Micro Local
Micro Local
Versions 6
Outline and Practicalities 7
Bibliography 159
6 CONTENTS
Versions
v1 30 November, 2021: L1 and compactification
v2 31 January, 2022: L2 started
v2b 4 February, 2022: Some corrections from Paige
v3 7 February, 2022: L3
v3a 8 February, 2022: L3 reorganized, correction from Zach
v3b 9 February, 2022: Asymptotic completeness added to L3
v3c 9 February, 2022: Finished L3
v4 14 February, 2022: Problems 1 included, part of L5
v5 16 February, 2022: L6
v5a 16 February, 2022: More of L6
v6 17 February, 2022: Corrections from Paige, L6
v7 19 February, 2022: Most of L7
v7a 19 February, 2022: More on L7
v7b 20 February, 2022: First version of Problems 2
v7c 22 February, 2022: More handwritten notes
v7d 24 February, 2022: Corrections from Paige
v8 26 February, 2022: L8
v9 2 March, 2022: L9
v9a 2 March, 2022: More on L9, start of L10, corrections from Paige, precision
in section on Elliptic symbols
v9b 4 March, 2022: Corrections, + version fixed
v9c 5 March, 2022: Revised propagation proof
v10 6 March, 2022: Smoothing operators for L10
v10a 7 March, 2022: Corrections
v10b 8 March, 2022: Bott periodicity statement
v11 9 March, 2022: L11
v11a 9 March, 2022: Minor
v12 9 March, 2022: L12 started
v12a 11 March, 2022: Collar nhbd
v12b 11 March, 2022: Isotropic section and corrections from Paige
v12c 14 March, 2022: Bott element
v13 15 March, 2022: L12/13
v13a 19 March, 2022: Hilbert space discussion added
v13b 20 March, 2022: Unipotent Grassmannian
v14 21 March, 2022: L14
v15 28 March, 2022: L15, corrections from Paige, Problems 3
v15a 28 March, 2022: Corrections K-theory
v15b 1 April, 2022: Corrections from Paige to Chapter 6
v16 6 April, 2022: Most of L17, L16 still to come
v17 7 April, 2022: Outline to end
v18 9 April, 2022: Most of L18
v18a 12 April, 2022: L18
v19 18 April, 2022: L19 in part
v19a 18 April, 2022: more of L19
v20 20 April, 2022: L20/21
v21 21 April, 2022: L21
v22 22 April, 2022: L22 – outline of proof of index theorem
OUTLINE AND PRACTICALITIES 7
the operator and the theorem gives a formula for it. One classical version of this
is the Riemann-Roch theorem for the ∂ operator on (line bundles over) a compact
Riemann surface. This already requires some effort to understand! There is a
one-dimensional real version of the theorem, due to Toeplitz, which states that the
index of an elliptic Toeplitz operator on the circle (the projection onto the Hardy
space, consisting of the functions smooth on and holomorphic on the interior of the
disk, of multiplication by a non-vanishing smooth function) is equal to (minus) the
winding number of the function.
You probably do know the Hodge theorem for a compact manifold without
boundary as the identification of the deRham cohomology with the space of har-
monic forms. For non-compact manifolds there is no simple generalization, rather
there are many corresponding to structures ‘at infinity’ (meaning near the bound-
ary).
Clearly, each of these theorems could easily expand to take the whole semes-
ter. Still I hope to show how they can be approached using pseudodifferential
operators and ‘quantization’. In fact an alternative title for this course might be
‘Smooth quantization’. So most of the time will be devoted to preparing the back-
ground material, specifically pseudodifferential operators on Rn , pseudodifferential
operators on a manifold, families of pseudodifferential operators and then rings of
pseudodifferential operators quantizing a Lie algebroid.
I plan to give 26 one-hour lectures in the 9:30-10:30 slot on Tuesdays and Thurs-
days and leave 20 minutes for questions and discussions (even short presentations
by students); if there is sufficient interest I will organize another ‘discussion’ time,
perhaps on Wednesdays in the afternoon. There will be notes for each topic (the
precise correspondence to the individual lectures will depend on various things),
which will include topics I will not have time to cover and will certainly include
further references – to books, lecture notes and papers. With any luck at least some
of the lectures should appear on my webpage before the beginning of the semester.
Problem sets: There will be approximately 5, every two weeks. Grading may
be by discussion with me.
Grades: Graduate students are expected to participate actively. That is what
‘A’ means to me. By this I mean that I expect people to attend lectures and to ask
questions. For undergraduates this course might be heavy lifting, it is for me, so
please talk to me by early in the semester at the latest. We can discuss what you
should expect. There are no exams.
Prerequisites: I will assume familiarity with manifolds and distributions, essen-
tially as in 18.155 but plan to review pretty much everything.
Why don’t I just follow a book or my earlier lecture notes? This probably
reflects some personal failing and general dissatisfaction with how things are done!
I find it difficult to think through things without seeing some other way of ap-
proaching them. If it is not to your taste, I am sorry but that is the way it is. I
may not get to all the results listed above, but I expect to at least get to the point
where they are all within reach and that is really what I want to do – try to put
these results in a general context that maybe encourages them to be exploited (i.e.
applied) and extended.
In the interim, feel free to contact me with questions or comments.
Richard Melrose, 17 November, 2021.
CHAPTER 1
157.42
the function pα (x), on the left. This is true in (1.3) as well but there the constants
commute with the 157.43
differentiation operators. Of course this is reflected in the fact
that the product (1.5) is not commutative.
Now, let’s concentrate Schwartz space. For this we have the Fourier transform
Z
157.4 (1.6) F : S (Rn ) −→ S (Rn ), F u(ξ) = û(ξ) = e−ix·ξ u(x)dx.
Rn
It is a linear isomorphism. We know that
157.5 (1.7) u ∈ S (Rn ) =⇒ F (Dα u)(ξ) = ξ α û(ξ).
The Fourier transform conjugates differentiation to multiplication. Whilst a mono-
mial such as ξ α is not in the Schartz space, but it does define an operator on it by
multiplication.
So the inverse Fourier transform, u(x) = (2π)−n eix·ξ û(ξ), allows us to write
R
Z
157.6 (1.8) Dα u(x) = (2π)−n eix·ξ ξ α û(ξ)dξ.
Rn
157.3
A linear partial differential
157.6 157.3
operator, (1.4), is given by a finite sum so we can
combine (1.8) with (1.4) and write
Z X
−n
157.7 (1.9) P u(x) = (2π) p(x, ξ)û(ξ)dξ, p(x, ξ) = pα (x)ξ α dξ.
Rn |α|≤m
n
Since û ∈ S (R ), the integral converges absolutely. If we just assume that the
coefficients are in C ∞ (Rn ) then the integral converges uniformly on compact subsets
in x ∈ Rn , with all its formal derivatives in x because of the obvious estimates
157.8 (1.10) |Dxγ p(x, ξ)| ≤ CK,γ (1 + |ξ|)m , x ∈ K b Rn , ξ ∈ Rn .
We can actually define the ‘standard’ space of pseudodifferential operators of
order m ∈ R by considering those functions a ∈ C ∞ (Rnx × Rnξ ) which satisfy the
symbol estimates
157.9 (1.11) |Dxγ Dξβ a(x, ξ)| ≤ Cβ,γ (1 + |ξ|)m−|β| , ∀ γ, β ∈ Nn0 .
157.7
Notice that p in (1.9) satisfies these estimates for an integer m if the coefficients
are in the space
∞
157.10 (1.12) C∞ (Rn ) = {f ∈ C ∞ (Rn ); sup |Dxγ f (x)| < ∞ ∀ γ},
consisting of the smooth functions with all derivatives
157.9
bounded.
m
The space of functions satifying estimates (1.11) is often written S1,0 as part
m
of a more general class of spaces Sρ,δ where the exponent m − |β| is replaced by
m − ρ|β| + δ|α|. Since we will not need the generalHormander-WeylCalculus
ρ, δ spaces – in fact there are
many variants of such estimates (see for instance [5]) – and we will already have
enough things to think about I will use157.9
the shorter notation, for the moment just
S m = S1,0
l
for the functions satisfying
157.9
(1.11).
It follows directly from (1.11) that if a ∈ S m then the direct generalization of
157.7
(1.9),
157.11 (1.13) Z
A(x, Dx )u = Au(x) = (2π)−n ∞
a(x, ξ)û(ξ)dξ =⇒ a : S (Rn ) −→ C ∞ (Rn ).
Rn
In fact much more is true:
1. PSEUDODIFFERENTIAL OPERATORS, MANIFOLDS AND COMPACTIFICATION 11
Then one can apply the more obvious fact the product is rapidly
decaying in ξ :
S m · S (Rnξ ) ⊂ S m−k ∀ k ∈ R. 157.46 (1.17)
157.11
The integral (1.13) is therefore convergent. Again, if you
like to be precise, you can see that
Sm ⊂ C ∞
0
(Rn ; L1 (Rnξ )), m < −n 157.47 (1.18)
−n− 1 n
since (1 + |ξ|) ∈ L (R ) if > 0. Now we can use standard
properties of Lebesgue (or improper Riemann) integrals to see
0
that Au ∈ C ∞ (Rn ) is a bounded continuous
157.11
function and the
same holds for all derivatives giving (1.13).
We should check a couple of other statements, towards The-
157.12
orem 1.1. First the stronger mapping property that
A : S (Rn ) −→ S (Rn ). 157.48 (1.19)
This is a matter of getting ‘decay’. Namely we need to show
that for any monomial and any derivative
xγ Dxα Au ∈ C ∞
0
(Rn ). 157.49 (1.20)
We can approach this one step at a time, asking just about xj Au.
Note that we can certainly multiply by xj but the operator xj A
is not in general in Ψm (Rn ) (for any m) since xj a(x, ξ) is not
1The ∗ in the header of the theorem is to indicate that I will not prove it immediately but a
full proof, and more, will follow later.
2Narrowed parts of a lecture are things I don’t expect to have time to cover.
12 1. PSEUDODIFFERENTIAL OPERATORS, MANIFOLDS AND COMPACTIFICATION
bounded
157.11
as |xj | → ∞ even for fixed ξ. However the integral in
(1.13) still converges rapidly in ξ for x in a compact set if we
replace a by xj a so
xj A : S (Rn ) −→ C ∞ (Rn ) 157.50 (1.21)
for instance.
157.50
157.51 Lemma 1.1. In the sense of operators (1.21)
[xj , A] = xj A − Axj ∈ Ψm−1 (Rn ). 157.52 (1.22)
Proof. We use ‘integration by parts’. Consider the opera-
tor Axj . The Fourier transform of xj u, u ∈ S (Rn ) is i∂ξj û so
Z
−n
Axj u = (2π) a(x, ξ)eix·ξ i∂ξj û(ξ)dξ = xj A(x, D)u + bj (x, D)u,
which I will denote Ψm
We will be most interested in a smaller space157.13 n
cl (R ) often
called the ring (with the composition property (1.14)) of ‘classical’ pseudodifferen-
tial operators where the symbols a have the additional property:
157.9
157.453 Definition 1.1. A symbol in the sense of (1.11) is classical (also ‘polyhomo-
geneous’) if there exists a sequence ai ∈ C ∞ (Rnx × (Rnξ \ {0})) of homogeneous
functions of degree m − i (in the ξ variables)
157.14 (1.25) ai (x, tξ) = tm−i a(x, ξ), t > 0, (x, ξ) ∈ Rn × (Rn \ {0})
such that for (any) cutoff χ ∈ C c∞ (Rnξ ) with χ = 1 near 0
N
X
157.15 (1.26) a(x, ξ) − (1 − χ(ξ))ai (x, ξ) ∈ S m−N −1 .
i=0
m
These ‘classical’ symbols form a filtered subring Scl
157.15
⊂ S m = S m . The relation-
ship (1.26) (see Problem set 1) is often written
X
157.33 (1.27) a' ai
i
and a is then said to have a complete asymptotic expansion. Such ‘asymptotic
summation’ (the existence of a given the ai ) is discussed below, it is closely related
to E. Borel’s ‘Lemma’ on Taylor series. There is no statement of convergence of the
157.15
series in (1.26) (although there is one lurking in the background) but you should be
157.15
able to see that the ai , assuming they exist are determined
157.11
by the relations ( 1.26).
When we insert such classical symbols in (1.13) (or if you prefer, restrict to
classical symbols) the resulting space of constitutes a filtered subring Ψm n
cl (R ) ⊂
1. MANIFOLDS WITH CORNERS 13
Ψm (Rn ) which for positive integral m includes the differential operators of order m
discussed above.
These two rings have many important properties but one of the most important
157.15
is that one can recover the terms ai in (1.26) from the operator A and the leading
term defines the principal symbol, σm , as a map
157.16 (1.28)
Ψm (Rn )
σm
/ {a0 ∈ C ∞ (Rn × (Rn \ {0}) homogeneous of degree m in ξ}.
cl
and this map is surjective, multiplicative and defines a short exact sequence
0
σm+m0 (A ◦ B) = σm (A)σm0 (B), A ∈ Ψm n m n
cl (R ), B ∈ Ψcl (R )
157.17 (1.29) Ψm−1
cl (Rn ) ,→Ψm n
cl (R ) −→
{a0 ∈ C ∞ (Rn × (Rn \ {0})) homogeneous of degree m in ξ}
Here I have stuck with a cumbersome notation for the homogeneous space which
will be refined below. There are similar exact sequences for the larger algebra
Ψ∗ (Rn ) but the principal symbol lies in a quotient space.
So, we want to prove all these things and a lot more! However, I do not want to
go there directly but rather map out the territory a bit first, in particular discussing
the ‘symbol spaces’ concretely. L1-end
V 0 ⊂ Rn such that the range V = V 0 ∩ [0, ∞)n . The coordinates on the coordinate
patch are the pull-backs of the coordinate functions xi on Rn .
To make clear what ‘C ∞ -related’ for two such coordinate patches means, we
need to define C ∞ (V ) (I will not bother with lower regularity than C ∞ ):
several useful ways to restate this condition but note how it fails for a ‘tear-shaped
region’ in the plane.
The C ∞ functions on M are those that are C ∞ in each coordinate patch, mean-
ing
157.38 (1.33) f ∈ C ∞ (M ) ⇐⇒ (F −1 )∗ (f U
) ∈ C ∞ (V ) for each coordinate patch.
This is equivalent to the same condition for any one compatible atlas.
The direct consequence of the ‘embedded’ requirement is that the boundary
hypersurfaces have defining functions:
157.39 (1.34) Hi ∈ M 1 (M ) =⇒ ∃ ρi ∈ C ∞ (M ), ρi ≥ 0, Hi = {ρi = 0},
d((F −1 )∗ ρi )(F (p)) 6= 0 ∀ p ∈ Hi for all coordinate patches containing p.
This last condition means that for each p ∈ Hi there is a coordinate patch containing
p in which ρi is a coordinate function.
If M̃ is a manifold without boundary, i.e. M̃1 = ∅, then M ⊂ M̃ is a(n embed-
ded) submanifold if M has a covering by coordinate patches of M̃ which restrict to
give it the structure of a manifold with corners.
157.40 Theorem 1.2. For any manifold with corners there exists a manifold without
boundary M̃ of the same dimension in which M is embedded as a submanifold; if
M is compact then M̃ can be taken to be compact.
Although there is no quite canonical way of constructing such an extension, M̃ ,
all the standard constructions of the tangent, cotangent, form bundles and other
bundles associated to the frame bundle, pass over to the case of a manifold with
corners in such a way that the restrictions for an extension of this type are canonical
157.41 (1.35) T M = T M̃ M
, T ∗ M = T ∗ M̃ M
etc.
However, there are important additional structures which arise from the boundary
faces as I will discuss later.
Why work in this degree of generality? Manifolds with corners are the smooth
(i.e. C ∞ ) analogue of smooth algebraic varieties with divisors and they occur for
similar reasons. One place manifolds with corners arise is through ‘compactifica-
tion’.
Remark 1. From now on the default meaning for ‘manifold’ is a manifold with
corners.
As you are no doubt aware, it is bad form to specify a class of objects, in this
case manifolds without specifying the ‘morphisms’ between them. For manifolds
without boundary the morphisms, forming a category, are just the smooth maps.
This also works when there are corners but it is not a good idea. The problem is
that a smooth map F : M −→ N in general does not ‘respect’ the boundary. For
instance a smooth curve, so M = (0, 1) might have a single point of tangency to a
boundary hypersurface.
What we want are b-maps, where as always the b- just stands boundary. Each
boundary hypersurface is specified by the ideal of functions which vanish on it;
this is a principal ideal because we are assuming boundary hypersurfaces to be
embedded. So if ρH is a defining function for H ∈ M 1 (M ) then
157.965 (1.36) I (H) = ρH C ∞ (M ).
16 1. PSEUDODIFFERENTIAL OPERATORS, MANIFOLDS AND COMPACTIFICATION
(3)
157.970 (1.40) F ∗I (H) contains a non-vanishing element.
The third case corresponds to F# (H) as in the second, with all entries 0. This is
more evident if one chooses a defining functiona ρ0H for H and ρK for the boundary
hypersufaces of M and then the three cases correspond to F ∗ ρ0H = 0,
F (H,K)
Y
157.971 (1.41) F ∗ ρ0H = a ρK# ,
K∈M 1 (M )
157.978 Lemma 1.2. For a real-valued element V ∈ Vb (M ) and each p ∈ M there exists
a maximal integral curve
157.982 Proposition 1.2. On a compact manifold the Lie algebra of the groupd of
diffeomorphisms is Vb (M ).
Note that this infinite-dimensional group is missing some properties of a Lie group,
for one the diffeomorphism obtained by integrating the Lie algebra do not contain
a neighbourhood of the identity. This can be obviated by considering parameter-
dependent vector fields.
If W is a t-depedent smooth vector field on M, so a vector field in Vb (U ) for an
157.978
open subset U ⊂ Rt × M such that V t = 0 then Lemma 1.2 yields a 1-parameter
family of diffeomorphisms, Gt defined on an appropriate open subset, and satisfying
d df
157.981 (1.48) (Gt )∗ f = (Gt )∗ (W f + ), f ∈ C ∞ (R × M ).
dt dt
This invariant formulation of the chain rule, and its generalizations, are used below
in various constructions.
2. Compactification
T1.Compactification
Although we will deal with non-compact manifolds, the ones that arise below
have some ‘structure at infinity’. One way to describe what this means is through
the notion of compactification.
Here, both M and M may have corners. As always when introducing a new
notion, we should specifiy when two compactifications are to be regarded as ‘the
same’.
18 1. PSEUDODIFFERENTIAL OPERATORS, MANIFOLDS AND COMPACTIFICATION
157.20 (1.49) =M 1
ι1
M e
ι2
!
M 1.
o
(1) The one-point compactification(s) given by a sphere V .
p
(2) The parabolic compactification gives a closed ball V .
R
(3) The radial compactification also given by a closed ball V = V .
2. COMPACTIFICATION 19
ι(t)
0
x
From the notation you can see that I have a preference for the radial compacti-
fication – I hope the discussion below shows why. Only the radial compactification
is really used subsequently.
These can all be constructed using variants of stereographic projection. So, let’s
start with V = Rn , i.e. choose a basis. We embed Rn into Rn+1 as the hyperplane
157.21 (1.50) Rn 3 x 7−→ (x, 1) ∈ P ⊂ Rn+1 .
In the first, case consider the the sphere So of radius 12 centred at (0, 12 ) and in the
second and third cases take the sphere SR of radius 1 centred at the origin. In both
these latter cases a point of Rn determines a unique line Lo (x) or LR (x) through
the image of x in P and the centre of the corresponding sphere then
157.22 (1.51)
Io : Rn −→ So , Io x is the other point in So ∩ Lo (x)
IR : Rn −→S+
R,
IR x is the other point in SR ∩ L1 (x) ⊂ S+
R = SR ∩ {xn+1 ≥ 0}
Ip : Rn −→Bp ⊂ Rn ,
Ip x is the projection of LR x onto the closed unit ball in Rn × {0}.
In all three cases the full orthogonal group O(n), acting on the first factor of Rn ×Rn
satisfies I• Ax = AL• x for all A ∈ O(n), effectifely reducing the discussion to the
case n = 1. Explicit formulæ for the maps are easily derived:
x 1
Io x = , ∈ So ⊂ Rn+1
+
1 + |x|2 1 + |x|2
x 1
157.23 (1.52) IR = 1 , 1 ∈ S+ n+1
R ⊂ R+
(1 + |x| ) 2 (1 + |x|2 ) 2
2
x n
Ip x = 1 ∈ R .
(1 + |x|2 ) 2
20 1. PSEUDODIFFERENTIAL OPERATORS, MANIFOLDS AND COMPACTIFICATION
1
Thus, for the radial compactification (1 + |x|2 )− 2 is a boundary defining function
and hence |x|−1 , which is a smooth function of it away from x = 0, is a defining
function near the boundary. It follows that
1 x
157.27 (1.53) {|x| > > 0} 3 x 7−→ , ∈ [0, 1) × Sn−1
|x| |x|
R
extends to a smooth product decomposition of Rn near the boundary. For the
parabolic compactification it follows similarly that
1 x
157.28 (1.54) {|x| > > 0} 3 x 7−→ , ∈ [0, 1) × Sn−1
|x|2 |x|
is a product decomposition near the boundary.
It can be seen directly that
x
157.24 (1.55) Io = SIo where S : So \ {(0, 1), (0, 0)} −→ So \ {(0, 1), (0, 0)},
|x|2
with S(y, yn ) = (y, −yn + 1)
is equatorial reflection on So . 157.23
In all cases it is clear either geometrically, or from the forumlæ (1.52), that the
action of O(n) extends smoothly from Rn to the compactification. Similarly the
scaling action by R+ , with generator on Rn
X ∂
157.25 (1.56) xi
i
∂xi
157.24
extends smoothly. For the one-point compactification this follows from (1.55) and
in the other two cases
tx x 1
157.26 (1.57) lim 1 = 1 and lim 1 = 0.
|x|→∞ (1 + t|x|2 ) 2 2
(|x| ) 2 |x|→∞ (1 + t|x|2 ) 2
Thus in all cases the action of the conformal group O(n)×R+ extends smoothly
to the compactification.
157.29 Proposition 1.3. The action of the general linear group extends smoothly
from Rn to the radial and parabolic compactifications, but not to the one-point
compactification; the translation action of Rn extends smoothly to the radial and
the one-point compactifications, but not to the parabolic compactification and there
are smooth surjective maps, which are not diffeomorphisms, giving a commutative
diagramme
157.30 (1.58)
GL(n, R) n Rn
Sn+
RO
IR
y Ip %
Sn0 o / Bnp
Io
O(n) n (R+ × Rn ) Rn GL(n, R).
3. COLLAR NEIGHBOURHOOD 21
3. Collar neighbourhood
Sect.collar
157.447 Remark. Edited by Paige Dote
This theorem provides a rather precise description of a neighbourhood of a
closed embedded submanifold, Y ⊂ M where M is an n-dimensional manifold.
The usual proof exploits the geodesic flow for a metric on M, but here we give a
related approach using the notation of a radial vector field for Y. This is very closely
Sternberg
related to the linearization theorem of Sternberg ([?]). We also carry it out in the
context of manifolds with corners.
A closed p-submanifold (p- if for ‘product’ it really is the natural notion of a
submanifold in the case of a manifold with corners) is a closed subset Y ⊂ M such
that for each point y ∈ Y , there exists adapted coordinates based at y in M in
terms of which Y is linear. Precisely, in a neighbourhood Uy ⊂ M of y, there exists
coordinate functions xi ≥ 0, yj ∈ C ∞ (Uy ) for i = 1, . . . , k and j = 1, . . . , n − k with
independent differentials and
157.448 (1.59) Y ∩ Uy = {x1 = · · · = xd0 = 0, y1 = · · · = yn0 −d0 ==}.
It follows that Y is itself a submanifold, with these as adapted coordinates for Y.
Note that if d0 > 0 here then Y is actually contained in a boundary face (assum-
ing it is connected, in any case one component is). Boundary faces themselves are
clearly p-submanifolds but quite a few natural submanifolds are not. For instance
the diagonal in M 2 , where ∂M 6= ∅, is not a p-submanifold. This turns out to be
rather significant.
We note the following
22 1. PSEUDODIFFERENTIAL OPERATORS, MANIFOLDS AND COMPACTIFICATION
for smooth coefficients ai,j,p and b . Pushing R formward under the inverse of the
157.961 i,k
scaling diffeomorphism (1.61), we obtain the one parameter family of vector fields
X X
157.450 (1.62) Rt = R0 + txi xj ai,j,p (tx)∂xp + txi bi,k (tx, y)∂yk .
i,j,p i,j
24 1. PSEUDODIFFERENTIAL OPERATORS, MANIFOLDS AND COMPACTIFICATION
Here I use the same letter for the operator and its Schwartz kernel – since the
Schwartz kernel theorem (which I will talk a little about later) shows that they
determine each other. 157.94
We can think of (2.1) in a couple of different ways – in general it is not a
convergent integral. We can make a (formal at this stage) linear change of variables
on R2n from (x, y) to (x, z), z = x − y and then
Z
−n
157.95 (2.2) A(x, y) = α(x, x − y) where α(x, z) = (2π) a(x, ξ)eiz·ξ dξ.
Rn
It is also the case that S (Rn ) is closed under convolution and we know that the
Fourier transform satisfies
157.59 (2.6) F (u ∗ v) = F (u)F (v), u, v ∈ S (Rn ).
The ring we are interested in is contained in
157.60 (2.7) C c−∞ (Rn ) + S (Rn )
157.59
for which the identity (2.6) still holds. Note that
157.62 (2.8) F (C c−∞ (Rn ) + S (Rn )) ⊂ C ∞ (Rn ) ∩ S 0 (Rn ).
So, we are looking for are some interesting spaces of smooth functions on the
dual Rn which are closed under multiplication. You might ask, in view of the
identification of the convolution kernels here with the inverse Fourier transforms of
symbols, why is there any problem at all? There isn’t a problem for convolution as
157.59
such because of (2.6) but recall that the Fourier transform does not ‘behave well’ on
say the space L∞ (Rn ). Of course the Fourier tranform maps this to a well-defined
linear subspace of the tempered distributions – which includes for instance the delta
functions at any point – but it is quite hard, in a certain sense I think impossible,
to give a ‘direct’ characterization of the Fourier image of L∞ and the same is true
for our symbols which are modeled on L∞ in the sense that they are defined by
bounds. We will in fact ‘sandwich’ the image between spaces characterized directly
(meaning without the Fourier tranform), but this still loses information which is
rather vital to us!
In the notes related to the first lecture, I discussed the radial compactification
of a real, finite-dimensional, vector space V, to a ball V . Ignoring all the niceties,
for Euclidean space, Rn with the standard Euclidean norm, we can identify the
complement of the origin with the product
x
157.63 (2.9) Rn \ {0} 3 x 7−→ |x|, = (r, ω) ∈ (0, ∞) × Sn−1 .
|x|
The inversion map r −→ 1/r is a diffeomorphism of (0, ∞) to itself ‘switching the
ends’. This allows us to add the sphere at infinity of Rn setting
Rn = Rn t [0, ∞) × Sn−1 /I
157.64 (2.10)
where
n 1 x
157.65 (2.11) I : R \ {0} 3 x 7−→ , ∈ (0, ∞) × Sn−1
|x| |x|
identifies the complement of the origin with the interior of the second part.
Thus Rn is a compact manifold with boundary ‘obtained by introducing in-
verted polar coordinates near infinity’. The interior is Rn and the boundary is ‘the
sphere at infinity’.
This immediately gives us a ring of functions on Rn , namely
157.66 (2.12) C ∞ (Rn ) ,→ C ∞ (Rn ).
I can write inclusion here for what is really the restriction from Rn to its interior
since this map is injective.
This is the space of ‘classical symbols on Rn of order zero’ which I will write as
157.67 (2.13) 0
Scl (Rn ) = C ∞ (Rn ).
I will approach the issue of characterizing this space precisely on Rn below.
1. SCHWARTZ KERNELS 27
T1.Compactification
As a consequence of the discussion of radial compactification in § 2, or directly,
we can see that the coordinate vector fields on Rn extend to be smooth on Rn . In
fact
157.78 Proposition 2.1. The coordinate vector fields on Rn extend to smooth vector
files on Rn and span, over C ∞ (Rn ), all the smooth vector fields which are of the
form
157.79 (2.14) ρW, W smooth and tangent to the boundary of Rn .
Here ρ ∈ C ∞ (Rn ) vanishes at the boundary.
P-Comp
See Project 1
0
157.80 Corollary 2. The space Scl (Rn ) consists of smooth functions which satisfy
the estimates
157.81 (2.15) sup |(1 + |ξ|)|α| ∂ξα a(ξ)| < ∞ ∀ α.
ξ∈Rn
0
Note that I do not say that this characterizes Scl (Rn ) = C ∞ (Rn ), because it
does not.
157.83 Definition 2.1. We denote the subspace of C ∞ (Rn ) of functions satisfying all
157.81
the estimates (2.15) by
157.84 (2.16) S 0 (Rn ) ⊃ Scl
0
(Rn ).
These are the ‘symbols with bounds’ containing the classical symbols.
More generally, consider the function
157.68 (2.17) (1 + |x|2 )z/2 on Rn , z ∈ C.
This is certainly smooth on Rn . It is rather clear that
157.69 (2.18) (1 + |x|2 )z/2 ∈ C ∞ (Rn ) iff z ∈ −N0 .
Indeed, in x 6= 0 it can be written
157.70 (2.19) t−z (1 + t2 )z/2 , t = 1/|x|.
This is smooth down to t = 0, the boundary of Rn , if and only if −z is a non-negative
integer.
We define the space of classical symbols of (complex) order z to be the products
157.71 (2.20) z
Scl (Rn ) = (1 + |x|2 )z/2C ∞ (Rn ) = (1 + |x|2 )z/2 Scl
0
(Rn ).
The space of symbols (with bounds) or real order m is similarly defined to be
157.85 (2.21) S m (Rn ) = (1 + |x|2 )m/2 S 0 (Rn ).
Why no complex order in the second case?
157.83
157.86 Exercise 2. Show that in terms of Definition 2.1
157.87 (2.22) (1 + |x|2 )is/2 ∈ S 0 (Rn ) ∀ s ∈ R.
This in turn implies that
z
157.88 (2.23) Scl (Rn ) ⊂ S Re z (Rn ) ∀ z ∈ C.
28 2. SYMBOLS AND CONORMAL DISTRIBUTIONS AT A POINT
Last time I talked about the symbol spaces S m (Rn ) and the space of distribu-
tions conormal at 0 defined as
m+ n
157.97 (2.33) IS 4
(Rn ; {0}) = F −1 (S m (Rn )).
Since we know that the symbol spaces form a filtered ring under multiplication we
deduce a corresponding result for convolution of the conormal spaces
0 M +M 0 − n
157.98 (2.34) ISM (Rn ; {0}) ∗ ISM (Rn ; {0}) = IS 4
(Rn ; {0}).
2. Homogeneous distributions
The quesiton naturally arises as to precisely what classical
conormal distributions ‘look like’. The answer is given, almost,
in terms of homogoneous distributions. Then the first question
is, What is a homogeneous distribution?
We know what a homogeneous function on Rn \ {0} is – it
is a function satisfying
u(tx) = tz u(x) ∀ t > 0, x ∈ Rn \ {0}. 157.983 (2.35)
We do not want to include the origin here since the degree z
could be negative, so the function need not be defined at 0. In
fact we are interested in smooth functions away from 0 which
are homogeneous. So this is straihgtforward
157.984 Lemma 2.2. The smooth functions on Rn \ {0} which are
homogeneous of degree z ∈ C and smooth outside the origin may
be identified with C ∞ (Sn−1 ) via
x
u(x) = v( )|x|z , v ∈ C ∞ (Sn−1 ). 157.985 (2.36)
|x|
When Re z > −n it follows that such a function is locally
integrable across the origin, and so defines a tempered distri-
bution. In this case we can express the condition ‘weakly’ by
noting
Z Z
u(tx)φ(x)dx = t−n u(x)φ(x/t)dx =⇒
Certainly, a ∈ S m (Rn ) if and only if a ∈ C ∞ (Rn ) and all these norms are finite.
Recall that a metric on a countably normed space, such as this, is defined by
X ku − vkm,N
157.100 (2.56) d(u, v) = 2−N .
1 + ku − vkm,N
N
157.100
So the topology is metric, generated by the open balls with respect to (2.56). I say
‘a metric’ because replacing the sequence 2−N by an positive, summable, sequence
gives the same topology.
157.114 Proposition 2.4. The spaces S m (Rn ) are Fréchet spaces, so complete with
157.100
respect to the translation-invariant distance (2.56).
If it matters to you, they are Montel spaces. They are not projective limits of
Hilbert spaces.
Proof. Convergence with respect to this distance is the same as convergence
with respect to each of the norms k · km,N (without
157.100
any uniformity in N ). Thus a
Cauchy sequence with respect to the metric (2.56) is Cauchy with respect to each
3. TOPOLOGY AND ASYMPTOTIC SUMMATION 33
of these norms and conversely. So all the derivatives converge locally uniformly and
with respect to the distance with the limit in the space.
There is a topology on S ∞ (Rn ) = m S m (Rn ) but I will leave you to figure it
S
out:
157.454 Exercise 3. Try to sort out (or look up) the inductive limit topology on
S ∞ (Rn ) defined by taking a set to be open if its intersection with each of the
S m (Rn ) is open and show that the inclusions S m (Rn ) −→ S ∞ (Rn ) are then con-
tinuous.
So the Fréchet topology on the symbol spaces induces a Fréchet topology
n
on ISM (Rn , {0}), since the Fourier transform identifies this with S M − 4 (Rn ). This
means we know what a continuous map into the conormal space (and also what a
smooth map into it) means.
Now to density. The intersection of the symbol spaces is the space of Schwartz
functions
\
157.101 (2.57) S (Rn ) = S −∞ (Rn ) = S m (Rn ) = S (Rn ).
m
157.102 Proposition 2.5. The ‘residual space’ S (Rn ) is dense in S m (Rn ) in the topol-
ogy of S m+ (Rn ) for any > 0. More precisely, there exist a sequence of ‘regularizing
operators’ which are linear maps
157.103 (2.58) Φk : S ∞ (Rn ) −→ S (Rn )
such that
157.104 (2.59) a ∈ S m (Rn ) =⇒ Φk a is bounded in S m (Rn )
and Φk a → a in the topology of S m+ (Rn ) ∀ > 0.
Proof. The Φk can be defined by cut-off. Take φ ∈ C c∞ (Rn ) with φ(ξ) = 1 in
{|ξ| < 1} and set
157.105 (2.60) Φk a(ξ) = φ(ξ/k)a(ξ) ∈ S (Rn ).
The difference
157.106 (2.61) (Id −Φk )a = (1 − φ(ξ/k))a(ξ)7 ∈ S m (Rn )
since 1 − φ(ξ/k) ∈ S 0 (Rn ).
Certainly 1 − φ(ξ/k) is uniformly bounded and the derivatives are
157.107 (2.62) ∂ξβ (1 − φ(ξ/k)) = −k −|β| (∂ β φ)(ξ/k), |β| > 0.
Since this function is supported in |ξ| < Ck, for some constant C, the product
satisfies
157.108 (2.63) sup |∂ξβ (1 − φ(ξ/k))| ≤ Cβ (1 + Ck)|β| k −|β| < ∞
ξ
This shows that 1 − φ(ξ/k) is bounded with respect to all the seminorms for
S 0 (Rn ). It follows that Φk a is bounded in S m (Rn ).
The seminorms on S (Rn ) on the difference 1−Φk , which has support in |ξ| > k,
have an extra factor of (1 + |ξ)−
4. Integration
157.109 Lemma 2.7. Integration in one of the variables, say the last, gives a continuous
linear map
Z Z
m n m− 14 n−1
157.110 (2.67) dxn : IS (R ) −→ IS (R ), F u(x)dxn = F (u) ξn =0 .
R R
Of course we can iterate this, integrating over k variables to get a conormal distri-
bution with order decreased by k/4.
5. Wavefrontset
The support of a function or distribution on Rn is defined by
[ {
supp(u) = {U ⊂ Rn ; U is open and u = 0 on U } . 157.115 (2.71)
This is really a notion defined for sheaves (the theory of which I will outline
below in case you have not seen it). We define a related notion for symbols
which is to do with the directions of growth at infinity.
If V ⊂ Sn−1 is open then the set
ξ
R+ V = {ξ ∈ Rn \ {0}; ∈V} 157.116 (2.72)
|ξ|
6. RESTRICTION 35
6. Restriction
157.110
What is this notion of wavefrontset good for? Notice in (2.67) that inte-
gration and restriction are dual under Fourier transform, at least in this
special case. In general we cannot expect to restrict a conormal distribution
to xn = 0 – for instance this is not reasonable for the delta function at the
origin. Dually, we cannot expect to integrate a symbol, it may just be too
large at infinity.
36 2. SYMBOLS AND CONORMAL DISTRIBUTIONS AT A POINT
xn =0
: {u ∈ ISM (Rn ; {0});{dxn , −dxn } ∩ WF(u) = ∅}
3
157.128 (2.88)
−→ I M + 4 (Rn−1 ; {0}).
7. Multiplicativity
One of the applications of the notion of wavefront – for general distributions
not just in the conormal case – is to provide conditions under which operations
extend to distributions.
R.11 Exercise 5. Show that if Γi ⊂ Rn \ {0}, i = 1, 2, are (relatively of course)
closed cones with the property
R.12 (2.89) ξi ∈ Γi , 1, 2 =⇒ ξ1 + ξ2 6= 0
8. ASYMPTOTIC COMPLETENESS 37
8. Asymptotic completeness
The main interest in symbols on Rn is their behaviour ‘at infinity’ (which is the
boundary of the radial compactification). This allows for a notion of ‘convergence’
which corresponds to the ‘asymptotic completeness’ in the following sense.
157.140 Theorem 2.2. If aj ∈ S mj (Rn ) is a sequence (we think of it as a series)
of symbols with mj → −∞ as j → ∞ then there exists a symbol a ∈ S M (Rn ),
M = sup mj such that for every k
X
157.141 (2.92) a− aj ∈ S M (k) , M (k) = sup mj
j>k
j≤k
Note that we are certainly not saying that the series on the right converges in
any sense (well people say it converges asymptotically, just meaning the order
mj → −∞).
I have been a little vague here about the range of j, usually one takes j ∈ N0 ,
so starting off at 0, but this is just a convention.
157.141
Proof. The ‘uniqueness’ (modulo S −∞ (Rn )) is immediate from (2.92) – given
two such ‘asymptotic sums’ a and a0 the difference satisfies
X X
157.153 (2.94) a0 − a = (a0 − aj ) − (a − aj ) ∈ S M (k) (Rn ) ∀ k =⇒
j≤k j≤k
a0 − a ∈ S −∞ (Rn ) = S (Rn ).
38 2. SYMBOLS AND CONORMAL DISTRIBUTIONS AT A POINT
For existence, I will assume, as discussed below, without loss of generality that
the mj are strictly decreasing, just to simplify notation.
I will use the ‘approximation’ operators (Id −Φl ) discussed above, where Φl is
multiplication by φ(ξ/l) for φ ∈ C c∞ (Rn ) equal to 1 near 0. So these are cutoffs near
infinity. Here l will vary with j so we are looking for a sequence of integers
157.144 (2.95) l(j) → ∞ in N.
What we want these integers to satisfy – they depend of course on the given sequence
aj – is
X
157.145 (2.96) k(Id −Φl(j) )aj kmk ,N < ∞ ∀ k, N.
j>k
in the strong sense that it converse absolutely with respect to each of the seminorms.
Now set
X
157.151 (2.102) a = a0 + (Id −Φl(j) )aj ∈ S m0 (Rn ).
j≥1
This is our asymptotic sum. To check this observe that the difference with a finite
sum can be written
X X X
157.152 (2.103) a− aj = − Φl(j) aj + (Id −Φl(j) )aj .
j≤k+1 j≤k+1 j>k+1
mk+1
The last sum here is in S and the finite sum is actually of compact support, so
157.141
in S −∞ (Rn ). The last term on the left is in the same space, S mk+1 (Rn ) so (2.92)
follows.
If we do not have a strictly decreasing sequence of orders, we can rearrange the
sequence so that the order is weakly decreasing and then sum up an finite sequences
of fixed order. This reduces the problem to the strictly decreasing case and, since
157.141
we have arranged absolute convergence, we recover (2.92) in general.
Except that the topology is a little dubious, we have shown that a series with
elements in S mj (Rn )/S −∞ (Rn ) ‘always converges’ if the mj → −∞. What this
really means is that there exist representatives of the elements in S mj (Rn ) such that
the series does converge and gives a well-defined limit in S sup mj (Rn )/S −∞ (Rn ). L3-end
L4
The smooth functions of ‘slow growth’ form a (not very pleasant) linear space
which I will denote by O (Rn ) which is a space of multipliers on S (Rn ). A smooth
function is an element ψ ∈ O (Rn ) if for each multiindex α ∈ Nn0 there exists mα
such that
157.154 (2.106) sup(1 + |ξ|)−mα |Dξα ψ(ξ)| < ∞.
ξ
n
The sum of the right does determine a unique element of S m− 4 /S −∞ (Rn ) =
n
S m− 4 (Rn )/S −∞ (Rn ) as we showed last time using asymptotic summation. If we
n
take a representative a ∈ S m− 4 (Rn ) – such as the actual Fourier transform of u –
then the terms in the infinite sum are of orders m − n4 − |α| so the sum ‘converges
asymptotically’.
Proof. Since we know that u ∈ ISm (Rn ; {0}) can be written as the sum of a
compactly supported term and one in S (Rn ) on which ψ ∈ O (Rn ) acts it suffices to
suppose that both u and ψ are compactly supported. The Taylor series expansion
X 1 X
157.458 (2.109) ψ= ∂xα xα ψ(0) + xα µα (x), µα ∈ C ∞ (Rn )
α!
|α|≤N |α|=N +1
m−|α|
The terms in the first sum are are in IS (Rn ; {0}) as the inverse Fourier trans-
forms of the
i|α| α |α|
157.460 (2.111) ∂ ψ(0)∂ξ û(ξ).
α! x
Similarly the terms in the second sum consists of products in
C c∞ (Rn ) · ISm−N −1 (Rn ; {0}).
157.460
Let v ∈ ISm (Rn ; {0}) be an asymptotic sum of this series (2.111) which can be
157.459
taken to have compact support. Then from (2.110)
X
157.461 (2.112) ψu − v = vN + µ0α (x)(xα u)µ0α (x), vN ∈ ISm−N −1 (Rn ; {0}).
|α|=N +1
10. Action of Ψ∗ on I ∗
157.464 Proposition 2.7. Pseudodifferential operators act on conormal distributions
giving a bilinear map
0 0
Ψm (Rn ) × ISm (Rn ; {0}) −→ ISm+m (Rn ; {0}) with
157.465 (2.114) X 1
σm+m0 (Au) ∼ Dxα ∂ξα (aσ(u)), A = QL (a)
α
α!
To take care of the ‘error terms’ below we need first consider a coarser mapping
property; here there is no need for precision as regards orders and regularity.
11. RADIAL COMPACTIFICATION AND SYMBOLS 41
mapping into the upper half-sphere (see picture below); the last equation
defines Rn . I take this as the definition of the radial compactification;
show that the embedding is given explicitly by
!
x 1
P1.5 (2.124) ι(x) = p ,p
1 + |x|2 1 + |x|2
1
and deduce that t = (1+|x|2 )− 2 ∈ C ∞ (Sn+ ) is a boundary defining function
(vanishes only on the boundary and has differential non-zero there).
(2) Derive the Taylor series of a ∈ Scl0
(Rn ) = C ∞ (Sn+ ) (this is the definition
of the space of classical symbols of order 0 from lectures) in the form
∞
X x √
P1.6 (2.125) |x|−k ak ( ), ak ∈ C ∞ (Sn−1 ), |x| > 2(⇐⇒ t < 2).
|x|
k=0
(9) Show that the space S 0 (Rn ) of ‘symbols with bounds’ is identified with
the space
P1.12 (2.132) {u ∈ L∞ (Rn ); V1 . . . VN u ∈ L∞ (Rn ) ∀ Vi ∈ Vb (Rn ) ∀ N }.
[Don’t get hung up on worrying about distributions on a manifold with
boundary, we will come back to this later.]
(10) Putting some of these things together show that
0
P1.13 (2.133) Scl (Rn ) ⊂ S 0 (Rn ).
(11) Deduce that an element a ∈ S 0 (Rn ) is in Scl
0
(Rn ) if and only if
X ξ
P1.15 (2.134) a∼ (1 − φ)(ξ)|ξ|−k ak ( )
|ξ|
k
where φ ∈ C c∞ (Rn ) is equal to one near 0 (to make everything smooth)
and the ak ∈ C ∞ (Sn−1 ).
CHAPTER 3
45
46 3. THE RING Ψ∗ (Rn )
So the only issue is the constant factor, but this does not affect the fact that
n n
F (L∗ u) ∈ S m− 4 (Rn ) if û ∈ S m− 4 (Rn ).
157.160
Now, this allows us to simplify the general case in Proposition 3.1. Namely we
can write F = LG where L is the Jacobian matrix of F at 0 and G : Ω −→ L−1 Ω0
still has G(0) = 0 and now has Jacobian equal to the identity at 0. Replacing G by
F again we can therefore assume that
X
157.165 (3.5) F (x) = x + xi xj Gij (x) in |x| < , Gij ∈ C ∞ .
ij
so that
d
157.170 (3.10) Vt ut + ut ∈ C ∞ ([0, 1]; C c∞ (Ω00 ));
dt
here Ω00 is some suitably 157.170
small open neighbourhood of 0.
The idea is to solve (3.10) by successive steps and the crucial point here is that
157.171 (3.11) Vt : Icm (Ω00 ; {0}) −→ Icm−1 (Ω00 ; {0}) ∀ m ∈ R.
157.156
This follows from Proposition 2.6 and the fact that
157.172 (3.12)
×xi : Icm (Ω00 ; {0}) −→ Icm−1 (Ω00 ; {0}), ∂xk : Icm (Ω00 ; {0}) −→ Icm+1 (Ω00 ; {0}).
2. LEFT/RIGHT INVARIANCE 47
Take RBM:Clarify
Z 1
157.174 (3.14) v0 = u, vj = Vt vj−1 , j ≥ 1.
t
Then, slightly generalizing the asymptotic summation result (see Problems 157.170
157.173
2) to
include the ‘parameter’ t ∈ [0, 1], we can find ut satisfying (3.13) and hence (3.10).
However, we do know that C c∞ is coordinate-invariant so we have proved the Propo-
sition.
The question arises as to what the full symbol of F ∗ u might be. The answer is
that it is not so simple to write out because of the iteration. We will deduce a few
things about this complicated formula below, but for the moment notice that the
‘prinicpal symbol’ is given by a relatively simple formula in terms of the Jacobian.
157.175 (3.15)
n n
σm (F ∗ u)(ξ) = | det L|−1 σm (u)((L−1 )t ξ) ∈ S m− 4 /S m−1− 4 (Rn ), L = D0 F.
157.175
157.176 Lemma 3.2. The transformation (3.15) is that of a density on T0∗ Rn .
2. Left/right invariance
For differential operators it is conventional to write the coefficients ‘on the left’
X X
157.178 (3.16) P = pα (x)Dxα , pL (x, ξ) = pα (x)ξ α .
|α|≤m |α|≤m
Here of course only a finite number of terms are non-zero. The formal power series
here are those of an exponential so we can write
157.182 (3.19) pL = exp(Dx · ∂x )pR , pR = exp(−Dx · ∂x )pL
to see that one is the inverse of the other.
Proof. Leibniz’ formula.
For pseudodifferential operators we can do ‘the same thing’ but it is then not
so clear that we get the same space of operators. For a differential operator with
coefficients written on the right we see, again using the Fourier inversion formula
on S that the operator is given by the formula
Z
157.183 (3.20) F (P u)(ξ) = e−iy·ξ pR (y, ξ)u(y)dy, u ∈ S (Rn ).
Rn
48 3. THE RING Ψ∗ (Rn )
∞
157.183
157.184 Proposition 3.2. If pR ∈ C ∞ (Rny ; S m (Rnξ )) the operator defined by (3.20) is
m n
157.181
an element of Ψ (R ) with left-reduced symbol, pL , given asymptotically by (3.18).
Proof. We proceed very much as in the proof above. Namely, the Schwartz
157.183
kernel of the operator P in (3.20) is
Z
−n
157.185 (3.21) P (x, y) = B(y, x − y), B(y, z) = (2π) pR (y, ξ)ei(x−y)·ξ dξ
Rn
where we may assume that pR is of very low order to ensure absolute convergence
of the integral (and sort the general case out by continuity). Thus the kernel is
given by introducing the coordinates y and z = x − y in R2n and taking the partial
inverse Fourier transform in z.
So this is very similar to the original ‘left-reduced’ formula except we have
switched x and y as the variable independent of z = x − y on R2n . Using the same
idea as above we can consider a 1-parameter family of ‘quantization maps’ including
RBM:Check QL is earlier left and right as extreme cases
Z
157.186 (3.22) Qt (at ) = (2π)−n at (tx + (1 − t)y, ξ)ei(x−y)·ξ dξ, t ∈ [0, 1]
Rn
and again allow a to vary smoothly with t. The full estimates we are considering
on a are therefore
157.187 (3.23) sup(1 + |ξ|)−m+|β| |∂tk ∂xα ∂ξ a(t, x, ξ)| < ∞ on [0, 1]t × Rnx × Rnξ .
So again, the claim is that the space of kernels, distributions that is, on R2n defined
157.186
by (3.22) is actually independent of t.
To see this we compute, as before, the derivative in t and note that it can be
written
d X d
157.188 (3.24) Qt (at ) = Qt (i ∂ξj ∂xj at + at )
dt j
dt
where the first term comes from the chain rule and integration by parts since
d
dt (tx + (1 − t)y) = x − y and x − y = −i∂ξ i(x − y) · ξ. So, now we want to choose
at so that
X d
157.189 (3.25) i ∂ξj ∂xj at + at is of order − ∞.
j
dt
157.189
In this case we can solve (3.25) explicitly by taking
X tk
157.192 (3.26) at (x, ξ) ∼ (Dx · ∂ξ )k a(x, ξ) = exp(tDx · ∂ξ )a
j
k!
in the sense of formal power series at t = 0.
If we choose at to be an asymptotic sum (uniform in the other variables) as in
157.192
(3.26) then the ‘error term’ is
d
157.193 (3.27) Qt (at ) = Qt (et (x, ξ)), sup(1 + |ξ|)−N |∂tk ∂xα ∂ξβ et | < ∞ ∀ k, α, β.
dt
So we can unload the last step in the proof on the following lemma.
L5-end
157.186
157.194 Lemma 3.4. For each t ∈ [0, 1] the quantization Qt , in (3.22) applied to the
residual symbols, which satisfy
157.195 (3.28) sup(1 + |ξ|)N |∂xα ∂ξβ a(x, ξ)| < ∞ ∀ N, α, β
gives the space of kernel of elements of Ψ−∞ (Rn ) are precisely those smooth func-
tions which satisfy
157.196 (3.29) sup(1 + |x − y|)N |∂xα ∂yβ A(x, y)| < ∞ ∀ N, α, β.
157.195
Proof. The functions satisfying (3.28) are exactly the elements of the space
∞
C∞ (Rnx ; S(Rnξ )). Consider the case t = 1, which is the ‘left quantization’ we started
with. Since the Fourier transform is an isomorphism of S (Rn ), the space of kernels
of elements of Ψ−∞ (Rn ) consists of the functions
157.198 (3.30) A(x, y) = B(x, x − y),
where B(x, z) is the (partial) inverse Fourier transform ξ −→ z
Z
157.199 (3.31) B(x, z) = (2π)−n eiz·ξ a(x, ξ)dξ =⇒ B ∈ C ∞ ∞
(Rnx ; S (Rnz )).
Thus, after this change of variable, the space of B’s satisfy the same estimates as
the symbols they are defined by
157.201 (3.32) sup(1 + |z|)N |∂xα ∂zβ B| < ∞.
x,z
Now we choose at as an asymptotic sum of the vj ’s. This goes beyond the earlier
summation because of the presence of the parmeters t ∈ [0, 1] and x ∈ Rn . What
we want to do is to ensure that the cutoff series
X
157.207 (3.39) Φn k v j
j>k
∞
should converge absolutely with respect to the seminorms of C ∞ ([0, 1]×Rn ; S mj (Rn )).
All the terms have lower order than this. The point is that there are still only a
countable number of norms, even though they now involve the supremum over
157.207
[0, 1] × Rn as well. So absolute convergence can be ensured for each series (3.39) by
choosing the nk large enough. Again this only involves a finite number of conditions
on each nk . 157.193
Once we choose at to be such an asymptotic sum then we get (3.27) and.
following the157.184
discussion of the residual terms above, the complete the proof of
Proposition 3.2.
We can assume that the symbols themselves are of order −∞ and use density. The
composite is then
Z
157.209 (3.42) (AB)v(x) = (2π)−n ei(x−y)·ξ aL (x, ξ)bR (y, ξ)v(y)dydξ
This is almost what we want, except the ‘amplitude’ in the integral depends ex-
plicitly on both x and y, as well as ξ. So we need to show that the kernel of the
composite
Z
157.210 (3.43) K(x, y) = (2π)−n ei(x−y)·ξ aL (x, ξ)bR (y, ξ)dξ ∈ Ψm+m (Rn ).
This can be proved by an argument very similar to the left/right reduction. I leave
the details to you!
2. LEFT/RIGHT INVARIANCE 51
Let’s return for a moment to the spaces of conormal distributions at the origin,
ISm (Rn ; {0}). We can easily define conormal distributions at another point simply
by translation. Thus, if p ∈ Rn ,
157.211 (3.44) ISm (Rn ; {p}) = {u ∈ S 0 (Rn ); u(x + p) = T−p
∗
u ∈ ISm (Rn )}.
This is made more convincing by the proof of coordinate invariance. Here Tq is
translation by q ∈ Rn , Tq x = x + q.
157.212 Exercise 6. Define, and the formulate (and prove) the coordinate-invariance
of, the spaces Icm (Ω, {p}) for p ∈ Ω ⊂ Rn open.
157.213 Lemma 3.5. The Schwartz kernels of elements of Ψm (Rn ) may be identified
with the space
∞ m+ n
157.214 (3.45) C∞ (Rny ; IS 4
(Rn ; {0})) by A(x, y) 7−→ A(x − y, y).
Proof.
Using earlier results we have another method. What we have shown above, in
left/right reduction is that the kernel of an element of Ψm (Rn )
If there is a little time left today I want to introduce another algebra of pseu-
dodifferential operators. This is a sign of things to come. I have been rather hard
∞
on the ‘coefficient ring’ C ∞ (Rn ) which is involved in the ring Ψm (Rn ). What is a
‘nicer’ possibility? The one I have in mind is the symbol space itself. We can easily
introduce the space of ‘symbol-valued symbols’ (in either direction)
These defining conditions give seminorms. Here there are two orders and differ-
entiation with respect to x lowers the second (but not the first) and differentiation
with respect to ξ lowers the first but not the second.
157.215
Directly from (3.46) we see that
∞
157.216 (3.48) k ≤ 0 =⇒ S m,k (Rn ; Rn ) ⊂ C ∞ (Rn ; S m (Rn )).
It is also the case that
(1 + |x|2 )k/2 ∈ S 0,k (Rn ; Rn ) and
157.217 (3.49) 0 0 0 0
S m,k (Rn ; Rn ) · S m ,k (Rn ; Rn ) = S m+m ,k+k (Rn ; Rn ).
Combining these two observations we see that
157.218 (3.50) (1 + |x|2 )−k/2 S m,k (Rn ; Rn ) ⊂ C ∞
∞
(Rn ; S m (Rn ).
52 3. THE RING Ψ∗ (Rn )
So, we can quantize these double symbols using left quantization and define
where we cut out the region where both |x| and |ξ| are less than nj . So
on the support of (1 − φ(x/nj )φ(ξ/nj )) either |x| > nj or |ξ| > nj but
the symbol lies in the space S m−j,k−j (Rn ; Rn ) with j > l. So this is small
in the symbol space S m−l,k−l (Rn ; Rn ) if we choose nj large enough. This
means we can make all the series converge by an appropriate choice of the
integers nj and then the error term and then the error term
d
Qt (at ) ∈ S (R2n ) 157.227 (3.58)
dt
is a residual operator in the new sense. So we win and we see that Qt (S m,k (Rn ; Rn ) =
Ψm,k n
sc (R ) for all t ∈ [0, 1].
3. ISOTROPIC ALGEBRA 53
sup(1 + |x|)−k+α
(1 + |y|) −k+|γ|
(1 + |ξ|) −m+|β|
|∂xα ∂yγ ∂ξβ b(x, y, ξ)| 157.229
< ∞. (3.60)
This defines a countably normed space. Notice that these are symbols ‘sep-
arately’ in all the variables, there is no joint decay.
Proof. We can think of the double symbols a(x, ξ)157.231 as special cases
of the triple symbols which are indpendent of y. Then (3.61) is standard
quantization, so the range certainly contains Ψm,k n
sc (R ). To see that it con-
tains nothing more, we can use the same deformation argument as above
and try to construct a family of triple symbols bt so that the intermediate
quantization maps
Z
Qt : bt −→ ei(x−y)·ξ bt (x, tx + (1 − t)y, ξ)dξ 157.232 (3.62)
Rn
3. Isotropic algebra
S.Isotropic MR2650633
There is a text on this by Parmeggiani [7].
157.468 Remark. Edited by Paige Dote.
∗,∗
The calculus Ψsc (Rn ), due to Shubin, is described above. Let me introduce
yet another algebra of pseudodifferential operators on Rn . This one, as we will see
later, has direct topological applications, whereas the scattering algebra is more of
geometric significance. The symbols considered are the ‘pure symbols’ on R2n =
Rnx × Rnξ ,
S m (R2n ) = {a ∈ C ∞ (R2n ); ∀α, β ∈ Nn0 , sup(1 + |x| + |ξ|)−m+|α|+|β| |∂xα ∂ξβ a| < ∞}.
So there is no difference in behavior between the x and ξ variables. That is what
‘isotropic’ is supposed to indicate here, meaning ‘the same in all directions’.
Rather than repeat the basic constructions again we can use the properties of
∗,∗
Ψsc (Rn ) in view of the following Lemma:
157.471 Lemma 3.6. The pure symbols and symbol-valued symbols are related by the
following:
m 2n m n m n
S (R ) ⊂ S (Rx ; S (Rξ )),
m≥0
m 2n m/2 n m/2 n
157.470 (3.63) S (R ) ⊂ S (R ; S (R )), m ≤ 0 .
−∞ 2n
−∞ n −∞
S (R ) = S (R ; S (Rn ))
157.470
This gives the leading estimates in the first two statements in (3.63) since a ∈
S m (R2n ) satisfies
|a(x, ξ)| ≤ (1 + |x|)m (1 + |ξ|)m , m ≥ 0
|a(x, ξ)| ≤ (1 + |x|)m/2 (1 + |ξ|)m/2 , m ≥ 0.
In S m (R2n ) an x-derivative corresponds to an extra decay factor of (1+|x|+|ξ|)−1 ≤
(1 + |x|)−1 and a ξ-derivative does as well but (1 + |x| + |ξ|)−1 ≤ (1 + |ξ|)−1 , giving
all of the ‘double’ symbol estimates. The last statement follows from the second
last.
Thus, using left quantization we can define
(3.64) Ψm n m 2n
iso (R ) = QL (S (R(x,ξ) ))
Since these are lectures, one can ask rather impertinent questions such as: Why
would one be interested in this? One practical answer is that the algebra is closely
related to the harmonic oscillator on Rn .
Definition 3.1 (Harmonic Oscillator). The harmonic oscillator, H, on Rn is
defined as X X
H = ∆ + |x|2 , where ∆ = Dx2i = − ∂x2i .
Indeed, H ∈ Ψ2iso (Rn ) is elliptic. I hope that somewhere below I will show how
one can use the isotropic algebra to prove Thom isomorphism in K-theory. This
is a very special case of the familiar Atiyah-Singer Index Theorem that I want to
describe below. The isotropic algebra is also an integral part of the proof, at least
for the proof I have in mind.
Now, for elliptic operators, such as H, in the isotropic calculus we can deduce
the existence of a parameterix just as is done microlocally in the standard case in
the next chapter and globally for manifolds later. It is significant that the error
here is compact.
157.475 Proposition 3.5. If A ∈ Ψm n
iso (R ) is elliptic then it has a two-sided parame-
−∞ n
terix modulo the ideal ΨS (R ).
Proof. By definition, A = QL (a) where a ∈ S m (R2n ) is globally elliptic. Thus
for some δ > 0
1
|a(x, ξ)| ≥ δ(|x| + |ξ|)m in |x| + |ξ| ≥ ,
δ
1−ϕ(x,ξ) −m 2n ∞ 2n
then b(x, ξ) = a(x,ξ) ∈ S (R ) if ϕ ∈ Cc (R ) is equal to 1 on the ball
|x| + |ξ| ≤ 1δ .
4. L2 boundedness
Project 2 P-Bounded
In this project I would like you to go through some ‘symbolic arguments’, giving
L2 boundedness of pseudodifferential operators.
4.1. Schur’s criterion. This is the same Schur as the lemma about irre-
ducibility, hence I just say ‘criterion’. This is quite a handy sufficient condition for
L2 boundedness in terms of the Schwartz kernel. It can be generalized to measure
spaces (and so manifolds), but for the moment let’s think about Rn . Then
Schur Proposition 3.6. If A : Rn −→ C is a Lebesgue measurable function which
satisfies
Z Z
P2.2 (3.68) sup |A(x, y)|dy, sup |A(x, y)|dx < ∞
x y
Proof. You might like to look it up, it is basically just a clever use of Schwarz
inequality.
Problem 2.1
Show that if A ∈ Ψm (Rn ) with m < −n then the Schwartz kernel is continuous
and satisfies
P2.3 (3.70) sup(1 + |x − y|)N |A(x, y)| < ∞ ∀ N.
x,y
give the same space of operators. For t = 1 this is ‘left’ quantization and
for t = 0 it is ‘right’ quantization where the kernel is written as
∞ m+ n
157.239 (4.8) A(x, y) = B(y, x − y), B ∈ C ∞ (Rny ; IS 4
(Rn )).
• Note that the case t = 21 is also of importance. It is called ‘Weyl quanti-
zation’ and means writing the kernel as
x+y ∞ m+ n
157.240 (4.9) A(x, y) = B( , x − y), B ∈ C ∞ (Rny ; IS 4 (Rn )).
2
It has some useful properties.
• The inverses of these quantization maps are the ‘total symbols’
∞
157.241 (4.10) σL , σR , σW : Ψm (Rn ) −→ C ∞ (Rny ; S m (Rn )).
• The right and left symbols are related asymptotically by
X 1
157.244 (4.11) σR (A) ∼ Dxα ∂ξα σL (A) = exp(Dx · ∂ξ )σL (A)
α
α!
where the exponential is to be formally expanded in Taylor series at 0.
157.248 Exercise 7. Derive a similar asymptotic relationships between σW
and σL .
• (Not discussed in lecture) The ‘formal’ (just meaning non-Hilbert space)
adjoint is defined for any continuous linear operator A : S (Rn ) −→ S 0 (Rn )
by duality
Z Z
157.242 (4.12) A∗ : S (Rn ) −→ S 0 (Rn ), (Au)vdx = A∗ vdx
Rn Rn
(where the distribution pairing is written as an integral). Then
∗
157.243 (4.13) : Ψm (Rn ) −→ Ψm (Rn ), σR (A∗ ) = σL (A)).
• The composition theorem with
X
157.246 (4.14) σL (A ◦ B) ∼ (∂ξα σL (A))(Dxα σL (B)).
α
As suggested in Lecture, check this for differential operators. In fact it is
enough
157.246
to take A = Dxγ and B = b(x) and apply Leibniz’ formula to get
(4.14).
• (Also not discussed at all). The elements of Ψm (Rn ) define by bounded
linear maps for any M on the standard Sobolev spaces
157.247 (4.15) A : H M (Rn ) −→ H M −m (Rn ).
Proof later. Here H M (Rn ) = F −1 ((1 + |ξ|)−M L2 (Rn )) is defined as usual
as the inverse Fourier transform of the weighted L2 spaces on the dual.
1. Ellipticity of symbols
S.EllipticSymbols
In the notes above the notion of ellipticity for elements of S m (Rn ) is discussed
(although I did not cover this in lectures). We want an extension of this idea
to C ∞ (RN ; S m (Rn )) (I have dropped both the boundedness assumptions on the
coefficents and the assumption that N = n since they are both irrelevant here).
¯ ∈ RN × (Rn \ {0}) if
Most importantly a symbol is said to be elliptic at (x̄, ξ)
‘it is as big as it can be’ in a cone around this point.
1. ELLIPTICITY OF SYMBOLS 61
Indeed to arrange this we just need to choose a smooth function on the sphere, ψδ
ξ ξ̄ ξ
which is equal to one in | |ξ| − ω| < δ/2, where ω = |ξ̄| , is positive on | |ξ| − ω| < δ/2
ξ
and has support in | |ξ| − ω| ≤ δ in terms of the distance on the sphere. Similarly
∞
choose ψ ∈ C c (R ) positive on |x| < 1, equal to 1 in |x| < 21 and supported in
N
Here WF0 (A) is called the ‘operator wavefront set’ of A for reasons that should
become clearer below – it is just a name.
Recall that we defined the principal sybmol of A to be σm (A) = [a] to be the
∞ ∞
equivalence class in C ∞ (Rn ; S m (Rn ))/C ∞ (Rn ; S m−1 (Rn )) and then it is also equal
to the equivalence class of the right-reduced symbol.
157.272 Lemma 4.3. The elliptic set only depends on σm (A) and WF0 (A) depends on
a modulo symbols of order −∞ and is also equal to the cone-support of the right
reduced symbol; for the product of operators
157.273 (4.36) Ellm+m0 (AB) = Ellm (A) ∩ Ellm0 (B), WF0 (AB) ⊂ WF0 (A) ∩ WF0 (B).
157.266
Proof. The first part follows directly from (4.31) and the fact that left- and
right-reduced symbols differ by a term of order m − 1. The last part is a little
more subtle, and depends on the formula for the asymptotic expansion of the right-
reduced sybmol in terms of the left-reduced symbol a.
X 1
157.274 (4.37) Dxα ∂ξα a.
α
α!
157.254 157.274
If a is rapidly decreasing in a truncated cone, as (4.20) then all the terms in (4.37)
are rapidly decaying in the same cone, because of the locality of differential oper-
ators. 157.273
It follows that any asymptotic sum is rapidly decreasing as well. The final
part (4.36) follows similarly.
Perhaps the most important construction associated to these definitions is ‘mi-
crolocal invertibility’ at elliptic points.
157.277 ¯ ∈ Ell(A), A ∈ Ψm (Rn ), there exists B ∈ Ψ−m (Rn )
Proposition 4.1. If (x̄, ξ)
such that
157.278 (4.38) ¯ ∈
(x̄, ξ) / WF0 (Id −AB) ∩ WF0 (Id −BA).
Proof. As the notation suggests, we start with
157.257
157.279 (4.39) B0 = QL (b), b as in Lemma4.2.
The properties of b mean that
ξ ξ¯
157.281 (4.40) WF0 (B0 ) ⊂ Cδ = {(x, ξ) ∈ Rn ξ(Rn \ {0}); |x − x̄| ≤ δ, k
− ¯ ≤ δ}
|ξ| |ξ|
where we are free to choose δ > 0. Then, from the product formula for symbols,
157.280 (4.41) B0 A = QL (cδ )) − E, E ∈ Ψ−1 (Rn ), WF0 (E) ⊂ Cδ .
Now, we can almost invert Id −E using the Neumann series. That is we can
choose
X
157.282 (4.42) F ∈ Ψ−1 (Rn ), F ∼ E k =⇒
k≥1
157.288 (4.49) ¯ ∈
(x̄, ξ) / WF(u) ⇐⇒
ξ¯
∃ φ ∈ C c∞ (Rn ), φ(x̄) 6= 0, ψ ∈ C c∞ (Sn−1 ), ψ( ¯ ) 6= 0 s.t.
|ξ|
ξ
|ψ( )F (φu)| ≤ CN (1 + |ξ|)−N ∀ N.
|ξ|
Proof. One way is straightforward, the other way depends on the construction
of microlocal invserses as above. 157.288
First, assume the right side of (4.49) holds – for some φ and ψ as indicated.
Then we can choose another cut-off χ ∈ C c∞ (Rn ) around zero in Rn and conclude
that
ξ
157.289 (4.50) (1 − χ(ξ))ψ( )F (φu)
|ξ|
3. WAVEFRONT SET OF A DISTRIBUTION 65
We have used the fact that A ∈ Ψm (Rn ) and WF0 (A) = ∅ =⇒ Au ∈ S (Rn )
if u ∈ Cc−∞ (Rn ). Note that it is not true that WF0 (u) = ∅ and u ∈ S 0 (Rn ) =⇒
AU ∈ S (Rn ). This is one of the ‘defects’ of the algebra Ψ∗ (Rn ).
Proof. The first statement is equivalent to saying that
(x, ξ) ∈ Ellm (A), (x, ξ) ∈
/ WF(Au) =⇒ (x, ξ) ∈
/ WF(u).
Again, we use the construction in Proposition ?? to find B ∈ Ψ−m (Rn ) with
/ WF0 (Id −BA). We can choose the cone-support of the symbol of B such
(x, ξ) ∈
that B(Au) ∈ S (Rn )m abd then it follows that (x, ξ) ∈
/ WF(u).
So far I have limited the definition of WF(u) to elements of C c−∞ (Rn ). This
is only because Ψm (Rn ) does not act on general distributions in C −∞ (Rn ). To
overcome this, we consider a smaller algebra consisting of the properly-supported
pseudodifferential operators.
A closed set S ⊂ R2n can be considered as a relation between subsets of Rn
−1
S ◦ U = πL (S ∩ πR (U ))
66 4. ELLIPTICITY AND WAVEFRONT SET
vanishes if A(x, y)u(y) vanishes (??). If A has proper support S = supp(A), then
ψ ∈ C c∞ (Rn ) with supp(ψ) ∩ (S · supp(u)) = ∅ satisfies
ψ(x)A(x, y)u(y) = 0 =⇒ supp(ψ) ∩ supp(Au) = ∅.
Thus, supp(Au) is compact if
A : C c∞ (Rn ) −→ C c∞ (Rn ).
We have assumed the same for the adjoint (and hence the transpose) from which
the remaining properties follow by duality. Then,
∞
WF(u) 63 (x, ξ) if ∃ A ∈ Ψm n n
p (R ), (x, ξ) ∈ Ellm (A), Au ∈ C (R ).
cone around ξ with ϕξ (x) 6= 0. In fact, we know from Proposition that this remains
true if we replace ϕξ by ϕϕξ for a fixed ϕ ∈ C c∞ (Rn ) around ξ with ϕ(x) 6= 0. The
cone of rapid decay does not decrease.
Now Sn−1 is compact and the cones of rapid decay of the ϕ dξ u correspond to an
open cone of S2n−1 . This has a finite subcover and it follows that we may choose
ϕ ∈ C c∞ (Rn ) with ϕ(x) 6= 0 and supp(ϕ) ⊂ {ϕξj = 1} for this finite cover. Then
ϕϕξj = ϕ and ϕu c is rapidly decreasing in all directions of Rn , so ϕu ∈ C c∞ (Rn ) and
hence x ∈ / singsupp u.
Remark 3. The scatting and157.494
isotropic algebras do have properly supported
subalgebras, but the analogue of (4.55) is not valid in these cases. We will see why
later.
Recall that we have insisted that WF(u) is a closed cone in Rn × (Rn \ {0}).
The results above show that this condition in in R2n is
WF(u) = (singsupp(u) × {0}) ∪ WF(u),
which gives a nice picture! L6-end
CHAPTER 5
Propagation of singularities
L7
Microlocal ellipticity, as discussed above, shows us that if u ∈ S 0 (Rn ) and
A ∈ Ψm (Rn ) then
157.297 (5.1) WF(u) ⊂ WF(Au) ∪ Char(A).
∞
In particular if Au ∈ C (Rn ) then WF(u) ⊂ Char(A).
To see what else we might be able to say, consider a simple case, where A = D1
is differention with respect to the first variable. Letting p = ξ1 be the principal
symbol we see that
157.298 (5.2) Char(D1 ) = {(x, ξ) ∈ Rn × (Rn \ {0}); ξ1 = 0}
is a smooth hypersurface. Any solution
157.299 (5.3) D1 u = 0 =⇒ u(x, x0 ) = v(x0 )
is independent of x1 , meaning that
Z
157.300 (5.4) u(φ) = v(ψ), ψ(x0 ) = φ(x1 , x0 )dx1 .
1. Hamiltonian mechanics
157.301
As we shall see, something like (5.5) can be proved much more generally. To
do so we use commutator methods. The idea here is that we try to get information
about solutions of P u = 0 where P ∈ Ψm (Rn ) by taking a commutator with a
‘test operator’ B ∈ Ψk (Rn ). I will explain this a bit more fully below, but for the
moment just recall that
69
70 5. PROPAGATION OF SINGULARITIES
where these are now homogeneous functions of orders m and k and we just ignore
157.302
the singularities at ξ = 0, meaing we work on Rn × (Rn \ {0}. Then (5.6) gives the
homogeneous principal symbol of [P, B].
Thw important point is that we can ‘recognize’ the formula, it is the Poisson
bracket of the symbols (ignoring the i). As you will know this comes from the
symplectic form on Rn × Rn = T ∗ Rn
X
157.304 (5.8) ω= dξi ∧ dxi
i
2. Hörmander’s Theorem
∗
157.308 Theorem 5.1. If P ∈ Ψm n
cl (R ) has real principal symbol then for any u ∈
0 n
S (R )
Check that the characteristic variety at each point is the cone τ 2 = |ξ|2 (with the
obvious notation) and that the integral curves of Hp within Char(P ) project into
Rn to light rays.
Restated the claim of the Theorem is that a maximally extended integral curve
of Hp in the open set (Rn × (Rn \ {0})) \ WF(P u) on which p vanishes either does
not meet WF(u) or lies completely within it.
3. THE HAMILTON VECTOR FIELD 71
157.327 (5.29) bδ,m,N (τ, y, Ξ) = (1 − Φ(Ξ))2 Ξ2m Φ2 (Ξ/N )bδ (τ, y),
1
where 0 ≤ Φ ∈ C c∞ (R), Φ(s) = 1 in |s| < , Φ(s) = 0 in |s| > 1 =⇒
2
Hp bδ,m,N = a2δ,m,N + eδ,m,N ,
aδ,M,n = (1 − Φ(Ξ))Ξm Φ(Ξ/N )aδ , eδ,m,N = (1 − Φ(Ξ))2 Ξ2m Φ2 (Ξ/N )eδ .
In fact we need to go a step further. Namely rather than the vector field Hp
what actually arises below is the operator with a zeroth order term
157.328 (5.30) Hp + f, f smooth, real-valued and homogenous of degree 0.
Then we replace bδ by
Z τ
157.329 (5.31) bδ (τ, y, Ξ) = ψ(τ )2 e−q(τ,y) eq(s,y) a2δ (s, y)ds
−δ
Z τ
q(τ, y) = f (s, y)ds.
0
The support properties are unchanged and now
157.330 (5.32) (Hp + f )bδ,m,N (τ, y, Ξ) = a2δ,m,N + eδ,m,N .
5. Proof of regularity
After all this preparation, let me add a few words about L2 boundedness –
which is in Problem set 2 – and uniformity before passing to the actual proof of
the statement above..
First uniformity. In the proof below, regularity for Au where A ∈ Ψm (Rn ) is ob-
tained by looking at an approximating sequence AN ∈ Ψ−∞ (Rn ) which is bounded
in Ψm (Rn ) and converges to A in Ψm+ (Rn ) for any > 0. This is constructed usual
sort of cutoff. Since WF0 (AN ) = ∅ for finite N we reserve the notation for the uni-
form operator wavefront set defined as previously in terms of the cone-support of
the symbol. So in this sense
157.342 (5.33) ¯ ∈
(x̄, ξ) / WF0 (A∗ ) = QL (a∗ ) ⇐⇒ (x̄, ξ)
¯ ∈/ conesupp(a∗ ) ⇐⇒
∞
cδ aN is bounded in C ∞ (Rn ; S −∞ (Rn )) as N → ∞.
157.260
Here cδ is a conic cut-off as in (4.24). It follows for instance that for two such
families
157.343 (5.34)
WF0 (A∗ ) ∩ WF0 (B∗ ) = ∅ =⇒ A∗ A0∗ ψ ∈ Ψ−∞ (Rn ) is bounded for ω ∈ C c∞ (Rn ).
So for instance A∗ A0∗ u is bounded in S (Rn ) if u ∈ C c−∞ (Rn ). Note that we do not
necessarily conclude that the product is bounded in Ψ−∞ (Rn ) since we have not
assumed any uniformity near infinity is space.
The basic result on L2 boundedness is that
157.344 (5.35) A ∈ Ψm (Rn ) : H M (Rn ) −→ H M −m (Rn ) ∀ M.
74 5. PROPAGATION OF SINGULARITIES
In the argument below we need a little more than this. Namely for sequences AN
as above bounded in Ψ2m (Rn ) if à ∈ Ψm (Rn ) is fixed and if
157.345 (5.36) WF0 (A∗ ) ⊂ Ellm (Ã) then |hψA∗ u, uiL2 | ≤ CkÃukL2 + CkukH m0 ,
ψ ∈ C c∞ (Rn ), u ∈ C c−∞ (Rn ) ∩ H m0 (Rn )
where the constants may depend on everything except N and u.
157.365 Lemma 5.2. If u ∈ C c−∞ (Rn ) and AN u is bounded in L2 (Rn ) then Au ∈
L (Rn ) and kAukL2 ≤ lim sup kAN ukL2 .
2
Proof. This follows from the fact that L2 (Rn ) is a Hilbert space. Thus the
norm boundedness of AN u implies that it has a weakly convergent subsequence
AN u ,→ v in L2 . It follows that
157.366 (5.37) hAu, φi = limhANi u, φi = hv, φi
for all φ ∈ S (Rn ) and hence Au = v as a distribution.
So, to the proof. We are in the setup discussed above, (x̄, ξ) ¯ is a non-radial
¯ ∈
characteristic point, (x̄, ξ) ¯ ∈
/ WF(P u) and exp(tHp )(x̄, ξ) / WF(u) for t ∈ (0, ) for
some > 0. We can always shrink . In the local coordinates in Rn × (Rn \ {0}) we
choose 0 < δ0 < /2 such that
157.367 (5.38) [−δ0 , ] × {|y| ≤ δ0 } ∩ WF(P u) = ∅, [/2, ] × {|y| ≤ δ0 } ∩ WF(u) = ∅.
We proceed to conclude from this that
157.368 (5.39) ¯ ∈
Am,δ u ∈ L2 (Rn ) ∀ δ < δ0 , ∀ m =⇒ A0 u ∈ C ∞ (Rn ) =⇒ (x̄, ξ) / WF(u)
¯
since A0 is elliptic at (x̄, ξ).
Set Bδ,m,N = QL (bδ,m,N (τ, y, Ξ)) which is uniformly bounded in Ψ2m (Rn ) and
∗
the formal adjoint is Bδ,m,N = Bδ,m,N + Sδ,m,N where Sδ,m,N is uniformly bounded
2m−1
in Ψ (R ). Consider the L2 inner product
n
0
157.341 (5.49) kAδ,m,N u|2L2 ≤ hEδ,m,N u, uiL2 | + C
≤ Cδ,δ0 kAδ0 ,m− 12 ukL2 H m−/ha + Cδ,δ0 kukH m0 + C
giving the inductive step.
analysing the wavefront set of the (the Schwartz kernel of) the fundamental solu-
tion or parametrix. I will likely get around to including something on this as ‘the
L9-end calculus of wavefront sets’.
CHAPTER 6
77
78 6. SMOOTHING OPERATORS AND K-THEORY
In essence, Ψ−∞
S (Rn ) is an infinite-dimensional matrix algebra. To justify this
directly we need a ‘basis’ for S (Rn ); the standard one is the Hermite basis. This
consists of the eigenfunctions for the harmonic oscillator
n
X
157.350 (6.21) H= Dx2i + |x|2 , Heκ = (n + 2|κ|)eκ , κ ∈ Nn0 .
i=1
with the composition becoming ‘matrix multiplication’. That is, the ring of ‘rapidly
decreasing infinite matrices’.
157.371 Lemma 6.1. The algebras Ψ−∞
S (Rn ), for different n, are isomorphic (although
not naturally so).
Proof. Expansion in terms of the Hermite basis reduces Ψ−∞ S (Rn ) to ‘matri-
ces’ meaning rapidly decreasing maps
X
157.484 (6.25) a : Nn0 × Nn0 −→ C, (1 + |α| + |β|)N |a(α, β)|2 < ∞.
α,β
157.369 Lemma 6.2. The elements of Ψ−∞ S (Rn ) act as elements of the trace ideal of
compact operators on L2 (Rn ) (or any Sobolev space) and form a ‘corner’ (not an
ideal) in the bounded operators in the sense that if B ∈ B (L2 (Rn ))
157.370 (6.26) A1 , A2 ∈ Ψ−∞ (Rn ) =⇒ A1 BA2 ∈ Ψ−∞ (Rn ).
Proof. Schur’s criterion (as discussed in Problem set 2) shows that these are
bounded operators and the fact that any sequence in S (Rn ) bounded with respect
to the seminorms has a convergent subsequence in L2 (Rn ) shows that they are
compact operators.
That these operators are in the trace ideal follows from the fact that the diag-
onal operator, with respect to the Hermite basis
X
157.539 (6.27) Tk 3 u 7−→ (1 + |α|)−k hu, eα ieα
α
1
is in the trace ideal if (T ∗ T ) 4 is in Hilbert-Schmidt, which follows if
1
X
157.540 (6.28) k(T ∗ T ) 4 k2HS = (1 + |α|)−k < ∞.
α
All the elements in the series are in Ψ−∞ (Rn ) but convergence is in prin-
ciple only as bounded operators – in fact the Neumann series converges in
Ψ−∞ (Rn ), i. e. in S (R2n ). To see this, expand the definition of B to see
that X
B = −A + A2 + A( (−1)k Ak )A = −A + A2 + ABA. 157.392 (6.33)
k≥1
It follows from the corner property that ABA ∈ Ψ−∞ (Rn ) and that the
series for B actually converges in this sense.
Thus, G∞ (Rn ) contains a neighbourhood N of 0 ∈ Ψ−∞ (Rn ) and hence a
neighbourhood around any point of G∞ (Rn ).
3. Odd K-theory
The group G−∞
S (Rn ) provides an entry point to ‘complex K-theory’.
157.357 Definition 6.1. The odd K-theory, of a manifold M consists of the smooth
homotopy classes of smooth maps
157.358 (6.34) K 1 (M ) = {u : M −→ G∞
S ;
u = Id on M \ K for some K b M }/smooth homotopy.
A smooth homotopy between two elements u0 and u1 is a smooth map
are involutions.
The group G−∞S (Rn ) is a Fréchet manifold even though I have not defined the
meaning – simply because it is an open subset of the Fréchet space Ψ−∞
S (Rn ). It is
much less clear that this Grassmannian is a manifold. As a subset
157.548 (6.48) Υ−∞
S (R2 ; C) ⊂ Ψ−∞
S (Rn ; C2 ) is closed.
The ‘relative index’ function R-ind clearly takes the value k on the elements βk in
157.547
(6.47).
If the dimension of the null space of P P1 is smaller than the range then a similar
argument follows with P1 replaced by P1 ⊕ Qk where Qk is the same projection but
now on the second factor. 157.547
This proves that the involution is conjugate to one of the options in (6.47) and
then it follows that the relative trace is either k or −k in the two cases.
If you are prepared to swallow a bit of homotopy theory you can see why the
Grassmannian is relevant here. Consider the component containing β0 = β∞ :
157.560 (6.63) Υ−∞ n 2 −∞ n 2
0,S (R ; C ) = {β ∈ Υ0,S (R ; C ); Tr(β − β∞ ) = 0}.
These two subgroups commute with each other so we can also write the quotient
as
Υ−∞ n 2
0,S (R ; C ) ≡
(6.66)
157.562
GS (R ; C )/{Id} × GS (R ) /G−∞
−∞ n 2 −∞ n
S (Rn ) × {Id}.
and Υ−∞ n 2
0,S (R ; C ) is a classifying space for even K-theory.
Proof. Not given here but from the long-exact sequence of Serre the homotopy
groups
157.361
can be identified from the identification of the homotopy groups of G−∞ in
(6.40). These are proved below, using some of the properties of Υ−∞ n 2
0,S (R ; C ) – but
157.565
not (6.67)!
For > 0 this does not do very much but the family certainly becomes singular
at = 0. L10-end
Said another way, an element of Ψ−∞ n
sl,S (R ) is a family of operators for ∈ (0, 1]
L11
defined by
157.378 (6.71) A() = QL (a(; x, ξ)), a ∈ C ∞ ([0, 1]; S (R2n
x,η )).
Proof. Writing the composite family in terms of the kernels −n Bi (, x, x−y
)
of the factors shows that
x−y
A1 ◦ A2 has kernel −n D(, x, ) where
y−z z−y
Z
157.489 (6.73) D(, x, t) = −n B1 (, x, t + )B2 (, z, )dz
R n
y−z
Z
= B1 (, x, t − Z)B2 (, x + Z, Z)dZ, Z = .
Rn
The integrand here is smooth in with values in the Schwartz functions in the
variables x, t, Z – as follows from the fact that
157.490 (6.74) (1 + |x| + |t − Z| + |x + Z| + |Z|) is comparable to (1 + |x| + |t| + |Z|).
Thus the integral is also Schwartz. Expanding in Taylor series at = 0 gives the
Moyal formula below.
Alternatively this follows from our earlier results on Ψ0 (Rn ). Observe that
157.379 (6.75) (1 + |ξ|) ≤ (1 + |ξ|), ∈ (0, 1)
from which it follows that
157.380 (6.76) (0, 1) 3 −→ a(, x, ξ) ∈ S (Rn ; S 0 (Rn ))
is (uniformly) bounded. Indeed the derivatives satisfy
157.381 (6.77) ∂xα ∂ξβ a(; x, ξ) = |β| ∂xα ∂ηβ a(; x, η) η=ξ
=⇒
|∂xα ∂ξβ a(; x, ξ)| ≤ Cα,β |β|
(1 + |ξ|)−|β| ≤ Cα,β (1 + |ξ|)−|β| .
So our earlier composition result shows that uniformly for > 0
157.382 (6.78) A1 () ◦ A2 () is bounded in Ψ0 (Rn )
88 6. SMOOTHING OPERATORS AND K-THEORY
Moreover the left-reduced symbol of the composite is given by the asymptotic for-
mula
X
157.383 (6.79) A1 () ◦ A2 () = QL (c()), c() ∼ ∂ξα a1 (; x, ξ)Dxα a2 (; x, ξ)
α
X
= |α| ∂ηα a1 (; x, η)Dxα a2 (; x, η) η=ξ
.
α
Using Borel’s Lemma to sum the Taylor series gives c ∈ C ∞ ([0, 1]; S (R2n )) such
that
157.384 (6.80) X
c(; x, η) − |α| ∂ηα a1 (; x, η)Dxα a2 (; x, η) ∈ N +1C ∞ ([0, 1]; S (R2n )) ∀ N.
|α|≤N
From either proof we see that again ‘Moyal’s formula’ appears (I think it is
more historically legitimate here than before!)
Ai ∈ Ψ−∞ n ∞ 2n
sl,S (R ), Ai = QL (ai (, x, ξ)), ai ∈ C ([0, 1]; S (R )) =⇒
Equality here is in the sense of formal power series (i.e. Taylor series) at = 0. So
this formula tells you nothing about what happens for > 0 as is to be expected.
However, the symbol of the product is determined by this formula up to terms
vanishing to infinte order.
157.398 Lemma 6.4. If a ∈ C ∞ ([0, 1]; S (R2n )) vanishes to infinite order at = 0 then
the semiclassical family
157.399 (6.83) A() = QL (a(, x, xξ))
is simply a smooth map [0, 1] −→ Ψ−∞ (Rn ) vanishing to infinite order at = 0.
The analogue of the ‘principal symbol map’ for pseudodifferential operators is
played by the ‘semiclassical symbol’
A() = QL (a(, x, xξ)), > 0 =⇒
157.394 (6.84)
σsl : Ψ−∞ n 2n
sl,S (R ) 3−→ a(0, y, η) ∈ S (Rx,η )
157.394
157.395 Proposition 6.5. The semiclassical symbol map (6.84) gives a short exact se-
quence of algebras
Make sure you understand what a semiclassical smoothing operator ‘looks like’.
Its kernel is of the form
x−y
157.408 (6.86) A (x, y) = −n B(, x, ), B ∈ C ∞ ([0, 1] ; S (R2n )
so as ↓ 0 the kernel is ‘squashed’ around the diagonal.
There are semiclassical operators of finite order as well, I hope I will have time
to talk a little about them. A semiclassical differential operator on Rn might be of
the form
Xn
157.409 (6.87) P = 2 Dx2j + V (x).
i=1
Notice that if you set = 0 only the zeroth order term survives. This is not the
semiclassical symbol which is instead the ‘rescaled’ symbol including the lower order
term:
157.410 (6.88) σsl (P ) = |ξ|2 + V (x).
157.409
Operators like (6.87) arose in quantum mechanices where ' ~ is the ‘coupling
constant’ relating the frequency of spectral lines to the jump in energy between
electron shells which produces them. In practice ~ ' 1/127 is small and the idea
is to think of it as ‘very small’ and perturb from = 0 (I don’t like to use ~ here
since it is actually a constant!) Simply setting = 0 is a bad idea since most of
the problem disappears. What is happening here is that the problem is becoming
commutative as ↓ 0 because of the commutation condition that
157.427a (6.89) [∂xi , xk ] = δik
so one is turning on the non-commutative product as becomes positive.
There is a lot one can do with semiclassical operators – see for instance the
MR2952218
book of Zworski [9].
6. The group G∞
sl
Observe what the form that the semiclassical symbol takes in this more general
case. It is just a symbol with values in the smoothing operators
0 0
157.406 (6.94) σsl : Ψ−∞ n −∞
sl,S (R ; ΨS (Rn )) −→ S (R2n
x,ξ ; Ψ
−∞
(Rn ))
157.396
giving a short exact sequence just like (6.85)
157.407 (6.95)
Ψ−∞ n −∞ 0
(Rn )) / Ψ−∞ (Rn ; Ψ−∞ 0
(Rn )) / S (R2n ; Ψ−∞ (Rn0 )).
sl,S (R ; ΨS sl,S S
So the product formula for the symbol in this case is multiplicative in the variables
0
(x, ξ) but with values in the non-commutative algebra Ψ−∞S (Rn ).
157.404 Proposition 6.6. The semiclassical group
0 0
−∞
157.405 (6.96) G∞ n n n −∞
sl,S (R ; S (R )) = {A ∈ Ψsl,S (R ; ΨS (Rn ));
0
∃ B ∈ Ψ−∞ n −∞
sl,S (R ; ΨS (Rn )) with (Id +A)(Id +B) = Id = (Id +B)(Id +A)}
0
is open in Ψ−∞ n −∞
sl,S (R ; ΨS (Rn )).
157.356
Proof. This is similar too, but a bit more involved than, the proof of Lemma 6.3.
0
The topology on the semiclassical algebra Ψ−∞ n
sl,S (R ; Ψ
−∞
(Rn )) comes from its
0 157.411 157.399
identification with C ∞ ([0, 1]; S (R2n+2n ) ) through either (6.93) or (6.83). From the
latter it follows that A() defines a bounded family of operators on L2 (Rn ) with
a uniform norm bound being a continuous seminorm. So if A lies in a sufficently
small neighbourhood of 0 the family, in , of operators
157.412 (6.97) (Id −A())−1 = Id +B() exists for > 0
with B() uniformly bounded as ↓ 0. So we only need to show that
0
157.413 (6.98) B() ∈ Ψ−∞ n −∞
sl,S (R ; ΨS (Rn )).
Proceeding ‘symbolically’ we can see that the model problem, the existence of
the inverse
0
157.414 (6.99) (Id +σsl,S (A))−1 (y, η) = Id +B0 (x, η) ∈ G−∞
S (R2n ),
has a unique solution,157.356
for A in a possibly smaller neighbourhood of 0. This indeed
follows from Lemma 6.3. Once we know that the principal symbol can be inverted,
the existence of a formal power series inverse follows from the behaviour of the Moyal
0
product. That is, we can find a sequence of elements Bk ∈ Ψ−∞ n −∞
sl,S (R ; ΨS (Rn ))
such that for any p
p
X 0
157.415 (6.100) (Id +A()(Id +B0 ()) + k Bk () ∈ Id +p+1 Ψ−∞ n −∞
sl,S (R ; ΨS (Rn )).
k=1
Using Borel’s lemma again we can sum the series and so find a ‘parametrix’ B 0 ∈
0
Ψ−∞ n −∞
sl,S (R ; ΨS (Rn )) such that
157.416 (6.101)
0
(Id +A()(Id +B 0 ()) − Id, (Id +B 0 ())(Id +A()) − Id ∈ C˙ ∞ ([0, 1]; Ψ−∞
S (Rn+n ))
(which is the same as saying a semiclassical
157.392
error vanishing to all orders at = 0).
The corner identity as in (6.33) then shows that the difference of the inverse
and the parametric B − B 0 is uniformly bounded as a function of ∈ (0, 1] with
0
values in Ψ−∞
S (Rn+n ). A similar argument for the derivatives with respect to
6. THE GROUP G∞
sl 91
shows that this difference is actually smooth and vanishes to infinite order at = 0
157.413
so proving (6.98).
157.407
For the semiclassical group the symbol sequence (6.95) gives the map
0 0
157.418 (6.102) σsl : G∞ n −∞
sl,S (R ; ΨS (Rn )) −→ S (R2n ; G∞ (Rn ))
which it is important to note is surjective. In fact we want a stronger lifting property
157.417 Proposition 6.7. The semiclassical group has the lifting property that for any
0
smooth symbol map u : M −→ S (R2n ; G∞ (Rn )) of compact support (reducing to
the identity outside a compact set) there is a smooth map, also of compact support
−∞ 0
ũ : M −→ G∞ n
sl,S (R ; ΨS (Rn )) giving a commutative diagramme
(6.103) G∞ n −∞
(Rn ))
0 σsl
/ S (R2n ; G∞ (Rn0 ))
157.419 sl,S (R ; ΨS O
i
u
ũ
M.
Moreover any two such lifts are smoothly homotopic.
This is another symbolic argument, just a uniform version of the sur-
Proof. 157.418
jectivity of (6.102). RBM:Expand?
157.421
The way we are getting (6.104) can be 157.423
described as ‘turning on quantization’ by
writing R2n = Rnx ×Rnξ . Looking again at (6) we can think of R4 = R2x,ξ ×R2y,η where
we first turn on quantization in (x, ξ) for the first map, then turn on quantization
in (y, η) for the second map. So we conclude easily enough that this is the same
as doing both simultaneously in the sense that we get a commutative diagramme
forming the upper right triangle
'
K 1 (R2 × M ) / K 1 (M )
157.421
where the diagonal map is (6.104) for n = 2. Atiyah’s idea has two parts. The first
is that we get a further commutative triangle on the lower left where the roles of
the two quantizations are interchanged – the second set of variables are quantized
RBM:Expand or refer first, giving the same final result.
However, we can do a little more, namely we can rotate the variables, smoothly,
by looking at x cos θ + y sin θ and dually on the other variables. This gives us a
smooth family of double quantization maps starting at the first and finishing at
the second, with the sign reversed. By homotopy invariance these must all give the
same result.
The way we use this depends on a157.422
157.419
good understanding of the lifting map in
(6.103) as giving us a right inverse to (6.105) which we can write somewhat myste-
riously as
157.425 (6.107) K 1 (M ) 3 κ 7−→ β ⊗ κ ∈ K 1 (R2 × M )
where as above, β is the Bott element of K 0 (R2 ). 157.422
K 1 (R2 × M ) is mapped to zero in K 1 (M ) by (6.105), so
So, suppose that κ ∈157.424
along the right side of (6.106). By surjectivity κ comes from β1 ⊗ κ ∈ K 1 (R4 ×
157.424
M ) along the top line of (6.106), so lifting in the first variables. Reversing the
quantization as discussed above we see that
157.426 (6.108) κ = ±q2 (β1 ⊗ κ) = ±β1 ⊗ q2 (κ) = 0
so κ = 0 and we have injectivity.
This argument depends on seeing that quantization in one set of variables
commutes with the lifting map in a different set of variables which we will see in
L11-end the lifting construction.
L12/13
157.433
Proof. First make any choice B0 as in (6.116). Then, from the semiclassical
symbol calculus the principal semiclassical symbol of the square satisfies
157.436 (6.118) σsl ((β∞ + B0 )2 − Id) = σsl (β∞ bo − bo β∞ + b20 ) = 0.
Thus in fact
157.437 (6.119) (β∞ + B0 )2 = Id +E1 , E1 ∈ Ψ−∞ 2
sl,S (R; C ).
Proof.
K 1 (∂M ) o K 1 (M ) o K 1 (M ; ∂M )
The horizontal maps here are straightforward. Namely there is an inclusion
map
C c∞ (M \ ∂M ; G−∞ ) ,→ C c∞ (M ; G−∞ )
since the former are maps which are equal to the identity outside a compact set
and so can be extended (as the identity) up to the boundary. Similarly there is a
restriction map
|∂M
C c∞ (M ; G−∞ ) −→ C c∞ (∂M ; G−∞ )
and both these maps ‘descend’ through homotopies.
The end maps, the connecting homomorphisms, are as usual a little less obvious.
On the right we start from the group
C c∞ (R × ∂M ; G∞ ).
By a very simple version of the collar neighbourhood theorem a neighbourhood of
the boundary in M is diffeomorphic to a product
[0, 1) × ∂M ,→ M.
12. CHERN CHARACTER-BUILDING 97
We can take the radial compactification of R to [0, 1] and thereby convert the map
157.503
(11) into a map
R × ∂M −→ M.
The definition in terms of rapid decay then means that there is a pull-back map
157.505 (6.139) S (R; C ∞ (∂M ; G−∞ )) ,→ C c∞ (M ; G−∞ )
which descends to give the connection homorphism on the right.
This leaves the connecting homomorphism on the left. This is where Bott
periodicity comes in. We are starting from (the components of) C c∞ (M ; G−∞ ) on
the lower left. However, by Bott perioidicty – Theorem ?? – we can instead start
from C c∞ (R2 × M ; G−∞ ). Now, separating the factors of R2 into R × R we can apply
the argument of the preceeding paragraph, giving the connecting homomorphism
on the right, to R × M instead of M, and this gives the map on the left. There is
an orientation issue here but since it amounts to a sign the choise one makes does
not affect the main result, which is:
157.498
157.499 Theorem 6.2. The six-term sequence (6.138) is exact for any manifold with
boundary.
is the natural map from the tangent space at g to the tangent space at Id
(which is the Lie algebra).
(2) Then the formal product (g −1 dg)j is supposed to be the j-fold exterior
product, with composiition thrown in, so it is a j-multilinear, totally
antisymmetric map from j copies of the tangent space at g to the tangent
space at Id . P3.14
(3) This explains (12) in the sense that the trace functional results in each
term being a (2k + 1)-multilinear function on the tangent space at each
point, which is what a form should be.
(4) The trace property Tr([a, b]) = 0 shows that
Tr((g −1 dg)2k ) = 0
which explains why there are no even terms.
(5) The deRhan differential is easy to define on 1-forms, even with values in
an infinite dimensional space, show that
d(g −1 dg) = −g −1 dg ∧ g −1 dg
in an appropriate sense.
(6) Conclude that the terms in Ch(g) are all closed.
(7) Suppose [0, 1] 3 t 7−→ G−∞
S (Rn ) is a smooth curve, show that
d
Ch(gt ) = d Et(g)
dt
where Et is the ‘Eta’ or Chern-Simons form
∞
(−1)k−1 k!
−1 dg
X
−1 2k
Et(g) = Tr g (g dg)
(2πi)k (2k)! dt
k=0
defines a closed form in each even degree. [I am not sure I have the
normalizing constants correct, but here it does not matter.]
(12) Conclude, proceeding as above, that this defines a map
K 0 (M ) ⊗ C −→ HdR
ev
(M ).
Ψ−1 n −∞
(Rn )
0
/ Ψ0 (Rn ; Ψ−∞ 0
(Rn ) / C ∞ (S2n−1 ; Ψ−∞ 0
(Rn ).
iso (R ; ΨS iso S S
(4) This is an algebra without identity, so add Id to get a ring which we can
denote
0
Id +Ψ̇0iso (Rn ; Ψ−∞
S (Rn ).
Check that we now a multiplicative exact sequence
Id +Ψ−1 n −∞
(Rn )
0
/ Id +Ψ̇0 (Rn ; Ψ−∞ 0
(Rn ) / Id +S (R2n−1 ; Ψ−∞ 0
(Rn ).
iso (R ; ΨS iso S S
Show that Cl(R2n ), the complexified Clifford algebra (not the Clifford
algebra on the complexification!) acts on λ∗ Cn through the formula we
saw for the Hodge-Dirac operator
√ √
cl(ek + iek+n )f α = i 2fk ∧ f α , cl(ek − iek+n )f α = −i 2ι(fk )f α , k ≤ n
where the constants are for length normalization an here α is a strictly
increasing sequence in {1, . . . , n}. This specifies the action of all basis
elements and we see that
cl(ek + iek+n ) cl(ek − iek+n ) + cl(ek − iek+n ) cl(ek + iek+n ) = 2 Id,
cl(ek + iek+n )2 = 0 = cl(ek − iek+n )2
=⇒ cl(ek ) cl(ek0 ) + cl(ek0 ) cl(ek ) = 2δkk0 Id, k, k 0 = 1, . . . , 2n.
(3) If ei is an (oriented) orthonormal basis of V, let Ei be the corresponding
elements in the Clifford algebra so
Ei Ej + Ej Ei = δij Id .
Then show that the Clifford algebra has a ‘maximal element’
Z = in(2n−1) E1 E2 . . . E2n =⇒ Z 2 = Id, ZEi + Ei Z = 0.
P3.28
(4) Okay, now observe that (1) is now
P3.35 (6.140) µ(ζ) = cos(χ(|ζ|)Z + sin(χ(r))(ζ̂ · E∗ )
where E∗ = (E1 , . . . , E2n ) is thought of as a vector with values in matrices
and · is the inner product. P3.35
(5) Now observe that the definition (6.140) extends to V = R2n to give a
smooth family of involutive matrices. Here, as before χ(r) is decreasing
from π near 0 to 0 near ∞.
(6) Review the proof in the notes to check that there is a semiclassical family
of idempotents qunatizing µ (so now to smoothing operators on Rn ).
(7) For a bonus, show that the quantization is again has relative index 1
(assuming I got the signs right which would be a pleasant accident). You
might like to do this by quantizing in two variables repeatedly, so working
by induction over n.
CHAPTER 7
Operators on manifolds
C.OpMa L14
In this second half of the course I will, finally, get to the discussion of analysis
on manifolds. First what I hope is a reminder of the invariant description of kernels
of operators.
157.567 (7.2) C˙ ∞ (M ) ⊂ C ∞ (M )
⊂
C˙c∞ (M ) ⊂ C c∞ (M )
and does correspond to the Schwartz space at least in the sense that
157.568 (7.3) C˙ ∞ (Rn ) = S (Rn ).
Recall the change-of-variable formula for the integral of functions on Euclidean
space. If F : Ω0 −→ Ω is a diffeomorphism between open subsets of Rn and
f ∈ C c∞ (Ω) then
Z Z
∂F
157.569 (7.4) F ∗ f | det |dx = f (y)dy.
Ω0 ∂x Ω
So there is no invariant integral of functions on a manifold.
To integrate we need something that transforms with the absolute value of the
157.569
determinant of the Jacobian of the diffeomorphism that appears in (7.4). This
is provided by densities. Recall that on an n-manifold the maximal degree, n−,
forms at a point, Λnm M for m ∈ M, may be defined as the linear space of totally
antisymmetric multilinear maps
157.570 (7.5) Tm M × n factors × Tm M −→ C.
Equivalently if we let λn Tm M denote the totally antisymmetric part of the n-fold
∗
tensor product of Tm M with itself (so Λnm M = λn Tm M ) then Λnm M is the space
of linear maps
157.571 (7.6) µ : λn Tm M −→ C.
103
104 7. OPERATORS ON MANIFOLDS
2. Distributions
The existence of the integral on densities leads to consistent pairings
157.581 (7.13) C c∞ (M ) × C ∞ (M ; Ω)
O
? )R
C c∞ (M ) × C c∞ (M /
_ ; Ω) 3 (f, g) 5
fg ∈ C
C ∞ (M ) × C c∞ (M ; Ω)
3. VECTOR BUNDLES 105
Each of the spaces of smooth sections of Ω has a topology coming from the
isomorphism with the corresponding space of functions given by the choice of a
smooth positive section of Ω (and the fact that multiplication by a positive smooth
element f ∈ C ∞ (M ) is an isomorphism on both C c∞ (M ) and C ∞ (M )). This means
that we can define the spaces of distributions by
This is arranged so that these are ‘generalized functions’ with natural inclusions
3. Vector bundles
In fact we want to generalize this in several respects. First for sections of a
vector bundle. I am already assuming you know what a vector bundle is. Just for
completness sake let me remind you that a vector bundle (either real or complex)
over a manifold M is another manifold V (often called the total space of the vector
bundle) with a surjective smooth map
157.584 (7.16) π : V −→ M
such that each fibre Vm = π −1 (m) ⊂ V has a linear space structure which is ‘smooth
and locally trivial’. This means that each point m ∈ M has a neighbourhood Om
for which there is a diffeomorphism giving a commutative diagramme
π π1
$ z
Om
with F linear on each fibre. Here K stands for R or C in the real or complex case
respectively. One can always assume that the Om are coordinate patches on M
157.585
and then the maps (7.17) give a special atlas on V – the transition conditions are
automatic.
There are sections of any vector bundle just as for functions (and densities)
C c∞ (M ; V ) ,→ C ∞ (M ; V ) corresponding to smooth maps u : M −→ V with π ◦ u =
IdM .
The dual bundle, V 0 , of a vector bundle V is defined, as a set, as the union of
the duals of the fibres
G
157.586 (7.18) V0 = Vm0 .
m∈m
157.585
It has a smooth structure coming from the local trivializations (7.17) by replacing
the map Fm : Vm −→ KN by (F t )−1 : Vm0 −→ KN .
There is then a pointwise pairing
157.587 (7.19) C ∞ (M ; V ) × C ∞ (M ; V 0 ) −→ C ∞ (M )
106 7. OPERATORS ON MANIFOLDS
157.581
and this leads to pairings analogous to (7.13):
Z
hf, gim , f ∈ C c∞ (M ; V ), g ∈ C ∞ (M ; V 0 ⊗ Ω), or
f ∈ C ∞ (M ; V ), g ∈ C c∞ (M ; V 0 ⊗ Ω).
Then, noting that there are again appropriate topologies, the spaces of distribu-
tional sections are defined as the spaces of continuous linear maps
C c−∞ (M ; V ) = {u : C ∞ (M ; V 0 ⊗ Ω) −→ C; linear and continuous}
157.589 (7.20)
C −∞ (M ; V ) = {u : C c∞ (M ; V 0 ⊗ Ω) −→ C; linear and continuous}.
157.602 Remark 6. I will generally identify operators and their kernels, using the same
letter to denote both.
If M is not compact we also need to think again about the ‘calculus of supports’.
5. Smoothing operators
Suppose M is a compact manifold without boundary, so all the support annoy-
ances are absent.
157.597 Definition 7.1. The space of smoothing operators between sections of two
bundles, V, W over a compact manifold, M, is defined by the space of kernels
157.598 (7.27) Ψ−∞ (M ; V, W ) = C ∞ (M 2 ; πL
∗ ∗
W ⊗ πR (V 0 ⊗ Ω)).
When W = V we abbreviate the notation to Ψ−∞ (M ; V ).
157.599 Remark 7. Smoothing operators are characterized (for M compact) by the
two conditions that
A : C −∞ (M ; V ) −→ C ∞ (M ; W ) and
157.600 (7.28)
At : C −∞ (M ; W 0 ⊗ Ω) −→ C ∞ (M ; V 0 ⊗ Ω).
For M compact it is again the case that the smoothing operators on a fixed
bundle V – so with W = V – form an algebra. The composition written in terms
of kernels is
Z
157.601 (7.29) A◦B = A(·, m)B(m, ·).
M
Make sure you understand why the integral here makes sense and gives again a
smoothing operator.
157.603 Proposition∗ 7.2. The group of invertibles in the ring Id +Ψ−∞ (M ; V ) for
a vector bundle over a compact manifold is isomorphic (but not naturally so) to
G−∞
S (Rn ).
To prove this we need a decent basis for L2 (M ; V ) – we will get this as the
eigenbasis for any self-adjoint elliptic pseudodifferential operator acting on V.
157.604 Exercise 10. Give an appropriate definition for the space Υ−∞ (M ; V ⊗ C2 )
for any complex vector bundle, V, over a compact (connected) manifold M and
check that it is isomorphic to Υ−∞
S (Rn ; C2 ).
L14-end
L15
6. Cornormal distributions at the zero section
Let W be a real vector bundle (any complex vector bundle has an underlying
real bundle) over a manifold S. We proceed to introduce the space of conormal
distributions on W, the total space of the vector bundle, relative to S appearing
as the zero section of W. We will want this to be a module over S (W ) = C˙ ∞ (W )
and the elements should be singular only at the zero section. In the case that
W = S × RN we know what we want, namely
m− 4s
157.605 (7.30) ISm (S × RN ; S × {0}) = C ∞ (S; IS (RN )).
Here s = dim S and the only the shift in the order is questionable. We already
considered such a space (with restricted coefficients) in case that S = Rn and
108 7. OPERATORS ON MANIFOLDS
157.611 Proposition 7.3. Under fibre-wise inverse Fourier transform the space of Schwartz
conormal distributions with respect to the zero section of a vector bundle is globally
isomorphic to a space of symbolic densities
dim W dimfib W
157.612 (7.37) ISm (W ; 0W ) ∈ F −1 (C ∞ (S; S m+ 4 − 2 (W 0 ; ΩW 0 ))).
Proof.
157.620 Proposition 7.4. If W −→ S is a real vector bundle, U −→ S is a complex
vector bundle and
157.621 (7.38) F : D2 −→ D1
is a diffeomrophism between open neigbhourhoods of the zero section of W which
maps the zero identically to itself, then
157.622 (7.39)
F ∗ : {u ∈ I m (W, OW , U ); supp(u) b D1 } −→ {v ∈ I m (W, OW , U ); supp(u) b D2 }.
Proof. Since conormal distributions are smooth away from the zero section
we may shrink the domain D1 with D2 replaced by the image and the result remains
unchanged.
We also know that the conormal space is a module over C ∞ (S) so we can use
a partition of unity in S to localize to an open sets over each of which W is trivial.
So we may assume W = S × RN .
To see how much remains to be proved, consider the following factorization of
F – in a possibly smaller domain
157.623 (7.40) F = L ◦ F̃ .
Here G is the fibre-preserving diffeomorphism constructed from the differential of
F as follows. At each point s ∈ S the tangent space to W is the direct sum
157.624 (7.41) Ts W = TS S ⊕ RN
where the second part is the tangent space of the fibre. Since S is fixed pointwise
by F the differential maps is the identity on Ts S and so decomposes into the sum
of two linear maps
157.625 (7.42) F∗ RN 3 w 7−→ (L0 w, Lw) ∈ Ts S ⊕ Rn .
So L is a well-defined family of linear maps on RN . We already know the invariance
under such maps so it suffices to consider he remainder
157.623
term F̃ .
This defines the first, fibre-linear, factor in (7.40), and hence the second. By
construction this has differential at s ∈ S which is the sum of the identiy and the
‘off-diagonal’ term mapping Ts Ws to Ts S. This means that the map itself is of the
form, in any locall trivialization
157.626 (7.43) F̃ (s, v) = (s, w) + v · h(s) + E(s, v), E(s, v) = O(|v 2 ), v ∈ RN .
Then we can deform F̃ to the identity through the family of diffeormorphisms
157.627 (7.44) F̃ (s, v) = (s, w) + t(v · h(s) + E(s, v)), t ∈ [0, 1].
Now we can use the ‘homotopy method’ again. Very close to S this 1-parameter
family of diffeomorphism is generated by a 1-parameter family of vector fields
d ∗
157.628 (7.45) F u = Ft∗ (Vt u), Vt = vanishes at S and Vt vj = O(||v|)2 .
dt t
110 7. OPERATORS ON MANIFOLDS
9. Elliptic operators
EllipticM
Now we turn to the construction of a parametrix for a globally elliptic op-
erator A ∈ Ψm (M ; V, W ). Thus the principal symbol of A has a representative
a ∈ S m (T ∗ M ; π ∗ hom(V, W )) which is elliptic. This means there is an inverse
modulo a compactly supported error,
In the inductive step for Bk+1 ∈ Ψ−m−k−1 (M ; W, V ) to satisfy the first condition
157.639
in (7.55) we must have
157.646
Proof. Suppose ind(A) ≤ 0. By Lemma 7.4 dim (A) ≤ dim (A∗ ). Thus we
can choose a smoothing operator A0 : (A) −→ (A∗ ) which is injective. Then any
element of the null space of A+A0 , satisfies Au = −A0 u. These lie in complementary
spaces so both must vanish and hence u ∈ (A) and A0 u = 0 so u = 0. If ind(A) = 0
it follows that (A∗ ) ⊂ Ran(A + A0 ) so A + A0 is a bijection. If ind(A) ≥ 0 thus
argument applies to A∗ .
157.655 Lemma 7.6. For any real order m and any bundle V there is an elliptic and
invertible operator Qm ∈ Ψm (M ; V ) with diagonal, positive principal symbol.
Proof. Take a Riemann metric on m and consider the symbol |ξ|m Id valued
in hom(V ). Clearly this is elliptic. For the case m = 0 we can of course take the
identity. Suppose m < 0 then choose an operator Lm/2 ∈ Ψm/2 (M ; V ) with symbol
|ξ|m/2 Id . It is a compact operator on L2 (M ; V ). Taking an hermitian inner product
on V and a positive smooth density on M gives L2 (M 0 V ) an explicity Hilbert inner
product. Then the adjoint L∗m/2 ∈ Ψm/2 (M ; V ) has the same principal symbol
so Q0m = L∗m/2 Lm/2 ∈ Ψm (M ; V ) is selfadjoint with principal symbol |ξ|m Id .
It is elliptic of order 0 with (Q0m ) ⊂ C ∞ (M ; V ) finite dimensional. Choosing a
positive definite matrix on (Q0m ) and adding it to Q0m gives an invertible element
of Ψm (M ; V ) with the same principal symbol. For m > take Q−m .
The existence of such an element shows that it really suffices to consider oper-
ators of order 0 in the discussion of the index since if A ∈ Ψm (M ; V, W ) is elliptic
then so is A0 = AQ−m ∈ Ψ0 (M ; V, W ) and
157.656 (7.69) σ0 (A0 ) = σm (A)||ξ|−m .
Corresponding to the stability of the index of Fredholm operators on a fixed
Hilbert space it also follows that
157.654 (7.70) ind(A + E) = ind(A) if A ∈ Ψm (M ; V, W ) is elliptic and
0
E ∈ Ψm (M ; V, W ), m0 < m.
Indeed, the index of (A + E)Q−m = AQ−m + EQ−m is the same as that of A + E
and EQ−m is compact from L2 (M ; V ) to L2 (M ; W ) so the index is the same as
that of AQ−m and hence A.
Said a different way, we have proved that
157.657 Lemma 7.7. The index of and elliptic element A ∈ Ψm (M ; V, W ) is determined
by any reprentative of
157.658 (7.71) σm (A)|ξ|−m ∈ S 0 (T ∗ M ; π ∗ hom(V, W )).
and so defines a map
157.659 (7.72) ind : Ell(V, W ) = {a : S 0 (T ∗ M ; π ∗ hom(V, W ));
with an inverse modulo lower order terms} −→ Z
for any bundles V and W.
Of course the bundles certainly have to have the same rank before such an elliptic
symbol exists.
We can 157.659
think of the pseudodifferential algebra (modules really) as define the
index map (7.72) through ‘quantization of the symbol’.
114 7. OPERATORS ON MANIFOLDS
0
157.660 Lemma 7.8. If A1 ∈ Ψm (M ; V, W ) and A2 ∈ Ψm (M ; W, U ) are two elliptic
operators between bundles then the composite is elliptic and
157.653 (7.73) ind(A2 ◦ A1 ) = ind(A2 ) + ind(A1 ).
157.655
Proof. Using Lemma 7.6 We can replace A1 by AQ −m and A2 by Q−m0 A2
157.646
and reduce to the case that m = m0 = 0. From Lemma 7.4 we see that ind(A∗ ) =
− ind(A). So if A1 and A2 have opposite signs we can assume that ind(A1 ) ≥ 0
and ind(A2 ) ≤ 0 or else pass to the adjoint, which reverses the order. If A1 and
A2 have the same sign then we can assume that both are non-negative. So it
suffices to consider the case ind(A1 ) ≥ 0. Then, adding a finite rank surjective
smoothing operator mapping the null space of A onto a complement to its range
we can assume that A1 is surjective. Then the range of A2 ◦ A1 is the range of
A2 and its null space is inverse image of the null space of A2 under A1 , which has
157.653
dimension dim((A1 )) + dim((A2 ). Thus (7.73) holds.
157.659
157.661 Corollary 4. The index map (7.72) is additive under the product
157.662 (7.74) Ell(W, U ) ◦ Ell(V, W ) −→ Ell(V, U )
(when these spaces are non-empty).
One of the things that I set out to do was to prove that such an elliptic operator
has ‘Weyl asymptotics’. L17-end
φ πU
# |
U.
So this is just like a vector bundle except the fibres are manifolds, not vector
spaces, and correspondingly the ‘local trivializations’ F are fibre-preserving diffeo-
morphisms. Of course the most obvious case is a product
157.668 (7.78) X =Z ×Y
which is globally trivial as a fibre bundle. You are probably familiar with the Hopf
fibration
157.669 (7.79) S2n+1 −→ S2n
given by thinking of S2n+1 as the unit sphere in Cn+1 and then taking the fibres to
be given by the multiplicative action of the circle S ⊂ C.
The ‘fibre above y ∈ Y ’, meaning φ−1 (Y ), is diffeomorphic to the model fibre
Z and is often written Zy . There is in general no natural choice of the trivialization
157.667
map F in (7.77) so no natural diffeomorphism between Z and Zy . A fibration is
how we are to interpret the notion of a ‘smooth family of manifolds’.
157.670 Proposition 7.6. If φ : X −→ Y is a surjective smooth map between compact
manifolds then it defines a fibre bundle if and only if it is a submersion, i.e. F∗ :
Tx X −→ TF (x) Y is surjective for each x ∈ X.
The compactness, or at least some additional condition, is needed here. The sur-
jectivity is less of an issue, since if φ is a submersion between compact manifolds its
image if open and closed and hence is some union of components of Y – assuming
we are not requiring manifolds to be connected!
Proof. This is a form157.667
of the Implicit Function Theorem. For a submersion,
there is a local version of (7.77) near each point of X. Namely, the surjectivity of
the differential φ∗ is equivalent to the injectivity of φ∗ : TF (p) Y −→ Tp∗ X. So if
yi are local coordinates in U ⊂ Y near F (p) then the functions φ∗ yi defined on
φ−1 (U ) have independent differentials. So they can be completed to a coordinate
system, φ∗ yi , zk on X in some neighbourhood of p. Then the fibres localy are the
surfaces φ∗ yi =const. This needs to be globalized along φ−1 (φ(p)).
116 7. OPERATORS ON MANIFOLDS
In each of the coordinate patches constructed along the ‘base fibre’ Z(p) =
φ−1 (φ(p)), the fibre containing p, the choice of local coordinates gives vector fields,
(k) (k)
just the Vi = ∂yi in the local coordinates label by k, which satisfy φ∗ (Vi ) =
(k)
∂yi are the coordinate vector fields in the base. The Vi in different coordinate
patches will be different but if we take a partition of unity rk on X near Z(p)
(k)
and subordinate to the cover then the Vi = ρk Vi satisfy φ∗ (Vi ) = ∂yi . The
compactness of Z(p) means that this is only a finite sum and so the Vi are defined in
an open set φ−1 (U ) for some neighbourhood p ∈ U in Y. The Vi need not commute
but we can simply integrate successively with respect them starting at Z(p). So
for each q ∈ Z(p) integrating V1 gives a smooth surface in Y containing Z(p) and
projecting under φ onto the y1 access in U. Then integrating from this surface gives
a submanifold of one dimension higher projecting onto the y1 , y2 coordinate plane
in U, where this may need to be shrunk at each step. After dim Y steps we have
a smooth map from a neighbourhood of the form φ−1 (U 0 ) of Z(p) in Y to Z(p)
which
157.667
together with the projection to U 0 gives a diffeomorphism F as required in
(7.77).
157.672 Proposition 7.7. If V −→ X is a vector bundle over the157.667 total space of a fibre
bundle then there are local trivializations of the bundle as in (7.77) over which there
are bundle isomorphisms
157.673 (7.80) V
τ / VZ × U
φ−1 (U )
φ−1 (U )
F / Z ×U
φ πU
# {
U
where VZ is a fixed bundle over the model fibre.
Notice that there is some constructive confusion between the total space of a vector
bundle and the vector bundle itself (which includes the projection to the base).
Proof.
Now, the kernel of a pseudodifferential operator on Z acting from sections of
one bundle to another is a distribution on Z 2 . So we need to consider the product
of each fibre with itself.
157.671 Proposition 7.8. If φ : X −→ Y is a fibre bundle then X 2 3 (x, x0 ) −→
(φ(x), φ(x0 )) ∈ Y 2 is a fibre bundle as is the fibre product, which is the closed
embedded submanifold
157.674 (7.81) X [2] = {(x, x0 ) ∈ X 2 ; φ(x) = φ(x0 )}
and the diagonal is the closed embedded submanifold
157.675 (7.82) Diag = {(x, x) ∈ X 2 ; x ∈ X}.
Clearly the diagonal is diffeomorphic to X.
Proof.
12. FAMILIES OF PSEUDODIFFERENTIAL OPERATORS 117
Here πL , πR : X [2] −→ Y are the restrictions of the two projections from X 2 and
Ω is the bundle over Y of densities on the fibres of φ.
A more standard notion for this space would be Ψm (X/Y ; V, W ) where there
is no space X/Y but it is supposed to suggest a family acting on the fibres of X. I
will not use that notation because it does not really fit with various generalizations
below.
So, we need to see that this definition does give a linear space of operators as
157.676
in (7.83) and then that they form an algebra. Again we can do this by localization
or we can think globally. What we want to do is make sense of the formula
Z
157.679 (7.85) Av(x) = A(x, x0 )v(x0 ), v ∈ C ∞ (X; V ).
fib
∗
We can certainly pull v back to X [2] under πR to give the a section of πR V over
[2]
X which is independent of x – constant on the fibres of πL . Then the product
with the kernel makes sense and using the pairing of V 0 and V gives a section
∗ ∗
of πL W ⊗ πR Ωfib . This is singular only at the diagonal and the projection πL is
transversal to the diagonal. So, we can see that the integral is well-defined and
gives a smooth section of W over X.
The tangent bundle of X has a subbundle, Tfib X ,→ T X consisting of the
vectors tangent to the fibe at each point. For each y ∈ Y the restriction if this fibre
tangent bundle is therefore just the tangent bundle of the fibre
157.683 (7.86) Tφ X Zy
= T Zy .
The dual bundle to Tφ X which I will denote Tφ∗ X is then the collection of the fibre
cotangent bundles
157.684 (7.87) Tφ∗ X Zy
= T ∗ Zy .
157.678
The adjustment in the order in (7.84) is so that:
157.680 Proposition 7.9. There is a well-defined principal symbol map giving a short
exact sequence
∗ σ ∗
157.681 (7.88) Ψm−1
φ (X; V, W ) ,→ Ψm m m
φ (X; V, W ) −→ S (Tφ X; π hom(V, W ))
118 7. OPERATORS ON MANIFOLDS
consistent with the single operator case in the sense that for each y ∈ Y there is a
restriction map
157.682 (7.89) Ψm m
φ (X; V, W ) 3 A 7−→ A(y) ∈ Ψφ (Zy ; Vy , Wy ), σm (A(y)) = σm (A) T ∗ Zy
.
⊕ +
C ∞ (S ∗ M ; π ∗ iso(V1 ⊕ V2 , W1 ⊕ W2 ))
ind / Z.
(7.93) C ∞ (S ∗ M ; W, U ) × C ∞ (S ∗ M ; V, W ) / Z×Z
ind × ind
◦ +
C ∞ (S ∗ M ; V, U )
ind / Z.
Now, we need to relate this to K-theory. Consider all of the triple (V, W, a)
where V and W are complex vector bundles over M and a is an isomorphism
between them , which we can write as a ∈ C ∞ (S ∗ M ; π ∗ iso(V, W )). We impose
three relations on this biggish set of data.
157.690 (7.94)
Bundle isomorphism invariance: (V1 , W1 , a) ' (V2 , W2 , bae−1 )
if b ∈ C ∞ (M ; W1 , W2 ), e ∈ C ∞ (M ; V1 , V2 )
157.692 (7.95)
Homotopy invariance: (V1 , W1 , a1 ) ' (V0 , W0 , a0 ) if ∃
h ∈ C ∞ (S ∗ M ×[0, 1]; iso(V, W )) where V, W are bundles over M × [0, 1]
with V M ×{0}
= V0 , V M ×{1}
= V1 , W M ×{0}
= W0 , W M ×{1}
= W1
157.691 (7.96) Stability: (V ⊕ U, W ⊕ U, a ⊕ IdU ) ' (V, W, a) for any bundle U −→ M.
Each of these is an equivalence relation – for the first and last this is straightfor-
ward, for homotopy equivalence one needs to do a littl work in concatenating two
homotopies to get a smooth one. The trick is to replace an initial homotopy by its
13. FAMILIES INDEX THEOREM 119
pull back under a smooth bijective map [0, 1] −→ [0.1] which is constant to infinite
order at both ends. Then following one of these with another gives smooth objects
over [0, 2] which can be reparameterized to a homotopy.
157.709
Taking the inner product of (7.108) with δv shows that δv = 0 (since u is
closed) and hence that
157.712 (7.110) u = h + dv.
If157.710
v = dw then taking the inner product with u shows that u is in the null space of
157.710
(7.109). Conversely if u is in the null space u = dv. Thus the map (7.109) descends
to an isomorphism
∗
157.713 (7.111) HdR (M ; R) −→ null(ð).
157.714 Corollary 5. The restriction of the Hodge-Dirac operator to
ð+ : C ∞ (M Λev ) −→ C ∞ (M ; Λodd )
has index the Euler characteristic of M
Xn
157.715 (7.112) ind(ð+ ) = (−1)k dim HdR
k
(M ).
k=0
L19-end
On a Riemann manifold a Clifford module is a complex vector bundle W −→ M
with Hermitian metric together with a norm-preserving bundle map
157.716 (7.113) cl : T ∗ M −→ hom(W )
satisfying the condition
157.717 (7.114) cl(ξ) cl(η) + cl(η) cl(ξ) = 2hξ, ηi Id in hom(W ).
157.718 Lemma 7.9. A Clifford module has a unitary connection ∇ satisfying
157.719 (7.115) ∇v cl(ξ)w = cl(∇c i)w + cl(ξ)∇v w,
∀ v ∈ C ∞ (M ; T M ), ξ ∈ C ∞ (M ; T ∗ M ), w ∈ C ∞ (M ; W )
where ∇v ξ is the action of the Levi-Civita connection on the cotangent bundle.
Proof. As usual the idea is to construct such a connection locally and then
patch using a partition of unity.
Interpreting the connection as a differential operator ∇ : C ∞ (M ; W ) −→
C (M ; T ∗ ⊗ W ) and the Clifford action as a contraction map cl : T ∗ W ⊗ W −→ W
∞
local orthonormal basis gives a trivialization of the bundle consistent with these
identifications.
A Clifford module is then a vector bundle with a multiplicative bundle map
157.723 (7.119) Cl(M ) −→ hom(W ).
Such modules arise from a Spin structure or a Spin-C structure on M. Both of
these only make sense on an oriented manifold. They have to do with the groups
157.724 (7.120) Spin −C(2k) −→ Spin(2k) −→ SO(2k).
Here (for k > 1) the spin group is the universal cover of SO(2k). Since π1 (SO(2k) =
Z2 this is a double cover. One way to construct concretely is using the Clifford
algebra on R2k . An element of SO(2k) acts on the Clifford algebra157.721through its
action on T0∗ R2k = R2k which preserves the ideal generated by (7.117). Such
an algebra-preserving isomorphism of a matrix group is necessarily generated by
conjugation so for O ∈ SO(2k) there is an element LO ∈ Cl(R2k ) such that the
action of O is
157.725 (7.121) cl(Oξ) = LO cl(ξ)L−1
O .
15. Gerbes?
I do not expect to have time to discuss this in lectures. The idea is that this give
a rather systematic approach to Spin and Spin-C structures on manifolds. Since it
is relevant, at least as background, let me briefly recall the classification of real and
complex line bundles over a manifold – which is generalized by gerbes.
A real line bundle over a manifold M can be given a smooth family of fibre
metrics and so reduced to a principal Z2 bundle the ‘sphere’ in the line at each
point
157.728 (7.123) Z2 M̂
M.
Thus, M̂ is a double cover of M with the action of Z2 being to interchage the points
in each fibre.
Such a principal bundle is classified by H 1 (M ; Z2 ). This is most easily seen in
terms of Čech cohomology. To be brief about this, any open cover of a manifold
has a refinement to a ‘good’ open cover – one in which all the open sets and all
non-trivial finite intersections of them are contractible. A covering by small (with
radius below the injectivity radius) Riemannian balls satifies this.
So, one can find such a good open cover over each element of which, Ui , the
Z2 bundle has a section. Then over each intersection Uij = Ui ∩ Uj the relation
15. GERBES? 123
M.
Now the local trivialization over a good open cover give a map
157.734 (7.129) cij : Uij −→ S s.t. cij cjk cki = 1 over Ui ∩ Uj ∩ Uk .
This cocycle yields a cohomology class, the Chern class, c ∈ Ȟ 1 (M ; S) which 157.733
is the
same as Ȟ 2 (M ; Z). The triviality of c implies the existence of a section of (7.128)
and conversely. Moreover a prinicpal circle bundle can be constructed from a class
c ∈ Ȟ 2 (M ; Z) with this as Chern class.
The relationship between real line bundles, and their complexification (obtained
by tensoring with C) is the Bocksteim homomorphism
157.735 (7.130) H 1 (M ; Z2 ) −→ H 2 (M ; Z).
Gerbes, in particular bundle gerbes, are the next step up from line bundles. I
include a brief discussion of ‘lifting bundle gerbes’, specifically for spin and spinC
structures. L20
CHAPTER 8
Semiclassical quantization
1. Blow up
11.4.2022.1 Definition 8.1. If S ⊂ M is a closed embedded submanifold a blow-up of M
along S, also called the blow-up (actually the radial blow-up) of S, is a manifold
with boundary [M ; S] and smooth surjective map β : [M ; S] −→ M (the blow-down
map) with the properties
(1) β : [M ; S] \ ∂[M ; S] −→ M \ S is a diffeomorphism
(2) β : ∂[M ; S] −→ S is a sphere bundle
(3) If 0 ≤ q ∈ C ∞ (M ) vanishes precisely at S and exactly to second order then
β ∗ q = x2 is the square of a boundary defining function x ∈ C ∞ ([M ; S])
(4) The Lie algebra of smooth vector fields on M which are tangent to S lift
to span, over C ∞ ([M ; S]), the Lie algebra of vector fields tangent to the
boundary.
The ‘precise second order vanishing’ is the statement that the Hessian of q at
each point s ∈ S is postive definite as a quadratic form vi vj q on Ns S, the normal
bundle to S.
If you recall the notion of compactification from early in the course you will see
some similarity. However blow-up is much more functorial.
Why blow up submanigolds? There are several reasons (but it isn’t always a
good idea!).
11.4.2022.2 Theorem 8.1. Any closed embedded submanifold has a blow-up and any two
are naturally diffeomorphic.
Recall that the normal bundle to S, N S = TS M/T S parameterizes the vector
fields at S ‘pointing into N ’. As we shall see the boundary of the blow-up is the
corresponding sphere bundle SN S = (N S \ 0S )/R+ . So the definition of [M ; S]
involves gluing this onto M \ S to get a manifold with boundary.
Proof. The existence will use the collar neighbourhood theorem, discussed in
Sect.collar
§ 3.
We start with a simple case, namely M = W is a real vector space and S = {0}
is the origin. We can get a quadratic function q = |w|2 as the square of the length
for some Eucludean norm on W. Then we ‘introduce polar coordinates’. These are
‘singular coordinates’ (whatever that means) but correspond to a smooth map
11.4.2022.3 (8.1) β : [0, ∞) × (W \ 0)/R+ −→ [0, ∞) × {w ∈ W ; |w| = 1} 3 (r, ŵ) 7−→ rŵ ∈ W.
Here the first map is a diffeomorphism where the metric is used to identify the
quotient with the unit sphere. The inverse of β, restricted to W \ 0, is
11.4.2022.4 (8.2) w 7−→ (|w|, [w]) ∈ [0, ∞) × (W \ 0/R+ ).
125
126 8. SEMICLASSICAL QUANTIZATION
So we conclude that under change of norm from |w| to |w|0 the identity map on
W \ 0 lifts to the smooth diffeomorphism
|w|
11.4.2022.5 (8.3) β : [0, ∞) × (W \ 0/R+ ) 3 (x, [w]) 7−→ ( x, [w])
|w|0
which is the identity on the boundary. So the construction does give the set
11.4.2022.6 (8.4) (W \ 0/R+ ) t (W \ 0)
a unique topology and C ∞ structure.
Now, we should check that this has the desired properties. The first two con-
ditions are clear since β is clearly a diffeomorphism of (0, ∞) × (W \ 0)/R+ onto
W \ {0} and the boundary11.4.2022.5
is a sphere. The quadratic function corresponding to
the metric used to define (8.3) is |w|2 so the third condition holds for this metric
and any other quadratic function is the sum of some other Euclidean metric (its
Hessian) and a function vanising to third order at the boundary. The first part is
the product of |w|2 with a smooth function on the sphere so it is the square of a
defining function for the boundary and the higher order term is smooth (since β is
smooth and vanishes to third order at the boundary.
So, it remains to check the fourth property. The vector fields tangent to 0 are
those which vanish there and so the are of the form, in any linear coordinates,
X
157.736 (8.5) aij wi ∂wj , aij ∈ C ∞ (W ).
ij
The coefficient pull back to be smooth, so to show that these lift, i.e. extend
smoothly from x > 0 down to x = 0, it suffices to show this for the wi ∂wj .
These are homogeneous of degree 0 on W and under β radial scaling becomes
(x, [w]) −→ (tx, [w]), t > 0. Thus writtend in terms of the product decomposition
11.4.2022.3
(8.1)
157.737 (8.6) wi ∂wj = a([w])x∂x + V
where V is a smooth vector field and a is a smooth function on the sphere. Both
extend smoothly down to x = 0. The radial vector field w∂157.737
w lifts to x∂x and the
wi ∂wj span all vector fields on W away from 0 so the V in (8.6) span all the vector
fields on the sphere.
This completes the proof for the blow-up of 0 ∈ W. It is clear that the linear
map reversing one coordinate lifts to be smooth as the reflection in the sphere.
Then the smoothness of the lifts of the linear vector fields wi ∂wj , which form the
Lie algebra of GL(W ) shows that the action of this group on W \ {0} extends to
[W, {0}]. It follows that the action
157.738 (8.7) GL(W ) × [W, {0}] −→ [W, {0}]
is smooth.
Now we pass to the case of the zero section of a real vector bundle, U −→ M.
We define this as the union over the base of the blow-up of the fibres just defined,
with blow-down maps
G β
157.739 (8.8) [U, 0U ] = [Um , {0}] −→ U.
m∈M
157.739
to make (8.8) into a smooth fibre bundle with smooth map. Similarly the first
and third conditions follow directly. The vector fields tangent to 0U are given in a
spanned in a local trivialization by the
157.740 (8.9) ∂zi , wi ∂wj
where the zi are coodinates on M. All these vector fields have smooth lifts in the
product decompositon of [U ; 0U ] and so globally and clearly span the vector fields
tangent to the boundary.
Finally then we pass to the general case but this follows from the collar neigh-
bourhood theorem (or working locally if you prefer) which gives a diffeomorphism χ
from a neighbourhood Q0 of an embedded submanifold S ⊂ M to a neighbourhood
Q of the zero section of its normal bundle. This gives a C ∞ stucture to
157.741 (8.10) [M ; S] = (M \ S) ∪ SN S
as a manifold with boundary and by the lifting property it is independent of the
choice of χ.
The discussion above is for a closed embedded submanifold of a manifold with-
out boundary, of course it applies unchanged if M has a boundary, including cor-
ners, but S does not meet the boundary. If S does meet the boundary we need to
specify the meaning of ‘embedded’. What is needed for the existence of a collar
neighbourhood theorem is the following condition.
157.742 Definition 8.2. A subset S ⊂ M of a manifold with corners is a p-submanifold
(the ‘p-’ being for ‘product’) if at each point of S of codimension k there are
‘adapted’ local coordinates
157.743 (8.11) x1 , . . . , xk , y1 , . . . , yn−k in O ⊂ M
where the xi ≥ 0 are local boundary defining functions and
157.744 (8.12) S ∩ O = {xl+i = 0, i = 1 . . . , k − l, yj+p = 0, p = 1, . . . , n − k − j}.
Here S has codimension n − l − j.
157.745 Theorem 8.2. Any closed p-submanifold of a manifold with corners has a
blow-up which is unique up to natural diffeomorphism.
Proof. Maybe it is best to prove the collar neighbourhood theorem in this
context first!
So it is enough to assume that w has support in F −1 (O) where the fibre bundle
is trivial over O ⊂ N and then use a partition of unity; we can arrange that W
is trivial over O as well. Then the diffeomorphism invariance of integration (of
densities) means that we really are reduced to the case that F −1 (O) = Z×) and
we are integrating over Z.
Once we know this we can reverse the definition of push-forward and see that
for a fibration pull-back extends by continuity to
157.754 (8.22) F ∗ : C −∞ (N ; W ) −→ C −∞ (M ; F ∗ W ) and is injective.
Really the pull-back is ‘constant along the fibres of F ’.
We are interested in more refined version of this. In particular notice that if
S ⊂ N is a closed embedded submanifold then for a fibration
157.755 (8.23) F −1 (S) ⊂ N is closed and embedded.
157.754
157.756 Proposition 8.3. Pull back under a fibration, as in (8.22) defines a continu-
ous linear map
157.757 (8.24) F ∗ : I m (N, S; W ) −→ I m−d/4 (M ; F −1 (S); F ∗ W )
for any closed embedded submanifold of N of codimension d and any vector bundle
W over N.
Proof.
157.752
We can, and should, ask a similar question abovt the push-foward map (8.20).
So consider a closed, embedded submanifold D ⊂ M. There are already lots of
submanifolds of M as the total space of a fibration, namely the fibres. Recall
157.758 Definition 8.3. Two embedded submanifolds D and Z in a manifold M meet
transversally if at each point intersection
157.759 (8.25) p ∈ D ∩ Z =⇒ Tp D + Tp Z = Tp M ⇐⇒ Np∗ D ∩ N ∗ )pZ = {0}.
157.760 Lemma 8.1. The transversal intesection of two embedded submanifolds is an
embedded submanifold with codimension the sum of the codimenstions.
In particular two manifolds which do not intersect ‘intersect transverally’. For two
embedded submanifolds the notation for their intersection D t Z means that they
intersect transversally.
Proof. By definition near any point of D there are local defining functions
wi which vanish on DW with independent differentials. Similarly there are local
defining functions uj for Z. At a point of intersection the transversality condition
means that the wi and uj have independent differentials. Since they vanish precisely
on the intersection locally, it is an embedded submanifold.
157.762 Definition 8.4. We say that a smooth map F : M −→ N is transvesal to an
embedded submanifold D ⊂ M (or that the submanifold is transversal to the map)
if the differential F∗ : Tp D −→ TF (p) N is surjective for each p ∈ D.
157.761 Proposition 8.4. A submanifold D ⊂ M of the total space of a fibration is
transversal to the fibration if and only if it is transversal to each fibre; if D is closed
then
157.763 (8.26) F∗ : Icm (M, D; F ∗ W ⊗ Ω) −→ C c∞ (M ; W ⊗ Ω).
130 8. SEMICLASSICAL QUANTIZATION
Proof.
157.767
is a diffeomorphism. Now it follows from (8.32) that all three manifolds must have
the same codimension – since the intersection has codimension which is equal to
the sum of each157.768 pair.
To prove (8.34),
157.766
first note that away from D1 ∩ D2 the image of the product
map in (8.31) lies in the sum of I ∗ (M ; D1 , W1 ⊗ W2 ) and I ∗ (M, D2 ; W1 ⊗ W2 ) so,
157.761
after composing with the bundle map h, so Proposition 8.4 applies and shows that
the push-forward is smooth. Thus, it suffices to consider a small neighbourhood of
a point n ∈ S ⊂ N and its preimage. We may also assume that the elements of
Icm1 (M, D1 ; W1 ) and Icm2 (M, D2 ; W2 ) are supported near unique preimage n0 of n
in D1 ∩ D2 (by using a partition of unity and discarding smooth terms).
Now, it is convenient to observe that D1 and D2 and the fibration can be
brought to simultaneous normal form near such a point n0 ∈ D1 ∩ D2 . Let z =
(z 0 , z 00 ) be local coordinates in the base, N, near n in which S = {z 00 = 0}. Since
φ is assumed to be a diffeomorphism when restricted to D1 it follows that it is a
graph over N locally. So we can introduce additional variables y near n0 so that
(y, z 0 , z 00 ) form a coordinate system and
157.778 (8.36) D1 = {(y, z 0 , z 00 ); y = 0} near n0 .
Since D2 is also a graph over N near n0 it takes the form
157.779 (8.37) yi = Yi (z 0 , z 00 ) near n0 .
The codimension of D1 , the number of yi , is equal to the codimension of S ⊂ N,
i.e. the number of z 00 variables.
The assumtion that D1 ∩ D2 ⊂ φ−1 (S) means that Y (z 0 , 0) = 0 and the
transversality of D1 and D2 implies that
157.780 (8.38) dz00 Yi (z 0 , z 00 ) are linearly independent at (0, 0, 0)
(since the dz00 Yi = 0. These form a square matrix, so the Yi can be introduced
as new variables in the base, in place of the z 00 and defining S. This then is the
coordinate normal form
157.781 (8.39) D1 = {y = 0}, D2 = {z 00 − y = 0}, D3 = {z 00 = 0}.
Ignoring the bundles, it follows that the conormal distributions locally take the
form
157.782 (8.40) Z Z
00
−y)·η 0
u1 = (2π)−d a(z 0 , z 00 , η)eiy·η dη, u2 = (2π)−d b(z 0 , z 00 , η 0 )ei(z dη 0
Rd Rd
where we can assume that the symbols a and b are supported near (z 0 , z 00 ) = 0. The
push-forward of the product is then
Z
157.783 (8.41) u1 (z, y)u2 (z, y)
Z
00 0
−2d
= (2π) a(z 0 , z 00 , ξ + η 0 )b(z 0 , z 00 , η 0 )ez ·η eiy·ξ dηdξdy, ξ = η − η 0 .
R3d
Formally at least the ξ, y double integral can be interpreted as a Fourier/inverse
Fourier transform which evaluates the integrand at ξ = 0 giving
Z
00 0
−2
157.784 (8.42) φ∗ (u1 u2 ) = (2π) a(z 0 , z 00 , η 0 )b(z 0 , z 00 , η 0 )ez ·η dη 0 ∈ I ∗ (N, S).
R3d
To justify these last step we use continuity in the symbol topology as usual.
132 8. SEMICLASSICAL QUANTIZATION
So the bundle involved here is precisely the fibre tangent bundle φ T X ⊂ T X with
the anchor map being the natural inclusion (so of constant rank).
The case of immediate interest is the semiclassicl Lie algebroid which is closely
related to Vφ . Namely, take a manifold M and consider
157.790 (8.47) M [1, sl] = M × [0, 1] −→ [0, 1].
On this space we consider fibre vector fields which in addition vanish at the bound-
ary = 0. Clearly in local coordinates, zi , on M this means the smooth vector fields
of the form
X
157.791 (8.48) ai (, z)∂zj .
i
These clearly form a Lie algebra, but what is the bundle V ? We have to construct
157.791
it. The form of (8.48) makes it rather clear that we have a local basis with elements
157.792 (8.49) ∂zi .
sl
So these give a basis for the bundle T M (which is a bundle over M [1, sl] not M
despite my notation). You might struggle a bit to think of ∂zi as an ‘entity’ rather
L21-end than the product of and ∂zi but that is precisely what is involved here.
L22
7. THE ATIYAH-SINGER INDEX THEOREM 133
b ∈ C ∞ (Tfib
∗ X; hom(CN )) is such that b : Ran(P ) −→ Ran(P ) is an iso-
1 2
∗ ∗ X) with P = 1 (β + Id) the
morphism over Sfib X (the boundary of Tfib i 2 i
positive projections of the involutions. This is the semiclassical data con-
sidered above the equivalence classes under homotopy and stability again
form the group K 0 (T ∗ X. Semiclassical quantization
157.795 fib
results in the same
i.data
index map (8.53) when the data reduces to that in 1.
∗
i.escl constant and b = Id exhausts K 8 (Tfib
(7) The data where β2 is157.795 X) so we can
get the index map (8.53) from semiclassical smoothing operators – were
β1 is constant near infinity.
(8) Now the main part of the proof of the index theorem is the embedding of
φ in a trivial fibration Rk × Y for some (largish) k :
X / Rk × Y
ι×φ
157.796 (8.54) Z Rk
φ
π2
{
Y.
This follows by choosing an embedding ι : X −→ Rk and then into Rk × Y
by adding the map φ.
(9) Now we think of X as a submanifold of Rk ×Y and contemplate its normal
bundle – so an open neighbourhood of the image. Each fibre Zy of X is
embedded in Rk so the full normal bundle is the bundle over X which
over Zy is N Zy ⊂ Rk × {y}. So we actually have a ‘tower’ of fibrations
157.797 (8.55) Rd NX
ψ
Z X
φ
Y
(10) The main idea in this proof by Atiyah-Singer (they had another one too,
using cobordism) is that we can ‘extend’ the data and quantization from
∗
Tfib X to
∗
157.798 (8.56) Tψψ N X = N X ⊕ N ∗ X −→ X.
157.797
The fibres of ψ in (8.55) are non-compact but of course they are real vector
157.798
spaces. So the fibewise cotangent bundle in (8.56) has fibres the sum of a
vector space and its dual.
∗
(11) So, we can find a family of involutions on Tψψ N X which are costant
outside a compact set and quantize (semiclassically) to have index a trivial
one-dimensional bundle over X.
(12) Then we take what is essentially the tensor product of this ‘Bott element’
(well, better to say a ‘Thom’ element) with the original family over X to
be a family for φ ◦ ψ : N X −→ Y which quantizes to the have the same
i.escl
image as a given element 7. We do this by quantizing in two steps.
(13) Now, we can arrange the support of the Thom element to be very close
to the zero section and thereby move it, using a collar map, to Rk × Y
9. THE DIRAC CASE 135
as a family. This is ‘excision’ and gives the same index for the extended
family.
(14) The index map from Rk ×Y to Y is an isomorphism – again this is explicit
Bott periodicity.
(15) Finally then we have an extended construction of the index map giving a
commutative diagram
ind
ind
' v ind
K 0 (Y )
(16) The final step then is to see that this diagram is the definition of the
push-forward in K-theory, the ‘Gysin’ map in this context.
8. Index formula
Of course this is not the end of the story, quite apart from the fact that there
are a few gaps in the argument – which I try to fill in below. The index in K-theory
as above precisely captures the obstruction to an ellipitic family (or semiclassical
family) have a smoothing perturbation which makes it invertible. In the case that
Y is a point K 0 (Y ) = Z and we can look for a formula for the actual ‘numerical
index’. In the case of a family we can look for a simpler obstruction to perturbative
invertibility, corresponding to the image of the the index in (let’s say deRham)
cohomology under the Chern character
157.800 (8.58) Ch : K 0 (Y ) −→ H ev (X; R).
Either of these is the index formula.
157.801 Theorem 8.4. The index of the image in cohomology is
Z
157.802 (8.59) Ch ◦ ind = φ∗ (Ch([(V, W, a]) ∧ Td) = Ch([(V, W, a]) ∧ Td
for arbitrary smooth coefficients. This means that there is a vector bundle b T M
over M with sections precisely these vector fields
157.808 (9.4) Vb (M ) = C ∞ (M ; b T M ).
Let’s think a little about the structure of the vector bundle b T M, since I am
asserting it is, in context, the appropriate replacement for the ‘ordinary’ tangent
bundle T M (to which is it isomorphic – just not naturally so). Over the interior
of M there is not much to say since these two bundles are naturally isomorphic.
Since the elements of Vb (M ) are smoth vector fields there is a completely natural
smooth vector bundle map (the anchor map of the Lie algebroid Vb (M ))
b
157.899 (9.5) T M −→ T M.
Over the boundary this has corank 1 – there is a 1-dimensional null space since
the vector field (in local coordinates) x∂x vanishes in the ordinary sense at the
boundary. So there is a 1-dimensional subbundle
b
157.900 (9.6) N ∂M ⊂ b T∂M M.
137
138 9. MANIFOLDS WITH BOUNDARY
In fact this is a canonically trivial subbundle. The element x∂x is actually (at a
boundary point) defined independently of coordinates. Indeed it satisfies
for any defining function ρ. This just reflects the fact that any other defining func-
tion ρ = a(y)x + O(x2 ).
In this behavour b T∂M M is ‘reversed’ from T∂M .M The latter has T ∂M as a
subbundle, with the quotient being N ∂M, the normal bundle. The former has a
(trivial) subbundle with quotient naturally T ∂M.
Let’s go a little further with this analysis of tangency. The primitive ideal of
functions vanishing at the boundary, I ∂ , generated by x, leads to a Lie ideal
This means that the space of sections of b T∂M M as a bundle over ∂M is itself a
Lie algebra. This is clear enough in local coordinates. So there is actually a Lie
algebra map
I have perhaps not emphasized enough uthat the Lie algebra structure of V (M )
in the boundaryless case, or Vb (M ) here, is what leads to the properties of the
differential operators, in particular that the leading part defines the (polynomial)
symbol map. This is fair warning that we should expect something similar at the
boundary for our, yet to be defined, b-pseudodifferential
157.903
operators. It will be the
‘indicial operator’ and arises precisely because (9.9) is a map of Lie algebras. We
can even guess it should take values in the pseudodifferential operators on the
boundary but with ‘an extra parameter’.
So what we want to find is an algebra of operators, say on C ∞ (M ), which
include the vector fields Vb (M ) and multliplication by C ∞ (M ) and which away
from the boundary should reduce to ordinary pseudodifferential operators. One
can approach this as for the semiclassical calculus. Writing, informally, a pseudo-
differential operator in terms of symbols we can try to replace a symbol
where (ξ, η) are the dual variables to (x, y). This does work but there are significant
issues involved.
Proceeding formally we can plug such a symbol into the inverse Fourier trans-
form and then change variables as for the semiclassical calculus (ignoring isses of
domains and convergence)
Z
0 0
157.867 (9.11) a(x, y, xξ, η)ei(x−x )ξ+i(y−y ) dξdη
Z
x−x0 0 dτ
= a(x, y, xξ, η)ei( x )τ +i(y−y ) dη, τ = xξ.
x
So, from this point of view what we need to do is to find a space on which (x−x0 )/x
is smooth, at least where it is finite. This we can do by an appropriate blow-up.
1. THE B-GENERALIZED PRODUCTS 139
The idea of this blow-up is to introduce a space of kernels which is much easier
to describe there than directly on M 2 . Note that one version of the Schwartz kernel
theorem states that, on a manifold with corners, continuous linear operators
157.816 (9.18) A : C˙ ∞ (M ) −→ C −∞ (M ) = (C˙ ∞ (M ; Ω))0
are identified with distributions
157.817 (9.19) A ∈ C −∞ (M 2 ; πR
∗
Ω).
157.816
The image space in (9.18) is the space of extendible distributions as is the space
of kernels on M 2 (apart from the density factor). These distributions are defined
on any compact manifold with corners X as the dual of C˙ ∞ (X; Ω) (they are the
analogue of tempered distributions on Rn with which they are identified for the
radial compactification).
157.818 Lemma 9.1. The blow-down map gives an isomorphism
157.820 (9.20) β ∗ : C˙ ∞ (M 2 ) −→ C˙ ∞ (M [2, b])
and in consequence the 157.816
Schwartz kernel theorem also identifies the space of contin-
uous linear operators (9.18) with
157.819 (9.21) A ∈ C −∞ (M [2, b]; πR
∗
Ω)
The point is that we are not actually changing the space of extendible distri-
butions by passing from M 2 to M [2, b], what is changing is the space of smooth
functions (which is getting bigger) and the space of conormal distributions with
respect to the diagonal (which is a pain to define on M 2 ).
Now, before going on let’s check that the passage to M [2, b] does ‘resolve’ the
Lie algebroid Vb (M ) in an appropriate sense. The vector fields in Vb (M ) acting
in M 2 are tangent to the corner x = 0 = x0
on the left (or the right) factor of M157.806
0
since they annihilate
157.807
x and satisfy (9.2). Thus they lift to be smooth on M [2, b].
157.815
In terms of (9.3) and the local coordinates (9.17), the ∂yi lift unchanged whereas
157.821 (9.22) x∂x = s∂s .
This may not seem like much of an improvement! However, the lifted diagonal is
at s = 1 so this vector field does not vanish there and we see:
157.822 Lemma 9.2. The elements of Vb (M ) lifted to M [2, b] from the left (or right)
factor of M in M 2 are transversal to the lifted diagonal Diagb ⊂ M [2, b] and so
the normal bundle to this submanifold is identified with b T M.
This is a minimal requirement for ‘resolution’ of Vb (M ).
As well as the ‘stretched double space’ there are similar replace-
ments for the higher products M k . I will invoke the stretched
triple space below. let me continue to assume that the bound-
ary of M is connected – if it has more than one component you
should proceed component by component, not thinking of inter-
action between the components which are ‘far apart’.
We can see that the boundary faces of M k consist of prod-
ucts where in each factor we have either ∂M or M. Now arrange
these in order of increasing dimension – starting at (∂M )k – and
then blow them up, one after another
M [k, b] = [M k ; (∂M )k ; M k−1 (M k ), . . . , M 2 (M k )] 157.868 (9.23)
1. THE B-GENERALIZED PRODUCTS 141
157.874
corners of M, as follows from (9.25), and can be described much
more precisely. 157.873
In particular, Proposition 9.2 must apply to the projection
maps
πL,R β : M [2, b] −→ M. 157.876 (9.27)
157.871
Here it is essy to check Proposition 9.1 by hand. Away from ff
is just locally one of the projection from M 2 to M for
the map157.872
which (9.24) certainly holds. Near the front face we have one of
the two coordinate systems
x x0
s = 0 , x0 , y, y 0 or t = , x, y, y 0 . 157.877 (9.28)
x x
The left projection, to (x, y) therefore becomes either
(s, x0 , y, y 0 ) 7−→ (x = sx0 , y) or (x, y) 157.878 (9.29)
157.872
depending on the point in ff both of which satisfy (9.24).
157.872
It follows directly from (9.24) that a simple b-fibration is
locally a fibration in the interior and also at boundary points of
codimension one. This allows us to deduce
157.879 Lemma 9.5. For a compact manifold with corners
∞
(πL )∗ : Ψm
b (M ) −→ C (M ). 157.880 (9.30)
157.804
This gives a direct proof of the mapping property (8.51).
3. The b-calculus
157.816
The most basic operator in (9.18) is the identity operator. We know that in
local coordinates the kernel of this is the Dirac ‘function’ at the diagonal
157.823 (9.31) δ(x − x0 )δ(y − y 0 ).
This is, as it must be, a distributional section of the right density bundle
∗
157.824 (9.32) Id ∈ I 0 (M 2 ; πR Ω).
To see that this makes invariant sense observe that it must be possible to pair the
delta ‘function’ with an element of C ∞ (M 2 ; πL
∗
Ω) since
∗ ∗
Ω(M 2 ) = πL Ω ⊗ πR Ω.
Say using a partition of unity to localize, we need to be able to make sense of the
distributional pairing, written formally as an integral
Z Z
157.826 (9.33) δ(x − x0 )δ(y − y 0 )ψ(x, y, x0 , y 0 )|dxdy| = ψ(x, y, x, y)|dxdy|
M2 M
157.824
which is indeed invariantly defined – so this is what (9.32) actually means.
Now, what happens when we look at the lift of this kernel to M [2, b] as an
extendible distribution – these form a subspace of the distributions on the interior.
It is certainly still supported at the diagonal,
157.815
so it is only a question of what it
looks like in the new coordinates say (9.17). The homogeneity of delta means it
becomes
157.827 (9.34) δ(x − x0 )δ(y − y 0 ) = (x0 )−1 δ(s − 1)δ(y − y 0 ).
We can absorb the extra singular factor of x0 into the measure to see that
∗
157.828 (9.35) Id ∈ I 0 (M [2, b], Diagb ; πR Ωb )
where Ωb is the density bundle coming from b T M so in fact in a natural way it has
a basis near the boundary
dx
157.829 (9.36) |dy| =⇒ Ωb = x−1 Ω.
x
This leads us to the definition of b-pseudodifferential operators through their
kernels.
157.830 Definition 9.1 (‘Small’ b-calculus). The space of b-pseudodifferential oper-
ators on a compact manifold with boundary, acting between sections of vector
bundles V and W is
157.831 (9.37) Ψm m
b (M ; V, W ) = {A ∈ I (M [2, b], Diagb ;
∗ ∗
πL W ⊗ πR (V 0 ⊗ Ωb )); A ≡ 0 at both lifted boundaries}.
I have not actually defined the conormal space here but it is exactly the re-
striction of the usual conormal space if one extends 157.831
across the boundary. These
distributions are smooth away from the diagonal so (9.37) makes sense since these
‘old’ boundaries do not meet the diagonal. Locally such a kernel, with the b-density
removed, just looks like
A(x, y, s, y − y 0 ) smooth in (x, y) and conormal at s = 1, y − y 0 = 0.
144 9. MANIFOLDS WITH BOUNDARY
distributions and of smooth functions we can assume more about A and then use
continuity. Namely we can suppose that the kernel A is continuous (we could
assume smoothness) and has support actually disjoint from the two ‘old boundaries’.
Then the coordiantes t = x0 /x, x y and y 0 are admissible over the support after
localization
dx0
157.847 (9.44) A = A(x, y, t, y − y 0 )| 0 dy 0 |.
x
The action of U on u is then the integral
dx0
Z
157.848 (9.45) Au = A(x, y, t, y − y 0 )u(x0 , y 0 )| 0 dy 0 |
x0 ≤ x
where u has compact support down to x0 = 0. This integral certainly exists for
x > 0 since then the integrand has compact supprt. We can change the variable of
integration from x0 to t and see that
Z
dt
157.849 (9.46) Au = A(x, y, t, y − y 0 )u(x/t, y 0 )| dy 0 |.
[0,∞) t
Now the integral exists (by fiat the support in t is in [C, 1/C] for some finite C)
157.836
and the integrand is smooth in x and y, so the result is C ∞ and we have (9.40)157.831
under these assumptions on A. In fact we can pass to unrestricted A as in (9.37)
since these kernels vanish rapidly at t = 0 and t = ∞ and 157.836
the only singularity is
conormal at t = 1 – across which we are integrating. Thus (9.40) follows in general
and as a bonus we see that, as anticipated above
157.836
157.850 Corollary 6. Restriction to the boundary in (9.40) gives
Au ∂M
= (A∂ )u , A∂ ∈ Ψm (∂M ; V, W ),
∂M
157.851 (9.47) 0 Z
0 dx 0
A = Ã(x, y, t, y − y )| 0 dy | =⇒ A∂ = A(0, t, y, y − y 0 )dt|dy 0 |
x R+
so the kernel of A∂ is the integral over the fibres of β : ff −→ (∂M )2 of the kernel
of A.
We still need to prove that the product of two b-pseudodifferential operators is
b-pseudodifferential.
Thus everything is very much as in the boundaryless case except that we have
much more structure at the boundary. You might like to reflect on the similarity
to the behaviour of the semiclassical calculus here.
algebra Vb (M ) – these of course define elements
For a moment return to the Lie157.836
of Ψ1b (M ). Certainly they satisfy (9.40) but also
157.837 (9.48) (V u) ∂M
= (V ∂M
)u ∂M
, u ∈ C ∞ (M ).
They are ‘localized at the boundary.’ However we know that the analogous state-
ment for the semiclassical calculus holds but misses important structure at the
boundary. Much the same happens here.
The Collar Neighbourhood Theorem reminds us that a neighbourhood of a
submanifold looks like a neighbourhood of the zero section of its normal bundle.
For the boundary this translates to mean that a model for M near the boundary
is the inward-point half of the normal bundle
157.838 (9.49) N + ∂M = {v ∈ T∂M ; vx ≥ 0}/T ∂M.
146 9. MANIFOLDS WITH BOUNDARY
The choice of a boundary defining function gives a positive section and hence triv-
ialization
157.839 (9.50) N + ∂M ←→ [0, ∞)dx × ∂M
where dx defines a linear functional on the fibres of T∂M M which vanishes on the
subbundle T ∂M.
Let’s pass to the radial compactification of N + ∂M = I ×∂M where I is a closed
interval, thought of as the radial compactification of [0, ∞) – there is no natural
trivialization
157.839
but any choice of boundary defining function provides one through
(9.50). The fibre R+ action extends smoothly to the compactification (fixing the
two boundaries {0} × ∂M and {∞} × ∂M ).
157.840 Proposition 9.3. There is a natural multiplicative map, the ‘indicial map’ to
R+ -invariant operators on the normal bundle, giving a short exact sequence
(9.51) xΨm / Ψm (M ; V, W ) I / Ψm+ (N + ∂M ; V∂M , W∂M ).
157.841 b (M ; V, W ) b R
Proof. Let’s choose a boundary defining function, rather than try to do things
invariantly – in the end nothing will depend on this choice. Then the model at the
boundary is
157.842 (9.52) N + ∂M = I × ∂M, I = [0, ∞]x .
The space on which the kernels for the b-pseudodifferential operators are defined is
therefore
157.843 (9.53) I[2, b] = [I 2 , {0} × {0}, {∞} × {∞}].
If you consider the R+ action on both factors starting at a point in the interior of
I[2, b] you will see that it is an open interval but the closure is smooth
157.841
up to both
the front faces. An R+ -invariant operator, as on the right of (9.51) corresponds
to a kernel which is constant under this action. So in fact it is determined by
its restriction to either of the front faces. Nothing much happens at the other
boundaries since everything is required to vanish to infinite order there.
So, the invariant operators are determined uniquely by their restrictions to the
front face ff 0 over x = 0. However this face for the model space is precisely the
same as the face ff(M [2, b]) – canonically diffeomorphic to it (independent of the
157.841
choice of defining function). So the indicial map in (9.51) is restriction of the kernel
to the front face, and then its null space consists of kernels that vanish there. You
might object that x + x0 , not x = (x + x0 )/(1 + s) is the defining function for the
front face, but the kernel vanish to infinite order where s → ∞ so we can just as
well divide by x.
This does not 157.841
explain the multiplicativity of the sequence but it follows that
the null space in (9.51) is pretty clearly an ideal, so the quotient is an algebra – it
is a question of what the product is! 157.850
To approach this reconsider Corollary 6. We have already noted that the space
157.831
of kernels in (9.37) is a module over C ∞ (M [2, b]) but more is true because of the
assumption of rapid vanishing at the boundary hypersurfaces other than ff . The
quotient of defining functions from the left and right, x/x0 , is smooth except at
one these two hypersurfaces and the rapid vanishing of the kernels there quashes
the singularity from x0 = 0. In fact the same is true for any power of this quotient,
3. THE B-CALCULUS 147
It follows (even without knowing the multiplicative property for the operators)
that
157.855 (9.56) (AB)∂ = A∂ B∂
157.851
and we see from (9.47) that
Z
−iz
157.856 (9.57) iz
(x Ax )∂ = A(0, t, y, y − y 0 )t−iz dt|dy 0 |
R+
first variable x – since this reduces to the Fourier transform on log x ∈ R. Thus in
fact
Z
dx
157.885 (9.60) v(x, y) −→ vM (z, y) = v(x, y)xiz
R x
extends from C˙ ∞ ([0, ∞] × ∂M ) to an isomorphism
157.886 (9.61) L2b ([0, ∞] × ∂M ) −→ L2 (Rz × ∂M ), z ∈ R.
157.856
The fact that the transformed operator in (9.57) is a family of pseudodifferential
operators in Ψ0 (∂M ) which is bounded as a function of z ∈ R shows that it is
157.886
bounded on the image of (9.61) so in fact the R+ -invariant operators are bounded
157.886
on the space on the left in (9.61). For the invariant operators we can again localize
near and away from the boundary and deduce that the part localized near the
boundary is bounded on L2b ([0, ∞]157.841
× ∂M ).
In view of the exact sequence (9.51) it suffices to consider elements of xΨ−∞
b (M )
(using the preceding argument). The extra vanishing at ff shows that this follows
directly from Schur’s criterion.
Of course we really want boundedness on Sobolev spaces, but the ones we want
here are the xs Hbm (M ) which we need to define. If we work in a fixed product
157.885
decomposition near the boundary we can use the Mellin isomorphism (9.60) as we
would for Euclidean space and define
5. Hodge theorems
157.857
Rather than the ‘regular metrics’ as in (9.63) I want to consider two classes
of metrics which are known in the geometric literature as ‘cylindrical end’ and
‘asymptotically locally Euclidean’ metrics. In fact the precise definition of these
terms is a bit vague, so instead I will use the following notation.
150 9. MANIFOLDS WITH BOUNDARY
Now you may see why I have not needed to develop the theory of ‘scattering
pseudodifferential operators’ to handle this case since in fact
1 ∗ ð∂ −x∂x − (k − d − 1)
157.896 (9.80) ðsc = xR, R ∈ Ψb (M ; Λ M ), I(R) = .
x∂x − k −ð∂
This means that it is b-pseudodifferential operators which are relevant for the null
space of ðsc . It is a different matter if you wish to discuss the spectral theory of
this operator (which I feel you should want to do)– which is indeed really scattering
theory.
157.888
157.897 Theorem 9.3. For a scattering metric (9.73) on a compact manifold with
boundary of dimension n there is a natural isomorphism
1
157.898 (9.81) {u ∈ x 2 n L2b (M ; Λkb M ); ðsc u = 0} −→
H
k
(M, ∂M ) k < 21 n
1 1
2n n
Im HdR (M, ∂M ) −→ Hb2 (M ) k = 12 n
H k (M ) k > 12 n.
157.898
The L2 space in (9.81) is the metric L2 space for gsc . Of course the ‘middle dimen-
sional case’ can only occur if n is even.
Try it out for the Euclidean metric on M = Rn . It follows that there is no
L2 null space at all! This corresponds to the fact that here are no L2 harmonic
forms on Rn – their coefficients would be harmonic functions which would mean
they decay at infinity. 157.892
Let’s think about a strategy for proving Theorem 9.2. First we need to get
some way to approach the deRham cohomology in this setting.
157.904 Proposition 9.5. For > 0 the cohomology of the deRham complex
Proof. The important point is the sufficiency, well the necessity is important
but only to know!
Proof. We are really working with the spectral theory of ð∂ here. First notice
157.920
that any solution of (9.92) has a conormal singularity at the ‘diagonal’ appearing
on the right. Morevover, using the symbol map, we can construct a parameterix
satisfying
157.922 (9.94) I(ðb )K0 = δ(s − 1)δ(y − y 0 ) − E(s, y, y 0 ), E ∈ C c∞ (ff(M [2, b]).
So it remains to ‘solve away’ the error term E which has support in the interior of
ff .
We ‘know’ (I hope) that the expansion of a smooth function such as E (valued
here in 2 × 2 matrices acting on Λ∗ ∂M ) in the y variable in terms of the eigenbasis
of ð∂M converges rapidly. So on each eigenspace, with eigenvalue λi ∈ Spec(ð∂M )
we wish to solve
λi −s∂s
157.924 (9.95) Ki0 (s, y 0 ) = Ei (s, y 0 ).
s∂s −λi
This ordinary differential, and R+ -invariant, equation has a unique solution
which vanishes near s = 0 and any two solutions differ by an element of the 2-
dimensional null space which is spanned by
u u
xλ i and x−λi
−u u
where u is the eigenvector.
As s = 1/t → ∞ the chosen solution is in the null space. Thus we can arrange
157.921
(9.93) for this one term by adding the approriate element of the null space. Sum-
ming over the eigenexpansion
157.922
gives rapid convergence and 157.921
hence we do in fact find
a unique solution to (9.94) with the desired behaviour, (9.93), with respect to a
given w ∈ R. There can only be equality there if the wright w is equal to one of the
eigenvalues but since these form a discrete set the kernel satisfies
157.926 (9.96) K ∈ xw− Hb∞ near {s = 0} and K ∈ x−w+ Hb∞ near {t = 0}.
154 9. MANIFOLDS WITH BOUNDARY
The idea of course is that this kernel is to be the restriction to ff(M [2, b]) of the
parametrix for ðb . Now it should be clear why we cannot get157.912 a parametrix in the
space Ψ−1
b (M ; Λ∗
b ) – because the solution we have found to ( 9.86) does not decay
157.830
rapidly at the boundaries of the front face as required by Definition 9.1.
157.928 Definition 9.3. Given a pair of index sets E = (E L , E R ) we define
where N is the number operator on boundary forms. Again these are generated
by the spectrum of ð∂ but the relationship is more complicated than in the b-case
above.
Each harmonic k-form, γ, on the boundary, in the null space of ð∂ , generates
a two-dimensional space on which I(R) acts, spanned by
dx
157.955 (9.112) x−k γ, ∧ x−k γ.
x2
These each correspond to an solution of I(R)v = 0, namely
γ dx γ
157.956 (9.113) xk , ∧ xn/2−k k .
xk x2 x
The first can be the leading term of a square-integrable form only if k > n/2 and
the second only if k < n/2.
To claify square-integrability we need to analyze the other indicial roots. An
eigenform for ð∂ with non-zero eigenvector corresponds to a pair of forms, a coclosed
(k−1)-form hk−1 and a closed k-form uk with duk−1 = λuk and δuk = λek−1 . These
generate a 4-dimensional bundle invariant under I(R) spanned by
uk−1 dx uk uk dx uk−1
157.957 (9.114) k−1
, 2 ∧ k , k and 2 k−1 .
x x x x x x
Since the 2-dimensional space spanned by the first two elements maps into that
spanned by the second two and conversely, the null space of I(R) is conatained in
these two subspaces.
In the two cases the indicial roots correspond to a null vector of
λ −iz + (n − k − 1)) λ −iz + (n − k)
157.958 (9.115) and
iz − k + 1 −λ iz − k −λ
which occur when
Proof. The main point here is that ellipticity alone does imply that the indi-
cial operator I(A) is invertible as an R+ -convolution operator such a set of weights.
The main part of this is that the indicial family
157.934 (9.119) ˆ
I(A) : C −→ Ψm (∂M ; V, W )
which is entire, has a meromorphic inverse – so with a discrete set of poles
157.935 (9.120) b -Spec(A) ⊂ C s.t. zj ∈ b -Spec(A), |zj | → ∞ =⇒ | Re zj | → ∞.
So this means there are only finiteley many poles in any strip Re z| < R.
This in turn is a form of ‘analytic Fredholm theory’. For a holomorphic family
of elliptic operators, such as we have here, defined on a connected open set, the
inverse is meromorphic there is one point at which the operator is invertible. In
157.935
this case the existence of such a point, and the bound on the set of poles in (9.120)
is a consequence of
157.936 Lemma 9.11. The indicial family I(A)(t+iτ ) for t ∈ R is a semiclassical family
down to = ±1/τ as τ → ±∞ in R.
Proof.
Then the parametrix construction above generalizes to yield the Fredholm prop-
157.933
ery (9.118) for
157.937 (9.121) w∈
/ Re b -Spec(A).
157.933
So the Fredholm condition for (9.118) for w in an open set with discrete comple-
ment. The index is necessarily constant on the open sets but changes as w crosses
an end-point. In fact there is a multiplicity function corresponding to the algebraic
ˆ
multiplicity of the residues at the poles of I(A),
157.938 (9.122) rank : b -Spec(A) −→ Z
and the change of the index in passing from one interval to the next is the sum
of the multiplicity of the points in b -Spec(A) with real part corresponding to the
end-point. It is elementary to see that the index is a decreasing
APS
function of w.
For Dirac operators Atiyah, Patodi and Singer ([?]) gave a formula for the
index which applies in this case. This is extensively discussed, from the present
tapsit
point of view, in [?]. You might ask, is there a formula for the index in this general
b-pseudodifferential case? The answer of course is yes!
If you have survived this far, you would certainly be tempted to ask: Is there
a families index theorem? There is, at least there is a families index formula for
the Dirac case in the literature and this can be extended to the case of families of
Fredholm b-pseudodifferential operators. However, unlike the case of one operator
discussed above, for an elliptic family of b-pseudodifferentia operators on the fibres
of a fibre bundle (so of course the fibres are compact manifolds with boundary)
there is an obstruction to this forming a Fredholm family.
157.939 Proposition 9.9. The family of indicial operators of an elliptic family of b-
pseudodifferential operators on the fibres of a compact fibre bundle define an index
class in K 1 the vanishing of which is a necessary and sufficient condition for the
existence of a Fredholm family with the same symbol.
158 9. MANIFOLDS WITH BOUNDARY
CoSLG [1] Pierre Albin, Panagiotis Dimakis, Richard Melrose, and David Vogan, Compactification of
semi-simple lie groups, 2019.
deConcini-Procesi [2] Corrado De Concini and Claudio Procesi, The invariant theory of matrices, University Lecture
Series, vol. 69, American Mathematical Society, Providence, RI, 2017. MR 3726879
MR1269107 [3] Alain Grigis and Johannes Sjöstrand, Microlocal analysis for differential operators, London
Mathematical Society Lecture Note Series, vol. 196, Cambridge University Press, Cambridge,
1994, An introduction. MR 1269107
Hormander-pseudo [4] L. Hörmander, Pseudo-differential operators, Comm. Pure Appl. Math. 18 (1965), 501–517.
Hormander-WeylCalculus [5] , The Weyl calculus of pseudo-differential operators, Comm. Pure Appl. Math. 32
(1979), 359–443.
iml [6] R. B. Melrose, Introduction to microlocal analysis, Lecture notes.
MR2650633 [7] Alberto Parmeggiani, Spectral theory of non-commutative harmonic oscillators: an introduc-
tion, Lecture Notes in Mathematics, vol. 1992, Springer-Verlag, Berlin, 2010. MR 2650633
MR0273463 [8] M. A. Šubin, Pseudodifferential operators in Rn , Dokl. Akad. Nauk SSSR 196 (1971), 316–319.
MR 0273463
MR2952218 [9] Maciej Zworski, Semiclassical analysis, Graduate Studies in Mathematics, vol. 138, American
Mathematical Society, Providence, RI, 2012. MR 2952218
159