Essential Parabolic Structures and Their Infinitesimal Automorphisms
Using the theory of Weyl structures, we give a natural generalization of the notion of essential conformal structures and conformal Killing fields to arbitrary parabolic geometries. We show that a parabolic structure is inessential whenever the automorphism group acts properly on the base space. As...
Saved in:
Date: | 2011 |
---|---|
Main Author: | |
Format: | Article |
Language: | English |
Published: |
Інститут математики НАН України
2011
|
Series: | Symmetry, Integrability and Geometry: Methods and Applications |
Online Access: | http://dspace.nbuv.gov.ua/handle/123456789/146856 |
Tags: |
Add Tag
No Tags, Be the first to tag this record!
|
Journal Title: | Digital Library of Periodicals of National Academy of Sciences of Ukraine |
Cite this: | Essential Parabolic Structures and Their Infinitesimal Automorphisms / J. Alt // Symmetry, Integrability and Geometry: Methods and Applications. — 2011. — Т. 7. — Бібліогр.: 14 назв. — англ. |
Institution
Digital Library of Periodicals of National Academy of Sciences of Ukraineid |
irk-123456789-146856 |
---|---|
record_format |
dspace |
spelling |
irk-123456789-1468562019-02-12T01:24:11Z Essential Parabolic Structures and Their Infinitesimal Automorphisms Alt, J. Using the theory of Weyl structures, we give a natural generalization of the notion of essential conformal structures and conformal Killing fields to arbitrary parabolic geometries. We show that a parabolic structure is inessential whenever the automorphism group acts properly on the base space. As a corollary of the generalized Ferrand-Obata theorem proved by C. Frances, this proves a generalization of the ''Lichnérowicz conjecture'' for conformal Riemannian, strictly pseudo-convex CR, and quaternionic/octonionic contact manifolds in positive-definite signature. For an infinitesimal automorphism with a singularity, we give a generalization of the dictionary introduced by Frances for conformal Killing fields, which characterizes (local) essentiality via the so-called holonomy associated to a singularity of an infinitesimal automorphism. 2011 Article Essential Parabolic Structures and Their Infinitesimal Automorphisms / J. Alt // Symmetry, Integrability and Geometry: Methods and Applications. — 2011. — Т. 7. — Бібліогр.: 14 назв. — англ. 1815-0659 2010 Mathematics Subject Classification: 53B05; 53C05; 53C17; 53C24 DOI:10.3842/SIGMA.2011.039 http://dspace.nbuv.gov.ua/handle/123456789/146856 en Symmetry, Integrability and Geometry: Methods and Applications Інститут математики НАН України |
institution |
Digital Library of Periodicals of National Academy of Sciences of Ukraine |
collection |
DSpace DC |
language |
English |
description |
Using the theory of Weyl structures, we give a natural generalization of the notion of essential conformal structures and conformal Killing fields to arbitrary parabolic geometries. We show that a parabolic structure is inessential whenever the automorphism group acts properly on the base space. As a corollary of the generalized Ferrand-Obata theorem proved by C. Frances, this proves a generalization of the ''Lichnérowicz conjecture'' for conformal Riemannian, strictly pseudo-convex CR, and quaternionic/octonionic contact manifolds in positive-definite signature. For an infinitesimal automorphism with a singularity, we give a generalization of the dictionary introduced by Frances for conformal Killing fields, which characterizes (local) essentiality via the so-called holonomy associated to a singularity of an infinitesimal automorphism. |
format |
Article |
author |
Alt, J. |
spellingShingle |
Alt, J. Essential Parabolic Structures and Their Infinitesimal Automorphisms Symmetry, Integrability and Geometry: Methods and Applications |
author_facet |
Alt, J. |
author_sort |
Alt, J. |
title |
Essential Parabolic Structures and Their Infinitesimal Automorphisms |
title_short |
Essential Parabolic Structures and Their Infinitesimal Automorphisms |
title_full |
Essential Parabolic Structures and Their Infinitesimal Automorphisms |
title_fullStr |
Essential Parabolic Structures and Their Infinitesimal Automorphisms |
title_full_unstemmed |
Essential Parabolic Structures and Their Infinitesimal Automorphisms |
title_sort |
essential parabolic structures and their infinitesimal automorphisms |
publisher |
Інститут математики НАН України |
publishDate |
2011 |
url |
http://dspace.nbuv.gov.ua/handle/123456789/146856 |
citation_txt |
Essential Parabolic Structures and Their Infinitesimal Automorphisms / J. Alt // Symmetry, Integrability and Geometry: Methods and Applications. — 2011. — Т. 7. — Бібліогр.: 14 назв. — англ. |
series |
Symmetry, Integrability and Geometry: Methods and Applications |
work_keys_str_mv |
AT altj essentialparabolicstructuresandtheirinfinitesimalautomorphisms |
first_indexed |
2025-07-11T00:47:15Z |
last_indexed |
2025-07-11T00:47:15Z |
_version_ |
1837309472212516864 |
fulltext |
Symmetry, Integrability and Geometry: Methods and Applications SIGMA 7 (2011), 039, 16 pages
Essential Parabolic Structures
and Their Infinitesimal Automorphisms
Jesse ALT
School of Mathematics, University of the Witwatersrand,
P O Wits 2050, Johannesburg, South Africa
E-mail: jesse.alt@wits.ac.za
URL: http://sites.google.com/site/jmaltmath/
Received November 02, 2010, in final form April 11, 2011; Published online April 14, 2011
doi:10.3842/SIGMA.2011.039
Abstract. Using the theory of Weyl structures, we give a natural generalization of the
notion of essential conformal structures and conformal Killing fields to arbitrary parabolic
geometries. We show that a parabolic structure is inessential whenever the automorphism
group acts properly on the base space. As a corollary of the generalized Ferrand–Obata
theorem proved by C. Frances, this proves a generalization of the “Lichnérowicz conjec-
ture” for conformal Riemannian, strictly pseudo-convex CR, and quaternionic/octonionic
contact manifolds in positive-definite signature. For an infinitesimal automorphism with
a singularity, we give a generalization of the dictionary introduced by Frances for conformal
Killing fields, which characterizes (local) essentiality via the so-called holonomy associated
to a singularity of an infinitesimal automorphism.
Key words: essential structures; infinitesimal automorphisms; parabolic geometry; Lichné-
rowicz conjecture
2010 Mathematics Subject Classification: 53B05; 53C05; 53C17; 53C24
1 Introduction
1.1 Motivation from conformal geometry
Let (M, c) be a smooth, n-dimensional semi-Riemannian conformal manifold. For any choice
of semi-Riemannian metric g from the equivalence class c defining the conformal structure, we
have the obvious inclusion of the group of isometric diffeomorphisms of (M, g) in the group
of conformal diffeomorphisms of (M, c), Isom(M, g) ⊆ Conf(M, c). At the infinitesimal level
of vector fields, we have the corresponding inclusion of Killing fields in the conformal vector
fields, KVF(M, g) ⊆ CVF(M, c), which is obvious from the definitions: KVF(M, g) := {X ∈
X(M) | LXg = 0}; and CVF(M, c) := {X ∈ X(M) | ∃λ ∈ C∞(M) s.t.LXg = λg}.
A conformal diffeomorphism ϕ ∈ Conf(M, c) is called essential if ϕ is not an isometry of any
metric g ∈ c, and the conformal structure (M, c) is essential if Isom(M, g) is a proper subgroup
of Conf(M, c) for all representatives g ∈ c. Similarly, a conformal vector field X ∈ CVF(M, c)
is called essential if there is no representative g ∈ c for which X ∈ KVF(M, g). It is a fact –
although not necessarily obvious from the preceding definitions – that there are compact and
non-compact essential conformal structures in all dimensions n ≥ 2 and all signatures (p, q),
which moreover admit essential conformal vector fields. The standard compact example is
given by the conformal “Möbius sphere” (Sp,q, c) of any signature (p, q) (also called the Einstein
universe – these are the conformally flat homogeneous models of conformal geometry, which in
Riemannian signature are just the standard n-spheres equipped with the conformal class of the
round metric), while the standard non-compact example is Rp+q equipped with the conformal
class of the flat metric of signature (p, q). In fact, as a result of the following well-known
mailto:jesse.alt@wits.ac.za
http://sites.google.com/site/jmaltmath/
http://dx.doi.org/10.3842/SIGMA.2011.039
2 J. Alt
theorem, giving a positive answering to the so-called Lichnérowicz conjecture, we know that in
Riemannian signature these two examples are the only essential structures:
Theorem A (Ferrand–Obata for n ≥ 3). If (M, c) is an essential Riemannian conformal struc-
ture of dimension n ≥ 2, then it is conformally diffeomorphic to the n-dimensional sphere with
the round metric, or to n-dimensional Euclidean space.
For compact manifolds, this theorem was proven by M. Obata and J. Lelong-Ferrand in
the late 1960’s and early 1970’s. A proof for the non-compact case, announced in 1972 by
Alekseevski, was later discovered to be incomplete, and a complete proof was first given in 1994
by Ferrand (cf. [7, 9] and references therein). Recently, a corresponding result was proven at
the infinitesimal level by C. Frances [10] (note that this theorem does not simply follow from an
application of the Ferrand–Obata theorem, because the conformal vector fields are not assumed
to be complete):
Theorem B (Frances). Let (M, c) be a conformal Riemannian manifold of dimension n ≥ 3,
endowed with a conformal vector f ield X which vanishes at x0 ∈ M . Then either: (1) There
exists a neighborhood U of x0 on which X is complete, generates a relatively compact flow in
Conf(U, c), and is inessential on U , i.e. X ∈ KVF(U, g) for some g ∈ c|U ; or (2) There is
a conformally flat neighborhood U of x0, and X is essential on each neighborhood of x0.
1.2 Organization of the text and summary of main results
One direction of research into how these results do (or do not) generalize to other settings is
to consider the analogous questions for pseudo-Riemannian metrics, where essential conformal
structures turn out to be much more prevalent (cf. [9] for a survey). The aim of the present text
is to introduce natural generalizations of the notion of essential structure, and the corresponding
notion at the infinitesimal level, to the category of parabolic geometries. After this, we estab-
lish a generalization of Theorem A to a class of geometries which have been called “rank one
parabolic geometries”: conformal Riemannian structures; strictly pseudo-convex CR structures
of hypersurface type; positive-definite quaternionic contact structures; and octonionic contact
structures (cf. [2]). In fact, once our general definitions have been introduced and some basic
properties established, we only have to prove the easy part of this generalized Theorem A, the
difficult part having been taken care of in [8]; in the CR case, similar results to [8] were ob-
tained in [13] and [14]; in the quaternionic contact case, see [12]. Finally, we establish some local
properties of essential infinitesimal automorphisms, which generalize essential conformal vector
fields.
We begin in Section 2.1 with a review of relevant tools from parabolic geometry, in particular
the notions of Weyl structures introduced in [4], which are then used to generalize the notions
of essentiality to arbitrary parabolic geometries, cf. Definition 2.1. Next, we turn to some basic
properties of essential automorphisms in Section 2.2, establishing equivalent characterizations
which will be required in the proofs of the main results.
In Section 3 we establish a basic global result in the general parabolic setting: a parabolic
structure is essential only if the action of the automorphism group Aut(G, ω) on M is non-
proper (cf. Proposition 3.2, which is a generalization of a result proven in the conformal case by
Alekseevski in [1]). With this, we may apply the main theorem of [8] to prove a Lichnérowicz
theorem for rank one parabolic geometries, confirming the conjecture formulated in Section 2.2
of [9] for these geometries:
Theorem 1.1. Let (G → M,ω) be a regular rank one parabolic geometry, with M connected.
If this parabolic structure is essential, then M is geometrically isomorphic to either the compact
homogeneous model G/P or the noncompact space G/P\{eP}.
Essential Parabolic Structures and Their Infinitesimal Automorphisms 3
In Section 4 we establish some local properties of essential infinitesimal automorphisms which
generalize some of the results of [10]. We begin in Section 4.1 by recalling a result characterizing
infinitesimal automorphisms of arbitrary Cartan geometries (G →M,ω) via an identity involving
the curvature of ω. This identity was established in [3] for parabolic geometries and carries over
without difficulty to general Cartan geometries. Next, we show that the local study of essential
infinitesimal automorphisms amounts to studying their singularities, since any infinitesimal
automorphism of a parabolic geometry is inessential in some neighborhood of any point x such
that X(x) 6= 0 (cf. Proposition 4.2). Then, we apply the identity reviewed in Section 4.1 to
prove a generalization of results of [10], which give a “dictionary” relating essentiality of an
infinitesimal automorphism near a singularity x0 to properties of its holonomy ht at x0, a one-
parameter subgroup of P which is determined up to conjugacy (cf. Definition 4.2; it should be
emphasized that this notion of “holonomy” of the singularity of an infinitesimal automorphism
is distinct from the holonomy of, e.g., a Cartan connection which is common in the literature).
The main local result can be stated as:
Theorem 1.2. Let (G → M,ω) be a parabolic geometry of type (G,P ) and X ∈ inf(G, ω)
an infinitesimal automorphism with singularity at x0 ∈ M . Then X is inessential in some
neighborhood U of x0 if and only if the holonomy ht of X at x0 is up to conjugacy a subgroup
of Ker(λ) ⊂ G0 (equivalently, if and only if ω(X(u0)) ∈ Ker(λ′) ⊂ g0 for some u0 ∈ Gx0).
Already in conformal geometry, this result is of some interest because it can be used to
determine whether a conformal vector field is locally essential from looking at the adjoint tractor
it determines. We expect that the generalization to arbitrary parabolic geometries will be useful
in trying to generalize Theorem B to the other rank one parabolic geometries.
2 Essential automorphisms: basic definitions and properties
2.1 Background on parabolic geometries and their Weyl structures
Let us begin by recalling the definitions of parabolic geometries and their Weyl structures (the
latter, introduced by A. Čap and J. Slovák in [4], will be central to our notion of essential
parabolic structures). Parabolic geometries are certain types of Cartan geometries, which are
very general: given a closed subgroup P of a Lie group G, a Cartan geometry of type (G,P ) (or
modelled on the homogeneous space G/P ) is given by a principal P bundle π : G →M , equipped
with a Cartan connection ω. That is, ω ∈ Ω1(G, g) satisfies:
R∗p(ω) = Ad
(
p−1
)
◦ ω, for all p ∈ P ; (1)
ω(X̃) = X, for any X ∈ p, X̃ its fundamental vector field on G; (2)
ω(u) : TuG → g is a linear isomorphism for all u ∈ G. (3)
A Cartan geometry of type (G,P ) is a parabolic geometry if G is a real or complex semi-
simple Lie group, and P ⊂ G is a parabolic subgroup as in representation theory – at the level
of Lie algebras, this means, for g complex semi-simple, that p must contain a Borel (maximal
solvable) subalgebra of g; for g real semi-simple, the complexification p(C) must contain a Borel
subalgebra of g(C). (For a more detailed discussion of the basic properties of parabolic subgroups
and parabolic geometries, the reader is referred to [5]. Here we only attempt to cite some of the
key facts which are germane to the subsequent text.) In particular, in the parabolic setting the
Lie algebra g of G has an induced |k|-grading for some natural number k, so g = g−k ⊕ · · · ⊕ gk
with [gi, gj ] ⊆ gi+j and the subalgebra g− = g−k ⊕ · · · ⊕ g−1 is generated by g−1. The Lie
algebra of the parabolic subgroup P is the parabolic subalgebra p = g0 ⊕ · · · ⊕ gk, which has
Levi decomposition p = g0⊕ p+ with g0 reductive and p+ = g1⊕ · · · ⊕ gk the nilradical of p. At
4 J. Alt
the group level, we have a reductive subgroup G0 ⊂ P whose Lie algebra is g0, and P ∼= G0nP+
where P+ = exp(p+) is a normal, nilpotent subgroup of P globally diffeomorphic to p+ via the
exponential map.
The above-stated properties of the parabolic pair (G,P ) and their Lie algebras are used to
identify the following important geometric structures associated to a parabolic geometry (G →
M,ω) of type (G,P ). The orbit space G0 := G/P+ of the P+-action on G defines a G0-principal
bundle π0 : G0 → M , while by definition we also have a P+-principal bundle π+ : G → G0.
The filtration of g by Ad(P )-invariant submodules gi = gi ⊕ · · · ⊕ gk descends to a filtration
of g/p ∼= g− which is invariant under the quotient representation Ad : P → Gl(g/p), and thus
determines a filtration of the tangent bundle on the base space, TM = T−kM ⊃ · · · ⊃ T−1M ,
via the isomorphism
TM ∼=ω G ×Ad(P ) g/p,
(which holds in general for Cartan geometries) and setting T iM ∼=ω G×Ad(P )g
i/p. Furthermore,
the Cartan connection ω descends to G0 to identify it as a reduction to G0 of the structure
group of the associated graded tangent bundle gr(TM) = gr−k(TM) ⊕ · · · ⊕ gr−1(TM), where
gri(TM) = T iM/T i+1M .
The data (M, {T iM},G0) – consisting of a smooth manifold M , a filtration {T iM} of its
tangent bundle which satisfies rk(T iM) = dim(gi/p), and a reduction G0 of the structure group
of gr(TM) to G0 –, is called an infinitesimal flag structure of type (g, P ). The flag structure is
regular if the Lie bracket of vector fields on M respects the filtration, i.e. [Γ(T iM),Γ(T jM)] ⊂
Γ(T i+jM), and if the alternating bilinear form thus induced on gr(TM) gives it a point-wise
Lie algebra structure isomorphic to g−. When the flag structure is induced by a parabolic
geometry of type (G,P ), this regularity assumption can be related to an equivalent regularity
condition that the curvature of the Cartan connection ω have strictly positive homogeneity
(cf. 3.1.8 of [5]). A fundamental theorem of parabolic geometry states that, for any regular
infinitesimal flag structure of type (g, P ), there exists a regular parabolic geometry of type
(G,P ) which induces it. This parabolic geometry is uniquely determined up to isomorphism by
a normalisation condition on the curvature of the Cartan connection, except for a number of
parabolic types (G,P ) where the reduction of gr(TM) to G0 provides no additional information
and an extra geometric structure is needed for uniqueness (e.g. projective structures, where an
additional choice of an equivalence class of connections is needed to fix the structure). In these
cases, we will assume that the extra geometric structure is included when we speak of the “regular
infinitesimal flag structure”. We thus identify the geometric structure of an infinitesimal flag
structure with the regular, normal parabolic geometry inducing it.
In [4], Čap and Slovák define a Weyl structure for any parabolic geometry (G →M,ω) of type
(G,P ), to be a G0-equivariant section σ : G0 → G of the P+-principal bundle π+ : G → G0. We
denote the set of Weyl structures by Weyl(G, ω). By Proposition 3.2 of [4], global Weyl structures
always exist for parabolic geometries in the real (smooth) category, and they exist locally in the
holomorphic category. Considering the pull-back of the Cartan connection, σ∗ω, the |k|-grading
of g gives a G0-invariant decomposition into components, σ∗ω = σ∗ω−k + · · · + σ∗ωk, and by
the observation that σ commutes with fundamental vector fields (i.e. Tuσ(X̃(u)) = X̃(σ(u))
for X ∈ g0 and X̃ denoting the fundamental vector fields of X on G0 and G) and the defining
properties of the Cartan connection, it follows that σ∗ωi is horizontal for all i 6= 0, and that
σ∗ω0 defines a principal G0 connection on G0 →M (cf. 3.3 of [4]). In particular, we see that the
pair (G0 →M,σ∗ω≤) defines a Cartan geometry of type (P ∗, G0), where P ∗ ∼= exp(g−) oG0 is
the subgroup of G containing G0 with Lie algebra p∗ = g− ⊕ g0, and the Cartan connection is
given by
σ∗ω≤ = σ∗ω−k + · · ·+ σ∗ω0 ∈ Ω1(G0, p
∗).
Essential Parabolic Structures and Their Infinitesimal Automorphisms 5
One reason Weyl structures are very useful for studying a parabolic geometry, is that they
are in fact determined by very simple induced geometric objects, namely by the R+-principal
connections they induce on certain ray bundles associated to G0. Fix an element Eλ in the
center of the reductive Lie algebra g0 such that ad(Eλ) acts by scalar multiplication on each
grading component gi of g (for example the grading element E, which always exists and satisfies
ad(E)|gi = i·). Then there is a unique representation λ : G0 → R+ satisfying λ′(A) = B(Eλ, A)
for all A ∈ g0, B the Killing form, and hence an associated R+-principal bundle Lλ → M . We
have Lλ ∼= G0/Ker(λ) so let us denote the projection πλ : G0 → Lλ. For any Weyl structure σ, the
1-form λ′ ◦ σ∗ω0 ∈ Ω1(G0) induces a R+-principal connection σλ on Lλ. After introducing these
objects and studying their properties in Section 3 of [4], Čap and Slovák prove the fundamental
result that the correspondence σ 7→ σλ defines a bijective correspondence between the set of
Weyl structures and the set of principal connections on Lλ (cf. Theorem 3.12 of [4]).
In particular, this fact makes it possible to define certain distinguished classes of Weyl struc-
tures: A Weyl structure σ is closed if the induced R+-principal connection σλ has vanishing
curvature; it is exact if σλ is a trivial connection induced by a global trivialisation of the scale
bundle Lλ → M . Čap and Slovák prove that closed and exact Weyl structures always exist
(in the smooth category), and the spaces of closed and exact Weyl structures are affine spaces
over the closed, respectively over the exact, 1-forms on M . Assuming a scale bundle Lλ to be
fixed, we denote the set of exact Weyl structures by Weyl(G, ω), which is naturally identified
with the set of global sections of Lλ, and note that this set is non-empty (cf. Proposition 3.7
of [4]). Equivalently, an exact Weyl structure σ is characterized by the existence of a holonomy
reduction of the G0-principal connection σ∗ω0 to the subgroup Ker(λ) ⊂ G0 (cf. Sections 3.13,
3.14 of [4]). We will denote this reduction by r : G0 ↪→ G0, and the corresponding reduction of G
to the structure group Ker(λ) by
σ := σ ◦ r : G0 → G.
Thus an exact Weyl structure determines a Cartan geometry (G0 → M,σ∗ω≤) of type
(P ∗,Ker(λ)) for P ∗ ∼= exp(g−) o Ker(λ) the subgroup of G containing Ker(λ) with Lie al-
gebra p∗ := g−⊕Ker(λ′). In the conformal case, exact Weyl structures correspond to metrics in
the conformal equivalence class, while a general Weyl structure is given by a Weyl connection,
i.e. a torsion free connection which preserves the conformal equivalence class.
2.2 Definition and basic properties of essential structures
Now we are ready to define essential parabolic structures and essential infinitesimal auto-
morphisms. For now, let us take the following definitions for automorphisms, respectively
infinitesimal automorphisms, of a Cartan geometry. For a Cartan geometry (G → M,ω) of
arbitrary type (G,P ), an automorphism Φ ∈ Aut(G, ω) is a P -principal bundle morphism of G
such that Φ∗ω = ω. An infinitesimal automorphism X ∈ inf(G, ω) is given by X ∈ X(G), such
that (Rp)∗X = X and the Lie derivative satisfies LXω = 0.
Note that when (G, ω) is a parabolic geometry, we get naturally induced G0-bundle morphisms
Φ0 : G0 → G0, R+-bundle morphisms Φλ : Lλ → Lλ, and diffeomorphisms ϕ : M → M for the
Φ ∈ Aut(G, ω). These are induced thanks to the P -equivariance of Φ, by using commutativity
with the appropriate projections, i.e. the defining identities are Φ0◦π+ = π+◦Φ, Φλ◦πλ = πλ◦Φ0
and ϕ ◦ π = π ◦Φ. Hence, we get a natural action of the automorphism group Aut(G, ω) on the
set of (exact) Weyl structures, by defining, for Φ ∈ Aut(G, ω), σ ∈Weyl(G, ω),
Φ∗σ := Φ−1 ◦ σ ◦ Φ0 : G0 → G,
and for σ ∈Weyl(G, ω) with corresponding global scale sσ ∈ Γ(Lλ),
Φ∗sσ := Φ−1
λ ◦ sσ ◦ ϕ : M → Lλ.
6 J. Alt
Using the defining relation π+ ◦ Φ−1 = Φ−1
0 ◦ π+ for Φ−1
0 and the corresponding relation bet-
ween Φ−1
λ and ϕ−1, we can verify that Φ∗σ ∈ Γ(G → G0) and Φ∗sσ ∈ Γ(Lλ →M); furthermore,
Φ∗σ is G0-equivariant by the equivariance properties of σ, Φ0 and Φ−1; i.e. Φ∗σ ∈ Weyl(G, ω)
and Φ∗sσ ∈Weyl(G, ω).
Similarly, for an infinitesimal automorphism X ∈ X(G), we get naturally induced vector fields
X0 ∈ X(G0), Xλ ∈ X(Lλ) and X ∈ X(M) (with the first two being invariant with respect to
the appropriate right-actions). For an arbitrary point of M , we may choose ε > 0 sufficiently
small so that ΦX,t, Φ−1
X,t, ΦX0,t, etc. all exist for −ε ≤ t ≤ ε, and so on this interval we have
a well-defined family
Φ∗X,tσ := Φ−1
X,t ◦ σ ◦ ΦX0,t ∈Weyl(G, ω),
for any σ ∈Weyl(G, ω), which is differentiable in t at t = 0, so we can define the Lie derivative
LXσ := (d/dt)|t=0Φ∗X,tσ; this is an element of the vector space on which the space of Weyl
sections (an affine space) is modeled, i.e. it can be thought of as a 1-form on M . For sσ ∈
Weyl(G, ω), the Lie derivative LXsσ is defined analogously, and it can be identified with an
exact 1-form on M .
Definition 2.1. For (G → M,ω) a parabolic geometry of type (G,P ), Φ ∈ Aut(G, ω) an
automorphism of the geometry, and σ ∈ Weyl(G, ω) a Weyl structure, Φ is an automorphism
of σ, written Φ ∈ Aut(σ), if and only if Φ∗σ = σ (equivalently, Φ◦σ = σ◦Φ0). If σ ∈Weyl(G, ω)
is an exact Weyl structure corresponding to sσ ∈ Γ(Lλ), then Φ is an exact automorphism of σ,
written Φ ∈ Aut(σ), if and only if Φ∗sσ = sσ (equivalently, Φλ ◦ sσ = sσ ◦ϕ). An automorphism
Φ ∈ Aut(G, ω) is essential if it is not an exact automorphism of any exact Weyl structure
σ ∈Weyl(G, ω). We call (G, ω) an essential parabolic structure if Aut(σ) $ Aut(G, ω) for every
exact Weyl structure σ. We call a regular infinitesimal flag structure M = (M, {T iM},G0)
of type (g, P ) an essential structure if the regular, normal parabolic geometry inducing it is
essential.
An infinitesimal automorphism X ∈ inf(G, ω) is an infinitesimal automorphism of σ, written
X ∈ inf(σ), if and only if LXσ = 0. For an exact Weyl structure σ ∈ Weyl(G, ω), X is an
exact infinitesimal automorphism of σ, written X ∈ inf(σ), if and only if LXsσ = 0. An
infinitesimal automorphism is essential if it is not an exact infinitesimal automorphism for any
exact Weyl structure. ForM a regular infinitesimal flag structure as above, we say a vector field
X ∈ X(M) is an essential infinitesimal automorphism ofM if it lifts to an essential infinitesimal
automorphism of the canonical parabolic geometry inducing M.
Remark 2.2. Charles Frances has pointed out to us the definition of essential parabolic struc-
ture given in Section 2.2 of [9], which did not make explicit use of Weyl structures. That
definition turns out to be almost equivalent to the one above, cf. Lemma 2.1, except that
in some cases the semi-simple part of the reductive group G0 can be properly contained in
Ker(λ). For example, in the case of strictly pseudoconvex CR structures, G0
∼= R+ × U(n)
and Ker(λ) ∼= U(n) for an appropriate choice of scale λ. An exact Weyl structure is seen to be
equivalent to a choice of pseudo-hermitian form for the CR structure, and so the above definition
of exact structure amounts to requiring that the group of CR transformations preserving any
pseudo-hermitian form is always properly contained in the group of CR transformations.
Remark 2.3. Definition 2.1 recovers the classical definition of essentiality when the regular
infinitesimal flag structure is given by a conformal semi-Riemannian structure (M, c) of signature
(p, q). In that case, G0 is just the conformal group R+×O(p, q), G0 is the bundle of frames which
are semi-orthonormal with respect to some metric g ∈ c, and the choice of scale representation,
λ : R+ ×O(p, q)→ R+, λ : (s,A) 7→ s−1,
Essential Parabolic Structures and Their Infinitesimal Automorphisms 7
identifies Lλ ∼= G0/Ker(λ) with the ray bundle Q → M of metrics in the conformal class,
with the standard R+-action given by gx.s := s2gx for any g ∈ c and x ∈ M corresponding to
gx ∈ Q. Exact Weyl structures thus correspond to choices of a metric in the conformal class, and
a conformal diffeomorphism ϕ which uniquely corresponds to an automorphism Φ ∈ Aut(G, ω) of
the canonical conformal Cartan geometry is essential in the sense of Definition 2.1 if Φ /∈ Aut(σ)
for all exact Weyl structures σ, i.e. if ϕ fails to preserve all metrics in the conformal class.
Remark 2.4. Let M = (M, {T iM},G0) denote a regular infinitesimal flag structure of some
parabolic type (g, P ), if necessary including the extra geometric data required so that the regular,
normal parabolic geometry of type (G,P ) inducing it is unique up to isomorphism. If we
need to distinguish this parabolic geometry from others of the same type, we will use the
notation (G, ωnc) to signify the canonical (normal Cartan) geometry. In this setting, we can
define an automorphism of the structure in terms of M: An automorphism of the regular
infinitesimal flag structure, ϕ ∈ Aut(M), is a diffeomorphism ϕ ∈ Diff(M) which satisfies:
(i) ϕ∗(T
i
xM) ⊆ T iϕ(x)M for all x ∈ M and all −k ≤ i ≤ −1; and (ii) the induced bundle
map gr(ϕ) (which as a consequence of (i) is a lift of ϕ defined on the bundle F(gr(TM)) of
frames of the associated graded tangent bundle) preserves G0 as a subbundle of F(gr(TM))
(and hence gr(ϕ) restricts to a G0-bundle morphism Φ0 of G0). We can identify Aut(M) with
Aut(G, ωnc) since by uniqueness of (G, ωnc) up to isomorphism, ϕ (and Φ0) lift to a unique P -
bundle morphism Φ of G preserving ωnc under pullback. Thus, we can think of an automorphism
ϕ ∈ Aut(M) as including as well the automorphism Φ ∈ Aut(G, ωnc) and the induced G0-bundle
morphism Φ0 = gr(ϕ)|G0 of G0.
Lemma 2.1. Let Φ ∈ Aut(G, ω) be an automorphism of a parabolic geometry, let σ be a Weyl
structure, and let a scale bundle Lλ →M be fixed. The following are equivalent:
(i) Φ ∈ Aut(σ);
(ii) For the induced bundle morphism Φ0 : G0 → G0, we have Φ0 ∈ Aut(G0, σ
∗ω≤);
(iii) Φλ preserves the scale bundle connection σλ ∈ Ω1(Lλ): Φ∗λσ
λ = σλ.
If σ is exact, then Φ ∈ Aut(σ) if and only if the induced bundle morphism Φ0 preserves the
sub-bundle G0 ⊂ G0 and the restriction satisfies (Φ0)|G0 ∈ Aut(G0, σ
∗ω≤).
For X ∈ inf(G, ω), X ∈ inf(σ) if and only if X0 ∈ inf(G0, σ
∗ω≤) and, for σ exact, X ∈ inf(σ)
if and only if the restriction of X0 to G0 is tangent to G0 and induces an element of inf(G0, σ
∗ω≤).
Proof. (i) ⇒ (ii), since by Φ ◦ σ = σ ◦ Φ0 and Φ ∈ Aut(G, ω) we have Φ∗0(σ∗ω) = σ∗ω,
and in particular Φ∗0(σ∗ω≤) = σ∗ω≤. And (ii) ⇒ (iii), since Φ∗0(σ∗ω≤) = σ∗ω≤ implies that
Φ∗0(σ∗ω0) = σ∗ω0. In particular, Φ∗0(λ′ ◦ σ∗ω0) = λ′ ◦ Φ∗0(σ∗ω0) = λ′ ◦ σ∗ω0. Hence, the R+-
bundle morphism Φλ and the R+-principal connection σλ, induced on Lλ by Φ0 and λ′ ◦ σ∗ω0,
respectively, satisfy: Φ∗λσ
λ = σλ.
We now show (iii) ⇒ (i): Consider the R+-principal connection (Φ∗σ)λ ∈ Ω1(Lλ). It is
induced by:
λ′ ◦ (Φ∗σ)∗ω0 = λ′ ◦
((
Φ−1 ◦ σ ◦ Φ0
)∗
ω0
)
= λ′ ◦
(
Φ∗0σ
∗(Φ−1
)∗
ω0
)
= λ′ ◦ (Φ∗0σ
∗ω0) = Φ∗0(λ′ ◦ σ∗ω0).
So (Φ∗σ)λ = Φ∗λσ
λ, which equals σλ by assumption (iii). Thus, by Theorem 3.12 of [4] (cf.
discussion in Section 2.1), the Weyl structures Φ∗σ and σ are equal, showing that (i) holds.
To see the final statement of the lemma, let σ be an exact Weyl structure and let us denote
by sσ ∈ Γ(Lλ) the global scale which induces the trivial connection σλ ∈ Ω1(Lλ). That is, for
any point p = sσ(x).r ∈ Lλ, for x ∈M , r ∈ R+, we have the decomposition
TpLλ = (Rr)∗((sσ)∗(TxM))⊕ Rζ1(p),
8 J. Alt
for ζ1 the fundamental vector field on Lλ of the vector 1 ∈ R; the value of σλ on a tangent
vector v ∈ TpLλ is given by the coefficient of ζ1(p) determined by this decomposition. Then the
holonomy reduction of (Lλ, σλ) to the trivial structure group is given by sσ(M) ⊂ Lλ, and the
reduction of (G0, σ
∗ω0) to Ker(λ) is given by
G0 = π−1
λ (sσ(M)) ⊂ G0.
Hence for x ∈M and u ∈ (G0)x, we have πλ(u) = sσ(x) and Φ0(u) ∈ G0 if and only if πλ(Φ0(u)) =
sσ(π0(Φ0(u))). But πλ(Φ0(u)) = Φλ(πλ(u)) = Φλ(sσ(x)) and π0(Φ0(u)) = ϕ(π0(u)) = ϕ(x).
Thus, Φ0(u) ∈ G0 if and only if Φλ(sσ(x)) = sσ(ϕ(x)). But if Φ ∈ Aut(σ), then clearly Φλ
preserves the induced (trivial) connection σλ ∈ Ω1(Lλ), so by (iii) ⇔ (ii) shown above we
have Φ∗0(σ∗ω≤) = σ∗ω≤ and since Φ0 preserves G0 the corresponding identity follows for the
restriction and σ∗ω≤.
The statements for X ∈ inf(G, ω) are proven in the same manner. �
Remark 2.5. In particular, it follows from the proof of Lemma 2.1 that any exact automor-
phism Φ of an exact Weyl structure σ is also an automorphism of σ, i.e. Φ ∈ Aut(σ) ⇒ Φ ∈
Aut(σ) as one would hope.
On the other hand, the converse does not hold: If Φ ∈ Aut(σ) for an exact Weyl structure σ,
the requirement that Φ0(G0) ⊂ G0 is necessary to guarantee that an automorphism Φ of an exact
Weyl structure σ is in fact an exact automorphism. An instructive example is the conformal
structure induced by the Euclidean metric on Rn: The diffeomorphism given by dilation by
a positive constant r is always an automorphism of the exact Weyl structure corresponding to
the Euclidean metric. However, for r 6= 1, this diffeomorphism is not an isometry, hence not an
exact automorphism. We are grateful to Felipe Leitner for bringing this to our attention, which
led us to modify an earlier version of Definition 2.1.
3 Lichnérowicz theorem for rank one parabolic geometries
In this section, we establish Theorem 1.1 for the so-called rank one parabolic geometries. These
are parabolic geometries of types (G,P ) such that the homogeneous model G/P is the boundary
of a real rank one symmetric space G/K (for K a maximal compact subgroup). The bound-
ary G/P is diffeomorphic to the sphere of dimension dim(G/K) − 1, while the symmetric
spaces G/K are given by the hyperbolic spaces of real, complex and quaternionic type in all ap-
propriate real dimensions, and the hyperbolic Cayley plane in real dimension 16. As mentioned
earlier, the rank one parabolic geometries associated to the corresponding types are, respec-
tively, conformal Riemannian structures, strictly pseudo-convex partially-integrable CR struc-
tures, quaternionic contact structures (of positive-definite signature), and octonionic contact
structures. The key result needed to prove this theorem is the following, proved by C. Frances
(Theorem 3 of [8]), which generalizes theorems of Ferrand [7] and Schoen [13] in the cases of
conformal Riemannian and strictly pseudo-convex CR structures:
Theorem 3.1 (Frances, [8]). Let (G →M,ω) be a regular rank one parabolic geometry, with M
connected. If Aut(G, ω) acts improperly on M , then M is geometrically isomorphic to either the
compact homogeneous model G/P or the noncompact space G/P\{eP}.
Here, “geometrically isomorphic” means there is a diffeomorphism of M onto the space in
question, which is covered by a morphism of Cartan bundles which pulls back the Maurer–Cartan
connection to ω. The assumption that the parabolic geometries are of rank one type is key for
the argument in [8], because it allows the author to exploit the so-called “north-south dynamics”
on the homogeneous model G/P . Theorem 1.1 now follows as a result of Theorem 3.1 and the
Essential Parabolic Structures and Their Infinitesimal Automorphisms 9
following proposition, which generalizes a result of [1] and whose proof follows the same line of
argumentation:
Proposition 3.2. If (G → M,ω) is an essential parabolic structure, then Aut(G, ω) acts im-
properly on M .
Proof. Fix a bundle of scales Lλ →M for (G, ω). Assume that Aut(G, ω) acts properly on M
and let us show that the parabolic structure is not essential. By definition, it suffices to construct
a global scale s : M → Lλ which is Aut(G, ω)-invariant, i.e. such that Φλ ◦ s = s ◦ϕ holds for all
Φ ∈ Aut(G, ω) and Φλ : Lλ → Lλ, ϕ : M →M the induced diffeomorphisms.
We construct this Aut(G, ω)-invariant scale s using classical properties of proper group ac-
tions. The so called “tube theorem” (alias “slice theorem”, cf. e.g. Theorem 2.4.1 in [6]) guaran-
tees the following, for a C∞-action of a Lie group H on a manifold M which is proper at x ∈M :
There exists a H-invariant neighborhood U of x on which the H-action is equivalent to the left
H-action on the quotient space H ×K B – for K ⊂ H a compact subgroup and B a K-invariant
neighborhood of 0 in a K-module V – given by h1.[h2, b] = [h1h2, b] for hi ∈ H, b ∈ B and [h, b]
the equivalence class of (h, b) ∈ H ×B under the left K-action k.(h, b) := (h.k−1, k.b). Starting
from a choice of global scale s0 : M → Lλ, and letting H = Aut(G, ω), e ∈ H the identity
automorphism and Φ ∈ H arbitrary, set:
sU ([e, b]) :=
∫
Ψ∈K
(Ψλ)−1(s0([Ψ, b]))dΨ; (4)
sU ([Φ, b]) := Φλ(sU ([e, b])). (5)
One verifies that this gives a well-defined local section sU : U → Lλ|U , which involves checking
that for (e, b) ∼ (Φ, b′) (i.e. for Φ ∈ K and b = ϕ(b′)) the values sU ([e, b]) given by (4) and
sU ([Φ, b′]) given by (5), agree. This follows by unwinding the definitions, and using a bi-invariant
Haar measure dΨ on the compact group K. And since [Φ, b] = Φ.[e, b] corresponds to the point
ϕ(x′) for x′ ' [e, b], the defining equation (5) automatically gives us the invariance property,
sU ◦ ϕ = Φλ ◦ sU .
We go from local Aut(G, ω)-invariant sections sU : U → LλU to a global Aut(G, ω)-invariant
scale sinv ∈ Γ(Lλ) as follows: First, note that Lλ admits a global section s ∈ Γ(Lλ) and hence
a global trivialisation Lλ ∼=s M ×R+ by identifying s(x).r 's (x, r) for any x ∈M and r ∈ R+.
This in turn induces a bijection rs : Γ(Lλ) → C∞(M,R+). For any Φ ∈ Aut(G, ω), define
the smooth function ϕs ∈ C∞(M,R+) by the identity Φλ(s(x)) = s(ϕ(x)).ϕs(x). Then we see
that a Φ-invariant scale s′ ∈ Γ(Lλ) corresponds, under the bijection rs, to a smooth function
r′ = rs(s
′) ∈ C∞(M,R+) which satisfies r′(ϕ(x)) = ϕs(x)r′(x).
Next, note that since Aut(G, ω) acts properly on M , there exists a covering {Uα} of M by
Aut(G, ω)-invariant open sets as above (so they all admit invariant local scales sα : Uα → LλUα
and we denote rα := rs(sα) ∈ C∞(Uα,R+)); and (applying Theorem 6 of [1]) there exists
a partition of unity {fi} for M which is subordinate to {Uα} (so supp(fi) ⊂ Uα(i) for all i) and
the fi are Aut(G, ω)-invariant. Now we define sinv ∈ Γ(Lλ) via rinv = rs(sinv) ∈ C∞(M,R+)
and the formula
rinv(x) =
∑
i
fi(x)rα(i)(x)
for all x ∈M . This is well-defined, smooth, and from the Aut(G, ω)-invariance of the fi, together
with the Aut(G, ω)-invariance of the local sections sα, we compute that rinv(ϕ(x)) = ϕs(x)rinv(x)
for all x ∈M and all Φ ∈ Aut(G, ω), i.e. sinv ∈ Γ(Lλ) is a Aut(G, ω)-invariant global scale. �
10 J. Alt
4 Proof of local results
A key reference for the study of infinitesimal automorphisms of parabolic geometries is [3].
In that text, A. Čap generalized to arbitrary parabolic geometries a bijective correspondence
between conformal vector fields and adjoint tractors (sections of the associated bundle to the
canonical Cartan bundle, G → M , induced by the adjoint representation on g) satisfying an
identity involving the Cartan curvature, which was first discovered by A.R. Gover in [11]. More-
over, the text of Čap relates this general bijective correspondence to the first splitting operator
of a so-called curved BGG-sequence for the parabolic geometry, cf. Theorem 3.4 of [3]. The
curvature identity of [3] extends without difficulty to general infinitesimal automorphisms of
Cartan geometries. This allows us to apply this fundamental identity to the Cartan geometries
(G0, σ
∗ω≤), which we do in Section 4.2 to establish a general “dictionary” between essentiality
of an infinitesimal automorphism near a singularity, and the so-called holonomy associated to
such a singularity (cf. Definition 4.2; again, this should not be confused with the holonomy of
the Cartan connection).
4.1 Background results on infinitesimal automorphisms
We recall some general notions, mainly following the development of [3] (cf. also 1.5 of [5]), in
the setting of a general Cartan geometry (G →M,ω) of type (G,P ) (for now not assumed to be
of parabolic type). For any representation ρ : P → Gl(V ), we have the associated vector bundle
V (M) := G ×ρ V . The smooth sections of such a bundle are identified with P -equivariant,
V -valued smooth functions on G in the standard manner, and we will simply treat them as such:
Γ(V (M)) =
{
f ∈ C∞(G, V ) | f(u.p) = ρ(p−1)(f(u))
}
=: C∞(G, V )P .
For the most part, the important associated bundles we are dealing with are tractor bundles,
which for our purposes simply means that the representation (ρ, V ) is the restriction to P of
a G-representation ρ̃ : G → Gl(V ). And the primary tractor bundle is the adjoint bundle
induced by the restriction of the adjoint representation Ad : G → Gl(g) to P , which we will
denote by A = A(M) if there is no danger of confusion about which Lie algebra g is meant,
and otherwise by g(M). Note that the Lie bracket [· , ·]g of g, by Ad(P )-invariance, determines
an algebraic bracket on fibers of A as well as on sections, which we denote with curly brackets
{· , ·} : A×A → A. Also, note that there is a natural projection Π : A → TM , induced by the
projection g→ g/p and the isomorphism TM ∼=ω G ×Ad(P ) g/p.
The Cartan connection determines an identification of right-invariant vector fields
X(G)P = {X ∈ X(G) |X(u.p) = (Rp)∗(X(u))},
with sections of the adjoint bundle. Namely, to X ∈ X(G) we associate a function sX ∈ C∞(G, g)
defined by sX(u) := ω(X(u)); conversely, to a function s ∈ C∞(G, g), associate Xs ∈ X(G)
defined by Xs(u) = ω−1
u (s(u)). The property (3) of a Cartan connection insures that both
maps are well-defined, they are inverse, and by property (1) of ω these maps restrict to an
isomorphism X(G)P ∼=ω Γ(A). Similarly, the Cartan connection ω induces natural identifications
Ωk(G; g)
P ∼=ω Ωk(M ;A) of the horizontal, Ad(P )-equivariant g-valued k-forms on G with the
A-valued k-forms on M . For a tractor bundle V (M), the identification X(G)P ∼=ω Γ(A) yields
two kinds of differentiation of smooth sections with respect to adjoint tractors:
Definition 4.1. The invariant differentiation or fundamental D-operator of V (M) is the map
DV : Γ(V (M))→ Γ(A∗ ⊗ V (M)) defined, for any s ∈ Γ(A) and any v ∈ Γ(V (M)), by:
DV
s v := Xs(v).
Essential Parabolic Structures and Their Infinitesimal Automorphisms 11
The tractor connection of V (M) is the map ∇V : Γ(V (M)) → Γ(A∗ ⊗ V (M)) defined, for any
s ∈ Γ(A) and any v ∈ Γ(V (M)), by:
∇Vs v := DV
s v + (dρ̃ ◦ s) ◦ v. (6)
Recall that ρ̃ : G → Gl(V ) is the G-representation which ρ is a restriction of, given by the
definition of a tractor bundle. In fact, the quantity defined by (6) only depends on the equiva-
lence class [s] of s under the quotient A/p(M) = G ×Ad(P ) g/p
∼= TM , and we identify the
tractor connection with a covariant derivative on V (M):
∇V : Γ(V (M))→ Γ(T ∗M ⊗ V (M)).
The curvature tensor of a Cartan connection is the g-valued two-form on G defined, for
any u ∈ G and v, w ∈ TuG, by the structure equation Ωω(v, w) := dω(v, w) + [ω(v), ω(w)]. The
curvature tensor is horizontal and P -equivariant, and we may equivalently consider the curvature
function κω ∈ C∞(G; Λ2(g/p)∗ ⊗ g)P induced by κω(u)(X,Y ) := Ωω(ω−1
u (X), ω−1
u (Y )) for any
u ∈ G and X,Y ∈ g. The following identity was proven for parabolic geometries in [3] and the
proof carries over without substantive changes to Cartan geometries of general type (cf. also
Lemma 1.5.12 in [5]):
Lemma 4.1. Let X 'ω sX for X ∈ X(G)P and sX ∈ Γ(A). Then for LXω ∈ Ω1(G; g)
P
we
have, under the identification Ω1(G; g)
P ∼=ω Ω1(M ;A):
LXω 'ω ∇AsX + Π(sX)yκω. (7)
4.2 Holonomy and essentiality of infinitesimal automorphisms
Let us return now to the setting of a (regular, normal) parabolic geometry (G → M,ω) of
type (G,P ) and the corresponding regular infinitesimal flag structure M = (M, {T iM},G0)
of type (g, P ). As was noted for automorphisms of (G, ωnc) in Remark 2.4, we may deter-
mine an infinitesimal automorphism X ∈ inf(G, ω) by conditions on the underlying vector
field X ∈ X(M), which just amount to imposing the same conditions for the locally defined
diffeomorphisms given by flowing along X, i.e. we must have [X,Γ(T iM)] ⊆ Γ(T iM) for all
−k ≤ i ≤ −1, and the condition that the local flows of X determine local bundle maps of
F(gr(TM)) which preserve G0 as a subbundle. In the different examples of parabolic geome-
tries, this translates into more geometric language. For example, in the conformal case, the
former condition is trivial, while the latter condition amounts to requiring the conformal Killing
equation, LXg = λg for any g ∈ c and some λ = λ(X) ∈ C∞(M). In the case of CR structures,
the conditions are that [X,Γ(H)] ⊆ Γ(H) for H ⊂ TM the codimension one contact distribution
defining the CR structure, and that LXJ = 0 for J the almost complex structure on H. We
write X ∈ inf(M) and consider the lift X ∈ inf(G, ωnc) to be implicitly included.
Using Lemma 4.1 and Lemma 2.1, we obtain a bijection between infinitesimal automorphisms
X ∈ inf(G, ω) and adjoint tractors sX ∈ Γ(A) satisfying ∇AsX + Π(sX)yκω = 0. In the present
setting, this gives us a bijection between X ∈ inf(M) and such sX, and moreover it is easy
to verify that Π(sX) = X ∈ X(M). Now, denote by X0 ∈ X(G0)G0 the invariant vector field
induced, via projection by π+, by X ∈ X(G)P . For any Weyl structure σ : G0 → G, the induced
Cartan connection σ∗ω≤ gives us an isomorphism denoted X(G0)G0 ∼=σ Γ(Aσ), where we denote
with Aσ = p∗(M) the adjoint bundle of the Cartan geometry (G0 → M,σ∗ω≤). Let us write
X0 'σ sX0 for the adjoint tractor corresponding to X0 ∈ X(G0)G0 . If we further denote by
∇σ : Γ(Aσ) → Γ(T ∗M ⊗ Aσ) the corresponding adjoint tractor connection, then Lemma 4.1
12 J. Alt
tells us that X ∈ inf(M) is an infinitesimal automorphism of some Weyl structure σ : G0 → G
if and only if
∇σsX0 +Xyκσ
∗ω≤ = 0.
At present, we are only interested in local properties of infinitesimal automorphisms, viz the
question if some neighborhood of a given point can be found on which X is inessential. The
following proposition shows that the points x ∈ M for which the answer could be “no”, must
all be singularities of the vector field, i.e. X(x) = 0 (where, incidentally, the above identity
simplifies to (∇σsX0)(x) = 0).
Proposition 4.2. Let M = (M, {T iM},G0) be a regular infinitesimal flag structure of type
(g, P ), let X ∈ inf(M), and let x ∈ M . If X(x) 6= 0, then there exists a neighborhood U of x
such that the restriction of X to U is inessential.
Proof. Take a neighborhood U of x on which flow-box coordinates for the flow of X can be
introduced:
U = {(x0, ϕX,t(x0)) |x0 ∈M0 ∩ U, −ε < t < ε},
where M0 is some locally defined hypersurface transversal to the integral curves ϕX,t(x0) of X,
which are defined for the interval given. This can be done since we may first restrict an open
neighborhood of x on which X is non-vanishing. But this provides all the features needed
to transfer the argument used in the proof of Proposition 3.2 to establish the existence of
local Aut(G, ω)-invariant sections sU : U → Lλ|U to the current context: Namely, the formula
s(x0, ϕX,t(x0)) := ΦXλ,t(s(x0)) gives a well-defined local scale s : U → Lλ|U which by construction
is invariant under the flows ϕX,t for t sufficiently small. �
From now on, let us fix a singularity x ∈M of an infinitesimal automorphism X ∈ inf(M).
We also choose a point u ∈ G in the fiber over x, and let u0 = π+(u) ∈ G0, likewise in the fiber
over x. The remaining text is aimed at relating the local essentiality of X near x, to invariant
properties of the holonomy of X at x, which is a one-parameter subgroup ht ⊂ P :
Definition 4.2 (cf. [10], Section 6). Given X, x and u as above, the holonomy htu of X at x
with respect to u is defined, for t sufficiently small, as follows: Let ΦX,t(u), the integral curve
of X through u, be defined for t ∈ (−ε, ε). Since X projects to X, X(u′) is tangent to Gx for all
u′ ∈ Gx, and hence all ΦX,t(u) lie in Gx. Then htu ∈ P is defined by:
ΦX,t(u) =: u.htu.
Since ht+su = htuh
s
u whenever both are defined, htu = exp(tXh,u) for some Xh,u ∈ p, and we
define htu via this identity for all t ∈ R.
Recall that, by definition, sX(u) = ω(X(u)). Also, since htu = exp(tXh,u), we have Xh,u =
(d/dt)|t=0h
t
u. By definition, the integral curve ΦX,t(u) satisfies Φ′X(0) = X(u). Hence, sX(u) =
ω((d/dt)|t=0(u.htu)), and so by property (2) of the Cartan connection ω, we have
Xh,u = sX(u).
In particular, this implies the following equivariance properties of Xh,u and htu with respect to
a change of the base point u ∈ Gx, so it makes sense to speak of the holonomy ht of X at x as
a conjugacy class of one-parameter groups in P :
Xh,u.p = Ad(p−1)(Xh,u), ∀ p ∈ P ;
htu.p = p−1htup, ∀ p ∈ P.
The first part of relating essentiality of X near x to its holonomy, is the following:
Essential Parabolic Structures and Their Infinitesimal Automorphisms 13
Proposition 4.3. Let X ∈ inf(M) have a singularity x ∈ M , and let u ∈ Gx be as above.
If X ∈ inf(σ) for any locally defined Weyl structure σ, then X has holonomy htu conjugate
under P to a one-parameter subgroup of G0 (equivalently, sX(u.p) = Ad(p−1)(sX(u)) ∈ g0 for
some p ∈ P ).
Proof. Assume that σ is any locally defined Weyl structure in a neighborhood of the point x
withX ∈ inf(σ). From Lemma 2.1, it follows that X0 ∈ inf(G0, σ
∗ω≤), i.e. we have LX0σ
∗ω≤ = 0.
Then note that the identity (7) from Lemma 4.1 simplifies, for u ∈ Gx and u0 := π+(u) ∈ (G0)x,
to give us the following two identities:
(∇AsX)(u) = 0; (8)
(∇σsX0)(u0) = 0. (9)
To prove the claim in the proposition, we compute the identity (8) in terms of σ, to show that (9)
implies sX(σ(u0)) ∈ g0. Since σ(u0), u ∈ Gx, therefore sX(σ(u0)) = Ad(p−1)(sX(u)) ∈ g0 for
some p ∈ P as claimed.
For the computation, note that in general, a Weyl structure σ allows us to identify a section
s ∈ Γ(A) with s ◦ σ ∈ C∞(G0, g)G0 ∼= ⊕ki=−kC∞(G0, gi)
G0 . We will write [s]σ = (sσ−k, . . . , s
σ
k) ∈
⊕ki=−kC∞(G0, gi)
G0 .
Now consider any Y ∈ TxM , and let Y0 be a local right-invariant vector field on G0 around
u0 ∈ (G0)x, projecting onto Y at x, and let Y be a local right-invariant vector field on G which
extends the vector field σ∗Y0. Then by the chain rule, we have Y(s)(σ(u′0)) = Y0([s]σ)(u′0)
for u′0 near u0 in G0. Now compute from the definition, for s = sX as above:(
∇AY s
)
(σ(u0)) = Y(s)(σ(u0)) + {ω ◦Y, s}(σ(u0)) = Y0([s]σ)(u0) + {σ∗ω ◦Y0, [s]
σ}(u0).
Now we translate the last line into vector notation, where the top, middle and bottom com-
ponents correspond, respectively, to the projection onto p+, g0 and g− (denoted, as usual, with
a subscript). Note that from Π(s) = X, and since X(x) = 0, we have sσ−(σ(u0)) = 0, and so we
get the following reformulation of the left-hand side of (8): Y0(sσ+)(u0)
Y0(sσ0 )(u0)
Y0(sσ−)(u0)
+
{σ∗ω(Y0), [s]σ}+(u0)
{ω0(Y0), sσ0}(u0) + {ω−(Y0), sσ+}0(u0)
{ω−(Y0), sσ0}(u0) + {ω−(Y0), sσ+}−(u0)
. (10)
On the other hand, let us compute the identity (9). The section sX0 ∈ C∞(G0, p
∗)G0 is given
by
sX0(u′0) = σ∗ω≤(X0(u′0)) = ω≤(σ(u′0))(σ∗(X0(u′0))).
Using the facts that σ is a section of π+ : G → G0, and that X projects onto X0 via π+, it follows
that X(σ(u′0))− σ∗(X0(u′0)) lies in the kernel of Tσ(u′0)π+. In particular, this means we have:
ω≤(σ(u′0))(σ∗(X0(u′0))) = ω≤(X(σ(u′0))),
or equivalently, sX0 = sσ− + sσ0 . Using this, a similar calculation to the one above gives:
(∇σY sX0)(u0) =
(
Y0(sσ0 )(u0)
Y0(sσ−)(u0)
)
+
(
{ω0(Y0), sσ0}(u0)
{ω−(Y0), sσ0}(u0)
)
. (11)
Comparing the g0-components of (10) and (11), we see that if both terms vanish, we must
have
{ω−(Y0), sσ+}0(u0) := prg0([ω−(Y0)(u0), sσ+(u0)]) = 0.
14 J. Alt
But we have g− = {ω−(Y0)(u0) |Y ∈ TxM}, and from this it follows that sσ+(u0) = 0, by using
the properties of |k|-graded semi-simple Lie algebras: We have the grading element E ∈ g0,
which satisfies [E, Yj ] = jYj for all Yj ∈ gj . Now using the Ad-invariance of the Killing form B,
we have, for any Y ∈ g−, and 0 < j ≤ k:
B([Y, sσj (u0)], E) = B(Y, [sσj (u0), E]) = −jB(Y, sσj (u0)),
and since B induces an isomorphism gj ∼= (g−j)
∗, this vanishes for all Y ∈ g− only if sσj (u0) = 0
for all j > 0, i.e. only if s(σ(u0)) ∈ g0, which is the claim of the proposition. �
Next we prove the converse to Proposition 4.3:
Proposition 4.4. Let X ∈ inf(M) have a singularity x ∈ M , and let u ∈ Gx, u0 = π+(u) ∈
(G0)x. If the holonomy htu is conjugate under P to a one-parameter subgroup of G0 (equivalently,
sX(u.p) = Ad(p−1)(sX̃(u)) ∈ g0 for some p ∈ P ), then X is an infinitesimal automorphism of
some local Weyl structure σ around x.
Proof. We will need the following “exponential coordinates” on G and M around u and x
respectively, which are induced by the Cartan connection ω: For any Y ∈ g, denote by Ŷ the
vector field on G which is determined by the identity, ω(Ŷ (u′)) = Y for all u′ ∈ G. Defining
Wu := {Y ∈ g |ϕŶ ,t(u) is defined for 0 ≤ t ≤ 1},
then there exist an open neighborhood Vu of 0 ∈ g and an open neighborhood Vu of u ∈ G such
that the exponential map expωu defined on Wu is a diffeomorphism of Vu onto Vu, where by
definition:
expωu : Y 7→ expω(u, Y ) := ϕŶ ,1(u).
Restricting Vu if necessary to a smaller neighborhood of zero, we get a diffeomorphism
expωu := π ◦ expωu : V−u
≈→ Ux,
where Ux is a neighborhood of x in M and V−u := Vu ∩ g−. Furthermore, for V −u := expωu(V−u ),
the restriction of the projection π gives a diffeomorphism of V −u onto Ux.
These exponential coordinates can obviously be used to define a local Weyl structure over Ux,
since they give a local section of π : G → M on Ux. This gives us the local trivialisations
π−1(Ux) ∼= V −u ×P and π−1
0 (Ux) ∼= π+(V −u )×G0. Then we simply define σ : π−1
0 (Ux)→ π−1(Ux)
by
σ : π+(u′).g0 7→ u′.g0; (12)
for any u′ ∈ V −u , g0 ∈ G0. This is by definition a G0-equivariant local section of π+ : G → G0,
that is a Weyl structure. We will now show, assuming htu ⊂ G0, that the local flows of X ∈ X(G)
commute with σ, i.e. we have ΦX,t ◦ σ = σ ◦ ΦX0,t on π−1
0 (Ux), for t sufficiently small so that
both sides exist. By Definition 2.1, this implies X ∈ inf(σ).
To show the commutativity of the local flows of X with the Weyl structure σ given by (12), we
need the following general equivariance relation for an infinitesimal automorphism in exponential
coordinates (cf. the proof of Proposition 4.2 of [10]):
ΦX,t(expω(u, Y )) = expω(u,Ad(htu)(Y )).htu. (13)
The identity (13) is based on the observation that we have [X, Ŷ ] = 0 for any infinitesimal
automorphism, and any Y ∈ g (this follows immediately from the defining equation, LXω = 0,
Essential Parabolic Structures and Their Infinitesimal Automorphisms 15
of an infinitesimal automorphism). Hence, the flows commute, ΦX,t◦ΦŶ ,s = ΦŶ ,s◦ΦX,t whenever
both sides are defined, which together with equivariance of ω may be used to show that both
sides of (13) are given as the endpoint of the same integral curve through ΦX,t(u) = u.htu.
Now consider an arbitrary point π+(u′).g0 ∈ π−1
0 (Ux), where u′ = expω(u, Y ) ∈ V −u for
Y ∈ V−u . Then σ(π+(u′).g0) := u′.g0, and we have, by P -equivariance of ΦX,t and (13):
(ΦX,t ◦ σ)(π+(u′).g0) = ΦX,t(u
′.g0) = expω(u,Ad(htu)(Y )).htug0.
But since htu ∈ G0, we have Ad(htu)(Y ) ∈ g− and for t sufficiently small we may also assume
Ad(htu)(Y ) ∈ Vu by continuity, so expω(u,Ad(htu)(Y )) ∈ V −u . We also have htug0 ∈ G0, so G0-
equivariance of π+ gives π+(ΦX,t(u
′.g0)) = π+(expω(u,Ad(htu)(Y ))).htug0, and hence combining
the above gives:
(ΦX,t ◦ σ)(π+(u′).g0) = ΦX,t(u
′.g0) = (σ ◦ π+ ◦ ΦX,t)(u
′, g0).
Finally, since ΦX0,t is defined on G0 via the relation (ΦX0,t ◦ π+) = (π+ ◦ΦX,t), this shows that
(ΦX,t ◦ σ) = (σ ◦ ΦX0,t) on π−1
0 (Ux). �
4.3 Proof of Theorem 1.2
There are two parts to establishing Theorem 1.2, and they are extensions of the arguments in
the proofs of Propositions 4.3 and 4.4.
First, we establish the following claim, which is a strengthening of Proposition 4.3: If X
is inessential in some neighborhood of x, then its holonomy htu is conjugate under P to a
one-parameter subgroup of Ker(λ) ⊂ G0 for a choice of scale representation λ : G0 → R+
(equivalently, sX(u.p) = Ad(p−1)(sX(u)) ∈ Ker(λ′) ⊂ g0 for some p ∈ P ).
The proof of this claim is analogous to the proof of Proposition 4.3. If σ is exact, then we have
the holonomy reduction G0 ⊂ G0 to structure group Ker(λ) ⊂ G0, and denoting the resulting
reduction by σ : G0 → G, the condition that X is an exact infinitesimal automorphism of σ is
(using Lemma 2.1), in addition to the requirements computed in the proof of Proposition 4.3,
that X0(u) ∈ TuG0 ⊂ TuG0 for all u ∈ G0 and that the resulting vector field X0 on G0 is an
infinitesimal automorphism of the Cartan connection σ∗ω≤. This gives(
∇σsX0
)
(u0) = 0,
where ∇σ denotes the tractor connection on the adjoint tractor bundle associated to G0 by
the adjoint representation on g− ⊕ Ker(λ′). Then by the same considerations, if we write
sσ0 (u0) = sσ0 (u0) + z(u0)Eλ, for u0 ∈ (G0)x, then this condition implies that
0 = [ω−(Y0(u0)), z(u0)Eλ] =
k∑
j=1
jz(u0)ω−j(Y0(u0)),
for all Y ∈TxM , which can only happen if z(u0) = 0. Hence we must have s(σ(u0))∈Ker(λ′)⊂g0.
Second, we make the following, additional claim under the setting of Proposition 4.4: If the
holonomy htu of X is conjugate under P to a one-parameter subgroup of Ker(λ) ⊂ G0 for a choice
of scale representation λ : G0 → R+ (equivalently, sX(u.p) = Ad(p−1)(sX(u)) ∈ Ker(λ′) ⊂ g0
for some p ∈ P ), then X is inessential on some neighborhood of x.
This claim is also proved in the same way as the claim of Proposition 4.4. The considerations
of that proof also show that the local, equivariant section σ of π+ : G → G0 over Ux which was
constructed using the exponential map, can be restricted to a map σ : (G0)Ux → GUx where the
sub-bundle (G0)Ux ⊂ (G0)Ux is just given by π+(V −u )×Ker(λ) in the local trivialization, giving
a locally exact Weyl structure. Finally, since htu is conjugate to a one-parameter subgroup of
16 J. Alt
Ker(λ), it can be arranged that the restriction of X0 to (G0)Ux is always tangent to this sub-
bundle, so X is locally inessential by Lemma 2.1: Namely, changing u ∈ Gx if necessary, we
may assume that htu ⊂ Ker(λ) and (equivalently) ω(X(u)) ∈ Ker(λ′). For any u′ ∈ V −u :=
expu(V−u ) ⊂ GUx , the identity (13) can be differentiated to show that
X(u′) =
d
dt |t=0
(
expω(u,Ad(htu)(Y )).htu
)
,
for some Y ∈ V−u . Since htu ⊂ Ker(λ), this shows that X(u′) is the sum of a vector tangent to V −u
with a vertical vector corresponding to an element of Ker(λ′) ⊂ g0. Thus, X0(π+(u′)) ∈ TG0
and hence this holds for restriction of X0 to (G0)|Ux , since any point in the latter space is given
by π+(u′).k for some u′ ∈ V −u and some k ∈ Ker(λ).
Acknowledgements
I am grateful to Charles Frances for discussions about the methods used in [8] and [10], and to
Felipe Leitner for useful comments on an earlier version of the text. The anonymous referees
made many helpful criticisms and suggestions; in particular, I am grateful for the suggestion
to reformulate an earlier (equivalent) version of Definition 2.1 in terms of the action of an
automorphism on the set of Weyl structures.
References
[1] Alekseevski D.V., Groups of conformal transformations of Riemannian spaces, Sb. Math. 18 (1972), 285–301.
[2] Biquard O., Métriques d’Einstein asymptotiquement symétriques, Astérisque (2000), no. 265 (English transl.:
Asymptotically symmetric Einstein metrics, SMF/AMS Texts and Monographs, Vol. 13, American Mathe-
matical Society, Providence, 2006).
[3] Čap A., Infinitesimal automorphisms and deformations of parabolic geometries, J. Eur. Math. Soc. 10
(2008), 415–437, math.DG/0508535.
[4] Čap A., Slovák J., Weyl structures for parabolic geometries, Math. Scand. 93 (2003), 53–90,
math.DG/0001166.
[5] Čap A., Slovák J., Parabolic geometries. I Background and general theory, Mathematical Surveys and Mono-
graphs, Vol. 154, American Mathematical Society, Providence, RI, 2009.
[6] Duistermaat J.J., Kolk J.A.C., Lie groups, Universitext, Springer-Verlag, Berlin, 2000.
[7] Ferrand J., The action of conformal transformations on a Riemannian manifold, Math. Ann. 304 (1996),
277–291.
[8] Frances C., Sur le groupe d’automorphismes des géométries paraboliques de rang un, Ann. Sci. École Norm.
Sup. (4) 40 (2007), 741–764 (English version: A Ferrand–Obata theorem for rank one parabolic geometries,
math.DG/0608537).
[9] Frances C., Essential conformal structures in Riemannian and Lorentzian geometry, in Recent Developments
in Pseudo-Riemannian Geometry, ESI Lectures in Mathematics and Physics, ESI Lect. Math. Phys., Eur.
Math. Soc., Zürich, 2008, 231–260.
[10] Frances C., Local dynamics of conformal vector fields, arXiv:0909.0044.
[11] Gover A.R., Laplacian operators and Q-curvature of conformally Einstein manifolds, Math. Ann. 336 (2006),
311–334, math.DG/0506037.
[12] Ivanov S., Vassilev D., Conformal quaternionic contact curvature and the local sphere theorem, J. Math.
Pures Appl. (9) 93 (2010), 277–307, arXiv:0707.1289.
[13] Schoen R., On the conformal and CR automorphism groups, Geom. Funct. Anal. 5 (1995), 464–481.
[14] Webster S.M., On the transformation group of a real hypersurface, Trans. Amer. Math. Soc. 231 (1977),
179–190.
http://dx.doi.org/10.1070/SM1972v018n02ABEH001770
http://dx.doi.org/10.4171/JEMS/116
http://arxiv.org/abs/math.DG/0508535
http://arxiv.org/abs/math.DG/0001166
http://dx.doi.org/10.1007/BF01446294
http://dx.doi.org/10.1016/j.ansens.2007.07.003
http://dx.doi.org/10.1016/j.ansens.2007.07.003
http://arxiv.org/abs/math.DG/0608537
http://arxiv.org/abs/0909.0044
http://dx.doi.org/10.1007/s00208-006-0004-z
http://arxiv.org/abs/math.DG/0506037
http://dx.doi.org/10.1016/j.matpur.2009.11.002
http://dx.doi.org/10.1016/j.matpur.2009.11.002
http://arxiv.org/abs/0707.1289
http://dx.doi.org/10.1007/BF01895676
http://dx.doi.org/10.2307/1997877
1 Introduction
1.1 Motivation from conformal geometry
1.2 Organization of the text and summary of main results
2 Essential automorphisms: basic definitions and properties
2.1 Background on parabolic geometries and their Weyl structures
2.2 Definition and basic properties of essential structures
3 Lichnérowicz theorem for rank one parabolic geometries
4 Proof of local results
4.1 Background results on infinitesimal automorphisms
4.2 Holonomy and essentiality of infinitesimal automorphisms
4.3 Proof of Theorem 1.2
References
|