We show that every compactly supported calibrated integral current with connected \(C^{3,\alpha}\) boundary is the unique solution to the oriented Plateau problem for its boundary data. This is proved as a consequence of
the boundary regularity theory for area-minimizing currents and classical unique continuation principles adapted to the minimal surface system.
The Plateau problem asks whether a given boundary bounds a minimal surface with least area. In the early 1930’s, Douglas [1] and Radó [2] gave the positive answer for Jordan curves in \(\mathbb{R}^3\), and their work was generalized to Riemannian manifolds in 1948 by Morrey [3]. Later, Federer and Fleming [4] introduced the notion of
area-minimizing integral currents1, a well-known generalization of minimal surfaces, to formulate and address the oriented Plateau problem. Due to its broad
applicability across all dimensions and codimensions, Federer and Fleming’s theory marks a major success in establishing existence results.
A natural question is whether solutions to the oriented Plateau problem are unique. In general, the uniqueness of solutions to this problem cannot be guaranteed, even with smooth boundary data (see e.g. [5], [6]). Fortunately, uniqueness is known to be a generic property. Morgan [7] first showed that almost every curve in \(\mathbb{R}^3\) bounds a unique area-minimizing surface. In [8], [9], he extended this result to elliptic integrands in arbitrary dimensions and, in the special case of area-minimizing flat chains modulo 2, to all codimensions. These
results are confined to the Euclidean setting or require uniform convexity assumptions on the boundary, as the proofs rely on Allard’s boundary regularity theorem [10].
More recently, Caldini, Marchese, Merlo, and Steinbrüchel [11]—based on the boundary regularity theory developed by De Lellis, De Philippis, Hirsch, and
Massaccesi [12]—established the generic uniqueness of area-minimizing integral currents in arbitrary ambient manifolds, all dimensions and codimensions, and
without the assumption of convexity on the boundary. However, given a specific boundary, it remains unclear whether it uniquely bounds an area-minimizing integral current. Uniqueness can be proved only in restricted cases, such as for surfaces in \(\mathbb{R}^3\) bounded by special Jordan curves [2], [13], for hypersurfaces in
\(\mathbb{R}^n\) satisfying suitable conditions [14], [15], or for graphical
hypersurfaces over convex domains [16].
In this paper, we prove uniqueness in the oriented Plateau problem under a natural geometric condition, without imposing any restrictions on dimension and codimension. Specifically, we show that every compactly supported calibrated
integral current with connected \(C^{3,\alpha}\) boundary uniquely solves the oriented Plateau problem for its boundary data. Our result also recovers several known cases—for instance, \(C^{3,\alpha}\) graphical hypersurfaces over convex domains, since they are calibrated (see 2). As in [11], the conclusion follows from the boundary regularity theory in [12]. However, our method
differs in that it relies on unique continuation principles for elliptic operators. Unless otherwise stated, we make the following assumptions:
Assumption 1. Fix \(n \geq 3\) and \(1 \leq k \leq n-1\). Let \(W \subset \subset U \subset \mathbb{R}^n\) be open connected sets. We
will always assume that \(\Gamma\subset W\) is a \((k-1)\)-dimensional closed oriented connected embedded \(C^{3,\alpha}\) submanifold of \(\mathbb{R}^n\), that \(T \in \mathcal{I}_{k, c}(\mathbb{R}^n)\) (i.e. integral \(k\)-current in \(\mathbb{R}^n\) with compact
support) is area-minimizing in \(U\) with \(\mathop{\mathrm{spt}}T \subset W\), and that \(\partial T = [[\Gamma]]\).
We can now state our main theorem, which is a special case of 27.
Theorem 2 (Uniqueness for the Oriented Plateau Problem). Let \(U = \mathbb{R}^n\) and suppose that \(T \in \mathcal{I}_{k,c}(\mathbb{R}^n)\) and \(\Gamma\) are as in 1. Let \(\varphi\in \mathcal{E}^k(\mathbb{R}^n)\) (i.e. smooth \(k\)-form in \(\mathbb{R}^n\)) be a calibration. Assume that \(T\) is calibrated by \(\varphi\) in \(\mathbb{R}^n\). If \(T^\prime \in \mathcal{I}_{k, c}(\mathbb{R}^n)\) is globally area-minimizing with \(\partial T^\prime = [[\Gamma]]\), then \(T^\prime = T\) in \(\mathbb{R}^n\).
The idea is to derive 2 as a consequence of uniqueness in the Cauchy problem for elliptic operators. Let \(\Omega\subset \mathbb{R}^k\) be a
bounded \(C^{1,1}\) domain and let \(L : H^1(\Omega; \mathbb{R}^{n-k}) \rightarrow \mathcal{L}^2(\Omega; \mathbb{R}^{n-k})\) be the second-order principally diagonal2 elliptic partial differential operator: \[(Lu(\mathbf{x}))^\sigma= \partial_i(a^{ij}(\mathbf{x})u_{x^j}^\sigma(\mathbf{x})) +
b_\gamma^{\sigma s}(\mathbf{x}) u_{x^s}^\gamma(\mathbf{x}) + c_\gamma^\sigma(\mathbf{x}) u^\gamma(\mathbf{x}) \text{ for } \sigma= 1,\ldots, n-k, \label{L1}\tag{1}\] where \(A := (a^{ij})_{1 \leq i,j\leq k} \in
C^{0,1}(\overline{\Omega}; \mathbb{R}^{k\times k})\) is a positive definite symmetric matrix with eigenvalues in \([\lambda, \Lambda]\) for \(0 < \lambda < \Lambda < \infty\)
and \(b_\gamma^{\sigma s}, c_\gamma^\sigma\in \mathcal{L}^\infty(\Omega)\). The Cauchy problem asks whether there exists a unique \(u \in H^1(\Omega; \mathbb{R}^{n-k})\) solving
\[\label{cauchy1} \begin{cases} Lu(\mathbf{x}) = 0 \text{ in } \Omega\\ u(\mathbf{x}) = 0 \text{ and } \partial_\nu u(\mathbf{x}) = 0 \text{ on } \Gamma, \end{cases}\tag{2}\]
where \(\Gamma\) is a relatively open portion of \(\partial\Omega\), \(\nu\) is the unit outer normal to \(\Gamma\), and
\(L\) is given by 1 . Since any solution \(u \in H^1(\Omega; \mathbb{R}^{n-k})\) to 2 has zero trace on \(\Gamma\), it is \(C^{1,\alpha}\) up to \(\Gamma\) by elliptic regularity so that \(u\) and \(\partial_\nu u\) continuously vanish along \(\Gamma\). By boundary Carleman estimates, if \(u\) solves 2 then \(u \equiv 0\) in \(\Omega\) (see [17], [18] and [19]).
Since uniqueness in the Cauchy problem follows from Carleman estimates, it is intimately tied to the unique continuation properties of elliptic operators. For this reason, it is often called unique continuation from Cauchy data. We say that a
function \(u \in H^1(\Omega; \mathbb{R}^{n-k})\) has the strong unique continuation property (SUCP) in \(\Omega\) if \(u \equiv 0\) whenever there
is a point \(\mathbf{x}_0 \in \Omega\) at which \(u\) vanishes to infinite order [20]. The SUCP
for solutions \(u\) to \(Lw = 0\) in \(\Omega\) with \(L\) in 1 was first proved by Aronszajn,
Krzywicki, and Szarski in [18] using Carleman estimates (see [19] and [20]–[22]). The condition that \(L\) is principally diagonal is necessary, since
there are counterexamples to SUCP and uniqueness in 2 when we allow for leading-order coupling (see e.g. [23]). Moreover, the results in
[18]–[20], [22] are essentially sharp, because for each \(\alpha \in (0,1)\) there are scalar divergence form and non-divergence form elliptic operators with \(C^{0,\alpha}\) leading coefficient functions for which SUCP fails (see [23], [24]).
The first step in the proof of 2 is to prove uniqueness in the Cauchy problem for Lipschitz stationary graphs. Let \(\Omega\subset \mathbb{R}^k\)
be a domain (not necessarily bounded) with Lipschitz boundary and suppose \(u \in C_{\text{loc}}^{0}(\overline{\Omega}; \mathbb{R}^{n-k}) \cap C_{\text{loc}}^{0,1}(\Omega; \mathbb{R}^{n-k})\). Then its graph \(\Sigma_u\) is a \(k\)-dimensional Lipschitz submanifold of \(\mathbb{R}^n\) with boundary \[\partial\Sigma_u :=
\{(\mathbf{x},u(\mathbf{x})) : \mathbf{x}\in \partial\Omega\subset \mathbb{R}^k\} \subset \mathbb{R}^n.\] Define \[F: \mathbb{R}^{(n-k)\times k} \rightarrow \mathbb{R}\text{ by } F(\mathbf{p}) := \sqrt{\det( I +
\mathbf{p}^T\mathbf{p})}.\] Then the \(k\)-dimensional surface area of \(\Sigma_u\) is given by the formula \[\label{areafunctional} \mathcal{A}(Du) := \int_{\Sigma_u} \, d\mathcal{H}^k = \int_\Omega\sqrt{\det g(Du(\mathbf{x}))}\, d\mathbf{x}= \int_\Omega F(Du(\mathbf{x})) \, d\mathbf{x},\tag{3}\] where \(\mathcal{H}^k\) is the \(k\)-dimensional Hausdorff measure and the induced metric \(g\) is \(\mathcal{H}^k\)-a.e. defined by
\[g(Du(\mathbf{x})) := I + Du(\mathbf{x})^TDu(\mathbf{x}) = (\delta_{ij} + u_{x^i}(\mathbf{x}) \cdot u_{x^j}(\mathbf{x}))_{1 \leq i,\,j \leq k}. \label{metric}\tag{4}\] The
submanifold \(\Sigma_u\) can be identified with a stationary varifold \(V_u:= \underline{v}(\Sigma_u, 1)\) if and only if \(u\) is a stationary
solution to the minimal surface system (MSS) in \(\Omega\): \[\begin{cases} \partial_i(F_{p_i^\gamma}(Du(\mathbf{x}))u_{x^j}^\gamma(\mathbf{x}) -
F(Du(\mathbf{x}))\delta_{ij}) = 0 \text{ for } j = 1,\ldots, k \\ \partial_i(F_{p_i^\sigma}(Du(\mathbf{x}))) = 0 \text{ for } \sigma= 1,\ldots, n-k, \end{cases} \label{MSS}\tag{5}\] where \(DF(\mathbf{p}) :=
(F_{p_i^\sigma}(\mathbf{p}))_{1 \leq i \leq k}^{1 \leq \sigma \leq n-k}\) is viewed as a map \(\mathbb{R}^{(n-k) \times k} \rightarrow \mathbb{R}^{(n-k) \times k}\). The first equation in 5 is
called the inner variation system for \(\mathcal{A}\), while the second is called the outer variation system, or Euler–Lagrange system, for \(\mathcal{A}\).
When \(u \in C_{\text{loc}}^2(\Omega; \mathbb{R}^{n-k}) \cap C_{\text{loc}}^{0}(\overline{\Omega}; \mathbb{R}^{n-k})\), the MSS is equivalent to the following quasi-linear elliptic system in non-divergence form:
\[g^{ij}(Du(\mathbf{x}))u_{x^i x^j}^\sigma(\mathbf{x}) = 0 \text{ for each } \sigma = 1,\ldots, n-k \text{ in } \Omega, \label{MSS2}\tag{6}\] where we have set \(g^{-1} := (g^{ij})_{1 \leq i,j \leq k}\). Note that \(g^{-1}\) is positive definite with eigenvalues in \([\lambda, \Lambda] \subset (0,\infty)\) for some
numbers \(\lambda, \Lambda\) depending on the Lipschitz constant for \(u\). Any solution \(u\) solving 6 is called a classical
solution to the MSS. When \(k = n-1\) (i.e. \(\Sigma_u\) has codimension one), the outer variation equation is the minimal surface equation (MSE), which is the quasi-linear
divergence form scalar equation with coefficients \[a^{ij}(Du(\mathbf{x})) := \frac{1}{\sqrt{1 + |Du(\mathbf{x})|^2}}\Bigg(\delta_{ij} - \frac{u_{x^i}(\mathbf{x}) u_{x^j}(\mathbf{x})}{\sqrt{1 +
|Du(\mathbf{x})|^2}}\Bigg). \label{MSE}\tag{7}\]
Now, a straightforward computation using the fundamental theorem of calculus (FTC) shows that the difference of two solutions to 6 (or 7 ) solves a homogeneous equation for a principally diagonal operator of
the form 1 on a neighborhood of \(\Gamma\) in \(\Omega\). Applying this along with the SUCP and partial regularity for Lipschitz stationary solutions to the MSS [25], we obtain uniqueness in the Cauchy problem for 5 .
Proposition 3 (Uniqueness in the Cauchy Problem). Suppose \(u \in C_{\text{loc}}^{0}(\overline{\Omega}; \mathbb{R}^{n-k}) \cap C_{\text{loc}}^{0,1}(\Omega; \mathbb{R}^{n-k})\) solves 5 in a Lipschitz domain \(\Omega\) (not necessarily bounded) and that \[u = \phi \text{ and } \partial_\nu u = \Phi \text{ on }
\Gamma\label{cauchy3}\qquad{(1)}\] on a relatively open strictly convex \(C^{2,\alpha}\) patch \(\Gamma\subset \partial\Omega\), where \(\nu\)
is the unit outer normal to \(\Gamma\), \(\phi \in C_{\text{loc}}^{2,\alpha}(\Gamma; \mathbb{R}^{n-k}\)), and \(\Phi \in
C_{\text{loc}}^{1,\alpha}(\Gamma;\mathbb{R}^{(n-k)\times k})\). Let \(v \in C_{\text{loc}}^{0}(\overline{\Omega}; \mathbb{R}^{n-k}) \cap C_{\text{loc}}^{0,1}(\Omega; \mathbb{R}^{n-k})\) be another solution to 5 in \(\Omega\) with Cauchy data ?? . Then \(u \equiv v\) in \(\Omega\). When \(n - k = 1\), we can drop
strict convexity of \(\Gamma\).
Remark 4. The convexity assumption on \(\Gamma\) can be removed and we may assume \(\Gamma\) is \(C^{1,1}\) if we know \(u,v \in C_{\text{loc}}^{1,1}\) up to \(\Gamma\) in \(U \cap \Omega\) a priori, where \(U\) is a neighborhood \(U\) of \(\Gamma\) in \(\mathbb{R}^n\).
Step 2 in the proof of 2 is to formulate and prove unique continuation from Cauchy data in the general setting of area-minimizing integral currents. However, the difficulty lies
in regularity: In general, area-minimizing currents tend to exhibit singularities. When the codimension is one, it is well known that, away from its boundary, the support of an area-minimizing integral \(k\)-current in
Euclidean space is an analytic hypersurface outside of a relatively closed subset of Hausdorff dimension at most \(k-7\). We refer to this set as the interior singular set. In higher codimension, Almgren [26] showed that the interior singular set of area-minimizing integral \(k\)-currents in Euclidean space has Hausdorff dimension at most
\(k-2\).
Boundary regularity is more subtle. In codimension one, several results are known (see e.g. [10], [27]). On
the other hand, in higher codimension our understanding remained limited for a long time. Only recently did De Lellis, De Philippis, Hirsch, and Massaccesi [12]
establish the first general boundary regularity theorem without restrictions on the codimension or the ambient manifold. With the help of their regularity theory, we are able to prove unique continuation from Cauchy data for compactly supported
area-minimizing integral currents. In the setting of area-minimizing currents, this can be stated as follows:
Proposition 5 (Unique Continuation from Cauchy Data). Assume \(U\subset \mathbb{R}^n\) and \(W \subset \subset U\) are open connected sets. Let \(T \in \mathcal{I}_{k,c}(\mathbb{R}^n)\) and \(\Gamma\subset W\) be as in 1. Suppose that \(T^\prime \in \mathcal{I}_{k,c}(\mathbb{R}^n)\) is also area-minimizing in \(U\) with \(\mathop{\mathrm{spt}}T^\prime \subset W\), and \(\partial T^\prime = [[\Gamma]]\). If the approximate tangent spaces for \(\mathop{\mathrm{spt}}T\) and \(\mathop{\mathrm{spt}}T^\prime\) agree along a relatively
open patch \(\Gamma^\prime \subset \Gamma\), then \(T = T^\prime\).
Due to the boundary regularity, we can choose a point \(p \in \Gamma^\prime\) such that, near \(p\), both \(\mathop{\mathrm{spt}}T\) and \(\mathop{\mathrm{spt}}T^\prime\) can be represented as Lipschitz stationary graphs over a domain whose boundary contains a relatively open \(C^{3, \alpha}\) portion. Moreover, their defining
functions satisfy the same Cauchy data, so 3 and 4 together imply that \(\mathop{\mathrm{spt}}T\) and \(\mathop{\mathrm{spt}}T^\prime\) agree in a neighborhood of \(p\). Since the interior regular sets of \(T\) and \(T^\prime\) are analytic, a unique continuation argument (i.e. 19) shows that \(T\) and \(T^\prime\) share the same regular part. Incorporating a discussion of the multiplicities yields the conclusion of 5.
The third and final step is to show that, for a given boundary \(\Gamma\subset \mathbb{R}^n\) satisfying the hypotheses in 1, the existence of a
calibrated current \(T \in \mathcal{I}_{k,c}(\mathbb{R}^n)\) with \(\partial T = [[\Gamma]]\) forces any other area-minimizing current \(T^\prime \in
\mathcal{I}_{k,c}(\mathbb{R}^n)\) with \(\partial T^\prime = [[\Gamma]]\) to share the same Cauchy data in the sense of 5 along a relatively
open portion of \(\Gamma\). The argument follows from a couple of observations. The first is that \(T^\prime\) is also calibrated by \(\varphi\) (26). Then, the first cousin principle (24) together with the boundary
regularity theory imply that the true tangent spaces of \(\mathop{\mathrm{spt}}T\) and \(\mathop{\mathrm{spt}}T^\prime\) coincide on a relatively open portion of \(\Gamma\). This allows us to apply 5, from which uniqueness follows. We conclude this discussion with a remark on general ambient manifolds.
Remark 6. With only cosmetic changes, 5, 2, and all the other lemmas/theorems in this
paper concerning area-minimizing currents continue to hold when the ambient manifold \(M\) is complete without boundary, analytic, and satisfies \(H_k(M; \mathbb{R}) = 0\).
The organization of this paper is as follows. Section 2 introduces the basic notation and definitions concerning area-minimizing and calibrated integral currents. In Section 3, we briefly recall the interior regularity theory for Lipschitz stationary
solutions to the MSS 5 , and then present the proof of 3. We also record certain boundary regularity results from [12] and use them to prove 5. Following this, we discuss the generalization to area-minimizing currents with
unbounded support or higher multiplicity. Section 4 contains a general version of 2 (i.e. 27) and the
sharpness of the assumptions in 27. Section 5 discusses 6 in depth. Finally, in Appendix A we prove
the following folklore result: If a compactly supported area-minimizing integral current is extendable in a suitable sense, then it is unique. We compare this with 2.
Acknowledgement 1. The authors thank Rick Schoen for bringing the questions addressed in this paper to their attention, as well as Jacek Rzemieniecki for pointing out the first cousin principle (i.e. 24). The first author thanks Connor Mooney for many helpful discussions related to this paper, along with Zihui Zhao for an informative conversation at the RMMC Summer School, 2025, that
inspired Section 3.1. The second author thanks Nick Edelen for his many insightful comments and Chung-Jun Tsai for pointing out the obstruction for general ambient spaces. He also thanks National Taiwan University, where part of this work was carried out,
for their hospitality. The first author was supported by C. Mooney’s NSF CAREER Grant DMS-2143668 as well as NSF RTG Grant DMS-2342135.
For fixed \(n \in \mathbb{N}\), we define \(\Lambda^1(\mathbb{R}^n) := (\mathbb{R}^n)^*\). That is, \(\Lambda^1(\mathbb{R}^n)\) is the space of linear
functionals \(\ell: \mathbb{R}^n \rightarrow \mathbb{R}\). For each \(k \geq 2\), we will write \(\Lambda^k(\mathbb{R}^n)\) for the space of alternating
\(k\)-linear functions (i.e. \(k\)-covectors). The space of smooth \(k\)-forms on \(\mathbb{R}^n\) is denoted by \(\mathcal{E}^k(\mathbb{R}^n) \mathrel{\vcenter{:}}= C^\infty(\mathbb{R}^n; \Lambda^k(\mathbb{R}^n))\). We will write \(\mathcal{D}^k(\mathbb{R}^n)\) for the space of smooth \(k\)-forms with compact support, which is a subspace of \(\mathcal{E}^k(\mathbb{R}^n)\) but equipped with the usual locally convex topology. The space of \(k\)-vectors is denoted \(\Lambda_k(\mathbb{R}^n)\). A \(k\)-vector is called simple if \(\xi = v_1 \wedge v_2 \wedge\cdots
\wedge v_k\), where \(v_i \in \Lambda_1(\mathbb{R}^n)\) for all \(i = 1, \ldots, k\) and \(\{v_1,\ldots, v_k\}\) forms a linearly independent set.
Note that we can identify the oriented linearly independent set \(\langle v_1, \ldots, v_k \rangle\) with an oriented \(k\)-plane in \(\mathbb{R}^n\)
passing through the origin. In the same vein, such a \(k\)-plane in \(\mathbb{R}^n\) can be represented by the simple \(k\)-vector formed as a wedge product
of all elements in its oriented basis. That is, a simple \(k\)-form is equivalent to an oriented \(k\)-plane passing through the origin, up to a scalar factor. We will therefore write \[G(k, \mathbb{R}^n) := \{\xi \in \Lambda_k(\mathbb{R}^n): |\xi| = 1 \text{ and } \xi \text{ is simple}\}\] for the Grassmanian of oriented \(k\)-planes in \(\mathbb{R}^n\) through the origin.
The comass norm of \(\varphi\in \mathcal{\mathcal{E}}^k(\mathbb{R}^n)\) at \(p \in \mathbb{R}^n\) is given by \[||\varphi||_p := \sup_{\xi \in G(k,\mathbb{R}^n)} \langle \varphi_p, \xi \rangle, \label{cmassp}\tag{8}\] where \(\langle \cdot, \cdot \rangle\) is the dual pairing3 for \(\Lambda^k(\mathbb{R}^n)\) and \(\Lambda_k(\mathbb{R}^n)\). The comass norm of \(\varphi\in \mathcal{E}^k(\mathbb{R}^n)\), denoted \(||\varphi||\), is the supremum in 8 over \(p \in \mathbb{R}^n\). We say \(\varphi\) has comass one if \(||\varphi||=1\).
The space \(\mathcal{D}_k(\mathbb{R}^n)\) will represent the continuous dual space of \(\mathcal{D}^k(\mathbb{R}^n)\) (i.e. the \(k\)-currents).
For any \(T \in \mathcal{D}_k(\mathbb{R}^n)\), we define \(\partial T \in \mathcal{D}_{k-1}(\mathbb{R}^n)\) (i.e. the boundary of \(T\)) by \(\partial T(\varphi) := T(d\varphi) \text{ for all } \varphi\in \mathcal{D}^{k-1}(\mathbb{R}^n)\). For \(T \in \mathcal{D}_k(\mathbb{R}^n)\), the mass of \(T\) is given by \[\mathbb{M}(T) := \sup_{\varphi\in \mathcal{D}^k(\mathbb{R}^n)}\{T(\varphi): ||\varphi|| \leq 1 \}.\] If \(U \subset \mathbb{R}^n\) is an
open set, we can define the mass of \(T\) in \(U\), denoted by \(\mathbb{M}_U(T)\), by instead taking the supremum over those \(\varphi\) with \(\mathop{\mathrm{spt}}\varphi\subset U\). We say \(T\) has finite mass if \(\mathbb{M}(T) <
\infty\), and \(T\) has locally finite mass if \(\mathbb{M}_W(T) < \infty\) for every open \(W \subset\subset \mathbb{R}^n\). If both
\(T\) and \(\partial T\) have (locally) finite mass, then \(T\) is called (locally) normal.
Definition 7 (Integral Current). If \(T \in \mathcal{D}_k(\mathbb{R}^n)\), we say that \(T\) is a locally rectifiable integral \(k\)-current (abbreviated as integral \(k\)-current) if \(T\) is locally normal and \[T(\varphi) = \int_{M_T} \langle
\varphi_p ,\xi_T(p) \rangle \theta_T(p) \, d\mathcal{H}^k(p) \text{ for all } \varphi\in \mathcal{E}^k(\mathbb{R}^n),\] where \(M_T \subset \mathbb{R}^n\) is a \(k\)-rectifiable Borel
set, \(\theta_T \in \mathcal{L}_{\text{loc}}^1(M_T; d\mathcal{H}^k)\) is a positive integer valued function on \(\mathbb{R}^n\), and \(\xi_T: M_T \rightarrow
\Lambda_k(\mathbb{R}^n)\) is a Borel measurable function such that for \(\mathcal{H}^k\)-a.e. \(p \in M_T\) we have \(\xi_T(p) = e_1 \wedge\cdots \wedge
e_k\), where \(\{e_1,\ldots,e_k\}\) is an orthonormal basis for the approximate tangent space \(T_p M_T\). It is common to write \(T = \underline{\tau}(M_T,
\theta_T, \xi_T)\).
Remark 8. Federer and Fleming’s boundary rectifiability theorem (cf. [28]) shows that if \(T\) is an
integral \(k\)-current, then \(\partial T\) is an integral \((k-1)\)-current.
The function \(\theta_T\) is the multiplicity function for \(T\) and \(\xi_T\) is called the orientation for \(T\). We will write \(\|T\| \mathrel{\vcenter{:}}= \mathcal{H}^k \llcorner \theta_T \llcorner M_T\) for the mass measure of \(T\). Note that \(\mathbb{M}_U(T) = \|T\|(U)\) for every Borel set \(U \subset \mathbb{R}^n\). Moreover, \(\mathop{\mathrm{spt}}T = \mathop{\mathrm{spt}}\|T\|\), where \(\mathop{\mathrm{spt}}\|T\|\) is the support of \(\|T\|\) in the usual sense of Radon measures. We shall write \(\mathcal{I}_k(\mathbb{R}^n)\) for the set of all
integral \(k\)-currents. The subset of currents in \(\mathcal{I}_k(\mathbb{R}^n)\) with compact support is denoted \(\mathcal{I}_{k,c}(\mathbb{R}^n)\).
Example 1 (Lipschitz Submanifolds). Suppose \(\Sigma\subset \mathbb{R}^n\) is a \(k\)-dimensional oriented embedded Lipschitz submanifold with \(\mathcal{H}^k(\Sigma) + \mathcal{H}^{k-1}(\partial \Sigma) < \infty\) and \(\mathcal{H}^k\)-a.e. defined orientation \(\xi_\Sigma: \Sigma\rightarrow \Lambda_k(
\mathbb{R}^n)\). Then there is a unique integral \(k\)-current \([[\Sigma]] \in \mathcal{I}_k(\mathbb{R}^n)\) associated to \(\Sigma\) defined by
taking \(M_T = \Sigma\), \(\theta_T = \theta_\Sigma\equiv 1\), and \(\xi_T = \xi_\Sigma\) in 7.
Federer and Fleming show that the homology of the chain complex of currents with compact support is related to the singular homology [4]. As an important
consequence, we have:
Lemma 9. If \(T \in \mathcal{I}_{k, c}(\mathbb{R}^n)\) satisfies \(\partial T = 0\), then \(T = \partial S\) for some \(S \in \mathcal{I}_{k+1, c}(\mathbb{R}^n)\). If \(T\) is a normal \(k\)-current with compact support and satisfies \(\partial T =
0\), then \(T = \partial S\) for some normal \((k+1)\)-current \(S\) with compact support.
Of particular interest in the present paper are the area-minimizing integral \(k\)-currents.
Definition 10. Let \(U \subset \mathbb{R}^n\) be an open set. An integral \(k\)-current \(T \in \mathcal{I}_k(\mathbb{R}^n)\) is
area-minimizing in \(U\) if for every open \(W \subset\subset U\) we have \[\mathbb{M}_W(T) \leq \mathbb{M}_W(T + S) \text{ for all } S \in
\mathcal{I}_{k, c}(\mathbb{R}^n) \text{ with } \partial S = 0 \text{ and } \mathop{\mathrm{spt}}S \subset W.\] We say \(T\) is globally area-minimizing if \(U =
\mathbb{R}^n\).
Throughout the paper, we will often consider the case when \(T\) is a globally area-minimizing integral current and has non-zero boundary \(\partial T := [[\Gamma]]\), where \(\Gamma\subset \mathbb{R}^n\) satisfies 1. If \(T\) has compact support, then \(T\) has the
least mass among all compactly supported integral currents with boundary \([[\Gamma]]\). If \(T\) is not compactly supported, then \(T\) has the least mass
among all “compact perturbations” of its support.
Let \(M \subset \mathbb{R}^n\) be a \(k\)-rectifiable Borel set and \(\theta\in \mathcal{L}^1_{loc}(M; d\mathcal{H}^k)\). A rectifiable \(k\)-varifold\(\underline{v}(M, \theta)\) is the equivalence class of all pairs \((M^\prime, \theta^\prime)\), where \(M^\prime\) is \(k\)-rectifiable with \(\mathcal{H}^k(M \Delta M^\prime) = 0\) and \(\theta= \theta^\prime\)\(\mathcal{H}^k\)-a.e. on \(M \cap M^\prime\). In particular, given an integral current \(T\), the varifold associated with \(T\)
is just the equivalence class \(\underline{v}(M_T, \theta_T)\), i.e. forgetting the orientation \(\xi_T\). A rectifiable varifold is called stationary if the first variation
vanishes. Loosely speaking, stationary rectifiable varifolds can be thought of as Lipschitz minimal submanifolds (not necessarily oriented). We refer the interested reader to [28] for more on rectifiable varifolds.
Let \(U \subset \mathbb{R}^n\) be an open set, and \(p \in U\). For \(T \in \mathcal{I}_{k}(\mathbb{R}^n)\) that is area-minimizing in \(U\) and satisfies \(\mathop{\mathrm{spt}}T \subset U\), the density of \(T\) at \(p\) is defined by
\[\Theta_T(p) = \lim_{r \rightarrow 0} \frac{\|T\|(B^n_r(p))}{\omega_k r^k} \label{dense}\tag{9}\] whenever the latter limit exists. Here, \(\omega_k\) denotes the volume of the unit \(k\)-ball. Whenever \(p \notin \mathop{\mathrm{spt}}(\partial T)\), the existence of this limit is always guaranteed
by the monotonicity formula (cf. [28]), since the varifold associated with an area-minimizing current is stationary (cf. [28]). Thus, area-minimizing integral currents can be heuristically thought of as oriented Lipschitz minimal submanifolds having the least surface area (i.e. mass) with respect to
their boundaries. The monotonicity formula also shows that \(\Theta_T(p)\) is an upper semi-continuous function of \(p\). We finally note that \(\Theta_T(p) =
\theta_T(p)\) for \(\mathcal{H}^k\)-a.e. \(p \in \mathop{\mathrm{spt}}T\) (cf. [28]).
Therefore, \(T\) has the canonical representation\[\label{cano} T = \underline{\tau}(\mathop{\mathrm{spt}}T, \Theta_T, \xi_T).\tag{10}\]
Let \(\varphi\in \mathcal{E}^k(\mathbb{R}^n)\) have comass one. For \(p \in \mathbb{R}^n\), we write \[\mathcal{G}_p(\varphi) \mathrel{\vcenter{:}}= \{\xi \in
G(k, \mathbb{R}^n): \langle \varphi_p, \xi \rangle = 1\}\] for the collection of planes where \(\varphi_p\) achieves its maximum. Such \(\mathcal{G}_p(\varphi)\) is called the
contact set of \(\varphi\) at \(p\). We also define \[\mathcal{G}(\varphi) \mathrel{\vcenter{:}}= \bigcup_{p \in \mathbb{R}^n}
\mathcal{G}_p(\varphi).\]
Definition 11. A smooth \(k\)-form \(\varphi\in \mathcal{E}^k(\mathbb{R}^n)\) is called a calibration if \(\varphi\) has comass
one and \(d \varphi= 0\). Let \(U \subset \mathbb{R}^n\) be an open set. An integral \(k\)-current \(T \in
\mathcal{I}_k(\mathbb{R}^n)\) is said to be calibrated by \(\varphi\) in \(U\) if \(\xi_T(p) \in \mathcal{G}_p(\varphi)\) for \(\|T\|\)-a.e. \(p \in U\).
Calibrated currents achieve their mass when acting on the calibration, and are area-minimizing.
Lemma 12 (cf. [29]). Let \(\varphi\in \mathcal{E}^k(\mathbb{R}^n)\) be a calibration and \(T \in \mathcal{I}_k(\mathbb{R}^n)\). Then \[(T \llcorner W)(\varphi) \leq \mathbb{M}_W(T)\] for every open \(W \subset\subset \mathbb{R}^n\), and the equality
holds if and only if \(T\) is calibrated by \(\varphi\) in \(W\).
Lemma 13 (cf. [29]). Fix \(1 \leq k \leq n-1\). Let \(U \subset
\mathbb{R}^n\) be an open set, and let \(\varphi\in \mathcal{E}^k(\mathbb{R}^n)\) be a calibration. Suppose that \(T \in \mathcal{I}_k(\mathbb{R}^n)\) is calibrated by \(\varphi\) in \(U\). Then \(T\) is area-minimizing in \(U\).
There are many examples of calibrations and calibrated currents, the easiest example being the current associated to the coordinate plane \(\mathbb{R}^k \times \{0\} \subset \mathbb{R}^{k} \times \mathbb{R}^{n-k}\)
calibrated by \(\varphi(\mathbf{x}) := dx^1 \wedge \cdots \wedge dx^k\), \(\mathbf{x}:= (x^1, \ldots, x^n) \in \mathbb{R}^n\). We provide a few historically important examples, but refer the
reader to [29] for a full treatment.
Example 2 (Lipschitz Stationary Graphs). Let \(\Omega\subset \mathbb{R}^{n}\) be a bounded \(C^{2}\) mean convex domain. Suppose \(u \in
C^\infty(\Omega) \cap C^{0}(\overline{\Omega})\) is a Lipschitz stationary solution to the MSE 7 , whose existence for fixed continuous boundary data is guaranteed by [30]. Then the volume form on its graph \(\Sigma_u\) is given by \[dvol_{\Sigma_u} = \frac{(-1)^{n}}{\sqrt{1 + |Du|^2} } dx^1 \wedge \cdots \wedge dx^n + \sum_{i =
1}^{n} \frac{(-1)^i u_{x^i}}{\sqrt{1 + |Du|^2} } \, dx^1 \wedge \cdots \wedge \widehat{dx^i} \wedge \cdots \wedge dx^n \wedge dy.\] The \(n\)-form \(\varphi(\mathbf{x},y)\) defined by
extending the right-hand side of the equality above to \(U := \Omega\times \mathbb{R}\) is a calibration and \([[\Sigma_u]]\) is calibrated by \(\varphi\) in
\(U\). That \(\varphi\) has comass one can be checked readily, and \(d\varphi\equiv 0\) in \(U\) since \(u\) solves 7 in \(\Omega\). It follows that if \(u\) solves 7 , then \([[\Sigma_u]]\) is area-minimizing in \(U\). If \(\Omega\) is convex, we can conclude that \([[\Sigma_u]]\) is globally
area-minimizing [16] by the convex hull property[16]. A
topological hypothesis on the boundary is necessary to ensure global uniqueness [31]. Since there are high-codimension stationary solutions to 5 over the unit ball which are not stable critical points of \(\mathcal{A}\) (see e.g. [32]), this example does
not extend to 5 .
Example 3 (Complex Submanifolds in \(\mathbb{C}^n\)). We say that \(\Sigma\subset \mathbb{C}^n\) is a \(k\)-dimensional complex
submanifold (real dimension \(2k\)), with or without boundary, if \(T_p \Sigma\) is a \(k\)-dimensional complex vector subspace of \(\mathbb{C}^n\) for each \(p \in \Sigma\). Using the identification \(\mathbb{C}^n \cong \mathbb{R}^n + \sqrt{-1} \, \mathbb{R}^n\), it is common to identify
\(\mathbb{C}^n\) with \(\mathbb{R}^{2n}\) in the coordinates \[\mathbf{z}:= (z^1,\ldots, z^n) = (x^1, \ldots, x^n, y^1, \ldots, y^n), \text{ } z^k := x^k +
\sqrt{-1} \, y^k.\] If we define the symplectic form \[\omega(\mathbf{z}) := \frac{1}{2 \sqrt{-1}} \sum_{j = 1}^n dz^j \wedge d \overline{z}^j =\sum_{j = 1}^n dx^j \wedge dy^j,
\label{symplectic}\tag{11}\] then \[\varphi(\mathbf{z}) := \frac{\omega^k(\mathbf{z})}{k!} := \frac{1}{k!}\underbrace{\omega(\mathbf{z}) \wedge \cdots \wedge \omega(\mathbf{z})}_{k \text{ times}}\] is a
closed complex \(k\)-form. By Wirtinger’s inequality (cf. [33]), \(\varphi\) is a calibration,
and \([[\Sigma]]\) is calibrated by \(\varphi\) whenever \(\Sigma\) is a \(k\)-dimensional complex submanifold of \(\mathbb{C}^n\). Common examples include the graph current \([[\Sigma_f]]\) in \(\mathbb{C}^2\) where \(f: \Omega\subset
\mathbb{C}\rightarrow \mathbb{C}\) is holomorphic and \(\Omega\) is a domain or, more generally, \([[Z(f)]]\) where \(Z(f)\) is the zero set of a
holomorphic function \(f: \Omega\times \mathbb{C}\rightarrow \mathbb{C}\) having no critical points in \(Z(f)\).
Example 4 (Special Lagrangian Submanifolds in \(\mathbb{C}^n\)). As in the previous example, we identify \(\mathbb{C}^n \cong \mathbb{R}^{2n}\) and let \(\omega\) be the symplectic form 11 . A real \(n\)-dimensional submanifold \(\Sigma\subset \mathbb{C}^n\) is called a Lagrangian
submanifold if \(\omega\big|_\Sigma= 0\). Let \(d\mathbf{z}\mathrel{\vcenter{:}}= dz^1 \wedge\cdots \wedge dz^n = \mathop{\mathrm{Re}}(d\mathbf{z})+ \mathop{\mathrm{Im}}(d\mathbf{z})\)
be a holomorphic \(n\)-form on \(\mathbb{C}^n\). A Lagrangian submanifold \(\Sigma\subset \mathbb{C}^n\) is called special if, in addition, \(\mathop{\mathrm{Im}}(d \mathbf{z})\big|_\Sigma= 0\). In this case, \(\varphi(\mathbf{z}) = \mathop{\mathrm{Re}}(d \mathbf{z})\) (as a real \(n\)-form) is a
calibration, and \([[\Sigma]]\) is calibrated by \(\varphi\) whenever \(\Sigma\) is a special Lagrangian submanifold in \(\mathbb{C}^n\).
For any \(C^2\) function \(u: \Omega\subset \mathbb{R}^n \rightarrow \mathbb{R}\) defined on a domain \(\Omega\), the graph \(\Sigma_{Du} \subset \mathbb{R}^{2n} \cong \mathbb{C}^n\) of \(D u\) is a Lagrangian submanifold in \(\mathbb{C}^n\) (cf. [29]). Any such \(u\) is called the potential function of \(\Sigma_{Du}\). If, in addition, \(u\) satisfies the special Lagrangian equation\[\sum_{i = 1}^n \arctan(\lambda_i(D^2u)) = \vartheta,\] where \(\vartheta \in (-\frac{n \pi}{2}, \frac{n
\pi}{2})\) is a constant and \(\lambda_i(D^2 u)\) denote the eigenvalues of the Hessian matrix \(D^2 u\), then \(\Sigma_{Du}\) is a special Lagrangian
submanifold in \(\mathbb{C}^n\), and \(\vartheta\) is called the phase of \(\Sigma_{Du}\).
Example 5 (Lawson–Osserman Cone). Take \(\mathscr{H}: \mathbb{S}^3 \rightarrow \mathbb{S}^2\) to be the Hopf map: \[\mathscr{H}(z^1,z^2) = (2\overline{z}^1 z^2, |z^1|^2 -
|z^2|^2)\] for \((z^1,z^2) \in \mathbb{C}^2\) with \(|z^1|^2 + |z^2|^2 = 1\). After making the identifications \(\mathbb{C}^2 \cong \mathbb{R}^4\) and
\(\mathbb{C} \times \mathbb{R}\cong \mathbb{R}^3\), we define the Lawson–Osserman cone (LOC), denoted by \(\mathcal{C}\), to be the graph of \(\mathscr{L}:
\mathbb{R}^4 \rightarrow \mathbb{R}^3\) given by \[\mathscr{L}(\mathbf{x}) := \frac{\sqrt{5}}{2}|\mathbf{x}|\mathscr{H}\left(\frac{\mathbf{x}}{|\mathbf{x}|}\right).
\label{LOC}\tag{12}\] It is a 4-dimensional stationary cone in \(\mathbb{R}^7\) with an isolated singularity at the origin, with associated integral \(4\)-current \([[\mathcal{C}]]\). The LOC was the first example of a singular Lipschitz stationary graph (cf. [32]). In [29], it was shown that the LOC is calibrated by the coassociative \(4\)-form on \(\mathbb{R}^7\): \[\varphi(\mathbf{x}) := dx^{1234} - dx^{67} \wedge(dx^{12}- dx^{34})+ dx^{57} \wedge(dx^{13} + dx^{24}) - dx^{56} \wedge(dx^{14} - dx^{23}),\] where \(dx^{i_1 i_2 \ldots i_k}:= dx^{i_1} \wedge dx^{i_2}
\wedge \cdots \wedge dx^{i_k}\) for \(1 \leq i_1 < i_2 < \cdots < i_k \leq 7\).
Before proving 3, we recall the interior regularity theory for Lipschitz stationary solutions to the MSS 5 , which follows from the theory for
outer stationary maps (i.e. solutions to the outer variation system). Assume only that \(u: \Omega\subset \mathbb{R}^k \rightarrow \mathbb{R}^{n-k}\) is outer stationary and \(u \in
C_{\text{loc}}^1(\Omega; \mathbb{R}^{n-k})\). Fix \(\mathbf{x}_0 \in \Omega\) and choose \(r > 0\) so that \(B_r^k(\mathbf{x}_0) \subset \subset
\Omega\). An important fact is that \(\mathcal{A}\) is strongly rank-one convex when restricted to Lipschitz functions [34] Lemma 6.7: the lemma is stated for \(k = 2\) but generalizes to arbitrary\(k\). Therefore, we can differentiate the equation in \(B_r^k(\mathbf{x}_0)\) to obtain a quasi-linear system with \(C^{0}\) coefficients satisfying the Legendre–Hadamard ellipticity condition[35]. The classical \(W^{2, p}\) theory for linear elliptic systems (see e.g. [35],
[36]) shows that \(u \in C^{1, \alpha}(B_r^k(\mathbf{x}_0); \mathbb{R}^{n-k})\).4
Morrey’s Theorem [36] then gives \(u \in C^\omega(B_r^k(\mathbf{x}_0); \mathbb{R}^{n-k})\) (i.e. the analytic functions). As a
result, \(C^1\) stationary solutions to 5 are analytic in the interior. In general, this cannot be improved due to the LOC. The outer variation system, 5 , and 6 are equivalent at regular points. In particular, when \(n-k = 1\) the MSS reduces to the MSE since Lipschitz solutions are analytic in the interior by the De Giorgi–Nash–Moser Theorem [37].
Since Lipschitz solutions to the MSE are analytic and solve a uniformly elliptic divergence form scalar equation, the SUCP and uniqueness in the Cauchy problem are immediate when \(n-k = 1\), provided enough boundary
regularity is assumed. This is not the case for high-codimension Lipschitz stationary graphs since there exist examples which are singular (see e.g. [32], [38], [39]). Hence, the set of interior regular points for a Lipschitz stationary graph is not apriori known to be connected. This is not an issue when \(k \leq 3\), since all such stationary graphs are analytic in the interior (see e.g. [40]–[42]).
In [25], it was shown that the interior singular set of a \(k\)-dimensional Lipschitz stationary solution to 5 in \(\mathbb{R}^n\) is relatively closed in the domain \(\Omega\subset \mathbb{R}^k\) and has Hausdorff dimension at most \(k-4\).
Denote this set \(\mathop{\mathrm{Sing}}(u) \subset \Omega\) for a given Lipschitz stationary solution to 5 on \(\Omega\). The following lemma is essential in our
argument as it implies that \(\mathop{\mathrm{Reg}}(u) \mathrel{\vcenter{:}}= \Omega\setminus \mathop{\mathrm{Sing}}(u)\) is connected, open, and dense in \(\Omega\). An elegant proof by
Mooney can be found in [43].
Lemma 14. Let \(\Omega\subset \mathbb{R}^k\) be open and connected. Suppose that \(K \subset \Omega\) is a relatively closed subset satisfying \(\mathcal{H}^{k-1}(K) = 0\). Then \(\Omega\setminus K \subset \Omega\) is relatively open, dense, and path connected.
The final ingredient is the boundary regularity theory. When \(n-k = 1\), we can apply boundary Schauder estimates. For general \(n,k\), we need Allard’s boundary regularity theory in
[10] which requires strict convexity of \(\partial\Omega\). See [32] for a proof.
Theorem 15 (Boundary Regularity). Assume \(\Omega\subset \mathbb{R}^k\) is a \(C^{2,\alpha}\) strictly convex domain and that \(\phi :
\partial\Omega \rightarrow \mathbb{R}^{n-k}\) is locally \(C^{2,\alpha}\). Suppose \(u \in C_{\text{loc}}^{0}(\overline{\Omega}; \mathbb{R}^{n-k}) \cap C_{\text{loc}}^{0,1}(\Omega;
\mathbb{R}^{n-k})\) solves the Dirichlet problem for 5 with boundary data \(\phi\). Then there is a neighborhood \(U\) of \(\partial\Omega\) in \(\mathbb{R}^n\) such that \(u \in C_{\text{loc}}^{2,\alpha}(U \cap \overline{\Omega}; \mathbb{R}^{n-k})\).
The utility of 15 is that it gives us some wiggle room near \(\partial\Omega\) to set up a Cauchy problem 2 with \(L\) as in 1 along regions of \(\partial\Omega\) with enough regularity. Suppose \(u \in C_{\text{loc}}^{0}(\overline{\Omega}; \mathbb{R}^{n-k})
\cap C_{\text{loc}}^{0,1}(\Omega; \mathbb{R}^{n-k})\), \(u\) solves 5 in a domain \(\Omega\) (not necessarily bounded) with Lipschitz boundary and that \(u\) has Cauchy data ?? . Let \(v \in C_{\text{loc}}^{0}(\overline{\Omega}; \mathbb{R}^{n-k}) \cap C_{\text{loc}}^{0,1}(\Omega; \mathbb{R}^{n-k})\) be any other solution to 5 in \(\Omega\) with Cauchy data ?? .
Proof of 3. Define \(w:= u - v\) in \(\Omega\). We only focus on the case of arbitrary \(k,n\), since the computation in the scalar case is standard (see e.g. [44]). Fix \(\mathbf{x}_0 \in
\Gamma\) and choose \(r >0\) small enough that \(\overline{B_r^k(\mathbf{x}_0)} \cap \partial\Omega\subset \Gamma\) is \(C^{2,\alpha}\) and has
connected interior in \(\Gamma\). Set \[\Omega_r := B_r^k(\mathbf{x}_0) \cap \Omega.\] While the MSS 5 is a diagonal divergence form system, it is not clear from the
divergence form structure that \(w\) solves a principally diagonal system \(Lw = 0\) with \(L\) in 1 . To side-step this issue, we use
the boundary regularity 15. Indeed, for \(r\) small enough we can ensure \(u,v \in C^\omega(\Omega_r;\mathbb{R}^{n-k}) \cap
C^{2,\alpha}(\overline{\Omega_r}; \mathbb{R}^{n-k})\). Hence, \(u\) and \(v\) both solve the diagonal non-divergence form system 6 in \(\Omega_r\) for this \(r\). A standard computation using the FTC (see e.g. [43]) shows that \(w\) satisfies a principally diagonal system of the form 1 in \(\Omega_r\), where \(a^{ij} \in C^{1,\alpha}(\overline{\Omega_r})\), \(b_\gamma^{\sigma s} \in C^{0,\alpha}(\overline{\Omega_r})\), and \(c_\gamma^\sigma\equiv 0\).
Now, \(U := \mathop{\mathrm{Reg}}(u) \cap \mathop{\mathrm{Reg}}(v)\) is open, dense, and connected in \(\Omega\) by 14 and [25], since \(U = \Omega\setminus (\mathop{\mathrm{Sing}}(u) \cup \mathop{\mathrm{Sing}}(v))\)
where \(\mathop{\mathrm{Sing}}(u) \cup \mathop{\mathrm{Sing}}(v)\) is closed and \[\mathcal{H}^{k-1}\big(\mathop{\mathrm{Sing}}(u) \cup \mathop{\mathrm{Sing}}(v)\big) = 0\] by the
subadditivity of the \(\mathcal{H}^{k-1}\)-measure. The computation in the first paragraph applied to \(w := u -v\) in \(\Omega_r\) shows \(w \equiv 0\) in \(\Omega_r\) since \(w\) solves a Cauchy problem 2 in \(\Omega_r\), and the SUCP for
analytic stationary graphs implies \(u = v\) on \(U\). As \(u\) and \(v\) are continuous, this forces \(u \equiv v\) on \(\Omega\). 4 is evident from the proof provided, since the only issue is the regularity of the
coefficient functions near the boundary for the system \(w\) solves. ◻
A simple example comes from applying 3 to the LOC.
Example 6. The function \(\mathscr{L}\) in 12 is the unique solution to ?? in \(B_1^4(0) \subset \mathbb{R}^4\) with \(u =
\frac{\sqrt{5}}{2} \mathscr{H}\) and \(Du = \frac{\sqrt{5}}{2}D\mathscr{H}\) on \(\partial B_1^4(0)\). It is unknown whether \(\mathscr{L}\) is the
unique solution to the Dirichlet problem for 5 on \(B_1^4(0)\) with boundary data \(\mathscr{H}\) on \(\partial B_1^4(0)\).
Existence and uniqueness of solutions to the Dirichlet problem for the MSS are not guaranteed, even when \(\Omega= B_1^k(0)\) and we assume \(C^\infty\) boundary data (see e.g. [32], [38]). However, we will see that 3
can be used to establish uniqueness in the Plateau problem.
We formulate and prove the unique continuation property from Cauchy data for area-minimizing integral currents using 3, the interior regularity theory of Almgren in [26] (see also [45], [46], [47]), and the boundary regularity theory of De Lellis, De Philippis, Hirsch, and Massaccesi in [12]. Going forward, we let \(\Gamma\) and \(T\) be as in 1 unless
otherwise stated.
We will need to distinguish between interior regular points and boundary regular points for \(T\) since we would like to locally reduce the problem to the case of Lipschitz stationary graphs using the regularity theory
for area-minimizing integral currents. Before proving 5, we summarize the key points from the regularity theory beginning with the definition of interior regular points.
Definition 16 (Interior Regular Point). A point \(p \in \mathop{\mathrm{spt}}T \setminus \Gamma\) is called an interior regular point for \(T\) if there is an
\(r > 0\) such that \(B_r^n(p) \cap \Gamma= \emptyset\) and there is a \(k\)-dimensional connected oriented embedded \(C^1\) submanifold \(\Sigma\subset B_r^n(p)\) without boundary in \(B_r^n(p)\) such that \(T \llcorner B_r^n(p) = Q[[\Sigma]]\),
where \(Q \in \mathbb{N}\) (so \(\mathop{\mathrm{spt}}(T \llcorner B_r^n(p)) = \Sigma\)). The set of interior regular points is denoted \(\mathop{\mathrm{Reg}}_i(T)\), and the set of interior singular points is \(\mathop{\mathrm{Sing}}_i(T) := \mathop{\mathrm{spt}}T \setminus (\Gamma\cup \mathop{\mathrm{Reg}}_i(T))\).
Remark 17. Due to 9 and 10 , we note that \(\theta_T(p) = \Theta_T(p) = Q \in \mathbb{N}\) when \(p \in
\mathop{\mathrm{Reg}}_i(T)\). Such \(Q\) may vary depending on \(p\); however, the Constancy theorem (cf. [28]) shows that \(\Theta_T\) is locally constant on \(\mathop{\mathrm{Reg}}_i(T)\). Therefore, \(\Theta_T\) is constant on
each connected component of \(\mathop{\mathrm{Reg}}_i(T)\).
Remark 18. If \(T \in \mathcal{I}_k(\mathbb{R}^n)\) is area-minimizing in \(U \subset \mathbb{R}^n\) and \(p \in \mathop{\mathrm{Reg}}_i(T
\llcorner U)\), then there is an \(r > 0\) such that \(B_r^n(p) \subset \subset U\), \(B_r^n(p) \cap \Gamma= \emptyset\), and \(T \llcorner B_r^n(p) = Q[[\Sigma]]\). In addition, \(Q[[\Sigma]]\) is stationary when identified with its corresponding varifold \(V_\Sigma:= \underline{v}(\Sigma,
Q)\). Choosing \(r\) smaller if necessary, \(\Sigma\) can be represented as the graph \(\Sigma= \Sigma_u\) of some analytic function \(u :\Omega \subset \mathbb{R}^k \rightarrow \mathbb{R}^{n-k}\), where \(\Omega\) is a bounded, simply connected domain. Therefore, \(T = Q[[\Sigma]]\) for an
analytic \(k\)-dimensional minimal submanifold \(\Sigma\) without boundary in a neighborhood of any point in \(\mathop{\mathrm{Reg}}_i(T\llcorner U)\).
When combined with a covering argument, analyticity yields the following identity lemma for connected interior regular sets. The requirement of a shared boundary is essential. We leave the details of the proof to the reader.
Lemma 19. Suppose that \(S, S^\prime \in \mathcal{I}_k(\mathbb{R}^n)\) are such that
\(\partial S = \partial S^\prime\).
\(\mathop{\mathrm{Reg}}_i(S)\), \(\mathop{\mathrm{Reg}}_i(S^\prime)\) are analytic and connected.
\(\mathop{\mathrm{Reg}}_i(S)\) and \(\mathop{\mathrm{Reg}}_i(S^\prime)\) make contact at a point \(x_0 \in \mathop{\mathrm{Reg}}_i(S) \cap
\mathop{\mathrm{Reg}}_i(S^\prime)\) of infinite order (see [20], [21] for a precise
definition).
Then \(\mathop{\mathrm{Reg}}_i(S) = \mathop{\mathrm{Reg}}_i(S^\prime)\) as oriented analytic submanifolds.
Defining boundary regular points is more delicate. The following definition depends on the boundary regularity assumed in 1.
Definition 20 (Boundary Regular Point). A point \(p \in \Gamma\) is called a boundary regular point for \(T\) if there is an \(r >
0\) and a \(k\)-dimensional connected oriented embedded \(C^{3, \alpha}\) submanifold \(\Sigma\subset B_r^n(p)\) without boundary in \(B_r^n(p)\) such that \(\mathop{\mathrm{spt}}(T \llcorner B_r^n(p)) \subset \Sigma\). The set of boundary regular points is denoted \(\mathop{\mathrm{Reg}}_b(T)\)
while the set of boundary singular points is \(\mathop{\mathrm{Sing}}_b(T) := \Gamma\setminus \mathop{\mathrm{Reg}}_b(T)\).
Fix \(p \in \mathop{\mathrm{Reg}}_b(T)\) (so \(p \in \Gamma)\) and let \(\Sigma\) be as in 20. Choose \(r >0\) so small that \(B^n_r(p) \cap \Sigma\) is diffeomorphic to a \(k\)-dimensional disk. Then the Constancy
Theorem implies:
\(\Gamma\cap B^n_r(p) \subset \Sigma\) and divides \(\Sigma\) into two disjoint \(k\)-dimensional oriented connected embedded \(C^{3, \alpha}\) submanifolds \(\Sigma^{\pm}\) with \(\partial\Sigma^\pm = \pm \Gamma\).
There is a natural number \(Q \in \mathbb{N}\) such that \[T \llcorner B^n_r(p) = Q[[\Sigma^+]] + (Q - 1)[[\Sigma^-]].\]
The number \(Q\) is the multiplicity of \(T\) at\(p \in \mathop{\mathrm{Reg}}_b(T)\). The density of \(T\)
at \(p \in \mathop{\mathrm{Reg}}_b(T)\) is \[\Theta_T(p) := Q - \frac{1}{2},\] and coincides with the definition 9 by [12]. The points \(p\) at which \(Q = 1\) are called density \(\frac{1}{2}\)
points, or one-sided points. The term “one-sided" comes from the fact that \(T \llcorner B^n_r(p) = [[\Sigma^+]]\). In other words, \(T\) can be locally identified with \(\Sigma^+\) near \(p\). If \(Q > 1\), we say that \(p\) is a two-sided point. See [12] for a helpful illustration of one-sided and two-sided boundary regular points. Since we intend to utilize PDE techniques, we need conditions under which a relatively open subset of
one-sided regular points exists along \(\Gamma\).
The question of the existence of one-sided boundary regular points for suitably regular \(\Gamma\) dates back to Almgren in his Big Regularity Paper [26]. A classic result of Hardt and Simon [27] says that if \(k = n - 1\), then \(\mathop{\mathrm{Reg}}_b(T) = \Gamma\). However, without additional hypotheses we cannot be sure that a given point \(p \in \mathop{\mathrm{Reg}}_b(T)\) is one-sided for general \(1 \leq k \leq n-1\). The first positive results on the existence of one-sided regular points in the high-codimension case were proved recently by De Lellis, De Philippis, Hirsch, and Massaccesi [12]. We record the most relevant.
Theorem 21 (cf. [12]). Assume \(U\subset \mathbb{R}^n\) and \(W \subset \subset
U\) are open connected sets. Suppose that \(T \in \mathcal{I}_{k,c}(\mathbb{R}^n)\) and \(\Gamma\subset W\) satisfy 1. Then:
\(\mathop{\mathrm{Reg}}_b(T)\) is open and dense in \(\Gamma\).
Every point in \(\mathop{\mathrm{Reg}}_b(T)\) is one-sided.
\(\mathop{\mathrm{Reg}}_i(T)\) is connected.
If \(p \in \mathop{\mathrm{Reg}}_b(T)\), then there is an \(r > 0\) such that \(T \llcorner B_r^n(p) = [[\Sigma_u]]\) for some stationary
solution \(u \in C^{3, \alpha}(\overline{\Omega}; \mathbb{R}^{n-k}) \cap C^\omega(\Omega; \mathbb{R}^{n-k})\) to 5 , where \(\Omega\subset \mathbb{R}^k\) is a bounded,
simply connected, \(C^{3, \alpha}\) domain.
\(\theta_T(p) = \Theta_T(p) = 1\) for all \(p \in \mathop{\mathrm{Reg}}_i(T)\) and \(\mathbb{M}(T) =
\mathcal{H}^k(\mathop{\mathrm{Reg}}_i(T))\).
Proof of 5. Fix \(p \in \mathop{\mathrm{Reg}}_b(T) \cap \mathop{\mathrm{Reg}}_b(T^\prime) \cap \Gamma^\prime\) and choose \(r > 0\) so that \(B^n_r(p) \subset \subset U\) and \(\Gamma^\prime \cap B^n_r(p)\) consists of one-sided regular boundary points for \(T\) and \(T^\prime\). Such a ball \(B^n_r(p)\) can be chosen due to 21(1) applied to
\(T\) and \(T^\prime\), along with the openness of \(\Gamma^\prime \subset \Gamma\). Let \(\Sigma\subset B^n_r(p)\) be as in
20 (so \(\mathop{\mathrm{spt}}(T \llcorner B^n_r(p))= \Sigma^+)\). Since \(T\llcorner B^n_r(p) = [[\Sigma^+]]\),
\(\partial\Sigma^+ = \Gamma\cap B^n_r(p)\), and \(T\) is area-minimizing in \(U\), \(\Sigma^+\) is a \(C^{3,\alpha}\) minimal submanifold in \(B^n_r(p)\). By 21(4), we can choose \(r\)
smaller if necessary so that \(\Sigma^+\) is the graph of a stationary solution \(u \in C^\omega(\Omega; \mathbb{R}^{n-k}) \cap C^{3,\alpha}(\overline{\Omega};\mathbb{R}^{n-k})\) to 5 on some simply connected domain \(\Omega\subset \mathbb{R}^k\) with \(C^{3,\alpha}\) boundary.
Now, the current \(T^\prime\) also satisfies the hypotheses of 21 so it inherits the same regularity as \(T\).
Therefore, we can choose \(r\) smaller once again to ensure \(T^\prime \llcorner B^n_r(p) = [[\Xi^+]]\) for some \(C^{3,\alpha}\) minimal submanifold \(\Xi \subset B^n_r(p)\) as in 20 which satisfies \(\partial\Xi^+ = \Gamma\cap B^n_r(p)\). By assumption, \(\partial\Sigma^+ = \partial\Xi^+\) and \(T_q \Sigma^+ = T_q \Xi^+\) whenever \(q \in \Gamma^\prime \cap B^n_r(p)\) since these planes are precisely the
approximate tangent spaces for \(\mathop{\mathrm{spt}}T\) and \(\mathop{\mathrm{spt}}T^\prime\) at \(q\). We can therefore assume that \(\Xi^+\) can be represented as the graph of a function \(v: \overline{\Omega} \rightarrow \mathbb{R}^{n-k}\) with the same regularity as \(u\) satisfying 5 . Let \[\Gamma^\star := \psi^{-1}(\Gamma^\prime \cap B_r^n(p)) \subset \partial\Omega, \text{ where } \psi(x) := (x, u(x)).\] Then \(\Gamma^\star\) is the projection of
\(\Gamma^\prime \cap B^n_r(p)\) onto \(\partial\Omega\) and \(u = v\) on \(\Gamma^\star \subset \partial\Omega\). Moreover,
as the tangent planes to \(\Sigma^+\) and \(\Xi^+\) coincide with each other along the points in \(\Gamma^\prime \cap B_r^n(p)\), we see that \(Du = Dv\) on \(\Gamma^\star\). In particular, \(\partial_\nu u = \partial_\nu v\) on \(\Gamma^\star\), where \(\nu\) is the unit outer normal to \(\Gamma^\star\). We conclude that \(u\) and \(v\) have the same Cauchy data along the
relatively open \(C^{3,\alpha}\) boundary patch \(\Gamma^\star \subset \partial\Omega\), and thus \(u \equiv v\) throughout \(\overline{\Omega}\) by 3 and 4. Together with 21(3), 18, and 19, this
implies that \(\mathop{\mathrm{Reg}}_i(T) = \mathop{\mathrm{Reg}}_i(T^\prime)\) as oriented submanifolds. In fact, \(T = T^\prime\) as integral \(k\)-currents, since 21(5) gives \(\Theta_T \equiv 1 \equiv \Theta_{T^\prime}\) on \(\mathop{\mathrm{Reg}}_i(T) = \mathop{\mathrm{Reg}}_i(T^\prime)\), and \(\mathcal{H}^k(\mathop{\mathrm{Sing}}_i(T) \cup \mathop{\mathrm{Sing}}_i(T^\prime)) = 0\) by Almgren’s interior regularity
theorem. ◻
Remark 22. Let \(U \subset \mathbb{R}^n\) be an unbounded open connected set. As 21(1) (see also [12]) still holds, 5 can of course be extended to integral currents \(T \in
\mathcal{I}_k(\mathbb{R}^n)\) with possibly unbounded supports and unbounded boundaries \(\Gamma\) which are area-minimizing in \(U\) if we know in advance that
There is an open set of one-sided regular points in \(\mathop{\mathrm{Reg}}_b(T)\).
\(\mathop{\mathrm{Reg}}_i(T)\) is connected.
\(\Theta_{T} \equiv 1\) on \(\mathop{\mathrm{Reg}}_i(T)\).
By 17, the third condition follows from the first two.
Remark 23. When \(T\) has higher multiplicity on the boundary, i.e. \(\partial T = Q [[\Gamma]]\) for some \(Q \in \mathbb{N}\setminus
\{1\}\), Fleschler and Resende [48] shows that \(\mathop{\mathrm{Reg}}_b(T)\) is open and dense in \(\Gamma\). In particular, when \(\Gamma\) is connected and both \(\Gamma\) and \(T\) are compactly supported, \(\mathop{\mathrm{Reg}}_b(T)\) contains only one-sided points (cf. [48]). Therefore, it seems possible to obtain a similar result as 5 if we know in advance that \(\mathop{\mathrm{Reg}}_i(T)\) is connected and \(\Theta_T \equiv Q\) on \(\mathop{\mathrm{Reg}}_i(T)\). As in the previous remark, the last condition should hold automatically if the first one holds.
4Uniqueness of compactly supported calibrated currents↩︎
We are ready to prove 2. This is accomplished by showing that the hypotheses of 5 are met for any such
area-minimizing currents. We start with two general lemmas that are only related to calibrations. Suppose \(e_1,\ldots, e_n\) is an orthonormal basis for \(\mathbb{R}^n\). Let \(1 \leq k \leq n\). For each \(i = 1,\ldots, k\), we define \[e_i \, \lrcorner \, (e_1 \wedge \cdots \wedge e_k) \mathrel{\vcenter{:}}= (-1)^{i-1} e_1 \wedge \cdots
\wedge\widehat{e_i} \wedge \cdots \wedge e_k.\] Hence, for each \(i\) we have \[e_i \wedge(e_i \, \lrcorner \, (e_1 \wedge \cdots \wedge e_k)) = e_1 \wedge\cdots \wedge e_k.\] The
following lemma has been adapted from [49], and is a consequence of the fact that a comass one \(k\)-form \(\varphi\in \mathcal{E}^k(\mathbb{R}^n)\) achieves its pointwise maximum on its contact set. Since the proof is standard, we omit the details.
Lemma 24 (First Cousin Principle). Suppose \(\varphi \in \mathcal{E}^k(\mathbb{R}^n)\) has comass one and let \(e_1,\ldots,e_n\) be an orthonormal basis for \(\mathbb{R}^n\). For each \(i = 1,\ldots, n\) and \(j = 1,\ldots, n-k\), set \[\begin{align} \xi &:= e_1 \wedge \cdots \wedge e_k
\in G(k,\mathbb{R}^n)\\ \xi_{ij} &:= e_{k+j} \wedge (e_i \, \lrcorner \, (e_1 \wedge \cdots \wedge e_k)) \in G(k,\mathbb{R}^n).
\end{align}\] If \(\left\langle{\varphi_p},{\xi}\right\rangle = 1\) for some \(p \in \mathbb{R}^n\), then \(\left\langle{\varphi_p},{\xi_{ij}}\right\rangle =
0\) for all \(i\) and \(j\).
For fixed \(i,j\), we call \(\xi_{ij}\) a first cousin of\(\xi\). The next two lemmata allow us to locally set up a Cauchy problem for the MSS
if it is known that a calibrated current exists for a given \(C^{3,\alpha}\) boundary.
Lemma 25 (Intersecting Planes Lemma). Let \(\varphi \in \mathcal{E}^k(\mathbb{R}^n)\) have comass one, fix \(p \in \mathbb{R}^n\), and suppose \(\{e_1, \ldots, e_{k-1}\} \subset \mathbb{R}^n\) is an orthonormal set. If \[\eta := e_1 \wedge \cdots \wedge e_{k-1} \in G(k-1,\mathbb{R}^n)\] and \(v \in
\mathbb{S}^{n-1} \cap \eta^\perp\) makes \[\eta_v := e_1 \wedge \cdots \wedge e_{k-1} \wedge v \in \mathcal{G}_p(\varphi), \label{etav}\qquad{(2)}\] then \(v\) is unique. In particular, if for some \(p \in \mathbb{R}^n\) the oriented planes \(\xi_1,\xi_2 \in \mathcal{G}_p(\varphi)\) satisfy \(\xi_1 \cap \xi_2 \in G(k-1,\mathbb{R}^n)\), then \(\xi_1 = \xi_2\).
Proof. Let \(\eta\) be as in the statement of the lemma. Suppose that \(v_1, v_2 \in \mathbb{S}^{n-1} \cap \eta^\perp\) have been chosen so that \(\eta_{v_1}, \eta_{v_2} \in \mathcal{G}_p(\varphi)\). Write \(v_2 := c_1 v_1 + c_2 v_1^\perp\), where \(v_1^\perp \in \eta_{v_1}^\perp\) and \(c_i \in \mathbb{R}\) for each \(i = 1, 2\). Set \(e_k := v_1\) and choose \(e_{k+1},\ldots, e_{n}\) so that \(\{e_1,\ldots, e_n\}\) is an orthonormal basis for \(\mathbb{R}^n\). Then \[v_1^\perp = \sum_{j = 1}^{n-k} a_{j} e_{k+j}\] for some \(\{a_j \in \mathbb{R}\}_{j=1}^{n-k}\). Since \(\eta_{v_1}, \eta_{v_2} \in \mathcal{G}_p(\varphi)\), 24
shows \[1 = \left\langle{\varphi_p},{\eta_{v_2}}\right\rangle = c_1 + c_2 \sum_{j = 1}^{n-k} a_{j} \left\langle{\varphi_p},{\xi_{kj}}\right\rangle = c_1,\] because \(\xi_{kj}\) are first
cousins of \(\eta_{v_1}\). Since \(v_2 \in \mathbb{S}^{n-1}\), we see that \(v_2 = v_1\). This proves the first statement, and the second is immediate from
the first. ◻
Given a calibrated integral current with compact support, all the area-minimizing compactly supported integral currents which have the same boundary must also be calibrated by the same form. This is proved using 12 and 9.
Lemma 26. Assume \(U\subset \mathbb{R}^n\) and \(W \subset \subset U\) are open sets (not necessarily connected). Let \(\varphi\in
\mathcal{E}^k(\mathbb{R}^n)\) be a calibration. Suppose that \(T \in \mathcal{I}_{k, c}(\mathbb{R}^n)\) is calibrated by \(\varphi\) in \(U\) with
\(\mathop{\mathrm{spt}}T \subset W\). If \(T^\prime \in \mathcal{I}_{k, c}(\mathbb{R}^n)\) is area-minimizing in \(U\) with \(\partial T^\prime = \partial T\) and \(\mathop{\mathrm{spt}}T^\prime \subset W\), then \(T^\prime\) is also calibrated by \(\varphi\) in \(U\).
Proposition 27. Assume \(U\subset \mathbb{R}^n\) and \(W \subset \subset U\) are open connected sets. Suppose that \(T \in
\mathcal{I}_{k,c}(\mathbb{R}^n)\) and \(\Gamma\subset W\) are as in 1. Let \(\varphi\in
\mathcal{E}^k(\mathbb{R}^n)\) be a calibration. Assume that \(T\) is calibrated by \(\varphi\) in \(U\). If \(T^\prime \in
\mathcal{I}_{k, c}(\mathbb{R}^n)\) is area-minimizing in \(U\) with \(\partial T^\prime = [[\Gamma]]\) and \(\mathop{\mathrm{spt}}T^\prime \subset
W\), then \(T^\prime = T\).
Proof. According to 26, \(T^\prime\) is also calibrated by \(\varphi\) in
\(U\). By 21(1) and (2), there exist \(p \in \Gamma\) and \(r > 0\) such that
\(\Gamma_r := B^n_r(p) \cap \Gamma\subset \mathop{\mathrm{Reg}}_b(T) \cap \mathop{\mathrm{Reg}}_b(T^\prime)\) consists of one-sided regular boundary points. Let \(\Sigma\) and \(\Xi\) be \(k\)-dimensional smooth submanifolds of \(\mathbb{R}^n\) satisfying \(\mathop{\mathrm{spt}}T \cap B^n_r(p) =
\Sigma^+\) and \(\mathop{\mathrm{spt}}T^\prime \cap B^n_r(p) = \Xi^+\), which can be done due to the discussion following 20. Then for all
\(q \in \Gamma_r\), the approximate tangent spaces for \(\mathop{\mathrm{spt}}T\) and \(\mathop{\mathrm{spt}}T^\prime\) at \(q\) match \(T_q \Sigma^+\) and \(T_q\Xi^+\), respectively. Moreover, \(T_q \Sigma^+, \, T_q \Xi^+ \in \mathcal{G}(\varphi)\) for
all \(q \in \Gamma_r\) by 26. Since \(\partial\Sigma^+ = \partial\Xi^+ = \Gamma_r\), the
intersection \(T_q \Sigma^+ \cap T_q \Xi^+\) must contain the \((k-1)\)-dimensional subspace \(T_q\Gamma_r\) for each \(q \in
\Gamma_r\). Then 25 shows that \(T_q \Sigma^+ = T_q \Xi^+\) for all \(q \in \Gamma_r\). Applying 5, we deduce that \(T \llcorner U = T^\prime \llcorner U\). Since \(\mathop{\mathrm{spt}}T \subset U\) and \(\mathop{\mathrm{spt}}T^\prime \subset U\), we conclude that \(T = T^\prime\). ◻
In the special case \(U = \mathbb{R}^n\), we simply choose \(W \subset \subset \mathbb{R}^n\) satisfying \(\mathop{\mathrm{spt}}T \cup
\mathop{\mathrm{spt}}T^\prime \subset W\) to obtain 2. As a result, we have proved uniqueness in the oriented Plateau problem for many important classes of area-minimizing
submanifolds (e.g. the examples in Section 2). While existence of solutions to the oriented Plateau problem is well understood, it is not clear when a calibrated current spans a given boundary. Thus, for now we must apply 2 case by case to determine uniqueness. Notice that 27 can be used to rule out the existence of calibrated currents spanning
a fixed boundary.
Corollary 28 (Non-Existence of Calibrated Currents). Let \(W \subset \subset U \subset \mathbb{R}^n\) be open connected sets and suppose \(\Gamma\subset W\) satisfies
the hypotheses in 1. Suppose there are distinct currents \(T,T^\prime \in \mathcal{I}_{k,c}(\mathbb{R}^n)\) with support in \(W\) such that:
Both \(T\) and \(T^\prime\) are area-minimizing in \(U\).
\(\partial T = [[\Gamma]] = \partial T^\prime\).
Then there is no \(S \in \mathcal{I}_{k,c}(\mathbb{R}^n)\) with \(\mathop{\mathrm{spt}}S \subset W\) satisfying \(\partial S = [[\Gamma]]\) that is
calibrated in \(U\).
Remark 29 (Sharpness of Hypotheses). The “compact support assumption" is crucial. The simplest counterexample is half-planes which have boundary equal to the axis but are at a non-zero angle: Consider \(\Sigma\mathrel{\vcenter{:}}= \{(x, 0, z) : x \geq 0, z \in \mathbb{R}\} \subset \mathbb{R}^3\) and \(\Sigma^\prime \mathrel{\vcenter{:}}= \{(0, y, z) : y \geq 0, z \in \mathbb{R}\} \subset
\mathbb{R}^3\). Note that \([[\Sigma]]\) is calibrated by \(dx \wedge dz\) in \(\mathbb{R}^3\) and \([[\Sigma^\prime]]\) is calibrated by \(dy \wedge dz\) in \(\mathbb{R}^3\), so they are both globally area-minimizing and share the same boundary \([[z\text{-axis}]]\). Though \(\Sigma\) and \(\Sigma^\prime\) are diffeomorphic as smooth manifolds with boundary, \([[\Sigma]]\)
and \([[\Sigma^\prime]]\) are certainly not the same as integral currents.
In \(\mathbb{R}^3\), the half-plane, half of the helicoid, and half of the Enneper surface (each positioned so that they share the same boundary line) serve as another counterexample: They are all globally
area-minimizing (see [50], [51], or [52]). However, they are neither equal as integral currents nor diffeomorphic to one another as smooth manifolds with boundary. Thus, the compact support assumption restricts the angle
the tangent planes at the boundary of a calibrated integral current can form with the support of the boundary current giving uniqueness, provided we have enough boundary regularity. Nonetheless, each of the aforementioned examples is the unique
area-minimizer with its Cauchy data by 5 and 22.
One can define calibrations on a general Riemannian manifold. A Riemannian manifold together with a calibration is called a calibrated manifold. There are many examples of calibrated manifolds (see e.g. [29]). However, in contrast to 13, a calibrated current in a calibrated manifold
can only be assumed to be homologically area-minimizing (cf. [29]). Nonetheless, all the results in Section 3.2–4 and Appendix A carry over to a general
ambient manifold \(M\) that is complete without boundary, analytic, and satisfies \(H_k(M; \mathbb{R}) = 0\).
The last two conditions are essential. For analyticity, our arguments heavily rely on the fact that \(C^1\) minimal submanifolds in an analytic ambient manifold are actually analytic by the bootstrapping argument (see
also 18), so that 19 holds. For the homological condition, a general ambient
manifold \(M\) may contain a closed \(k\)-dimensional submanifold \(\Sigma\) calibrated by a \(k\)-form \(\varphi\). For example, \(\Sigma= \mathbb{S}^n\) is a closed special Lagrangian submanifold in \(T^*\mathbb{S}^n\) with the Stenzel metric (see e.g. [53] or [54]). By chopping \(\Sigma\)
into two submanifolds in a suitable way, it gives counterexamples to 13 and shows the non-uniqueness in the oriented Plateau problem. However, by
assuming \(H_k(M; \mathbb{R}) = 0\), statements similar to 13 and 26 hold. The main reason is that we rule out the existence of closed calibrated \(k\)-dimensional submanifolds in \(M\)
by the universal coefficient theorem and de Rham’s theorem.
6Uniqueness of restrictions of global area-minimizers↩︎
In this section, we show that if \(T \in \mathcal{I}_{k, c}(\mathbb{R}^n)\) is a global area-minimizer arising as the restriction of another global area-minimizer \(T^\prime \in \mathcal{I}_{k,
c}(\mathbb{R}^n)\), then \(T\) is the unique global area-minimizer among all compactly supported integral currents with the same boundary as \(T\). The authors believe that it must be
known to experts in the field (particularly when stated as 32). Due to the lack of a precise reference, the proof is included for completeness.
Theorem 30. Let \(T \in \mathcal{I}_{k, c}(\mathbb{R}^n)\) be globally area-minimizing. Suppose that there exists a globally area-minimizing integral \(k\)-current
\(T^\prime \in \mathcal{I}_{k, c}(\mathbb{R}^n)\) with compact support such that \(T^\prime \llcorner \mathop{\mathrm{spt}}T = T\) and \(\mathop{\mathrm{spt}}T
\subset \mathop{\mathrm{spt}}T^\prime \setminus \mathop{\mathrm{spt}}(\partial T^\prime)\). Then \(T\) is the unique global area-minimizer among all compactly supported integral \(k\)-currents with the same boundary as \(T\).
Proof. Suppose that \(S \in \mathcal{I}_{k,c}(\mathbb{R}^n)\) is another globally area-minimizing integral \(k\)-current with compact support that satisfies \(\partial S = \partial T\). Then the assumptions on \(T^\prime\) show that \[\label{eq1} \mathbb{M}(T^\prime) \leq \mathbb{M}(T^\prime -
T + S) \leq \mathbb{M}(T^\prime - T) + \mathbb{M}(S) \leq \mathbb{M}(T^\prime - T) + \mathbb{M}(T) = \mathbb{M}(T^\prime).\tag{13}\] In other words, \(T^\prime - T + S\) is a global area-minimizing integral
\(k\)-current with compact support which has the same boundary as \(T^\prime\).
We next claim that \(\mathop{\mathrm{spt}}T^\prime\) and \(\mathop{\mathrm{spt}}(T^\prime - T + S)\) agree \(\mathcal{H}^k\)-a.e. on \(\mathop{\mathrm{spt}}T^\prime \setminus \mathop{\mathrm{spt}}T\). Suppose otherwise. Then there exists a relatively open, non-empty subset \(\mathscr{O}\) of \(\mathop{\mathrm{spt}}T^\prime \setminus \mathop{\mathrm{spt}}T\) such that \(S\) cancels out \(T^\prime - T\) on \(\mathscr{O}\). That is, \(S \llcorner \mathscr{O} = -(T^\prime - T)\llcorner \mathscr{O} = -T^\prime \llcorner \mathscr{O}.\) We then compute the mass \[\begin{align}
\mathbb{M}(T^\prime - T + S) &= \mathbb{M}(T^\prime - T + S + S\llcorner\mathscr{O} - S\llcorner\mathscr{O}) \\ &\leq \mathbb{M}(T^\prime - T - T^\prime \llcorner \mathscr{O}) + \mathbb{M}(S - S\llcorner\mathscr{O}) \\ &< \mathbb{M}(T' -
T) + \mathbb{M}(S) \\ &= \mathbb{M}(T^\prime).\end{align}\] This contradicts 13 , and hence the claim follows.
Near \(\mathop{\mathrm{Reg}}_i(T^\prime) \cap \partial T\), the aforementioned claim together with an argument similar to 19
applied to each connected component shows that \(\mathop{\mathrm{Reg}}_i(T^\prime) = \mathop{\mathrm{Reg}}_i(T^\prime - T + S)\). Moreover, by 17 and the assumption that \(\partial T = \partial S\), the densities agree on each connected component. In addition, Almgren’s interior regularity theorem implies that the singular sets
are \(\mathcal{H}^k\)-null. It follows that \(T^\prime = T'- T + S\) near \(\mathop{\mathrm{Reg}}_i(T^\prime) \cap \partial T\). In other words, \[T \llcorner \mathscr{U} = S \llcorner \mathscr{U}\] for some open neighborhood \(\mathscr{U}\) of \(\mathop{\mathrm{Reg}}_i(T^\prime) \cap \partial T\) in \(\mathop{\mathrm{spt}}T\).
Finally, we claim that every connected component of \(\mathop{\mathrm{Reg}}_i(T)\) and \(\mathop{\mathrm{Reg}}_i(S)\) must intersect \(\mathscr{U}\)
non-trivially. If not, we may assume that there is a connected component \(M\) of \(\mathop{\mathrm{Reg}}_i(T)\) such that \(M \cap \mathscr{U} =
\emptyset\). Then a cutoff argument—essentially the same as that used in the third passage of the proof of [55], with \(d-5\) and \(p < 5\) replaced by \(k-2\) and \(p < 2\), respectively—shows that \(\partial(T
\llcorner M) = 0\). This violates the assumption that \(T\) is area-minimizing. The same reasoning applies to \(S\), and thus the claim follows. With this claim established, the
unique continuation argument in the previous paragraph can therefore be extended to the whole \(\mathop{\mathrm{Reg}}_i(T^\prime) \cap \mathop{\mathrm{spt}}T\), yielding \[T \llcorner
(\mathop{\mathrm{Reg}}_i(T^\prime) \cap \mathop{\mathrm{spt}}T) = S \llcorner (\mathop{\mathrm{Reg}}_i(T^\prime) \cap \mathop{\mathrm{spt}}T).\] As \(\mathop{\mathrm{Sing}}_i(T^\prime)\) is \(\mathcal{H}^k\)-null, the proof is complete. ◻
Remark 31. In contrast to 27, 30 does not require regularity or connectivity of
\(\partial T\).
Let \(C \subset \mathbb{R}^n\) be a \(k\)-dimensional cone (e.g. Simons’ cone or the LOC). Then \(C\) is called regular if \(C \setminus \{p\} = \mathop{\mathrm{Reg}}_i(C)\) for some \(p \in \mathbb{R}^n\). For such a \(p\), we call \(C \cap
\mathbb{S}^{n-1}(p)\) its link. We have the following corollary for area-minimizing regular cones:
Corollary 32. Let \(C\) be a \(k\)-dimensional regular cone in \(\mathbb{R}^n\), and suppose that \([[C]]\) is globally area-minimizing in \(\mathbb{R}^n\). Then \([[C]] \llcorner B^n_1(p)\) is the unique global area-minimizer bounded by its link.
32 proves the corresponding claim in the fourth paragraph of Section 7 in [39].
J. Douglas. Solution of the problem of Plateau. Trans. Amer. Math. Soc., 33(1):263–321, 1931.
[2]
T. Radó. On Plateau’s problem. Ann. of Math. (2), 31(3):457–469, 1930.
[3]
C. B. Morrey, Jr. The problem of Plateau on a Riemannian manifold. Ann. of Math. (2), 49:807–851, 1948.
[4]
H. Federer and W. H. Fleming. Normal and integral currents. Ann. of Math. (2), 72:458–520, 1960.
[5]
J. C. C. Nitsche. Contours bounding at least three solutions of Plateau’s problem. Arch. Rational Mech. Anal., 30:1–11, 1968.
[6]
F. Morgan. A smooth curve in \(R\sp{4}\) bounding a continuum of area minimizing surfaces. Duke Math. J., 43(4):867–870, 1976.
[7]
F. Morgan. Almost every curve in \(\textbf{R}\sp{3}\) bounds a unique area minimizing surface. Invent. Math., 45(3):253–297,
1978.
[8]
F. Morgan. Generic uniqueness for hypersurfaces minimizing the integral of an elliptic integrand with constant coefficients. Indiana Univ. Math. J., 30(1):29–45, 1981.
[9]
F. Morgan. Measures on spaces of surfaces. Arch. Rational Mech. Anal., 78(4):335–359, 1982.
[10]
W. K. Allard. On the first variation of a varifold: boundary behavior. Ann. of Math. (2), 101:418–446, 1975.
[11]
G. Caldini, A. Marchese, A. Merlo, and S. Steinbrüchel. Generic uniqueness for the Plateau problem. J. Math. Pures Appl. (9), 181:1–21, 2024.
[12]
C. De Lellis, G. De Philippis, J. Hirsch, and A. Massaccesi. On the boundary behavior of mass-minimizing integral currents. Mem. Amer. Math. Soc., 291(1446):v+166, 2023.
[13]
J. C. C. Nitsche. A new uniqueness theorem for minimal surfaces. Arch. Rational Mech. Anal., 52:319–329, 1973.
[14]
F.-H. Lin. Uniqueness and nonuniqueness of the Plateau problem. Indiana Univ. Math. J., 36(4):843–848, 1987.
[15]
V. Nandakumaran and G. Székelyhidi. Uniqueness in the plateau problem near quadratic cones, 2025. arXiv preprint: 2509.15503.
[16]
R. M. Hardt. Uniqueness of nonparametric area minimizing currents. Indiana Univ. Math. J., 26(1):65–71, 1977.
[17]
V. Adolfsson and L. Escauriaza. domains and unique continuation at the boundary. Comm. Pure Appl. Math., 50(10):935–969, 1997.
[18]
N. Aronszajn, A. Krzywicki, and J. Szarski. A unique continuation theorem for exterior differential forms on Riemannian manifolds. Ark. Mat., 4:417–453, 1962.
[19]
N. Aronszajn. A unique continuation theorem for solutions of elliptic partial differential equations or inequalities of second order. J. Math. Pures Appl. (9), 36:235–249,
1957.
[20]
N. Garofalo and F.-H. Lin. Monotonicity properties of variational integrals, \(A_p\) weights and unique continuation. Indiana Univ.
Math. J., 35(2):245–268, 1986.
[21]
N. Garofalo and F.-H. Lin. Unique continuation for elliptic operators: a geometric-variational approach. Comm. Pure Appl. Math., 40(3):347–366, 1987.
[22]
H. Koch and D. Tataru. Carleman estimates and unique continuation for second-order elliptic equations with nonsmooth coefficients. Comm. Pure Appl. Math., 54(3):339–360,
2001.
[23]
A. Pliś. On non-uniqueness in Cauchy problem for an elliptic second order differential equation. Bull. Acad. Polon. Sci. Sér. Sci. Math. Astronom. Phys., 11:95–100,
1963.
[24]
K. Miller. Nonunique continuation for uniformly parabolic and elliptic equations in self-adjoint divergence form with Hölder continuous coefficients. Arch. Rational
Mech. Anal., 54:105–117, 1974.
[25]
B. Dimler. Partial regularity for Lipschitz solutions to the minimal surface system. Calc. Var. Partial Differential Equations, 62(9):Paper No. 260, 30, 2023.
[26]
F. J. Almgren, Jr. Almgren’s big regularity paper, volume 1 of World Scientific Monograph Series in Mathematics. World Scientific Publishing Co., Inc., River Edge, NJ,
2000. \(Q\)-valued functions minimizing Dirichlet’s integral and the regularity of area-minimizing rectifiable currents up to codimension 2, With a preface by Jean E. Taylor and Vladimir Scheffer.
[27]
R. Hardt and L. Simon. Boundary regularity and embedded solutions for the oriented Plateau problem. Ann. of Math. (2), 110(3):439–486, 1979.
[28]
L. Simon. Introduction to geometric measure theory. NTU Lectures, 2018.
[29]
R. Harvey and H. B. Lawson, Jr. Calibrated geometries. Acta Math., 148:47–157, 1982.
[30]
E. Giusti. Minimal surfaces and functions of bounded variation, volume 80 of Monographs in Mathematics. Birkhäuser Verlag, Basel, 1984.
[31]
R. Hardt, C. Lau, and F. Lin. Non-minimality of minimal graphs. Indiana University Mathematics Journal, 36(4):849–855, 1987.
[32]
H. B. Lawson, Jr. and R. Osserman. Non-existence, non-uniqueness and irregularity of solutions to the minimal surface system. Acta Math., 139(1-2):1–17, 1977.
[33]
H. Federer. Geometric measure theory, volume Band 153 of Die Grundlehren der mathematischen Wissenschaften. Springer-Verlag New York, Inc., New York, 1969.
[34]
R. Tione. Minimal graphs and differential inclusions. Comm. Partial Differential Equations, 46(6):1162–1194, 2021.
[35]
M. Giaquinta and L. Martinazzi. An introduction to the regularity theory for elliptic systems, harmonic maps and minimal graphs, volume 11 of Appunti. Scuola Normale
Superiore di Pisa (Nuova Serie) [Lecture Notes. Scuola Normale Superiore di Pisa (New Series)]. Edizioni della Normale, Pisa, second edition, 2012.
[36]
C. B. Morrey, Jr. Multiple integrals in the calculus of variations, volume Band 130 of Die Grundlehren der mathematischen Wissenschaften. Springer-Verlag New York,
Inc., New York, 1966.
[37]
D. Gilbarg and N. S. Trudinger. Elliptic partial differential equations of second order. Classics in Mathematics. Springer-Verlag, Berlin, 2001. Reprint of the 1998
edition.
[38]
X. Xu, L. Yang, and Y. Zhang. Dirichlet boundary values on Euclidean balls with infinitely many solutions for the minimal surface system. J. Math. Pures Appl. (9),
129:266–300, 2019.
[39]
B. Dimler. Minimal submanifolds with multiple isolated singularities, 2025. arXiv preprint: 2409.20327.
[40]
J. L. M. Barbosa. An extrinsic rigidity theorem for minimal immersions from \(S\sp{2}\) into \(S\sp{n}\). J. Differential Geometry, 14(3):355–368, 1979.
[41]
D. Fischer-Colbrie. Some rigidity theorems for minimal submanifolds of the sphere. Acta Math., 145(1-2):29–46, 1980.
[42]
Q. Han, N. Nadirashvili, and Y. Yuan. Linearity of homogeneous order-one solutions to elliptic equations in dimension three. Comm. Pure Appl. Math., 56(4):425–432, 2003.
[43]
C. Mooney. Partial regularity for singular solutions to the Monge-Ampère equation. Comm. Pure Appl. Math., 68(6):1066–1084, 2015.
[44]
T. H. Colding and W. P. Minicozzi, II. A course in minimal surfaces, volume 121 of Graduate Studies in Mathematics. American Mathematical Society, Providence, RI,
2011.
[45]
C. De Lellis and E. Spadaro. Regularity of area minimizing currents I: gradient \(L^p\) estimates. Geom. Funct. Anal.,
24(6):1831–1884, 2014.
[46]
C. De Lellis and E. Spadaro. Regularity of area minimizing currents II: center manifold. Ann. of Math. (2), 183(2):499–575, 2016.
[47]
C. De Lellis and E. Spadaro. Regularity of area minimizing currents III: blow-up. Ann. of Math. (2), 183(2):577–617, 2016.
[48]
I. Fleschler and R. Resende. On the regularity of area minimizing currents at boundaries with arbitrary multiplicity, 2025. arXiv preprint: 2410.04566.
[49]
F. R. Harvey. Spinors and calibrations, volume 9 of Perspectives in Mathematics. Academic Press, Inc., Boston, MA, 1990.
[50]
B. White. Half of Enneper’s surface minimizes area. In Geometric analysis and the calculus of variations, pages 361–367. Int. Press, Cambridge, MA, 1996.
[51]
J. Pérez. Stable embedded minimal surfaces bounded by a straight line. Calc. Var. Partial Differential Equations, 29(2):267–279, 2007.
[52]
N. Edelen and Z. Wang. A Bernstein-type theorem for minimal graphs over convex domains. Ann. Inst. H. Poincaré C Anal. Non Linéaire, 39(3):749–760, 2022.
[53]
M. B. Stenzel. Ricci-flat metrics on the complexification of a compact rank one symmetric space. Manuscripta Math., 80(2):151–163, 1993.
[54]
C.-J. Tsai and M.-T. Wang. Mean curvature flows in manifolds of special holonomy. J. Differential Geom., 108(3):531–569, 2018.
[55]
N. Edelen, L. Spolaor, and B. Velichkov. A strong maximum principle for minimizers of the one-phase bernoulli problem. Indiana Univ. Math. J., 73:1061–1096, 2024.
That is, a linear system having no leading-order coupling.↩︎
Formally, \(\langle \varphi_p,\xi \rangle := \varphi_p(\xi)\) for all \(\xi \in \Lambda_k(\mathbb{R}^n)\).↩︎
A simple illustration of the argument for uniformly convex functionals is outlined at the beginning of [35], though a similar argument works for
strongly rank-one convex functionals.↩︎