Lang LocalCurrents
Lang LocalCurrents
Urs Lang∗
November 7, 2008; revised May 5, 2010
Abstract
Ambrosio and Kirchheim presented a theory of currents with finite
mass in complete metric spaces. We develop a variant of the theory that
does not rely on a finite mass condition, closely paralleling the classi-
cal Federer–Fleming theory. If the underlying metric space is an open
subset of a Euclidean space, we obtain a natural chain monomorphism
from general metric currents to general classical currents whose image
contains the locally flat chains and which restricts to an isomorphism
for locally normal currents. We give a detailed exposition of the slic-
ing theory for locally normal currents with respect to locally Lipschitz
maps, including the rectifiable slices theorem, and of the compactness
theorem for locally integral currents in locally compact metric spaces,
assuming only standard results from analysis and measure theory.
Contents
Introduction 2
1 Preliminaries 7
2 Metric currents 10
4 Mass 22
5 Normal currents 31
6 Slicing 38
References 61
∗
Supported by the Swiss National Science Foundation.
1
2 U. Lang
Introduction
Currents in the sense of geometric measure theory, linear functionals on spaces
of differential forms, were introduced by G. de Rham in 1955 for use in the
theory of harmonic forms [12]. A few years later, H. Federer and W. H. Flem-
ing devised the class of rectifiable currents, generalized oriented surfaces with
integer multiplicities, and the subclass of integral currents, whose boundary
is of the same type. Their fundamental paper [15] from 1960 also furnished
the compactness theorem for integral currents and thereby a solution to the
Plateau problem for surfaces of arbitrary dimension and codimension in Eu-
clidean spaces. The theory of currents then rapidly developed into a powerful
apparatus in the calculus of variations. Federer’s monograph [14] gives a com-
prehensive account of the state of the subject prior to 1970. Since then, the
theory has been extended in various directions and has found numerous ap-
plications in geometric analysis and Riemannian geometry, far beyond pure
area minimization problems.
A breakthrough was achieved in 2000, when L. Ambrosio and B. Kirch-
heim [3], following ideas of E. De Giorgi [7], presented a theory of currents in
complete metric spaces. This elegant approach employs (m+1)-tuples of real-
valued Lipschitz functions in place of differential m-forms and provides some
new insights even if the ambient space is Euclidean. Ambrosio and Kirchheim
discovered new proofs of the boundary rectifiability theorem and the closure
theorem for rectifiable currents, valid in any complete metric space. As an
application, they obtained existence results for generalized Plateau problems
in compact metric spaces and certain Banach spaces. The theory of metric
currents has been further developed in [8], [32], [34], and some geometric ap-
plications have been found [33], [35]. Various interactions with other areas
have emerged and should be further explored. In this context we just mention
the functions of bounded higher variation [21], [8], [9], [28], [25], the recent
progress on (generalized) flat chains [37], [38], [18], [1], the scans from [16],
[10], [11], the differentiation theory on metric measure spaces [4], [22], [5],
and the derivations from [30], [31]. We refer to [19] for an excellent survey of
some of these and further related topics.
The metric currents considered by Ambrosio and Kirchheim have finite
mass by definition. This a priori assumption plays a crucial role in their
development of the theory, in particular it is used to derive the properties that
qualify the functionals under consideration as analogues of classical currents.
Here we present a somewhat different approach, strongly inspired by the work
of Ambrosio and Kirchheim, but not relying on a finite mass condition. In
addition, we switch to local formulations. At a later stage, this enables us to
discuss metric currents with locally finite mass and locally rectifiable currents.
For this purpose it is appropriate to assume the underlying metric space to
be locally compact. However, once some basic properties are established, the
theory readily extends to a theory of currents with locally compact support
in arbitrary metric spaces. Furthermore, in the case of finite mass, it is also
Local currents in metric spaces 3
T (f, gπ1 , π2 , . . . , πm ) = T (f g, π1 , . . . , πm ) + T (f π1 , g, π2 , . . . , πm ).
We also obtain a chain rule, a special case of which states that if (f, π) =
(f, π1 , . . . , πm ) ∈ D m (X) and g = (g1 , . . . , gm ) ∈ [C 1,1 (Rm )]m , i.e., the partial
derivatives of gi are locally Lipschitz, then
for all (f, g1 , . . . , gm ) ∈ Cc∞ (U )×[C ∞ (U )]m . This makes the aforesaid guiding
principle rigorous. Some more properties of these comparison maps Cm are
mentioned further below.
In Sect. 3 we define the support spt(T ) ⊂ X of a metric current T and
discuss the boundary and push-forward operators. For a classical m-current
T̄ , the boundary ∂ T̄ is the (m−1)-current satisfying ∂ T̄ (φ) = T̄ (dφ) for every
(m − 1)-form φ. Correspondingly, the boundary ∂T ∈ Dm−1 (X) of a metric
current T ∈ Dm (X) verifies
for all (f, π1 , . . . , πm−1 ) ∈ D m−1 (X) and for all σ such that σ = 1 on some
neighborhood of spt(f ). We have ∂ ◦ ∂ = 0, and, in case X = U is an open
set in Rn , ∂ ◦ Cm = Cm−1 ◦ ∂, so that the Cm form a chain map. The push-
forward F# T̄ of a general classical current T̄ is defined for every smooth map
F whose restriction to the support of T̄ is proper, and considerable efforts
are required to extend the definition, for particular classes of currents, to
locally Lipschitz maps. Given a metric current T ∈ Dm (X) and another
locally compact metric space Y , the push-forward F# T ∈ Dm (Y ) is defined
for every locally Lipschitz map F : D → Y such that spt(T ) ⊂ D ⊂ X and
F |spt(T ) is proper. In case D = X, we have
F# T (f, π1 , . . . , πm ) = T (f˜, π1 ◦ F, . . . , πm ◦ F )
(T b B)(f, π) = T (χB f, π)
for every open set V ⊂ X with compact closure, then there is a subse-
quence Tn(1) , Tn(2) , . . . that converges weakly to some T ∈ Dm (X), i.e.,
limi→∞ Tn(i) (f, π) = T (f, π) for every (f, π) ∈ D m (X). Since MV is lower
semicontinuous with respect to weak convergence, and also ∂Tn(i) → ∂T
weakly, the limit T is locally normal. In case X = U is an open set in Rn , the
restriction of the comparison map Cm to the vector space of locally normal
currents is an isomorphism onto the space of classical locally normal currents.
An important technique in the theory of currents consists in relating in-
formation on the structure of a current T to properties of lower dimensional
slices of T in the level sets of a function. In Sect. 6 we discuss slicing of a
locally normal current T ∈ Dm (X) with respect to a locally Lipschitz map
π : X → Rk , where 1 ≤ k ≤ m. Let Tπ ∈ Dm−k (X) be the locally normal
current satisfying
for all (f, g) ∈ D m−k (X). If spt(T ) is separable, then for almost every y ∈ Rk
there is a locally normal current in Dm−k (X) with support in π −1 {y}∩spt(T ),
denoted by hT, π, yi, such that
Z
hT, π, yi(f, g) dy = Tπ (f, g)
Rk
for all (f, g) ∈ D m−k (X). Moreover, for every Borel set B ⊂ X,
Z
khT, π, yik(B) dy = kTπ k(B).
Rk
The first identity also holds more generally if f is a bounded Borel function
with compact support.
6 U. Lang
By combining this result with the compactness theorem for locally normal
currents mentioned earlier, we obtain the compactness theorem for locally
integral currents.
Further results and applications will be discussed elsewhere.
1 Preliminaries
We now fix the notation and collect some basic facts from analysis and mea-
sure theory. A few more prerequisites will be discussed in individual sections.
Given a point x in a metric space X = (X, d), B(x, r) := {y : d(x, y) ≤ r}
and U(x, r) := {y : d(x, y) < r} denote the closed and open ball, respectively,
with center x and radius r.
for every set A ⊂ X. Then it is also true that if B is a ν-measurable set with
ν(B) < ∞, and if > 0, then ν(B \ K) < for some compact set K ⊂ B
(cf. [14, 2.2.5]).
For m ∈ N, we denote by
αm := L m (B(0, 1))
µ(B(x, r))
Mµ (x) := sup .
r>0 αm r m
for all s > 0. In particular, Mµ (x) < ∞ for L m -almost every x ∈ Rm . Note
also that Mµ is lower semicontinuous on Rm .
We further recall that, given a function u ∈ L1 (Rm ), L m -almost every
x ∈ Rm is a Lebesgue point of u, i.e.,
Z
1
lim |u(y) − u(x)| dy = 0.
r→0 αm r m B(x,r)
1.5 Smoothing
We shall use the following basic facts regarding smoothing. Let η ∈ Cc∞ (Rm )
R a mollifier, so that spt(η) ⊂ U(0, 1),1 η(−z)
be = η(z) ≥ 0 for all z ∈ Rm , and
m
Rm η(z) dz = 1. Recall that for g ∈ Lloc (R ), the convolution defined by
Z Z
(η ∗ g)(x) := η(z)g(x − z) dz = η(x − z)g(z) dz
Rm Rm
Dk (η ∗ g) = η ∗ Dk g, (1.5)
2 Metric currents
From now on, unless otherwise stated, X will always denote a locally compact
metric space. We write A b X if A ⊂ X and the closure of A is compact.
We let D(X) be the algebra of all f ∈ Lip(X) whose support spt(f ) is
compact; these will serve as test functions. For every compact set K ⊂ X
Local currents in metric spaces 11
and every constant l ≥ 0 we put LipK,l (X) := {f ∈ Lipl (X) : spt(f ) ⊂ K},
so that D(X) is the union of all LipK,l (X). Then we equip D(X) with a
locally convex vector space topology τ with respect to which
fj → f in D(X) (2.1)
if and only if all fj belong to some fixed LipK,l (X) and fj → f pointwise
on X for j → ∞, which implies that fj → f uniformly on X. Explicitly,
this topology τ is given as follows. Let β be the collection of all absolutely
convex sets W ⊂ D(X) with the following property: For every pair (K, l)
and every f ∈ W ∩ LipK,l (X), there is an > 0 such that g ∈ W whenever
g ∈ LipK,l (X) and kf − gk∞ < . (Recall that W is absolutely convex if and
only if sf + tg ∈ W for all f, g ∈ W and s, t ∈ R with |s| + |t| ≤ 1.) Then
β is a local base of τ at 0, thus τ is the collection of all unions of sets of
the form f + W , where f ∈ D(X) and W ∈ β. (See e.g. [27, p. 152] for the
corresponding construction in classical distribution theory.)
Similarly, we equip Liploc (X) with a locally convex vector space topology
with respect to which
πj → π in Liploc (X) (2.2)
if and only if for every compact set K ⊂ X there is a constant lK such that
Lip(πj |K ) ≤ lK for all j and πj → π pointwise, hence uniformly, on K for
j → ∞. For a fixed compact set K ⊂ X, let βK be the collection of all
absolutely convex sets W ⊂ Liploc (X) with the following property: For every
l ≥ 0 and every π ∈ W with Lip(π|K ) ≤ l, there is an > 0 such that ρ ∈ W
whenever ρ ∈ Liploc (X), Lip(ρ|K ) ≤ l, and k(π − ρ)|K k∞ < . The union
of all βK forms a local base of the topology of Liploc (X). The topologies of
D(X) and Liploc (X) will only be used through (2.1) and (2.2).
We define the spaces
which will serve as substitutes for the spaces of compactly supported m-forms.
The guiding principle is that
(f j , π1j , . . . , πm
j
) → (f, π1 , . . . , πm ) in D m (X) (2.4)
Dm (X).
We endow Dm (X) with the locally convex weak topology with respect to which
Tn → T if and only if
Tn (f, π1 , . . . , πm ) → T (f, π1 , . . . , πm )
Proof. In view of conditions (1) and (3), the function T can be extended to
D m (X) so that
Definition 2.3. For T ∈ Dm (X) and (u, v) ∈ Liploc (X)×[Liploc (X)]k , where
m ≥ k ≥ 0, we define the current T b (u, v) ∈ Dm−k (X) by
T (f, π1 , . . . , πm ) = 0.
T (f, gh, π2 , . . . , πm ) = T (f g, h, π2 , . . . , πm ) + T (f h, g, π2 , . . . , πm ).
14 U. Lang
Proof. To prove (1), it suffices to show that if T ∈ D2 (X) and (f, π) ∈ D 1 (X),
then T (f, π, π) = 0. For k ∈ Z, let ρk : R → R be the piecewise affine 1-
P function with ρk |[2k,2k+1] = 1 and spt(ρk ) = [2k − 1, 2k + 2]. Note
Lipschitz
that k∈Z ρk = 1. Let σ, σ̄ : R → R denote the piecewise affine 4-Lipschitz
functions such that σ|spt(ρk ) = 2k for k even and σ̄|spt(ρk ) = 2k for k odd.
Then X
T (f, σ ◦ π, σ̄ ◦ π) = T ((ρk ◦ π)f, σ ◦ π, σ̄ ◦ π);
k∈Z
note that (ρk ◦ π)f = 0 for almost all k since π|spt(f ) is bounded. By the strict
locality property (2.5), each summand is zero because σ ◦π or σ̄ ◦π is constant
on spt(ρk ◦ π) for k even or odd, respectively. Hence T (f, σ ◦ π, σ̄ ◦ π) = 0.
In the above definitions of the functions ρk , σ, σ̄ we may equally well replace
the unit by 1/j, for j ∈ N. The argument then shows that
T (f, σj ◦ π, σ̄j ◦ π) = 0,
where σj (s) = σ(js)/j and σ̄j (s) = σ̄(js)/j for s ∈ R. Letting j tend to
∞, we obtain T (f, π, π) = 0 by the continuity of T , since σj ◦ π → π and
σ̄j ◦ π → π in Liploc (X).
For the proof of (2), it suffices to show that if T ∈ D1 (X) and (f, g) ∈
D (X), then T (f, g 2 ) = 2T (f g, g). Let ρk , σ, σ̄ be defined as above. Since
1
σ ◦ g|spt(ρk ◦g) = 2k for k even and σ̄ ◦ g|spt(ρk ◦g) = 2k for k odd, and since
(ρk ◦ g)f = 0 for almost all k, the multilinearity of T and (2.5) give
T (f, (σ ◦ g)(σ̄ ◦ g))
X
= T ((ρk ◦ g)f, (σ ◦ g)(σ̄ ◦ g))
k∈Z
X X
= 2kT ((ρk ◦ g)f, σ̄ ◦ g) + 2kT ((ρk ◦ g)f, σ ◦ g)
k even k odd
X
= 2kT ((ρk ◦ g)f, σ ◦ g + σ̄ ◦ g)
k∈Z
= T ((τ ◦ g)f, (σ + σ̄) ◦ g)
P
for the piecewise affine 2-Lipschitz function τ := k∈Z 2kρk , which satisfies
τ |[2k,2k+1] = 2k for k ∈ Z. Rescaling by the factor 1/j, as in the proof of (1),
we obtain the identity
T (f, (σj ◦ g)(σ̄j ◦ g)) = T ((τj ◦ g)f, (σj + σ̄j ) ◦ g),
where τj (s) = τ (js)/j for s ∈ R. Taking the limit for j → ∞ we conclude
that T (f, g 2 ) = T (gf, 2g).
We now deduce a chain rule, which subsumes both the alternating prop-
erty and the case g = h of Proposition 2.4(2). For an open set U ⊂
Rn , C 1,1 (U ) denotes the space of all g ∈ C 1 (U ) with partial derivatives
D1 g, . . . , Dn g ∈ Liploc (U ). For n ≥ m ≥ 1, we let Λ(n, m) be the set of
all strictly increasing maps λ : {1, . . . , m} → {1, . . . , n}.
Local currents in metric spaces 15
In case m = n, we obtain
for all (f, g) ∈ D(U ) × [C 1,1 (U )]m . We now arrive at the first family of
examples of metric currents, corresponding to the integration of a simple
m-form over U ⊂ Rm .
[u](f j , g j ) − [u](f, g)
m
[u](f, g1 , . . . , gi−1 , hji , gi+1
j
X
j j j
= [u](f − f, g ) + , . . . , gm ).
i=1
Since the sequence (det(Dg j ))j∈N is bounded in L∞ (V ), the first term on the
right side clearly tends to 0 for j → ∞. Now consider the summand for i = 1;
the other summands are treated similarly. Since uf ∈ L1 (V ), we want to
show that Z
lim v det(D(hj1 , g2j , . . . , gm
j
)) dx = 0 (2.12)
j→∞ V
If hj1 , g2j , . . . , gm
j
∈ C 2 (V ), then d(whj1 dg2j ∧ . . . ∧ dgm
j
R
V ) = 0 and hence
Z Z
w dhj1 ∧ dg2j ∧ ... ∧ j
dgm =− hj1 dw ∧ dg2j ∧ . . . ∧ dgm
j
,
V V
The definition is justified by the next lemma, whose proof employs Lip-
schitz partitions of unity. This is made possible by the fact that the functions
in D(X) have compact support. In [3], the support of a current T is defined
as the support of the associated finite Borel measure kT k.
(1) The support spt(T ) equals the set of all x ∈ X such that for every > 0
there exists an (f, π) ∈ D m (X) with spt(f ) ⊂ B(x, ) and T (f, π) 6= 0.
Proof. Let Σ be the set described in (1). Suppose that x 6∈ spt(T ). There
is a closed set C with the property stated in Definition 3.1 such that x 6∈ C.
Then there is an > 0 such that T (f, π) = 0 whenever spt(f ) ⊂ B(x, ). This
shows that x 6∈ Σ, so Σ ⊂ spt(T ).
Next we prove that T (f, π) = 0 whenever spt(f ) ∩ Σ = ∅. Since spt(f ) is
a compact subset of S X \ Σ, there exist finitely many open balls U1 , . . . , UN
such that spt(f ) ⊂ N k=1 Uk and T (g, π) = 0 whenever {g 6= 0} ⊂ Uk for
some k. Decomposing f by means of a Lipschitz partition of unity (ρk )N k=1
on spt(f ) with {ρk 6= 0} ⊂ Uk we see that T (f, π) = 0. As Σ is closed, this
shows in particular that spt(T ) ⊂ Σ.
For (2), let βj be the function defined after (2.5), j ∈ N. If f |spt(T ) = 0,
then spt(βj ◦ f ) ∩ spt(T ) = ∅. The argument of the previous paragraph then
shows that T (βj ◦ f, π) = 0 for all j, thus T (f, π) = 0 by the continuity of T .
To prove (3), by the linearity, locality, and the alternating property of T
it suffices to show that T (f, π1 , . . . , πm ) = 0 if π1 = 0 on K := spt f |spt(T ) .
Then spt(βj ◦ π1 ) ∩ K = ∅ for βj as above, j ∈ N. For fixed j, since K is
compact, there is a function σ ∈ D(X) such that σ|K = 1 and βj ◦ π1 = 0 on
some neighborhood of spt(σ). Then (1 − σ)f |spt(T ) = 0, hence
T (f, βj ◦ π1 , π2 , . . . , πm ) = T (σf, βj ◦ π1 , π2 , . . . , πm ) = 0
Proof. For every compact set K ⊂ A, every l ≥ 0, and every c > 0 there exist
a compact set K 0 ⊂ X containing K, an l0 ≥ l, and an operator
E : f ∈ LipK,l (A) : kf k∞ ≤ c → LipK 0 ,l0 (X)
cf. Sect. 1.1. This has the required properties, with K 0 = spt(σ) and l0 =
l+c Lip(σ). Now the result follows easily from Lemma 2.2 and Lemma 3.2.
T ∈ Dm (X) (3.2)
for (f, π1 , . . . , πm−1 ) ∈ D m−1 (X), where σ ∈ D(X) is any function such that
σ = 1 on {f 6= 0} ∩ spt(T ).
Proof. Let (f, g) ∈ D m−k−1 (X), and choose σ ∈ D(X) with σ|spt(f ) = 1.
Then
((∂T )b (u, v))(f, g) = ∂T (uf, v, g)
= T (σ, uf, v, g)
= T (σf, u, v, g) + T (σu, f, v, g)
= T (f, u, v, g) + (−1)k T (σu, v, f, g)
= (T b (1, u, v))(f, g) + (−1)k (T b (u, v))(σ, f, g)
= (T b (1, u, v))(f, g) + (−1)k ∂(T b (u, v))(f, g);
the third step uses Proposition 2.4(2), the fourth 2.4(1).
Local currents in metric spaces 21
F# T (f, π1 , . . . , πm ) := TA (f ◦ F, π1 ◦ F, . . . , πm ◦ F )
If m ≥ 1, then
∂(F# T ) = F# (∂T ). (3.9)
To see this, let (f, π) ∈ D m−1 (Y ), and choose σ ∈ D(Y ) such that σ|spt(f ) = 1;
then
this is consistent with the above definition in case D is locally compact and
F is proper. When F ∈ Liploc (X, Y ) and F |spt(T ) is proper, it follows that
for h(x) := u(x)f (F (x)) det(Dπ(F (x))) sign(det(DF (x))). By the change of
variables formula (cf. [14, Theorem 3.2.3(2)] or [13, 3.3.3], the case n = m),
Z X Z
F# [u](f, π) = h(x) dy = v(y)f (y) det(Dπ(y)) dy
Rm Rm
x∈F −1 {y}
= [v](f, π).
4 Mass
We now define the mass of a metric current. Our approach is inspired by both
the classical definition (recalled in (5.2)) and [3, Proposition 2.7]. Currents
with locally finite mass will be of particular interest.
Definition 4.1 (mass). For T ∈ Dm (X), m ≥ 0, and every open set V ⊂ X,
we define the mass MV (T ) of T in V as the least number M ∈ [0, ∞] such
that X
T (fλ , π λ ) ≤ M
λ∈Λ
Furthermore, we define
kT k(A) = MA (T ). (4.2)
Proof. Let (Tk )k∈N be a Cauchy sequence in (Mm (X), M). For every > 0
there exists an index k such that
|T (f j , π j ) − T (f, π)|
≤ |T (f j , π j ) − Tk (f j , π j )| + + |Tk (f, π) − T (f, π)| ≤ 3.
24 U. Lang
Hence T is a current.
P Whenever Λ is a finite set, (fλ , π λ ) ∈ D(X)×[Lip1 (X)]m
for λ ∈ Λ, and λ∈Λ |fλ | ≤ 1, we have
X
(Tk − Tl )(fλ , π λ ) ≤ M(Tk − Tl ) ≤
λ∈Λ
(for the last step, see e.g. [26, Theorem 8.16]). Now let T ∈ Mm,loc (X),
m ≥ 1. Let first (f, π) ∈ D(X) × [Lip1 (X)]m , and consider Tπ := T b (1, π) ∈
D0 (X). Then kTπ k ≤ kT k, thus Tπ ∈ M0,loc (X), and
Z Z
|T (f, π)| = |Tπ (f )| ≤ |f | dkTπ k ≤ |f | dkT k.
X X
Finally, given (f, π) ∈ D m (X), m
exists π̃ ∈ [Lip(X)] such that π̃ = π on
there
spt(f ) and Lip(π̃i ) = Lip πi |spt(f ) for i = 1, . . . , m. Then T (f, π) = T (f, π̃)
by the strict locality property (2.5), and the result follows.
26 U. Lang
Recall from Sect. 1.2 that we denote by Bc∞ (X) the algebra of all bounded
Borel functions f : X → R such that spt(f ) is compact. From Theorem 4.3(4)
it follows that every T ∈ Mm,loc (X), m ≥ 0, naturally extends to a function
To see this, note that D(X) is dense in L1 (kT k) (since Cc (X) is, see e.g. [26,
3.14], and since every element of Cc (X) is a uniform limit of a sequence of
Lipschitz functions, cf. Sect. 1.1). Hence, whenever f ∈ Bc∞ (X) ⊂ L1 (kT k)
and U b X is a neighborhood of spt(f ), there is a sequence (gk )k∈N in D(X)
such that gk → f in L1 (kT k) and spt(gk ) ⊂ U for all k. By Theorem 4.3(4),
for every π ∈ [Liploc (X)]m , (T (gk , π))k∈N is a Cauchy sequence whose limit is
independent of the choice of (gk )k∈N . Then T (f, π) is defined to be this limit.
theorem. For the second term Ron the right side, given > 0, choose g ∈ D(X)
such that spt(g) ⊂ U and lm X |f − g| dkT k < /3. For j sufficiently large,
|T (g, π j ) − T (g, π)| < /3 by the continuity of T , hence
Definition 4.5. For T ∈ Mm,loc (X) and (u, v) ∈ Bloc∞ (X) × [Lip (X)]k ,
loc
where m ≥ k ≥ 0, we define T b (u, v) ∈ Mm−k,loc (X) by the same equation
as in Definition 2.3. For a Borel set B ⊂ X,
T b B := T b χB .
(1) For all (f, π) ∈ Bc∞ (Y ) × [Liploc (Y )]m and any σ ∈ Bc∞ (X) such that
σ = 1 on {f ◦ F 6= 0} ∩ spt(T ),
F# T (f, π) = T (σ(f ◦ F ), π ◦ F ).
Hence F# T ∈ Mm,loc (Y ).
To prove (1), fix π ∈ [Liploc (Y )]m and ρ ∈ D(Y ), ρ ≥ 0. Choose τ ∈ D(X)
such that τ = 1 on {ρ ◦ F 6= 0} ∩ spt(T ), and denote by Φ the set of all
f ∈ Bc∞ (Y ) such that |f | ≤ ρ and F# T (f, π) = T (τ (f ◦ F ), π ◦ F ). It
follows from Theorem 4.4(2) that Φ is a Baire class. By (3.12), Φ contains all
f ∈ D(Y ) with |f | ≤ ρ and therefore consists of all f ∈ Bc∞ (Y ) with |f | ≤ ρ,
cf. Sect. 1.2. In view of (4.7), this gives (1).
For (2), suppose (f, π) ∈ D(Y ) × [Lip1 (Y )]m , and let σ be the character-
istic function of F −1 (B) ∩ {f ◦ F 6= 0} ∩ spt(T ). By (1) and Theorem 4.4(4),
In particular,
kT k({x}) = |T (χ{x} )| (4.12)
for every x ∈ X.
The following simple example illustrates the σ-finiteness assumption in the
lemma. Let R be any uncountable discrete space, and equip X = R × (−1, 1)
with the metric d defined by
(
0 0 |s − s0 | if r = r0 ,
d((r, s), (r , s )) :=
1 if r 6= r0 .
Note that X is locally compact. Let T ∈ M1,loc (X) be the current satisfying
Z
dπ
T (f, π) = f (r, s) (r, s) dH 1 (r, s)
X ds
Proof of Lemma 4.7. We show the first part. By Theorem 4.4(4), the inequal-
ity always holds. To see that kT k(B) is the least number with this property,
let > 0, and choose an open set V such that B ⊂ V and kT k(V \ B) ≤ .
Note that this is possible by the assumption on B. Let α < kT k(V ). Then
there exist Λ and (fλ , π λ ) as in the definition of MV (T ) such that
X X X
α≤ T (fλ , π λ ) = T (χB fλ , π λ ) + T (χV \B fλ , π λ )
λ∈Λ λ∈Λ λ∈Λ
≤ kT k(B) + .
Furthermore, by (4) and Lemma 4.7, for every Borel set B ⊂ X, kT k(B) is
the least number such that
X
T (fλ , π λ ) ≤ kT k(B) (4.14)
λ∈Λ
whenever Λ is a finite set, (fλ , π λ ) ∈ B ∞ (X) × [Lip1 (X)]m , and λ∈Λ |fλ | ≤
P
χB .
Suppose now, for the remaining part of this section, that X is an arbitrary
metric space. Combining the extension just described with the discussion
of (3.2), we obtain a corresponding normed space (Mm (X), M) of currents
with finite mass and locally compact support in X. By definition, an element
T of Mm (X) is a functional on B ∞ (XT ) × [Lip(XT )]m for some closed and
locally compact set XT ⊂ X. However, T may now equally well be viewed as
a functional on B ∞ (X) × [Lip(X)]m (compare (3.1) and Proposition 3.3). A
current
T ∈ Mm (X)
is then a function as in (4.13) satisfying (1)–(4) (now for arbitrary X), where
kT k is a finite Borel regular outer measure that is concentrated on its locally
compact support spt(kT k) and characterized by (4.14). Since kT k is finite,
spt(kT k) is separable (cf. [14, Theorem 2.2.16]) and hence also σ-compact.
In contrast to Proposition 4.2, the space (Mm (X), M) of all such T , where
M(T ) = kT k(X), is no longer complete in general.
We now denote by (MAK m (X), M) the Banach space of all m-currents in
the sense of Ambrosio and Kirchheim, viewed as functionals on B ∞ (X) ×
[Lip(X)]m , cf. [3, Theorem 3.5]. A current
T ∈ MAK
m (X)
Local currents in metric spaces 31
5 Normal currents
We turn to the chain complex of normal currents. In the first part of this
section we prove the compactness theorem for locally normal metric currents.
Then we compare metric currents in an open set U ⊂ Rn with classical cur-
rents, in particular we establish an isomorphism for locally normal currents.
Definition 5.1 (normal current). For T ∈ Dm (X) and every open set V ⊂
X, define
NV (T ) := MV (T ) + MV (∂T )
if m ≥ 1 and NV (T ) := MV (T ) if m = 0, and let N(T ) := NX (T ). The
vector space
Nm,loc (X)
of m-dimensional locally normal currents in X consists of all T ∈ Dm (X)
such that NV (T ) < ∞ for all open sets V b X. An m-dimensional normal
current in X is an element of
for every open set V ⊂ X. Together with (4.10), this shows that T b (u, v) ∈
Nm−k,loc (X). By Lemma 4.6 and (3.9), push-forwards of locally normal cur-
rents are locally normal.
The following result corresponds to [3, Proposition 5.1].
(2) For all (f, g), (f˜, g̃) ∈ D(X) × [Lip1 (X)]m ,
Z
˜
T (f, g) − T (f , g̃) ≤ |f − f˜| dkT k
X
m
X Z Z
+ Lip(f ) |gi − g̃i | dkT k + |f ||gi − g̃i | dk∂T k .
i=1 spt(f ) X
Proof. For the proof of (1) we assume m = 1. Let (f, g) ∈ D 1 (X), and choose
σ ∈ D(X) with σ|spt(f ) = 1. By the product rule,
hence Z Z
|T (f, g)| ≤ Lip(f ) |g| dkT k + |f g| dk∂T k.
spt(f ) X
Lemma 5.2 yields the following proposition, which will be used repeatedly
in the sequel.
|Tn (f, g) − Tn (f˜, g̃)| ≤ R(f, g, f˜, g̃)N(Tn ) ≤ R(f, g, f˜, g̃)M.
Fix (f, g) ∈ Lip1 (X) × [Lip1 (X)]m for the moment. Given > 0, there is
(f˜, g̃) ∈ F × F m such that R(f, g, f˜, g̃)M ≤ , and there is an index n0 such
that |Tn (f˜, g̃)−Tn0 (f˜, g̃)| ≤ for all n, n0 ≥ n0 . Then |Tn (f, g)−Tn0 (f, g)| ≤ 3
for all n, n0 ≥ n0 , so (Tn (f, g))n∈N is a Cauchy sequence. Define T : Lip1 (X)×
[Lip1 (X)]m → R such that
for all (f, g) ∈ Lip1 (X) × [Lip1 (X)]m . It follows that |T (f, g) − T (f˜, g̃)| ≤
R(f, g, f˜, g̃)M for all (f, g), (f˜, g̃) ∈ Lip1 (X) × [Lip1 (X)]m . Now it is clear
that T extends uniquely to a current T ∈ Dm (X), and Tn → T weakly. By
the lower semicontinuity of mass, T ∈ Nm (X).
for all φ ∈ D̄ m (U ), cf. [14, 4.1.5 and p. 349] and [29, 26.4–26.8]. The seminorm
N̄ on D̄m (U ) and the space N̄m,loc (U ) of locally normal currents are defined
in analogy to Definition 5.1. For a compact set K ⊂ U , N̄m,K (U ) denotes
the set of all T ∈ D̄m (U ) with spt(T ) ⊂ K and N̄(T ) < ∞, cf. [14, p. 358].
The flat seminorm of a form φ ∈ D̄ m (U ) relative to a compact set K ⊂ U
is given by
F̄K (φ) = sup supx∈K kφ(x)k, supx∈K kdφ(x)k ,
Note that F̄K (∂T ) ≤ F̄K (T ) for m ≥ 1. If F̄K (T ) < ∞, then spt(T ) ⊂ K. If
T ∈ D̄m (U ) and spt(T ) ⊂ K, then
for all S ∈ D̄m+1 (U ) with spt(S) ⊂ K, and equality holds for at least one such
S. The F̄K -closure of N̄m,K (U ) in D̄m (U ) is denoted by F̄m,K (U ). The space
F̄m (U ) of flat chains with compact support in U is the union of all F̄m,K (U ).
The space F̄m,loc (U ) of locally flat chains consists of all T ∈ D̄m (U ) such that
T b σ ∈ F̄m (U ) for every σ ∈ Cc∞ (U), cf. [14, 4.1.12].
Finally, we note that the set T ∈ F̄m,K (U ) : M̄(T ) < ∞ equals the
M̄-closure of N̄m,K (U ) in D̄m (U ), cf. [14, 4.1.17], and
F̄m,K (U ) = R + ∂S :R ∈ F̄m,K (U ), M̄(R) < ∞,
S ∈ F̄m+1,K (U ), M̄(S) < ∞ , (5.3)
For currents with finite mass and compact support in Rn this result was
proved by Ambrosio and Kirchheim in [3, Theorem 11.1]. They conjectured
that the image under Cm of {T ∈ Mm (Rn ) : spt(T ) is compact} coincides
with the space of m-dimensional flat chains with finite mass and compact
support in Rn . In view of (4) and the many analogous properties of Dm (U )
and F̄m,loc (U ), one may similarly ask whether Cm (Dm (U )) = F̄m,loc (U ).
Proof. In case m > n, Dm (U ) = {0} by (2.10), and also D̄m (U ) = {0}. Thus
Cm is the trivial map in this case.
In case m = 0, given T ∈ D0 (U ), C0 (T ) is the functional satisfying
C0 (T )(f ) = T (f ) for all f ∈ D̄ 0 (U ) = Cc∞ (U ). The continuity property
of T implies that C0 (T ) is sequentially continuous on D̄ 0 (U ). This yields
C0 (T ) ∈ D̄0 (U ) (cf. [27, Theorem 6.6]).
P let T ∈ Dm (U ), 1 ≤ m ≤ n. To define∞Cm (T ), write φ ∈ D̄ (U ) as
Now m
X
Cm (T )(φ) = T (φλ , πλ(1) , . . . , πλ(m) ) = T (f, g1 , . . . , gm )
λ∈Λ(n,m)
n
Hence M̄V (Cm (T )) ≤ m MV (T ). This yields (2).
We prove (3). From (1) and (2) it follows that Cm (T ) ∈ N̄m,loc (U ) if and
only if T ∈ Nm,loc (U ). Hence, since Cm is injective, it suffices to construct
a map C̄m : N̄m,loc (U ) → Dm (U ) such that Cm (C̄m (T̄ )) = T̄ for all T̄ ∈
N̄m,loc (U ). Let T̄ ∈ N̄m,loc (U ). We first observe that whenever (f, g), (f˜, g̃) ∈
Cc∞ (U ) × [C ∞ (U ) ∩ Lip1 (U )]m , then
Z
˜
T̄ (f dg1 ∧ . . . ∧ dgm ) − T̄ (f dg̃1 ∧ . . . ∧ dg̃m ) ≤ |f − f˜| dkT̄ k
U
m
X Z Z
+ Lip(f ) |gi − g̃i | dkT̄ k + |f ||gi − g̃i | dk∂ T̄ k ; (5.4)
i=1 spt(f ) U
this is just the “classical” analogue of Lemma 5.2(2). To define C̄m (T̄ ), let
(f, g) ∈ D(U ) × [Lip1 (U )]m , and choose a sequence ((f k , g k ))k∈N in Cc∞ (U ) ×
[C ∞ (U )∩Lip1 (U )]m such that f k → f in D(U ) and gik → gi locally uniformly
for i = 1, . . . , m. It follows from (5.4) that (T̄ (f k dg1k ∧ . . . ∧ dgm
k ))
k∈N is a
Cauchy sequence whose limit is independent of the choice of the sequence
((f k , g k ))k∈N . Put
Let (f˜, g̃) ∈ D(U ) × [Lip1 (U )]m be another such tuple. By choosing the
approximating sequences appropriately, we see that (5.4) holds in the limit,
i.e., C̄m (T̄ )(f, g) − C̄m (T̄ )(f˜, g̃) is less than or equal to the expression on
the right side of (5.4). Now it is clear that C̄m (T̄ ) extends to a current
C̄m (T̄ ) ∈ Dm (U ) satisfying
for all (f, g) ∈ Cc∞ (U )×[C ∞ (U )]m . As the left side of this last equality equals
Cm (C̄m (T̄ ))(f dg1 ∧ . . . ∧ dgm ), we have Cm (C̄m (T̄ )) = T̄ .
It remains to prove (4). First we observe that for every compact set
K ⊂ U , the restriction of C̄m to N̄m,K (U ) naturally extends to a map
38 U. Lang
from the set T̄ ∈ F̄m,K (U ) : M̄(T̄ ) < ∞ , which equals the M̄-closure of
N̄m,K (U ) in D̄m (U ), into Mm (U ). Since the restriction of C̄m is 1-Lipschitz
with respect to M̄ and M, this follows by the completeness of (Mm (U ), M)
(cf. Proposition 4.2). For every T̄ ∈ F̄m,K (U ) with M̄(T̄ ) < ∞, we still
have Cm (C̄m (T̄ )) = T̄ . Next, let T̄ ∈ F̄m,K (U ). Choose R̄ ∈ F̄m,K (U ) with
M̄(R̄) < ∞ and S̄ ∈ F̄m+1,K (U ) with M̄(S̄) < ∞ such that T̄ = R̄ + ∂ S̄,
cf. (5.3), and put C̄m (T̄ ) := C̄m (R̄) + ∂(C̄m+1 (S̄)) ∈ Dm (U ). Then
Cm (C̄m (T̄ )) = Cm (C̄m (R̄)) + ∂ Cm+1 (C̄m+1 (S̄)) = T̄ .
Since Cm is injective, this identity also shows that C̄m (T̄ ) is well-defined.
Finally, suppose that T̄ ∈ F̄m,loc (U ). Given (f, g1 , . . . , gm ) ∈ D m (U ), choose
σ ∈ Cc∞ (U ) such that σ = 1 on some neighborhood of spt(f ) and put
C̄m (T̄ )(f, g1 , . . . , gm ) := C̄m (T̄ b σ)(f, g1 , . . . , gm ). It is readily checked that
this yields a well-defined C̄m (T̄ ) ∈ Dm (U ) with Cm (C̄m (T̄ )) = T̄ .
for φ ∈ D̄ 2 (R4 ). Note that these currents have boundary zero. To compute
the mass, use the following inequality: If v, w are vectors in R4 with Euclidean
norm |v|, |w| ≤ 1, then
6 Slicing
We now develop the slicing theory for locally normal m-currents in X with
respect to a locally Lipschitz map π : X → Rk , where 1 ≤ k ≤ m. The slices
are currents of dimension m − k in the level sets of π. General references
are [14, 4.2.1 and 4.3.1–5], [29, 28.6–28.10] for classical currents and [3, pp. 31–
36] for metric currents of finite mass. We first treat the case k = 1, which
suffices for most geometric applications. The general case will be relevant in
Sect. 7 and in Theorems 8.4 and 8.5.
Local currents in metric spaces 39
respectively.
Note that since T ∈ Mm,loc (X) and ∂T ∈ Mm−1,loc (X), the restrictions
T b B and (∂T )b B are defined for every Borel set B ⊂ X. Moreover, by
Theorem 4.4(1), T b B + T b (X \ B) = T . If (kT k + k∂T k)(π −1 {s}) = 0, then
T b π −1 {s} = 0 and (∂T )b π −1 {s} = 0, thus hT, π, s−i = hT, π, s+i. It follows
that if spt(T ) is separable, then
for every s ∈ R.
There is a useful characterization of hT, π, s+i as a weak limit. For this
we approximate the characteristic function χ{π>s} by a family (us,δ )δ>0 in
Liploc (X) such that
hT, π, s+i − T b (1, us,δ ) = (∂T )b (χ{π>s} − us,δ ) − ∂(T b (χ{π>s} − us,δ )).
It follows that
hT, π, s+i = lim T b (1, us,δ ) (6.4)
δ→0+
(3) For every kT b (1, π)k-measurable set B ⊂ X with kT b (1, π)k(B) < ∞,
Z
khT, π, s+ik(B) ds = kT b (1, π)k(B).
R
(4) If spt(T ) is separable, then hT, π, s+i ∈ Nm−1,loc (X) for almost every
s ∈ R. If π|spt(T ) is proper, or if T ∈ Nm (X) and π ∈ Lip(X), then
hT, π, s+i ∈ Nm−1 (X) for almost every s ∈ R.
Proof. For every δ > 0, fix a non-decreasing function γδ ∈ C 1,1 (R) such that
γδ |(−∞,0] = 0 and γδ |[δ,∞) = 1. Let s ∈ R. Define γs,δ (t) := γδ (t − s) for t ∈ R,
and put us,δ := γs,δ ◦ π. By Theorem 2.5 (chain rule),
0 0
T b (1, us,δ ) = T b (γs,δ ◦ π, π) = (T b (1, π))b (γs,δ ◦ π).
Now we let δ tend to 0. We know that then T (f, us,δ , g) → hT, π, s+i(f, g)
for every s ∈ R. Moreover, since |us,δ | ≤ 1, Lemma R 5.2(1) yields the
uniform bound |T (f, us,δ , g)| ≤ Lip(f )kT k(spt(f )) + X |f | dk∂T k. Hence,
s 7→ hT, π, s+i(f, g) is a bounded Borel function with compact support, and
Z ∞ Z ∞ Z
lim T (f, us,δ , g) ds = hT, π, s+i(f, g) ds = hT, π, s+i(f, g) ds.
δ→0+ a a R
Local currents in metric spaces 41
R∞ R t−a
On the other hand, the functions t 7→ a γs,δ (t) ds = −∞ γδ (r) dr are 1-
Lipschitz for all δ > 0 and converge uniformly to t 7→ max{t − a, 0}, as δ → 0.
It follows that
Z ∞
lim T f, us,δ ds, g = T (f, π − a, g) = T (f, π, g) = (T b (1, π))(f, g).
δ→0+ a
for the lower integral is a direct consequence of (2). This shows (3) for every
open set V with kT b (1, π)k(V ) < ∞. Since every compact set K ⊂ X is a
difference of two such open sets, the same identity holds for all compact sets,
and we obtain (3) by approximation.
Finally, (4) follows easily from (3), or just (6.5), together with the corre-
sponding result for ∂hT, π, s+i = −h∂T, π, s+i in case m ≥ 2.
for every y ∈ R. (We could equally well arrange that hT, π, yi = (hT, π, y−i +
hT, π, y+i)/2, as in [14, 4.3.4], by choosing a more symmetric definition of
hT, π, yi.) Before proving the existence of slices of locally normal currents in
the case k > 1 we discuss a few properties that can be derived directly from
the definition.
Let y = (y1 , . . . , yk ) ∈ Rk , and suppose that hT, π, yi ∈ Dm−k (X) exists.
For δ > 0, consider the cube
so that
spt(hT, π, yi) ⊂ π −1 {y} ∩ spt(T ). (6.8)
If m > k, then (∂T )b (1, γy,δ ◦ π) = (−1)k ∂(T b (1, γy,δ ◦ π)) by (3.7), hence
(1) For L k -almost every y ∈ Rk , the slice hT, π, yi ∈ Dm−k (X) exists and
is locally normal.
Reasoning as in the proof of Theorem 6.2 and using again the σ-compactness
of spt(T ) we obtain
Z
kTy k(B) dy = kT b (1, π)k(B) (6.11)
Rk
44 U. Lang
for all (f, g) ∈ Bc∞ (X)×[Liploc (X)]m−k . Given φ ∈ Bloc ∞ (Rk ), we may replace
Now we show that for almost every y ∈ Rk , the slice hT, π, yi exists as
an element of Nm−k,loc (X) and coincides with Ty . By (6.12) and (6.11), this
will complete the proof of the theorem.
We first assume that X is compact, so that T is normal and π is Lipschitz.
Then we fix a countable set F ⊂ Lip1 (X) that is dense in Lip1 (X) with
respect to k · k∞ . Put Tπ := T b (1, π) ∈ Nm−k (X) for the moment, and let
y ∈ Rk and δ > 0. Using (6.7), (4.7), and Theorem 4.4 we get
1 1
= (−1)k k (∂T )b (1, π) b π −1 (C(y, δ)) = k (∂Tπ )b π −1 (C(y, δ)).
δ δ
For every Borel set C ⊂ Rk , put
if m > k, and µ(C) := kTπ k(π −1 (C)) if m = k. This defines a finite Borel
measure µ on Rk . Let Mµ : Rk → [0, ∞] be the maximal function of µ. Then
1
N T b (1, γy,δ ◦ π) = k µ(C(y, δ)) ≤ αk k k/2 Mµ (y).
δ
Local currents in metric spaces 45
The following theorem will be used in the proof of Theorem 8.5 (rectifiable
slices).
In case k 0 = 1 the result is clear from the preceding proof. Our argument
for the general case follows [14, Theorem 4.3.5].
0
Now (6.16) with M in place of N shows that hT, (π, π 0 ), (y, ·)i(f, g) ∈ L1 (Rk )
and, in conjunction with the bounded convergence theorem, that Aδ0 (y, y 0 )
satisfies the continuity condition for metric currents. Hence Aδ0 (y, y 0 ) is a
current; since
Z
1
N(Aδ0 (y, y 0 )) ≤ 0 k N hT, (π, π 0 ), (y, z 0 )i dz 0 ,
(6.17)
(δ ) C(y0 ,δ0 )
it is normal by (6.16). Now let δ, δ 0 > 0 and (y, y 0 ) ∈ D1 . For all (f, g) ∈
0
D m−k−k (X), we obtain
by the same argument as for (6.15), and by Fubini’s theorem. Note that the
left side of this identity is continuous in (y, y 0 ), moreover the limit for δ → 0
and for fixed δ 0 > 0 is equal to hT, π, yi(f, γy0 ,δ0 ◦ π 0 , g). Let D2 be the Borel
set of all (y, y 0 ) ∈ D1 where the limit of (6.18) equals Aδ0 (y, y 0 )(f, g) for all
0 0
rational numbers δ 0 > 0 and (f, g) ∈ F × F m−k−k . Then L k+k (D1 \ D2 ) =
0, and for every (y, y 0 ) ∈ D2 , the equality
0
for all (y, y 0 ) ∈ D3 and (f, g) ∈ F × F m−k−k . Combining (6.19), (6.17), and
0
(6.20), we have supδ0 >0 N hT, π, yib (1, γy0 ,δ0 ◦ π ) < ∞, so it follows from
Proposition 5.3 (convergence criterion) that
We conclude this section with a result that will be used in the proof of
Theorem 8.9 (closure theorem).
Proposition 6.6 (slicing convergent sequences). Suppose m ≥ 1, (Tn )n∈N
is a sequence in Nm,loc (X) that converges weakly to some T ∈ Nm,loc (X),
each spt(Tn ) is separable, and supn NV (Tn ) < ∞ for every open set V b X.
Let π ∈ Liploc (X). Then for L 1 -almost every s ∈ R there is a subsequence
(Tn(i) )i∈N such that
hTn(i) , π, si → hT, π, si
weakly and supi NV (hTn(i) , π, si) < ∞ for every open set V b X.
Compare [36, p. 208] and [3, Proposition 8.3].
Proof. We assume without loss of generality that X is compact.
We first show the existence of a subsequence (Tn(i) )i∈N such that for all but
countably many s ∈ R, hTn(i) , π, si → hT, π, si weakly for i → ∞. Consider
the measures µn on R defined by µn (B) := (kTn k + k∂Tn k)(π −1 (B)) for
all Borel sets B ⊂ R. Note that supn µn (R) < ∞. Choose a subsequence
(µn(i) )i∈N that converges weakly∗ to some finite Borel measure µ on R (see
e.g. [13, Sect. 1.9]). Now let s ∈ R be such that µ({s}) = 0. Suppose
(f, g) ∈ D(X) × [Lip1 (X)]m , and |f | ≤ 1. Given > 0, choose δ > 0 such
that kT k({s < π < s+δ}) < /3 and µ([s, s+δ]) < /3, and let us,δ ∈ Lip(X)
be given as in (6.4). Then
(T b χ{π>s} )(f, g) − (T b us,δ )(f, g) = |T ((χ{π>s} − us,δ )f, g)|
≤ kT k({s < π < s + δ}) < /3.
Since lim supi→∞ µn(i) ([s, s + δ]) ≤ µ([s, s + δ]), it follows similarly that
(Tn(i) b χ{π>s} )(f, g) − (Tn(i) b us,δ )(f, g) ≤ µn(i) ([s, s + δ]) < /3
for all sufficiently large i. Moreover, since Tn(i) → T weakly,
(Tn(i) b us,δ )(f, g) − (T b us,δ )(f, g) < /3
for i sufficiently large. Combining these estimates we conclude that
Tn(i) b χ{π>s} → T b χ{π>s}
weakly for i → ∞, and a completely analogous argument shows that
(∂Tn(i) )b χ{π>s} → (∂T )b χ{π>s}
weakly. By (6.6), hTn(i) , π, si → hT, π, si weakly for i → ∞.
Applying Theorem 6.2(3) and (6.3) we obtain
Z Z
lim inf N(hTn , π, si) ds ≤ lim inf N(hTn , π, si) ds
R n→∞ n→∞ R
≤ Lip(π) supn N(Tn ) < ∞.
Hence, for L 1 -almost every s ∈ R, there is a subsequence (Tn(i) )i∈N such that
supi N(hTn(i) , π, si) < ∞. Together with the first part of the proof this yields
the result.
48 U. Lang
holds for all ϕ ∈ Cc1 (A, Rm ), which implies that VA (u) = A |∇u| dx. In par-
R
for ϕ ∈ Cc1 (U, Rm ) or even ϕ ∈ Lipc (U, Rm ) = [D(U )]m . The Radon measure
|Du| is characterized by |Du|(A) = VA (u) for all open sets A ⊂ U .
Functions of bounded variation will be used through the following known
fact (cf. the proof of [3, Lemma 7.3] and the references there), which expresses
a Lipschitz property in terms of the maximal function of the variation mea-
sure.
Proof. SupposeR first that u ∈ C 1 (Rm ) in addition, and let r > 0. Since
1
u(z) − u(0) = 0 h∇u(tz), zi dt for z ∈ Rm , it follows that
1
|u(z) − u(0)|
Z Z Z
dz ≤ |∇u(tz)| dt dz
B(0,r) |z| B(0,r) 0
Z 1Z
= |∇u(tz)| dz dt
0 B(0,r)
Z 1 Z
1
= |∇u(z)| dz dt
0 tm B(0,tr)
≤ αm rm M|Du| (0).
|u(z) − u(x)|
Z
dz ≤ αm rm M|Du| (x)
B(x,r) |z − x|
and define A0 similarly with x0 in place of x. The above estimate implies that
L m (A), L m (A0 ) < αm rm /c and thus
The next result relates normal currents of the type described in Propo-
sition 2.6 (standard example) to functions of bounded variation, cf. [29, Re-
mark 26.28] and [3, Theorem 3.7].
for every f ∈ Bc∞ (U ), where u(x) := hT, π, xi(σ) for any σ ∈ D(U ) with
σ(x) = 1 (note that spt(hT, π, xi) ⊂ {x}). By (2.11) and a smoothing argu-
ment it follows that T = [u]. Together with (1), this proves (2).
by (4.5), Lemma 4.6(2), and (4.10). Since fk → f in L1 (kT k), by the com-
pleteness of L1 (Rm ) there is a function u ∈ L1 (Rm ) such that uk → u
in L1 (Rm ). By passing to the M-limit on either side of the identity
π# (T b fk ) = [uk ], for k → ∞, we conclude that π# (T b f ) = [u]. Now, to
see that (1) implies (2), apply this procedure to the characteristic function
f of a Borel set B b X and an approximating sequence (fk )k∈N in D(X).
To show that (2) implies (1), choose (fk )k∈N such that each fk is a finite
linear combination of characteristic functions of compact sets, and such that
fk → f ∈ D(X) in L1 (kT k).
52 U. Lang
(1) If condition (1) or (2) of Lemma 7.3 holds for some π ∈ Lip(X, Rm ),
then
kT b (1, π)k(π −1 (N ) ∩ B) = 0
whenever N ⊂ Rm is a Borel set with L m (N ) = 0 and B ⊂ X is a Borel
set that is σ-finite with respect to kT k.
(2) If condition (1) or (2) of Lemma 7.3 holds for all π ∈ Lip(X, Rm ), then
kT k(A) = 0 for every set A ⊂ X with H m (A) = 0.
Together with Theorem 7.2, this shows that for every T ∈ Nm,loc (X), kT k
is absolutely continuous with respect to H m (cf. [3, Theorem 3.9]).
Proof. We assume that condition (1) of Lemma 7.3 holds for some π ∈
Lip(X, Rm ). Suppose N ⊂ Rm is a bounded Borel set with L m (N ) = 0.
Let f ∈ D(X), and choose σ ∈ D(X) with σ|spt(f ) = 1. By assumption,
π# (T b f ) = [uf ] for some uf ∈ L1 (Rm ). We abbreviate Tπ := T b (1, π); note
that Tπ ∈ M0,loc (X). Using Lemma 4.6(1) we obtain
We conclude this section with two results for locally normal currents that
relate the previous discussion to the slicing theory.
for all Borel sets B ⊂ Rm . Note that the finite Borel measure µ so defined
is independent of f . Now we choose a countable subset F of {f ∈ D(X) :
0 ≤ f ≤ 1, Lip(f ) ≤ 1} with the following property: Whenever x ∈ X and
0 < r ≤ 1, there is an f ∈ F such that
3
f (x) ≥ r, f ≤ r, spt(f ) ⊂ U(x, r). (7.2)
4
54 U. Lang
By Theorem 6.4, Theorem 7.5, and Lemma 7.1 there exists a Borel set N ⊂
Rm with L m (N ) = 0 such that whenever y, y 0 ∈ Rm \ N and f ∈ F , the slice
Ty := hT, π, yi exists as an element of M0 (X), Ty (f ) = uf (y), Mµ (y) < ∞,
and
By Proposition 7.4(1),
kT b (1, π)k(π −1 (N )) = 0.
The next theorem corresponds to [3, Theorem 4.5] in the case of finite
mass. A similar characterization of classical rectifiable currents is given in [14,
Theorem 4.1.28].
56 U. Lang
H m E \ i Fi (Ki ) = 0.
S
(8.2)
S∈D
Let (f, π) m (X), and denote by χ and χ the characteristic functions of
i
i i
B and i B , respectively. Using Theorem 4.4(2) we get
X X
T (f, π) = T (χf, π) = T (χi f, π) = T i (f, π).
i i
Local currents in metric spaces 57
We now supplement the slicing theory for locally normal currents T with
the following two theorems, providing criteria for the rectifiability of T in
terms of the rectifiability of slices, and vice-versa. We refer to [38], [20] for
similar statements in the context of classical currents, and to [3, Theorem 8.1]
for a corresponding result for normal metric currents.
Proof. For the proof of (1), we first consider the case k = m, so that
π ∈ Liploc (X, Rm ). Choose a countable family U of open sets U b X such
that for every x ∈ spt(T ) and > 0 there is a U ∈ U with x ∈ U ⊂ U(x, ),
and such that the union of finitely many elements of U is in U . By as-
sumption, for every U ∈ U there exists a function uU ∈ L1 (Rm , Z) such that
π# (T b U ) = [uU ]. Thus, for L m -almost every y, hT, π, yi ∈ M0,loc (X), and,
by Theorem 7.5,
hT, π, yi(χU ) = uU (y) ∈ Z
for all U ∈ U . It follows that for every such y, hT, π, yi(χK ) ∈ Z for all
compact sets K ⊂ X. Lemma 8.2 then shows that hT, π, yi ∈ I0,loc (X).
Now suppose 1 ≤ k < m, and let π ∈ Liploc (X, Rk ). Choose a countable
set P ⊂ Lip1 (X) as in Theorem 8.4. Fix ρ ∈ P m−k for the moment. We
know from Theorem 6.5 (iterated slices) and the result in the case k = m
that
hT, π, yi, ρ, z = hT, (π, ρ), (y, z)i ∈ I0,loc (X)
for L m -almost every (y, z) ∈ Rk × Rm−k . Hence, there is a set Nρ ⊂ Rk with
L k (Nρ ) = 0 such that for every y ∈ Rk \ Nρ , hT, π, yi ∈ Nm−k,loc (X) and
hT, π, yi, ρ, z ∈ I0,loc (X) for all z ∈ Rm−k \ Ny,ρ
0 , where L m−k (N 0 ) = 0.
y,ρ
Suppose y ∈ Rk \ ρ∈P m−k Nρ , and put Ty := hT, π, yi. For each ρ ∈ P m−k ,
S
by (6.3) and the induction hypothesis. As this holds for every π ∈ Lip(X),
Theorem 8.5(2) implies that ∂T ∈ Im,loc (X).
60 U. Lang
References
[1] Adams, T., Flat chains in Banach spaces. J. Geometric Anal. 18 (2008),
no. 1, 1–28.
[2] Ambrosio, L., Fusco, N., Pallara, D., Functions of bounded variation
and free discontinuity problems. Oxford Mathematical Monographs. The
Clarendon Press, Oxford University Press, New York, 2000.
[3] Ambrosio, L., Kirchheim, B., Currents in metric spaces. Acta Math. 185
(2000), no. 1, 1–80.
[5] Cheeger, J., Kleiner, B., On the differentiability of Lipschitz maps from
metric measure spaces to Banach spaces. Inspired by S. S. Chern, 129–
152, Nankai Tracts Math., 11. World Sci. Publ., Hackensack, NJ, 2006.
[6] Courant, R., Hilbert, D., Methoden der mathematischen Physik, Band
I. Zweite verbesserte Auflage. Berlin, Springer, 1931.
[8] De Lellis, C., Some fine properties of currents and applications to distri-
butional Jacobians. Proc. Roy. Soc. Edinburgh Sect. A 132 (2002), no.
4, 815–842.
[10] De Pauw, Th., Hardt, R., Size minimization and approximating prob-
lems. Calc. Var. Partial Differential Equations 17 (2003), no. 4, 405–442.
[11] De Pauw, Th., Hardt, R., Application of scans and fractional power
integrands. Variational problems in Riemannian geometry, 19–31, Progr.
Nonlinear Differential Equations Appl., 59. Birkhäuser, Basel, 2004.
[13] Evans, L. C., Gariepy, R. F., Measure theory and fine properties of
functions. Studies in Advanced Mathematics. CRC Press, Boca Raton,
FL, 1992.
[14] Federer, H., Geometric measure theory. Die Grundlehren der mathema-
tischen Wissenschaften, 153. Springer, New York, 1969.
62 U. Lang
[15] Federer, H., Fleming, W. H., Normal and integral currents. Ann. of
Math. (2) 72 (1960), 458–520.
[16] Hardt, R., Rivière, T., Connecting topological Hopf singularities. Ann.
Sc. Norm. Super. Pisa Cl. Sci. (5) 2 (2003), no. 2, 287–344.
[19] Heinonen, J., Nonsmooth calculus. Bull. Amer. Math. Soc. (N.S.) 44
(2007), no. 2, 163–232.
[20] Jerrard, R. L., A new proof of the rectifiable slices theorem. Ann. Sc.
Norm. Super. Pisa Cl. Sci. (5) 1 (2002), no. 4, 905–924.
[22] Keith, S., A differentiable structure for metric measure spaces. Adv.
Math. 183 (2004), no. 2, 271–315.
[23] Kirchheim, B., Rectifiable metric spaces: local structure and regularity
of the Hausdorff measure. Proc. Amer. Math. Soc. 121 (1994), no. 1,
113–123.
[25] Mucci, D., Fractures and vector valued maps. Calc. Var. Partial Differ-
ential Equations 22 (2005), no. 4, 391–420.
[26] Rudin, W., Real and complex analysis. Third edition. McGraw-Hill, New
York, 1987.
[30] Weaver, N., Lipschitz algebras. World Scientific Publishing, River Edge,
NJ, 1999.
[31] Weaver, N., Lipschitz algebras and derivations. II. Exterior differentia-
tion. J. Funct. Anal. 178 (2000), no. 1, 64–112. Erratum, J. Funct. Anal.
186 (2001), no. 2, 546.
[33] Wenger, S., Filling invariants at infinity and the Euclidean rank of
Hadamard spaces. Int. Math. Res. Not. 2006, Art. ID 83090, 33 pp.
[34] Wenger, S., Flat convergence for integral currents in metric spaces. Calc.
Var. Partial Differential Equations 28 (2007), no. 2, 139–160.
[35] Wenger, S., Gromov hyperbolic spaces and the sharp isoperimetric con-
stant. Invent. Math. 171 (2008), no. 1, 227–255.
[36] White, B., A new proof of the compactness theorem for integral currents.
Comment. Math. Helv. 64 (1989), no. 2, 207–220.
[37] White, B., The deformation theorem for flat chains. Acta Math. 183
(1999), no. 2, 255–271.
[38] White, B., Rectifiability of flat chains. Ann. of Math. (2) 150 (1999), no.
1, 165–184.
Urs Lang
[email protected]
Department of Mathematics
ETH Zurich
8092 Zurich
Switzerland