The Volume Measure for Flat Connections as
Limit of the Yang–Mills Measure
Ambar N. Sengupta
Department of Mathematics
Louisiana State University
Baton Rouge, LA 70803, USA
e-mail:
[email protected]
Abstract
We prove that integration over the moduli space of flat connections
can be obtained as a limit of integration with respect to the Yang–
Mills measure defined in terms of the heat kernel for the gauge group.
In doing this we also give a rigorous proof of Witten’s formula for the
symplectic volume of the moduli space of flat connections. Our proof
uses an elementary identity connecting determinants of matrices along
with a careful accounting of certain dense subsets of full measure in
the moduli space. 2000MSC: 81T13
1
Introduction
Summary and Brief Background
We work with a closed, oriented surface Σ of genus g ≥ 2, and
a compact, connected, semisimple Lie group G equipped with a biinvariant metric. The space A of all connections on a principal G–
bundle over Σ has a natural symplectic structure which is preserved
by the pullback action ω 7→ φ∗ ω of the group G of bundle automorphisms φ. The moment map turns out to be J : ω 7→ Ωω , where
Ωω denotes the curvature of any connection ω. In this setting, the
Marsden-Weinstein procedure can be carried out rigorously [19] and
produces a symplectic structure Ω on the smooth strata of J −1 (0)/G.
Since J −1 (0) is the set of connections with zero curvature, J −1 (0)/G
1
is the moduli space of flat connections. This space, along with the
symplectic structure Ω on it, is of interest from many different points
of view (as attested to by the collection [14]). To be precise, J −1 (0)/G
is not, in general, a smooth manifold but there is a subset M0g (arising
from points of J −1 (0) of “minimal” isotropy) which is a manifold and
Ω is a symplectic structure on M0g .
In this paper we
• give a rigorous proof of Witten’s formula ([24, formula (4.72)])
volΩ (M0g ) = |Z(G)| vol(G)2g−2
1
α (dim α)2g−2
P
(1)
for the symplectic volume of the moduli space M0g of flat connections, for a compact, semisimple gauge group G, over a closed
oriented surface of genus g ≥ 2 (terminology, notation, and hypotheses are explained in detail later in this introduction; note
also that M0g is actually a subset of the full moduli space of flat
connections), and
• prove Forman’s theorem [8, Theorem 1] that Wilson loop expectations in the quantum Yang-Mills theory converges to the
corresponding symplectic integrals.
We will keep things as self-contained as reasonably possible and no
knowledge of the moduli space of flat connections is actually necessary
to understand the technical content of this paper. Indeed we shall work
with a standard realization of M0g as a finite-dimensional manifold.
Our proof has two main ingredients:
(i) a determinant identity (Proposition 1),
(ii) careful accounting of certain dense subsets of full measure in the
moduli space M0g where nice properties hold.
Witten [24] determined the symplectic volume of the moduli space
of flat connections in several different ways. One way involves the limit
of the partition function of the quantum Yang-Mills theory over the
surface. It is this approach, involving the heat kernel on the structure
(gauge) group, that we follow here. Forman used this approach and
Witten’s volume formula to prove the convergence of the Wilson loop
expectations. Liu [15, 16] used Forman’s approach along with other
ideas to study the symplectic volume and related integrals. We refer
to the collection [23], and the bibliography therein, for other works
concerning the symplectics of the moduli space of flat connections.
2
In the present paper we restrict our attention to the moduli space
of flat connections without distinguishing between bundles of different
topological type. The methods used here should extend to bundles of
specified topology and also to the case of surfaces with boundary but
this is not carried out here.
The limiting result we prove can be reformulated to give the limit
of the discrete Yang-Mills measure for cell-complexes but we do not
describe how this is done and deal only with the case where the surface of genus g is obtained by appropriate pasting of 1–cells on the
boundary of a single 2–cell. (The method is described in the proof of
[18, Lemma 8.5].)
We use, in several places, the existence of appropriate dense subsets. We give either proofs or exact references to proofs, when we state
or use such density results. It is widespread practice in the literature
on this subject to state or use without clear justification results concerning certain subsets of the moduli space of flat connections which
are assumed to be dense and, implicitly, of full measure. But much
of the technical difficulty in proving the volume formula lies in taking
proper account of such subsets (which need also to be of full measure)
and so we have strived to be careful about this issue. (I am thankful
to the anonymous referee for stressing the necessity of having sets of
full measure.)
Statement of Results
We work with a compact, connected, semisimple Lie group G,
whose Lie algebra LG is equipped with an Ad–invariant inner-product.
The heat kernel on G is a function Qt (x), for t > 0 and x ∈ G, satisfyt (x)
ing the heat equation ∂Q∂t
= 21 ∆G QRt (x), where ∆G is the Laplacian
on G, and the initial condition limt↓0 G f (x)Qt (x) dx = f (e) for every
continuous function f on G, where e is
the identity in G and dx is
R
Haar measure on G of unit total mass G dx = 1.
For any integer g ≥ 1, let Kg : G2g → G be the product commutator map given by
−1 −1
−1
Kg : G2g → G : (a1 , b1 , ..., ag , bg ) 7→ b−1
g ag bg ag ...b1 a1 b1 a1
(2)
The subset Kg−1 (e), where e is the identity in G, of G2g will be of
special interest to us. The group G acts by conjugation on G2g . If
A ⊂ G2g is preserved by this action, denote by A0 the set of all points
on A where the isotropy is Z(G), the center of G. The quotient
Mg = Kg−1 (e)/G
3
(3)
is identifiable in a standard way with the moduli space of flat G–
connections over a closed, connected, oriented two–dimensional manifold of genus g, but we shall not need any detail of this. The subset
M0g = Kg−1 (e)0 /G
(4)
(when non-empty) has a manifold structure and on M0g there is a
natural symplectic form Ω. Let volΩ be the volume form corresponding
to this symplectic structure; i.e. volΩ =
Our main result is:
1 d/2
,
d! Ω
where d = dim M0g .
Theorem 1 Suppose g ≥ 2. Let f be a continuous G–conjugationinvariant function on G2g , and f˜ the function induced on M0g =
Kg−1 (e)0 /G. Then
vol(G)2−2g
f (x)Qt (Kg (x)) dx =
lim
t↓0 G2g
|Z(G)|
Z
Z
M0g
f˜ dvolΩ
(5)
where the integration on the left is with respect to unit mass Haar
measure, the integration on the right is with respect to the symplectic
volume measure, |Z(G)| is the number of elements in the center Z(G)
of G, and vol(G) is the volume of G with respect to the Riemannian
structure on G given by the Ad–invariant metric on LG.
The integral on the left in (5) arises from integration with respect
to the Yang–Mills measure in the euclidean quantum field theory of
the Yang–Mills field on a compact oriented surface of genus g. We
shall not need this; for more details see [17] or the review [23].
Setting f = 1 leads, after some computation (detailed in (42) below) to Witten’s formula ([24, formula (4.72)]) for the symplectic volume of the moduli space M0g :
volΩ (M0g ) = |Z(G)| vol(G)2g−2
X
α
1
(dim α)2g−2
(6)
where α runs over all irreducible representations of G.
In essence, equation (5) for f = 1 is one of the approaches used by
Witten in [24] to determine the symplectic volume of moduli space.
For general f , Theorem 1 was proven by Forman in [8] using Witten’s volume formula (in fact, this is also what we shall do, but we shall
also prove the volume formula (6)). For G = SU (2) and G = SO(3),
the result was proven in [21].
4
What we shall prove in this paper is actually the limit formula :
f (x)
dvol(x),
t↓0
|dKg (x)∗ |
(7)
for any continuous function f on G2g , where the linear map dKg (x)∗ :
LG → (LG)2g is the adjoint of the derivative (LG)2g → LG : H 7→
Kg (x)−1 Kg′ (x)(xH), and dvol is Riemannian volume measure on the
submanifold Kg−1 (e)0 ⊂ G2g . The known result (34) then implies (5).
The main difficulty in proving (7) lies in taking proper care of the
critical points of Kg and it is to this technical issue that most of the
work in this paper is devoted.
Now we give a quick definition of the symplectic structure Ω.
It will be useful to think of G2r as a subset of G4r via the map
lim
Z
G2g
1−2g
f (x)Qt (Kg (x)) dx = vol(G)
Z
Kg−1 (e)0
Φ : G2r → G4r
−1
−1 −1
: (a1 , b1 , ..., ar , br ) 7→ (a1 , b1 , a−1
1 , b1 , ..., ar , br , ar , br )
For any 1 ≤ i ≤ 4r, and x ∈ G4r , we write
fi = Ad(xi−1 ...x1 ) : LG → LG
with f1 being the identity map. Next let Ω̃ be the 2–form on G4r
specified by
Ω̃x (xH, xH ′ ) =
1 X
−1
−1
ǫij hfi−1
Hi , fj−1
Hj′ iLG
2 1≤i,j≤4r
′ ) ∈ (LG)4r , and ǫ equals 1
where H = (H1 , ..., H4r ), H ′ = (H1′ , .., H4r
ij
for i < j, equals −1 if i > j, and is 0 if i = j. Finally, define
Ω = Φ∗ Ω̃ a 2–form on G2r
(8)
The quotient space M0g = Kg−1 (e)0 /G, if non-empty, has a unique
smooth manifold structure for which the quotient map q : Kg−1 (e)0 →
Kg−1 (e)0 /G is a submersion. The restriction of Ω to Kg−1 (e)0 drops
down to a 2–form Ω on M0g = Kg−1 (e)0 /G:
q ∗ Ω = Ω|Kg−1 (e)0
(9)
It was shown in [11, 12] (with more details in [19]) that Ω is a symplectic form on M0g , and, as proved in [19], is induced Marsden-Weinsteinstyle from the Atiyah-Bott symplectic structure [1] on the space of all
connections.
5
Other Remarks
We take this opportunity to correct in this paper Corollary 3.2 and
Lemma 4.4(ii) of [22]. The correct forms involve sets of full measure
and we have stated the correct result here as Proposition 7. It is this
form, using sets of full measure, which is useful both for the results of
[22] and for our results here. I am very grateful to an anonymous referee for pointing out this error which was present in an earlier version
of this paper.
It should be noted that what we compute is the volume of M0g
and not of the full moduli space Mg . The latter is not, in general, a
smooth manifold but is believed to be the union of symplectic manifolds, called symplectic strata, of different dimensions, these manifolds
corresponding to the different isotropy groups for the action of G on
Kg−1 (e). Volumes of all the strata have been calculated for G = SU (2)
and G = SO(3) in [21].
2
Summary of technical tools
In this section we collect together some results, proven elsewhere,
which we will need.
2.1
A determinant identity
Let V and W be finite dimensional real inner-product spaces, and
A : V → W a linear map. If A : V → W (6= 0) is a linear isomorphism
onto its image A(V ), then by the determinant of A we shall mean
det A =
the determinant of a matrix of A relative
(10)
to orthonormal bases in V and A(V ).
If ker(A) 6= {0}, or if V = {0}, then we define det(A) = 0.
Thus det A is determined up to a sign ambiguity, and | det A| is
independent of the choice of bases.
Let A : V → W and B : W → Z be linear maps between finite
dimensional inner-product spaces. If A is an isomorphism onto W or
if B is an isometry (in which case | det B| = 1 unless W = {0}) then
| det(BA)| = | det(B)|| det(A)|
Consideration of matrices shows that
det A|(ker A)⊥ = det (A∗ |Im A)
6
(11)
The following is a slightly sharpened form of Proposition 2.1 of
[22]. (It is this sharper statement which was used in [22].)
Proposition 1 Let X, Y (6= {0}) be finite dimensional vector spaces
equipped with inner-products, and let V be a subspace of X, and Z a
subspace of Y . Let L1 : X → Z and L2 : X → Y be linear maps such
that
L1 |V ⊥ = 0 and L2 |V = 0
(12)
Let
L = L1 + L2 ,
(13)
and N = ker(L). Then :
(i) there exists a
I : N ⊕ N⊥ → V ⊕ V ⊥
unitary isomorphism
and a
linear isomorphism
J : Z ⊕ Y → Z ⊕ Y,
with | det J| = 1,
such that
J (L1 |V ) ⊕ (L2 |V ⊥ ) I = (L1 |N ) ⊕ (L|N ⊥ )
(14)
(ii) The maps L1 |V : V → Z and L2 |V ⊥ : V ⊥ → Y are both surjective if and only if L1 |N : N → Z and L|N ⊥ : N ⊥ → Y are both
surjective.
(iii) The following equality of determinants holds :
| det L∗1 || det L∗2 | = | det(L1 |N )∗ || det L∗ |
(15)
Here L∗1 : Z → X, L∗2 : Y → X, (L1 |N )∗ : Z → N and L∗ : Y →
X.
Since the statement is slightly sharper than the one in [22] (where
this sharper form is used) we include the full proof, though it is almost
identical to that given in [22].
Proof. (i) Let
I : N ⊕ N ⊥ → V ⊕ V ⊥ : (a, b) 7→ (a + b)V , (a + b)V ⊥ ,
7
wherein the subscripts signify orthogonal projections onto the corresponding subspaces. Since N ⊕ N ⊥ ≃ X ≃ V ⊕ V ⊥ isometrically, by
means of (x, y) 7→ x + y, I corresponds to the identity map on X and
is thus a unitary isomorphism.
Let Ll : Y → N ⊥ ⊂ X, be a linear left-inverse for the injective
map L|N ⊥ ; thus Ll L(b) = b for every b ∈ N ⊥ . Next define
J = J2 J1 : Z ⊕ Y → Z ⊕ Y,
where
J1 : Z ⊕ Y → Z ⊕ Y : (a, b) 7→ J1 (a, b) = (a, a + b)
J2 : Z ⊕ Y → Z ⊕ Y : (a, b) 7→ J2 (a, b) = (a − L1 Ll b, b).
It is clear that both J1 and J2 are injective. Moreover, they are also
surjective, because for any (z, y) ∈ Z ⊕ Y , J1 (z, y − z) = (z, y) and
J2 (z +L1 Ll y, y) = (z, y); note that z +L1 Ll y ∈ Z because L1 (X) ⊂ Z.
So J1 and J2 are isomorphisms and hence so is J.
By considering matrix representations for J1 and J2 , we have
| det J1 | = | det J2 | = 1, and so
| det J| = | det J2 | | det J1 | = 1
(16)
For any (a, b) ∈ N ⊕ N ⊥ , we have :
J (L1 |V ) ⊕ (L2 |V ⊥ ) I(a, b) = J L1 (a + b)V , L2 (a + b)V ⊥
= J L1 (a + b), L2 (a + b)
= J2 L1 (a + b), L(a + b)
= J2 L1 (a + b), L(b)
=
=
L1 (a + b) − L1 Ll L(b), L(b)
L1 (a), L(b) .
This proves equation (14), and part (i).
(ii) follows directly from (i).
(iii) Since L1 |V ⊥ = 0 it follows that L∗1 (Z) ⊂ V . Similarly,
∗
L2 (Y ) ⊂ V ⊥ and L∗ (Y ) ⊂ N ⊥ . So, with appropriately restricted
codomains (for instance we are taking L∗1 : Z → V instead of L∗1 :
Z → X) :
8
(L1 |V )∗ = L∗1 ,
(L2 |V ⊥ )∗ = L∗2 ,
and (L|N ⊥ )∗ = L∗
In view of this, we may take adjoints in equation (14) to obtain :
I ∗ (L∗1 ⊕ L∗2 )J ∗ = (L1 |N )∗ ⊕ L∗ , as maps Z ⊕ Y → N ⊕ N ⊥
wherein again some of the operators are taken with restricted codomains.
Taking determinants (which, by our definition, is not affected by restriction of codomains), and using the determinant of products given
in (11), and the fact that | det J| equals 1, we obtain the determinant
formula (15). QED
We will use the preceding proposition in a specific context. Let
G be a compact, connected, semisimple Lie group with Lie algebra
LG equipped with an Ad–invariant metric. Let g1 and g2 be positive
integers, and g = g1 + g2 . We have the product commutator maps
Kgi : G2gi → G and Kg : G2g → G specified through (2). Let xi ∈ G2gi
and x = (x1 , x2 ). Define
C1 : G2g → G : (x1 , x2 ) 7→ Kg1 (x1 ),
C2 : G2g → G : (x1 , x2 ) 7→ Kg2 (x2 )
Then Kg (x) = C2 (x)C1 (x) and we have the derivative maps
Kg (x)−1 dKg (x) : Tx G2g → LG,
Ci (x)−1 dCi (x) : Tx G2g → LG
which are related by
Kg (x)−1 dKg (x) = C1 (x)−1 dC1 (x) + Ad(C1 (x)−1 )C2 (x)−1 dC2 (x)
We will apply Proposition 1 with
X = (LG)2g ≃ (LG)2g1 ⊕ (LG)2g2 ,
V = (LG)2g1 ⊕ 0
and
L1 = C1 (x)−1 dC1 (x),
L2 = Ad(C1 (x)−1 )C2 (x)−1 dC2 (x)
Specializing Proposition 1 to this situation gives us the following
result:
Proposition 2 Let x = (x1 , x2 ) ∈ G2g1 × G2g2 . Then Kgi is submersive at xi , for both i = 1 and i = 2, if and only if Kg is submersive at
x and C1 |Kg−1 (e) : Kg−1 (e) → G is submersive at x. Furthermore,
| det dKg (x)∗ | det [dC1 (x)| ker dKg (x)]∗
= | det dC1 (x)∗ | | det dC2 (x)∗ |
= | det dKg1 (x1 )∗ | | det dKg2 (x2 )∗ |
(17)
9
2.2
A disintegration formula
The following disintegration formula, proved in Proposition 3.1 of [22]
will be useful. (The formula (19) is proved for vastly more general K
by Federer [7]).
Proposition 3 Let K : M → N be a smooth mapping between Riemannian manifolds. Let NK = K (M \ CK ), where CK is the set of
points where K is not submersive, i.e. the rank of dK is less than
dim N . Assume that CK 6= M . Suppose φ is a continuous function of
compact support on M . Let vol denote Riemannian volume measure.
(For example, on the submanifold K −1 (h) \ CK ⊂ M , for h ∈ NK ,
which is given the metric induced from M . If dim K −1 (h) = 0, the
Riemannian volume is understood to be counting measure.)
If φ vanishes in a neighborhood of CK , then
h 7→
Z
K −1 (h)\CK
φ dvol
is continuous on NK
(18)
and
Z
M
φ dvol =
Z
NK
"Z
K −1 (h)\CK
#
φ
dvol
| det(dK)∗ |
dvol(h)
(19)
In our application, every open subset U of M can be expressed
as the union of a sequence of open subsets Un with compact closure,
and there is a sequence of continuous functions 0 ≤ φ1 ≤ φ2 ≤ ... ≤
φn ↑ 1U , where φn is 0 outside Un . Then, for f any continuous non–
negative function on M , using φn f in place of f in (19), and letting
n → ∞, monotone convergence shows that (19) holds for f 1U in place
of φ, if U is any open subset of M \ CK . In particular, (19) holds for
1U in place of φ and hence, if vol(M ) < ∞, also for 1U −V for any open
sets U, V ⊂ M \ CK .
2.3
Some dense sets of full measure
We shall describe some useful subsets which are dense and of full
measure in appropriate sets.
The group G acts by conjugation on G2r :
G × G2r → G2r : (h, x) 7→ hxh−1 = (hx1 h−1 , ..., hx2r h−1 )
where x = (x1 , ..., x2r ).
10
(20)
Semisimplicity of the compact group G (i.e. that the center G is
finite) is important in the following. We equip G with an Ad–invariant
metric.
Proposition 4 Let G be a compact, connected, semisimple Lie group
and T a maximal torus in G, acting on G by conjugation. Then the
set of points in G where the isotropy is Z(G) is an open set of full
measure.
By “full measure” we mean a measurable set whose complement has
measure zero. In particular, an open set of full measure is automatically dense since the measures under consideration assign positive
measure to non-empty open sets.
Proof. Under the adjoint action of the compact abelian group T ,
the Lie algebra LG splits up as a direct sum of LT and two-dimensional
spaces R1 , ..., Rk on each of which T acts by ‘rotations.’
The compact Lie group G, equipped with the Ad-invariant metric
on the Lie algebra LG, is a complete Riemannian manifold. We shall
use a result concerning the exponential map for such manifolds.
For each unit vector u ∈ LG let δ(u) be the infimum of all real
numbers r > 0 such that the distance of exp(ru) from the identity e
is r. Now let B be the subset of LG consisting of 0 and all v 6= 0
such that |v| < δ(v/|v|), and let W = exp(B). Then it is known (see,
for instance, Chavel [5, Theorem 3.2 and Proposition 3.1]) that B is
open, W is an open set of full measure in G, and
B → W : v 7→ exp(v) is a diffeomorphism onto W
(21)
For any t ∈ T , the conjugation map G → G : x 7→ txt−1 is an isometric
isomorphism and so the function δ is invariant under the adjoint action
of T on LG. Therefore, Ad(t)B = B for all t ∈ B.
Let
W 0 = exp(W ′ )
(22)
where W ′ is the subset of B consisting of all points of the form v =
vLT + v1 + · · · + vk , with vLT ∈ LT and each vi ∈ Ri being non-zero:
W ′ = {vLT + v1 + · · · + vk | vLT ∈ LT , each vi ∈ Ri and vi 6= 0} (23)
Suppose t ∈ T commutes with x ∈ W 0 . We know that x = exp(v)
for a unique v ∈ W ′ . Moreover, since exp is injective on B and
Ad(t)v ∈ B, the relation txt−1 = x implies that Ad(t)v = v. Since
11
Ad(t) preserves each subspace Ri , whose direct sum along with LT
is LG, it follows that Ad(t)vi = vi for each i ∈ {1, ..., k}. Since T
acts on the two-dimensional spaces Ri by rotations and fixes the nonzero vector vi , this means that in fact Ad(t) is actually the identity
on each Ri . Therefore, Ad(t) is the identity on all of LG and so
t ∈ Z(G). Thus the T –isotropy at each point of W 0 is Z(G). Now
W ′ is clearly an open subset of full measure in B, and so, since exp is
a diffeomorphism on B, it follows that W 0 is a subset of full measure
in W . Since W is of full measure in G, we conclude that W 0 is of full
measure in G.
By a general result of transformation group theory (Bredon [3,
Theorem 4.3.1 and Corollary 6.2.5], Kawakubo [10, Theorem 4.27],
Bourbaki [2, IX.96,No.4,Thèoréme 2]) for compact Lie groups acting
on connected manifolds, the set of points of minimal isotropy is (dense
and) open in the whole space. QED
We apply this to show that the conjugation action of G on Gr has
minimal isotropy Z(G) on a set of full measure:
Proposition 5 Let G be a compact, connected, semisimple Lie group,
and k any integer ≥ 2. For the conjugation action of G on Gk , the
subset on which the isotropy group is Z(G) is a dense open set of full
measure in G.
Proof. As noted earlier, the set of points of minimal isotropy (for a
compact Lie group acting on a connected manifold) is open, being a
consequence of a general result on transformation groups (Bourbaki
[2, IX.96,No.4,Thèoréme 2]). So we focus on the measure theoretic
issue.
Since Gk = G2 × Gk−2 , it will suffice to prove the result for k = 2.
Let U be the subset of G2 consisting of all points where the isotropy
group of the conjugation action of G is Z(G). The subset G0 of G
which consists of points which generate maximal tori is of full measure in G (see, for example, Bröcker and tom Dieck [4, Theorem
IV.2.11(ii)]). If x ∈ G0 then the preceding lemma implies that for
almost every y ∈ G the isotropy group at (x, y) is Z(G) (any element
which commutes with x lies in the maximal torus generated by x;
see, for example, [4, Theorem IV.2.3(i)]). So, by Fubini’s theorem,
(G0 × G) ∩ U is of full measure in G2 . So U is of full measure in
G2 . QED
The preceding result has the following consequence:
12
Proposition 6 For any integer r ≥ 1 and compact, connected semisimple group G, the critical points of the mapping Kr : G2r → G form a
set of measure 0 in G2r .
Proof. There is a remarkable relationship, stated below in (32), between the the derivative dKr and the istropy of the conjugation action
of G on G2r . The relation (32) implies that at any critical point x of
Kr the isotropy group {g ∈ G : gxg −1 = x} has a non-trivial Lie algebra, and so, in particular, the isotropy group is not equal to Z(G).
The preceding proposition then implies that the set of critical points
of Kr is contained in a set of measure 0. QED
Next we show that almost every point on almost every level set
Kr−1 (h) is a point of isotropy Z(G). For this we use the important
fact that the product commutator map Kr : G2r → G is surjective.
This is proven in [20, Proposition 4.2.4] and uses semisimplicity of G
(as I learnt later, this result also appears in Bourbaki [2, Lie IX.33
Corollaire to Proposition 10]).
Proposition 7 For any integer r ≥ 1, let Ur0 be the subset of G2r
where the isotropy of the conjugation action of G is Z(G). Then for
almost every h ∈ G the set Kr−1 (h) ∩ Ur0 is of full measure in Kr−1 (h).
Proof. Let Ur be the set of all non-critical points of Kr . Then
Ur0 ⊂ Ur
(24)
because of the striking relation (32) between the behavior of dKr and
the isotropy of the conjugation action. The mapping Kr |Ur : Ur →
G is an open mapping. We have the co-area/disintegration formula
giving the volume of any open set A ⊂ Ur :
vol(A) =
Z
Kr (Ur )
"Z
Kr−1 (h)∩A
#
dvol
dvol
| det(dKr )∗ |
(25)
where vol always denotes Riemannian volume arising, in our situation,
from any choice of Ad-invariant metric on G. Since vol(G2r ) < ∞,
the formula (25) holds when A is the difference of open sets. Taking
A to be the set Ur − Ur0 of measure 0, it follows that for almost every
h ∈ Kr (Ur ) the set Kr−1 (h)∩Ur0 is of full measure in Kr−1 (h)∩Ur . Now
Kr (Ur ) contains all regular values of Kr : here we use the surjectivity
of Kr which assures that every regular value of Kr is in fact a value of
Kr . Moreover, by Sard’s theorem, the set of all regular values of Kr is
13
a set of full measure in G, and, furthermore, note that Kr−1 (h) ⊂ Ur for
any regular value h of Kr . Thus almost every point h ∈ G satisfies the
condition that Kr−1 (h)∩Ur0 is of full measure in Kr−1 (h)∩Ur = Kr−1 (h).
QED
Using the notation from the preceding result we also have:
Proposition 8 For g1 , g2 ≥ 1 and g = g1 + g2 , let
Kg−1 (e)0,0 = (Ug01 × Ug02 ) ∩ Kg−1 (e)
(26)
and let Ci : G2g1 × G2g2 → G : (x1 , x2 ) 7→ Kgi (xi ), for i ∈ {1, 2}.
Then Kg−1 (e)0,0 is not empty and the set
def
U12 = C1 (Kg−1 (e)0,0 ) = C2 (Kg−1 (e)0,0 ) = Kg1 (Ug01 ) ∩ Kg2 (Ug02 ) (27)
is a dense open subset of full measure in G.
Proof. Let Di be the set of all regular values of Kgi . If h ∈ Di is in the
complement of Kgi (Ug0i ) then Kg−1
(h) ∩ Ug0i = ∅, while, by surjectivity
i
−1
of Kgi , the level set Kgi (h) is a non-empty closed submanifold of
G2gi and so has positive volume. So by the preceding result, the set
of all such elements h has measure 0. Thus Kgi (Ug0i ) ∩ Di is of full
measure in Di . By Sard’s theorem, Di is a set of full measure in G,
and so Kgi (Ug0i ) has full measure in G. Since Kgi is submersive on Ug0i
it follows that the image Kgi (Ug0i ) is an open subset of G. So the sets
Kgi (Ug0i ), for ∈ {1, 2}, are open sets of full measure on G and hence
so is their intersection
U = Kg1 (Ug01 ) ∩ Kg2 (Ug02 )
The relation
Kr (br , ar , ..., b1 , a1 ) = Kr (a1 , b1 , ..., ar , br )−1
(28)
shows that
Kr (Ur0 ) = Kr (Ur0 )−1
and so
U = U −1
Let h ∈ U . Then there is, for i = 1, 2, an xi ∈ Ug0i with Kg1 (x1 ) = h
and Kg2 (x2 ) = h−1 . Then x = (x1 , x2 ) is a point in Kg−1 (e)0,0 whose
14
image under C1 is h and whose image under C2 is h−1 . This, together
with the inversion property (28) implies
Ci (Kg−1 (e)0,0 ) ⊃ U
for i = 1, 2.
Conversely, suppose h ∈ C1 (Kg−1 (e)0,0 ). This means that there is
a point (x1 , x2 ) ∈ Kg−1 (e)0,0 with C1 (x1 , x2 ) = h. Since Kg (x1 , x2 ) =
C2 (x2 )C1 (x1 ), it follows that C2 (x2 ) = h−1 . The condition (x1 , x2 ) ∈
Kg−1 (e)0,0 says also that xi ∈ Ug0i , for i = 1, 2, and so h ∈ Kg1 (Ug01 ) and
h−1 ∈ Kg2 (Ug02 ). The inversion property (28) then implies that h ∈ U .
The argument works if we start with h ∈ C2 (Kg−2 (e)0,0 ). QED
2.4
Facts about Ω and Ω
The compact, semisimple group G acts by conjugation on G2g . Let
Ug0 be the set of all points where the isotropy is Z(G). Clearly, this is
carried into itself by the conjugation action. Moreover, Ug0 is a dense
open subset of full measure in G2g , as we have shown.
Let Kg−1 (e)0 = Ug0 ∩ Kg−1 (e), the set of points on Kg−1 (e) where
the isotropy group of the conjugation action of G is Z(G), and assume
that it is non-empty (Proposition 8 implies that this is so when g ≥ 2).
Let Kg−1 (e)0 be the set of points x in Kg−1 (e) where Kg is submersive, i.e. dKg (x) : Tx G2g → TKg (x) G is surjective. It is a consequence
of Theorem 2(v) below that Kg−1 (e)0 is a subset of Kg−1 (e)0 .
Then Kg−1 (e)0 , being a level set of a smooth submersion Kg |Ug0 :
0
Ug → G, is a smooth submanifold of G2g .
The quotient
M0g = Kg−1 (e)0 /G
being a quotient of a smooth manifold by a compact Lie group, having
the same isotropy subgroup Z(G) everywhere, is a smooth manifold
(sections 16.14.1 and 16.10.3 in [6]).
The conjugation action of the group G on G2r , gives for any x =
(x1 , ..., x2r ) ∈ G2r the orbit map
γx : G → G2r : h 7→ hxh−1 = (hx1 h−1 , ..., hx2r h−1 )
(29)
The derivative at x of the product commutator map Kg : G2g → G
is, technically, a map Tx G2g → TKg (x) G, but by means of appropriate
left translations to the identity we shall sometimes view it as a map
15
(LG)2g → LG and sometimes as (LG)2g → TKg (x) G. Its adjoint,
relative to the given Ad–invariant metric on LG, is then a linear map
dKg (x)∗ : LG → (LG)2g
(30)
Recall from (8) the 2–form Ω on G2g .
We summarize some facts about Ω, γ, and Kg :
Theorem 2 Let g ≥ 1 and assume that Kg−1 (e)0 is not empty. Then:
(i) there is a unique smooth manifold structure on M0g = Kg−1 (e)0 /G
such that the quotient map q : Kg−1 (e)0 → Kg−1 (e)0 /G is a submersion
(ii) there is a unique smooth 2–form Ω on Kg−1 (e)0 /G such that
q ∗ (Ω) = Ω|Kg−1 (e)0
(iii) the 2–form Ω is closed and non–degenerate, i.e. it is symplectic
on M0g (Proposition IV.E in [11] and [12, Proposition 3.3])
(iv) Ω satisfies the “moment map” formula
Ωx (xY, γx′ H) = hY, dKg (x)∗ Hi(LG)2g
(31)
for all x ∈ Kg−1 (e), H ∈ LG and Y ∈ (LG)2g ([11, Proposition
IV.G])
(v) for any x = (x1 , ..., x2g ) ∈ G2g , the kernel of γx′ : LG → (LG)2g
equals the kernel of dKg (x)∗ : LG → (LG)2g :
ker γx′ = ker dKg (x)∗ = {H ∈ LG : Ad(x1 )H = · · · = Ad(x2g )H = H}
(32)
([11, Proposition IV.C] and also in [9])
(vi) If x ∈ Kg−1 (e)0 then
|Pfaff(Ωq(x) )| =
| det γx′ |
| det dKg (x)∗ |
(33)
where the Pfaffian is, as usual, the square root of the determinant of the matrix of Ωq(x) relative to an orthonormal basis ([12,
Proposition 3.3])
(vii) If f is a measurable function on Kg−1 (e)0 , invariant under the
conjugation action of G, and f˜ the induced function on M0g =
Kg−1 (e)0 /G then
Z
M0g
f˜ dvolΩ =
1
vol(G/Z(G))
16
Z
Kg−1 (e)0
f
dvol,
| det dKg∗ |
(34)
whenever either side is defined, where volΩ is symplectic volume
for the symplectic structure Ω, while vol by itself always denotes
Riemannian volume. (Essentially [12, Proposition 3.5] or by part
(vi) and [22, Lemma 3.4].)
2.5
An application
We shall “prefabricate” a result that will go into the proof of Theorem
1.
Let g1 , g2 be positive integers and g = g1 + g2 . Let Kg−1 (e)0,0 the
subset of Kg−1 (e) consisting of all points (x1 , x2 ) ∈ G2g1 × G2g2 such
that the isotropy of the G–conjugation action on Ggi is Z(G) at xi ,
for i = 1, 2. We have the maps Ci : G2g → G specified by
C1 (x1 , x2 ) = Kg1 (x1 ),
C2 (x1 , x2 ) = Kg2 (x2 )
Recall from Proposition 8 (equation (27)) that
def
U12 = C1 (Kg−1 (e)0,0 ) = C2 (Kg−1 (e)0,0 ) = Kg1 (Ug01 ) ∩ Kg2 (Ug02 )
is an open subset of full measure in G.
Let Di be the set of all regular values of Kgi . By Sard’s theorem,
Di is a subset of full measure in G. The maps Kgi being surjective,
Di is contained in the image of Kgi . (The set Di is also open in G.)
The inversion relation
Kr (br , ar , ..., b1 , a1 ) = Kr (a1 , b1 , ..., ar , br )−1
(35)
implies that Di = Di−1 . Therefore,
def
D = D1 ∩ D2−1
(36)
is also a subset of full measure in G.
Proposition 9 The following disintegration formula holds:
Z
Kg−1 (e)0,0
dvol
=
| det dKg∗ |
vol(G)
Z "Z
D
Kg−1
(h)
1
(37)
dvol(x1 )
| det dKg1 (x1 )∗ |
# "Z
Kg−1
(h−1 )
2
#
dvol(x2 )
dh
| det dKg2 (x2 )∗ |
where dh is the unit-mass Haar measure on G and vol(G) is the volume
of G with respect to the given Ad-invariant metric on the Lie algebra
of G.
17
Proof. Let Ur0 be the subset of G2r consisting of all points where
the isotropy of the conjugation action of G is Z(G). Then Ur0 is a
non-empty (in fact, dense) open subset of G2r (this is a special case
of a general theorem on group actions: Bredon [3, Theorem 4.3.1 and
Corollary 6.2.5], Kawakubo [10, Theorem 4.27], Bourbaki [2, IX.96,
No.4, Thèoréme 2]). By Theorem 2(v), the map Kg : G2g → G is
a submersion at every point in Ug01 × Ug02 , and so, Kg−1 (e)0,0 , being a
level set of a submersion, is a smooth submanifold of G2g .
From Proposition 2 it follows that C1 |Kg−1 (e)0,0 is submersive at
every point. Therefore, by the disintegration formula in Proposition
3, we have
Z
Kg−1 (e)0,0
dvol
=
| det dKg∗ |
vol(G)
Z
U12
(38)
"Z
C1−1 (h)∩Kg−1 (e)0,0
#
dvol
dh
| det dKg∗ | | det(dC1 | ker dKg )∗ |
Next we use the determinant identity from Proposition 2 to obtain:
Z
Kg−1 (e)0,0
dvol
=
| det dKg∗ |
vol(G)
Z
U12
"Z
(39)
C1−1 (h)∩Kg−1 (e)0,0
#
dvol(x1 , x2 )
dh
| det dKg1 (x1 )∗ | | det dKg2 (x2 )∗ |
Now the identity map
(h−1 )0 : (x1 , x2 ) 7→ (x1 , x2 )
(h)0 × Kg−1
C1−1 (h) ∩ Kg−1 (e)0,0 → Kg−1
2
1
is an isometry (the metric on the left is inherited from that on G2g ).
So we have
Z
Kg−1 (e)0,0
dvol
=
| det dKg∗ |
vol(G)
Z
U12
"Z
Kg−1
(h)0
1
(40)
dvol(x1 )
| det dKg1 (x1 )∗ |
# "Z
Kg−1
(h−1 )0
2
Since both U12 and D are subsets of full
measure in G, the integration
R
·
·
·
dh
above
can
be
replaced
by
·
U12
D · · dh. Finally, by Proposition
0 is of full measure in K −1 (c) for almost every c, and
7, the set Kg−1
(c)
gi
i
so we obtain the desired formula (37). QED
R
18
#
dvol(x2 )
dh
| det dKg2 (x2 )∗ |
2.6
A heat-kernel integral and its limit
If X1 , ...., Xd is an orthonormal basis of the Lie algebra of G, and α
P
an irreducible representation of G then di=1 α∗ (Xi )2 is of the form
−Cα I, where Cα is a scalar (‘Casimir’ ) and I is the identity operator
on the representation space of α. The heat kernel Qt has a standard
character expansion:
Qt (x) =
X
(dim α)e−Cα t/2 χα (x),
α
where χα is the character of the representation α.
The following is a very useful formula:
Z
−1 −1
−1
Qt (hb−1
g ag bg ag . . . b1 a1 b1 a1 ) da1 . . . dbg =
X e−Cα t/2 χα (h)
,
(dim α)2g−1
(41)
where the sum is over all inequivalent irreducible representations α
of G. This can be verified using : (i) the identity (see Ex 4.17.3 in
Bröcker and tom Dieck [4])
G2g
α
Z
G
χα (aba−1 c) da = (dim α)−1 χα (b)χα (c),
(ii) repeated application of standard convolution properties of characters, and (iii) commuting integral and a series sum. Integral and sum
can be commuted because
X
e−Cα t/2 (dim α)
α
Z
|χα (· · ·)| d · · · ≤
X
e−Cα t/2 (dim α)2 = Qt (e) < ∞
α
Formula (41) is due to Witten ([24, equation (2.51)]) who determined it in his exact evaluation of the partition function of two–
dimensional quantum Yang–Mills theory (the heat-kernel was not used
explicitly in [24]).
P
It is known (Knapp[13, Lemma 10.3]) that α (dim1 α)k < ∞ for
k ≥ 2. So, for g ≥ 2, using dominated convergence in (41) gives
lim
t↓0
Z
G2g
−1 −1
−1
Qt (hb−1
g ag bg ag . . . b1 a1 b1 a1 ) da1 . . . dbg =
X
α
χα (h)
,
(dim α)2g−1
(42)
Proposition 10 The limit formula (42) continues to hold, with the
P
limit limt↓0 and the sum α being both in the L2 (G, dh)–sense.
19
Proof. Let k = 2g − 1, and dα = dim α. Then
X e−Cα t
α
dkα
χα −
X 1
α
dkα
χα
2
L2 (G)
X (e−tCα − 1)2
=
d2k
α
α
which,for t > 0, is bounded, term by term, by the convergent series
1
α d2k . QED
P
α
3
Evaluation of Limits
With notation and assumptions as before, let
Kg−1 (h)0
def
=
the set of all non-critical points of
Kg :
G2g
→ G which lie on
(43)
Kg−1 (h)
for any h ∈ G.
A point x ∈ G2g is a non-critical point of Kg if and only if the
isotropy group at x of the conjugation action of G on G2g is discrete, an
observation immediate from Theorem 2 (v). Therefore, in particular,
Kg−1 (e)0 ⊂ Kg−1 (e)0
(44)
If g ≥ 2 then, by Proposition 8 (also Proposition III B of [11]), Kg−1 (e)0
is not empty and hence also Kg−1 (e)0 6= ∅.
As a consequence of the disintegration formula, we have the following result (mentioned in [11, section IV]):
Lemma 1 Suppose g is an integer ≥ 2. Let f be a continuous function on G2g which is 0 in a neighborhood of the critical points of Kg .
Then
lim
t↓0
Z
G2g
f (x)Qt (Kg (x)) dx = vol(G)1−2g
Z
Kg−1 (e)0
f
dvol
| det dKg∗ |
(45)
Proof Let C be the set of all critical points of Kg . Then the complement G2g \ C is open and the image Kg (G2g \ C) is an open subset
of G of full measure (by Sard’s theorem, since it contains all regular
values of the surjective map Kg ) and hence is also dense in G. By
Proposition 3 we have the disintegration
Z
G2g
f (x)Qt (Kg (x)) dx = vol(G)−2g
Z
20
Kg (G2g \C)
F (h)Qt (h) dvol(h),(46)
where
def
F (h) =
Z
Kg−1 (h)0
f
dvol
| det dKg∗ |
(47)
is a continuous function of h ∈ Kg (G2g \ C).
The identity e belongs to Kg (G2g \ C) since Kg−1 (e)0 6= ∅. Moreover, F (h) is 0 when h is outside the compact set Kg (support(f )) ⊂
Kg (G2g \ C). So F extends to a continuous function on G, 0 outside
Kg (G2g \ C). So, remembering that the Riemannian volume on G is
vol(G) times the normalized Haar mass dh,
Z
G2g
f (x)Qt (Kg (x)) dx = vol(G)1−2g
Z
G
F (h)Qt (h) dh
(48)
and, by the initial–condition property of the heat kernel Qt , this approaches the limit
1−2g
vol(G)
1−2g
F (e) = vol(G)
f
dvol
| det dKg∗ |
Z
Kg−1 (e)0
as t ↓ 0. QED
Things are much easier when we deal with a regular value of Kg :
Lemma 2 Let r be any integer ≥ 1, f a continuous function on G2r ,
and c a regular value of Kr : G2r → G. Then
lim
t↓0
Z
G2r
f (x)Qt (Kr (x)c−1 ) dx = vol(G)1−2r
Z
Kr−1 (c)
f
dvol
| det dKr∗ |
(49)
Proof. The argument is essentially the same as in the preceding
lemma, but we no longer have to worry about critical points of Kr
since there are none on Kr−1 (c).
Let U and V be neighborhoods of c, with V ⊂ U , and U consisting
only of regular values of Kr . Let φ be a continuous function on G,
with 0 ≤ φ ≤ 1 everywhere, equal to 1 on V and equal to 0 outside
U . Let ψ = 1 − φ. Then f = (φ ◦ Kr )f + (ψ ◦ Kr )f , and
Z
G2r
f (x)ψ(Kr (x))Qt Kr (x)c−1 dx
≤
|f |sup sup Qt (yc−1 )
y∈G\V
→ 0,
as t ↓ 0
by a uniform-limit property of the heat kernel Qt as t ↓ 0.
21
On the other hand, the integrand in
Z
G2r
f (x)φ(Kr (x))Qt (Kr (x)c−1 ) dx
is 0 near the critical points of Kr . Note also that φ(Kr (x)) = 1
when x ∈ Kr−1 (c), and Kr−1 (c) contains no critical point of Kr . So,
by Proposition 3 and the argument used in Lemma 1, as t ↓ 0, this
integral approaches the limit
vol(G)1−2r
Z
Kr−1 (c)
f
dvol
| det dKr∗ |
Combining all these observations, we obtain the desired result. QED
The preceding result is essentially present in Forman [8].
4
Proof of the main result
Let g be a positive integer. Recall that Kg−1 (e) ⊂ G2g . The set of
points on Kg−1 (e) where dKg (x) : Tx G2g → TKg (x) G is surjective is
denoted Kg−1 (e)0 . The set of points on Kg−1 (e) where the isotropy
group of the G–conjugation action is Z(G) is denoted Kg−1 (e)0 .
Now suppose g1 and g2 are positive integers with g = g1 + g2 .
We denote by Kg−1 (e)0,0 the subset of Kg−1 (e) consisting of all points
(x1 , x2 ) ∈ G2g1 × G2g2 such that the isotropy of the G–conjugation
action on Ggi is Z(G) at xi , for i = 1, 2. Thus
(c)0
(c−1 )0 × Kg−1
Kg−1 (e)0,0 = ∪c∈G Kg−1
2
1
(50)
The subset Ug0i of G2gi where the isotropy group is Z(G) is (dense and)
open in G2gi , as proven in Proposition 5. So
Kg−1 (e)0,0 = (Ug01 × Ug02 ) ∩ Kg−1 (e) = (Ug01 × Ug02 ) ∩ Kg−1 (e)0
is an open subset of Kg−1 (e)0 .
Theorem 3 For any integer g ≥ 2, and integers g1 , g2 ≥ 1 with
g = g1 + g2 ,
Z
Kg−1 (e)0
dvol
| det dKg∗ |
dvol
| det dKg∗ |
Z
dvol
=
−1
0,0
| det dKg∗ |
Kg (e)
=
Z
Kg−1 (e)0
= vol(G)2g−2 lim
t↓0
22
Z
G2g
(51)
(52)
Qt (Kg (x)) dx (53)
Proof. If f is a continuous function on G2g , with 0 ≤ f ≤ 1, which is
0 in a neighborhood of the critical points of Kg then
vol(G)1−2g
Z
Kg−1 (e)0
f dvol
| det dKg∗ |
Z
= lim
t↓0
G2g
Z
≤ lim
t↓0
G2g
f (x)Qt (Kg (x)) dx(54)
Qt (Kg (x)) dx
The right side was noted in (42) to be finite. Taking appropriate f ,
with f = 1 at distances beyond 1/n from the critical points of Kg ,
and then letting n → ∞ we have, by dominated convergence,
1−2g
vol(G)
Z
Kg−1 (e)0
dvol
≤ lim
t↓0
| det dKg∗ |
Z
G2g
Qt (Kg (x)) dx
(55)
Next, observing that
Kg (x1 , x2 ) = Kg2 (x2 )Kg1 (x1 )
for x1 ∈ Gg1 and x2 ∈ Gg2 , and using the convolution property of the
heat kernel
Z
G
Qt (ac)Qs (c−1 b) dc = Qt+s (ab) = Qt+s (ba)
we have
Z Z
−1
G2g1
G
Qt (Kg1 (x1 )c
) dx1
Z
G2g2
Qt (cKg2 (x2 )) dx2 dc
=
Z
G2g
(56)
Q2t (Kg (x)) dx
Then
lim
Z
t→0 G2g
lim
Qt (Kg (x)) dx =
Z Z
t→0 G
=
Z
G
G2g1
lim
Z
Qt (Kg1 (x1 )c−1 ) dx1
t→0 G2g1
···
lim
Z
t→0 G2g2
Z
g2
G
···
Qt (cKg2 (x2 )) dx2 dc
dc
(57)
because of the L2 (G, dc)–convergence of the limits limt→0 noted in
Proposition 10.
Let Di be the set of all regular values of Kgi : G2gi → G, and
def
D = D1 ∩ D2 ,
23
(58)
which, as we have already noted in the context of (36), is a dense open
subset of full measure in G.
R
R
Since D is of full measure in G, we can replace G · · · dc by D · · · dc
on the right side in (57). Then using the limit value computed in
Lemma 2 we have
lim
Z
t→0 G2g
Qt (Kg (x)) dx =
2−2g
vol(G)
Z "Z
D
Kg−1
(c)
1
(59)
dvol
| det dKg∗1 |
Z
Kg−1
(c−1 )
2
#
dvol
dc
| det dKg∗2 |
Now inserting
our “prefabricated” piece Proposition 9, we see that the
R
integral D [· · ·] dc on the right side in (59) is equal to
[vol(G)]−1
Z
Kg−1 (e)0,0
dvol
| det dKg∗ |
Combining this with (55), we write
1−2g
vol(G)
Z
Kg−1 (e)0
dvol
| det dKg∗ |
≤ lim
Z
t→0 G2g
=
Z
Qt (Kg (x)) dx
Kg−1 (e)0,0
(60)
dvol
vol(G)1−2g
| det dKg∗ |
Since Kg−1 (e)0,0 ⊂ Kg−1 (e)0 , it follows that the inequalities in (60) are
equalities. QED
Since the middle integral in (60) is finite so are the others. As
consequence, we have
Corollary 1 For any integer g ≥ 2, the sets Kg−1 (e)0,0 and Kg−1 (e)0
open, dense subsets of full measure in Kg−1 (e)0 .
Now we are ready for
Proposition 11 For any integer g ≥ 2 and any continuous function
f on G2g ,
lim
t↓0
Z
G2g
f (x)Qt (Kg (x)) dx = vol(G)1−2g
24
Z
Kg−1 (e)0
f
dvol
|dKg∗ |
(61)
Proof. We have proved this (in Lemma 1) when f is zero near the
critical points of Kg . We have also proved this for f = 1 in Theorem
3. Now by Proposition 6, the set Ug of non-critical points of Kg is of
full measure in G2g , and so
Z
G2g
f (x)Qt (Kg (x)) dx =
Z
Ug
f (x)Qt (Kg (x)) dx
Since Kg−1 (e)0 is a subset of Ug , the task reduces to proving a limiting
result for integrals over Ug , given that the limiting formula holds for
continuous functions of compact support as well as for the constant
function 1. The proof is finished by using Lemma 3 below (take X to
be Ug , which is an open subset of G2g ). QED
Lemma 3 Let µt , for t ≥ 0, be finite Borel measures on a locally
compact Hausdorff space X such that limt↓0 µt (X) = µ0 (X) and
lim
t↓0
Z
X
f dµt =
Z
X
f dµ0
for every continuous function f of compact support in X. Assume
that X is the union of a countable collection of compact sets. Then
lim
t↓0
Z
X
f dµt =
Z
X
f dµ0
for every bounded continuous function f on X.
Proof. Let ǫ > 0.
Since X is the union of a countable number of compact sets, and
µ0 (X) < ∞, there is a compact set K ⊂ X for which
µ0 (K c ) < ǫ
By local compactness there is an open set U ⊃ K with compact closure
U , and, by Urysohn’s lemma, there is a continuous function Φ with
1K ≤ Φ ≤ 1U
First we demonstrate that lim supt↓0 µt (U ) is < ǫ. For s > 0 we have
c
µs (U ) = µs (X) − µs (U ) ≤ µs (X) −
25
Z
X
Φ dµs
and so, for any t > 0,
Z
c
sup µs (U ) ≤ sup µs (X) − inf
0<s≤t
0<s≤t X
0<s≤t
Φ dµs
which implies
c
lim sup µt (U ) ≤ lim sup µt (X) − lim inf
t↓0
t↓0
t↓0
= µ0 (X) −
Z
X
Z
X
Φ dµt
Φ dµ0
< µ0 (K c ) < ǫ
Now choose an open set V ⊃ U with compact closure V , and a continuous function ψ with
i.e. 1V c ≤ ψ ≤ 1U c
1U ≤ 1 − ψ ≤ 1V ,
(62)
Let f be a continuous function on X and write it as
f = ψf + (1 − ψ)f
Since (1 − ψ)f is continuous and of compact support,
lim
t↓0
Z
X
R
Now we must bound
|
(1 − ψ)f dµt =
X
Z
ψf dµt −
X
R
X
Z
X
(1 − ψ)f dµ0
ψf dµ0 . To this end, we have
c
f ψ dµt | ≤ |f |sup µt (U )
for all t ≥ 0.
Combining all this, we have
lim sup |
t↓0
Z
X
f dµt −
Z
X
h
c
c
i
f dµ0 | ≤ lim sup |f |sup µt (U ) + |f |sup µ0 (U )
t↓0
≤ 2|f |sup ǫ
and since ǫ > 0 is arbitrary, this is all we needed. QED
Finally, we can turn to
Proof of Theorem 1 Let f be a continuous function on G2g , invariant under the conjugation action of G, and f˜ the function induced on
26
M0g = Kg−1 (e)/G. Then
lim
t↓0
Z
G2g
f (x)Qt (Kg (x)) dx
1−2g
= vol(G)
Z
Kg−1 (e)0
1−2g vol(G)
= vol(G)
|Z(G)|
Z
f
dvol
|dKg∗ |
(by equation (61))
f˜ dvolΩ
(by Theorem 2(vii))
M0g
which is what we had set out to prove. QED
Acknowledgements. I am thankful to Jeff Mitchell for comments
and the reference [5]. Parts of this paper were written while I was
visiting the Indian Statistical Institute, Kolkata, and the University
of Bonn, and I thank these institutions and my hosts K.B. Sinha
and S. Albeverio, respectively, at these institutions. I also thank the
Alexander von Humboldt Foundation for support during my visits in
Germany and the US National Science Foundation for grants DMS
9800955 and DMS 0201683. I am most grateful to an anonymous
referee for a very careful examination of this paper, for pointing out
a serious gap in an earlier version of this paper and for several useful
comments.
References
[1] M. Atiyah and R. Bott, The Yang-Mills Equations over Riemann
Surfaces, Phil. Trans. R. Soc. Lond. A 308, 523-615 (1982)
[2] N. Bourbaki: Groupes et Algebrès de Lie, Chapitre 9. Masson,
1982.
[3] G. Bredon: Introduction to Compact Transformation Groups.
Academic Press (1972).
[4] T. Bröcker and T. tom Dieck: Representations of Compact Lie
Groups. Springer-Verlag, 1985.
[5] I. Chavel: Riemannian geometry: A modern introduction. Cambridge University Press, 1993.
[6] J. Dieudonne: Treatise on Analysis, Vols III and IV, transl. I.G.
Macdonald, Academic Press, New York and London, 1972.
[7] H. Federer: Geometric Measure Theory, Springer-Verlag, New
York, 1969.
27
[8] R. Forman: Small volume limits of 2-d Yang-Mills, Commun.
Math. Phys. 151, 39-52 (1993)
[9] W. Goldman, The Symplectic Nature of Fundamental Groups of
Surfaces, Adv. Math. 54, 200-225 (1984)
[10] K. Kawakubo: The Theory of Transformation Groups. Oxford
University Press, 1991.
[11] C. King and A. Sengupta, An Explicit Description of the Symplectic Structure of Moduli Spaces of Flat Connections, J. Math.
Phys. (Special Issue on Topology and Physics) 35, 5338-5353
(1994)
[12] C. King and A. Sengupta: The Semiclassical Limit of the Two Dimensional Quantum Yang-Mills Model, Journal of Mathematical
Physics (Special Issue on Topology and Physics) 35, 5354-5361
(1994)
[13] A.W. Knapp : Representation theory of Semisimple Groups: An
Overview Based on Examples. Princeton University Press.
[14] Deformation Quantization of Singular Symplectic Quotients,
(2001) edited by N. P. Landsman, M. Pflaum and M. Schlichenmaier. Progress in Mathematics, vol. 198. Publisher: Birkhäuser.
[15] K. F. Liu: Heat Kernel and Moduli Space,
Letters 3, 743-762 (1996)
Math. Research
[16] K. F. Liu: Heat Kernel and Moduli Spaces II, Math. Research
Letters 4, 569-588 (1997)
[17] A. Sengupta: Gauge Theory on Compact Surfaces, Memoirs of
the Amer. Math. Soc. 126 number 600 (1997)
[18] A. Sengupta, Yang-Mills on Surfaces with Boundary :Quantum Theory and Symplectic Limit, Commun. Math. Phys., 183
(1997), 661-705.
[19] A. Sengupta: The Moduli Space of Yang-Mills Connections over
a Compact Surface, Reviews in Mathematical Physics 9, 77-121
(1997)
[20] A. Sengupta: A Yang-Mills Inequality for Compact Surface, Infinite Dimensional Analysis, Quantum Probability, and Related
Topics 1, 1-16 (1998)
[21] A. Sengupta: The Moduli Space of Flat SU(2) and SO(3) Connections over Surfaces, J. Geom. Phys. 28, 209-254 (1998)
28
[22] A. Sengupta: Sewing Symplectic Volumes for Flat Connections
over Compact Surfaces J. Geom. Phys. 32, 269-292 (2000)
[23] A. Sengupta: The Yang-Mills Measure and Symplectic Structure
on Spaces of Connections, in [14]
[24] E. Witten: On Quantum Gauge Theories in Two Dimensions,
Commun. Math. Phys. 141, 153-209 (1991)
[25] E. Witten: Two Dimensional Quantum Gauge Theory revisited,
J. Geom. Phys. 9, 303-368 (1992)
29