You are on page 1of 42

DIFFERENTIAL FORMS ON ORBIFOLDS WITH CORNERS

JAKE P. SOLOMON AND SARA B. TUKACHINSKY

Abstract. We give a detailed account of differential forms and currents on orbifolds with
arXiv:2011.10030v1 [math.SG] 19 Nov 2020

corners, the pull-back and push-forward operations, and their fundamental properties. We
work within the formalism where the category of orbifolds with corners is obtained as a
localization of the category of étale proper groupoids with corners. Constructions and
proofs are formulated in terms of the structure maps of the groupoids, avoiding the use of
orbifold charts. The Fréchet space of differential forms on an orbifold and the dual space of
currents are shown to be independent of which étale proper groupoid is chosen to represent
the orbifold.

Contents
1. Introduction 2
1.1. Statement of results 2
1.2. Outline 3
1.3. Acknowledgements 3
2. Orientations and integration 4
2.1. Manifolds with corners 4
2.2. Orientation conventions 5
2.3. Integration properties 7
3. Orbifolds with corners 9
3.1. Étale proper groupoids with corners 9
3.2. The weak 2-category of orbifolds with corners 11
4. Differential forms on orbifolds with corners 12
4.1. Main definitions 12
4.2. Clean subsets 15
4.3. Comparing differential forms and cleanly supported differential forms 20
5. Integration properties 24
5.1. Fiber products of orbifolds 24
5.2. Integration properties in EPG 28
5.3. Integration properties in Orb 35
6. Currents 37
References 42

Date: November 2020.


2020 Mathematics Subject Classification. 20L05, 58A10 (Primary) 18E35, 58A25 (Secondary).
Key words and phrases. Differential form, integration over fibers, orbifold, étale proper groupoid, bicategory,
category of fractions.
1
1. Introduction
1.1. Statement of results. In this paper, we give a detailed account of differential forms
and currents on orbifolds with corners, the pull-back and push-forward operations, and their
fundamental properties. We work within the formalism where the bicategory of orbifolds
with corners is obtained as a localization of the 2-category of étale proper groupoids with
corners. To each orbifold with corners X , we associate a differential graded algebra A∗ (X ) of
differential forms on X . We endow A∗ (X ) with the structure of a Fréchet space compatible
with the differential graded algebra structure. To any morphism of orbifolds with corners
f : X → Y, we associate a pull-back morphism
f ∗ : A∗ (Y) → A∗ (X ),
and to any relatively oriented proper submersion f : X → Y, we associate a push-forward
morphism
f∗ : A∗ (X ) → A∗ (Y).
The pull-back and push-forward morphisms are continuous with respect to the Fréchet topol-
ogy. Our definitions of differential forms, pull-back, and push-forward, for orbifolds with
corners generalize the usual ones for manifolds with corners.
The following theorem is our main result. We refer the reader to Section 2 for the notions
of strongly smooth maps, fiberwise boundary, and relative orientations.
Theorem 1.
(a) Let f, g : X → Y be morphisms of orbifolds with corners and let α : f ⇒ g be a
2-morphism. Then f ∗ = g ∗ . If f and g are relatively oriented proper submersions,
and α is relatively oriented, then also f∗ = g∗ .
(b) Let g : X → Y, f : Y → Z, be morphisms of orbifolds with corners. Then
g ∗ ◦ f ∗ = (f ◦ g)∗.
If f, g, are relatively oriented proper submersions, then
f∗ ◦ g∗ = (f ◦ g)∗ .
(c) Let f : X → Y be a relatively oriented proper submersion of orbifolds with corners,
and let α ∈ A∗ (Y), β ∈ A∗ (X ). Then
f∗ (f ∗ α ∧ β) = α ∧ f∗ β.
(d) Let
p
X ×Z Y / Y
q ✈ g
v~ ✈✈✈✈✈
 f 
X / Z
be a weak pull-back diagram of orbifolds with corners, where f and g are strongly
smooth and g is a relatively oriented proper submersion. Equip q with the pull-back
relative orientation and let α ∈ A∗ (Y). Then
q∗ p∗ α = f ∗ g∗ α.
2
(e) Let f : X → Y be a strongly smooth relatively oriented proper submersion of orbifolds
with corners with dim X = s, and let ξ ∈ At (X ). Then

d(f∗ ξ) = f∗ (dξ) + (−1)s+t f ∂ v X ∗ ξ,

where ∂ v X is the fiberwise boundary with respect to f, with the induced relative ori-
entation.

We define the space of currents on an orbifold with corners X to be the continuous dual of
the topological vector space of compactly supported differential forms A∗c (X ) equipped with
the weak∗ topology. It follows from Theorem 1 (a) and (b) that equivalent orbifolds with
corners have isomorphic spaces of differential forms and currents.
From Theorem 1 (a) and (b), we also obtain Theorem 2 below. Let Orb denote the weak 2-
category of orbifolds, and let Orbpr denote the subcategory of Orb with the same objects and
relatively oriented proper submersions for morphisms. Let Fdg denote the 2-category where
objects are differential graded Fréchet spaces, morphisms are continuous maps commuting
with the differential and preserving the grading up to a shift, and only identity 2-morphisms.
Let Fdga denote the 2-category with objects Fréchet differential graded algebras, morphisms
continuous differential graded algebra homomorphisms, and only identity 2-morphisms.

Theorem 2. Differential forms give rise to two homomorphisms of weak 2-categories,

F∗ : Orb → Fdgaop , F∗ : Orbpr → Fdg,

given on objects X of Orb by

F∗ (X ) = F∗ (X ) = A∗ (X),

and on morphisms f : X → Y in Orb by

F∗ (f ) = f ∗ , F∗ (f ) = f∗ .

Of course, any 2-morphism in Orb is mapped by both F∗ and F∗ to the identity 2-morphism.

1.2. Outline. In Section 3, we recall the definition of orbifolds with corners and morphisms
thereof. We discuss the notions of submersions, proper morphisms, refinements, and rela-
tively oriented morphisms. In Section 4 we define the algebra of differential forms and the
space of differential forms with clean support on an orbifold and construct explicit isomor-
phisms between the two. We define the operations of pullback and of push-forward along
a proper submersion, and show that for a refinement the two operations are inverse to one
another. Section 5 is devoted to a detailed proof of Theorem 1. Section 6 defines currents
and currents relative to the boundary and proves an analog of Theorem 1 in that context.

1.3. Acknowledgements. We are grateful to an anonymous referee who asked us to provide


a reference for the results in this paper. We are grateful to P. Giterman, O. Kedar, and A.
Yuval for helpful conversations. The first author was partially supported by ERC starting
grant 337560 and ISF Grant 569/18. The second author was supported by NSF grant No.
DMS-163852.
3
2. Orientations and integration
2.1. Manifolds with corners. For M a manifold with corners, the boundary ∂M is defined
as in [5, Definition 2.6]. So, ∂M is again a manifold with corners and comes with a natural
map iM : ∂M → M. We say a map is smooth if all its partial derivatives, including one-sided
derivatives, exist and are continuous at every point. We say a map is strongly smooth if it is
strongly smooth in the sense of [6, Definition 2.1(e)]. We say a smooth map is a submersion
if it is locally diffeomorphic to a projection of a direct product of manifolds with corners
to one factor. In particular, a smooth map induces a smooth map of the boundary, which
we call the restriction to the boundary, the restriction of a strongly smooth map to the
boundary is also strongly smooth, and the restriction of a submersion to the boundary is
also a submersion. As usual, diffeomorphisms are smooth maps with a smooth inverse.
Remark 2.1. A map that is smooth in our terminology is called weakly smooth in [5, Defini-
tion 3.1] and [6, Definition 2.1(a)]. Our definition of strongly smooth is identical to [6, Def-
inition 2.1(e)] but corresponds to smooth in [5, Definition 3.1]. Our notion of submersion
corresponds to a horizontal submersion in the terminology of [13, Definition 19(a)]. For
strongly smooth maps it coincides with the submersion of [5, Definition 3.2(iv)].
Transversality of strongly smooth maps of manifolds with corners is defined as in [5,
Definition 6.1], so any strongly smooth map is tranverse to a strongly smooth submersion.
Fiber products of transverse strongly smooth maps exist by [5, Theorem 6.4]. Throughout
the paper, we assume without further comment that all maps are smooth.
2.1.1. Vertical and horizontal boundary. Let M, N, be smooth manifolds with corners, and
let f : M → N be a strongly smooth map. Following [5, Section 4], we decompose ∂M into
horizontal and vertical parts with respect to f , as follows.
Denote by S 1 (N) the set of all points of depth 1 in N. Recall that β is a local boundary
component at y ∈ N if β is a connected component of S 1 (N) such that y ∈ β. Thus,
 
M f ×iN ∂N = x, (y, β) | f (x) = y = iN (y, β) .
Let (x, (y, β)) ∈ M f ×iN ∂N and let V ⊂ N be a neighborhood of y such that there is a
boundary defining function b : V → [0, ∞) at (y, β) in the sense of [5, Definition 2.14]. By
the strongly smooth assumption, either b ◦ f ≡ 0 or b ◦ f : f −1 (V ) → [0, ∞) is a boundary
defining function at x for a unique local boundary component. Therefore, the subset
 
Ξ := x, (y, β) | b ◦ f is boundary defining at x ⊂ M f ×iN ∂N
comes with an inclusion
ξ : Ξ −→ ∂M
and we write
∂M = ∂ v M ⊔ ∂ h M,
∂ h M = im(ξ), ∂ v M = ∂M \ im(ξ).
By [5, Proposition 4.3], ∂ v M, ∂ h M, are closed and open in ∂M and thus are manifolds with
corners. We denote the restrictions of iM by
ivM : ∂ v M −→ M, ihM : ∂ h M −→ M.
These come with canonical relative orientations induced by restricting oiM .
4
f
Remark 2.2. Our definition of ∂ v M is ∂+ M in the notation of [5].
Remark 2.3. In the special case when f is
`a strongly smooth proper submersion, ∂ v recovers
v −1
the fiberwise boundary, that is, ∂ M = y∈N ∂(f (y)).
2.2. Orientation conventions. For a diffeomorphism f : M → N of oriented manifolds
with corners, we define sgn(f ) to be 1 if f preserves orientation and −1 otherwise. We use
similar notation for isomorphisms of oriented vector spaces. We follow [5, Convention 7.2(a)]
for orienting the boundary and [5, Convention 7.2(b)] for orienting fiber products, as detailed
in the following.
To orient boundary, let M be an oriented orbifold with corners. Let p ∈ ∂M and let B
be a basis for Tp ∂M. Let N ∈ TiM (p) M be the outward-pointing normal at p. We say B is
positive if N ◦ B is a positive basis for TiM (p) M.
To orient fiber products, let M, N, and P, be oriented orbifolds with corners. Let f : M →
N and g : P → N be transverse smooth maps, and consider the following pull-back diagram.
p
M ×N P / P (1)
q g
 f 
M / N
Let (m, p) ∈ M × P with f (m) = g(p). By the transversality assumption,
F := dfm ⊕ −dgp : Tm M ⊕ Tp P −→ Tf (m) N
is surjective, and by definition of fiber product, there is a natural isomorphism

ψ : T(m,p) (M ×N P ) −→ ker(F ).
So, we have a short exact sequence
ψ F
0 −→ T(m,p) (M ×N P ) −→ Tm M ⊕ Tp P −→ Tf (m) N −→ 0, (2)
and splitting gives an isomorphism

ϕ : Tm M ⊕ Tp P −→ T(m,p) (M ×N P ) ⊕ Tf (m) N. (3)
We take the orientation on T(m,p) (M ×N P ) to be the one that makes sgn(ϕ) = (−1)dim P ·dim N .
For a submersion of manifolds with corners h : Q → S and y ∈ S, we orient the fiber
h−1 (y) by identifying it with the fiber product {y} ×S Q.
The preceding orientation conventions determine the signs in Proposition 2.6 below as well
as Stokes’ theorem, Proposition 2.7.
Remark 2.4. Our convention for orientation of fiber products agrees with that of [2] in case
f is a submersion.
Let M be a manifold with corners and let M f → M denote the orientation cover considered
as a Z/2 principal bundle. Let f : M → N be a map of manifolds with corners. The relative
orientation bundle of f is the Z/2 principle bundle
f, f ∗N
Lf := HomZ/2 (M e ) → M.
Alternatively, we can view Lf as the quotient of the frame bundle of the line bundle
Hom(det T M, f ∗ det T N) ≃ det T M ∨ ⊗ det T N
5
by the multiplicative group R>0 . A relative orientation of f is a section of Lf . So, ori-
entations on M and N induce a relative orientation of f , but f may be relatively oriented
even if neither M nor N are orientable. Let g : N → Q be another map of manifolds with
corners. Observe that Lg◦f = Lf ×Z/2 f ∗ Lg . So, relative orientations of and og of f and
g respectively induce a relative orientation og ◦ of of g ◦ f. Also, a local diffeomorphism
f : M → N comes with a canonical relative orientation, which we denote by ofc . At a
point m ∈ M, the canonical relative orientation is given as follows. Let v1 , . . . , vk ∈ Tm M
be a basis, let u1 = dfm (v1 ), . . . , uk = dfm (vk ), be the corresponding basis of Tf (m) N, and let
v1∗ , . . . , vk∗ , be the dual basis. Then,
(ofc )m = [v1∗ ∧ · · · ∧ vk∗ ⊗ u1 ∧ · · · ∧ uk ].
If g : N → Q is also a local diffeomorphism, then ogc ◦ ofc = og◦f c . The boundary map
iM : ∂M → M comes with the relative boundary orientation oiM as follows. Let
m ∈ im(iM ) and β a local boundary component at m. Let N ∈ Tm M the outward pointing
normal, u1 , . . . , un−1 ∈ T(m,β) (∂M) a basis, and u∗1, . . . , u∗n−1 , the dual basis. Then
(oiM )(m,β) = [u∗1 ∧ · · · ∧ u∗n−1 ∧ N ∧ u1 ∧ · · · ∧ un−1].
Relative orientations pull-back to fiber products as follows. Consider pull-back diagram (1).
The isomorphism ϕ from (3) induces an isomorphism

det(Tm M) ⊗ det(Tp P ) −→ det(T(m,p) (M ×N P )) ⊗ det(Tf (m) N),
which in turn induces an isomorphism

det(T(m,p) (M ×N P ))∨ ⊗ det(Tm M) −→ det(Tf (m) N) ⊗ det(Tp P )∨.
Composing with the canonical isomorphism

det(Tf (m) N) ⊗ det(Tp P )∨ −→ det(Tp P )∨ ⊗ det(Tf (m) N)
given by y ⊗ v 7→ (−1)dim P ·dim N v ⊗ y, we obtain an isomorphism

det(T(m,p) (M ×N P ))∨ ⊗ det(Tm M) −→ det(Tp P )∨ ⊗ det(Tf (m) N). (4)
Letting (m, p) vary in M ×N P, we obtain an isomorphism of Z/2 principle bundles
ϕ̄ : Lq → f ∗ Lg .
So, given a relative orientation og of g, and letting f −1 og denote the corresponding section
of f ∗ Lg , we define the pull-back relative orientation by
f ∗ og = ϕ̄−1 (f −1 og ).
When the relevant manifolds are oriented, the relative boundary orientation and the pull-back
relative orientation agree with the relative orientations induced from the orientations of the
boundary and the fiber product respectively. Similarly, in the situation of diagram (1), given
a relative orientation of of f, we can define the transpose pull-back relative orientation
t ∗ f
g o of p so that when M, N and P, are oriented, the relative orientation tg ∗ of is induced by
the orientations of M ×N P and P.
Lemma 2.5. Suppose the map g in the pull-back diagram (1) is a local diffeomorphism.
Then, the map q is also a local diffeomorphism and oqc = f ∗ ogc .
6
Proof. It is well known that if g is a local diffeomorphism, then so is q. We check the equality
of orientations pointwise. Let r : Tm M ⊕ Tp P → T(m,p) (M ×N P ) satisfy r ◦ ψ = Id . Any
such r gives rise to an isomorphism ϕ as in (3) given by
ϕ(u, v) = r(u, v) ⊕ F (u, v), u ∈ Tm M, v ∈ Tp P.
The main step in the proof is to compute
det ϕ : det(Tm M ⊕ Tp P ) → det(T(m,p) (M ×N P ) ⊕ Tf (m) N).
The orientation does not depend on the choice of r, so we choose r to facilitate the compu-
tation. Since g is a local diffeomorphism, we can define G : Tm M → Tp P by
G(u) = (dgp )−1 ◦ dfm (u),
and take r(u, v) = r(u) = (u, G(u)). Let u1 , . . . , uk ∈ Tm M and v1 , . . . , vl ∈ Tp P be bases.
Since g is local diffeomorphism, y1 = dgp (v1 ), . . . , yl = dgp (vl ) ∈ Tf (m) N is a basis. So,
det ϕ(u1 ∧ · · · ∧ uk ∧ v1 ∧ · · · ∧ vl ) = (5)
= ϕ(u1 ) ∧ · · · ∧ ϕ(uk ) ∧ ϕ(v1 ) ∧ · · · ∧ ϕ(vl )
= (r(u1 ) ⊕ dfm (u1)) ∧ · · · ∧ (r(uk ) ⊕ dfm (uk )) ∧ (−dgp (v1 )) ∧ · · · ∧ (−dgp (vl ))
= (−1)l r(u1 ) ∧ · · · ∧ r(uk ) ∧ y1 ∧ · · · ∧ yl .
Observe that w1 = r(u1), . . . , wk = r(uk ) ∈ T(m,p) (M ×N P ) is a basis, and let w1∗ , . . . , wk∗ ∈
T(m,p) (M ×N P )∨ denote the dual basis. Let v1∗ , . . . , u∗l ∈ Tp P ∨ denote the dual basis of
v1 , . . . , vl . Since
dim P · dim N + l = l2 + l ≡ 0 (mod 2),
it follows from equation (5) that the isomorphism (4) is given by
w1∗ ∧ · · · ∧ wk∗ ⊗ u1 ∧ · · · ∧ uk 7→ v1∗ ∧ · · · ∧ vl∗ ⊗ y1 ∧ · · · ∧ yl . (6)
Since dgp (vi ) = yi and dq(m,p) (wi ) = ui , the products v1∗ ∧ · · · ∧ vl∗ ⊗ y1 ∧ · · · ∧ yl and
w1∗ ∧ · · · ∧ wk∗ ⊗ u1 ∧ · · · ∧ uk represent the canonical orientations of g and q respectively. 
2.3. Integration properties. Let f : M → N be a relatively oriented proper submersion
with fiber dimension rdim f = r. Denote by f∗ : A∗ (M) → A∗ (N)[−r] the push-forward of
forms along f , that is, integration over the fiber. We will need the following properties of f∗
formulated in [7, Section 3.1].
Proposition 2.6.
(1) Let M be oriented, let f : M → pt and let α ∈ Am (M) with compact support. Then
(R
M
α, m = dim M,
f∗ α =
0, otherwise.
(2) Let g : P → M, f : M → N, be proper submersions with relative orientations og and
of respectively. Then, equipping f ◦ g with the relative orientation of ◦ og , we have
f∗ ◦ g∗ = (f ◦ g)∗ .
(3) Let f : M → N be a relatively oriented proper submersion, and let α ∈ A∗ (N),
β ∈ A∗ (M). Then
f∗ (f ∗ α ∧ β) = α ∧ f∗ β.
7
(4) Let
p
M ×N P / P
q g
f  
M /N
be a pull-back diagram, where f and g are strongly smooth and g is a proper submer-
sion. Let og be a relative orientation of g and equip q with the pull-back orientation
f ∗ og . Let α ∈ A∗ (P ). Then
q∗ p∗ α = f ∗ g∗ α.
Properties (1),(3), and (4), uniquely determine f∗ . Furthermore, we have the following
generalization of Stokes’ theorem. We have corrected the sign of [7, Section 3.1].
Proposition 2.7 (Stokes’ theorem). Let f : M → N be a strongly smooth proper submersion
with relative orientation of and dim M = s, and let ξ ∈ At (M). Then

d(f∗ ξ) = f∗ (dξ) + (−1)s+t f ◦ ivM ∗ ξ,
where (f ◦ ivM )∗ is taken with respect to the induced relative orientation of ◦ oiM .
Below are some consequences of the properties of integration above that we will need.
Lemma 2.8. Consider the pull-back diagram (1) with f, g, strongly smooth and transverse,
and let of be a relative orientation of f. Consider also the following fiber product diagram.

P ×N M / M
p̂ f
 g 
P /N

Let θ : M ×N P → P ×N M be the canonical map, so p̂ ◦ θ = p. Then, the pull-back relative


orientation g ∗ of of p̂ and the transpose pull-back relative orientation tg ∗ of of p are related as
follows:
g ∗of ◦ oθc = (−1)rdim f ·rdim g · tg ∗ of .
Proof. This follows from an argument similar to the proof of Proposition 7.5(a) of [5] given
in Remark 7.6(ii). 
We will use the following variation of [11, Lemma 5.4].
Lemma 2.9. Let f : M → M be a diffeomorphism and let α ∈ A∗ (M). Considering the
push-forward along f with respect to the canonical relative orientation, we have
f ∗ α = f∗−1 α.
Proof. We begin by showing that f∗ 1 = 1. It suffices to show g ∗ f∗ 1 = 1 for g : P → M the
inclusion of an arbitrary point. Indeed, the fiber product P g ×f M is the fiber f −1 (g(P )),
which consists of a Rsingle point with positive orientation. So, Proposition 2.6 (1) and (4)
imply that g ∗f∗ 1 = f −1 (g(P )) 1 = 1. Therefore, using Proposition (2.6) (2) and (3), we obtain
f ∗ α = f∗−1 f∗ (f ∗ α ∧ 1) = f∗−1 (α ∧ f∗ 1) = f∗−1 α.

8
Lemma 2.10. Let
p
M ×N P / P
q g
 f 
M /N

be a pull-back diagram, where f and g are strongly smooth and f is a proper submersion. Let
of be a relative orientation of f and equip p with the transpose pull-back orientation tg ∗ of .
Let α ∈ A∗ (M). Then
p∗ q ∗ α = (−1)rdim f ·rdim g g ∗ f∗ α.
Proof. This follows from Lemmas 2.8 and 2.9. 

3. Orbifolds with corners


In this section we summarize relevant background on orbifolds.

3.1. Étale proper groupoids with corners. A groupoid is a small category G in which
each morphism is an isomorphism. A groupoid X is étale with corners of dimension n
if the object set X0 and morphism set X1 are manifolds with corners of dimension n and
the source and target maps s, t : X1 → X0 are local diffeomorphisms. It follows that the
composition, inverse and identity maps,
m : X1 s ×t X1 → X1 , i : X1 → X1 , e : X0 → X1 ,
are also local diffeomorphisms. An étale groupoid with corners X is proper if the map
s × t : X1 → X0 × X0 is proper.
Some important examples of étale proper groupoids with corners arise as follows.
Example 3.1. Let M be a manifold with corners. We obtain an étale proper groupoid X M
with objects X0M = M, morphisms X1M = M and source and target maps s = t = IdM .
Example 3.2. More generally, let M be a manifold with corners and let U = {Uα }α∈A be an
open cover of M. For α, β ∈ A, let iαβ : Uα ∩ Uβ → Uα denote the inclusion. One obtains
an étale proper groupoid X U with object and morphism spaces and source and target maps
given by
a a a a
X0U = Uα , X1U = Uα ∩ Uβ , s= iαβ , t= iβα .
α∈A α,β∈A α,β∈A α,β∈A

Example 3.3. Generalizing Example 3.1 in a different direction, let M be a manifold with
corners and let G be a discrete group. Consider a proper action
ϕ : G × M → M.
One obtains an étale proper groupoid with corners X ϕ with objects X0ϕ = M, morphisms
X1ϕ = G × M, source map given by s(g, m) = m and target map given by t(g, m) = ϕ(g, m).
A functor of étale proper groupoids with corners F : X → Y is called smooth (resp.
strongly smooth) if the underlying maps on objects F0 : X0 → Y0 and on morphisms
F1 : X1 → Y1 are smooth (resp. strongly smooth).
9
Example 3.4. Consider the étale proper groupoid with corners X U associated to a cover
U = {Uα }α∈A of a manifold with corners M as in Example 3.2. Let U ′ = {Uα′ }α∈A′ be a
refinement of U and let f : A′ → A be a map such that Uα′ ⊂ Uf (α) . For α, β ∈ A′ , let
f
jαf : Uα′ → Uf (α) and jαβ : Uα ∩ Uβ → Uf (α) ∩ Uf (β) denote the inclusions. Then, we have a
U′
smooth functor F : X → X U given by
f

a a f
F0f = jαf , F1f = jαβ .
α∈A′ α,β∈A′

Let F, G : X → Y be two smooth functors. A natural transformation α : F ⇒ G is


called smooth if the underlying map α : X0 → Y1 is smooth. We denote by EPG the
2-category of étale proper groupoids with corners, with morphisms given by smooth functors
and 2-morphisms given by smooth natural transformations. Observe that all 2-morphisms
are in fact isomorphisms.
To an étale proper groupoid with corners X , we associate the topological space |X |, which
is the set of isomorphism classes of objects equipped with the quotient topology induced
from X0 . It follows from the properness of X that |X | is Hausdorff; see Lemma 4.22 below.
We say that a smooth functor of étale proper groupoids with corners F : X → Y is proper
if the induced map |F | : |X | → |Y| is proper. In particular, we say that X is compact if
|X | is compact. In the examples above,
|X M | = |X U | = M, |X ϕ | = M/G, |F f | = IdM .
We say that a morphism F : X → Y in EPG is a submersion if F0 and hence F1 are
submersions. We say that a morphism F : X → Y in EPG is a local diffeomorphism
if F0 and hence also F1 are local diffeomorphisms. We say that morphisms F : X → Z,
G : Y → Z, are transverse if F0 and G0 are transverse. In particular, any strongly smooth
morphism is transverse to a strongly smooth submersion.
An orientation of an étale proper groupoid with corners X is an orientation on the
space of objects X0 and an orientation on the space of morphisms X1 such that the local
diffeomorphisms s, t : X1 → X0 are orientation preserving. We will also use the notion of
a relative orientation of a smooth functor F : X → Y. Recall the notation of Section 2.2
for relative orientations of smooth maps of manifolds with corners. A relative orientation
of a smooth functor F : X → Y is a pair of relative orientations oF0 and oF1 of F0 and F1
respectively, such that
osc ◦ oF1 = oF0 ◦ osc , otc ◦ oF1 = oF0 ◦ otc . (7)
In the case F is a local diffeomorphism, the canonical relative orientations of F0 and F1 give
what we call the canonical relative orientation of F. Let F, G : X → Y be relatively
oriented smooth functors and let α : F ⇒ G be a smooth natural transformation. Observe
that the underlying map α : X0 → Y1 satisfies s ◦ α = F and t ◦ α = G. Thus, a relative
orientation of α is a relative orientation oα of the underlying map α : X0 → Y1 such that
osc ◦ oα = oF0 , otc ◦ oα = oG0 . (8)
Example 3.5. Suppose F is a smooth functor and let oF1 and oF2 be two relative orientations
of F. Let α : F ⇒ F be the identity natural transformation. Then α is relatively oriented if
and only if oF1 = oF2 .
10
Recall the discussion of the boundary of manifolds with corners from Section 2.1. The
boundary of an étale proper groupoid with corners X is the étale proper groupoid with
corners ∂X together with a morphism iX : ∂X → X defined as follows. The object and
morphisms spaces of ∂X are given by
(∂X)j = ∂Xj , j = 0, 1,
and the morphism iX is given by (iX )j = iXj . Observe that a local diffeomorphism of
manifolds with corners f : M → N induces a local diffeomorphism of the boundaries
∂f : ∂M → ∂N. Thus, the structure maps s, t, m, i, and e, of X , being local diffeomor-
phisms, induce structure maps on ∂X .
Some references on étale proper groupoids are [8, 9]. Étale proper groupoids with corners
are examples of polyfolds [4].
3.2. The weak 2-category of orbifolds with corners. A smooth functor of étale proper
groupoids with corners F is called a refinement if F is an equivalence of categories and
a local diffeomorphism. For example, the functor F f given in Example 3.4 is a refinement.
Observe that a refinement need not have a smooth weak inverse, so a refinement is generally
not an equivalence in the 2-category EPG. Following Zernik [13], we define the weak 2-
category of orbifolds with corners Orb to be the weak 2-category of fractions of EPG
with respect to refinements as defined by Pronk [10]. In particular, in the category Orb
a refinement is an equivalence. Zernik [13] verifies that refinements satisfy the necessary
conditions for the category of fractions to be defined.
More explicitly, we can describe Orb as follows. The objects of Orb are the objects of
EPG. A morphism f : X → Y in Orb is a pair of morphisms R : X ′ → X and F : X ′ → Y
R F
in EPG with R a refinement. We abbreviate f : X ← X ′ → Y or simply f = F |R. Let
R F S G
f : X ← X ′ → Y and g : X ← X ′′ → Y be morphisms X → Y in Orb. A 2-morphism
α : f ⇒ g in Orb is given by a pair of refinements T1 : X ′′′ → X ′ and T2 : X ′′′ → X ′′ and a
pair of 2-morphisms α1 : R ◦ T1 ⇒ S ◦ T2 and α2 : F ◦ T1 ⇒ G ◦ T2 in EPG as illustrated by
the following diagram:

= X ❇❚❚❚❚
③ ❇❇ ❚❚❚
③③ ❇❇ ❚❚❚❚F
③③③ ❇ ❚❚❚❚
③③ T1 R ❇❇ ❚❚❚❚
′′′
❚❚)
X ❉ ⇓ α1 X ⇓ α2 ❥5 Y
❉❉ ⑤>
❥❥❥ ❥

❉❉T2 S ⑤⑤⑤ ❥❥❥ ❥
❉❉ ⑤⑤ ❥❥❥❥
❉! ⑤⑤❥❥❥❥❥❥ G

X ′′
A detailed account of compositions of morphisms of different types can be found in [10]. We
discuss the composition of 1-morphisms in greater detail in Section 5.3.2. One can identify
isomorphic morphisms in EPG to obtain a (1-)category, which coincides with the category of
fractions with respect to refinements as defined by Gabriel-Zisman [3]. This is the approach
taken by [9] to define the category of orbifolds, and is also used in the context of polyfolds [4].
One disadvantage of this approach is that morphisms in the category of fractions are not local.
That is, one cannot construct a morphism by gluing together morphisms on each member of
a covering that agree on the intersections. See [8]. In the weak 2-category approach, one can
construct a morphism by gluing together morphisms on each member of a covering if one
specifies 2-morphisms on intersections that satisfy a cocycle condition on triple intersections.
11
Observe that every refinement is a strongly smooth proper submersion. Moreover, a
refinement carries a canonical relative orientation since it is a local diffeomorphism. Thus,
R F
we say that a morphism f : X → Y in Orb given by a diagram X ← X ′ → Y is relatively
oriented (resp. strongly smooth, proper, a submersion) if the morphism F in EPG is
relatively oriented (resp. strongly smooth, proper, a submersion).

4. Differential forms on orbifolds with corners


4.1. Main definitions. The present section gives the definitions of differential forms on
orbifolds with corners and the push-forward and pull-back operations. The definitions rely
on a series of lemmas, which we formulate here and prove in Sections 4.2 and 4.3. Our
approach is based on ideas from [1] and [13].
For a manifold with corners M, denote by A∗ (M) the Fréchet space of smooth differential
forms on M with the C ∞ topology. Note that the exterior derivative d and the pull-back
f ∗ along a smooth map f are linear and continuous. For a proper submersion f , the push-
forward f∗ is linear and continuous as well. Let X be an object of EPG. The differential
forms on X are given by
A∗ (X ) := ker(s∗ − t∗ : A∗ (X0 ) → A∗ (X1 )).
In other words, the differential forms on X are the invariants of the action of isomorphisms
on A∗ (X0 ). It is a closed subset of A∗ (X0 ) and thus is also a Fréchet space.
Example 4.1. Consider X M as in Example 3.1. Then, A∗ (X M ) recovers the usual differential
forms A∗ (M).
Example 4.2. Consider X ϕ as in Example 3.3. Then, A∗ (X ) recovers the G-invariant differ-
ential forms A∗ (M)G .
A subset C ⊂ X0 is called clean if t|s−1 (C) is proper, or equivalently, if s|t−1 (C) is proper.
A subset C ⊂ X1 is called clean if s(C), t(C), are clean.
Lemma 4.3. Let C ⊂ X1 be clean. Then s|C and t|C are proper.
Lemma 4.4. A closed subset of a clean subset C ⊂ Xi is clean for i = 0, 1.
Lemma 4.5. A finite union of clean subsets of Xi is clean for i = 0, 1.
Remark 4.6. If X is compact, it follows from Lemma 4.24 below that a set C ⊂ Xi , i = 0, 1,
is clean if and only if it is compact. However, when X is not compact, clean sets are
significantly more flexible. For example, Lemma 4.11 below on the existence of partitions
of unity holds only for functions with clean support. This leads us to consider differential
forms with clean support.
Let A∗cl (X0 ), A∗cl (X1 ), denote the locally convex subspaces of A∗ (X0 ), A∗ (X1 ), consisting
of differential forms with clean support on X0 , X1 , respectively. Below, the push-forward
of differential forms by any of the structure maps of an étale proper groupoid with corners,
which are by definition local diffeomorphisms, is always taken with respect to the canonical
relative orientation of Section 2.2. We define the differential forms with clean support
on X by
A∗cl (X ) := coker(s∗ − t∗ : A∗cl (X1 ) → A∗cl (X0 )).
12
Being a quotient of A∗cl (X0 ) by a linear subspace, A∗cl (X ) is a topological vector space. A
priori it is not known to be Hausdorff, but in Corollary 4.13 below we conclude that it is
in fact Fréchet. The push-forward operations s∗ , t∗ , are well defined by Lemma 4.3, and
they take differential forms with clean support to differential forms with clean support by
Lemma 4.4. Moreover the difference of two differential forms with clean support is again
cleanly supported by Lemma 4.5. The differential forms with clean support on X are the
coinvariants of the action of isomorphisms on A∗cl (X0 ).
For the following two lemmas, let F : X → Y be a morphism in EPG.
Lemma 4.7. Let Ci ⊂ Xi be clean subsets. If F is proper, then the restrictions
Fi |Ci : Ci → Yi
are proper maps.
Lemma 4.8.
(1) The pull-back map F0∗ : A∗ (Y0 ) → A∗ (X0 ) carries A∗ (Y) to A∗ (X ).
(2) If F is a relatively oriented proper submersion, the push-forward map
(F0 )∗ : A∗cl (X0 ) → A∗cl (Y0 )
carries im(s∗ − t∗ ) to im(s∗ − t∗ ).
In light of Lemma 4.8, we define
F ∗ : A∗ (Y) → A∗ (X ),
by F ∗ α := F0∗ α. If F is a relatively oriented proper submersion, we define
F∗ : A∗cl (X ) → A∗cl (Y)
by F∗ [α] := [(F0 )∗ α] for α ∈ Acl (X0 ). The operations F ∗ and F∗ are again linear and
continuous.
Lemma 4.9. Let F, G : X → Y be morphisms in EPG and suppose there exists a 2-
morphism α : F ⇒ G. Then F ∗ = G∗ . If F and G are relatively oriented proper submersions,
and α is relatively oriented, then F∗ = G∗ .
Next, for an object X of EPG, we construct maps between the differential forms A∗ (X )
and the cleanly supported differential forms A∗cl (X ). By the definition of clean, for α ∈
A∗cl (X0 ), the maps s|supp(t∗ α) and t|supp(s∗ α) are proper, so s∗ t∗ α and t∗ s∗ α are well-defined.
Lemma 4.10.
(1) For α ∈ A∗cl (X0 ), we have t∗ s∗ α = s∗ t∗ α.
(2) For α ∈ im(s∗ − t∗ : A∗cl (X1 ) → A∗cl (X0 )), we have
s∗ t∗ α = 0.
(3) For α ∈ A∗cl (X0 ), we have
s∗ t∗ α ∈ A∗ (X ).
Let
J : A∗cl (X ) → A∗ (X )
be given by J([α]) := t∗ s∗ α, which is well-defined by Lemma 4.10. Observe that J is linear
and continuous.
13
A partition of unity for an orbifold X is a smooth function ρ : X0 → [0, 1] with clean
support such that
t∗ s∗ ρ = 1.
Lemma 4.11. For any object X of EPG there exists a partition of unity.
Given a partition of unity ρ, we define a linear continuous map
K : A∗ (X ) → A∗cl (X )
by K(α) := [ρα].
Lemma 4.12. The maps J and K are inverse to one another. In particular, K does not
depend on the choice of the partition of unity.
Corollary 4.13. A∗cl (X ) is continuously isomorphic to A∗ (X ) and therefore is Fréchet.
In light of the preceding, it is natural to make the following definitions.
Definition 4.14. Let F : X → Y be a relatively oriented proper submersion in EPG. We
define the push-forward operation
F∗ : A∗ (X ) −→ A∗ (Y)
by
F∗ := J ◦ F∗ ◦ K. (9)
Assume now that X is oriented, let pt denote the étale proper groupoid associated with the
point via Example 3.1, and let F : X → pt be the unique smooth functor. Let ξ ∈ A∗ (X )
such that π(supp(ξ)) ⊂ |X | is compact. We define
Z
ξ := F∗ ξ.
X

Remark 4.15. In the situation of the preceding definition, let ρ : X0 → [0, 1] be a partition
of unity on X . Using Jpt = Id and Proposition 2.6(1), we have
Z Z
ξ = F∗ ξ = (F0 )∗ (Kξ) = (F0 )∗ (ρξ) = ρξ.
X X0

To pass from the category EPG to the category Orb, we need the following lemma.
Lemma 4.16. If F : X → Y is a refinement, then F ∗ : A∗ (Y) → A∗ (X ) is an isomorphism
with inverse F∗ taken with respect to the canonical relative orientation of F.
The main definition of the paper is the following.
Definition 4.17. Let X be an object of Orb. We define the differential forms A∗ (X ) the
same as for X considered as an object in EPG. Integration is also defined the same as for
X considered as an object in EPG. Let f : X → Y be the morphism in Orb given by the
diagram
R F
X ← X ′ → Y.
We define the pull-back operation
f ∗ : A∗ (Y) → A∗ (X )
14
by
f ∗ α := R∗ ◦ F ∗ α.
If f is a relatively oriented proper submersion, we define the push-forward operation
f∗ : A∗ (X ) → A∗ (Y)
by
f∗ α := F∗ ◦ R∗ α.
Thus defined, the operations f ∗ and f∗ are linear and continuous, being a composition of
such.
4.2. Clean subsets.
4.2.1. Clean subsets and properness. Let X be an object of EPG.
Lemma 4.18. A clean subset C ⊂ X0 is closed.
Proof. Observe that the image of the identity map e : X0 → X1 is closed. Moreover, we
have C = s(e(X0 ) ∩ t−1 (C)). Since s|t−1 (C) is proper and e(X0 ) is closed, it follows that
s|t−1 (C)∩e(X0 ) is proper. So, its image C is closed. 
Lemma 4.19. Let Z be a topological space that is sequential and Hausdorff, let f : Z → Y
be a continuous map, and let C ⊂ Z. If f |C : C → Y is proper, then C is closed.
Proof. By definition of a sequential space, C is closed if the limit of each convergence sequence
in C belongs to C. Since a sequence together with its limit is a compact set, it suffices to
show for each compact subset K ⊂ Z that C ∩ K is closed. Since Z is Hausdorff, K is closed
in Z, so K ∩ C is closed in C and f |C∩K is proper. So, C ∩ K = (f |C∩K )−1 (f (K)) is compact
and therefore closed as desired. 
Lemma 4.20. Let f : M → N be a continuous map of metrizable spaces. Then f is proper
if and only if for every sequence pi ∈ M such that f (pi ) is convergent, possibly after passing
to a subsequence, there exists p ∈ M such that pi → p.
Proof. We first prove the ‘if’ part of the lemma. For K ⊂ N compact we prove that f −1 (K)
is compact. Indeed, since M is a metrizable space, it suffices to show that any sequence
pi ∈ f −1 (K) has a convergence subsequence. But f (pi ) ∈ K, so possibly after passing to a
subsequence, f (pi ) is convergent. By assumption, after possibly passing to a subsequence
again, there exists p ∈ M such that pi → p, which shows that f −1 (K) is compact as desired.
Conversely, assume f is proper and let pi ∈ M be a sequence such that f (pi ) converges
to q. Then, the set K = {f (pi )}∞ i=1 ∪ {q} is compact by the definition of convergence,
and pi ∈ f −1 (K). Since f is proper, f −1 (K) is compact, and possibly after passing to a
subsequence, there exists p ∈ f −1 (K) such that pi → p. 
Lemma 4.21. The projection π : X0 → |X | is an open map.
Proof. Indeed, let U ⊂ X0 be open. Then, since s is a local diffeomorphism and hence an
open map, it follows that
π −1 (π(U)) = s(t−1 (U))
is open. By the definition of the quotient topology, it follows that π(U) is open. 
Lemma 4.22. The space |X | is Hausdorff.
15
Proof. The quotient |X | is defined by the equivalence relation im(s × t) ⊂ X0 × X0 . By
properness of X , the relation is closed. It now follows from Lemma 4.21 that |X | is Hausdorff.

Lemma 4.23. The space |X | is metrizable.
Proof. It follows from Lemma 4.21 that |X | inherits the locally compact and second countable
properties from X. Since X is a proper groupoid, |X | is Hausdorff. As a locally compact
Hausdorff space, |X | is regular. By the Urysohn metrization theorem, a second countable
regular Hausdorff space is metrizable. 
Lemma 4.24. A subset C ⊂ X0 is clean if and only if π|C : C → |X | is proper.
Proof. First, suppose C is clean. We prove that π|C is proper. Let xi ∈ C be a sequence
such that π(xi ) → y ∈ |X |. By Lemmas 4.20 and 4.23, it suffices to prove that possibly after
passing to a subsequence, there exists x ∈ C such that xi → x. Indeed, by Lemma 4.21 we
can find a sequence x′i ∈ X0 such that π(x′i ) = π(xi ) and x′i → x′ ∈ X0 . Since π(x′i ) = π(xi ),
we can choose a sequence zi ∈ X1 such that s(zi ) = xi and t(zi ) = x′i . Again invoking
Lemma 4.20, since t|s−1 (C) is proper, possibly after passing to a subsequence, there exists
z ∈ X1 such that zi → z. Thus, choosing x = s(z), continuity of s implies xi → x. It follows
from Lemma 4.18 that x ∈ C as desired.
Conversely, suppose π|C : C → |X | is proper and let K ⊂ X0 be compact. We show that
s−1 (C) ∩ t−1 (K) is compact. Indeed,
s(s−1 (C) ∩ t−1 (K)) = C ∩ s(t−1 (K)) = C ∩ π −1 (π(K)),
which is compact by properness of π|C . On the other hand, since π|C is proper, π(C) is
closed, and by continuity π −1 (π(C)) is also closed. Thus,
t(s−1 (C) ∩ t−1 (K)) = K ∩ t(s−1 (C)) = K ∩ π −1 (π(C))
is compact as a closed subset of a compact set. Since the product of compact sets is compact,
it follows that (s × t)(s−1 (C) ∩ t−1 (K)) is compact. Since s × t : X1 → X0 × X0 is proper
by the definition of EPG, we conclude that (s × t)−1 ((s × t)(s−1 (C) ∩ t−1 (K)) is compact.
Lemma 4.19 implies that C is closed, and K is closed because it is compact, so s−1 (C)∩t−1(K)
is closed by continuity of s and t. Since
s−1 (C) ∩ t−1 (K) ⊂ (s × t)−1 ((s × t)(s−1 (C) ∩ t−1 (K)),
it follows that s−1 (C) ∩ t−1 (K) is compact as a closed subset of a compact set. 
Remark 4.25. One can prove Lemma 4.24 without using metrizability of X0 , X1 , |X |, but in-
stead using the local compactness of X0 . However, given the potential interest in generalizing
the results of this paper to infinite dimensional étale proper groupoids such as polyfolds, it
seemed preferable to avoid using local compactness in an essential way. For polyfolds, at least
when the orbit space is paracompact, the metrizability assumptions hold. See Theorems 2.2
and 7.2 of [4].
Lemma 4.26. A clean subset C ⊂ X1 is closed.
Proof. By definition, the sets s(C) and t(C) are clean. By Lemma 4.24 the restrictions π|s(C)
and π|t(C) are proper. Since the product of proper maps is proper, (π × π)|s(C)×t(C) is proper.
Since the map s×t : X1 → X0 ×X0 is proper, it follows that (π×π)◦(s×t)|C : C → |X |×|X |
is proper. Thus, Lemma 4.19 implies that C is closed. 
16
Proof of Lemma 4.3. Since C is clean, s(C) is clean, so t|s−1 (s(C)) is proper. Since C is closed
by Lemma 4.26, also t|s−1 (s(C))∩C is proper. But s−1 (s(C)) ∩ C = C, so t|C is proper. A
similar argument shows that s|C is proper. 

Proof of Lemma 4.4. For i = 0, this follows from the definition since the restriction of a
proper map to a closed set is proper. We turn to the case i = 1. Let D ⊂ C ⊂ X1 be a
closed subset. By Lemma 4.3, it follows that s(D), t(D), are closed. Since s(D) ⊂ s(C) and
t(D) ⊂ t(C), the case i = 0 implies that s(D), t(D), are clean, which means that D is clean
as desired. 

Proof of Lemma 4.5. In the case i = 0, this follows from the fact that if the restriction of a
map to each of a finite collection of sets is proper, then so is its restriction to their union.
The case i = 1 follows from the definition and the case i = 0. 

4.2.2. Clean subsets and morphisms.

Proof of Lemma 4.7. First we prove that F0 |C0 : C0 → Y0 is proper. Since Y0 is Hausdorff, a
compact subset K ⊂ Y0 is closed. So, it suffices to prove that the closed subset
(F0 |C0 )−1 (K) ⊂ C0
is contained in a compact set. This follows from the properness of |F |, Lemma 4.24, and the
commutativity of the following diagram:
F 0 | C0
C0 / Y0
π|C0 π
 |F | 
|X | / |Y|.

To prove that F1 |C1 : C1 → Y1 is proper, we consider the following diagram:


F 1 | C1
C1 / Y1
s×t|C1 s×t
 F0 |s(C1 ) ×F0 |t(C1 ) 
s(C1 ) × t(C1 ) / Y0 × Y0 .

By the part of the lemma that we have already proved, the map F0 |s(C1 ) × F0 |t(C1 ) is proper.
The map s × t is proper by the definition of EPG. Its restriction to C1 is proper because
C1 is closed by Lemma 4.26. So, the commutativity of the diagram implies that F1 |C1 is
proper. 
Lemma 4.27. Let F : X → Y be a morphism in EPG and let Ci ⊂ Xi be clean subsets. If
F is proper, then Fi (Ci ) ⊂ Yi is clean.
Proof. First we prove that F0 (C0 ) ⊂ Y0 is clean. Indeed, by Lemma 4.24 it suffices to prove
the map π|F0 (C0 ) : F0 (C0 ) → |Y| is proper. Since |Y| is Hausdorff, a compact subset K ⊂ |Y|
−1
is closed. So, it suffices to prove that the closed set π|F0 (C0 ) (K) ⊂ F0 (C0 ) is contained in
17
a compact set. This follows from the properness of |F |, Lemma 4.24, and the commutativity
of the following diagram:
F 0 | C0
C0 / F0 (C0 )
π|C0 π|F0 (C0 )
 |F | 
|X | / |Y|.
To prove that F1 (C1 ) ⊂ Y1 is clean, we need to show that s(F1 (C1 )) and t(F1 (C1 )) are clean.
Indeed, s(F1 (C1 )) = F0 (s(C1 )), which is clean by what we have already proved, and the
same argument works for t(F1 (C1 )). 
Proof of Lemma 4.8. To prove part (1), observe that for α ∈ A∗ (Y0 ) we have
s∗ F0∗ α − t∗ F0∗ α = F1∗ (s∗ α − t∗ α) = 0.
We turn to the proof of part (2). By Lemma 4.7 the maps Fi : Xi → Yi are proper when
restricted to clean subsets, so they give rise to push-forward maps on differential forms with
clean support. Lemmas 4.27 and 4.4 imply that if α ∈ A∗cl (Xi ), then the push-forward (Fi )∗ α
has clean support and so belongs to A∗cl (Yi ). Finally, recalling the definition (7) of a relative
orientation for F , by Proposition 2.6 (2) for α ∈ A∗cl (X1 ) we have
(F0 )∗ (s∗ α − t∗ α) = s∗ (F1 )∗ α − t∗ (F1 )∗ α.
So, (F0 )∗ carries im(s∗ − t∗ ) to im(s∗ − t∗ ) as desired. 
For the next two lemmas, let F, G : X → Y be morphisms in EPG and let α : F ⇒ G be
a 2-morphism.
Lemma 4.28. Suppose F is proper. If C ⊂ X0 is clean, then α|C : C → Y1 is proper.
Proof. Let K ⊂ Y1 be compact. We show that α−1 (K) ∩ C is compact. Indeed, s(K) is
compact as the continuous image of a compact set, so Lemma 4.7 implies that F0−1 (s(K))∩C
is compact. Since s ◦ α = F0 , it follows that α−1 (K) ⊂ F0−1 (s(K)). Thus, keeping in mind
Lemma 4.18, we see that α−1 (K) ∩ C is compact as a closed subset of a compact set. 
Lemma 4.29. Suppose F and G are proper. If C ⊂ X0 is clean, then α(C) ⊂ Y1 is clean.
Proof. We have s(α(C)) = F0 (C) and t(α(C)) = G0 (C). So the claim follows from Lemma 4.27.

Proof of Lemma 4.9. We have α : X0 → Y1 with s ◦ α = F0 and t ◦ α = G0 . So, if η ∈ A∗ (Y),
then
F0∗ η − G∗0 η = (s ◦ α)∗ η − (t ◦ α)∗ η = α∗ (s∗ η − t∗ η) = 0.
On the other hand, by Lemma 4.28, the map α : X0 → Y1 is proper when restricted to a
clean subset, so it gives rise to a push-forward map on differential forms with clean support.
Lemmas 4.29 and 4.4 imply that if ξ ∈ A∗cl (X0 ), then the push-forward α∗ ξ has clean support
and so belongs to A∗cl (Y1 ). Finally, recalling the definition (8) of a relative orientation for α,
by Proposition 2.6 (2) if ξ ∈ A∗cl (X0 ), then
(F0 )∗ ξ − (G0 )∗ ξ = (s ◦ α)∗ ξ − (t ◦ α)∗ ξ = (s∗ − t∗ )α∗ ξ.
So,
F∗ [ξ] − G∗ [ξ] = [(s∗ − t∗ )α∗ ξ] = 0.
18


4.2.3. Clean subsets and groupoid composition.

Lemma 4.30. Let Y, Z, W, be topological spaces with W Hausdorff, and let p1 : Y ×W Z → Y


and p2 : Y ×W Z → Z denote the projections. If A ⊂ Y ×W Z is closed and there exist compact
subsets B1 ⊂ Y and B2 ⊂ Z such that pi (A) ⊂ Bi for i = 1, 2, then A is compact.

Proof. Since W is Hausdorff, Y ×W Z ⊂ Y × Z is closed. So, A is closed in Y × Z. Since A


is contained in the compact set B1 × B2 and A is closed, it follows that A is compact. 

The following lemma is familiar from algebraic geometry. We provide a proof in the purely
topological context for the reader’s convenience.

Lemma 4.31. Consider the following fiber square.

p2
Y ×W Z / Z
p1 g
 f 
Y / W

Suppose Z and W are Hausdorff. If f is proper, then p2 is proper.

Proof. Let K ⊂ Z be compact. We prove that p−1 2 (K) is compact. By Lemma 4.30, it
suffices to show that p−1
2 (K) is closed and p (p
i 2
−1
(K)) is compact for i = 1, 2. Indeed, since Z
is Hausdorff, K is closed, and hence p2 (K) is closed. Clearly p2 (p−1
−1
2 (K)) = K is compact.
Finally, p1 (p−1
2 (K)) = f −1
(g(K)) is compact since f is proper. 

Lemma 4.32. Consider the fiber square

p2
X1 ×X0 X1 / X1
p1 t
 
s
X1 / X0

as well as the composition map m : X1 ×X0 X1 → X1 . Let C ⊂ X1 be clean. Then,


(1) p2 |p−1
1 (C)
is proper;
(2) m|p1 (C) is proper;
−1

(3) s|m(p−1
1 (C))
is proper.

Proof. Keeping in mind Lemma 4.3, part (1) is a special instance of Lemma 4.31. To prove
part (2), we show that for any compact K ⊂ X1 , the preimage

(m|p−1
1 (C)
)−1 (K) = m−1 (K) ∩ p−1
1 (C)
19
is compact. Consider the following commutative diagram.

m
p−1
1 (C) ⊂ X1 ×X0 X1
/X ⊃K
♣ 1 ❂❂
✾✾ ♣
p1 ssss ✾✾ t ♣♣
♣ ❂❂
ss ✾✾ p ♣♣♣♣ ❂❂
ss ✾✾ 2 ♣ ❂❂
y s
s t ✾✾ x♣♣♣ ❂❂s
C ⊂ X1✾ ✾✾
/ X0 ❂❂
✾✾ ✾✾ ❂❂
✾✾ ✾✾ ❂❂
✾✾ s ✾ ❂❂
✾✾ 
s
✾✾ t−1 (s(C)) ⊂ X1 / X0
✾✾ s
✾✾ t ssss
✾✾ sss
 ysss
X0

By Lemma 4.30, it suffices to show m−1 (K) ∩ p−1 −1 −1


1 (C) is closed and pi (m (K) ∩ p1 (C)) is
contained in a compact set for i = 1, 2. On the one hand, the restriction t|C is proper by
Lemma 4.3. So, (t|C )−1 (t(K)) is a compact set, and it contains p1 (m−1 (K) ∩ p−1 1 (C)) by
the commutativity of the diagram. On the other hand, since C is clean, so is s(C), and conse-
quently s|t−1 (s(C)) is proper. Observe also that t−1 (s(C)) = p2 (p−1 −1
1 (C)). So, (s|t−1 (s(C)) ) (s(K))
is a compact set and it contains p2 (m−1 (K) ∩ p−1 1 (C)) by the commutativity of the diagram.
Finally, it follows from Lemma 4.26 that m (K) ∩ p−1
−1
1 (C) is closed.
−1
We now prove part (3). Since m|p−1 1 (C)
is proper, m(p 1 (C)) is closed. Furthermore, since
C is clean, so is t(C), and consequently s|t−1 (t(C)) is proper. By commutativity of the diagram,
m(p−1 −1
1 (C)) ⊂ t (t(C)), so s|m(p−1 1 (C))
is proper as the restriction of a proper map to a closed
subset. 

4.3. Comparing differential forms and cleanly supported differential forms. Recall
that the push-forward of differential forms by any of the structure maps of an étale proper
groupoid with corners, which are by definition local diffeomorphisms, is always taken with
respect to the canonical relative orientation as defined in Section 2.2.

Proof of Lemma 4.10. To prove part (1), consider the following diagram.

X1 ❇
⑤ O ❇❇
s ⑤⑤⑤ ❇❇t
⑤⑤ ❇❇
}⑤
⑤ ❇!
X0 a❇ i i
=
X0
❇❇ ⑤⑤
❇❇ ⑤⑤
❇ ⑤
t ❇❇ ⑤
 ⑤⑤ s
X1

By Lemma 2.9, since i ◦ i = Id, we have i∗ = i∗ . Thus, by Proposition 2.6 (2) for α ∈ A∗cl (X0 )
we have
t∗ s∗ α = t∗ i∗ t∗ α = t∗ i∗ t∗ α = s∗ t∗ α.
20
Next, we prove part (2). Suppose η ∈ A∗cl (X1 ) and α = s∗ η − t∗ η. The following diagram
commutes.
m
X1 ×X0 X1 / X1
s ❑❑
p1 ss ❑❑ p2
ss ❑❑
s ❑❑
ss ❑❑
y s
s %
X1 ❑❑ X1 ❇ s
❑❑❑ ss ❇❇
❑s❑ t sss ❇❇s
❑❑❑ s s ❇❇
❑% ss ❇! 
ys
s
X0 X0
We equip p2 with the canonical relative orientation of a local diffeomorphism, which co-
incides with pull-back orientation s∗ otc by Lemma 2.5. So, keeping in mind Lemma 4.32,
Proposition 2.6 implies that
s∗ t∗ s∗ η = s∗ m∗ p∗1 η. (10)
On the other hand, consider the local diffeomorphism
q : X1 ×X0 X1 → X1 ×X0 X1
given by q(x, y) = (i(x), m(x, y)). Equip q with the canonical relative orientation. Observe
that q ◦ q = Id, so q is in fact a diffeomorphism and Lemma 2.9 gives q ∗ = q∗ . The following
diagram commutes.
p1 m
XO 1 o X1 ×X0 X1 / X1 ❇
O ❇❇
❇❇s
❇❇

i i q q
>
X0
⑤⑤⑤

⑤⑤ s
 p1  ⑤⑤
m
X1 o X1 ×X0 X1 / X1

So, Proposition 2.6 (2) gives


s∗ m∗ p∗1 i∗ η = s∗ m∗ p∗1 i∗ η = s∗ m∗ q ∗ p∗1 η = s∗ m∗ q∗ p∗1 η = s∗ m∗ p∗1 η. (11)
Combining equations (10) and (11), we conclude that
s∗ t∗ α = s∗ t∗ s∗ η − s∗ t∗ t∗ η = s∗ t∗ s∗ η − s∗ t∗ s∗ i∗ η = s∗ m∗ p∗1 η − s∗ m∗ p∗1 i∗ η = 0,
as desired.
The proof of part (3) is similar. Equipping p1 with the canonical relative orientation of a
local diffeomorphism, we derive from the above diagrams that
s∗ t∗ s∗ α = (p1 )∗ m∗ s∗ α, (p1 )∗ m∗ s∗ α = i∗ (p1 )∗ m∗ s∗ α.
So, it follows from part (1) that

s∗ s∗ t∗ α − t∗ s∗ t∗ α = s∗ t∗ s∗ α − t∗ t∗ s∗ α =
= s∗ t∗ s∗ α − i∗ s∗ t∗ s∗ α = (p1 )∗ m∗ s∗ α − i∗ (p1 )∗ m∗ s∗ α = 0.

21
Lemma 4.33. Let F : X → Y be a proper local diffeomorphism in EPG that is fully faithful
as a functor of the underlying groupoids. Equipping F with the canonical relative orientation,
the following diagram commutes.
F∗
A∗ (X ) o A∗ (Y)
O O
J J
F∗
A∗cl (X ) / A∗cl (Y)

Proof. By definition, F is fully faithful if and only if the following diagram is Cartesian.
F1
X1 / Y1
t×s t×s
 F0 ×F0

X0 × X0 / Y0 × Y0
In other words, we have a canonical diffeomorphism
X1 ≃ (X0 × X0 ) ×Y0 ×Y0 Y1 .
Furthermore, we have a canonical diffeomorphism
(X0 × X0 ) ×Y0 ×Y0 Y1 ≃ (X0 ×Y0 Y1 ) ×Y0 X0 .
Thus, we obtain the following commutative diagram, in which both squares are Cartesian.
X1 ❚❚❚❚
❚❚❚❚
❚❚❚s❚
≀ ❚❚❚❚
t  ❚❚❚❚
q2 ❚)
(X0 ×Y0 Y1 ) ×Y0 X0 / X0

q1 F0
  s◦p2 
X0 o p1 X0 ×Y0 Y1 /
❥❥❥5
Y0
❥❥❥
s ❥❥❥
F0 p2
❥❥❥❥❥❥❥
  ❥❥❥❥❥

Y0 o t
Y1

All maps in the diagram are local diffeomorphisms, and we equip them with their canonical
relative orientations. So, for α ∈ A∗cl (X0 ) Lemma 2.5, Proposition 2.6, and Lemma 2.9, give
F0∗ t∗ s∗ (F0 )∗ α = (p1 )∗ p∗2 s∗ (F0 )∗ α = (p1 )∗ (s ◦ p2 )∗ (F0 )∗ α = (p1 )∗ (q1 )∗ q2∗ α = t∗ s∗ α,
as desired. 
Lemma 4.34. Suppose F : X → Y is a refinement. If ρ is a partition of unity for X , then
F∗ ρ is a partition of unity for Y.
Proof. The claim is equivalent to JF∗ ρ = 1. It follows from Lemma 4.33 that
F ∗ JF∗ ρ = Jρ = 1.
22
So, it suffices to show that if α ∈ A∗ (Y) satisfies F ∗ α = 1, then α = 1. Indeed, consider the
following fiber product.
p2
X0 ×Y0 Y1 / Y1
p1 t
 F0

X0 / Y0
Since F is an equivalence of categories, it is in particular essentially surjective, which means
that the map
s◦p2
X0 ×Y0 Y1 −→ Y0
is surjective. Since α ∈ A∗ (Y), we have s∗ α = t∗ α, so
1 = p∗1 1 = p∗1 F0∗ α = p∗2 t∗ α = p∗2 s∗ α = (s ◦ p2 )∗ α.
Since s ◦ p2 is a surjective local diffeomorphism, the claim follows. 
Proof of Lemma 4.11. It is shown in [1, pp. 14-15] that there exists a refinement F : X ′ → X
such that there exists a partition of unity ρ′ for X ′ . Lemma 4.34 asserts that ρ = F∗ ρ′ is a
partition of unity for X . 
Proof of Lemma 4.12. First we prove that J ◦ K = Id . Indeed, for α ∈ A∗ (X ), we have
J ◦ K(α) = t∗ s∗ (ρα) = t∗ (s∗ρ s∗ α) = t∗ (t∗ α s∗ρ) = α t∗ s∗ρ = α.
Next, we prove that J is injective. More specifically, suppose α ∈ Acl (X0 ) and s∗ t∗ α = 0. By
Lemma 4.11 choose a partition of unity ρ ∈ A∗cl (X0 ). We claim that α = (t∗ − s∗ )(s∗ρ t∗ α),
so [α] = 0 ∈ A∗cl (X ) = A∗cl (X0 )/ im(s∗ − t∗ ). Indeed,
(t∗ − s∗ )(s∗ ρ t∗ α) = t∗ (t∗ α s∗ρ) − s∗ (s∗ρ t∗ α) = αt∗ s∗ ρ − ρs∗ t∗ α = α · 1 − ρ · 0 = α.
Since J ◦ K = Id, it follows that J is surjective. So, J is invertible, and hence K = J −1 . 
Lemma 4.35. If F : X → Y is a refinement, then the following diagram commutes.
F∗
A∗ (X ) o A∗ (Y)
K K
 F∗

A∗cl (X ) / A∗cl (Y)

Proof. By Lemma 4.11 choose a partition of unity ρ ∈ A∗cl (X ). By Lemma 4.34 the push-
forward F∗ ρ ∈ A∗cl (Y) is a partition of unity. So, for α ∈ A∗ (Y) Proposition 2.6(3) gives
F∗ KF ∗ α = F∗ (ρF ∗ α) = (F∗ ρ)α = Kα.

Proof of Lemma 4.16. Lemma 4.12 and Lemma 4.33 give
F ∗ ◦ (J ◦ F∗ ◦ K) = (F ∗ ◦ J ◦ F∗ ) ◦ K = J ◦ K = Id .
On the other hand, Lemma 4.12 and Lemma 4.35 give
(J ◦ F∗ ◦ K) ◦ F ∗ = J ◦ (F∗ ◦ K ◦ F ∗ ) = J ◦ K = Id .
The lemma follows. 
23
5. Integration properties
5.1. Fiber products of orbifolds. In a discussion of properties of differential forms, it is
useful for us to describe an explicit construction of fiber products in EPG. To that end, we
recall the definition of a weak fiber products in a general 2-category C as in [12, Remark 2.2].
Let F, G, H : X → Y be morphisms in C such that there are 2-morphisms α : F ⇒ G and
β : G ⇒ H. Then we denote by
β◦α:F ⇒H
the (vertical) composition 2-morphism that attaches to any object x ∈ X0 the morphism
(β ◦ α)x := βx ◦ αx : F (x) −→ H(x).
Let F, H : X → Y and G, L : Y → Z be morphisms such that there are 2-morphisms
α : F ⇒ H and β : G ⇒ L. Then we denote by
β∗α:G◦F ⇒ L◦H
the (horizontal) composition 2-morphism that attaches to any object x ∈ X0 the morphism
(β ∗ α)x := L1 (αx ) ◦ βF (x) = βH(x) ◦ G1 (αx ) : G(F (x)) −→ L(H(x)).
In particular,
(IdG ∗α)x = G1 (αx ), (β ∗ IdF )x = βF (x) .
Let F : X → Z and G : Y → Z be morphisms C. A weak fiber product of F, G, is a
quadruple (P, A1 , A2 , α) where P is a 0-cell, A1 : P → X and A2 : P → Y are morphisms,
and α : G ◦ A2 ⇒ F ◦ A1 is an invertible 2-morphism,
A2
P / Y (12)
α q
A1 |t qq G
 
F
X / Z,

such that the two properties below are satisfied.


I. For any triple (D, B1 : D → X , B2 : D → Y), if there exists an invertible 2-morphism
α′ : G ◦ B2 ⇒ F ◦ B1 , then there exists a triple (U : D → P, β1 : B1 ⇒ A1 ◦ U, β2 :
B2 ⇒ A2 ◦ U) such that βj are invertible and
(α ∗ IdU ) ◦ (IdG ∗β2 ) = (IdF ∗β1 ) ◦ α′ : G ◦ B2 ⇒ F ◦ A1 ◦ U.

D❅ B2
❅ U
❅ β2 ✂
❅ | ✂ A2 $
C; P Y
β1
/
B1
⑦⑦ α′ q
A1 |t qqαq G
  F

X / Z

24
II. For any triple (D, U : D → P, U ′ : D → P), if there exist invertible 2-morphisms
γj : Aj ◦ U ⇒ Aj ◦ U ′ for j = 1, 2, such that
(α ∗ IdU ′ ) ◦ (IdG ∗γ2 ) = (IdF ∗γ1 ) ◦ (α ∗ IdU ) : G ◦ A2 ◦ U ⇒ F ◦ A1 ◦ U ′ ,
then there exists a unique invertible γ : U ⇒ U ′ such that
γj = IdAj ∗γ, j = 1, 2.

❧ ❥ D✬ ❘ ❖

s ♦ ✩ ❑
A1 ◦U ⑦ ❅ A2 ◦U
γ
A1 ◦U ′ ks ❴❴ ✤ A2 ◦U ′ ✹

✚U ✰
✓ ❅❅❅❅❅γ1 U′ γ2 ⑧⑧
❅ # 
✗ ⑧

✚ P ▼▼ { ⑧⑧⑧⑧ ✩
qq ▼ ✦
✥✢ q qqqq ▼▼▼▼

qq ▼ ✣
✧
~xqqqq A1 A2 ▼▼▼
▼& ✜
X ❴❴
ks ❴❴❴ ❴ Y
▼▼▼ α
▼▼▼ F G qqqqq
▼▼▼ qq
▼▼▼
▼& qqqqq
xq
Z
Remark 5.1. In a bicategory, the weak fiber product is defined similarly, except the formulae
in properties I.-II. involve extra 2-morphisms that compensate for the lack of associativity.
See, e.g., Remark 2.2 in [12]. The precise expressions will not be required for our purposes.
To construct a weak fiber product in Orb, we follow the argument of [13, Lemma 26]. As a
first step, we describe a candidate weak fiber product in EPG. For this, consider morphisms
F : X → Z and G : Y → Z, and define a groupoid P by
P 0 = X 0 F 0 × s Z 1 t × G0 Y 0 ,
P1 = X1 s◦F1 ×s Z1 t ×s◦G1 Y1 ,
with the structure maps defined below. Let f = (f1 , f2 , f3 ) ∈ P1 . Write
x1 = s(f1 ), z1 = f2 , y1 = t(f1 ), x2 = s(f3 ), y2 = t(f3 ),
and take z2 so that the following diagram commutes.
z1
F0 (x1 ) / G0 (y1 )
F1 (f1 ) G1 (f3 )
 z2

F0 (x2 ) ❴ ❴ ❴ ❴ ❴ ❴/ G0 (y2 )
Define the source and the target of f by
s(f ) := (x1 , z1 , y1), t(f ) := (x2 , z2 , y2). (13)
It is immediate from definition that these indeed belong to P0 :
F0 (s(f1 )) = F0 (x1 ) = s(z1 ), G0 (s(f3 )) = G0 (y1) = t(z1 ),
F0 (t(f1 )) = F0 (x2 ) = s(z2 ), G0 (t(f3 )) = G0 (y2 ) = t(z2 ).
25
For p = (x, z, y) ∈ P0 , define the identity morphism by
e(p) := (e(x), z, e(y)). (14)
To verify that e(p) ∈ P1 , note that s ◦ F1 (e(x)) = F0 ◦ s(e(x)) = F0 (x) = s(z), and similarly
s ◦ G1 (e(y)) = t(G1 (e(y))) = t(z). For p, q, r ∈ P0 and f : p → q, g : q → r ∈ P1 , write
f = (f1 , f2 , f3 ), g = (g1 , g2 , g3 ), and define inverse and composition by
i(f ) := (i(f1 ), G1 (f3 ) ◦ f2 ◦ i(F1 (f1 )), i(f3 )), m(g, f ) := (m(g1 , f1 ), f2 , m(g3 , f3 )). (15)
To verify the images are indeed in P1 , compute
s(F1 (i(f1 )) = s(i(F1 (f1 ))) = s(G1 (f3 ) ◦ f2 ◦ i(F1 (f1 ))),
s(G1 (i(f3 ))) = t(G1 (f3 )) = t(G1 (f3 ) ◦ f2 ◦ i(F1 (f1 ))),
s ◦ F1 (m(g1 , f1 )) = F0 (s(f1 )) = s(f2 ),
s ◦ G1 (m(g3 , f3 )) = G0 (s(f3 )) = s(G1 (f3 )) = t(f2 ).
Lemma 5.2. Let F : X → Z and G : Y → Z be transverse strongly smooth morphisms in
EPG. Then the weak fiber product P := X F ×G Y exists in EPG with
P 0 = X 0 F 0 × s Z 1 t × G0 Y 0 ,
P1 = X1 s◦F1 ×s Z1 t ×s◦G1 Y1 ,
and structure maps defined via (13), (14), and (15).
Proof. By [5, Theorem 6.4] the topological spaces P0 , P1 , are manifolds with corners, and it
is immediate that the structure maps make P into an étal proper groupoid. Recall that in
EPG all 2-morphisms are automatically invertible.
Consider the specified P with A1 , A2 , projections on the first and third component, re-
spectively. We construct a 2-morphism α : (G ◦ A2 ) ⇒ (F ◦ A1 ) to complete diagram (12)
as follows. For any p = (x, z, y) ∈ P0 , define
αp := i(z) : (G ◦ A2 )0 (p) = G0 (y) −→ (F ◦ A1 )0 (p) = F0 (x).
It is immediate from definition that with α thusly defined, for any f : p → q ∈ P1 the
diagram
αp
G0 (A20 (p)) / F0 (A10 (p))
G1 (A21 (f )) F1 (A11 (f ))
 αq 
G0 (A20 (q)) / F0 (A10 (q))
commutes. Thus, (P, A1 , A2 , α) is a weak fiber product of F, G.
If (D, B1 , B2 , α′ ) are given as in property I., define U : D → P by
U0 (d) := (B10 (d), i(αd′ ), B20 (d)), d ∈ D0 ,
ˆ := (B11 (d),
U1 (d) ˆ
ˆ i(α′ ˆ ), B21 (d)), dˆ ∈ D1 .
s(d)

It is easy to see that the image of Uj is indeed in Pj for j = 0, 1 and that Bj = Aj ◦ U. Take
βj := IdBj , j = 0, 1. Then relation we need to prove becomes α ∗ IdU = α′ . To verify it, note
that for any d ∈ D0 the definition of α gives
(α ∗ IdU )d = αU (d) = α(B10 (d),i(α′d ),B20 (d)) = i(i(αd′ )) = αd′ ,
26
as desired.
To verify property II., let (D, U, U ′, γ1 , γ2 ) be a quintuple that satisfies the assumptions.
Set

γ := γ1 × i(α ∗ IdU ) × γ2 .

Thus defined, γ satisfies γj = IdAj ∗γ for j = 1, 2. Since for any d ∈ D0 the following square
is Cartesian,

A21
HomP (U0 (d), U0′ (d)) / HomY (A20 (U0 (d)), A20 (U0′ (d)))
G1

A11 HomZ (G0 (A20 (U0 (d), G0 (A20 (U0′ (d))))
α◦−◦i(α)
 F1

HomX (A10 (U0 (d)), A10 (U0′ (d))) / HomZ (F0 (A10 (U0 (d))), F0(A10 (U0′ (d)))),

the morphism γd is uniquely determined by its projections Aj1 (γd ) = (IdAj ∗γ)d , j = 1, 2. So,
γ is unique. 

In the situation of Lemma 5.2, consider the weak pull-back diagram

A2
X F ×G Y / Y
α ♠♠
A1 ♠♠♠♠ G
 rz ♠♠ 
F
X / Z.

Given a relative orientation oG = (oG0 , oG1 ) of G, we define the pull-back relative ori-
entation F ∗ oG of A1 as follows. Consider the following diagram in which all squares are
Cartesian.

A20

B01 B02 %
X0 F0 ×s Z1 t ×G0 Y0 / Z 1 t × G0 Y 0 / Y0
C01 T0 G0
 S0
 
t
A10 X0 F 0 × s Z 1 / Z1 / Z0
C02 s
!  F0

X0 / Z0
27
1 C2 1
Let oC0 := S0∗ t∗ oG0 and let oA10 := oc 0 ◦ oC0 . Consider also the following diagram in which
all squares are Cartesian.
A21

B11 B12 %
X1 s◦F1 ×s Z1 t ×s◦G1 Y1 / Z1 t ×s◦G1 Y1 / Y1
C11 T1 s◦G1
 S1
 
t
A11 X1 s◦F1 ×s Z1 / Z1 / Z0
C12 s
!  s◦F1

X1 / Z0
1 C2 1
Let oC1 := S1∗ t∗ (osc ◦ oG1 ) and let oA11 := oc 1 ◦ oC1 . Then, F ∗ oG := (oA10 , oA11 ). Similarly,
given a relative orientation oF of F we can define the transpose pull-back orientation G∗ oF
of A2 .
The following is true by [12, Corollary 0.3 and Theorem 0.2].
Lemma 5.3. Let f = F |R : X ← X ′ → Z, g = G|Q : Y ← Y ′ → Z be transverse strongly
smooth morphisms in Orb. Let (P b = X ′ × Y ′ , Â1 , Â2 , α̂) be a weak fiber product in EPG
F G
as described in Lemma 5.2. Then a weak fiber product P = X f ×g Y exists in Orb that is
given by
P0 := Pb0 = X0′ F0 ×s Z1 t ×G0 Y0′ , P1 := Pb1 = X1′ s◦F1 ×s Z1 t ×s◦G1 Y1′ ,
with the projections
a1 := (R ◦ Â1 )| IdP : P −→ X , a2 := (Q ◦ Â2 )| IdP : P −→ Y,
and the 2-morphism α := IdP ∗α̂.
We can illustrate this construction via a commutative diagram in EPG:
Id Q◦Â2
X f ×g Y o X f ×g Y / YO
O O✤
Id Id
✤ Q

Id Â2
X f ×g Y o❴ ❴ ❴ ❴ ❴ ❴ X ′ F ×G Y ′ / Y′
R◦Â1 Â1 G
  
R F
X o X′ / Z
5.2. Integration properties in EPG. The objective of this section is to prove integra-
tion properties for differential forms on étale proper groupoids, particularly the analogue
Theorem 1. Note that the property analogous to Theorem 1(a) is covered by Lemma 4.9
above.
5.2.1. Property (b). The claim on composition of pullbacks is immediate from definition. For
push forward, we detail as follows.
Lemma 5.4. Let G : X → Y, F : Y → Z, be relatively oriented proper submersions in
EPG. Then F∗ ◦ G∗ = (F ◦ G)∗ .
28
Proof. Let ξ ∈ A∗ (X ). By Lemma 4.12 and Proposition 2.6(2) applied to F0 , G0 , we have
F∗ G∗ ξ = J(F0 )∗ KJ(G0 )∗ Kξ = J(F0 )∗ (G0 )∗ (Kξ) = J(F0 ◦ G0 )∗ (Kξ) = (F ◦ G)∗ ξ.


5.2.2. Property (c). First we establish two basic results.


Lemma 5.5. A∗cl (X ) is a module over A∗ (X ).
Proof. Let ξ ∈ A∗ (X ) and η ∈ im(s∗ −t∗ ) ⊂ A∗cl (X0 ). We need to show that ξ∧η ∈ im(s∗ −t∗ ).
Indeed, take ζ such that η = (s∗ − t∗ )(ζ). Then
ξ ∧ η = ξ ∧ s∗ ζ − ξ ∧ t∗ ζ
= s∗ (s∗ ξ ∧ ζ) − t∗ (t∗ ξ ∧ ζ),
and since s∗ ξ = t∗ ξ,
= (s∗ − t∗ )(s∗ ξ ∧ ζ).

Lemma 5.6. For ζ ∈ A∗ (X ) and η ∈ A∗cl (X0 ), we have J[ζ ∧ η] = ζ ∧ J[η].
Proof. By assumption, s∗ ζ = t∗ ζ. Therefore,
J[ζ ∧ η] = t∗ s∗ (ζ ∧ η)
= t∗ (s∗ ζ ∧ s∗ η)
= t∗ (t∗ ζ ∧ s∗ η),
which, by Proposition 2.6(3),
= ζ ∧ t∗ s∗ η = ζ ∧ J[η].

We are now ready to prove the integration property.
Lemma 5.7. Let F : X → Y be a relatively oriented proper submersion in EPG and let
ξ ∈ A∗ (Y), η ∈ A∗ (X ). Then F∗ (F ∗ ξ ∧ η) = ξ ∧ F∗ η.
Proof. Compute
F∗ (F ∗ ξ ∧ η) = JF∗ K(F0∗ ξ ∧ η)
= JF∗ [F0∗ ξ ∧ ρη]
= J[(F0 )∗ (F0∗ ξ ∧ ρη)]
which, by Proposition 2.6(3) applied to F0 ,
= J[ξ ∧ (F0 )∗ (ρη)],
and by Lemma 5.6,
= ξ ∧ J[(F0 )∗ (ρη)] = ξ ∧ F∗ η.

29
5.2.3. Property (d). We start with a series of three background lemmas. They lead up to
Lemma 5.11, which establishes a property of partitions of unity, which in turn is useful for
describing the push-forward of differential forms on fiber products.
Lemma 5.8. Let f : X → W , g : Y → W , and s, t : Z → W be continuous maps between
topological spaces such that s × t : Z → W × W is proper. Let X f ×s Z t ×g Y be the fiber
product in Top. Let A ⊂ X and B ⊂ Y be compact subsets. Then A f ×s Z t ×g B is compact.
Proof. First, note that A f ×s Z t ×g B is a closed subset of the direct product X × Z × Y .
Next,
A f ×s Z t ×g B = {(a, z, b) ∈ A × Z × B|f (a) = s(z), t(z) = g(b)}
⊂ {(a, z, b) ∈ A × Z × B|s(z) ∈ f (A), t(z) ∈ g(B)}
= A × (s × t)−1 (f (A) × g(B)) × B.
By assumption, A and B are compact. By continuity, f (A) × g(B) is compact, and by
properness, (s × t)−1 (f (A) × g(B)) is compact. Thus, the last line above is a compact subset
of X × Z × Y . So, A f ×s Z t ×g B is compact as a closed subset of a compact set. 
Lemma 5.9. Let F : X → Z, G : Y → Z, be transverse strongly smooth morphisms in
EPG and let P be the weak fiber product given by Lemma 5.2. If A ⊂ X0 and B ⊂ Y0 are
clean, then A F0 ×s Z1 t ×G0 B ⊂ P0 is clean.
Proof. Write for short C := A F0 ×s Z1 t ×G0 B ⊂ P0 . Denote by sD , tD , the source and target
maps of D = X , Y, P. Denote by πj the projection from P to the jth component for
j = 1, 2, 3, with a second index to indicate the projection of objects of P vs. morphisms
(e.g., π10 : P0 → X0 and π11 : P1 → X1 ). Let K ⊂ P0 be a compact subset. We show that
t−1 −1
P (K) ∩ sP (C) is compact.
First, by Lemma 4.18, A and B are closed. Thus, C and therefore t−1 −1
P (K) ∩ sP (C) is
closed as well. Next,
t−1 −1
P (K)∩sP (C) ⊂
 
⊂ π11 (t−1 −1 −1 −1
P (K)) s◦F1 ×s Z1 t ×s◦G1 π31 (tP (K)) ∩ sX (A) s◦F1 ×s Z1 t ×s◦G1 sY (B)
 
= π11 (t−1
P (K)) ∩ s −1
X (A) s◦F × s Z 1 ×
t s◦G1 π31 (t−1
P (K)) ∩ s −1
Y (B)
 1 
−1 −1 −1
⊂ tX (π10 (K)) ∩ sX (A) s◦F1 ×s Z1 t ×s◦G1 tY (π30 (K)) ∩ s−1 Y (B) .
By continuity, π1 (K) and π3 (K) are compact. From cleanness of A and B and by Lemma 5.8,
it follows that the last expression above is compact. In total, t−1 −1
P (K)∩sP (C) is compact. 

Let X, Y, Z, W, V, be manifolds with corners and let h : Z → V , k : W → V , f : X → Z,


g : Y → W , be strongly smooth maps with k ⋔ h and f, g, proper submersions. Let of , og ,
be relative orientations for f, g, respectively. Denote by
f × g : X h◦f ×k◦g Y → Z h ×k W
the induced map, and by
Y
πX : X h◦f ×k◦g Y → X, πYX : X h◦f ×k◦g Y → Y,
πZW : Z h ×k W → Z, Z
πW : Z h ×k W → W,
30
the projections. Equip f × g with the induced relative orientation defined as follows. Note
that
IdX ×g : X h◦f ×k◦g Y −→ X h◦f ×k W, f × IdW : X h◦f ×k W −→ Z h ×k W,
satisfy
f × g = (f × IdW ) ◦ (IdX ×g).
Consider the following pull-back diagrams
X
πY
f ×IdY
X h◦f ×k Y ≃ X f ×πZY (Z h ×k Y ) / Z h ×k Y, X h ×k◦g Y ≃ (X h ×k W ) πX ×g Y / Y
W
Y
πX Y
πZ g
IdX ×g
 f   X
πW 
X / Z X h ×k W / W.

Equip f × IdY with the transpose pull-back relative orientation t (πZY )∗ of and equip IdX ×g
X ∗
with the pull-back orientation (πW ) og . The induced relative orientation of f × g is defined
t Y ∗ f X ∗
to be the composition (πZ ) o ◦ (πW ) og .
Lemma 5.10. For all α ∈ A∗ (X) and β ∈ A∗ (Y ),
Y ∗
(f × g)∗ ((πX ) α ∧ (πYX )∗ β) = (−1)⋆ (πZW )∗ f∗ α ∧ (πW
Z ∗
) g∗ β,
with ⋆ = rdim f · (|β| + rdim k ◦ g).
Proof. By Proposition 2.6(2), it is enough to prove the claim for the case when one of the
maps is the identity, since then
Y ∗
(f × g)∗ ((πX ) α ∧ (πYX )∗ β) = (f × IdW )∗ ◦ (IdX ×g)∗ ((πX
Y ∗
) α ∧ (πYX )∗ β) =
Y ∗ Z ∗
= (f × IdW )∗ ((πX ) α ∧ (πW ) g∗ β) = (−1)rdim f ·(|β|+rdim g+rdim k) (πZW )∗ f∗ α ∧ (πW
Z ∗
) g∗ β.
Assume first that W = Y and g = IdY . Consider the following diagrams
f ×IdY f ×IdY
X h◦f ×k Y ≃ X f ×πZY (Z h ×k Y ) / Z h ×k Y X h◦f ×k Y / Z h ×k Y
■■ πX Z
πY ✇
■■ Y ✇✇
Y
πX Y
πZ ■■
■■ ✇✇✇

 f  ■$ {✇✇
X / Z, Y .

By Lemma 2.10 for the first diagram and commutativity of the second,
Y
(πZY )∗ f∗ α = (−1)rdim f ·rdim πZ (f × IdY )∗ (πX
Y ∗
) α, (πYX )∗ β = (f × IdY )∗ (πYZ )∗ β.
By Proposition 2.6(3),
Y ∗
(f × g)∗ ((πX ) α ∧ (πYX )∗ β) = (f × IdY )∗ ((πX
Y ∗
) α ∧ (f × IdY )∗ (πYZ )∗ β)
= (−1)|α||β| (f × IdY )∗ ((f × IdY )∗ (πYZ )∗ β ∧ (πX
Y ∗
) α)
= (−1)|α||β| (πYZ )∗ β ∧ (f × IdY )∗ (πX
Y ∗

Y ∗ α|)
= (−1)|β|(|α|+|(f ×IdY )∗ (πX ) Y ∗
(f × IdY )∗ (πX ) α ∧ (πYZ )∗ β
Y
= (−1)|β|(|α|+|f∗α|)+rdim f ·rdim πZ (πZY )∗ f∗ α ∧ (πYZ )∗ (IdY )∗ β.
31
The case f = IdX is proved similarly, except no signs arise. The diagrams

X
πY IdX ×g
X h ×k◦g Y ≃ (X h ×k W ) πX ×g Y / Y X h ×k◦g Y / X h ×k W
W ■■ πY W ✈
■■ X πX ✈✈
IdX ×g g ■■
■■ ✈✈✈

 X
πW  ■$ z ✈

X h ×k W / W, X .

give

Y ∗
(f × g)∗((πX ) α ∧ (πYX )∗ β) = (IdX ×g)∗ ((IdX ×g)∗ (πX
W ∗
) α ∧ (πYX )∗ β)
W ∗
= (πX ) α ∧ (IdX ×g)∗ (πYX )∗ β
W ∗ X ∗
= (πX ) (IdX )∗ α ∧ (πW ) g∗ β.

Lemma 5.11. Let (12) be a weak pull-back diagram in EPG with P = X F ×G Y as in


Lemma 5.2. Let ρX , ρY , be partitions of unity on X , Y, respectively. Then ρ := A∗10 ρX · A∗20 ρY
is a partition of unity on P.

Proof. By Lemma 5.9, the map ρ : P0 → [0, 1] is cleanly supported. It is left to verify that
(tP )∗ (sP )∗ ρ = 1.
Write sD , tD , for the source and target maps of D for D = X , Y, Z, P. Let π1 : P → X ,
π2 : P → Z1 , and π3 : P → Y, be the projection maps. We use a second index to indicate
the projection of objects of P vs. morphisms, e.g., π10 : P0 → X0 and π11 : P1 → X1 . Then

(tP )∗ (sP )∗ (π10


∗ ∗
ρX ∧ π30 ∗ ∗
ρY ) = (tP )∗ (π11 ∗ ∗
sX ρX ∧ π31 sY ρY ),

= (tP )∗ (π11 (s∗X ρX ) ∧ π21
∗ ∗
1 ∧ π31 (s∗Y ρY )),

by Lemma 5.10 applied repeatedly to tP = tX × IdZ1 ×tY ,


= π10 ((tX )∗ s∗X ρX ) ∧ π20
∗ ∗
1 ∧ π30 ((tY )∗ s∗Y ρY )
= 1.

We are now ready to prove the integration property.

Lemma 5.12. Let (12) be a weak pull-back diagram in EPG, where G is a proper submersion
with relative orientation oG . Equip A1 with the pull-back relative orientation F ∗ oG . Let ξ ∈
A∗ (Y). Then (A1 )∗ (A2 )∗ ξ = F ∗ G∗ ξ.
32
Proof. Writing explicitly the weak fiber product as in Lemma 5.2, we get the following
commutative diagram,
A20

B01 B02 %
X0 F0 ×s Z1 t ×G0 Y0 / Z 1 t × G0 Y 0 / Y0 (16)
C01 T G0
  
S t
A10 X0 F 0 × s Z 1 / Z1 / Z0
C02 s
!  F0

X0 / Z0
where each of the three squares is Cartesian in the category of manifolds with corners, and
the right vertical arrow in each is a submersion, while G0 and T are also proper. The relative
orientation of s is the canonical one, the relative orientations of T, C01 , and C02 , are given by
2 1
pull-back, and oA10 := oC0 ◦ oC0 .
Let ρX and ρY be partitions of unity on X , Y, respectively. In view of Lemma 5.11, we
take ρ := (A10 )∗ ρX · (A20 )∗ ρY : P0 → [0, 1], a partition of unity on P. For ξ ∈ A∗ (Y), we
have
(A1 )∗ A∗2 ξ = (A1 )∗ A∗20 ξ
= J[(A10 )∗ (ρ · A∗20 ξ)]
= J[(A10 )∗ (A∗10 ρX · A∗20 ρY · A∗20 ξ)]
= J[(A10 )∗ (A∗10 ρX · A∗20 (ρY · ξ))]
By Proposition 2.6(3),
= J[ρX · (A10 )∗ A∗20 (ρY · ξ)]
= JK((A10 )∗ A∗20 (ρY · ξ))
= (A10 )∗ A∗20 (ρY · ξ)
= (C02 )∗ (C01 )∗ (B01 )∗ (B02 )∗ (ρY · ξ)
by Proposition 2.6(4),
= (C02 )∗ S ∗ T∗ (B02 )∗ (ρY · ξ)
= F0∗ s∗ t∗ (G0 )∗ (ρY · ξ)
by Lemma 4.10(1),
= F0∗ t∗ s∗ (G0 )∗ (ρY · ξ)
by Lemma 4.10(2),
= F ∗ J(G0 )∗ K(ξ)
= F ∗ G∗ ξ.

33
5.2.4. Property (e). For X , Y, in EPG and a strongly smooth F : X → Y, decompose ∂X
into vertical and horizontal parts ∂X = ∂ v X ⊔ ∂ h X , where
(∂ v X )j = ∂ v Xj , (∂ h X )j = ∂ h Xj , j = 0, 1.
They come with maps
ivX : ∂ v X −→ X , ihX : ∂ h X −→ X , (17)
and the structure of an étale proper groupoids induced by restricting the structure maps
of ∂X . Furthermore, the above maps come with relative orientations coming from the
iX iX
compositions ∂ v X ֒→ ∂X → X and ∂ h X ֒→ ∂X → X.
Lemma 5.13. If ρ is a partition of unity on X , then ρ∂ := i∗X ρ is a partition of unity on
∂X . For a strongly smooth F : X → Y, the map ρv∂ := (ivX )∗ ρ is a partition of unity on ∂ vX .
Proof. Denote by s∂ , t∂ , the source and target maps of ∂X . Denote by i0 and i1 the inclusions
i1 : ∂X1 ֒→ X1 and i0 : ∂X0 ֒→ X0 , respectively. Consider the pull-back diagrams
i1 i1
∂X1 / X1 ∂X1 / X1
s∂ s t∂ t
 i0
  i0

∂X0 / X0 , ∂X0 / X0 .
By the commutativity of the diagrams, Proposition 2.6(4), and the assumption t∗ s∗ ρ = 1,
we get
(t∂ )∗ (s∂ )∗ (i∗0 ρ) =(t∂ )∗ i∗1 s∗ ρ
=i∗0 t∗ s∗ ρ
=1.
A similar argument goes through for (iv∂ )∗ ρ. 
As a first step in proving the integration property on A∗ (X ), note that its analogue holds
on A∗cl (X ):
Lemma 5.14. Let F : X → Y be a strongly smooth relatively oriented proper submersion in
EPG, let m := dim X , and let ξ ∈ Akcl (X ). Then d(F∗ ξ) = F∗ (dξ) + (−1)m+k (F |∂ vX )∗ (ξ|∂ vX ).
Proof. This is immediate from Stokes’ theorem, Proposition 2.7, for F0 . 
To deduce the result for A∗ (X ), we note the following two lemmas.
Lemma 5.15. The exterior derivative d descends to A∗cl (X ).
Proof. Since s, t, are local diffeomorphisms, their fibers are zero dimensional. By Stokes’
theorem, this implies s∗ , t∗ , commute with d, and so s∗ − t∗ is a chain map. It follows that
d descends to the cokernel, A∗cl (X ). 
Lemma 5.16. The exterior derivative d commutes with J, K.
Proof. Since t is a local diffeomorphism, and as such has no fiberwise boundary, d commutes
with J = t∗ s∗ . Since K is inverse to J, it follows that d commutes with K as well:
d◦J = J ◦d =⇒ d=J ◦d◦K =⇒ K ◦ d = d ◦ K.

34
We are now ready to prove the integration property.
Lemma 5.17. Let F : X → Y be a strongly smooth relatively oriented proper submersion in
EPG, let m := dim X , and let ξ ∈ Ak (X ). Then d(F∗ ξ) = F∗ (dξ) + (−1)m+k (F |∂ vX )∗ (ξ|∂ vX ).
Proof. Applying Lemma 5.14 to Kξ and using Lemma 5.16, we get
d((F0 )∗ (Kξ)) =(F0 )∗ (d(Kξ)) + (−1)m+k (F0 |∂ vX0 )∗ ((Kξ)|∂ vX )
=(F0 )∗ (K(dξ)) + (−1)m+k (F0 |∂ vX0 )∗ ((Kξ)|∂ vX ).
Applying J on the left and using Lemma 5.16 again we get
d(J(F0 )∗ (Kξ)) = J(F0 )∗ (K(dξ)) + (−1)m+k J(F0 |∂ vX0 )∗ ((Kξ)|∂ vX ),
Note that Lemma 5.13 implies (Kξ)|∂ vX = [ρ|∂ v X · ξ|∂ v X ], where ρ is a partition of unity on
X . Thus, we get
dF∗ (ξ) = F∗ (dξ) + (−1)m+k (F |∂ vX )∗ (ξ|∂ vX ),
as desired. 
5.3. Integration properties in Orb. The objective of this section is to prove Theorem 1.
Note that with our convention (9), Definition 4.17 reads, for f = F |R,
f ∗ = R∗ F ∗ , f∗ = F∗ R∗ .
5.3.1. Proof of Theorem 1(a). Write f = F |R, g = G|Q, and let the following diagram
represent the 2-morphism α : f ⇒ g.

X ❚
③= ❇❇❇❚❚❚❚❚❚
③③③ ❇❇ ❚❚❚❚F
③③ ❇❇ ❚❚❚❚
③③ T1 R ❇ ❚❚❚❚
′′′
❚❚)
X ❉ ⇓ α 1 X ⇓ α2 Y
❥❥❥❥
> 5
❉❉ ⑤ ⑤ ❥❥ ❥
❉❉T2 Q ⑤⑤
❉❉ ⑤⑤ ❥❥❥❥❥G ❥❥❥❥
❉! ⑤
⑤ ❥❥❥❥
′′ ❥
X
By Lemmas 4.16, 5.4, and 4.9, we have
g ∗ = Q∗ G∗ = Q∗ (T2 )∗ T2∗ G∗ = R∗ (T1 )∗ T1∗ F ∗ = R∗ F ∗ = f ∗ .
Similarly,
g∗ = G∗ Q∗ = G∗ (T2 )∗ T2∗ Q∗ = F∗ (T1 )∗ T1∗ R∗ = F∗ R∗ = f∗ .
5.3.2. Composition of morphisms. For the next two parts of Theorem 1, we describe more
explicitly how composition works in Orb.
Lemma 5.18. Let F : X → Z, G : Y → Z, be transverse strongly smooth morphisms in
EPG, and let P := X ×Z Y be their weak fiber product described in Lemma 5.2. Let A1 , A2 ,
be the projections as in diagram (12). Then
(1) If G0 is a local diffeomorphism, then so is A10 .
(2) If G is essentially surjective, then so is A1 .
(3) If G is fully faithful, then so is A1 .
(4) If G is a refinement, then so is A1 .
Proof.
35
(1) Consider diagram (16). Since G0 is a local diffeomorphism, so is T and therefore also
C01 . Since s is a local diffeomorphism, so is C02 . Therefore, A10 = C02 ◦ C01 is a local
diffeomorphism.
(2) Let x ∈ X0 . By assumption on G, there exist z ∈ Z1 and y ∈ Y0 such that s(z) =
F0 (x) and t(z) = G0 (y). Then p := (x, z, y) ∈ P0 and A10 (p) = x. Thus, A10 is
surjective and in particular A1 is essentially surjective.
(3) Let p, q ∈ P0 . Denote by
αpq : HomZ ((G0 ◦ A20 )(p), (G0 ◦ A20 )(q)) −→ HomZ ((F0 ◦ A10 )(p), (F0 ◦ A10 )(q))
the map given by
αpq (f ) := αq ◦ f ◦ i(αp ).
Then the following square is Cartesian:
A21
HomP (p, q) / HomY (A20 (p), A20 (q))
G1

A11 HomZ (G0 ◦ A20 (p), G0 ◦ A20 (q))
αpq
 F1

HomX (A10 (p), A10 (q)) / HomZ (F0 ◦ A10 (p), F0 ◦ A10 (q)).
From the assumption on G and the invertibility of α, the right column is a bijection.
Therefore, the left column is a bijection.
(4) This is a combination of the preceding three properties.

Q G R F
Let f : X ← X ′ → Y and g : Y ← Y ′ → Z be morphisms in Orb. Taking the weak fiber
product in EPG of G and R, we get the following diagram:
X ′ G ×R Y ′ (18)
t ■■
tt ■■
tt ■■
tt ■
tt R ′ G′ ■■■
t
z α $ ′
X′ ❏ ⇒ Y ❆
Q ⑥⑥
⑥ ❏❏
❏❏ G ttt ❆❆
⑥ ❏❏ R tt ❆❆F
⑥⑥ ❏❏ tt ❆❆
⑥ ❏ tt ❆
~⑥ g ❏% ytt f
X /Y / Z.

By Lemma 5.18, the projection R′ is a refinement. Thus, so is Q ◦ R′ , and we can define


f ◦ g to be (F ◦ G′ )|(Q ◦ R′ ).
Remark 5.19. The weak fiber product offers a natural choice of squares required for the
definition of composition in [10, Section 2.2], as remarked in [13, Lemma 23].
5.3.3. Proof of Theorem 1(b). Write g = G|Q, f = F |R, and f ◦ g = (F ◦ G′ )|(Q ◦ R′ ), as in
diagram (18) above. By Lemmas 5.4 and 5.12,
(f ◦ g)∗ = (Q ◦ R′ )∗ (F ◦ G′ )∗ = Q∗ R∗′ (G′ )∗ F ∗ = Q∗ G∗ R∗ F ∗ = g ∗ f ∗ .
Similarly,
(f ◦ g)∗ = (F ◦ G′ )∗ (Q ◦ R′ )∗ = F∗ G′∗ (R′ )∗ Q∗ = F∗ R∗ G∗ Q∗ = f∗ g∗ .
36
5.3.4. Proof of Theorem 1(c). Write f = F |R. Then,
f∗ (f ∗ ξ ∧ η) = F∗ R∗ (R∗ F ∗ ξ ∧ η)
= F∗ (R∗ R∗ F ∗ ξ ∧ R∗ η),
by Lemma 4.16,
= F∗ (F ∗ ξ ∧ R∗ η)
by Lemma 5.7,
= ξ ∧ F∗ R∗ η
= ξ ∧ f∗ η.

R F Q G
5.3.5. Proof of Theorem 1(d). Write f : X ← X ′ → Z and g : Y ← Y ′ → Z. By Lemma 5.3,
we may write p = (Q ◦ Â2 )| Id and q = (R ◦ Â1 )| Id. Applying Lemma 5.12 to the smaller
square in
Id Q◦Â20
X0′ ×Z0 Z1 ×Z0 Y0′ o X0′ ×Z0 Z1 ×Z0 Y0′ / YO 0
O
Id Q
Â20
X0′ ×Z0 Z1 ×Z0 Y0′ X0′ ×Z0 Z1 ×Z0 Y0′ / Y0′
R◦Â10 Â10 G
  
R F
X0 o X0′ / Z0
we get
f ∗ g∗ = R∗ F ∗ G∗ Q∗ = R∗ (Â10 )∗ Â∗20 Q∗ = q∗ p∗ .
R F
5.3.6. Proof of Theorem 1(e). Write f : X ← X ′ → Y. Note that R restricts to a refinement
R|∂ vX ′ : ∂ vX ′ → ∂ vX , so f ∂ vX = (F ∂ vX ′ )|(R ∂ vX ′ ). Applying Lemma 5.17 to R∗ ξ we get

f∗ (dξ) + (−1)s+t f ∂ vX ∗ ξ = F∗ R∗ (dξ) + (−1)s+t (F ∂ vX ′ )∗ (R ∂ vX ′ )∗ (ξ ∂ vX )

= F∗ (dR∗ ξ) + (−1)s+t (F v ′ )∗ (R∗ ξ) v ′
∂ X ∂ X
= d(F∗ R∗ ξ)
= d(f∗ ξ).

6. Currents
Let X be an object of Orb. We denote by A∗c (X ) ⊂ A∗ (X ) the locally convex subspace
of differential forms with compact support on X , meaning,
A∗c (X ) = {ξ ∈ A∗ (X ) | π(supp(ξ)) ⊂ |X | is compact} .
Thus, for f : X → Y a proper morphism, the pull-back f ∗ : A∗c (Y) → A∗c (X ) is well-defined.
We need f to be proper to guarantee it preserves compact support. If, on the other hand,
f is a relatively oriented submersion, the push-forward f∗ : A∗c (Y) → A∗c (X ) is well-defined
even without the properness assumption, because the restriction to the support of the given
form is always proper.
37
Denote by Ak (X ) the space of currents of cohomological degree k, that is, the continu-
ous dual space of the compactly supported differential forms Adim c
X −k
(X ) with the weak*
topology. When X is oriented, differential forms are identified as a subspace of currents by
ϕ : Ak (X ) ֒→ Ak (X ),
Z
ϕ(γ)(η) = γ ∧ η, η ∈ Adimc
X −k
(X ).
X

The exterior derivative extends to currents via dζ(η) = (−1)1+|ζ| ζ(dη). In particular, if
∂X = ∅, we have dϕ(γ) = ϕ(dγ). Currents are a bimodule over differential forms with
(η ∧ ζ)(γ) := (−1)|η|·|ζ|ζ(η ∧ γ), γ ∈ A∗c (X ), η ∈ A∗ (X ), ζ ∈ A∗ (X ),
and
(ζ ∧ η)(γ) := ζ(η ∧ γ), γ ∈ A∗c (X ), η ∈ A∗ (X ), ζ ∈ A∗ (X ).
Let f : X → Y be a proper morphism in Orb. Set rdim f := dim Y − dim X . Define the
push-forward
f∗ : Ak (X ) → Ak−rdim f (Y) (19)
by
(f∗ ζ)(η) = (−1)m·rdim f ζ(f ∗η), η ∈ Am
c (Y).

So, when f is a submersion, f∗ ϕ(α) = ϕ(f∗ α). Similarly, for f : X → Y a relatively oriented
submersion, define the pull-back
f ∗ : Ak (Y) → Ak (X ) (20)
by
(f ∗ ζ)(η) = ζ(f∗ η), η ∈ Am
c (Y).

We can now formulate an analogue of Theorem 1 for currents:


Proposition 6.1.
(i) Let f, g : X → Y be morphisms of orbifolds with corners and let α : f ⇒ g be
a 2-morphism. If f, g, are proper, then f∗ = g∗ : Ak (X ) → Ak−rdim f (Y). If f, g,
are relatively oriented submersions and α is relatively oriented, then also f ∗ = g ∗ :
A∗ (Y) → A∗ (X ).
(ii) Let g : X → Y, f : Y → Z, be proper morphisms of orbifolds with corners. Then
(f∗ ◦ g∗ )(ζ) = (f ◦ g)∗ζ, ∀ζ ∈ A∗ (X ).
If f, g, are relatively oriented submersions, then
(g ∗ ◦ f ∗ )(ζ) = (f ◦ g)∗ ζ, ∀ζ ∈ A∗ (Z).
(iii) Let f : X → Y be a proper morphism of orbifolds with corners. Then
f∗ (f ∗ η ∧ ζ) = η ∧ f∗ ζ, ∀ η ∈ A∗ (Y), ζ ∈ A∗ (X ).
If in addition f is a relatively oriented submersion, then
f∗ (f ∗ ζ ∧ η) = ζ ∧ f∗ η, ∀ η ∈ A∗ (X ), ζ ∈ A∗ (Y).
38
(iv) Let
p
X ×Z Y / Y
q ✈ g
~v ✈✈✈✈✈
 f 
X / Z
be a weak pull-back diagram of orbifolds with corners, where f and g are strongly
smooth, f is a relatively oriented submersion and g is proper. Equip p with the
transpose pull-back relative orientation and let ζ ∈ A∗ (Y). Then
q∗ p∗ ζ = f ∗ g∗ ζ.
Proof.
(i) By definition, rdim f = rdim g. By Theorem 1(a), for all γ ∈ A∗c (X ), we have
(f∗ ζ)(γ) = (−1)|γ|·rdim f ζ(f ∗γ) = (−1)|γ|·rdim g ζ(g ∗γ) = (g∗ ζ)(γ).
Similarly, for all γ ∈ A∗c (Y),
(f ∗ ζ)(γ) = (−1)|γ|·rdim f ζ(f∗γ) = (−1)|γ|·rdim g ζ(g∗γ) = (g ∗ ζ)(γ).
(ii) By Theorem 1(b), for all γ ∈ A∗c (X ) we have

(f ◦ g)∗ (ζ)(γ) = (−1)|γ|·rdim(f ◦g) ζ((f ◦ g)∗ γ) = (−1)|γ|·(rdim f +rdim g) ζ(g ∗f ∗ γ) =


= (−1)|γ|·rdim f (g∗ ζ)(f ∗γ) = (f∗ g∗ ζ)(γ).
Similarly, for all γ ∈ A∗c (Z),
(f ◦ g)∗(ζ)(γ) = ζ((f ◦ g)∗ γ) = g ∗ f ∗ ζ(γ).
(iii) For all γ ∈ A∗c (Y), we have
f∗ (f ∗ η ∧ ζ)(γ) =(−1)|γ|·rdim f (f ∗ η ∧ ζ)(f ∗γ)
=(−1)|γ|·rdim f +|η|·|ζ|ζ(f ∗ η ∧ f ∗ γ)
=(−1)|γ|·rdim f +|η|·|ζ|+(|η|+|γ|)·rdim f (f∗ ζ)(η ∧ γ)
=(−1)|η|·|ζ|+|η|·rdim f +|η|(|ζ|−rdim f ) (η ∧ f∗ ζ)(γ)
=(η ∧ f∗ ζ)(γ).
If f is a relatively oriented submersion,
f∗ (f ∗ ζ ∧ η)(γ) =(−1)|γ|·rdim f (f ∗ ζ ∧ η)(f ∗ γ)
=(−1)|γ|·rdim f (f ∗ ζ)(η ∧ f ∗ γ)
=(−1)|γ|·rdim f ζ(f∗ (η ∧ f ∗ γ))
=(−1)|γ|·rdim f +|η||γ| ζ(f∗(f ∗ γ ∧ η))
=(−1)|γ|·rdim f +|η||γ| ζ(γ ∧ f∗ η)
=(−1)|γ|·rdim f +|η||γ|+|γ||f∗η| ζ(f∗η ∧ γ)
=(ζ ∧ f∗ η)(γ).
39
(iv) Using the orbifold analog of Lemma 2.10 and rdim g = rdim q, we get
(q∗ p∗ ζ)(γ) =(−1)|γ|·rdim q (p∗ ζ)(q ∗ γ)
=(−1)|γ|·rdim q ζ(p∗ q ∗ γ)
=(−1)|γ|·rdim q+rdim f ·rdim g ζ(g ∗f∗ γ)
=(−1)|γ|·rdim q+rdim f ·rdim g+|f∗ γ|·rdim g (f ∗ g∗ ζ)(γ)
=(f ∗ g∗ ζ)(γ).

We proceed to formulate a version of Theorem 1(e) for currents. For this, we need a
notion of restriction to the boundary. However, restricting a current to the boundary in
general requires conditions on its wavefront set. To formulate the analogue of Theorem 1(e)
in elementary terms, we consider only currents that come from differential forms by the
inclusion ϕ. However, as noted above, ϕ commutes with d only if ∂X = ∅. Otherwise, a
correction is needed that depends on the boundary, in the spirit of Stokes’ theorem. To avoid
the issue, we introduce the following modification of the complex of currents.
Define differential forms relative to the boundary by
A∗c (X , ∂X ) = {η ∈ A∗c (X ) | i∗X η = 0},
and denote by Ak0 (X ) the dual space of Adim
c
X −k
(X , ∂X ). Consider the inclusion
ψ : A∗ (X ) −→ A∗0 (X ).
This now produces an operation that commutes with d even for orbifolds with non-empty
boundary:
Lemma 6.2. d ◦ (ψ ◦ ϕ) = (ψ ◦ ϕ) ◦ d.
Proof. For η ∈ A∗ (X ) and γ ∈ A∗c (X , ∂X ), we have
d ◦ (ψ ◦ ϕ)(η)(γ) =(−1)1+|η| (ψ ◦ ϕ)(η)(dγ)
Z
1+|η|
=(−1) η ∧ dγ
X
Z Z
=− d(η ∧ γ) + dη ∧ γ
X X
Z
= dη ∧ γ
X
=(ψ ◦ ϕ)(dη)(γ).

The last thing we need before formulating a version of Stokes’ theorem for currents is the
following observation.
Lemma 6.3. Let f : M → N be a strongly smooth map of manifolds with corners and let
γ ∈ A∗ (N) such that i∗N γ = 0. Then (ihM )∗ f ∗ γ = 0.
40
Proof. With the notation of Section 2.1.1, we have the following commutative diagram.
ξ ih
M
Ξ / ∂hM / M
π2 f
 iN 
∂N / N
By [5, Proposition 4.2], Ξ can be given the structure of a smooth manifold with corners
such that ξ is a covering map and π2 is smooth. By assumption, π2∗ i∗N γ = 0, and therefore
ξ ∗ (ihM )∗ f ∗ γ = 0. Since ξ is a surjective submersion, it follows that (ihM )∗ f ∗ γ = 0. 
Remark 6.4. In the situation of Lemma 6.3, it is not true in general that (ivM )∗ f ∗ γ = 0. Thus
pull-back of differential forms does not preserve differential forms relative to the boundary.
So, it is better to work with usual differential forms when defining pull-back and push-
forward.
In the following, we consider only oriented orbifolds so that the inclusion ϕ : A∗ (X ) →

A (X ) is defined. More generally, one should consider differential forms with coefficients in
a local system. Denote by ϕ∂ v the inclusion
ϕ∂ v : A∗ (∂X ) −→ A∗ (∂ v X ).
Proposition 6.5. Let f : X → Y be a strongly smooth proper morphism of oriented orbifolds
with corners with dim X = s, and let η ∈ At (X ). Then
ψ(f∗ (ϕ(dη))) = dψ(f∗ (ϕ(η))) + (−1)s+t ψ((f ◦ ivX )∗ (ϕ∂ v )((ivX )∗ η)).
dim X −|η|−1
Proof. For γ ∈ Ac (Y, ∂Y), compute
ψ(d(f∗ ϕ(η)))(γ) =(−1)1+|η|+rdim f (f∗ ϕ(η))(dγ)
=(−1)1+|η|+rdim f +rdim f ·(1+|γ|) ϕ(η)(f ∗ dγ)
Z
1+|η|+rdim f ·|γ|
=(−1) η ∧ d(f ∗ γ)
Z X Z
1+rdim f ·|γ| ∗

=(−1) d(η ∧ f γ) − dη ∧ f ∗ γ
ZX ZX

=(−1)1+rdim f ·|γ| (η ∧ f ∗ γ) − dη ∧ f ∗ γ .
∂X X
By Lemma 6.3,
Z Z
1+rdim f ·|γ| ∗

=(−1) (η ∧ f γ) − dη ∧ f ∗ γ
∂v X X
1+rdim f ·|γ| ∗

=(−1) ϕ (η|∂v ∂v X )(f γ| ∂v X ) − ϕ(dη)(f ∗ γ)
=(−1)1+rdim f ·|γ|+|γ|·(rdim f +1) (f |∂ v X )∗ ϕ∂ v (η|∂ v X )(γ)−
− (−1)1+rdim f ·|γ|+|γ|·rdim f f∗ ϕ(dη)(γ)
=(−1)1+|γ| (f |∂ v X )∗ ϕ∂ v (η|∂ v X )(γ) + f∗ ϕ(dη)(γ)
=(−1)dim X +|η| (f |∂ v X )∗ ϕ∂ v (η|∂ v X )(γ) + f∗ ϕ(dη)(γ).

41
References
[1] K. Behrend, Cohomology of stacks, Intersection theory and moduli, ICTP Lect. Notes, XIX, Abdus
Salam Int. Cent. Theoret. Phys., Trieste, 2004, pp. 249–294.
[2] K. Fukaya, Y.-G. Oh, H. Ohta, and K. Ono, Lagrangian intersection Floer theory: anomaly and obstruc-
tion. Parts I,II, AMS/IP Studies in Advanced Mathematics, vol. 46, American Mathematical Society,
Providence, RI; International Press, Somerville, MA, 2009.
[3] P. Gabriel and M. Zisman, Calculus of fractions and homotopy theory, Ergebnisse der Mathematik und
ihrer Grenzgebiete, Band 35, Springer-Verlag New York, Inc., New York, 1967.
[4] H. Hofer, K. Wysocki, and E. Zehnder, Polyfold and Fredholm Theory, arXiv e-prints (2017),
arXiv:1707.08941, arXiv:1707.08941.
[5] D. Joyce, On manifolds with corners, Advances in geometric analysis, Adv. Lect. Math. (ALM), vol. 21,
Int. Press, Somerville, MA, 2012, pp. 225–258.
[6] D. Joyce, A generalization of manifolds with corners, Adv. Math. 299 (2016), 760–862,
doi:10.1016/j.aim.2016.06.004.
[7] R. Kupferman and J. P. Solomon, A Riemannian approach to reduced plate, shell, and rod theories, J.
Funct. Anal. 266 (2014), no. 5, 2989–3039, doi:10.1016/j.jfa.2013.09.003.
[8] E. Lerman, Orbifolds as stacks?, Enseign. Math. (2) 56 (2010), no. 3-4, 315–363,
doi:10.4171/LEM/56-3-4.
[9] I. Moerdijk, Orbifolds as groupoids: an introduction, Orbifolds in mathematics and physics (Madi-
son, WI, 2001), Contemp. Math., vol. 310, Amer. Math. Soc., Providence, RI, 2002, pp. 205–222,
doi:10.1090/conm/310/05405.
[10] D. A. Pronk, Etendues and stacks as bicategories of fractions, Compositio Math. 102 (1996), no. 3,
243–303.
[11] J. P. Solomon, Involutions, obstructions and mirror symmetry, arXiv e-prints (2018),
arXiv:1810.07027.
[12] M. Tommasini, Weak fiber products in a bicategory of fractions, arXiv e-prints (2014), arXiv:1412.3295.
[13] A. N. Zernik, Moduli of open stable maps to a homogeneous space, arXiv e-prints (2017),
arXiv:1709.07402.

Institute of Mathematics, Hebrew University, Givat Ram, Jerusalem, 91904, Israel


Email address: jake@math.huji.ac.il

School of Mathematics, Institute for Advanced Study, Princeton, NJ 08540, USA


Email address: sarabt1@gmail.com

42

You might also like