The Dirac Spectrum
The Dirac Spectrum
Editors:
J.-M. Morel, Cachan
F. Takens, Groningen
B. Teissier, Paris
Nicolas Ginoux
ABC
Nicolas Ginoux
NWF I -Mathematics
University of Regensburg
Universitätsstraße 31
93040 Regensburg
Germany
nicolas.ginoux@mathematik.uni-regensburg.de
This overview is based on the talk [101] given at the mini-workshop 0648c
“Dirac operators in differential and non-commutative geometry”, Mathe-
matisches Forschungsinstitut Oberwolfach. Intended for non-specialists, it
explores the spectrum of the fundamental Dirac operator on Riemannian
spin manifolds, including recent research and open problems. No background
in spin geometry is required; nevertheless the reader is assumed to be famil-
iar with basic notions of differential geometry (manifolds, Lie groups, vector
and principal bundles, coverings, connections, and differential forms). The
surveys [41, 132], which themselves provide a very good insight into closed
manifolds, served as the starting point. We hope the content of this book
reflects the wide range of findings on and sometimes amazing applications of
the spin side of spectral theory and will attract a new audience to the topic.
vii
Acknowledgements
ix
Contents
xi
xii Contents
Bibliography . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 145
Index . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 155
Introduction
xiii
xiv Introduction
In this chapter we define spin structures, spinors, the Dirac operator and
discuss the properties we need further on. Unless explicitly mentioned all
objects (manifolds, bundles, sections) will be assumed smooth in the whole
survey. For the thorough treatment of spin or spinc groups, spin or spinc
structures on vector bundles, representation theory of Clifford algebras and
Dirac operators on arbitrary semi-Riemannian Clifford modules we refer to
[63, 88, 173].
ξ
0 −→ Z2 −→ Spinn −→ SOn −→ 1,
η×ξ η M
Proof of Proposition 1.1.3: First recall that, for any Lie group G, there exists
a bijection between the set of equivalence classes of G-principal bundles over
some manifold N and the set H 1 (N, G) (which, if G is abelian, is the first
Čech-cohomology group with coefficients in G), see e.g. [173, App. A]. In
particular the set of two-fold coverings of - i.e., of Z2 -bundles over - N can
be identified with H 1 (N, Z2 ).
For n = 1 the result is a trivial consequence of this observation, since in
that case H 2 (M, Z2 ) = 0 and a spin structure is a 2-fold covering of the
manifold itself, hence R has exactly one and the circle 2 spin structures, see
also Example 1.4.3.1 below for a more precise description.
Assume for the rest of the proof n ≥ 2. First note that, from their
definition, the spin structures on (M n , g) coincide with the 2-fold cover-
ings of SO(T M ) which are non-trivial on each fibre of the projection map
SO(T M ) −→ M . This follows essentially from the standard lifting prop-
erty of maps through coverings. From the remark above, the set of 2-fold
coverings of SO(T M ) can be identified with H 1 (SO(T M ), Z2 ). Now the
1.1 Spin group and spin structure 3
π∗ ι∗ w
0 −→ H 1 (M, Z2 ) −→ H 1 (SO(T M ), Z2 ) −→ H 1 (SOn , Z2 ) −→ H 2 (M, Z2 )
Notes 1.1.4
1. In particular the existence of a spin structure does not depend on the
metric or the orientation of a given manifold. Actually, if the manifold
M is oriented, spin structures can be defined independently of any metric
(declare them to be non-trivial 2-fold coverings of the bundle of oriented
frames of T M ). In an equivalent way, a spin structure for a given met-
ric canonically induces a spin structure for another one. For a detailed
discussion of this point we refer to [173, Chap. 2], [10] and to [62].
2. Not every orientable manifold is spin. On surfaces the spin condition is
equivalent to the vanishing of the mod 2 reduction of the Euler class, thus is
fulfilled for orientable surfaces. In dimension 3, the second Stiefel-Whitney
class is the square of the first one [173, p.86], hence any 3-dimensional
orientable manifold is spin. The simplest counter-example comes up in
dimension 4: the complex projective plane CP2 is not spin (even if it
canonically carries a so-called spinc structure as a Kähler manifold). Indeed
a complex manifold is spin if and only if the mod 2 reduction of its first
Chern class vanishes, see [173, App. D].
3. However any simply-connected manifold has a unique spin structure as
soon as it is spin, since in that case H 1 (M, Z2 ) = 0.
4. It was first noticed by J. Milnor [193] that different spin struc-
tures may provide equivalent principal Spinn -bundles. For instance, the
2-dimensional torus has 4 different spin structures (see Example 1.4.3.2),
all of which are equivalent as principal Spin2 -bundles. This is due to the
fact that, for the torus, H 1 (M, Z2 ) = H 1 (M, Z) ⊗ Z2 , see [173, p.84].
From now on, each time we assume that a manifold is spin we shall implicitly
mean that a spin structure is fixed on it.
4 1 Basics of spin geometry
x · y + y · x = −2 x, y
1 (1.1)
1.2 Spinor bundle and Clifford multiplication 5
for all x, y ∈ V . Since we do not want to deal with Clifford algebras in detail
we just recall their most important properties for our purpose (see e.g. [173,
Chap. 1]):
1. The Clifford algebra is the smallest associative algebra with unit contai-
ning V and satisfying the Clifford relations.
2. The Clifford algebra of the n-dimensional Euclidean space is linearly iso-
morphic to its exterior algebra Λ∗ Rn , the Clifford product being then
given by
x· x ∧ − x (1.2)
for every x ∈ Rn .
3. If V is real then the complexification of the Clifford algebra of (V, · , ·
)
coincides with the (complex) Clifford algebra of (V ⊗ C, · , ·
C ) (where
· , ·
C is complex bilinear).
4. The Clifford algebra of C2n with its canonical complex bilinear form is
isomorphic to the algebra of all complex 2n × 2n matrices, and that of
C2n+1 to two copies of this algebra [173, Tab. I p.28].
Property 4 implies in particular the existence of exactly two and one
irreducible non-trivial representations of the Clifford algebra of Cn for n
odd and even respectively. Now Spinn can be identified with the set of
even Clifford products of unit vectors of Rn (this can actually be used
as definition of Spinn ), in particular sits in the Clifford algebra of Cn .
After restriction onto Spinn both Clifford-algebra-representations turn out
to become equivalent for n odd whereas the unique one splits into two
inequivalent equally dimensional representations of Spinn for n even. The
statement on their dimensions easily follows.
The simplest way to distinguish δn+ from δn− consists in looking at the action
of the so-called complex volume form of Rn and which is defined for every
n by
ωnC := i[ 2 ] e1 · . . . · en
n+1
(1.3)
for any positively-oriented orthonormal basis (e1 , . . . , en ) of Rn . The complex
volume form does in general not lie in Spinn but in the complex Clifford
algebra of Cn and it can be shown that
for n even whereas δn (ωnC ) = IdΣn or −IdΣn for n odd (in the latter case both
possibilities can occur).
ΣM := Spin(T M ) ×δn Σn ,
− −
where, for n even, Σn := Σ+ n ⊕ Σn and δn := δn ⊕ δn .
+
Σ+ M := Spin(T M ) ×δ+ Σ+
n
n
such that
X ·(Y · ϕ) + Y ·(X · ϕ) = −2g(X, Y )ϕ (1.5)
∇X
Σ
(Y · ϕ) = (∇X Y )· ϕ + Y · ∇X
Σ
ϕ
where ω = 1≤j1 <...<jp ≤n ωj1 ,...,jp e∗j1 ∧ . . . ∧ e∗jp in a local orthonormal ba-
sis {ej }1≤j≤n of T M . Moreover the Clifford algebra being associative (see
Property 1 in the proof of Proposition 1.2.1 above), we shall in the following
forget about the parentheses and write X ·Y · ϕ instead of X ·(Y · ϕ).
The spinor bundle comes with a natural Hermitian inner product which
together with Clifford multiplication exist and are unique in some sense.
1.2 Spinor bundle and Clifford multiplication 7
1
n
∇ϕα = g(∇ej , ek )ej · ek · ϕα , (1.6)
4
j,k=1
1
n
∇
RX,Y ϕ= g(RX,Y ej , ek )ej · ek · ϕ, (1.7)
4
j,k=1
∇ S
RX,Y ϕ= (X · Y − Y · X) · ϕ, (1.8)
8
n
1
∇
ej · RX,e ϕ= Ric(X) · ϕ, (1.9)
j=1
j
2
n
(1.7) 1
n
∇
ej · RX,e ϕ = g(RX,ej ek , el )ej · ek · el · ϕ
j=1
j
4
j,k,l=1
1 n
= − g(Rej ,ek X, el )ej · ek · el · ϕ
4
j,k,l=1
1 n
− g(Rek ,X ej , el )ej · ek · el · ϕ
4
j,k,l=1
1.3 The Dirac operator 9
1
n
= − g(RX,ej ek , el )(ek · el · ej − ek · ej · el ) · ϕ,
4
j,k,l=1
with
(1.5)
ek · el · ej − ek · ej · el = −ek · ej · el − 2δjl ek + ej · ek · el + 2δjk el
(1.5)
= 2ej · ek · el + 4δjk el − 2δjl ek .
We deduce that
n
n
1
n
∇
3 ej · RX,e ϕ=− g(RX,ej ej , el )el · ϕ + g(RX,ej ek , ej )ek · ϕ
j=1
j
2
j,l=1 j,k=1
n
1
n
= g(Ric(X), el )el · ϕ + g(Ric(X), ek )ek · ϕ
2
l=1 k=1
3
= Ric(X) · ϕ,
2
which is the result.
n
Dϕ := ej · ∇ej ϕ
j=1
for every ϕ ∈ Γ(ΣM ), where {ej }1≤j≤n is any local orthonormal basis of
TM.
D = D+ ⊕ D− ,
Proof : In even dimension the Clifford action of the complex volume form ωnC
is a non-trivial parallel involution of ΣM anti-commuting with the Clifford
multiplication with vectors (i.e., X · ωnC = −ωnC · X for every X ∈ T M ), so
that
D(ωnC · ϕ) = −ωnC · Dϕ (1.10)
We come to the properties of the Dirac operator which are fundamental for
the further study of its spectrum. First we need a formula computing the
commutator of the Dirac operator with a function. For technical reasons we
include into the next lemma the computation of commutators or anticommu-
tators involving the Dirac operator and which we shall need in the following.
i)
D(f ϕ) = grad(f ) · ϕ + f Dϕ, (1.11)
iii)
D2 (f ϕ) = f D2 ϕ − 2∇grad(f ) ϕ + (Δf )ϕ, (1.13)
where Δ := δd = −tr(Hessg (·)) denotes the scalar Laplace operator on
(M n , g).
Proof : Fix a local orthonormal basis {ej }1≤j≤n of T M . We compute:
n
D(f ϕ) = ej · ∇ej (f ϕ)
j=1
n
= ej · (ej (f )ϕ + f ∇ej ϕ)
j=1
= grad(f ) · ϕ + f Dϕ,
n
D(ξ · ϕ) = ej ·∇ej (ξ · ϕ)
j=1
n
n
= ej ·(∇ej ξ) · ϕ + ej ·ξ·∇ej ϕ
j=1 j=1
(1.5) n n
n
= ej ·(∇ej ξ) · ϕ − ξ · ej ·∇ej ϕ − 2 g(ξ, ej )∇ej ϕ
j=1 j=1 j=1
(1.2) n n
= ( ej ∧ ∇ej ξ )·ϕ − ( ej ∇ej ξ )·ϕ
j=1 j=1
n
n
−ξ · ej ·∇ej ϕ − 2 g(ξ, ej )∇ej ϕ
j=1 j=1
and
(1.11)
D2 (f ϕ) = D(df ·ϕ + f Dϕ)
(1.12)
= −df ·Dϕ − 2∇grad(f ) ϕ + (d + δ)df ·ϕ + df ·Dϕ + f D2 ϕ
= f D2 ϕ − 2∇grad(f ) ϕ + (Δf )ϕ,
where we have identified 1-forms with vector fields through the metric g.
This concludes the proof.
12 1 Basics of spin geometry
T ∗ M −→ End(ΣM )
ξ
−→ ξ
·,
n
Dϕ, ψ
= ej · ∇ej ϕ, ψ
j=1
n
=− ∇ej ϕ, ej · ψ
j=1
n
= −ej ( ϕ, ej · ψ
) + ϕ, ∇ej ej · ψ
j=1
n
+ ϕ, ej · ∇ej ψ
j=1
= div(Vϕψ ) + ϕ, Dψ
, (1.14)
(g ⊗ IdC )(Vϕψ , X) := ϕ, X · ψ
whole Hilbert space L2 (ΣM ), which is defined as the completion of the space
Γc (ΣM ) := {ϕ ∈ Γ(ΣM ) | supp(ϕ) is compact} w.r.t. (· , ·) := M · , ·
vg . The
concept needed here is that of essential self-adjointness: one has to show that
the closure of the operator in L2 (ΣM ) is self-adjoint. For D this is possible
as soon as the underlying Riemannian manifold is complete.
Proof : The proof presented here is based on the talk [43]. Let dom(·) de-
note the domain of definition of an operator. By construction the operator
D is densely defined in L2 (ΣM ) and one can set dom(D) := Γc (ΣM ).
As a consequence it admits a unique closure D∗∗ , whose graph is the
closure of that of D. Proposition 1.3.4 implies that D - thus D∗∗ - is
symmetric in L2 (ΣM ). Its adjoint D∗ is defined on {ψ ∈ L2 (ΣM ) | ϕ
→
(Dϕ, ψ) is bounded on (Γc (ΣM ), · )}. Since the topological dual of L2 is
L2 itself, dom(D∗ ) = {ψ ∈ L2 (ΣM ) | ϕ
→ (Dϕ, ψ) ∈ L2 (ΣM )}. Considering
D at the distributional level, one deduces from the formal self-adjointness of
D (Proposition 1.3.4) that
It remains to show that dom(D∗ ) = dom(D∗∗ ), i.e., the inclusion “⊂”. This
is equivalent to proving that Ker(D∗ − iεId) = 0 for both ε ∈ {±1}, see e.g.
[233, Cor. VII.2.9]. Let ψ ∈ dom(D∗ ) with D∗ ψ = iεψ, then (D∗ ψ, ϕ) =
iε(ψ, ϕ) for every ϕ ∈ Γc (ΣM ), i.e., (ψ, Dϕ) = iε(ψ, ϕ). Since D is formally
self-adjoint this is equivalent to Dψ = iεψ in the distributional sense. The
operator D − iεId being elliptic, general elliptic theory (see e.g. [173, Thm.
4.5 p.190]) implies that ψ is actually smooth. Let now (ρk )k>0 be a sequence
of smooth compactly supported [0, 1]-valued functions converging pointwise
to 1 and with |grad(ρk )| ≤ k1 for all k. Such a sequence exists because of the
completeness assumption on (M n , g). The preceding identity together with
Lemma 1.3.3 provide
iε(ψ, ρk ψ) = (Dψ, ρk ψ)
= (ψ, D(ρk ψ))
(1.11)
= (ψ, ρk Dψ) + (ψ, grad(ρk ) · ψ)
= (ρk ψ, Dψ) + (ψ, grad(ρk ) · ψ)
= −iε(ψ, ρk ψ) + (ψ, grad(ρk ) · ψ),
that is, 2iε(ψ, ρk ψ) = (ψ, grad(ρk ) · ψ), whose l.h.s. tends to 2iεψ2 and
whose r.h.s. tends to 0 when k goes to ∞. Therefore ψ = 0, QED.
14 1 Basics of spin geometry
Note 1.3.6 It seems at first glance that the completeness assumption on the
metric enters the proof in a very weak manner and one could therefore think
about getting rid of it. This is unfornutately hopeless. Consider for example
M :=]0, +∞[ with standard metric and canonical spin structure. Its Dirac op-
erator is D = i dtd
(the Clifford multiplication of e1 = 1 ∈ R can be identified
with that of i on Σ1 = C). Set dom(D) := Γc (ΣM ) = Cc∞ (]0, +∞[, C), then
the Dirac operator remains symmetric in L2 (ΣM ). However a simple com-
putation shows that the kernel of D∗ − iεId in the space of distributions is
Reεt . Since et ∈/ L2 (ΣM ) one has Ker(D∗ − iId) = 0, however e−t ∈ L2 (ΣM )
so that Ker(D∗ − iId) = Re−t . Therefore D is not essentially self-adjoint in
L2 (ΣM ). Actually the fact that Ker(D∗ ∓ iId) do not have the same dimen-
sion imply the non-existence of self-adjoint extensions of D in L2 (ΣM ) (see
[233, Thm. VII.2.10]).
Another example is M :=]0, 1[ with the same metric and spin structure:
d
again D = i dt is not essentially self-adjoint, but this time it has infinitely
many self-adjoint extensions. Indeed the exponential et belongs in that case
to L2 (ΣM ) so that Ker(D∗ − iεId) = Reεt for both ε ∈ {±1}. More-
over Ker(D∗ ∓ iId) have the same dimension and it can be shown that
there exists an S1 -parametrized family of self-adjoint extensions of D, see
[233, Ex. VII.2.a)].
We summarise the spectral properties of the Dirac operator on closed
manifolds. Further basics on non-compact Riemannian spin manifolds are
given in Section 7.1.
Theorem 1.3.7 Let (M n , g) be a closed Riemannian spin manifold and de-
note by Spec(D) the spectrum of its Dirac operator D, then the following
holds:
i) The set Spec(D) is a closed subset of R consisting of an unbounded dis-
crete sequence of eigenvalues.
ii) Each eigenspace of D is finite-dimensional and consists of smooth sec-
tions.
iii) The eigenspaces of D form a complete orthonormal decomposition of
L2 (ΣM ), i.e.,
L2 (ΣM ) = Ker(D − λId).
λ∈ Spec(D)
and analogously
e
j,k · ∇e
j,k
= χk ej,k · ∇ej,k
j,k j,k
= χk D
k
= D.
ej,k · ϕ), e
(D( j,k · ϕ) = − (
ej,k · Dϕ, e
j,k · ϕ)
j,k j,k
− 2(∇e
j,k
ϕ, e
j,k · ϕ)
ej,k · ϕ, e
+ ((d + δ) j,k · ϕ)
= −n(Dϕ, ϕ) + 2(Dϕ, ϕ)
− ej,k ∧ d
(( ej,k ) · ϕ, ϕ)
j,k
= nmϕ2 ,
From Corollary 2.1.5 below, the Dirac spectrum of the n-dimensional real
projective space RPn endowed with its round metric of sectional curvature 1
and one of its both spin structures is { n2 + n1 + 2k, − n2 − n2 − 2k, k ∈ N},
where n1 is the mod 2 reduction of n−3 n+1
4 and n2 that of 4 (here the multi-
plicities are not taken into account). Thus the symmetry property of Spec(D)
from Theorem 1.3.7.iv) breaks in dimension n ≡ 3 (4).
Eigenvectors for D associated to the eigenvalue 0 are called harmonic
spinors. Parallel spinors are harmonic but the converse is false in general.
Moreover, unlike that of harmonic forms, the number of linearly independent
harmonic spinors generally varies under a change of metric, see Section 6.2.
Turning to the square of the Dirac operator, its principal symbol is given by
ξ
→ −g(ξ, ξ)Id, which is exactly that of the rough Laplacian. The difference
between both must therefore be a linear differential operator of order at most
one. In fact, it turns out to be a very simple curvature expression.
S
D 2 = ∇∗ ∇ + Id, (1.15)
4
Proof : Fix a local orthonormal basis {ej }1≤j≤n of T M , then using the com-
patibility relations as well as (1.5) one has, for any ϕ ∈ Γ(ΣM ),
n
D2 ϕ = ej · ∇ej (ek · ∇ek ϕ)
j,k=1
1.3 The Dirac operator 17
n
= ej · ∇ej ek · ∇ek ϕ + ej · ek · ∇ej ∇ek ϕ
j,k=1
n
n
=− ej · ek · ∇∇ejek ϕ + ej · ek · ∇ej ∇ek ϕ
j,k=1 j,k=1
n
= (∇∇ejej − ∇ej ∇ej )ϕ
j=1
+ ej · ek · (∇ej ∇ek − ∇ek ∇ej − ∇∇ejek + ∇∇ekej )ϕ
1≤j<k≤n
= ∇∗ ∇ϕ − ej · ek · (∇[ej ,ek ] − [∇ej , ∇ek ])ϕ
1≤j<k≤n
1n
= ∇∗ ∇ϕ − ej · ek · Re∇j ,ek ϕ.
2
j,k=1
In particular the square of the Dirac operator coincides with the rough Lapla-
cian acting on spinors as soon as the scalar curvature of the underlying
manifold vanishes. This provides an answer to Dirac’s original question on
Euclidean space.
Applications of the Schrödinger-Lichnerowicz formula to eigenvalue es-
timates are discussed in Chapters 3 and 4. Vanishing theorems can also
be obtained combining the Schrödinger-Lichnerowicz formula with the cel-
ebrated Atiyah-Singer index theorem [28], stating that the topological and
the analytical index of any elliptic linear differential operator coincide. In the
case of the Dirac operator, the index theorem reads as follows.
Proposition 1.3.10 Let the spin structure be fixed and denote by D the
Dirac operator of M for the conformal metric g := e2u g, with u ∈ C ∞ (M, R).
Then there exists a unitary isomorphism between the spinor bundle of (M n , g)
and that of (M n , g) (denoted in the whole text by ϕ
→ ϕ) such that
D(e− ϕ) = e−
n−1 n+1
2 u 2 u Dϕ (1.16)
1 X(u)
∇X ϕ = ∇X ϕ − X · grad(u) · ϕ − ϕ. (1.17)
2 2
Since {e−u ej }1≤j≤n is a local o.n.b. of T M for g as soon as {ej }1≤j≤n is one
for g, we deduce that
n
Dϕ = e−2u ej · ∇ej ϕ
j=1
n
(1.17) 1 ej (u)
= e−2u ej · ∇ej ϕ − ej · grad(u) · ϕ − ϕ
j=1
2 2
1.4 Spinors on hypersurfaces and coverings 19
n
1 ej (u)
= e−u ej · ∇ej ϕ − ej · ej · grad(u) · ϕ − ej · ϕ
j=1
2 2
n−1
= e−u (Dϕ + grad(u) · ϕ). (1.18)
2
Hence
n − 1 − n−1 u
D(e− e−u (D(e−
n−1 n−1
2 u ϕ) = 2 u ϕ) + e 2 grad(u) · ϕ)
2
(1.11) n − 1 − n−1 u −u
e 2 e grad(u) · Dϕ + e− 2 u e−u Dϕ
n−1
= −
2
n − 1 − n−1 u −u
+ e 2 e grad(u) · ϕ
2
e− 2 u Dϕ,
n+1
=
| =
where, for n odd, the two copies of ΣM correspond to the splitting ΣM M
| ⊕ Σ− M
Σ+ M | . Moreover this isomorphism can be chosen so as to satisfy
M M
20 1 Basics of spin geometry
C
ωn · ϕ if n is even
M
iν · ϕ = 0 Id (1.19)
Id 0 , if n is odd
X · ϕ if n is even
X ·ν·ϕ= M (1.20)
(X · ⊕ − X · )ϕ if n is odd
M M
and
X ϕ = ∇X ϕ + A(X) · ν · ϕ
∇ (1.21)
2
for all X ∈ T M and ϕ ∈ Γ(ΣM | ), where ω C is the complex volume form
M n
(see (1.3) and A := −∇ν denotes the Weingarten endormorphism field of ι.
In particular the fundamental Dirac operators D and D of M and M
res-
pectively are related through
Dϕ ν ϕ + ν · (D2 − nH )ϕ,
=ν·∇ (1.22)
2
where
D if n is even
D2 :=
D ⊕ −D if n is odd
1
and H := n tr(A) is the mean curvature of ι.
, it is spin. This can be seen as
Proof : Since M has trivial normal bundle in M
a consequence of a more general result [194] or, alternatively, in the follwing
way: the pull-back of Spin(T M )| to SO(T M ) over the map “completion
M
by ν”
)|
SO(T M ) −→ SO(T M M
(e1 , . . . , en ) −→ (e1 , . . . , en , ν)
D2 (ν · ϕ) = −ν · D2 ϕ. (1.23)
)
Γ × Spin(T M )
Spin(T M
Id×η η
M
)
Γ × SO(T M )
SO(T M
∼ Γ\ΣM
ΣM =
Γ(ΣM ) ∼ ) |
= {ϕ ∈ Γ(ΣM ϕ(γ · x) = γ · ϕ(x) ∀x ∈ M, ∀γ ∈ Γ},
(1.24)
. In particular, the
where we also denote by “ · ” the action of Γ on ΣM
eigenvectors of the Dirac operator on M identify with those of the Dirac
operator on M satisfying (1.24).
is simply-connected, then the spin structures on M stand in one-
iii) If M
to-one correspondence with Hom(Γ, Z2 ).
22 1 Basics of spin geometry
) ×δ Σn ) = \ΣM
)) × Σ ∼ \(Spin(T M
ΣM = (Γ\Spin(T M .
δn n = Γ n Γ
For the same reason the Hermitian inner product of ΣM remains preserved
by Γ - hence the above identification can be assumed to be unitary - and so
does the Clifford multiplication: for all x ∈ M , X ∈ Tx M , ϕ ∈ Σx M and
γ ∈ Γ, γ · (X · ϕ) = (dγ(X)) · (γ · ϕ). The equivariance condition follows from
this observation. In the case where M is simply-connected one has
Examples 1.4.3
1. Let M := R and Γ := 2πZ acting on M by translations. Since Spin(T R) =
R × Spin1 = R × {±1} there are only two possible lifts of the Γ-action to
the spin level which are determined by the image (−1)δ of the generator
2π of Γ in {±1}, where δ ∈ {0, 1}. We call the spin structure induced by
δ = 0 the trivial spin structure on S1 = 2πZ\R and the one induced by
δ = 1 the non-trivial one.
2. More generally, consider a lattice Γ ⊂ M := Rn of rank n ≥ 1 and the
corresponding torus M := Γ\R with flat metric. The action of Γ by
n
odd. On the round sphere both bundles SO(T M ) and Spin(T M ) canoni-
cally identify to SOn+1 and Spinn+1 respectively. Therefore, the existence
of a lift of the action of Γ to the spin level is equivalent to that of a group ho-
momorphism : Γ −→ Spinn+1 such that ξ ◦ is the inclusion Γ ⊂ SOn+1 .
If this is fulfilled then from Proposition 1.4.2.iii) there are as many spin
structures on Γ\S as there are group homomorphisms Γ −→ {±1}. For
n
the eigenvalues lying in the interval [c, +∞[ (resp. ]c, +∞[, ] − ∞, c[, ] − ∞, c],
[c, d]). Since BgAPS = π≥β , we have, for β ≤ 0 and all ϕ, ψ ∈ Γ(ΣM ) satisfying
BgAPS (ϕ|∂M ) = BgAPS (ψ|∂M ) = 0:
(ϕ, ν · ψ)∂M = π<β (ϕ), π<β (ν · ψ)
∂M
(1.23)
= (π<β (ϕ), ν · π>−β (ψ)
∂M
0
= 0.
from which one deduces that ψ = 0 hence ϕ = 0. This shows the ellipticity
of the mgAPS boundary condition.
Let now ϕ, ψ ∈ Γ(ΣM ) satisfying BmgAPS (ϕ|∂M ) = BgAPS (ψ|∂M ) = 0,
then with the notations introduced above for the gAPS boundary condition
and for β ≤ 0 one has:
2(ϕ, ν · ψ)∂M = {Id + ν·}ϕ, {Id + ν·}ν · ψ
∂M
= π<β ({Id + ν·}ϕ), π<β ({Id + ν·}ν · ψ)
∂M
(1.23)
= π<β ({Id + ν·}ϕ), ν · π>−β ({Id + ν·}ψ )
∂M
0
= 0,
which shows that the spectrum of D under the mgAPS boundary condition
is real and concludes the proof.
Chapter 2
Explicit computations of spectra
In this chapter we present the few known closed Riemannian spin mani-
folds whose Dirac spectrum - or at least some eigenvalues - can be explicitly
computed.
1
n
∗
± 2π|γ + δj γj∗ |, γ ∗ ∈ Γ∗ ,
2 j=1
where Γ∗ := {θ ∈ (Rn )∗ | θ(Γ) ⊂ Z} is the dual lattice and (γ1∗ , . . . , γn∗ ) the
basis of Γ∗ dual to (γ1 , . . . , γn ). Furthermore, if non-zero, the eigenvalue pro-
vided by γ ∗ has multiplicity 2[ 2 ]−1 . In case δ1 = . . . = δn = 0 the multiplicity
n
[n ]
of the eigenvalue 0 is 2 2 .
Beware that the multiplicities add if the corresponding eigenvalues
n are
equal. Thus, the multiplicity of the eigenvalue ±2π|γ ∗ + 12 j=1 δj γj∗ |
n
is always atleast 2[ 2 ] (even for n = 1): if it is non zero, then
γ ∗ := −γ ∗ − j=1 δj γj∗ ∈ Γ∗ provides the same eigenvalue and γ ∗ = γ ∗ .
n
∇X φγ,l = X(φγ,l )
= 2iπθγ (X)e2iπθγ σl
= 2iπθγ (X)φγ,l ,
n
Dφγ,l = ek · ∇ek φγ,l
k=1
n
= 2iπ θγ (ek )ek · φγ,l
k=1
= 2iπθγ · φγ,l .
Notes 2.1.2
1. For n = 1, Theorem 2.1.1 reads as follows: the Dirac spectrum of the
circle S1 (L) of length L > 0 (for L = 2π we just write S1 ) and the δ-spin
structure, where δ ∈ {0, 1}, is 2π L ( 2 + Z). Furthermore, each eigenvalue is
δ
simple.
2. It is remarkable that the kernel of the Dirac operator of (Tn , gflat ) is not
reduced to 0 for the (0, . . . , 0)-spin structure (called sometimes the triv-
ial spin structure) whereas it is for all other ones. Flat tori are thus the
most simple-minded examples of closed manifolds with non-zero harmonic
spinors for some spin structure. Moreover, as already noticed in the proof of
Theorem 2.1.1, the kernel of D actually consists of parallel spinors. There-
n
fore, (Tn , gflat ) admits a 2[ 2 ] -dimensional space of parallel spinors for the
trivial spin structure and no non-zero one otherwise. The reader interested
in basic results as well as the classification of complete Riemannian spin
manifolds with parallel spinors should refer to Section A.4.
Only few spectra of closed flat manifolds are known, although such manifolds
are always covered by flat tori as a consequence of Bieberbach’s theorems.
This is due to the high complexity of the groups involved, which makes
the search for equivariant eigenvectors very technical. Up to now, only di-
mension 3 (F. Pfäffle [206]) and some particular cases in higher dimensions
(R. Miatello and R. Podestá [188]) have been handled using representation-
theoretical methods, see Section 2.2.
A rather different technique leads to the Dirac spectrum of round spheres.
Theorem 2.1.3 (S. Sulanke [226], see also [38, 72, 228, 229])
Consider, for n ≥ 2, the round sphere M = Sn := {x ∈ Rn+1 , |x| = 1} with
its canonical metric g of constant sectional curvature 1 and its canonical
spin structure.
32 2 Explicit computations of spectra
Then the spectrum of the Dirac operator is {±( n2 + k), k ∈ N} and each
n n + k − 1
eigenvalue ±( n2 + k) has multiplicity 2[ 2 ] · .
k
Proof : We present here C. Bär’s proof [38, Sec. 2], which has the advantage
to get to the result in a very elementary way. It is based on the knowledge of
the spectrum of the scalar Laplacian and on the trivialization of the spinor
bundle of Sn through either − 12 - or 12 -Killing spinors, see Example A.1.3.2.
Let ϕ be a non-zero 2ε -Killing spinor on (Sn , can), where ε ∈ {±1}. Then
n ∞ n
Dϕ = 2ε j=1 ej · ej · ϕ = − nε 2 ϕ, so that, for every f ∈ C (S , R),
(1.13)
D2 (f ϕ) = f D2 ϕ − 2∇grad(f ) ϕ + (Δf )ϕ
n2
= f ϕ − εgrad(f ) · ϕ + (Δf )ϕ
4
2
(1.11) n
= f ϕ − ε(D(f ϕ) − f Dϕ) + (Δf )ϕ
4
n2 n
= ( − )f ϕ − εD(f ϕ) + (Δf )ϕ,
4 2
2
that is, ((D + 2ε Id)2 − 14 Id)(f ϕ) = ( n4 − n2 )f ϕ + (Δf )ϕ, or, equivalently,
ε n−1 2
(D + Id)2 (f ϕ) = (Δf + ( ) f )ϕ. (2.2)
2 2
The spectrum of the scalar Laplacian Δ on (Sn , can) is given by (see e.g.
[74]) {k(n + k − 1) | k ∈ N}, where the eigenvalue k(n + k − 1) appears with
n + k − 1
n+k−1 ·
multiplicity n+2k−1
k
. Since the spinor bundle ΣSn of Sn is tri-
vialized by 2ε -Killing spinors one deduces from (2.2) that, if {fk }k∈N denotes
a L2 -orthonormal basis of eigenfunctions of Δ on Sn and {ϕj }1≤j≤2[ n2 ] a tri-
vialization of ΣSn through a pointwise orthonormal basis, then {fk ϕj | k ∈ N,
n
1 ≤ j ≤ 2[ 2 ] } provides a complete L2 -orthonormal basis of L2 (ΣSn ) made out
of eigenvectors of (D + 2ε Id)2 associated to the eigenvalues ( n−1 2
2 + k) with
n n + k − 1
k ∈ N, each of those having multiplicity 2[ 2 ] · n+2k−1
n+k−1 · . Therefore
k
the spectrum of D is contained in − 2 ± ( 2 + N), where the multiplicities
ε n−1
ε n−1 ε n−1
Ker((D + Id)2 − ( + k)2 Id) = Ker(D + ( − − k)Id)
2 2 2 2
ε n−1
Ker(D + ( + + k)Id)
2 2
that, for every k ∈ N,
n + 2k − 1 n+k−1
m(λ±
k) + m(λ±
−k−1 )
[n
=22] · · .
n+k−1 k
n+k−1
Next we show by induction on k that m(λ±
k) =2 · [n
2] .
k
− n
For k = 0, both λ+ 0 and λ0 have multiplicity 2
[2]
because, as we have seen
ε
above, 2 -Killing spinors are eigenvectors for D associated to the eigenvalue
±
− nε
2 . Alternatively the eigenvalues λ−1 = ∓( 2 − 1) cannot appear because
n
+∞
n
F± (z) := m(±( + k))z k .
2
k=0
Theorem 2.1.4 (C. Bär [38]) For n ≥ 3 odd, let (M n , g) := (Γ\S , can)
n
1 χ∓ ((γ)) − χ± ((γ)) · z
F± (z) = ,
|Γ| det(IdRn+1 − z · γ)
γ∈Γ
± ±
where χ± := tr(δn+1 ) : Spinn+1 −→ C is the character of δn+1 .
The proof of Theorem 2.1.4 relies on a similar formula for the Laplace
eigenvalues by A. Ikeda, see [38, Sec. 3]. As an application of Theorem 2.1.4,
one obtains the Dirac spectrum of real projective spaces. We keep the nota-
tions of Example 1.4.3.3.
n n−3 n n+1
{ + k, k ∈ N ∩ (δ + + 2Z)} ∪ {− − k, k ∈ N ∩ (δ + + 2Z)},
2 4 2 4
n n + k − 1
each eigenvalue corresponding to k having multiplicity 2[ 2 ] · .
k
±
Proof : On the one hand, χ± ((Id)) = tr(δn+1
n−1
(1)) = 2 2 , on the other hand,
(1.4) n+1
= ±(−1)δ+ 4 tr(IdΣ± n
)
n+1 n−1
= ±(−1)δ+ 4 2 2 ,
1 χ∓ ((γ)) − χ± ((γ)) · z
F± (z) =
2 det(IdRn+1 − z · γ)
γ∈Γ
= +
2 (1 − z)n+1 (1 + z)n+1
2.2 Spectrum of some other homogeneous spaces 35
n−1
1 (−1)δ+ 4
n+1
=2 2 ∓
2(1 − z)n 2(1 + z)n
n+1
n−1
+∞
1 ∓ (−1)k+δ+ 4 n+k−1
=2 2 zk ,
2 k
k=0
Let us first introduce a few notations and recall basic facts. Denote by
M := G/H an n-dimensional homogeneous space and by g (resp. h) the
Lie algebra of G (resp. of H). In that case the existence as well as the
set of spin structures on M can be read off the isotropy representation of
M , which is defined as the Lie-group-homomorphism α : H → GL(g/h) in-
duced by the restriction of the adjoint map Ad of G onto H. It is indeed
well-known that M carries a homogeneous Riemannian metric (i.e., a metric
invariant under the left G-action) or an orientation if and only if α(H) is
compact or connected respectively. Assuming both conditions α becomes a
map H −→ SO(g/h). The existence of a homogeneous spin structure (i.e., a
spin structure on which the left G-action on SO(T M ) lifts) on M is then
of α into Spin(g/h), the set of spin structures
equivalent to that of a lift α
standing then in one-to-one correspondence with that of such α ’s, that is,
with the set of group-homomorphisms H −→ {±1}, see [34, Lemma 1]. For
example, if H is connected, then there can exist only one homogeneous spin
structure on M . Moreover, if G is simply-connected, then all spin structures
are obtained in such a way.
To describe the Dirac operator on M , one has to look at the left action of
G onto the space of sections of ΣM , which can be identified with the space of
equivariant maps G −→ Σn . This left action induces a unitary representation
of G onto the space of L2 sections on M , which can be split into irreducible
and finite-dimensional components since G is compact. The Dirac operator
preserves that splitting and can be determined with the help of the following
result based on the Frobenius reciprocity principle (see reference in [34]).
Theorem 2.2.1 (see e.g. [34]) Let M := G/H be an n-dimensional Rie-
mannian homogeneous spin manifold with G compact and simply-connected.
36 2 Explicit computations of spectra
1. The space L2 (M, Σα M ) splits under the unitary left action of G into a
direct Hilbert sum
Vγ ⊗ HomH (Vγ , Σn )
γ∈G
n
Dγ (A) = − ek · A ◦ de γ(Xk )
k=1
⎛ ⎞
n
+⎝ βi ei + αijk ei · ej · ek ⎠ · A, (2.3)
i=1 i<j<k
and
1
n
βi := [Xj , Xi ]p , Xj
2 j=1
1
αijk := ( [Xi , Xj ]p , Xk
+ [Xj , Xk ]p , Xi
+ [Xk , Xi ]p , Xj
)
4
then the round sphere Sn is the total space of an S1 -bundle called the Hopf-
n−1 n−1
fibration Sn −→ CP 2 , where CP 2 is the complex projective space of
n−1
complex dimension 2 . In particular one may decompose the round metric
on Sn along the S1 -fibres and their orthogonal complement; we denote this
decomposition by can = gS1 ⊕ g⊥ .
Theorem 2.2.2 (C. Bär [39]) For n = 2m + 1 ≥ 3, let M n := Sn with
Berger metric gt = tgS1 ⊕ g⊥ for some real t > 0. Then the Dirac operator of
(M n , gt ) has the following eigenvalues:
m + k
i) 1 m+1
t( 2 + k) + tm
2 , k ∈ N, with multiplicity .
k
1 m + k
ii) (−1)m+1 t(
m+1
2 + k) + tm
2 , k ∈ N, with multiplicity .
k
iii) (−1)j 2t ± [ 2t (m − 1 − 2j) + 1t (a1 − a2 + m−1
2 − j)] + 4(m + a1 − j)
2
(1 + a2 + j), a1 , a2 ∈ N, j ∈ {0, 1 . . . , m − 1}, with multiplicity
(m+a1 )!(m+a2 )!(m+1+a1 +a2 )!
a1 !a2 !m!j!(m−1−j)!(m+a1 −j)(1+a2 +j) .
Γ = < A1 , B1 , . . . Ag , Bg , X1 , . . . , Xr |
X1m1 = . . . = Xrmr = X1 . . . Xr A1 B1 A−1 −1 −1 −1
1 B1 . . . Ag Bg Ag Bg = 1 > .
Theorem 2.2.3 (J. Seade and B. Steer [218]) Let M := PSL2 (R)/Γ,
where Γ is a co-compact Fuchsian subgroup of signature (g; m1 , . . . , mr )
38 2 Explicit computations of spectra
mj −1
2k − 1 r
1 sin((2k − 1)mj π)
Vol(M ) − lπ
.
2π j=1
mj sin( m )
l=1 j
iii) − 2t ± (2n − 1)2 (1 + t−2 ) − (2k − 1)2 , k ≥ 1, n > k, with the same
multiplicity
as in ii).
iv) − 2t ± (2n − 1)2 (1 + t−2 ) − s2 , n ∈ Z, s ∈ Λ, where Λ is some countable
subset of ] − 1, 1[∪iR.
The set Λ depends on representation-theoretical data and cannot be ex-
plicited in general. For t = 1, the metric gt has constant negative sectional
curvature. This is up to the knowledge of the author the only compact hy-
perbolic manifold whose Dirac spectrum has been computed.
To summarise, we reproduce - with small changes - the list in [42] of all
homogeneous spaces of which Dirac spectrum has already been computed:
space description references
G simply-connected compact Lie groups [83]
Rn/Zn flat tori [86]
R3/Γ 3-dim. (flat) Bieberbach manifolds [206]
Rn/Γ some n-dim. (flat) Bieberbach manifolds [188]
Sn round spheres [226]
Sn/Γ spherical spaceforms [38]
S2m+1 spheres with Berger metric [148, 39]
S3/Z 3-dim. lens spaces with Berger metric [34]
k
S3/Q S through the group of quaternions,
3
[103]
8
with Berger metric
H 3/Γ 3-dim. Heisenberg manifolds [15]
PSL2 (R)/Γ (Γ Fuchsian) [218]
CP2m+1 complex projective spaces [68, 69, 222]
HPm quaternionic projective spaces [66, 189]
OP2 Cayley projective plane [234]
Gr2 (R2m ) real 2-Grassmannians [224, 225]
Gr2p (R2m ) real 2p-Grassmannians [219]
Gr2 (Cm+2 ) complex 2-Grassmannians [190]
G2/SO - [219, 220]
4
2.3 Small eigenvalues of some symmetric spaces 39
Theorem 2.3.1 has been applied by J.-L. Milhorat in [192] to compute the
smallest eigenvalue λ1 (D2 ) for the following symmetric spaces (where S de-
notes the scalar curvature of M and Ep the exceptional simple Lie group of
rank p):
m2 +6m−4 m2 +6m−4
Spinm+4/(Spin · Spin )
m 4
4m m(m+2) · m
2 = m(m+2) · S
4
(m even)
E6/(SU · SU )
6 2
40 41
6 = 41
30 · S
4
E7/(Spin
12 · SU2 ) 64 95
9 = 95
72 · S
4
E8/(E · SU )
7 2
112 269
15 = 269
210 · S
4
The quotient S4 has been each time factorized out in order to compare the
n
dimension-depending coefficient standing before with n−1 , which is the cor-
responding one in Friedrich’s inequality (3.1).
Chapter 3
Lower eigenvalue estimates
on closed manifolds
The most general sharp lower bound for the Dirac spectrum has been proved
by T. Friedrich in [85] and is now known under the name “Friedrich’s in-
equality”. For the concept of Killing spinor we refer to Section A.1.
1 ∗ 1 ∗
n n
∇ϕ = ∇ϕ + ej ⊗ ej · Dϕ − e ⊗ ej · Dϕ,
n j=1 n j=1 j
=P ϕ ∈ Ker(μ) ∈ Ker(μ)⊥
that is,
n n
|Dϕ| −
2 2
S|ϕ| vg = |P ϕ|2 vg . (3.3)
M 4(n − 1) n−1 M
Choose ϕ to be a non-zero eigenvector for D associated to the eigenvalue λ.
From |P ϕ|2 ≥ 0 one straightforward obtains the inequality (3.1).
If (3.1) is an equality for some eigenvalue λ then (3.3) implies P ϕ = 0
for any non-zero eigenvector ϕ for D associated to λ, hence any such ϕ is
a (necessarily real) Killing spinor on (M n , g). Conversely, if (M n , g) carries
a non-zero α-Killing spinor ϕ, then since M is compact α must be real.
Moreover, on the one hand ϕ is an eigenvector for D associated to the
eigenvalue −nα, on the other hand we know from Proposition A.4.1 that the
scalar curvature of (M n , g) must be S = 4n(n − 1)α2 , in particular it must
be non-negative. Therefore
such
a ϕ must be an eigenvector for D associated
to the eigenvalue 4(n−1) or − 4(n−1)
nS nS
. This shows the equivalence in the
limiting-case and concludes the proof.
Another method for the proof of (3.1), which is actually T. Friedrich’s in [85],
relies on the modified connection
X ψ := ∇X ψ + λ X · ψ
∇
n
for every X ∈ T M , where Dψ = λψ: Compute |∇ψ| 2 (which plays the role
of |P ψ| above), integrate and apply the Schrödinger-Lichnerowicz formula.
2
Alternatively but still along the same idea, √ it can be directly deduced from
the Cauchy-Schwarz inequality that |Dϕ| ≤ n|∇ϕ| for every section ϕ, from
which (3.1) follows.
As a consequence of Theorem 3.1.1, if the scalar curvature S of (M n , g) is
positive then its Dirac operator has trivial kernel - whatever the spin structure
is. This had been already noticed by A. Lichnerowicz in [175] where he had
obtained as a straightforward application of (3.2) the following estimate:
1
λ2 ≥ inf (S). (3.4)
4 M
3.2 Improving Friedrich’s inequality in presence of a parallel form 43
It follows from (3.4) combined with the Atiyah-Singer index theorem [28]
(see Theorem 1.3.9) that a Riemannian manifold with positive scalar cur-
vature must have vanishing topological index. In particular, if the manifold
has non-vanishing A-genus, then it cannot carry any Riemannian manifold
with positive scalar curvature. In other words, there exists a topological ob-
struction to the existence of metrics with positive scalar curvature on closed
spin manifolds, at least in even dimensions. The reader interested in further
results in that topic - such as Gromov-Lawson’s work - should refer to [173]
or to [124]. The existence of Riemannian metrics for which the Dirac kernel
is non-zero is discussed in Section 6.2. Moreover, we mention another closely
related application of the Atiyah-Singer index theorem to geometry via the
Schrödinger-Lichnerowicz formula, namely to the so-called scalar curvature
rigidity issue which asks for the possibility of increasing the scalar curvature
without shrinking the distances of a given metric on a fixed background man-
ifold. For example this is not possible on the round sphere (M. Llarull [180,
Thm. B]) nor on any connected closed Kähler manifold with non-negative
Ricci curvature (S. Goette and U. Semmelmann [109, Thm. 0.1]), we refer to
[110] for the case of symmetric spaces and references.
Although it requires the non-negativity of S to be non-trivial, Friedrich’s
inequality (3.1) provides fine information of geometrical nature on the Dirac
spectrum. Indeed S stands for a very weak curvature invariant of a given
Riemannian manifold. This shows for example a difference of behaviour with
other differential operators such as the scalar Laplacian Δ: by a result of A.
Lichnerowicz [174], any non-zero eigenvalue λ of Δ satisfies
n
λ≥ inf (Ric),
n−1 M
where Ric is the Ricci curvature tensor of (M n , g), which is a stronger cur-
vature invariant. In case inf M (S) ≤ 0 Friedrich’s inequality (3.1) can be
improved in different ways using various techniques, see Sections 3.3 to 3.7.
Besides, (3.1) is sharp since e.g. M := Sn (n ≥ 2) admits non-zero Killing
spinors, see Example A.1.3.2. For the classification of Riemannian spin man-
ifolds carrying non-zero real Killing spinors we refer to Theorems A.4.2 and
A.4.3 in Appendix A.
O. Hijazi [128] and A. Lichnerowicz [176, 177] noticed that equality in (3.1)
cannot hold on those M admitting a non-zero parallel k-form for some k ∈
{1, . . . , n − 1}. This suggests (3.1) could be enhanced under this assumption.
44 3 Lower eigenvalue estimates on closed manifolds
The idea of proof for Theorems 3.2.1, 3.2.4 and 3.2.6 can be summarised as
follows (see [64]): given any eigenvector ϕ of D to the eigenvalue λ, decompose
|∇ϕ|2 in a sharper way than for the proof of Friedrich’s inequality, using the
splitting of ΣM induced by the Clifford action of the parallel form.
n−1
λ2 ≥ inf (S), (3.5)
4(n − 2) M
Proof : Let ξ be the dual vector field to the harmonic 1-form of constant
length. We may assume that g(ξ, ξ) = 1 on M . Define the following Penrose-
like operator
1 1
TX ϕ := ∇X ϕ + (X − g(X, ξ)ξ) · Dϕ − (ng(X, ξ) + X · ξ·)∇ξ ϕ
n−1 n−1
n
|T ϕ|2 = |Tej ϕ|2
j=1
1 n
= |∇ϕ|2 + |Dϕ|2 + |∇ξ ϕ|2
n−1 n−1
3.2 Improving Friedrich’s inequality in presence of a parallel form 45
2
− (|Dϕ|2 + e ( ξ · Dϕ, ∇ξ ϕ
))
n−1
2
− (n|∇ξ ϕ|2 − e ( ξ · ∇ξ ϕ, Dϕ
))
n−1
2
− e ( ξ · ∇ξ ϕ, Dϕ
))
n−1
1 n 2
= |∇ϕ|2 − |Dϕ|2 − |∇ξ ϕ|2 + e ( ξ · ∇ξ ϕ, Dϕ
)).
n−1 n−1 n−1
Now we can express the last term on the r.h.s. through the other ones, a trick
due to the authors of [202]: namely, since ξ is assumed to be harmonic, i.e.,
closed and co-closed, the identity (1.12) reads D(ξ · ϕ) = −ξ·Dϕ − 2∇ξ ϕ and
hence
1
e ( ξ · ∇ξ ϕ, Dϕ
) = |∇ξ ϕ|2 + (|Dϕ|2 − |D(ξ · ϕ)|2 ),
4
from which we obtain
1 n−2 1
|T ϕ|2 = |∇ϕ|2 − |Dϕ|2 − |∇ξ ϕ|2 + (|Dϕ|2 − |D(ξ · ϕ)|2 ).
n−1 n−1 2(n − 1)
But choosing ϕ to be eigen for D for the smallest (in absolute value) eigen-
value λ, the min-max principle (see Lemma 5.0.2) implies
|D(ξ · ϕ)| vg ≥ λ
2 2
|ξ · ϕ|2 vg
M
M
= λ2 |ϕ|2 vg
M
= |Dϕ|2 vg ,
M
hence
n−2 2 1 n−2
( λ − inf (S)) |ϕ|2 vg ≥ |T ϕ|2 vg + |∇ξ ϕ|2 vg (3.6)
n−1 4 M M M n−1 M
λ
∇X ϕ = − (X − g(X, ξ)ξ) · ϕ (3.7)
n−1
1 n
∇
Ric(ξ) · ϕ = ej · Rξ,e ϕ
2 j=1
j
λ
n
= (g(ξ, ∇ej ξ) − g(ej , ∇ξ ξ))ej · ξ
n − 1 j=1
+ (g(ξ, ξ)∇ej ξ − g(ej , ξ)∇ξ ξ) · ej · ϕ
λ n
= (−2∇ξ ξ · ξ + ∇ej ξ · ej ) · ϕ.
n−1 j=1
The last sum vanishes because of (1.2) together with ξ being closed and co-
closed. On the other hand dξ = 0 means that g(∇X ξ, Y ) − g(∇Y ξ, X) = 0 for
all X, Y ∈ T M , hence for X = ξ one obtains - using once again g(ξ, ξ) = 1
- that g(∇ξ ξ, Y ) = 0 for all Y ∈ T M , i.e., ∇ξ ξ = 0. This shows Ric(ξ) = 0.
From Bochner’s formula for the Laplace operator on 1-forms (see e.g. [173,
Cor. 8.3 p.156]) one deduces that ∇ξ = 0, i.e., that ξ is parallel.
We now prove the limiting-case in Theorem 3.2.1. If ξ is parallel then the
universal cover of M must be a Riemannian product of the form R × N .
W.r.t. the pull-back spin structure the lift of ϕ to R × N also satisfies (3.7)
provided ξ is replaced by ∂t ∂
. Since each {t} × N sits totally geodesically in
3.2 Improving Friedrich’s inequality in presence of a parallel form 47
R × N the Gauss-type formula (1.21) implies that the induced spinor field
on N is a real Killing spinor for one of the constants ± n−1
λ
.
carries the trivial spin structure (i.e., the spin structure induced by the triv-
ial lift of the Z3 -action to the spin level, see Proposition 1.4.2). In fact (3.5)
is an equality if and only if there exists a π1 (M )-equivariant solution - in the
sense of (1.24) - to (3.7) on the universal cover R × N of M .
Although the (real) Killing-spinor-equation is completely understood (see
Theorems A.4.2 and A.4.3), the list of all local Riemannian products on
which (3.5) is sharp is not entirely known. B. Alexandrov, G. Grantcharov
and S. Ivanov have shown in [5] that, under this assumption and if n = 7 is
odd, then M is diffeomorphic - but not necessarily isometric - to S1 × Sn−1 .
It is moreover important to note that the hypothesis in Theorem 3.2.2 on
the length being constant cannot be removed: C. Bär and M. Dahl showed in
[46] that in dimension n ≥ 3 Friedrich’s inequality (3.1) cannot be improved
with the help of topological assumptions. Namely there exists on any given
compact spin manifold M n admitting a metric with positive scalar curvature
a smooth family of Riemannian metrics (gt )t>0 with Sgt ≥ n(n − 1) and
n2 n2
≤ λ1 (Dg2t ) ≤ + t,
4 4
where Dgt stands for the Dirac operator to the metric gt on M . In other words,
one can get as close as one wants to the equality in Friedrich’s inequality
(3.1) on any such manifold. Note that the set of compact spin manifolds
with positive first Betti number and admitting a metric with positive scalar
curvature is non-empty since it contains e.g. S1 × Sn−1 , n ≥ 3.
The generalization of Theorem 3.2.1 to locally reducible Riemannian man-
ifolds was achieved by B. Alexandrov, extending earlier work by E.C. Kim
[154]:
k
TM = Tj ,
j=1
k
1
|T ϕ|2 = |∇ϕ|2 − |D[j] ϕ|2 .
j=1
nj
k
On the other hand, it is an exercise to show that D2 = j=1 D[j] 2
and that
D[j] is formally self-adjoint, so that, after integration and application of the
Schrödinger-Lichnerowicz formula (1.15), one obtains
nk k
nk − nj nk
Dϕ2 = T ϕ2 + D[j] ϕ2 + (Sϕ, ϕ). (3.9)
nk − 1 j=1
nj (nk − 1) 4(nk − 1)
Ω(X, Y ) := g(J(X), Y )
for all X ∈ T M if n
2 is odd and a non-zero section ψ of ΣM satisfying
2
D ψ = λ2 ψ
∇X ψ = − n1 (X + iJ(X)) · Dψ
(3.12)
Ω · ψ = −2iψ
Ω · Dψ =0
for all X ∈ T M if n
2 is even.
Proof : We follow the proof given in [131, 133], see also [223, Sec. 3] or [150].
We may assume that S > 0 on M (otherwise the estimate is trivial). Set, for
every X ∈ T M , p± (X) := 12 (X ∓ iJ(X)) ∈ T M ⊗ C. In the whole proof we
shall redenote m := n2 . Given a pointwise orthonormal basis (e1 , . . . , en ) of
T M such that ej+m = J(ej ) for every 1 ≤ j ≤ m, define zj := p+ (ej ) and
z j := p− (ej ) for all 1 ≤ j ≤ m. Then (z1 , . . . , zm ) and (z 1 , . . . , z m ) are bases
of T 1,0 M := p+ (T M ) and T 0,1 M := p− (T M ) respectively satisfying
zj · zk = −zk · zj , z j · z k = −z k · z j , zj · z k + z k · zj = −δjk
50 3 Lower eigenvalue estimates on closed manifolds
m
ΣM = Σr M. (3.13)
r=0
(r) 1 1
TX ϕ := ∇X ϕ + p− (X) · D+ ϕ + p+ (X) · D− ϕ
2(r + 1) 2(m − r + 1)
n
n
p+ (ej )·p− (ej )· = iΩ·−mId and p− (ej )·p+ (ej )· = −iΩ·−mId,
j=1 j=1
n n
in particular j=1 p+ (ej ) · p− (ej ) · ϕ = −2rϕ and j=1 p− (ej ) · p+ (ej ) · ϕ =
−2(m − r)ϕ. We deduce for the norms that
n
|T (r) ϕ|2 = |Te(r)
j
ϕ|2
j=1
n
= |∇ej ϕ|2
j=1
3.2 Improving Friedrich’s inequality in presence of a parallel form 51
n
1 1
+ |p− (ej ) · D+ ϕ|2 + |p (e ) · D− ϕ|2
2 + j
j=1
4(r + 1)2 4(m − r + 1)
n
1
+2 e ∇ej ϕ, p− (ej ) · D+ ϕ
j=1
2(r + 1)
n
1
+2 e ∇ej ϕ, p+ (ej ) · D− ϕ
j=1
2(m − r + 1)
n
1
+2 e p− (ej ) · D+ ϕ, p+ (ej ) · D− ϕ
j=1
4(r + 1)(m − r + 1)
= |∇ϕ|2
1 n
− p+ (ej ) · p− (ej ) · D+ ϕ, D+ ϕ
1 n
− p− (ej ) · p+ (ej ) · D− ϕ, D− ϕ
that is,
1 1
|T (r) ϕ|2 = |∇ϕ|2 − |D+ ϕ|2 − |D− ϕ|2 . (3.14)
2(r + 1) 2(m − r + 1)
S 1
T (r) ψ2 = (D2 ψ, ψ) − ( ψ, ψ) − Dψ2
4 2(r + 1)
2r + 1 2 1
=( λ − inf (S))ψ2 ,
2(r + 1) 4 M
2(r+1)
from which one deduces that λ2 ≥ 4(2r+1) inf M (S). The r.h.s. of that inequal-
ity decreases with r, so that it is bounded from below by the corresponding
expression for r = m−12 in case m is odd and for r = m−2 2 in case m is even.
Inequality (3.10) follows.
Assume now (3.10) to be an equality for some eigenvalue λ. If inf M (S) = 0
then (M n , g) has a non-zero parallel spinor, as already proved in Theorem
3.1.1. If inf M (S) > 0, then for any eigenvector ψ for D associated to the
eigenvalue λ, one has on the one hand ψ = ψ m−1 + ψ m+1 if m is odd and
2 2
ψ = ψ m−2 + ψ m2 + ψ m+2 if m is even, on the other hand T (r) ψr = 0 for
2 2
r = m±1 2 and r = m±22 for m odd and even respectively. Redenoting ψ m−1
2
by ψ and ψ m+1 by φ, we obtain (3.11) in case m is odd. If m is even then
2
redenoting ψ m−2 by ψ one obtains (3.12). Conversely, mimiking the proof
2
of Proposition A.4.1, it is elementary to show that, if (3.11) is satisfied by
some non-zero (ψ, φ), then ψ + φ is an eigenvector for D associated to the
2
eigenvalue λ and S = 4nλ n
n+2 , therefore (3.10) is an equality. Similarly, if 2 is
even and (3.12) is satisfied by some non-zero ψ, then the scalar curvature of
2
(M n , g) is equal to 4(n−2)λ
n , therefore (3.10) is an equality.
A pair of spinors (ψ, φ) satisfying (3.11) for some non-zero real number λ is
called a real Kählerian Killing spinor. As for Killing spinors (see Proposition
A.4.1), it is not too difficult to show that, if a non-zero real Kählerian Killing
spinor exist on a given complete Kähler spin manifold and associated to
some (non-zero) real λ, then this manifold has odd complex dimension and is
Einstein with positive scalar curvature (in particular it is closed). However,
the precise classification of those Kähler spin manifolds carrying non-trivial
real Kählerian Killing spinors is more technical, even if it turns out to provide
simpler results. The idea to achieve it, due to A. Moroianu [197], can be
summarised as follows: Show the existence of a suitable S1 -bundle over such
a manifold where the pull-back of the real Kählerian Killing spinor induces a
non-zero real Killing spinor; then show that, among the possible holonomies
listed in C. Bär’s classification (Theorem A.4.3), only those associated to a
so-called regular 3-Sasaki structure can occur on that S1 -bundle. We refer to
[197] for details and mention that, before [197] was published, partial results
had been obtained by K.-D. Kirchberg [156, 157, 158] and O. Hijazi [131],
see references in [197].
The even-complex-dimensional case turns out to be more involved since
the underlying manifold is no more Einstein. In dimension n = 4, argu-
ments from complex geometry and based on Kirchberg’s work [159, Thm. 15]
3.2 Improving Friedrich’s inequality in presence of a parallel form 53
Beware that, in case n2 is even, the Kähler manifold must not be holomorphi-
cally isometric to the quotient Γ\N × R endowed with the Kähler structure
2
n + 12
λ2 ≥ S, (3.15)
4(n + 8)
where S is the scalar curvature of M . Moreover, this inequality is an equality
for some eigenvalue λ if and only if (M n , g) is isometric to the quaternionic
n
projective space HP 4 .
Here it should be noticed that every quaternionic Kähler manifold of
even quaternionic dimension is spin whereas only the quaternionic projective
space is spin if n4 is odd; moreover, every quaternionic Kähler manifold is
Einstein, hence has constant scalar curvature, see references in [164].
Sketch of proof of Theorem 3.2.6: We follow the proof detailed in [164], which
relies on the representation theory of Sp1 × Spk . Denote n4 by m. A quater-
nionic structure on (M n , g) is given by a triple (I, J, K) of parallel orthogonal
endomorphisms of T M with I 2 = J 2 = K 2 = −IdT M and IJ = −JI = K.
Each of those endormorphisms is a Kähler structure on T M with associated
Kähler form, so that one may define the so-called fundamental form
Ω := ΩI ∧ ΩI + ΩJ ∧ ΩJ + ΩK ∧ ΩK
n n
|Dψ|2 − S |ψ|2 vg = |P ψ|2 vg ≥ 0.
M 4(n − 1) n−1 M
implies in turn
λe−u
∇X ψ = − X ·ψ
n
for every X ∈ T M . Elementary computations as in the proof of Proposition
A.4.1 but carried out on (M n , g) (see e.g. [134, Prop. 5.12]) show that
necessarily du = 0, thus u is constant and therefore ϕ is a real Killing spinor
on (M n , g). The converse statement follows from the characterization of the
equality case in (3.1). This concludes the proof.
2πχ(M 2 )
λ2 ≥ , (3.17)
Area(M 2 , g)
56 3 Lower eigenvalue estimates on closed manifolds
where μ1 denotes the first eigenvalue of the scalar conformal Laplace op-
erator 4 n−1
n−2 Δ + S. Moreover, (3.18) is an equality for some eigenvalue λ
of D if and only if (M n , g) carries a non-zero real Killing spinor.
Proof : We deduce both (3.18) and (3.17) from (3.16) and from the following
transformation formula for scalar curvature after conformal change of the
metric:
Se2u = S + 2(n − 1)Δu − (n − 1)(n − 2)|grad(u)|2 , (3.19)
for g := e2u g and u ∈ C ∞ (M, R). This is applied to a conformal metric g for
which Se2u is constant on M .
Svg
i) Let u0 ∈ C ∞ (M, R) solve Δu0 = 2Area(M M
2 ,g) − 2 (such a solution
S
4
where S is the scalar curvature of (M n , g := h n−2 g). Thus, choosing
4 4
g 0 := h0 g on M n , one obtains Sh0 = μ1 , which together with Theorem
n−2 n−2
Another proof of (3.18) involving Kato type inequalities can be found in [71].
Inequality (3.18) improves Friedrich’s inequality (3.1) for n ≥ 3 since ob-
viously μ1 ≥inf (S). It also proves the existence in dimension n ≥ 3 of an
M
explicit conformal lower bound for the spectrum of D2 :
Corollary 3.3.3 (O. Hijazi [130]) For any Riemannian metric g on a
closed n(≥ 3)-dimensional spin manifold M n ,
2 n
2
λ1 (DM,g )Vol(M, g) n ≥ Y (M, [g]), (3.20)
4(n − 1)
2
where λ1 (DM,g ) denotes the smallest non-negative eigenvalue of D2 associ-
ated to the metric g and Y (M, [g]) is the Yamabe invariant of M w.r.t. the
conformal class of g.
Proof : Recall that the Yamabe invariant of M n w.r.t. [g] is the conformal
invariant defined by
(4 n−1
M n−2 Δg f + Sg f )f vg
Y (M, [g]) := ∞
inf 2n n−2
.
f ∈C (M,R)\{0} ( M f n−2 vg ) n
2n n−2 2
Hölder’s inequality gives M f 2 vg ≤ ( M f n−2 vg ) n ·Vol(M n , g) n . Assuming
Y (M, [g]) > 0 (otherwise (3.20) is trivially satisfied), one obtains
2
μ1 Vol(M n , g) n ≥ Y (M, [g]),
Note however that (3.18) is not itself conformal. We also mention that inequa-
lity (3.18) can be combined with lower bounds of μ1 to provide an estimate
2
of λ1 (DM,g ) in terms of the total Q-curvature in dimension n = 4 and of the
first eigenvalue of the so-called Branson-Paneitz operator in dimension n ≥ 5,
see [145, Sec. 4]. The a priori existence of a qualitative conformal lower bound
for the Dirac spectrum was proved independently by J. Lott [181] using the
boundedness of particular Sobolev embeddings. More precisely, if the Dirac
operator of a given closed Riemannian spin manifold (M n , g) is invertible,
then there exists a positive constant c depending only on the conformal class
of g such that [181, Prop. 1]
2
2
λ1 (DM,g )Vol(M, g) n ≥ c (3.21)
where λ1 (ΔS2 ,g ) denotes the smallest positive eigenvalue of the scalar Laplace
operator Δ on (S2 , g). For lower bounds in higher genus, where (3.17) is
trivial, see Section 3.6.
As noticed in the proof of Corollary 3.3.2, one has from the Gauss-Bonnet
2πχ(M 2 ) 2 M2
Svg
Theorem in dimension 2 the following identity: Area(M 2 ,g) = 4(2−1) Area(M 2 ,g) .
The main idea to prove inequality (3.1) was to split the spinorial Levi-Civita
connection in a clever way so as to make the term which is dropped off after
integration and application of the Schrödinger-Lichnerowicz formula as small
as possible. This led to the introduction of the Penrose operator. In an equiv-
alent way, this means deforming the Levi-Civita connection in the direction
of IdT M , i.e., defining TX ϕ := ∇X ϕ + f X · ϕ for some real or complex-valued
function f to be fixed later, see T. Friedrich’s method of proof in Section 3.1.
O. Hijazi’s idea for the following result is to introduce a different Penrose-
like operator, deforming the Levi-Civita connection in the direction of some
symmetric 2-tensor tensor T ψ associated to an eigenvector ψ:
Theorem 3.4.1 (O. Hijazi [132]) Let λ be an eigenvalue of the fundamen-
tal Dirac operator on a closed n(≥ 2)-dimensional closed Riemannian spin
manifold (M n , g) and ψ be a non-zero eigenvector for D to the eigenvalue λ.
Then
S
λ ≥ inf
2
+ |T | ,
ψ 2
(3.23)
Mψ 4
where T ψ (X, Y ) := 12 e X · ∇Y ψ + Y · ∇X ψ, |ψ|
ψ
2
for all X, Y ∈ T M
and Mψ := {x ∈ M | ψ(x) = 0}. If furthermore (3.23) is an equality then ψ
solves
∇X ψ = −T ψ (X) · ψ (3.24)
3.4 Improving Friedrich’s inequality with the energy-momentum tensor 59
for all X ∈ T M .
on ΣM (and outside the
Proof : Define the following modified connection ∇
zero set of ψ, of which measure vanishes) by
X ψ := ∇X ψ + T ψ (X) · ψ
∇
n
2=
|∇ψ| e ψ|2
|∇ j
j=1
n
= |∇ej ψ|2 + |T ψ (ej )|2 |ψ|2 + 2e ∇ej ψ, T ψ (ej ) · ψ
j=1
n
= |∇ψ|2 + |T ψ |2 |ψ|2 − 2 T ψ (ej , ek )e ek · ∇ej ψ, ψ
j,k=1
= |∇ψ| − |T | |ψ|
2 ψ 2 2
since from its definition the tensor T ψ is symmetric. Integrating and applying
the Schrödinger-Lichnerowicz formula (1.15) leads straightforward to the
inequality, of which limiting-case implies ∇ψ = 0. This concludes the proof.
n
ψ
ψ
trg (T ) = e ej · ∇ej ψ,
j=1
|ψ| 2
ψ
= e Dψ,
|ψ|2
= λ,
λ2
so that |T ψ |2 = |(T ψ )0 |2 + n and
60 3 Lower eigenvalue estimates on closed manifolds
n n
λ2 ≥ inf S+ |(T ψ )0 |2 ,
Mψ 4(n − 1) n−1
≥0
Theorem 3.5.1 (T. Friedrich and K.-D. Kirchberg [96]) Any eigen-
value λ of D on an n(≥ 2)-dimensional closed Riemannian spin manifold
(M n , g) with divergence-free curvature tensor, vanishing scalar curvature and
nowhere-vanishing Ricci-curvature satisfies:
1 inf M |Ric|2
λ2 > ,
4 n−1 inf |Ric| − κ
n M 0
be found among the following families, see [96, Ex. 1-4] and [160, Ex. 4.1
& 4.2]: (local) Riemannian products of Einstein manifolds, warped prod-
ucts of S1 with an Einstein manifold with positive scalar curvature, warped
products on Riemannian surfaces, conformally flat manifolds. Note however
that Einstein manifolds themselves or manifolds whose Ricci tensor vanishes
somewhere cannot be handled by Theorem 3.5.1. This was the motivation
of T. Friedrich and K.-D. Kirchberg for obtaining a lower bound involving
the Weyl tensor only. The best result in this direction was obtained by K.-D.
Kirchberg, generalizing an earlier one by T. Friedrich and himself [95, Thm.
3.1]:
Theorem 3.5.2 (K.-D. Kirchberg [161]) Any eigenvalue λ of D on
an n(≥2)-dimensional closed Riemannian spin manifold (M n , g) with
divergence-free Weyl-tensor and μ > 0 satisfies:
1 n 4ν0 2
λ ≥
2
(2n − 1) inf (S) + inf (S)2 + ( ) , (3.25)
8(n − 1) M M n−1 μ
C. Bär’s inequality (3.17) does not give any information on the spectrum
of D on compact Riemannian surfaces with nonpositive Euler characteristic,
i.e., with positive genus. Estimates on such surfaces have to depend on the
choice of spin structure, as the example of the 2-torus already shows: for its
trivial spin structure (i.e., for the spin structure coming from the trivial lift of
the lattice-action to the spin level, see Proposition 1.4.2) it admits harmonic
spinors - for flat hence any metrics because of (1.16) - but not for any other
spin structure [86].
The first estimate to have been proved is a qualitative one and dates back
to J. Lott’s work [181] providing lower bounds for general conformally covari-
ant elliptic self-adjoint linear differential operators. In the case of surfaces it
62 3 Lower eigenvalue estimates on closed manifolds
Sketch of proof of Theorem 3.6.1: The proof of the inequality combines the
following steps. First one chooses a flat metric g0 := e2u g in the conformal
class of g. Using the min-max principle (see e.g. Lemma 5.0.2) it can be easily
proved that λ2 ≥ e2 max(u) λ20 , where λ0 > 0 is the smallest Dirac eigenvalue
(in absolute value) on (T2 , g0 ) and for the same spin structure. Now the Dirac
spectrum of (T2 , g0 ) for any spin structure is explicitly known (see Theorem
2.1.1), in particular the following equality holds
Theorem 3.6.3 (B. Ammann and C. Bär [17]) Let M be a closed Rie-
mannian surface of positive genus g with spin structure whose Arf-invariant
equals 1. Then any eigenvalue λ of D satisfies
2 π 1
|λ| ≥ · − . (3.27)
2g + 1 Area(M ) δ(M )
Although the lower bound need this time not be positive there exist examples
for which it is: as above, consider an ε-tubular neighbourhood Mε of a closed
plane curve with exactly g − 1 intersections and such that, w.r.t. any allowed
spin structure, δ(Mε ) ∼ cst (fix for instance the diameter equal to 1ε ). Then
ε→0 ε
the lower bound is asymptotic to 2g+1 2
· Area(M
π
ε)
for ε → 0. In the case
where g = 1 the k-dependent expression in the r.h.s. of (3.26) is for k = 2
greater than the r.h.s. of (3.27), so that (3.26) is better than (3.27).
Combining Theorems 3.6.2 and 3.6.3 with the extrinsic upper bound (5.19)
for the smallest Dirac eigenvalue for surfaces embedded in R3 one obtains
a lower bound of the Willmore functional, see [17, Thm. 7.1]. Besides, we
mention that Theorems 3.6.2 and 3.6.3 can be extended to complete surfaces
with finite area [17, Thm. 8.1].
√
, g) denotes the space of ±
for c < 0, where K± √c (M c
2
2 -Killing spinors on
, g).
(M
the spinorial Levi-Civita connection
Proof in the case c = 0: Denote by ∇
. The Schrödinger-Lichnerowicz formula for the Dirac operator D
of M of
, g) and elementary computations as in Section 1.3 show that, for any
(M
ϕ ∈ Γ(ΣM),
2
|Dϕ| = 2 ϕ, ϕ
+ div (V )
e D M
(1.15)
= e ∇ ϕ
+ S |ϕ|2 + div (V )
∗ ∇ϕ,
4 M
= 2 + S |ϕ|2 + div (V + W ),
|∇ϕ| (3.29)
4 M
where
V and Ware the vector fields on
M defined by the relations g(V, X) :=
ϕ
and g(W, X) := e ∇
e X · Dϕ, X ϕ, ϕ
for all X ∈ T M
respectively
S 2 n 2
|ϕ| − |Dϕ| = −|Pϕ|2 − divM
(V + W ),
4 n+1
where π≥0 : Γ(Σ) → Γ(Σ) denotes the L2 -orthogonal projection onto the
eigenspaces of D2 to nonnegative eigenvalues, see Section 1.5 and Chapter 4.
Since S ≥ 0 and H ≥ 0 the identity (3.30) with ϕ := φ implies
! #
S 2 M S 2 n 2
0≤ |φ| vg = |φ| − |Dφ| vgM
M 4
M 4 n + 1
nH 2
≤ D2 φ, φ
− |φ| vg
M 2
nH
≤ D2 π≥0 φ, π≥0 φ
− |π≥0 φ|2 vg
2
M
nH
= (|λ| − )|ψ|2 vg
M 2
⊂M
In case M n+1 (c), where Mn+1 (c) is a spaceform with constant curvature
c ≤ 0, Gauss’ equations imply in particular ( n2 )2 (H 2 + c) ≥ 4(n−1)
n
S, hence
√
(3.28) improves (3.1) under the supplementary assumption H ≥ −c.
There exists a conformal version of (3.28) in terms of the so-called Yamabe
relative invariant, see [144].
The characterization of the equality case in (3.28) provides a short proof
of Alexandrov’s theorem (see reference in [142]) on constant mean curvature
embedded hypersurfaces in the Euclidean and hyperbolic spaces respectively
[142, Thm. 8]:
Theorem 3.7.2 (A.D. Alexandrov) Every closed embedded hypersurface
with constant mean curvature in Rn+1 or Hn+1 is a round geodesic hyper-
sphere.
Proof in the case c = 0: Let M be such a hypersurface. It is embedded so
that on the one hand it bounds a compact domain M (in particular it is
orientable hence spin, see Proposition 1.4.1); on the other hand, it can be
shown that necessarily H ≥ 0 by a result of S. Montiel and A. Ros (see
reference in [142]). Moreover the assumption H constant implies that (3.28)
is an equality, in which case every non-zero eigenvector of D2 associated to
the eigenvalue nH
2 must be the restriction onto M of a parallel spinor on M
(Theorem 3.7.1). But considering the spinor field
x −→ ϕx := νx · φ + Hx · φ,
D2 ϕ = D2 (ν · φ) + HD2 (x · φ)
(1.19)
= −ν · D2 φ + HD2 (x · φ)
(1.22) nH ν (x · φ) − ν · D(x
· φ))
= −ν · D2 φ + H( x·φ−∇
2
68 3 Lower eigenvalue estimates on closed manifolds
nH nH
= − ν · φ + H( x · φ − ν · φ + (n + 1)ν · φ)
2 2
nH
= (ν · φ + Hx · φ)
2
nH
= ϕ,
2
X ϕ = −A(X) · φ + HX · φ.
0=∇
Since φ has no zero on M one deduces that A = HIdT M , i.e., that M must
be totally umbilical in M hence in Rn+1 . This concludes the proof for c = 0.
For c < 0 we refer to [140].
The study of the Dirac operator on compact spin manifolds M with bound-
ary was initiated by Atiyah, Patodi and Singer [27] in the search for index
theorems on such manifolds, see e.g. [59] which is a comprehensive reference
on the subject. In order to be able to talk about eigenvalues of the Dirac oper-
ator of M in this context, elliptic boundary conditions have to be introduced
as we have seen in Section 1.5. Following [82] a whole bunch of spectral prop-
erties of the Dirac operator have recently been proved using varied boundary
conditions, leading sometimes to very beautiful geometric results, such as
those already presented in Section 3.7. In this chapter we mainly show that,
under any of the four boundary conditions introduced in Section 1.5, some
kind of Friedrich’s inequality (3.1) holds on M although the lower bound is
not always attained according to the boundary condition chosen. For readers
interested in more details and references we suggest [139].
From (1.22) and the definition of G it can be easily proved that D2 G = GD2 ,
so that, if ψ := ϕ|∂M then
4.2 Case of the CHI boundary condition 71
D2 ψ, ψ
vg∂M = D2 (ν · Gψ), ψ
vg∂M
∂M ∂M
(1.23)
= − ν · D2 (Gψ), ψ
vg∂M
∂M
= − ν · G(D2 ψ), ψ
vg∂M
∂M
= − D2 ψ, ν · Gψ
vg∂M
∂M
= − D2 ψ, ψ
vg∂M ,
∂M
that is, ∂M D2 ψ, ψ
vg∂M = 0. Formula (3.30) together with the assumption
H ≥ 0 imply
S 2 n−1
|ϕ| − |Dϕ| vg ≤ 0,
2
M 4 n
which leads to (4.1).
If inequality (4.1) is an equality, then (3.30) implies on the one hand
that ϕ must be a Killing spinor to the real Killing constant − nλ (it is an
eigenvector of D lying in the kernel of the Penrose operator P of (M n , g), see
Appendix A) and on the other hand H = 0, i.e., the boundary ∂M must be
minimal in M . Moreover, f := Gϕ, ϕ
defines a smooth real function on M
whose differential is given on any X ∈ M by
X(f ) = ∇X (Gϕ), ϕ
+ Gϕ, ∇X ϕ
λ λ
= G(− X · ϕ), ϕ
− Gϕ, X · ϕ
n n
λ
= − ( Gϕ, X · ϕ
+ X · ϕ, Gϕ
)
n
2λ
= − e ( Gϕ, X · ϕ
) .
n
Hence the Hessian of f evaluated on any X, Y ∈ T M is given by
2λ
Hess(f )(X, Y ) = − e ( ∇X (Gϕ), Y · ϕ
+ Gϕ, Y · ∇X ϕ
)
n
2λ2
= 2 e ( G(X · ϕ), Y · ϕ
+ Gϕ, Y · X · ϕ
)
n
2λ2
= e ( Gϕ, X · Y · ϕ
+ Gϕ, Y · X · ϕ
)
n2
4λ2
=− Gϕ, ϕ
g(X, Y ),
n2
72 4 Lower eigenvalue estimates on compact manifolds with boundary
2
i.e., Hess(f ) = − 4λ
n2 f g. On the other hand neither λ nor the function f
vanish, since (1.26) implies
Dϕ, Gϕ
− ϕ, D(Gϕ)
vg = ψ, ν · Gψ
vg∂M ,
M ∂M
i.e., 2λf = ∂M |ψ|2 vg∂M , which does not vanish because of the unique
continuation property mentioned above. Hence Reilly’s characterization of
the hemisphere (see reference in [139]) implies that (M n , g) is isometric to a
n
hemisphere with radius 2|λ| . Since the standard sphere for its canonical spin
structure and metric with constant sectional curvature 1 carries non-zero
− 12 and 12 -Killing spinors (see e.g. Examples A.1.3.2) the other implication
is trivial.
= − D2 (ψ), iν · ψ
vg∂M
∂M
= − D2 ψ, ψ
vg∂M ,
∂M
that is, ∂M D2 ψ, ψ
vg∂M = 0. Formula (3.30) together with the assumption
H ≥ 0 imply
S 2 n−1
|ϕ| − |Dϕ|2 vg ≤ 0,
M 4 n
which leads to (4.2).
If (4.2) is an equality, then again ϕ must be a − nλ -Killing spinor on
(M n , g). Since in that case m(λ) > 0 (see Section 1.5), one deduces from
Proposition A.4.1 that λ ∈ iR∗+ and hence S = 4 n−1 2
n λ < 0, contradiction.
Therefore (4.2) is always a strict inequality.
Moreover, (4.3) is an equality if and only if the mean curvature of the bound-
ary is constant and (M n , g) admits a non-zero imaginary Killing spinor.
Besides, as for the CHI boundary condition, there exists a Hijazi-type
conformal lower bound for the Dirac spectrum under MIT bag boundary
condition, which was proved by S. Raulot in [211, 212] but which is never
sharp.
74 4 Lower eigenvalue estimates on compact manifolds with boundary
(1.23)
π<β (ψ − ν · ψ) = π<β ψ − ν · π>−β ψ
= π≤−β ψ − π[β,−β] ψ − ν · π>−β ψ
= ν · π≥β ψ − π[β,−β] ψ − ν · π>−β ψ
= ν · π[β,−β] ψ − π[β,−β] ψ,
4.4 Case of the mgAPS boundary condition 75
In this chapter we turn to the very different game of looking for upper eigen-
value bounds for the fundamental Dirac operator. We concentrate on two
methods for obtaining them. The first method, due to C. Vafa and E. Witten
[230], consists in comparing D to another Dirac-type operator D of which
kernel is non trivial for index-theoretical reasons, then in estimating the zero-
order difference D −D by geometric quantities. As a result, those bound from
above a topologically determined number of eigenvalues of D. This method
was applied to prove Theorems 5.1.1, 5.2.1 and 5.2.2. The second method
relies on the min-max principle, which is a general variational principle char-
acterizing eigenvalues of self-adjoint elliptic operators, see e.g. [74, pp.16-17]:
Another way to obtain upper bounds consists in comparing the Dirac spectra
for different metrics and applying the min-max principle. The remarkable
property of conformal covariance of the Dirac operator allows this method to
work. The first result is this direction is due to J. Lott [181, Prop. 3].
Proof : Fix g0 ∈ [g] with Sg0 < 0 on M n . For some u ∈ C ∞ (M n , R), let
g = e2u g0 ∈ [g] with Sg < 0. As in Proposition 1.3.10, we denote by ϕ
→ ϕ
the unitary isomorphism Σg0 M −→ Σg M . Choose ψ0 to be a non-zero eigen-
vector for Dg0 associated to the eigenvalue λ1 (Dg0 ) and set ψ := e− 2 u ψ0 .
n−1
We now estimate the quotient of integrals on the r.h.s. with the help of (3.19).
Namely (3.19) reads
where we have used Δg0 (e−u ) = −e−u Δg0 u − e−u |gradg0 (u)|2g0 . We deduce
that
e−u vg0 supM (−Sg ) M −Sg0 e−u vg0
M ≤ ·
eu vg0 inf M (−Sg0 ) −Sg eu vg0
M M
supM (−Sg ) e−u |gradg0 (u)|2g0 vg0
= 1 + n(n − 1) M
inf M (−Sg0 ) M g
S eu vg0
supM (−Sg )
≤ ,
inf M (−Sg0 )
On surfaces, Lott’s estimate (5.2) only applies if the genus is at least 2. For
genus 0 or 1 the use of (3.19) can be avoided through the fact that eigenvectors
associated to the lowest Dirac eigenvalue have constant length:
80 5 Upper eigenvalue bounds on closed manifolds
iii) The smallest positive eigenvalue of DT22 ,g on T2 endowed with trivial spin
structure satisfies
T2
e−3u {λ2 (DT22 ,g0 ) + |gradg0 (u)|2g0 }vg0
λ2 (DT22 ,g ) ≤ . (5.6)
T2
e−u vg0
1
DM,g ψ = e−u (DM,g0 ψ0 + gradg0 (u) · ψ0 )
2
−u 1
= e (λ1 (DM,g0 )ψ + gradg0 (u) · ψ0 ),
2
where we have denoted by “·” the Clifford multiplication on Σg0 M . We deduce
that
1
|DM,g ψ|2 = e−2u λ1 (DM,g
2
)|ψ0 |2 + |gradg0 (u) · ψ0 |2
0
4
+ λ1 (DM,g0 )e( ψ0 , gradg0 (u) · ψ0
1
= e−2u 2
λ1 (DM,g ) + |gradg0 (u)|2g0 .
0
4
DM,g ψ = e−
3u
2 DM,g0 ψ0
= λ1 (DM,g0 )e−
3u
2 ψ0 ,
which is (5.5).
iii) Set ψ := e− 2 ψ0 . On the one hand,
3u
e− 2
3u
(1.16)
e−u (DT2 ,g0 (e− 2 ψ0 ) +
3u
DT2 ,g ψ = gradg0 (u) · ψ0 )
2
(1.11) −u
3 3u
− e− 2 gradg0 (u) · ψ0 + e− 2 DT2 ,g0 ψ0
3u
= e
2
e− 2
3u
+ gradg0 (u) · ψ0
2
= e−u · e− 2 (λ2 (DT2 ,g0 )ψ0 − gradg0 (u) · ψ0 ).
3u
We deduce that
T2
|DT2 ,g ψ|2 vg
λ2 (DT22 ,g ) ≤
T2
|ψ|2 vg
82 5 Upper eigenvalue bounds on closed manifolds
T2
e−2u · e−3u {λ2 (DT22 ,g0 ) + |gradg0 (u)|2g0 }vg
=
T2
e−u vg0
T2
e−3u {λ2 (DT22 ,g0 ) + |gradg0 (u)|2g0 }vg0
= ,
T2
e−u vg0
λ1 (DS22 ,g )Area(S2 , g) − 4π ≤
1
inf 2 |gradg0 (uΦ )|2g0 vg0 , Φ∗ g0 = e−2uΦ g , (5.7)
Φ∈Conf(S ,[g0 ]) 4 S2
3 1
2
lim λ1 (DM ) ∈ [2, + ln(2)] and 2
lim λ1 (DM )≤ . (5.8)
a→0 a
2 a→∞ a
4
Both estimates provide much sharper upper bounds as C. Bär’s one (5.19) in
terms of the averaged total squared mean curvature.
In the case of the 2-torus and with the notations of Theorem 2.1.1, the
smallest eigenvalue of DT22 ,g w.r.t. the (δ1 , δ2 )-spin structure and flat metric is
not greater than 4π 2 | 12 (δ1 γ1∗ + δ2 γ2∗ )|2 , so that by (5.5) it satisfies [2, Thm. 4]
−u
2 e vg
λ1 (DT22 ,g )Area(T2 , g) ≤π 2
|δ1 γ1∗ + δ2 γ2∗ |2 T u 0 . (5.9)
T2
e vg 0
For both the (1, 0)- and (0, 1)-spin structures (however not for the (1, 1)-one)
these estimates enhance those obtained from (5.19) below.
In higher dimensions another general upper bound in terms of the sectional
curvature can be obtained from the min-max principle. In the following the-
orem we denote by Br (p) the geodesic ball of radius r > 0 around some point
p ∈ M and, provided r is smaller than the injectivity radius radinj (M n , g) of
the Riemannian manifold (M n , g), by 0 < μ1 (Br (p)) ≤ μ2 (Br (p)) ≤ . . . the
spectrum of the scalar Laplace operator with Dirichlet boundary condition
on Br (p). For any x ≥ 0 and t > 0 we also define
⎧ √
⎪ 1−cos( xt)
⎨ √x sin(√xt) if x > 0
fx (t) :=
⎪
⎩ t
2 if x = 0.
2 n−2
+ (n − 1)(1 + [ ])(ρ1 + ρ2 )fκ+ρ2 (r) , (5.10)
3 2
where the second infimum ranges over r ∈ ]0, min{radinj (M n , g), √κ+ρ π
2
}[.
Moreover, (5.10) is an equality for (M n , g) = (Sn , can), j = 1 and ρ1 = ρ2 = 0.
Sketch of proof of Theorem 5.1.4: First consider a geodesic ball Br (p) of
radius 0 < r < radinj (M n , g). Since Br (p) is convex its spinor bundle can
be trivialized by a pointwise orthonormal family ϕ1 , . . . , ϕN . Let {fj }j≥1
be a Hilbert basis of L2 (Br (p), C) made out of eigenfunctions for the scalar
Laplace operator Δ with Dirichlet boundary condition on Br (p), in particular
Δfj = μj (Br (p))fj with fj |∂Br (p) = 0 holds. Then {fj ϕk }1≤j,1≤k≤N is a
Hilbert basis of L2 (ΣBr (p)). Since DB
2
r (p)
is of Laplace type (see Schrödinger-
Lichnerowicz’ formula (3.2)) one may talk about its eigenvalues with respect
to the Dirichlet boundary condition and the min-max principle also applies.
84 5 Upper eigenvalue bounds on closed manifolds
The second step in the proof, which is the main and the most technical
one, consists in estimating the supremum on the r.h.s. by geometric data.
This can be done essentially by controlling the growth of the pointwise norm
along geodesics and applying Rauch’s comparison theorem, we refer to [36,
Lemma 1] and [36, Sec. 4]. The third and last step consists in comparing
the eigenvalues of D2 with those of DB 2
r (p)
subject to Dirichlet boundary
condition: the monotonicity principle (see e.g. [74, Cor. 1 p.18]) implies that
In this section we assume the existence of some map from the manifold M
to another manifold and want to derive upper eigenvalue estimates in terms
of geometric invariants associated to this map. The first situation which has
been studied is the case where there exists a map of sufficiently high degree
from M into the round sphere of same dimension.
k−1
2 −1
n
deg(ι) ≥ 1 + 2 mj
j=1
for some positive integer k, where mj is the multiplicity of the jth eigenvalue
of D2 . Then the kth eigenvalue λk of D satisfies
n
|λk | ≤ 2 2 −1
n
· max dx ι.
2 x∈M
1
R(ι) := S − 2ric(ν,
ν) .
n(n − 1)
| ), then
in any local orthonormal basis {ej }1≤j≤n of T M : let ϕ ∈ Γ(ΣM M
n n
D2 ϕ
(1.21)
= e ϕ − 1
ej · ν · ∇ ej · ν · A(ej ) · ν · ϕ
j
j=1
2 j=1
= + nH ϕ,
−ν · Dϕ (5.12)
2
so that
(1.23)
+ D2 ( nH
D22 ϕ = ν · D2 (Dϕ) ϕ)
2
(1.11)
+n
= ν · D2 (Dϕ) (grad(H) · ν · ϕ + HD2 ϕ)
2
(5.12)
= (D + n grad(H) · ν · ϕ + nH D2 ϕ
+ nH ν·)Dϕ
2 2 2
2 2
2ϕ + n H n
= D ϕ + grad(H) · ν · ϕ. (5.13)
4 2
n+1 , g) one has, in any local orthonormal
Since ψ is a twistor-spinor on (M
basis {ej }1≤j≤n of T M ,
n
2ψ
D
= D( e ψ)
ej · ∇ j
j=1
1
n
= −
D(ej · ej · Dψ)
n + 1 j=1
n
= D(Dψ)
n+1
n 1
n
(A.4)
= j ) · ψ + S ej · ej · ψ)
(− ej · Ric(e
n − 1 j=1 2 4n
n S 1 S
= ψ + ν · Ric(ν) ·ψ− ψ
n−1 2 2 4
n 1 1
ν))ψ + ν · Ric(ν) T·ψ
= (S − 2ric(ν,
n−1 4 2
n n(n − 1) 1 T
= R(ι)ψ + ν · Ric(ν) · ψ
n−1 4 2
2
n n T · ψ,
= R(ι)ψ + ν · Ric(ν) (5.14)
4 2(n − 1)
88 5 Upper eigenvalue bounds on closed manifolds
T := n ric(ν,
where we denoted by Ric(ν) ej )ej the tangential projection
j=1
of Ric(ν). Combining (1.13) (which obviously holds for D22 instead of D2 ),
(5.13) and (5.14) we obtain, for every f ∈ C ∞ (M, R),
2 2
2 ψ + n H ψ + n grad(H) · ν · ψ)
D22 (f ψ) = f (D
4 2
grad(f ) ψ − A(grad(f ))
−2(∇ · ν · ψ) + (Δf )ψ
2
n2 2 nf nf T·ψ
= (H + R(ι))f ψ + grad(H) · ν · ψ + ν · Ric(ν)
4 2 2(n − 1)
2
+ grad(f ) · DM
ψ + A(grad(f )) · ν · ψ + (Δf )ψ. (5.15)
n+1
We deduce that
n2 2 2f
e( D22 (f ψ), f ψ
) = (H + R(ι))f 2 |ψ|2 + e( grad(f ) · DM ψ, ψ
)
4 n+1
+f (Δf )|ψ|2
n2 2
= (H + R(ι))f 2 |ψ|2 − g(f grad(f ), grad(|ψ|2 ))
4
+f (Δf )|ψ|2 ,
The case n = 1 - the “baby case” - should not be of interest since the
spectrum of S1 for both spin structures is explicitly known (see Theorem
= R2 or S2 -
2.1.1). However similar results turn out to hold - at least for M
and to follow from very elementary geometric properties of plane or space
curves:
5.2 Extrinsic upper bounds 89
π2
λ1 (D2 ) =
L2
π2
≤ 2 n2c
L
L
π2 1
= 2 · 2( H(t)dt)2
L 4π 0
L
1
≤ · L · H 2 (t)dt
4L2 0
L
1
= H 2 (t)dt,
4L 0
which proves the inequality. The equality holds if and only if H is constant
and |nc | = 1, hence we obtain the statement in that case.
• Case κ = 1: Let μc be the bridge number and H be the curvature of c as
2 2
space curve. From H = H + 1 and
90 5 Upper eigenvalue bounds on closed manifolds
L
1
μc ≤ H(t)dt
2π 0
π2 2
λ1 (D2 ) ≤ μ
L2 c
L
1
≤ ( H(t)dt) 2
4L2 0
L
1 2 (t)dt
≤ H
4L 0
L
1
= (H 2 (t) + 1)dt,
4L 0
which shows the inequality. As before the equality only holds if H is constant
and μc = 1, hence we obtain the statement in that case and conclude the
proof.
Proof : Both results follow directly from Theorem 5.2.3: considering the
expression inside the infimum of the r.h.s. of inequality (5.11) one just has
1
to choose f to be a non-zero constant in the first case and to be |ψ| in the
second one.
Corollary 5.2.5 provides the following estimates which were proved by C. Bär
[40, Cor. 4.2 & 4.3] for hypersurfaces of the Euclidean space Rn+1 or of the
round sphere Sn+1 and by N. Ginoux [97, 99, 98] for hypersurfaces of the
hyperbolic space Hn+1 : let Mn+1 (κ) := Rn+1 if κ = 0, Sn+1 if κ = 1 and
Hn+1 if κ = −1, then for any closed hypersurface M of M n+1 (κ) carrying
the induced metric and spin structure,
5.2 Extrinsic upper bounds 91
n2
λ1 (D2 ) ≤ (H 2 + κ)vg (5.19)
4Vol(M ) M
if κ ≥ 0,
n2
λ1 (D2 ) ≤ (max (H 2 ) − 1) (5.20)
4 M
and
n2 1
λ1 (D ) ≤
2
(H − 1)vg +
2
inf |d(ln(|ψ|))|2 vg
4Vol(M ) M Vol(M ) ψ∈ K ± i M
2
ψ=0
(5.21)
if κ = −1 (the space K i refers here to the space of ± i -Killing spinors on
±2 2√
Hn+1 ). Indeed M n+1 (κ) carries at least one non-zero ± κ -Killing spinor ψ
2
(which is in particular a twistor-spinor and vanishes nowhere), see e.g. [56, 63]
and Examples A.1.3. Moreover R(ι) = κ and, if κ ≥ 0, then Proposition A.4.1
implies that |ψ| is constant on M n+1 (κ).
The inequalities (5.19)√
and (5.20) actually hold for higher eigenvalues of
D2 , since the space of ± 2κ -Killing spinors on Mn+1 (κ) is 2[ n+1
2 ] -dimensional
and the upper bound in (5.19) or (5.20) does not depend on the Killing spinor
ψ, see [40, 99].
The inequalities (5.19), (5.20) and (5.21) are equalities for all geodesic
hyperspheres in M n+1 (κ), see [40], [99] and [98] respectively. For κ ≤ 0 the
question remains open whether those are the only hypersurfaces enjoying this
property. For κ = 1, generalized Clifford tori in Sn+1 as well as minimally
embedded S /Q8 (where Q8 denotes the finite group of quaternions) in S4
3
also satisfy the limiting-case in (5.19) and it is conjectured that this actually
holds for every homogeneous hypersurface in the round sphere, see [100, 103].
For n = 2 the upper bound in (5.19) is nothing but the so-called Willmore
functional of the immersion M → M 3 (κ). Combining the estimate (5.19)
with lower bounds of the Dirac spectrum (see Section 3.6) B. Ammann [9]
and C. Bär [40] proved the Willmore conjecture (“ M H 2 vg ≥ 2π 2 for every
embedded torus M in R3 ”) for particular metrics.
Note 5.2.6 Could (5.18), (5.20) and (5.21) be enhanced in
n2
λ1 (D2 ) ≤ (H 2 + R(ι))vg ? (5.22)
4Vol(M ) M
holds.
Chapter 6
Prescription of eigenvalues
on closed manifolds
From its definition the Dirac spectrum a priori depends on the metric, the spin
structure and of course the underlying manifold. In a very formal manner,
the Dirac spectrum can be thought of as a functor from the category of
closed Riemannian spin manifolds to that of real discrete sequences with
closed image and unbounded on both sides. In this chapter, we investigate
this functor, in particular its injectivity and surjectivity. In other words, does
its Dirac spectrum determine a given Riemannian spin manifold? For a given
real sequence as above, is there a Riemannian spin manifold whose spectrum
coincides with this sequence? If the answer to the former question is definitely
negative (Section 6.1), only partial results have been found regarding the
latter in the case where the whole spectrum is replaced by a finite set of
eigenvalues. In this situation one has to distinguish between the eigenvalue
0 - whose associated eigenvectors are called harmonic spinors - and the other
ones. We shall see in Section 6.2 that, in dimension n ≥ 3 (the case of
surfaces must be handled separately), the metric can in general be modified
so as to make 0 a Dirac eigenvalue, whereas generic metrics just have as many
harmonic spinors as the Atiyah-Singer-index theorem forces them to do. If
the finite set of real numbers does not contain 0 (and is symmetric about 0 if
n ≡ 3 (4)), then it is always the lower part of the Dirac spectrum of a given
metric on a fixed spin manifold (Section 6.3).
manifold. A sharper insight in the formula shows for example that, in di-
mension 4, the Euler characteristic of the manifold is determined by its Dirac
spectrum as soon as the scalar curvature is assumed to be constant [54]. How-
ever, the Dirac spectrum in general determines neither the isometry class nor
the topology of the manifold. To illustrate this, we describe different fami-
lies of examples, evolving from the “simplest” to the most sophisticated ones
where even the explicit knowledge of the Dirac spectrum is not needed.
The founding result for isospectrality issues is indisputably J. Milnor’s
famous one-page-long article (see reference in [15]), where the author de-
scribes two Laplace-isospectral but non-isometric 16-dimensional tori. The
idea is the following. The spectrum of the scalar Laplace operator on a flat
torus Γ\R is {4π 2 |γ ∗ |2 , γ ∗ ∈ Γ∗ }, where we keep the notations of Theorem
n
are not isometric. From Theorem 2.1.1, the Dirac spectrum of a flat torus
with trivial spin structure is {±2π|γ ∗ |, γ ∗ ∈ Γ∗ }, hence the same argument
shows that the Dirac spectra of Γ1 \R and Γ2 \R also coincide, at least for
16 16
Theorem 6.1.1 (C. Bär [38]) For n ≥ 3 odd let Γ1 , Γ2 ⊂ SOn+1 be finite
subgroups acting freely on Sn . For j = 1, 2 let j : Γj −→ Spinn+1 be a group
homomorphism such that ξ ◦j is the inclusion map Γj ⊂ SOn+1 and consider
the induced spin structure on Γj \S . Assume the existence of a bijective map
n
Corollary 6.1.2 (C. Bär [38]) Let n ≡ 3 (8), n ≥ 19. Let a be a prime
number with a ≡ 1 ( n+1 n+1 2
4 ), let b := ( 4 ) and let r be chosen such that its
n+1 ×
mod a class is of order 4 in (Za ) . Then there exist two Dirac isospectral
non-isometric spaceforms diffeomorphic to Γ(a, b, r)\S .
n
and set Γj,h := Span(Bj,h · ( e2n + Id), e1 + Id, . . . , en + Id). The subgroup
Γj,h of the isometry group of Rn is orientation-preserving and co-compact,
therefore the quotient
:= Γj,h \R
n n
Mj,h
is a compact orientable flat manifold with holonomy group Z2 . Moreover it
n
can be shown that Mj,h is spin and that its first homology group with integer
coefficients is H1 (Mj,h ) = Zn−j−h ⊕ (Z2 )h [187, Prop. 4.1]. In particular, the
n
j+h
Fk+ := {Mj,h
n
| ≡ k (2)}.
2
covering ξ. This assignment fixes exactly two spin structures since with those
assumptions the lift of en + Id must be (−1)k , see [187, Prop. 4.2] for details.
Theorem 6.1.3 (R. Miatello and R. Podestá [188]) For a fixed inte-
ger n ≥ 3 consider the Mj,h
n
’s defined above.
i) All elements of the family {(Mj,h n
, ε0 ) | Mj,h ∈ F0+ } are pairwise Dirac
isospectral closed flat Riemannian spin manifolds, which are pairwise non-
homeomorphic as soon as the pairs (j, h) are different.
ii) If n ≡ 3 (4) then all elements of the family {(Mj,hn
, ε1 ) | Mj,h
n
∈ F1+ } are
pairwise Dirac isospectral closed flat Riemannian spin manifolds, which
are pairwise non-homeomorphic as soon as the pairs (j, h) are different.
Theorem 6.1.4 (B. Ammann and C. Bär [15]) Let M := Γ\G and
M := Γ \G be spin homogeneous spaces where Γ, Γ are co-compact lattices
in the simply-connected Lie group G. If the right-representations of G onto
L2 (ΣM ) and L2 (ΣM ) are equivalent, then for any left-invariant metric on
G, the manifolds M and M are Dirac isospectral.
The independence of the spectrum on the left-invariant metric is a very
strong statement. It allows in particular to produce continuous families of
isospectral metrics, the spin structures staying fixed. Of course the real
difficulty consists in applying Theorem 6.1.4, i.e., in picking groups G so
that the equivalence of the G-representations is satisfied. For nilpotent Lie
groups that are strictly non-singular (i.e., for every z in the center of G
and x in its complement, there exists a y with xyx−1 y −1 = z), this condition
simplifies in terms of group theoretical data [15, Thm. 5.3]. The nilpotent Lie
groups chosen by R. Gornet provide concrete examples where this criterion
is fulfilled, we refer to [15, Thm. 5.6] for details and references.
Theorem 6.1.5 (B. Ammann and C. Bär [15]) There exist in dimen-
sions 7 and 8 a continuous family of Dirac isospectral non-isometric closed
Riemannian spin manifolds. Each manifold inside a family is diffeomorphic
to the same quotient of some nilpotent Lie group by a co-compact lattice.
6.2 Harmonic spinors 97
for any u ∈ C ∞ (M, R). It can for example be deduced from this fact combined
with Theorem 2.1.3 that, whatever the metric chosen, the 2-sphere S2 does
not carry any non-zero harmonic spinor (there exists only one conformal class
as well as one spin structure on S2 ). Alternatively this straightforward follows
from Bär’s inequality (3.17) since the lower bound 2πχ(M 2 ) is a topological
invariant which is positive for M 2 = S2 .
On closed surfaces the presence of a quaternionic structure on the spinor
bundle, which commutes with the Dirac operator (see e.g. [101, Lemma 1]),
forces the number d to be even. On the other hand, d is bounded from above
independently of the metric and spin structure: N. Hitchin proved that d ≤
2 ] for every closed Riemann surface M of genus g and, if g ≤ 2 then
2[ g+1 2
d does not even depend on the conformal class [148, Prop. 2.3]. The case of
S2 has just been discussed. For the 2-dimensional torus T2 it can be seen as
a consequence of the conformal property above and of Theorem 2.1.1, which
implies that T2 has a 2-dimensional space of harmonic spinors for the trivial
spin structure and no non-zero one otherwise. If g = 2 the independence
on the conformal class follows from [148, Prop. 2.3] or, alternatively, from
the following argument: the Atiyah-Singer-index theorem implies that d2 ≡
α(M ) (2), where α(M ) is the α-genus of M , see (6.1) below. Either α(M ) = 1
and it follows from Hitchin’s upper bound that d = 2 as soon as g ≤ 4, or
α(M ) = 0 and, for g ≤ 2, one has d = 0. For g = 3, 4 and α(M ) = 0 both
possibilities d = 0 and d = 4 occur [20].
In higher genus the picture is more complex. As a consequence of [19, Thm.
1.1], for any given spin structure, there exists a conformal class for which d
coincides with the lower bound provided by the Atiyah-Singer-index theorem,
hence for which d ∈ {0, 2}. Next one can ask whether d can be made maximal.
To answer this question, it is more convenient to study the variations of d in
terms of the spin structure, the conformal class being fixed. Recall first that,
98 6 Prescription of eigenvalues on closed manifolds
where α(M ) ∈ Z2 is the α-genus of M (see e.g. [45, Sec. 3] for a definition).
It was first conjectured by C. Bär [39] that (6.1) is an equality for generic
metrics, i.e., metrics belonging to some subset which is open in the C 1 - and
dense in the C ∞ -topology in the space of all Riemannian metrics. In dimen-
sions n = 3 and 4 perturbation methods combined with the formula linking
spinors for different metrics [62] suffice in order to prove the conjecture to
hold true (S. Maier [185]). If the underlying manifold is assumed to be simply-
connected, then C. Bär and M. Dahl [45] proved the conjecture to hold true
for all dimensions n ≥ 5. Based on bordism theory, their argument consists of
the three following steps. First, the conjecture holds true for some generators
of the spin bordism ring. Second, any closed spin manifold is spin bordant to
a spin manifold where the conjecture holds true. By a theorem of Gromov and
Lawson, any n(≥ 5)-dimensional closed simply-connected manifold which is
spin bordant to a spin manifold can be obtained from it by surgeries of codi-
mension at least 3. It remains in the last step to show that the conjecture
survives to surgeries of codimension at least 3. A set of generators is given by
spin manifolds admitting metrics of positive scalar curvature (for which we
already know that Ker(D) = 0) as well as products of some of the irreducible
manifolds of Mc.K. Wang’s classification (see Theorem A.4.2), on which it
can be relatively easily shown that the conjecture holds true. The second step
is a pure argument of bordism theory using a theorem by S. Stolz (see [45]
for references) and the explicit set of generators described above. The crucial
step is the last one, which can be deduced from the following theorem.
Theorem 6.2.2 (C. Bär and M. Dahl [45]) Let (M n , g) be a closed
Riemannian spin manifold. Let M be obtained from M by surgery of codi-
mension at least 3. Let ε > 0 and L > 0 with ±L ∈ / Spec(D).
Then there exists a Riemannian metric g̃ on M such that the Dirac eigen-
n , g̃) in ] − L, L[ differ at most by ε.
values of (M n , g) and (M
Coming back to the conjecture, it has been proved by B. Ammann, M. Dahl
and E. Humbert [19] to hold true in full generality. Their argument relies on
a very fine generalization of the surgery theorem (Theorem 6.2.2) to codi-
mension greater than 1, we refer to [19] for a detailed proof.
At what seems to be the opposite side there have appeared since
N. Hitchin’s pioneering article [148] several results showing the existence in
certain dimensions n ≥ 3 of metrics with lots of non-zero harmonic spinors.
Already computing the Dirac spectrum on S3 with Berger metric (and canon-
ical spin structure), N. Hitchin noticed [148] that, for every N ∈ N, there
exists a metric on S3 admitting at least N linearly independent harmonic
spinors. Furthermore he constructed with the help of differential topological
methods metrics with non-zero harmonic spinors on all closed spin mani-
folds M n of dimension n ≡ 0, 1, 7 (8). Extending the computation of the
Dirac spectrum to all Berger spheres C. Bär showed [39] the existence of
such metrics on all closed spin manifolds M n with n ≡ 3, 7 (8). His proof
is based on the following simple ideas: first, for any closed odd-dimensional
100 6 Prescription of eigenvalues on closed manifolds
In the last section we have seen that, if n ≥ 3 and n ≡ 0, 1, 3, 7 (8), then any
n-dimensional closed spin manifold admits a metric for which the kernel of the
corresponding Dirac operator is non-trivial. In other words, the eigenvalue 0
can be always prescribed in those dimensions. What about prescribing the
6.3 Prescribing the lower part of the spectrum 101
rest of the spectrum? Although this question remains open, at least the lower
part of the spectrum can be fixed. Note that, in dimension n ≡ 3 (4), Theorem
1.3.7.iv) imposes a priori this part to be symmetric about the origin.
Theorem 6.3.1 (M. Dahl [79]) For n ≥ 3 let M n be any closed spin man-
ifold. Let L > 0 (resp. m ≥ 1) be real (resp. integral) and l1 , . . . , lm be
non-zero real numbers. Then the following holds:
i) If n ≡ 3 (4) and −L < l1 < . . . < lm < L, then there exists a Riemannian
metric on M n such that Spec(D, g) ∩ (] − L, L[\{0}) = {l1 , . . . , lm }, where
each of those eigenvalues is simple.
ii) If n ≡ 3 (4) and 0 < l1 < . . . < lm < L, then there exists a Riemannian
metric on M n such that Spec(D, g) ∩ (] − L, L[\{0}) = {±l1 , . . . , ±lm },
where each of those eigenvalues is simple.
The proof of Theorem 6.3.1 relies on techniques similar to those used for
the construction of metrics with harmonic spinors. Roughly speaking, after
possibly rescaling a fixed given metric on M , one has to add sufficiently
many spheres of different sizes, each having one of the li ’s as single and
simple eigenvalue in ] − L, L[. The surgery theorem (Theorem 6.2.2) ensures
that the resulting manifold has the desired eigenvalues modulo a small error.
We refer to [79, Sec. 4] for a detailed proof.
Chapter 7
The Dirac spectrum on non-compact
manifolds
In this chapter we investigate the less familiar situation where the underlying
manifold is non-compact. A good but somewhat not up-to-date reference is
[42, Sec. 8.2]. The Dirac operator of a non-compact Riemannian spin man-
ifold has in general no well-defined spectrum, even if its square does as a
non-negative operator (Section 7.1). If the Riemannian manifold is complete
then its Dirac spectrum - which is well-defined - is composed of a discrete part
and of a disjoint union of intervals. After giving examples where the Dirac
spectrum can be explicitly computed (Section 7.2), we survey the different
situations where the geometry or particular analytical properties of the un-
derlying manifold produce either gaps in their Dirac spectrum (Section 7.3)
or the non-existence of one of both spectral components (Section 7.4).
We have seen in Proposition 1.3.5 that the Dirac operator is essentially self-
adjoint as soon as the underlying Riemannian spin manifold is complete. This
provides the existence of a canonical self-adjoint extension of D, which makes
its spectral theory somewhat easier, see below. In case the manifold is not
complete, there exists no canonical such extension as we have already seen in
Note 1.3.6. However, since the square of D is symmetric and non-negative on
its domain Γc (ΣM ) = {ϕ ∈ Γ(ΣM ) | supp(ϕ) compact}, it admits a canonical
non-negative self-adjoint extension, as was noticed by C. Bär in [44, Sec. 2]:
Theorem 7.2.1 The Dirac operator on (Rn , can) has no point spectrum and
its continuous spectrum is R.
1 x
(D − |λ|Id)ψj,
λ
= χ( + 3je1 ) · (D − |λ|)(ei
λ,x ϕλ, )
φλj, j
1 x
+ grad(χ)( + ke1 ) · ei
λ,x ϕλ, ,
j j
with
n
∂ i
λ,x
D(ei
λ,x ϕλ, ) = el · (e ϕλ, )
∂xl
l=1
n
= iλl el · (ei
λ,x ϕλ, )
l=1
so that
1 x
(D − |λ|Id)ψj,
λ
= grad(χ)( + ke1 ) · ei
λ,x ϕλ, .
jφλj, j
n
A simple transformation formula provides φλj, = j 2 χL2 , from which
12
1 x
(D − |λ|Id)ψj,
λ
= n |grad(χ)( + ke1 )|2 dx1 . . . dxn
j · j 2 χL2 Rn j
106 7 The Dirac spectrum on non-compact manifolds
1 n
= n · j 2 grad(χ)L2
j · j χL2
2
grad(χ)L2
=
jχL2
The second family where the spectrum can be explicitly described is that of
hyperbolic spaces, whose Dirac spectra are given in the following table (note
however that 0 is never a Dirac eigenvalue on RHn as claimed in [65, Cor.
4.6] for n even):
In general not much can be said on both components of the Dirac spectrum.
However, as in the compact case, a spectral gap about 0 occurs as soon as
the scalar curvature is bounded below by a positive constant, even if the
underlying manifold is not complete:
n 2
min (M , [g], ) ≥
λ+ min(σ(Lg )) · Vol(M, g) n ,
n 2
inf (7.3)
4(n − 1) g∈[g]
Vol(M,g)<∞
every complete g ∈ [g] with finite volume [115, Lemma 3.3]. As an example,
if (N n−1 , h) is any closed Riemannian spin manifold with positive scalar
curvature and n ≥ 5, then the Riemannian product (N × R, h ⊕ dt2 ) endowed
with the product spin structure satisfies the assumptions of Theorem 7.3.4,
therefore (7.3) holds [115, Ex. 4.1], where the r.h.s. can be shown to be
positive (see reference in [115]). However the cases n = 3, 4 remain open.
There are particular situations where one of both components of the Dirac
spectrum can be excluded out of geometric considerations. This kind of ques-
tion has attracted a lot of attention in the last years. We choose to present
here five different settings with sometimes non-empty mutual intersection.
Two of them deal with manifolds with cusps. A manifold with cusps can
be written as the disjoint union of a compact manifold with non-empty
boundary together with cusps, which are Riemannian manifolds of the form
(]0, ∞[×N, dt2 ⊕gt ) for some smooth 1-parameter-family of Riemannian met-
rics on the manifold N .
First consider an oriented complete n-dimensional hyperbolic manifold
(recall that a metric is called hyperbolic if it has constant sectional curvature
−1). Assume it to have finite volume. Then M can be shown to possess a
finite number of cusps of the form (]0, ∞[×N n−1 , dt2 ⊕e−2t gflat ) for some flat
metric on some closed manifold N n−1 (see reference in [41]). If M is spin,
then any spin structure on M induces a spin structure on each cusp and
hence on each N n−1 -factor. We call the spin structure trivial along a cusp
if the Dirac operator of the corresponding N has non-zero kernel and non-
trivial otherwise. Since by the decomposition principle (Proposition 7.1.3) the
essential spectrum of D is unaffected by perturbations on a compact subset,
it can be only influenced by the geometry of the cusps. As a striking fact, it
turns out to depend only on the spin structure on M , where one obtains the
following dichotomy as shown by C. Bär [41, Thm. 1]:
dx2
n−1 n−1
dxdyj
g = a00 (x, y) 4 + a0j (x, y) 2 + aij (x, y)dyi dyj ,
x j=1
x i,j=1
The particular form of the metric near the boundary in Theorem 7.4.4
allows furthermore the existence of a nice Weyl’s asymptotic estimate for the
eigenvalue counting function [204, Thm. 3].
In the radically different situation where the curvature is non-negative, one
still may guarantee the point spectrum to be empty or almost empty. How-
ever the hypotheses needed are much stronger. Let (M n , g) be geodesically
starshaped w.r.t. some point x0 . Assume the scalar curvature of (M n , g) to
be non-negative and that particular parts of the sectional curvature remain
pointwise pinched in [0, c(n)r 2 ], where r is the distance function from x0 and
c(n) is a positive constant depending explicitly on n. Then σd (D) = ∅ or {0}
as shown by S. Kawai [153, Thm. 3].
Chapter 8
Other topics related
with the Dirac spectrum
We outline the main topics in relation with the spectrum of Dirac operators
that have been left aside in this overview.
(n − 1)2
λ1 (DS2n ,g ) ≥ 2
4fmax
n n−1+k
and there are at most 2[ 2 ] · eigenvalues of DS2n ,g in the interval
k
2
( n−1
2 +k) ( n−1 +k+1)2
[ 2
fmax , 2 f2 [, for every nonnegative integer k.
max
The proof of Theorem 8.1.1 relies on the following arguments: the SOn -
action allows a dense part of (Sn , g) to be written as a warped product of
Sn−1 with an interval. On this dense part the eigenvalue problem on (Sn , g)
translates into a singular nonlinear differential equation of first order with
boundary conditions at both ends. The rest of the proof involves Sturm-
Liouville theory, we refer to [166] for details.
N. Ginoux, The Dirac Spectrum, Lecture Notes in Mathematics 1976, 113
DOI 10.1007/978-3-642-01570-0 8,
c Springer-Verlag Berlin Heidelberg 2009
114 8 Other topics related with the Dirac spectrum
Note that the inequality in Theorem 8.1.1 is not sharp for the standard
2
metric on Sn since λ1 (DS2n ,can ) = n4 (see Theorem 2.1.3). However Theorem
8.1.1 provides sharp asymptotical eigenvalue estimates in the two follow-
ing situations. First consider the cylinder C n (L) :=]0, L[×Sn−1 with half
n-dimensional spheres glued at both ends. Obviously C n (L) admits an iso-
metric SOn -action for which fmax = 1, in particular Theorem 8.1.1 implies
2
that λ1 (DC2 n (L) ) ≥ (n−1)
4 . On the other hand, C n (L) sits in Rn+1 by con-
struction; now C. Bär’s upper bound (5.19) in terms of the averaged total
squared mean curvature is not greater than
so that [166]
(n − 1)2
lim λ1 (DC2 n (L) ) = .
L→∞ 4
x2
For the 2-dimensional ellipsoid Ma := {x ∈ R3 | x21 + x22 + a23 = 1} (where
a > 0) the maximal length of S1 -orbits is 2π, so that by Theorem 8.1.1 the
2
inequality λ1 (DM a
) ≥ 14 holds. Combining this with the upper bound (5.8)
provides [165]
2 1
lim λ1 (DM )= .
a→∞ a
4
The technique of separation of variables used in the proof of Theorem 8.1.1
also provides a lower eigenvalue bound on warped product fibrations over
S1 in terms of the Dirac eigenvalues of the fibres, see [167, Thm. 2]. As for
the case of higher dimensional fibres over arbitrary base manifolds, the only
family which has been considered so far is that of warped products with fibre
Sk with k ≥ 2, where decomposing the Dirac operator into block operator
matrices provides similar results to those of Theorem 8.1.1, see [169].
Another natural but completely different way to study the Dirac eigen-
values consists in comparing them with those of other geometric operators.
Hijazi’s inequality (3.18) is already of that kind since μ1 is the smallest eigen-
value of the conformal Laplace operator. As for spectral comparison results
between the Dirac and the scalar Laplace operators, the first ones were proved
by M. Bordoni. They rely on a very nice general comparison principle between
two operators satisfying some kind of Kato-type inequality. The estimate
which can be deduced reads as follows.
Theorem 8.1.2 (M. Bordoni [60]) Let 0 = λ0 (Δ) < λ1 (Δ) ≤ λ2 (Δ) ≤ . . .
be the spectrum of the scalar Laplace operator Δ on a closed n(≥ 2)-
dimensional Riemannian spin manifold (M n , g). Then for any positive
integer N [60, Prop. 4.20]
8.1 Other eigenvalue estimates 115
n λk (Δ)
λ2N (D2 ) ≥ inf (S) + , (8.1)
4(n − 1) M
n
2(2[ 2 ] + 1)2
N
where k = [ n ].
2[ 2 ] +1
n n λ1 (Δ)
[ inf (S), inf (S) + [,
4(n − 1) M 4(n − 1) M
n
2(2[ 2 ] + 1)2
Inequality (8.2) is optimal and sharp for M 2 = S2 : indeed for any Rieman-
λ1 (Δ )
nian metric g one has λ1 (DS22 ,g ) ≥ S ,g 2
2 as a straightforward consequence
of Bär’s inequality (3.17) and Hersch’s inequality (3.22). Moreover, (8.2) com-
pletes [1] where I. Agricola, B. Ammann and T. Friedrich prove the existence
of a 1-parameter family (gt )t≥0 of S1 -invariant Riemannian metrics on T2 for
which, in the same notations as just above, λ1 (ΔT2 ,gt ) < λ1 (DT2 2 ,gt ) for any
t ≥ 0, where T2 is endowed with its trivial spin structure. The inequality
λk (ΔT2 ,g ) ≥ λk (DT2 2 ,g ) for k large enough and for particular metrics g on T2
with trivial spin structure has been proved independently by M. Kraus [168].
In the case where the manifold sits as a hypersurface in some spaceform,
the best known result is the following.
Theorem 8.1.4 (C. Bär [40]) Let (M n , g) be isometrically immersed into
Rn+1 or Sn+1 and carry the induced spin structure, then [40, Thm. 5.1]
n2
λN (D2 ) ≤ sup(H 2 ) + κ + λ[ N −1 (Δ) (8.3)
4 M 2μ ]
116 8 Other topics related with the Dirac spectrum
1 2
μj := λj (D2 ) + n sup(H 2 ) − inf (S) ,
4 M M
where H and S are the mean and the scalar curvature of M respectively.
Then for any k ≥ 1
k
4
(μk+1 − μj )(μk+1 − (1 + )μj ) ≤ 0. (8.4)
j=1
n
4
k
1
μk+1 ≤ (1 + ) μj ,
k n j=1
Theorem 8.3.1 implies the existence of a uniform lower eigenvalue bound for
the Dirac operator in the following family: there exists an ε = ε(n, K, d) > 0
such that on every n-dimensional closed Riemannian spin manifold (M n , g)
with |Ksec (M n , g)| < K, diam(M n , g) < d, S(M n , g) > −ε and which is
not spin diffeomorphic to a nilmanifold with trivial spin structure the rth
eigenvalue of D2 satisfies
λr (D2 ) ≥ ε.
118 8 Other topics related with the Dirac spectrum
The choice for r, which looks a priori curious, is actually optimal since the
product of a so-called K3-surface with a torus carries exactly r − 1 linearly
independent parallel spinors, see [25, Ex. (2) p.411]. The proof of Theorem
8.3.1 makes use of an approximation result by U. Abresch (see reference in
[25]) in an essential way, we refer to [25, Sec. 7] for details.
Under the supplementary assumption of a lower bound on the volume, the
metric can even be shown to stay near to some with parallel spinors.
Theorem 8.3.2 (B. Ammann and C. Sprouse [25]) Let K, d, V, δ be
positive real constants, n ≥ 2 be an integer, r := 1 if n = 2, 3 and
n
r := 2[ 2 ]−1 + 1 if n ≥ 4. Then there exists an ε = ε(n, K, d, V, δ) > 0
such that for every n-dimensional closed Riemannian spin manifold (M n , g)
with
|Ksec (M n , g)| < K, diam(M n , g) < d, S(M n , g) > −ε, Vol(M n , g) > V
2 1,α
and λr (DM n ,g ) < ε, the metric g is at C -distance at most δ to a metric
admitting a non-zero parallel spinor.
The proof of Theorem 8.3.2 relies on a similar general eigenvalue pinch-
ing valid for arbitrary rough Laplacians on arbitrary vector bundles due
to P. Petersen (see reference in [25]) and on the Schrödinger-Lichnerowicz
formula (3.2). Petersen’s method can also be applied to the rough Laplacian
associated to the deformed covariant derivative X
→ ∇X + ρX· and in this
case it provides the following:
Theorem 8.3.3 (B. Ammann and C. Sprouse [25]) Let K, d, V, ρ, δ be
positive real constants, n ≥ 2 be an integer, r := 1 if n = 2, 3 and r :=
n
2[ 2 ]−1 + 1 if n ≥ 4. Then there exists an ε = ε(n, K, d, V, ρ, δ) > 0 such that
for every n-dimensional closed Riemannian spin manifold (M n , g) with
Theorem 8.3.4 (A. Vargas [231]) The conclusion of Theorem 8.3.3 holds
with
3 if n = 6 or n ≡ 1 (4)
r := n+9
4 if n ≡ 3 (4).
8.4 Spectrum of other Dirac-type operators 119
Up to now we have concentrated onto the fundamental (or spin) Dirac opera-
tor on a spin manifold. As already mentioned at the beginning of Chapter 1,
Dirac-type operators may be defined in the more general context where a so-
called Clifford bundle [173, Sec. II.3] is at hand. Roughly speaking, a Clifford
bundle is given by a Hermitian vector bundle together with a covariant deriva-
tive and on which the tangent bundle acts by Clifford multiplication such that
all three objects (Hermitian metric, covariant derivative and Clifford multi-
plication) are compatible with each other in the sense of Definition 1.2.2 and
Proposition 1.2.3. The associated Dirac operator is defined as the Clifford
multiplication applied to the covariant derivative. One may add a zero-order
term and obtain a so-called Dirac-Schrödinger operator. In this section we dis-
cuss spectral results in relation with the spinc Dirac operator, with twisted
Dirac-Schrödinger operators, with Dirac operators associated to particular
geometrically relevant connections, with the basic Dirac operator and in the
pseudo-Riemannian setting.
First, the concept of spin structure may be weakened to that of spinc
structure, whose structure group is the spinc group Spincn := Spinn × S /Z2 .
1
We note however that little has been done in the spinc context in comparison
with the spin one.
If the underlying space is again our familiar spin manifold (M n , g) and if we
choose an arbitrary Riemannian or Hermitian vector bundle E over M , then
the tensor product bundle ΣM ⊗ E carries a canonical Clifford multiplication
(extend the Clifford multiplication by the identity on the second factor). If
we endow E with a metric covariant derivative, then we obtain a structure
of Clifford bundle and an associated Dirac operator called Dirac operator of
120 8 Other topics related with the Dirac spectrum
E
M twisted with E. This operator is usually denoted by DM . For example,
the Euler operator d + δ can be seen as the Dirac operator of M twisted with
ΣM : this follows essentially from (1.2) and may actually be stated without
any spin structure on M [173, Sec. II.6]. Another prominent example is the
Dirac operator of a spin submanifold twisted with the spinor bundle of its
normal bundle (where the latter is assumed to be spin). Various studies have
been devoted to the spectrum of twisted Dirac operators, therefore we restrict
ourselves to a few ones which we hope to be representative. We include all
that concerns Dirac-Schrödinger operators, since in that case the zero order
term mainly translates the upper or lower bounds by a constant.
Let first E be as above, M be closed and f be a smooth real function
n by κ1 the smallest eigenvalue on M of the pointwise linear
on M . Denote
operator k,l=1 ek · el · ReEk ,el , where RE is the curvature tensor of the chosen
covariant derivative on E (and (ek )1≤k≤n is a local o.n.b. of TM ). If the
inequalities n(S + κ1 ) > (n − 1)f 2 > 0 hold on M , then any eigenvalue λ of
the Dirac-Schrödinger operator DM E
− f acting on Γ(ΣM ⊗ E) satisfies [105,
Prop. 4.1]
1 n 2
λ2 ≥ inf (S + κ1 ) − |f | . (8.5)
4 M n−1
Inequality (8.5), which can be deduced from a clever choice of modified covari-
ant derivative, stands for the analog of Friedrich’s inequality in this context,
see [105] for other kinds of estimates and references to earlier works on that
topic (such as [196]). In the particular case where n = 4, f = 0, E is ar-
bitrary and carries a selfdual covariant derivative, the estimate (8.5) can be
enhanced using the decomposition ΣM = Σ+ M ⊕ Σ− M and the vanishing of
one half of the auxiliary curvature term computed from RE : H. Baum proved
[52, Thm. 2] that
1
λ2 ≥ inf (S)
3 M
E
for any eigenvalue λ of DM , which is exactly Friedrich’s inequality (3.1) for
the eigenvalues of the spin Dirac operator.
Staying in dimension 4, if the spin manifold (M 4 , g) carries a Hermitian
structure J (i.e., an orthogonal complex structure on TM ) then one is led
0,∗
to the Dirac operator twisted with√ E = ΣM = Λ TM which is nothing
else than the Dolbeault operator 2(∂ + ∂). Although Kirchberg’s inequality
(3.10) does not apply, sharp lower bounds for the eigenvalues of the Dolbeault
operator are still available: B. Alexandrov, G. Grantcharov and S. Ivanov
proved [6, Thm. 2] that
1
λ2 ≥ inf (S)
6 M
√
for any eigenvalue λ of 2(∂ + ∂). Beware that equality cannot occur for a
non-flat Kähler metric because of (3.10). The proof of that inequality relies
on Weitzenböck formulas and the clever choice of twistor operators asso-
ciated to a canonical one-parameter-family of connections, we refer to [6]
8.4 Spectrum of other Dirac-type operators 121
for details. Besides, we mention that upper eigenvalue bounds for particular
twisted Dirac operators have been obtained in [52], [40] and [101].
From the point of view of geometers investigating the integrability of par-
ticular G-structures, there exists another interesting family of Dirac-type
1 1
operators which are usually denoted by D 3 and defined by D 3 := Dg + T4 ·,
where T is some given 3-form and Dg is the spin Dirac operator on the
Riemannian spin manifold (M n , g). For example if (M n , g) is a so-called re-
1
ductive homogeneous space then D 3 is the so-called Kostant Dirac operator
1
(see reference in [3]); if (M n , g) is a Hermitian manifold then D 3 coincides
with the Dolbeault-operator defined just above. In case T is the charac-
teristic torsion of a 5-dimensional closed spin Sasaki manifold with scalar
curvature bounded from below, the use of suitable deformations of the con-
nection by polynomials of the torsion form allowed I. Agricola, T. Friedrich
1
and M. Kassuba to prove the following estimates of any eigenvalue λ of (D 3 )2
[3, Thm. 4.1]:
1 √
(1 + 1 inf M (S))2 if − 4 < S ≤ 4(9 + 4 5)
16 4
λ ≥ √
5 inf M (S) if S ≥ 4(9 + 4 5).
16
where q ≥ 2 stands for the codimension of the foliation and S tr for its
transversal scalar curvature. In case the normal bundle of the foliation carries
122 8 Other topics related with the Dirac spectrum
At this point one should beware that the mass endomorphism of (Sn , [can])
vanishes and that this does not characterize the round sphere since flat tori
also have vanishing mass endomorphism. We refer to [23] for the details. For a
generalization of the Bär-Hijazi-Lott invariant to manifolds with non-empty
boundary we refer to [212, 214].
124 8 Other topics related with the Dirac spectrum
a metric with constant sectional curvature −1). Those sequences only exist
in dimensions 2 and 3 and, provided the convergence respects the spin struc-
tures in some sense, the limit manifold must have discrete Dirac spectrum in
dimension 3 whereas it may have continuous spectrum in dimension 2, see
references in [208] where a precise description of hyperbolic degenerations is
recalled. In case the limit manifold M is assumed to have discrete Dirac spec-
trum, F. Pfäffle proved the convergence of the Dirac spectrum of (Mj )j∈ N in
the following sense [208, Thm. 1.2] (see also [207]): For all ε > 0 and Λ ≥ 0,
there exists an N ∈ N such that for all j ≥ N the real number Λ lies neither
in the spectrum of D nor in that of DMj , both Dirac operators DMj and D
have only discrete eigenvalues and no other spectrum in [−Λ, Λ], they have
the same number m of eigenvalues in [−Λ, Λ] which can be ordered so that
(j)
|λk − λk | ≤ ε holds for all 1 ≤ k ≤ m.
The diameter of the converging sequence of degenerating hyperbolic mani-
folds cannot be controlled since the limit-manifold must have a finite number
of so-called cusps, which by definition are unbounded. The third context to
have been considered precisely deals with the situation where both the diam-
eter and the sectional curvature of the converging sequence are assumed to
remain bounded. In that case J. Lott proved the following very general result
[183]. Consider a sequence (gj )j∈ N of bundle metrics on the total space of a
spin fibre bundle M over a base spin manifold B. Assume the fibre length to
go to 0 as j tends to ∞ while both the diameter and the sectional curvature
of (M, gj ) remain bounded. Then the Dirac spectrum of (M, gj ) converges
in the sense just above to that of some differential operator of first order on
B which can be explicitly constructed. Since a precise formulation and the
discussion of the results would require too many details we recommend the
introduction of [183].
8.7 Eta-invariants
As we have seen in Theorem 1.3.7, the Dirac spectrum of any closed n-dimen-
sional Riemannian spin manifold is symmetric w.r.t. the origin in dimension
n ≡ 3 (4). To measure the asymmetry of the Dirac spectrum in case n ≡ 3 (4),
Atiyah, Patodi and Singer introduced [27] the so-called η-invariant of D which
is defined by η(D) := η(0, D), where, for every s ∈ C with e(s) > n,
sgn(λj )
η(s, D) := .
|λj |s
λj =0
Although this section has more to do with physics as with the Dirac spectrum,
we include it because on the one hand the proofs of the results presented
involve simple spinorial techniques as already used above, and on the other
hand positive mass theorems nowadays play a central role in many other
topics of global analysis such as the Yamabe problem. A good but not up-to-
date reference for that topic is [125].
A positive mass theorem (sometimes called positive energy theorem) is a
two-fold statement reading roughly as follows: Let (M n , g) be a Riemannian
manifold which is asymptotic to a model manifold (in a sense that must be
precised) and some of which curvature invariant satisfies a pointwise inequal-
ity, then some asymptotic geometric invariant called its mass also satisfies a
similar inequality and, if this latter inequality is an equality, then the whole
128 8 Other topics related with the Dirac spectrum
manifold is globally isometric to the original model manifold. To fix the ideas
we concentrate from now on onto the original positive mass theorem as proved
by R. Schoen and S.-T. Yau [215, 216] and independently by E. Witten [235]
in the spinorial setting, in particular we leave aside all recent developments
in what has become a whole field of research at the intersection between
mathematics and general relativity, see e.g. [236] for references.
Let (M n , g) be a Riemannian manifold of dimension n ≥ 3. Call it asymp-
totically flat of order τ ∈ R if there exists a compact subset K ⊂ M , a positive
real number R and a diffeomorphism M \ K −→ {x ∈ Rn , |x| > R} such
that the pushed-out metric fulfills: gij − δij = O(|x|−τ ), ∂xijk = O(|x|−τ −1 )
∂g
∂2g
ij
and ∂xk ∂x l
= O(|x|−τ −2 ) as |x| → ∞, for all 1 ≤ i, j, k, l ≤ n. Given such a
manifold (M n , g), set
n
1 ∂gij ∂gii
m(g) := · lim ( − )νj dA,
16π r→∞ Sr i,j=1 ∂xi ∂xj
where ν denotes here the outer unit normal to Sr = ∂Br . The miracle
in Witten’s proof happens here: it can be easily shown that the bound-
ary term Sr ∇ν ψ, ψ
dA is asymptotic to m(g) times some finite positive
constant c as r goes to ∞. After passing to the limit one is left with
m(g) = c( M (|∇ψ|2 + S4 |ψ|2 )vg , which implies m(g) ≥ 0. The equality
m(g) = 0 requires ψ to be parallel for any ψ constructed this way, in par-
ticular the spinor bundle of (M n , g) must be trivialized by parallel spinors,
from which the identity (M n , g) = (Rn , can) can be deduced. An alternative
spinorial proof but with supplementary assumptions on the dimension or the
Weyl tensor has been given by B. Ammann and E. Humbert [21] using the
Green’s operators associated to the Dirac operator, see Section 8.5.
Appendix A
The twistor and Killing spinor equations
P ψ = 0, (A.1)
131
132 Appendix A The twistor and Killing spinor equations
) (and
positive if n is odd), gives a − 2i -Killing spinor. We deduce from Proposition
A.4.1.2 below that there are no other Killing spinors on (H n , can) than
those constructed and from Proposition A.2.1.3.b) that, if n ≥ 3, then the
space of twistor-spinors is exactly the direct sum of both spaces (for n = 2
see Proposition A.2.3).
P = e 2 ◦ P ◦ e− 2 ,
u u
u
where P := Pg . In particular e 2 ψ is a twistor-spinor on (M n , g).
2. If S denotes the scalar curvature of (M n , g) then
nS
D2 ψ = ψ. (A.3)
4(n − 1)
3. If n ≥ 3 then:
a) for every X ∈ T M ,
n 1 S
∇X (Dψ) = − Ric(X) · ψ + X ·ψ . (A.4)
n−2 2 4(n − 1)
n
b) dim(Ker(P )) ≤ 2[ 2 ]+1 .
c) The zero-set of ψ is either discrete in M n or M n itself.
d) If (M n , g) is Einstein with S = 0 then ψ is the sum of two non-parallel
Killing spinors.
e) If |ψ| is a non-zero constant then (M n , g) is Einstein. Moreover either
S = 0 and ψ is parallel or S > 0 and ψ is the sum of two real non-
parallel Killing spinors.
4. If M n is closed then Ker(P ) is finite dimensional. In the case ψ = 0 if
furthermore S is constant then either S = 0 and ψ is parallel or S > 0
and ψ is the sum of two real non-parallel Killing spinors.
Proof :
1. For any ϕ ∈ Γ(ΣM ), f ∈ C ∞ (M ) and X ∈ T M one has
1
PX (f ϕ) = ∇X (f ϕ) + X · D(f ϕ)
n
(1.11)
= X(f )ϕ + f ∇X ϕ
1 f
+ X · grad(f ) · ϕ + X · Dϕ
n n
1
= X(f )ϕ + X · grad(f ) · ϕ + f PX ϕ. (A.5)
n
A.2 Elementary properties of twistor-spinors 135
1
P X ϕ = ∇X ϕ + X·Dϕ
n
1 X(u)
= ∇X ϕ − X · gradg (u) · ϕ − ϕ
2 2
e−u n−1
+ X·(Dϕ + gradg (u) · ϕ)
n 2
1 X(u)
= PX ϕ − X · gradg (u) · ϕ − ϕ, (A.6)
2n 2
so that
u u
u (A.5) e 2 e2 u
P X (e 2 ϕ) = X(u)ϕ + X·gradg (u)· ϕ + e 2 P X ϕ
2 2n
u u
(A.6) e 2 e2
= X(u)ϕ + X·gradg (u)· ϕ
2 2n
u 1 X(u)
+e 2 (PX ϕ − X · gradg (u) · ϕ − ϕ)
2n 2
u
= e 2 PX ϕ,
which shows 1.
2. Let X, Y ∈ Γ(T M ), then
1 1
∇X ∇Y ψ = − ∇X Y · Dψ − Y · ∇X Dψ,
n n
from which we deduce
∇
RX,Y ψ = ∇[X,Y ] ψ − [∇X , ∇Y ]ψ
1
= (Y · ∇X Dψ − X · ∇Y Dψ). (A.7)
n
1
n
1
Ric(X) · ψ = (ej · ej · ∇X Dψ − ej · X · ∇ej Dψ)
2 n j=1
1
= (−(n − 2)∇X Dψ + X · D2 ψ). (A.8)
n
Hence
1
n
1
− Sψ = ej · Ric(ej ) · ψ
2 2 j=1
136 Appendix A The twistor and Killing spinor equations
1
n
(A.8)
= (−(n − 2)ej · ∇ej Dψ + ej · ej · D2 ψ)
n j=1
2(n − 1) 2
= − D ψ
n
which shows 2.
3. Assume n ≥ 3.
a) Coming back to (A.8) using (A.3) we obtain
1 n−2 S
Ric(X) · ψ = − ∇X Dψ + X ·ψ
2 n 4(n − 1)
X(|ψ|2 ) = 2e ( ∇X ψ, ψ
)
2
= − e ( X · Dψ, ψ
)
n
one has
2
Hess(|ψ|2 )(X, Y ) = − e ( Y · ∇X Dψ, ψ
+ Y · Dψ, ∇X ψ
)
n
(A.4) 1
= e( Y · Ric(X) · ψ, ψ
)
n−2
S
− e( Y · X · ψ, ψ
)
2(n − 1)(n − 2)
2
+ e( Y · Dψ, X · Dψ
)
n2
|ψ|2
= − ric(X, Y )
n−2
S|ψ|2 2|Dψ|2
+ + g(X, Y ). (A.9)
2(n − 1)(n − 2) n2
A.2 Elementary properties of twistor-spinors 137
2|Dψ|2
If ψp = 0 then Hess(|ψ|2 )p = n2 p gp . In the case ψ = 0 one must have
(Dψ)p = 0 (otherwise the ∇T -parallel section ψ ⊕Dψ would vanish at p
and hence identically), therefore the Hessian of |ψ|2 is positive definite
at p and the result follows.
d) If (M n , g) is Einstein then (A.4) becomes
n S S
∇X Dψ = (− + )X · ψ
n − 2 2n 4(n − 1)
S
=− X ·ψ
4(n − 1)
1 1
∇X ψ±1 = (∇X ψ ± ∇X Dψ)
2 λ
1 1 1 λ2
= − X · Dψ ± (− X · ψ)
2 n λ n
λ 1
= ∓ X · (ψ ± Dψ)
2n λ
λ
= ∓ X · ψ±1 ,
n
that is, (M n , g) is Einstein. Moreover from the latter equation the scalar
curvature of (M n , g) is then given by
4(n − 1) |Dψ|2
S= · ≥ 0. (A.10)
n |ψ|2
1
|∇ϕ|2 = |P ϕ|2 + |Dϕ|2 (A.11)
n
and
(A.11) 1 2
P ∗P = ∇∗ ∇ − D
n
(1.15) n−1 ∗ S
= ∇ ∇− Id (A.12)
n 4n
the operator P ∗ P is elliptic, hence its kernel is finite-dimensional. If fur-
thermore ψ = 0 and S is constant then integrating the Hermitian product
of (A.3) with ψ one obtains
nS
|ψ| vg =
2
D2 ψ, ψ
vg
4(n − 1) M
M
= |Dψ|2 vg ,
M
which shows S ≥ 0. On the other hand (A.3) already stands for the
limiting-case in T. Friedrich’s inequality (3.1), so that ψ must either be
parallel (in case S = 0) or the sum of two real Killing spinors (in case
S > 0). This shows 4. and concludes the proof.
1
PX ψ = X(f+ )ϕ+ + X(f− )ϕ− + X · (df+ · ϕ+ + df− · ϕ− ).
2
For the Kähler structure J associated to g and the orientation of M one has
however
2
X · Y · ϕ± = g(X, ej )g(Y, ek )ej · ek · ϕ
j,k=1
2
= g(X, ej )g(Y, ej )ej · ej · ϕ
j=1
+(g(X, e1 )g(Y, e2 ) − g(X, e2 )g(Y, e1 ))e1 · e2 · ϕ±
= −g(X, Y )ϕ − g(X, J(Y ))e1 · e2 · ϕ±
= (−g(X, Y ) ± ig(X, J(Y )))ϕ±
= −2g(X, p± (Y ))ϕ± ,
Note that Proposition A.2.3 together with Note A.2.2 imply in particular
that Proposition A.2.1.3.c) still holds in dimension n = 2, since holomorphic
and anti-holomorphic functions on a surface vanish either on a discrete subset
or identically.
140 Appendix A The twistor and Killing spinor equations
Proof : The statement on the zero set of ψ has been proved in Proposition
ψ
A.2.1.3.c). From Proposition A.2.1.1 the spinor φ := |ψ| is a twistor-spinor
n
on (M , g). In dimension n ≥ 3 since it has constant norm it is the sum of
two real Killing spinors (Proposition A.2.1.3.e)); furthermore
(1.18) n − 1 grad(|ψ|2 )
Dφ = |ψ|2 (Dφ − · φ)
2 |ψ|2
grad(|ψ|) 1 n − 1 grad(|ψ|2 )
= |ψ|2 (− ·ψ+ Dψ − · φ)
|ψ| 2 |ψ| 2 |ψ|2
n
= |ψ|Dψ − grad(|ψ|2 ) · φ,
2
Note that the equivalent statement in dimension n = 2 does not hold because
of Proposition A.2.3.
For general M W. Kühnel and H.-B. Rademacher proved that the Ricci-flat
1
metric |ψ| 4 g on M
n
\ Zψ is either flat or locally irreducible, more precisely:
following:
a) n = 2m ≥ 4, Hol = SUm and N = 2.
b) n = 4m ≥ 8, Hol = Spm and N = m + 1.
c) n = 7, Hol = G2 and N = 1.
d) n = 8, Hol = Spin7 and N = 1.
Theorem A.3.2, a proof of which can be found in the beautiful paper
[172], actually requires Mc.K. Wang’s classification of manifolds with non-
zero parallel spinors, see Theorem A.4.2 below. Besides, we mention that up
to now no example with reduced holonomy of type b), c) or d) has been
described (an example with Hol = SUm is constructed in [170]).
S
Ric(X) · ψ = (2(n − 2)α2 + )X · ψ
2(n − 1)
142 Appendix A The twistor and Killing spinor equations
spinor if and only if it is either the round sphere (Sn , can) (in which case
n n
(p, q) = (2[ 2 ] , 2[ 2 ] )) or one of the following:
a) (4m + 1)-dimensional Einstein-Sasaki, m ≥ 1, and in that case (p, q) =
(1, 1).
b) (4m+3)-dimensional Einstein-Sasaki but not 3-Sasaki, m ≥ 2, and in that
case (p, q) = (0, 2).
c) (4m+3)-dimensional 3-Sasaki, m ≥ 2, and in that case (p, q) = (0, m+2).
d) 6-dimensional nearly Kähler non-Kähler, and in that case (p, q) = (1, 1).
e) 7-dimensional with a nice 3-form φ satisfying ∇φ = ∗φ but not Sasaki,
and in that case (p, q) = (0, 1).
f) 7-dimensional Sasaki but not 3-Sasaki, and in that case (p, q) = (0, 2).
g) 7-dimensional 3-Sasaki, and in that case (p, q) = (0, 3).
For the definitions of 3-Sasaki structures and nice forms as well as the proof
of Theorem A.4.3 we refer to [37]. Parts of this classification had already been
obtained in [89, 129, 90, 91, 92, 117, 87]. As an interesting fact, two higher
eigenvalues of (n = 4m + 3)-dimensional 3-Sasaki manifolds can be explicitly
computed in terms of the scalar curvature:A. Moroianu showed [199] that on
such manifolds both − 4(n−1) nS
− 1 and nS
4(n−1) + 2 are Dirac eigenvalues
with multiplicities at least 3m and m respectively. The proof relies on a clever
combination of the Killing vector fields provided by the 3-Sasaki structure
and the Killing spinors.
In the last case (α ∈ iR∗ ) the classification turns out to rely on totally
different arguments. Studying in detail the level sets of the length function
of an imaginary Killing spinor H. Baum proved the following theorem, which
relies on Theorem A.4.2 but where the assumption π1 (M ) = 1 turns out not
to be necessary.
Theorem A.4.4 (H. Baum [50]) Let (M n , g) be an (n ≥ 2)-dimensional
connected complete Riemannian spin manifold without boundary. Then
(M n , g) admits a non-trivial α-Killing spinor with α ∈ iR∗ if and only if
it is isometric to a warped product of the form
(N × R, e4iαt h ⊕ dt2 ),
∇X ψ = αX · ψ
145
146 Bibliography
46. , The First Dirac Eigenvalue on Manifolds with Positive Scalar Curva-
ture, Proc. Amer. Math. Soc. 132 (2004), 3337–3344.
47. C. Bär, P. Gauduchon, A. Moroianu, Generalized cylinders in semi-Riemannian
and Spin geometry, Math. Z. 249 (2005), no. 3, 545–580.
48. C. Bär, P. Schmutz, Harmonic spinors on Riemann surfaces, Ann. Glob. Anal.
Geom. 10 (1992), no. 3, 263–273.
49. H. Baum, Spin-Strukturen und Dirac-Operatoren über pseudo-Riemannschen
Mannigfaltigkeiten, Dissertation A, Humboldt-Universität zu Berlin, 1980,
Teubner-Texte zur Mathematik 41 (1981), Teubner Verlag Leipzig.
50. , Complete Riemannian manifolds with imaginary Killing spinors, Ann.
Glob. Anal. Geom. 7 (1989), no. 3, 205–226.
51. , An upper bound for the first eigenvalue of the Dirac operator on compact
spin manifolds, Math. Z. 206 (1991), no. 3, 409–422.
52. , Eigenvalue estimates for Dirac operators coupled to instantons, Ann.
Glob. Anal. Geom. 12 (1994), no. 2, 193–209.
53. , A remark on the spectrum of the Dirac operator on pseudo-
Riemannian spin manifolds, preprint SFB 288 136 (1994), available at
http://www.mathematik. hu-berlin.de/∼baum
54. H. Baum, T. Friedrich, Spektraleigenschaften des Dirac-Operators. Die Funda-
mentallösung seiner Wärmeleitungsgleichung und die Asymptotenentwicklung
seiner Zeta-Funktion, J. Diff. Geom. 15 (1980), 1–26.
55. , Eigenvalues of the Dirac operator, twistors and Killing spinors on Rie-
mannian manifolds, Clifford algebras and spinor structures, 243–256, Math.
Appl. 321, Kluwer Acad. Publ., Dordrecht, 1995.
56. H. Baum, T. Friedrich, R. Grunewald, I. Kath, Twistor and Killing spinors
on Riemannian manifolds, Teubner-Texte zur Mathematik, Band 124, Teubner-
Verlag Stuttgart/Leipzig 1991.
57. F.A. Belgun, N. Ginoux, H.-B. Rademacher, A singularity theorem for twistor-
spinors, Ann. Inst. Fourier 57 (2007), no. 4, 1135–1159.
58. N. Berline, E. Getzler, M. Vergne, Heat kernels and Dirac operators,
Grundlehren Text Editions, Springer-Verlag, 2003.
59. B. Booß-Bavnbek, K.P. Wojciechowski, Elliptic boundary problems for Dirac
operators, Mathematics: Theory & Applications, Birkhäuser Boston (1993).
60. M. Bordoni, Spectral estimates for Schrödinger and Dirac-type operators on
Riemannian manifolds, Math. Ann. 298 (1994), no. 4, 693–718.
61. M. Bordoni, O. Hijazi, Eigenvalues of the Kählerian Dirac operator, Lett. Math.
Phys. 58 (2001), no. 1, 7–20.
62. J.-P. Bourguignon, P. Gauduchon, Spineurs, opérateurs de Dirac et variations
de métriques, Comm. Math. Phys. 144 (1992), no. 3, 581–599.
63. J.-P. Bourguignon, O. Hijazi, J.-L. Milhorat, A. Moroianu, A Spinorial approach
to Riemannian and Conformal Geometry, in preparation.
64. T. Branson, O. Hijazi, Vanishing theorems and eigenvalue estimates in Rieman-
nian spin geometry, Int. J. Math. 8 (1997), no. 7, 921–934.
65. U. Bunke, The spectrum of the Dirac operator on the hyperbolic space, Math.
Nachr. 153 (1991), 179–190.
66. , Upper bounds of small eigenvalues of the Dirac operator and isometric
immersions, Ann. Glob. Anal. Geom. 9 (1991), no. 2, 109–116.
67. , On the gluing problem for the η-invariant, J. Diff. Geom. 41 (1995),
no. 2, 397–448.
68. M. Cahen, A. Franc, S. Gutt, Spectrum of the Dirac operator on complex pro-
jective space P2q−1 (C), Lett. Math. Phys. 18 (1989), no. 2, 165–176.
69. , Erratum to: “The spectrum of the Dirac operator on complex projective
space P2q−1 (C)”, Lett. Math. Phys. 32 (1994), no. 4, 365–368.
70. M. Cahen, S. Gutt, L. Lemaire, P. Spindel, Killing spinors, Bull. Soc. Math.
Belg. Sér. A 38 (1986), 75–102.
148 Bibliography
96. , Eigenvalue estimates of the Dirac operator depending on the Ricci ten-
sor, Math. Ann. 324 (2002), no. 4, 799–816.
97. N. Ginoux, Opérateurs de Dirac sur les sous-variétés, thèse de doctorat,
Université Henri Poincaré, Nancy (2002).
98. , Reilly-type spinorial inequalities, Math. Z. 241 (2002), no. 3, 513–525.
99. , Une nouvelle estimation extrinsèque du spectre de l’opérateur de Dirac,
C. R. Acad. Sci. Paris Sér. I 336 (2003), no. 10, 829–832.
100. , Remarques sur le spectre de l’opérateur de Dirac, C. R. Acad. Sci. Paris
Sér. I 337 (2003), no. 1, 53–56.
101. , Dirac operators on Lagrangian submanifolds, J. Geom. Phys. 52 (2004),
no. 4, 480–498.
102. , A short survey on eigenvalue estimates for the Dirac operator on com-
pact Riemannian spin manifolds, Oberwolfach Report 53 (2006), 3140–3141.
103. , The spectrum of the Dirac operator on SU2 /Q8 , manuscr. math. 125
(2008), no. 3, 383–409.
104. N. Ginoux, G. Habib, A spectral estimate for the Dirac operator on Riemannian
flows, preprint (2009).
105. N. Ginoux, B. Morel, On eigenvalue estimates for the submanifold Dirac oper-
ator, Int. J. Math. 13 (2002), no. 5, 533–548.
106. S. Goette, Äquivariante eta-Invarianten homogener Räume, Doktorarbeit,
Shaker Verlag, Aachen, 1997.
107. , Equivariant eta-Invariants on Homogeneous Spaces, Math. Z. 232
(1999), no. 1, 1–42.
108. , Eta invariants of homogeneous spaces, arXiv:math.DG/0203269 (2002),
to appear in Pure Appl. Math. Q.
109. S. Goette, U. Semmelmann, Spinc Structures and Scalar Curvature Estimates,
Ann. Glob. Anal. Geom. 20 (2001), no. 4, 301–324.
110. , Scalar curvature estimates for compact symmetric spaces, Diff. Geom.
Appl. 16 (2002), no. 1, 65–78.
111. , The point spectrum of the Dirac operator on noncompact symmetric
spaces, Proc. Amer. Math. Soc. 130 (2002), no. 3, 915–923.
112. J. Gracia-Bondı́a, J.C. Várilly, H. Figueroa, Elements of noncommutative geom-
etry, Birkhäuser, 2001.
113. J.-F. Grosjean, E. Humbert, The first eigenvalue of Dirac and Laplace operators
on surfaces, arXiv:math.DG/0609493 (2006).
114. N. Große, On a conformal invariant of the Dirac operator on noncompact man-
ifolds, Ann. Glob. Anal. Geom. 30 (2006) no. 4, 407–416.
115. , The Hijazi inequality on conformally parabolic manifolds,
arXiv:0804.3878 (2008).
116. , On a spin conformal invariant on open manifolds, Doktorarbeit, Uni-
versität Leipzig, 2008.
117. R. Grunewald, Six-dimensional Riemannian manifolds with a real Killing
spinor, Ann. Glob. Anal. Geom. 8 (1990), no. 1, 43–59.
118. C. Guillarmou, S. Moroianu, J. Park, Eta invariant and Selberg Zeta function of
odd type over convex co-compact hyperbolic manifolds, arXiv:0901.4082 (2009).
119. G. Habib, Tenseur d’impulsion-énergie et Feuilletages, thèse de doctorat, Insti-
tut Élie Cartan - Université Henri Poincaré, Nancy (2006).
120. , Eigenvalues of the basic Dirac operator on quaternion-Kähler foliations,
Ann. Glob. Anal. Geom. 30 (2006), no. 3, 289–298.
121. , Eigenvalues of the transversal Dirac operator on Kähler foliations,
J. Geom. Phys. 56 (2006), no. 2, 260–270.
122. , Energy-Momentum Tensor on Foliations, J. Geom. Phys. 57 (2007),
no. 11, 2234–2248.
123. G. Habib, K. Richardson, A brief note on the spectrum of the basic Dirac oper-
ator, arXiv:0809.2406 (2008).
150 Bibliography
124. B. Hanke, T. Schick, Enlargeability and index theory, J. Diff. Geom. 74 (2006),
no. 2, 293–320.
125. M. Herzlich, Théorèmes de masse positive, Sém. Théor. Spec. Géom. Inst.
Fourier Grenoble 16 (1998), 107–126.
126. M. Herzlich, A. Moroianu, Generalized Killing spinors and conformal eigen-
value estimates for Spinc manifolds, Ann. Glob. Anal. Geom. 17 (1999), no. 4,
341–370.
127. O. Hijazi, Opérateurs de Dirac sur les variétés riemanniennes: minoration des
valeurs propres, thèse de doctorat, Ecole Polytechnique - Paris VI (1984).
128. , A conformal lower bound for the smallest eigenvalue of the Dirac op-
erator and Killing spinors, Commun. Math. Phys. 104, (1986) 151–162.
129. , Caractérisation de la sphère par les premières valeurs propres de
l’opérateur de Dirac en dimensions 3, 4, 7 et 8, C. R. Acad. Sci. Paris Sér.
I Math. 303 (1986), no. 9, 417–419.
130. , Première valeur propre de l’opérateur de Dirac et nombre de Yamabe,
C. R. Acad. Sci. Paris Sér. I Math. 313 (1991), no. 12, 865–868.
131. , Eigenvalues of the Dirac operator on compact Kähler manifolds, Comm.
Math. Phys. 160 (1994), no. 3, 563–579.
132. , Lower bounds for the eigenvalues of the Dirac operator, J. Geom. Phys.
16 (1995), 27–38.
133. , Twistor operators and eigenvalues of the Dirac operator, Quaternionic
structures in mathematics and physics (Trieste, 1994), 151–174, Int. Sch. Adv.
Stud. (SISSA), Trieste, 1998.
134. , Spectral properties of the Dirac operator and geometrical structures,
Proceedings of the Summer School on Geometric Methods in Quantum Field
Theory, Villa de Leyva, Colombia, July 12–30, (1999), World Scientific 2001.
135. O. Hijazi, J.-L. Milhorat, Minoration des valeurs propres de l’opérateur de Dirac
sur les variétés spin Kähler-quaternioniennes, J. Math. Pures Appl. (9) 74
(1995), no. 5, 387–414.
136. , Décomposition du fibré des spineurs d’une variété spin Kähler-
quaternionienne sous l’action de la 4-forme fondamentale, J. Geom. Phys. 15
(1995), no. 4, 320–332.
137. , Twistor operators and eigenvalues of the Dirac operator on compact
quaternion-Kähler spin manifolds, Ann. Glob. Anal. Geom. 15 (1997), no. 2,
117–131.
138. O. Hijazi, S. Montiel, Extrinsic Killing spinors, Math. Z. 244 (2003), no. 2,
337–347.
139. O. Hijazi, S. Montiel, A. Roldán, Eigenvalue Boundary Problems for the Dirac
Operator, Comm. Math. Phys. 231 (2002), no. 3, 375–390.
140. , Dirac operators on hypersurfaces of manifolds with negative scalar cur-
vature, Ann. Glob. Anal. Geom. 23 (2003), 247–264.
141. O. Hijazi, S. Montiel, F. Urbano, Spinc geometry of Kähler manifolds and the
Hodge Laplacian on minimal Lagrangian submanifolds, Math. Z. 253 (2006),
no. 4, 821–853.
142. O. Hijazi, S. Montiel, X. Zhang, Dirac Operator on Embedded Hypersurfaces,
Math. Res. Letters 8 (2001), 195–208.
143. , Eigenvalues of the Dirac Operator on Manifolds with Boundary, Com-
mun. Math. Phys. 221 (2001), 255–265.
144. , Conformal Lower Bounds for the Dirac Operator of Embedded Hyper-
surfaces, Asian J. Math. 6 (2002), no. 1, 23–36.
145. O. Hijazi, S. Raulot, Branson’s Q-curvature in Riemannian and Spin Geometry,
SIGMA 3 (2007), 119, 14 p.
146. O. Hijazi, X. Zhang, Lower bounds for the eigenvalues of the Dirac operator.
I. The hypersurface Dirac operator, Ann. Glob. Anal. Geom. 19 (2001), no. 4,
355–376.
Bibliography 151
147. , Lower bounds for the eigenvalues of the Dirac operator. II. The sub-
manifold Dirac operator, Ann. Glob. Anal. Geom. 20 (2001), no. 2, 163–181.
148. N. Hitchin, Harmonic spinors, Adv. Math. 14 (1974), 1–55.
149. P. Jammes, Extrema de valeurs propres dans une classe conforme, Sém. Théor.
Spec. Géom. Inst. Fourier Grenoble 24 (2007), 23–42.
150. M. Jardim, R.F. Leão, Survey on eigenvalues of the Dirac operator and geomet-
ric structures, Int. Math. Forum 3 (2008), no. 1-4, 49–67.
151. S.D. Jung, The first eigenvalue of the transversal Dirac operator, J. Geom. Phys.
39 (2001), no. 3, 253–264.
152. S.D. Jung, T.H. Kang, Lower bounds for the eigenvalue of the transversal Dirac
operator on a Kähler foliation, J. Geom. Phys. 45 (2003), no. 1-2, 75–90.
153. S. Kawai, On the point spectrum of the Dirac operator on a non-compact man-
ifold, J. Geom. Phys. 56 (2006), no. 9, 1782–1789.
154. E.C. Kim, Lower bounds of the Dirac eigenvalues on compact Riemannian spin
manifolds with locally product structure, arXiv:math.DG/0402427 (2004).
155.
, The A-genus and symmetry of the Dirac spectrum on Riemannian prod-
uct manifolds, Diff. Geom. Appl. 25 (2007), no. 3, 309–321.
156. K.-D. Kirchberg, An estimation for the first eigenvalue of the Dirac operator on
closed Kähler manifolds of positive scalar curvature, Ann. Glob. Anal. Geom. 4
(1986), no. 3, 291–325.
157. , Compact six-dimensional Kähler Spin manifolds of positive scalar cur-
vature with the smallest possible first eigenvalue of the Dirac operator, Math.
Ann. 282 (1988), no. 1, 157–176.
158. , The first eigenvalue of the Dirac operator on Kähler manifolds,
J. Geom. Phys. 7 (1990), no. 4, 449–468.
159. , Properties of Kählerian twistor-spinors and vanishing theorems, Math.
Ann. 293 (1992), no. 2, 349–369.
160. , A relation between the Ricci tensor and the spectrum of the Dirac op-
erator, preprint SFB 288 535 (2002), available at http://www-sfb288.math.
tu-berlin.de/Publications/Preprints.html
161. , Curvature dependent lower bounds for the first eigenvalue of the Dirac
operator, J. Geom. Phys. 50 (2004), no. 1-4, 205–222.
162. D. Koh, The η-invariant of the Dirac operator on the Berger spheres, Diplomar-
beit, Universität Hamburg, 2004.
163. W. Kramer, U. Semmelmann, G. Weingart, The first eigenvalue of the Dirac
operator on quaternionic Kähler manifolds, Comm. Math. Phys. 199 (1998),
no. 2, 327–349.
164. , Eigenvalue estimates for the Dirac operator on quaternionic Kähler
manifolds, Math. Z. 230 (1999), no. 4, 727–751.
165. M. Kraus, Lower bounds for eigenvalues of the Dirac operator on surfaces of
rotation, J. Geom. Phys. 31 (1999), no. 2-3, 209–216.
166. , Lower bounds for eigenvalues of the Dirac operator on n-spheres with
SO(n)-symmetry, J. Geom. Phys. 32 (2000), no. 4, 341–348.
167. , Eigenvalues of the Dirac operator on fibrations over S1 , Ann. Glob.
Anal. Geom. 19 (2001), no. 3, 235–257.
168. , Asymptotic estimates of Dirac and Laplace eigenvalues on warped prod-
ucts over S1 , manuscr. math. 112 (2003), no. 3, 357–373.
169. M. Kraus, C. Tretter, A new method for eigenvalue estimates for Dirac operators
on certain manifolds with S k -symmetry, Diff. Geom. Appl. 19 (2003), no. 1,
1–14.
170. W. Kühnel, H.-B. Rademacher, Conformal completion of U(n)-invariant Ricci-
flat Kähler metrics at infinity, Zeitschr. Anal. Anwendungen 16 (1997), no. 1,
113–117.
171. , Asymptotically Euclidean manifolds and twistor spinors, Comm. Math.
Phys. 196 (1998), no. 1, 67–76.
152 Bibliography
199. , Sur les valeurs propres de l’opérateur de Dirac d’une variété spinorielle
simplement connexe admettant une 3-structure de Sasaki, Stud. Cerc. Mat. 48
(1996), no. 1-2, 85–88.
200. , Kähler manifolds with small eigenvalues of the Dirac operator and a
conjecture of Lichnerowicz, Ann. Inst. Fourier 49 (1999), no. 5, 1637–1659.
201. A. Moroianu, S. Moroianu, The Dirac spectrum on manifolds with gradient con-
formal vector fields, J. Funct. An. 253 (2007), 207–219.
202. A. Moroianu, L. Ornea, Eigenvalue estimates for the Dirac operator and har-
monic 1-forms of constant length, C. R. Math. Acad. Sci. Paris 338 (2004),
561–564.
203. A. Moroianu, U. Semmelmann, Parallel spinors and holonomy groups, J. Math.
Phys. 41 (2000), no. 4, 2395–2402.
204. S. Moroianu, Weyl laws on open manifolds, Math. Ann. 340 (2008), no. 1, 1–21.
205. R. Petit, Spinc -structures and Dirac operators on contact manifolds, Diff. Geom.
Appl. 22 (2005), no. 2, 229–252.
206. F. Pfäffle, The Dirac spectrum of Bieberbach manifolds, J. Geom. Phys. 35
(2000), no. 4, 367–385.
207. , Eigenwertkonvergenz für Dirac-Operatoren, Doktorarbeit, Universität
Hamburg, Shaker-Verlag, Aachen 2003.
208. , Eigenvalues of Dirac operators for hyperbolic degenerations, manuscr.
math. 116 (2005), no. 1, 1–29.
209. H.-B. Rademacher, Generalized Killing spinors with imaginary Killing function
and conformal Killing fields, Global differential geometry and global analysis
(Berlin, 1990), 192–198, Lecture Notes in Math. 1481, Springer, Berlin, 1991.
210. S. Raulot, Optimal eigenvalues estimate for the Dirac operator on domains with
boundary, Lett. Math. Phys. 73 (2005), no. 2, 135–145.
211. , The Hijazi inequality on manifolds with boundary, J. Geom. Phys. 56
(2006), no. 11, 2189–2202.
212. , Aspect conforme de l’opérateur de Dirac sur une variété à bord, thèse
de doctorat, Université Henri Poincaré, Nancy (2006).
213. , Rigidity of compact Riemannian spin manifolds with boundary, Lett.
Math. Phys. 86 (2008), no. 2, 177–192.
214. , On a spin conformal invariant on manifold with boundary, Math. Z.
261 (2009), no. 2, 321–349.
215. R. Schoen, S. T. Yau, On the proof of the positive mass conjecture in general
relativity, Comm. Math. Phys. 65 (1979), no. 1, 45–76.
216. , Proof of the positive mass theorem II, Comm. Math. Phys. 79 (1981),
no. 2, 231–260.
217. E. Schrödinger, Diracsches Elektron im Schwerefeld, Sitzungsber. Preuß. Akad.
Wiss., Phys.-Math. Kl. (1932), no. 11-12, 105–128.
218. J. Seade, B. Steer, A note on the eta function for quotients of PSL2 (R) by
co-compact Fuchsian groups, Topology 26 (1987), no. 1, 79–91.
219. L. Seeger, The Dirac operator on oriented Grassmann manifolds and G2 /SO(4),
Preprint, Math. Inst. Bonn, 1997.
220. , The spectrum of the Dirac operator on G2 /SO(4), Ann. Glob. Anal.
Geom. 17 (1999), no. 4, 385–396.
221. , Metriken mit harmonischen Spinoren auf geradedimensionalen
Sphären, Shaker Verlag, Aachen, 2001.
222. S. Seifarth, U. Semmelmann, The spectrum of the Dirac operator on the complex
projective space P 2m−1 (C), Preprint SFB 288, 1993.
223. U. Semmelmann, A short proof of eigenvalue estimates for the Dirac operator on
Riemannian and Kähler manifolds, Differential Geometry and its applications,
proceedings, Brno (1998), 137–140.
224. H. Strese, Über den Dirac-Operator auf Grassmann-Mannigfaltigkeiten, Math.
Nachr. 98 (1980), 53–59.
154 Bibliography
155
156 Index
Vol. 1791: M. Knebusch, D. Zhang, Manis Valuations Vol. 1813: L. Ambrosio, L. A. Caffarelli, Y. Brenier,
and Prüfer Extensions I: A new Chapter in Commutative G. Buttazzo, C. Villani, Optimal Transportation and its
Algebra (2002) Applications. Martina Franca, Italy 2001. Editors: L. A.
Vol. 1792: D. D. Ang, R. Gorenflo, V. K. Le, D. D. Trong, Caffarelli, S. Salsa (2003)
Moment Theory and Some Inverse Problems in Potential Vol. 1814: P. Bank, F. Baudoin, H. Föllmer, L.C.G.
Theory and Heat Conduction (2002) Rogers, M. Soner, N. Touzi, Paris-Princeton Lectures on
Vol. 1793: J. Cortés Monforte, Geometric, Control and Mathematical Finance 2002 (2003)
Numerical Aspects of Nonholonomic Systems (2002) Vol. 1815: A. M. Vershik (Ed.), Asymptotic Combi-
Vol. 1794: N. Pytheas Fogg, Substitution in Dynamics, natorics with Applications to Mathematical Physics.
Arithmetics and Combinatorics. Editors: V. Berthé, S. St. Petersburg, Russia 2001 (2003)
Ferenczi, C. Mauduit, A. Siegel (2002) Vol. 1816: S. Albeverio, W. Schachermayer, M. Tala-
Vol. 1795: H. Li, Filtered-Graded Transfer in Using Non- grand, Lectures on Probability Theory and Statistics.
commutative Gröbner Bases (2002) Ecole d’Eté de Probabilités de Saint-Flour XXX-2000.
Vol. 1796: J.M. Melenk, hp-Finite Element Methods for Editor: P. Bernard (2003)
Singular Perturbations (2002) Vol. 1817: E. Koelink, W. Van Assche (Eds.), Orthogonal
Vol. 1797: B. Schmidt, Characters and Cyclotomic Fields Polynomials and Special Functions. Leuven 2002 (2003)
in Finite Geometry (2002) Vol. 1818: M. Bildhauer, Convex Variational Problems
Vol. 1798: W.M. Oliva, Geometric Mechanics (2002) with Linear, nearly Linear and/or Anisotropic Growth
Vol. 1799: H. Pajot, Analytic Capacity, Rectifiability, Conditions (2003)
Menger Curvature and the Cauchy Integral (2002) Vol. 1819: D. Masser, Yu. V. Nesterenko, H. P. Schlick-
Vol. 1800: O. Gabber, L. Ramero, Almost Ring Theory ewei, W. M. Schmidt, M. Waldschmidt, Diophantine
(2003) Approximation. Cetraro, Italy 2000. Editors: F. Amoroso,
Vol. 1801: J. Azéma, M. Émery, M. Ledoux, M. Yor U. Zannier (2003)
(Eds.), Séminaire de Probabilités XXXVI (2003) Vol. 1820: F. Hiai, H. Kosaki, Means of Hilbert Space
Vol. 1802: V. Capasso, E. Merzbach, B. G. Ivanoff, Operators (2003)
M. Dozzi, R. Dalang, T. Mountford, Topics in Spatial Vol. 1821: S. Teufel, Adiabatic Perturbation Theory in
Stochastic Processes. Martina Franca, Italy 2001. Editor: Quantum Dynamics (2003)
E. Merzbach (2003)
Vol. 1822: S.-N. Chow, R. Conti, R. Johnson, J. Mallet-
Vol. 1803: G. Dolzmann, Variational Methods for Crys- Paret, R. Nussbaum, Dynamical Systems. Cetraro, Italy
talline Microstructure – Analysis and Computation 2000. Editors: J. W. Macki, P. Zecca (2003)
(2003)
Vol. 1823: A. M. Anile, W. Allegretto, C. Ringhofer,
Vol. 1804: I. Cherednik, Ya. Markov, R. Howe, G.
Mathematical Problems in Semiconductor Physics.
Lusztig, Iwahori-Hecke Algebras and their Representa-
Cetraro, Italy 1998. Editor: A. M. Anile (2003)
tion Theory. Martina Franca, Italy 1999. Editors: V. Bal-
doni, D. Barbasch (2003) Vol. 1824: J. A. Navarro González, J. B. Sancho de Salas,
C ∞ – Differentiable Spaces (2003)
Vol. 1805: F. Cao, Geometric Curve Evolution and Image
Processing (2003) Vol. 1825: J. H. Bramble, A. Cohen, W. Dahmen, Mul-
Vol. 1806: H. Broer, I. Hoveijn. G. Lunther, G. Vegter, tiscale Problems and Methods in Numerical Simulations,
Bifurcations in Hamiltonian Systems. Computing Singu- Martina Franca, Italy 2001. Editor: C. Canuto (2003)
larities by Gröbner Bases (2003) Vol. 1826: K. Dohmen, Improved Bonferroni Inequal-
Vol. 1807: V. D. Milman, G. Schechtman (Eds.), Geomet- ities via Abstract Tubes. Inequalities and Identities of
ric Aspects of Functional Analysis. Israel Seminar 2000- Inclusion-Exclusion Type. VIII, 113 p, 2003.
2002 (2003) Vol. 1827: K. M. Pilgrim, Combinations of Complex
Vol. 1808: W. Schindler, Measures with Symmetry Prop- Dynamical Systems. IX, 118 p, 2003.
erties (2003) Vol. 1828: D. J. Green, Gröbner Bases and the Computa-
Vol. 1809: O. Steinbach, Stability Estimates for Hybrid tion of Group Cohomology. XII, 138 p, 2003.
Coupled Domain Decomposition Methods (2003) Vol. 1829: E. Altman, B. Gaujal, A. Hordijk, Discrete-
Vol. 1810: J. Wengenroth, Derived Functors in Functional Event Control of Stochastic Networks: Multimodularity
Analysis (2003) and Regularity. XIV, 313 p, 2003.
Vol. 1811: J. Stevens, Deformations of Singularities Vol. 1830: M. I. Gil’, Operator Functions and Localiza-
(2003) tion of Spectra. XIV, 256 p, 2003.
Vol. 1812: L. Ambrosio, K. Deckelnick, G. Dziuk, Vol. 1831: A. Connes, J. Cuntz, E. Guentner, N. Hig-
M. Mimura, V. A. Solonnikov, H. M. Soner, Mathemat- son, J. E. Kaminker, Noncommutative Geometry, Mar-
ical Aspects of Evolving Interfaces. Madeira, Funchal, tina Franca, Italy 2002. Editors: S. Doplicher, L. Longo
Portugal 2000. Editors: P. Colli, J. F. Rodrigues (2003) (2004)
Vol. 1832: J. Azéma, M. Émery, M. Ledoux, M. Yor Vol. 1858: A.S. Cherny, H.-J. Engelbert, Singular
(Eds.), Séminaire de Probabilités XXXVII (2003) Stochastic Differential Equations (2005)
Vol. 1833: D.-Q. Jiang, M. Qian, M.-P. Qian, Mathemati- Vol. 1859: E. Letellier, Fourier Transforms of Invariant
cal Theory of Nonequilibrium Steady States. On the Fron- Functions on Finite Reductive Lie Algebras (2005)
tier of Probability and Dynamical Systems. IX, 280 p, Vol. 1860: A. Borisyuk, G.B. Ermentrout, A. Friedman,
2004. D. Terman, Tutorials in Mathematical Biosciences I.
Vol. 1834: Yo. Yomdin, G. Comte, Tame Geometry with Mathematical Neurosciences (2005)
Application in Smooth Analysis. VIII, 186 p, 2004. Vol. 1861: G. Benettin, J. Henrard, S. Kuksin, Hamilto-
Vol. 1835: O.T. Izhboldin, B. Kahn, N.A. Karpenko, nian Dynamics – Theory and Applications, Cetraro, Italy,
A. Vishik, Geometric Methods in the Algebraic Theory 1999. Editor: A. Giorgilli (2005)
of Quadratic Forms. Summer School, Lens, 2000. Editor: Vol. 1862: B. Helffer, F. Nier, Hypoelliptic Estimates and
J.-P. Tignol (2004) Spectral Theory for Fokker-Planck Operators and Witten
Vol. 1836: C. Nǎstǎsescu, F. Van Oystaeyen, Methods of Laplacians (2005)
Graded Rings. XIII, 304 p, 2004. Vol. 1863: H. Führ, Abstract Harmonic Analysis of Con-
Vol. 1837: S. Tavaré, O. Zeitouni, Lectures on Probabil- tinuous Wavelet Transforms (2005)
ity Theory and Statistics. Ecole d’Eté de Probabilités de
Vol. 1864: K. Efstathiou, Metamorphoses of Hamiltonian
Saint-Flour XXXI-2001. Editor: J. Picard (2004)
Systems with Symmetries (2005)
Vol. 1838: A.J. Ganesh, N.W. O’Connell, D.J. Wischik,
Big Queues. XII, 254 p, 2004. Vol. 1865: D. Applebaum, B.V. R. Bhat, J. Kustermans,
Vol. 1839: R. Gohm, Noncommutative Stationary Pro- J. M. Lindsay, Quantum Independent Increment Pro-
cesses. VIII, 170 p, 2004. cesses I. From Classical Probability to Quantum Stochas-
Vol. 1840: B. Tsirelson, W. Werner, Lectures on Probabil- tic Calculus. Editors: M. Schürmann, U. Franz (2005)
ity Theory and Statistics. Ecole d’Eté de Probabilités de Vol. 1866: O.E. Barndorff-Nielsen, U. Franz, R. Gohm,
Saint-Flour XXXII-2002. Editor: J. Picard (2004) B. Kümmerer, S. Thorbjønsen, Quantum Independent
Vol. 1841: W. Reichel, Uniqueness Theorems for Vari- Increment Processes II. Structure of Quantum Lévy
ational Problems by the Method of Transformation Processes, Classical Probability, and Physics. Editors: M.
Groups (2004) Schürmann, U. Franz, (2005)
Vol. 1842: T. Johnsen, A. L. Knutsen, K3 Projective Mod- Vol. 1867: J. Sneyd (Ed.), Tutorials in Mathematical Bio-
els in Scrolls (2004) sciences II. Mathematical Modeling of Calcium Dynam-
Vol. 1843: B. Jefferies, Spectral Properties of Noncom- ics and Signal Transduction. (2005)
muting Operators (2004) Vol. 1868: J. Jorgenson, S. Lang, Posn (R) and Eisenstein
Vol. 1844: K.F. Siburg, The Principle of Least Action in Series. (2005)
Geometry and Dynamics (2004) Vol. 1869: A. Dembo, T. Funaki, Lectures on Probabil-
Vol. 1845: Min Ho Lee, Mixed Automorphic Forms, ity Theory and Statistics. Ecole d’Eté de Probabilités de
Torus Bundles, and Jacobi Forms (2004) Saint-Flour XXXIII-2003. Editor: J. Picard (2005)
Vol. 1846: H. Ammari, H. Kang, Reconstruction of Small Vol. 1870: V.I. Gurariy, W. Lusky, Geometry of Müntz
Inhomogeneities from Boundary Measurements (2004) Spaces and Related Questions. (2005)
Vol. 1847: T.R. Bielecki, T. Björk, M. Jeanblanc, M. Vol. 1871: P. Constantin, G. Gallavotti, A.V. Kazhikhov,
Rutkowski, J.A. Scheinkman, W. Xiong, Paris-Princeton Y. Meyer, S. Ukai, Mathematical Foundation of Turbu-
Lectures on Mathematical Finance 2003 (2004) lent Viscous Flows, Martina Franca, Italy, 2003. Editors:
Vol. 1848: M. Abate, J. E. Fornaess, X. Huang, J. P. M. Cannone, T. Miyakawa (2006)
Rosay, A. Tumanov, Real Methods in Complex and CR Vol. 1872: A. Friedman (Ed.), Tutorials in Mathemati-
Geometry, Martina Franca, Italy 2002. Editors: D. Zait- cal Biosciences III. Cell Cycle, Proliferation, and Cancer
sev, G. Zampieri (2004) (2006)
Vol. 1849: Martin L. Brown, Heegner Modules and Ellip- Vol. 1873: R. Mansuy, M. Yor, Random Times and En-
tic Curves (2004) largements of Filtrations in a Brownian Setting (2006)
Vol. 1850: V. D. Milman, G. Schechtman (Eds.), Geomet- Vol. 1874: M. Yor, M. Émery (Eds.), In Memoriam Paul-
ric Aspects of Functional Analysis. Israel Seminar 2002- André Meyer - Séminaire de Probabilités XXXIX (2006)
2003 (2004) Vol. 1875: J. Pitman, Combinatorial Stochastic Processes.
Vol. 1851: O. Catoni, Statistical Learning Theory and Ecole d’Eté de Probabilités de Saint-Flour XXXII-2002.
Stochastic Optimization (2004) Editor: J. Picard (2006)
Vol. 1852: A.S. Kechris, B.D. Miller, Topics in Orbit
Vol. 1876: H. Herrlich, Axiom of Choice (2006)
Equivalence (2004)
Vol. 1853: Ch. Favre, M. Jonsson, The Valuative Tree Vol. 1877: J. Steuding, Value Distributions of L-Functions
(2004) (2007)
Vol. 1854: O. Saeki, Topology of Singular Fibers of Dif- Vol. 1878: R. Cerf, The Wulff Crystal in Ising and Percol-
ferential Maps (2004) ation Models, Ecole d’Eté de Probabilités de Saint-Flour
Vol. 1855: G. Da Prato, P.C. Kunstmann, I. Lasiecka, XXXIV-2004. Editor: Jean Picard (2006)
A. Lunardi, R. Schnaubelt, L. Weis, Functional Analytic Vol. 1879: G. Slade, The Lace Expansion and its Applica-
Methods for Evolution Equations. Editors: M. Iannelli, tions, Ecole d’Eté de Probabilités de Saint-Flour XXXIV-
R. Nagel, S. Piazzera (2004) 2004. Editor: Jean Picard (2006)
Vol. 1856: K. Back, T.R. Bielecki, C. Hipp, S. Peng, Vol. 1880: S. Attal, A. Joye, C.-A. Pillet, Open Quantum
W. Schachermayer, Stochastic Methods in Finance, Bres- Systems I, The Hamiltonian Approach (2006)
sanone/Brixen, Italy, 2003. Editors: M. Fritelli, W. Rung- Vol. 1881: S. Attal, A. Joye, C.-A. Pillet, Open Quantum
galdier (2004) Systems II, The Markovian Approach (2006)
Vol. 1857: M. Émery, M. Ledoux, M. Yor (Eds.), Sémi- Vol. 1882: S. Attal, A. Joye, C.-A. Pillet, Open Quantum
naire de Probabilités XXXVIII (2005) Systems III, Recent Developments (2006)
Vol. 1883: W. Van Assche, F. Marcellàn (Eds.), Orthogo- Vol. 1910: V.D. Milman, G. Schechtman (Eds.), Geo-
nal Polynomials and Special Functions, Computation and metric Aspects of Functional Analysis. Israel Seminar
Application (2006) 2004-2005 (2007)
Vol. 1884: N. Hayashi, E.I. Kaikina, P.I. Naumkin, Vol. 1911: A. Bressan, D. Serre, M. Williams,
I.A. Shishmarev, Asymptotics for Dissipative Nonlinear K. Zumbrun, Hyperbolic Systems of Balance Laws.
Equations (2006) Cetraro, Italy 2003. Editor: P. Marcati (2007)
Vol. 1885: A. Telcs, The Art of Random Walks (2006) Vol. 1912: V. Berinde, Iterative Approximation of Fixed
Vol. 1886: S. Takamura, Splitting Deformations of Dege- Points (2007)
nerations of Complex Curves (2006) Vol. 1913: J.E. Marsden, G. Misiołek, J.-P. Ortega,
Vol. 1887: K. Habermann, L. Habermann, Introduction to M. Perlmutter, T.S. Ratiu, Hamiltonian Reduction by
Symplectic Dirac Operators (2006) Stages (2007)
Vol. 1888: J. van der Hoeven, Transseries and Real Dif- Vol. 1914: G. Kutyniok, Affine Density in Wavelet
ferential Algebra (2006) Analysis (2007)
Vol. 1889: G. Osipenko, Dynamical Systems, Graphs, and Vol. 1915: T. Bıyıkoǧlu, J. Leydold, P.F. Stadler,
Algorithms (2006) Laplacian Eigenvectors of Graphs. Perron-Frobenius and
Vol. 1890: M. Bunge, J. Funk, Singular Coverings of Faber-Krahn Type Theorems (2007)
Toposes (2006) Vol. 1916: C. Villani, F. Rezakhanlou, Entropy Methods
Vol. 1891: J.B. Friedlander, D.R. Heath-Brown, for the Boltzmann Equation. Editors: F. Golse, S. Olla
H. Iwaniec, J. Kaczorowski, Analytic Number Theory, (2008)
Cetraro, Italy, 2002. Editors: A. Perelli, C. Viola (2006) Vol. 1917: I. Veselić, Existence and Regularity Prop-
Vol. 1892: A. Baddeley, I. Bárány, R. Schneider, W. Weil, erties of the Integrated Density of States of Random
Stochastic Geometry, Martina Franca, Italy, 2004. Editor: Schrödinger (2008)
W. Weil (2007) Vol. 1918: B. Roberts, R. Schmidt, Local Newforms for
Vol. 1893: H. Hanßmann, Local and Semi-Local Bifur- GSp(4) (2007)
cations in Hamiltonian Dynamical Systems, Results and Vol. 1919: R.A. Carmona, I. Ekeland, A. Kohatsu-
Examples (2007) Higa, J.-M. Lasry, P.-L. Lions, H. Pham, E. Taflin,
Vol. 1894: C.W. Groetsch, Stable Approximate Evalua- Paris-Princeton Lectures on Mathematical Finance 2004.
tion of Unbounded Operators (2007) Editors: R.A. Carmona, E. Çinlar, I. Ekeland, E. Jouini,
Vol. 1895: L. Molnár, Selected Preserver Problems on J.A. Scheinkman, N. Touzi (2007)
Algebraic Structures of Linear Operators and on Function Vol. 1920: S.N. Evans, Probability and Real Trees. Ecole
Spaces (2007) d’Été de Probabilités de Saint-Flour XXXV-2005 (2008)
Vol. 1896: P. Massart, Concentration Inequalities and Vol. 1921: J.P. Tian, Evolution Algebras and their Appli-
Model Selection, Ecole d’Été de Probabilités de Saint- cations (2008)
Flour XXXIII-2003. Editor: J. Picard (2007) Vol. 1922: A. Friedman (Ed.), Tutorials in Mathematical
Vol. 1897: R. Doney, Fluctuation Theory for Lévy BioSciences IV. Evolution and Ecology (2008)
Processes, Ecole d’Été de Probabilités de Saint-Flour Vol. 1923: J.P.N. Bishwal, Parameter Estimation in
XXXV-2005. Editor: J. Picard (2007) Stochastic Differential Equations (2008)
Vol. 1898: H.R. Beyer, Beyond Partial Differential Equa- Vol. 1924: M. Wilson, Littlewood-Paley Theory and
tions, On linear and Quasi-Linear Abstract Hyperbolic Exponential-Square Integrability (2008)
Evolution Equations (2007) Vol. 1925: M. du Sautoy, L. Woodward, Zeta Functions
Vol. 1899: Séminaire de Probabilités XL. Editors: of Groups and Rings (2008)
C. Donati-Martin, M. Émery, A. Rouault, C. Stricker Vol. 1926: L. Barreira, V. Claudia, Stability of Nonauto-
(2007) nomous Differential Equations (2008)
Vol. 1900: E. Bolthausen, A. Bovier (Eds.), Spin Glasses Vol. 1927: L. Ambrosio, L. Caffarelli, M.G. Crandall,
(2007) L.C. Evans, N. Fusco, Calculus of Variations and Non-
Vol. 1901: O. Wittenberg, Intersections de deux Linear Partial Differential Equations. Cetraro, Italy 2005.
quadriques et pinceaux de courbes de genre 1, Intersec- Editors: B. Dacorogna, P. Marcellini (2008)
tions of Two Quadrics and Pencils of Curves of Genus 1 Vol. 1928: J. Jonsson, Simplicial Complexes of Graphs
(2007) (2008)
Vol. 1902: A. Isaev, Lectures on the Automorphism Vol. 1929: Y. Mishura, Stochastic Calculus for Fractional
Groups of Kobayashi-Hyperbolic Manifolds (2007) Brownian Motion and Related Processes (2008)
Vol. 1903: G. Kresin, V. Maz’ya, Sharp Real-Part Theo- Vol. 1930: J.M. Urbano, The Method of Intrinsic Scaling.
rems (2007) A Systematic Approach to Regularity for Degenerate and
Vol. 1904: P. Giesl, Construction of Global Lyapunov Singular PDEs (2008)
Functions Using Radial Basis Functions (2007) Vol. 1931: M. Cowling, E. Frenkel, M. Kashiwara,
Vol. 1905: C. Prévôt, M. Röckner, A Concise Course on A. Valette, D.A. Vogan, Jr., N.R. Wallach, Representation
Stochastic Partial Differential Equations (2007) Theory and Complex Analysis. Venice, Italy 2004.
Vol. 1906: T. Schuster, The Method of Approximate Editors: E.C. Tarabusi, A. D’Agnolo, M. Picardello
Inverse: Theory and Applications (2007) (2008)
Vol. 1907: M. Rasmussen, Attractivity and Bifurcation Vol. 1932: A.A. Agrachev, A.S. Morse, E.D. Sontag,
for Nonautonomous Dynamical Systems (2007) H.J. Sussmann, V.I. Utkin, Nonlinear and Optimal
Vol. 1908: T.J. Lyons, M. Caruana, T. Lévy, Differential Control Theory. Cetraro, Italy 2004. Editors: P. Nistri,
Equations Driven by Rough Paths, Ecole d’Été de Proba- G. Stefani (2008)
bilités de Saint-Flour XXXIV-2004 (2007) Vol. 1933: M. Petkovic, Point Estimation of Root Finding
Vol. 1909: H. Akiyoshi, M. Sakuma, M. Wada, Methods (2008)
Y. Yamashita, Punctured Torus Groups and 2-Bridge Vol. 1934: C. Donati-Martin, M. Émery, A. Rouault,
Knot Groups (I) (2007) C. Stricker (Eds.), Séminaire de Probabilités XLI (2008)
Vol. 1935: A. Unterberger, Alternative Pseudodifferential Vol. 1958: M.C. Olsson, Compactifying Moduli Spaces
Analysis (2008) for Abelian Varieties (2008)
Vol. 1936: P. Magal, S. Ruan (Eds.), Structured Popula- Vol. 1959: Y. Nakkajima, A. Shiho, Weight Filtrations
tion Models in Biology and Epidemiology (2008) on Log Crystalline Cohomologies of Families of Open
Vol. 1937: G. Capriz, P. Giovine, P.M. Mariano (Eds.), Smooth Varieties (2008)
Mathematical Models of Granular Matter (2008) Vol. 1960: J. Lipman, M. Hashimoto, Foundations of
Vol. 1938: D. Auroux, F. Catanese, M. Manetti, P. Seidel, Grothendieck Duality for Diagrams of Schemes (2009)
B. Siebert, I. Smith, G. Tian, Symplectic 4-Manifolds Vol. 1961: G. Buttazzo, A. Pratelli, S. Solimini,
and Algebraic Surfaces. Cetraro, Italy 2003. Editors: E. Stepanov, Optimal Urban Networks via Mass Trans-
F. Catanese, G. Tian (2008) portation (2009)
Vol. 1939: D. Boffi, F. Brezzi, L. Demkowicz, R.G. Vol. 1962: R. Dalang, D. Khoshnevisan, C. Mueller,
Durán, R.S. Falk, M. Fortin, Mixed Finite Elements, D. Nualart, Y. Xiao, A Minicourse on Stochastic Partial
Compatibility Conditions, and Applications. Cetraro, Differential Equations (2009)
Italy 2006. Editors: D. Boffi, L. Gastaldi (2008) Vol. 1963: W. Siegert, Local Lyapunov Exponents (2009)
Vol. 1940: J. Banasiak, V. Capasso, M.A.J. Chap- Vol. 1964: W. Roth, Operator-valued Measures and Inte-
lain, M. Lachowicz, J. Miȩkisz, Multiscale Problems in grals for Cone-valued Functions and Integrals for Cone-
the Life Sciences. From Microscopic to Macroscopic. valued Functions (2009)
Bȩdlewo, Poland 2006. Editors: V. Capasso, M. Lachow- Vol. 1965: C. Chidume, Geometric Properties of Banach
icz (2008) Spaces and Nonlinear Iterations (2009)
Vol. 1941: S.M.J. Haran, Arithmetical Investigations. Vol. 1966: D. Deng, Y. Han, Harmonic Analysis on
Representation Theory, Orthogonal Polynomials, and Spaces of Homogeneous Type (2009)
Quantum Interpolations (2008) Vol. 1967: B. Fresse, Modules over Operads and Functors
Vol. 1942: S. Albeverio, F. Flandoli, Y.G. Sinai, SPDE in (2009)
Hydrodynamic. Recent Progress and Prospects. Cetraro, Vol. 1968: R. Weissauer, Endoscopy for GSP(4) and the
Italy 2005. Editors: G. Da Prato, M. Röckner (2008) Cohomology of Siegel Modular Threefolds (2009)
Vol. 1969: B. Roynette, M. Yor, Penalising Brownian
Vol. 1943: L.L. Bonilla (Ed.), Inverse Problems and Imag-
Paths (2009)
ing. Martina Franca, Italy 2002 (2008)
Vol. 1970: M. Biskup, A. Bovier, F. den Hollander, D.
Vol. 1944: A. Di Bartolo, G. Falcone, P. Plaumann,
Ioffe, F. Martinelli, K. Netočný, F. Toninelli, Methods of
K. Strambach, Algebraic Groups and Lie Groups with
Contemporary Mathematical Statistical Physics. Editor:
Few Factors (2008)
R. Kotecký (2009)
Vol. 1945: F. Brauer, P. van den Driessche, J. Wu (Eds.), Vol. 1971: L. Saint-Raymond, Hydrodynamic Limits of
Mathematical Epidemiology (2008) the Boltzmann Equation (2009)
Vol. 1946: G. Allaire, A. Arnold, P. Degond, T.Y. Hou, Vol. 1972: T. Mochizuki, Donaldson Type Invariants for
Quantum Transport. Modelling, Analysis and Asymp- Algebraic Surfaces (2009)
totics. Cetraro, Italy 2006. Editors: N.B. Abdallah, Vol. 1973: M.A. Berger, L.H. Kauffmann, B. Khesin,
G. Frosali (2008) H.K. Moffatt, R.L. Ricca, De W. Sumners, Lectures on
Vol. 1947: D. Abramovich, M. Mariño, M. Thaddeus, Topological Fluid Mechanics. Editor: R.L. Ricca (2009)
R. Vakil, Enumerative Invariants in Algebraic Geo- Vol. 1974: F. den Hollander, Random Polymers: École
metry and String Theory. Cetraro, Italy 2005. Editors: d’Été de Probabilités de Saint-Flour XXXVII – 2007
K. Behrend, M. Manetti (2008) (2009)
Vol. 1948: F. Cao, J-L. Lisani, J-M. Morel, P. Musé, Vol. 1975: J.C. Rohde, Cyclic Coverings, Calabi-Yau
F. Sur, A Theory of Shape Identification (2008) Manifolds and Complex Multiplication (2009)
Vol. 1949: H.G. Feichtinger, B. Helffer, M.P. Lamoureux, Vol. 1976: N. Ginoux, The Dirac Spectrum (2009)
N. Lerner, J. Toft, Pseudo-Differential Operators. Quan-
tization and Signals. Cetraro, Italy 2006. Editors: L.
Rodino, M.W. Wong (2008) Recent Reprints and New Editions
Vol. 1950: M. Bramson, Stability of Queueing Networks, Vol. 1702: J. Ma, J. Yong, Forward-Backward Stochas-
Ecole d’Eté de Probabilités de Saint-Flour XXXVI-2006 tic Differential Equations and their Applications. 1999 –
(2008) Corr. 3rd printing (2007)
Vol. 1951: A. Moltó, J. Orihuela, S. Troyanski, Vol. 830: J.A. Green, Polynomial Representations of
M. Valdivia, A Non Linear Transfer Technique for GLn , with an Appendix on Schensted Correspondence
Renorming (2009) and Littelmann Paths by K. Erdmann, J.A. Green and
Vol. 1952: R. Mikhailov, I.B.S. Passi, Lower Central and M. Schoker 1980 – 2nd corr. and augmented edition
Dimension Series of Groups (2009) (2007)
Vol. 1953: K. Arwini, C.T.J. Dodson, Information Geo- Vol. 1693: S. Simons, From Hahn-Banach to Monotonic-
metry (2008) ity (Minimax and Monotonicity 1998) – 2nd exp. edition
Vol. 1954: P. Biane, L. Bouten, F. Cipriani, N. Konno, (2008)
N. Privault, Q. Xu, Quantum Potential Theory. Editors: Vol. 470: R.E. Bowen, Equilibrium States and the Ergodic
U. Franz, M. Schuermann (2008) Theory of Anosov Diffeomorphisms. With a preface by
Vol. 1955: M. Bernot, V. Caselles, J.-M. Morel, Optimal D. Ruelle. Edited by J.-R. Chazottes. 1975 – 2nd rev.
Transportation Networks (2008) edition (2008)
Vol. 1956: C.H. Chu, Matrix Convolution Operators on Vol. 523: S.A. Albeverio, R.J. Høegh-Krohn, S. Maz-
Groups (2008) zucchi, Mathematical Theory of Feynman Path Integral.
Vol. 1957: A. Guionnet, On Random Matrices: Macro- 1976 – 2nd corr. and enlarged edition (2008)
scopic Asymptotics, Ecole d’Eté de Probabilités de Saint- Vol. 1764: A. Cannas da Silva, Lectures on Symplectic
Flour XXXVI-2006 (2009) Geometry 2001 – Corr. 2nd printing (2008)
LECTURE NOTES IN MATHEMATICS 123
Edited by J.-M. Morel, F. Takens, B. Teissier, P.K. Maini
1. Lecture Notes aim to report new developments in all areas of mathematics and their
applications - quickly, informally and at a high level. Mathematical texts analysing new
developments in modelling and numerical simulation are welcome.
Monograph manuscripts should be reasonably self-contained and rounded off. Thus
they may, and often will, present not only results of the author but also related work
by other people. They may be based on specialised lecture courses. Furthermore, the
manuscripts should provide sufficient motivation, examples and applications. This clearly
distinguishes Lecture Notes from journal articles or technical reports which normally are
very concise. Articles intended for a journal but too long to be accepted by most journals,
usually do not have this “lecture notes” character. For similar reasons it is unusual for
doctoral theses to be accepted for the Lecture Notes series, though habilitation theses may
be appropriate.
2. Manuscripts should be submitted either online at www.editorialmanager.com/lnm to
Springer’s mathematics editorial in Heidelberg, or to one of the series editors. In general,
manuscripts will be sent out to 2 external referees for evaluation. If a decision cannot yet
be reached on the basis of the first 2 reports, further referees may be contacted: The author
will be informed of this. A final decision to publish can be made only on the basis of the
complete manuscript, however a refereeing process leading to a preliminary decision can
be based on a pre-final or incomplete manuscript. The strict minimum amount of material
that will be considered should include a detailed outline describing the planned contents
of each chapter, a bibliography and several sample chapters.
Authors should be aware that incomplete or insufficiently close to final manuscripts
almost always result in longer refereeing times and nevertheless unclear referees’ recom-
mendations, making further refereeing of a final draft necessary.
Authors should also be aware that parallel submission of their manuscript to another
publisher while under consideration for LNM will in general lead to immediate rejection.
3. Manuscripts should in general be submitted in English. Final manuscripts should contain
at least 100 pages of mathematical text and should always include
– a table of contents;
– an informative introduction, with adequate motivation and perhaps some historical re-
marks: it should be accessible to a reader not intimately familiar with the topic treated;
– a subject index: as a rule this is genuinely helpful for the reader.
For evaluation purposes, manuscripts may be submitted in print or electronic form (print
form is still preferred by most referees), in the latter case preferably as pdf- or zipped
ps-files. Lecture Notes volumes are, as a rule, printed digitally from the authors’ files.
To ensure best results, authors are asked to use the LaTeX2e style files available from
Springer’s web-server at:
ftp://ftp.springer.de/pub/tex/latex/svmonot1/ (for monographs) and
ftp://ftp.springer.de/pub/tex/latex/svmultt1/ (for summer schools/tutorials).
Additional technical instructions, if necessary, are available on request from:
lnm@springer.com.
4. Careful preparation of the manuscripts will help keep production time short besides en-
suring satisfactory appearance of the finished book in print and online. After acceptance
of the manuscript authors will be asked to prepare the final LaTeX source files and also
the corresponding dvi-, pdf- or zipped ps-file. The LaTeX source files are essential for
producing the full-text online version of the book (see
http://www.springerlink.com/openurl.asp?genre=journal&issn=0075-8434 for the exist-
ing online volumes of LNM).
The actual production of a Lecture Notes volume takes approximately 12 weeks.
5. Authors receive a total of 50 free copies of their volume, but no royalties. They are entitled
to a discount of 33.3% on the price of Springer books purchased for their personal use, if
ordering directly from Springer.
6. Commitment to publish is made by letter of intent rather than by signing a formal contract.
Springer-Verlag secures the copyright for each volume. Authors are free to reuse material
contained in their LNM volumes in later publications: a brief written (or e-mail) request
for formal permission is sufficient.
Addresses:
Professor J.-M. Morel, CMLA,
École Normale Supérieure de Cachan,
61 Avenue du Président Wilson, 94235 Cachan Cedex, France
E-mail: Jean-Michel.Morel@cmla.ens-cachan.fr
Professor F. Takens, Mathematisch Instituut,
Rijksuniversiteit Groningen, Postbus 800,
9700 AV Groningen, The Netherlands
E-mail: F.Takens@rug.nl
Professor B. Teissier, Institut Mathématique de Jussieu,
UMR 7586 du CNRS, Équipe “Géométrie et Dynamique”,
175 rue du Chevaleret,
75013 Paris, France
E-mail: teissier@math.jussieu.fr