem with a competing non- local term. II. The general case, Commun. Pure Appl. Math. 67 (2014), 1974–1994.

[17] T. OHTA, K. KAWASAKI: Equilibrium morphologies of block copolymer melts, Macromolecules 19 (1986), 2621–2632.

Riferimenti

Documenti correlati

Solution proposed by Roberto Tauraso, Dipartimento di Matematica, Universit`a di Roma “Tor Vergata”, via della Ricerca Scientifica, 00133 Roma,

It remains to show that the partition exists when n is multiple of 8 and n ≥ 16.. For any non-negative integer x, we have a partition of {1

It is then clearly preferable to consider the majority of the g1ven values rather than to choose an arbitrary set, composed of the smallest number of discrete values necessary

One compound can have a higher affinity for the stationary phase with a particular mobile phase and will present a higher retention time than another metabolite with a lower

I n their retrospective follow-up study of all patients attending a hospital anticoagulant clinic for atrial fi- brillation, Copland and colleagues 1 found that the elderly

Urakawa: Rigidity of transversally biharmonic maps between foliated Riemannian manifolds, to appear in Hokkaido Math.. [36] Ye-Lin Ou and Liang Tang: The generalized Chen’s

For instance, it was proved that a finite nilpotent group has modular subgroup lattices if and only if every two of its subgroups are permutable.. The full description of finite

Govil [4] extended the inequality (1.3) to an inequality involving moduli of all the zeros of a complex polynomial rather than restricting only to the zero of maximum