introduction z ⇢z - applied mathematics and statisticsprashant/papers/2016_joca_weighted_tv... ·...

30
WEIGHTED TV MINIMIZATION AND APPLICATIONS TO VORTEX DENSITY MODELS P. ATHAVALE, R. L. JERRARD, M. NOVAGA, AND G. ORLANDI Abstract. Motivated by models arising in the description of Bose-Einstein condensa- tion, we consider total variation minimization problems in which the total variation is weighted by a function that may degenerate near the domain boundary, and the fidelity term contains a weight that may be both degenerate and singular. We develop a general theory for a class of such problems, with special attention to the examples arising from physical models. 1. Introduction We consider the weighted Total Variation functional J (u)= Z f (x)|Du| := sup ⇢Z u div(f (x)) dx : f 2 C 1 c (; R n ), || 1 , (1.1) defined for maps u 2 L 1 loc (), R n open bounded. We always assume that the weight f is positive and continuous in and possibly vanishing on some portion of @ . We are interested in the analysis of functionals of the type I λ (u)= J (u)+ λ 2 Z |u - u 0 | 2 g(x) dx (1.2) with λ > 0, g(x) > 0 a measurable weight function, and u 0 2 L 2 g := L 2 (,g(x)dx). Notice that, when f (x)= g(x) 1, I λ corresponds to the well-known Rudin-Osher- Fatemi Total Variation based denoising model [27]. Anisotropic versions of TV functionals and applications to active contour and edge detection have been studied in [17]. Related homogeneous functionals for the description of landslides have been studied [22, 14], and vector-valued generalizations can also be considered, such as J (u)= kT · Duk where T is a tensor (see for example [13] for the case T · P = trace (P ), corresponding to J (u)= kdiv uk, which is related to the Hele-Shaw model). Due to the variety of applications we are interested in a systematic study of properties of minimizers, in particular criteria for the minimizer being nontrivial, the characterization of the coincidence set, and conditions yielding the presence of flat zones and/or free boundaries. We are particularly interested in functionals of the type of I λ that arise as reduced models in superconductivity and superfluidity (namely Bose-Einstein condensation) in axisymmetric 3-D domains (see e.g. [10, 11]), suited to describe the distribution of vor- ticity that takes place in certain asymptotic regimes. Indeed, it is proved in [10, 11] that in certain scaling limits, a Bose-Einsten condensate may be described by a velocity field 1

Upload: lamtuong

Post on 05-Apr-2018

213 views

Category:

Documents


0 download

TRANSCRIPT

WEIGHTED TV MINIMIZATION AND APPLICATIONS TO VORTEXDENSITY MODELS

P. ATHAVALE, R. L. JERRARD, M. NOVAGA, AND G. ORLANDI

Abstract. Motivated by models arising in the description of Bose-Einstein condensa-tion, we consider total variation minimization problems in which the total variation isweighted by a function that may degenerate near the domain boundary, and the fidelityterm contains a weight that may be both degenerate and singular. We develop a generaltheory for a class of such problems, with special attention to the examples arising fromphysical models.

1. Introduction

We consider the weighted Total Variation functional

J(u) =

Z

⌦f(x)|Du| := sup

Z

⌦u div(f(x)⇠) dx : f⇠ 2 C1

c (⌦;Rn), |⇠| 1

, (1.1)

defined for maps u 2 L1loc(⌦), ⌦ ⇢ Rn open bounded. We always assume that the weight

f is positive and continuous in ⌦ and possibly vanishing on some portion of @⌦.We are interested in the analysis of functionals of the type

I�(u) = J(u) +�

2

Z

⌦|u� u0|2g(x) dx (1.2)

with � > 0, g(x) > 0 a measurable weight function, and u0 2 L2g := L2(⌦, g(x)dx).

Notice that, when f(x) = g(x) ⌘ 1, I� corresponds to the well-known Rudin-Osher-Fatemi Total Variation based denoising model [27]. Anisotropic versions of TV functionalsand applications to active contour and edge detection have been studied in [17]. Relatedhomogeneous functionals for the description of landslides have been studied [22, 14], andvector-valued generalizations can also be considered, such as J(u) = kT · Duk whereT is a tensor (see for example [13] for the case T · P = trace (P ), corresponding toJ(u) = kdiv uk, which is related to the Hele-Shaw model).

Due to the variety of applications we are interested in a systematic study of properties ofminimizers, in particular criteria for the minimizer being nontrivial, the characterizationof the coincidence set, and conditions yielding the presence of flat zones and/or freeboundaries.

We are particularly interested in functionals of the type of I� that arise as reducedmodels in superconductivity and superfluidity (namely Bose-Einstein condensation) inaxisymmetric 3-D domains (see e.g. [10, 11]), suited to describe the distribution of vor-ticity that takes place in certain asymptotic regimes. Indeed, it is proved in [10, 11] thatin certain scaling limits, a Bose-Einsten condensate may be described by a velocity field

1

2 P. ATHAVALE, R. L. JERRARD, M. NOVAGA, AND G. ORLANDI

v : ⌦ ! R3, where ⌦ is a bounded, open subset of R3, that minimizes a functional of theform

E(v) :=Z

D⇢TF

|r⇥ v|+ 1

2|v � �|2

.

Here

• ⇢TF is a nonnegative integrable function known as the “Thomas-Fermi density”,typically smooth in the interior of its support, corresponding to the limiting con-densate density in the asymptotic regime under consideration.

• D := {x 2 R3 : ⇢TF (x) > 0} is the limiting region occupied by the condensate• � is a given vector field that models forcing terms.• r⇥v is interpreted as the vorticity. In general it is only a measure, and

R

D |r⇥v|must be understood as a total variation.

The model case occurs when ⇢TF (x) = (constant�quadratic)+ and �(x) = ↵(�x2, x1, 0)for some ↵ 2 R, corresponding to a quadratic trapping potential and forcing via rotationaround the x3 axis. In [10] it is noted that for this �, if ⇢ is rotationally symmetric aboutthe x3-axis (that is, ⇢TF has the form ⇢TF (x) = ⇢TF (r, x3) for r =

p

x21 + x22), then theminimizer has the form v⇤ = w⇤(r, x3)(

�x2r2

, x1r2, 0), for w⇤ minimizing the reduced energy

Ered↵ (w) :=

Z

⌦⇢TF

|Dw|+ (w � ↵r2)2

r

dr dz, (1.3)

where⌦ := {(r, z) 2 (0,1)⇥ R : ⇢TF (r, z) > 0}.

Here, vortex lines are level sets of w⇤, and “flat spots” – sets of positive measure on whichw⇤ is constant — represent regions in which the flow is irrotational. If ⌦ is bounded, thenthis can be written as a functional of the form I�, with coe�cient functions f = ⇢TF , g =2⇢TF /r.

All our results are formulated with this application in mind. In particular, we willalways assume that

f 2 C(⌦) \ L1(⌦), f > 0 in ⌦, (1.4)

g 2 L1loc(⌦),

1

g2 L1

loc(⌦), g > 0 in ⌦. (1.5)

Note that we do not assume that f, g are bounded away from zero, or that g is integrable.(We require g 2 L1(⌦) only once, in Lemma 4.2, and it will turn out that other mitigatingfactors make the results of this lemma available for functionals of the form (1.3), evenwhen g = ⇢TF /r is not integrable.) The second half of this paper focuses on the specificexample (1.3), for which the corresponding f, g described above satisfy (1.4), (1.5).

The book [1] is a good general reference for mathematical issues relating to vorticesin Bose-Einstein condensates. Most prior work on 3d condensates has dealt with anasymptotic regime in which the number of vortex filaments is bounded. As proposed in[3], then justified first in [21] and in greater generality in [24], in this limit the geometryof vortex filaments is governed by a limiting functional that is closely related to what oneobtains by considering in E above only the terms that are homogeneous of degree 1 withrespect to v. Studies of this limiting functional are carried out in [2] and [25].

WEIGHTED TV MINIMIZATION AND APPLICATIONS TO VORTEX DENSITY MODELS 3

Plan of the paper: in Section 2 we revisit the theory of weighted BV functions andweighted TV functionals so as to include possibly degenerate weights, and we studygeneral properties of minimizers such as existence, uniqueness and the Euler-Lagrangeequation.

In Section 3 we investigate properties of level sets of minimizers. Our main result inthis section is Proposition 3.1, where we show a uniform lower bound on the measureof superlevel sets. As a consequence, for bounded minimizers, the maximal level set isflat and has positive measure. The proof is based on a suitable weighted isoperimetricinequality which may be of independent interest.

In Section 4, under an additional regularity assumption on the weights, and using someideas from [25], it is shown that minimizers are in fact of class BV (Lemma 4.1), and indimension n = 2 level sets meet orthogonally the boundary (Lemma 4.2).

In Section 5 we apply the previous theory to Bose-Einstein condensation models (forwhich g /2 L1(⌦)). Our main result in this section is Theorem 5.1, where we analyticallyderive properties of minimizers and their level sets, which represent vortex lines in theconsidered models. In particular, our results imply that in the limit in which condensatesare described by the functional Ered

↵ (·) from (1.3), a condensate in the ground state alwayspossesses a zone that vortex filaments do not penetrate, and in which the correspondingflow is thus irrotational.

In Section 6 we show some numerical simulations, and eventually in the appendix weprove the weighted isoperimetric inequality used in Proposition 3.1.

Acknowledgments. The authors wish to thank Mauro Bonafini for fruitful discussionsabout the numerics, and Ivan Salgado for a careful reading of the paper. The authorsare also grateful to the Centro de Giorgi, where part of this work was carried out. R.J.was partially supported by the National Science and Engineering Research Council ofCanada under operating Grant 261955. M.N. was partially supported by the ItalianCNR-GNAMPA and by the University of Pisa via Grant PRA-2015-0017.

2. Weighted TV regularization – general results for rough weights

2.1. Preliminaries on BV functions. The natural space for minimization of I� is theweighted L2 space

H := L2g(⌦) := {u : ⌦ ! R measurable : kuk < 1}

with the norm and inner product

kuk := hu, ui1/2, hu, vi :=Z

⌦g u v dx. (2.1)

Note that the weight g is not explicitly indicated in the notation for k · k or h·, ·i. Withminimization problems in mind, we always work in the space H, even when discussingthe pure weighted total variation J(·), without the quadratic term. Thus, all definitionsthat follow (subgradient, Legendre-Fenchel transform, etc.) are understood with respectto the H = L2

g inner product.We also explicitly point out that, in this spirit, the domain of J is always understood

to be a subset of H:DomJ := {u 2 H : J(u) < 1}. (2.2)

4 P. ATHAVALE, R. L. JERRARD, M. NOVAGA, AND G. ORLANDI

Let ⌦ be an open subset of Rn. A function u 2 L1(⌦) whose gradient Du in the senseof distributions is a (vector-valued) Radon measure with finite total variation in ⌦ iscalled a function of bounded variation. The class of such functions will be denoted byBV (⌦). The total variation of Du on ⌦ is defined as

|Du|(⌦) = sup

Z

⌦u div z dx : z 2 C1

c (⌦;RN ), |z(x)| 1 8x 2 ⌦

, (2.3)

where for a vector v = (v1, . . . , vn) 2 Rn we set |v|2 :=Pn

i=1 v2i , and is denoted by

|Du|(⌦) or by R⌦ |Du|. The map u ! |Du|(⌦) is L1loc(⌦)-lower semicontinuous. BV (⌦)

is a Banach space when endowed with the norm kuk :=R

⌦ |u| dx+ |Du|(⌦).Analogously, we define BVloc(⌦) as the space of functions u 2 L1

loc(⌦) such that|Du|(U) < +1 for any U b ⌦.

A measurable set E ✓ ⌦ is said to be of finite perimeter in ⌦ if (2.3) is finite when u issubstituted with the characteristic function �E of E. The perimeter of E in ⌦ is definedas P (E,⌦) := |D�E |(⌦). We denote by Ln and Hn�1, respectively, the n-dimensionalLebesgue measure and the (n� 1)-dimensional Hausdor↵ measure in Rn.

In case E is a set with finite perimeter, one can define its “reduced boundary” @⇤E asthe set of points where there exists the limit

lim⇢!0

D�E(B⇢(x))

|D�E |(B⇢(x))= ⌫E(x) 2 Sn�1.

Then, one has the representation

D�E = ⌫E(x)|D�E | and |D�E | = Hn�1x @⇤E .

The first equality just follows from Besicovitch’s derivation theorem for Radon measures,while the second from a careful study of the reduced boundary, see Section 3 in [7]. Inparticular, it follows that for any open set A ⇢ ⌦,

P (E,A) = Hn�1(@⇤E \A) .

If E ✓ Rn is a measurable set and x 2 Rn, we define the (Lebesgue) upper density ofE at x by

D(E, x) := lim sup⇢!0

|E \B⇢(x)||B⇢(x)| ,

while the density, when it exists, is simply the limit lim⇢!0 |E \ B⇢(x)|/|B⇢(x)|. Thenit can also be shown that up to Hn�1-negligible sets, the reduced boundary @⇤E of a setwith finite perimeter coincides with the set of points where E has density exactly 1/2,and with the set of points where E has density neither 0 nor 1.

Given u 2 BV (⌦), we define

u+(x) := inf{t : D({u > t}, x) = 0} and u�(x) := sup{t : D({u < t}, x) = 0}.We say that u is approximately continuous at x 2 ⌦ if and only if u+(x) = u�(x). Theset of points where u is not approximately continuous is usually called the jump set ofu and is denoted by Ju. If u is the characteristic �E of a set with finite perimeter, then

WEIGHTED TV MINIMIZATION AND APPLICATIONS TO VORTEX DENSITY MODELS 5

@⇤E ⇢ J�E and Hn�1(J�E \ @⇤E) = 0. On the other hand, almost any level {u > t} of aBV function has finite perimeter, and there holds the co-area formula

|Du|(⌦) =

Z +1

�1P ({u > t},⌦) dt . (2.4)

For a comprehensive treatment of functions of bounded variation, we refer the readerto [7].

2.2. Basic theory for J and I�. In this section we assume that ⌦ is a bounded, opensubset of Rn, and that f, g satisfy (1.4), (1.5). We collect here some basic results aboutJ and I�, see e.g. [17, 4, 15, 16]. We first characterize the domain of the functional J :the following useful fact is a special case of the general results established in [6].

Lemma 2.1. If u 2 L1loc(⌦), then J(u) < 1 if and only if

u 2 BVloc(⌦), and f is |Du|-integrable.

We now recall a version of the coarea formula. The statement we give (in particularour convention for defining the sub/super level sets Et) is more or less forced on us bythe fact that if w 2 H and g 62 L1(⌦), then �{w>t} 62 H for t < 0.

Lemma 2.2 (Coarea Formula). Assume that w 2 DomJ , and for t 2 R define

Et = Et(w) := {x 2 ⌦ : sign(t)(w(x)� t) > 0}. (2.5)

Then �Et 2 DomJ for a.e. t 2 R, and

J(w) =

Z

⌦f |Dw| =

Z 1

�1

Z

⌦\@⇤Et

fdHn�1

dt =

Z 1

�1J(�Et)dt .

Proof. Since |w| > |t| in Et, it is clear that �2Et

t�2w2, and hence that �Et 2 H forevery nonzero t. Next, Lemma 2.1 implies that w 2 BVloc, and then the standard coareaformula implies that the set {x 2 ⌦ : w(x) > t} has locally finite perimeter for a.e. t, andthat

Z

⌦f |Dw| =

Z 1

�1

Z

⌦\@⇤{w>t}fdHn�1

!

dt .

Then the thesis follows from the facts that @⇤{w > t} \ ⌦ = @⇤{w t} \ ⌦ for every t,and that @⇤{w t} \ ⌦ = @⇤{w < t} \ ⌦ for every t such that the level set {w = t} hasmeasure zero, which holds for a.e. t. ⇤

Similarly, we will sometimes use a “layer-cake decomposition” in the formZ

⌦'(x)q(w(x), x) dx =

Z

Rsign(t)

Z

Et

'(x)qt(t, x) dx dt (2.6)

6 P. ATHAVALE, R. L. JERRARD, M. NOVAGA, AND G. ORLANDI

where w 2 H and ' 2 L21/g, and q(t, x) is a Lipschitz function (often q(t, x) = t) such

that q(0, x) = 0. Indeed, in this caseZ

⌦'(x)q(w(x), x) dx =Z

{w>0}'(x)

Z

�{0<t<w(x)} qt(t, x) dt dx�Z

{w<0}'(x)

Z

�{0>t>w(x)} qt(t, x) dt dx,

and then (2.6) follows by Fubini’s Theorem.

Notice now that for a general convex positively 1-homogeneous function J on a Hilbertspace H we have, for any u 2 DomJ ,

⌘ 2 @J(u) , J(v)� J(u) � h⌘, v � ui for any v 2 DomJ, (2.7)

where h·, · i denotes the inner product in H. Testing with v = �u for � > 0, we obtain

⌘ 2 @J(u) ) J(u) = h⌘, ui . (2.8)

We deduce from (2.7) the equivalent characterization

⌘ 2 @J(u) , J(u) = h⌘, ui and h⌘, vi J(v) for any v 2 DomJ . (2.9)

The following lemma, a small modification of results of Anzellotti [8], will be neededfor a characterization of the subgradient @J .

Lemma 2.3. For any v 2 DomJ ⇢ L2g and ⇠ 2 L1(⌦;Rn) such that div(f⇠) 2 L2

1/g, we

define a distribution (f⇠ ·Dv) byZ

⌦(f⇠ ·Dv)' := �

Z

⌦f⇠ ·r' v �

Z

⌦div(f⇠)v', for ' 2 C1

c (⌦) (2.10)

Then (f⇠ ·Dv) is well-defined and satisfies

Z

⌦(f⇠ ·Dv)'

k'k1 k⇠k1Z

Af |Dv| if supp(') ⇢ A open, (2.11)

and hence is a Radon measure with finite total variation.

Proof. This is proved in [8, Theorem 1.5] if f = g = 1. In the present situation, ourstanding assumptions (1.4), (1.5) and the fact that div(f⇠) 2 L2

1/g guarantee that the

integrals on the right-hand side of (2.10) are well-defined. Then (2.11) is clear if v issmooth, and the general case follows, as in [8], by a density argument. This is straight-forward since we only require approximation of v 2 DomJ by smooth functions in theinterior of ⌦, and this is clearly possible in view of (1.4) and (1.5). ⇤

Lemma 2.4. If u 2 DomJ , then ⌘ 2 @J(u) if and only if there exists a vector field

⇠ 2 L1(⌦;Rn) such that k⇠k1 1, div(f⇠) 2 L21/g,

Z

⌦(f⇠ ·Dv) = h⌘, vi =

Z

⌦g ⌘ v for all v 2 DomJ, (2.12)

and

Z

⌦(f⇠ ·Du) = J(u) =

Z

⌦f |Du| . (2.13)

WEIGHTED TV MINIMIZATION AND APPLICATIONS TO VORTEX DENSITY MODELS 7

Moreover, if Du is an integrable vector field, then ⇠ = Du|Du| a.e. in supp (Du) \ ⌦.

Notice that, by taking v 2 C1c (⌦) ⇢ DomJ , we get that

(2.12) =) div(f⇠) = �⌘g in ⌦.

Formally, (2.12) also implies the boundary conditionZ

@⌦vf⇠ · ⌫ = 0 for all v 2 DomJ ⇢ L2

g.

One cannot immediately conclude that f⇠ ·⌫ = 0 on @⌦, even formally or in a weak sense,without further assumptions on g, since membership in H may force every v 2 DomJ tovanish to high order on some subset of @⌦. On the other hand, the boundary conditionf⇠ · ⌫ = 0 trivially holds on portions of @⌦ where f vanishes.

Proof. Fix ⌘ 2 @J(u), and define the Banach space

Mf (⌦;Rn) = Mf := {Rn-valued measures ~µ on ⌦ : k~µkMf :=

Z

⌦f d|~µ| < 1}.

If v 2 DomJ , then Dv 2 Mf , and

|h⌘, vi| J(v) = kDvkMf . (2.14)

Thus the linear functional Dv 7! h⌘, vi extends, by the Hahn-Banach Theorem, to alinear functional � : Mf ! R satisfying

�(Dv) = h⌘, vi =

Z

⌦g ⌘ v for v 2 DomJ (2.15)

and

|�(~µ)| Z

⌦f d|~µ| for all µ 2 Mf . (2.16)

Let �0 denote the restriction of � to L1f (⌦;Rn). From (2.16) we deduce that there exists

a vector field ⇠, with k⇠k1 1, such that

�0( ) = �( ) =

Z

⌦f(x)⇠(x) · (x) for all 2 L1

f (⌦;Rn). (2.17)

By combining (2.15) and (2.17), we see that if v 2 C1c (⌦) ⇢ DomJ , then the distri-

bution div(f⇠) satisfiesZ

1pgdiv(f⇠)

(pgv) := �

Z

⌦f⇠ ·Dv = ��0(Dv) CkvkH = CkpgvkL2(⌦).

Since this holds for all v 2 C1c (⌦), and since g is bounded below on compact subsets of

⌦, it follows that div(f⇠) 2 L21/g. As a result, �(Dv) =

R

⌦ div(f⇠)v for all compactlysupported v 2 DomJ .

8 P. ATHAVALE, R. L. JERRARD, M. NOVAGA, AND G. ORLANDI

Next, for v 2 DomJ and ' 2 C1c (⌦), note that D(v') = vr'+ 'Dv, from which it

follows that v' 2 DomJ . Thus, by the linearity of �,

�('Dv) = �(D(v'))� �0(v r')= �

Z

⌦div(f⇠) v'�

Z

⌦f⇠ ·r'v

=

Z

⌦(f⇠ ·Dv)'.

Since (f⇠ ·Dv) is a finite measure, and in view of (2.16), this identity continues to holdfor ' 2 C(⌦). In particular, we may set ' = 1 to find that

Z

⌦(f⇠ ·Dv) = �(Dv) = h⌘, vi for v 2 Dom J.

Thus (2.12) holds, and (2.13) follows from (2.8).Conversely, if (2.12) and (2.13) are satisfied, then it follows from (2.11) that

h⌘, vi J(v) for all v 2 DomJ, h⌘, ui = J(u).

As noted in (2.9), this implies that ⌘ 2 @J(u).Finally, if u 2 W 1,1 then f⇠ ·Du 2 L1(⌦). So in this case (2.12), (2.13) imply that

Z

⌦f (|Du|�Du · ⇠) =

Z

supp(Du)f (|Du|�Du · ⇠) = 0

which, since k⇠k1 1, implies the final conclusion of the lemma. ⇤

A consequence of (2.13) is the following

Lemma 2.5. Let u 2 DomJ and define Et = Et(u) as in (2.5). Then ⌘ 2 @J(u) if andonly if sign(t)⌘ 2 @J(�Et) for a.e. t 2 R.

Proof. Assume that ⌘ 2 @J(u) and fix ⇠ as in Lemma 2.4. Then, denoting by �Et thecharacteristic function of Et, Lemmas 2.2 (the coarea formula) and 2.4 imply

Z

RJ(�Et) dt = J(u) =

Z

⌦(f⇠ ·Du) . (2.18)

On the other hand, fixing a sequence of compactly supported smooth functions 'k % �⌦,we haveZ

⌦(f⇠ ·Du) = lim

k

Z

⌦(f⇠ ·Du)'k

= limk

Z

Rsign(t)

Z

⌦(f⇠ ·D�Et)'k dt =

Z

Rsign(t)

Z

⌦(f⇠ ·D�Et) dt.

The second equality follows from (2.10) and (2.6), and the last equality follows from(2.11), (2.18), and the Dominated Convergence Theorem. Combining these identitiesyields

Z

Rdt

J(�Et) � sign(t)

Z

⌦(f⇠ ·D�Et)

= 0.

WEIGHTED TV MINIMIZATION AND APPLICATIONS TO VORTEX DENSITY MODELS 9

Since |⇠| 1, by (1.1) we have, for any t 2 R, J(�Et) � sign(t)R

⌦(f⇠ ·D�Et), and hencefor (2.18) to hold we must have, for a.e. t 2 R,

J(�Et) =

Z

⌦(f(sign(t)⇠) ·D�Et) . (2.19)

In view of (2.12), (2.13), this states that sign(t)⌘ 2 @J(�Et). The proof of the converseis very similar. ⇤

2.3. Properties of minimizers of I�. We have the following

Proposition 2.6. Given f, g satisfying (1.4), (1.5), � > 0 and u0 2 H, there exists a

unique minimizer w 2 DomJ of I�. This minimizer is characterized by

�(u0 � w) 2 @J(w). (2.20)

In addition, if u0 2 L1(⌦), then

ess inf

⌦u0 ess inf

⌦w ess sup

⌦w ess sup

⌦u0 . (2.21)

Proof. Note that I�(0) =�2ku0k2 < 1, so DomI� is nonempty. The functional I� is a

coercive, convex, weakly lower semicontinuous functional on H, so that the existence ofa minimizer w 2 Dom J is obtained by direct methods in the calculus of variations (see[7]). Uniqueness is a consequence of the strict convexity of the functional I�.

If we temporarily write Q(u) = 12ku� u0k2, then the Euler-Lagrange equation for the

minimizer of I� is

0 2 @I�(w) = @(J + �Q)(w) = @J(w) + �@Q(w),

where the final equality follows from standard convex analysis considerations, see for ex-ample [19], Proposition I.5.6. Upon rewriting this, we obtain the necessary and su�cientcondition (2.20).

The uniform bound (2.21) follows by noticing that, for any v 2 BV (⌦),

I�(ess infu0 _ v ^ ess supu0) I�(v).

2.4. Dual formulation. Equation (2.20) is equivalent to

w 2 @J⇤ (�(u0 � w)) , (2.22)

where J⇤, the convex conjugate to J , is given by

J⇤(v) =

0 if kvk⇤ 1+1 otherwise,

(2.23)

andkvk⇤ = sup { hv, ui : u 2 H, J(u) 1} . (2.24)

We may view the norm k ·k⇤ related to the dual convex function J⇤ as a weighted W�1,1

norm. Very much as in the poof of Lemma 2.4, the Hahn-Banach theorem implies thevariational characterization

kvk⇤ = inf

k⇠k1 :

Z

⌦(f⇠ ·Dv) = h⌘, vi for all v 2 DomJ

, (2.25)

10 P. ATHAVALE, R. L. JERRARD, M. NOVAGA, AND G. ORLANDI

where the infimum is taken among vector fields ⇠ 2 L1(⌦;Rn) such that div(f⇠) 2 L21/g.

The following variant of a standard fact illustrates the impact of the (non)integrabilityof g.

Lemma 2.7. Assume that u0 2 H = L2g, let w be the minimizer of I�, and define

a = a⇤ :=

(

R

g(x)u0(x)dx/R

g(x)dx if g 2 L1(⌦)

0 if g 62 L1(⌦).(2.26)

Then

w is constant () w = a⇤ () � �c =1

ku0 � a⇤k⇤ . (2.27)

Moreover, if g 2 L1(⌦), thenZ

g(u0 � w) = 0. (2.28)

The minimizer w does not need to satisfy (2.28) if g 62 L1 (when the integral makessense, which is not guaranteed). This can be seen from the concrete problems consideredin Section 5, in which u0 is positive and belongs to L1

g, but ku0k⇤ < 1, so that w = 0 isthe minimizer for su�ciently small values of �.

If n � 3, there exist u0 2 H such that ku0 � a⇤k⇤ = +1, and hence w is nonconstantfor all � > 0. In the unweighted case, this is essentially equivalent to the statement thatL2 does not embed in W�1,1 for n � 3, and the presence of weights satisfying (1.4), (1.5)does not change things. Whether or not such u0 exist in 2-dimensional domains dependson the behavior of f and g near @⌦.

Proof. It follows from (2.22) and (2.23) that

�ku0 � wk⇤ 1, (2.29)

since @J⇤(�(u0 � w)) = ; if J⇤(�(u0 � w)) = +1.If g is integrable, then every constant function c belongs to L2

g. Then (2.29) and thedefinition (2.24) of the k·k⇤ norm imply that hu0�w, ci ku0�wk⇤ for every c, and hencethat hu0 � w, 1i = 0, which is (2.28). It then follows from (2.28) that if w is constant,then the constant must equal a⇤.

On the other hand, if g 62 L1, then no nonzero constant belongs to H, and so the onlypossible constant minimizer is a⇤ = 0.

In either case, the definition of �c and the characterization (2.23) of J⇤ imply that if� > �c then J⇤(�(u0 � a⇤)) = +1. Then @J⇤(�(u0 � a⇤)) = ;, and w = a⇤ does notsatisfy (2.22).

On the other hand, if � �c then J⇤(�(u0 � a⇤)) = 0, and it follows that a⇤ 2@J⇤(�(u0�a⇤)). This is clear if a⇤ = 0, since J⇤ attains its global minimum at �(u0�a⇤).If a⇤ 6= 0, then necessarily g 2 L1(⌦), and then one can check that every constant functionbelongs to @J⇤(�(u� a⇤)). Indeed, for any v 2 H and any a 2 R

J⇤(�(u0 � w)) + ha, v � �(u0 � a⇤)i = ha, vi J⇤(v).

WEIGHTED TV MINIMIZATION AND APPLICATIONS TO VORTEX DENSITY MODELS 11

Remark 2.8. Using the language of di↵erential forms, and fixing a 1-form !0 such thatd!0 = 0 (i.e. !0 is closed), (f!0)N = 0 on @⌦ \K and v = d⇤(f ·!0), we may reformulatethe variational problem (2.25) as follows

kvk⇤ = inf

d⇤

f+ !0

1, 2 C1(⌦;⇤2(Rn)), N = 0 on @⌦ \K

. (2.30)

The explicit derivation of (2.30) from (2.25) can be found in [13, Section 2].We observe that Problem (2.30) can be interpreted as a (weighted) vector-valued analog

of the Aronsson problem [9]. It would be interesting to give further equivalent descrip-tions of problem (2.30) in a game-theoretical framework, as the tug-of-war description ofAronsson problem (see e.g. [12], [26]).

2.5. Level set formulation. We shall say that a set E ⇢ ⌦ is admissible if �E 2 DomJ .For an admissible E ⇢ ⌦, we have

P (E) = J(�E) =

Z

@⇤E\⌦f(x) dHn�1, A(E, t) = sign(t)

Z

E(u0 � t)g(x) dx .

and

I�,t(E) = P (E)� �A(E, t)

Note that A(E, t) = sign(t)hu0 � t�E ,�Ei, which makes sense for an admissible set E.

Lemma 2.9. If w 2 DomJ , there holds

I�(w)� I�(0) =

Z

RI�,t(Et) dt .

Proof. The thesis follows from the coarea formula (Lemma 2.2) and the “layer-cake de-composition”, see (2.6), applied to

Z

g

2(w � u0)

2 �Z

g

2u20 =

Z

⌦g(x)q(w(x), x) dx for q(t, x) =

1

2t2 � tu0(x),

where Et = Et(w) is defined in (2.5). ⇤

Proposition 2.10. A function w 2 DomJ minimizes I� if and only if, for a.e. t, theset Et = Et(w) minimizes I�,t among all admissible E ⇢ ⌦.

Proof. It is immediate from Lemma 2.9 that if Et minimizes I�,t for a.e. t, then wminimizes I�.

Conversely, notice first that from the Euler-Lagrange equation (2.20) for I� and Lemma2.5 we have sign(t)�(u0 � w) 2 @J(�Et) for a.e. t 2 R. Hence, if �F 2 DomJ we have

sign(t)

Z

⌦�(�F � �Et)(u0 � w)g dx J(�F )� J(�Et) = P (F )� P (Et).

12 P. ATHAVALE, R. L. JERRARD, M. NOVAGA, AND G. ORLANDI

Moreover, (2.5) implies that sign(t)(t� w)(�F � �Et) � 0, so we deduce that

�A(F, t)� �A(Et, t) = sign(t)

Z

⌦�(�F � �Et)(u0 � t)g dx

sign(t)

Z

⌦�(�F � �Et)(u0 � w)g dx P (F )� P (Et) , (2.31)

and the conclusion follows. ⇤

Remark 2.11. From the preceeding proof we deduce that if F minimizes I�,t thennecessarily we must have equality in (2.31) and hence the minimizer w satisfies

(t� w(x)) · (�F (x)� �Et(x)) = 0 for a.e. x 2 ⌦.

This implies that Et ⇢ F ⇢ Et := {x : sign(t)(w(x) � t) � 0}. The sets Et and Et arecalled respectively the minimal and the maximal solution of the minimization problemmin I�,t. In particular, since for a.e. t 2 R the level set {w(x) = t} has zero measure, Et

is the unique minimizer of I�,t for a.e. t 2 R.

Remark 2.12. By (2.20) and (2.9) we deduce that

h�(u0 � w),�F i =Z

F�(u0 � w)g(x) dx

Z

@⇤Ff(x) dHn�1 = P (F ) (2.32)

whenever �F 2 DomJ , with equality for F = Et, for a.e. t 2 R, i.e.

sign(t)

Z

Et

�(u0 � w)g(x) dx =

Z

@⇤Et

f(x) dHn�1 = P (Et). (2.33)

One may wonder if for F satisfying (2.33) we necessarily have that F minimizes I�,t.Since sign(t)(t� w)�Et 0 it follows from (2.33) that for a.e. t 2 R,

min I�,t = I�,t(Et) = P (Et)� �A(Et, t) = sign(t)

Z

Et

�(t� w)g(x) dx 0 . (2.34)

Moreover, min I�,t = 0 if and only if t � ess sup⌦w > 0 or t ess inf⌦w < 0.

Remark 2.13. Since I�,t(E) I�,s(E) for every E, when 0 < t < s, we have

min I�,t min I�,s for any 0 < t < s, (2.35)

Moreover, still assuming s > t > 0, we have Es ⇢ Et and

ı(s)� ı(t)(2.34)= �

Z

Et\Es

(t� w)g(x) dx�Z

Es

(t� s)g(x) dx ,

where ı(⌧) = min I�,⌧ , so that

lims!t+

ı(s)� ı(t)

s� t=

Z

Et

g(x) dx , limt!s�

ı(s)� ı(t)

s� t=

Z

Esg(x) dx ,

where Es =T

`<sE` = {w � s}. Hence the map ı possesses bounded left and rightderivatives at every point. Flat level sets of w correspond to points of non-di↵erentiabilityfor ı. Similar considerations apply to t < s < 0, with signs reversed.

WEIGHTED TV MINIMIZATION AND APPLICATIONS TO VORTEX DENSITY MODELS 13

3. Extremal level sets of the minimizer

When (2.27) is violated, the minimizer w is nontrivial, and hence one may wonderwhether w has “flat spots” – that is, whether any level sets of w have positive measure,as happens in the classical obstacle problem when the constraint is saturated or also forrelated models (see e.g. [14]).

The answer to this question requires more assumptions on the boundary behaviour off and g with respect to those that we have made up to now. In particular, we will stillassume (1.4) and (1.5), and we will also require that

f(x) � C⇢m(x), |u0(x)| g(x) C⇢(x)m✓ , (3.1)

for some positive integer m 2 N and some ✓ 2 (0, 1] to be specified below, where

⇢(x) := dist(x, @⌦) for x 2 ⌦. (3.2)

We continue to allow non-integrable g’s, but we require that the possible blow-up of gnear @⌦ is compensated by the decay of u0.

Proposition 3.1. Assume that ⌦ ⇢ Rnis a bounded open connected set with smooth

boundary, and let w minimize I�.If f, g, u0 satisfy (3.1) for some m 2 N and ✓ 2 (n+m�1

n+m , 1], then there exists some

constant a > 0, depending on f, g, u0, n,⌦,�, such that

Ln(Et) � a for every t such that Ln(Et) > 0. (3.3)

In particular, w 2 L1(⌦), and if either of t± := ess sup⌦w±is positive, then the

corresponding level set E± := {x 2 ⌦ : w(x) = ±t±} has measure at least a.Finally, if (3.1) holds with ✓ = n+m�1

n+m , then w 2 L1(⌦).

Remark 3.2. Proposition 3.1 remains valid if there exists h : Rn�1 ! R such that

⌦ = {(x1, x0) 2 R⇥ Rn�1 : 0 < x1 < h(x0)}, |rh(x0)| > 0 where h(x0) = 0,

and in addition f > 0 on {x 2 @⌦ : x1 = 0}. Indeed, in the proof of the proposition, thesmoothness of @⌦ is only used in invoking the weighted relative isoperimetric inequality(see Proposition 3.4 below), and this result is still valid under the hypotheses describedhere.

Remark 3.3. If u0 � 0, so that w � 0 and t� = 0, then Proposition 3.1 does not ingeneral say whether the level set {w = ess inf⌦w} is flat.

However, if g(x) C⇢m✓ (which in particular implies that g is integrable) and u0 2L1, then both extremal level sets of w are always flat. Indeed, the hypotheses of theproposition are still satisfied by

I�(u) := J(u) +�

2

Z

⌦g|u� u0|2, for u0 := u0 � c

for any constant c. Moreover, if w minimizes I�, then w := w�c clearly minimizes I�. Bya suitable choice of c, one can arrange that u0 0 in ⌦, and hence that ess inf⌦ w < 0.It follows that the level set {w = ess inf⌦ w} = {w = ess inf⌦w} is flat.

14 P. ATHAVALE, R. L. JERRARD, M. NOVAGA, AND G. ORLANDI

On the other hand, in the model case discussed in the next section, the results in [2]imply that {w = ess inf⌦w} may have either positive or zero measure, depending on thechoice of the parameters. Note that in this case, we have g 62 L1(⌦).

The main tool in the proof of Proposition 3.1 is the following weighted relative isoperi-metric inequality.

Proposition 3.4. Assume that ⌦ ⇢ Rnis a bounded open connected set with smooth

boundary.

Then, for every m 2 N there exists a constant C = C(⌦,m, n) such that for every set

E ⇢ ⌦ of locally finite perimeter,

min

Z

E⇢m(x) dx,

Z

⌦\E⇢m(x) dx

!

C(⌦,m, n)

Z

@⇤E\⌦⇢m(x)dHn�1

n+mn+m�1

. (3.4)

We mention that weighted isoperimetric problems have been widely studied in recentyears (see for instance [18] and references therein).

The proof of Proposition 3.4 is deferred to the Appendix. For now, we accept theresult and use it to prove that positive and negative extremal level sets of the minimizerare flat.

Proof of Proposition 3.1. We first claim that there exists a constant a1 > 0 such that

if A(E, t) a1 and 0 <

Z

E⇢m dx 1

2

Z

⌦⇢m dx, then I�,t(E) > 0. (3.5)

Note that under these conditions, our hypotheses on f and the relative isoperimetricinequality (3.4) imply that

0 <

Z

E⇢m dx C

Z

@⇤E\⌦f dHn�1

n+mn+m�1

= CP (E)n+m

n+m�1 . (3.6)

For simplicity, we will prove (3.5) for t > 0; the argument for t < 0 is essentiallythe same. First, we may assume that A(E, t) > 0, since if A(E, t) 0, then I�,t(E) �P (E) > 0 by (3.6).

Next, for an admissible E with A(E, t) > 0, we compute

A(E, t) =

Z

E(u0 � t)g dx

Z

E(u0 � t)+g dx C

Z

E⇢✓m

(u0 � t)+

u0dx

using our assumption (3.1) on g. Thus by Holder’s inequality and (3.6),

A(E, t) C

Z

E⇢m dx

◆✓

Z

(u0 � t)+

u0

11�✓

dx

!1�✓

c(t)P (E)�

for

� =✓(n+m)

n+m� 1> 1 , c(t) = C

Z

(u0 � t)+

u0

11�✓

dx

!1�✓

.

It follows that

I�,t(E) �

A(E, t)

c(t)

�1/�

� �A(E, t).

WEIGHTED TV MINIMIZATION AND APPLICATIONS TO VORTEX DENSITY MODELS 15

Since c(t) CLn(⌦)1�✓ for all t, is is easy to find a1 > 0 such that I�,t(E) > 0 if0 < A(E, t) a1, proving (3.5).

Next, fix a > 0 such that if Ln(E) < a, then the hypotheses of (3.5) are satisfied. If0 < t < t+, we deduce from (2.34) that I�,t(Et) 0, and then from (3.5) that Ln(Et) � a,proving (3.3). It easily follows that w 2 L1 and that Ln(E+) = Ln(\0<t<t+Et) � a.

Finally, if ✓ = n+m�1n+m , then the above arguments show that if A(Et, t) > 0, then

0 � It,�(Et) � A(Et, t)

1

c(t)� �

.

By the Dominated Convergence Theorem, there exists some T � 0 such that c(t) 12�

whenever t � T . As a result, if t � T , then A(Et, t) 0, which implies that I�,t(Et) � 0,and hence (again by Remark 2.12) that t � t+ = ess sup⌦w. This proves that t+ T .Since similar considerations apply to t� = ess sup⌦w�, it follows that w 2 L1(⌦). ⇤

4. BV and boundary behavior for smooth weights

In this section we impose further decay and smoothness conditions on the weight f , aswell as compatible decay conditions on g, involving the minimizer w as well as the datau0.

Lemma 4.1. Assume that f = !mfor some positive integer m and ! 2 C2(⌦) such that

|r!|2 + ! � c1 > 0 in ⌦, ! = 0 on @⌦. (4.1)

Let w minimize I� and assume that there exists some constant C1 such that

|g(w � u0)| C1f Hn�1a.e. in ⌦. (4.2)

Then w 2 BV (⌦), and there exists some constant C, depending on c1, C1, k and k!kC2,

such that

Z

⌦|Dw| CJ(w), and Hn�1(@⇤Et \ ⌦) C

Z

@⇤Et\⌦fdHn�1

for a.e. t.

Assumption (4.2) is in general hard to verify, since it requires knowing something aboutthe minimizer w. However,

• it is satisfied if g Cf and u0 2 L1.• If g does not decay at @⌦, and in particular if g 62 L1(⌦), it can also be provedto hold for certain u0. Indeed, we prove in Section 5 that this is the case for theproblems related to Bose-Einstein condensation.

The proof is an adaptation to our setting of an argument from Section 3 of [25], themain di↵erence being simply that we use the first variation formula and an approximationargument in place of the direct ODE arguments of [25].

Proof. Consider V 2 C1c (⌦;Rn), and let

�h(x) := x+ hV (x)

Then there exists h0 > 0 such that if |h| < h0, then �h is a di↵eomorphism of ⌦ ontoitself which fixes a neighborhood of @⌦.

16 P. ATHAVALE, R. L. JERRARD, M. NOVAGA, AND G. ORLANDI

Recall from Proposition 2.10 that Et = Et(w) minimizes I�,t(·) = P (·)��A(·, t) amongall admissible sets F , that is, sets such that �F 2 DomJ . Consider any t for which thisholds, and for |h| < h0 let Eh

t := �h(Et). Since �h is C1, standard theory implies that Eht

is a set of locally finite perimeter in ⌦. Moreover, Eht coincides with Et in a neighborhood

of @⌦, and thus is easily seen to be admissible. Then the minimality of E�,t implies that

0 =d

dh

h=0

I�,t(Eht )

=

Z

@⇤Et

h

f(�ij � ⌫i⌫j)Vixj

+ V ifxi

i

dHn�1 � � sign(t)

Z

@⇤Et

g(u0 � t)V · ⌫ dHn�1 ,

where ⌫(x) denotes the approximate outer unit normal vector to Et at x 2 @⇤Et. Indeed,the first integral on the right-hand side is the first variation of the weighted perimeterP (·), see for example [23], chapter 17, and the second integral on the right-hand side isthe first variation of A(·, t).

Let now V = q(!)r!, for some smooth function q such that

0 q 1, q0 � 0, q = 0 in a neighborhood of 0.

Since q � !, and hence V , are compactly supported, we deduce thatZ

@⇤Et

q(!)m!m�1|r!|2dHn�1 = �Z

@⇤Et

!mq0(!)(I � ⌫ ⌦ ⌫) : (r! ⌦r!) dHn�1

�Z

@⇤Et

!mq(!)(I � ⌫ ⌦ ⌫) : r2! dHn�1

� � sign(t)

Z

@⇤Et

q(!)g(u0 � t)r! · ⌫ dHn�1.

For a.e. t, it follows from (4.2) that |g(u0 � t)| C1f = C1!m almost everywhere on@⇤Et. Using this and the fact that q0 � 0, we deduce that

Z

@⇤Et

q(!)m!m�1|r!|2dHn�1 �Z

@⇤Et

!mq(!)(I � ⌫ ⌦ ⌫) : r2! dHn�1

� � sign(t)

Z

@⇤Et

q(!)g(u0 � t)⌫ ·rf dHn�1

k!kC2 (1 + �C1)

Z

@⇤Et

!m.

Letting q % �R+ , we conclude thatZ

@⇤Et\⌦!m�1|r!|2dHn�1 C

Z

@⇤Et\⌦!m dHn�1.

for C = C(k!kC2 , C1,�), for a.e. t. Since |r!|2 + ! � c1 > 0 in ⌦, it follows thatZ

@⇤Et\⌦!m�1 dHn�1 1

c1

Z

@⇤Et\⌦!m�1(|r!|2 + !)dHn�1 C

Z

@⇤Et\⌦!m.

WEIGHTED TV MINIMIZATION AND APPLICATIONS TO VORTEX DENSITY MODELS 17

Repeating essentially same argument, with V replaced successively by q(!)!�jr! forj = 1, . . . ,m� 1, we find that

Z

@⇤Et\⌦!m�j�1 dHn�1 C

Z

@⇤Et\⌦!m�j dHn�1, 0 j m� 1

for a.e. t. Putting together these estimates, we conclude that Hn�1(@⇤Et) CP (Et) fora.e. t. The estimate

R

⌦ |Dw| CJ(w) then follows from the coarea formula. ⇤

The next result is due to Montero and Stephens (see Section 4 of [25]) in a slightlydi↵erent setting. For this we specialize to n = 2, and we impose still stronger assumptionson f, g, u0.

Lemma 4.2. Assume that n = 2, that the hypotheses of Lemma 4.1 are satisfied, and

that in addition

g 2 W 1,1(⌦), u0 2 W 1,1(⌦).

Then for a.e. t, @Et is a union of C2curves meeting orthogonally @⌦ .

We recall the argument from [25]; this will be needed later when we argue that theresult still applies to certain models of Bose-Einstein condensates for which g 62 L1(⌦).

Proof. Fix some t such that Et minimizes I�,t and H1(@⇤Et) < 1 and such that for� := g(u0 � t), we have |�| C1f almost everywhere on @⇤Et. (These conditionsare satisfied a.e. by Proposition 2.10, Lemma 4.1 and (4.2).) The minimality of Et,together with standard regularity results, implies that @Et \ ⌦ is a countable union ofnonintersecting C2-curves, each of which must have finite arclength. Let � : (a, b) ! ⌦be an arclength parametrization of one such curve. The Euler-Lagrange equation for I�,tmay be written (along �) in the form

� �f(�(s))�0(s)�0 +rf(�(s))� ��(�(s))⌫(s) = 0

where ⌫(s) is the outer unit normal to @Et at �(s). Suppose that �(s) approaches @⌦ ass & a or s % b; we consider the former case for concreteness, and we assume that a = 0.Then by integrating the above equation from 0 to s, taking the inner product with ⌫(s),and rewriting, we find (since f = 0 on @⌦) that

0 =1

sf(�(s))�0(s) · ⌫(s)

= ⌫(s) ·rf(�(s)) +1

s

Z s

0⌫(s) ·

h

�rf(�(t))�rf(�(s))�� ��(�(t))⌫(t)

i

dt .

Using this and hypotheses (4.1), (4.2) one can verify that ⌫(s) ·r!(�(s)) ! 0 as s & a,which implies that the curve meets @⌦ orthogonally.

5. Applications to quantum fluids

As mentioned in the Introduction, in certain limits the velocity field in rotationallyforced Bose-Einstein condensates is described by a particular instance of the functionalI� from (1.2), where � = 2 and f, g,⌦ have the form

⌦ := {(r, z) 2 (0,1)⇥ R : ⇢TF (r, z) > 0}, f = ⇢TF , g(r, z) =1

r⇢TF . (5.1)

18 P. ATHAVALE, R. L. JERRARD, M. NOVAGA, AND G. ORLANDI

Here ⇢TF : R2 ! R is a smooth function known as the “Thomas-Fermi density” anddetermined, in physical experiments, by the form of a potential that is used to confinethe condensate. We will assume that

⇢TF + |r⇢TF | > 0 on @⌦, ⇢TF is C2, ⌦ is bounded. (5.2)

The rotational forcing is encoded in the function u0, which has the form

u0(r, z) = ↵r2 (5.3)

where ↵ > 0 is a parameter.The following results about this functional are mostly immediate consequences of the

general theory developed above.

Theorem 5.1. Let I↵(·) denote I�(·) for � = 2 and f, g, u0 as in (5.1), (5.3), where ⇢TF

satisfies (5.2) and either

@⇢TF

@r(r, z) 0 in ⌦ (5.4)

or

⌦ is a bounded subset of (0,1)⇥ R. (5.5)

Then for every ↵ > 0, there exists a unique minimizer w↵ 2 DomJ of I↵.

(1) w↵ is the unique element of H = L2g satisfying the equivalent conditions:

• the Euler-Lagrange inclusion (2.20);• the dual formulation (2.22);• the level set formulation from Proposition 2.10.

(2) w↵ is constant (corresponding to a vortex-free condensate) if and only if

↵ 1

2kr2 � c⇤k⇤ for c⇤ =

(

0 if (5.4) holdsR

⌦ gr2dr dz/R

⌦ gdr dz if (5.5) holds

where the k · k⇤ norm is defined in (2.24).(3) If we write I↵t (E) :=

R

@⇤E f dH1 � 2R

E g(↵r2 � t), then

kw↵kL1 = inf{t > 0 : I↵t (E) � 0 for all admissible E ⇢ ⌦}. (5.6)

(4) The level set {x 2 ⌦ : w↵ = max⌦w↵} has positive measure.

(5) If (5.5) holds, then the level set {x 2 ⌦ : w↵ = inf⌦w↵} has positive measure.

(6) If (5.4) holds, then 0 w↵ u0. As a result, w↵ = @w↵@⌫ = 0 on @⌦ \ {r = 0}.

(7) w↵ 2 BV (⌦), and @w↵@⌫ = 0 on @⌦ \ {r > 0} in the sense that a.e. level set of

w↵ meets @⌦ orthogonally.

Remark 5.2. Both the hypotheses (5.4), (5.5) are motivated by physical experiments.Indeed, assumption (5.4) covers any convex trapping potential that is symmetric aboutthe vertical axis. In particular, this includes the model case of a quadratic trappingpotential, corresponding to ⇢TF (r, z) = ⇢0 � r2

a2� z2

b2for positive constants ⇢0, a, b. Note

that when (5.4) holds, necessarily g 62 L1(⌦).

The second case (5.5) is relevant to toroidal Bose-Einstein condensates.

WEIGHTED TV MINIMIZATION AND APPLICATIONS TO VORTEX DENSITY MODELS 19

Remark 5.3. Recall that the level set formulation states that for a.e. t > 0, the level setEt = {w > t} of w↵ minimizes the functional I↵t appearing in conclusion (3). It followsthat @Et is analytic in ⌦ and satisfies the equation

div(⇢TF ⌫) = ⇢TF +r⇢TF · ⌫ = 2⇢TF

↵r � t

r

(5.7)

where ⌫ and denote respectively the exterior unit normal and the curvature of @Et.(This equation appears already in a di↵erent form in the proof of Lemma 4.2 above.)

Proof of Theorem 5.1. In case (5.5), f and g are smooth in ⌦, and both are comparableto the distance ⇢(·) to @⌦, so all the hypotheses of every result in Sections 2 - 4 aresatisfied (with m = 1 for Section 3), and all the conclusions above follow, after notingthat conclusion (3) is a direct consequence of the level set formulation.

If (5.4) holds, then necessarily ⇢TF > 0 on {(r, z) 2 @⌦ : r = 0}, and the form (5.1) of gimplies that g 62 L1(⌦). In this case, the results of Section 2 are still available, and theseimply conclusions (1) - (3) of the theorem. Moreover, it is clear that |uo|g = ↵r⇢TF C⇢TF = Cf , so the hypotheses of Proposition 3.1 (see also Remark 3.2) are satisfied, andit follows that conclusion (4) holds.

We next prove conclusion (6). The inequality w↵ � 0 follows by noticing that I↵(0 _w) I↵(w) for all w 2 DomJ . The inequality w↵ ↵r2 is equivalent to the inclusions

Et ⇢ H↵t for a.e. t > 0, (5.8)

where we set H↵t = {↵r2 > t}. To prove (5.8) it is su�cient to show that

I↵t (E \H↵t ) I↵t (E) for any admissible E ⇢ ⌦. (5.9)

In fact, recalling that@⇢TF@r 0 and ↵r2 t in E \H↵

t , we haveZ

@⇤(E\H↵t )⇢TF

Z

@⇤E⇢TF and

Z

E\H↵t

2⇢TF

r(↵r2 � t)

Z

E

2⇢TF

r(↵r2 � t)

for any admissible E, which gives (5.9).It follows that the hypotheses of Lemma 4.1 are satisfied (in particular condition (4.2)

is a consequence of conclusion (6)), and hence that w 2 BV (⌦).Finally, we prove that @w

@⌫ = 0 on @⌦ \ {r > 0}. It su�ces to show that for a.e. t > 0,the conclusions of Lemma 4.2 hold for @⇤Et. The lemma as stated does not immediatelyapply here, since it assumes that g 2 W 1,1(⌦). But the proof of Lemma 4.2 only requiresg to be Lipschitz near @⇤Et, and this is the case here by (5.8), so the conclusion follows.

We now illustrate a consequence of conclusion (5.6) above.

Lemma 5.4. Assume the hypotheses of Theorem 5.1 and assume that

@⇢TF@r 0 in ⌦.

For t > 0, if there exists a 1-form ⇣t such that

(1) d⇣t = �2⇢TF (↵r � tr )dr ^ dz in ⌦

(2) ⇣t = 0 on {(r, z) 2 @⌦ : r > 0}(3) ⇢TF � |⇣t| > 0 in H↵

t = {(r, z) 2 ⌦ : ↵r2 > t}then kw↵k1 t.

20 P. ATHAVALE, R. L. JERRARD, M. NOVAGA, AND G. ORLANDI

Proof. Properties of ⇣t and Stokes Theorem imply that for any nonempty admissible E,

I↵t (E) =

Z

@⇤E⇢TF dH1 +

Z

@⇤E⇣t �

Z

@⇤E(⇢TF � |⇣t|)dH1 > 0,

where @⇤Et is oriented in the standard way in the integralR

@⇤E ⇣t. The final inequality

above uses conclusion (6) of Theorem 5.1. It follows that Et, as the minimizer of I↵t ,must be the empty set, and hence that kw↵k1 t by (5.6). ⇤

On the other hand, (5.6) also implies lower bounds for kw↵kL1 , such as

if i↵t (�) := It↵

⌦ \ {r > �}⌘

< 0 for some � > 0, then kw↵k1 � t. (5.10)

Example 5.5. We now consider the model case arising from a quadratic trapping po-tential:

⇢TF = ⇢0 � r2

a2� z2

b2, ⇢0, a, b positive constants. (5.11)

In this case, a 1-form ⇣ satisfying the conditions of Lemma 5.4 is given by ⇣t = ⌘t(r, z) dz,where

⌘t :=a2↵

2⇢2TF

� t

2R2 ln

aR

r

� ⇢TF

, where R = R(z) :=

⇢0 � z2

b2

12

.

(5.12)

Since ⌘t = 0 on {(r, z) 2 @⌦ : r > 0}, and @⌘t

@r < 0 in H↵t , we see that ⌘t > 0 in H↵

t .Thus

kw↵k1 t(↵) := infn

t 2 [ 0,↵a2]�

(⇢TF � ⌘t)(r, z) > 0 in H↵t

o

, (5.13)

for ⌘t as in (5.12).On the other hand, a direct computation shows that the function i↵t (·) appearing in

(5.10) is

i↵t (�) =4b

3Z3 � 8b

15↵a2Z5 + t

"

8b

3⇢3/20 tanh�1

Z

⇢0

� 8b

9Z3 � 8b

3⇢0Z

#

,

where Z = Z(�) := (⇢0 � �2

a2)1/2. Thus (5.10) becomes

kw↵k1 � t(↵) := supn

t 2 [ 0,↵a2]�

min� 2(0, ap⇢0)

i↵t (�) < 0o

(5.14)

for it↵ displayed above.Both t(↵) and t(↵) are easy computed numerically with considerable accuracy. In the

next section we report the results of some such computations, and we compare the resultsto numerical solutions of the minimization problem. These results show that the upperand lower bounds for kw↵k1 given by (5.13) and (5.14) are often surprisingly close toeach other.

WEIGHTED TV MINIMIZATION AND APPLICATIONS TO VORTEX DENSITY MODELS 21

Remark 5.6. The t = 0 functional I↵0 (·) determines whether or not kw↵k1 > 0, andalso determines the curve around which vortices first appear, at the critical value for ↵at which w↵ becomes nontrivial. Following Lemma 5.4 and (5.12), we have

I↵0 (E) =

Z

@⇤E⇢TF dH1 +

1

2↵a2

Z

@⇤E⇢2TF

dz.

This functional is considered in [2], where it is shown that if b > a in (5.11), then forevery ↵ > 0, the minimizer is either ⌦ or the empty set, whereas if a > c0 · b, where c0 isan explicit constant, and ↵ is not too large, then ⌦ does not minimize I0.

6. Numerical results

Since the functionals studied in this paper are weighted versions of the Rudin-Osher-Fatemi denoising model [27], we can compute minimizers by adapting techniques devel-oped in the image processing literature. In the computations reported below, we havefollowed a modified version of a primal-dual algorithm [28], delineated in Algorithm 1below, to compute minimizers of the functional I↵, from Theorem 5.1, in the model caseof a quadratic trapping potential, in which ⇢TF is given by (5.11). We recall that thealgorithm numerically solves the Euler-Lagrange equations (2.20), see also Lemma 2.4,which in the present case may be written formally as as

�r div⇣

⇢TF

Du

|Du|⌘

+ 2⇢TF (u� ↵r2) = 0. (6.1)

These are augmented with boundary conditions from conclusions (6) and (7) of Theo-rem 5.1. Since the problem is formally overdetermined on @⌦ \ {r = 0}, we only enforcethe Dirichlet condition u = 0 there.

We performed computations for the lower and upper limits of kw↵k1 found in (5.13)and (5.14) above, and we compared these to kw↵k1 for an approximation w↵, obtainedusing Algorithm 1 below, of the minimizer w↵.

In Tables 1 and 2 below we show the results of these computations, for a range of valuesof the parameter ↵ and for two di↵erent choices of the parameters a and b appearing inthe definition (5.11) of ⇢TF . Numerical minimizers w↵ for several choices of parametersare displayed in Figures 1 and 2. The computations show clearly that as the rotationalforcing increases, the size of the vortex-free zone (the “flat spot”) decreases but does notvanish.

22 P. ATHAVALE, R. L. JERRARD, M. NOVAGA, AND G. ORLANDI

Figure 1. Minimizer w↵ for a pancake-shaped condensate, with param-eters as in Table 1, for ↵ = 20, 40, 80 from top to bottom.Left: vertical axis is w↵/↵. Note the change in vertical scale.Right: level sets of minimzers.

WEIGHTED TV MINIMIZATION AND APPLICATIONS TO VORTEX DENSITY MODELS 23

Figure 2. Minimizer w↵ for a cigar-shaped condensate, with parametersas in Table 2, for ↵ = 20, 40, 80 from top to bottom.Left: vertical axis is w↵/↵. Note the change in vertical scale.Right: level sets of minimzers.

24 P. ATHAVALE, R. L. JERRARD, M. NOVAGA, AND G. ORLANDI

Algorithm 1: Modified Zhu-Chan primal-dual algorithm.

Data: Given ↵ and TOL (tolerance for stopping)

Initialize: k = 0, u0 = ↵r2, p0 = ru0

|ru0| ;

while kuk � uk�1k1 > TOL do

Choose time discretization parameters ⌧1, and ⌧2;

(we follow [28] and choose ⌧1 = 0.2 + 0.08k, ⌧2 =0.5�5/(15+k)

⌧1.) ;

pk+1 = pk � ⌧1�2⇢

r ruk�

;

pk+1 = pk+1

max{ |pk+1| , 1 } ;

uk = uk�1 + ⌧2⇣

2⇢ (↵r2 � uk) + r div (⇢ pk)⌘

;

Set uk(r = 0) = 0;

k = k + 1;

end

↵ t (↵)↵

kw↵k1 t (↵)↵

5 0.140 0.146 0.2110 0.321 0.320 0.3920 0.485 0.490 0.5440 0.6185 0.620 0.6680 0.722 0.723 0.76160 0.799 0.800 0.82320 0.855 0.856 0.87640 0.8967 0.896 0.911280 0.926 0.925 0.932560 0.9477 0.946 0.95

Table 1. Numerical approximations of t (↵)↵ , kw↵k1, and t (↵)↵ for various

values of ↵. The experiments are performed for pancake-shaped regionsi.e. ⇢0 = 1, a = 1, and b = 0.5.

WEIGHTED TV MINIMIZATION AND APPLICATIONS TO VORTEX DENSITY MODELS 25

↵ t (↵)↵

kw↵k1 t (↵)↵

20 0.0350 0.038 0.0540 0.0803 0.085 0.1080 0.1211 0.125 0.14160 0.1546 0.158 0.17320 0.1802 0.183 0.19640 0.1997 0.201 0.211280 0.2137 0.214 0.222560 0.2242 0.224 0.23

Table 2. Numerical approximations of t (↵)↵ , kw↵k1, and t (↵)↵ for various

values of ↵. The experiments are performed for cigar-shaped regions i.e.⇢0 = 1, a = 0.5, and b = 1.

Appendix A. A weighted relative isoperimetric inequality

In this Appendix we present the proof of the weighted relative isoperimetric inequalitywhich we used in Section 3.

Proof of Proposition 3.4. We divide the proof into five steps.Step 1. We first note that it su�ces to show that there exists some c0 > 0 such that

Z

E⇢m(x) dx C

Z

@⇤E\⌦⇢m(x)dHn�1

n+mn+m�1

if

Z

E⇢m(x) dx c0. (A.1)

Indeed, for any 0 < c0 12

R

⌦ ⇢m, if we define

c1(c0,⌦) := inf

Z

@⇤E\⌦⇢mdHn�1 : c0

Z

E⇢m 1

2

Z

⌦⇢m�

,

then standard arguments imply that the infimum is attained. The fact that ⌦ is smoothand connected implies that

c1(c0,⌦) > 0,

and this fact and (A.1) together imply (3.4).Step 2. For � > 0 to be fixed below, we will write

⌦� := {x 2 ⌦ : ⇢(x) < �}, ⌦� := {x 2 ⌦ : ⇢(x) > �},We will chose � small enough that ⌦� is connected and @⌦� is smooth, and moreover ⌦�

consists of a finite number of connected components, say ⌦i for i = 1, . . . , k, one for eachcomponent of @⌦. (We will also impose a couple of additional smallness conditions on �below, see (A.8), (A.9).) We will write

E� = E \ ⌦�, E� = E \ ⌦�, Ei = E \ ⌦i, i = 1, . . . , k.

26 P. ATHAVALE, R. L. JERRARD, M. NOVAGA, AND G. ORLANDI

In view of (A.1), we may assume thatZ

E⇢m 1

2min

Z

⌦1

⇢m, . . . ,

Z

⌦k

⇢m,

Z

⌦�⇢m�

. (A.2)

Since @⌦� is smooth, it follows from (A.2) and the relative isoperimetric inequality in ⌦�

(see for example [20] section 4.5.2) thatZ

E�⇢m C|E�| C

h

Hn�1(@⇤E� \ ⌦�)in/n�1 C

Z

@⇤E\⌦�⇢mdHn�1

�n/n�1

where all constants may depend on � as well as n,⌦,m. IfR

E� ⇢m > 12

R

E ⇢m, then (A.1)

follows easily, since thenR

E ⇢m C(

R

E� ⇢m)✓ for ✓ := n+mn+m�1

n�1n 2 (0, 1). So we may

assume that this inequality fails, i.e. thatZ

E�

⇢m � 1

2

Z

E⇢m. (A.3)

Once we assume this, in view of (A.2) it is enough to prove the following weighted relativeisoperimetric inequality in ⌦i for every i = 1, . . . , k:

if

Z

Ei

⇢m 1

2

Z

⌦i

⇢m, then

Z

Ei

⇢m Ci

Z

@⇤Ei\⌦i

⇢mdHn�1

n+mn+m�1

. (A.4)

Indeed, once this is know, then for E satisfying (A.2) and (A.3), we haveZ

E⇢m 2

X

i

Z

Ei

⇢m CX

i

Z

@⇤Ei\⌦i

⇢mdHn�1

n+mn+m�1

C

X

i

Z

@⇤E\⌦i

⇢mdHn�1

!

n+mn+m�1

C

Z

@⇤E\⌦⇢mdHn�1

n+mn+m�1

.

Step 3. We now introduce auxiliary functions that will be used in the proof of (A.4).We will extend ⇢ to the complement of ⌦ by setting

⇢(x) :=

(

dist(x, @⌦) if x 2 ⌦

�dist(x, @⌦) if not.

Thus the extended ⇢ is the signed distance to @⌦. We recall some standard facts aboutthe signed distance function. First, because ⌦ is bounded and @⌦ smooth, ⇢ is smoothin

N� := {x 2 Rn : |⇢(x)| �},for all su�ciently small � > 0, and in this set,

⇡(x) := x� ⇢(x)r⇢(x) is the (unique) closest point in @⌦ to x.

In addition,

r⇢(x) = ⌫(⇡(x)), where ⌫(y) = the inward unit normal to @⌦ at y 2 @⌦.

In particular,

|r⇢|2 = 1, and by di↵erentiating, r2⇢r⇢ = 0. (A.5)

WEIGHTED TV MINIMIZATION AND APPLICATIONS TO VORTEX DENSITY MODELS 27

For (x, y) 2 N� ⇥ Rm we will write

⇢(x, y) := (⇢2(x) + |y|2)1/2 = the (unsigned) distance from (x, y) to @⌦⇥ {0}⌦� := {(x, y) : ⇢(x, y) < �}.

Note also that for (x, y) 2 ⌦�, the closest point in @⌦⇥{0} to (x, y) is ⇡(x, y) := (⇡(x), 0).Now we define F : ⌦� ! ⌦� by

F (x, y) = ⇡(x) + ⇢(x, y)r⇢(x) = x+ [⇢(x, y)� ⇢(x)]r⇢(x).We note for future reference that F�1(x) is a m-dimensional sphere of radius ⇢(x) forevery x 2 ⌦�; This can be seen by inspecting the definition of F , or it can be deducedfrom the formula

F�1(x) = {(⇡(x) + sr⇢(x), y) : s2 + |y|2 = ⇢(x)2}, (A.6)

which is not hard to check.Step 4. It su�ces to prove (A.4) for Ei that @Ei \ ⌦i is smooth, since if we know

(A.4) for smooth sets, then the general case follows by a rather standard approximationargument (which the weights do not greatly complicate). It is convenient to write

E+i = Ei, E�

i = ⌦i \ Ei, E±i = F�1(E±

i ) ⇢ F�1(⌦i) =: ⌦i.

Note that @Ei is smooth in ⌦i away from @⌦i ⇥ {0}, and that @Ei = F�1(@Ei). Also,⌦i has smooth boundary, so the relative isoperimetric inequality guarantees that there issome C = C(⌦i) < 1 such that

min±

Hn+m(E±i ) C

h

Hn+m�1(@Ei \ ⌦i)i

n+mn+m�1

. (A.7)

We will show below that, if � is small enough, then

JF 2 in ⌦� , (A.8)

where JF denotes the Jacobian of F ; and that, writing F@ := F |@E1

2 Jn�1F@ in @E, (A.9)

where Jn�1F@ denotes the (n � 1)-dimensional Jacobian, whose relevant properties arerecalled below. We will first use these two facts to give the proof of (A.4), and then wegive the details of (A.8) and (A.9).

With this notation, we use the coarea formula to computeZ

E+i

⇢m(A.2)= min

±

Z

E±i

⇢m(x)dx(A.6)= c(m)min

±

Z

E±i

Hm(F�1(x))dx

= c(m)min±

Z

E±i

JF (x, y) dx dy(A.8) Cmin

±Hn+m(E±

i ),

Similarly, using (more sophisticated) the coarea formula from Federer [20], section 3.2.22,

Hn+m�1(@Ei \ ⌦�)(A.9) C

Z

@Ei\⌦�

JF@ dHn�1 = C

Z

@Ei\⌦⇢m(x)dHn�1.

28 P. ATHAVALE, R. L. JERRARD, M. NOVAGA, AND G. ORLANDI

Combining these two inequalities with (A.7), we deduce that (A.4) holds.Step 5. To complete the proof of the proposition, we verify (A.8) and (A.9). We fix

(x, y) 2 ⌦� and write M := DF (x, y). Then a computation shows that M = M0 + M1

where M0,M1 are n⇥ (n+m) matrices defined (in block form) by

M0 = (Idn + (⇢

⇢� 1)r⇢⌦r⇢,r⇢⌦ y

⇢), M1 = (⇢� ⇢)(r2⇢, 0).

This means that for V = (v, w) 2 Rn ⇥ Rm (understood as a column vector),

M0V = v +

(⇢

⇢� 1)(v ·r⇢) + w · y

r⇢. (A.10)

One easily checks from the definitions that that |⇢� ⇢| � in ⌦�, so

|M1V | � kr2⇢k1 |v| = C�|V |, (A.11)

It also follows directly that

kerM0 = {(�r⇢(x), w) : �⇢(x) + w · y = 0},and can also check from (A.10) that

M0 is a linear isometry of (kerM0)? onto Rn. (A.12)

Indeed, if we fix an orthonormal basis {v1, . . . , vn} of Rn such that v1 = r⇢(x), then thevectors defined by

V1 =1

⇢(x, y)(⇢(x)r⇢(x), y), Vi = (vi, 0) for i = 2, . . . , n,

form an orthonormal basis for (kerM0)? that satisfies M0Vi = vi for all i. We now letVn+1, . . . , Vn+m be an orthonormal basis for kerM0, so that {Vi}n+m

i=1 is an orthonormalbasis for Rn+m. The Jacobian is

JF (x, y) =

0

@

X

1i1<...<inn+m

|MVi1 ^ . . . ^MVin |21

A

1/2

(A.11)=

0

@

X

1i1<...<inn+m

|M0Vi1 ^ . . . ^M0Vin |21

A

1/2

+O(�) = 1 +O(�).

Thus (A.8) follows by a suitable choice of �. To verify (A.9), we compute Jn�1F@(x, y)for some aribitrary (x, y) 2 @E. By definition,

Jn�1F@(x, y) =

0

@

X

1i1<...<in�1n+m�1

|MVi1 ^ . . . ^MVin�1 |21

A

1/2

where {Vi}n+m�1i=1 form an orthonormal basis for T(x,y)@E. Rn+m which must contain

kerM0(x, y), since this is just the tangent space at (x, y) to F�1(F (x, y)) ⇢ @E. ThusT(x,y)@E has the form

T(x,y)@E = kerM0 � L, for some n� 1-dim. subspace L ⇢ kerM?0 ,

WEIGHTED TV MINIMIZATION AND APPLICATIONS TO VORTEX DENSITY MODELS 29

And according to (A.12), M0 is a linear isometry of L onto its image. So we can choosea basis {Vi}n+m�1

i=1 such that M0Vi = 0 for i � n and {M0Vi}n�1i=1 form an orthonormal

set. Then as above we find that Jn�1F@(x, y) = 1 +O(�), and we conclude by a suitablechoice of �.

References

[1] A. Aftalion, Vortices in Bose-Einstein condensates. Progress in Nonlinear Di↵erential Equations andtheir Applications, 67. Birkhauser, Boston, MA, 2006.

[2] A. Aftalion, R. L. Jerrard, Shape of vortices for a rotating Bose-Einstein condensate. Physical ReviewA, 66(2):023611/1–023611/7, 2002.

[3] A. Aftalion and T. Riviere, Vortex energy and vortex bending for a rotating Bose-Einstein condensate,Phys. Rev. A 64:043611, 2001.

[4] F. Alter, V. Caselles, A. Chambolle, A characterization of convex calibrables sets in RN , Math.Annalen, 332:329–366, 2005.

[5] W. K. Allard, On the first variation of a varifold, Ann. of Math. (2) 95:417?491, 1972.[6] M. Amar, G. Bellettini, A notion of total variation depending on a metric with discontinuous coe�-

cients, Annales de l’institut Henri Poincare(C) Analyse non lineaire, 11:91–133, 1994.[7] L. Ambrosio, N. Fusco, D. Pallara, Functions of Bounded Variation and free discontinuity problems.

Oxford University Press, 2000.[8] G. Anzellotti, Pairing between measures and bounded functions and compensated compactness. Ann.

Mat. Pura Appl., 135:293–318, 1983.[9] G. Aronsson, Extension of functions satisfying Lipschitz conditions. Ark. Mat., 6:551–561, 1967.

[10] S. Baldo, R. L. Jerrard, G. Orlandi, H. M. Soner, Convergence of Ginzburg-Landau functionals in3-d superconductivity. Archive Rat. Mech. Analysis, 2015(3):699–752, 2012.

[11] S. Baldo, R. L. Jerrard, G. Orlandi, H. M. Soner, Vortex density models for superconductivity andsuperfluidity. Comm. Math. Phys., 318(1):131–171, 2013

[12] E. N. Barron, L. C. Evans, R. Jensen, The infinity Laplacian, Aronsson’s equation and their gener-alizations. Trans. Amer. Math. Soc., 360(1):77–101, 2008.

[13] A. Briani, A. Chambolle, M. Novaga, G. Orlandi, On the gradient flow of a one-homogeneous func-tional. Confluentes Mathematici, 3(4):617–635, 2011.

[14] G. Carlier, M. Comte, On a weighted total variation minimization problem. J. Funct. Anal.,250(1):214–226, 2007.

[15] V. Caselles, A. Chambolle, M. Novaga, Total Variation in imaging. In Handbook of MathematicalMethods in Imaging, Springer, 1016–1057, 2011.

[16] A. Chambolle, V. Caselles, D. Cremers, M. Novaga and T. Pock, An introduction to Total Variationfor Image Analysis. In Theoretical Foundations and Numerical Methods for Sparse Recovery, DeGruyter, Radon Series Comp. Appl. Math., 9:263–340, 2010.

[17] V. Caselles, G. Facciolo and E. Meinhardt, Anisotropic Cheeger sets and applications. SIAM J.Imaging Sci., 2(4):1211–1254, 2009.

[18] G. De Philippis, G. Franzina and A. Pratelli, Existence of isoperimetric sets with densities “convergingfrom below” on RN . J. Geometric Anal., to appear.

[19] I. Ekeland, R. Temam, Convex analysis and variational problems. SIAM, 1999.[20] H. Federer, Geometric Measure Theory Springer, New York, 1969.[21] Robert L. Jerrard, Local minimizers with vortex filaments for a Gross-Pitaevsky functional. ESAIM

Control Optim. Calc. Var. 13(1): 35-71, 2007.[22] I. Ionescu, T. Lachand-Robert Generalized Cheeger sets related to landslides. Calc. Var. Partial

Di↵erential Equations, 23(2):227–249, 2005.[23] F. Maggi, Sets of finite perimeter and geometric variational problems. An introduction to geometric

measure theory. Cambridge Studies in Advanced Mathematics, 135. Cambridge University Press,Cambridge, 2012.

30 P. ATHAVALE, R. L. JERRARD, M. NOVAGA, AND G. ORLANDI

[24] J. A. Montero, Hodge decomposition with degenerate weights and the Gross-Pitaevskii energy. J.Funct. Anal. 254(7):926–1973, 2008.

[25] A. Montero, B. Stephens, On the geometry of Gross-Pitaevskii vortex curves for generic data. Pro-ceedigs of the A.M.S., 141(11):3871-3881, 2012.

[26] Y. Peres, O. Schramm, S. She�eld, D. Wilson, Tug-of-war and the infinity Laplacian. J. Amer. Math.Soc., 22(1):167–210, 2009.

[27] L. Rudin, S. J. Osher and E. Fatemi, Nonlinear total variation based noise removal algorithms.Physica D, 60:259–268, 1992.

[28] M. Zhu and T. Chan, An e�cient primal-dual hybrid gradient algorithm for total variation imagerestoration. CAM Report 08-34, Mathematics Department, UCLA, 2008.

Department of Mathematics, University of Toronto, Toronto, Ontario, CanadaE-mail address: [email protected], [email protected]

Dipartimento di Matematica, Universita di Pisa, Pisa, ItalyE-mail address: [email protected]

Dipartimento di Informatica, Universita di Verona, Verona, ItalyE-mail address: [email protected]