Home Partial regularity for mass-minimizing currents in Hilbert spaces
Article
Licensed
Unlicensed Requires Authentication

Partial regularity for mass-minimizing currents in Hilbert spaces

  • Luigi Ambrosio EMAIL logo , Camillo De Lellis and Thomas Schmidt
Published/Copyright: June 3, 2015

Abstract

Recently, the theory of currents and the existence theory for Plateau’s problem have been extended to the case of finite-dimensional currents in infinite-dimensional manifolds or even metric spaces; see [Acta Math. 185 (2000), 1–80] (and also [Proc. Lond. Math. Soc. (3) 106 (2013), 1121–1142], [Adv. Calc. Var. 7 (2014), 227–240] for the most recent developments). In this paper, in the case when the ambient space is Hilbert, we provide the first partial regularity result, in a dense open set of the support, for n-dimensional integral currents which locally minimize the mass. Our proof follows with minor variants [Indiana Univ. Math. J. 31 (1982), 415–434], implementing Lipschitz approximation and harmonic approximation without indirect arguments and with estimates which depend only on the dimension n and not on codimension or dimension of the target space.

Award Identifier / Grant number: ERC Grant agreement GeMeThnES no. 246923

Funding statement: The research leading to these results has received funding from the European Research Council under the European Community’s Seventh Framework Programme (FP7/2007–2013): ERC Grant agreement GeMeThnES no. 246923.

A Cartesian products and homotopy construction

In the following we will be concerned with Cartesian products of currents, which arise by a slight variant of the cone construction as presented in [5, Definition 10.1] (compare also with the classical construction in [22]), and we will work in the Hilbert space × (with inner product (t1,z1),(t2,z2):=t1t2+z1,z2). In this regard we use the isometric embeddings

it:×given by it(z):=(t,z),

and for a Lipschitz function χ:× we write Dtχ for its partial derivative with respect to the first variable. Given any σ-finite Borel measure μ on this derivative exists (1μ)-a.e. in × and is bounded by Lip(χ). Consequently, for 1-a.e. t the composition (Dtχ)it is defined μ-almost-everywhere. This last observation, applied with χ=ψj and μ=T, makes the following definition well-posed:

Lemma A.1 (Cartesian products).

Consider an (n-1)-current TMn-1(H). Then the specification

(0,1×T)(φi=1ndψi):=j=1n(-1)j-101((it)T)(φDtψji=1ijndψi)d1(t)

for (φ,ψ)Lipb(R×H)×Lip(R×H)n defines a metric functional T with

(A.1)0,1×Tn(1[0,1])T.

Moreover:

  1. If T𝐍n-1() is normal, then 0,1×T𝐍n(×) is a normal n-dimensional current with

    (A.2)(0,1×T)=(i1)T-(i0)T-0,1×(T).
  2. If T is (integer-)rectifiable, then also 0,1×T is an (integer-)rectifiable current.

Proof.

Clearly, 0,1×T is linear in φ and each ψi and is thus a metric functional. To prove (A.1) we consider (φ,ψ)Lipb(×)×Lip(×)n with Lip(ψi)1 for all i{1,2,,n}. Then from the definitions of the Cartesian product, the pushforward, and the mass, combined with the Lipschitz bound Lip(ψiit)1, we readily deduce

|(0,1×T)(φi=1ndψi)|n01|φit|dTd1(t)=n[0,1]×|φ|d(1T).

By multilinearity and again by the definition of mass this implies (A.1) and in particular 𝐌(0,1×T)<.

If T is a normal current, following the proof of [5, Proposition 10.2] one can show that 0,1×T satisfies the continuity axiom and the formula (A.2). Indeed, we will not discuss the adaption of the respective arguments, as the required changes[10] are mostly notational ones. By applying (A.1) with T in place of T we obtain 𝐌(0,1×T)<; then (A.2) allows to conclude 𝐌((0,1×T))<. As the locality axiom is easily verified, we thus have 0,1×T𝐍n(×).

Finally, for Tn-1() we denote by (ST,θT,T) a corresponding triplet as in Section 2.4 with countably (n-1)-rectifiable ST, and we write T=i=1n-1Ti with Borel functions Ti:ST. As candidates for the multiplicity and the orientation of 0,1×T we consider ϑ:[0,1]×ST(0,) with ϑ(t,z):=θT(z) and τ:[0,1]×STΛn(×) with τ(t,z):=(1,0)i=1n-1(0,Ti(z)). We moreover fix (φ,ψ)Lipb(×)×Lip(×)n, and we recall that the inner product of n-vectors is given by the determinant in (2.1); by Laplace expansion of this determinant we then get

i=1nD[0,1]×STψi(t,z),τ(t,z)
  =j=1n(-1)j-1D[0,1]×STψj(t,z),(1,0)i=1ijnD[0,1]×STψi(t,z),i=1n-1(0,Ti(z))
  =j=1n(-1)j-1Dtψj(t,z)i=1ijnDST(ψiit)(z),T(z)

for all (t,z)[0,1]×ST. We multiply this equality with φ(t,z) and ϑ(t,z)=θT(z) and integrate with respect to n-1 in z. Using additionally the representation (2.9) for the (n-1)-current T, we infer

{t}×STφi=1nD[0,1]×STψi,τϑdn-1
  =j=1n(-1)j-1ST(φit)(Dtψjit)i=1ijnDST(ψiit),TθTdn-1
  =j=1n(-1)j-1((it)T)(φDtψji=1ijndψi)

for all t[0,1]. Now we integrate the resulting equality also in t, on the left-hand side we involve the coarea formula [6, Theorem 9.4] with area factor 1 on the countably n-rectifiable set [0,1]×ST, and on the right-hand side we exploit the definition of 0,1×T. We then find

[0,1]×STφi=1nD[0,1]×STψi,τϑdn=(0,1×T)(φi=1ndψi).

All in all, we have thus shown that 0,1×T is the current which is induced by the triplet ([0,1]×ST,ϑ,τ) in the sense of (2.9). By [5, Theorem 9.1] this current is rectifiable.

Finally, for TIn-1() the multiplicity θT can be chosen -valued. Consequently, the above function ϑ is -valued, and in conclusion we get 0,1×TIn(×) in this situation. ∎

Remark A.2.

The following two observations concern the rectifiability part of the preceding argument. They are scarcely relevant for our purposes, but may still be worth pointing out:

  1. For rectifiable currents T one can actually improve the estimate (A.1) – by similar arguments as above and the representation of mass in (2.11) – to the equality

    0,1×T=(1[0,1])T.

    This situation may be compared to [22, 2.10.45, 3.2.23].

  2. The rectifiability of 0,1×T can alternatively be proved by the following shorter and more elementary reasoning provided that T is normal and rectifiable: By definition T is concentrated on a countably (n-1)-rectifiable set S, and then (A.1) implies that 0,1×T is concentrated on the countably n-rectifiable set [0,1]×S. By [5, Theorem 3.9][11] this property already implies the rectifiability of 0,1×T.

We now consider the pushforward H(0,1×T) under a homotopy H:[0,1]×. We remark that for H(t,z)=tz+(1-t)z0 this pushforward becomes a cone over T with vertex z0, and we get back the original cone construction of [5, Definition 10.1] as a special case. However, in the following we will make a different choice of H, which will lead to a proof of the following lemma.

Lemma A.3 (Homotopy retraction on a graph).

Consider a current TNn-1(H) with T0, an n-plane π in H, and a Lipschitz function f:pπ(sptT)(Spanπ). If we have

(A.3)K:=supsptT|qπ-fpπ|<,

then there exists another current VNn(H) with sptV(pπ)-1(pπ(sptT)) and

(A.4)V=T-F(pπ)T,𝐌(V)nK(1+Lip(f))n-1T(Graphf).

Additionally, if T is (integer-)rectifiable, then also V can be chosen (integer-)rectifiable.

Proof.

We set

V:=H(0,1×T),

where 0,1×T is defined in Lemma A.1, and where H:[0,1]×sptT is the Lipschitz homotopy given by

H(t,z):=tz+(1-t)F(pπ(z)).

Then the image of H is contained in (pπ)-1(pπ(sptT)), and the claimed inclusion of the support of V follows at once. From the interchangeability of and H, equation (A.2), and T0 we moreover deduce that V is normal with

V=H(0,1×T)=(Hi1)T-(Hi0)T=T-F(pπ)T.

To get the mass estimate in (A.4) we first notice Lip(Hit)1+Lip(f) for t[0,1]. Moreover, we have

Lip(H(,z))=|z-F(pπ(z))|=|qπ(z)-f(pπ(z))|for all zsptT,

which implies that |Dt(χH)||qπ-fpπ| holds (1T)-a.e. on [0,1]×sptT for every fixed χLip() with Lip(χ)1. Next we employ the definitions of pushforward, product, and mass together with the preceding observations in the following estimate[12] for (φ,ψ)Lipb()×Lip()n with Lip(ψi)1 for all i{1,2,,n}:

|V(φi=1ndψi)|=|(0,1×T)((φH)i=1nd(ψiH))|
j=1n01|T((φHit)(Dt(ψjH)it)i=1ijnd(ψiHit))|d1(t)
(1+Lip(f))n-1j=1n01sptT|φHit||Dt(ψjH)it|dTd1(t)
n(1+Lip(f))n-101sptT|φHit||qπ-fpπ|dTd1(t).

As qπ-fpπ vanishes on Graphf and is elsewhere controlled by (A.3), we finally get

|V(φi=1ndψi)|nK(1+Lip(f))n-1[0,1]×(sptTGraphf)|φH|d(1T).

We have thus shown

VnK(1+Lip(f))n-1H((1[0,1])(T(sptTGraphf))),

and in particular the total mass estimate in (A.4) follows.

Finally, the remaining claim about conservation of (integer-)rectifiability follows from the fact that this property is preserved under both the product construction of Lemma A.1 and the pushforward operation. ∎

B Monotonicity formula and density results

In this appendix, we denote by C1(,) the space of Fréchet differentiable maps Φ: such that the derivative DΦ is continuous from to the space of linear maps in endowed with the operator norm. We write Id for the identity map in , and we use the notation

divTΦ(z):=traceT(z)(DΦ(z)),

where the right-hand side is defined in (2.6).

Proposition B.1 (First variation).

Let ΩH be open, let TIn(H) be locally minimizing in Ω, and consider ΦC1(H,H), with support at a positive distance from HΩ, such that DΦ is bounded. Then we have

(B.1)divTΦdT=0.

Proof.

Given L: linear and continuous, we denote by Jn(L,T) the n-dimensional Jacobian in (2.4) with the n-plane π given by T. We write Ψε:=Id+εΦ and notice that, for ε small enough, Ψε is injective; in addition, the area formula and the local minimality of T give

𝐌(T)𝐌((Ψε)T)=Jn(DΨε,T)dT.

Using (2.5) (and the rule for the derivative of the square root), we can differentiate with respect to ε to obtain (B.1). ∎

Proof of (2.17).

For zΩ and 0<η<σ<dist(z,Ω) we will show that

T(z+𝐁η)ηnT(z+𝐁σ)σn.

We can assume, without loss of generality, T(z+𝐁η)=T(z+𝐁σ)=0. Under this extra assumption, an easy approximation argument shows that (B.1) still holds for all vector fields Φ of the form Φ(w)=χ(|w-z|)(w-z), with χ:[0,)[0,) Lipschitz, with support in [0,dist(z,Ω)), whose derivative χ(t) has at most jump discontinuities at t=η and t=σ. Now we specifically insert in (B.1) the vector field

Φ(w):=[min{η-n,|w-z|-n}-σ-n]+(w-z),

whose support is the closure of z+𝐁σ. Then, the claim follows at once, when we use divTΦn[η-n-σ-n] on z+𝐁η and divTΦ-nσ-n on (z+𝐁σ)(z+𝐁η). ∎

Proposition B.2.

Consider TIn(H). Then, for T-a.e. zH there hold

(B.2)limr0r-nz+𝐁r|T-T(z)|2dT=0,
(B.3)limr0Tz,r=Θ(T,z)nSpanT(z),

where Iz,r(w):=(w-z)/r, Tz,r:=(Iz,r)T and the latter convergence is understood in duality with bounded continuous functions with bounded support.

Proof.

Let us first reduce the proof to the case when T=Fθ for some Lipschitz map F:n and θL1(n), with θ{0}n-a.e. in n and F bi-Lipschitz on θ-1(). Indeed, thanks to [5, Theorem 4.5], we can write any integer-rectifiable T with finite mass as an 𝐌-convergent series of currents Ti of this form with pairwise disjoint measure-theoretic supports Si. Since for any i it holds

T=Tin-a.e. on Si,Θ*n(jiTj,z)=0n-a.e. on Si

(for the second statement, see for instance [22, 2.10.18 (2)]), we see that both (B.2) and (B.3) for T at n-a.e. point of Si follow from (B.2) and (B.3) for Si.

So, let us assume T=Fθ for some Lipschitz map F:n and some θL1(n) such that θ{0} holds n-a.e. in n and such that F is bi-Lipschitz on θ-1(). We denote by e1,,en the canonical basis of n. Using the area formula [6, Theorem 5.1], one can check by a straightforward computation that (B.2) holds at a point z=F(x), with T(z) equal to the normalization of i=1nDF(x)ei, provided that we have Θ*n(T,z)< and that xθ-1() satisfies

(B.4)limr01θ(x+𝐁rn)x+𝐁rn|DF-DF(x)|dθ=0.

Here, DF exists n-a.e. on n by the Rademacher theorem [9, Theorem 5.11.1], (B.4) holds for θ-a.e. xn, and all in all the preceding conditions are satisfied for T-a.e. z; thus, we arrive at the claim about (B.2).

In connection with (B.3), we consider z=F(x) with xθ-1() such that F is classically Fréchet-differentiable at x with SpanT(z) equal to the image of DF(x) and with

limr0r-nx+𝐁rn[|Jn(DF)-Jn(DF(x))|+|θ-θ(x)|]dn=0.

When we set Fr(y):=(F(x+ry)-F(x))/r and θr(y):=θ(x+ry), then Fr(y) converges to DF(x)y locally uniformly in yn, on the level of the Jacobians Jn(DFr) converges to Jn(DF(x)) in Lloc1(n), and also θr converges to θ(x) in Lloc1(n). Furthermore, we have

(Iz,r)T=(Fr)θr,

and with the help of the preceding convergences we conclude

limr0φdTz,r=limr0nφ(Fr)Jn(DFr)θrdn
=θ(x)Jn(DF(x))nφ(DF(x)y)dn(y)
=θ(x)SpanT(z)φdn

for every bounded continuous function φ: with bounded support. As a side benefit of this convergence we also get Θn(T,z)=limr0Tz,r(𝐁1)/ωn=θ(x). Similar as above, the assumed conditions on z=F(x) hold true for n-a.e. xθ-1() and thus for T-a.e. z, so that we arrive at the claim regarding (B.3). ∎

References

[1] F. J. Almgren, Existence and regularity almost everywhere of solutions to elliptic variational problems among surfaces of varying topological type and singularity structure, Ann. Math. 87 (1968), 321–391. 10.2307/1970587Search in Google Scholar

[2] F. J. Almgren, Almgren’s big regularity paper. Q-valued functions minimizing Dirichlet’s integral and the regularity of area-minimizing rectifiable currents up to codimension 2, World Scientific, Singapore 2000. Search in Google Scholar

[3] L. Ambrosio, Metric space valued functions of bounded variation, Ann. Sc. Norm. Super. Pisa Cl. Sci. (4) 17 (1990), 439–478. Search in Google Scholar

[4] L. Ambrosio, N. Fusco and D. Pallara, Functions of bounded variation and free discontinuity problems, Oxford University Press, Oxford 2000. 10.1093/oso/9780198502456.001.0001Search in Google Scholar

[5] L. Ambrosio and B. Kirchheim, Currents in metric spaces, Acta Math. 185 (2000), 1–80. 10.1007/BF02392711Search in Google Scholar

[6] L. Ambrosio and B. Kirchheim, Rectifiable sets in metric and Banach spaces, Math. Ann. 318 (2000), 527–555. 10.1007/s002080000122Search in Google Scholar

[7] L. Ambrosio and T. Schmidt, Compactness results for normal currents and the Plateau problem in dual Banach spaces, Proc. Lond. Math. Soc. (3) 106 (2013), 1121–1142. 10.1112/plms/pds066Search in Google Scholar

[8] G. Anzellotti and M. Giaquinta, Convex functionals and partial regularity, Arch. Ration. Mech. Anal. 102 (1988), 243–272. 10.1007/BF00281349Search in Google Scholar

[9] V. I. Bogachev, Gaussian measures, American Mathematical Society, Providence 1998. 10.1090/surv/062Search in Google Scholar

[10] E. Bombieri, Regularity theory for almost minimal currents, Arch. Ration. Mech. Anal. 78 (1982), 99–130. 10.1007/BF00250836Search in Google Scholar

[11] A. P. Calderón and A. Zygmund, On the existence of certain singular integrals, Acta Math. 88 (1952), 85–139. 10.1007/BF02392130Search in Google Scholar

[12] E. De Giorgi, Frontiere orientate di misura minima, Seminario di Matematica, Sc. Norm. Super. Pisa, 1960–1961. Search in Google Scholar

[13] C. De Lellis and E. Spadaro, Center manifold: A case study, Discrete Contin. Dyn. Syst. 31 (2011), 1249–1272. 10.3934/dcds.2011.31.1249Search in Google Scholar

[14] C. De Lellis and E. Spadaro, Q-valued functions revisited, Mem. Am. Math. Soc. 211 (2011), Paper No. 991. 10.1090/S0065-9266-10-00607-1Search in Google Scholar

[15] C. De Lellis and E. Spadaro, Regularity of area minimizing currents I: Gradient Lp estimates, Geom. Funct. Anal. 24 (2014), 1831–1884. 10.1007/s00039-014-0306-3Search in Google Scholar

[16] C. De Lellis and E. Spadaro, Regularity of area minimizing currents II: Center manifold, preprint (2013), http://arxiv.org/abs/1306.1191. 10.4007/annals.2016.183.2.2Search in Google Scholar

[17] C. De Lellis and E. Spadaro, Regularity of area minimizing currents III: Blow-up, preprint (2013), http://arxiv.org/abs/1306.1194. 10.4007/annals.2016.183.2.3Search in Google Scholar

[18] T. De Pauw and R. Hardt, Rectifiable and flat G chains in a metric space, Am. J. Math. 134 (2012), 1–69. 10.1353/ajm.2012.0004Search in Google Scholar

[19] F. Duzaar and K. Steffen, Optimal interior and boundary regularity for almost minimizers to elliptic variational integrals, J. reine angew. Math. 546 (2002), 73–138. 10.1515/crll.2002.046Search in Google Scholar

[20] L. C. Evans, Partial differential equations, American Mathematical Society, Providence 1998. Search in Google Scholar

[21] L. C. Evans and R. F. Gariepy, Measure theory and fine properties of functions, CRC Press, Boca Raton 1992. Search in Google Scholar

[22] H. Federer, Geometric measure theory, Springer, Berlin 1969. Search in Google Scholar

[23] H. Federer, The singular sets of area minimizing rectifiable currents with codimension one and of area minimizing flat chains modulo two with arbitrary codimension, Bull. Am. Math. Soc. 76 (1970), 767–771. 10.1090/S0002-9904-1970-12542-3Search in Google Scholar

[24] D. Gilbarg and N. S. Trudinger, Elliptic partial differential equations of second order, 2nd ed., Springer, Berlin 1983. Search in Google Scholar

[25] E. Giusti, Minimal surfaces and functions of bounded variation, Birkhäuser, Basel 1984. 10.1007/978-1-4684-9486-0Search in Google Scholar

[26] E. Giusti, Direct methods in the calculus of variations, World Scientific Publishing, Singapore 2003. 10.1142/5002Search in Google Scholar

[27] R. L. Jerrard, A new proof of the rectifiable slices theorem, Ann. Sc. Norm. Super. Pisa Cl. Sci. (5) 1 (2002), 905–924. Search in Google Scholar

[28] R. L. Jerrard and H. M. Soner, Rectifiability of the distributional Jacobian for a class of functions, C. R. Math. Acad. Sci. Paris 329 (1999), 683–688. 10.1016/S0764-4442(00)88217-8Search in Google Scholar

[29] R. L. Jerrard and H. M. Soner, Functions of bounded higher variation, Indiana Univ. Math. J. 51 (2002), 645–677. 10.1512/iumj.2002.51.2229Search in Google Scholar

[30] B. Kirchheim, Rectifiable metric spaces: Local structure and regularity of the Hausdorff measure, Proc. Am. Math. Soc. 121 (1994), 113–123. 10.1090/S0002-9939-1994-1189747-7Search in Google Scholar

[31] U. Lang, Local currents in metric spaces, J. Geom. Anal. 21 (2011), 683–742. 10.1007/s12220-010-9164-xSearch in Google Scholar

[32] E. R. Reifenberg, On the analyticity of minimal surfaces, Ann. Math. (2) 80 (1964), 15–21. 10.2307/1970489Search in Google Scholar

[33] T. Schmidt, Partial regularity for degenerate variational problems and image restoration models in BV, Indiana Univ. Math. J. 63 (2014), 213–279. 10.1512/iumj.2014.63.5174Search in Google Scholar

[34] R. Schoen and L. Simon, A new proof of the regularity theorem for rectifiable currents which minimize parametric elliptic functionals, Indiana Univ. Math. J. 31 (1982), 415–434. 10.1512/iumj.1982.31.31035Search in Google Scholar

[35] J. Simons, Minimal varieties in Riemannian manifolds, Ann. Math. (2) 88 (1968), 62–105. 10.2307/1970556Search in Google Scholar

[36] I. Tamanini, Regularity results for almost minimal oriented hypersurfaces in n, Quaderni del Dipartimento di Matematica dell’Università di Lecce (1984), no. 1. Search in Google Scholar

[37] F. A. Valentine, A Lipschitz condition preserving extension for a vector function, Am. J. Math. 67 (1945), 83–93. 10.2307/2371917Search in Google Scholar

[38] S. Wenger, Isoperimetric inequalities of Euclidean type in metric spaces, Geom. Funct. Anal. 15 (2005), 534–554. 10.1007/s00039-005-0515-xSearch in Google Scholar

[39] S. Wenger, Plateau’s problem for integral currents in locally non-compact metric spaces, Adv. Calc. Var. 7 (2014), 227–240. 10.1515/acv-2012-0018Search in Google Scholar

[40] B. White, Rectifiability of flat chains, Ann. Math. (2) 150 (1999), 165–184. 10.2307/121100Search in Google Scholar

[41] B. White, The deformation theorem for flat chains, Acta Math. 183 (1999), 255–271. 10.1007/BF02392829Search in Google Scholar

Received: 2013-5-27
Revised: 2015-1-10
Published Online: 2015-6-3
Published in Print: 2018-1-1

© 2018 Walter de Gruyter GmbH, Berlin/Boston

Downloaded on 25.9.2025 from https://www.degruyterbrill.com/document/doi/10.1515/crelle-2015-0011/html
Scroll to top button