Home On the area-preserving Willmore flow of small bubbles sliding on a domain’s boundary
Article
Licensed
Unlicensed Requires Authentication

On the area-preserving Willmore flow of small bubbles sliding on a domain’s boundary

  • Jan-Henrik Metsch ORCID logo EMAIL logo
Published/Copyright: February 20, 2024

Abstract

We consider the area-preserving Willmore evolution of surfaces ϕ that are close to a half-sphere with a small radius, sliding on the boundary S of a domain Ω while meeting it orthogonally. We prove that the flow exists for all times and keeps a “half-spherical” shape. Additionally, we investigate the asymptotic behavior of the flow and prove that for large times the barycenter of the surfaces approximately follows an explicit ordinary differential equation. Imposing additional conditions on the mean curvature of S, we then establish convergence of the flow.

MSC 2020: 53E40; 35G31; 47J07

Communicated by Guofang Wang


A Expansion

For the duration of this section, we will write u ( s ) := u ^ t ( s ) , p := ξ ( t ) , g ~ ( s ) := g ~ p , s , and d μ [ u , g ~ ] := d μ f u * g ~ . We consider the map

z ( s ) := I i [ u ( s ) , g ~ ( s ) ] = 𝕊 + 2 W [ u ( s ) , g ~ ( s ) ] P H [ u ( s ) , g ~ ( s ) ] C i [ u ( s ) , g ~ ( s ) ] 𝑑 μ [ u ( s ) , g ~ ( s ) ]

and prove that z ( 0 ) = z ( 0 ) = 0 and z ′′ ( 0 ) = - 3 i H S ( p ) . This implies the identities for y t , i ( s ) = - s z ( s ) .

First, we note that, as W [ 0 , δ ] = 0 , H [ 0 , δ ] = 2 and C i [ 0 , δ ] = - 3 2 π ω i , we may conclude z ( 0 ) = 0 and

z ( 0 ) = - 3 2 π 𝕊 + 2 ( D 1 W [ 0 , δ ] u s | s = 0 + D 2 W [ 0 , δ ] g ~ s | s = 0 ) ω i d μ 𝕊 2 .

As u ( 0 ) is even,

D 1 W [ 0 , δ ] u ( 0 ) = - Δ ( Δ + 2 ) u ( 0 )

is also even and the first integral vanishes. For the second integral, we recall Lemma 6.1 to obtain that D 2 W [ 0 , δ ] g ~ ( 0 ) is an even function and we deduce that the second integral also vanishes. It remains to compute z ′′ ( 0 ) . Again using W [ 0 , δ ] = 0 , we get

(A.1)

z ′′ ( 0 ) = d 2 d s 2 | s = 0 𝕊 + 2 W [ s u ( 0 ) , δ + s g ~ ( 0 ) ] ( P H C i ) [ s u ( 0 ) , δ + s g ~ ( 0 ) ] d μ [ s u ( 0 ) , δ + s g ~ ( 0 ) ]
+ 𝕊 + 2 ( D 1 W [ 0 , δ ] u ′′ ( 0 ) + D 2 W [ 0 , δ ] g ~ ′′ ( 0 ) ) ( P H C i ) [ 0 , δ ] 𝑑 μ 𝕊 2 .

We claim that the first line vanishes. First, note that by Lemma 6.1 the integrand in the first line is an odd function. We now need to check that the derivative of the measure at s = 0 is even[1]. For this we use the following two formulas from [2, Lemma 1[2] and the proof of Lemma 7]:

D 1 d μ [ 0 , δ ] φ = 2 φ d μ 𝕊 2 ,
D 2 d μ [ 0 , δ ] q = 1 2 tr 𝕊 2 q .

Inserting g ~ ( 0 ) from the proof of Lemma 6.1, we see that

D 1 d μ [ 0 , δ ] u ( 0 ) = 2 u ( 0 ) ,
D 2 d μ [ 0 , δ ] g ~ ( 0 ) = tr 3 g ~ ( 0 ) - g ( 0 ) ( ω , ω ) = - 2 h a b ( p ) ω a ω b ω 3

are both even. Thus the first line in equation (A.1) vanishes and we must only compute the last line. By definition,

W [ u , g ~ ] = 1 2 ( Δ g H + | h 0 | 2 H + Ric g ~ ( ν ~ , ν ~ ) H ) .

As g ~ ( s ) is the pullback of a flat metric, we may drop the last term. Next, we note that h 0 [ 0 , δ ] = 0 , and hence D 1 | h 0 | 2 [ 0 , δ ] = 0 and similarly D 2 | h 0 | 2 [ 0 , δ ] = 0 . Using H [ 0 , δ ] = 2 , we get

( D 1 Δ [ 0 , δ ] φ ) H [ 0 , δ ] = ( D 2 Δ [ 0 , δ ] q ) H [ 0 , δ ] = 0

for arbitrary φ, q. Combining these considerations, we deduce

(A.2) z ′′ ( 0 ) = 1 2 𝕊 + 2 Δ ( D 1 H [ 0 , δ ] u ′′ ( 0 ) + D 2 H [ 0 , δ ] g ~ ′′ ( 0 ) ] ) - 3 2 π ω i d μ 𝕊 2 .

Using [2, Lemma 7] and the standard result for the variation of H, we get

(A.3) D 2 H [ 0 , δ ] φ = - ( Δ + | h 0 | 2 ) φ = - ( Δ + 2 ) φ ,
(A.4) D 2 H [ 0 , δ ] q = - 1 2 tr 𝕊 2 ν ~ q + tr 𝕊 2 q ( ν ~ , ) + q ( ν ~ , ν ~ ) - tr 𝕊 2 q .

We apply (A.4) to q = g ~ ′′ ( 0 ) . Denoting the second fundamental form of S in the chart f [ p , ] by h i j , we can use formula (2.2) to get ( * -entries are to be inferred from symmetry)

g ~ ′′ ( 0 ) = [ 2 h 1 a h 1 b x a x b 2 h 1 a h 2 b x a x b 1 h a b x a x b * 2 h 2 a h 2 b x a x b 2 h a b x a x b * * 0 ] .

Inserting into (A.4) and repeatedly using tr 𝕊 2 A = tr 3 A - A ( ω , ω ) gives

(A.5) 𝕊 + 2 Δ D 2 H [ 0 , δ ] g ~ ′′ ( 0 ) ω i 𝑑 μ 𝕊 2 = 𝕊 + 2 Δ ( 6 a h k j ω k ω j ω a ω 3 - 2 j h j a ω a ω 3 ) ω i 𝑑 μ 𝕊 2 = - 3 π i H ( p ) .

Next, we must consider

(A.6) 𝕊 + 2 Δ D 1 H [ 0 , δ ] u ′′ ( 0 ) 𝑑 μ 𝕊 2 = - 𝕊 + 2 Δ ( Δ + 2 ) u ′′ ( 0 ) ω i 𝑑 μ 𝕊 2 = 𝕊 + 2 Δ u ′′ ( 0 ) η ω i 𝑑 μ 𝕊 2 .

To evaluate this integral, we must exploit the boundary condition B 0 [ u ( s ) , g ~ ( s ) ] = 0 and differentiate twice.

Lemma A.1.

The following identities hold:

D 1 B [ 0 , δ ] φ = ( φ η , - η ( Δ + 2 ) φ ) ,
D 2 B [ 0 , δ ] g ~ ′′ ( 0 ) = ( a h b c ω a ω b ω c , 10 a h b c ω a ω b ω c - 2 a h a b ω b ) .

Denote arbitrary even functions from 𝕊 + 2 to by E. Once Lemma A.1 is proven we may combine it with Lemma 6.1 to get

(A.7)

0 = D 1 B 0 [ 0 , δ ] u ′′ ( 0 ) + D 2 B 0 [ 0 , δ ] g ′′ ( 0 ) + d 2 d s 2 B 0 [ s u ( 0 ) , δ + s g ~ ( 0 ) ]
= ( u ′′ ( 0 ) η + a h b c ω a ω b ω c + E , - η ( Δ + 2 ) u ′′ ( 0 ) + 10 a h b c ω a ω b ω c - 2 a h a b ω b + E ) .

Combining both components of equation (A.7) gives

Δ u ′′ ( 0 ) η = 12 a h b c ω a ω b ω c - 2 a h a b ω b + E .

Inserting into equation (A.6) and combining with equation (A.5) yields

1 2 𝕊 + 2 Δ ( D 1 H [ 0 , δ ] u ′′ ( 0 ) + D 2 H [ 0 , δ ] g ~ ′′ ( 0 ) ) 𝑑 μ 𝕊 2 = 2 π i H S ( p ) .

Multiplying with - 3 2 π and recalling equation (A.2) shows z ′′ ( 0 ) = - 3 i H S ( p ) . It remains to prove Lemma A.1.

Proof.

Let ω 0 𝕊 + 2 and let ϕ : U 2 ϕ ( U ) 𝕊 + 2 be a parameterization near ω 0 so that g i j := i ϕ , j ϕ satisfies g i j ( ω 0 ) = δ i j and k g i j ( ω 0 ) = 0 . Put f ϵ ( ω ) := ( 1 + ϵ φ ( ω ) ) ω , 𝔤 ( μ ) := δ + μ q with q := g ~ ′′ ( 0 ) , and let ν ~ ( ϵ , μ ) denote the inner normal of f ϵ with respect to 𝔤 ( μ ) (so ν ~ ( 0 , 0 ) = - ω ). Then, at ω 0 ,

(A.8) d d ϵ | ϵ , μ = 0 ν ~ = i φ i ϕ and d d μ | ϵ , μ = 0 ν ~ = q ( ϕ , i ϕ ) i ϕ + 1 2 q ( ϕ , ϕ ) ϕ .

The second formula is proven in [2, Lemma 7]. To get the first one, first note that ν ~ ( ϵ , 0 ) , ν ~ ( ϵ , 0 ) = 1 implies that ϵ ν ~ ( 0 , 0 ) must be tangential. Hence,

ν ~ ϵ | ϵ , μ = 0 = i ϕ , ν ~ ϵ | ϵ , μ = 0 i ϕ = - ν ~ ( 0 , 0 ) , ϵ | ϵ = μ = 0 i f ϵ i ϕ = i φ i ϕ ,

where we used that ν ~ ( 0 , 0 ) = - ϕ . For g ~ = δ , we have ν ~ 2 = e 3 and h ~ 2 = 0 . Using equations (A.3) and (A.8), we get

D 1 B 0 [ 0 , δ ] φ = ( ν ~ ϵ | ϵ , μ = 0 , e 3 , η D 1 H [ 0 , δ ] φ ) = ( ω 3 , φ , - ( Δ + 2 ) φ η )

at ω 0 . As ω 3 = e 3 = η , the first identity follows.

To establish the second formula, we use [2, Lemma 3], H [ 0 , δ ] = 2 , ν ~ ( 0 , 0 ) = - ϕ and note h ~ i j 2 0 for μ = 0 to get

(A.9) D 2 B [ 0 , δ ] 𝔤 ( 0 ) = ( d d μ | μ = 0 ν ~ ( 0 , μ ) , e 3 𝔤 ( μ ) 33 , η D 2 H [ 0 , δ ] 𝔤 ( 0 ) + 2 d h ~ 2 μ | μ = 0 ( ϕ , ϕ ) ) .

Moreover, ν ~ ( 0 , 0 ) = - ω e 3 and 𝔤 ( 0 ) = q . Using equation (A.8) then gives

(A.10) d d μ | μ = 0 ν ~ ( 0 , μ ) , e 3 𝔤 ( μ ) 33 = q ( ϕ , i ϕ ) i ϕ 3 = q μ ν ω μ ω ν , ω 3 = q μ 3 ω μ

at ω 0 . As q ( t ω ) = t 2 q ( ω ) , we obtain ν ~ q = - 2 q . Using equation (A.4) then gives

D 2 H [ 0 , δ ] q = 3 q ( ν ~ , ν ~ ) + tr 3 q ( ν ~ , ) .

Inserting ν ~ = - ω , we may write

tr 3 q ( x , ) = μ q μ ν x ν and q ( ν ~ , ν ~ ) = q μ ν x μ x ν .

As we take η , we only need terms that contain precisely one x 3 . Hence at ω 0 ,

(A.11) D 2 ( H η ~ ) [ δ ] q = η ( 6 a h i j ω i ω j ω a ω 3 - 2 j h j a ω a ω 3 ) = 6 a h i j ω i ω j ω a - 2 j h j a ω a .

Finally, we must linearize

h ~ i j 2 = 𝔤 α β Γ i j α ν ~ 2 β .

As Γ i j α vanishes for μ = 0 , we must only take the derivatives of the Christoffel symbols into account. Using 3 q = 0 , an easy computation then shows

(A.12) D 2 ( H h ~ 2 ( ν ~ , ν ~ ) ) q = 2 ω i ω j D Γ i j 3 [ δ ] q = ( 2 i q 3 j - 3 q i j ) ω i ω j = 4 a h i j ω i ω j ω a .

Equations (A.9)–(A.12) imply the second formula in the lemma. ∎

B Solution of the elliptic problem

Alessandroni and Kuwert [2] study the elliptic problem (2.9) for ϕ 𝒮 ( λ , θ ) by making the ansatz

ϕ p , λ = F λ [ p , f u ] .

For given p S , they first derive a solution to the elliptic problem with prescribed barycenter C [ ϕ ] = p . This is achieved by applying the diffeomorphism F λ [ p , ] and studying the equation in 3 with the background metric g ~ = g ~ p , λ :

{ P K [ u , g ~ ] W [ u , g ~ ] = 0 , B 0 [ u , g ~ ] = 0 , C [ u , g ~ ] = 0 , A [ u , g ~ ] = 2 π .

It is shown[3] that the unique solution u ( λ ) C 4 , γ ( 𝕊 + 2 ) depends regularly on λ. In fact, it is shown that, for Ω C m and l = m - 1 6 , we have g ~ p , λ G 0 l ( δ ; σ 0 ) for λ λ 0 ( Ω , σ 0 ) and that u C l - 4 ( [ 0 , λ 1 ) , C 4 , γ ( 𝕊 + 2 ) ) for some small λ 1 ( Ω ) > 0 . Putting φ := u ( 0 ) and using Lemma 6.1, we see that

{ Δ ( Δ + 2 ) φ = even , φ η , Δ φ η = even , - 3 2 π 𝕊 + 2 φ ω i = 0 , 2 𝕊 + 2 φ = - d d λ | λ = 0 A [ 0 , g ~ p , λ ] .

The uniqueness of this linear problem implies that φ is even, and hence u - C 4 , γ C λ 2 . Thus ϕ p , λ 𝒮 - ( λ , K ) for suitable K. Finally, in [2, Section 3], a suitable choice for p is derived that makes ϕ p , λ a solution of (2.9).

C Parabolic Schauder theory and regularity of meta maps

The following definitions are taken from [7]. Let Ω n be a bounded domain, let T > 0 and let Ω T := [ 0 , T ] × Ω . For ( l , α ) 0 × 0 n , we abbreviate D l , α u ( t , x ) := t l D x α u ( t , x ) . For m 0 , let

C m ( Ω ¯ T ) := { u : Ω ¯ T D l , α u  is defined in  int Ω T  for  4 l + | α | m  and continuous on  Ω ¯ T } ,
u C m ( Ω ¯ T ) := 4 l + | α | m D l , α C 0 ( Ω ¯ T ) .

For u C m ( Ω ¯ T ) and γ ( 0 , 1 ) , we define the temporal and spatial-Hölder seminorms

[ u ] γ s := sup 4 l + | α | = m sup ( x , t ) ( y , t ) | D l , α u ( x , t ) - D l , α u ( y , t ) | | x - y | γ ,
[ u ] γ t := sup 0 < m + γ - 4 l - | α | 4 sup ( x , t ) ( x , s ) | D l , α u ( x , t ) - D l , α u ( x , s ) | | s - t | m + γ - 4 l - | α | 4 ,

and put

u C m , γ ( Ω ¯ T ) = u C m ( Ω ¯ T ) + [ u ] γ s + [ u ] γ t .

Then the parabolic Hölder space C m , γ ( Ω ¯ T ) is defined by

C m , γ ( Ω ¯ T ) := { u C m ( Ω ¯ T ) u C m , γ ( Ω ¯ T ) < } .

We also refer to this space as C m , [ m / 4 ] - , γ , where [ ] - denotes the floor function. The definition of boundary spaces such as C 3 , 0 , γ ( [ 0 , T ] × Ω ) is analogue and also covered in [7]. Lifting the definitions onto a manifold M works as usual and is, e.g., covered in [7] for the case when M = U for a sufficiently regular domain U.

Improved Schauder estimates.

All Schauder theory except for the decay estimate used, e.g., in Theorem 3.3 are standard and may be derived by following Simon’s scaling argument [17]. To prove the decay, consider a function φ X T satisfying φ ( t , ) X 0 (recall equation (6.5)) for all t [ 0 , T ] and

(C.1) { φ ˙ + 1 2 Δ ( Δ + 2 ) φ = 0 on  [ 0 , T ] × 𝕊 + 2 , φ ( 0 , ) = ψ 0 on  { 0 } × 𝕊 + 2 .

Following [2, Lemma 4], we see that, for all u 0 X 0 , we have

𝕊 + 2 u 0 Δ ( Δ + 2 ) u 0 𝑑 μ 𝕊 2 λ 2 ( λ 2 - 2 ) 𝕊 + 2 u 0 2 = 24 𝕊 + 2 u 0 2 𝑑 μ 𝕊 2 ,

where λ 2 = 6 is the second non-zero eigenvalue of - Δ . Let μ := 12 and φ μ ( t , ω ) := e μ t φ ( t , ω ) . Then φ μ has zero boundary conditions along 𝕊 + 2 and solves

{ φ ˙ μ + 1 2 Δ ( Δ + 2 ) φ μ - μ φ μ = 0 on  [ 0 , T ] × 𝕊 + 2 , φ μ ( 0 , ) = ψ 0 on  { 0 } × 𝕊 + 2 .

Standard Schauder theory then gives the a priori estimate

(C.2) φ μ C T 4 , 1 , γ C ( 𝕊 + 2 , γ ) ( ψ 0 C 4 , γ + sup t [ 0 , T ] φ μ ( t ) L 2 ( 𝕊 + 2 ) ) .

Using the fact that φ ( t , ) X 0 for all t [ 0 , T ] and remembering (C.1), we compute

d d t 1 2 𝕊 + 2 φ μ 2 𝑑 μ 𝕊 2 = - 1 2 𝕊 + 2 φ μ Δ ( Δ + 2 ) φ μ 𝑑 μ 𝕊 2 + μ 𝕊 + 2 φ μ 2 𝑑 μ 𝕊 2 ( 6 ( 2 - 6 ) 2 + μ ) 𝕊 + 2 φ μ 2 𝑑 μ 𝕊 2 = 0 .

This implies

sup t [ 0 , T ] φ μ ( t , ) L 2 ( 𝕊 + 2 ) 2 = ψ 0 L 2 ( 𝕊 + 2 ) 2 ,

and therefore

e μ T φ ( T , ) C 4 , γ = φ μ ( T , ) C 4 , γ φ μ C T 4 , 1 , γ C ( 𝕊 + 2 , γ ) ψ 0 C 4 , γ .

Recalling μ = 12 , we obtain

φ ( T ) C 4 , γ C ( 𝕊 + 2 , γ ) e - 12 T ψ 0 C 4 , γ .

Regularity of meta maps.

Let U be a bounded domain and let V be a bounded convex domain. Put

A := C 0 , γ 4 ( [ 0 , T ] , C l ( V , ) ) , B := C 0 , 0 , γ ( [ 0 , T ] × U , V ) , C := C 0 , 0 , γ ( [ 0 , T ] × U ) .

For g A and f B , put ( g f ) ( t , x ) := g ( t , f ( t , x ) ) and consider the meta map

T : A × B C , ( g , f ) g f .

It is easy to see that T is well defined if l 1 . Moreover, T is even continuous if l 2 . As an example, we show that [ g f - g f ~ ] γ t is small if f f ~ . Indeed,

Δ := | g ( t , f ( t , x ) ) - g ( t , f ~ ( t , x ) ) - g ( s , f ( s , x ) ) + g ( s , f ~ ( s , x ) ) |
| g ( t , f ( t , x ) ) - g ( t , f ~ ( t , x ) ) - g ( t , f ( s , x ) ) + g ( t , f ~ ( s , x ) ) |
+ | g ( t , f ( s , x ) ) - g ( t , f ~ ( s , x ) ) - g ( s , f ( s , x ) ) + g ( s , f ~ ( s , x ) ) |
| g ( t , λ f ( t , x ) + ( 1 - λ ) f ~ ( t , x ) ) - g ( t , λ f ( s , x ) + ( 1 - λ ) f ~ ( s , x ) ) | λ = 0 λ = 1 |
+ | g ( t , λ f ( s , x ) + ( 1 - λ ) f ~ ( s , x ) ) - g ( s , λ f ( s , x ) + ( 1 - λ ) f ~ ( s , x ) ) | λ = 0 λ = 1 | .

The convexity of V and the chain rule then readily imply

Δ ( D 2 g C 0 [ f - f ~ ] γ t + D 2 2 g C 0 f - f ~ C 0 ( [ f ] γ t + [ f ~ ] γ t ) + [ D 2 g ] γ t f - f ~ C 0 ) | t - s | γ 4 ,

which shows the continuity of T. Similarly, one shows that for T C k it suffices if l 2 + k , as differentiating g f with respect to f produces D g f , which must be continuous in f. Similar arguments allow one to study g f C 4 , 1 , γ ( [ 0 , T ] × U ) for f C 4 , 1 , γ ( [ 0 , T ] × U , V ) . The corresponding meta map is of class C k as long as l 6 + k , as four additional derivatives of g are required.

D The Riemannian barycenter

Let g ~ G 0 l ( δ ; ϵ ) with l 2 . This section serves the purpose of constructing the Riemannian barycenter of the immersion f u : 𝕊 + 2 ( 3 , g ~ ) for small enough u. The analysis follows the construction in [2], but changes the euclidean projection π 2 used in [2] to the Riemannian projection π 2 [ g ~ , ] . We recall D r := { x 2 | x | < r } and Z r := D ¯ r × [ - r , r ] .

Lemma D.1.

For ϵ > 0 small enough, the following is true: For each p Z 3 / 2 , there exists a unique x D 2 , denoted by π R 2 [ g ~ , p ] , such that g ~ x ( p - x , R 2 ) = 0 . The map

π 2 : G 0 l ( δ ; ϵ ) × Z 3 2 D 2 , ( g ~ , p ) x ,

is of class C l .

Proof.

First, we prove uniqueness. Fix a point p = ( p , p 3 ) Z 3 / 2 and consider the map

Φ : D 2 ( 0 ) 2 2 , f ( q ) := p - ( δ - g ~ q ) ( p - q , e i ) e i .

For small ϵ > 0 , it is easy to show that Φ : D 2 D 2 defines a contraction which implies the existence and uniqueness. To check the regularity, we consider the C l -map

Φ ~ : Z 3 2 × D 2 × G ϵ 2 , f ~ ( p , q , g ~ ) := g ~ q ( p - q , e i ) e i .

Given any p 0 = ( p 0 , p 0 3 ) Z 3 / 2 , we note Φ ~ ( p 0 , p 0 , δ ) = 0 and D 2 Φ ~ ( p 0 , p 0 , δ ) = - id 2 . The regularity of the local and hence global inverse follows from the implicit function theorem. ∎

We briefly motivate the modified definition of the Riemannian barycenter. For an immersion f u : 𝕊 + 2 ( 3 , g ~ ) , we wish to define the Riemannian barycenter x 2 by the implicit equation

I := 𝕊 + 2 ( exp x g ~ ) - 1 ( f u ( ω ) ) 𝑑 μ g ( ω ) ! g ~ x 2 .

Note that, by definition, I T x 3 , and hence it is sensible to demand orthogonality with respect to g ~ x . We may reformulate the condition as g ~ x ( x + I - x , 2 ) = 0 , which implies π 2 [ g ~ , x + I ] = x . Hence we study zeros of the map

X : U ϵ × G ϵ × D ϵ 2 , X [ u , g ~ , x ] := π 2 [ g ~ , x + 𝕊 + 2 ( exp x g ~ ) - 1 ( f u ( ω ) ) 𝑑 μ g ( ω ) ] - x ,

which reduces to X (we include the prime to distinguish the map from [2] from the one studied here) from [2] if we replace π 2 [ g ~ , ] with π 2 [ δ , ] as is used there. In particular, both maps are identical if g ~ = δ is inserted. Repeating the analysis from [2], we get the following theorem.

Theorem D.2 (The two-dimensional barycenter).

There exist ϵ > 0 and ρ > 0 such that, for u C 1 ( S + 2 , R ) and g ~ C 2 ( Z 2 , M 3 ( R ) ) satisfying u C 1 < ϵ and g ~ - δ C 2 < ϵ , there exists a unique point x = C [ u , g ~ ] D ρ ( 0 ) R 2 such that X [ u , g ~ , C [ u , g ~ ] ] = 0 and

| x | C ( u C 1 ( 𝕊 + 2 ) + g ~ - δ C 2 ) .

As a map G 0 l × C 4 , γ ( S + 2 ) R 2 the map C is of class C l - 1 .

Proof.

As X [ , δ , ] = X [ , δ , ] , the proof from [2] carries over. ∎

As X [ , δ , ] = X [ , δ , ] , we conclude the same explicit formula for C [ u , δ ] as is given in [2]:

C [ u , δ ] = π 2 ( f u 𝑑 μ f u * δ ) .

Following [2], we now compute the L 2 -gradient of C i . The only difference in the analysis that is required is to also take the derivative of the projection into account. Let f ( ϵ ) be a variation of f u with f ˙ ( 0 ) = φ ν ~ , where ν ~ is the inner normal of f u . We set

g := f u * g ~ , x ( ϵ ) := C [ f ( ϵ ) , g ~ ] , I [ g ~ , x , f ] := 𝕊 + 2 ( exp x g ~ ) - 1 ( f ) 𝑑 μ g ,

and compute

(D.1)

0 = d d ϵ | ϵ = 0 π 2 [ g ~ , x + 𝕊 + 2 ( exp x g ~ ) - 1 ( f ( ϵ ) ) d μ f ( ϵ ) * g ~ ] - d x d ϵ | ϵ = 0
= D 2 π 2 [ g ~ , x + I [ g ~ , x , f u ] ] d d ϵ | ϵ = 0 ( x + 𝕊 + 2 ( exp x g ~ ) - 1 ( f ( ϵ ) ) d μ f ( ϵ ) * g ~ ) - d x d ϵ | ϵ = 0 .

We investigate the derivative:

(D.2)

d d ϵ | ϵ = 0 𝕊 + 2 ( exp x g ~ ) - 1 ( f ( ϵ ) ) d μ f ( ϵ ) * g ~ = 1 μ g ( 𝕊 + 2 ) 2 𝕊 + 2 φ H d μ g 𝕊 + 2 ( exp x g ~ ) - 1 ( f u ( ω ) ) d μ g
+ 𝕊 + 2 D ( ( exp x g ~ ) - 1 ) ( f u ( ω ) ) φ ν ~ - ( exp x g ~ ) - 1 ( f u ( ω ) ) H φ d μ g
+ 𝕊 + 2 D x ( ( exp x g ~ ) - 1 ) ( f u ( ω ) ) d μ g d x d ϵ | ϵ = 0 .

Inserting (D.2) into (D.1) and dropping the arguments on I for spatial reasons gives

d x d ϵ | ϵ = 0 = - ( D 2 π 2 [ g ~ , x + I ] ( 𝕊 + 2 D x ( ( exp x g ~ ) - 1 ) ( f u ( ω ) ) d μ g + id 2 ) - id 2 ) - 1
× D 2 π 2 [ g ~ , x + I ] ( 𝕊 + 2 φ H d μ g 𝕊 + 2 ( exp x g ~ ) - 1 ( f u ( ω ) ) d μ g
+ 𝕊 + 2 D ( ( exp x g ~ ) - 1 ) ( f u ( ω ) ) φ ν ~ - ( exp x g ~ ) - 1 ( f u ( ω ) ) H φ d μ g ) .

As the Riemannian barycenter is invariant under reparameterization, the L 2 -gradient of C i is normal along f u . We may therefore multiply the actual L 2 -gradient with ν ~ to obtain a scalar function which we denote by C i [ u , g ~ ] . By definition, we then have

d d ϵ | ϵ = 0 C i [ f ( ϵ ) , g ~ ] = 𝕊 + 2 C i [ u , g ~ ] g ~ ( f ϵ | ϵ = 0 , ν ~ ) d μ g .

The analysis above gives the explicit formula

i = 1 2 C i [ u , g ~ ] e i = - 1 μ g ( 𝕊 + 2 ) ( D 2 π 2 [ g ~ , x + I ] ( 𝕊 + 2 D x ( ( exp x g ~ ) - 1 ) ( f u ( ω ) ) 𝑑 μ g + id 2 ) - id 2 ) - 1
× D 2 π 2 [ g ~ , x + I ] ( 𝕊 + 2 ( exp x g ~ ) - 1 ( f u ) 𝑑 μ g H + D ( ( exp x g ~ ) - 1 ) ( f u ( ω ) ) - ( exp x g ~ ) - 1 ( f u ( ω ) ) H ) .

Specializing to g ~ = δ gives

C i [ u , δ ] = - 1 μ g ( 𝕊 + 2 ) π 2 ( 𝕊 + 2 ( exp x δ ) - 1 ( f u ) 𝑑 μ g H + D ( ( exp x δ ) - 1 ) ( f u ) - ( exp x δ ) - 1 ( f u ) H ) i .

Note that

𝕊 + 2 ( exp x δ ) - 1 ( f u ) 𝑑 μ g δ 2

by the definition of x = C [ u , δ ] . This eliminates the first term, and thus establishes the same formula that is derived in [2]. In particular, we recover

C i [ 0 , δ ] = - 3 2 π ω i .

A note on regularity.

We consider C i as a map

C i : G T l × C 4 , 1 , γ ( [ 0 , T ] × 𝕊 + 2 ) C 0 , γ 4 ( [ 0 , T ] , 2 )

defined on a neighborhood of ( δ , 0 ) . We use the fact that the map ( x , v , g ~ ) exp x g ~ v 3 defined on a suitably large neighborhood of ( 0 , 0 , δ ) 2 × 3 × G 0 l is of class C l - 1 . This is proven in [2, Appendix 4]. Combined with the explicit formula for C i , we get that C i is of class C l - 7 as long as l 7 .

Application.

We now define C [ ϕ p , u λ ] by applying Theorem D.2 to f u with the pullback metric g ~ p , λ . This gives a point x = C [ u , g ~ p , λ ] 2 and we wish to define

C [ ϕ ] := F λ [ p , x ] .

We must now prove that this definition does not depend on p and the chosen frame.

Proof.

Abbreviate g ~ := g ~ p , λ . Rewriting the defining equation for x gives

x = C [ f u , g ~ ] if and only if 𝕊 + 2 ( exp x g ~ ) - 1 ( f u ( ω ) ) 𝑑 μ f u * g ~ 2  with respect to  g x ,

This, in turn, is equivalent to

(D.3) 𝕊 + 2 ( exp x g ~ ) - 1 ( f u ( ω ) ) 𝑑 μ f u * g ~ = k ν ~ 2 ( x ) for some  k .

We note that F λ [ p , ] is an isometry up to the scaling factor λ 2 in the definition of g ~ p , λ . Applying the differential ( D F λ [ p , x ] ) - 1 to (D.3), we obtain

x = C [ f u , g ~ ] if and only if ( D F λ [ p , x ] ) - 1 [ 𝕊 + 2 ( exp x g ~ ) - 1 ( f u ( ω ) ) 𝑑 μ f * g ~ ] = λ k N S ( F λ [ p , x ] ) .

This is equivalent to

𝕊 + 2 ( exp F [ p , x ] δ ) - 1 ( F λ [ p , f u ( ω ) ] ) 𝑑 μ ϕ * δ = λ k N S ( F λ [ p , x ] ) ,

which in turn is equivalent to

𝕊 + 2 ( f u ( ω ) - F [ p , x ] ) 𝑑 μ ϕ * δ = λ N S ( F λ [ p , x ] ) .

For λ small enough, we see that x = C [ f u , g ~ ] is equivalent to

(D.4) π S 𝕊 + 2 ϕ ( ω ) 𝑑 μ ϕ * δ = F λ [ p , x ] ,

where π S is the nearest point projection onto S. As the left-hand side of equation (D.4) is independent of the choices for p and the frame, the right-hand side must be too. ∎

An immediate consequence is that for ϕ = F λ [ p , f u ] we have C [ ϕ ] = p if and only if C [ f u , g ~ p , λ ] = 0 . We close by proving that we may always parameterize ϕ over its barycenter.

Theorem D.3.

There exist λ 0 > 0 , θ 0 > 0 such that for λ λ 0 each ϕ S ( λ , θ 0 ) may be parameterized over its barycenter C [ ϕ ] . That is, there exists a unique graph function u ~ C 4 , γ ( S + 2 ) such that

ϕ = F λ [ C [ ϕ ] , f u ~ ] .

Proof.

As | φ [ p , λ x ] | C λ 2 , we derive

F λ [ p , Z 2 ] = im F λ [ p , ] B 3 2 λ ( p )

for small enough λ. Next, we have the following claim.

Claim.

There exist ϵ , δ , λ > 0 such that F λ [ p , x ] im F λ [ q , ] for p , q S and x Z 2 satisfying d ( p , q ) < λ δ and | x | 1 + ϵ .

Proof of the claim. For small enough λ, we have im F λ [ q , ] B ( 3 / 2 ) λ ( q ) . Now if | x | < 1 + ϵ , then

| F λ [ p , x ] - q | | p - q | + λ | x | + | φ [ q , λ x ] | λ δ + λ ( 1 + ϵ ) + C λ 2 ,

which implies the claim for sufficiently small choices of ϵ, δ and λ.

Now suppose that ϕ 𝒮 ( ϵ , λ ) . Then we may write ϕ = F λ [ p , f u ] for some p S and u C 4 , γ < ϵ . Let q S denote the barycenter of ϕ. By definition, q = C [ ϕ ] = F λ [ p , C [ f u , g ~ p , λ ] ] and, using the estimate from Theorem D.2, we get | C [ f u , g ~ p , λ ] | < δ ( ϵ , λ ) with δ 0 as λ , ϵ 0 . This gives d ( p , q ) C λ δ ( ϵ , λ ) . As | f u ( ω ) | < 1 + ϵ , we may use the claim for small enough ϵ and λ to define u ~ by

f u ~ ( ω ) = ( F λ [ q , ] ) - 1 ( F λ [ p , f u ( ω ) ] ) .

The uniqueness is easily established. ∎

E The constraint space

Let g ~ C 5 ( Z 2 , M 3 ( ) ) be a metric that is close to the euclidean metric g ~ - δ C 5 ϵ . Also, consider a function u C 4 , γ ( 𝕊 + 2 ) that satisfies u C 4 , γ ϵ . We consider the L 2 -gradients of the area functional A [ u , g ~ ] and the barycenter components C i [ u , g ~ ] and put

(E.1) K [ u , g ~ ] := span { A [ u , g ~ ] , C i [ u , g ~ ] } .

In [2, Appendix], the following formulas are derived for g ~ = δ and u = 0 :

(E.2) A [ 0 , δ ] = - 2 and C i [ 0 , δ ] = - 3 2 π ω i .

There it is also shown that the maps ( g ~ , u ) A [ u , g ~ ] , C i [ u , g ~ ] C 0 , γ ( 𝕊 + 2 ) are of class C 2 . Hence we obtain that, for ϵ small enough, the following quantities are well defined:

ψ 0 [ u , g ~ ] := A [ u , g ~ ] A [ u , g ~ ] L 2 ( f u * g ~ ) and ψ i [ u , g ~ ] := C i [ u , g ~ ] C i [ u , g ~ ] L 2 ( f u * g ~ ) .

Clearly, ( ψ μ [ u , g ~ ] ) μ = 0 2 constitutes a generating system for K [ u , g ~ ] . Moreover, (E.2) implies that for u = 0 and g ~ = δ the functions ( ψ [ 0 , δ ] ) μ = 0 2 even provide an L 2 ( 𝕊 + 2 ) -orthonormal basis of K [ 0 , δ ] . Let

A μ ν [ u , g ~ ] := ψ μ [ u , g ~ ] , ψ ν [ u , g ~ ] L 2 ( f u * g ~ ) .

As A μ ν [ 0 , δ ] = δ μ ν , we obtain that, for ϵ small enough, ( ψ μ [ u , g ~ ] ) μ = 0 2 constitutes a basis of K [ u , g ~ ] and that the matrix A μ ν [ u , g ~ ] is invertible. As a consequence, we may write the L 2 ( f u * g ~ ) -projection of any X onto the space K [ u , g ~ ] as

(E.3) P K [ u , g ~ ] ( X ) := μ , ν = 0 2 A μ ν [ u , g ~ ] ψ μ [ u , g ~ ] , X L 2 ( f u * g ~ ) ψ ν [ u , g ~ ] .

We denote the complementary projection onto the L 2 ( f u * g ~ ) -orthogonal complement K [ u , g ~ ] by P K [ u , g ~ ] . As u and g ~ are usually clear from the context, we often drop them in the notation and we simply write P K and P K .

The metric g ~ p , λ satisfies g ~ p , λ C n - 1 C ( Ω ) λ if Ω is of class C n . Thus we obtain that, for λ λ 0 ( Ω , ϵ ) , we may always achieve g ~ - δ C 5 < ϵ .

F Computation of metric derivative

We abbreviate f := f u , g ~ ( t ) := g ~ ξ ( t ) , λ and denote the L 2 ( f * g ~ ) scalar product by , . We also put

ϕ ( t ) := F λ [ ξ ( t ) , f ( t ) ] .

We have to compute D 2 C i [ u , g ~ ] g ~ ˙ as well as D 2 A [ u , g ~ ] g ~ ˙ .

Barycenter.

We begin by investigating the barycenter. As C i [ f ( t ) , g ~ ( t ) ] = 0 , we may write

(F.1) D 2 C [ f ( t ) , g ~ ( t ) ] g ~ ˙ ( t ) = - C i [ f ( t ) , g ~ ( t ) ] , g ~ ( f ˙ ( t ) , ν ~ ) .

The geometric object that evolves in time is ϕ. Its evolution is then decomposed into two parts. The function f only represents the evolution of the ‘sphere-shape’, while the movement of the barycenter is not contained in f. However, to exploit that C [ ϕ ( t ) ] = ξ ( t ) , we must take the full evolution of ϕ into account. For that purpose, we fix t and consider small time displacements ϵ from t. We may then write

ϕ ( t + ϵ ) = F λ [ ξ ( t ) , h ( t , ϵ ) ] ,

where h ( t , 0 ) = f ( t ) . Unlike f, the quantity h encodes the full evolution of ϕ, as t is a fixed time and the dynamical variable is now ϵ. Expressing ξ ( t + ϵ ) = C [ ϕ ( t + ϵ ) ] inside the chart centered at ξ ( t ) as

ξ ( t + ϵ ) = f [ ξ ( t ) , ( ξ 1 ( t + ϵ ) , ξ 2 ( t + ϵ ) )

and differentiating at ϵ = 0 gives

(F.2)

λ - 1 ξ ˙ i ( t ) = d d ϵ | ϵ = 0 λ - 1 ξ i ( t + ϵ )
= d d ϵ | ϵ = 0 C i [ h ( t , ϵ ) , g ~ ξ ( t ) , λ ]
= C i [ f ( t ) , g ~ ( t ) ] , g ~ ( h ϵ | ϵ = 0 , ν ~ ) .

Here ν ~ is the inner normal of h. Next, we must relate ϵ h ( t , 0 ) to f ˙ ( t ) . For that purpose, we use the relation

F λ [ ξ ( t + ϵ ) , f ( t + ϵ ) ] = ϕ ( t ) = F λ [ ξ ( t ) , h ( t , ϵ ) ]

to obtain

(F.3) D 2 F λ [ ξ ( t ) , h ( t , 0 ) ] h ϵ | ϵ = 0 = d d ϵ | ϵ = 0 F λ [ ξ ( t ) , h ( t , ϵ ) ] = d d ϵ | ϵ = 0 F λ [ ξ ( t + ϵ ) , f ( t + ϵ ) ] = D 1 F λ [ ξ ( t ) , f ( t ) ] ξ ˙ ( t ) + D 2 F λ [ ξ ( t ) , f ( t ) ] f ˙ ( t ) .

Remembering h ( t , 0 ) = f ( t ) , multiplying with ( D 2 F λ [ ξ ( t ) , f ( t ) ] ) - 1 , recalling X from equation (2.22), and inserting into equations (F.1) and (F.2) yields

(F.4)

D 2 C i [ f ( t ) , g ~ ( t ) ] = - C i [ f ( t ) , g ~ ( t ) ] , g ~ ( f ˙ ( t ) , ν ~ ) (by (F.1))
= - C i [ f ( t ) , g ~ ( t ) ] , g ~ ( h ϵ | ϵ = 0 , ν ~ ) - X
= - 1 λ ξ ˙ i ( t ) + C i [ f ( t ) , g ~ ( t ) ] , X (by (F.2)) .

Area.

We copy the derivation from above. This time we use that A [ f ( t ) , g ~ ( t ) ] = 2 π . Taking the derivative then gives

(F.5) D 2 C [ f ( t ) , g ~ ( t ) ] g ~ ˙ ( t ) = - C i [ f ( t ) , g ~ ( t ) ] , g ~ ( f ˙ ( t ) , ν ~ ) .

Next, we must use that A [ ϕ ( t ) ] = 2 π λ 2 and again use h ( t , ϵ ) for small ϵ to obtain

0 = A [ f ( t ) , g ~ ( t ) ] , g ~ ( h ϵ | ϵ = 0 , ν ~ ) .

We reuse equation (F.3) to derive the analogue of equation (F.4) given by

(F.6) D 2 A [ f ( t ) , g ~ ( t ) ] g ~ ˙ = A [ f ( t ) , g ~ ( t ) ] , X .

Conclusion.

We may now substitute equations (F.4) and (F.6) into the definition of τ in equation (2.17) to get

(F.7) τ [ u , g ~ ] = - μ , ν = 0 3 A μ ν [ u , g ~ ] D 2 C μ [ u , g ~ ] g ~ ˙ C μ [ u , g ~ ] L 2 ( f u * g ~ ) C ν [ u , g ~ ξ , λ ] C ν [ u , g ~ ] L 2 ( f u * g ~ ) (by (2.17))
(F.8) = - μ , ν = 0 3 A μ ν [ u , g ~ ] C μ [ u , g ~ ] , X C μ [ u , g ~ ] L 2 ( f u * g ~ ) C ν [ u , g ~ ξ , λ ] C ν [ u , g ~ ] L 2 ( f u * g ~ ) (by (F.4), (F.6))
+ ν = 0 2 i = 1 2 A i ν [ u , g ~ ] ξ ˙ i λ C i [ u , g ~ ] L 2 ( f u * g ~ ) C ν [ u , g ~ ξ , λ ] C ν [ u , g ~ ] L 2 ( f u * g ~ )
= - P K ( X ) + ν = 0 2 i = 1 2 A i ν [ u , g ~ ] ξ ˙ i λ C i [ u , g ~ ] L 2 ( f u * g ~ ) C ν [ u , g ~ ξ , λ ] C ν [ u , g ~ ] L 2 ( f u * g ~ ) .

Acknowledgements

The author would like to thank Ernst Kuwert for the suggestion of this interesting topic and the helpful guidance as well as Marius Müller for the many helpful discussions.

References

[1] S. Agmon, A. Douglis and L. Nirenberg, Estimates near the boundary for solutions of elliptic partial differential equations satisfying general boundary conditions. I, Comm. Pure Appl. Math. 12 (1959), 623–727. 10.1002/cpa.3160120405Search in Google Scholar

[2] R. Alessandroni and E. Kuwert, Local solutions to a free boundary problem for the Willmore functional, Calc. Var. Partial Differential Equations 55 (2016), no. 2, Article ID 24. 10.1007/s00526-016-0961-3Search in Google Scholar

[3] N. D. Alikakos and A. Freire, The normalized mean curvature flow for a small bubble in a Riemannian manifold, J. Differential Geom. 64 (2003), no. 2, 247–303. 10.4310/jdg/1090426942Search in Google Scholar

[4] A. Bahri and J.-M. Coron, The scalar-curvature problem on the standard three-dimensional sphere, J. Funct. Anal. 95 (1991), no. 1, 106–172. 10.1016/0022-1236(91)90026-2Search in Google Scholar

[5] G. Bellettini and G. Fusco, Some aspects of the dynamic of V = H - H ¯ , J. Differential Equations 157 (1999), no. 1, 206–246. 10.1006/jdeq.1998.3626Search in Google Scholar

[6] S. Eberle, B. Niethammer and A. Schlichting, Gradient flow formulation and longtime behaviour of a constrained Fokker–Planck equation, Nonlinear Anal. 158 (2017), 142–167. 10.1016/j.na.2017.04.009Search in Google Scholar

[7] S. D. Ejdel’man, Parabolic equations, Partial Differential Equations VI, Springer, Berlin (1994), 203–316. 10.1007/978-3-662-09209-5_3Search in Google Scholar

[8] N. Ikoma, A. Malchiodi and A. Mondino, Area-constrained Willmore surfaces of small area in Riemannian three-manifolds: An approach via Lyapunov-Schmidt reduction, Regularity and Singularity for Partial Differential Equations with Conservation Laws, RIMS Kôkyûroku Bessatsu B63, Research Institute for Mathematical Sciences, Kyoto (2017), 31–50. Search in Google Scholar

[9] N. Ikoma, A. Malchiodi and A. Mondino, Embedded area-constrained Willmore tori of small area in Riemannian three-manifolds I: minimization, Proc. Lond. Math. Soc. (3) 115 (2017), no. 3, 502–544. 10.1112/plms.12047Search in Google Scholar

[10] N. Ikoma, A. Malchiodi and A. Mondino, Foliation by area-constrained Willmore spheres near a nondegenerate critical point of the scalar curvature, Int. Math. Res. Not. IMRN 2020 (2020), no. 19, 6539–6568. 10.1093/imrn/rny203Search in Google Scholar

[11] H. Karcher, Riemannian center of mass and mollifier smoothing, Comm. Pure Appl. Math. 30 (1977), no. 5, 509–541. 10.1002/cpa.3160300502Search in Google Scholar

[12] T. Lamm and J. Metzger, Small surfaces of Willmore type in Riemannian manifolds, Int. Math. Res. Not. IMRN 2010 (2010), no. 19, 3786–3813. 10.1093/imrn/rnq048Search in Google Scholar

[13] T. Lamm, J. Metzger and F. Schulze, Foliations of asymptotically flat manifolds by surfaces of Willmore type, Math. Ann. 350 (2011), no. 1, 1–78. 10.1007/s00208-010-0550-2Search in Google Scholar

[14] M. Mattuschka, The Willmore flow with prescribed area for a small bubble in a Riemannian manifold, PhD thesis, University of Freiburg, 2018. Search in Google Scholar

[15] A. Mondino, Some results about the existence of critical points for the Willmore functional, Math. Z. 266 (2010), no. 3, 583–622. 10.1007/s00209-009-0588-6Search in Google Scholar

[16] F. Pacard and X. Xu, Constant mean curvature spheres in Riemannian manifolds, Manuscripta Math. 128 (2009), no. 3, 275–295. 10.1007/s00229-008-0230-7Search in Google Scholar

[17] L. Simon, Schauder estimates by scaling, Calc. Var. Partial Differential Equations 5 (1997), no. 5, 391–407. 10.1007/s005260050072Search in Google Scholar

[18] G. Simonett, The Willmore flow near spheres, Differential Integral Equations 14 (2001), no. 8, 1005–1014. 10.57262/die/1356123177Search in Google Scholar

[19] A. Stahl, Regularity estimates for solutions to the mean curvature flow with a Neumann boundary condition, Calc. Var. Partial Differential Equations 4 (1996), no. 4, 385–407. 10.1007/BF01190825Search in Google Scholar

[20] R. Ye, Foliation by constant mean curvature spheres, Pacific J. Math. 147 (1991), no. 2, 381–396. 10.2140/pjm.1991.147.381Search in Google Scholar

Received: 2023-03-08
Accepted: 2023-08-23
Published Online: 2024-02-20
Published in Print: 2024-10-01

© 2024 Walter de Gruyter GmbH, Berlin/Boston

Articles in the same Issue

  1. Frontmatter
  2. Another proof of the existence of homothetic solitons of the inverse mean curvature flow
  3. A Weierstrass extremal field theory for the fractional Laplacian
  4. Minimizing movements for anisotropic and inhomogeneous mean curvature flows
  5. A singular Yamabe problem on manifolds with solid cones
  6. Uniqueness for volume-constraint local energy-minimizing sets in a half-space or a ball
  7. Limit of solutions for semilinear Hamilton–Jacobi equations with degenerate viscosity
  8. Monotonicity of entire solutions to reaction-diffusion equations involving fractional p-Laplacian
  9. Hierarchy structures in finite index CMC surfaces
  10. No breathers theorem for noncompact harmonic Ricci flows with asymptotically nonnegative Ricci curvature
  11. Wolff potentials and local behavior of solutions to elliptic problems with Orlicz growth and measure data
  12. Asymptotic analysis of single-slip crystal plasticity in the limit of vanishing thickness and rigid elasticity
  13. On the regularity of optimal potentials in control problems governed by elliptic equations
  14. Sobolev embeddings and distance functions
  15. Effective quasistatic evolution models for perfectly plastic plates with periodic microstructure: The limiting regimes
  16. On the area-preserving Willmore flow of small bubbles sliding on a domain’s boundary
  17. Sobolev contractivity of gradient flow maximal functions
  18. Discrete approximation of nonlocal-gradient energies
  19. Symmetry and monotonicity of singular solutions to p-Laplacian systems involving a first order term
  20. Flat flow solution to the mean curvature flow with volume constraint
Downloaded on 10.9.2025 from https://www.degruyterbrill.com/document/doi/10.1515/acv-2023-0023/html
Scroll to top button