ArticlePDF Available

Motion of hypersurfaces by Gauss curvature

Authors:

Abstract

We consider n-dimensional convex Euclidean hypersurfaces moving with normal velocity proportional to a positive power α of the Gauss curvature. We prove that hypersurfaces contract to points in finite time, and for α ∈ (1/(n +2], 1/n] we also prove that in the limit the solutions evolve purely by homothetic contraction to the final point. We prove existence and uniqueness of solutions for non-smooth initial hypersurfaces, and develop upper and lower bounds on the speed and the curvature independent of initial conditions. Applications are given to the flow by affine normal and to the existence of non-spherical homothetically contracting solutions.
PACIFIC JOURNAL OF MATHEMATICS
Volume 195 No. 1 September 2000
Pacific Journal of Mathematics 2000 Vol. 195, No. 1
Pacific
Journal of
Mathematics
Volume 195 No. 1 September 2000
PACIFIC JOURNAL OF MATHEMATICS
http://www.pjmath.org
Founded in 1951 by
E. F. Beckenbach (1906–1982) F. Wolf (1904–1989)
EDITORS
Vyjayanthi Chari
Department of Mathematics
University of California
Riverside, CA 92521-0135
chari@math.ucr.edu
Robert Finn
Department of Mathematics
Stanford University
Stanford, CA 94305-2125
finn@math.stanford.edu
Kefeng Liu
Department of Mathematics
University of California
Los Angeles, CA 90095-1555
liu@math.ucla.edu
V. S. Varadarajan (Managing Editor)
Department of Mathematics
University of California
Los Angeles, CA 90095-1555
pacific@math.ucla.edu
Darren Long
Department of Mathematics
University of California
Santa Barbara, CA 93106-3080
long@math.ucsb.edu
Jiang-Hua Lu
Department of Mathematics
The University of Hong Kong
Pokfulam Rd., Hong Kong
jhlu@maths.hku.hk
Sorin Popa
Department of Mathematics
University of California
Los Angeles, CA 90095-1555
popa@math.ucla.edu
Sorin Popa
Department of Mathematics
University of California
Los Angeles, CA 90095-1555
popa@math.ucla.edu
Jie Qing
Department of Mathematics
University of California
Santa Cruz, CA 95064
qing@cats.ucsc.edu
Jonathan Rogawski
Department of Mathematics
University of California
Los Angeles, CA 90095-1555
jonr@math.ucla.edu
PRODUCTION
pacific@math.berkeley.edu
Paulo Ney de Souza, Production Manager Silvio Levy, Senior Production Editor Nicholas Jackson, Production Editor
SUPPORTING INSTITUTIONS
ACADEMIA SINICA, TAIPEI
CALIFORNIA INST. OF TECHNOLOGY
CHINESE UNIV. OF HONG KONG
INST. DE MATEMÁTICA PURA E APLICADA
KEIO UNIVERSITY
MATH. SCIENCES RESEARCH INSTITUTE
NEW MEXICO STATE UNIV.
OREGON STATE UNIV.
PEKING UNIVERSITY
STANFORD UNIVERSITY
UNIVERSIDAD DE LOS ANDES
UNIV. OF ARIZONA
UNIV. OF BRITISH COLUMBIA
UNIV. OF CALIFORNIA, BERKELEY
UNIV. OF CALIFORNIA, DAVIS
UNIV. OF CALIFORNIA, IRVINE
UNIV. OF CALIFORNIA, LOS ANGELES
UNIV. OF CALIFORNIA, RIVERSIDE
UNIV. OF CALIFORNIA, SAN DIEGO
UNIV. OF CALIF., SANTA BARBARA
UNIV. OF CALIF., SANTA CR UZ
UNIV. OF HAWAII
UNIV. OF MONTANA
UNIV. OF NEVAD A, RENO
UNIV. OF OREGON
UNIV. OF SOUTHERN CALIFORNIA
UNIV. OF UTAH
UNIV. OF WASHINGTON
WASHINGTON STATE UNIVERSITY
These supporting institutions contribute to the cost of publication of this Journal, but they are not owners or publishers and have no respon-
sibility for its contents or policies.
See inside back cover or www.pjmath.org for submission instructions.
Regular subscription rate for 2006: $425.00 a year (10 issues). Special rate: $212.50 a year to individual members of supporting institutions.
Subscriptions, requests for back issues from the last three years and changes of subscribers address should be sent to Pacific Journal of
Mathematics, P.O. Box 4163, Berkeley, CA 94704-0163, U.S.A. Prior back issues are obtainable from Periodicals Service Company, 11
Main Street, Germantown, NY 12526-5635. The Pacific Journal of Mathematics is indexed by Mathematical Reviews, Zentralblatt MATH,
PASCAL CNRS Index, Referativnyi Zhurnal, Current Mathematical Publications and the Science Citation Index.
The Pacific Journal of Mathematics (ISSN 0030-8730) at the University of California, c/o Department of Mathematics, 969 Evans Hall,
Berkeley, CA 94720-3840 is published monthly except July and August. Periodical rate postage paid at Berkeley, CA 94704, and additional
mailing offices. POSTMASTER: send address changes to Pacific Journal of Mathematics, P.O. Box 4163, Berkeley, CA 94704-0163.
PUBLISHED BY PACIFIC JOURNAL OF MATHEMATICS
at the University of California, Berkeley 94720-3840
A NON-PROFIT CORPORATION
Typeset in L
A
T
E
X
Copyright ©2006 by Pacific Journal of Mathematics
PACIFIC JOURNAL OF MATHEMATICS
Vol. 195, No. 1, 2000
MOTION OF HYPERSURFACES BY GAUSS CURVATURE
Ben Andrews
We consider n-dimensional convex Euclidean hypersurfaces
moving with normal velocity proportional to a positive power
α of the Gauss curvature. We prove that hypersurfaces con-
tract to points in finite time, and for α (1/(n + 2], 1/n] we
also prove that in the limit the solutions evolve purely by ho-
mothetic contraction to the final point. We prove existence
and uniqueness of solutions for non-smooth initial hypersur-
faces, and develop upper and lower bounds on the speed and
the curvature independent of i niti al conditions. Applications
are given to the flow by affine normal and to the exi stence of
non-spherical homothetically contracting solutions.
1. Introduction.
Motivation for the study of hypersurfaces moving by their Gauss curvature
comes from several sources:
1.1. Tumbling stones. W.J. Firey introduced the Gauss curvature flow
in 1974, as a model of the wearing process undergone by a p ebble on a beach
[Fi]. Consider a stone which occupies an open, bounded convex region of
R
n+1
at time t = 0. The stone tumbles, and collides with a hyperplane (the
beach) with random orientation. We assume for simplicity that the amount
of material removed in a collision at a point x of the stone depends only on
the normal direction ν
x
(thus allowing some anisotropy in the material of
the stone). The number of collisions with a region B of the surface of the
stone is proportional to the measure of the set ν(B) = {ν
x
: x B} S
n
of normal directions to B. This is equal to
R
B
K
x
dH
n
(x), where K
x
is the
Gauss curvature of the hypersurface at x. The rate at which the stone wears
away at a point x is given by ρ(ν
x
)K
x
for some positive function ρ on S
n
,
and we have the evolution equation
(1) ˙x = ρ(ν
x
)K
x
ν
x
.
1.2. Affine geometry: Inner parallel surfaces and the affine normal
flow. K. Leichtweiss introduced the notion of inner parallel surfaces for a
convex body in affine geometry in the paper [Le]. Given a convex region Ω,
the idea is to construct a family of related regions P
t
Ω, by a procedure which
is well-defined in the setting of affine geometry that is, if we perform
1
2 BEN ANDREWS
an area-preserving affine transformation L to get a new region LΩ, then
P
t
(LΩ) = L(P
t
Ω).
The procedure is as follows: For each direction z S
n
, there exists a
unique supporting hyperplane H
z
= {hz, yi = h(z)} to with normal di-
rection z pointing outward from Ω. The notion of parallel hyperplanes is
well-defined in affine geometry, as is the notion of the volume of a region.
Hence we can choose a unique hyperplane H
z,t
parallel to H
z
such that the
volume of the part of between H
z,t
and H
z
is equal to t
(n+2)/2
(in the case
where is the region above a paraboloid, this choice of exponent ensures
that H
z,t
moves at constant speed). Equivalently, we can define a function
h
t
(z) by the requirement
(2) Vol ({y : h
t
(z) hy, zi h(z)}) = t
(n+2)/2
.
Then we define P
t
to be the convex set
(3)
\
zS
n
{y R
n+1
: hz, yi h
t
(z)}.
In contrast to the corresponding situation in Euclidean geometry, this
procedure does not define a semi-group: If we begin with a region Ω, con-
struct the regions P
t
Ω, and use them to construct the regions P
τ
P
t
Ω, then
these are not in general given by P
t
0
for any t
0
. To remedy this we consider
P
n
t/n
Ω, obtained by following the above construction repeatedly over small
intervals, and take the limit n of infinitesimally small steps to obtain
a region
˜
P
t
Ω. This defines a deformation which is clearly well-defined in
affine geometry, and satisfies the semi-group property
˜
P
t
˜
P
τ
=
˜
P
t+τ
.
To find an explicit description of this deformation in the case where
is smooth and strictly convex, we consider the regions P
t
in the limit of
small t: Fix z, and choose coordinates for R
n+1
such that e
1
, . . . , e
n
span the
supporting hyperplane H
z
of Ω, and the supporting point is at the origin.
Then M = is locally a graph in these coordinates:
x
n+1
=
1
2
n
X
i,j=1
h
ij
x
i
x
j
+ O(|x|
3
),
where h
ij
is the second fundamental form at the supporting point. There ex-
ists a volume-preserving linear transformation which fixes the e
n+1
direction
and brings M locally to the form
x
n+1
=
1
2
K
1/n
n
X
i=1
x
2
i
+ O(|x|
3
)
where K = det h
ij
is the Gauss curvature at the supporting point. Then
Vol ({y : h(z) d hy, zi h(z)})
= 2
n/2
ω
n
K
1/2
d
(n+2)/2
+ O(d
(n+3)/2
)
MOTION OF HYPERSURFACES BY GAUSS CURVATURE 3
where ω
n
is the volume of the unit ball in R
n
, and hence the requirement
(2) implies that
h
t
(z) = h(z)
K
1
n+2
2
n
n+2
ω
2
n+2
n
t + O(t
3/2
).
It follows that the limiting deformation is given by the equation
(4) ˙x = c
n
K
1
n+2
x
ν
x
.
This evolution equation is the simplest invariant flow in affine differential
geometry; up to reparametrisation it is the motion of a hypersurface in the
direction of its affi ne normal vector. This has been considered in [ST1, ST2]
for the case of convex curves in the plane, and in [A4] more generally. For
nonconvex curves results were recently obtained in [AST].
1.3. Image analysis. Many fundamental problems in image analysis have
been approached using geometric flows: An image represented by a grey-
scale density function u can be processed to remove noise by smoothing the
level sets of u with a parabolic flow. Various candidates have been con-
sidered, but in [AGLM] axioms were proposed which included the natural
requirement of affine invariance. This leads to the evolution Equation (4).
In the case of nonconvex hypersurfaces this is no longer parabolic, and var-
ious authors (see [AGLM], [CS], [NK]) have considered the generalization
(for two-dimensional surfaces)
˙x = (sgnH) max{K, 0}
1/4
ν
where H is the mean curvature. Applications of plane curve evolution equa-
tions to image analysis and computer vision are described in [AGLM],
[OST], and [ST1]-[ST3].
1.4. Gradient flows of the mean width. The width of a convex region
in a direction z S
n
is defined by
w(z) = sup
y
1
,y
2
hy
1
y
2
, zi = h(z) + h(z)
where h(z) = sup
y
hy, zi is the support function of Ω. The mean width
V
1
(Ω, ϕ) with resp e ct to a measure ϕdµ on S
n
is given (up to a constant
factor) by integrating the width over all directions z S
n
:
V
1
(Ω) =
Z
S
n
w(z)ϕ(z)(z) =
Z
S
n
h(z)(ϕ(z) + ϕ(z))(z)
=
Z
Kh ˜ϕ dH
n
,
4 BEN ANDREWS
where ˜ϕ(z) = ϕ(z) + ϕ(z). The first variation formula for the mean width
can be calculated as follows: Consider a smooth family
t
of convex regions,
with support functions h
t
(z) such that
t
h
t
(z)
t=0
= f(z),
then
d
dt
V
1
(Ω
t
)
t=0
=
Z
S
n
˜ϕ(z)f(z) =
Z
0
Kf ˜ϕ dH
n
.
We consider the flow of steepest descent of the mean width in L
p
spaces
on that is, we seek that variation f for which V
1
(Ω
t
, ϕ) decreases
fastest amongst all variations with the same L
p
norm
R
|f|
p
σ(ν) dH
n
1/p
(σ is a positive smooth function on S
n
): By the older inequality we have
for p > 1
Z
0
Kf ˜ϕdH
n
=
Z
0
Kf
˜ϕ
σ
σdH
n
Z
0
|f|
p
σdH
n
1/p
Z
0
K ˜ϕ
σ
p
p1
σdH
n
!
11/p
with equality if and only if f = c (K ˜ϕ/σ)
1/(p1)
. The flow of steepes t descent
is therefore
(5) ˙x = ρ(ν
x
)K
1/(p1)
x
ν
x
where ρ = ( ˜ϕ/σ)
1/(p1)
is a smooth positive function on S
n
.
1.5. Evolving hypersurfaces and degenerate fully nonlinear PDE.
The evolution equations derived above are included in a large class of par-
abolic evolution equations for hypersurfaces which have been considered
before. Simplest in this class is the mean curvature flow, in which a hy-
persurface moves in the direction of its inward normal with speed given
by the mean curvature. Huisken [Hu] showed that convex hypersurfaces
moving under such equations contract to points in finite time, and that the
hypersurfaces become s pherical in shape in the process . This argument has
since been extended to many processes where convex hypersurfaces move
with speeds given by homogeneous degree one, concave or convex monotone
symmetric functions of the principal curvatures: Chow considered flows by
the nth root of the Gauss curvature [Ch1] and the square root of the scalar
curvature [Ch2], and the author has considered a general class of s uch evo-
lution equations [A1]. Corresponding results for flows where the speed has
other positive degrees of homogeneity in the curvature seem much harder to
prove. The author has treated the special case of flow by the power 1/(n+2)
of the Gauss curvature, which is the flow by affine normal [A4]. Tso [Ts]
MOTION OF HYPERSURFACES BY GAUSS CURVATURE 5
and Chow [Ch1] have shown that hypersurfaces moving with speed equal to
any positive power of the Gauss curvature contract to points in finite time.
The Gauss curvature flows form a convenient class of examples of par-
abolic equations with varying degeneracy: For large α they become more
degenerate, and for small α they become singularly parabolic. Intermediate
values of α are singular in some situations and degenerate in others. The
precise effect of such degeneracy or singularity on the regularity of solutions
is extremely complicated. In particular, it would be interesting to know
how irregular solutions can be, how the regularity estimates depend on time
(particularly where the initial solution is highly irregular), and how solutions
behave in the neighbourhood of degenerate or singular regions.
There are several other important families of PDE for which similar ques-
tions can be asked in particular, natural families of parabolic equations
with varying degeneracy include the porous medium equations
˙u = ∆(|u|
m1
u),
and the p-harmonic heat flows
˙u = ·
|∇u|
p2
u
,
for which there is also a natural generalization to p-harmonic maps between
Riemannian manifolds. The Gauss curvature flows can be considered a
geometric analogue of the porous medium equations.
In the case of curves in the plane, more complete results are known:
Gage [Ga1]-[Ga2] and Hamilton [GH] showed that convex curves contract
to points in finite time and become round under the curve shortening flow
(where the speed of motion equals the curvature), and Grayson [Gr] ex-
tended this by showing that any embedded curve eventually becomes convex.
This was extended to include anisotropic analogues of the curve-shortening
flow by Gage [Ga3] in the convex case, and by Oaks [Oa] for nonconvex
curves. The author considered equations of varying degeneracy in the con-
vex case [A2], [A8], and obtained optimal estimates on the regularity of
solutions, including their initial behaviour. The particular case of the affine
normal flow has also been extended to nonconvex curves [AST].
2. The result.
Our main aim in this paper is to prove results about the regularity and
limiting behaviour of solutions of the Gauss curvature flows of the form
(6)
dx
dt
= ρ(ν(x))K(x)
α
ν(x),
where α is in the range (1/(n + 2), 1/n]. These particular exponents arise
as follows in the proof: We first prove (in Section 4 of the paper) that the
solutions of Gauss curvature flows have isoperimetric ratio bounded as long
6 BEN ANDREWS
as they exist, provided α is greater than the critical value 1/(n + 2). This
value is sharp the flow with ρ 1 and α = 1/(n + 2) is the affine normal
flow, for which solutions converge to ellipsoids of arbitrary eccentricity [A4],
and so the isoperimetric ratio tends to stay bounded but does not generally
improve; in a separate paper [A9] we prove that for exponents smaller than
this (or equal to this if ρ is nonconstant) there are solutions which have
isoperimetric ratios approaching infinity. Second, we prove (in Sections 5
and 6) that if the hypersurface has bounded isoperimetric ratio, and α
1/n, then a short time later the moving hypersurfaces are strictly convex
and have bounded curvature. The exponent 1/n is again sharp, as there are
solutions of the Gauss curvature flow for any α > 1/n which remain non-
strictly convex and are not C
in fact any initial convex hypersurface
which includes a planar piece will behave this way. This phenomenon was
first noted by Richard Hamilton for the case α = 1 [Ha1]. We describe such
behaviour more fully in Section 12 of this paper.
By combining these results, we obtain the following:
Theorem 1. For any open bounded convex region
0
, any smooth positive
function ρ : S
n
R, and any α (1/(n + 2), 1/n], there exists a family of
embeddings x : S
n
× [0, T ) R
n+1
satisfying (6), unique up to composition
with an arbitrary time-independent diffeomorphism, such that M
t
= x(S
n
, t)
converges in Hausdorff distance to the boundary of the region
0
as t ap-
proaches zero. x is smooth and strictly convex for t > 0 and converges to a
point p R
n+1
as t approaches T . Furthermore, the hypersurfaces
˜
M
t
=
Vol(S
n
)
Vol(M
t
)
1/(n+1)
(M
t
p)
converge in C
as t approaches T , to a smooth, strictly convex limit hyper-
surface
˜
M
T
for which hx, νi = (ν)K
α
for some c > 0.
This m eans that the evolving hypersurface contracts to a point, and
asymptotically approaches a solution which evolves purely by homotheti-
cally scaling about this limiting point.
We also have the following generalisation for smaller α:
Theorem 2. The result of Theorem 1 also holds for a solution of (6) with
any α (0, 1/n], provided the isoperimetric ratios of the evolving hypersur-
faces remain bounded.
We deduce in Section 7 the first part of Theorem 1, that solutions contract
to points in finite time (in fact we prove this for any α > 0). This was proved
previously for isotropic cases by Tso [Ts] and Chow [Ch1]. In Section 8 we
prove that s olutions exist starting from singular initial hypersurfaces, and
immediately become smooth and strictly convex if α 1/n note that we
make no regularity assumptions about the initial hypersurface, other than
MOTION OF HYPERSURFACES BY GAUSS CURVATURE 7
those implied by its convexity. We also prove in Theorem 15 the existence of
unique viscosity solutions for α > 1/n, although this is not required for the
proof of Theorems 1 and 2. In Section 9 we digress from the main argument
of the pap er to apply the regularity and convexity estimates in a simple
new proof of the convergence theorem for the affine normal flow. In Section
11 of the paper we use Theorem 2 to deduce the existence of non-spherical
homothetic solutions of Equation (6) for constant ρ and suitable α between
0 and 1/(n + 2).
We remark that Urbas [U2] has considered noncompact solutions of iso-
tropic equations of the form (6), in particular proving the e xistence of solu-
tions which evolve by homothetically expanding or translating.
3. Notation and preliminaries.
The inradius r
of an open convex region is the supremum of the radii
of all balls contained in it, and the circumradius r
+
is the infinum of the
radii of balls containing it. In this paper we refer to the ratio r
/r
+
as the
isoperimetric ratio of the body.
For a convex region with boundary given by a smooth embedding x : M
R
n+1
, we have an outward unit normal vector field ν : M S
n
R
n+1
,
which we use to define the Weingarten map W
x
: T
x
M T
x
M by the
formula
W(u) = D
u
ν T
ν(x)
S
n
' T
x
M
for any x M and u T
x
M. The eigenvalues λ
1
, . . . , λ
n
of W(x) are the
principal curvatures of M at x. The elementary symmetric functions E
j
of
these are defined by
(7) E
j
=
1
n
k
X
1i
1
<i
2
<···<i
j
n
j
Y
k=1
λ
i
k
!
.
In particular, K = E
n
is the Gauss curvature, and H = E
1
is the mean
curvature.
The covariant derivative on the hypersurface is given by the formula
u
v = D
u
v + hW(u), viν.
We will find it convenient at some points in this paper to describe an
open convex region R
n+1
in terms of its support function h : S
n
R,
defined by
(8) h(z) = sup
x
hx, zi.
The support function completely describes the region in particular,
can be recovered from h via the expression
=
zS
n
y R
n+1
: hz, yi < h(z)
.
8 BEN ANDREWS
Alternatively, in the case where M is strictly convex and smooth, the support
function can be used to define a canonical embedding ¯x of S
n
with image
equal to M:
(9) ¯x(z) = h(z)z +
¯
i
h(z)¯g
ij
¯
j
z.
This has the property that the outward normalto M at the point ¯x(z) is
equal to z, for each z S
n
.
The Weingarten map can also be recovered directly from h:
hW
1
(u), vi =
¯
u
¯
v
h + hu, vih
where
¯
is the covariant derivative on S
n
, and we identify T
x
M and T
ν(x)
S
n
.
For convenience we will denote by r
ij
the corresponding symmetric bilinear
form, the eigenvalues of which are the principal radii of curvature r
i
= λ
1
i
,
i = 1, . . . , n:
(10) r
ij
=
¯
i
¯
j
h + ¯g
ij
h.
r
ij
satisfies a Codazzi-type identity:
¯
k
r
ij
=
¯
k
¯
i
¯
j
h + ¯g
ij
¯
k
h(11)
=
¯
i
¯
k
¯
j
h +
¯g
kj
¯
i
h ¯g
ij
¯
k
h
+ ¯g
ij
¯
k
h
=
¯
i
r
kj
.
Differentiating (11), commuting derivatives, and applying (11) again to the
result, we obtain a version of the Simons’ identity for the second derivatives
of the second fundamental form [Si]:
(12)
¯
(i
¯
j)
r
kl
=
¯
(k
¯
l)
r
ij
+ ¯g
ij
r
kl
¯g
kl
r
ij
where the brackets denote symmetrisation.
For convenience, we will denote by S
k
the kth eleme ntary symmetric
function of the eigenvalues of r
ij
. In particular, S
n
= K
1
. S
k
may be
considered as a function of the components of the matrix r
ij
, and we denote
by
˙
S
ij
k
and
¨
S
ijpq
k
the first and second derivatives:
˙
S
ij
k
=
S
k
r
ij
and
¨
S
ijpq
k
=
2
S
k
r
ij
r
pq
.
˙
S
k
is a positive definite symmetric bilinear form provided r is positive defi-
nite, and S
1/k
k
is a concave function of the components of r for k = 1, . . . , n
[Mi].
MOTION OF HYPERSURFACES BY GAUSS CURVATURE 9
The support function allows the degenerate parabolic system of equation
(6) to be re-written as a parabolic scalar equation (see [U1], [A1], [A4]):
(13)
dh(z)
dt
= ρ(z) det
¯
2
h + Idh
α
.
In particular, this implies the existence of a smooth solution of Equation (6)
for a short time for any smooth, strictly convex initial hypersurface.
In a region where a family of hypersurfaces moving under Equation (6) can
be represented as graphs x
n+1
= u
t
(x
1
, . . . , x
n
) for some convex functions u
t
,
we can work with an equivalent scalar parab olic equation for the functions
u
t
:
(14)
t
u = ˜ρ(Du)
det D
2
u
α
(1 + |Du|
2
)
α(n+2)1
2
where
˜ρ(Du) = ρ
P
n
i=1
e
i
D
i
u e
n+1
p
1 + |Du|
2
!
.
We note some elementary features of Eq. (6): First, the speed of motion
is given by a homogeneous function of the curvatures, and this homogeneity
leads to a scaling property of solutions. Spe cifically, if x : M ×[0, T ] R
n+1
satisfies Eq. (6), then for each λ > 0 another solution x
λ
: M ×[0, λ
1+
T ]
R
n+1
is given by
(15) x
λ
(p, t) = λx(p, λ
(1+)
t).
This also implies corresponding scaling invariance properties for the solu-
tions of Equations (13) and (14).
A second important property of solutions of Eq. (6) is the comparison
principle: If {M
(i)
t
}, i = 1, 2 are two families of smooth, strictly convex hy-
persurfaces moving under Eq. (6), and M
(1)
0
M
(2)
0
= , then M
(1)
t
M
(2)
t
=
for all t > 0 in the common interval of existence. A local version also holds:
If {M
(i)
t
}, i = 1, 2 are families of smooth, strictly convex hypersurfaces with
boundary, M
(1)
0
M
(2)
0
= , and M
(1)
t
M
t
(2) = M
(2)
t
M
t
(1) = for
t [0, T ], then M
(1)
t
M
(2)
t
= for t [0, T ].
4. Monotone quantities and diameter bounds.
In this section we prove that whenever α > 1/(n + 2), the evolving hyper-
surfaces have bounded isoperimetric ratios for as long as the solution exists.
The main to ol used here is an integral estimate known as the entropy esti-
mate, which was proved for the case α = 1 by Chow [Ch1], and for other α
by the author [A3].
10 BEN ANDREWS
We define an integral quantity Z
ρ,α
for any given α and ρ by
Z
ρ,α
= Vol(M)
n/(n+1)
1
R
S
n
ρ
Z
M
ρK
α
dH
n
1/(α1)
if α 6= 1, and
Z
ρ,1
= Vol(M)
n/(n+1)
exp
1
R
S
n
ρ
Z
M
ρK log K dH
n
if α = 1. For convenience, we also denote by Z
\,α
the same quantity with
ρ 1.
Theorem 3. For any smooth, strictly convex solution of Equation (6),
d
dt
Z
ρ,α
0
with equality if and only if the equation hx, νi = (ν)K
α
holds for some
c > 0 and some choice of origin in R
n+1
.
This integral bound will be combined with the following estimate to de-
duce isoperimetric ratio bounds for solutions of the flow:
Theorem 4. For any smooth, strictly convex hypersurface M
n
in R
n+1
,
r
+
(M)
r
(M)
C(α, ρ)Z
(n+1)β(α)
ρ,α
for some positive constant β(α), provided α > 1/(n + 2).
Proof. We begin with a bound in terms of Z
\,α
:
Following [Ha2], we begin by obtaining a lower bound on the n-dimen-
sional areas of projections of M onto hyperplanes: Given a direction z
0
S
n
,
the area of the projection on to the plane with normal z
0
is given by
A
z
0
=
1
2
Z
S
n
|hz, z
0
i|S
n
dµ.
We apply the older inequality to bound this from below, as follows:
A
z
0
=
1
2
Z
S
n
|hz, z
0
i|S
n
1
2
Z
S
n
S
1α
n
1/(1α)
Z
S
n
|hz, z
0
i|
11
α/(α1)
provided α (0, 1). The integral
R
S
n
|hz, z
0
i|
β
is bounded for β > 1.
Hence for α (1/2, 1) we have
A
z
0
C
Z
S
n
S
1α
n
1/(1α)
= CV
n/(n+1)
Z
1
\,α
.
For α 1 this inequality still holds, because by the older inequality
Z
\,α
is increasing in α.
MOTION OF HYPERSURFACES BY GAUSS CURVATURE 11
Finally, we consider the case where α (1/(n + 2), 1/2]: The older
inequality gives
Z
\,α
Z
(n+1)(3/4α)
4(1α)(3n+2)
\,1/(n+2)
Z
α1/(n+2)
(1α)(34/(n+2))
\,3/4
.
The affine isoperimetric inequality (see [B], §26 and §73, [Sa] or [A4], The-
orem 7.1) implies that Z
\,1/(n+2)
1. Hence by applying the bound on A
z
0
in terms of Z
\,3/4
, which we know from the cases treated above, we have
A
z
0
CV
n/(n+1)
Z
1
\,3/4
CV
n/(n+1)
Z
(1α)(34/(n+2))
α1/(n+2)
\,α
.
Hence for each α (1/(n+2), ) we have A
z
0
CZ
β(α)
\,α
for some constant
C and some positive exponent β(α).
Next we deduce a bound on the maximum width of M (the largest dis-
tance between parallel supporting hyperp lanes ): Let z
0
be the normal di-
rection of a pair of parallel supporting hyperplanes for M at maximal sep-
aration. T hen the points of contact of M with these two planes are joined
by a segment with length equal to the maximum width of M, and which
is entirely contained in M. Choosing the origin to be at the centre of this
segment, we have h(z
0
) = h(z
0
) = w
+
/2 where w
+
is the maximum width
of M, and h(z) |hz, z
0
i|h(z
0
) for all z S
n
. But then the enclos ed volume
of M is computed by:
Vol(M) =
1
n + 1
Z
S
n
hS
n
1
n + 1
Z
S
n
h(z
0
)|hz, z
0
i|S
n
=
w
+
2(n + 1)
A
z
0
.
Hence w
+
2(n + 1)Vol(M)/A
z
0
CZ
β(α)
\,α
Vol(M)
1/(n+1)
.
Note that Vol(M) w
w
n
+
where w
is the minimum width of M, since
M in contained between (n + 1) pairs of parallel planes in any set of or-
thonormal directions; and, in particular, in the case where one of the pairs
of planes is at minimal separation. Then the separation of all the other pairs
is bounded by w
+
. It follows that w
V w
n
+
CZ
(α)
\,α
Vol(M)
1/(n+1)
.
This gives a bound on the ratio of the minimum and maximum widths of
the hypersurface, and this is sufficient to bound the isoperimetric ratio (see
for example [A1], Lemma 5.4).
Finally, we consider the anisotropic cases ρ 6= const.: For α < 1, Z
ρ,α
is
comparable to Z
\,α
:
inf
S
n
ρ
1/(1α)
R
S
n
ρ
Z
\,α
Z
ρ,α
sup
S
n
ρ
1/(1α)
R
S
n
ρ
Z
\,α
.
For α 1 the desired inequality res ults from the monotonicity of Z
ρ,α
as a
function of α, a consequence of the older inequality.
12 BEN ANDREWS
We have shown in particular that for α > 1/(n +2), any solution of Equa-
tion (6) with smooth, strictly convex initial data has uniformly bounded
isoperimetric ratio on the entire interval of its existence.
5. Displacement and speed bounds.
In this section we prove that the ratio of the maximum and minimum values
of the speed remains uniformly bounded for as long as the solution exists.
We will first deduce upper bounds on the displacement of the hypersur-
faces, by using spheres enclosed within M
0
as barriers:
Theorem 5. For any α > 0 and smooth positive ρ, and any smooth, strictly
convex solution {M
t
}
t>0
of Eq. (13),
h(z, t) h(z, 0) C
r
+
(M
0
)
r
(M
0
)
t
1
1+
for all t (0, C
0
r
(M
0
)
1+
] in the interval of existence of the solution,
where C and C
0
depend only on α and ρ.
Proof. Choose the origin at the centre of a ball of radius r
(M
0
) enclosed
by M
0
. Fix z S
n
, and define for each ε (0, 1] a sphere
S
ε
= {y R
n+1
: |y (1 ε)¯x(z)| = εr
(M
0
)}.
Then S
ε
is contained in the convex hull of ¯x(z) and B
r
(M
0
)
(0), so by con-
vexity is contained in M
0
.
Any family of spheres of the form S
r(t)
(p) with p R
n+1
and
r(t) =
r(0)
1+
sup ρ(1 + )t
1/(1+)
satisfies h˙x, νi ρK
α
, and hence act as barriers for solutions of Eq. (6),
by the comparison principle.
This gives an estimate on the support function of M
t
in direction z: S
ε
produces a barrier which shrinks to its centre at time ε
1+
r
(M)
1+
/((1+
) sup ρ), and we have
h(z, t) h(z, 0)
((1 + ) sup ρt)
1
1+
r
(M)
h(z, 0) C(α, ρ)
r
+
(M
0
)
r
(M
0
)
t
1
1+
.
Our next estimate is a speed bound, which we prove using the maximum
principle applied to the evolution equations for the spee d and the support
function. The proof is related to that given in [Ts] for the isotropic case of
Eq. (1).
MOTION OF HYPERSURFACES BY GAUSS CURVATURE 13
Theorem 6. For any smooth, strictly convex solution {M
t
}
[0,T ]
of Eq. (13)
with R
r
(M
t
) r
+
(M
t
) R
+
for t [0, T ],
ρ(z)S
α
n
C(n, α, ρ)
R
+
R
+
R
t
1+
!
.
Proof. From the definition (10) and the evolution Equation (13) we obtain
(16)
t
r
ij
=
¯
i
¯
j
(ρS
α
n
) + ¯g
ij
ρS
α
n
.
Since S
n
= det r
ij
, this implies
(17)
t
ρS
α
n
= αρS
(1+α)
n
˙
S
kl
n
¯
k
¯
l
(ρS
α
n
) + ¯g
kl
ρS
α
n
.
We also have
t
h(z, t) = ρS
α
n
(18)
= αρS
(1+α)
n
˙
S
kl
n
¯
k
¯
l
h + ¯g
kl
h
(1 + )ρS
α
n
.
Combining Eqs. (17) and (18), we obtain for q =
ρS
α
n
hR
/2
t
q = αρS
(1+α)
n
˙
S
kl
n
¯
k
¯
l
q +
2αρS
(1+α)
n
˙
S
kl
n
h R
/2
¯
k
h
¯
l
q
q
2
(αR
H/2 (1 + ))
where H =
P
n
i=1
r
1
i
= nS
n1
/S
n
nS
1/n
n
. By the maximum principle,
this implies the following inequality for Q = sup
S
n
q:
dQ
dt
Q
2
C(n, α, ρ)R
R
1
+
Q
1
(1 + )
and we deduce
Q max
C(α)
R
R
+
, C
0
(α)R
1+
R
1
1+
+
t
1+
.
From the definition of Q and the estimate h 2R
+
, we have
ρS
α
n
max
(
C(n, α, ρ)R
, C
0
(n, α, ρ)
R
+
R
1+
t
1+
)
.
We now proceed to obtain lower bounds on the speed and displacement.
It is in this estimate that we require α 1/n. The argument combines
barrier arguments with a Harnack inequality (proved for isotropic Gauss
curvature flows by Chow [Ch3] and for more general flows by the author
[A7]).
14 BEN ANDREWS
The Harnack estimate can be stated as follows for any solution of Equation
(13) (see [A7], Theorem 5.6):
Theorem 7. For any smooth, strictly convex solution of (13) on S
n
×[0, T ),
d
dt
ρS
α
n
t
+1
0
everywhere on S
n
× (0, T ).
Our lower speed estimate is the following:
Theorem 8. For α < 1/n the following holds for any smooth, strictly con-
vex solution of Eq. (13):
h(z, t) h(z, 0) C(ρ, n, α)r
+
(M
0
)
2
1
t
1
1
and
ρ(z)S
n
(z, t)
α
C
0
(ρ, n, α)r
+
(M
0
)
2
1
t
1
for 0 < t < C
00
(n, ρ, α)r
+
(M
0
)
1+
. For α = 1/n we have instead the
estimates for each γ > 0
h(z, t) h(z, 0) C(n, ρ, γ)r
+
(M
0
)
1+2γ
t
γ
e
C
0
(ρ,n,γ)r
+
(M
0
)
2n
t
n
and
ρ(z)S
n
(z, t)
1/n
¯
C(n, ρ, γ)r
+
(M
0
)
2γ1
t
γ
e
¯
C
0
(ρ,n,γ)r
+
(M
0
)
2n
t
n
for 0 < t <
¯
C
00
(n, ρ, γ)r
+
(M
0
)
2
.
Proof. For n = 1 these estimates are proved in [A2], Theorem II2.4. Suppose
n 2.
In the case α < 1/n it suffices to use large spheres as barriers: Fix
z S
n
. Then M
0
is enclosed in a hemispherical region obtained by in-
tersecting the sphere of radius 2r
+
(M
0
) centred at ¯x(z) with the half-space
{y R
n+1
: hy, zi h(z)}. For any ε < 2r
+
(M) this hemispherical region is
enclosed by the sphere S
ε
of radius (ε
2
+4r
+
(M)
2
)/(2ε) centred at the point
¯x(z)(4r
+
(M)
2
ε
2
)/(2ε)z (this sphere is chosen to have support function in
direction z equal to h(z) + ε). We consider the evolution of these spheres for
suitably small ε: Since ρ inf
S
n
ρ, a sphere of radius r evolves in time t to be
contained inside a sphere of radius
r
1+
(1 + ) inf ρt
1/(1+)
about
the same centre, which is enclosed by the sphere of radius r inf ρr
t.
In particular, this applies for each of the sphere S
ε
, and by the compari-
son principle M
t
is also enclosed by this smaller sphere. This implies the
inequality
h(z, t) h(z, 0) ε (inf ρ)
ε
2
+ 4r
+
(M)
2
2ε
t
ε
1 inf ρ(4r
+
(M)
2
)
ε
1
t
.
MOTION OF HYPERSURFACES BY GAUSS CURVATURE 15
In particular, cho osing
ε =
t inf ρ
2
1+2
r
+
(M
0
)
2
1/(1)
we obtain
h(z, t) h(z, 0)
inf ρ
2
1+2
r
+
(M
0
)
2
1/(1)
t
1/(1)
for t C(ρ, n, α)r
+
(M
0
)
1+
.
This estimate on the change in the support function can be converted
to an estimate on the speed using the Harnack estimate from Theorem 7:
Applying the estimate on the time interval [t/2, t], we have
Cr
+
(M
0
)
2
1+
t
1
1
h(z, t/2) h(z, t)
= ρ(z)
Z
t
t/2
S
n
(z, τ )
α
t/2 sup
τ[t/2,t]
ρS
α
n
.
Theorem 7 then gives
t
+1
ρ(z)S
n
(z, t)
α
(t/2)
+1
sup
[t/2,t]
ρS
α
n
and hence
ρ(z)S
n
(z, t)
α
C
0
r
+
(M
0
)
2
1+
t
1/(1)1
.
In the case 1/n the sphere barriers are not sufficient, and we instead work
with graphical barriers. The displacement bound is a consequence of the
following:
Lemma 9. Suppose α = 1/n. If M
0
has bounded isoperimetric ratio, and
lies in the region x
n+1
0 and within the ball kxk R, then M
t
lies inside
the region
x
n+1
C
1
R
1+2γ
t
γ
e
C
2
R
2n2
(r2R)
2
t
n
+ e
C
2
R
2n2
(r+2R)
2
t
n
for 0 t C
3
R
2
, where r
2
=
P
n
i=1
x
2
i
, and C
1
, C
2
and C
3
are constants
depending only on n, ρ, and γ.
Proof. We will show that the boundary of the region described is a graphical
subsolution of the evolution Equation (14). In the special case of a radially
symmetric function, we have
K =
u
00
(u
0
)
n1
r
n1
(1 + (u
0
)
2
)
(n+2)/2
,
16 BEN ANDREWS
and so
(19)
˙u ˜ρ(Du)K
1/n
p
1 + |Du|
2
˙u inf ρ(u
00
)
1/n
u
0
r
11/n
1 + (u
0
)
2
1/n
.
A direct computation shows that the function
u(r, t) = C
1
R
12γ
t
γ
e
C
2
R
2n2
(r2R)
2
t
n
+ e
C
2
R
2n2
(r+2R)
2
t
n
makes the right-hand side of (19) non-positive on the region r < R, t
C
3
R
2
, for any γ 0, where C
1
, C
2
, and C
3
depend on n, ρ and γ. Since the
boundary of this region cannot intersect the hypersurface M
t
, the compari-
son principle applies.
This gives the bounds in the theorem, since we can rotate and translate
the solution to bring the initial supporting hyperplane to the hyperplane
x
n+1
= 0, with M
0
satisfying the conditions of Lemma 9 with R = 2r
+
(M
0
).
Thus h(z, 0) = 0. For positive sufficiently small t, Lemma 9 gives
h(z, t) = inf
M
t
x
n+1
u(0, t)
as required.
Similar barriers can also be constructed for each α < 1/n.
The speed bound follows using Theorem 7 as for the previous cases.
We remark that the estimate for α < 1/n does not rely at all on the
particular structure of the Gauss curvature flows the same result holds
for any strictly parabolic flow with speed homogeneous of degree less than
1 in the curvatures.
6. Curvature control.
In this section we prove that the ratio of the maximum and minimum princi-
pal curvatures remains bounded throughout the evolution, given the upper
and lower speed bounds of the previous section. Our argument is an appli-
cation of the parabolic maximum principle to the evolution equation for the
curvature.
In the case n = 1 the speed bounds above and below already give complete
control on the curvatures. For the rest of this section we assume n 2.
Theorem 10. Suppose h : S
n
× [0, T ] is a solution of Eq. (13) for which
the isoperimetric ratio is bounded and the speed is bounded above and below
that is, there exist constants C
1
, C
2
such that
0 < C
1
S
n
(z, t) C
2
for every z S
n
and t [0, T ]. Then there exist positive const ants C
3
and
C
4
depending only on C
1
, C
2
, ρ, n and α such that
λ
i
(z, t) min{C
3
t
n1
, C
4
}
MOTION OF HYPERSURFACES BY GAUSS CURVATURE 17
for all i {1, . . . , n}, z S
n
, and t [0, T ].
Proof. We begin by computing the evolution equation for the matrix r
ij
=
¯
i
¯
j
h + ¯g
ij
h under Eq. (13):
t
r
ij
=
¯
i
¯
j
ρS
α
n
¯g
ij
ρS
α
n
= αρS
(1+α)
n
˙
S
kl
n
¯
i
¯
j
r
kl
α(1 + α)ρS
(2+α)
n
¯
i
S
n
¯
j
S
n
+ αρS
(2+α)
n
¨
S
klmn
n
¯
i
r
kl
¯
j
r
mn
¯g
ij
ρS
α
n
S
α
n
¯
i
¯
i
ρ
+ αS
(1+α)
n
¯
i
ρ
¯
j
S
n
+ αS
(1+α)
n
¯
j
ρ
¯
i
S
n
.
In the first term here we apply the identity (12), to yield:
t
r
ij
= αρS
(1+α)
n
˙
S
kl
n
¯
k
¯
l
r
ij
α(1 + α)ρS
(2+α)
n
¯
i
S
n
¯
j
S
n
(20)
+ αρS
(2+α)
n
¨
S
klmn
n
¯
i
r
kl
¯
j
r
mn
+ ( 1)ρS
α
n
¯g
ij
αρS
(1+α)
n
˙
S
kl
n
¯g
kl
r
ij
+ S
α
n
¯
i
¯
j
ρ + αS
(1+α)
n
¯
i
ρ
¯
j
S
n
+ αS
(1+α)
n
¯
j
ρ
¯
i
S
n
.
We wish to obtain an upper bound for the eigenvalues of r
ij
, so the second
term on the first line and the last term on the second line are good terms
since they are negative. The first term of the first line is an elliptic operator,
and so is non-positive at a point and direction where a maximum eigenvalue
occurs. The first term of the second line we estimate using the concavity
of the nth root of the determinant as a function of the components of the
matrix, which is equivalent to the inequality
(21)
¨
S
klmn
n
n 1
nS
n
˙
S
kl
n
˙
S
mn
n
ξ
kl
ξ
mn
0
for any symmetric matrix ξ. Finally, the first term on the last line is
bounded, and the other two terms of the last line can be estimated in terms
of the good second term of the first line:
¯
i
ρ
¯
i
S
n
Cε
¯
i
S
n
2
+ Cε
1
for any ε > 0. Combining these estimates, we obtain
(22)
t
r
ij
αρS
(1+α)
n
˙
S
kl
n
¯
k
¯
l
r
ij
+ CS
α
n
¯g
ij
αρS
(1+α)
n
˙
S
kl
n
¯g
kl
r
ij
.
The last term here will allow us to obtain an estimate independent of
initial data: We have S
kl
n
¯g
kl
= S
n1
/S
n
, and the Newton inequalities [Mi]
give
(23) S
n1
S
n2
n1
n
S
1
n1
1
CS
n2
n1
n
r
1
n1
max
.
18 BEN ANDREWS
Now work at a p oint and time where a maximum eigenvalue is attained,
and suppose i = j and e
i
is the eigenvector of r with the largest eigenvalue.
Then the first term in the evolution equation is negative, and
r
max
t
CS
α
n
¯g
ij
CS
(
n
n1
+α)
n
r
n
n1
max
CS
α
n
¯g
ij
CS
n
n1
n
r
n
n1
max
.
Given the bound below on S
α
n
, the bracket is negative provided r
max
is
sufficiently large. For the same reason the coefficient in front of the bracket
does not become small, and we have for r
max
sufficiently large
d
dt
r
max
Cr
n/(n1)
max
.
The result now follows by the parabolic maximum principle and comparison
with the solution of the ordinary differential equation du/dt = Cu
n/(n1)
.
Next we observe that this automatically provides an upper bound on the
principal curvatures:
Proposition 11. Let W be a positive definite symmetric matrix for which
W εId and detW C. Then W Cε
(n1)
Id.
Proof. Number the eigenvalues λ
1
of W in ascending order: λ
1
λ
2
···
λ
n
. Then
λ
n
=
K
λ
1
λ
2
. . . λ
n1
K
λ
n1
1
,
where K = λ
1
. . . λ
n
= det W.
In particular, the upper speed bound of Theorem 6 and the lower curva-
ture bound of Theorem 10 imply an upper curvature bound:
Corollary 12. Un der the conditions of Theorem 10 there exist constants
C
5
and C
6
such that
W max{C
5
t
(n1)
2
, C
6
}Id.
7. Convergence to a point.
In this section we prove that any solution of Eq. (6) with a smooth, strictly
convex initial hypersurface converges to a point in finite time. In the special
case of isotropic flows (ρ 1) this was proved by K.S. Chou [Ts] for α = 1
and by Ben Chow [Ch1] for other α. While we only need the result for
α 1/n, we give a proof which works for larger α as well.
Theorem 13. For any α > 0 and positive ρ C
(S
n
), and any smooth,
strictly convex hypersurface M
0
R
n+1
, the h ypersurfaces M
t
given by the
solution of Eq. (6) exist for a finite time T and converge in Hausdorff dis-
tance to p R
n+1
as t approaches T .
MOTION OF HYPERSURFACES BY GAUSS CURVATURE 19
Proof. The maximal time of existence must be finite: By the comparison
principle, if M
0
is enclosed by a sphere S
n
r(0)
(q) for some r > 0 and q R
n+1
,
then for all t in the interval of existence, M
t
is enclosed by the sphere S
n
r(t)
(q),
where
r(t) =
r(0)
1+
(inf
S
n
ρ)(1 + )t
1/(1+)
.
r(t) converges to zero in finite time, and M
t
cannot exist beyond this time.
Consider again the estimate (22) for the evolution of the curvature. We
also have the evolution equation
t
h = ρS
α
n
(24)
= αρS
(1+α)
n
˙
S
kl
n
¯
k
¯
l
h + αρS
(1+α)
n
˙
S
kl
n
¯g
kl
h (1 + )ρS
α
n
.
Combining these, we obtain
t
(r
ij
+ Ah¯g
ij
) αρS
(1+α)
n
˙
S
kl
n
¯
k
¯
l
(r
ij
+ Ah¯g
ij
)
+ (C A(1 + ))S
α
n
¯g
ij
+ (Ah r
ij
) αρS
(1+α)
n
˙
S
kl
n
¯g
kl
.
Choose A = C/(1 + ), so that the last term of the first line vanishes.
Also note that since the hypersurfaces are contracting, we have h h
0
=
sup
S
n
h(z, 0) as long as the solution exists. Therefore we have, writing q
ij
=
r
ij
+ Ah¯g
ij
,
t
q
ij
αρS
(1+α)
n
˙
S
kl
n
¯
k
¯
l
q
ij
αρS
(1+α)
n
˙
S
kl
n
¯g
kl
(q
ij
2h
0
¯g
ij
)
and hence by the parabolic maximum, the maximum eigenvalue of q
ij
is
decreasing if it is larger than 2h
0
. Since the initial hypersurface is smooth
and strictly convex, q
ij
is bounded at t = 0. Therefore we have a uniform
bound on q
ij
and hence r
ij
throughout the interval of e xistence.
Suppose the inradius of the hypersurfaces M
t
do not converge to zero
that is, the solution exists for a maximal time interval [0, T ), but there is
some ball of positive radius that remains enclosed by the solution through-
out. By the argument in Section 5, the speed remains bounded throughout
the interval of existence.
By Proposition 11 this also implies a bound on the curvature, so that
the hypersurfaces remain uniformly smooth and strictly convex on the time
interval [0, T ). It follows that there exists a subsequence of times {t
k
} con-
verging to T such that M
t
k
converges in C
to a smooth, strictly convex
limit M
T
. Furthermore, the C
convergence implies that all time deriva-
tives converge, so that in fact M
t
approaches M
T
in C
as t approaches
20 BEN ANDREWS
T . Hence we have a smooth solution on [0, T ], and the short-time exis-
tence result implies that this can be extended beyond T , contradicting the
assumption that T was maximal.
Therefore the inradius converges to zero. Since r
ij
is uniformly bounded,
this also implies that the circumradius converges to zero, and the hypersur-
faces converge to a point.
Theorem 13 implies a bound below on the time of existence of solutions
with smooth and strictly convex initial data, in terms of the inradius of the
initial hypersurface, ρ, and α: Any sphere which is initially enclosed by the
hypersurface acts as a barrier, preventing the hypersurface from contracting
to a point too quickly.
8. Short-time existence.
In this section we prove the following:
Theorem 14. For any positive ρ C
(S
n
), α 1/n, and open bounded
convex region
0
R
n+1
, there exists a smooth, strictly convex solution
x
t
: S
n
× (0, T ) of Eq. (6) which converges to M
0
=
0
in Hausdorff
distance as t approaches zero. Any other such solution y
t
: S
n
× (0, T
0
) is
given by x
t
ϕ for some smooth diffeomorphism ϕ of S
n
.
At the end of the section we also prove the existence and uniqueness of
viscosity solutions for arbitrary convex initial data and arbitrary α > 0.
8.1. Existence. In order to construct a solution which approaches M
0
at
the initial time, we consider a family of smooth, strictly convex hypersurfaces
M
(ε)
0
which approach M
0
in Hausdorff distance as ε approaches zero. By
Theorem 13, for each ε > 0 there exists a unique solution M
(ε)
t
of Eq. (6)
with initial condition M
(ε)
0
, which converges to a point in finite time T
ε
> 0.
Then we have
d
H
(M
0
, M
(ε)
t
) d
H
(M
0
, M
(ε)
0
) + d
H
(M
(ε)
0
, M
(ε)
t
)
d
H
(M
0
, M
(ε)
0
) + Ct
1/(1+)
.
By Theorem 10 and Corollary 12, the hypersurfaces M
(ε)
t
satisfy bounds
above and below on the principal curvatures, uniformly in ε over every
compact subset of S
n
× (0, T ). It follows from the regularity theory for
solutions of uniformly parabolic equations concave in the second derivatives
([K], Theorem 5.5) that there are also bounds on all higher derivatives of the
curvatures, uniformly in ε over every compact subset of S
n
×(0, T ). It follows
from the Arzela-Ascoli theorem that there exists a sequence ε
k
approaching
zero such that {M
(ε
k
)
t
} converges in C
to a family of hypersurfaces {M
t
}
satisfying the same bounds. In particular, {M
t
} satisfies Eq. (6) on S
n
×
MOTION OF HYPERSURFACES BY GAUSS CURVATURE 21
(0, T ), M
t
is smo oth and strictly convex for each t > 0, and d
H
(M
0
, M
t
)
Ct
1/(1+)
.
8.2. Uniqueness. Suppose we have two solutions {M
(1)
t
} and {M
(2)
t
} of
Eq. (6), both converging to M
0
in Hausdorff distance as t 0, and denote
by h
(i)
t
the corresponding support functions. Fix ε > 0. Then there exists
t
0
(ε) > 0 such that |h
(i)
t
(z) h
0
(z)| < ε for i = 1, 2 and all z S
n
and
t (0, t
0
(ε)). Choose a smooth, strictly convex hypersurface M
(ε)
0
with
support function h
(ε)
0
such that h
0
(z) 2ε < h
(ε)
0
(z) < h
0
(z) ε. Without
loss of generality we assume that the origin is at the centre of a ball of radius
r
(M
0
) enclosed by M
0
. Then r
(M
0
) h
0
(z) 2r
+
(M
0
). It follows that
h
(ε)
0
(z) < h
0
(z) ε < h
(i)
t
(z) < h
0
(z) + ε <
1 +
3ε
r
(M
0
) 2ε
h
(ε)
0
for i = 1, 2 and all z S
n
and t (0, t
0
(ε)). The comparison principle and
the scaling property given by Eq. (15) then imply
h
(ε)
τ
(z) < h
(i)
t+τ
(z) < (1 + λ)h
(ε)
(1+λ)
(1+)
τ
(z)
for all τ 0 for which these all exist, where λ =
3ε
r
(M
0
)2ε
. Consequently,
h
(2)
t+τ
(z) h
(1)
t+τ
(z)
(1 + λ)h
(ε)
(1+λ)
(1+)
τ
(z) h
(ε)
τ
(z)
λh
(ε)
(1+λ)
(1+)
τ
(z) +
h
(ε)
(1+λ)
(1+)
τ
(z) h
(ε)
τ
(z)
2r
+
(M
0
)λ + C
1 (1 + λ)
(1+)
τ
1/(1+)
Cλ + C(λτ)
1/(1+)
.
Here we used Theorem 6 to obtain the second-last line. Now take t 0.
Since ε > 0 is arbitrary and C independent of ε, we have for each τ > 0 and
z S
n
h
(2)
τ
(z) = h
(1)
τ
(z).
Note that the proof presented here does not rely strongly on the particular
structure of the evolution Equation (6). In particular, the uniqueness argu-
ment is valid for any flow by a monotone, positively homogeneous function
of curvature, since the bound on the change in the support function given in
Theorem 6 also holds for all such evolution equations. The existence argu-
ment requires a speed bound and regularity estimates independent of initial
data.
We now proceed to the case α > 1/n: In this case (as we show in Section
12) one c annot expect to produce smooth solutions from arbitrary convex
initial hypersurfaces. Instead we will work with a weaker notion of solution:
A family of convex regions {
t
}
0<t<T
is called a viscosity solution of Eq. (6)
22 BEN ANDREWS
if the following conditions hold: First, for any smooth, strictly convex hy-
persurface M
0
contained in
t
0
for some t
0
(0, T ), the hypersurfaces M
t
given by solving (6) are contained in
t
0
+t
for all t [0, T t
0
) in the domain
of existence of the M
t
. Second, for any smooth, strictly convex hypersurface
M
0
which encloses
t
0
for some t
0
(0, T ), the hypersurfaces M
t
enclose
t
0
+t
for all t [0, T t
0
).
Theorem 15. For any smooth positive ρ C
(S
n
), α > 0, and any open
bounded convex region
0
R
n+1
, there exists a unique viscosity solution
{
t
}
0<t<T
which converges to
0
is Hausdorff distance as t approaches zero.
t
converges to a point as t approaches T .
Proof. We use the same construction as presented in the proof of Theorem
14, producing a solution {M
(ε)
t
} for each ε > 0, with M
(ε)
0
approaching
0
in Hausdorff distance as ε approaches zero. We specify further that M
(ε)
0
is
contained in
0
for all ε > 0, and is increasing in ε.
For ε sufficiently small, we can choose an origin for R
n+1
and radii R >
r > 0 such that the ball B
r
(0) is enclosed by all of the hypersurfaces M
(ε)
0
,
and the ball B
R
(0) contains all of the hypersurfaces M
(ε)
0
. By the comparison
principle, there exists δ > 0 such that the ball B
r/2
(0) is enclosed by all the
hypersurfaces M
(ε)
t
for t [0, δ]. The hypersurfaces also remain enclosed by
the ball B
R
(0).
It follows that the support functions h
(ε)
t
(z) are uniformly Lipschitz: By
the Formula (9), we have |¯x|
2
= h
2
+ |
¯
h|
2
, and re-arrangement gives
|
¯
h|
2
|¯x|
2
R
2
, which is a uniform Lipschitz bound.
Furthermore, the displacement bound and the speed bound of Theorem
6 show that h
(ε)
t
is older continuous in t, uniformly in ε and z, and also
uniformly Lipschitz on compact subsets of (0, δ). Therefore h
(ε)
(z, t) is a
older continuous function on S
n
× [0, δ], uniformly in ε. By the Arzela-
Ascoli theorem, there exists a sequence ε
k
approaching z ero which converges
to a limit h(z, t) satisfying the same e stimate s. By the Blaschke selection
theorem, each of the functions h
t
= h(., t) is the support function of a convex
region
t
, and the same argument as in the proof of Theorem 14 shows that
t
approaches
0
in Hausdorff distance as t approaches zero.
We need to prove that the family {
t
} is a viscosity solution. The first
condition is easily checked: If M
0
0
is contained within
0
, then M
0
0
is also
enclosed by M
(ε)
0
for ε > 0 sufficiently small. By the comparison principle,
the resulting solution M
0
t
is enclosed by M
(ε)
t
for t > 0, and also M
(ε)
t
is
increasing in ε and converges to
t
as ε approaches zero. Therefore M
0
t
is
contained in
t
for t > 0.
MOTION OF HYPERSURFACES BY GAUSS CURVATURE 23
The second condition also follows easily: Any hypersurface M
0
0
which
encloses
0
also encloses all of the hypersurfaces M
(ε)
0
, and so by the com-
parison principle M
0
t
encloses M
(ε)
t
for ε > 0 and t > 0, and so also encloses
the limit
t
.
The uniqueness statement in Theorem 15 follows exactly as in the proof
of uniqueness in Theorem 14, and the same argument shows that the regions
t
converge to a point.
9. Application: The affine normal flow.
In this section we apply the speed and curvature bounds of the previous
sections to give a short proof of the following theorem:
Theorem 16. Let α = 1/(n + 2) and ρ 1. For any convex open region
0
R
n+1
there exists a smooth family of strictly convex embeddings x
t
:
S
n
R
n+1
satisfying Eq. (6) for which the Ha usdorff distance between
M
t
= x
t
(S
n
) and M
0
approaches zero as t 0. Any other such solution {˜x
t
}
is related to {x
t
} by composition with a time-independ ent diffeomorphism.
M
t
converges to a point p R
n+1
as t approaches a finite time T , and
˜
M
t
=
Vol(S
n
)
Vol(M
t
)
1/(n+1)
(M
t
p)
converges in C
to an ellipsoid centred at the origin.
This theorem was proved in the case of smooth, strictly c onvex M
0
in
[A4]. The results of Section 6 allow us to give a proof which works also
for singular initial hypersurfaces. The argument is also considerably simpler
because it avoids the com plicated third-derivative estimate which was the
key to the proof in [A4]. On the other hand, we use the result that elliptic
affine hyperspheres are ellipsoids, which was not necessary for the proof in
[A4].
Proof. By Section 8, we have a unique solution of Eq. (6) with the given
initial condition. Since this is smooth and strictly convex for t > 0, the
result of Theorem 13 implies that this solution converges to p R
n+1
in
finite time T .
In the proof we use the fact that the evolution equation is invariant under
the action of the special affine group: If {M
t
} is a family of hypersurfaces
moving under Eq. (6), then {L(M
t
)} is also such a family, for any special
affine transformation L.
Fix t [0, T ). There exists a special affine transformation L
t
such that
L
t
(M
t
) has
C
Vol(M
t
)
1/(n+1)
Vol(S
n
)
1/(n+1)
h
L
t
(M
t
)
(z) C
+
Vol(M
t
)
1/(n+1)
,
for some absolute constants C
±
.
24 BEN ANDREWS
We consider the solution
˜
h given by the scaling relation (15):
˜
h(z, τ ) =
Vol(S
n
)
Vol(M
t
)
1/(n+1)
h
L
t
z, t +
Vol(S
n
)
Vol(M
t
)
2/(n+2)
τ
!
,
where h
L
is the support function of the convex body obtained by apply-
ing the special affine transformation L to the body with support func-
tion h. Then C
˜
h(z, 0) C
+
. By the comparison principle we also
have
1
2
C
˜
h(z, τ ) C
+
for τ [0, δ], where δ = ( C
)
2(n+1)/(n+2)
(1
2
2(n+1)/(n+2)
)(n + 2)/2(n + 1). Hence on the interval [δ/2, δ] there are
uniform speed and displacement bounds (by Theorem 6), a uniform lower
bound on the speed (by Theorem 8), uniform bounds above and below on
the principal curvatures (by Theorem 10 and Corollary 12), and uniform
bounds on all higher derivatives of the curvature (by Theorem 5.5 of [K]).
It follows that the original solution satisfies uniform bounds on all quan-
tities which are both scaling invariant and special-affine invariant, on the
time interval
t +
1
2
CVol(M
t
)
2/(n+2)
, t + CVol(M
t
)
2/(n+2)
, for some abso-
lute constant C. Since t is arbitrary, we have such bounds on the entire
interval [T /2, T ).
Therefore there exists a sequence {t
k
} approaching T , and a sequence of
special affine transformations {L
k
}, such that the hyp e rsurfaces {L
k
(
˜
M
t
k
)}
converge in C
to a smooth, strictly convex limit
˜
M
T
.
Suppose
˜
M
T
does not satisfy the condition K
1/(n+2)
= chx, νi for some
c > 0 and some choice of origin. Then the time derivative of Z
\,1/(n+2)
on
˜
M
T
is strictly negative , by Theorem 3. By the C
convergence and scaling,
there exists k
0
, δ > 0, and ε > 0 such that whenever k k
0
we have
Z
\,1/(n+2)
M
t
k
+δVol(M
t
k
)
2/(n+2)
Z
\,1/(n+2)
(M
t
k
) ε.
But since Z
\,1/(n+2)
is non-increasing, this would imply Z
\,1/(n+2)
(
˜
M
t
k
)
−∞ as k , which is impossible. Therefore
˜
M
T
satisfies the required
condition.
By Theorem 1 of [Ca], a smooth, strictly convex hypersurface satisfies
the condition K
1/(n+2)
= chx, νi if and only if it is an ellipsoid.
The stronger convergence statements in the Theorem follow by consid-
ering the linearization of the evolution Equation (13) about the space of
ellipsoids a direct calculation shows that this space is strictly linearly
stable, so Proposition 9.2.3 of [Lu] and a scaling argument implies that M
t
converges in C
to the ellipsoid
˜
M
T
after rescaling. The details of this
argument are given for a related evolution equation in [A10], Propositions
40-41.
MOTION OF HYPERSURFACES BY GAUSS CURVATURE 25
Corollary 17. For any open bounded convex region
0
R
n+1
, the follow-
ing generalised affine isoperimetric inequality holds:
lim
t0
Z
\,
1
n+2
(Ω
t
) 1
with equality if and only if
0
is an ellipse.
In effect this result allows a definition of the affine surface area for convex
hypersurfaces which may be non-smooth or non-strictly convex, in such a
way that the affine isoperimetric inequality remains true. A related exten-
sion of the affine surface area has been given in [Le].
10. Proof of the main Theorems.
In this section we complete the proofs of Theorems 1 and 2. The proof is
similar to that presented in the special case of the affine normal flow in the
previous section, but is somewhat simpler because we do not require the
machinery of affine corrections to obtain bounded isoperimetric ratios.
Section 8 provides us with a unique solution {M
t
} of Eq. (6) for a short
time, and Theorem 13 ensures that this solution remains smooth until it
converges to a point p R
n+1
in finite time T . In the case α > 1/(n +
2), Theorem 4 gives a uniform bound on the isoperimetric ratios of the
hypersurfaces M
t
throughout the interval [0, T ]. In the case described in
Theorem 2, we also have such an estimate by hypothesis.
Theorems 6, 8 and 10 and Corollary 12 therefore imply uniform bounds
above and below on the principal curvatures of the rescaled hypersurfaces
{
˜
M
t
} defined by rescaling to fixed enclosed volume about the point p. The-
orem 5.5 of [K] and Schauder estimates ([Li], Theorem 4.9) imply uniform
bounds on all higher derivatives of curvature. It follows that there exists a
subsequence of times {t
k
} approaching T for which the hypersurfaces
˜
M
t
k
converge in C
to a smooth, strictly convex limit
˜
M
T
. By the same argu-
ment as that in Section 9,
˜
M
t
satisfies the equation hx, νi = cρK
α
for some
c > 0. The convergence in C
of
˜
M
t
to
˜
M
T
as t approaches T follows from
Theorem 2 of [A6].
11. Application: Non-trivial homothetic solutions.
In this section we give an application of Theorem 2 to prove the existence
of homothetically contracting solutions of the isotropic flows
t
x = K
α
ν
for s ufficiently small α. The idea is to consider the evolution of hypersurfaces
close to the sphere S
n
, possessing suitable symmetries. Precisely, our result
is the following:
26 BEN ANDREWS
Theorem 18. Let Γ be a proper subgroup of SO(n + 1) such that for every
z S
n
, the orbit of z under Γ spans R
n+1
(that is, the inclusion of Γ in
SO(n + 1) is an irreducible representation). Let λ be the smallest eigenvalue
corresponding to a non-trivial Γ-invariant spherical harmonic ϕ. Then for
α (0, 1/(λn)) there exists a non-spherical, Γ-symmetric, smooth, strictly
convex hypersurface satisfying the identity hx, νi = K
α
.
Proof. The linearization of the normalised isotropic Equation (6) about the
sphere solution h = 1 is given by
t
η = α(∆ + n)u + u
where is the Laplacian on S
n
. In particular, for α (0, 1/(λ n)) the
h 1 solution is strictly unstable in the direction ϕ. By [Lu], Theorem
9.1.3, there exists a Γ-symmetric solution {M
t
} of Eq. (6) which converges
to h 1 as t −∞ and diverges exponentially from h 1.
We observe that Γ-symmetry of a convex hypersurface implies a bound
on the isoperimetric ratio:
Lemma 19. For any Γ satisfying the conditions of Theorem 18, there exists
a const ant C such that every Γ-symmetric convex hypersurface M R
n+1
satisfies r
+
(M)/r
(M) C.
Proof. If this is not the case, then we can find a sequence of Γ-symmetric
convex hypersurfaces M
k
such that r
+
(M
k
) = 1 and r
(M
k
) 1/k. By
the Blaschke selection theorem ([Sc], Theorem 2.5.14) we can choose a sub-
sequence M
k
0
which converges in Hausdorff distance to a limit M
which
is again Γ-symmetric but has r
(M
) = 0 and r
+
(M
) = 1. It follows
that M
is contained in a lower-dimensional subspace of R
n+1
. But this
is impossible, because there exists x M
with |x| = 1; by Γ- symm etry
all of the point g(x) are in M
, but these are not contained in any such
sub-space.
Therefore we can apply Theorem 2 to the solution {M
t
}, obtaining C
convergence to a limit M
T
satisfying the required identity (possibly after
scaling to ensure c = 1). M
T
is Γ-symm etric, and has Z strictly less than
that for the sphere solution, since Z has strictly decreased along the solution
{M
t
}. Therefore M
T
is non-spherical.
Corollary 20. In the case n = 1, for each k 3 and each α (0, 1/(k
2
1)) there exists a non-circular strictly convex smooth curve C
k
with k-fold
symmetry which contracts homothetically under the flow
t
x = κ
α
n
where κ is the curvature, and n the unit normal.
MOTION OF HYPERSURFACES BY GAUSS CURVATURE 27
Proof. For k 3 the subgroup Γ
k
generated by rotation through 2π/k satis-
fies the conditions of Theorem 18. The first spherical harmonics symmetric
under Γ
k
are cos(kθ) and sin(), with corresponding eigenvalue λ = k
2
.
Corollary 21. In the case n = 2, there exists a homothetically contracting
solution of Eq. (6) with tetrahedral symmetry provided α (0, 1/10); there
exists one with octahedral symmetry provided α (0, 1/18); and there exists
one with icosahedral symmetry provided α (0, 1/40).
Proof. In this case the only subgroups satisfying the required condition are
the s ymmetry groups of the platonic solids. There are three such groups,
since the dual solids have the same symmetry group. The first tetrahedrally
symmetric spherical harmonic is given by the restriction of the function
xyz to S
2
, and the corresponding eigenvalue is 12. The first octahedrally
symmetric spherical harmonic is x
4
+ y
4
+ z
4
3x
2
y
2
3x
2
z
2
3y
2
z
2
, and
the corresponding eigenvalue is 20. Finally, the first icosahedrally symmetric
spherical harmonic is
231z
6
315z
4
(x
2
+ y
2
) + 105z
2
(x
2
+ y
2
)
2
5(x
2
+ y
2
)
3
+ 42zx
5
420zx
3
y
2
+ 210zxy
4
,
and the corresponding eigenvalue is 42.
Corollary 22. For n 3, there exists a non-spherical homothetically con-
tracting solution with the symmetry of a regular (n + 2)-simplex for α
(0, 1/(2(n + 3))), and one with the symmetries of a regular hypercube for
α (0, 1/3(n + 4)).
Proof. The function
P
i
x
4
i
(6/n)
P
i6=j
x
2
i
x
2
j
has the symmetry of a regular
hypercube in R
n+1
, and its restriction to S
n
is a spherical harmonic with
eigenvalue 4(n + 3).
For all n 1 there exists a cubic homogeneous harmonic polynomial u
n
on R
n+1
with the symmetries of a regular simplex. These are given by the
recursive definitions
u
1
(x
1
, x
2
) = x
3
2
3x
2
1
x
2
u
n+1
(x
1
, . . . , x
n+2
) = u
n
(x
1
, . . . , x
n+1
)
+ β
n
x
3
n+2
3
n + 1
x
2
1
+ ··· + x
2
n+1
where β
1
=
2 and
β
k+1
= β
k
s
(k + 1)
3
k
2
(k + 3)
.
The restriction of u
n
to S
n
is a spherical harmonic with the required sym-
metries, and the corresponding eigenvalue is 3(n + 2).
28 BEN ANDREWS
In [A5] we prove that the only homothetically contracting solutions of
flows by positive powers of curvature are those given in Corollary 20. For
n = 2 we expect that there are many more solutions which are not described
by Corollary 21. In particular, for small α there should be many solutions
symmetric under each of the platonic symmetry groups, and there should
also be solutions symmetric under some other subgroups of SO(3), such as
the subgroup of rotations about a fixed axis, and its discrete subgroups. In
the case n 3 there are of course many more examples of suitable subgroups
Γ which we have not mentioned explicitly in Corollary 22.
12. Hypersurfaces with planar or cyl indrical parts.
In this section we demonstrate that the result of Theorem 1 no longer holds
for any α > 1/n. Specifically, we show that any solution starting from a
hypersurface containing a planar region must still contain a planar region for
small positive times. We also consider the behaviour of hypersurfaces which
contain regions which are cylindrical or loc ally have the form M
nk
× R
k
for some k > 0.
Theorem 23. Suppose M
0
is a compact convex hypersurface, and F
0
a sub-
set of M
0
which has the form N
nk
0
×U
k
0
, where N
0
is a smooth convex hy-
persurface in R
n+1k
and U
0
is an open subset of R
k
. Then for any smooth
positive function ρ on S
n
and any α > 1/k, the viscosity solution {M
t
}
starting from M
0
contains an open subset F
t
of F
0
for t > 0 sufficiently
small.
Conversely, suppose
0
in a bounded open convex set in R
n+1
for which
σ
nk
= inf
x
0
sup
P,Γ
κ : y
0
hx y, P
i
1
2
κ|π
Γ
(x y)|
2
> 0
where the supremum is over all supporting hyperplanes P of
0
which con-
tain x and all n k-dimensional affine subspaces Γ of P through x, and P
is the unit normal to P which points outward from
0
. Then for any ρ and
any α (0, 1/k] the viscosity solution {
t
} of (6) with initial condition
0
is smooth a nd strictly convex for t > 0 and remains so until it contracts to
a point.
When
0
is smooth, σ
nk
is strictly positive if and only if E
nk
(W) is
strictly positive, or equivalently if and only if the sum of the smallest k + 1
principal curvatures is strictly positive at every point.
Proof. To prove the first part of the Theorem, we will construct barriers with
cylindrical symmetry, desc ribed by embeddings of the form ϕ : S
nk
×B
k
R
(0)
with ϕ(z, x) = (u(|x|)z, x) R
n+1k
× R
k
' R
n+1
. We consider the case
where u is concave and decreasing in |x|. The metric and second fundamental
MOTION OF HYPERSURFACES BY GAUSS CURVATURE 29
form are given by
g
z
i
z
j
= u
2
¯g
ij
g
x
i
z
j
= 0
g
x
i
x
j
= δ
ij
+
x
i
x
j
|x|
2
(u
0
)
2
and
W
z
j
z
i
=
δ
j
i
u
p
1 + (u
0
)
2
W
x
j
z
i
= 0
W
x
j
x
i
=
u
00
(1 + (u
0
)
2
)
3/2
x
i
x
j
|x|
2
u
0
|x|
p
1 + (u
0
)
2
δ
j
i
x
i
x
j
|x|
2
.
The Gauss curvature is given by the expression
K = (1)
k
u
00
(u
0
)
k1
u
nk
|x|
k1
(1 + (u
0
)
2
)
(n+2)/2
and the unit normal is
ν =
z u
0
ˆx
p
1 + (u
0
)
2
where ˆx = x/|x|.
It follows that any function u(r, t) satisfying the inequality
(25) ˙u sup ρ(u
00
)
α
(u
0
)
α(k1)
u
α(nk)
|x|
α(k1)
1 + (u
0
)
2
1α(n+2)
2
has a cylindrical graph which acts as an inner barrier for convex solutions
of Eq. (6). We proceed to construct such barriers.
Lemma 24. For given k, ρ, and α > 1/k there exist constants c
1
, c
2
and
c
3
such that for any λ > 0 and u
0
> 0, R > 0, the function
u = u
0
1 c
1
(k, α)
|x|
R
+ c
2
(ρ, k, α)
u
0
R
2
t
u
1+
0
1
k+11
k1
+
is C
2
and satisfies (25) on the region
|x| 2R, 0 t c
3
(ρ, k, α)u
1+
0
R
u
0
2
,
provided α (1/k, 1/(k 1)). If α 1/(k 1) then u is not C
2
but acts as
a barrier for smooth, strictly convex solutions of Eq. (6), hence also for the
viscosity solutions constructed in Theorem 15.
30 BEN ANDREWS
Proof. The case α (1/k, 1/(k 1)) follows by direct computation. In
the case α 1/(k 1) the same computation gives the result except when
|x| = R c
2
(u
0
/R)
2
R/u
1+
0
t, where u is not C
2
. A smooth convex
hypersurface lying outside the graph of u and meeting the graph at such a
point must have K = 0, so the barrier condition is verified.
To use these barriers in the comparison principle, we need to check that
the viscosity solution {M
t
} stays away from the boundary of the barrier
produced in Lemma 24. Fix x in the interior of M
0
and ε > 0 small, and
choose u
0
smaller than the smallest radius of curvature of N
0
, and place
the origin at the point x (u
0
+ ε)ν
x
. Choose R sufficiently small so that
the distance from x to the boundary of F
0
is at least 3R. Let
˜
M
0
be the
hypersurface given by
v
u
u
t
nk
X
i=1
x
2
i
= u
0
1 c
1
(k, α)
q
P
n
i=nk+1
x
2
i
R
1
k+11
k1
+
for
P
n
i=nk+1
x
2
i
4R
2
. Then
˜
M
0
lies entirely within M
0
. Furthermore, for
each y
˜
M
0
, the sphere B
c(ρ,k,α) min{u
0
,R}
(y) is enclose d by M
0
for some
constant c. Define
˜
M
t
to be the hypersurface
v
u
u
t
nk
X
i=1
x
2
i
= u
0
1 c
1
q
P
n
i=nk+1
x
2
i
R
+ c
2
u
0
R
2
t
u
1+
0
1
k+11
k1
+
for
P
n
i=nk+1
x
2
i
4R
2
. For t < c(ρ, n, k, α) min{u
0
, R}
1+
we have
˜
M
t
[
y
˜
M
0
B
r(t)
(y)
where r(t)
1+
= (c min{u
0
, R})
1+
(1 + ) sup ρt. The comparison
principle implies that each of these balls is enclosed by M
t
, and therefore
that
˜
M
t
does not me et M
t
. It follows by the comparison principle that
˜
M
t
remains entirely enclosed by M
t
on this time interval, and therefore that
h(ν
x
, t) h(ν
x
, t) ε for all ε > 0, so that x has not moved during this time
interval.
Now we proceed to the second part of the Theorem, the case of α 1/k.
This also proceeds using barrier constructions. Let x
0
. Since σ
nk
> 0,
there exists a hyperplane P supporting
0
at x and an n k-dimensional
subspace Γ of P such that
0
is contained in the region
hx y, P
i
1
4
σ
nk
|π
Γ
(x y)|
2
.
MOTION OF HYPERSURFACES BY GAUSS CURVATURE 31
0
is also clearly contained in the ball B
2r
+
(Ω
0
)
(x). We now construct bar-
riers:
Lemma 25. Fix k, ρ, α (0, 1/k) and constants R > 0 and u
0
> 0. If
every y
0
satisfies
(26) |y| R,
nk
X
i=1
y
2
i
u
2
0
then for 0 < t c(ρ, k, n, α) min{u
0
, R}
1+
we have for every y = (Y, η)
t
R
n+1k
× R
k
' R
n+1
,
|Y | u
0
ct
1
1
R
α1
1
u
α(nk)
1
0
(4R |η|)
α(k+1)
1
+ (4R + |η|)
α(k+1)
1
where c depends only on ρ, k, n, and α. If α = 1/k and every y
0
satisfies
(26), then for 0 < t c(ρ, k, n) min{u
0
, R}
1+n/k
and y = (Y, η)
t
we have
|Y | u
0
c
1
R exp
c
2
R
2k
u
nk
0
t
k
cosh
c
2
R
2k1
u
nk
0
|η|t
k
where c
1
and c
2
depend only on ρ, k, and n.
Proof. A direct calculation shows that these regions are given by cylindrical
graphs satisfying the inequality
(27) ˙u inf ρ(u
00
)
α
(u
0
)
α(k1)
u
α(nk)
|x|
α(k1)
1 + (u
0
)
2
1α(n+2)
2
which therefore act as outer barriers for solutions of Eq. (6).
It follows that the supp ort function at every point must change: If we
fix z S
n
, and place a cylindrical barrier outside the hypersurface M
0
and
passing through x, then we have
h(z, t) h(z, 0) ct
1
1
R
1+
1
u
α(nk)
1
0
for α < 1/k, and
h(z, t) h(z, 0) c
1
R exp
c
2
R
2k
u
nk
0
t
k
when α = 1/k. Theorem 7 then gives lower bounds on the speed ρS
α
n
uniformly in z for each t > 0, by the same argument as given in the proof of
Theorem 8. Theorem 10 implies bounds b e low on all the principal curvatures
at each positive time, and Corollary 12 gives upper bounds on the principal
curvatures. Theorem 5.5 of [K] implies that the solution hypersurface is
smooth and strictly convex for sufficiently small positive time s, and the
result of Theorem 23 follows from Theorem 13.
32 BEN ANDREWS
References
[AGLM] L. Alvarez, F. Guichard, P.L. Lions and J.M. Morel, Axioms and fundamental
equations of image processing, Arch. Rat. Mech. Anal., 123 (1993), 199-257.
[A1] B. Andrews, Contraction of convex hypersurfaces in Euclidean space, Calc. Var.
& P.D.E., 2 (1994), 151-171.
[A2] , Evolving convex curves, Calculus of Variations and P.D.E., 7 (1998),
315-371.
[A3] , Entropy estimates for evolving hypersurfaces, Comm. Analysis and Ge-
ometry, 2 (1994), 53-64.
[A4] , Contraction of convex hypersurfaces by their affine normal, J. Differen-
tial Geometry, 43 (1996), 207-230.
[A5] , Classification of limiting shapes for isotropic curve flows, preprint in
preparation.
[A6] , Monotone quantities and unique limits for evolving convex hypersurfaces,
Int. Math. Res. Not., (1997), 1001-1031.
[A7] , Harnack inequalitites for evolving hypersurfaces, Math. Z., 217 (1994),
179-197.
[A8] , Instability of limiting shapes for curves evolving by curvature, preprint.
[A9] , Instability and non-convergence in Gauss curvature flows, preprint in
preparation.
[A10] , The affine curve-lengthening flow, Journal f¨ur die reine und angewandte
Mathematik, 506 (1999), 43-83.
[AST] S. Angenent, G. Sapiro & A. Tannenbaum, On the affine heat equation for non-
convex curves, preprint, 1997 (37 pages).
[B] W. Blaschke, Vorlesungen ¨uber Differentialgeometrie II, Verlag von Julius
Springer, Berlin, 1923.
[Ca] E. Calabi, Complete affine hyperspheres I, Ist. NAZ Alta Mat. Sym.Mat., X
(1972), 19-38.
[CS] V. Caselles and C. Sbert, What is the best causal scale-space for 3D images?,
SIAM J. Appl. Math., 56 (1996), 1199-1246.
[Ch1] B. Chow, Deforming convex hypersurfaces by the nth root of the Gaussian cur-
vature, J. Differential Geometry, 22 (1985), 117-138.
[Ch2] , Deforming hypersurfaces by the square root of the scalar curvature, In-
vent. Math., 87 (1987), 63-82.
[Ch3] , On Harnack’s inequality and entropy for the Gaussian curvature flow,
Comm. Pure Appl. Math., 44 (1991), 469-483.
[Fi] W.J. Firey, On the shapes of worn stones, Mathematika, 21 (1974), 1-11.
[Ga1] M.E. Gage, An isoperimetric inequality with applications to curve shortening,
Duke Math. J ., 50 (1983), 1225-1229.
[Ga2] , Curve shortening makes convex curves circular, Invent. Math., 76
(1984), 357-364.
[Ga3] , Evolving plane curves by curvature in relative geometries, Duke Math.
J., 72 (1993), 441-466.
MOTION OF HYPERSURFACES BY GAUSS CURVATURE 33
[GH] M.E. Gage and R.S. Hamilton, The heat equation shrinking convex plane curves,
J. Differential Geometry, 23 (1986), 69-96.
[Gr] M. Grayson, The heat equation shrinks embedded closed curves to round points,
J. Differential Geometry, 26 (1987), 285-314.
[Ha1] R. Hamilton, Worn stones with flat sides, in ‘A tribute to Ilya Bakelman’ (College
Station, TX 1993), 69-78; Discourses Math. Appl., 3, Texas A&M Univ., College
Station, TX, 1994.
[Ha2] , Remarks on the entropy and Harnack estimates for the Gauss curvature
flow, Comm. Anal. Geom., 2 (1994), 155-165.
[Hu] G. Huisken, Contraction of convex hypersurfaces by their mean curvature, J.
Differential Geometry, 20 (1984), 237-268.
[K] N.V. Krylov, Nonlinear elliptic and parabolic equations of second order, D. Rei-
del, 1987.
[Le] K. Leichtweiss, On inner parallel bodies in the equiaffine geometry, Analysis and
geometry (Eds. B. Fuchssteiner and W.A.J. Luxemburg), BI-Verlag, Mannheim,
1992, 113-123.
[Li] G. M. Lieberman, Second order parabolic differential equations, World Scientific,
Singap ore, 1996.
[Lu] A. Lunardi, Analytic Semigroups and Optimal Regularity in Parabolic Problems,
Birk-h¨auser, Basel, 1995.
[Mi] D.S. Mitrinowic, Analytic inequalities, Springer-Verlag, Berlin-Heidelberg-New
York, 1970.
[NK] P. Neskovic and B. Kimia, Three-dimensional shape representation from
curvature-dependent surface evolution, Technical report LEMS-128, Divison of
Engineering, Brown University, 1993.
[Oa] J. Oaks, Singularities and self-intersections of curves evolving on surfaces, Indi-
ana Univ. Math. J., 43 (1994), 959-981.
[OST] P. Olver, G. Sapiro and A. Tannenbaum, Invariant geometric evolutions of sur-
faces and volumetric smoothing, SIAM J. Appl. Math., 57 (1997), 176-194.
[Sa] L. Santal´o, Un invariante afin para los cuerpos convexos del espacio de n dimen-
siones, Portugal. Math., 8 (1949), 155-161.
[ST1] G. Sapiro and A. Tannenbaum, On invariant curve evolution and image analysis,
Indiana Univ. Math. J., 42 (1993), 985-1009.
[ST2] , Affine invariant scale-spaces, Internat. J. Comput. Vision, 11 (1993),
25-44.
[ST3] , On affine plane curve evolution, J. Funct. Anal., 119 (1994), 79-120.
[Sc] R. Schneider, Convex bodies: The Brunn-Minkowski Theory, Encyclopedia of
Mathematics, 44, Cambridge University Press, 1993.
[Si] J. Simons, Minimal varieties in Riemannian manifolds, Ann. Math., 88 (1968),
62-105.
[Ts] K. Tso, Deforming a hypersurface by its Gauss-Kronecker curvature, Comm.
Pure Appl. Math., 38 (1985), 867-882.
[U1] J.I.E. Urbas, An expansion of convex surfaces, J. Differential Geometry, 33
(1991), 91-125.
34 BEN ANDREWS
[U2] , Complete noncompact self-similar solutions of Gauss curvature flows I.
Positive powers, Math. Ann., 311 (1998), 251-274.
Received Octobe r 28, 1998. This research was partially supported by an Australian Re-
search Council Q EII Fellowship.
Australian National University
A.C.T. 0200
Australia
E-mail address: andrews@maths.anu.edu.au
PACIFIC JOURNAL OF MATHEMATICS
Vol. 195, No. 1, 2000
SUMMATION OF FORMAL SOLUTIONS OF A CLASS OF
LINEAR DIFFERENCE EQUATIONS
B.L.J. Braaksma, B.F. Faber, and G.K. Immink
Dedicated to the memory of W.A. Harris, Jr
We consider difference equations y(s+1) = A(s)y(s), where
A(s) is an n × n-matrix meromorphic in a neighborhood of
with det A(s) 6≡ 0. In general, the formal fundamental so-
lutions of this equation involve gamma-functions which give
rise to the critical variable s log s and a level 1
+
. We show
that, under a mild condition, formal fundamental matrices of
the equation can b e summed uniquely to analytic fundamental
matrices represented asymptotically by the formal fundamen-
tal solution in appropriate domai ns.
The method of proof is analogous to a method used to
prove multi-summability of formal solutions of ODE’s. Start-
ing from analytic lifts of the formal fundamental matrix in
half planes, we construct a sequence of increasingly precise
quasi-functions, each of which is determined uniquely by its
predecessor.
1. Introduction.
This paper is concerned with summability of formal solutions of linear ho-
mogeneous difference equations. We consider the system
y(s + 1) = A(s)y(s),(1.1)
where s is a complex variable, y(s) C
n
, and A(s) an n × n-matrix, mero-
morphic at infinity, det A(s) 6≡ 0. For some p N Equation (1.1) has a
formal fundamental matrix solution of the form
ˆ
Y(s) =
ˆ
H(s)s
Λs
e
G(s)
s
L
,(1.2)
with
ˆ
H(s) End(n, C[[s
1/p
]]), det
ˆ
H(s) 6≡ 0, Λ =
L
m
j=1
λ
j
I
j
where λ
j
1
p
Z
and I
j
is the n
j
× n
j
-identity matrix, G(s) =
L
m
j=1
g
j
(s)I
j
where g
j
(s) 0
or g
j
(s) is a polynomial in s
1/p
of degree at most p with g
j
(0) = 0, and
L =
L
m
j=1
L
j
, L
j
= c
j
I
j
+ N
j
with c
j
C and N
j
an n
j
× n
j
-nilpotent
matrix, and with n
1
+ n
2
+ · · · + n
m
= n.
35
36 B.L.J. BRAAKSMA, B.F. FABER, AND G.K. IMMINK
The purpose of this pap er is to sum the entries of
ˆ
H(s) on certain un-
bounded domains D in order to obtain uniquely characterizable analytic
fundamental matrix solutions
Y(s) = H(s)s
Λs
e
G(s)
s
L
, with
H(s)
ˆ
H(s), s on D.
(1.3)
Any solution of (1.1) on D can be written as Y(s)P (s) where P (s) is a
1-periodic C
n
-valued function.
If the factor s
Λs
does not appear in the formal fundamental matrix, i.e.,
if all λ
j
’s vanish, the formal fundamental matrix resembles that of a homo-
geneous linear differential system. Formal power series solutions of mero-
morphic differential equations can be summed by means of a method known
as multisummation. With such an equation one can associate so-called ‘lev-
els’, positive rational numbers k
1
, . . . ,k
r
, and corresponding ‘critical vari-
ables’ s
k
1
, . . . ,s
k
r
, which play a crucial part in the summation process.
Multi-summation is a particular case of accelero-summation (see [Eca87]),
involving only elementary accelerations. There exist various equivalent defi-
nitions of multisummability (see Definition 2). It can be formulated in terms
of Borel and Laplace transforms (cf. [MR91]), or in a more abstract way
(cf. [MR92]). In [Bal94] Balser presented yet another definition. Multi-
summability of solutions of both linear and nonlinear meromorphic differen-
tial equations has been proved both by using Borel-Laplace techniques (se e
[Bra91] and [Bra92]) and in a way based on the definition of Malgrange
and Ramis (see [BBRS91], [RS94], [Bal94], [Tov96], and [BIS]).
Two of the most important features that distinguish linear difference
equations from linear differential equations are:
(i) The solution space of a homogeneous linear difference equation is lin-
ear over the 1-periodic functions instead of C-linear as in the case of
homogeneous linear differential equations.
(ii) The occurrence of the factor s
Λs
, that does not appear in formal solu-
tions of differential equations.
If the factor s
Λs
does not appear in the formal fundamental matrix, or,
more generally, if all λ
j
’s are equal, then all entries of
ˆ
H(s) are multi-
summable in all but at most a countably infinite numbe r of directions. This
was shown in [BF96] by means of Borel-Laplace techniques in the spirit of
the work of Ecalle [Eca85]. With the same techniques multisummability
of formal solutions of a class of non-linear difference equations was proved
there.
If not all λ
j
’s are equal, some of the entries of
ˆ
H(s) may not be multi-
summable in any direction. This is due to the fact that, in this case, one of
the critical variables is s log s, which is not a rational power of s.
Following Ecalle (cf. [Eca85]), one might set out to sum the formal so-
lutions by accelero-summation, using Borel and Laplace transforms. For a
SUMMATION METHOD FOR DIFFERENCE EQUATIONS 37
particular class of linear difference equations, accelero-summability of the
formal solutions was established in [Imm]. Ecalle’s method involves the
study of a convolution equation, obtained from the equation satisfied by
ˆ
H
by means of a formal Borel transformation in the variable s log s, which does
not look very inviting. In the present paper we take a different approach,
similar to the method employed in [BIS] to sum formal solutions of lin-
ear differential equations (cf. Theorem 13). Our starting point is the ‘main
asymptotic existence theorem’ for difference equations (Theorem 6), which
says that
ˆ
H(s) can be lifted on half planes in C
, bounded by the real or
imaginary axis, to an analytic matrix H(s) such that (1.3) defines an analytic
fundamental matrix Y(s) of the difference equation. With the equation we
associate certain levels 0 < k
1
< · · · < k
r
= 1, that can be extracted from
the formal fundamental solution, as well as a level 1
+
if not all λ
j
’s are equal
(see Definitions 3 and 5). We choose a covering of a neighbourhood of in
C
by appropriate half planes and, on each half plane a fundamental sys-
tem of (1.1) represented asymptotically by the formal fundamental system
(1.2). In several steps, modifying the solutions by exponentially small func-
tions of increasing order at each subsequent step, we construct a sequence
of so-called k
j
-precise quasi-functions, j = 1, . . . , r. If the e quation does not
possess a level 1
+
, this procedure yields the multi-sum, or (k
1
, . . . , k
r
)-sum
of the formal solution (Theorem 13).
If the equation does possess level 1
+
, the final step is more delicate than
the preceding ones. This is due to the relative ‘closeness’ of the levels 1 and
1
+
and the transcendental nature of the critical variable s log s. In order to
end up with a unique sum, we need to consider domains that are strictly
smaller than half planes, but sufficiently large to exclude the existence of flat
solutions of the difference equation satisfied by
ˆ
H, with a dominant factor
of the form s
(λ
i
λ
j
)s
, with λ
i
6= λ
j
. Here we shall consider domains of the
type {s C
| arg s ((h 1)π, (h + 1)π), (1)
h
<{s(log s + )} > 1},
with θ R, h Z (cf. Figures 2-5). On the union of two such domains with
the same h we can define a sum H(s) of
ˆ
H(s) if a certain generic condition
is satisfied (cf. Section 7). By means of (1.3) we obtain a unique analytic
fundamental matrix of the difference equation (Theorem 18).
In order to illustrate the particular properties of difference equations with
level 1
+
, we e nd this introduction with a simple example.
Example 1. Consider the equation
h(s + 1) as
1
h(s) = s
1
with a R, a > 0(1.4)
which can be transformed into the matrix equation
y
1
y
2
(s + 1) =
a/s 1/s
0 1
y
1
y
2
(s).
38 B.L.J. BRAAKSMA, B.F. FABER, AND G.K. IMMINK
(1.4) belongs to a class of equations that was discussed in [BH75] and in
great detail by Ecalle in [Eca85, §3.6] and later by Immink in [Imm]. It
has a unique formal solution
ˆ
h =
P
n1
h
n
s
n
. Let ˆu(t) :=
P
n1
h
n
(n1)!
t
n1
,
the formal Borel transform of
ˆ
h. The power series ˆu formally satisfies the
convolution equation
e
t
u(t) a(1 u)(t) = 1.
This equation has the unique analytic solution
u(t) = e
ta
e
ae
t
.
Thus ˆu(t) coincides with the Taylor series at t = 0 of this function and is
actually a convergent power series which extends to a holomorphic function
on C. The convergence of ˆu implies that
ˆ
h is 1-Gevrey. By using Cauchy’s
formula for the coefficients in a convergent Taylor series, one may derive the
more precise estimate
|h
n
| K
n
log n
n
A
n
, n 2, for some K, A > 0.
This type of estimate is typical of difference equations possessing a level 1
+
(cf. [Imm88]).
The function u(t) = e
ta
e
ae
t
is bounded in the left half plane and, conse-
quently,
ˆ
h(s) is 1-summable in all directions in (
π
2
,
3π
2
). The 1-sum h
l
(s) is
analytic on the sector 0 < arg s < 2π and is a solution of (1.4).
In the right half plane u(t) grows faster than exponentially of any order
on the horizontal strips {t C | <t > 0, =t (π/2, π/2) mod 2π}, but on
the strips
{t C | <t > 0, =t (π/2, 3π/2) mod 2π}
it decreases faster than exponentially of any order. Hence, the functions
h
r,n
(s) :=
Z
C
n
e
st
u(t)dt, n Z,(1.5)
with C
n
a path from 0 to ++, θ (π/2+2, 3π/2+2) (see Figure 1)
are well defined and satisfy (1.4).
O
π/2 + 2
3π/2 + 2
C
n
Figure 1. Contour C
n
.
SUMMATION METHOD FOR DIFFERENCE EQUATIONS 39
The functions h
r,n
all have the asymptotic expansion
ˆ
h as s ,
π
2
<
arg s <
π
2
, uniformly on closed subsectors (cf. also [BH75]). However, it
can be shown that the h
r,n
are not 1-sums. See also [vdPS97, Chapter 11].
In order to characterize these solutions by means of their asymptotic
behaviour, we have to consider this asymptotic behaviour on regions other
than sectors, namely regions of the form
D(θ) := {s C | <{s(log s + )} > 1}, θ R,
see Figures 2-5.
-10
-5 5
10
-10
-5
5
10
Figure 2. Region D(0).
-10
-5 5
10
-10
-5
5
10
Figure 3. Region D(
π
2
).
-10
-5 5
10
-10
-5
5
10
Figure 4. Region D(π).
-10
-7.5 -5
-2.5 2.5
5 7.5
10
-10
-7.5
-5
-2.5
2.5
5
7.5
10
Figure 5. The regions ‘rotate’
clockwise with increasing θ.
40 B.L.J. BRAAKSMA, B.F. FABER, AND G.K. IMMINK
Proposition. For any θ (2, 2(n + 1)π), there exist K, A > 0 such that
h
r,n
(s)
N1
X
n=1
h
n
s
n
KA
N
(N!)|s|
N
, s D(θ), N N.(1.6)
This proposition has been proved by Borel-Laplace methods in [Fab97].
According to a theorem by Immink in [Imm96], h
r,n
is uniquely determined
by the ab ove property.
2. Preliminaries.
By Argz we denote the principal argument of z C\{0}; we take Argz
(π, π]. The Riemann surface of the logarithm will be denoted by C
.
For α, β R we denote by S(α, β) the open sector {s C
| α < arg s <
β}, and by S[α, β] the closed sector {s C
| α arg s β}. Similarly,
S[α, β) and S(α, β] denote half-open sectors. For µ Z we define
H
µ
:= S((µ 1)π/2, (µ + 1)π/2); H
µ
:= S[(µ 1)π/2, (µ + 1)π/2].
Throughout this paper, by an upper half plane, a fourth quadrant, etcetera,
we understand a lift of the upper half plane, the fourth quadrant, etcetera,
from the complex plane to the Riemann surface of the logarithm. A sector
will always be a sector of C
with vertex at the origin.
By definition, a neighbourhood of in a sector S (S not necessarily
open) is an op e n subset U of S, such that, for any closed subsector S
0
of S
with aperture π, we can find s
0
S such that s
0
+ S
0
U. In particular,
Re
iµπ/2
+ H
µ
, µ Z, R > 0, is a neighbourhood of both in H
µ
and in
H
µ
.
If we write f(s) = O(g(s)) or f(s) = o(g(s)) as s on a sector S, we
mean that f and g are functions defined on a neighbourhood U of in S,
and that the O or o relation holds uniformly, as s , on the intersection
of U and any closed subsector of S.
Similarly, if
ˆ
f(s) =
P
j0
a
j
s
j/p
where p > 0, and if S is a s ector, then
f(s)
ˆ
f(s), s on S, means the following: f is an analytic function on
a neighbourhood U of in S and for any closed subsector S
0
S and any
N N, we can find positive constants R and C
S
0
,R,N
, such that
f(s)
N1
X
j=0
a
j
s
j/p
C
S
0
,R,N
|s|
N/p
, s S
0
U, |s| > R.(2.1)
The set of such functions f with an asymptotic expansion
ˆ
f on S as above
will be denoted by A(S).
SUMMATION METHOD FOR DIFFERENCE EQUATIONS 41
In accordance with the above, when we write f(s) 0, s on S, we
mean that f(s) = o(s
N
), s on S, for any N N.
Suppose ˆu(s) =
P
m
j=0
ˆ
h
j
(s)(log s)
j
with
ˆ
h
j
C[[s
1/p
]] for j = 0, . . . , m.
If we write u(s) ˆu(s), s on S, we mean that there exist analytic
functions h
j
, j = 0, . . . , m, on a neighborhood of in S such that u(s) =
P
m
j=0
h
j
(s)(log s)
j
, with h
j
(s)
ˆ
h
j
(s), s on S.
If f A(S) such that (2.1) holds and if there exist k > 0, and K
S
0
,R
, A
S
0
,R
> 0 such that, for each N,
C
S
0
,R,N
K
S
0
,R
A
N
S
0
,R
Γ
N
pk
,
then we call f a k-Gevrey function on S with respect to the family
1
p
N
0
,
and we write f A
(1/k)
(S). Note that
1
p
N
0
is an example of a ‘convenient
family’, ac cording to the terminology introduced by Malgrange in [Mal95].
Any f A
(1/k)
(S) has an asymptotic expansion
ˆ
f(s) =
P
j=0
a
j
s
j/p
with the a
j
satisfying
|a
j
| KA
j
Γ
j
pk
, j > 0,
for some positive K and A. Such a formal series
ˆ
f will be called a Gevrey
series (in s
1
) of order 1/k with respect to the family
1
p
N
0
, and
C[[s
1/p
]]
1
pk
denotes the set of such series.
In the sequel all Gevrey functions and Gevrey series will be with respect
to the family
1
p
N
0
with p as in (1.2), and we omit the references to this
family in our notations.
A function f defined on a neighbourhood of in a sector S is exponen-
tially small of order k > 0 on S if for any closed subsector S
0
of S there
exists a p os itive constant c such that f(s) = O(e
c|s|
k
), s on S
0
. If
this holds for all positive c then f is said to be supra-exponentially small of
order k on S. The set of all analytic functions on a neighbourhood of in
S which are exponentially or supra-exponentially small of order k on S will
be denoted by A
≤−k
(S) and A
<k
(S) respectively. If S = S(α, β) then we
will also write these latter sets as A
≤−k
(α, β) and A
<k
(α, β). Similarly if
S = S(α, β] etc. If f and g both are in A
(1/k)
(S), and f and g have the
same asymptotic expansion, then it can be shown that their difference f g
is in A
≤−k
(S) (cf. [Mal95]).
Let l > 0 and S be an open sector. Let {S
i
}
iI
be a covering of S
consisting of open sectors and let f
(i)
A(S
i
), i I, such that f
(i
1
)
f
(i
2
)
A
≤−l
(S
i
1
S
i
2
) for any i
1
, i
2
I with S
i
1
S
i
2
6= . These data determine
an l-precise quasi-function on S. We identify two such sets of data
({f
(i)
}
iI
; {S
i
}
iI
) and ({g
(j)
}
jJ
; {
˜
S
j
}
jJ
) if f
(i)
g
(j)
A
≤−l
(S
i
˜
S
j
)
where i I, j J such that S
i
˜
S
j
6= . They define the same l-precise
42 B.L.J. BRAAKSMA, B.F. FABER, AND G.K. IMMINK
quasi-function on S and we write (A/A
≤−l
)(S) for the set of these l-precise
quasi-functions on S. Indeed, if we identify the interval I = (a, b) in R
and the sector S(a, b), then A and A
≤−k
, k > 0, can be considered as
sheaves on R and then A/A
≤−l
is the quotient sheaf. Also elements of
(A/A
≤−l
)
n
(S)[log s] with n N will be called l-precise quasi-functions on
S.
All elements in a representative {f
(i)
}
iI
of f (A/A
≤−l
)
n
(S)[log s] have
the same asymptotic expansion
ˆ
f and this expansion is independent of the
chosen representative. Therefore we may write f(s)
ˆ
f(s), s on S,
without causing confusion.
Similarly if l > k > 0 we define (A
(1/k)
/A
≤−l
)(S) as the set of l-precise
quasi-functions f which have representatives f
i
on S
i
as above with f
i
A
(1/k)
(S
i
) for all i I.
If
ˆ
f C[[s
1/p
]]
1
pk
then there exists a unique f (A
(1/k)
/A
≤−k
)(C
p
)
such that f
ˆ
f, where C
p
denotes the Riemann surface of s
1/p
. This f will
be denoted by T
1
ˆ
f (cf. [MR92, Cor. (1.8)]).
Let f (A/A
≤−l
)
n
(S)[log s]. The ‘restriction’ of f to an open subsector
S
0
of S, denoted by f|
S
0
, is defined as follows: Suppose {f
(i)
}
iI
is a rep-
resentative of f with respe ct to a covering {S
i
}
iI
of S. Then f|
S
0
is the
element of (A/A
≤−l
)
n
(S
0
)[log s] defined by {f
(i)
|
S
i
S
0
}
iI
, where f
(i)
|
S
i
S
0
is the restriction of f
(i)
to a neighb o urhood of in S
i
S
0
.
Definition 2. Let 0 < k
1
< · · · < k
r
, and let
ˆ
f C[[s
1/p
]]
1/(pk
1
)
. Fur-
thermore, let S
1
. . . S
r
be a nested sequence of open sectors, where
S
i
has ape rture larger than π/k
i
, i = 1, . . . , r and S
1
has ape rture at most
2. We say that
ˆ
f is (k
1
, . . . , k
r
)-summable on (S
1
, . . . , S
r
) if there exist
f
i
(A
(1/k
1
)
/A
≤−k
i+1
)(S
i
), i = 1, . . . , r 1, and f
r
A
(1/k
1
)
(S
r
), such that
f
i
|
S
i+1
f
i+1
mod A
≤−k
i+1
, i = 0 , . . . , r 1 where f
0
= T
1
ˆ
f. We call f
r
the (k
1
, . . . , k
r
)-sum of
ˆ
f on (S
1
, . . . , S
r
), and we have f
r
(s)
ˆ
f(s), s
on S
r
.
According to the ‘relative Watson lemma’ ([MR92, Prop. (2.1)]) f
i+1
is completely determined by f
i
and S
i+1
, i = 0, . . . , r 1. Hence the
(k
1
, . . . , k
r
)-sum of
ˆ
f on (S
1
, . . . , S
r
) is uniquely defined. We may
extend the definition of multisummability in an obvious way to the case
that
ˆ
f is an n-vector or an n × n-m atix with elements in C[[s
1/p
]]
1/(pk
1
)
.
Definition 3. Let f(s) = s
ds
e
2πisb+q(s)
s
γ
, with d
1
p
Z, b C, q(s) iden-
tically zero or a polynomial in s
1/p
without constant term and of degree at
most p1, and γ C. We will say that f(s) is of level 1
+
if d 6= 0, of le vel
1 if d = 0, b 6= 0, of level k with k {
1
p
, . . . ,
p1
p
} if d = b = 0, q(s) 6≡ 0 and
q(s) has degree pk in s
1/p
, and of level 0 if d = b = 0 and q(s) 0.
SUMMATION METHOD FOR DIFFERENCE EQUATIONS 43
Let f be of level k (0, 1], so d = 0, 2πisb + q(s) 6≡ 0. A closed interval
[σπ/k, σ] will be called a Stokes interval of level k of f if f A
≤−k
(σ
π/k, σ). So if k = 1 then σ π Argb mod 2π whereas if 0 < k < 1 and
q(s) = ωs
k
+ o(s
k
), s , ω 6= 0, then
3
2
π Arg ω mod 2π.
If f is of level 1
+
we will, in Section 7, associate with it a certain Stokes
number. This number is connected with curves that separate regions of
growth from regions of decay. All these curves have the limiting directions
π
2
mod π.
If f is of level 1
+
, we have f(s) = exp(ds log s(1 + o(1))), s + with
d 6= 0, and so it grows or decays faster than exponentially of order 1 on R
+
,
but slower than any higher exponential order.
Definition 4. For any (i.e., not necessarily open) sector S we will write
f A
≤−1
+
(S) to express that f is analytic on a neighbourhood U of in
S, and that for any c losed subsector S
0
of S, there exists a positive constant
c (dep e nding on S
0
) such that f(s) = O(e
c|s| log |s|
), uniformly as s on
S
0
U.
We define 1
+
-precise quasi-functions by replacing l by 1
+
in the definition
of l-precise quasi-functions above.
So, for e xample, e
ds log s
A
≤−1
+
(H
0
) if d <0. And if p(s)=
P
j0
p
j
e
2πisj
is an analytic 1-periodic function on {s C|=s > R} for some R > 0, then
p(s)e
ds log s
A
≤−1
+
(H
0
H
1
) if d < 0.
With Equation (1.1) we associate levels, and with each level certain Stokes
intervals or numbers. For this purpose we rewrite the formal fundamental
matrix solution (1.2) as follows:
ˆ
Y(s) =
ˆ
U(s)F(s),(2.2)
where
ˆ
U(s) =
ˆ
H(s)s
N
, N =
L
m
j=1
N
j
and F(s) = s
Λs
e
G(s)
s
C
, C =
L
m
j=1
c
j
I
j
.
The columns ˆy
l
(s) (l = 1, . . . , n) of
ˆ
Y(s) form a formal fundamental
system of solutions {ˆy
l
}
n
l=1
and we have
ˆy
l
(s) = f
l
(s)ˆu
l
(s), f
l
(s) = s
d
l
s
e
2πisb
l
+q
l
(s)
s
γ
l
,(2.3)
where ˆu
l
(s) C
n
[[s
1/p
]][log s] is the l-th column of
ˆ
U(s), and, furthermore,
if 0 < l (n
1
+ . . . + n
j1
) n
j
, then d
l
= λ
j
, b
l
C and q
l
(s) 0 or q
l
(s)
is a polynomial in s
1/p
without constant term and of degree at most p 1
such that 2πisb
l
+ q
l
(s) = g
j
(s) and γ
l
= c
j
. Without loss of generality, we
may assume that <b
l
[0, 1), l = 1, . . . , n.
We use the following abbreviations (cf. (2.3)): f
ml
:= f
m
f
1
l
, d
ml
:=
d
m
d
l
, b
ml
:= b
m
b
l
, q
ml
:= q
m
q
l
, γ
ml
:= γ
m
γ
l
. We write κ
ml
for
the level of f
ml
.
44 B.L.J. BRAAKSMA, B.F. FABER, AND G.K. IMMINK
Definition 5. The levels of Equation (1.1) are the levels of the functions
e
2πisj
f
ml
(s), j Z, m, l {1, . . . , n}. Let k {
1
p
, . . . ,
p1
p
, 1}. The Stokes
intervals of level k of the equation are the Stokes intervals of level k of the
functions e
2πisj
f
ml
(s), j Z, m, l {1, . . . , n}.
Taking j = 0 we see that all the κ
ml
are levels of the equation. Moreover,
0 and 1 always are levels of the equation (take m = l and then j = 0 and
j 6= 0, respectively). By 0 < k
1
< · · · < k
r
= 1 we denote the increasing
sequence of levels of the equation in the interval (0, 1]. If =b
ml
6= 0 for some
m and l then there are infinitely many Stokes directions (endpoints of Stokes
intervals) Arg(b
ml
+ j) mod π, j Z which cluster at 0 mod π.
The following theorem is the counter part in the theory of linear differ-
ence equations of the ‘main asymptotic existence’ theorem in the theory of
differential equations.
Theorem 6. Let l {1, . . . , n} and ˆy
l
(s) = f
l
(s)ˆu
l
(s) be a formal solution
of (1.1) of the form (2.3) with ˆu
l
(s) C
n
[[s
1/p
]][log s].
Then for any µ Z there exists an analytic solution y
l
(s) = f
l
(s)u
l
(s) of
the equation such that u
l
(s) ˆu
l
(s), s on H
µ
.
A proof of this theorem (for the case that no logarithmic terms appear
in ˆu(s)) can be found in [vdPS97]. It is based on the so-called quadrant
theorem, already stated by Birkhoff and Trjitzinsky in [BT33], but made
rigorous by Immink in [Imm91].
3. Two auxiliary lemmas.
The following lemma gives information on the relation between two funda-
mental systems of solutions of Equation (1.1), which have the same asymp-
totic behaviour at on some sector.
Lemma 7. Suppose we have two fundamental matrix solutions of Equation
(1.1), Y = UF and Y
1
= U
1
F, such that U(s)
ˆ
U(s) and U
1
(s)
ˆ
U(s) for
s on an open sector S, with F(s) and
ˆ
U(s) as in (2.2). Let u
l
and u
l,1
be the l-th column of U and U
1
, respectively.
Then there exist analytic 1-periodic functions p
lm
, m = 1, . . . , n, on a
neighbourhood of in S, such that
u
l
u
l,1
=
n
X
m=1
p
lm
f
ml
u
m
.
Moreover,
p
lm
(s)f
ml
(s) 0, s on S, m {1, . . . , n}.
If
u
l
u
l,1
(A
≤−k
)
n
(S) for some k > 0 (including k = 1
+
),
SUMMATION METHOD FOR DIFFERENCE EQUATIONS 45
then
p
lm
f
ml
A
≤−k
(S), m {1, . . . , n}.
Proof. Let
e
U := U U
1
. Then
e
UF = UFP, or, equivalently, FPF
1
= U
1
e
U,
for some 1-periodic analytic matrix function P = (P
ml
), on a neighbourhood
of in S. From the diagonal form of the matrix F it follows that if p
lm
is
the element in the m-th row and l-th column of P, then
p
lm
f
ml
= (m-th row of U
1
)(u
l
u
l,1
).
As U(s)
ˆ
U(s) we have U
1
(s)
ˆ
U
1
(s) = s
N
ˆ
H
1
(s). Hence any entry
of U
1
is of order O(s
µ
(log s)
ν
), s for some µ, ν Z. Since u
l
(s)
u
l,1
(s) 0 as s on S, we thus find that
p
lm
(s)f
ml
(s) 0, s on S, m {1, . . . , n}.
Similarly, we see that p
lm
f
ml
A
≤−k
(S), if u
l
u
l,1
(A
≤−k
)
n
(S) for some
k > 0, including k = 1
+
.
The next lemma yields more information on the asymptotic behaviour of
the functions p
lm
f
ml
in the previous lemma.
Lemma 8. Let S := S(α
1
, α
2
) be an open sector with 0 < α
2
α
1
π.
Let p(s) be an analytic, 1-periodic function on a neighbourhood of in
S, and let f be a function of level k {0, 1/p, . . . , 1, 1
+
} as in Section 2:
f(s) = s
ds
e
2πisb+q(s)
s
γ
with <b [0, 1). Assume g(s) := p(s)f(s) 0,
s on S. Let H be an upper or lower half plane in C
which has a
nonempty intersection with S.
Then g A
≤−k
(S) if k > 0 and g A
≤−1
(H) if k = 0. If ν denot es
some integer we have:
1) If k = 0: If α
1
< νπ < α
2
then p = g = 0.
2) If 0 < k < 1:
Then there exists c C such that g cf A
≤−1
(H) and cf
A
≤−k
(S). If α
1
< νπ < α
2
then p(s) = c. If g A
<k
(S) then
g A
≤−1
(H).
3) If k = 1:
If p 6= 0 then with H corresponds an integer N such that p(s)
p
N
e
2πiNs
as |=s| on H where p
N
6= 0. If b R
then g
A
≤−1
(H) and if moreover α
1
< νπ < α
2
then p = g = 0. If α
1
νπ α
2
and (1)
ν
=b < 0 then p = g = 0.
Next suppose
(I) α
1
νπ α
2
and (1)
ν
=b > 0.
(II) (α
1
, α
2
) (β
1
, β
2
) where (β
1
, β
2
) does not contain a Stokes interval
of e
2πisj
f(s) of level 1 for any j Z.
Then there exist analytic 1-periodic functions p
+
and p
such that
p = p
+
+ p
and p
+
f A
≤−1
(α
1
, β
2
) and p
f A
≤−1
(β
1
, α
2
). If
46 B.L.J. BRAAKSMA, B.F. FABER, AND G.K. IMMINK
α
1
= νπ then p
+
f A
≤−1
[νπ, β
2
) and similarly if α
2
= νπ then
p
f A
≤−1
(β
1
, νπ].
4) If k = 1
+
:
If α
1
< (ν +
1
2
)π < α
2
, then p = g = 0. If νπ α
1
< α
2
(ν +
1
2
)π or (ν
1
2
)π α
1
< α
2
νπ then g A
≤−1
+
[νπ, α
2
) and
g A
≤−1
+
(α
1
, νπ] respectively. If f 6∈ A
≤−1
+
(S) then p = g = 0.
5) If k 1 and g A
<1
(S) then g = 0.
Proof. We will give the proof for the cases (i) 2 < α
1
< α
2
< (2h + 1)π,
(ii) α
1
= 2 and (iii) α
1
< 2 < α
2
for some h Z. The other cases can
be treated similarly.
We may choose H = H
4h+1
. So p(s) and g(s) are analytic on H for
=s > R for some R > 0. Put z = e
2πis
and P (z) := p(s). Then |z| = e
2π=s
and we have expansions
P (z) =
X
j=−∞
p
j
z
j
if 0 < |z| < e
2πR
, p(s) =
X
j=−∞
p
j
e
2πisj
if =s > R.
(3.1)
In case (iii) p is an entire function. So then (3.1) holds with p
j
= 0 if j < 0
and R may be replaced by −∞. We now treat separately the different cases
of the lemma.
Ad 1) We have f (s) = s
γ
. Hence p(s) = s
γ
g(s) 0 as s on
S. The 1-periodicity of p(s) then implies that p(s) 0 as =s , so
P (z) 0 as z 0. Therefore p
j
= 0 if j 0 and p, g A
≤−1
(H). In case
(iii) also p(s) 0 as =s −∞, so P (z) 0 as z . Hence P = 0 and
so g = p = 0.
Ad 2) For any closed sector S
1
(S H) and any ρ > R, there exist
positive constants K and a such that
|p(s)| = |g(s)f (s)
1
| K exp(a|s|
k
), if s S
1
and =s ρ.
From the 1-periodicity of p(s) and the fact that k (0, 1) it follows that
|p(s)| K exp(a
0
=s), ∀=s ρ, for some a
0
(0, 2π), if we choose ρ suffi-
ciently large. This implies P (z) = o(z
1
), z 0. Hence p
j
= 0 if j < 0
in (3.1). With c := p
0
we get p(s) c A
≤−1
(H) and g cf A
≤−1
(H).
If f 6∈ A
≤−k
(S H) then c = 0 since otherwise g is unbounded in a neigh-
bourhood of in S. So in cases (i) and (ii) we have cf A
≤−k
(S). Also
c = 0 if g A
<k
(S) and therefore g A
≤−1
(H). In case (iii) we have
moreover P (z) = o(z) as z and so P (z) = p(s) c. If c 6= 0 then
g = cf A
≤−k
(S) as g 0 in S.
Ad 3) The fact that f is of level 1 implies that d = 0, b 6= 0.
We have p(s) = g(s)f(s)
1
= O(1) exp(2πis(b + o(1))) as s on
S H and therefore P (z) = O(z
b+o(1)
) as z 0. So if p 6= 0 then there
exists N Z such that p
j
= 0 for all j < N and p
N
6= 0 in (3.1). Then
SUMMATION METHOD FOR DIFFERENCE EQUATIONS 47
p(s) = p
N
e
2πiNs
(1+o(1)) and g(s) = p
N
e
2πi(b+N+o(1))s
(1+o(1)) as =s .
As g(s) 0 as s in S it follows that g A
≤−1
(S H) where S H = S
in cases (i) and (ii). Moreover, if =b = 0 then b+N > 0 and so g A
≤−1
(H).
In cases (ii) and (iii) we see that if p 6= 0 and =b < 0 then g(s) on
arg s = 2 + for sufficiently small p ositive. Hence p = g = 0 if =b < 0.
In case (iii) similar reasoning as above leads to p
j
= 0 for all j > M
with some M Z, M N, g(s) = e
2πi(b+M+o(1))s
(p
M
+ o(1)) as =s −∞
and g A
≤−1
(α
1
, 2). In particular, if =b = 0 and p
N
6= 0 6= p
M
, then
the fact that g 0 in S(2 , 2 + ) for some > 0 implies that
N + b > 0 > M + b in contradiction with M N, and therefore p = g = 0.
So in case (iii) we have =b 6= 0 if p 6= 0 and so =b > 0. Consequently g is
exponentially small of order 1 in S(2 , 2 + ) for some > 0. Thus
g A
≤−1
(S).
Next consider the case that (I) and (II) are satisfied. Now ν = 2h and
only cases (ii) and (iii) with =b > 0 have to be considered. Let σ
j
:=
(2h + 1)π Arg(b + j) for all j Z. Then S(σ
j
π, σ
j
) is a maximal sector
where e
2πisj
f(s) is exponentially small of order 1 and the behaviour of g on
S H implies that σ
N
π 2 < α
2
σ
N
. If σ
N
β
2
then we see that
g A
≤−1
[2, β
2
). Next suppose σ
N
< β
2
. As σ
j
increases monotonically
from 2 to (2h + 1)π as j increases from −∞ to +, there exists A Z
such that σ
A
< β
2
σ
A+1
and A N . The condition on Stokes intervals
now implies that σ
A
π β
1
.
Let p
(s) :=
P
A
j=N
p
j
e
2πisj
. Then p
(s) = O(e
2πisA
) on the lowe r
half plane and p
(s) = O(e
2πisN
) on the upper half plane. So p
f is
exponentially small of order 1 for arg s (σ
A
π, 2] and for arg s
[2, σ
N
). Hence p
f A
≤−1
(β
1
, α
2
). Furthermore, p
+
:= p p
=
P
j=A+1
p
j
e
2πisj
= O(e
2πis(A+1)
) on the upper half plane and as σ
A+1
β
2
we see that p
+
f A
≤−1
[2, β
2
). Moreover, p
+
f = gp
f A
≤−1
(α
1
, α
2
)
and we conclude that p
+
f A
≤−1
(α
1
, β
2
).
Ad 4) Now f (s) = exp{ds(log s + O(1))}, s with d 6= 0.
Since <(s log s) = <s log |s| =s arg s, we have p(s) = g(s)f(s)
1
=
O(exp{−d<s log |s| + O(s)}), as s on S. As <s/=s is a nonzero
constant on any ray arg s = ψ 6∈
π
2
Z we see that if S contains such a ray on
which d<s > 0 then p(s) = O(1) exp(N|=s|) for any N N and therefore
p = g = 0. In particular, if f 6∈ A
≤−1
+
(S) then there exists a ray where
d<s > 0, so p = g = 0.
It is now sufficient to consider the case that S belongs to a right half
plane and d < 0. Choose ε with 0 < ε < (α
2
α
1
)/2 and ε < α
2
/2 .
Let ψ (2, α
2
2ε). For any s with arg s = ψ there exists s
H with
s s
N and arg s
(α
2
ε, α
2
ε/2) if =s is sufficiently large. Then
<s
< =s cot(α
2
ε). Thus
p(s) = p(s
) = O(1) exp[d cot(α
2
ε)=s log =s + O(s)].
48 B.L.J. BRAAKSMA, B.F. FABER, AND G.K. IMMINK
From this and =s log =s = |s| sin ψ log(|s| sin ψ) = |s|(sin ψ log |s|+ O(1)) we
conclude that |p(s)| K
1
exp{−d cot( α
2
ε) sin ψ|s| log |s| + K
2
|s|} if =s
is sufficiently large where K
1
and K
2
are some positive constants. Using
g(s) = O(1)p(s) exp[−|ds|(cos ψ) log |s| + O(s)] and cos ψ (sin ψ) cot(α
2
ε) = sin(α
2
ε ψ)/ sin(α
2
ε) > sin ε/ sin(α
2
ε) =: c
ε
> 0 we see that
g(s) = O(1) exp(c
ε
|ds| log |s|)
if =s is sufficiently large and arg s (2, α
2
2ε). Thus we see that
g A
≤−1
+
[2, α
2
). In case (iii) we get similarly g A
≤−1
+
(α
1
, 2].
Furthermore, the 1-periodic function p is bounded on any bounded strip
parallel to the real axis intersected with S. Thus we obtain g A
≤−1
+
(S)
in case (iii).
Ad 5) If g(s) A
<1
(S) and k 1 we deduce p(s) = g(s)/f(s)
A
<1
(S). T herefore p(s) = O(1)e
c|s|
as s for all c > 0. So P (z) =
O(z
j
) as z 0 for all j. Hence P (z) 0 and so p = g = 0.
4. A Gevrey property of solutions
Proposition 9. Let k
1
be the lowest positive level of (1.1). Then the ele-
ments of
ˆ
H(s) are Gevrey series of order 1/k
1
. There exist fundamental
matrices Y
(µ)
(s) = H
(µ)
(s)s
Λs
e
G(s)
s
L
of (1.1) such that H
(µ)
(s) is a matrix
of k
1
-Gevrey functions on H
µ
with H
(µ)
(s)
ˆ
H(s) on H
µ
for all µ Z.
For any µ
0
Z a representative of T
1
ˆ
H (cf. definition of T
1
in Sec-
tion 2) on the covering {H
µ
| µ = µ
0
, . . . , µ
0
+ 4p 1} of C
p
is given by
{H
(µ)
| µ = µ
0
, . . . , µ
0
+ 4p 1}.
Let S be an open sector of aperture at most π and let ˆu
l
be given by (2.3)
for l = 1, . . . , n. Assume that f
l
v
l
is a solution of (1.1) such that v
l
ˆu
l
on
S for l = 1, . . . , n. Then v
l
(A
(1/k
1
)
)
n
(S)[log s]. Moreover, {f
l
v
l
}
n
l=1
is a
fundamental set of solutions of equation (1.1).
Proof. To prove the last statement, let V be the matrix with v
l
as l-th
column. Then V
ˆ
U on S, where
ˆ
U as in (2.2). As det
ˆ
U 6≡ 0, we also
have det V 6≡ 0. Thus Y := VF is a matrix solution of Equation (1.1) and
det Y 6≡ 0, i.e., it is a fundamental matrix solution.
According to Theorem 6 and the last statement of the proposition under
consideration we have fundamental matrices Y
(µ)
(s) = H
(µ)
(s)s
N
F(s), µ =
µ
0
, . . . , µ
0
+ 4p 1, with H
(µ)
(s)
ˆ
H(s), s on H
µ
,
ˆ
H(s) as in (1.2).
Since e
2i
H
µ
= H
µ+4p
we define H
(µ
0
+4p)
(s) = H
(µ
0
)
(se
2i
), s H
µ
0
+4p
.
Then H
(µ
0
+4p)
(s)
ˆ
H(se
2i
) =
ˆ
H(s), s on H
µ
0
+4p
. If s H
µ
0
+4p
,
ζ := se
2i
H
µ
0
, then s + 1 = (ζ + 1)e
2i
and
Y
(µ
0
+4p)
(s) := H
(µ
0
+4p)
(s)s
N
F(s)
= H
(µ
0
)
(ζ)(ζe
2i
)
N
F(ζe
2i
) = Y
(µ
0
)
(ζ)P(ζ),
SUMMATION METHOD FOR DIFFERENCE EQUATIONS 49
where P(ζ) = e
2i(ζΛ+L)
, with Λ and L as in (1.2), is a 1-periodic matrix
function, and det P (ζ) 6= 0. Hence Y
(µ
0
+4p)
(s) is a fundamental matrix.
Next we prove that the entries of H
(µ)
(s) are in A
(1/k
1
)
(H
µ
), µ = µ
0
, . . . ,
µ
0
+4p. As the half planes H
µ
, µ = µ
0
, . . . , µ
0
+4p1, cover a neighbourhood
of on the Riemann surface of z
1/p
, it is, by [MR92, Theorem 1.6], suffi-
cient to prove that the entries of H
(µ+1)
(s) H
(µ)
(s) are exponentially small
of order k
1
on H
µ
H
µ+1
, µ = µ
0
, . . . , µ
0
+4p1. If we denote by u
(µ)
l
(s) the
l-th column of H
(µ)
(s)s
N
, l = 1, . . . , n, µ = µ
0
, . . . , µ
0
+4p, this is equivalent
to proving that the differences u
(µ+1)
l
u
(µ)
l
are in (A
≤−k
1
)
n
(H
µ
H
µ+1
).
We have
u
(µ+1)
l
u
(µ)
l
=
n
X
m=1
p
lm
f
ml
u
(µ)
m
,
for some 1-periodic functions p
lm
on a neighbourhood of in H
µ
H
µ+1
.
Since u
(µ+1)
l
(s)u
(µ)
l
(s) 0, as s on H
µ
H
µ+1
, we have p
lm
(s)f
ml
(s)
0, s on H
µ
H
µ+1
, for m = 1, . . . , n, according to Lemma 7. Lemma 8
now yields that p
lm
f
ml
A
≤−k
1
(H
µ
H
µ+1
), m = 1, . . . , n, so
u
(µ+1)
l
u
(µ)
l
(A
≤−k
1
)
n
(H
µ
H
µ+1
).
Applying [MR92, Theorem 1.6], we conclude that
u
(µ)
l
(A
(1/k
1
)
)
n
(H
µ
)[log s], l = 1, . . . , n, µ = µ
0
, . . . , µ
0
+ 4p 1,
H
(µ)
is a k
1
-Gevrey function on H
µ
and the elements of
ˆ
H(s) are Gevrey se ries
of order 1/k
1
. Moreover, it follows that {H
(µ)
| µ = µ
0
, . . . , µ
0
+ 4p 1} is
a representative of T
1
ˆ
H.
Finally we prove the statement concerning the functions v
l
(s). It is suf-
ficient to consider the case that S (H
µ
0
H
µ
0
+1
). There exist 1-periodic
functions ˜p
lm
analytic on a neighbourhood of in S H
µ
0
, such that
v
l
u
(µ
0
)
l
=
n
X
m=1
˜p
lm
f
ml
u
(µ
0
)
m
.
Since v
l
(s) u
(µ
0
)
l
(s) 0, as s on S H
µ
0
, Lemma 7 and Lemma 8
now tell us that ˜p
lm
f
ml
A
≤−k
1
(S H
µ
0
), hence
v
l
u
(µ
0
)
l
(A
≤−k
1
)
n
(S H
µ
0
).
The same holds with µ
0
replaced by µ
0
+ 1. It follows that v
l
(A
(1/k
1
)
)
n
(S)[log s], what had to be proven.
50 B.L.J. BRAAKSMA, B.F. FABER, AND G.K. IMMINK
5. Refinement of chains of solutions.
Consider the fundamental matrix Y
(µ)
(s) of Proposition 9 for µ Z. Let
its c olumns be denoted by {f
l
u
(µ)
l,0
} for l = 1, . . . , n as in (2.3). Then u
(µ)
l,0
(A
(1/k
1
)
)
n
(H
µ
)[log s], u
(µ)
l,0
ˆu
l
on H
µ
and {u
(µ)
l,0
| µ = µ
0
, . . . , µ
0
+ 4p
1} represents the k
1
-precise quasi-function corresponding to T
1
ˆu
l
on the
covering {H
µ
| µ = µ
0
, . . . , µ
0
+ 4p 1} of C
p
for any µ
0
Z.
In this section we show how these k
1
-precise quasi-functions can be refined
to k
2
-precise quasi-functions with representatives {u
(µ)
l,1
} on an open sector
S(α
1
, β
1
) with aperture > π/k
1
such that (α
1
, β
1
) does not contain a Stokes
interval of level k
1
of the equation and such that {f
l
u
(µ)
l,1
}
n
l=1
again is a
fundamental system of (1.1) with the same asymptotic expansion as before.
This will be done by expressing the differences f
l
(u
(µ)
l,0
u
(µ+1)
l,0
) in terms of
suitable fundamental systems and distributing the terms that are f
l
times
an exponentially small factor of order k
1
, over u
(µ)
l,0
and u
(µ+1)
l,0
. The same
method can be applied to proceed from k
j
-precise quasi-functions u
(µ)
l,j1
to k
j+1
-precise quasi-functions u
(µ)
l,j
corresponding to solutions f
l
u
(µ)
l,j1
and
f
l
u
(µ)
l,j
of (1.1).
Proposition 10. Let 0 < k
1
< · · · < k
r
= 1 be the levels in (0, 1] of (1.1).
Let j {1, . . . , r}, and define k := k
j
. If j < r, then k
0
:= k
j+1
, otherwise
k
0
:= 1
+
. Let (α, β) be an open interval of length > π/k not containing a
Stokes interval of level k of (1.1). Let M and N be the integers, such that
(M 1)
π
2
α < M
π
2
< N
π
2
< β (N + 1)
π
2
. Define Γ
µ
:= H
µ
S(α, β)
for µ = M, . . . , N.
Suppose that we have fundamental systems of solutions {f
l
u
(µ)
l
}
n
l=1
on Γ
µ
for µ = M, . . . , N which satisfy for l = 1, . . . , n:
(i) u
(µ)
l
(A
(1/k
1
)
)
n
µ
)[log s], u
(µ)
l
(s) ˆu
l
(s), s on Γ
µ
,
(ii) u
(µ+1)
l
u
(µ)
l
(A
≤−k
)
n
(H
µ
H
µ+1
).
Then there exist fundamental systems of solutions {f
l
˜u
(µ)
l
}
n
l=1
for µ =
M, . . . , N such that for l = 1, . . . , n:
˜u
(µ)
l
u
(µ)
l
(A
≤−k
)
n
µ
), if µ {M, . . . , N }(5.1)
and
˜u
(µ+1)
l
˜u
(µ)
l
(A
≤−k
0
)
n
(H
µ
H
µ+1
), if µ {M, . . . , N 1}.(5.2)
Moreover, for each l {1, . . . , n} the family of functions {˜u
(µ)
l
}
N
µ=M
defines
an element ˜u
l
in (A
(1/k
1
)
/A
≤−k
0
)
n
(α, β)[log s], which is uniquely determined
by the properties of the ˜u
(µ)
l
mentioned above.
SUMMATION METHOD FOR DIFFERENCE EQUATIONS 51
We prove the proposition subsequently for the cases k (0, 1) and k = 1.
Proof for k (0, 1).
We introduce the following sets:
St
(µ, l) = {m | κ
ml
= k, f
ml
A
≤−k
(α, (µ+1)π/2)} if µ {M 1, . . . , N
1};
St
+
(µ, l) = {m | κ
ml
= k, f
ml
A
≤−k
(µπ/2, β)} if µ {M, . . . , N}. Obvi-
ously
St
(µ + 1, l) St
(µ, l), and St
+
(µ 1, l) St
+
(µ, l).(5.3)
Because of the assumption that β α > π/k the two sets St
(µ, l) and
St
+
(µ, l) are disjoint. Since f
jl
= f
jm
f
ml
the following transitivity relation
holds:
j St
(µ, m) m St
(µ, l) j St
(µ, l).(5.4)
Finally, let St(µ, l) := St
(µ, l) St
+
(µ, l), µ = M, . . . , N 1. If µ
{M, . . . , N 1} then
κ
ml
= k and f
ml
A
≤−k
(H
µ
H
µ+1
) m St(µ, l).(5.5)
We only give the proof that the left statement implies the right one since the
converse is trivial. The left-hand side implies that f
ml
has a Stokes interval
[σ π/k, σ] c ontaining (µπ/2, (µ + 1)π/2). Because of the assumptions of
the proposition we have either α < σπ/k < β σ or σ π/k α < σ < β
and therefore m St(µ, l).
For k < 1 the first statement of Proposition 10 is an easy consequence of
the following two lemmas.
Lemma 11. Under the assumptions of Proposition 10 with k < 1 there exist
fundamental systems {f
l
u
(µ)
l,1
}
n
l=1
, µ = M, . . . , N, satisfying:
(i) u
(µ)
l,1
u
(µ)
l
(A
≤−k
)
n
µ
);
(ii) u
(µ+1)
l,1
u
(µ)
l,1
=
P
mSt
(µ,l)
c
(µ)
m
f
ml
u
(µ)
m,1
+ ψ
(µ)
l,1
, where c
(µ)
m
C and
ψ
(µ)
l,1
(A
≤−k
0
)
n
(H
µ
H
µ+1
).
Proof. The proof goes by induction on µ. Define u
(M)
m,1
:= u
(M)
m
, m =
1, . . . , n. Next assume u
(M)
m,1
, . . . , u
(µ)
m,1
have been defined for all m = 1, . . . , n
and some µ {M, . . . , N 1}. In the remaining part of this section m will
always be understood to be in {1, . . . , n}. Fix l {1, . . . , n}.
From condition (i) of the prop osition it follows that u
(µ)
m,1
ˆu
m
on Γ
µ
.
Thus the functions f
m
u
(µ)
m,1
, m St
(µ, l), together with the functions
f
m
u
(µ+1)
m
, m 6∈ St
(µ, l), form a fundamental s ystem of solutions according
52 B.L.J. BRAAKSMA, B.F. FABER, AND G.K. IMMINK
to Proposition 9. Hence, there exist 1-pe riodic analytic functions p
(µ)
m
, p
(µ+1)
m
,
1 m n, such that
(5.6) u
(µ+1)
l
u
(µ)
l,1
=
X
mSt
(µ,l)
p
(µ)
m
f
ml
u
(µ)
m,1
+
X
mSt
+
(µ,l)
p
(µ+1)
m
f
ml
u
(µ+1)
m
+
X
m6∈St(µ,l)
p
(µ+1)
m
f
ml
u
(µ+1)
m
.
We have u
(µ+1)
l
u
(µ)
l,1
= u
(µ+1)
l
u
(µ)
l
+ u
(µ)
l
u
(µ)
l,1
(A
≤−k
)
n
(H
µ
H
µ+1
),
and thus (by Lemma 7) we may conclude that each term in the sums on
the right-hand side of (5.6) belongs to this set. Next we apply Lemma 8 to
these terms, and find:
If m St
(µ, l) (resp. m St
+
(µ, l)): Then f
ml
is of level k < 1,
and it is an element of A
≤−k
(H
µ
H
µ+1
). So there exist complex
constants c
(µ)
m
(resp. d
(µ+1)
m
) such that (p
(µ)
m
(s) c
(µ)
m
)f
ml
(s) (resp.
(p
(µ+1)
m
(s) d
(µ+1)
m
)f
ml
(s)) belong to A
≤−1
(H), where H is the upper
or lower half plane containing H
µ
H
µ+1
.
If m 6∈ St(µ, l): Then κ
ml
< k or κ
ml
= k with f
ml
6∈ A
≤−k
(H
µ
H
µ+1
)
or κ
ml
> k with in all cases p
(µ+1)
m
f
ml
A
≤−k
(H
µ
H
µ+1
). In the
last case Lemma 8 implies p
(µ+1)
m
f
ml
A
≤−k
0
(H
µ
H
µ+1
). This also
follows in the first two cases from Lemma 8-2 with c = 0.
So we have
u
(µ+1)
l
u
(µ)
l,1
=
X
mSt
(µ,l)
c
(µ)
m
f
ml
u
(µ)
m,1
+
X
mSt
+
(µ,l)
d
(µ+1)
m
f
ml
u
(µ+1)
m
+ ψ
(µ)
l,1
,
with ψ
(µ)
l,1
a function in (A
≤−k
0
)
n
(H
µ
H
µ+1
). Obviously, if we define
u
(µ+1)
l,1
:= u
(µ+1)
l
X
mSt
+
(µ,l)
d
(µ+1)
m
f
ml
u
(µ+1)
m
,
then u
(µ+1)
l,1
satisfies the requirements.
So, if we have constructed u
(λ)
m,1
for m {1, . . . , n} and λ {M, M +
1, . . . , µ} (µ N 1), then we can construct u
(µ+1)
l,1
for each l {1, . . . , n}
and the lemma follows by induction on µ.
We next refine the solutions of the previous lemm a to solutions which
satisfy (5.1) and (5.2) in Proposition 10.
Lemma 12. Let k < 1. Suppose the assumptions of Proposition 10 hold,
and furthermore, assume that u
(µ+1)
l
u
(µ)
l
=
P
mSt
(µ,l)
c
(µ)
lm
f
ml
u
(µ)
m
+ψ
(µ)
l
,
for some constants c
(µ)
lm
and a function ψ
(µ)
l
(A
≤−k
0
)
n
(H
µ
H
µ+1
).
SUMMATION METHOD FOR DIFFERENCE EQUATIONS 53
Then there exist fundamental systems {f
l
˜u
(µ)
l
}
n
l=1
, µ = M, . . . , N , such
that
(i) ˜u
(µ)
l
u
(µ)
l
=
P
mSt
(µ,l)
˜c
(µ)
lm
f
ml
u
(µ)
m
for some constants ˜c
(µ)
lm
;
(ii) ˜u
(µ+1)
l
˜u
(µ)
l
(A
≤−k
0
)
n
(H
µ
H
µ+1
).
Proof. This lemma can also be proven by induction, but this time we start
from the other end of the covering {Γ
µ
}
N
µ=M
of S(α, β): define ˜u
(N)
m
:= u
(N)
m
for all m {1, . . . , n}. Next suppose ˜u
(N)
m
, . . . , ˜u
(µ+1)
m
have been defined
and possess the properties of the lemma for all m {1, . . . , n} and some
µ {M, . . . , N 1}. Let l {1, . . . , n}. We have
˜u
(µ+1)
l
u
(µ)
l
= ˜u
(µ+1)
l
u
(µ+1)
l
+ u
(µ+1)
l
u
(µ)
l
=
X
mSt
(µ+1,l)
˜c
(µ+1)
lm
f
ml
u
(µ+1)
m
+
X
mSt
(µ,l)
c
(µ)
lm
f
ml
u
(µ)
m
+ ψ
(µ)
l
.
Furthermore,
u
(µ+1)
m
u
(µ)
m
=
X
jSt
(µ,m)
c
(µ)
mj
f
jm
u
(µ)
j
+ ψ
(µ)
m
.
From these two relations and properties (5.3) and (5.4) we obtain
˜u
(µ+1)
l
u
(µ)
l
=
X
mSt
(µ,l)
˜c
(µ)
lm
f
ml
u
(µ)
m
+
˜
ψ
(µ)
l
,
where the ˜c
(µ)
lm
are constants in C and
˜
ψ
(µ)
l
(A
≤−k
0
)
n
(H
µ
H
µ+1
).
If we define
˜u
(µ)
l
:= u
(µ)
l
+
X
mSt
(µ,l)
˜c
(µ)
lm
f
ml
u
(µ)
m
,
then ˜u
(µ)
l
satisfies the requirements of the lemma. Again the lemma follows
by induction on µ.
The first statement of Proposition 10 follows from the previous lemmas
and the last statement is a direct consequence of the relative Watson lemma
referred to after De finition 2. We could also prove the uniqueness directly,
without reference to this lemma, along the same lines as in [BIS].
Proof for k = 1.
Recall that the Stokes intervals of level 1 of Equation (1.1) with for-
mal solution (2.3) are, by definition, the Stokes intervals of the functions
e
2πisj
f
ml
(s), j Z, l, m {1, . . . , n}, that are of level 1. Taking l = m and
54 B.L.J. BRAAKSMA, B.F. FABER, AND G.K. IMMINK
j 6= 0, we find that [(h1)π, ] is a Stokes interval of level 1 of the equation
for any h Z. Hence, due to the assumption that ( α, β) does not contain a
Stokes interval of level 1, the sector S(α, β) does not contain a lift of both the
positive and the negative real axis. From this and α < Mπ/2 < N π/2 < β
it follows that N M 2 and at least one of the integers M and N has to
be odd.
We will prove the proposition for the cases M = 1 and N = 0 or N = 1,
that is α [π, π/2), β (0, π]. The other cases can be proven similarly.
Let l {1, . . . , n}. There exist 1-periodic analytic functions p
(1)
m
(s),
m = 1, . . . , n, such that
u
(0)
l
u
(1)
l
=
n
X
m=1
p
(1)
m
f
ml
u
(1)
m
.(5.7)
By assumption (ii) of Proposition 10, we have u
(0)
l
u
(1)
l
(A
≤−1
)
n
(π/2, 0).
So, by Lemma 7, p
(1)
m
f
ml
A
≤−1
(π/2, 0), m {1, . . . , n}. From Lemma 8
we conclude that
p
(1)
m
f
ml
A
≤−1
(π, 0) if κ
ml
< 1,(5.8)
and
p
(1)
m
f
ml
A
≤−1
+
(π/2, 0) if κ
ml
= 1
+
.(5.9)
If κ
ml
= 1 and p
(1)
m
6= 0 then according to Lemma 8-3 we have =b
ml
0.
Moreover, this lemma tells us that
p
(1)
m
f
ml
A
≤−1
(π, 0) if κ
ml
= 1, =b
ml
= 0.(5.10)
Let t
1
, . . . , t
ν
denote the numbers =b
h
, h {1, . . . , n} in decreasing order
of magnitude. We will use induction on τ {1, . . . , ν}.
If =b
l
= t
1
, then =b
ml
0 for all m {1, . . . , n}. So if κ
ml
= 1 and
p
(1)
m
6= 0 then we know already that =b
ml
0 and so =b
ml
= 0 and (5.10)
applies. Therefore if =b
l
= t
1
we define
˜u
(1)
l
= u
(1)
l
+
X
κ
ml
1
p
(1)
m
f
ml
u
(1)
m
, ˜u
(0)
l
= u
(0)
l
.
Then from (5.7), (5.8), (5.9) and (5.10) it follows that f
l
˜u
(µ)
l
are solutions
of (1.1) with
˜u
(µ)
l
u
(µ)
l
(A
≤−1
)
n
µ
) if µ {−1, 0};
˜u
(0)
l
˜u
(1)
l
(A
≤−1
+
)
n
(π/2, 0).
(5.11)
Next let τ {2, . . . , ν}, =b
l
= t
τ
and suppose that for all m I(l) :=
{m {1, . . . , n}|κ
ml
= 1, =b
ml
> 0} the functions ˜u
(1)
m
and ˜u
(0)
m
have
SUMMATION METHOD FOR DIFFERENCE EQUATIONS 55
already been defined such that (5.11) holds with l replaced by m. We have
u
(0)
l
u
(1)
l
=
X
m6∈I(l)
˜p
(1)
m
f
ml
˜u
(1)
m
+
X
mI(l)
˜p
(1)
m
f
ml
˜u
(1)
m
,
for some 1-periodic functions ˜p
(1)
m
(s) analytic on a neighbourhood of in
H
1
. As before we have
˜p
(1)
m
f
ml
A
≤−1
(π/2, 0).(5.12)
Now we have analogues of (5.8), (5.9) and (5.10), and so the first sum can
be written as ϕ
(1)
l
+ ψ
(1)
l
such that ϕ
(1)
l
(A
≤−1
)
n
(H
1
) and ψ
(1)
l
(A
≤−1
+
)
n
(π/2, 0) and f
l
ϕ
(1)
l
and f
l
ψ
(1)
l
are solutions of (1.1).
Next consider the case that m I(l) and ˜p
(1)
m
6= 0. Then according to
Lemma 8-3 there exist analytic 1-periodic functions p
±
m
such that ˜p
(1)
m
=
p
m
+ p
+
m
and p
m
f
ml
A
≤−1
(α, 0) and p
+
m
f
ml
A
≤−1
(π/2, β). Now define
˜u
(1)
l
:= u
(1)
l
+ ϕ
(1)
l
+
X
mI(l)
p
m
f
ml
˜u
(1)
m
,
˜u
(0)
l
:= u
(0)
l
X
mI(l)
p
+
m
f
ml
˜u
(0)
m
.
Then ˜u
(0)
l
˜u
(1)
l
= ψ
(1)
l
+
P
mI(l)
p
+
m
f
ml
(˜u
(1)
m
˜u
(0)
m
) (A
≤−1
+
)
n
(π/2, 0)
and it follows that the functions ˜u
(µ)
l
satisfy (5.11). By induction (5.11) fol-
lows for all l. So in case N = 0 the proposition has be en proved.
Next suppose that N = 1, so β > π/2. If p
(1)
m
(s), m = 1, . . . , n, are the
1-periodic functions analytic on a neighbourhood of in the upper half
plane such that
u
(1)
l
˜u
(0)
l
=
n
X
m=1
p
(1)
m
f
ml
u
(1)
m
,
then p
(1)
m
f
ml
A
≤−1
(0, π/2). If κ
ml
< 1 then as before p
(1)
m
f
ml
A
≤−1
(H
1
).
Next suppose κ
ml
= 1 and p
(1)
m
6= 0. Then by Lemma 8-3 there exists an
integer N such that p
(1)
m
(s) = p
N
e
2πisN
(1 + o(1)) as =s with p
N
6= 0
and therefore e
2πisN
f
ml
(s) A
≤−1
(0, π/2). So there exists a Stokes interval
[σ
m
π, σ
m
] of e
2πisN
f
ml
which contains (0, π/2). Now σ
m
π π/2 > α
and therefore σ
m
> β. Hence p
(1)
m
f
ml
A
≤−1
(0, β). Moreover, if κ
ml
= 1
+
,
then p
(1)
m
f
ml
A
≤−1
+
(0, π/2) according to Lemma 8. Thus, if we define
˜u
(1)
l
:= u
(1)
l
P
κ
ml
1
p
(1)
m
f
ml
u
(1)
m
, then
˜u
(1)
l
u
(1)
l
(A
≤−1
)
n
(0, β); ˜u
(1)
l
˜u
(0)
l
(A
≤−1
+
)
n
(0, π/2).
56 B.L.J. BRAAKSMA, B.F. FABER, AND G.K. IMMINK
The fundamental systems {f
l
˜u
(µ)
l
}
n
l=1
, µ = 1, 0, 1, thus obtained satisfy
(5.1) and (5.2).
The uniqueness property of Proposition 10 is an immediate consequence
of a more general form of the relative Watson lemma by Malgrange and
Ramis, that can be found in [BIS].
6. Equations without level 1
+
.
Theorem 13 has already been stated and proven in [BF96], but here we
present a new proof.
Theorem 13. Let 0 < k
1
< · · · < k
r
= 1 be the levels in (0, 1] of Equation
(1.1), and suppose that this equation does not contain a level 1
+
(i.e., d
ml
=
0, m, l {1, . . . , n}). Let
ˆ
H(s)s
Λs
e
G(s)
s
L
be a formal fundamental matrix
as in (1.2).
Let S
i
= S(α
i
, β
i
), i = 1, . . . , r, be a sequence of open sectors such that
S
1
. . . S
r
, S
1
has aperture less than 2, S
i
has aperture larger than
π/k
i
and (α
i
, β
i
) does not contain a Stokes interval of level k
i
, i = 1, . . . , r.
Then
ˆ
H is (k
1
, . . . , k
r
)-summable on (S
1
, . . . , S
r
) with sum H
r
such that
H
r
(s)s
Λs
e
G(s)
s
L
is an analytic fundamental matrix of (1.1).
Proof. Define M
j
, N
j
, j = 1, . . . , r, to be the integers such that (M
j
1)π/2 α
j
< M
j
π/2 < N
j
π/2 < β
j
(N
j
+ 1)π/2 and let Γ
j,µ
:= S
j
H
µ
,
µ = M
j
, . . . , N
j
, j = 1, . . . , r. Also, let S
0
:= C
p
be the Riemann surface
of s
1/p
, Γ
0
:= H
µ
, µ = M
0
, . . . , N
0
where M
0
:= M
1
, N
0
:= M
0
+ 4p 1.
By Proposition 9 we have a representative {H
(µ)
(s)}
N
0
µ=M
0
of T
1
ˆ
H(s) on
the covering {H
µ
}
N
0
µ=M
0
of C
p
such that H
(µ)
(s)s
Λs
e
G(s)
s
L
is an analytic
fundamental matrix of (1.1). To show that the columns
ˆ
h
l
of
ˆ
H are multi-
summable we have to construct h
l,j
(A
(1/k
1
)
/A
≤−k
j+1
)
n
(S
j
), j = 0, . . . , r
such that h
l,j
|
S
j+1
h
l,j+1
mod A
≤−k
j+1
, j = 0, . . . , r if k
r+1
= .
Let U
(µ)
0
(s) := H
(µ)
(s)s
N
so that U
(µ)
0
(s)F(s) is a fundamental matrix of
(1.1) (cf. (2.2)) and let ˜u
(µ)
l,0
denote the lth column of U
(µ)
0
. The construction
mentioned above is equivalent to the construction of functions {˜u
(µ)
l,j
}
N
j
µ=M
j
for j = 1, . . . , r and l = 1, . . . , n such that:
(1) {f
l
˜u
(µ)
l,j
}
n
l=1
is a fundamental system of Equation (1.1),
(2) {˜u
(µ)
l,j
}
N
j
µ=M
j
represents a k
j+1
-precise quasi-function
˜u
l,j
(A
(1/k
1
)
/A
≤−k
j+1
)
n
(S
j
)[log s], j = 1, . . . , r,
(3) ˜u
l,j1
|
S
j
˜u
l,j
mod (A
≤−k
j
)
n
, j = 1, . . . , r.
Suppose we have constructed ˜u
l,i
for i = 0, . . . , j 1, for some j
{1, . . . , r}. Then we can apply Prop os ition 10 with α = α
j
, β = β
j
, and
SUMMATION METHOD FOR DIFFERENCE EQUATIONS 57
with u
(µ)
l
= ˜u
(µ)
l,j1
|
Γ
j,µ
, µ = M
j
, . . . , N
j
. Defining ˜u
(µ)
l,j
:= ˜u
(µ)
l
, l = 1, . . . , n,
µ = M
j
, . . . , N
j
we see that properties (1), (2) and (3) are satisfied for i = j
as well. So they are satisfied for all j {1, . . . , r}.
We have ˜u
(µ+1)
l,r
˜u
(µ)
l,r
(A
≤−1
+
)
n
r,µ
Γ
r,µ+1
). We also have ˜u
(µ+1)
l,r
˜u
(µ)
l,r
=
P
n
m=1
p
m
f
ml
˜u
(µ)
m,r
for some 1-periodic analytic functions p
m
. Lemma 7
now tells us that each p
m
f
ml
A
≤−1
+
µ,r
Γ
µ+1,r
), and then it follows
from Lemma 8 and the fact that the equation has no level 1
+
, that p
m
= 0,
for all m. Hence, the functions ˜u
(µ)
l,r
, µ = M
r
, . . . , N
r
, are the restrictions of
an analytic function ˜u
l,r
(A
(1/k
1
)
)
n
(S
r
)[log s].
7. Equations with level 1
+
.
In this section we will consider Equation (1.1) under the assumption that
there does exist a pair (m, l) such that d
m
6= d
l
in the notation of (2.3); that
is, the equation possesses the level 1
+
. We will show in this section that
we can still assign a uniquely characterizable fundamental system Y(s) with
asymptotic expansion
ˆ
Y(s) in appropriate regions of the Riemann surface of
the logarithm, provided <b
l
6= <b
m
if b
l
6= b
m
(cf. notation in (2.3)).
Before we state the main result of this paper (Theorem 18), we need to
define the Stokes numbers of level 1
+
of the equation. The Stokes number
of a function f of level 1
+
of the form
f(s) = exp(ds log s + 2πibs + q(s) + γ log s), with d 6= 0,(7.1)
occurring in formal solutions of equations possessing a level 1
+
, is associated
with curves that separate regions of growth from regions of decay of f . We
have
<{ds log s + 2πibs} = d<s log |s| (d arg s + 2π<b)=s 2π=b<s,
and therefore the main contribution to |f(s)| comes from exp[<{s(d log s +
2πi<b)}]. Let h be an even integer if d < 0 and an odd integer if d > 0.
Then f A
≤−1
+
(H
2h
). Moreover, f be haves as an exponential function
of order 1 in vertical strips, f becomes exponentially large on any open
sector containing H
2h
and, if S
+
(h) := S((h
1
2
)π, (h +
1
2
)π] and S
(h) :=
S[(h
1
2
)π, (h +
1
2
)π) then it is easily verified that
f A
≤−1
(S
±
) iff ± (h + 2<b/d) < 1/2,
where the upper (lower) signs belong together.
Let {f
j
u
j
}
n
j=1
be a fundamental system of (1.1) such that u
j
ˆu
j
as
s on S
±
(h). Assume d
ml
= d
m
d
l
< 0 and h is even. Then there exists
N
±
Z such that ±{h + 2(<b
ml
+ N
±
)/d
ml
} < 1/2. Then e
2πisN
±
f
ml
A
≤−1
(S
±
) and therefore the solutions f
l
u
l
and f
l
u
l
+ e
2πisN
±
f
m
u
m
have the
same asymptotic behaviour on S
±
(h). So in this case it is not possible to
58 B.L.J. BRAAKSMA, B.F. FABER, AND G.K. IMMINK
characterize fundamental systems Y(s) by their asymptotic behaviour
ˆ
Y(s)
on S
±
(h).
In order for a fundamental system to be in some way uniquely determined
by its asymptotic expansion in a sector of C
, which contains an open right
or left half plane, this sector therefore should contain the closure of this
half plane. However, such fundamental systems do not exist in general (see
[vdPS97, Chapter 11]).
Hence we have to characterize fundamental systems by their asymptotic
behaviour in a more complicated type of region. This region should contain
a neighbourhood of in som e half plane H
2h
but not a neighborhood of
in H
2h
. In the case d = (1)
h+1
a suitable region is given by
Definition 14. For θ R, h Z, we define
D(h; θ) := {s C
| (h 1)π < arg s < (h + 1)π;
(1)
h
<{s(log s + )} > 1}.
If h Z and θ
1
, θ
2
R then D(h; θ
1
, θ
2
) := D(h; θ
1
) D(h; θ
2
).
We denote the boundary of D(h; θ) by C(h; θ). We have D(h; θ)
S
1
j=1
H
2h+j
and se
ihπ
D(h; θ) s D(0; θ + ). Similarly with D
replaced by C. Details on C(0; θ) can be found in [Imm84] and [Imm91].
We have <s = O(=s/ log |s|) as s on C(h; θ). This implies that
arg s ±π/2 mod 2π as =s ±∞, s C(h; θ).
If h Z and θ R, then the following properties are easily established.
Let
˜
θ < θ. As (1)
h
<{s(log s + i
˜
θ)} = (1)
h
<{s(log s + )}
(1)
h
(
˜
θ θ)=s, it follows that:
i) D(h;
˜
θ) H
2h1
D(h; θ) H
2h1
, and that
ii) D(h;
˜
θ) H
2h+1
D(h; θ) H
2h+1
.
One could say that the regions D(h; θ) ‘rotate’ (modulo some de-
formation) clockwise with increasing θ.
D(h; θ) H
2h
is a neighbourhood of in H
2h
. However, since for any
s
0
H
2h
, s
0
+ H
2h
6⊂ D(h; θ), the intersection D(h; θ) H
2h
is not a
neighbourhood of in H
2h
.
Let D := D(h; θ
1
, θ
2
), for some h Z and θ
1
< θ
2
. A set U D is called
a neighbourhood of in D, if, for any θ (θ
1
, θ
2
), there exists s
0
D
such that s
0
+ D(h; θ) U.
Suppose g is an analytic function on a neighbourhood U of in D, and
suppose there exist a k > 0 and a series ˆg(s) =
P
n0
g
n
s
n/p
such that for
any θ (θ
1
, θ
2
) we have
g(s)
N1
X
n=0
g
n
s
n/p
KA
N
Γ
N
pk
|s|
N/p
,(7.2)
SUMMATION METHOD FOR DIFFERENCE EQUATIONS 59
s D(h; θ) U, N N,
for some positive constants K and A, which only depend on θ. Then we call
g a k-Gevrey function on D (with respect to
1
p
N
0
), and the set of these
functions is denoted by A
(1/k)
(D).
Concerning 1-Gevrey functions on D we have the following theorem by
Immink (cf. [Imm96]):
Theorem 15. Suppose θ
1
, θ
2
R, θ
1
< θ
2
, and h Z. Let D :=
D(h; θ
1
, θ
2
).
If g A
(1)
(D), then g is uniquely determined by its asymptotic expansion
ˆg.
In [Imm96] it is shown that g is already uniquely determined by its
asymptotic expansion if (7.2) with k = 1 holds on D(h; θ) for one θ (θ
1
, θ
2
).
Let f be a function of level 1
+
given by (7.1) and θ
1
< 2π<b/d < θ
2
.
Let h Z be odd if d > 0 and even if d < 0. Then there exist positive
constants K and c such that |f(s)| Ke
c|s|
for s D(h; θ
1
) D(h; θ
2
) and
|s| sufficiently large. From this it follows that (7.2) holds on this set with
g := f, g
n
= 0, and k = 1. This is not true if
2π<b
d
6∈ [θ
1
, θ
2
]. Therefore we
introduce the following definition:
Definition 16. Let f(s) be the function of level 1
+
given by (7.1). Then
we call
2π<b
d
its Stokes number. The Stokes numbers of level 1
+
of
the equation (1. 1) are the Stokes numbers of the functions e
2πijs
f
ml
(s),
l, m {1, . . . , n}, j Z, that are of level 1
+
. That is, they are given by the
expression
2π
d
ml
(<b
ml
+ j), j, l, m as above, d
ml
6= 0.
Suppose θ
1
< θ
2
and α < β. Let D := D(h; θ
1
, θ
2
) and S := S(α, β).
Assume D S 6= . We define a neighbourhood of in D S to be an
open set U in D S, such that for any θ, γ, δ satisfying θ
1
< θ < θ
2
and
α < γ < δ < β, there exists s
0
D S such that s
0
+ (D(h; θ) S(γ, δ))
U. We write A
≤−1
(D S) for the set of functions that are analytic on a
neighbourhood of in D S, and exponentially small of order 1, as s ,
uniformly on D(h; θ) S(γ, δ), for any θ, γ, δ as above.
We extend Definition 2 of multisummability as follows:
Definition 17. Let 0 < k
1
< · · · < k
r1
< k
r
= 1, and define k
r+1
= 1
+
.
Let S
1
. . . S
r
be a nested sequence of sectors S
i
with aperture > π/k
i
,
i = 1 , . . . , r, aperture of S
1
at most 2 and assume S
r
S((h
1
2
)π
ε, (h+
1
2
)π+ε), for some h Z and some ε > 0. Finally, let D := D(h; θ
1
, θ
2
)
for some θ
1
< θ
2
.
A formal power series
ˆ
f C[[s
1/p
]]
1/(pk
1
)
is called (k
1
, . . . , k
r
, 1
+
)-
summable on (S
1
, . . . , S
r
, D) with (k
1
, . . . , k
r
, 1
+
)-sum f A
(1/k
1
)
(D),
60 B.L.J. BRAAKSMA, B.F. FABER, AND G.K. IMMINK
if there exist quasi-functions f
i
(A
(1/k
1
)
/A
≤−k
i+1
)(S
i
), i = 1, . . . , r, satis-
fying:
f
i
|
S
i+1
f
i+1
mod A
≤−k
i+1
, i = 0, . . . , r 1 with f
0
= T
1
ˆ
f;
f
r
has a representative {f
r,ω
}
ω
with respect
to a covering {S
r,ω
}
ω
of S
r
with open sectors S
r,ω
such that f
r,ω
f A
≤−1
(D S
r,ω
), ω .
Let g be another function such that f
r,ω
g A
≤−1
(D S
r,ω
), ω Ω.
Then f g A
≤−1
(D S
r,ω
), ω Ω, hence, f g A
≤−1
(D S
r
) =
A
≤−1
(D). Theorem 15 implies f = g and it follows that f is uniquely deter-
mined by f
r
and D. By the relative Watson lemma ([MR92, Prop. (2.1)])
f
i
is uniquely determined by f
i1
and S
i
, i = r, . . . , 1. Hence the sum f is
uniquely determined by
ˆ
f and (S
1
, . . . , S
r
, D).
Similarly to Definition 2 we extend this definition to the case that
ˆ
f is an
n-vector or n × n-matrix with elements in C[[s
1/p
]]
1/(pk
1
)
.
The main result of this paper is the following theorem.
Theorem 18. Let k
1
< · · · < k
r
= 1 be the sequence of positive levels 1
of (1.1). Suppose that 1
+
is a level of (1.1) and that <b
l
6= <b
m
if b
l
6= b
m
where b
l
is defined below (2.3). Let
ˆ
H(s)s
Λs
e
G(s)
s
L
be a formal fundamental
matrix as in (1.2).
Let S
i
= S(α
i
, β
i
), i = 1, . . . , r, be a sequence of open sectors, such that
S
1
. . . S
r
and β
1
α
1
2, β
i
α
i
> π/k
i
and (α
i
, β
i
) does not
contain a Stokes interval of level k
i
, i = 1, . . . , r of (1.1). Moreover, suppose
that (h1)π < α
r
< (h
1
2
)π and (h+
1
2
)π < β
r
< (h+1)π, for some h Z.
Let
˜
θ, θ R,
˜
θ < θ, such that (
˜
θ, θ) does not contain a Stokes number of
level 1
+
and define D = D(h;
˜
θ, θ).
Then
ˆ
H is (k
1
, . . . , k
r
, 1
+
)-summable on (S
1
, . . . , S
r
, D) with (k
1
, . . . ,
k
r
, 1
+
)-sum H such that (1.3) defines an analytic fundamental matrix of
(1.1).
For the proof of this theorem we will use the following lemma which
extends the results of Lemma 8-4.
Lemma 19. Let h Z and Q
±
be the quadrant H
2h
H
2h±1
. Here and in
the following the upper signs belong together and so do the lower signs. Let f
be given by (7.1). Define θ
j
:=
2π
d
(<b + j), j Z, and D := D(h; θ
N1
, θ
N
),
for some N Z.
Suppose that p(s) 6≡ 0 is a 1-periodic analytic function on a neighbourhood
of in H
2h±1
such that p(s)f(s) A
≤−1
+
(Q
±
).
Then there exists a 1-periodic function p
(s), such that p
(s) is analytic
on a neighbourhood of in H
2h±1
, p
+
(s) := p(s) p
(s) is an entire
SUMMATION METHOD FOR DIFFERENCE EQUATIONS 61
function, and
p
(s)f(s) A
≤−1
(D H
2h±1
),
p
+
(s)f(s) A
≤−1
(D H
2h1
) A
≤−1
+
(H
2h
).
Proof. We will give the details of the proof for the lower sign and h is even.
The other cases can be proven in a similar way.
Now Q
is a fourth quadrant and H := H
2h1
is a lower half plane.
According to Lemma 8-4 we have p(s)f(s) A
≤−1
+
((h
1
2
)π, ], and,
since p(s) 6≡ 0 also f(s) A
≤−1
+
((h
1
2
)π, ). Therefore d < 0, f(s)
A
≤−1
+
((h
1
2
)π, (h +
1
2
)π) and θ
N
< θ
N1
.
We have an expansion for p(s) as in (3.1) for =s < ρ for some ρ > 0.
Let p
(s) :=
P
jN 1
p
j
e
2πisj
. Then p
(s) is analytic for =s < ρ and
p
(s) = e
2πis(N1)
O(1), s on
H. Moreover, p
+
(s) =
P
jN
p
j
e
2πisj
is
an entire function.
First consider p
(s)f(s). The properties of p
imply
p
(s)f(s) A
≤−1
+
((h
1
2
)π, ] A
≤−1
+
(Q
).(7.3)
For ε > 0, let S
ε
:= S((h
1
2
)π ε, (h
1
2
)π + ε). In order to prove
that p
(s)f(s) A
≤−1
(D H) it is sufficient to s how, that with any θ
(θ
N
, θ
N1
) we c an find positive constants K, c and ε, such that
|p
(s)f(s)| Ke
c|s|
, s D(h; θ) S
ε
, =s < ρ.(7.4)
As p
(s)f(s) = O(1) exp(ds log s + 2πis(b + N 1) + o(s)), s on H, it
is sufficient to prove that
<{ds log s + 2πis(b + N 1)} c|s|, s D(h; θ) S
ε
,(7.5)
for some c > 0.
So let θ (θ
N
, θ
N1
). On D(h; θ) we have:
<{ds log s + 2πis(b + N 1)}
= d<{s(log s + ) + is(
2π
d
(b + N 1) θ)}
< d[=s(θ
N1
θ) +
2π
d
<s=b]
= d sin(arg s)(θ
N1
θ)|s|
1 +
2π=b
d(θ
N1
θ)
cot(arg s)
.
We have d sin(arg s)(θ
N1
θ) > 0 on H. Furthermore, there exists an ε > 0
such that
4π=b
d(θ
N1
θ)
cot(arg s)
< 1, s S
ε
. Thus (7.5) and (7.4) follow.
Next consider p
+
(s)f(s). As p(s)f(s), p
(s)f(s) A
≤−1
+
((h
1
2
)π, ],
also p
+
(s)f(s) belongs to this set. Moreover, f(s) A
≤−1
+
(H
2h
) and
|p
+
(s)| Ke
2πN=s
, =s ρ for som e K > 0, and therefore p
+
(s)f(s)
A
≤−1
+
(H
2h
). To prove the lemma it now suffic es to prove that p
+
f
62 B.L.J. BRAAKSMA, B.F. FABER, AND G.K. IMMINK
A
≤−1
(D H
2h+1
), hence that for any θ (θ
N
, θ
N1
) we can find positive
constants K, c and ε, such that
|p
+
(s)f(s)| Ke
c|s|
, s D(h; θ) S((h +
1
2
)π ε, (h +
1
2
)π + ε).
A proof of this inequality runs along the same lines as that of (7.4).
The following proposition extends the results of Proposition 10.
Proposition 20. Let k
1
< · · · < k
r
be the positive levels 1 of equation
(1.1) and assume that (1.1) has a level 1
+
.
Let M be odd. Assume α < M π/2, β > (M + 2)π/2 such that I =
(α, β) does not contain a Stokes interval of level 1 of (1.1). Define Γ
µ
:=
H
µ
S(α, β) for µ = M, M + 1, M + 2. Assume
˜
θ, θ R,
˜
θ < θ, such
that (
˜
θ, θ) does not contain any Stokes number of level 1
+
of (1.1). Define
D := D(
1
2
(M + 1);
˜
θ, θ).
Assume we have fundamental systems {f
l
u
(µ)
l
}
n
l=1
, µ = M, M + 1, M + 2
such that for all l {1, . . . , n}:
(i) u
(µ)
l
(A
(1/k
1
)
)
n
µ
)[log s], u
(µ)
l
ˆu
l
on Γ
µ
if µ = M, M + 1, M + 2,
(ii) u
(µ+1)
l
u
(µ)
l
(A
≤−1
+
)
n
µ
Γ
µ+1
) if µ = M, M + 1.
Then there exists a fundamental system {f
l
˜u
l
}
n
l=1
of equation (1.1) such
that for all l {1, . . . , n}:
˜u
l
u
(µ)
l
(A
≤−1
)
n
(D H
µ
) if µ = M, M + 2,(7.6)
˜u
l
u
(M+1)
l
(A
≤−1
+
)
n
(H
M+1
).(7.7)
Moreover, for each l {1, . . . , n} the function ˜u
l
(A
(1/k
1
)
)
n
(D)[log s]
is uniquely determined by these properties.
Remark. We can find α and β satisfying the above conditions if and only
if [νπ π/2, νπ + π/2] is not a Stokes interval of level 1 for any ν Z. This
corresponds to the condition that <b
l
6= <b
m
if b
l
6= b
m
.
Proof. Throughout the proof l, m {1, . . . , n}. Let h := (M + 1)/2, so h is
an integer. We will write m l if f
ml
A
≤−1
+
(H
2h
), and I(l) := {m | m
l}. The relation gives a partial ordering on {1, . . . , n}. We will prove the
proposition by induction with respect to this partial ordering.
Let p
(µ)
m
, µ = M, M + 1, be the 1-periodic analytic functions such that
u
(µ+1)
l
u
(µ)
l
=
n
X
m=1
p
(µ)
m
f
ml
u
(µ)
m
, µ = M, M + 1.(7.8)
As u
(µ+1)
l
u
(µ)
l
(A
≤−1
+
)
n
µ
Γ
µ+1
) by assumption, it follows from
Lemma 7 that each of the summands must belong to this set. If m 6∈ I(l)
then e
d
ml
s log s
is unbounded on H
2h
, so <d
ml
s is positive on H
2h
. Therefore
SUMMATION METHOD FOR DIFFERENCE EQUATIONS 63
f
ml
6∈ A
≤−1
+
µ
Γ
µ+1
) for µ = M, M + 1 and Lemma 8 tells us that
p
(µ)
m
= 0. So in (7.8) we only need to sum over m I(l).
First let l be s uch that I(l) = . Then u
(µ)
l
is independent of µ {M, M +
1, M + 2} and ˜u
l
= u
(µ)
l
satisfies (7.6) and (7.7).
Now let l be such that ˜u
m
have been defined, and satisfy (7.6) and (7.7)
for all m I(l). The functions f
m
˜u
m
, m I(l), together with the functions
f
m
u
(M+1)
m
, m 6∈ I(l), form a fundamental system of solutions. With the aid
of Lemmas 7 and 8 we may conclude as ab ove that
u
(M+1)
l
u
(M)
l
=
X
mI(l)
˜p
(M)
m
f
ml
˜u
m
,
for some 1-periodic analytic functions ˜p
(M)
m
on a neighbourhood of in
Q := Γ
M
Γ
M+1
, and ˜p
(M)
m
f
ml
A
≤−1
+
(Q). According to Lemma 19
the functions ˜p
(M)
m
(s) can be written as p
(M)
m
(s) + p
(M)
m+
(s) with p
(M)
m
f
ml
A
≤−1
(D H
M
) and p
(M)
m+
f
ml
A
≤−1
(D H
M+2
) A
≤−1
+
(H
M+1
). We
define
u
l,1
:= u
(M)
l
+
X
mI(l)
p
(M)
m
f
ml
˜u
m
,
so that u
l,1
u
(M)
l
(A
≤−1
)
n
(D H
M
) and
u
l,1
u
(M+1)
l
=
X
mI(l)
p
(M)
m+
f
ml
˜u
m
A
≤−1
n
(D H
M+2
)
A
≤−1
+
n
(H
M+1
).
From these relations and assumption (ii) it follows that u
(M+2)
l
u
l,1
(A
≤−1
+
)
n
M+1
Γ
M+2
) and as above we find that there exist 1-periodic
functions p
(M+1)
m
(s), p
(M+1)
m+
(s), m I(l), such that
u
(M+2)
l
u
l,1
=
X
mI(l)
p
(M+1)
m
f
ml
˜u
m
+ p
(M+1)
m+
f
ml
˜u
m
,
where the first term of each summand belongs to (A
≤−1
)
n
(D H
M+2
), and
the second one to (A
≤−1
)
n
(DH
M
)(A
≤−1
+
)
n
(H
M+1
). Hence, if we define
˜u
l
:= u
(M+2)
l
X
mI(l)
p
(M+1)
m
f
ml
˜u
m
,
then
˜u
l
= u
l,1
+
X
mI(l)
p
(M+1)
m+
f
ml
˜u
m
,
64 B.L.J. BRAAKSMA, B.F. FABER, AND G.K. IMMINK
and it is easy to verify that ˜u
l
satisfies (7.6) and (7.7). The uniqueness of
˜u
l
follows from Theorem 15.
Theorem 18 can be proved similarly to Theorem 13, with the aid of Propo-
sitions 9, 10 and 20.
References
[Bal94] W. Balser, From Divergent Power Series to Analytic Functions, Lecture Notes
in Mathematics 1582, Springer Verlag, Heidelberg, 1994.
[BBRS91] W. Balser, B.L.J. Braaksma, J.-P. Ramis and Y. Sibuya, Multisummability of
formal power series solutions of linear ordinary differential equations, Asympt.
Anal., 5 (1991), 27-45.
[BT33] G.D. Birkhoff and W.J. Trjitzinsky, Analytic theory of singular difference equa-
tions, Acta Math., 60 (1933), 1-89.
[Bra91] B.L.J. Braaksma, Multisummability and Stokes multipliers of linear meromor-
phic differential equations, J. Diff. Eq., 92 (1991), 45-75.
[Bra92] , Multisummability of formal power series solutions of nonlinear mero-
morphic differential equations, Ann. Inst. Fourier, 42 (1992), 517-540.
[BF96] B.L.J. Braaksma and B.F. Faber, Multisummability for some classes of differ-
ence equations, Ann. Inst. Fourier, 46(1) (1996), 183-217.
[BH75] B.L.J. Braaksma and W.A. Harris, Jr., On an open problem in the theory of
linear difference equations, Nieuw Archief voor Wiskunde (3), XXIII (1975),
228-240.
[BIS] B.L.J. Braaksma, G.K. Immink and Y. Sibuya, The Stokes phenomenon in
exact asymptotics, Pacific J. Math., 187 (1999), 13-50.
[Eca85] J. Ecalle, Les fonctions esurgentes, III, Publ. Math. Orsay, 1985.
[Eca87] , L’acc´el´eration des fonctions esurgentes, Manuscript, 1987.
[Fab97] B.F. Faber, Differen ce equations and summability, Revista del Seminario
Iberoamericano de Matem´aticas, V (1997), 53-63.
[Imm84] G.K. Immink, Asymptotics of Analytic Difference Equations, Lecture Notes in
Mathematics, 1085, Springer Verlag, Heidelberg, 1984.
[Imm88] , Asymptotic expansions with error bounds for solutions of difference
equations of ‘level 1
+
,
´
Equations diff´erentielles dans le champ complexe, Vol. 1,
I.R.M.A., Strasbourg, (1988), 35-60.
[Imm91] , Reduction to canonical forms and the Stokes phenomenon in the theory
of linear difference equations, SIAM J. Math. Anal., 22 (1991), 238-259.
[Imm] , On the summability of the formal solutions of a class of inhomogeneous
linear difference equations, Funkcial. Ekvac., 39(3) (1996), 469-490.
[Imm96] , A particular type of summability of divergent power series, with an
application to linear difference equations, Preprint, 1996.
[Mal95] B. Malgrange, Sommation des eries divergentes, Expo. Math., 13 (1995), 163-
222.
[MR92] B. Malgrange and J.-P. Ramis, Fonctions multisommables, Ann. Inst. Fourier,
41 (1992), 353-368.
SUMMATION METHOD FOR DIFFERENCE EQUATIONS 65
[MR91] J. Martinet and J.-P. Ramis, Elementary acceleration and multisummability,
Ann. Inst. H. Poincar´e, Phys. Th´eor., 54 (1991), 331-401.
[RS94] J.-P. Ramis and Y. Sibuya, A new proof of multisummability of formal solutions
of non-linear meromorphic differential equations, Ann. Inst. Fourier, 44 (1994),
811-848.
[Tov96] A.I. Tovbis, Multisummable solutions of nonlinear ordinary differential equa-
tions, J. Diff. Eq., 124 (1996), 359-377.
[vdPS97] M. van der Put and M.F. Singer, Galois theory of difference equations, Lecture
Notes in Mathematics, 1666, Springer Verlag, Heidelberg, 1997.
Received October 22, 1998.
University of Groningen
Department of Mathematics, P.O.Box 800
9700 AV Groningen
The Netherlands
E-mail address: B.L.J.Braaksma@math.rug.nl
University of Groningen
Department of Mathematics, P.O.Box 800
9700 AV Groningen
The Netherlands
E-mail address: bernard@math.rug.nl
University of Groningen
Department of Econometrics, P.O.Box 800
9700 AV Groningen
The Netherlands
E-mail address: G.K.Immink@eco.rug.nl
PACIFIC JOURNAL OF MATHEMATICS
Vol. 195, No. 1, 2000
APPROXIMATION OF RECURRENCE IN NEGATIVELY
CURVED METRIC SPACES
Charalambos Charitos and Georgios Tsapogas
For metric spaces with curvature less than or equal to χ,
χ < 0, it is shown that a recurrent geodesic is approximated
by closed geodesics. A counter example is provided for the
converse.
1. Introduction and preliminaries.
In hyperbolic geometry it has been shown lately that many geometric prop-
erties are determined by the distance function on the space itself rather than
the differential structure. It is shown in this work that, partially, this is the
case with the notion of recurrence. For complete hyperbolic manifolds, a
recent result of Aebischer, Hong and McCullough (see [1]) states that a ge-
odesic is recurrent if and only if it is approximated by closed geodesics. We
show that, in metric spaces with curvature less than or equal to χ, χ < 0,
recurrent geodesics are approximated by closed geodesics (see Theorem 2
below). The proof of the converse statement crucially depends on the man-
ifold structure, in particular on the fact that two geodesics coincide if they
do so on an open interval. Hence, the converse statement fails in our context
due to the bifurcation property of geodesics. A counterexample exhibiting
this failure is provided in Section 4 below. A geodesic γ is called recurrent
if there exists a sequence {t
n
} R, t
n
such that t
n
γ γ as t
n
.
Convergence in this definition is me ant to be uniform convergence on com-
pact sets which, in fact, induces the topology on the space GX consisting
of all (local) isometries R X when X is (not) simply connected. R acts
on GX by right translations, namely, (t, g) tg, where tg : R X is the
geodesic defined by tg (s) = g (s + t) , s R. This action is simply the ge-
odesic flow. The notion of convergence in the above definition is analogous
to the tangential condition which defines recurrence in the manifold case.
We use the notion of approximation given in Definition 6 below which was
introduced in [1] in order to characterize recurrent geodesics in hyperbolic
manifolds.
X will always denote a locally compact, complete, geodesic m etric space
with curvature less than or equal to χ, χ < 0. Recall that a geodesic metric
space is said to have curvature less than or equal to χ if each x X has
a neighborhood V
x
such that every geo des ic triangle of perimeter strictly
67
68 C. CHARITOS AND G. TSAPOGAS
less than
2π
χ
(=+ when χ 0) contained in V
x
satisfies CAT (χ)
inequality (see [11] for definitions and basic properties). We will denote
the metric by d (·, ·) and will use the same letter to denote distance when
the metric space to which we refer is understood. All curves are assumed
to be parametrized by arclength. A geodesic segment in X is an isometry
c : I X, where I is a closed interval in R. A geodesic in X is a map
c : R X such that for each closed interval I R, the map c |
I
: I X is
a geodesic segment. A local geodesic segment (usually called geodesic arc)
in X is a map c : I X such that for each t I there is an ε > 0 such
that c |
[tε,t+ε]I
: [t ε, t + ε] I X is a geodesic segment. Similarly, a
local geodesic R X is defined. A closed geodesic in X is a local geodesic
c : R X which is a periodic map.
Definition 1. An oriented geodesic g in X is said to be approximated by
closed geodesics if, for every ε > 0 and every x Im g, there exists a closed
oriented geodesic c such that for some point y Im c,
d (c (t + t
y
) , g (t + t
x
)) < ε
for all t [0, period (c)] , where t
x
, t
y
R with x = g (t
x
) and y = c (t
y
) .
The following theorem is the main result of this paper.
Theorem 2. Let X be a locally compact, complete, geodesic metric space
which has curvature less than or equal to χ, χ < 0. If a geodesic or geodesic
ray in X is recurrent, then it is approximated by closed geodesics.
The proof of Theorem 2 uses the notion of quasi-geodesic and its stability
properties. We will closely follow notation and terminology app earing in [8,
Ch. 3] where we refer the reader for first definitions and basic prope rties of
quasi-geodesics. Here we only recall the following definition.
Definition 3. Let f : [a, b] X be a continuous map with −∞ a b
+ and λ, κ, L real numbers with λ 1, κ 0, L > 0.
f is a (λ, κ, L) quasi-geodesic if for every subinterval [a
0
, b
0
] of [a, b] sat-
isfying
length f

a
0
, b
0

L,
the following inequality holds
length f

a
0
, b
0

λd
f
a
0
, f
b
0

+ κ.
The next proposition is a well know fact for CAT (χ) spaces. We include
a short proof of it, since it is difficult to find exact reference (when X is a
geometric polyhedron this result follows from [3, p. 403]).
Proposition 4. Let M be a complete geodesic space satisfying CAT (χ)
inequality with χ < 0. Every local geodesic segment in M is a geodesic seg-
ment.
RECURRENCE IN NEGATIVELY CURVED SPACES 69
Proof. Let δ : [0, L] M be a local geodesic segment in M, L > 0. Set
l = sup
t [0, L]
δ |
[0,t]
is a geodesic segment
.
Apparently, l > 0 and by completeness of M, δ |
[0,l ]
is a geodesic segment
joining δ (0) with δ (l). Assuming the conclusion is not true, i.e., l < L, let ε
be a p ositive number such that δ |
[l ε,l+ε]
is a geodesic segment. Denote by
[δ (0) , δ (l + ε)] the geodesic segment in M joining δ (0) with δ (l + ε). Since
δ |
[0,l +ε]
is not the geodesic segment joining δ (0) with δ (l + ε) ,
d (δ (0) , δ (l + ε)) < d (δ (0) , δ (l)) + d (δ (l) , δ (l + ε)) .(1)
The points δ (0) , δ (l ) and δ (l + ε) define a geodesic triangle in M. Denote
by =
δ (0), δ (l ), δ (l + ε)
the corresponding comparison triangle which
is non-degenerate by inequality (1). Choose points B on δ |
[0,l ]
and B
0
on
δ |
[l ,l+ε]
such that d (B, δ (l )) = d (B
0
, δ (l)) = ε
0
< ε and denote by B and
B
0
the corresponding points on the comparison triangle. Then by (1) the
angle of at δ (l) is smaller than π and therefore
d
B, B
0
< d
B, δ (l)
+ d
δ (l), B
0
= 2ε
0
.
By comparison, d (B, B
0
) d
B, B
0
so we obtain
d
B, B
0
< d (B, δ (l)) + d
δ (l) , B
0
.
This contradicts the fact that δ |
[l ε
0
,l +ε
0
]
is a geodesic segment.
Let
e
X be the universal cover of X and p :
e
X X the projection map.
e
X becomes a metric space as follows: Given ex, ey
e
X choose any curve
ec : [a, b]
e
X with ec (a) = ex and ec (b) = ey and define the distance from ex to
ey to be the length of the unique length minimizing curve in the homotopy
class of pec with endpoints fixed. For the existence of the length minimizing
curve see [10]. This distance function is a metric on
e
X which inherits the
properties of X, namely,
e
X becomes a complete geodesic locally compact
(hence, proper) me tric space. π
1
(X) acts on
e
X and the action commutes
with p. As the projection p is a local isometry, it follows that π
1
(X) acts on
e
X by local isometries. Using the fact that
e
X is geodesic and Proposition 4,
it is routine to show that π
1
(X) acts on
e
X by isometries. In addition,
e
X
has curvature less than or equal to χ, χ < 0 and, by a theorem of Gromov
(see for example [11, p. 325]),
e
X satisfies CAT (χ) inequality.
GX is by definition the space of all local geodesics R X and, by Propo-
sition 4 above, G
e
X is the space consisting of all global geodesics R
e
X.
The topology on these spaces is uniform convergence on compact sets. The
boundary
e
X can be defined using either equivalence classes of sequences
or, equivalence classes of geodesic rays. The local compactness assumption
on X implies that
e
X is proper and hence the two definitions coincide (see
70 C. CHARITOS AND G. TSAPOGAS
[8, Ch. 2]). We will be using them interchangeably. For any two distinct
points ξ, η in
e
X there exists a unique, up to parametrization, (oriented)
geodesic g with g (−∞) = ξ and g () = η (se e for example [5, Prop. 2]).
We need the following lem ma which asserts that the projection of a point
onto a geodesic always exists.
Lemma 5. Let g be a geodesic in G
e
X (or a geodesic segment) and x
0
a
point in
e
X. There exists a unique real number s such that g (s) realizes the
distance of x
0
from Im g, i.e., dist (x
0
, Im g) = d (x
0
, g (s)) .
Proof. We may assume that x
0
/Im g. Existence is apparent. Assume that
s 6= s
0
are two such numbers. The points g (s) , g (s
0
) and x
0
define a non-
degenerate geodesic triangle in
e
X and denote by =
g (s), g (s
0
), x
0
the
corresponding comparison triangle. is an equilateral triangle in the unique
complete simply connected Riemannian 2-manifold of constant sectional cur-
vature χ. Hence, the angles of at g (s) and g (s
0
) are each less than π/2.
Therefore, there exists a point g (t) on the side of opposite to x
0
such
that d
x
0
, g (t)
< d
x
0
, g (s)
= d
x
0
, g (s
0
)
. By CAT (χ) inequality,
d (x
0
, g (t)) d (x
0
, g (s)) , a contradiction.
Remark 1. If c GX is a closed geodesic and x
0
X, the same argument
applied to a lifting ec of c shows that there exists a unique point B Im c
such that d (x
0
, B) = dist (x
0
, Im c) .
Remark 2. Set
2
e
X =
n
(ξ, η)
e
X ×
e
X : ξ 6= η
o
and let ρ : G
e
X
2
e
X
be the fiber bundle given by ρ (g) = (g (−∞) , g (+)). Since for any two
distinct points ξ, η in
e
X there exists a unique (oriente d) geodesic g with
g (−∞) = ξ and g () = η (see for example [5, Prop. 2]), the fiber of ρ is
R. Moreover, this bundle is trivial (see for example [4, Th. 4.8]). To define
a trivialization, let x
0
be a base point and let
H : G
e
X
2
e
X × R(2)
be the trivialization of ρ with respect to x
0
defined by
H (g) = (g (−∞) , g (+) , s)
where s is the real number provided by Lemma 5. Note that the composite
of the geodesic flow R × G
e
X G
e
X with H is given by the formula
(ξ
1
, ξ
2
, s) (ξ
1
, ξ
2
, s + t)
for all (ξ
1
, ξ
2
)
2
e
X and s R.
RECURRENCE IN NEGATIVELY CURVED SPACES 71
2. Recurrent geodesics.
Definition 6. A geodesic γ in X is called recurrent if there exists a sequence
{t
n
} R, t
n
such that t
n
γ γ as t
n
.
For a recurrent geodesic γ in X there exists a sequence of closed (in
fact, piece-wise geodesic) curves {γ
n
}
nN
, associated to γ as follows: Fix
a convex neighborhood U of γ (0) , i.e., a neighborhood which satisfies the
following property: For all x, y U there exists a unique geodesic segment
with endpoints x and y lying entirely in U. Such a neighborhood exists (see
for example [2]). If {t
n
} is the sequence given by Definition 6 above and
ε
n
= d (γ (0) , γ (t
n
)), let K N such that γ (t
n
) U for all n K. Define
γ
n
, n K to be the curve
γ
n
: [0, t
n
+ ε
n
] X(3)
with γ
n
(t) = γ (t) t [0, t
n
] and γ
n
|
[t
n
,t
n
+ε
n
]
the unique geodes ic segment
in U joining γ (t
n
) with γ (0) . Note that t
n
+ ε
n
is the period of the closed
curve γ
n
. In the sequel, we will refer to these closed curves by writing
γ
n
, n N but it will always b e implicit that n is large enough so that γ
n
are
defined.
Using the following lemma, we may assume that given a recurrent geodesic
γ, the associated closed curves {γ
n
}
nN
are not homotopic to a point.
Lemma 7. Given a recurren t geodesic γ there exists M N such that each
closed curve γ
n
, n N associated to γ is not homotopic to a point, provided
n M.
Proof. Let eγ be a lift of γ to the universal cover
e
X of X parametrized so that
eγ (0) projects to γ (0) = γ
n
(0) . The curve γ
n
|
[0,t
n
]
is a local geodesic segment
and, by Proposition 4, its lift fγ
n
|
[0,t
n
]
to
e
X starting at eγ (0) is a geodesic
segment. Moreover, γ
n
|
[t
n
,t
n
+ε
n
]
and its lift fγ
n
|
[t
n
,t
n
+ε
n
]
to
e
X starting at
eγ
n
(t
n
) are both geodesic segments. We have
d (eγ
n
(t
n
+ ε
n
) , eγ
n
(0)) d (eγ
n
(t
n
) , eγ
n
(0)) d (eγ
n
(t
n
+ ε
n
) , eγ
n
(t
n
))
= t
n
ε
n
.
Since ε
n
0 and t
n
as n , we may choose M N such that
eγ
n
(t
n
+ ε
n
) , eγ
n
(0) are distinct for all n M. Therefore, fγ
n
|
[0,t
n
+ε
n
]
, which
is the lift of the closed curve γ
n
starting at eγ (0) = fγ
n
(0) , has distinct
endpoints and, therefore, γ
n
, n M is not homotopic to a point.
The following proposition shows that the lifts (to the universal cover
e
X)
of the closed curves γ
n
associated to a recurrent geodesic γ are, for n large
enough, quasi-geodesics with arbitrarily large L. Recall that a CAT (χ)
space is a δhyperbolic space in the sense of Gromov (see for example [11,
72 C. CHARITOS AND G. TSAPOGAS
Sec. 2]). This applies to the universal covering
e
X, since it satisfies CAT (χ)
inequality globally. Let δ denote the hyperbolicity constant of the space
e
X.
Proposition 8. Let γ be a recurrent geodesic in X and {γ
n
}
nN
the as-
sociated closed curves. For every L > 0, there exists N N such that all
lifts fγ
n
: R
e
X of γ
n
with n N are (λ, κ, L) quasi-geodesics provided
κ > 16δ and λ = 1, where δ is the h yperbolicity const ant of
e
X.
Proof. Let γ be a recurrent geodesic and L > 0 be given. The sequence {t
n
}
given by Definition 6 converges to infinity. Moreover, ε
n
= d (γ (0) , γ (t
n
))
0 and t
n
+ ε
n
= period (γ
n
) also converges to infinity as n . Hence, we
may choose N such that
t
n
+ ε
n
> L and ε
n
<
1
2
(κ 16δ) for all n N.(4)
Let now [a, b] be any interval with b a < L (cf. Definition 3). For each
n N there exists an integer k
n
such that
fγ
n
([a, b]) fγ
n
([(k
n
1) (t
n
+ ε
n
) , k
n
(t
n
+ ε
n
) + t
n
]) .(5)
Denote by [fγ
n
(a) , fγ
n
(b)] the unique geodesic segment in
e
X joining fγ
n
(a)
with fγ
n
(b) and set
A
n
:= fγ
n
(k
n
(t
n
+ ε
n
) ε
n
)
y
k
n
:= fγ
n
(k
n
(t
n
+ ε
n
)) .
The distance of any point on [fγ
n
(a) , fγ
n
(b)] from fγ
n
([a, b]) is bounded by a
number which depends on the hyperbolicity constant δ of the space
e
Xand on
the number of geodesic segments which constitute fγ
n
([a, b]) , see [8, Lemma
1.5, p. 25]. In our case here, fγ
n
([a, b]) consists of at most three geodesic seg-
ments (since the right hand side of inclusion (5) above consists of 3 geodesic
segments) and the bound is 8δ. Hence we have
d (y
k
n
, [fγ
n
(a) , fγ
n
(b)]) 8δ.(6)
By Lemma 5, let B
n
be the point on [fγ
n
(a) , fγ
n
(b)] which realizes the dis-
tance in the left hand side of inequality 6. Assume that neither fγ
n
(a) nor
fγ
n
(b) lies on the geodesic segment [A
n
, y
k
n
] . Then we have the following
triangle inequalities
d (fγ
n
(a) , A
n
) d (fγ
n
(a) , B
n
) + d (B
n
, y
k
n
) + d (y
k
n
, A
n
)
d (y
k
n
, fγ
n
(b)) d (y
k
n
, B
n
) + d (B
n
, fγ
n
(b))
which, after employing the fact that d (A
n
, y
k
n
) = ε
n
, become
length fγ
n
([a, b]) = d (fγ
n
(a) , A
n
) + d (A
n
, y
k
n
) + d (y
k
n
, fγ
n
(b))
2ε
n
+ 2d (y
k
n
, B
n
) + d (fγ
n
(a) , fγ
n
(b))
by inequality (6)
2ε
n
+ 2 · 8δ + d (fγ
n
(a) , fγ
n
(b))
by inequality (4)
κ + d ( fγ
n
(a) , fγ
n
(b)) .
RECURRENCE IN NEGATIVELY CURVED SPACES 73
The case where fγ
n
(a) and/or fγ
n
(b) lies on [A
n
, y
k
n
] is treated similarly.
Corollary 9. For n N sufficiently large, the isometry of
e
X in π
1
(X)
which corresponds to the homotopy class of the closed curve γ
n
is hyperbolic.
Proof. It suffices to show that each fγ
n
: R
e
X determines exactly two
boundary points fγ
n
(−∞), fγ
n
(+) . By Lemma 8 there exists an M N
such that fγ
n
is a quasi-isometry for all n M. Each such fγ
n
induces a
map R
e
X which is a homeomorphism onto its image, see [8, Th. 2.2,
p. 35]. As R consists of two distinct points, fγ
n
(−∞), fγ
n
(+)
e
X are
also distinct for all n M.
It now follows that a recurrent geodesic γ in X as well as each of the
(oriented) closed curves γ
n
, n M (cf. Lemma 7 and Corollary 9 above)
determine exactly two boundary points in
e
X denoted by eγ (−∞), eγ (+)
and fγ
n
(−∞), fγ
n
(+) respectively. We need the following lemma concern-
ing these boundary points. Recall that
e
X
e
X is a compact space which is
metrizable (see [8, p. 134]), and we will denote such metric by d
e
X
e
X
.
Lemma 10. fγ
n
(−∞) eγ (−∞) and fγ
n
(+) eγ (+) as n .
Proof. As above, let ε
n
= length (Im γ
n
) t
n
so that t
n
+ ε
n
is the pe-
riod of γ
n
. We first show that fγ
n
(+) eγ (+) . Consider the sequence
fγ
n
(k (t
n
+ ε
n
)) , k N which converges to fγ
n
(+) as k . Thus, there
exists k
n
N such that
d
e
X
e
X
(fγ
n
(k
n
(t
n
+ ε
n
)) , fγ
n
(+)) < 1/n.(7)
Now consider the sequences y
n
:= fγ
n
(k
n
(t
n
+ ε
n
)) and x
n
:= eγ (t
n
), n N.
Since x
n
eγ (+), by inequality (7) above it is enough to show that the
sequences {x
n
} and {y
n
} represent the same element in
e
X or, in other
words, that the hyperbolic product (x
n
, y
n
)
x
0
with respect to the base point
x
0
:= eγ (0) converges to + as n +. For the notion of hyperbolic
product of sequences and their equivalence, see [8].
The stability property of quasi-geodesics states (see Corollary 1.10 of [8,
p. 31]) that given any two numbers κ 0 and λ 1, there exists a constant
C depending on λ, κ and on the hyperbolicity constant δ of the space such
that if L is bigger than 2C then every (λ, κ, L)-quasi-geodesic f : [a, b]
e
X
lies within a Cneighborhood of the geodesic segment [f (a) , f (b)] . By
choosing λ = 1, κ > 16δ where δ is the hyperbolicity constant of the space
e
X and L > 2C we obtain, by Proposition 8 above, a natural number N
such that all fγ
n
: R
e
X with n N are (λ, κ, L)-quasi-geodesics. In
particular, fγ
n
: [0, k
n
(t
n
+ ε
n
)]
e
X are (λ, κ, L)-quasi-geodesics for all
n N. Therefore, by Corollary 1.10 of [8, p. 31] as explained above,
d
x
n
, x
0
n
< C n N
74 C. CHARITOS AND G. TSAPOGAS
where x
0
n
denotes the projection of x
n
on the geodesic segment [eγ (0) , y
n
]
(cf. Lemma 5). Hence,
(x
n
, y
n
)
x
0
=
1
2
(d (x
n
, x
0
) + d (y
n
, x
0
) d (x
n
, y
n
))
1
2
(d (x
0
n
, x
0
) C + d (y
n
, x
0
) d (x
0
n
, y
n
) C)
= (x
0
n
, y
n
)
x
0
C
= d (x
0
, x
0
n
) C.
Apparently, d (x
0
, x
0
n
) as n + and, hence, (x
n
, y
n
)
x
0
as
required.
In order to show that fγ
n
(−∞) eγ (−∞) we work in a similar manner:
The se quence fγ
n
(k (t
n
+ ε
n
)) , k N converges to fγ
n
(−∞) as k .
Hence, there exists k
n
N such that d
e
X
e
X
fγ
n
(k
n
(t
n
+ ε
n
)) ,
fγ
n
(−∞)
< 1/n. As before, sequences {y
n
} and {x
n
} are defined by y
n
:=
fγ
n
(k
n
(t
n
+ ε
n
)) and x
n
:= eγ (t
n
), n N. Then we use the same argu-
ments to show that the hyperbolic product (x
n
, y
n
)
x
0
with respect to the
base point x
0
:= eγ (0) converges to + as n +.
3. Proof of main theorem.
Let γ be a recurrent geodesic, ε > 0 and x Im γ be given. We may assume
that x = γ (0) . Let {t
n
} be the sequence given by Definition 6 and {γ
n
} the
sequence of the associated closed curves given by formula (3) above. For
each n N, there exists a unique closed geodesic c
n
in the free homotopy
class of γ
n
. The number t
n
+ ε
n
is the period of γ
n
and let s
n
denote the
period of c
n
(apparently, s
n
< t
n
+ ε
n
). Let B
n
be the projection of γ (0)
onto Im c
n
, i.e., d (γ (0) , B
n
) = d (γ (0) , Im c
n
) . Such a point exists and is
unique by Remark 1 following Lemma 5. Lift γ to an isome try eγ : R
e
X
with a base point eγ (0) satisfying p (eγ (0)) = γ (0) , where p :
e
X X is
the universal covering map. Lift each c
n
to an isometry ec
n
: R
e
X and
parametrize it so that ec
n
(0) is a point
f
B
n
satisfying
d
f
B
n
, eγ (0)
= d (B
n
, γ (0)) and p
f
B
n
= B
n
.
For the reader’s convenience, we have gathered all the above notation in
Figure 1.
Since
p (fγ
n
(t
n
+ ε
n
)) = p (eγ (0)) = γ (0)
and γ
n
, c
n
are homotopic, the isometry φ
n
of
e
X which translates ec
n
(in the
positive direction) satisfies
φ
n
(eγ (0)) = fγ
n
(t
n
+ ε
n
) .
RECURRENCE IN NEGATIVELY CURVED SPACES 75
X
X
X
X
X
X
X
X
X
X
X
X
X
X
X
X
X
X
X
X
X
X
X
X
X
X
h
h
h
h
h
h
h
h
h
h
h
h
h
h
h
h
h
h
h
h
h
h
h
h
h
h
h
h
J
J
J
J
J
s s ss
s s s s
s
s
s
ec
n
eγ
eγ(0)
ec
n
(0) =
f
B
n
eγ(s) eγ(t
n
) eγ(t
n
+
n
)
fγ
n
(t
n
+
n
)
D
n
F
n
F
0
n
ec
n
(s)
ec
n
(s
n
) = φ
n
f
B
n
Figure 1.
Moreover,
d
fγ
n
(t
n
+ ε
n
) , φ
n
f
B
n

= d
φ
n
(eγ (0)) , φ
n
f
B
n

= d
eγ (0) ,
f
B
n
.
(8)
We now proceed to show that given ε > 0, there exists N N such that for
all n N
d (eγ (s) , ec
n
(s)) < ε s [0, s
n
] .(9)
Recall that s
n
is the period of c
n
and s
n
< t
n
+ ε
n
= period (γ
n
) . Using
Lemma 10 and the fact that γ
n
, c
n
are homotopic for all n large enough, we
have that ec
n
(+) eγ (+) and ec
n
(−∞) eγ (−∞) . Let H : G
e
X
2
e
X ×R be the trivilization of the fiber bundle G
e
X
2
e
X with respect to
the base point x
0
= eγ (0) . This homeomorphism was described in Remark
2 following Lemma 5. By the choice of parametrization for each ec
n
made
above
i.e., ec
n
(0) =
f
B
n
, we have that H
1
( ec
n
(−∞) , ec
n
(+) , 0) = ec
n
.
Moreover, H
1
(eγ (−∞) , eγ (+) , 0) = eγ and, thus, ec
n
eγ uniformly on
compact sets. Observe that such convergence is weaker than property (9).
However, it implies, in particular, that dist (eγ (0) , Im ec
n
) 0 as n .
Hence, we may choose N N such that
d
eγ (0) ,
f
B
n
< ε/5, for all n N.(10)
Moreover, we may choose N such that, in addition, the following inequality
is satisfied
ε
n
= d (eγ (t
n
) , eγ (t
n
+ ε
n
)) < ε/5, for all n N.(11)
To show inequality (9), let s [0, s
n
] be arbitrary and let D
n
(resp. F
n
) be the point on the geodesic segment [eγ (0) , fγ
n
(t
n
+ ε
n
)]
76 C. CHARITOS AND G. TSAPOGAS
resp.
h
eγ (0) , φ
n
f
B
n
i
whose distance from eγ (0) is equal to s. Then,
d (eγ (s) , ec
n
(s)) d (eγ (s) , D
n
) + d (D
n
, F
n
) + d (F
n
, ec
n
(s))
d (eγ (s) , D
n
) + d (D
n
, F
n
) + d
F
n
, F
0
n
+ d
F
0
n
, ec
n
(s)
where F
0
n
is the point on
h
f
B
n
, φ
n
f
B
n
i
satisfying
d
F
n
, φ
n
f
B
n

= d
F
0
n
, φ
n
f
B
n

.
By comparison (see for example [12, Prop. 29]) we have
d (eγ (s) , D
n
) d (eγ (t
n
+ ε
n
) , fγ
n
(t
n
+ ε
n
)) 2ε
n
d (D
n
, F
n
) d
eγ (t
n
+ ε
n
) , φ
n
f
B
n

d
F
n
, F
0
n
d
eγ (0) ,
f
B
n
d
F
0
n
, ec
n
(s)
d
eγ (0) , φ
n
f
B
n

d
f
B
n
, φ
n
f
B
n

< d
eγ (0) ,
f
B
n
.
Combining the above inequalities with inequalities (8), (10) and (11), we
obtain property (9) which completes the proof of the existence of a sequence
of closed geodesics approximating a given recurrent geodesic.
Remark. Let Γ be a discrete group of isometries of a locally compact, com-
plete geodesic metric space Y satisfying CAT (χ) inequality, χ < 0. The
notion of controlled concentration points in the limit set of Γ can be defined
as follows. ξ Γ is a controlled concentration point if it admits a neighbor-
hood U containing ξ with the following property: For every neighborhood
V of ξ there exists an element γ Γ such that γ (U) V and ξ γ (V ) .
Following [1], one can show that ξ is a controlled concentration point if and
only if there exists a sequence of {φ
n
} of distinct elements of Γ such that
φ
n
(ξ) ξ and φ
n
(0) η with η 6= ξ. The proof in this more general
setting is identical with the one provided in [1] except that the convergence
property used there, namely, φ
n
(x) η for all x Y Y, is provided
in our case by Proposition 7.2 in [6, Ch. 1]. The latter property for ξ is
equivalent to the existence of a recurrent geodesic γ with γ (+) = ξ and
γ (−∞) = η. Hence we obtain the following connection between recurrent
geodesics and controlled concentration points which also holds for manifolds
(see [1]).
Theorem 11. Let Y be a locally compact, complete geodesic metric space Y
satisfying CAT (χ) inequality, χ < 0 and Γ a discrete group of isometries
of Y. A limit point ξ Y is a controlled concentration point if and only if
γ (+) = ξ for some recurrent geodesic γ in Y.
RECURRENCE IN NEGATIVELY CURVED SPACES 77
4. Construction of a counterexample.
As it was mentioned in the introduction, approximation by closed geodesics
does not imply recurrence. The following example demonstrates the exis-
tence of a geodesic in a CAT (χ) , χ < 0 space which is not recurrent but
can be approximated by closed geodesics in the sense of Definition 1. Let X
be the union of two hyperbolic cylinders identified along a (convex) geodesic
strip bounded by two geodesic segments (see Figure 2). We may adjust the
geometry of X so that the unique simple closed geodesic in each cylinder,
denoted by c
1
and c
2
, have a common image in the geodesic strip, namely,
the geodesic segment indicated by letters A and B in Figure 2. Using Cor. 5
of [2] and the fact that the geodesic strip is a convex closed subset it follows
that X is a CAT (χ) space with χ < 0.
Figure 2.
Let ω
1
and ω
2
be the periods of c
1
and c
2
respectively and assume that
c
1
and c
2
are parametrized so that c
1
(0) = c
2
(0) = B and clockwise i.e.,
c
1
(s) = c
2
(s) for all s [0, d (A, B)] . Define γ : R X as follows:
γ (t) = c
1
(t) , for t [0, ω
1
]
γ (t) = c
2
(t) , for t (−∞, 0] [ω
1
, +) .
It is apparent that γ can be approximated by closed geodesics in the sense
of Definition 1. We pro cee d to show that γ is not recurrent by showing that,
γ and are not close in the compact open topology for any positive real
s. For this it suffice s to show that there e xists ε > 0 and a compact M R
78 C. CHARITOS AND G. TSAPOGAS
such that for any positive s R,
d
(t
0
) , γ (t
0
)
ε for some t
0
M.(12)
For simplicity, we may assume that d (A, B) = ω
1
/2 = ω
2
/4. Pick ε <
d (A, B) /2 and cho os e a compact M R containing the real numbers 0 and
3ω
1
/4. Let s be arbitrary positive real. If
d
γ (s) , γ (0)
= d
(0) , γ (0)
ε
then Equation (12) is satisfied for the number t
0
= 0. If d
γ (s) , γ (0)
< ε
then for t
0
= 3ω
1
/4 we have
d
(t
0
) , γ (t
0
)
= d
3ω
1
4
, c
1
3ω
1
4

>
ω
1
4
=
d (A, B)
2
> ε.
This completes the proof that γ is not recurrent and, therefore, approxima-
tion by closed geodesics does not imply recurrence.
References
[1] B. Aebicher, S. Hong and McCullough, Recurrent geodesics and controlled concentra-
tion points, Duke Math. J., 75(3) (1994), 759-774.
[2] W. Ballman, Singular spaces of non-positive curvature, in Sur les groups hyper-
boliques d’apr´es Gromov (Seminaire de Berne), ´edit´e par E. Ghys et P. de la Harpe,
(a paraitre chez Birkh¨auser), 1990.
[3] M.R. Bridson, Geodesics and geometry in metric simplicial complexes, in Group The-
ory from a Geometrical Viewpoint, (ICTP, Trieste, Italy, March 26 - April 6, 1990),
E. Ghys and A. Haefliger eds. (1991).
[4] Ch. Champetier, Petite simplification dans les groupes hyperboliques, Ann. Fac. Sci.
Toulouse, VI. Ser., Math., 3(2) (1994), 161-221.
[5] C. Charitos, Closed geodesics in ideal polyhedra of dimension 2, Rocky Mountain
Journal of Mathematics, 26(1) (1996), 507-521.
[6] M. Coornaert, Sur les groupes proprement discontinus d’isom´e tries des espaces hy-
perboliques au sens de Gromov, Th`ese U.L.P., Publication de l’IRMA.
[7] , Measures de Patterson-Sullivan sur le bord d’un espace hyperbolique au sens
de Gromov, Pacific J. Math., 159(2) (1993), 241-270.
[8] M. Coornaert, T. Delzant and A. Papadopoulos, G´eom´etrie et th´eorie des groupes,
Lecture Notes in Mathematics, 1441, Springer-Verlag, (1990).
[9] M. Gromov, Hyperbolic groups, in ‘Essays in Group Theory’, MSRI Publ., 8, Springer
Verlag, (1987), 75-263.
[10] , Structures etriques pour les vari´et´es riemanniennes, written with J. La-
fontaine and P. Pansu, Cedic Fernand Nathan, Paris, 1981.
[11] F. Paulin, Constructions of hyperbolic groups via hyperbolization of polyhedra, in
Group Theory from a Geometrical Viewp oint, (ICTP, Trieste, Italy, March 26 - April
6, 1990), E. Ghys and A. Haefliger eds. (1991).
RECURRENCE IN NEGATIVELY CURVED SPACES 79
[12] M. Troyanov, Espaces `a courbure egative et group´es hyperboliques in Sur les groups
hyperboliques d’apr`es Gromov (Seminaire de Berne), ´edit´e par E. Ghys et P. de la
Harp e, (a paraitre chez Birkh¨auser), 1990.
Received September 23, 1998 and revised May 27, 1999.
Agricultural University of Athens
75 Iera Odos
Athens 11855
Greece
E-mail address: bakis@auadec.aua.gr
University of t he Aegean
Karlovassi
Samos 83200
Greece
E-mail address: gtsap@aegean.gr
PACIFIC JOURNAL OF MATHEMATICS
Vol. 195, No. 1, 2000
COHOMOLOGY OF SINGULAR HYPERSURFACES
Bernard M. Dwork
Professor Dwork passed away on May 9, 1998 after a long
illness. The manuscript was completed a few days earlier, and
was submitted to the Pacific Journal of Mathematics follow-
ing his express desire. It is a testimony to his dedication to
mathematics even during his last ill ness - Managing Editor.
Part I.
Our object is to extend earlier work [D1] on singular hypersurfaces defined
over an algebraic number field to singular hypersurfaces defined over function
fields in characteristic zero.
A key role will be played by the results of Bertolin [B1] which in turn is
based upon the Transfer Theorem of Andr´e–Baldassarri–Chiarellotto [DGS,
Theorem VI 3.2].
Let h(A, x) be the generic form of degree d in n+1 variables x
1
, . . . , x
n+1
.
Thus letting F
0
=
u N
n+1
n+1
P
i=1
u
i
= d
,
h(A, x) =
X
u∈F
0
A
u
x
u
where the symbols {A
u
}
u∈F
0
are algebraically independent over Q.
Let E
i
= x
i
x
i
(1 i n + 1), h
i
= E
i
h. Let R(A) be the resultant of
{h
1
, h
2
, . . . , h
n
, h}.
Let V be an absolutely irreducible subvariety of the discriminant locus,
R(A) = 0. Let k be the field of definition of V .
Let be a suitable universal domain in characteristic zero, and let L
be
the ring of all formal sums
L
=
(
ξ
=
X
u∈F
C
u
1
x
u
C
u
π
u
0
)
where F = {u = (u
0
, u
1
, . . . , u
n+1
) | du
0
= u
1
+ · · · + u
n+1
} and where
π
p1
= p, p a rational prime. (Thus π need not be in Ω.)
For λ V , λ rational over we write
D
i,λ
= γ
(E
i
+ πx
0
h
i
(λ, x)) 1 i n + 1
81
82 BERNARD M. DWORK
an endomorphism of L
where γ
is the projection operator
γ
x
v
=
(
0 if any v
i
1
x
v
if all v
i
0.
For each integer , let K
(`)
λ
be the set of all ξ
L
such that ξ
is annihilated
by all monomials of degree in {D
i,λ
}
1in+1
.
In the following, ord refers to a rank one valuation of Ω.
For b R, b > 0, let L
(b) =
n
P
u∈F
C
u
1
x
u
inf
u
(ord C
u
+ u
0
b) > −∞
o
.
Let Γ be an indeterminate and consider the polynomial h(λ, x)+Γh(A, x).
Let R(λ, Γ, A) be the resultant of
E
1
(h(λ, x) + Γh(A, x)), . . . , E
n+1
(h(λ, x) + Γh(A, x))
and write
R(λ, Γ, A) = Γ
e
(ρ
0
(λ, A) + Γρ
1
(λ, A) + Γ
2
ρ
2
(λ, A) + · · · ),
where ρ
0
(λ, A) 6= 0. The key result of the research of Bertolin [B1, Theorem
3.11] states that:
Theorem 1.
K
(`)
λ
L
(τ(n, d, e, ) ord ρ
0
(λ, A) + ε)
for all ε > 0. Here τ (n, d, e, ) depends only on n, d, e and and is indepen-
dent of the coefficients of h(λ, X).
Remark. Be rtolin obtains estimates independent of . The estimate given
here depends upon but is simpler to state. The slight error in [B1, Theorem
3.11] is corrected in [B2].
Corollary 1. If λ V and ρ
0
(λ, A) 6= 0, then K
(`)
λ
L
(ε) for all and
all ε > 0 and for all but a finite set of valuations (depending on λ).
Corollary 2. For λ V with ρ
0
(λ, A) 6= 0, dim K
(`)
λ
is independent of λ.
Proof. We choose a valuation v of k(λ) such that (extending the valuation
of k(λ) to k(λ, A) via the Gauss norm relative to A)
|ρ
0
(λ, A)|
v
= 1
|λ|
v
1.
By the Lemma of Appendix B, we may choos e a generic point λ
0
of V over
k so close to λ v-adically that |λ λ
0
|
v
< 1 and hence |ρ
0
(λ
0
, A)|
v
= 1. Thus
K
(`)
λ
and K
(`)
λ
0
lie in L
(ε) (v-adically) for all ε > 0 and hence T
λ,λ
0
= γ
exp
πX
0
(h(λ
0
, x) h(λ, x)) is an isomorphism between K
(`)
λ
and K
(`)
λ
0
as vector
spaces over Ω.
COHOMOLOGY OF SINGULAR HYPERSURFACES 83
Part II: Koszul compl ex.
In earlier work [D1, Theorem 19.2] we discussed the (cohomological) Koszul
complex of D
1
(0)
, . . . , D
n+1
(0)
operating on K
()
λ
(0)
=
[
`=1
K
(`)
λ
(0)
where λ
(0)
is algebraic over Q. We denote by H
(s)
(K
()
λ
(0)
) the s-th cohomology group
of this complex. We showed:
Theorem 2.
dim H
(s)
K
()
λ
(0)
< .
We also showed [D1, Theorem 17.1] that this dimension can be bounded
in terms of d and n alone.
Note. Equation 19.4 of [D1] is stated without proof. This gap will b e
filled in Appendix A.
Corollary 3. For λ
0
V, dim H
(s)
(K
()
λ
0
) < and if ρ
0
(λ
0
, A) 6= 0, then
dim H
(s)
(K
()
λ
0
) is independent of λ
0
.
Proof. We choose λ
(0)
algebraic over Q such that λ
(0)
V and ρ
0
(λ
(0)
, A) 6=
0. We choose a valuation v such that
ρ
0
λ
(0)
, A
v
= 1,
λ
(0)
v
1
and then choose a generic point λ of V ove r k in as in the proof of Corollary
2. Then T
λ
(0)
provides an isomorphism of K
(s)
λ
(0)
with K
(s)
λ
which induces an
isomorphism of H
(s)
(K
()
λ
(0)
) with H
(s)
(K
()
λ
) for all s. This shows finiteness
for λ generic.
If ρ
0
(λ
0
, A) 6= 0, then by the same argument choosing λ generic close to
λ
0
we c onclude that dim H
(s)
(K
()
λ
0
) = dim H
(s)
(K
()
λ
). If ρ
0
(λ
0
, A) = 0,
then λ
0
lies in a proper subvariety of V and we may use induction on the
dimension.
Notation. For B = {1, 2, . . . , n + 1} and W a vector space over k(λ) let
F
s
(W ) = Hom (
V
s
B, W ).
Corollary 4. For large enough (depending upon V ) and λ a generic point
of V over k,
H
s
K
()
λ
' ker
δ
s+1
, F
s
K
(`)
λ
.
F
s
K
(`)
λ
δ
s,λ
F
s1
K
()
λ

.
(For definition of δ
s
see [D1].)
Proof. There is a natural injection of the right hand space into the left–hand
one induced by the inclusion K
(`)
λ
K
()
λ
. The left–hand space is of finite
dimension and so the mapping is surjective.
84 BERNARD M. DWORK
We now give H
(s)
(K
()
λ
) the structure of a differential module when
viewed as a vector space over k(λ). Let λ
1
, . . . , λ
t
be a transcendence basis
over k of k(λ). Viewing λ
t+1
, λ
t+2
etc. as dependent variables we define for
1 i t
σ
i
= γ
λ
i
πx
0
h
λ
i
.
These operators commute with {D
j,λ
}
1jn+1
and hence induce a set of
commuting operators on H
(s)
(K
()
λ
). If λ
(1)
V , then horizontal elements
are obtained by applying T
λ
(1)
to H
(s)
(K
()
λ
(1)
).
Theorem 3. Let λ
(1)
be a generic point of V . We consider all extensions to
k(λ
(1)
) of valuations of k whose restriction to k(λ
(1)
1
, . . . , λ
(1)
t
) is given by the
Gauss norm of that field relative to λ
(1)
1
, . . . , λ
(1)
t
. For almost all such valu-
ations the horizontal elements converge for |(λ
1
, . . . , λ
t
)(λ
(1)
1
, . . . , λ
(1)
t
)| <
1.
Corollary 5. If k is an algebraic number field, then H
(s)
(K
()
λ
) is a G-
module.
Appendix A.
Let k be a field of characteristic zero and let f (x
1
, . . . , x
n+1
) be a form of
degree d in n + 1 variables. If is an extension of k, let us write L
for the
ring of all polynomials in x
0
, x
1
, . . . , x
n+1
of the form
X
du
0
=u
1
+···+u
n+1
C
u
π
u
0
x
u
C
u
.
We define D
i
= E
i
+ πx
0
f
i
, E
i
= x
i
x
i
, f
i
= E
i
f. The D
i
are commuting
endomorphisms of L
and likewise by restricting to L
k
we obtain commuting
endomorphisms of that ring.
Let L
(resp: L
k
) be the adjoint space of L
(resp: L
k
) and K
(`)
(resp:
K
(`)
k
) the set of all ξ
L
(resp: L
k
) annihilated by all forms in {D
1
, D
2
, ...,
D
n+1
} of degree , where D
i
= γ
(E
i
+ πx
0
f
i
).
Again let K
()
(resp: K
()
k
) be the union
[
`=1
K
(`)
resp :
[
`=1
K
(`)
k
.
Finally we define H
(s)
(K
()
) (resp: H
(s)
(K
()
k
)) to be the s-th cohomol-
ogy group of the (cohomological) Koszul complex of D
1
, . . . , D
n+1
operating
on K
()
(resp: K
()
k
).
Theorem.
(i) K
()
= K
()
k
COHOMOLOGY OF SINGULAR HYPERSURFACES 85
(ii) H
(s)
(K
()
) = H
(s)
(K
()
k
) .
This was stated without proof as Equation (19.4) of [D1].
Proof. We first show for <
(iii) K
(`)
= K
(`)
k
Ω.
We know [D1, Lemma 7.2] that dim
k
K
(`)
k
< , dim
K
(`)
< . We may
view each element of K
(`)
as an –tuple (z
1
, z
2
, . . . ) indexed by a countable
set I. Indeed ξ
K
(`)
implies ξ
=
P
u
C
u
/x
u
. The sum being over all u such
that du
0
= u
1
+ · · · + u
n+1
. Here C
u
= π
u
0
C
u
with C
u
Ω. Identifying the
{C
u
} with the {z
i
}, the condition that ξ
K
(`)
is equivalent to an infinite
set of conditions
X
t
j,i
z
i
= 0 for all j J.
Here t
j,i
k, t
j,i
= 0 for almost all i, for each fixed j. For ξ
K
(`)
k
we have the same set of conditions. Following a suggestion by Wan, by
elementary operations on the rows of the matrix {t
j,i
} the finite dimension
of the subspace is given by the number of zero columns in the reduced
echelon form. The echelon form is the same for the equation over as over
k. I t follows that indeed dim
k
K
(`)
k
< dim K
(`)
< and both are then
equal and K
(`)
= K
(`)
k
Ω. The first assertion now follows.
For a vector space W we write F
s
(W ) = Hom (
V
s
B, W ) with B =
{1, 2, . . . , n + 1}. Then ξ
F
s
(K
()
) implies ξ
=
P
η
i
ξ
i
a finite sum
with ξ
i
F
s
(K
()
k
) and {η
i
} a finite set of elements of linearly indep e n-
dent over k.
If δ
s+1
ξ
= 0 then by linear independence δ
s+1
ξ
i
= 0 and so
ker(δ
s+1
, F
s
(K
()
)) = ker(δ
s+1
, F
s
(K
()
k
)) Ω.
Also
δ
s
F
s1
K
()
= δ
s
F
s1
K
()
k
.
The theorem now follows from the following well known proposition.
Proposition. Let U be a subspace of a linear k space W . Then
W /U ' (W/U) .
Appendix B. Approximation by generic points.
Lemma. Let the origin O be on an irreducible affine variety V defined over
a field k of characteristic zero. Let be a universal domain complete under
a rank one valuation. Then there exists a generic point of V rational over
which is as close as you please to the origin.
We first show the lemma holds if V is a curve in A
n
.
86 BERNARD M. DWORK
Proof. Let P = (x
1
, . . . , x
n
) be a generic point of V over k. Then R =
k[x
1
, . . . , x
n
] has a specialization into k given by (x
1
, . . . , x
n
) 7→ O and hence
there exists a place p of k(V ) with center O. Letting T be a uniformizing
parameter of p, each coordinate x
i
as element of k(V )
p
, the completion at
p of k(V ), is represented as a power series
x
i
= a
i1
T + a
i2
T
2
+ · · · + k
0
[[T ]]
where k
0
is the residue class field at p of k(V ), a finite extension of k. This
series may have zero radius of convergence in the metric of Ω, but if we
choose (as we shall) the uniformizing parameter, T , in k(V ) then the series
represents an algebraic function of T and hence by Eisenstein’s Theorem (or
more elementarily by Clark’s Theorem) the series has a non-trivial radius of
convergence.
Since P is a generic point, these series are not all constant. We think
of P (T ) as function of T for T restricted to a small disk D(0, r
) in Ω–
space. Trivially P (T ) 0 as T 0. We may suppose x
1
is a non–constant
function of T . The theory of Newton polygons shows that the image of
D(0, r
) under x
1
contains elements transcendental over k. This completes
the proof for dim V = 1.
We recall [H, Chapter I, Proposition 7.1]:
Proposition. If V is irreducible of dimension s in A
n
and H is a hyper-
surface not containing V then each irreducible component of H V has
dimension s 1.
Proof of Lemma. Letting V
0
= V we define inductively V
1
V
2
· · · by
the condition that V
j
be an irreducible component of V
j1
{x | x
j
= 0}
which contains the origin. Since
1 + dim V
j1
dim V
j
dim V
j1
, dim V
n
= {0}
there exists j such that V
j
is a curve on V passing through the origin.
We conclude there exists a curve V
0
on V passing through the origin.
Let k
0
k be a field of definition of V
0
. By our previous treatment of
curves there exists P V
0
, P as close as you please to O such that k
0
(P )
is of transcendence degree unity over k
0
. Let P
1
be a coordinate of P of
transcendence degree unity over k
0
.
Let L = k(P
1
), A be the ideal of all f k[x
1
, . . . , x
n
] which are zero
everywhere on V . If g L[x
1,
. . . , x
n
], g = 0 on V then g AL[x] and
hence for each automorphism τ of L/k we have g
τ
= 0 on V . In particular
x
1
P
1
cannot be zero on V as otherwise (x
1
P
1
)
τ
would also be zero on
V and hence P
1
P
τ
1
would be zero on V for every τ which is impossible as
there are nontrivial automorphisms of L/k.
Thus V does not lie in the hyperplane x
1
= P
1
and so the intersection has
an irreducible component W passing through P of dimension s1. Let k
00
be
COHOMOLOGY OF SINGULAR HYPERSURFACES 87
a field of definition of W , P W . By induction there exists Q W, Q P
as small as you please with transc deg k
00
(Q)/k
00
= s 1.
Clearly Q is as c lose as you please to O. It remains to show that s =
trans deg k(Q)/k.
Since Q
1
= P
1
, k(Q) k(P
1
). Hence
s trans deg k(Q)/k = trans deg k(Q)/k(P
1
) + trans deg k(P
1
)/k
trans deg k
00
(Q)/k
00
+ trans deg k
0
(P
1
)/k
0
(s 1) + 1 = s,
the two inequalities being based on
if k
0
k then trans deg k(P
1
)/k trans deg k
0
(P
1
)/k
0
if k
00
k(P
1
) then trans deg k(Q)/k(P
1
) trans deg k
00
(Q)/k
00
.
This completes the proof of the lemma.
Appendix C: (Generalization of Heaviside’s generalized
exponential functions).
In this article we examined the Koszul complex of {D
1
, . . . , D
n+1
} oper-
ating on K
()
λ
. In this appe ndix, we replace L
by
L
0
=
(
X
n∈F
0
A
u
1
x
u
A
u
π
u
0
)
and D
i,λ
= γ
(E
i
+ πx
0
h
i
(x
1
, x)) by D
0
i,λ
= E
i
+ πx
0
h
i
(λ, x). Here
F
0
= {(u
0
, u
1
, . . . , u
n+1
)
Z
n+2
du
0
= u
1
+ · · · + u
n+1
}.
Thus L
0
consists of formal Laurent series in {x
i
,
1
x
i
} i = 1, . . . , n + 1. We
note that L
0
is adjoint to L
0
, the ring of Laurent polynomials with support
in
du
0
= u
1
+ · · · + u
n+1
.
Let D
s
denote the ideal of all forms of degree s in D
1
, . . . , D
n+1
with
coefficients in k(λ). We assert that
L
0
= L + D
s
L
0
.
For s = 1 this follows by the proof of [D2, Lemma 9.7.1]. Assume the
formula valid for some given s then L
0
= L + D
s
(L + DL
0
) L + D
s+1
L
0
,
which completes the proof by induction.
This shows that the natural m apping of L into L
0
induces a surjection
L/D
s
L L
0
/D
s
L
0
. We recall that K
(s)
denotes the annihilator in L
of
D
s
L. Let K
0(s)
denote the annihilator in L
0
of D
s
L
0
. We now know that
the dimension of L
0
/D
s
L
0
is finite and hence the same holds for K
0(s)
. Thus
88 BERNARD M. DWORK
by duality the mapping of K
0(s)
into K
(s)
adjoint to the natural mapping is
injective. This adjoint mapping is γ
.
Conclusion. The mapping γ
maps K
0(s)
into K
(s)
injectively.
We now restrict our attention to the case where h(λ, x) Q[λ, x] and
consider λ
(0)
algebraic over Q. For b > 0, c R, w a finite valuation of Ω,
let L
0
(b, c) be the set of all formal Laurent series ξ
=
P
u∈F
0
B
u
1
x
u
such that
B
u
Ω, and ord(B
uv
) b(u
0
+ v
0
) + c for all u, v, F
0
. Let L
0
(b) =
S
cR
L
0
(b, c), a Banach space. For almost all valuations of Q(λ
(0)
) we have
a completely continuous mapping of L
0
(b
0
) (giving a valuation extending
that of Q(λ
(0)
)) defined by putting F (x) = exp π(x
0
h(λ
(0)
, x)x
q
0
h(λ
(0)
, x
q
))
where q is the order of the residue class field of Q(λ
(0)
) and writing
α
0∗
= F φ, φ : x
u
x
qu
L
0∗
(b
0
)
φ
L
0∗
(b
0
/q)
F
L
0∗
(b
0
/q) L
0∗
(b
0
)
where b
0
is chosen in [0,
p1
p
]. Here F means multiplication by F and the
last map is the inclusion map.
Letting L
(b) = γ
L
0∗
(b) we have the completely continuous endomor-
phism of L
(b
0
) (for almost all valuations of Q(λ
(0)
)) given by
α
= γ
F φ .
By the trace formula the two mappings have the same Fredholm determi-
nant. Defining
W
z
=
[
k
ker
(I zα
)
k
, L
(b
0
)
W
0∗
z
=
[
k
ker
(I zα
0∗
)
k
, L
0∗
(b
0
)
we conclude equality of dimensions and hence
γ
W
0∗
z
= W
z
.
Now K
(`)
λ
(0)
is covered by a union of spaces W
z
and hence by γ
(finite
union of spaces W
0∗
z
) which lies in γ
K
0(`
0
)
λ
(0)
for suitable
0
.
Conclusion. γ
gives a bijection of K
0()
λ
(0)
onto K
()
λ
(0)
provided λ
(0)
is alge-
braic over Q.
We propose to remove the restriction that λ
(0)
be algebraic. Again let
λ
(0)
V be an algebraic point, ρ
0
(λ
(0)
, A) 6= 0. Excluding a finite set of
primes of Q(λ
(0)
) we choose λ generic point of V close to λ
(0)
.
If ξ
K
0()
λ
(0)
, then γ
ξ
K
(`)
λ
(0)
for some and hence γ
ξ
is a finite
sum of elements of spaces W
z
and hence is the image under γ
of a finite
COHOMOLOGY OF SINGULAR HYPERSURFACES 89
sum of elem ents of spaces W
0∗
z
. But γ
is injective, and hence ξ
is a
sum of elements of spaces W
0∗
z
. Thus for almost all valuations of Q(λ
(0)
),
ξ
L
0∗
(b
0
), b
0
<
p1
p
. More precisely K
0()
λ
(0)
lies in L
0∗
(b
0
) for all b
0
> 0
and almost all primes of Q. Hence for almost all primes multiplication
by exp πx
0
h(λ, x) h(λ
(0)
, x)
provides an isomorphism T
0
λ
(0)
of K
0()
λ
(0)
with K
0()
λ
. On the other hand, T
λ
(0)
= γ
T
0
λ
(0)
gives an isomorphism
between K
()
λ
(0)
and K
()
λ
K
0()
λ
(0)
T
0
λ
(0)
K
0()
λ
y
γ
y
γ
K
()
λ
(0)
T
λ
(0)
K
()
λ
.
The horizontal arrows of this c ommutative diagram are isomorphisms.
The first vertical arrow is also an isomorphism. It follows that the second
vertical arrow is also an isomorphism. This completes the proof.
Note. The purpose of the argument involving λ
(0)
is to show that γ
is
injective on K
0()
λ
.
References
[B1] C. Bertolin, G-fonctions et cohomologie des hypersurfaces singuli`eres, Bull. Aus-
tralian Math. Soc., 55 (1997), 353-383.
[B2] , G-fonctions et cohomologie des hypersurfaces singuli`eres II, Bull. Aus-
tralian Math. Soc., 58 (1998), 189-198.
[D1] B. Dwork, B. On the zeta function of a hypersurface III, Ann. Math., 83 (1966),
457-519.
[D2] , Generalized hypergeometric functions, Clarendon Press, Oxford, 1990.
[DGS] B. Dwork, G. Gerotto and F. Sullivan, An Introduction to G-functions, Annals of
Math Studies, 133 (1994), Princeton Univ. Press.
[H] R. Hartshorne, Algebraic Geometry, Springer-Verlag, New York-Heidelberg-Berlin,
1977.
Received October 28, 1998.
Universita degli Studi di Pavoda
35131 Padova
Italy
E-mail address: baldassa@math.unipd.it
PACIFIC JOURNAL OF MATHEMATICS
Vol. 195, No. 1, 2000
RIEMANNIAN MANIFOLDS ADMITTING ISOMETRIC
IMMERSIONS BY THEIR FIRST EIGENFUNCTIONS
Ahmad El Soufi and Sa
¨
ıd Ilias
Given a compact manifold M, we prove that every critical
Riemannian metric g for the functi onal “first eigenvalue of
the Laplacian” is λ
1
-minimal (i.e., (M, g) can be immersed
isometrically in a sphere by its first eigenfunctions) and give
a sufficient condition for a λ
1
-minimal metric to be critical.
In the second part, we consider the case where M is the 2-
dimensional torus and prove that the flat metrics correspond-
ing to square and equilateral lattices of R
2
are the only λ
1
-
minimal and the only critical ones.
Introduction.
Many recent works concerning the spectrum of compact Riemannian man-
ifolds have pointed out the importance of a particular class of Riemannian
metrics which we called in [5] λ
1
-minimal. Recall that a metric g on a
compact m-dimensional manifold M is λ
1
-minimal if the eigenspace E
1
(g)
associated to the first nonzero eigenvalue λ
1
(g) of the Laplacian of g con-
tains a family f
1
, . . . f
k
of functions satisfying:
P
1ik
df
i
df
i
= g. It
follows from a well known result of Takahashi [8] that this last condition is
equivalent to the fact that the map f = (f
1
, . . . f
k
) is a minimal isometric
immersion from (M, g) into the Euclidean sphere S
k1
r
of radius r =
q
m
λ
1
(g)
.
The best known examples of λ
1
-minimal metrics are the standard met-
rics of rank one compact symmetric spaces (i.e., spheres and projective
spaces). More generally, any Riemannian irreducible homogeneous space is
λ
1
-minimal. Also, Yau [9] conjectured that a minimal embedded hypersur-
face of a Euclidean sphere, carrying the induced metric, must be λ
1
-minimal.
In [2], Berger showed that the λ
1
-minimality of a metric g is strongly re-
lated to the extremality of g for a spectral functional involving the k-smallest
eigenvalues of the Laplacian (where k is the multiplicity of λ
1
(g)). Recently,
Nadirashvili [7] considered the functional λ
1
: g 7→ λ
1
(g) defined on the set
of Riemannian metrics of given area on a compact surface M and showed
that the extremal metrics of this functional are λ
1
-minimal (here extremal-
ity is defined in a generalized sense because of the non-differentiability of
λ
1
).
91
92 AHMAD EL SOUFI AND SA
¨
ID ILIAS
In the first part of this paper we generalize Nadirashvili’s theorem to
higher dimensions (Theorem 1.1). We also give a sufficient condition for a
λ
1
-minimal metric to be extremal for λ
1
(Proposition 1.1).
Using results established by us in [4] about λ
1
-minimal metrics we deduce
that (Corollary 1.1), if g is an extremal metric of the λ
1
functional then:
(i) The multiplicity of λ
1
(g) is at least equal to m + 1 and equality holds
only for the standard metric of Euclidean spheres.
(ii) The restriction of the λ
1
functional to the conformal class of g achieves
its maximum at g. In particular, the λ
1
functional has no local minima.
(iii) The metric g is, up to dilatation, the unique extremal metric in its
conformal class.
(iv) If g is not isometric to the standard metric of a Euclidean sphere then
any conformal diffeomorphism of (M, g) is an isometry.
The second part of this paper deals with the classification of λ
1
-minimal
metrics and of the extremal metrics of the λ
1
functional. The only mani-
fold for which this classification was available is the 2-dimensional sphere.
Indeed, on S
2
the standard metric is (up to dilatation) the only one to
be λ
1
-minimal and the only extremal metric for λ
1
(this follows from the
uniqueness of the conformal class on S
2
and property (iii) above).
The m ain theorem of Section 2 (Theorem 2.1) states that in genus one
(i.e., on the torus T
2
) there exists, up to dilatation, exactly two λ
1
-minimal
metrics: The Clifford metric g
cl
and the equilateral metric g
eq
induced from
the Euclidean metric respec tively on R
2
/Z
2
and R
2
/Γ
eq
with Γ
eq
= Z(1, 0)
Z(1/2,
3/2). These two metrics are also the only extremal metrics for λ
1
(Corollary 2.2). Moreover, we prove that for each of them, the standard
embedding (in S
3
for g
cl
and S
5
for g
eq
) is, up to equivalence, the only full
(minimal) isometric immersion by the first eigenfunctions.
Note that a first step towards this classification was achieved by Mon-
tiel and Ros [6] who proved that the only minimal torus immersed in S
3
by its first eigenfunctions is the Clifford torus. They deduced that if the
aforementioned conjecture of Yau is true, then the Clifford torus is the only
minimally embedded torus in S
3
(Lawson’s conjecture).
1. Extremal metrics for the λ
1
functional.
Let M b e a compact smooth manifold of dimension m 2. Denote by
R
0
(M) the set of Riemannian metrics of volume 1 on M . For any g
R
0
(M), we denote by 0 < λ
1
(g) λ
2
(g) ··· λ
k
(g) ··· the increasing
sequence of eigenvalues of the Laplacian
g
of g. The functional:
λ
1
: R
0
(M) R
g 7→ λ
1
(g)
ISOMETRIC IMMERSIONS BY FIRST EIGENFUNCTIONS 93
is continuous but not differentiable in general. However, for any family
(g
t
)
t
of metrics, analytic in t, λ
1
(g
t
) has right and left derivatives w.r.t. t.
Indeed, if (g
t
)
t]δ,δ[
is such a family and if k is the multiplicity of λ
1
(g
0
),
then there exists k analytic families Λ
1,t
, . . . , Λ
k,t
of real numbers and k
analytic families of smooth functions u
1,t
, . . . , u
k,t
such that: i k and t,
g
t
u
i,t
= Λ
i,t
u
i,t
, Λ
i,0
= λ
1
(g
0
) and {u
1,t
, . . . , u
k,t
} is L
2
(g
t
)-orthonormal
(see [1] and [2] for details). Moreover, Berger [2] gave the following formula
for the derivative of Λ
i,t
:
d
dt
Λ
i,t
t=0
=
Z
M
hq(u
i
), hiν
g
0
,
where ν
g
0
is the Riemannian volume element of g
0
, u
i
= u
i,0
, h =
d
dt
g
t
t=0
,
h , i is the inner product induced by g
0
on the space S
2
(M) of symmetric
covariant 2-tensors of M and where for any u C
(M),
q(u) = du du +
1
4
g
0
(u
2
)g
0
.
From the continuity of λ
i
(g
t
) and Λ
i,t
w.r.t. t , we have for t small enough
{Λ
i,t
}
1ik
= {λ
i
(g
t
)}
1ik
and thus λ
1
(g
t
) = min
1ik
{Λ
i,t
}. This proves
the left and right differentiability of λ
1
(g
t
) and gives:
d
dt
λ
1
(g
t
)
t=0
+
= min
1ik
d
dt
Λ
i,t
t=0
= max
1ik
Z
M
hq(u
i
), hiν
g
0
,
and
d
dt
λ
1
(g
t
)
t=0
= max
1ik
d
dt
Λ
i,t
t=0
= min
1ik
Z
M
hq(u
i
), hiν
g
0
.
This suggests the following definition:
Definition 1.1. A metric g R
0
(M) is said to be extremal for the λ
1
functional if for any analytic deformation (g
t
)
t
R
0
(M), with g
0
= g, the
left and right derivatives of λ
1
(g
t
) at t = 0 have opposite signs, i.e.,
d
dt
λ
1
(g
t
)
t=0
+
0
d
dt
λ
1
(g
t
)
t=0
.
This last condition is equivalent to :
λ
1
(g
t
) λ
1
(g) + o(t) as t 0.
Hence our definition of extremality is a equivalent formulation of Nadi-
rashvili’s one [7].
The main result of this section is:
Theorem 1.1. If a Riemannian metric g R
0
(M) is extremal for λ
1
then
it is λ
1
-minimal.
94 AHMAD EL SOUFI AND SA
¨
ID ILIAS
In the 2-dimensional cas e this result was proved by Nadirashvili [7]. Some
of the arguments in our proof are inspired by his. However, the use of the
aforementioned result of Berger makes the proof of this theorem simpler and
more transparent.
Lemma 1.1. If a metric g R
0
(M) is extremal for λ
1
then for any h
S
2
0
(M) = {h S
2
(M);
R
M
tr
g
g
= 0} there exists u E
1
(g)\{0} such
that:
Z
M
hq(u), hiν
g
= 0.
Proof. Suppose that g is extremal for λ
1
and let h S
2
0
(M). We let,
for small t, g
t
=
g+th
V (g+th)
2/m
R
0
(M), where V (g + th) is the Riemann-
ian volume of g + th. Since
d
dt
V (g + th)
t=0
=
1
2
R
M
tr
g
g
= 0, we find
d
dt
g
t
t=0
= h. T he extremality condition implies that the quadratic form
u E
1
(g) 7→
R
M
hq(u), hiν
g
takes on both nonpositive and nonnegative
values, and therefore it admits at least one isotropic direction.
Proof of Theorem 1.1. Let K be the convex hull in S
2
(M) of {q(u), u
E
1
(g)}. The set K {g} is contained in a finite dimensional subspace of
S
2
(M). We claim that g K. Indeed, if g / K then, s ince K is a convex
cone, the Hahn-Banach theorem implies the existence of s S
2
(M) such
that:
Z
M
hs, giν
g
> 0 and for every l K\{0},
Z
M
hl, siν
g
< 0.
The 2-tensor ˜s = s
(
R
M
hs,giν
g
)
mV (g)
g belongs to S
2
0
(M) and, for any u
E
1
(g)\{0},
Z
M
hq(u), ˜siν
g
=
Z
M
hq(u), siν
g
1
mV (g)
Z
M
hs, giν
g
Z
M
|du|
2
ν
g
< 0.
By Lemma 1.1, this contradicts the extremality of g.
Thus g K and there exists w
1
, ...w
d
E
1
(g) such that:
g =
X
1id
q(w
i
) =
X
1id
dw
i
dw
i
+
1
4
X
1id
w
2
i
g
=
X
1id
dw
i
dw
i
+
1
2
λ
1
(g)w
2
i
|dw
i
|
2
g
.
ISOMETRIC IMMERSIONS BY FIRST EIGENFUNCTIONS 95
The traceless part of the last member of this equation must be zero. There-
fore,
X
1id
dw
i
dw
i
|dw
i
|
2
m
g
= 0,
and then:
λ
1
2
X
1id
w
2
i
= 1 +
m 2
2m
X
1id
|dw
i
|
2
.(1)
The λ
1
-minimality of g will follow from the fact that
P
1id
|dw
i
|
2
is con-
stant and equal to m. Indeed, set f =
P
1id
w
2
i
m
λ
1
(g)
. From (1) we
get:
(m 2)∆
g
f = 2(m 2)
λ
1
(g)
X
1id
w
2
i
X
1id
|dw
i
|
2
= 4λ
1
(g)f.
This implies that f = 0 (the Laplacian being a positive ope rator). Therefore
P
1id
w
2
i
=
m
λ
1
(g)
. Replacing in (1) we obtain
P
1id
|dw
i
|
2
= m.
In [4] we showed that λ
1
-minimal metrics satisfy certain remarkable con-
formal properties. Theorem 1.1 tells us that all these properties are still
true for extremal metrics:
Corollary 1.1. Let g R
0
(M) be an extremal metric for λ
1
.
(i) The multiplicity of λ
1
(g) satisfies: mult(λ
1
(g)) m+1, where equality
holds if and only if g is isometric to a standard metric of a Euclidean
sphere.
(ii) For any g
0
C
0
(g) = {g
0
R
0
(M) ; g
0
conformal to g} we have
λ
1
(g
0
) λ
1
(g), and equality holds if and only if g
0
is isometric to g.
In pa rticular, the functional λ
1
does not admit a local minimum in
R
0
(M).
(iii) The metric g is, up to isometry, the only extremal metric of λ
1
in
C
0
(g).
(iv) If (M, g) is not isometric to a Euclidean sphere then any conformal
diffeomorphism of (M, g) is an isometry.
The following is a converse to Theorem 1.1.
Proposition 1.1. Let g R
0
(M) and assume there exists an L
2
(g)-ortho-
normal basis {φ
1
, . . . , φ
k
} of E
1
(g) such that the 2-tensor
P
1ik
i
i
is proportional to g. Then g is extremal for λ
1
.
96 AHMAD EL SOUFI AND SA
¨
ID ILIAS
Proof. Let (g
t
)
t
R
0
(M) be a family of metrics analytic in t with g
0
= g
and set h =
d
dt
g
t
t=0
. With the same notation as above we have for small t:
X
1ik
λ
i
(g
t
) =
X
1ik
Λ
i,t
.
Therefore,
P
1ik
λ
i
(g
t
) is differentiable at t = 0 and
d
dt
X
1ik
λ
i
(g
t
)
t=0
=
d
dt
X
1ik
Λ
i,t
t=0
= trace Q
h
,
where Q
h
is the quadratic form defined on E
1
(g) by:
Q
h
(u) =
Z
M
hq(u), hiν
g
,
and where the trace of Q
h
is taken w.r.t the L
2
inner product induced by g.
Now
trace Q
h
=
X
1ik
Q
h
(φ
i
)
=
Z
M
*
X
1ik
i
i
, h
+
ν
g
+
1
4
X
1ik
Z
M
φ
2
i
, h
ν
g
.
Since
P
1ik
i
i
is proportional to g and
R
M
hg, hiν
g
=2
d
dt
V (g
t
)
t=0
=
0 we have
R
M
D
P
1ik
i
i
, h
E
ν
g
= 0. Moreover, by Takahashi’s theo-
rem
P
1ik
φ
2
i
is constant. Therefore, trace Q
h
=0 and
d
dt
P
1ik
λ
i
(g
t
)
t=0
= 0. The extremality of g then follows from the inequality λ
1
(g
t
)
1
k
P
1ik
λ
i
(g
t
) which is an equality at t = 0.
Remarks.
1) It is known that compact irreducible homogeneous Riemannian spaces
satisfy the hypothesis of Proposition 1.1 (see [8]). Thus, their standard
metrics are extremal for λ
1
.
2) We restricted ourselves to λ
1
. Nevertheless, the results of this para-
graph can be carried over to the case of higher eigenvalues.
2. λ
1
-minimal and extremal metrics on the torus.
Let (M, g) be an orientable compact surface of genus one endowed with a
Riemannian metric g. It is well known that there exists a lattice Γ of R
2
such that (M, g) is conformally equivalent to the torus
R
2
/Γ, g
Γ
, where
g
Γ
is the flat me tric induced from the Euclidean metric on R
2
. The Clifford
torus
T
2
cl
= R
2
/Γ
cl
, g
cl
= g
Γ
cl
with Γ
cl
= Z(1, 0) Z(0, 1), and the equi-
lateral torus
T
2
eq
= R
2
/Γ
eq
, g
eq
= g
Γ
eq
with Γ
eq
= Z(1, 0) Z(1/2,
3/2),
ISOMETRIC IMMERSIONS BY FIRST EIGENFUNCTIONS 97
each admit a natural homothetic minimal embedding into a sphere. Thes e
embeddings, denoted by φ
cl
and φ
eq
, are those induced on T
2
cl
and T
2
eq
from
˜
φ
cl
: R
2
S
3
, where
˜
φ
cl
(x, y) =
1
2
(exp 2x, exp 2y), and
˜
φ
eq
: R
2
S
5
,
where
˜
φ
eq
(x, y)=
1
3
exp 4y/
3, exp 2(x y/
3), exp 2(x + y/
3)
.
Theorem 2.1. Let (M, g) be a compact orientable surface of genus one and
suppose that there exists a full isometric immersion φ = (φ
1
, . . . , φ
n+1
) from
(M, g) in the n-dimensional unit sphere S
n
such that i n+1, φ
i
E
1
(g).
Then either:
(i) (M, g) is isometric to the normalized Clifford torus (T
2
cl
, 2π
2
g
cl
), n = 3
and φ is equivalent to φ
cl
, or
(ii) (M, g) is isometric to the normalized equilateral torus (T
2
eq
,
8π
2
3
g
eq
),
n = 5 and φ is equivalent to φ
eq
.
Recall that an immersion φ into S
n
is full if its image is not contained in a
great sphere of S
n
. Two immersions φ and ψ into S
n
are called equivalent if
there exists an isometry R of S
n
such that φ = R ψ. A direct consequence
of Theorem 2.1 is:
Corollary 2.1. A compact genus one orientable surface (M, g) is λ
1
-mini-
mal if and only if it is homothetic to (T
2
cl
, g
cl
) or (T
2
eq
, g
eq
).
As the metrics g
cl
and g
eq
trivially satisfy the hypothesis of Proposition
1.1 we have the following:
Corollary 2.2. Let M be a compact orientable surface of genus one. A
metric g on M is extremal for λ
1
if and only if (M, g) is homothetic to
(T
2
cl
, g
cl
) or (T
2
eq
, g
eq
).
The proof of Theorem 2.1 is base d on the following Propositions 2.1 and
2.2 which are valid in a more general setting.
Proposition 2.1. Let (M, g) be a n-dimensional compact Riemannian ho-
mogeneous manifold non homothetic to S
n
. If a metric g = fg
0
, conformal
to g
0
, is λ
1
-minimal, then f is constant on M.
Proof. As (M, g) is λ
1
-minimal non homothetic to S
n
then any conformal
diffeomorphism of (M, g) is an isometry (cf. [4]). It follows that any isome-
try of (M, g
0
) is also an isometry of (M, g). Thus the function f is invariant
under the isometry group of (M, g
0
). The result follows from the homogene-
ity of (M, g
0
).
Proposition 2.2. Let η
1
, η
2
, . . . , η
N
be N continuous functions on a do-
main of R
m
and assume that the N
2
functions: 2η
j
(1 j N), η
k
+ η
l
and η
k
η
l
(1 k < l N) are non-constant and mutually distinct modulo
2π. If φ = (φ
1
, . . . , φ
n+1
) is a map from to S
n
such that all its compo-
nents φ
i
are in the vector space generated by {cos η
j
, sin η
j
, 1 j N},
98 AHMAD EL SOUFI AND SA
¨
ID ILIAS
then there exists an isometry R of S
n
such that
R φ = (α
1
exp
j
1
, α
2
exp
j
2
, . . . , α
r
exp
j
r
, 0, . . . , 0) ,
where r (n + 1)/2, j
1
, . . . , j
r
{1, . . . , N} and α
1
, . . . , α
r
are positive
constants satisfying
P
1jr
α
2
j
= 1. In particular, R(φ(Ω)) S
1
(α
1
) ×
S
1
(α
2
) × ··· × S
1
(α
r
) × {0}.
The proof of this proposition is quite elementary and can be omitted.
Proof of Theorem 2.1. In view of Proposition 2.1, it suffices to consider the
case where the metric g is flat. It is well known that there exists (a, b) R
2
;
0 a
1
2
, b > 0 and a
2
+ b
2
1, such that (M, g) is homothetic to
T
2
a,b
= R
2
/Γ(a, b), g
ab
= g
Γ(a,b)
with Γ(a, b) = Z(1, 0) Z(a, b) (cf. [3]).
Now the existence of an isometric immersion from (M, g) into the unit sphere
by the first eigenfunctions implies that λ
1
(g) = 2. Since λ
1
(g
ab
) =
4π
2
b
2
,
(M, g) is in fact isometric to
T
2
a,b
,
2π
2
b
2
g
ab
. Let E
a,b
be the first eigenspace
of g
ab
and φ :
T
2
a,b
,
2π
2
b
2
g
ab
S
n
a full isometric immersion whose compo-
nents φ
i
E
a,b
.
If a
2
+ b
2
> 1 then the dimension of E
a,b
is 2 and there is no such φ.
If a
2
+ b
2
= 1 and (a, b) 6= (1/2,
3/2) then E
a,b
is generated by cos η
j
,
sin η
j
, j 2, with η
1
(x, y) =
2πy
b
and η
2
(x, y) = 2π
x
ay
b
. From
Proposition 2.2, it follows that n = 3 and, up to an isometry of S
3
,
φ has the form φ = (α
1
exp(
1
), α
2
exp(
2
)) with α
1
> 0, α
2
> 0
and α
2
1
+ α
2
2
= 1. As φ is isometric we deduce that a = 0, b = 1 and
α
1
= α
2
=
2/2. Thus (M, g) is isometric to (T
2
cl
, 2π
2
g
cl
) and φ is
equivalent to φ
cl
.
If (a, b) = (1/2,
3/2) then E
a,b
is generated by cos η
j
, sin η
j
, j 3,
where η
1
(x, y) = 4πy/
3, η
2
(x, y) = 2π
x
y
3
and η
3
(x, y) =
2π
x +
y
3
. As before n 5 and, up to isometry, φ = (α
1
exp(
1
),
α
2
exp(
2
), α
3
exp(
3
)) where α
1
, α
2
and α
3
are nonnegative con-
stants such that α
2
1
+ α
2
2
+ α
2
3
= 1. As φ is isometric we obtain
α
1
= α
2
= α
3
=
3/3. Thus φ is equivalent to φ
eq
.
References
[1] S. Bando and H. Urakawa, Generic properties of the eigenvalue of Laplacian for com-
pact Riemannian manifolds, ohoku Math. J., 35 (1983), 155-172.
[2] M. Berger, Sur les premi`eres valeurs propres des vari´et´es riemanniennes, Compositio.
Math., 26 (1973), 129-149.
ISOMETRIC IMMERSIONS BY FIRST EIGENFUNCTIONS 99
[3] M. Berger, P. Gauduchon and E. Mazet, Le spectre d’une vari´et´e riemannienne, Lec-
ture Notes in Math., Vol. 194, Springer, 1971.
[4] A. El Soufi and S. Ilias, Immersions minimales, premi`ere valeur propre du laplacien et
volume conforme, Math. Ann., 275 (1986), 257-267.
[5] , Majoration de la seconde valeur propre d’un op´erateur de Schr
¨
ødinger sur une
vari´et´e compacte et applications, J. of Funct. Anal., 103(2) (1992), 294-316.
[6] S. Montiel and A. Ros, Minimal immersions of surfaces by the first eigenfunctions and
conformal area, Invent. Math., 83 (1986), 153-166.
[7] N. Nadirashvili, Berger’s isoperimetric problem and minimal immersions of surfaces,
Geom. and Funct. Anal., 6 (1996), 877-897.
[8] T. Takahashi, Minimal immersions of riemannian manifolds, J. Math. Soc. Jap., 18
(1966), 380-385.
[9] S.-T. Yau, Problems section. Seminar on Differential Geometry, Ann. Math. Stud.,
Vol. 102, Princeton University Press, 1982.
Received October 20, 1998 and revised June 15, 1999.
Laboratoire de mathematiques et physique theo rique
Universite de Tours
Parc de Grandmont
37200 Tours
France
E-mail address: elsoufi@univ-tours.fr
Laboratoire de mathematiques et physique theo rique
Universite de Tours
Parc de Grandmont
37200 Tours
France
E-mail address: ilias@univ-tours.fr
PACIFIC JOURNAL OF MATHEMATICS
Vol. 195, No. 1, 2000
THE RUBINSTEIN–SCHARLEMANN GRAPHIC OF A
3-MANIFOLD AS THE DISCRIMINANT SET OF A
STABLE MAP
Tsuyoshi Kobayashi and Osamu Saeki
We show that Rubinstein–Scharlemann graphics for 3-mani-
folds can be regarded as the images of the singular sets
(: discriminant set) of stable maps from the 3-manifolds into
the plane. As applications of our understanding of the graphic,
we give a method for describing Heegaard surfaces in 3-mani-
folds by using arcs in the plane, and give an orbifold version
of Rubinstein–Scharlemann’s setting. Then by using this set-
ting, we show that every genus one 1-bridge position of a non-
trivial two bridge knot is obtained from a 2-bridge position in
a standard manner.
1. Introduction.
In this paper, we show that Rubinstein-Scharlemann graphics for 3-manifolds
can be regarded as the images of the singular sets (: discriminant set) of sta-
ble maps from the 3-manifolds into the plane, and as applications, we give
a method for describing Heegaard surfaces in 3-manifolds by using arcs in
the plane, and give an orbifold version of Rubinstein-Scharlemann’s setting.
Then by using this setting, we show that every genus one 1-bridge position
of a non-trivial two bridge knot is obtained from a 2-bridge position in a
standard manner.
In [18], Rubinstein-Scharlemann introduced a powerful machinery, which
is called a graphic, for studying Heegaard s plittings of 3-manifolds, and suc-
ceeded to obtain deep results on the Reidemeister-Singer distance of two
strongly irreducible Heegaard splittings of a 3-manifold. We note that Ru-
binstein and Scharlemann derived a graphic from two Heegaard splittings of
a 3-manifold via Cerf theory [5]. Then the purpose of this paper is to intro-
duce another way for understanding the graphic. That is, we show that we
can regard a graphic as the image of the singular set of a “stable map”(for
definition, see Sect. 3) from the 3-manifold into the plane R
2
(Theorem 4.2).
An immediate consequence of this is that we can regard a Heegaard sur-
face as the preimage of an arc in R
2
, and as an application of our understand-
ing, we first give a method for instructing a procedure for deforming one
Heegaard surface to the other by using the arcs as above (Proposition 5.4),
101
102 T. KOBAYASHI AND O. SAEKI
and describe how stabilization works in this setting (Proposition 5.6). In
[19], Rubinstein and Scharlemann give a generalization of results in [18]
for 3-manifolds with boundary. As the second application, we will give an-
other formulation for generalizing the idea in [18] for link spaces. In fact,
we will introduce an orbifold version of the Rubinstein-Scharlemann type
argument(Sect. 6), and, by using this, we show that any genus one 1-bridge
position of a 2-bridge knot is obtained from a 2-bridge position in a standard
manner (Theorem 8.2).
2. Rubinstein-Scharlemann graphic.
Throughout this paper, we work in the differential category, and for standard
terminology in 3-dimensional topology, we refer to [9], and [11].
In this section, we quickly review the setting of Rubinstein-Scharlemann’s
paper [18].
Let M be a closed orientable 3-manifold.
Definition 2.1. We say that a decomposition M = A
P
B is a (genus g)
Heegaard splitting of M if A, B are 3-dimensional genus g handlebodies in
M such that M = A B, A B = A = B = P . Then P is called a (genus
g) Heegaard surface of M.
Definition 2.2. A disk D properly embedded in a handlebody H is called
a meridian disk of H if D is an essential simple close d curve in H.
Definition 2.3. A Heegaard splitting M = A
P
B is stabilized, if there
are meridian disks D
A
, D
B
of A, B respectively such that D
A
and D
B
intersects transversely in a single point.
Remark 2.4. We note that a genus g Heegaard splitting M = A
P
B is
stabilized if and only if there exists a genus g 1 Heegaard splitting A
0
P
0
B
0
such that A
P
B is obtained from A
0
P
0
B
0
by adding a “trivial” handle.
Then we say that M = A
P
B is obtained from A
0
P
0
B
0
by a stabilization.
Definition 2.5. A Heegaard splitting M = A
P
B is reducible, if there
exist meridian disks D
A
, D
B
of A, B respectively such that D
A
= D
B
.
Definition 2.6. A Heegaard splitting M = A
P
B is weakly reducible, if
there exist meridian disks D
A
, D
B
of A, B res pectively such that D
A
D
B
= .
Remark 2.7. It is easy to see that if a Heegaard splitting M = A
P
B
is reducible then it is weakly reducible. And it is also easy to see that if
M = A
P
B is stabilized and is not a genus one Heegaard splitting of the
3-sphere S
3
, then it is reducible.
Remark 2.8. It is known, by Haken [8], that if M is reducible (that is,
if M is a connected sum of two 3-manifolds which are not S
3
), then any
Heegaard splitting of M is reducible.
RUBINSTEIN–SCHARLEMANN GRAPHIC 103
Remark 2.9. It is known, by Casson-Gordon [4], that if a Heegaard split-
ting M = A
P
B is weakly reducible, then either it is reducible, or M
contains an incompressible surface.
Setting of Rubinstein-Scharlemann graphic.
Let A
P
B, X
Q
Y be a pair of Heegaard splittings of a closed 3- manifold
M. Let Θ
A
, Θ
B
, Θ
X
, Θ
Y
be spines of A, B, X, Y respectively such that
(except for genus 0, or 1 Heegaard splittings) each vertex of Θ
A
, Θ
B
, Θ
X
,
Θ
Y
has valency 3 (see Figure 2.1). Note that for a genus 0 handleb ody (:the
3-ball) B
3
, we let the spine of B
3
be a point in IntB
3
, and for a genus 1
handlebody (:solid torus), we let the spine be a core circle of the solid torus.
Then M
A
Θ
B
) is homeomorphic to P ×(0, 1), where P ×{ε} is close to
Θ
A
, and P ×{1ε} is close to Θ
B
for a small ε > 0. We let P
s
be the s urface
in M corresponding to P ×{s}. Then, by regarding P
0
= Θ
A
, and P
1
= Θ
B
,
we obtain a continuous map H : P × I M such that H(P, s) = P
s
, and we
call this a sweep-out associated to A
P
B. Similarly we obtain a sweep-out
G : Q × I M associated to X
Q
Y , and set G(Q, t) = Q
t
.
Here we may suppose that Θ
A
Θ
B
and Θ
X
Θ
Y
, Θ
A
Θ
B
and G, and
Θ
X
Θ
Y
and H are in general positions. This implies that H(P × [0, ε])
H(P × [1 ε, 1]) and G”, G(Q × [0, ε]) G(Q × [1 ε, 1]) and H are in a
“standard” position. That is:
Regard G(Q×(0, 1))
G
1
Q×(0, 1)
proj.
(0, 1) as a height function.
Then except for a neighborhood of the maxima and minima (with
respect the height function) and vertices of Θ
A
, each component
of H(P × [0, ε]) Q
t
is a meridian disk intersecting Θ
A
in one
point (for a small ε > 0), and in the neighborhoods Q
t
looks as
in Figure 2.2. The same picture holds for the other pair.
Figure 2.1.
104 T. KOBAYASHI AND O. SAEKI
Figure 2.2.
Then, by Cerf [5], we see that for “generic” sweep-outs H, G we obtain
a stratification of Int(I × I) which consists of four parts below.
Regions: Each region is a component of the subset of Int(I × I) consist-
ing of values (s, t) such that P
s
and Q
t
intersect transversely, and this
is an open set.
Edges: Each edge is a component of the subset consisting of values
(s, t) such that P
s
and Q
t
intersect transverse ly except for one non-
degenerate tangent point. The tangent point is either a “center” or
a “saddle”. Edge is a 1-dimensional s ubset of Int(I × I), which is
monotonously increasing or decreasing.
Figure 2.3.
Figure 2.4.
Crossing vertices: Each crossing vertex is a component of the subset
consisting of points (s, t) such that P
s
and Q
t
intersect transversely
except for two non-degenerate tangent points. Crossing vertex is an
isolated point in Int(I × I). In a neighborhood of a crossing vertex,
four edges are coming in, where one can regard the crossing vertex as
RUBINSTEIN–SCHARLEMANN GRAPHIC 105
the intersection of two edges `
1
, `
2
with the signs of the slopes of `
1
and `
2
are either different or the same.
Figure 2.5.
Figure 2.6.
Birth-death vertices: Each birth-death vertex is a component of the
subset consisting of points (s, t) such that P
s
and Q
t
intersect trans-
versely except for a single degenerate tangent point. In particular,
there is a parametrization (λ, µ) of I × I such that P
s
= {(x, y, z)|z =
0}, and Q
t
= {(x, y, z)|z = x
2
+ λ + µy + y
3
}. A birth-death vertex is
an isolated point in Int(I × I), and in a neighborhood of a birth-death
vertex, two edges `
1
, `
2
are coming in, with one from c enter tangency,
the other from saddle tangency, and the signs of the slopes of `
1
and
`
2
the same.
106 T. KOBAYASHI AND O. SAEKI
Figure 2.7.
Let Γ be the union of edges and vertices above. By the above, Γ is a
1-complex in Int(I × I). Since we have assumed that H, G are standard in
a regular neighborhood of Θ
A
Θ
B
Θ
X
Θ
Y
, we see that the 1-complex
Γ naturally extends to (I × I). We abuse Γ to denote this 1-complex, and
we call Γ a graphic (obtained from the sweep-outs H, G).
Figure 2.8
Example 2.10. We show that there exist infinitely many 3-manifolds, and
a pair of Heegaard splittings, say A
P
B, X
Q
Y , of each 3-manifold such
that the corresponding graphic is as in Figure 2.9.
Figure 2.9.
Note that the picture admits 4-fold (Z
2
Z
2
) symmetry, and we will give
an explicit description of the Heegaard surfaces belonging to the lower-left
RUBINSTEIN–SCHARLEMANN GRAPHIC 107
quarter of I ×I, which can be naturally extended to the whole picture under
the above symmetry.
We may regard the Heegaard surface P as P
1/2
, and Q as Q
1/2
. Then
A and A Q
ε
(0 ε 1/2) look like as follows. (Here we suppose that
A admits a symmetry generated by ϕ
1
, ϕ
2
in Figure 2.10, where ϕ
1
is an
orientation preserving involution which changes the right side and the left
side of A with the fixed point set an arc properly embedded in A, and ϕ
2
is
an orientation reversing involution which changes the right side and the left
side of A with the fixed point set a disk properly embedded in A.)
A Q
0
(= A Θ
X
) is a 1-complex as in Figure 2.10. Then, when ε is
sufficiently small, A Q
ε
is the frontier of a regular neighborhood of the 1-
complex (the surface is homeomorphic to a torus with one hole). If we make
ε bigger, then we come to the point (1) of Figure 2.9 and simlutaneously
four points in the boundary of a torus Q
ε
tend to four directions as in (1)
of Figure 2.10.
Figure 2.10.
Then, by m aking ε bigger further, we come to the point (2) and, then, (3)
(: ε = 1/2), where the corresponding figures of AQ
ε
look as in Figure 2.11.
When we come to the point (2), the boundary component touches itself
simultaneously in two places. In the right side, a band is produced, and, in
the left side, the surface is boundary compressed when we pass the point (2).
Note that A Q
1/2
is a vertical surface (which is a disk with two holes),
which is located in the middle of A, that is, A Q
1/2
is invariant under ϕ
1
,
and ϕ
1
exchanges the components of A Q
1/2
.
Figure 2.11.
108 T. KOBAYASHI AND O. SAEKI
Figures 2.12 describes the deformations (4) (5). When we come to
(5) from (1) with passing the edge containing (4), a torus with one hole is
boundary compressed and becomes an annulus.
Figure 2.12.
Figure 2.13 describes the surface of (6). When we come to (6) from (5),
two points of a boundary component of an annulus tend to the right side
and touch and a band is produced, and simultaneously two points contained
in diferent boundary components of an annulus tend to the back of A and
the surface is boundary copressed.
Figure 2.13.
Figures 2.14 and 2.15 des cribe the surfaces of (7), (8), and (9). At (7)
Q A is an inessential disk properly embedded in A. When we come to (8)
from (7), two points of the boundary of a disk tend to the right side and
touch at a middle part of A, and simultaneously two points in the boundary
of a disk tend to the back of A and the surface is boundary comressed. As
a result, an inessential disk be come s a separating essential disk in A.
Figure 2.14.
RUBINSTEIN–SCHARLEMANN GRAPHIC 109
Figure 2.15.
Figure 2.16 describes (10). When we come to (10) from (7) with passing
an edge, an inessential disk is boundary compressed and becom es two non-
separating disks.
Figure 2.16.
Let A Q
1ε
= ϕ
1
(A Q
ε
) (0 ε 1/2), which gives the whole descrip-
tion of Q
t
(0 t 1) in A.
Let B be a copy of A, and f : A B the homeomorphism induced by the
identification, and φ = f|
A
: A B the corresponding homeomorphism.
Let F be the pattern on A induced by Q
t
’s. Le t M be the 3-manifold
obtained by attaching A to B by the homeomorphism φ ϕ
0
2
, where ϕ
0
2
:
A A is isotopic to ϕ
2
with ϕ
0
2
(F) = F, and ϕ
0
2
(`
2
) = `
1
, ϕ
0
2
(`
1
) = `
2
.
(Note that M is actually the connected sum of two S
2
× S
1
’s.) We note
that F is invariant under ϕ
2
, and, hence, the surfaces Q
t
(0 < t < 1) in A
are matched to the surfaces f ϕ
2
(Q
t
) (0 < t < 1) in B to make a system
of closed surfaces, say Q
t
again, in M. It is directly observed from the
pictures that Q
t
gives a sweep out G : Q × I M, and, by construction,
we immediately see that the corresponding graphic is as in Figure 2.9.
Then let `
1
, `
2
be the components of (A Q
1/2
) as in Figure 2.11, and
D
i
: A A (i = 1, 2) the Dehn twist along `
i
. For a pair of integers
(p, q), we let M
(p,q)
be the manifold obtained by attaching B to A by the
homeomorphism φ D
p
1
D
q
2
ϕ
2
: A B. Since there is a regular
neighborhood N(`
i
, A) such that F restricted to N(`
i
, A) is a foliation
by circles parallel to `
i
, we can arrange so that the configuration of the
sweep-outs H and G are respected in M
(p,q)
(and, hence, the corresponding
graphic is the same as above). It is easy to see that M
(p,q)
is a connected
110 T. KOBAYASHI AND O. SAEKI
sum of two lens spaces L(p, 1), and L(q, 1), which implies the existence of
infinitely many examples.
We note that it is easy to see from Figure 2.11, that the Heegaard surfaces
are isotopic in each of the examples.
Example 2.11. By using the arguments in Example 2.10, we show that
there exist infinitely many 3-manifolds, and a pair of Heegaard splittings,
say A
P
B, X
Q
Y , of each 3-manifold such that the corresponding graphic
is as in Figure 2.17.
Figure 2.17.
As in Figure 2.10, Q
ε
is a torus with one hole which is the frontier of a
regular neighborhood of a 1-complex for a small ε > 0. Four points in the
boundary of the torus with one hole tend to four directions according as the
expansion of the regular neighborhood.
Figure 2.18.
When we come to (3a) from (1) with passing (2), a band which goes
around the right handle twice is produced, and a slit in the surface which
goes around the left handle twice occurs to produce a boundary compression.
As a result, a torus with one hole becomes a disk with two holes.
RUBINSTEIN–SCHARLEMANN GRAPHIC 111
Figure 2.19.
Figure 2.20a.
Note that in (3b), Q
1/2
is invariant with respect to the involution ϕ
1
.
Figure 2.20b.
When we come to (6) from (3b) with passing (4), two boundary compres-
sions occur, and a disk w ith two holes becomes a separating essential disk
in A.
Figure 2.21.
112 T. KOBAYASHI AND O. SAEKI
In the following, we show pictures with turning back to front for the
convenience of drawing.
When we come to (5) from (3b) with passing an edge, the right band is
boundary compressed and a disk with one hole becomes an annulus.
When we come to (6) from (5) with passing an edge, an annulus is bound-
ary compressed to become a separating essential disk.
Figure 2.22.
Figure 2.23.
When we come to (6) from (1) with passing (7), two bands, one of which
goes around the left handle once and the other goes around the left handle
twice are produced.
Figure 2.24.
When we come to (8) from (1) with passing an edge, a torus with one
hole is boundary compressed and the punctured torus become an annulus.
When we come to (8) from (6) with passing an edge, a band which goes
around the left handle once is attached and the disk becomes an annulus.
RUBINSTEIN–SCHARLEMANN GRAPHIC 113
Figure 2.25.
When we come to (6) from (10) with passing (9), two points of the bound-
ary of an inessential disk tend to the right side and touch at a middle part
of A, and simultaneously two points in the boundary of a disk tend to the
back of A and touch to produce a boundary compression. As a result, an
inessential disk becomes a separating essential disk in A.
Figure 2.26.
When we come to (11) from (10) with passing an edge, an inessential disk
is boundary compressed and becomes two non-separating disks.
Figure 2.27.
Note that, by using the arguments in the previous example, we can show
that such 3-manifolds are obtained from A and B by pasting their boundaries
applying the Dehn twists along ` in Figure 2.20b. Note that, as a result of
this construction, we obtain 3-manifolds each of which is a union of two
Seifert fibered spaces with orbit space a disk with two exceptional fibers
of index two, and the exterior of a (2, 2n)-torus link (see [13]). Note that
except in one case (case n = 0) they are Haken manifolds.
114 T. KOBAYASHI AND O. SAEKI
3. Stable maps.
The purpose of this section is to show that any differentiable map from an n-
manifold into a surface can be deformed to an “excellent” (:stable) map, and
this assertion is an es sential part of this paper. In the following, manifolds
have countable basis and all manifolds and maps are assumed to be C
.
Let M be a connected n-dimensional manifold (possibly with boundary)
with n 2 and N a surface without boundary. For a smooth map f : M
N, S(f) denotes the singular set of f; i.e., S(f) is the set of the points in M
where the rank of the differential df is strictly less than 2. The discriminant
set is the image of the singular set, f(S(f )). We denote by C
(M, N) the
space of the smooth maps of M into N endowed with the Whitney C
topology (fine topology) (see [7, 10]).
Definition 3.1. A smooth map f : M N is stable if there exists a
neighborhood U of f in C
(M, N) such that for each g U there exist
diffeomorphisms H : M M and h : N N which make the following
diagram commutative.
M
f
N
H
y
y
h
M
g
N
Definition 3.2. Let f : M N be a proper smooth map of an n-dimensio-
nal manifold M (n 2) into a surface N without boundary. For an open
set U of N, we say that f is excellent on U, if f
1
(U) M = and
the following conditions are s atisfied: For all p f
1
(U), there exist local
coordinates (u, x, y
1
, · · · , y
n2
) centered at p and (X, Y ) centered at f(p)
such that f has one of the following forms:
L
0
) X f = u, Y f = x (p : regular point)
L
1
) X f = u, Y f = ±x
2
+
n2
X
i=1
±y
2
i
(p : fold point)
L
2
) X f = u, Y f = ux x
3
+
n2
X
i=1
±y
2
i
(p : cusp point);
and
G
1
) If p f
1
(U) is a cusp point, then f
1
(f(p)) S(f) = {p},
G
2
) f|
(S(f )f
1
(U)−{cusp points})
is an immersion with normal crossings.
RUBINSTEIN–SCHARLEMANN GRAPHIC 115
Note that it is well known that a prop e r smooth map f : M N of a
manifold M with M = is stable if and only if f is excellent on N (see
[7, 14]). The terminology “excellent”comes from [22, §2].
The main purpose of this section is to prove the following.
Theorem 3.3. Let f : M N be a proper smooth map. Suppose that F
is a closed 2-dimensional submanifold (possibly with boundary) of N such
that f
1
(Int F ) M. Furthermore we suppose that f is excellent on a
neighborhood of F and that f is transverse to F . Furthermore, let V
be an arbitrary open neighborhood of f in C
(M, N). Then there exists a
smooth map g : M N such that
(1) g V,
(2) g = f on f
1
(F ) = g
1
(F ),
(3) g is excellent on a neighborhood of N Int F .
Remark 3.4. In Theorem 3.3 the condition (2) is essential. In fact, if we
drop the condition (2), it has already been well known.
In the following, we use the following notation. For manifolds X and Y ,
J
k
(X, Y ) denotes the k-jet bundle over X×Y , i.e., J
k
(X, Y ) = {(x, y, j
k
(x))|
x X, y Y f : X Y : C
s.t. f(x) = y}. For integers s 1 and k 0,
J
k
s
(X, Y ) denotes the s-fold k-jet bundle ([7, p. 57]). We denote by X
(s)
the subset of X
s
= X × · · · × X (the s-fold product space of X) consisting
of the elements (x
1
, · · · , x
s
) such that x
i
6= x
j
for i 6= j. We denote by π :
J
k
s
(X, Y ) X
(s)
× Y
s
the canonical projection and by π
Y
: J
k
s
(X, Y ) Y
s
the natural projection to the target. We se t
s
Y
= {(y, · · · , y) Y
s
} and
d :
s
Y
Y is the natural identification map. Furthermore, for a smooth
map f : X Y , j
k
s
f : X
(s)
J
k
s
(X, Y ) denotes the s-fold k-jet of f . (For
details, see [7, Chapter II, §4].)
In order to prove Theorem 3.3, we need the following.
Proposition 3.5. Let f : X Y be a smooth map between manifolds
(Y need not be a surface). Let W be a submanifold of J
k
s
(X, Y ) such that
π
Y
(W )
s
Y
. Suppose that U is an open subset of Y and that V is an open
neighborhood of f in C
(X, Y ). Then there exists a smooth map g : X Y
such that
(1) g V,
(2) g = f on f
1
(U) = g
1
(U),
(3) j
k
s
g is transverse to W on W π
1
Y
(d
1
(Y U)).
Proof. Set W
0
= W π
1
Y
(d
1
(Y U)), which is an open submanifold of
W . Then there exists a countable family {W
r
}
r=1
of open sets of W
0
with
the following properties (a)-(f).
(a)
r=1
W
r
= W
0
.
116 T. KOBAYASHI AND O. SAEKI
(b) W
r
W
0
, where W
r
denotes the closure of W
r
in J
k
s
(X, Y ).
(c) W
r
is compact.
(d) There exist coordinate neighborhoods U
r,1
, · · · , U
r,s
in X and V
r,1
, . . . ,
V
r,s
in Y such that {U
r,i
}
s
i=1
are mutually disjoint and π(W
r
) U
r,1
× · · · ×
U
r,s
× V
r,1
× · · · × V
r,s
.
(e) U
r,i
is compact for 1
i s.
(f) V
r,i
U = for 1
i s.
Using this family {W
r
}
r=1
in the argument of [7, pro of of Theorem 4.13
(p. 58)] or [15, pp. 311-312], we see that there exists a smooth map g
r
:
X Y such that j
k
s
g
r
: X
(s)
J
k
s
(X, Y ) is transverse to W on W
r
and that
g
r
= f on f
1
(U) = g
1
r
(U) in an arbitrary neighborhood of f in C
(X, Y ).
Thus, putting
C
f,U
= {g C
(X, Y ) : g = f on f
1
(U) = g
1
(U)},
we see that
D
r
= C
f,U
{g C
(X, Y ) : j
k
s
g is transverse to W on W
r
}
is dense in C
f,U
. On the other hand, D
r
is open by [7, Lemma 4.14 (p. 57)].
The proposition is proved if we show that
r=1
D
r
is dense in C
f,U
. Thus we
have only to show that C
f,U
is a Baire space (see [7, Definition 3.2 (p. 44)]).
First note that C
f,U
is a closed subset of C
(X, Y ). Then by imitating the
proof of [7, Proposition 3.3 (p. 44)], we see easily that C
f,U
is a Baire space.
This completes the proof.
Let M and N be as in Theore m 3.3. We consider some submanifolds of the
(multi-)jet bundles as follows. For the jet bundle J
3
(M, N), we consider the
four submanifolds Σ
n1,0
, Σ
n,0
, Σ
n1,1,0
and Σ
n1,1,1,0
as defined in [1] (or [7,
p. 156, Sect. 5]. Note that their codimensions are equal to n 1, 2n, n and
n+1 respectively by [1, Theorem (6.2)]. For the multi-jet bundle J
3
2
(M, N),
we consider
S
1
2
= {(j
3
f(p), j
3
g(q)) : f(p) = g(q), j
3
f(p) Σ
n1,0
, j
3
g(q) Σ
n1,0
},
S
2
2
= {(j
3
f(p), j
3
g(q)) : f(p) = g(q), j
3
f(p) Σ
n1,0
, j
3
g(q) Σ
n1,1,0
},
S
3
2
= {(j
3
f(p), j
3
g(q)) : f(p) = g(q), j
3
f(p) Σ
n1,1,0
, j
3
g(q) Σ
n1,1,0
}.
For the multi-jet bundle J
3
3
(M, N), we consider
S
1
3
= {(j
3
f(p), j
3
g(q), j
3
h(r)) : f(p) = g(q) = h(r),
j
3
f(p), j
3
g(q), j
3
h(r) Σ
n1,0
}.
Note that S
1
2
, S
2
2
, S
3
2
and S
1
3
are easily seen to be submanifolds and that
their codimensions are equal to 2n, 2n + 1, 2n + 2 and 3n + 1 respectively.
For a smooth map f : M N, we have the following facts:
(1) j
3
f is transverse to Σ
n1,0
, Σ
n,0
, Σ
n1,1,0
and Σ
n1,1,1,0
if and only if
f exhibits only fold and cusp points as its singularities.
RUBINSTEIN–SCHARLEMANN GRAPHIC 117
(2) Suppose f satisfies (1). Then j
3
3
f is transverse to S
1
3
if and only if
f|
(S(f )−{cusp points})
has no multiple points of multiplicity greater than two.
(3) Suppose f satisfies (1) and (2). Then j
3
2
f is transverse to S
1
2
if and
only if f|
(S(f )−{cusp points})
is an immersion with normal crossings (see [7,
Proposition 5.6 (p. 158)]).
(4) Suppose f satisfies (1). Then j
3
2
f is transverse to S
2
2
and S
3
2
if and
only if for every cusp point p of f, we have f
1
(f(p)) S(f) = {p}.
Using the above facts, we obtain the following.
Lemma 3.6. Let f : M N be a proper smooth map of an n-dimensional
manifold M (n 2) into a surface N. For an open set U of N, f is excellent
on U if and only if f
1
(U) M = and the jets of f are transverse to
Σ
n1,0
, Σ
n,0
, Σ
n1,1,0
, Σ
n1,1,1,0
, S
1
2
, S
2
2
, S
3
2
and S
1
3
on the part corresponding
to f
1
(U).
Proof of Theorem 3.3. Set U = IntF . By Proposition 3.5 and Lemma 3.6,
we see that there exists a smooth map g : M N such that g V, g = f
on f
1
(U) = g
1
(U) and that g is excellent on N F . Since f and g are
continuous and N is Hausdorff, we see that g = f on the closure of f
1
(U).
However, we do not know if g is excellent on a neighborhood of N IntF.
This is be cause there is a possibility of a point in f
1
(N F ) being mapped
into F by g. In order to exclude this possibility, we modify the argument
as follows.
Since F is a closed submanifold of N , the set of maps of M into N
transverse to F forms an open set of C
(M, N) (see [7, Proposition 4.5 (p.
52)]). Thus we may assume that every map in the open set V is transverse
to F from the beginning. Furthermore, since the set of the proper maps
of M into N forms an open set (see [10, Theorem 1.5 (p. 38)]), we may
further assume that every element of V is a proper map. We may further
assume that each element of V maps M into F V by a similar reason,
where V is a closed neighborhood of F in N . Now suppose that g V.
Then, since g
1
(F ) is a closed regular submanifold of IntM, we see that
the closure of g
1
(U) is equal to g
1
(F ). Since f = g on the closure of
f
1
(U) and f
1
(U) = g
1
(U), we see that f = g on f
1
(F ) = g
1
(F ).
Combining the facts that g is excellent on N F and that f is excellent
on a neighborhood of F , we see that g is excellent on a neighborhood of
N IntF . This completes the proof of Theorem 3.3.
Remark 3.7. Results similar to Theorem 3.3 hold for some other dimension
pairs as well.
4. Graphic as the discriminant set.
Let Θ
A
, Θ
B
, Θ
X
, Θ
Y
, H, G be as is Section 2, where H, G may not be
generic. In this section, we first observe that we can obtain a smooth map
118 T. KOBAYASHI AND O. SAEKI
f : M
A
Θ
B
Θ
X
Θ
Y
) I × I from H and G, and we show, by
using Theorem 3.3, that f can be deformed to a map Φ which is excellent
in the exterior of a regular neighborhood of Θ
A
Θ
B
Θ
X
Θ
Y
by an
arbitrarily small deformation. Then we see that we can obtain sweep-outs
H
0
, G
0
associated to A
P
B, X
Q
Y respectively from Φ, which have
the feature “generic” in Sect. 2. Finally we observe that the corresponding
graphic is actually the discriminant set (for the definition, see Sect. 3) of Φ.
Let M be a closed 3-manifold. In this section we consider a smo oth map
f from M to the Euclidean space R
2
. Recall that S(f) denotes the set of
singular points (or singular set) of f. That is,
S(f) = {q M| rank (df
q
: T
q
M T
f(q)
R
2
) 1}.
Then as a special situation of Definition 3.2, we have:
Definition 4.1. Let f : M R
2
be a smooth map. For an open set U of
R
2
, we say that f is excellent on U if f
1
(U) M = , and the following
conditions are satisfied.
(1) For each point q S(f) there exist local coordinates (u, x, y) for q,
and (X, Y ) for f(q) such that:
(1-1) X f = u, Y f = x
2
+ y
2
, or
(1-2) X f = u, Y f = x
2
y
2
, or
(1-3) X f = u, Y f = y
2
+ ux x
3
.
(2) For each cusp q, f
1
(f(q)) S(f) = {q} (that is, the fiber which
contains q does not contain another singular point).
(3) f|
S(f )−{cusps}
is an immersion (possibly) with normal crossing (that
is, an immersion (possibly) with transverse self intersections).
We call a singular point of type (1-1) ((1-2) resp.) a definite fold (indefinite
fold resp.). Recall that a singular point of type (1-3) is c alled a cusp.
Figure 4.1.
RUBINSTEIN–SCHARLEMANN GRAPHIC 119
Figure 4.2.
Figure 4.3.
Figure 4.4.
Now we describe the relationship between graphic and excellent map. Let
M, A
P
B, X
Q
Y , Θ
A
, Θ
B
, Θ
X
, and Θ
Y
be as in Section 2. (Here we
suppose that Θ
A
Θ
B
and Θ
X
Θ
Y
are in general position.)
Let H, G be sweep-outs obtained from the Heegaard splittings A
P
B,
X
Q
Y respectively. We may suppose that H|
P ×(0,1)
: P × (0, 1) M
A
Θ
B
) and G|
Q×(0,1)
: Q × (0, 1) M
X
Θ
Y
) are smooth. Let
Φ : M
A
Θ
B
Θ
X
Θ
Y
) I × I be the map defined by:
() Φ(p) = (s, t) p P
s
Q
t
.
Since H|
P ×(0,1)
, G|
Q×(0,1)
are smooth maps, we see that Φ is also smooth.
Since {P
s
} and Θ
X
Θ
Y
, {Q
t
} and Θ
A
Θ
B
are generic, we see that:
120 T. KOBAYASHI AND O. SAEKI
(1) H (G resp.) is standard (see Sect. 2) in a small regular
neighborhood N
X
Θ
Y
) of Θ
X
Θ
Y
(N
A
Θ
B
) of Θ
A
Θ
B
resp.), and, hence, Φ is transverse to the frontier of a regular
neighborhood of (I × I) in I × I.
By Theorem 3.3, we see that we can deform Φ in M (N
A
Θ
B
)
N
X
Θ
Y
)) by an arbitrarily small deformation, to a map Φ
0
which is
excellent on the complement of the regular neighborhood of (I × I) in
I × I, and this together with (1) implies that Φ
0
is excellent on Int(I × I).
Since Φ
0
is obtained from Φ by a small deformation, we may suppose:
(2) pr
1
Φ
0
, pr
2
Φ
0
have no critical points, where pr
1
, pr
2
:
I × I I are the projections to the first, and second factors
respectively.
By condition (2), we see that, there exist sweep-outs H
0
, G
0
such that
H
0
(P × {s}) = Φ
01
({s} × I), and G
0
(Q × {t}) = Φ
01
(I × {t}). Note that
H
0
, G
0
are small deformations of H, G. By the definition of H
0
, G
0
, we
immediately have:
()
0
Φ
0
(p) = (s, t) p H
0
(P × {s}) G
0
(Q × {t}).
Then by Definition 4.1 (and the definition of the graphic in Sect. 2) we se e
that H
0
and G
0
are generic in the sense of Rubinstein and Scharlemann (see
Sect. 2) and, by comparing Definition 4.1 and the definition of the graphic
in Sect. 2, it is easy to see that the corresponding graphic is actually the
image of the singular set of Φ
0
on M
A
Θ
B
Θ
X
Θ
Y
), where the
image of a definite fold corresponds to a center tangency, the image of an
indefinite fold corresponds to a saddle tangency, and the image of a cusp
corresponds to a birth-death vertex.
Now we summarize the above results to give:
Theorem 4.2. Let H, G be as above. Then, by an arbitrarily small defor-
mation of H and G, we obtain sweep-outs H
0
and G
0
such that:
1. The above map Φ
0
(see ()
0
) is excellent on Int (I × I),
2. The maps H
0
and G
0
are generic. Hence we can obtain a graphic Γ
from H
0
and G
0
, and then Γ Int (I × I) is the discriminant set of the
excellent map Φ
0
|
M
A
Θ
B
Θ
X
Θ
Y
)
.
5. Isotopy, and stabilization of Heegaard surfaces.
Let f : M I × I( R
2
) be an excellent map obtained in the previous
section (which is denoted by Φ
0
there). Let α : I R
2
be an embedding of
the unit interval.
Definition 5.1. We say that α is transverse to f if α satisfies the following
two conditions:
RUBINSTEIN–SCHARLEMANN GRAPHIC 121
(1) α(I) (I × I), and α and (I × I) are transverse (i.e., a smooth
slight extension of α is transverse to smooth extensions of I × {0}, I × {1},
{0} × I, and {0} × I),
(2) for each pair (t, q) (0, 1) × M with α(t) = f(q), we have:
df
q
(T
q
M) +
t
(T
t
I) = T
f(q)
R
2
.
Then, by Definition 4.1, it is easy to see:
Lemma 5.2. Suppose that α satisfies Definition 5.1 (1). Then α is trans-
verse to f if and only if:
For q M with f(q) α(I), w e have either one of:
1. q is not a singular point of f,
2. q is a fold point which is not a normal crossing, and α is transverse
to the discriminant set at f(q),
3. q is a fold point which is mapped to a normal crossing, and α is trans-
verse to the two arcs (which are local images of the singular set), or
4. q is a cusp. In this case, there are two arcs in a neighborhood of
f(q) (which are local images of the singular set). Then there is a slight
smooth extensions of the arcs, say `
1
, `
2
such that α is transverse to
`
1
, and `
2
.
Figure 5.1.
Lemma 5.3. Suppose that α is transverse to f. Then f
1
(α(I)) is a 2-
dimensional submanifold of M.
Proof. By condition Definition 5.1 (2), we see that f
1
(α(0, 1)) is a 2-
dimensional proper s ubmanifold in M
A
Θ
B
Θ
X
Θ
Y
). Then by
condition Definition 5.1 (1), we see that f
1
(α[0, ε]) (f
1
(α[1 ε, 1]) resp.)
is a disk, which cap off f
1
(α(0, 1)) to make a closed surface f
1
(α(I)).
Proposition 5.4. Let α, β be arcs transverse to f. Suppose that α(I) is
deformed to β(I) through a sequence of moves of the following types.
(0) Ambient isotopy of I × I wh ich fixes (I × I) Γ setwise.
(1) Passing a crossing vertex as α
α
0
α
+
in Figure 5.2.
122 T. KOBAYASHI AND O. SAEKI
Figure 5.2.
(2) Passing a cusp as α
α
0
α
+
in Figure 5.3.
Figure 5.3.
(3) Passing a vertex in (I × I) as in Figure 5.4.
Figure 5.4.
(4) Passing a corner of (I × I) as in Figure 5.5.
Figure 5.5.
Then the surfaces f
1
(α(I)), and f
1
(β(I)) are isotopic in M.
Proof. First, we consider moves (0), (1), and (2). Suppose that α
0
(I) is
deformed to α
1
(I) through a sequence of moves (0), (1), and (2). By
Lemma 5.2, we see that there is a 1-parameter family of transverse arcs
RUBINSTEIN–SCHARLEMANN GRAPHIC 123
α
s
(0 s 1) from α
0
to α
1
. Then we obtain an isotopy of surfaces
f
1
(α
s
) in M.
Now we consider move (3). Note that the deformation (3) gives the isotopy
as in Figure 5.6.
Figure 5.6.
This shows that the deformation (3) gives mutually isotopic surfaces.
Finally we consider about move (4). Note that in a neighborhood of a
corner of I × I, P
s
and Q
t
are disjoint. Hence the deformation (4) obviously
gives equal surfaces.
Combining the above observations, we have the conclusion of the propo-
sition.
As a consequence of Propos ition 5.4. we have:
Corollary 5.5 (cf. Example 2.10). Suppose t hat the graphic obtained from
P and Q contains a region as in Figure 5.7. Then P and Q are isotopic in
M.
Figure 5.7.
Proof. Let α(t) = (t, ε), and β(t) = (ε, t) for a small ε > 0. It is easy to
see that α(I) is deformed to β(I) within the above region by applying the
deformations of Proposition 5.4.
For a stabilization of a Heegaard splitting, we have:
Proposition 5.6. Let α be an arc transverse to f such that f
1
(α(I)) is a
Heegaard surface. Suppose that a transverse arc α
0
is obtained by changing
124 T. KOBAYASHI AND O. SAEKI
α locally as in Figure 5.8. Then f
1
(α
0
(I)) is a Heegaard surface which is
a stabilization of Φ
1
(α(I)).
Figure 5.8.
Proof. We use the following picture (Figure 5.9) for the proof. The picture
corresponds to the point of the intersection of α(I) and the image of an
indefinite fold (here P
s
’s are represented by horizontal planes).
Figure 5.9.
Let P
i
be the subsurface of Φ
1
(α
0
(I)) corresponding to Φ
1
(α
i
), where
α
i
is as in Figure 5.8. It is directly observed from Figure 5.9 that each P
i
looks as in Figure 5.10.
Figure 5.10.
By summing up P
i
’s, we see that Φ
1
(α
0
(I)) is a stabilization of
Φ
1
(α(I)).
Corollary 5.7. Let P , and Q be the Heegaard surfaces as in Example 2.11.
Then P and Q become isotopic by applying one stabilization.
RUBINSTEIN–SCHARLEMANN GRAPHIC 125
Proof. We first take transverse arcs α, and β as in Figure 5.11. Then, by
Figure 5.11, we see that α(I) can be deformed to the arc in Figure 5.12,
by one application of the deformation of Proposition 5.6, and deformations
in Proposition 5.4. By reflecting the pictures in Figure 5.11 in the line
connecting right-bottom corner to left-top corner, we see that β(I) is also
deformed to the arc in Figure 5.12, and this gives the conclusion.
Figure 5.11.
Figure 5.12.
6. Orbifold version of Rubinstein-Scharlemann graphic.
In this section, we formulate an orbifold version of the Rubinstein-Scharle-
mann setting, and show that the local lab e lling scheme described in [ 18]
holds in this setting.
Let M be a compact 3-manifold, γ a union of mutually disjoint arcs or
simple close d curves properly embedded in M, F a surface properly em-
bedded in M, which is in general position with respect to γ, and `( F ) a
simple closed curve with ` γ = .
Definition 6.1. A surface D is a γ-disk, if D is a disk intersecting γ in at
most one transverse point.
Definition 6.2. We say that ` is γ-inessential if ` bounds a γ-disk in F .
We say that ` is γ-essential if it is not γ-inessential.
126 T. KOBAYASHI AND O. SAEKI
Let `
1
, `
2
( F) be simple closed curves with `
i
γ = (i = 1, 2).
Definition 6.3. We say that `
1
and `
2
are γ-parallel if `
1
`
2
bounds an
annulus A in F such that A γ = .
Definition 6.4. We say that D is a γ-compressing disk for F if; D is a
γ-disk; and D F = D, and D is γ-essential in F . The surface F is
γ-compressible if it admits a γ-compressing disk, and it is γ-incompressible
if it is not γ-compressible.
Let a be an arc properly embedded in F with a γ = .
Definition 6.5. We say that a is γ-inessential if there is a subarc b of F
such that b = a, and a b bounds a disk D in F such that D γ = .
We say that a is weakly γ-inessential if there is a subarc b of F such that
b = a, and a b bounds a γ-disk D in F .
Definition 6.6. Let F
1
, F
2
be surfaces embedded in M such that F
1
=
F
2
. We say that F
1
and F
2
are γ-parallel, if there is a submanifold N in M
such that (N, F
1
F
2
, N γ) is homeomorphic to (F
1
×I, F
1
×{1/2}, P ×I)
as a triple, where P is a union of points in Int(F
1
), and F
1
(F
2
resp.)
corresponds to the closure of the component of (F
1
× I) F
1
× {1/2}
containing F
1
× {0} (F
1
× {1} resp.).
The submanifold N is called a γ-parallelism between F
1
and F
2
.
We say that F is γ-boundary parallel if there is a subsurface F
0
in M
such that F and F
0
are γ-parallel.
Definition 6.7. Let F
1
, F
2
be mutually disjoint surfaces in M which are
in general position with respect to γ. We say that F
1
and F
2
are γ-isotopic
if there is an ambient isotopy φ
t
(0 t 1) of M such that; φ
0
= id
M
;
φ
1
(F
1
) = F
2
, and; φ
t
(γ) = γ for each t.
Genus g n-bridge position.
Let Γ = {γ
1
, . . . , γ
n
} be a system of mutually disjoint arcs properly em-
bedded in M.
Definition 6.8. We say that Γ is trivial if there exists a system of mutually
disjoint disks {D
1
, . . . , D
n
} in M such that (1) D
i
Γ = D
i
γ
i
= γ
i
, and
(2) D
i
M is an arc, say α
i
, such that α
i
= c`(D
i
γ
i
).
Example 6.9. Let β be a system of trivial two arcs in a 3-ball B. The pair
(B, β) is often refered as 2-string trivial tangle, or a rational tangle.
Let K be a link in a closed 3-manifold M. Let M = A
P
B be a genus
g Heegaard splitting. Then the next definition is borrowed from [6].
Definition 6.10. We say that K is in a (genus g) n-bridge position (with
respect to the Heegaard splitting A
P
B) if K A (K B resp.) is a system
of trivial n arcs in A (B resp.).
RUBINSTEIN–SCHARLEMANN GRAPHIC 127
In this paper, we abbreviate genus 0 n-bridge position to n-bridge posi-
tion.
Unknotting tunnel.
Let K be a knot in a closed 3-manifold M . A tunnel for K is an embedded
arc σ in S
3
such that σK = σ. We say that a tunnel σ for K is unknotting
if S
3
Int N (K σ, S
3
) is a genus two handlebody.
Orbifold setting.
Let K be a link in a closed 3-manifold M. We regard K as γ above. Let
L
1
, L
2
be a pair of mutually disjoint 1-complexes in M such that:
1. Each vertex of L
i
has valency zero, one or three,
2. (L
1
L
2
) K consists of the union of a (possibly, empty) sublink of
K, and a subset of the vertices of L
1
L
2
with valency one,
3. Let N be a regular neighborhood of L
1
L
2
, and E = c`(M N).
Then (E, E K) is homeomorphic to (P × (0, 1), P × (0, 1)), where P
is a closed surface, and P is a finite set of (possibly empty) points in
P .
Let A, B be the closures of the components of M (P × {1/2}), where
L
1
A, L
2
B. We say that A B is an orbifold Heegaard splitting
of (M, K). Then as in Sect. 2, we obtain a sweep-out H : P × I M.
Let R
1
, R
2
be another pair of 1-complexes satisfying the above conditions
(1), (2), and (3), and G : Q × I M the corresponding sweep-out. Then,
as in Theorem 4.2, we may suppose that we can obtain an excellent map
f : M (L
1
L
2
R
1
R
2
) R
2
from H and G such that the graphic
obtained from H and G is the discriminant set f(S(f)) I × I. Here we
note that we have to slightly generalize the definition of standard position
for a neighborhood of a valency one vertex (e.g. Figure 6.1), and it is easy
to see the procedures in Sect. 4 work under this situation.
Figure 6.1.
Example 6.11. Suppose that K is in a genus g n-bridge position with
respect to a Heegaard splitting A
P
B. Then, by adding n edges to each
of the appropriate spines of A and B, we can obtain 1-complexes satisfying
the above conditions.
128 T. KOBAYASHI AND O. SAEKI
Figure 6.2.
Example 6.12. Let K be a tunnel number one knot with unknotting tunnel
τ. Let L
1
= K τ. Let L
2
be a spine of the genus two handlebody c`(M
N(K τ)) with each vertex having valency three. Then L
1
, L
2
satisfies the
above conditions.
Definition 6.13. Let H, G be sweep-outs as above. We say that H and G
are K-comparable if f|
K(L
1
L
2
R
1
R
2
)
: K(L
1
L
2
R
1
R
2
) Int(I×I)
is an immersion (possibly) with normal crossing, and f(K (L
1
L
2
R
1
R
2
)) and f(S(f)) are in general position in Int(I × I).
Proposition 6.14. By an arbitrarily small deformation on K rel (L
1
L
2
R
1
R
2
) with respect to Whitney topology, we can arrange H and G to be
K-comparable.
Proof. Note that f
1
(f(S(f ))) has the structure of a simplicial complex with
dimension at most 2 (see Figure 6.3), and, hence, by an arbitrarily small
deformation of K with respect to Whitney topology we can arrange so that
K and f
1
(f(S(f ))) are in general position, that is, K and the 1-skeleton are
disjoint, and K and f
1
(f(S(f ))) intersect transversely in a finite number
of points. This shows that f (K) and f(S(f)) intersect transversely in a
finite number of points. Possibly f(K) may contain a crossing vertex of the
graphic f(S(f)). Then we further apply a small deformation to make f(K)
avoid crossing vertices and to make f|
K(L
1
L
2
R
1
R
2
)
an immersion with
normal crossing, and this gives the conclusion.
Figure 6.3.
RUBINSTEIN–SCHARLEMANN GRAPHIC 129
For a K-comparable pair H and G we can obtain a graphic Γ as in the
following manner.
Regions: A region is a component of the subset of Int(I × I) consisting
of values (s, t) such that P
s
and Q
t
intersect transversely, and K
(P
s
Q
t
) = .
Edges: An edge is a component of the subset consisting of values (s, t)
such that either:
(1) P
s
and Q
t
intersect transversely except for one non-degenerate
tangent point and K (P
s
Q
t
) = , or
(2) P
s
and Q
t
intersect transversely and K (P
s
Q
t
) consists of
one point.
By Definition 6.13, we see that edge is a 1-dimensional subset of
Int(I × I) which is monotonously increasing or decreasing.
Crossing vertices: A crossing vertex is a com ponent of the subset con-
sisting of points (s, t) such that either:
(1) P
s
and Q
t
intersect transversely except for two non-degenerate
points of tangency and K (P
s
Q
t
) = , or
(2) P
s
and Q
t
intersect transversely except for one non-degenerate
tangent point and K (P
s
Q
t
) consists of one point, or
(3) P
s
and Q
t
intersect transversely and K (P
s
Q
t
) consists of
two points.
Note that in this setting we may (as in Sect. 2) also regard a cross-
ing vertex to be a crossing point of two edges. This follows from
the same reason as in Section 2 (Case (1)), or from the condition
K and f
1
(f(S(f ))) are generic” (Case (2)), or from the condition
f|
K(L
1
L
2
R
1
R
2
)
is an immersion with normal crossings” (Case (3)).
Birth-death vertices: A birth-death vertex is a c omponent of the sub-
set consisting of points (s, t) such that P
s
and Q
t
intersect transversely
except for a single degenerate tangent point and K (P
s
Q
t
) = .
Labelling regions of the graphic.
Consider a region of the graphic I × I Γ. Then the K-isotopy class of
P
s
Q
t
in P
s
or Q
t
is independent of the choice of (s, t) in each region, and,
hence, we often abbreviate P
s
by P , and Q
t
by Q.
The purp ose of the rest of this section is to claim that the nature of the
(local) labelling schemes discussed in [18] holds also in our setting. We
assume that the reader is familiar with Sect. 4, 5 of [18].
Definition 6.15. We say that an orbifold Heegaard splitting A
P
B is
weakly K-reducible if there exist K-compressing disks D
A
, D
B
for P in A,
B respectively such that D
A
D
B
= . The orbifold Heegaard splitting
A
P
B is strongly K-irreducible if it is not weakly K-reducible.
130 T. KOBAYASHI AND O. SAEKI
Definition 6.16. Let (s, t) be a point in a region of I ×I Γ. (Hence, P Q
consists of a system of s imple closed curves in M disjoint from K.) Let C
K
P
(C
K
Q
resp.) be the subset of the simple closed curves which are K-essential
in P (Q res p.). Then the subset C
K
A
of C
K
P
is defined by:
C
K
A
= {c|c bounds a K-disk D in Q C
K
P
such that N(D, D) A},
where N(D, D) is a regular neighborhood of D in D.
Analogously C
K
B
( C
K
P
), and C
K
X
, C
K
Y
( C
K
Q
) are defined.
Lemma 6.17 (Lemma 4.3 of [18]). If c C
K
A
, then c bounds a K-disk in
A.
Proof. The proof is basically the same as Rubinstein-Scharlemann’s except
for the consideration on K. That is:
Let D be the K-disk which c bounds in Q, such that N (D, D) A. If
IntD P = , then D gives a desired K-disk. Suppose that IntD P 6= .
Let ∆( D) be an innermost disk. Since IntD C
K
P
= , we see that
bounds a K-disk
0
in P . For a proof of the next claim, see Appendix A-3.
Claim.
0
K = if and only if K = . Furthermore, if
0
K 6= ,
then and
0
are K-parallel, i.e.,
0
bounds a 3-ball D
3
such that
D
3
K is an unknotted arc.
By the claim, we see that we can apply cut and paste on D using and
0
to get a new disk D
0
with fewer intersections. By applying the argument
finitely many times, we obtain the desired disk.
As an immediate consequence of Lemma 6.17, we have:
Corollary 6.18 (Corollary 4.4 of [18]). If there exists a region such that
both C
K
A
and C
K
B
are non-empty, then A
P
B is weakly K-reducible.
In the rest of this section, we suppose:
M admits a 2-fold branched covering space p :
˜
M M along
K.
Lemma 6.19 (Lemma 4.5 of [18]). Suppose that C
K
P
= , C
K
Q
= , and
there exists a -reducing K-disk in A which intersects Q only in K-inessen-
tial simple closed curves. Suppose, moreover, that A contains a K-essential
curve of Q. Then either A
P
B is weakly K-reducible, or M is the 3-sphere
S
3
and K is a trivial knot.
Proof. By Appendix A-3, we may suppose, by K-isotopy, that P and Q are
disjoint, and that the -reducing K-disk D and Q are disjoint. Without loss
of generality, we may suppose that Y is contained in A. Now consider the
2-fold branched covers (along K)
˜
A,
˜
B,
˜
P ,
˜
X,
˜
Y ,
˜
D of A, B, P , X, Y , D
respectively. Note that, by the definition of an orbifold Heegaard splitting,
˜
A,
˜
B are handlebodies.
RUBINSTEIN–SCHARLEMANN GRAPHIC 131
Take a maximal compression body
˜
C of
˜
A Int
˜
Y for
˜
A. By the unique-
ness of maximal compres sion body, we may suppose, by applying Z
2
-equi-
variant cut and paste arguments as in the Proof of 10.3 of [9] or [12], that
˜
C is invariant under the cove ring translation τ . Let P
0
be a component of
the inner boundary of
˜
C.
If P
0
is a sphere, then M is S
3
(see Proof of [18, Lemma 4.5]), and, by
Z
2
-Smith Conjecture ([21], or [16]), K is a trivial knot in S
3
.
Suppose that P
0
is not a sphere. Note that P
0
is compressible in
˜
B
˜
C
since P
0
is contained in a handlebody
˜
X (see the proof of Lemma 4.5 of
[18]). Then note that
˜
B
˜
P
˜
C is a Heegaard splitting in the sense of Casson-
Gordon [4]. Hence, by [4], there exists a compressing disk D
0
(
˜
B
˜
C) for
P
0
such that D
0
˜
P consis ts of a circle (hence, D
0
˜
C is an annulus).
Now we show that we can have such D
0
which moreover is equivariant
with respect to τ. We may suppose, by general position argument, that D
0
and τ(D
0
) intersect transversely. Then, by isotopy, we may suppose that
each component of
˜
C (D
0
τ(D
0
)) is an essential arc in the annulus D
0
˜
C,
(and τ(D
0
)
˜
C). Then, by applying equivariant cut and paste arguments as
in the proof of 10.3 of [9] or [12] on D
0
, and τ(D
0
), we obtain an equivariant
compressing disk(s) D
B
. Since each component of
˜
C (D
0
τ (D
0
)) is an
essential arc of D
0
˜
C, we see that each component of D
B
intersects
˜
C in
an annulus, (hence, intersects
˜
P in a circle).
Then apply cut and paste arguments on D
B
˜
C and
˜
D to obtain a com-
pressing disk
˜
D
0
(
˜
C) for
˜
P such that
˜
D
0
(D
B
˜
C) = . Then, by
applying equivariant cut and paste arguments on
˜
D
0
, and τ(
˜
D
0
), we obtain
equivariant disk(s) D
A
(
˜
C) for
˜
P such that D
A
D
B
= . Then p(D
A
),
and p(D
B
˜
B) give weak K-reducibility of K.
Labelling scheme.
Now we mimic the procedures in [18, Section 5]. If C
K
A
(C
K
B
, C
K
X
, C
K
Y
resp.) is non-empty, then we label the region A (B, X, Y resp.). If C
K
P
and
C
K
Q
are both empty and A (B resp.) contains an K-essential curve of Q,
then we lab el the region b (a resp.), and if X (Y resp.) contains an essential
curve of P , then we label the region y (x resp.). By Corollary 6.18 we have:
Rule 1. If there exists a region with both labels A and B assigned, then
A
P
B is weakly K-reducible.
We obviously have:
Rule 2. No region can have both an upper case label and lower case label.
Next, we consider how labels change as one cross an edge of Γ.
Note that we have the following three pos sibilities.
1) The edge comes from center tangency.
132 T. KOBAYASHI AND O. SAEKI
In this case, the regions have exactly the same label.
2) The edge comes from saddle tangency.
In this case, the effect is banding two components of P Q, say c
0
and
c
1
, to make a simple closed curve, say c, or vice versa.
3) The edge comes from P Q K.
In this case the effect is that a component of P Q passes a puncture by
K on P, (and Q).
Note that situation 3) did not appear in Rubinstein-Scharlemann setting.
With this fact in mind, it is easy, by tracing the proof of [18, Corollary 5.1],
to see:
Rule 3 ([18, Corollary 5.1]). If both labels A and B appear in two adjacent
regions, then A
P
B is weakly K-reducible.
Then we have:
Rule 4 ([18, Corollary 5.2]). In adjacent regions of I × I Γ, labels a and
b (x and y resp.) cannot app ear.
Proof. Suppose that a and b occur opposite sides of an edge. Then argu-
ments in the proof of [18, Corollary 5.2] show that edge does not come from
saddle tangency. Then it is easy to see that this phenomena can occur only
in case when Q is a 2-sphere and QK consists of three points a
1
, a
2
, a
3
and
a component of the intersection P Q is changed from a circle separating a
1
and a
2
a
3
to a circle separating a
1
a
2
and a
3
. However this is impossible,
since Q K must consists of even number of points.
With tracing the proof of [18, Lemma 5.3] with consideration on K we
easily have:
Lemma 6.20 ([18, Le mma 5.3]). Suppose, in I × I Γ, a region labelled
A is adjacent to a region labelled with a lower case letter. Then the edge
represents either (1) a saddle tangency in which a band which is K-essential
in P and weakly K-inessential in Q is attached to an intersection curve
which is K-inessential in both P and Q, or (2) a passing of K which changes
an element of C
K
A
bounding a disk (in P) with two punctures by K into a
disk with one puncture by K.
Then we have:
Rule 5 ([18, Corollary 5.4]). Suppose, in I × I Γ, a region labelled A is
adjacent to a region labelled b. Then either A
P
B is weakly K-reducible,
or M
=
S
3
and K is a trivial knot.
Proof. We see, by Lemma 6.20, that A
P
B satisfies the assumption of
Lemma 6.19, and this gives the conclusion.
RUBINSTEIN–SCHARLEMANN GRAPHIC 133
In the following, the notation a stands for, as in [18], either a or A, and
similar for b, x, and y. With the above rules, we see that the arguments in
the proof of [18, Lemma 5.7] (it is easy to check that [18, Lemma 5.6] holds
in our setting since the new phenomenon is the situation 3) in the preceding
Rule 3) completely works in our setting to give:
Rule 6 ([18, Lemm a 5.7]). If all letters a, b, x, and y appear in quadrants
of a crossing vertex of Γ, then either two opposite quadrants are unlabelled,
or one of A
P
B, X
Q
Y is weakly K-reducible, or M
=
S
3
and K is a
trivial knot.
By using these rules, the arguments in the proof of [18, Proposition 5.9]
show (the difference here is the consideration on K, a possibility that three
edges may be joined to a vertex in (I × I) (see Figure 6.1)):
Proposition 6.21. Let A
P
B, X
Q
Y be orbifold Heegaard splittings for
(M, K) obtained from two bridge positions as in Example 6.11. Suppose that
A
P
B, X
Q
Y are strongly K-irreducible, and K is not a trivial knot in
S
3
. Then there is an unlabelled region in I × I Γ.
And, this together with Appendix A-3, and the arguments in the proof
of [18, Corollary 6.2] shows:
Corollary 6.22. Let A
P
B, X
Q
Y be as in Proposition 6.21. Then,
by applying K-isotopy, we may suppose that P and Q intersect non-empty
collection of simple closed curves which are K-essential in both P and Q.
7. 2-bridge position of a 2-bridge knot.
Let K be a non-trivial 2-bridge knot (that is, K is a non-trivial knot which
admits a genus 0 2-bridge position). In this section, we show that the
2-bridge positions of K are unique up to K-isotopy, which was originally
proved by Schubert [20].
Theorem 7.1. Let K be a non-trivial 2-bridge knot, and P , Q are 2-spheres
in S
3
which give 2-bridge positions of K. Then P is K-isotopic to Q, i.e.,
there is an ambient isotopy ϕ
t
(0 t 1) of S
3
such that (1) ϕ
t
(K) = K
(0 t 1), (2) ϕ
0
= id
S
3
, and (3) ϕ
1
(P ) = Q.
For the proof of Theorem 7.1, we prepare some lemmas, proofs of which
are given in Appendix B. (For the defiition of β-essential surface, see Defi-
nition 6.2.)
Lemma 7.2 (Appendix B-1). Let (B, β) be a 2-string trivial tangle. Let F
be a surface properly embedded in B. Suppose that F is β-essential. Then
F is a disk which is disjoint from β, and F separates the components of β.
134 T. KOBAYASHI AND O. SAEKI
Figure 7.1.
Recall that it is often said that (C, γ) is a rational tangle if (C, γ) is
homeomorphic to the 2-string trivial tangle (B, β) as a pair.
Lemma 7.3 (Appendix B-2). Let (B, β) be a 2-string trivial tangle, and F
a β-incompressible surface in B.
Then either (0) F is β-essential (see Lemma 7.2), (1) F is a β-boundary
parallel disk intersecting β in at most one point, (2) F is a β-boundary
parallel disk intersecting β in two points (and, hence, F separates (B, β)
into the parallelism and a rational tangle), or (3) F is β boundary parallel
annulus such that F β = .
Figure 7.2.
Lemma 7.4 (Appendix B-3). Let D be a β-compressible disk in B such
that D is β-essential in B, and D β consists of two points. Then D
separates (B, β) into two tangles (B
1
, β
1
), and (B
2
, β
2
), where (B
1
, β
1
) is
a rational tangle such that there is a β-essential disk D
0
in (B
1
, β
1
) with
D D
0
= . Moreover if (B
2
, β
2
) happens to be a rational tangle, then
(B
2
, β
2
) is a β-boundary parallelism for D.
Figure 7.3.
RUBINSTEIN–SCHARLEMANN GRAPHIC 135
Let A, B (X, Y resp.) be the closures of the components of S
3
P (S
3
Q
resp.).
Proposition 7.5. Every genus 0 Heegaard splitting of S
3
which gives a 2-
bridge position of K is strongly K-irreducible.
Proof. We give the proof for A
P
B. Assume that A
P
B is weakly K-
reducible, and let D
A
, D
B
be a pair of K-esssential disks in A, B respectively
such that D
A
D
B
= . Since P K consists of four points, we see that
D
A
and D
B
are parallel in P K, and this together with Lemma 7.2
implies that K is a 2-component trivial link, a contradiction.
Proof of Theorem 7.1. Note that, by Proposition 7.5, A
P
B, X
Q
Y are
strongly K-irreducible. Hence , by Corollary 6.22, we may suppose that
P and Q intersect non-empty collection of simple closed curve s which are
K-essential in both P and Q.
The proof is carried out by the induction on the number of the components
of P Q. The following Claims 1 and 2 give the first step of the induction.
Claim 1. If P Q consists of one component, then P and Q are K-isotopic.
Proof. Let D
A
= QA, D
B
= QB, D
X
= P X, and D
Y
= P Y . Then
D
A
, D
B
, D
X
, D
Y
are disks each of which intersects K in two points. Then
we have the following case s.
Case 1. The disks D
A
, D
B
, D
X
, D
Y
are K-incompressible in A, B, X, Y
respectively.
In this case, by Lemma 7.3 (2), we have:
(1) D
A
and D
X
are K-parallel in A or D
A
and D
Y
are K-parallel in A”,
(2) D
B
and D
X
are K-parallel in B or D
B
and D
Y
are K-parallel in B”,
(3) D
X
and D
A
are K-parallel in X or D
X
and D
B
are K-parallel in
X”, and
(4) D
Y
and D
A
are K-parallel in Y or D
Y
and D
B
are K-parallel in Y ”.
It is easy to see that the above 4 conditions imply either one of:
(1) D
A
and D
X
are K-parallel in A (and, X), and D
B
and D
Y
are K-
parallel in B (and, Y )”, or
(2) D
A
and D
Y
are K-parallel in A (and, Y ), and D
B
and D
X
are K-
parallel in B (and, X)”.
Since the argument is symmetric, we may suppose that (1) holds. Then,
by using the parallelisms, we can move D
A
to D
X
, and D
B
to D
Y
to give a
desired K-isotopy.
Case 2. One of the disks D
A
, D
B
, D
X
, or D
Y
is K-compressible in A, B,
X, or Y .
136 T. KOBAYASHI AND O. SAEKI
Without loss of generality, we may suppose that D
Y
is K-compressible in
Y , and the K-compressing disk is contained in B. Note that this implies that
D
B
is K-compressible in B. Note moreover that D
X
is K-incompressible
in X, since D
X
separates the boundary components of each component of
KX in X, and similarly D
A
is K-incompressible in A. Then, by Lemma 7.3
(2) and the last half of Lemma 7.4, we see that D
B
and D
X
are K-parallel
in B (and, X). Similarly we can show that D
A
and D
Y
are K-parallel in A
(and, Y ). Hence we can obtain a desired K-isotopy as in Case 1.
This completes the proof of Claim 1.
Claim 2. If P Q consists of two components, then P and Q are K-isotopic.
Proof. Let D
1
, A
0
, D
2
be the closures of the components of P (P Q) such
that D
1
and D
2
are disks, and A
0
is an annulus. Without loss of generality,
we may suppose that D
1
D
2
is contained in X, and A
0
is contained in Y .
Subclaim. Either D
1
or D
2
is K-boundary parallel in X.
Proof. If D
1
or D
2
is K-incompressible, then this immediately follows from
Lemma 7.3 (2). Suppose that D
1
and D
2
are K-compressible. Let B
3
1
be the
closure of the component of X D
1
which corresponds to B
2
in Lemma 7.4.
By exchanging suffix, if necessary, we may suppose that Int(B
3
1
) P =
. Without loss of generality, we may suppose that B
3
1
is contained in A.
Let D
1
= B
3
1
Q. Since D
1
separates the boundary components of each
component of K A in A, we see that D
1
is K-essential in A. Hence,
by Lemma 7.3 (2) and the last half of Lemma 7.4, we see that B
3
1
is a
K-parallelism.
Let B
3
1
be the parallelism between D
1
and X obtained in Subclaim.
Then we can push D
1
out of X along the parallelism, and we have the
conclusion by Claim 1.
This completes the proof of Claim 2.
Completion of Proof. Suppose that ](P Q) 3. Note that the components
of P Q are mutually K-parallel in P . Let D
1
, A
1
, . . . , A
m
, D
2
be the
closures of the components of P (P Q) such that D
1
, D
2
are disks and
A
1
, . . . , A
m
are annuli that are located on P in this order. Suppose that
D
1
or D
2
is K-boundary parallel in X or Y . Then, by using the arguments
in the proof of Claim 2, we can reduce ](P Q), to give the conclusion.
Suppose that D
1
and D
2
are not K-boundary parallel in X and Y . By
Lemma 7.3 (2), this implies that D
1
and D
2
are K-compressible.
Claim 3. Both D
1
and D
2
are contained in the closure of a component of
S
3
Q, say X. And each component of P Y is a K-incompressible annulus.
RUBINSTEIN–SCHARLEMANN GRAPHIC 137
Proof. Ass ume that D
1
is contained in X, and D
2
is contained in Y . By
applying K-compression on D
1
(D
2
resp.) we obtain a K-essential disk E
1
(E
2
resp.) in X (Y resp.) such that E
i
= D
i
. Note that E
1
(E
2
resp.)
separates the components of K X (K Y resp.). This implies that K
is a 2-component trivial link, a contradiction. Hence we may suppose that
D
1
and D
2
are contained in X. Let A
i
be a component of P Y . Assume
that A
i
is K-compressible in Y . Then, by applying K-compression on A
i
,
we obtain two K-essential disks in Y . Then, by the above argument, we see
that K is a 2-c omponent trivial link, a contradiction.
Claim 3 together w ith Lemma 7.3 (3) implies that each component of
P Y is a K-boundary parallel annulus in Y . Take an outermost one of
P Y , say A
j
, and push A
j
out of Y along the parallelism. This reduces
](P Q) by two, and we have the conclusion by the assumption of the
induction.
This completes the proof of Theorem 7.1.
8. Genus one 1-bridge position of a 2-bridge knot.
For a 2-bridge knot K we can obtain four genus one 1-bridge positions of K
as follows.
Let A
P
B be the Heegaard splitting which gives the 2-bridge position.
Then let a
1
, a
2
, b
1
, b
2
be the closures of the components of K P , where
a
1
a
2
(b
1
b
2
resp.) is contained in A (B resp.). Let T
1
= A N (b
1
, B),
α
1
= a
1
b
1
a
2
, T
2
= c`(B N(b
1
, B)), and α
2
= b
2
. Then each T
i
is a
solid torus and α
i
is a trivial arc in T
i
, and, hence, T
1
T
2
gives genus one
1-bridge position of K. Moreover, by using a
1
, a
2
, b
2
for b
1
, we can obtain
other three genus one 1-bridge positions of K.
Figure 8.1.
Remark 8.1. In [17], Morimoto-Sakuma study the isotopy classes of the
1-bridge positions ab ove.
138 T. KOBAYASHI AND O. SAEKI
We say that these genus one 1-bridge positions are standard.
The purpose of this section is to prove:
Theorem 8.2. Every genus one 1-bridge positions of a non-trivial 2-bridge
knot is standard.
First, we prepare some lemmas for the proof of Theorem 8.2, proofs of
which are given in Appendix C. Let α be a trivial arc in a solid torus T .
For (T, α), we have:
Lemma 8.3 (Appendix C-1). Let D be an α-compressing disk for T .
Then D is either:
(1) a meridian disk of T with Dα = . In this case, we obtain, by cutting
(T, α) along D, a 1-string trivial tangle,
(2) a meridian disk of T with D α consists of one point, and we obtain,
by cutting (T, α) along D, a 2-string trivial tangle, o r
(3) -parallel disk in T with Dα = . In this case, D cobounds a 1-string
trivial tangle in (T, α).
Figure 8.2.
Lemma 8.4 (Appendix C-2). Let D be an α-essential disk in T such that
D α consists of two points. Then there exists an α-compressing disk D
0
disjoint from D such that D
0
α consists of one point. Moreover, by cutting
(T, α) along D
0
, we obtain 2-string trivial tangle (B, β) such that D is a
β-incompressible disk in (B, β) (hence D is β-boundary parallel).
We note that the disk D in Lemma 8.4 is either separating or non-
separating in T .
Figure 8.3.
Lemma 8.5 (Appendix C-3). Let D
1
, D
2
be mutually disjoint non α-paral-
lel, α-essential disks such that D
i
α (i = 1, 2) consists of two points. Then
there exists an α-compressing disk D
0
for T disjoint from D
1
D
2
such
RUBINSTEIN–SCHARLEMANN GRAPHIC 139
that D
0
α consists of one point. Moreover each D
i
is non-separating in T ,
and, by cutting (T, α) along D
0
, we obtain 2-string trivial tangle (B, β), and
D
1
, D
2
are mutually non β-parallel, β-boundary parallel, β-incompressible
disks in (B, β).
Figure 8.4.
Lemma 8.6 (Appendix C-4). Let D be an α-compressible disk such that
D is α-essential in T , and D α consists of two points. Then there is a
disk in T such that D = D = γ an arc, and α = c`( γ).
Particularly, if D is separating in T , then D separates (T, α) into (T
0
, α
0
),
and (B
0
, α
00
) such that α
0
is a trivial arc in a solid torus T
0
. In this case,
if (B
0
, α
00
) happens to be a rational tangle, then (B
0
, α
00
) is an α-boundary
parallelism.
Figure 8.5.
In the rest of this section, we give a proof of Theorem 8.2. Let K be a
non-trivial 2-bridge knot. Let A
P
B be a genus 0 Heegaard splitting of
S
3
which gives a 2-bridge position of K, and X
Q
Y a genus one He egaard
splitting which gives a genus one 1-bridge position of K. Note that A B,
X Y give orbifold Heegaard splittings of (S
3
, K) (see Example 6.11).
Proposition 8.7. Exactly one of the following (1) or (2) holds.
1. X
Q
Y gives a standard genus one 1-bridge position.
2. X
Q
Y is strongly K-irreducible.
Proof. Suppose that X
Q
Y is weakly K-reducible. Let D
X
, D
Y
be a pair
of K-esssential disks in X, Y respectively such that D
X
D
Y
= . Since
H
1
(S
3
) = {0}, we see that either D
X
is separating in X or D
Y
is separating
in Y . Without loss of generality, we may suppose that D
X
is separating in
X.
140 T. KOBAYASHI AND O. SAEKI
Claim 1. The disk D
Y
is non-separating in Y .
Proof. Ass ume that both D
X
and D
Y
are separating in X and Y respec-
tively. By Lemma 8.3 (3), we see that the closure of a component of X D
X
(Y D
Y
resp.), say B
3
X
(B
3
Y
resp.), is a 3-ball such that K B
3
X
(K B
3
Y
resp.) is a trivial arc . Since D
X
D
Y
= , we see that D
X
and D
Y
are K-parallel in Q. This implies that B
3
X
B
3
Y
is a 3-ball, and B
3
X
B
3
Y
is
a disk intersecting K in two p oints. This shows that K is a connected sum
of trivial knots, and, hence, K is a trivial knot, a contradiction.
Claim 2. The disk D
Y
intersects K in one point.
Proof. Ass ume this does not hold, i.e., D
Y
K = . Let N(D
Y
) be a regular
neighborhood of D
Y
in Y . Let X
0
= X N(D
Y
), and Y
0
= c`(Y N (D
Y
)).
Then, by Lemma 8.3 (3), we see that X
0
is a 3-ball such that K X
0
is
a trivial arc. Moreover, by Lemma 8.3 (1), we see that Y
0
is a 3-ball such
that K Y
0
is a trivial arc. Hence we see that K is a trivial knot, a
contradiction.
Let N
0
(D
Y
) be a regular neighborhood of D
Y
in Y . Let X
00
= XN
0
(D
Y
),
and Y
00
= c`(Y N
0
(D
Y
)). Then, by Lemma 8.3 (3), we see that X
00
is a
3-ball such that K X
00
is a system of 2-string trivial arcs. Moreover, by
Lemma 8.3 (2), we see that Y
00
is a 3-ball such that K Y
00
is a system of
2-string trivial arcs. Hence X
00
Y
00
gives a 2-bridge position of K, and the
genus one 1-bridge position X Y is obtained from X
00
Y
00
in a standard
manner.
Conversely, suppose that X
Q
Y gives a standard genus one 1- bridge
position. Then it is clear there exist disks corresponding to D
X
, D
Y
above
in X, Y , and these disks give a weak K-irreducibility of X
Q
Y . This
together with the above shows that X
Q
Y gives a standard genus one
1-bridge position if and only if X
Q
Y is weakly K-irreducible.
This completes the proof of Proposition 8.7.
Then we prove:
Proposition 8.8. Suppose tha t P Q consists of non-empty collection of
simple closed curves which are K-essential in both P and Q. Then the genus
one 1-bridge position X
Q
Y of K is obtained from A
P
B in a standard
manner.
For the proof of Prop os ition 8.8, we prepare the following lemma, the
proof of which is left to the reader.
Lemma 8.9. Let T be a solid torus, and A an annulus properly embedded
in T such that each component of A is a longitude of T . Then there is a
homeomorphism h : (annulus) × I T such that h((annulus) × {1/2}) = A.
RUBINSTEIN–SCHARLEMANN GRAPHIC 141
Proof of Proposition 8.8. The proof is carried out by the induction of the
number of the components P Q. The following Claims 1 and 2 give the
first step of the induction.
Claim 1. If P Q consists of one component, then the genus one 1-bridge
position X
Q
Y of K is obtained from A
P
B in a standard manner.
Proof. Let D
X
= P X, and D
Y
= P Y . Then D
X
, D
Y
are disks properly
embedded in X, Y respectively, each of which intersect K in two points such
that D
X
= D
Y
. Since P is separating in S
3
, D
X
is separating in X, and
this shows that D
X
is contractible in Q. Let E be the disk in Q bounded
by D
X
(= D
Y
). Then E intersects K in two points. Without loss of
generality, we may suppose that E is contained in A. Since E separates
the boundary components of each component of K A, we see that E is
K-incompressible in A. Then, by Lemma 7.3(2) we have either one of the
following two cases.
Case 1. E is not K-parallel to one of D
X
or D
Y
in A.
In this case, we may suppose without loss of generality that E is not K-
parallel to D
X
. Then, by Lemma 7.3 (2), we see that E is K-parallel to D
Y
in Y , and, by Lemma 8.4, we see that there is a K-compressing disk D for
X in X such that D intersects K in one point and D D
X
= , and these
imply that the genus one 1-bridge position X
Q
Y of K is K-isotopic to a
genus one 1-bridge position which is obtained from A
P
B in a standard
manner by using the arc K (B X).
Figure 8.6.
Case 2. E is K-parallel to D
X
and D
Y
in A.
In this case, we consider the genus one surface F = Q B. Note that
F is a K-essential simple closed curve in P . Then, by Lemma 7.3, we see
that there is a K-compressing disk D for F in B. Without loss of generality,
we may suppose that D is contained in X. Then we have the following two
cases.
Case 2.1. D K = .
In this case, we obtain a K-compressing disk D
0
for B in B by com-
pressing F along D. By applying a slight isotopy, we may regard D
0
as a
142 T. KOBAYASHI AND O. SAEKI
K-compressing disk for X in X such that D
0
= E. Then, by Lemma 8.6,
we s ee that the arc K X is pushed into by an isotopy rel in X to an arc
contained in E. Then we further push the arc along the parallelism between
E and D
Y
to an arc contained in the disk D
Y
. We denote by K
0
the image
of K under this isotopy. Then K
0
is contained in B and K
0
B is an arc.
Since this isotopy does not move the component of K B contained in Y
(= c`(K
0
IntB)), c`(K
0
IntB) is a trivial arc in B. This implies that K
is a trivial knot, a contradiction.
Figure 8.7.
Case 2.2. D K consists of one point.
In this case, we obtain, by applying K-compression on F along D and
slight isotopy, a disk D
00
in B with D
00
= E, and D
00
intersects K in two
points.
Subclaim. D
00
E bounds a K-parallelism betwee n D
00
and E in X.
Proof. Ass ume not. Suppose that D
00
is K-compressible in X. Then, by
applying K-compression and slight isotopy, we obtain a K-compressing disk
E
0
in B such that E
0
= E. Then, by the argument of Case 2.1, we see that
K is a trivial knot, a contradiction. Suppose that D
00
is K-incompressible
in X. By the assumption, we see that E
0
D
X
bounds a rational tangle in
X which is not a K-parallelism between E
0
and D
X
. Then, by Lemma 7.3
(2), we see that E
0
and D
Y
must bound a K-parallelism in B. But this is
impossible, since D
Y
contains the boundary components of a component of
K B.
By the subclaim together with the arguments in Case 1, we see that the
genus one 1-bridge position X
Q
Y of K is K-isotopic to a genus one 1-
bridge position which is obtained from A
P
B in a standard manner by
using the arc K (B X).
RUBINSTEIN–SCHARLEMANN GRAPHIC 143
Figure 8.8.
This completes the proof of Claim 1.
Claim 2. If P Q consists of two components, then the genus one 1-bridge
position X
Q
Y of K is obtained from A
P
B in a standard manner.
Proof. Let D
1
, A
0
, D
2
be the closures of the components of P (P Q) such
that D
1
and D
2
are disks, and A
0
is an annulus. Without loss of generality,
we may suppose that D
1
D
2
is contained in X, and A
0
is contained in
Y . Since P is separating in S
3
, we see that both D
1
and D
2
are either
separating or non-separating in X.
Case 1. Both D
1
and D
2
are separating in X.
Let E be the disk in Q bounded by D
1
. By changing subscript, if
necessary, we may suppose that D
2
E = , i.e., D
1
is “outer” than D
2
.
Without loss of generality, we may suppose that E is contained in A. Then
we have the following case s.
Case 1.1. D
1
and E are K-parallel in X.
In this case, we can push D
1
along the parallelism out of X to make
P X = D
2
. Hence, we have the conclusion by Claim 1.
Case 1.2. D
1
and E are not K-parallel in X.
In this c ase , we first claim that D
1
is K-incompressible in X. Assume
that D
1
is K-compressible in X. Then, by Lemma 8.6, the c omponent of
K D
1
contained in X is isotopic rel to an arc in D
1
joining D
1
K. We
denote by K
0
the image of K under this isotopy. Then K
0
is contained in A
and K
0
A(= K
0
D
1
) is an arc, and c`(K
0
IntA) is a trivial arc in A (see
Claim 1, Case 2.1). This implies that K is a trivial knot, a contradiction.
Hence D
1
is K-incompressible in X.
Let B
3
be the 3-ball in X bounded by D
1
E. By Lemma 8.4, we see
that (B
3
, K B
3
) is a rational tangle. Ass ume that E is K-compressible
in A. Then by applying the last half of Lemma 7.4 to E in A, we see that
(B
3
, K B
3
) is a K-parallelism, c ontradicting the fact that D
1
and E are not
K-parallel in X. Hence E is K-incompressible in A. Then, by Lemma 7.3
(2), we see that E and c`(P D
1
) bounds a K-parallelism in A. However
144 T. KOBAYASHI AND O. SAEKI
this is impossible, since the boundary components of a component of K A
is contained in c`(P D
1
). This shows that Case 1.2 does not occur.
Case 2. Both D
1
and D
2
are non-separating in X.
In this case, we first note that, since D
i
(X K) (i = 1, 2) consists of two
points, the two points KQ are contained in a component of Q(D
1
D
2
).
Then, let R be the closure of the component of X (D
1
D
2
) which does
not contain K Q. Without loss of generality, we may suppose that R is
contained in A. We have the following cases.
Case 2.1. Both D
1
and D
2
are K-incompressible in X.
In this case, we have the following subcases.
Case 2.1.1. D
1
and D
2
are not K-parallel in X.
Figure 8.9.
By Lemma 8.5, we see that (R, K R) is a rational tangle. By Lemma 8.9,
we see that A
0
is parallel to the annulus R Q in Y , and, hence, P is K-
isotopic to R. Now let a be the comp onent of K B that is contained in
X. Then, by Lemma 8.5 again, we see that the torus obtained from R by
adding a tube along a is K-isotopic to Q. Hence we have seen that the genus
one 1-bridge position X
Q
Y of K is obtained from A
P
B in a standard
manner by using the arc a.
Case 2.1.2. D
1
and D
2
are K-parallel in X, and there exists a K-incompress-
ible disk D
0
in X such that D
0
intersects K in two points, D
0
is non-
separating in X, D
0
(D
1
D
2
) = , and D
0
is not K-parallel to D
1
(or
D
2
).
Figure 8.10.
RUBINSTEIN–SCHARLEMANN GRAPHIC 145
Let R
0
be the closure of the component of X (R D
0
) which does not
contain K Q. By exchanging subscript, if necessary, we may suppose that
R
0
R = D
2
. Let A
0
= R
0
Q, and D
00
= A
0
D
0
. We note that D
00
is a
disk properly embedded in B, which intersects K in two points. Since D
00
separates the boundary components of each component of K B in P , we
see that D
00
is K-incompressible. Moreover D
00
and D
2
are not K-parallel
in X, hence in B. Hence, by Lemma 7.3 (2), we see that D
00
and c`(P D
2
)
(= D
1
A
0
) are K-parallel in B. This shows that P is K-isotopic to R
0
.
Then, by the argument of Case 2.1.1, with regarding R
0
as R we see that
the genus one 1-bridge position X
Q
Y of K is obtained from A
P
B in a
standard manner.
Case 2.1.3. D
1
and D
2
are K-parallel in X, and there does not exist a
K-incompressible disk D
0
as in Case 2.1.2.
Figure 8.11.
In this case, it is easy to see that the argument of Case 2.1.1 works to
show that the genus one 1-bridge position X
Q
Y of K is obtained from
A
P
B in a standard manner.
Case 2.2. Either D
1
or D
2
is K-compressible in X.
Without loss of generality, we may suppose that there is a K-compressing
disk D for D
1
D
2
such that D D
2
.
By applying K-compression on D
2
, we obtain a compressing disk D for
X in X such that DK = . By applying a slight isotopy, we may suppose
that D
2
D = . By Lemma 8.9, we see that A
0
and R Q are parallel in
Y . Hence, by K-isotopy, we may suppose that P = R, (and A = R).
Then we have the following subcases.
Case 2.2.1. D is not contained in R.
146 T. KOBAYASHI AND O. SAEKI
Figure 8.12.
Since D
1
and D
2
are K-parallel, we see that D
1
is also K-compressible
in X. Then, by Lemma 8.6, we see that there is a disk in X such that
D
1
= D
1
= γ an arc, and (K X) = c`( γ). Hence, we
can move K by an isotopy along such that (K X) is moved to γ
(hence, the comp onent of K B which inters ects Y is not changed by this
isotopy). Since each component of K B is a trivial arc, this shows that K
is a trivial knot, a contradiction.
Case 2.2.2. D is contained in R.
Figure 8.13.
Let R
0
= c`(X R), and A
= R
0
X. Then A
is an annulus properly
embedded in B such that each component of A
is K-essential in B,
and A
intersects K in two points. By Lemma 7.3, we see that there is a
K-compressing disk D
0
for A
in B.
Then we claim that D
0
is contained in X. In fact, assume that D
0
is
contained in Y . Since H
1
(S
3
) = {0}, we see that D
0
bounds a disk E in
A
such that E K consists of two points, and D
0
K = . Let B
30
be the
3-ball in Y bounded by D
0
E, and B
3
= c `(X N(D)). By Lemma 8.3
(1), we see that K B
3
(= K X) is a trivial arc in B
3
. By Lemma 8.3 (3),
we see that K B
30
is a trivial arc in B
30
. Note that B
3
B
30
is a 3-ball,
and B
3
B
30
= E. This shows that K is a trivial knot, a contradiction.
Hence D
0
is contained in X. Since each component of A
separates the
boundary points of each component of K B in P , we see that D
0
K 6= ,
RUBINSTEIN–SCHARLEMANN GRAPHIC 147
and, hence, D
0
K consists of a point. By applying K-compression on A
along D
0
, we obtain two disks D
0
1
, D
0
2
in B such that D
0
1
= D
1
, and
D
0
2
= D
2
. Since D
0
i
(i = 1, 2) separates the boundary components of a
component of K B, D
0
i
is K-incompressible in B. Then, by Lemma 7.3
(2), D
i
D
0
i
(i = 1, 2) bounds a rational tangle in B. Let B
300
be the 3-ball
obtained from X by cutting along D
0
. We regard D
0
1
, D
0
2
as contained in
B
300
, and D
1
, D
2
are properly embedded in B
300
. By Lemma 8.3 (2), we
see that (B
300
, K B
300
) is a 2-string trivial tangle. Then we see that D
1
, D
2
are K-compressible, by the condition of Case 2.2. Hence, by the last half of
Lemma 7.4, we see that D
1
and D
0
1
(D
2
and D
0
2
resp.) are K-parallel. Let
a be the component of K B that is contained in X. Then, by the above
observations, we see that the ge nus one 1-bridge position X
Q
Y of K is
obtained from A
P
B in a standard manner by using a.
This completes the proof of Claim 2.
Completion of the proof of Proposition 8.8. Suppose that ](P Q) 3.
Note that the components of P Q are mutually K-parallel in P . Let D
1
,
A
1
, . . . , A
m
, D
2
be the closures of the components of P (P Q) such that
D
1
, D
2
are disks and A
1
, . . . , A
m
are annuli that are located on P in this
order. Then we have the following cases.
Case 1. Either D
1
or D
2
is non-separating in the solid torus.
Without loss of generality, we may suppose that D
1
is contained in X,
and is non-separating in X. Then A
1
is contained in Y , and, by Lemma 8.9,
there is a homeomorphism from A
1
× I to Y such that A
1
corresponds to
A
1
× {1/2}. Let U be the closure of the component of Y A
1
which does
not contain K Q.
Suppose that (IntU) P 6= . Then we can push the component of
(IntU) P along the parallelism U to X to reduce ](P Q), yet still have
at least two components A
1
.
Suppose that (IntU) P = . Then we can push A
1
along the parallelism
U to X to reduce ](P Q) by two.
In either case we have the conclusion by the assumption of the induction.
Case 2. Both D
1
and D
2
are separating in the solid torus.
If D
1
or D
2
is K-boundary parallel, then we can apply the assumption
of the induction by the argument as in Case 1. Hence we suppose that D
1
and D
2
are not K-boundary parallel. Let E
i
be the disk in Q bounded by
D
i
. Without loss of generality, we may suppose that E
1
E
2
, and D
1
is
contained in X. Then we have the following subcas es .
Case 2.1. D
1
is K-incompressible in X.
Let B
3
be the closure of the component of X D
1
such that B
3
=
D
1
E
1
. Then, by Lemma 8.4, (B
3
, K B
3
) is a rational tangle. Suppose
148 T. KOBAYASHI AND O. SAEKI
(IntB
3
) P 6= . We note that each component of c`((IntB
3
) P ) is an
annulus whose boundary components are parallel to D
1
in Q K. This
shows that each annulus is K-incompressible in B
3
. Hence, by Lemma 7.3
(3), we see that the closure of each component of (IntB
3
) P is boundary
parallel annulus. Hence, we can push them to Y to re duce ](P Q), and we
have the conclusion by the assumption of the induction. If (IntB
3
) P = ,
then by Lemma 7.3 (2), the last half of Lemma 7.4, and the assumption that
D
1
is not K-boundary parallel in X, we se e that c`(P D
1
) is K-isotopic
to E
1
rel D
1
. Hence, by the argument in Claim 1, we see that the genus
one 1-bridge position X
Q
Y of K is obtained from A
P
B in a standard
manner.
Case 2.2. D
1
is K-compressible in X.
We moreover have the following subcases .
Case 2.2.1. D
2
Y , and D
2
is K-incompressible in Y .
Let B
3
be the closure of the component of Y D
2
such that B
3
= D
2
E
2
.
In this case, s ince D
1
IntE
2
, we see that (IntB
3
) P 6= . Hence, by
the argument of Case 2.1 (for the case (IntB
3
) P 6= ), we can show that
the genus one 1-bridge position X
Q
Y of K is obtained from A
P
B in a
standard manner.
Case 2.2.2. D
2
Y , and D
2
is K-compressible in Y .
In this case, by K-compressing D
2
, we obtain a K-compressing disk D
0
2
for Y in Y such that D
0
2
= D
2
. Then, by Lemma 8.3 (3), we see that
K Y is rel isotopic to an arc α
Y
in E
1
. And we also see that K X is
rel isotopic to an arc α
X
in E
1
such that α
X
α
Y
= α
X
= α
Y
. Hence
K is a trivial knot, a contradiction.
Case 2.2.3. D
2
X.
Let T
0
and B
3
be the closures of the component of T D
1
such that T
0
is
a solid torus and B
3
is a 3-ball. Without loss of generality, we may suppose
that B
3
-side of D
1
is contained in B. Since D
1
is K-compressible, we see
that K T
0
is rel isotopic in T
0
to an arc α in D
1
, by an isotopy that
does not move K T
0
. Note that since D
2
X, c`(K T
0
) is a component
of K B, hence, a trivial arc in B. This shows that K is a trivial knot, a
contradiction.
This completes the proof of Proposition 8.8.
Proof of Theorem 8.2. Let A
P
B be a Heegaard splitting which gives a 2-
bridge position of K, and X
Q
Y a Heegaard splitting which gives the given
genus one 1-bridge position of K. By Proposition 8.7, it is enough to assume
that X
Q
Y is strongly K-irreducible for the proof of Theorem 8.2. Then, by
Proposition 7.5, and Corollary 6.22, we may suppose that P and Q intersect
RUBINSTEIN–SCHARLEMANN GRAPHIC 149
in non-empty collection of simple closed curves which are K-essential in both
P and Q. Then we have the conclusion by Proposition 8.8.
Appendix A.
Let γ be a system of trivial arcs in a handlebody H, and p :
˜
H H the
two fold branched cover of H along γ.
Let F be a surface properly embedded in H, which is in general position
with respect to γ. Then, by using Z
2
-equivariant loop theorem [12], we see
that:
Lemma A.1. F is γ-incompressible if and only if
˜
F (= p
1
(F )) is incom-
pressible.
Moreover, by using Z
2
-equivariant cut and paste argument as in [9, Proof
of 10.3], we see that:
Lemma A.2. A γ-incompressible surface F is γ-boundary compressible if
and only if
˜
F is boundary compressible.
By using Z
2
-Smith conjecture ([21], [16]) together with the Z
2
-equivariant
cut and paste argument and the irreducibility of H, we have:
Lemma A.3. A γ-incompressible surface F is γ-boundary parallel if and
only if
˜
F is boundary parallel. In particular, if F is a disk intersecting γ in
one point, and F bounds a disk D in H such that D intersects γ in one
point, then F is γ-boundary parallel (in fact, F and D are γ-parallel).
Appendix B.
Let (B, β) be a 2-string trivial tangle, and (
˜
B,
˜
β) the 2-fold branched cov-
ering space of B along β. Then
˜
B is a solid torus,
˜
β a system of two trivial
arcs in
˜
B, and the covering translation τ is a π-rotation along
˜
β (for details,
see [3, Chapter 12]).
Figure B-1.
150 T. KOBAYASHI AND O. SAEKI
We leave the proof of the next lemma to the reader.
Lemma B.0. Let F be an orientable incompressible surface properly em-
bedded in a solid torus. Then either:
1. F is a meridian disk,
2. F is a boundary parallel disk, or
3. F is a boundary parallel annulus.
Then, we s how:
Lemma B.1. Let D be a β-essential surface in B. Then D is a disk disjoint
from β, and D separates the compo nents of β.
Proof. Let
˜
D be the lift of D in
˜
B. By Lemmas A-1, A-3, we se e that
˜
B is
an essential surface in the solid torus
˜
B. By Lemma B-0, we see that
˜
D is
a meridian disk. Suppose that
˜
D
˜
β 6= . Then we see that
˜
D β consists
of a point, and, hence, D β consists of a point. However this implies that
D is β-inessential in B. Hence, by Lemmas A-3, and B-0, we see that D
is β-boundary parallel, a c ontradiction. Since D is β-essential in B, we
see that D separates the points β in B. This shows that D separates
the components of β.
Lemma B.2. Let F be a β-incompressible surface in B.
Then either (0) F is β-essential, (1) F is a β-boundary parallel disk inter-
secting β in at most one point, (2) F is a β-boundary parallel disk intersect-
ing β in two points and F separates (B, β) into the parallelism and a rational
tangle, or (3) F is a β-boundary parallel annulus such that F β = .
Proof. Let
˜
F be the lift of F in
˜
B. By Lemma A-1, we see that
˜
F is one
of (1), (2), or (3) of Le mm a B-0. It is easy to see that (1) ((2) resp.) of
Lemma B-0 corresponds to the conclusion (0) ((1) resp.). Suppose that
˜
F
is an incompressible annulus ((3) of Lemma B-0). Then it is easy to see
that we have conclusion (2) if
˜
F
˜
β 6= , and we have conclusion (3) if
˜
F
˜
β = .
Lemma B.3. Let D be a β-compressible disk in B such that D is β-
essential in B, a nd D β consists of two points.
Then D separates (B, β) into two tangles (B
1
, β
1
), and (B
2
, β
2
), where
(B
1
, β
1
) is a rational tangle such that there is a β-essential disk D
0
in
(B
1
, β
1
) with D D
0
= . Moreover if (B
2
, β
2
) happens to be a rational
tangle, then (B
2
, β
2
) is a β-boundary parallelism for D.
Proof. Let
˜
D be the lift of D in
˜
B. By Lemma A-1, we see that
˜
D is a
compressible annulus in
˜
B. Since D is β-essential, we see that, by com-
pressing
˜
D, we obtain two meridian disks, say D
1
and D
2
. Let B
3
1
, B
3
2
be
the closures of the components of
˜
B (D
1
D
2
). Then B
3
1
, B
3
2
are 3-balls,
RUBINSTEIN–SCHARLEMANN GRAPHIC 151
and the closure of a component of
˜
B
˜
D, say
˜
T , is obtained from one of B
3
1
or B
3
2
, say B
3
1
, by adding a 1-handle, and hence
˜
T is a solid torus, and there
is an equivariant compressing disk
˜
D
0
for
˜
D in
˜
T such that τ (
˜
D
0
)
˜
D
0
= .
Hence
˜
T gives a rational tangle, and p(
˜
D
0
) gives D
0
. Then the closure of
the other component of
˜
B
˜
D, say E, is obtained from B
3
2
by removing a
regular neighborhood of an arc properly embedded in B
3
2
, and taking the
closure. This show s that E is homeomorphic to the exterior of a knot in
S
3
by a homeomorphism such that E T is a regular neighborhood of
a meridian loop of the knot in E. Suppose that E is a solid torus, i.e.,
(E, K E) is a rational tangle. Then the knot is a trivial knot. Since
E T is a regular neighborhood of a meridian loop of the knot, we see
that E is a -parallelism for
˜
D, and by Lemma A-3, we see that (B
2
, β
2
) is
a β-boundary parallelism for D.
Appendix C.
Let α be a trivial arc in a solid torus T , and (
˜
T , ˜α) the 2-fold branched
covering space of T along α. Then
˜
T is a genus two handlebody, ˜α a 1-
string trivial arc in
˜
T , and the covering translation is a π-rotation along
˜α.
Figure C-1.
Lemma C.1. Let D be an α-compressing disk for T . Then D is either:
(1) a meridian disk of T with Dα = . In this case, we obtain, by cutting
(T, α) along D, a 1-string trivial tangle,
(2) a meridian disk of T with D α consists of one point, and we obtain
by cutting (T, α) along D, a 2-string trivial tangle, o r
(3) -parallel disk in T with Dα = . In this case, D cobounds a 1-string
trivial tangle in (T, α).
152 T. KOBAYASHI AND O. SAEKI
Proof. Let
˜
D be the lift of D in
˜
T . Then we have either
˜
D is a union of two
disks if D α = (Case 1), or
˜
D is a disk if D α consists of one point
(Case 2).
In Case 1, we have either each component of
˜
D is non-separating, or
separating, which correspond to the conclusions (1), (3) respectively. In Case
2, it is easy, by a homological argument, to see that
˜
D is non-separating,
and this gives conclusion (2)
Lemma C.2. Let D be a n α-essential disk in T such that D α consists
of tw o points. Then there exists an α-compressing disk D
0
for T such
that D
0
D = , and D
0
α consists of one point. Moreover, by cutting
(T, α) along D
0
, we obtain a 2-string trivial tangle (B, β) such that D is a
β-incompressible disk in (B, β) (hence D is β-boundary parallel).
Proof. Let
˜
D be the lift of D in
˜
T . By Lemma A-1, we see that
˜
D is an
essential annulus in
˜
T . Then it is easy to see that:
The annulus
˜
D is obtained from a meridian disk by attaching a band.
(For a pro of of this, see, for example, [13, Lemma 3.2].)
Let
˜
F = c`(
˜
T N(
˜
D), where N (
˜
D) is a regular neighborhood of
˜
D
in
˜
T . Note that
˜
F is compressible in
˜
T (in fact, slightly push off of gives
a compressing disk of
˜
F ). Hence, by Z
2
-equivariant loop theorem [12], we
have an equivariant compressing disk(s) G for
˜
F .
Claim 1. G consists of one disk, and, hence, G α consists of one p oint.
Proof. Ass ume that G consists of two disks D
1
, D
2
. Then we have the
following three cases.
Case 1. Each D
i
is separating in
˜
T .
In this case D
1
and D
2
are parallel and the closures of
˜
T G are two
solid tori T
1
, T
2
, and a 3-ball B, which is a parallelism between D
1
, and
D
2
. Note that ˜α is contained in B, and this shows that
˜
D is contained in
B, contradicting the incompressibility of
˜
D.
Figure C-2.
RUBINSTEIN–SCHARLEMANN GRAPHIC 153
Case 2. Each D
i
is non-separating, and G = D
1
D
2
is non-separating in
˜
T .
Since G
˜
D = , this contradicts the incompressibility of
˜
D.
Figure C-3.
Case 3. Each D
i
is non-separating, and G = D
1
D
2
is separating in
˜
T .
In this case, we see that D
i
(i = 1, 2) intersects
˜
β, a contradiction.
Figure C-4.
Hence, G is a disk, and G α consists of one point. Then:
Claim 2. G is non-separating in
˜
T .
Proof. Ass ume that G is separating. Then the closures of
˜
T G are solid
tori T
1
, T
2
with τ(T
i
) = T
i
. Moreover the fixed point set of τ |
T
i
is an arc
in T
i
. But, by using Z
2
-equivariant loop theorem, it is easy to see that such
τ|
T
i
does not exist.
By Claim 2, we see that we obtain a solid torus T
0
by cutting
˜
T along G,
and this shows that we obtain a 2-string trivial tangle (B, β), and obviously
D is β-incompressible.
Figure C-5.
154 T. KOBAYASHI AND O. SAEKI
Lemma C.3. Let D
1
, D
2
be mutually disjoint non α-parallel, α-essential
disks such that D
i
α (i = 1, 2) consists of two points. Then there exists
an α-compressing disk D
0
for T disjoint from D
1
D
2
such that D
0
α
consists of one point. Moreover each D
i
is non-separating in T , and, by
cutting (T, α) along D
0
, we obtain 2-string trivial tangle (B, β), and D
1
, D
2
are mutually non β-parallel, β-boundary parallel, β-incompressible disks in
(B, β).
Proof. Let
˜
D
i
be the lift of D
i
in
˜
T (i = 1, 2). By Lemma A-1, we see that
˜
D
i
is an incompress ible annulus in
˜
T . Let
˜
F = c`(
˜
T N (
˜
D
1
˜
D
2
,
˜
T )). The
argument in the proof of Lemma C-2 works in this case to s how that there
is an e quivariant compressing disk G for
˜
F such that G is non-separating,
and G intersects ˜α in one point. Then let
˜
T
0
be the solid torus obtained by
cutting
˜
T along G, and G
1
, G
2
the copies of G in
˜
T
0
. By Lemma B-0 (3),
we see that there are annuli A
1
, A
2
in
˜
T
0
such that A
i
and
˜
D
i
are parallel
(i = 1, 2). Since
˜
D
1
,
˜
D
2
are essential, we see that G
1
A
1
, G
2
A
2
. It is
easy to see that this gives the conclusion of Lemma C-3.
Lemma C.4. Let D be an α-compressible disk such that D is α-essential
in T , and D α consists of two points. Then there is a disk in T such
that D = D = γ an arc, and α = c`(γ). Particularly, if D
is separating in T , then D separates (T, α) into (T
0
, α
0
), and (B
0
, α
00
) such
that α
0
is a trivial arc in a solid torus T
0
. In this case, if (B
0
, α
00
) happens
to be a rational tangle, then (B
0
, α
00
) is an α-boundary parallelism.
Proof. Let
˜
D be the lift of D in
˜
T . By Lemma A-1, we see that
˜
D is a
compressible annulus in
˜
T . It is easy to se e that there is a compressing disk
˜
for
˜
D such that τ (
˜
∆) =
˜
∆, hence ˜α
˜
∆. Then the projection of
˜
gives
∆. Since D is β-essential, we see that, by compressing
˜
D along
˜
∆, we
obtain two meridian disks, say D
1
and D
2
, which are mutually parallel in
˜
T .
Suppose that D is separating in T . Then
˜
D is also separating in
˜
T , and the
closures of the components of
˜
B (D
1
D
2
) consist of B
3
and T
0
, where B
3
is a 3-ball, and T
0
is a solid torus. Then the closure of a component of
˜
T
˜
D,
say H, is obtained from T
0
by adding a 1-handle, and hence H is a genus two
handlebody. Now we consider the closure of the other component of
˜
T
˜
D,
say E. Then E is obtained from B
3
by removing a regular neighborhood of
an arc properly embedded in B
3
, and taking the closure. This shows that E
is homeomorphic to the exterior of a knot in S
3
by a homeomorphism such
that E T is a regular neighborhood of a meridian loop of the knot in E.
Suppose that E is a solid torus, i.e., (B
0
, α
00
) is a rational tangle. Then the
knot is a trivial knot. Since E T is a regular neighborhood of a meridian
loop of the knot, we see that E is a -parallelism for
˜
D, and by Lemma A-3,
we see that D is α-boundary parallel.
RUBINSTEIN–SCHARLEMANN GRAPHIC 155
References
[1] J.M. Boardman, Singularities of differentiable maps, Publ. Math. I. H. E. S., 33
(1967), 21-57.
[2] F. Bonahon and J.-P. Otal, Scindement s de Heegaard des espaces lenticulaires, Ann.
Sci. Ec. Norm. Sup., 16(4) (1983), 451-466.
[3] G. Burde and H. Zieschang, Knots, Walter de Gruyter, Berlin, New York, 1985.
[4] A. Casson and C. Gordon, Reducing Heegaard splittings, Topology and its Applica-
tions, 27 (1987), 275-283.
[5] J. Cerf, Sur les diff´eomorphismes de la sph`ere de dimension trois
4
= 0), Lecture
Notes in Math., 53 (1968), Springer-Verlag, Berlin and New York.
[6] H. Doll, A generalized bridge number for links in 3-manifolds, Math. Ann., 294 (1992),
701-717.
[7] M. Golubitsky and V. Guillemin, Stable mappings and their singularities, Graduate
Texts in Math., 14 (1973), Springer-Verlag, New York, Heidelberg, Berlin.
[8] W. Haken, Some results on surfaces in 3-manifolds, Studies in Modern Topology,
Math. Assoc. Amer., (1968), 39-98.
[9] J. Hempel, 3-manifolds, Princeton University Press, Princeton, N. J., 1976.
[10] M.W. Hirsch, Differential topology, Graduate Texts in Math., 33 (1976), Springer-
Verlag, New York, Heidelberg, Berlin.
[11] W. Jaco, Lectures on three-manifold topology, CBMS Regional Conf. Ser. in Math.,
43 (1980).
[12] P.K. Kim and J.L. Tollefson, Splitting the P.L. involutions of nonprime 3-manifolds,
Michigan Math. J., 27 (1980), 259-274.
[13] T. Kobayashi, Structures of the Haken manifolds with Heegaard splittings of genus
two, Osaka J. Math., 21 (1984), 437-455.
[14] H. Levine, Classifying immersions into R
4
over stable maps of 3-manifolds into R
2
,
Lect. Notes in Math., 1157 (1985), Springer-Verlag, Berlin, Heidelberg, New York,
Tokyo.
[15] J.N. Mather, Stability of C
mappings. V: Transversality, Adv. Math., 4 (1970),
301-336.
[16] J.W. Morgan and H. Bass (Eds.), The Smith Conjecture, Pure and Applied Math.,
112 (1984), Academic Press.
[17] K. Morimoto and M. Sakuma, On unknotting tunnels for knots, Math. Ann., 289
(1991), 143-167.
[18] H. Rubinstein and M. Scharlemann, Comparing Heegaard splittings of non-Haken
3-manifolds, Top ology, 35 (1996), 1005-1026.
[19] , Comparing Heegaard splittings-the bounded case, preprint.
[20] H. Schubert, Knoten mit zwei Br¨ucken, Math. Zeit., 66 (1956), 133-170.
[21] F. Waldhausen,
¨
Uber Involutionen der 3-Sph¨are, Topology, 8 (1969), 81-91.
156 T. KOBAYASHI AND O. SAEKI
[22] H. Whitney, On singularities of mappings of Euclidean spaces. I. Mappings of the
plane into the plane, Ann. of Math., 62 (1955), 374-410.
Received April 1, 1998 and revised June 24, 1999. The authors were supported by Grant-
in-Aid for Scientific Research, Ministry of Education, Science and Culture, Japan.
Department of Mathematics
Nara Women’s University
Kita-Uoya Nishimachi
Nara 630
Japan
E-mail address: tsuyoshi@cc.nara-wu.ac.jp
Department of Mathematics, Faculty of Science
Hiroshima University
Higashi-Hiroshima 739
Japan
E-mail address: saeki@top2.math.sci.hiroshima-u.ac.jp
PACIFIC JOURNAL OF MATHEMATICS
Vol. 195, No. 1, 2000
GEOMETRIC REALIZATIONS OF FORDY–KULISH
NONLINEAR SCHR
¨
ODINGER SYSTEMS
Joel Langer and Ron Perline
A method of Sym and Pohlmeyer, which produces geo-
metric realizations of many integrable systems, is applied to
the Fordy–Kulish generalized non-linear Schr¨odinger systems as-
sociated with Hermitian symmetric spaces. The resulting
geometric equations correspond to distinguished arclength-
parametrized curves evolving in a Lie algebra, generalizing
the localized induction model of vortex filament motion. A natu-
ral Frenet theory for such curves is formulated, and the general
correspondence between curve evolution and natural curvature
evolution is analyzed by means of a geometric recursion operator.
An appropriate specialization in the context of the symmet-
ric space SO(p + 2)/SO(p) × SO(2) yields evolution equations
for curves in R
p+1
and S
p
, with natural curvatures satisfying
a generalized mKdV system. This example is related to recent
constructions of Doliwa and Santini and illuminate s certain
features of the latter.
1. Introduction.
Shortly after it was discovered that the Korteweg-deVries equation could
be linearized via the spectral transform method [G-G-K-M], Shabat and
Zakharov [S-Z] showed that the method could also be applied to the (cubic)
non-linear Schr¨odinger equation,
(NLS)
t
= ψ
ss
+
1
2
|ψ|
2
ψ.
Almost concurrently, Hasimoto [Has] discovered the connection between
NLS and the localized induction equation (LIE), an idealized model of the
evolution of the curved centerline of a thin vortex tube in a three-dimensional
ideal fluid. (See [Bat] for a derivation, and [Ric] for the history of this equa-
tion, also known as the Betchov-Da Rios equation.) Denoting this evolving
centerline by γ(s, t) (where s is arclength along the curve and t is time), the
curve evolution in this model is described by
(LIE) γ
t
= γ
s
× γ
ss
= κB,
where κ(s) is the curvature and B the binormal. Recall that along a space
curve, the Frenet frame {T, N, B} satisfies the equations T
s
= κN, N
s
=
157
158 JOEL LANGER AND RON PERLINE
κT + τ B, B
s
= τN. The LIE-NLS connection is this: If a curve γ with
curvature κ and torsion τ evolves according to LIE, then the associated
complex curvature function, ψ = κe
i
R
s
τ(u)du
, evolves according to NLS.
In view of the LIE-NLS correspondance, it is not surprising that LIE
manifests familiar integrability characteristics, but in geometric form: Soli-
ton solutions, a hierarchy of conserved Hamiltonians in involution, a recur-
sion operator generating the corresponding infinite sequence of commuting
Hamiltonian vectorfields the localized induction hierarchy.
The Hamiltonian nature of LIE itself was introduced by Marsden and
Weinstein in [M-W]; the equation’s Poisson geometry was further elucidated
in [L-P1]; Yasui and Sasaki developed the structure of LIE in the setting of
hereditary operator, Hamiltonian pairs, and master symmetries [S-Y].
Other recent papers have addressed a variety of closely related geomet-
ric topics of a more concrete nature, including: Knotted soliton curves of
constant torsion [C-I]; planar, spherical, and constant torsion-preserving
curve evolution [L], [L-P3, L-P4]; integrable variational problems for curves
[Lan-S1, Lan-S2, Lan-S3]; pseudospherical surfaces and Weingarten sys-
tems [Per1, Per2], evolution of immersed Riemann surfaces in R
3
pre-
serving the Willmore integral [G-L]. It is by now clear that the localized
induction hierarchy is a rich source of examples and structure in the clas sical
differential geometry of curves and surfaces.
Here we consider natural generalizations of the LIE hierarchy in higher
dimensional spaces. Our starting point is the Fordy-Kulish [F-K] construc-
tion of a generalized nonlinear Schr¨odinger equation (gNLS) (with spectral
problem) associated to a Hermitian symmetric Lie algebra g. We apply
a technique due to Sym [Sym] and Pohlmeyer [Pohl], differentiation with
respect to the spectral parameter, which produces geometric realizations of
many integrable systems. By this route, we arrive at a generalized LIE hi-
erarchy for distinguished arclength-parametrized curves evolving in g, the
first three terms of which are:
γ
t
= γ
s
,
γ
t
= [γ
s
, γ
ss
],
γ
t
=
γ
sss
+
3
2
[γ
ss
, [γ
s
, γ
ss
]]
.
Here, [ , ] is the Lie bracket in g, and the subscript s denotes derivative by
a curve parameter which is unit speed with respect to the Cartan-Killing
form on g. In this setting, a direct generalization of Hasimoto’s result is
proved (Theorem 3), establishing the correspondence between the above
curve evolution equations and evolution of natural curvatures by equations
in the gNLS hierarchy; in particular, a curve evolving by the second order
flow, gLIE, has curvatures satisfying gNLS.
GEOMETRIC REALIZATIONS OF FORDY–KULISH SYSTEMS 159
Interestingly, as in the three-dimensional case , the odd-order flows are
more amenable to geometrically meaningful reductions. In fact, by an ad
hoc reduction, in the class of symmetric spaces SO(p + 2)/SO(p) × SO(2),
we were able to fully realize our original goal; namely, we obtain geometric
evolution equations applicable to arbitrary smooth curves in E
n
and S
n
. For
the third order flow, our equations take the form
γ
t
=
γ
sss
+
3
2
kγ
ss
k
2
γ
s
=
1
2
k
2
T +
X
i
(u
i
)
s
U
i
!
.
Here, u
1
, . . . , u
n1
are curvatures belonging to a natural frame T, U
1
, . . . ,
U
n1
. We show (Theorem 5) that the corresponding natural curvature vec-
tor, u = (u
1
, . . . , u
n1
), satisfies the vector modified Kortew eg-deVries equa-
tion:
(mKdV) u
t
=
u
sss
+
3
2
|u|
2
u
s
.
Note that these simple equations for γ and u are given, finally, without
reference to a Lie algebra.
We now describe the contents of the paper. Section 2 is a review of
the Fordy-Kulish construction of generalized NLS equations. In Section 3,
we apply the Sym-Pohlmeyer geometrization procedure in the Fordy-Kulish
setting, and develop a natural Frenet theory for the resulting curves. In
Section 4, we introduce the geometric recursion operator for the generalized
LIE hierarchy, and derive key variation formulas. Section 5 treats the special
class of Hermitian symmetric spaces mentioned above, and the reduction
yielding curve evolutions in Euclidean spaces and spheres. We note that
our constructions in the latter case are related to recent work of Doliwa
and Santini [D-S]; in fact, our investigation developed out of an effort to
better understand their equations. Since the completion of our paper, we
have learned from Chuu-Lian Terng of her own work (with K. Uhlenbeck)
on generalizations of LIE [T-U1, T-U2].
2. The Fordy-K ulish generalizations of NLS.
Following a standard framework in the theory of integrable systems, the
nonlinear soliton equations arise as c ompatibility conditions for an overde-
termined linear system
(LS) φ
s
= (λA + Q)φ, φ
t
= V φ.
This system involves two independent variables, s (“position”) and
t (“time”), and a scalar λ, the spectral parameter. The eigenfunction φ(s, t; λ)
has values in a Lie group G, while U(s, t; λ) = λA + Q(s, t) and V (s, t; λ)
have values in the Lie algebra g of G. Here Q is the potential, which
is meant to evolve isospectrally, hence the lack of λ-dependence. For the
160 JOEL LANGER AND RON PERLINE
Fordy-Kulish generalized NLS equations, g is taken to be the compact real
form of a complex semi-simple Lie algebra g
C
; in fact, g is required to be
a Hermitian symmetric Lie algebra, and A, Q, are specific to the struc-
ture of g. To recall briefly some of the relevant features of this structure,
g has a decomposition as a vector space sum, g = k m, of a com-
pact subalgebra k and complement m, satisfying the bracket conditions
[k, k] k, [m, m] k, and [k, m] m. Also, k is associated with a sp ec ial
element A in h, a Cartan subalgebra of g; namely, k is the commutator
algebra of A: k = kernel(ad
A
) = {B g : [B, A] = 0}. Further, J = ad
A
satisfies J
2
|
m
= Id, i.e., J is a complex structure on m . Such an element
A is fixed to form (LS) and Q is required to be an m-potential, that is,
Q(s, t) m for all t. The set of m-potentials Q(s) is clearly a vector space;
we will refer to a tangent vectorfield W as a m-field. Some of the above will
be explained more explicitly, as required for specializations, below. (Also,
see [F-K], [Hel] for more details.)
Cross-differentiating the equations in (LS) gives the zero curvature con-
dition U
t
V
s
+ [U, V ] = 0 or
(ZC1) Q
t
= V
s
[λA + Q, V ].
With the aim of finding V in terms of Q, such that the compatibility
condition (ZC1) is satisfied, a polynomial ansatz is invoked: V =
P
n
j=0
P
(j)
(s, t)λ
nj
. (Our indexing convention reverses the order of [F-K].)
The strategy here is to substitute this express ion for V into (ZC1), set the
coefficients of powers of λ equal to zero, then solve recursively for the P
(j)
and, finally, obtain a nonlinear PDE for the m-field Q (from the λ
0
term).
To carry this out requires the decomposition of g given above. Namely,
each P
(j)
is decomposed as P
(j)
= P
(j)
m
+ P
(j)
k
, with P
(j)
m
m and P
(j)
k
k.
Then, using the above bracket conditions and J
2
= Id, one obtains the
equations:
P
(0)
m
= 0,
P
(j)
m
= J(
s
P
(j1)
m
[Q, P
(j1)
k
]), j = 1, . . . , n,
s
P
(j)
k
= [Q, P
(j)
m
], j = 0, . . . , n,
Q
t
=
s
P
(n)
m
[Q, P
(n)
k
].
Note the first and third of these equations imply P
(0)
is necessarily a
constant in k we will take the “obvious choice” P
(0)
k
= P
(0)
= A. Also,
a choice of “constant of integration is made, at each stage, as we are ap-
parently required to compute an antiderivative to obtain P
(j)
k
. An essential
(and remarkable) feature of the recursion scheme is that the antiderivative
is explicit, and is polynomial in Q and its derivatives; here, we simply il-
lustrate this point with the important case n = 2. The required terms are
readily generated in the order: P
(1)
m
= Q, P
(1)
k
= 0, P
(2)
m
= [A, Q
s
], P
(2)
k
=
GEOMETRIC REALIZATIONS OF FORDY–KULISH SYSTEMS 161
1
2
[Q, [Q, A]] (the last antiderivative following from the Jacobi identity). The
induced evolution on the m-field Q is then given by
(gNLS) Q
t
= JQ
ss
1
2
ad
3
Q
A ,
which we will refer to as the Fordy-Kulish generalized NLS equation. (In
[F-K], the NLS equations are displayed componentwise, rather than in vec-
tor notation.)
The above recursion scheme may be more compactly described by intro-
ducing a recursion operator
˜
R =
˜
R
Q
which takes m-fields
˜
X to m-fields:
˜
R
˜
X = (
s
ad
Q
1
s
ad
Q
)J
˜
X.
(Appropriate specification of the antiderivative
1
s
depends on the context.)
Defining
˜
X
(j)
= JP
(j)
m
, we can now write the recursion scheme and nonlinear
equations as
˜
X
(1)
= JQ,
˜
X
(2)
=
˜
R
˜
X = Q
s
,
˜
X
(j+1)
=
˜
R
˜
X
(
j
)
, j = 1, 2, 3, . . . ,
Q
t
=
˜
X
(j+1)
, j = 0, 1, 2, . . . .
The last of these equations defines the (j + 1)rst term in the Fordy-Kulish
NLS hierarchy, the above NLS equation being the third term.
It will be useful to have a concrete (and particularly simple) example at
hand for illustrating the main ideas in the next few sections; thus, we begin
our:
Running example.
For the “classical NLS”, we take g = su(2). We use the basis A =
i
2
σ
3
,
B =
i
2
σ
1
, C =
i
2
σ
2
, where σ
1
, σ
2
, σ
3
are the Pauli matrices
σ
1
=
0 1
1 0
!
, σ
2
=
0 i
i 0
!
, σ
3
=
1 0
0 1
!
.
The bracket relations [A, B] = C, [B, C] = A, [C, A] = B, imply that
k = span(A) and m = span(B, C) define a Hermitian symmetric Lie algebra
structure on g = su(2). Writing Q = bB + cC, and plugging into the
generalized NLS equation yields:
Q
t
= J
Q
ss
+
1
2
|Q|
2
Q
,
where |Q|
2
= (b
2
+ c
2
). In this case, we can identify m with the complex
numbers (Q with ψ = b + ic), and then J coincides with multiplication by
i. Using this identification (and a time reversal), we obtain exactly (NLS),
given in the introduction.
162 JOEL LANGER AND RON PERLINE
We conclude this section by recording some useful general identities, to
be used in later s ections. First of all, we note that the above treatment of
the overdetermined linear system LS did not fully reflect the dependence of
φ on the three variables, s, t, and λ. Introducing W(s, t, λ) = φ
λ
φ
1
, we
write down the augmented linear system
φ
s
= Uφ, φ
t
= V φ, φ
λ
= Wφ,
with corresponding compatibility conditions
(ZC2) U
t
V
s
+ [U, V ] = V
λ
W
t
+ [V, W ] = W
s
U
λ
+ [W, U] = 0.
Secondly, the geometric objects related to LS will be expressed in terms
of conjugates of U, V , and W , for which we will use the following notational
shorthand: For B, C g, we write {B} = φ
1
Bφ, and {B, C} = {[B, C]}.
The following lemma (whose proof involves straighforward differentiation)
states analogues of a standard principle of rigid body mechanics (with “time”
being s, t, or λ): absolute velocity = relative velocity + tran sferred velocity
([Arn], p. 128).
Proposition 1. For any g-field B(s, t, λ),
i) {B}
s
= {B
s
} + {B, U},
ii) {B}
t
= {B
t
} + {B, V }, and
iii) {B}
λ
= {B
λ
} + {B, W }.
Finally, combining Proposition 1 with (ZC2) yields at once six simple
identities; three of these we will use, so we collect them in:
Proposition 2. For U, V , and W as above, we have
{V }
s
= {U
t
}, {W }
t
= {V
λ
}, and {W }
s
= {U
λ
}.
3. Sym-Pohlmeyer curves.
Throughout this section, we “freeze time” in the definitions of the previous
section. In other words, we consider a time-independent potential Q(s), and
suppose φ = φ(s; λ) satisfies the linear system φ
s
= Uφ = (λA + Q)φ, for
each value of the “parameter” λ. Setting W (s, λ) = φ
λ
φ
1
, we consider the
g-valued function
γ(s, λ) = {W } = φ
1
φ
λ
.
By Proposition 2, we have γ
s
= {W }
s
= {U
λ
} = {A}. If K is the Cartan-
Killing form on g, K(B, C) = tr(ad
B
ad
C
), then by Ad-invariance of K,
K(γ
s
, γ
s
) = K(A, A) = constant. In fact, K(A, A) = tr((ad
A
)
2
) = d,
where d = dim(m). Therefore, γ will be an arclength-parameterized curve
in g with respect to the rescaled form h, i =
1
d
K. Henceforth, we refer to
any curve in the one parameter family γ(s, λ) as a Sym-Pohlmeyer curve,
and denote by T the unit tangent vector T = γ
s
= {A}.
GEOMETRIC REALIZATIONS OF FORDY–KULISH SYSTEMS 163
To develop a Frenet theory for such curves, we first use Propos ition 1
to obtain an expression for the curvature normal κN of a Sym-Pohlmeyer
curve,
κN = T
s
= {A}
s
= {A
s
} + {A, λA + Q} = {A, Q} = {
˜
Q}.
We refer to the m-field Q itself as a curvature coefficients vector. Next,
we fix a basis for g of the form A
1
= A, A
2
, . . . , A
c
, B
1
, . . . , B
d
, where the
A
i
span k and the B
j
span m. Since the Killing form K is definite (g is
compact), we can further specify the basis to be orthonormal with respect
to h, i. The curvature normal vector is now expressible as
κN =
d
X
j=1
κ
j
B
j
=
d
X
j=1
κ
j
N
j
where N
j
= {B
j
}.
Next, we w rite the derivatives of the N
j
,
(N
j
)
s
= {B
j
}
s
= {B
j
, U } = {B
j
, λA + Q},
as a linear combination of themselves and the vectors T
i
= {A
i
}, i = 1, . . . , c.
Finally, we write the derivatives
(T
i
)
s
= {A
i
}
s
= {A
i
, Q}, i = 2, . . . , c.
At this point, we have a closed system of Frenet equations for the c + d =
dim(g) frame vectors T
i
, N
j
, involving only the curvature functions κ
j
, the
spectral parameter λ, and the structure constants of g.
Running example.
For Sym-Pohlmeyer curves in (su(2);
1
2
K)
=
(R
3
; h, i) with curva-
ture vector Q = bB + cC, the curvature normal is given by T
s
= κN =
c{B} + b{C} = κ
1
N
1
+ κ
2
N
2
, with κ
1
= c, κ
2
= b, N
1
= {B}, N
2
= {C}.
Our Frenet system is completed by the two equations (N
1
)
s
= κ
1
T λN
2
,
and (N
2
)
s
= κ
2
T + λN
1
. For λ = 0, this is none other than the natural
Frenet system for curves in R
3
(see, e.g., [Bis]). For a general value of the
constant λ, such a system may be thought of as inertial, in that the rigid
body defined by {T, N
1
, N
2
} (identifying s with time) has constant tangen-
tial component of angular velocity. The relationship to the classical Frenet
system can be written κ
1
+
2
= κe
, and N
1
+ iN
2
= (N + iB)e
, where
θ =
R
s
τ(u) + λdu; also, κ
2
= κ
2
1
+ κ
2
2
and τ = κ
2
(κ
1
(κ
2
)
s
κ
2
(κ
1
)
s
) λ.
While κ, τ and {T, N, B} are uniquely defined along a regular space curve γ
(with κ 6= 0), the curvatures κ
1
, κ
2
and frame vectors N
1
, N
2
are determined
(given λ) only up to multiplication by a complex unit, e
this freedom
corresponds to the choice of antiderivative in the above formulas. Aside from
this difference, the natural Frenet theory resulting from these definitions is
similar to the classical Fundamental Theorem for space curves. In partic-
ular, the set of unit speed curves Γ = {γ : R 7→ R
3
} can be parametrized
164 JOEL LANGER AND RON PERLINE
by the following data: initial position γ(0), initial frame T (0), N
1
(0), N
2
(0),
and shape κ
1
(s), κ
2
(s). For a given curve γ, this data is unique up to choice
of arclength parameter s, real parameter λ, and S
1
-parameter e
.
How does the Sym-Pohlmeyer construction fit together with the above
parametrization of Γ? Since Q = κ
2
B κ
1
C and λ are explicitly part
of the construction, it suffices to discuss the initial data γ(0), and T (0),
N
1
(0), N
2
(0). Writing T = Ad
φ
1
A, N
1
= Ad
φ
1
B, N
2
= Ad
φ
1
C, we see
that the initial frame is determined by the initial condition on φ, via the
adjoint representation of SU (2). In fact, the two-to-one homomorphism
Ad : SU (2) 7→ SO(3) implies all initial frames are achieved (twice) as
φ(0) varies over SU (2). Next, allow the initial condition on φ to depend
on λ, and regard φ(0, λ) as an arbitrary curve in SU (2). Since φ
1
φ
λ
de-
scribes the usual trivialization of the tangent bundle T SU (2), it follows that
γ(0, λ
0
) = φ
1
φ
λ
|
(0
0
)
is an arbitrary point in su(2)
=
R
3
. In conclusion,
the Sym-Pohlmeyer curves are precisely the unit speed curves in R
3
, and
the correspondence between the Sym-Pohlmeyer construction and Γ is fully
described.
Remarks.
1) The above example is prototypical in some, but not all respects. In
general, Sym-Pohlmeyer curves constitute a very special subclass of the unit
speed curves in g
=
R
c+d
the latter cannot all be described by only d
curvatures. In fact, the tan gent indicatrix T (s) of a regular curve in R
c+d
can be any smooth curve in the unit sphere S
c+d1
R
c+d
, whereas a Sym-
Pohlmeyer curve has tangent of the form T = {A} = Ad
φ
1
A. Now the
Ad-orbit of A can be identified with the Hermitian symmetric space G/K
(K having Lie algebra k). Thus, the tangent indicatrix of a Sym-Pohlmeyer
curve lies in G/K S
c+d1
g. In special cases , the above procedure may
produce a closed system with fewer than (c + d) frame vectors - this will be
true of our main construction of Section 5 - and the situation may resemble
the example more closely.
2) In the general case, it is reasonable to refer to k
1
, . . . , k
d
as natural
curvatures (though this term will have a more special meaning in Section
5). Note that the non-uniqueness of natural curvatures is described by
the group K (SU(1) = {e
} in the example). Spec ifically, suppose φ
s
=
(λA + Q)φ, and c onsider the Sym-Pohlmeyer curve γ = φ
1
φ
λ
. Now let
ϕ = φ
0
φ, where φ
0
K is any constant element. Then γ = ϕ
1
ϕ
λ
, and ϕ
satisfies the linear syste m ϕ
s
= (λA + Ad
φ
0
Q)ϕ, as is easily checked. So the
“gauge transformation” φ 7→ ϕ = φ
0
φ leaves the curve γ unchanged, while
transforming the natural curvatures according to Q 7→ Ad
φ
0
Q.
GEOMETRIC REALIZATIONS OF FORDY–KULISH SYSTEMS 165
4. The recursion operator and variation formulas.
Next we “un-freeze” time, and apply the above constructions to time-de-
pendent potentials Q(s, t), obtaining two-parameter families of unit speed
curves γ(s, t, λ). The t-derivatives of these will b e called Sym-Pohlmeyer
(variation) fields. Note that Proposition 2 gives a formula for such vector-
fields:
γ
t
= {W}
t
= {V
λ
}.
Also, V satisfies the zero curvature equation, V
s
= U
t
+ [U, V ] = Q
t
+
[λA + Q, V ]. Comparing k-components gives (V
k
)
s
= [Q, V
m
], i.e., the k-
component of V is determined by the m-c omponent of V , according to
V
k
=
1
s
[Q, V
m
]. Differentiation of this equation by λ shows that, similarly,
(V
λ
)
k
is determined by (V
λ
)
m
: (V
λ
)
k
=
1
s
[Q, (V
λ
)
m
]. It is convenient to
introduce an operator K which takes m-fields to k-fields:
K(B
m
) =
1
s
[Q, B
m
].
Thus any Sym-Pohlmeyer field is of the form Y = {K(B
m
) + B
m
}, for some
m-field B
m
(modulo integration constant in K). Now the ab ove definitions
easily imply the following formula for the arclength derivative of such a
Sym-Pohlmeyer field:
Y
s
= {(B
m
)
s
} + {K(B
m
), Q} + λ{B
m
, A} = {C
m
},
where C
m
is an m-field.
Remark. The result just obtained has the following (partial) interpretation
in the context of curve geometry (in a Riemannian m anifold). Suppose
γ(s, t) is any one-parameter family of arclength parametrized curves, and
let X be the vectorfield X =
t
γ. Then X
s
(the covariant derivative of X
with respect to the unit tangent T ) has no tangential component; in fact,
the condition for a vectorfield X to be locally arclength preserving ([L-P2])
is hX
s
, T i = 0. Of course, a Sym-Pohlmeyer field satisfies this condition:
hT, Y
s
i = h{A}, {C
m
}i = hA, C
m
i = 0, since k and m are orthogonal with
respect to the Killing form. In the special case c = 1, the Sym-Pohlmeyer
fields are exactly the locally arclength preserving vectorfields, while for c > 1,
the Sym-Pohlmeyer vectorfields form a strictly smaller class of vectorfields.
For a Sym-Pohlmeyer curve γ in a Hermitian symmetric Lie algebra g,
we now define three ope rators on vectorfields Y = {B} along γ.
(i) renormalization operator:
P({B}) = {K(B
m
) + B
m
} = {
1
s
[Q, B
m
] + B
m
};
(ii) geometric recursion operator:
RY = −P([T,
s
Y ]);
166 JOEL LANGER AND RON PERLINE
(iii) intertwining operator:
Z(Y ) = ad
A
(Ad
φ
Y ).
The next lemma explains the nomenclature for Z:
Proposition 3. For Y a Sym-Pohlmeyer field,
ZRY = (
˜
R λ)ZY.
Proof. Using the above computation of Y
s
, we have
ZRY = −ZP([T, Y
s
])
= −Z({A, (B
m
)
s
+ [K(B
m
), Q] + λ[B
m
, A]})
= −Z({J((B
m
)
s
+ [K(B
m
), Q]) + λB
m
})
= J(J((B
m
)
s
ad
Q
1
s
ad
Q
B
m
) + λB
m
)
= (
s
ad
Q
1
s
ad
Q
)B
m
λJB
m
= (
s
ad
Q
1
s
ad
Q
)J
2
B
m
JλB
m
= (
˜
R λ)JB
m
= (
˜
R λ)ZY.
Next, we consider a Sym-Pohlmeyer variation, γ(s, t, λ) = φ
1
φ
λ
= {W},
and the corresponding Sym-Pohlmeyer field (infinitesimal variation) X =
γ
t
= {V
λ
}.
Proposition 4. RX = {V +
˜
A} , for some constant
˜
A k.
Proof.
RX = −P

T,
s
t
γ

= −P([T, T
t
])
= −P([T, {A}
t
]) = −P([T, {A, V }])
= −P({A, [A, V ]}) = −P({J
2
V })
= P({V
m
}) = {K(V
m
) + V
m
} = {V +
˜
A}.
This last step uses K(V
m
) = V
k
+
˜
A, as observed above, with the arbitrary
“integration constant”
˜
A k explicitly displayed here.
Theorem 1. Variation of curvatures formula:
The time variation of the “curvature coefficients vector” Q induced by a
Sym-Pohlmeyer field X = γ
t
= {V
λ
} is given by
Q
t
= ZR
2
X + [Q,
˜
A].
In the “gauge term”, [Q,
˜
A],
˜
A k is a constant.
Proof. Combining Propositions 2 and 4, we have
ZR
2
(γ
t
) = ZR({V +
˜
A}) = −ZP([T, {V +
˜
A}
s
])
= −ZP([T, {Q
t
} + {
˜
A, Q}]) = −ZP({A, Q
t
+ [
˜
A, Q]})
GEOMETRIC REALIZATIONS OF FORDY–KULISH SYSTEMS 167
= (ad
A
)
2
(Q
t
+ [
˜
A, Q]) = Q
t
+ [
˜
A, Q] .
The term [Q,
˜
A] can be interpreted as follows. As explained in Remark
2 of the previous section, the non-uniqueness of natural curvatures for a
given curve corresponds to the se t of transformations Q 7→ Ad
φ
0
Q, where
φ
0
K is a constant. For a curve evolving in time t, φ
0
should be treated
as a function of t as well (with initial value φ
0
|
t=0
= Id). Differentiation
of φ
0
with respect to t results in the term [(φ
0
)
t
|
t=0
, Q] in the infinitesimal
variation of Ad
φ
0
Q.
The appearance of the square of the recursion operator in this formula
suggests that between the curve γ and curvature coefficients vector Q, there
is an intermediate object whose variation ought to be considered in this
context. The appropriate intermediate object is a Sym-Pohlmeyer frame
T
i
= {K
i
}, i = 1, . . . , c, , N
j
= {M
j
}, j = 1, . . . , d, as considered above.
Theorem 2. Variation of Frames formula:
If B g is constant, then the time variation of F = {B} induced by a
Sym-Pohlmeyer field X = γ
t
is given by
F
t
= [F, RX] + {
˜
A, B},
for some constant
˜
A k; i.e., RX is essentially the “ Darboux vector” for
any Sym-Pohlmeyer frame along γ.
Proof. Using Propositions 1 and 4, we compute
F
t
= {B, V } = [{B}, {V }] = [F, RX] + {
˜
A, B}.
We are now in a position to geometrize the Fordy-Kulish NLS hierarchy,
the first few terms of which we list here for convenience:
˜
X
(1)
= JQ,
˜
X
(2)
= Q
s
,
˜
X
(3)
= JQ
ss
1
2
ad
3
Q
A,
˜
X
(n+1)
=
˜
R
˜
X
(n)
.
For a Sym-Pohlmeyer curve γ with curvature vector Q and with λ = 0,
let vector fields X
n
be defined along γ according to:
X
(0)
= {A} = T,
X
(1)
= {Q} = [γ
s
, γ
ss
],
X
(2)
=
1
2
[Q, [Q, A]] + [A, Q
s
]
=
γ
sss
+
3
2
[γ
ss
, [γ
s
, γ
ss
]]
,
X
(n+1)
= RX
(n)
.
168 JOEL LANGER AND RON PERLINE
Now consider the curve evolution equation γ
t
= X
(n)
. By Theorem 1 and
Proposition 3, we can write the corresponding curvature evolution as
Q
t
= ZR
2
X
(n)
+ [Q,
˜
A] = ZR
(n+1)
X
(1)
+ [Q,
˜
A]
=
˜
R
(n+1)
ZX
(1)
+ [Q,
˜
A] =
˜
R
(n+1)
˜
X
(1)
+ [Q,
˜
A]
=
˜
X
(n+2)
+ [Q,
˜
A].
We summarize this result (suppressing the gauge term, [Q,
˜
A]) as:
Theorem 3. Evolution of a Sym-Pohlmeyer curve (with λ = 0) by γ
t
=
X
(n)
corresponds to curvature evolution by Q
t
=
˜
X
(n+2)
. In particular, the
generalized LIE,
(gLIE) γ
t
= [γ
s
, γ
ss
],
corresponds to the curvature evolution by gNLS, Q
t
= JQ
ss
1
2
ad
3
Q
A (the
analogue of Hasimoto’s result).
Running example.
The geometric recursion operator for curves in R
3
can be written: RX =
−P(T ×
s
X). Here, × is the cross product in R
3
, and the reparameteriza-
tion operator P turns an arbitrary vectorfield along γ, Y = fT + gU + hV ,
into a locally arclength preserving vectorfield, PY =
R
s
(κ
1
g + κ
2
h)ds T +
gU + hV , by changing only the tangential component. Using the identifi-
cation of m with the complex plane, the operator Z may be regarded as a
simple isomorphism betwee n normal vectorfields Y = gN
1
+ hN
2
and com-
plex functions Z(Y ) = i(g + ih). On the other hand, Q has already been
identified with the complex function ψ = b + ic = κ
2
1
. Using these def-
initions, the infinitesimal variation of ψ induced by the vectorfield X = γ
t
may be written as:
t
ψ = ZR
2
X + irψ, r a real constant. This differ-
ential formula for the Hasimoto transformation easily implies, e.g., that if
γ(s, t) evolves by LIE,
t
γ = κB = κ
2
N
1
+ κ
1
N
2
, then ψ (with λ = 0)
evolves according to NLS. [There is a minor difference b e tween the formulas
discussed here and those of [L-P1, L-P2, L-P3]. In those references, ψ
was the “complex curvature function” κ
1
+
2
= i(b + ic) mentioned in the
introduction. The differential formula was written in terms of this ψ (with a
minor difference in the definition of Z); if
t
γ = κB, then NLS is satisfied
by the latter ψ as well (NLS being i-equivariant).]
5. Evolution of curves in R
p+1
and S
p
.
We have seen that the geometric realization of the Fordy-Kulish NLS hier-
archy is a sequence of evolution equations on the space of Sym-Pohlmeyer
curves in a (real compact) Lie algebra g. The Sym-Pohlmeyer curves have
curvature vectors Q which are m-valued. As stated in Sec tion 3, Sym-
Pohlmeyer curves in general form a proper subset of the set of all arc-length
GEOMETRIC REALIZATIONS OF FORDY–KULISH SYSTEMS 169
parameterized curves in the Lie algebra g. In this section we describe a spe-
cific instance of our constructions which allows for a more complete anal-
ysis and full geometric interpretation. We will consider the Lie algebra
g = so(p + 2) with subalgebra k = so(p) so(2) corresponding to the Her-
mitian symmetric space BDI. After describing the relevant structure and
commutation relations in appropriate detail, we give an explicit formula for
the the operator
˜
R
2
restricted to a distinguished subspace of m-fields.We
then show that the Sym-Polhmeyer curves associated with appropriately re-
stricted curvature functions can be naturally considered as corresponding
to all curves in the Euclidean space R
p+1
, and the geometric realizations of
the terms in an associated mKdV hierarchy appear as quite natural evolu-
tion equations on curves. We explicitly compute the first non-trivial term,
and show that it induces curvature evolution corresponding to a particu-
larly simple coupled mKdV system. This mKdV system is a rather special
reduction of a system which fits into the general framework of [F-K], [A-F]
(though BDI is an exceptional case in the framework of [A-F]).
We consider so(p + 2) lying in gl(p + 2, R). We have gl(p + 2, R) commu-
tation relations [e
j,k
, e
l,m
] = δ
k,l
e
j,m
δ
m,j
e
l,k
, 1 j, k, l, m p + 2, where
e
j,k
is the matrix with 1 in the jth row, kth column, zero otherwise. Setting
f
i,j
= e
i,j
e
j,i
, we can express so(p+ 2) commutation relations in the form
[f
j,k
, f
l,m
] = δ
j,m
f
k,l
+ δ
k,l
f
j,m
δ
j,l
f
k,m
δ
k,m
f
j,l
.
As it turns out, in addition to the natural notation for the so(p+2) basis,
{f
i,j
}, 1 i < j p + 2, it will be c onvenient to have a notation adapted
to a particular decomposition of so(p + 2); thus we define
A = f
1,2
,
X
j
= f
1,j+2
, j = 1, . . . , p,
Y
k
= f
k+2,2
, k = 1, . . . , p,
K
m,n
= f
m+2,n+2
, m, n = 1, . . . , p.
The so(p + 2) commutation relations now take the form:
[A, X
j
] = Y
j
, [A, Y
j
] = X
j
, [A, K
m,n
] = 0,
[X
j
, Y
k
] = δ
j,k
A,
[X
j
, X
k
] = [Y
j
, Y
k
] = K
j,k
,
[X
j
, K
m,n
] = δ
j,m
X
n
δ
j,n
X
m
,
[Y
j
, K
m,n
] = δ
j,m
Y
n
δ
j,n
Y
m
,
[K
j,k
, K
l,m
] = δ
j,m
K
k,l
+ δ
k,l
K
j,m
δ
j,l
K
k,m
δ
k,m
K
j,l
.
Now consider the following subspaces of g = so(p + 2):
k = span{A} span{K
m,n
},
m
x
= span{X
j
}, m
y
= span{Y
k
},
170 JOEL LANGER AND RON PERLINE
and
m = m
x
m
y
.
Part of the structure implicit in these definitions is summarized in
Proposition 5.
i) g = k m, and k = so(2) so(p);
ii) k is the commutator subalgebra of A in g;
iii) J = ad
A
|
m
satisfies J
2
= I.
In particular, g admits a Hermitian symmetric Lie algebra structure (as
defined in Section 2).
Remark. The proposition does not fully capture all the relevant structure
of g for the geometric considerations to follow. In this connection it should
be noted that the same Hermitian symmetric Lie algebra g = k m arises
as a byproduct of the standard construction of so(p + 2) as compact real
form of so(p+ 2, C). However, a different X, Y -decomposition of m appears,
which lacks the required properties; specifically, the X, Y -bracket relations
are not as simple as above.
Next, recall the recursion operator
˜
R =
˜
R
Q
(first introduced in Section
2) which takes an m-field
˜
X to an m-field
˜
R
˜
X (and which depe nds on the
m-potential Q). Henceforth, we adopt the following:
Specialization. Q is an m
x
-valued potential (more briefly, an m
x
-potential)
and
˜
X is an m
x
-field.
The fact that this specialization is preserved by
˜
R
2
is an immediate con-
sequence of the following:
Proposition 6. Let Q =
P
k
u
k
(s)X
k
, and
˜
X =
P
m
x
m
(s)X
m
. Then
i)
˜
R
˜
X =
P
k
s
x
k
+ u
k
P
l
1
s
(u
l
x
l
)
Y
k
;
ii)
˜
R
2
˜
X =
P
k
2
s
x
k
+
P
l
(
s
(u
k
1
s
(u
l
x
l
))
+u
l
1
s
(u
l
s
x
k
u
k
s
x
l
))
X
k
.
Proof. The proof is by straightforward c omputation; however, we include
it, since (i) depends on the nice bracket formula [X
j
, Y
k
] = δ
j,k
A, and (ii) in-
volves a noteworthy cancellation. Consider
˜
R
˜
X = (
s
J +ad
Q
1
s
ad
Q
J)
˜
X.
The first term,
s
J
˜
X, can immediately be written as
P
k
s
x
k
Y
k
. The
second term can be rewritten as
ad
Q
1
s
ad
Q
J
˜
X =
X
k,l,m
u
k
1
s
(u
l
x
m
)[X
k
, [X
l
, Y
m
]]
=
X
k,l,m
u
k
1
s
(u
l
x
m
)[X
k
, δ
l,m
A]
GEOMETRIC REALIZATIONS OF FORDY–KULISH SYSTEMS 171
=
X
k,l,m
u
k
1
s
(u
l
x
m
)δ
l,m
Y
k
=
X
k
u
k
X
l
1
s
(u
l
x
l
)Y
k
.
Summing these two terms gives the desired formula (i).
To prove (ii), write
˜
R
˜
R
˜
X = (
s
J +ad
Q
1
s
ad
Q
J)
˜
R
˜
X, and note that the
first term is
s
J
˜
R
˜
X =
s
(
P
k
(
s
x
k
+u
k
P
l
1
s
(u
l
x
l
)) X
k
); this accounts
for the first two terms of (ii). Next, using [X
i
, [X
j
, X
k
]] = [X
i
, K
j,k
] =
δ
i,k
X
j
δ
i,j
X
k
, one computes
ad
Q
1
s
ad
Q
J
˜
R
˜
X
=
X
i,j,k
u
i
1
s
u
j
s
x
k
+ u
k
X
r
1
s
(u
r
x
r
)
!!
(δ
i,k
X
j
δ
i,j
X
k
)
=
X
l,j
u
l
1
s
u
j
s
x
l
+ u
l
X
r
1
s
(u
r
x
r
)
!!
X
j
X
l,k
u
l
1
s
u
l
s
x
k
+ u
k
X
r
1
s
(u
r
x
r
)
!!
X
k
=
X
l,k
u
l
1
s
u
k
s
x
l
+ u
l
X
r
1
s
(u
r
x
r
)
!
u
l
s
x
k
+ u
k
X
r
1
s
(u
r
x
r
)
!!
X
k
=
X
l,k
u
l
1
s
(u
l
s
x
k
u
k
s
x
l
)X
k
,
which is the last te rm of formula (ii).
We return now to the Fordy-Kulish NLS hierarchy. The second evolution
equation, Q
t
=
˜
X
(2)
= Q
s
evidently preserves the space of m
x
-potentials.
In fact, writing Q =
P
i
u
i
(s)X
i
, the resulting evolution for the components
u
i
is given by (u
i
)
t
= (u
i
)
s
.
From the proposition, it now follows that the evolution Q
t
=
˜
X
(4)
=
˜
R
2
˜
X
(2)
also preserves the space of m
x
-potentials, and the same is true for
all of the even evolution equations Q
t
=
˜
X
(2n)
. In particular, we can apply
formula (ii) to the m
x
-field
˜
X
(2)
=
P
i
(u
i
)
s
X
i
, obtaining
˜
X
(4)
=
X
k
3
s
u
k
+
X
l
(
s
(u
k
1
s
(u
l
s
u
l
))
172 JOEL LANGER AND RON PERLINE
+ u
l
1
s
(u
l
2
s
u
k
u
k
2
s
u
l
))
X
k
.
Substituting
1
2
u
l
2
for
1
s
(u
l
s
u
l
), and u
l
s
u
k
u
k
s
u
l
for
1
s
(u
l
2
s
u
k
u
k
2
s
u
l
), we obtain a simpler expression for
˜
X
(4)
:
˜
X
(4)
=
X
k
3
s
u
k
+
3
2
X
l
u
l
2
s
u
k
!
X
k
.
It follows that the evolution equation Q
t
=
˜
X
(4)
in terms of the components
u
i
is a modified Korteweg-deVries system:
(mKdVS) (u
i
)
t
=
3
s
u
i
+
3
2
X
l
u
l
2
s
u
i
!
, i = 1, . . . , p.
Remark. So far, we have seen that the space of m
x
-fields (along m
x
-
potentials) is preserved by the operator
˜
R
2
; as a consequence, the even
terms in the NLS hierarchy induce evolution equations on the space of m
x
-
potentials. It turns out that a corresponding result holds in the general
setting of Hermitian symmetric Lie algebras, with the X, Y-decomp osition of
m mentioned in the previous remark. For reasons already given, one cannot
generally expect such simple formulas and equations corresponding to those
just presented. But from our point of view, the most important difference
with the present case shows up in the construction of Sym-Pohlmeyer curves
from curvature data Q. Note the key role of the X, Y -bracket relations in
the following
Proposition 7. Consider an m
x
-potential Q =
P
i
u
i
X
i
with associated
Sym-Pohlmeyer curve γ (with λ = 0). The unit tangent vectorfield T =
{A} and vectorfields U
i
= {Y
i
}, i = 1, . . . , p , defined along γ, satisfy the
following closed, linear system:
T
s
=
X
i
u
i
U
i
,
(U
i
)
s
= u
i
T.
Proof. We have
T
s
= {A}
s
= {A
s
} + {[A, Q]}
=
("
A,
X
i
u
i
X
i
#)
=
X
i
u
i
{Y
i
} =
X
i
u
i
U
i
;
(U
i
)
s
= {Y
i
}
s
= {(Y
i
)
s
} + {[Y
i
, Q]}
GEOMETRIC REALIZATIONS OF FORDY–KULISH SYSTEMS 173
=
Y
i
,
X
j
u
j
X
j
=
X
j
u
j
{[Y
i
, X
j
]}
=
X
j
u
j
δ
i,j
{A} = u
i
{A} = u
i
T.
Let Ψ be a (p + 1) × (p + 1) matrix which is the fundamental solution to
the matrix differential equation
Ψ
s
=
0 u
1
· · · u
p
u
1
0 · · · 0
.
.
.
.
.
.
.
.
.
u
p
0 · · · 0
Ψ,
Ψ(s
0
) = I
p+1
= (p + 1) × (p + 1) identity matrix.
Then by the fundamental theorem of differential equations, we can express
the moving frame T (s), U
i
(s) as
T (s)
U
1
(s)
.
.
.
U
p
(s)
= Ψ
T (s
0
)
U
1
(s
0
)
.
.
.
U
p
(s
0
)
.
It follows that the Sym-Pohlmeyer curve γ g actually lies in the
affine space γ(s
0
) + R
p+1
, where R
p+1
is here identified as the span of
T (s
0
), U
i
(s
0
), i = 1, . . . , p. Moreover, the equations given in the propo-
sition are the natural Frenet equations (see Running example in Section 3)
for a curve in R
p+1
with curvatures u
i
(s), i = 1, . . . , p and natural frame
T, U
i
, i = 1, . . . , p . Thus, we conclude that the Sym-Pohlmeyer curve γ
may be regarded as a general space curve in R
p+1
.
Since the natural Frenet equations are not so well-known, we take a mo-
ment to indicate some of their geometric significance. To begin with, the
conclusion just reached dep e nds on an analogue of the classical Fundamental
Theorem of Curve Theory (for curves in n-dimensional Euclidean space). In
particular, every smooth curve γ in R
p+1
satisfies the above system for some
choice of curvature functions u
i
(s), i = 1, . . . , p , and is uniquely dete rmined,
up to congruence, by these functions. (The converse statement differs a bit
from that of the classical theorem, in that the natural curvatures u
i
are
uniquely determined by a curve γ only after the frame T (s
0
), U
i
(s
0
) has
been specified at some initial point γ(s
0
).) Note that T
s
= κN =
P
i
u
i
U
i
,
174 JOEL LANGER AND RON PERLINE
implies κ
2
=
P
i
(u
i
)
2
, so one can recover the standard (first) curvature from
natural curvatures.
What’s more important, the curvatures u
i
measure the sphericity of a
curve in R
p+1
. In particular, suppose that, for s ome j, u
j
= c
j
= constant 6=
0 . Then (γ(s) + (1/c
j
)U
j
)
s
= T + (1/c
j
)c
j
(T ) = 0, so γ lies on a p-
dimensional sphere of radius 1/c
j
. Further, U
j
is the inward pointing unit
normal to the sphere along γ, and the remaining frame vectors determine a
natural Frenet system along the spherical curve γ in the sense of covariant
differentiation in the sphere. Namely,
T
T =
X
i6=j
u
i
U
i
,
T
U
i
= u
i
T, i 6= j.
More generally, if u
i
1
= c
1
, u
i
2
= c
2
, . . . , u
i
l
= c
l
, then γ lies on a (p+1l)-
dimensional sphere of radius (c
2
1
+ c
2
2
. . . c
2
l
)
1/2
. In the exceptional case,
u
i
1
= u
i
2
= . . . u
i
l
= 0, γ(s) lies on a (p + 1 l)-plane. This corresponds
to the case in classical Frenet theory in which the last l curvatures vanish
here the order of the curvatures matters.
As an application of the above discussion, we are now in a position to give
a purely geometric version of our earlier variation formula, in the context of
Euclidean and spherical curves.
Theorem 4. Variation of curvatures formula:
Let M be Euclidean space of dimension d = (p + 1), or a round sphere of
dimension d = p. Denote by G the (constant) scalar curvature of M (so
G = 0 or G =
1
r
2
in the case of a sphere of radius r). Let γ
t
= X =
αT +
P
d1
i=1
x
i
U
i
describe a variation of a curve in M through unit speed
curves, where U
i
, i = 1, . . . , (d 1), is a natural frame along γ(s, t). Then
the induced variation of the associated natural curvatures u
i
is given by
(u
i
)
t
= (
2
s
+ G)x
i
+
s
(αu
i
) +
d1
X
l=1
(u
l
1
s
(u
l
s
x
i
u
i
s
x
l
))
+
X
j
c
i,j
u
j
, i = 1, . . . , (d 1).
In the gauge term
P
j
c
i,j
u
j
, the c
i,j
are constants with c
i,j
= c
j,i
.
Proof. Combine Theorem 1 with Propositions 3 and 6, after taking account
of the above discussion.
We now merge the last few topics and consider the geometric evolution of
curves in R
p+1
and S
p
corresponding to the above mKdV system. According
to Theorem 3, the curve evolution γ
t
= X
(2)
corresponds to the curvature
evolution Q
t
=
˜
X
(4)
. In the present context, a fully geometric interpretation
GEOMETRIC REALIZATIONS OF FORDY–KULISH SYSTEMS 175
of this result is possible; note that the following theorem (like the previous
one) is formulated entirely in terms of curve geometry no Lie algebras!
Theorem 5. Motion of a curve γ(s, t) in R
p+1
by the geometric evolution
equation
γ
t
=
3
2
κ
2
T + T
ss
=
1
2
k
2
T +
X
i
(u
i
)
s
U
i
!
,
corresponds to curvature evolution by the mKdV system
(u
i
)
t
=
3
s
u
i
+
3
2
|u|
2
s
u
i
, i = 1, . . . , p.
Here, the functions u
i
are natural curvatures and k
2
= |u|
2
=
P
i
(u
i
)
2
the
squared curvature of γ. In particular, if on e of the natural curvatures u
j
is
initially constant along γ, then this condition is preserved, and γ evolves on
a sphere.
Proof. The first statement is easily obtained from the previous theorem, by
direct computation. Alternatively, in view of Theorem 3, it suffices to note
that in the present context,
X
(2)
=
1
2
[Q, [Q, A]] + [A, Q
s
]
=
1
2
X
i,j
u
i
u
j
[X
i
, [X
j
, A]] +
X
i
(u
i
)
s
[A, X
i
]
=
(
1
2
X
i
(u
i
)
2
A +
X
i
(u
i
)
s
Y
i
)
=
1
2
X
i
(u
i
)
2
T +
X
i
(u
i
)
s
U
i
!
=
3
2
κ
2
T + T
ss
.
It is also evident from the form of the coupled mKdV equations that the
condition u
j
= constant is preserved in time, so the last statement follows.
Remark. In the case k 6= 0, the ve ctorfield X
(2)
is readily expressed in
terms of the standard Frenet frame:
X
(2)
=
3
2
κ
2
T + T
ss
=
1
2
κ
2
T + κ
s
N + κτ B
.
Here, τ and B are, respectively, the second curvature and second normal (in
three dimensions, the torsion and binormal).
Expressed in this form, this vectorfield appeared in [L-P1] as the “next”
term (above X
1
= kB) in the loc alized induction hierarchy, and the connec-
tion to the (complex) mKdV equation was discussed. The planar version of
176 JOEL LANGER AND RON PERLINE
this curve evolution and the connection to mKdV were considered in [G-P]
and in [L-P3]. In [L-P4], the authors showed that the even terms of the
R
3
localized induction hierarchy preserve curves lying on the sphere S
2
, ex-
pressing these vectorfields in terms of the natural curvatures and frames. In
[D-S], Doliwa and Santini show that X =
1
2
κ
2
T + κ
s
N + κτB describes an
evolution of curves on S
3
, and discuss corresponding curve evolution equa-
tions in spheres of arbitrary dimension. Our curve evolution equations (of
this section) apparently coincide w ith those of [D-S] (though our approach
is considerably different). On the other hand, the corresponding curvature
evolutions of [D-S] are increasing complicated as dimension of the sphere in-
creases, and do not generally bear a close resemblance to the familiar (scalar)
mKdV equation; we attribute this to their use of standard Frenet systems,
rather than the natural Frenet systems we have employed here. Finally, our
general formalism suggests a new perspective on the main conclusions of
[D-S], regarding the characterization of integrable curve dynamics (as will
be discussed in a future paper).
In this work we have described a Lie-theoretic construction of generaliza-
tions of the localized induction hierarchy and, in a special case, have shown
how curvature evolution equations related to (but simpler in form than)
those of Doliwa and Santini may be extracted as a subhierarchy. These
geometric realizations of the Fordy-Kulish NLS systems have a structure
remarkably similar to the R
3
LIE equations studied previously. Given the
known relations between the localized induction hierarchy and classical geo-
metric constructions, it is not unreasonable to expect that the geom etric
realizations of the Fordy-Kulish NLS systems will have similar interesting
relations to geometry.
References
[Arn] V. Arnold, Mathematical Methods of Mechanics, Springer-Verlag, 1980.
[A-F] C. Athorne and A. Fordy, Generalized KdV and mKdV eq uations associated
with symmetric spaces, J. Phys. A, 20 (1987), 1377-1386.
[Bat] G.K. Batchelor, An Introduction to Fluid Dynamics, Cambridge University
Press, 1967.
[Bis] R. Bishop, There is more than one way to frame a curve, American Mathe-
matical Monthly, (March 1975), 246-251.
[C-I] A. Calini and T. Ivey, acklund transformations and knots of constant torsion,
J. Knot Theory and its Ramifications, 7(6) (1998), 719-746.
[D-S] A. Doliwa and P.M. Santini, An elementary geometric characterization of the
integrable motions of a curve, Physics Letters A, 185 (1994), 373-384.
[F-K] A.P. Fordy and P. Kulish, Nonlinear Schr¨odinger equations and simple Lie
algebras, Communications in Mathematical Physics, 89 (1983), 427-443.
GEOMETRIC REALIZATIONS OF FORDY–KULISH SYSTEMS 177
[G-L] O. Garay and J. Langer, Taimanov’s motion of surfaces and acklund trans-
formations for curves, J. Conformal Geometry and Dynamics, 3 (1999), 37-49.
[G-G-K-M] C.S. Gardner, J.M. Greene, M.D. Kruskal and R.M. Miura, Method for solving
the Korteweg-deVries equation, Physical Review Letters, 19 (1967), 1095-
1097.
[G-P] R. Goldstein and D. Petrich, The Korteweg-de Vries hierarchy as dynamics of
closed curves in the plane, Physical Review Letters, 67(23) (1991), 3203-3206.
[Has] H. Hasimoto, A soliton on a vortex filament, Journal of Fluid Mechanics, 51
(1972), 477-485.
[Hel] S. Helgason, Differential geometry, Lie groups, and symmetric spaces, Aca-
demic Press, New York, 1978.
[L] J. Langer, Recursion in curve geometry, New York J. Math., 5 (1999), 25-51.
[L-P1] J. Langer and R. Perline, Poisson geometry of the filament equation, J. Non-
linear Sci., 1 (1991), 71-93.
[L-P2] , The filament equation, the Heisenberg model, and the non-linear
Schr¨odinger equation, in W. Shadwick, P. Krishnaprasad, T. Ratiu, eds.,
Fields Institute Communications, Mechanics Day, AMS, Providence, RI, 1996,
181-188.
[L-P3] , The planar filament equation, in W. Shadwick, P. Krishnaprasad, T.
Ratiu, eds., Fields Institute Communications, Mechanics Days, AMS, Provi-
dence, RI, 1996, 171-180.
[L-P4] , Local geometric invariants of integrable evolution equations, J. Math.
Phys., 35 (1994), 1732-1737.
[Lan-S1] J. Langer and D. Singer, Knotted elastic curves in R
3
, J. London Math. Soc.,
30(2) (1984), 512-520.
[Lan-S2] , Liouville integrability of geometric variational problems, Comm.
Math. Helv., 69 (1994), 272-280.
[Lan-S3] , Lagrangian aspects of the Kirchoff elastic rod, SIAM Review, 38
(1996), 605-618.
[M-W] J. Marsden and A. Weinstein, Coadjoint orbits, vortices, and Clebsch variables
for incompressible fluids, Physica, 7D (1983), 305-323.
[Per1] R. Perline, Localized induction equation and pseudospherical surfaces, Journal
of Physics A, 27 (1994), 5335-5344.
[Per2] , Localized induction hierarchy and Weingarten systems, Physics Let-
ters A, 220 (1996), 70-74.
[Pohl] K. Pohlmeyer, Integrable Hamiltonian systems and interactions through qua-
dratic constraints, Comm. Math. Phys., 46 (1976), 207-221.
[Ric] R.L. Ricca, Rediscovery of the Da Rios equations, Nature, 352 (1991), 561-
562.
[S-Y] N. Sasaki and Y. Yasui, Differential geometry of the vortex filament equation,
J. Geom. Phys., 28(1-2) (1998), 195-207.
[S-Z] A. Shabat and V. Zakharov, Exact theory of two-dimensional self-focusing and
one-dimensional self-modulation of waves in nonlinear media, Soviet Physics
JETP, 34 (1972), 62-69.
178 JOEL LANGER AND RON PERLINE
[Sym] A. Sym, Soliton surfaces and their applications, in ‘Geometrical aspects of
the Einstein equations and integrable systems’, Lecture Notes in Physics, 239
(1985), 154-231.
[T-U1] C.L. Terng and K. Uhlenbeck, Poisson actions and scattering theory for inte-
grable systems, preprint.
[T-U2] , Schr¨odinger flows on Grassmanians, preprint math.dg/9901086.
Received September 2, 1998 and revised March 16, 1999.
Dept. of Mathematics
Case Western Reserve University
Cleveland OH 44106
E-mail address: jxl6@po.cwru.edu
Dept. of Mathematics and Comp. Sci.
Drexel University
Philadelphia PA 19104
E-mail address: rperline@mcs.drexel.edu
PACIFIC JOURNAL OF MATHEMATICS
Vol. 195, No. 1, 2000
SEIBERG–WITTEN INVARIANTS FOR 3-MANIFOLDS
IN THE CASE b
1
= 0 OR 1
Yuhan Lim
In this note we give a detailed exposition of the Seiberg-
Witten invariants for cl osed oriented 3-manifolds paying par-
ticular attention to the case of b
1
= 0 and b
1
= 1. These are
extracted from the moduli space of solutions to the Seiberg-
Witten equations which depend on choices of a Riemannian
metric on the underlying manifold as well as certain pertur-
bation terms in the equations. In favourable circumstances
this moduli space is finite and naturally oriented and we may
form the algebraic sum of the points. Given any two sets
of choices of metric and perturbation which are connected
by a 1-parameter family, we analyse in detail the singular-
ities which may develop in the interpolati ng moduli space.
This leads then to an understanding of how the algebraic sum
changes. In the case b
1
= 0 a topological invariant can be
extracted with the addition of a suitabl e counter-term, which
we identify (this idea is attributed to Donaldson). In the
case b
1
= 1 a topological invariant is defined which depends
only on cohomological information related to the perturba-
tion term. We prove a ‘wall-crossing’ formula which tells us
how the invariant changes with different choices of this per-
turbation. Throughout we pay careful attenti on to genericity
statements and the issue of orientations and signs in all the
relations. The equivalence of t his invariant in the case of an
integral homology sphere with the Casson invariant is treated
in Lim, 1999 (see also works of Nicolescu, preprint). The
equivalence with Reidemeister Torsion in the case b
1
> 0 is a
result of Meng & Taubes, 1996. Some related material is in
Marcolli, 1996, Froyshov, 1996 and in the survey Donaldson,
1996. Taubes, 1990 contains the originating construction in
this article in the context of flat SU (2)-connections.
1. The Seiberg-Witten Invariants.
We denote by Y an oriented 3-manifold. Let g be a Riemannian metric on
Y and P Y a spin-c structure (see for example [LM]). Denote by S Y
the associated positive spinor bundle, i.e., S = P ×
%
C
2
where % : spin
c
(3)
End
C
(C
2
) is the irreducible representation of the complex Clifford algebra
179
180 YUHAN LIM
Cl(R
3
) with %(dy) = +1, where dy is the oriented volume form on Y . Let
ξ : spin
c
(3) = spin(3)×
1}
U(1) U(1) be the map which takes the square
of the second factor. Then the bundle L = L
P
= P ×
ξ
C Y is the
determinant line bundle of P ; it is clearly a U(1)-bundle. We remark that a
classical fact is w
2
(Y ) = 0 and this gives the existence of a spin-c structure
P Y with c
1
(L) any given class in 2H
2
(Y ; Z) H
1
(Y ; Z).
For a pair (A, Φ) consis ting of a U(1)-connection on L and a section of S
the π-perturbed Seiberg-Witten equations (SW
π
) read:
F
A
=
1
4
σ, Φ) + ω, D
A+α
Φ = 0.
Here F
A
denotes the curvature 2-form of A and D
A
the Dirac operator
coupled to A. The term σ(·, ·) is a certain symmetric bilinear form S
Λ
2
(iR) (see Section 2 for details). The perturbation π consists of a pair
(α, ω) where α
1
(iR) and ω
2
(iR), = 0.
U(1) embeds into spin
c
(3) by the map ι(z) = [Id, z]. Given a smooth
map g : Y U (1) C, we have an automorphism of P given by the rule
p 7→ (g). This induces, by pulling back, the action (A, Φ) g(A, Φ) =
(A + 2g
1
dg, g
1
Φ). If (A, Φ) is a Seiberg-Witten solution, then g(A, Φ)
is also a solution. Thus the set of Seiberg-Witten solutions is invariant
under the above automorphisms of S. The automorphism g called a gauge
transformation and the group of all gauge transformations is the gauge group
denoted G.
We will consider the set of L
2
2
-SW
π
-solutions modulo gauge equivalence
(details in Sec. 2). The set of such solutions will be denoted by Z
π
(P ; g)
or Z
π,g
(P ). When the underlying metric is understood we shall omit it and
simply write Z
π
(P ). Fixing a value of k 1, denote by P
k
the space of all
perturbations π of class C
k
.
Theorem 1. For π from an open dense subset of P
k
the irreducible part
Z
π
(P ; g) of Z
π
(P ; g) is a finite set of points and these are naturally oriented.
Let #Z
π
(P ; g) denote the algebraic sum, assuming π as above. Then:
(i) if b
1
(Y ) > 1, #Z
π
(P ; g) is independent of g and π
(ii) if b
1
(Y ) = 1, #Z
π
(P ; g) depends only on the component of H
2
(Y ; R)\
{c
1
(L)
R
} in which [
i
2π
ω] lies in, ω being the 2-form component of π
(iii) if b
1
(Y ) = 0, #Z
π
(P ; g) is independent of π and g after the addition
of a counter-term ζ(π, g) which is a combination of the spectral invari-
ants of Atiyah-Patodi-Singer. #Z
π
(P ; g) + ζ(π; g) takes values in Z if
H
1
(Y ; Z) = {0} and Z
h
1
8|H
1
(Y ;Z)|
i
if H
1
(Y ; Z) 6= {0}.
The exact expression for the counter-term ζ(π, g) is in Proposition 17.
For b
1
(Y ) = 1 the formula for the change in #Z
π
(P ; g) when we cross the
‘wall’ in H
2
(Y ; R) defined by {c
1
(L)
R
} is given in Corollary 20.
SEIBERG-WITTEN INVARIANTS ... 181
Let spin
c
(Y ) denote the equivalence classes of spin-c structures on Y .
There is a well-defined map spin
c
(Y ) 2H
2
(Y ; Z) which sends a represen-
tative P to c
1
(L
P
).
Corollary 2. Let Y be connected. The Seiberg-Witten equations define an
oriented diffeomorphism invariant τ of Y in the following form:
(i) if b
1
(Y ) > 0, τ : spin
c
(Y ) Z
(ii) if b
1
(Y ) = 1, let spin
c
(Y ) be the set of pairs ([P ], U) where U a com-
ponent of H
2
(Y ; R)\{c
1
(L
P
)
R
}. Then τ : spin
c
(Y ) Z
(iii) if b
1
(Y ) = 0 and H
1
(Y ; Z) = {0} there is a unique spin-c structure
and so τ Z
(iv) if b
1
(Y ) = 0 and H
1
(Y ; Z) 6= {0}, then τ : spin
c
(Y ) Z
h
1
8|H
1
(Y ;Z)|
i
.
In the subsequent se ctions we work up to a proof of Theorem 1. Sec-
tion 2 discusses the framework for defining the moduli space and their first
properties. Section 3 looks at generic properties. The details of the proof of
Theorem 1 are in Section 4.
Addendum. The reviewer has brought to the attention of the author of
an alternative exposition of some of the material in this article, in [Ch].
2. The Moduli Space.
Throughout this section Y denotes a closed oriented 3-manifold with Rie-
mannian metric g and P Y a fixed spin-c structure.
2.1. The Basic Set-up.
As in the usual gauge theory set-up we work with the following spaces.
(For more details see, for instance, [M].) Let C(P ) denote the space of
pairs (A, Φ) consisting of a L
2
2
connection A on L and Φ a L
2
2
- section of
S Y . This forms a Hilbert manifold. Let G denote the space of L
2
3
gauge
transformations of S i.e., L
2
3
maps g : Y S
1
C. This forms a Hilbert
Lie group. G acts on C(P ) by g(A, Φ) = (A + 2g
1
dg, g
1
Φ). This action is
smooth with Hausdoff quotient B(P ).
A pair (A, Φ) is irreducible if Φ is not identically 0. Otherwise it is called
reducible. G acts freely on C
(P ), the open set of irreducibles and its quotient
is denoted by B
(P ). The projection map C
(P ) B
(P ) forms a principle
G-bundle. At a reducible (A, 0), for which we simply write A, the stabilizer
of G is exactly those gauge transformations g for which dg = 0; thus the
stabilizer is identified with U (1) C, the constant gauge transformations.
Let
p
k
(iR) denotes the p-forms on Y of class L
2
k
, and Γ
k
(S) the sections
of S of class L
2
k
. Since C(P ) is an affine space modelled on the vector space
1
2
(iR) × Γ
2
(S) the tangent space at any point is c anonically identified with
the vector space itself. On the other hand the tangent space to the identity
of G is identified with
0
3
(iR).
182 YUHAN LIM
The derivative of the map g 7→ g(A, Φ) at the identity is given by γ 7→
(2, γΦ). The tangent bundle of C(P ) carries a natural Riemannian met-
ric which is the L
2
-inner product on
1
2
(iR)×Γ
2
(S). This inner product is in-
variant under complex multiplication in Γ
2
(S). By taking the L
2
-orthogonal
to the image of the derivative of the gauge group action we obtain a slice
(A, Φ) + X
A,Φ
for the action at (A, Φ). X
A,Φ
is defined as
{(a, φ) | 2d
a = ihiΦ, φi}
1
2
(iR) × Γ
2
(S).
If Φ = 0 then X
A,Φ
reduces to ker d
× Γ
2
(S). The stabilizer of A pre-
serves X
A
and acts by z(a, φ) = (a, z
1
φ); therefore the stabilizer acts as
the opposite complex structure on Γ
2
(S). If Φ 6= 0 then a sufficiently small
neighbourhood N of zero in X
A,Φ
models an open set for the gauge equiva-
lence class [(A, Φ)] in B(P ). If Φ = 0 then the same is true except that we
should take N/U(1) instead.
The symmetric bilinear form σ, Ψ) Λ
2
(iR) for us will be defined as
the adjoint of Clifford multiplication, that is defined by the condition that
for all ω Λ
2
(iR),
hω · Ψ, Φi = hω, σ, Ψ)i.
The represe ntation c : Λ
2
(iR) End
C
(S) given by Clifford multiplication is
an isomorphism onto its image which is the trace-free Hermitian symmetric
endomorphisms of S. If we identify Λ
2
(iR) with its image under c then
σ, Ψ) =
1
2
Ψ
+ Ψ Φ
hΦ, Ψi
R
Id).
In the formula expressions of the form v w
mean the endomorphism
v w
(u) = vhu, wi
C
. We remark that this formula assumes the convention
that if τ is unit length in Λ
2
(R) then c() is to be unit length in End
C
(S).
Fix a perturbation term π = (α, ω) of class C
k
, k 1 (we assume this
from now on). To set up the moduli space we define the SW
π
-section s =
s
π,g
:
0
2
(iR) × C(P )
1
1
(iR) Γ
1
(S) by
s(η, A, Φ) =
F
A
1
4
σ, Φ) ω
+ 2, D
A+α
Φ ηΦ
.
Since we will want to vary the perturbation term later, we introduce the the
Banach s pace Q
k
of C
k
-sections of Λ
1
(Y ) iR and the Banach space
k
of
closed C
k
-sections of Λ
2
(Y ) iR. Then our perturbations π are from the
space P
k
which is Q
k
×
k
.
From the definition of the SW
π
-section, it would seem that the zeros
might capture a much larger set than the SW
π
-solutions themselves; but as
the following Lemma shows this is only so in a minor way.
Lemma 3. Let s(η, A, Φ) = 0. If Φ 6= 0 then η = 0 and (A, Φ) is a SW
π
-
solution. If Φ = 0 then η = constant and (A, 0) is a SW
π
-solution.
SEIBERG-WITTEN INVARIANTS ... 183
Proof. We claim that the vector ((F
A
1
4
σ, Φ) ω), D
A+α
Φ) is L
2
-
orthogonal to (2, ηΦ). Before we show this we recall some useful identi-
ties: (i) if η
0
k
(iR) then D
A+α
(ηΦ) = · Φ + ηD
A+α
Φ (ii) the Clifford
action of a
1
k
(iR) is equal to the action of a (iii) if η
0
k
(iR) then
hηΦ, Ψi = −hΦ, ηΨi. To prove the claim we compute:
h2, (F
A
1
4
σ, Φ) ω)i
L
2
hηΦ, D
A+α
Φi
L
2
= h2η, d
(F
A
ω)i
L
2
h∗2,
1
4
σ, Φ)i
L
2
hηΦ, D
A+α
Φi
L
2
=
1
2
h∗ · Φ, Φi
L
2
hηΦ, D
A+α
Φi
L
2
=
1
2
h · Φ, Φi
L
2
hηΦ, D
A+α
Φi
L
2
=
1
2
hD
A+α
(ηΦ), Φi
L
2
1
2
hηD
A+α
Φ, Φ i
L
2
hηΦ, D
A+α
Φi
L
2
=
1
2
hD
A+α
(ηΦ), Φi
L
2
+
1
2
hD
A+α
Φ, ηΦi
L
2
hηΦ, D
A+α
Φi
L
2
= 0.
The last step follows from D
A+α
being self-adjoint. Therefore s(η, A, Φ) = 0
if and only if (A, Φ) is a SW
π
- solution and (2, ηΦ) = 0.
Thus we identify the space of Seiberg-Witten solutions with
s
1
(0) {0} × C(P )
and the Seiberg-Witten moduli space Z
π
(P ) is the quotient by G of this,
where G acts only on the C(P ) factor. The local structure of the moduli
space near a solution (η, A, Φ) s
1
(0) is determined by the elliptic complex
associated to the map s:
0
3
(iR)
δ
0
A,Φ
0
2
(iR)
1
2
(iR) Γ
2
(S)
δ
1
η,A,Φ
1
1
(iR) Γ
1
(S)
where
δ
0
A,Φ
(γ) = (0, 2, γΦ)
δ
1
η,A,Φ
(ξ, a, φ) = ((da
1
2
σ, φ)) + 2, D
A+α
φ +
1
2
a · Φ ξΦ ηφ).
Since we will be interested only in the case where η = 0 we will in subsequent
notation omit it when that is understood. Thus we let H
i
A,Φ
, (i = 0, 1, 2)
denote the cohomology of the complex when η = 0.
Lemma 4. Let (A, Φ) be a SW
α,ω
-solution.
(i) If Φ 6= 0 then H
0
A,Φ
= 0, H
1
A,Φ
= H
2
A,Φ
(ii) If Φ = 0 then H
0
A,Φ
= H
0
(iR), H
1
A,Φ
= H
0
A,Φ
H
2
A,Φ
, H
2
A,Φ
= H
1
(iR)
H
A+α
.
Note. In this article H
k
(iR) always denotes the pure imaginary harmonic
forms of degree k and H
A
the kernel of D
A
.
184 YUHAN LIM
Proof. A direct computation shows that the L
2
-adjoint of δ
1
A,Φ
is given by
δ
1
A,Φ
(b, ψ) = (2d
b ihiΦ, ψi, δ
1
A,Φ
(a, ψ)). The adjoint of δ
0
A,Φ
on the other
hand, is δ
0
A,Φ
(η, a, φ) = 2d
a ihiΦ, φi. There fore
H
1
A,Φ
= ker δ
0
A,Φ
ker δ
1
A,Φ
= {(ξ, a, φ) | 2d
a ihiΦ, φi = 0, da
1
2
σ, φ) = 0,
D
A+α
φ +
1
2
a · Φ = 0, = 0, ξΦ = 0}
H
2
A,Φ
= ker δ
1
A,Φ
= {(b, ψ) | 2d
b ihiΦ, ψi = 0, db
1
2
σ, ψ) = 0,
D
A+α
ψ +
1
2
b · Φ = 0}.
The remainder of the proof follows easily.
The local structure of the moduli space may now be deduced by the
Kuranishi argument. Let (0, A, Φ) s
1
(0). If Φ 6= 0, then a neighbourhood
of [(0, A, Φ)] in Z
π
(P ) is modelled on the zeros of a map (the obstruction
map) Ξ : H
1
A,Φ
H
2
A,Φ
. If Φ = 0 then the same is true except that Ξ is
S
1
-equivariant and we should take Ξ
1
(0)/S
1
.
If (A, Φ) is a regular solution, i.e., H
2
A,Φ
= {0}, and Φ 6= 0 then Ξ
1
(0) is
exactly one point. Thus a regular irreducible solution is isolated. Therefore
we have:
Proposition 5. If Z
π
(P ) consists solely of gauge equivalence classes of reg-
ular solutions then Z
π
(P ) is a discrete set, i.e., every point is isolated.
2.2. Compactness and Regularity.
Proposition 6. Fix π P
k
, k 1.
(i) If (A, Φ) is a SW
π
-solution then (A, Φ) is gauge equivalent to a SW
π
-
solution of class L
p
k+1
, p 2.
(ii) Let {(A
i
, Φ
i
)}
i=1
be a sequence of SW
π
-solutions.
Then there is a subsequence {i
0
} {i} and gauge transformations {g
i
0
} such
that {g
i
0
(A
i
0
, Φ
i
0
)} converges in L
2
k
to a L
2
k
SW
π
-solution. In particular, this
converges in C(P ), and therefore B(P ).
Proof. The proof is due to [KM]. We include it here for completeness. The
Bochner formula for the Dirac operator reads (see for instance [LM])
D
A+α
D
A+α
Φ =
A+α
A+α
Φ +
1
4
κΦ +
1
2
F
A+α
· Φ,
κ being scalar curvature. We also have Kato’s inequality
1
2
|Φ|
2
h∇
A+α
A+α
Φi.
SEIBERG-WITTEN INVARIANTS ... 185
If (A, Φ) is a SW
α,ω
-solution then D
A+α
Φ = 0 and
hF
A+α
· Φ, Φi =
1
4
|σ, Φ)|
2
+ h(ω + ) · Φ, Φi
=
1
8
|Φ|
4
+ h(ω + ) · Φ, Φi.
Applying Kato’s inequality to the Bo chner formula we obtain
1
2
|Φ|
2
1
4
κ|Φ|
2
1
8
|Φ|
4
+ |ω + |.|Φ|
2
.
At a maximum for Φ, |Φ|
2
0. If this is non-zero we obtain
|Φ|
2
max
Y
(2κ + 8|ω + |, 0).
Since ω and α are in C
1
, we obtain a uniform pointwise bound on the spinor
component of any SW
α,ω
-solution.
Let us prove (ii). Suppose that (A
i
, Φ
i
) is a given sequence of SW
α,ω
-
solutions. Choose a fixed reference smooth connection A, and write A
i
=
A + a
i
. Then after a gauge transformation we may assume that da
i
= 0 and
the harmonic component of a
i
is uniformly bounded, since the component
group of maps Y U(1) is H
1
(Y ; Z), and thus H
1
(Y ; iR)/H
1
(Y ; iZ) is
compact. Let ˆa
i
be the L
2
-component of a
i
which is L
2
-pependicular to the
harmonic forms. Since the harmonic forms are C
, a
i
ˆa
i
must lie in C
k
for
every k. The SW
α,ω
equations together with the uniform pointwise bound
on Φ
i
gives (by ellipticity) a uniform L
p
1
bound for the ˆa
i
. Applying this
to the equation for Φ
i
this gives again by ellipticity a uniform L
p
1
bound on
the Φ
i
. Circulating inductively we terminate with uniform L
p
k+1
bounds on
both ˆa
i
and Φ
i
, since α and ω are assumed to be in C
k
. The uniform bound
holds for all p 2. Therefore the sequence (a
i
, Φ
i
) is uniformly bounded in
L
p
k+1
. By Rellich’s theorem a subsequence of (a
i
, Φ
i
) converges in L
p
k
. For
p sufficiently large, L
p
k
L
2
k+1
; thus the sequence converges in C(P ), since
the underlying topology in L
2
2
.
The proof of (i) follows from the preceding by applying it to the constant
sequence.
2.3. Reducible Solutions.
When Φ = 0, the Seiberg-Witten equations reduce to a single equation
for the connection A: F
A
= ω. If ω = 0 then the reducible (up to gauge
equivalence) is identified with the moduli space of flat U(1)-connections on
L. Thus a necessary condition is that c
1
(L)
R
= 0. If this is so, then by by
a well-known fact in differential geometry, the gauge equivalence classes of
flat connections is completely determined by the holonomy representation
of π
1
(Y ) and is therefore topologically a product U (1) × · · · × U(1) where
the number of factors equals b
1
(Y ). In particular if b
1
(Y ) = 0 then the
186 YUHAN LIM
reducible is exactly one p oint. (Note: When b
1
(Y ) = 0, L admits only one
flat connection, up to gauge.)
Lemma 7. The equation F
A
= ω has a solution if and only if the real coho-
mology classes [F
A
] = [ω] or equivalently [
i
2π
ω] = c
1
(L)
R
. If the latter holds,
then the space of equivalence classses of reducible solutions is topologically
U(1) × · · · × U(1) where the number of factors equals b
1
(Y ), and in the case
b
1
(Y ) = 0, a single point.
Proof. As explained above the c ondition [F
A
] = [ω] is necessary. For suffi-
ciency, let A
0
be such that [F
A
0
] = [ω]. Then we only need to solve for a in
F
A
0
+a
= ω which is equivalent to da = ω F
A
0
. Since ω F
A
0
is exact such
an a can be found. Assuming solutions exist, let F denote the space of all
A’s such that F
A
= 0. Then F + a describes all the solutions to F
A
= ω.
Therefore the space of reducible solutions up to gauge are topologically the
same as in the case ω = 0.
2.4. Orientation.
Suppose Z
π
(P ) consists only of regular points. It is clear that Z
π
(P )
is orientable. We want to produce a procedure for inducing a global ori-
entation. The fundamental elliptic complex can be combined into a single
operator L
η,A,Φ
:
0
2
(iR)
1
2
(iR) Γ
2
(S)
0
1
(iR)
1
1
(iR) Γ
1
(S),
L
η,A,Φ
= δ
1
η,A,Φ
+ δ
0
η,A,Φ
. A direct computation verifies that this operator is
formally self-adjoint.
Let Λ = det Ind{L
η,A,Φ
}. Let g G and (a, φ) ker L
η,A,Φ
. Then
(a, g
1
φ) ker L
η,g(A,Φ)
. Therefore the action of G lifts to an action on Λ.
Note that if Φ = 0 then the stabilizer U(1) maps the fibre of Λ at A = (A, 0)
back to itself by the identity. Hence Λ decends to a line bundle
ˆ
Λ over
0
2
(iR) × B(P ).
Proposition 8. The real line bundle
ˆ
Λ is trivial.
Proof. We need to show that Λ posseses a G-equivariant trivialization. The
substitution of (1 ε)Φ, 0 ε 1, for Φ and (1 ε)η for η in the definition
of δ
1
η,A,Φ
and δ
0
η,A,Φ
defines a homotopy of L
η,A,Φ
to an operator L
0
η,A,Φ
given
by L
0
η,A,Φ
(ξ, a, φ) = (da + 2, D
A
φ). This homotopy is G-equivariant. We
have ker L
0
η,A,Φ
= H
0
(iR) H
1
(iR) H
A
= coker L
0
η,A,Φ
. This family has
a trivial determinant, and this proves Λ is G-equivariantly trivial.
Notice that the homotopy given in the proof is the identity over {0} ×
C
Red
(P ). Thus over this set det Ind{L
η,A,Φ
} is the determinant of the index
of a constant family {L
dRham
} tensored with the complex family {D
A
}.
Since a complex family is canonically oriented, we may ignore it. The kernel
and cokernel of L
dRham
are H
0
(iR) H
1
(iR) and by identifying them with
each other we obtain a trivialization of det Ind{L
η,A,Φ
} over {0} × C
Red
(P ).
SEIBERG-WITTEN INVARIANTS ... 187
This orients det Ind{L
η,A,Φ
} over all of
0
2
(iR) × C(P ). This is the natural
orientation of
ˆ
Λ.
Consider the trivial real line bundle R over
0
2
(iR) × B
(P ). Then
ˆ
Λ has
the property that over the open set O of
0
2
(iR) × B
(P ) defined by the
condition that ker L
η,A,Φ
= 0, there is a canonical isomophism h :
ˆ
Λ|
O
=
R|
O
.
Proposition 9. Let
ˆ
Λ have the natural orientation described above. As-
sume Z
π
(P ) consists only of regular points. Then the following rule defines
an orientation ε : Z
π
(P ) 1}. Let x Z
π
(P ). Denote by o(x, R) the
canonical orientation of R|
x
and o(x,
ˆ
Λ) the orientation induced by
ˆ
Λ via
the isomorphism h above. Then
ε(x) =
1 if o(x, R) = o(x,
ˆ
Λ)
1 otherwise.
3. Generic Properties.
Let {g(t)}, t I
ε
= (ε, 1 + ε) be a 1-parameter family of metrics on Y . In
this section we examine the parameters (π, t) in P × I
ε
for which the moduli
space Z
π
(P ; g(t)) consists solely of regular points.
In order to understand how the geometric structures and operators
changes with the 1-parameter family of metrics it will be useful to b e able to
work with a single reference underlying metric, spin-c structure and model
for the spinors. Fix an underlying metric which we take to be g, let P
SO
be
the corresponding oriented orthonormal frame bundle and h an automor-
phism of T Y . h induces an automorphism h
of P
GL
+
, the component of
positively oriented frames of the frame bundle of Y . The image of P
SO
in
P
GL
+
under h describes the orthonormal frame bundle of another metric.
Conversely, the positive orthonormal frame bundle of any other metric can
be recovered on this way. Call the second metric g
0
.
h can be lifted to an isomorphism between P , the spin-c structure for g
and P
0
, the spin-c structure for g
0
. This, in turn, induces a fibrewise isometry
ˆ
h between the corresponding spinor bundles S and S
0
. By changing
ˆ
h to
e
u
ˆ
h where u is a smooth function on Y we can arrange it so that e
u
ˆ
h gives
an isometry between Γ(S) and Γ(S
0
) with respect to their L
2
-norms.
Given the 1-parameter family g(t) the construction of e
u
ˆ
h above can be
carried out smoothly in the parameter t, taking for instance g = g(0) to be
the reference metric. Therefore using these isomorphisms as identifications
we may assume (A, Φ) etc. for every t is defined on a fixed reference bundle.
The Dirac operator now depends also on t, and we denote this as D
g(t)
A
. It
is always self-adjoint with respect to the reference spinor bundle. Further
information regarding the relation between the Dirac operator for different
metrics can be found in [B], [BG] and [H].
188 YUHAN LIM
3.1. Singular Locus of Dirac Operators.
In this section we discuss the singular locus for certain families of Dirac
operators, i.e., the parameters for which the Dirac operator is singular. This
will be crucial for us later.
Suppose b
1
(Y ) = 0. Since H
2
(Y ; iR) = 0, we have a bounded right
inverse d
1
:
k
1
2
(iR) for the op erator d. Let θ be a fixed C
flat
connection on the determinant L. Then for any given ω, A = θ + d
1
(ω)
solves F
A
= ω. Define {D(α, ω, t)} to be the family of Dirac operators
D : P × I
ε
Fred
0
(Γ(S)), D(α, ω, t) = D
g(t)
θ+α+d
1
(ω)
.
Here Fred
0
denotes the Banach space of Fredholm operators of index zero.
In the case b
1
(Y ) = 1, we shall also define a family as follows: This time
we keep the metric fixed, so we drop it from the notation. Let A
0
be a fixed
C
connection on L and denote by ω
0
its curvature. Fix a choice of non-zero
a
0
H
1
(iR) such that
i
4π
a
0
defines a generator for H
1
(Z). (The choice of
constants here is so that a
0
is the class of a gauge change 2g
1
dg.) Then the
set {A
0
+ ta
0
| t [0, 1)}, parametrizes all the reducible SW
α,ω
0
solutions
up to gauge equivalence. Let E
k
, k 2, denote the exact forms in
k
.
Then on E
k
we can as before define a bounded inverse d
1
: E
k
1
2
(iR).
Given ω E
k
then A = A
0
+ ta
0
+ d
1
(ω) solves F
A
= ω
0
. Thus the set
{A
0
+ ta
0
+ d
1
(ω
0
) | t [0, 1)} parameterizes up to gauge equivalence all
the reducible SW
α,ω+ω
0
-solutions, ω E
k
. Define the family {D(α, ω, t)} by
D : Q × E
k
× I
ε
Fred
0
(Γ(S)), D(α, ω, t) = D
A
0
+ta
0
+α+d
1
(ω)
.
Proposition 10. Let N be the subset of P ×I
ε
consisting of all (α, ω, t) for
which D(α, ω, t) is singular. Similiarly define the subset K of Q × E
k
× I
ε
for D(α, ω, t). Then N and K are nowhere dense closed subspaces.
Proof. We prove only the case for N . The other is done similiarly. Let B
be the unit L
2
-ball in Γ
2
(S). Let V P × I
ε
× B be the vector bundle
whose fibre at (π, φ) is the real L
2
-orthogonal to φ in Γ
1
(S). By evaluating
D(α, ω, t) on φ B we obtain a section, call it D, of V . We claim this section
is transverse to the zero section. Let D(α
0
, ω
0
, t
0
)φ
0
= 0. Let ψ V
α
0
0
,t
0
0
be L
2
-orthogonal to the derivative dD at (α
0
, ω
0
, t
0
, φ
0
). By varying φ
0
in
the tangent direction δφ we find dD(δφ) = D(α
0
, ω
0
, t
0
)δφ; thus ψ must
also satisfy D(α
0
, ω
0
, t
0
)ψ = 0 (since D(α, ω, t) is self-adjoint). On the other
hand, by varying α
0
, dD(δα) = δα ·
t
0
φ
0
and if ψ is L
2
-pependicular to this
then ψ = ifφ
0
for some real function f. The condition D(α
0
, ω
0
, t
0
)ψ = 0
then leads to df ·
t
0
φ = 0; but since φ
0
(x) 6= 0 on an open set it must be
that df = 0, and so f is a constant. Finally ψ being in V is necessarily
L
2
-orthogonal to φ
0
; thus f = 0. Hence ψ = 0 and transversality holds
and the zeros of D defines a smooth infinite dimensional submanifold M of
P × I
ε
× B. The projection map p : M P × I
ε
is proper since the kernel
SEIBERG-WITTEN INVARIANTS ... 189
of the Dirac operator is always finite dimensional. Applying the Sard-Smale
theorem we can concude that there exists an open dense set O in P ×I
ε
with
the property that D(α, ω, t), (α, ω, t) O has nullity 1 over the reals. But
since this is a complex linear operator and self-adjoint, D(α, ω, t) must be
non-singular. The proposition now follows.
Let π
i
= (α
i
, ω
i
), i = 0, 1, be given perturbations. Denote by {π(t)} =
{(α(t), ω(t))}, t I
ε
the 1-parameter family of pe rturbations defined by
π(t) = (1 t)π
0
+
1
. For a fixed value of π, D(π(t) + π, t) defines a 1-
parameter family of Dirac operators. We use the notation {D
π
(t)} for this 1-
parameter family. We call this family transverse (for the choice of π) if D
π
(0)
and D
π
(1) are non-singular and the family has transverse spectral flow, as
t varies over [0, 1]. Transverse spectral also includes the condition that
multiple zero-eigenvalues do not occur as t varies. For the case b
1
(Y ) = 1,
fix a value of α
0
. In a similar way we have a 1-parameter family {D
π
(t)}
obtained by considering D((α
0
, ω
0
) + π, t) for a fixed π. Transversality is
defined in the same way as before.
Proposition 11. Suppose {D
0
(t)} has the property that D
0
(t) is non-singu-
lar for t = 0, 1. Then there are arbitarily small π such that {D
π
(t)} is a
transverse family. A similiar statement holds for {D
π
(t)}.
Proof. We shall only prove the case of {D
0
(t)}. The other case is handled
by ess entially the same argument. Let x = (π
0
, t
0
) P × I, I = (0, 1).
Consider the map
G : Γ
2
(S) × P × I Γ
2
(S), G(φ, π, t) = D(π, t)φ.
The differential of G at (0, π
0
, t
0
) is given by
dG(δφ, δω, δt) = D(π
0
, t
0
)δφ.
Then ker dG = H
t
0
P R and coker dG = H
t
0
. Here H
t
0
denotes the kernel
of D(π(t
0
)+π
0
, t
0
) (acting on Γ
2
(S)). By the implicit function theorem there
is a neighbourhood V of (0, x) H
t
0
× P × I and a unique smooth map
f : V H
t
0
such that for (φ, π, t) V ,
(I Π)G(φ + f(φ, π, t), π, t) = 0, Π = L
2
-projection onto H
t
0
.(1)
Note that the linear extension of f in the φ variable continues to satisfy (1)
so we may take V to be of the form H
t
0
× W, W a neighbourhood of x.
Because of (1) the injective/surjective properties of D(π, t), (π, t) W , are
completely determined by the finite-dimensional operator finite dimensional
operator H
t
0
H
t
0
,
T (π, t)φ = ΠG(φ + f(φ, π, t), π, t).
We claim T (π, t) is self-adjoint with respect to the (real C-invariant) L
2
-
inner product on H
t
0
. We introduce the notation , ·i
H
t
0
, ψ, φ H
t
0
, to
190 YUHAN LIM
denote the L
2
inner product on H
t
0
and , ·i
Γ(S)
the L
2
-inner product on
Γ(S). We compute:
hT (π, t)φ, ψi
H
t
0
= hG(ψ + f(ψ, π, t), π, t), φi
Γ(S)
= hG(ψ + f(ψ, π, t), π, t), φ + f(φ, π, t)i
Γ(S)
since G(ψ + f(ψ, π, t), π, t) H
t
0
, f(φ, π, t) H
t
0
= hψ + f(ψ, π, t), G(φ + f(φ, π, t), π, t)i
Γ(S)
by self-adjointness of D(π, t)
= hψ, T (π, t)φi
H
t
0
.
This proves the claim.
Since T (π, t) is complex linear and self-adjoint with respect to the real
inner product, it is Hermitian with resp ec t to the natural complex extension
of the L
2
-inner product on H
t
0
. Let Herm(H
t
0
) denote the (real) vector
space of Hermitian transformations on H
t
0
. The determinant function det :
Herm(H
t
0
) R C and det
1
(0) is a closed subvariety of codimension 1
in Herm(H
t
0
). Introduce the notation
N
(k)
= {l Herm(H
t
0
) | dim ker(l) k}.
Lemma 12. Suppose dim
C
(H
t
0
) > 0. The derivative (dT )
π
0
,t
0
|
Q
of T re-
stricted to Q has non-trivial image in Herm(H
t
0
). If dim
C
H
t
0
2 then this
image is of dimension 2.
First let us show that the image of (dT )
π
0
,t
0
is non-trivial. The derivative
at (π
0
, t
0
) is computed to be
dT
π
0
,t
0
(δ α, δt)φ = Π(δα + d
1
δω) ·
t
0
φ, π = (α, ω).
Suppose that hdT
π
0
,t
0
(δα)φ, φi
L
2
= 0 for all δα Q. Since
hδα ·
t
0
φ, φi
L
2
=
Z
b
Y
hδα, σ
t
0
(φ, φ)i.
This implies σ
t
0
(φ(y), φ(y)) = 0 for all y Y and thus φ = 0. Thus the
image of dT
π
0
,t
0
is non-trivial.
Assume now that dim
C
(H
t
0
) 2. Let δα be such that dT
π
0
,t
0
(δα) 6= 0.
Let φ
1
, . . . , φ
n
be an complex orthonormal basis for Herm(H
t
0
) such that
dT
π
0
,t
0
(δα) is diagonal with respect to this basis. Thus hδα · φ
i
, φ
j
i
L
2
,C
= 0
for i 6= j. Since the φ
k
are harmonic spinors, unique c ontinuation implies
that there is a open set in Y on which φ
i
6= φ
j
, i 6= j on this open set, in
particular say at the point y Y . We can find a δα
0
with support in an
arbitarily sm all neighbourhood of y such that
R
b
Y
hδα
0
· φ
i
, φ
j
i
C
6= 0, i 6= j.
Thus dT
π
0
,t
0
(δα
0
) is independent of dT
π
0
,t
0
(δα). This shows that the image
is at least 2-dimensional. This proves the Lemma.
SEIBERG-WITTEN INVARIANTS ... 191
Consider for each (α, ω) W,
τ
α,ω
(t) = T (α, ω, t), |t t
0
| < ε
where ε > 0 is chosen so that (α, ω, t) W . Clearly τ
α,ω
defines a path in
Herm(H
t
0
).
By an open cover argument the following exists: (1) a finite set {t
1
, . . . , t
n
}
[0, 1] together with open neighbourhoods V
i
of t
i
in R such that {V
i
}
n
i
covers [0, 1] (2) an open neighbourhood W
0
W of 0 P (3) maps
T
i
: W
0
× I
i
Herm(H
t
0
) as in the preceding which preserves the in-
jectivity/surjectivity properties of D(α, ω, t), (α, ω, t) W
0
× I
i
. We denote
the corresponding paths by τ
i
α,ω
(t).
Let N = max
i
{dim
C
(H
t
i
)}. Suppose N > 1. Let j be such that
dim(H
t
j
) = N. Note that N
(N)
= {0} Herm(H
t
j
). Thus any non-
zero element in Herm(H
t
0
) lies in N
(k)
, 0 k < N . Then by the above
Lemma we can find a sufficiently small perturbation (α, 0) W
0
so that
τ
i
α,0
(t) N
(k
0
)
, 0 k
0
< N , t I
i
and for i 6= j, τ
i
α,0
(t) Herm(H
t
i
)
for all t I
i
. Thus we establish that there is an arbitarily small α so that
τ
i
α,0
(t) N
(k)
with 0 k < N for every i. Repeating the above con-
struction over but with W
0
taken to be an open neighbourhood of (α, 0)
instead, we inductively prove that we can find an arbitarily small α
0
so that
we have N = 1. The perturbation argument in this case makes each path
τ
i
α
0
,0
transverse to N
(1)
= {0} Herm(H
t
0
)
=
R.
Let us show that {D
α
0
,0
(t)} is a transverse family. Suppose at s, τ
i
α
0
,0
(s) =
0. Let φ H
t
i
be unit length. Let λ(t) be the 1-parameter family of
eigenvalues satisfying D
α
0
,0
(t)φ = λ(t)φ for t close to s. Thus we have
hD
α
0
,0
(t)φ, φi
L
2
= λ(t)hφ, φi
L
2
.
Differentiating this equation with respect to t and evaluating at t = s gives
hdT
π
0
,t
0
(0, 0, 1)φ, φi
L
2
= λ
0
(s).
The left hand term is simply the velocity of τ
i
α
0
,0
at t = s and transversality
means this is non-zero. Thus λ
0
(s) 6= 0 and we have transverse spectral flow
at t = s.
3.2. The Parameterized Moduli Space.
As before we assume the 1-parameter families {g(t)}, t I
ε
. Define the
parametrized Seiberg-Witten section to be the map
˜s :
0
2
(iR) × C(P ) × P × I
ε
1
1
(iR) Γ
1
(S),
˜s(η, A, Φ, π, t) = s
π,g(t)
(η, A, Φ).
Then the parameterized moduli space is Z(P ) = ˜s
1
(0)/G (with G acting
only on the C(P ) factor). There is the projection map p : Z(P ) P × I
ε
and clearly p
1
(π, t) = Z
π
(P ; g(t)).
192 YUHAN LIM
Proposition 13. The irreducible part Z
(P ) of the parameterized moduli
space is smooth Hilbert space manifold and the projection map p|
Z
(P )
:
Z
(P ) P × I
ε
is a smooth Fredhom map of index zero.
If b
1
(Y ) = 0 we determined in Prop. 10 that p
1
(π) contains no reducibles
if we are off the set N P × I
ε
. If b
1
(Y ) > 0 we observed in Lemma 7 that
there are no reducibles in p
1
(α, ω) if and only if [
i
π
ω] is not an integral
class in cohomology. Thus:
Lemma 14. Let q : P
k
× I
ε
H
2
(Y ; R), q(α, ω, t) = [
i
2π
ω] and set W =
q
1
(c
1
(L
R
)). Then p
1
(π, t) has reducibles if and only if (π, t) W. W is
a closed nowhere dense subset of of codimension equal to b
1
(Y ).
We remark that in the case b
1
(Y ) > 0 the condition of regularity (i.e.,
H
2
A,Φ
= {0}) for reducible solutions can never be satisfied. This is because
when Φ = 0, H
2
A,Φ
reduces to H
2
(iR)H
A
. So regularity implies the absence
of reducibles in this case.
Corollary 15. p|
Z
(P )
: Z
(P ) P × I
ε
is proper over P × I
ε
\W where
(i) W = N if b
1
(Y ) = 0 (ii) W = W if b
1
(Y ) > 0. Therefore for an open
dense set O P × I
ε
, p
1
(z), z O, is a finite set of regular points.
Proof. The properness assertion is the content of Proposition 6 and the
fact that a regular reducible point in the case b
1
(Y ) = 0 is necessarily
isolated, by the Kuranishi local model. The Sard-Smale theorem then gives
the ‘open dense set’ statement since regularity of an irreducible solution
(A, Φ) is equivalent to the derivative d˜s at (0, A, Φ) being surjective. (Note:
without the properness assertion we can only conclude regularity on a Baire
set.)
Proof of Proposition 13. We have to show that the derivative d˜s is surjective
at every point (0, A
0
, Φ
0
, π
0
, t
0
) ˜s
1
(0) for which Φ
0
6= 0. Let (b, ψ) lie in
the cokernel of ds
0,A
0
,Φ
0
, i.e., (b, ψ) H
1
A,Φ
, thus
(i) db =
1
2
σ, ψ), (ii) D
A
ψ +
b
2
· Φ = 0, (iii) 2d
b = ihiΦ, ψi.(2)
Suppose (b, ψ) is L
2
-orthogonal to the image of d˜s. The Proposition is proven
as soon as we can show (b, ψ) = 0. If δω
k
, then d˜s(δω) = (δω, 0). Thus
b must be L
2
orthogonal to all the co-closed forms; this implies that b must
be closed. Then from (i) we obtain the condition σ
t
0
0
, ψ) = 0. Working
at a point, the kernel of the transformation v 7→ σ
t
0
(w, v) is of dimension 1
and it is easy to check that σ
t
0
(w, iw) = 0. Therefore ψ = ifΦ
0
for some
real valued function f. Putting this into (ii) of (2) we obtain
0 = D
g(t
0
)
A
0
(ifΦ) +
b
2
·
t
0
Φ
= (idf +
b
2
) · Φ (since D
g(t
0
)
A
0
Φ
0
= 0).
SEIBERG-WITTEN INVARIANTS ... 193
Hence we obtain the pointwise condition idf +
b
2
= 0 on the open dense set
O where Φ
0
6= 0. By continuity it holds on all of Y . Substituting into (iii)
of (2) we get the equation
4∆f = −|Φ|
2
f.
Taking the pro duct with f and integrating we obtain:
Z
Y
4|df|
2
+ |Φ|
2
|f|
2
= 0.
Thus f = 0 on O and therefore Y and we finally obtain (b, ψ) = 0. Finally
the index zero assertion follows directly from Lemma 4.
4. Proof of Theorem 1.
Y is assumed to be a closed oriented 3-manifold with Riemannian metric
g and spin-c structure P Y . According to Corollary 15 applied to the
constant family {g(t) = g}, we may choose a perturbation π from an open
dense set in P
k
(k 3) such that Z
π,g
(P ) consists of a finite set of regular
points, i.e., the cohomology H
2
A,φ
is trivial at these points. For this π, if
b
1
(Y ) = 0 there is a unique isolated reducible (up to gauge equivalence) and
if b
1
(Y ) > 0 there are no reducibles. Z
π,g
(P ) is then naturally oriented by
our conventions (Proposition 9) and we can form the algebraic sum
#Z
π,g
(P ).
To prove the claimed invariance properties of #Z
π,g
(P ), let g
0
, g
1
be two
metrics on Y and let π
i
be two perturbations which satisfy the above with
respect to g
i
. We want to relate #Z
π
0
,g
0
(P ) and #Z
π
1
,g
1
(P ). Consider
the 1-parameter family of metrics {g(t)} = {(1 t)g
0
+ tg
1
} defined for
t I
ε
= (ε, 1 + ε). Thus as in Section 3 we have a parameterized moduli
space Z(P ) and projection map p : Z(P ) P
k
× I
ε
.
We consider a smooth path σ : [0, 1] P
k
× I
ε
, σ(0) = (π
0
, g
0
), σ(1) =
(π
1
, g
1
). We introduce the notation Z
σ
(P ) for the σ-parametrized mo duli
space {(x, t) | x p
1
(σ(t))}. If a portion of σ misses the ‘singular’ sets N
or W of Sec.3.1 and is transverse p, then that portion of Z
σ
(P ) consists
purely of regular points and therefore is a smooth arc. This is oriented in
the following way. The local deformation theory of Z
σ
(P ) is described by
an elliptic complex of the form
0
3
(iR)
0
2
(iR)
1
2
(iR) Γ
2
(S) R
1
1
(iR) Γ
1
(S).
Therefore the orientation is determined by looking at the ‘wrapped up’ op-
erator
L
η,A,Φ,t
:
0
2
(iR)
1
2
(iR) Γ
2
(S) R
0
1
(iR)
1
1
(iR) Γ
1
(S).
194 YUHAN LIM
The orientation of the regular irreducible points of Z
σ
(P ) is determined by
an orientation of the determinant of the index of the family {L
η,A,Φ,t
}. We
have the short exact sequence of 2-step complexes:
0
0
2
(iR)
1
2
(iR) Γ
2
(S)
L
η,A,Φ
0
1
(iR)
1
1
(iR) Γ
1
(S) 0
?
?
y
?
?
y
0
0
2
(iR)
1
2
(iR) Γ
2
(S) R
L
η,A,Φ,t
0
1
(iR)
1
1
(iR) Γ
1
(S) 0
?
?
y
?
?
y
0 R 0 0.
This gives rise to a canonical isomorphism
h : ker L
η,A,Φ,t
coker L
η,A,Φ
ker L
η,A,Φ
R coker L
η,A,Φ,t
.(3)
An orientation for det Ind{L
η,A,Φ
} defines an orientation for det Ind{L
η,A,Φ,t
}
according to this rule: Choose an orientation for coker L
η,A,Φ
. Then an ori-
entation of ker L
η,A,Φ
is determined, since det Ind{L
η,A,Φ
} is oriented. Now
given an orientation of coker L
η,A,Φ,t
, then ker L
η,A,Φ,t
is oriented so that h is
an orientation-preserving isomorphism, where the domain and range spaces
are given the product orientation in the order written in (3). With this ori-
entation convention, if Z
σ
consists entirely of regular irreducible points and
if compact then its boundary is precisely Z
σ(1)
(P ) Z
σ(0)
(P ), as oriented
spaces.
The proof of Theorem 1 in the case b
1
(Y ) > 1 can now be eas ily estab-
lished. By Lemma 14 σ may be chosen to be disjoint from the subset of W
for which Z
σ
(P ) has reducibles. Furthermore σ can be assume to be trans-
verse to the projection p : Z(P ) P
k
× I
ε
. Thus Z
σ
(P ) defines a smooth
compact oriented cobordism betwee n Z
π
0
,g
0
(P ) and Z
π
1
,g
1
(P ). This proves
the invariance of #Z
π,g
(P ) in this case.
This argument extends to the cases b
1
(Y ) = 0, 1 provided (π
0
, 0) and
(π
1
, 1) can be connected by a path which missed the ‘bad’ sets N , W of
Sect. 3.1, Lemmma 14 respectively. However this is not generally true, as
we shall describe below.
4.1. The case b
1
(Y ) = 0.
The argument in the case b
1
(Y ) > 1 may fail here due to the presence of
a reducible (unique up to gauge) solution in each Z
σ(t)
(P ). The reducible
stratum of Z
σ
(P ) is an arc which under p projects diffeomorphically onto I
ε
.
The path σ may meet the subset N of Prop. 10 and singularities may occur
in Z
σ
. Choose σ to be the path defined by the family {g(t)} and the family of
perturbations {π(t)} = {(t1)π
0
+
1
}, t I
ε
. Fix θ a flat connection on L.
Writing π(t) = (α(t), ω(t)), in the notation of Sec. 3.1, the reducible solution
up to equivalence in Z
σ(t)
(P ) is given by θ(t) = θ + d
1
(ω(t)). Furthermore
the associated family of Dirac operators {D
0
(t)} determine the cohomology
SEIBERG-WITTEN INVARIANTS ... 195
group H
2
θ(t)
. According to Prop. 11, by an arbitarily small perturbation π we
can make this a transverse family {D
π
(t)}. This corresponds to deformation
of σ as σ + π which in turn is induced by perturbations (π
0
+ π, 0) and
(π
1
+ π, 0) of the end-points of σ. If π is sufficiently small then #Z
π
i
+π
(P )
coincides with #Z
π
i
(P ); thus without loss we may absorb π and simply
assume transversality for π = 0. Then spectral flow for {D(t)} = {D
0
(t)}
occurs at exactly the values of t where σ(t) meets N .
Proposition 16. Assume {D(t)} is a transverse family and σ is transverse
to the projection p : Z
σ
(P ) P
k
×I
ε
away from N . Let σ N = {σ(t
i
)}
n
i=1
.
Then for each t
i
there is a open neighbourhood N
i
of θ(t
i
) such that:
(i) Z
σ
(P )\ N
i
is a smooth compact 1-manifold with boundary
(ii) Z
σ
(P ) N
i
is diffeomorphic to the zeros of the map R × R
+
R,
(t, ξ) 7→
(iii) Z
σ+π
(P ) N
i
= 0 × R
+
and the orientation of 0 × R
+
is ε
i
ξ
where
ε
i
is the sign of the spectral flow of {D(t)} at t
i
.
An immediate consequence of this is the formula
#Z
σ(1)
(P ) #Z
σ(0)
(P ) = SF{D(t)}
where ‘SF’ on the right denotes the total spectral flow as t varies from 0 to
1. Notice that the left-hand term is actually independent of choice of path,
therefore the spectral-flow term only depends on the end-points of the path.
(In fact, it is possible to verify this directly as well.)
To define an invariant in this case it is necessary to intro duce a counter-
term which should be a function of ω and g which has the same change
as #Z
π,g
(P ) as we cross from one connected component of P
k
× I
ε
\N to
another. Such a function can be obtained from the spectral invariants of
[APS].
Proposition 17. Assume b
1
(Y ) = 0. Let (α, ω, g) be given. Let θ be the
unique (up to gauge) flat connection on L and let a be defined by the condi-
tion d
a = 0, da = ω. Define
ζ(α, ω, g) =
1
8
η (d d|
even
, g)
+
1
2
dim
C
ker D
g
θ+a+α
+ η
D
g
θ+a+α

+
1
32π
2
Z
Y
(a + α) d(a + α)
where η denotes the Atiyah-Patodi-Singer spectral invariant of the associated
operator. Then:
(i) ζ(ω, g) lies in Z
h
1
8|H
1
(Y ;Z)|
i
; if in addition H
1
(Y, Z) = 0 it lies in Z
(ii) given the path σ as Prop. 16, we have ζ(σ(1)) ζ(σ(0)) = SF {D(t)}.
196 YUHAN LIM
Thus we see that the combination
#Z
π,g
(P ) + ζ(π, g)
defines a topological invariant in the case b
1
(Y ) = 0.
Proof of Proposition 17. Every spin-c structure on Y is obtained by ten-
soring a spin structure on Y with a complex line bundle. By a Theorem
of Milnor every spin Y is the oriented spin boundary of an oriented spin
4-manifold X with b
1
(X) = 0. Every complex line bundle over Y can be
extended over X; therefore we may assume a spin-c structure P
0
X which
induces the given P Y . We may also assume X to have a metric which
near the boundary which is a product Y ×[0, ε) of an interval and the metric
on Y and with orientation dy dt.
We may extend the connection θ ove r L(P
0
) = det(P
0
) as the connection
Θ, a as ˆa and α as ˆα over X. These extensions can be taken to be products
over Y × [0, ε). The index theorem of [APS] applied to the Dirac operator
D
g
Θ+ˆa+ˆα
over X associated to P
0
gives:
Index D
g
Θ+ˆa+ˆα
=
Z
X
exp
1
2
c
1
+ ˆa + ˆα)
ˆ
A
1
2
dim
C
ker D
g
θ+a+α
+ η
D
g
θ+a+α

.
Here
c
1
+ ˆa + ˆα) =
i
2π
F
Θ+ˆa+ˆα
and
ˆ
A is the
ˆ
A-polynomial in the Pontrjagin classes. On the other hand
consider the signature operator on X. This has index
sig(X) =
Z
X
L η(d d|
even
, g).
L is the Hirzebruch L-polynomial in the Pontrjagin classes. Since
exp
1
2
c
1
(B)
= 1 +
1
2
c
1
(B) +
1
8
c
1
(B) c
1
(B) + . . . ,
ˆ
A = 1
1
24
p
1
+ . . . ,
L = 1 +
1
3
p
1
+ . . . ,
the above index formulas give
1
8
η(d d|
even
, g) +
1
2
dim
C
ker D
g
θ+a
+ η(D
g
θ+a
)
+
1
32π
2
Z
Y
(a + α) d(a + α)
SEIBERG-WITTEN INVARIANTS ... 197
=
1
8
Z
X
c
1
(Θ) c
1
(Θ)
1
8
sig(X) Index D
g
Θ+ˆa+ˆα
.
If Y is an integral homology s phere then L is trivial and we may choose
its extension over X as the trivial bundle; therefore Θ in this case may be
assumed trivial. Furthermore the intersection form on X is then unimodular
so sig(X) is divisible by 8. Thus ζ is an integer. When Y is not an inte-
gral homology sphere then the term
R
X
c
1
(Θ) c
1
(Θ) depends only on the
topological type of the extension of L over X. It can be identified with the
Z
h
1
|H
1
(Y ;Z)|
i
-intersection product of the class [c
1
(Θ)] H
2
(X; Z)/torsion
with itself. Therefore ζ takes values in Z
h
1
8|H
1
(Y ;Z)|
i
.
The term η(dd|
even
, g) depends continuously on g whereas according
to [APS]
1
2
η(D
g
θ+a+α
) jumps by the spectral flow. Thus ζ has the correct
behaviour as we cross components of P
k
× I
ε
\N , as claimed.
Proof of Proposition 16. The proof of the proposition relies on a detailed
understanding of the Kuranishi local mo del at the singular points [θ(t
i
)].
Without loss of generality we assume that there is only one value t = t
1
where {D(t)} is singular. The Seiberg-Witten section which gives Z
σ
(P ) is
of the form ˆs :
0
2
(iR) × C(P ) × R
1
1
(iR) Γ
1
(S),
ˆs(η, A, Φ, t) =
t
(F
A
1
4
σ
t
, Φ) ω(t)), D
g(t)
A+α(t)
Φ ηΦ
where the t in the notation denotes a dependence on t. (Note: just as in
Sec. 3 we work with a fixed P Y with respect to a basepoint metric.)
The linearization of ˆs at η = 0, A = θ(t
1
), Φ = 0, t = t
1
is given by
dˆs(δη, δa, δφ, δt) =
t
1
(d(δa) + ω
0
(t
1
)δt), D(t
1
)δφ
.
Let X
θ(t
1
),0
be the slice of the gauge group action on C(P ) at (0, θ(t
1
), 0)
(Sec. 2). Then
ker (dˆs) (Ω
0
2
(iR) × X
θ(t
1
),0
× R) = H
0
(iR) H
θ(t
1
)
R
t
coker (dˆs) = H
θ(t
1
)
.
Here R
t
= span{(d
1
(ω
0
(t
1
)), 1)} in the
1
2
(iR) R factor. The Kuranishi
obstruction map then takes the form
Ξ : iR × H
θ(t
1
)
× R
t
U H
θ(t
1
)
.
This gives ˆs
1
(0)/G near (0, θ(t
1
), 0, t
1
) as Ξ
1
(0)/S
1
. A direct verification
shows that Ξ
1
(0) (iR × 0 × R
t
) U . The subset (iR × 0 × 0) U consists
of ‘virtual’ Seiberg-Witten solutions and thus should be ignored to get the
Seiberg-Witten moduli space proper (Lemma 3). The subset (0×0×R
t
)U
are the reducible solutions near (0, θ(t
1
), 0, t
1
). Our assumption on σ being
transverse to p away from W means that the closure of the irreducible part
198 YUHAN LIM
of Ξ
1
(0)/S
1
, is a compact 1-manifold with boundary except possibly at
(0, 0, 0). Furthermore Ξ
1
(0) U (R × 0 × R
t
(0, 0, 0)) = .
By construction the derivative of Ξ at (0, 0, 0) is the zero map. We aim
to compute the second derivative: This will give us the quadratic approxi-
mation to Ξ which will be sufficient for our purposes. In the following, we
identify R
t
with R via t 7→ (d
1
(ω
0
(t
1
)), 1)t.
Claim 18. The second derivative of Ξ at (0, 0, 0) is given by
D
2
Ξ(δη, δφ, δt) = cδtδφ δηδφ
where c is a non-zero real constant and has the same sign as that of the
spectral flow of {D(t)} at t = t
1
.
To prove the claim: the obstruction map Ξ is constructed as a c omposition
of the form x 7→ Π ˆs(x + f(x)) where x O iR × H
θ(t
1
)
× R
t
and
f : O (iR × H
θ(t
1
)
× R
t
)
is given by the implicit function theorem. As
such its derivative at 0 is the zero map. It is then seen that D
2
Ξ is given
by Π D
2
ˆs. This is given by the expression
D
2
Ξ(δη, δφ, δt) = Π
D
0
(t
1
)(δφ)
δηδφ.
The map δφ 7→ Π(D
0
(t
1
)δφ) defines a He rmitian transformation on H
θ(t
1
)
=
C with respec t to the complex L
2
-inner product. Thus it is multiplication
by a real constant c = hD
0
(t)v, vi, v being of unit length. Our assumption
on σ was that at t = t
1
a single eigenvalue λ(t), |tt
1
| < δ, for D(t) changed
from negative to positive or vice-versa. In the first case the spectral flow is
+1 and in the latter 1. We have a 1-parameter family of unit eigenvectors
v(t), |t t
1
| < δ, such that
D(t)v(t) = λ(t)v(t).
Differentiating this equation at t = t
1
and taking the inner product with
v(t
1
) we obtain using self-adjointness of D(t),
hD
0
(t
1
)v(t
1
), v(t
1
)i = λ
0
(t
1
).
Thus the sign of the spectral flow is seen to be same as that of λ
0
(t
1
). This
proves the Claim.
To continue the proof of Proposition 16: by the Claim, Ξ(η, φ, t) is ap-
proximated up to second order by
(ct η)φ.
The zeros of the quadratic approximation fall into two branches: 0 × 0 × R
t
and 0 × H
θ(t
1
)
× 0. We claim that these two branches gives a complete
picture of the zeros of Ξ near (0, 0, 0). To get a clearer picture, consider
restricting the φ variable to the real span of a fixed non-zero vector in H
θ
0
.
Call this map
e
Ξ. Then Ξ
1
(0)/S
1
=
e
Ξ
1
(0)/ ± 1. Since
˜
Ξ(η, 0, t) = 0 we
may factor out the branch {φ = 0} by setting
˜
Ξ(η, φ, t) = φ.Θ(η, φ, t). Using
SEIBERG-WITTEN INVARIANTS ... 199
Claim 18 the linearization of Θ is seen to be dΘ(δη, δt, δφ) = (cδt δη). We
have ker dΘ = {δt = δη = 0}. Invoking the Implicit Function Theorem
we see that near (0, 0, 0), Θ
1
(0) is a smooth arc tangent to {φ = 0} at
(0, 0, 0). This dem onstrates the claimed local structure near the singular
point. For later we note the following: the implicit function theorem gives
a map G : H
θ(t
1
)
R
t
= {0} × R
t
iR × R
t
such that the closure of the
irreducible part of Ξ
1
(0) is given by graph(G) = {(0, φ, G(φ)) | φ H
θ(t
1
)
}.
What remains is to determine the orientation of the above arc of irre-
ducible solutions. We have an orientation of det Ind{L
η,A,Φ,t
} determined
by that of det Ind{L
η,A,Φ
} according to the map h of (3). (See discussion
following there.) Since we are working at a point where Φ = 0, the orienta-
tion of det Ind{L
η,A,Φ
} is determined by the kernel and cokernel of L
θ(t
1
),0
;
namely H
0
(iR)H
θ(t
1
)
. The long e xact sequence inducing h takes the form
0 H
0
(iR) H
θ(t
1
)
ker L
θ(t
1
),0,t
1
κ
R H
0
(iR) H
θ(t
1
)
β
coker L
θ(t
1
),0,t
1
0.
The map κ in the sequence sends the subspace R
t
isomorphically onto the
the target space. This isomorphism sends (1, d
1
(ω
0
(t
1
)))r 7→ r, r R.
Choose an orientation of coker L
θ(t
1
),0
= H
0
(iR) H
θ(t
1
)
; then s ince β in
the sequence is an isomorphism, our orientation convention dictates that
coker L
θ(t
1
),0,t
1
has the orientation induced by β, and the orientation on
ker L
θ(t
1
),0,t
1
is the product orientation ker L
θ(t
1
),0
R
t
, where R
t
is oriented
via κ.
In order to determined orientations in the local Kuranishi picture cor-
rectly we shall need to combine the obstruction map Ξ with a local slice
condition coming from the S
1
-action, which is the inverse of complex mul-
tiplication on the H
θ(t
1
)
factor. The set graph(G) Ξ
1
(0) represents
the closure of the S
1
-orbits of the irreducible solutions near (0, 0, 0). Let
0 6= v H
θ(t
1
)
. Then the linearization of the S
1
-action at (0, v, G(v)) is
a map iR iR × H
θ(t
1
)
× R
t
, δγ 7→ (0, (δγ)v, 0). The adjoint of this
map sends (δη, δφ, δt) 7→ ihiv, δφi. Therefore a further local description
for Ξ
1
(0)/S
1
near (0, v, G(v)) is the zeros of the map
χ : iR × H
θ(t
1
)
× R
t
iR × H
θ(t
1
)
, χ(η, φ, t) = (ihiv, φi, Ξ(η, φ, t)).
In what follows, we may for simplicity assume G = 0; the result for general
G is obtained by working sufficiently close to (0, 0, 0) where graph(G) is
approximated to arbitarily high order by 0 × H
θ(t
1
)
× 0. With this assumed,
the irreducible zeros of χ is the set of positive multiples of (0, v, 0). The
normal bundle to 0×H
θ(t
1
)
×0 at (0, 0, 0) is iR×0 ×R
t
. This is mapped via
dΘ isomorphically onto H
θ(t
1
)
. If we pull-back the complex orientation by
dΘ, then the induced orientation on iR×0×R
t
is ηδt. By continuity this
is carried to the point v as the same orientation. The last remaining direction
200 YUHAN LIM
to the normal bundle of χ
1
(0) at v is given by (0, iv, 0). This is mapped by
to (ihiv, ivi, 0) = (i, 0). Let us take the product orientation on iR ×
H
θ(t
1
)
for the range space of χ, in this order (the final answer is independent
of this choice); then our orientation convention dictates that the domain
space of χ is oriented in the order iR × H
θ(t
1
)
× R
t
. Then the pull-back
of the orientation on iR × H
θ(t
1
)
to the normal bundle of χ
1
(0) at v is
δθ η δt, where δθ is the (0, iv, 0) direction. Let δr be the direction
given by v, and εδr the induced orientation of χ
1
(0) near v. Then we
require that the orientation on χ
1
(0) followed by the orientation in the
normal direction equals the orientation on iR × H
θ(t
1
)
× R
t
, that is
εδr (δθ η δt) = δη δr δθ δt.
This shows the induced orientation on χ
1
(0) as r, as claimed.
4.2. The case b
1
(Y ) = 1.
This case is similiar but with some slight differences to b
1
(Y ) = 0. Any
path σ which connects (π
0
, g
0
) to (π
1
, g
1
) in P
k
× I
ε
may cross the codi-
mension 1 subset W. At the points where σ meets W the corresponding
Z
σ(t)
(P ) will admit an S
1
’s worth of reducibles, otherwise Z
σ(t)
(P ) contains
no reducibles. As following our notation conventions, σ(t) in components is
(π(t), g(t)) or (α(t), ω(t), g(t)).
We may by general transversality arguments assume that σ meets W
transversely and orthogonally and transverse to the projection p : Z
(P )
P
k
×I
ε
away from W. Let {t
i
} b e the finite set of values for which σ(t
i
) W.
To simplify matters even more, since W the preimage of a set in P
k
we may
assume near W that σ lies in the subset P
k
× {t
i
}. Hence for values near t
i
,
the metric represented by σ is unchanging.
We can always find connections A
i
such that F
A
i
= ω(t
i
). Using the
value α(t
i
), we can as in Sec. 3.1 form the 1-parameter family of operators
{D
i
π
i
(s)}. By Prop. 11 we can make this family transverse by an arbitarily
small perturbation π
i
. This perturbation can be achieved by a perturbation
of σ, supported for values of t near t
i
, and maintaining the original properties
of σ. Thus we can assume {D
i
0
(s)} is a transverse family and we drop the
‘0’ subscript notation. Finally let s
i,j
be the values of s for which {D
i
(s)}
has spectral flow. Denote by A
i,j
the connection A
i
+ s
i,j
a
0
.
A technical issue which will be significant is the orientation of the family
{D
i
(s)}. Looking back at the definition in Sec. 3.1 we see that this involved
a certain choice of a non-zero element a
0
in H
1
(iR). We shall make a
specific choice for each i. The assumption that σ meets W orthogonally
means in particular that the derivative ω
0
(t
i
) is L
2
-orthogonal to the exacts.
Thus d
(ω
0
(t
i
)) = 0 so ω
0
(t
i
) is closed. For a chosen i we now make the
convention that the a
0
should be a positive multiple of [ ω
0
(t
i
)].
SEIBERG-WITTEN INVARIANTS ... 201
Proposition 19. For each (i, j) there is a open neighbourhood N
i,j
of [A
i,j
]
such that (i) Z
σ
(P )\ N
i,j
is a smooth compact 1-manifold with boundary
(ii) N
i,j
is diffeomorphic to the zeros of the map R × R
+
R, (s, ξ) 7→
with Z
σ
(P ) N
i,j
= 0 × R
+
(iii) the orientation of 0 × R
+
is ε
i,j
ξ
where
ε
i,j
is the sign of the spectral flow of {D
i
(s)} at s
i,j
, as s varies from 0 to 1.
Proof. This largely proceeds in the manner of the cas e b
1
(Y ) = 0. We
continue to use notation introduced there. Again, without loss, we may
assume σ meets W exactly once, at t = t
1
. We consider again the map ˆs of
the case b
1
(Y ) = 0. Computed at η = 0, A = A
1,j
, Φ = 0, this time we find
ker (dˆs) (Ω
0
2
(iR) × X
A
1,j
× R) = iR H
1
(iR) H
A
1,j
coker (dˆs) = H
A
1,j
.
We have the obstruction map Ξ : iR × H
1
(iR) × H
A
1,j
H
A
1,j
whose
second derivative at (0, 0, 0) is
D
2
Ξ(δη, δa, δφ) =
1
2
Π(δa · δφ) δηδφ.
Then if we let δa = ω
0
(t
1
)δt, the term
1
2
Π(δa · δφ) =
1
2
δtΠ( ω
0
(t
1
) · δφ) = cδtδφ
where c is a non-zero real constant with the same sign of the spectral flow
{D
i
(s)} at s
1,j
. As in the b
1
= 0 case, the irreducible zeros of Ξ are mo delled
by the subset 0 × 0 × H
A
1,j
and the reducible zeros by 0 × H
1
(iR) × 0.
Let us now deal with the orientations in this case. Looking at the long
exact sequence inducing h of (3) we see
0 H
0
(iR) H
1
(iR) H
A
1,j
=
ker L
A
1,j
,0,t
1
R
κ
H
0
(iR) H
1
(iR) H
A
1,j
β
coker L
A
1,j
,0,t
1
0.
We note that κ(t) = ω
0
(t
1
)t maps isomorphically onto the H
1
(iR) factor,
and coker L
A
1,j
,0,t
1
is H
0
(iR) H
1
(iR). β is the obvious projection. Let us
assume the canonical orientations on H
0
(iR), H
A
1,j
, and the orientation on
H
1
(iR) induced by κ, which is given by ω
0
(t
1
). Finally choose the product
orientation (in the order indicated) on H
0
(iR) H
1
(iR) H
A
1,j
. Then
ker L
A
1,j
,0,t
1
is identically oriented and coker L
A
1,j
,0,t
1
is oriented according
to the order H
0
(iR) H
A
1,j
.
Let v H
A
1,j
. Then combining the slice condition with Ξ gives the moduli
space near (0, 0, v) (as before we may assume G = 0) as the zeros of the map
χ : iR × H
1
(iR) × H
A
1,j
iR × H
A
1,j
,
χ(η, a, φ) = (ihiv, φi, Ξ(η, a, φ)).
202 YUHAN LIM
As before the zeros of χ are the positive multiples of (0, 0, v). The pull-back
of the orientation on the target space onto the normal bundle of χ
1
(0) is
given by δθ η δa where δθ is the angular coordinate on H
A
1,j
. Letting
r be the direction determined by v and εδr the induced orientation, then we
require
εδr (θ δη δa) = δη δa δr δθ
which gives the induced orientation on χ
1
(0) as r.
As mentioned b efore, Z
σ(t)
(P ) admits reducible solutions exactly σ(t)
W. This corresponds to when [
i
2π
ω(t)] coincides with the c lass c
1
(L)
R
. Let
U denote a connected component of H
1
(Y ; R){c
1
(L)
R
}. Then if our path
σ has the property that [
i
2π
ω(t)] U for all t, then #Z
σ(0)
(P ) = #Z
σ(0)
.
Therefore #Z
α,ω,g
(P ) is an integer-valued function depending only on the
choice of U. Denote this function as τ(U).
We think of {c
1
(L)
R
} as a ‘wall’ in H
2
(Y ; R). Then as we cross this wall
τ changes. This change can be determined from the previous proposition to
give a ‘wall-crossing’ formula.
Corollary 20. Let a H
2
(Y ; Z)/torsion be an indivisible class and let
c
1
(L)
R
= 2na. Let U
±
be the component of H
2
(Y ; R){c
1
(L)
R
} containing
(2n ± 1/2)a. Then
τ(U
+
) τ(U
) = n.
Proof. Take (π
0
, g
0
) and (π
1
, g
1
) which define the values τ(U
+
) and τ (U
)
respectively. Choose our connecting path σ with properties as used for as
for Prop. 19. Without loss, We may suppose that σ cros ses W exactly once,
say at t = t
1
. We now follow the notation and ideas in the proof of Prop. 19.
According to Prop. 19 We need then to compute the total spectral flow of
the family {D
1
(s)} as s varies from 0 to 1. The orientation of this family
is determined by ω
0
(t
1
). We shall choose a to be consistent with this
orientation, but the statement of the corollary is actually independent of
this choice. Take a positive multiple ω of ω
0
(t) such that with [
i
2π
ω] = 2a.
Thus A
1
ωs, 0 s < 1 parameterizes all the reducibles in Z
σ(t
1
)
(P ).
We may deform the family {D
1
(s)} preserving self-adjointness to the family
{D
g(t
1
)
A
1
−∗ωs
}, s [0, 1]. Thus it suffices to compute the spectral flow for this
family. Notice that there is a gauge transformation g such that g(A
1
) = A
1
ω, or equivalently g
1
dg = −∗ω. A theorem of [APS] says that the spectral
flow of the Dirac operators {D
g(t
1
)
A
1
−∗ωs
} is equivalent to computing the index
of a Dirac operator D
(4)
A
on Y ×S
1
with a spin-c structure obtained by taking
the product P ×[0, 1] over Y ×[0, 1] and identifying via g : P ×{1} P ×{0}.
A is a connection which is in temporal gauge and coincides with A ωs on
L × {s}. (Remark: We follow the orientation c onventions of [APS] closely.
SEIBERG-WITTEN INVARIANTS ... 203
In particular Y × S
1
has the product orientation dy ds where dy is the
orientation form on Y and s the real coordinate on S
1
thinking of it as
R/Z.) Denote the resulting dete rminant on Y × S
1
by L
0
. The index of
D
(4)
A
is given by
1
8
hc
2
1
(L
0
), [Y × S
1
]i +
1
8
sig(Y × S
1
).
To compute the first term, we notice F
A
= d(A
1
ωs) = F
A
1
ωds. Then
hc
2
1
(L
0
), [Y × S
1
]i =
Z
Y ×[0,1]
i
2π
F
A
i
2π
F
A
=
1
4π
2
Z
Y ×[0,1]
(F
A
1
ωds) (F
A
1
ωds)
=
1
4π
2
(2)
Z
Y ×[0,1]
F
A
1
ωds
=
1
4π
2
(2)
Z
Y ×[0,1]
ωds
=
1
4π
2
(2n)
Z
Y
ω ω
Z
1
0
ds
=
1
4π
2
(2n)4πia · PD(4πia)
= 8n.
Here ‘PD’ denotes Poincare Duality. Since sig(Y × S
1
) = 0, the index of
D
(4)
A
is n and the corollary follows.
References
[APS] M.F. Atiyah, V.K. Patodi and I.M. Singer, Spectral asymmetry and Riemannian
geometry I, II, III, Math. Proc. Camb. Phil. Soc., 77 (1975), 43-69; 78 (1975),
405-32; 79 (1976), 71-99.
[B] J-P. Bourgiugnon, Spinors, Dirac operators, and changes of metric, in Differential
Geometry: Geometry in mathematical physics and related topics (Los Angeles CA,
1990), 41-44, Proc. Sympos. Pure Math., 54, part 2, Amer. Math. Soc., Providence,
RI, 1993.
[BG] J-P. Bourguignon and P. Gauduchon, Spineurs, operateurs de Dirac et variations
de metriques, Comm. Math. Phys., 144(3) (1992), 581-599.
[Ch] W. Chen, Casson invariant and Seiberg-Witten gauge theory, Turkish J. Math., 21
(1997), 61-81.
[D] S.K. Donaldson, The Seiberg-Witten equations and 4-manifold topology, Bull.
A.M.S., 33(1) (1996), 45-70.
[Fy] K.A. Froyshov, The Seiberg-Witten equations and 4-manifolds with boundary,
Math. Res. Lett., 3 (1996), 373-390.
204 YUHAN LIM
[H] O. Hijazi, A conformal lower bound for the smallest eigenvector of the Dirac oper-
ator and Killing spinors, Comm. Math. Phys., 104(1) (1986), 151-162.
[KM] P. Kronheimer and T. Mrowka, The genus of embedded surfaces in the projective
plane, Math. Res. Let., 1(6) (1994), 797-808.
[LM] H.B. Lawson and M-L. Michelsohn, Spin Geometry, Princeton Univ. Press, 1989.
[L] Y. Lim, The Equivalence of Seiberg-Witten and Casson Invariants for Homology
3-spheres, Math. Res. Lett., 6(6) (1999), 631-643.
[Mar] M. Marcolli, Seiberg-Witten-Floer theory and Heegaard splittings, Int. J. Math.,
7(5) (1996), 671-696.
[MT] G. Meng and C.H. Taubes, SW = Milnor Torsion, Math. Res. Lett., 3(5) (1996),
661-674.
[M] J. Morgan, The Seiberg-Witten equations and applications to the topology of smooth
four-manifolds, Princeton Math. Notes, 44, Princeton Univ. Press, 1996.
[N] L. Nicolaescu, Lattice points, Dedkind-Rademacher sums and a conjecture of
Kronheimer-Mrowka, preprint.
[T] C.H. Taubes. Casson’s invariant and gauge theory, J. Diff. Geom., 31 (1990), 547-
599.
Received February 1, 1998 and revised May 14, 1999.
University of Ca lifornia
Santa Barbara, CA 93106
E-mail address: ylim@gauss.math.ucsb.edu
PACIFIC JOURNAL OF MATHEMATICS
Vol. 195, No. 1, 2000
VIRTUAL HOMOLOGY OF SURGERED TORUS BUNDLES
Joseph D. Masters
Let M be a once-punctured torus bundle over S
1
with
monodromy h. We show that, under certain hypotheses on
h, “most” Dehn-fillings of M (in some cases all but finitely
many) are virtually Z-representable. We apply our results to
show that even surgeries on the figure eight knot are virtually
Z-representable.
1. Introduction.
Embedded incompressible surfaces are fundamental in the study of 3-mani-
folds. Accordingly, the following conjecture of Waldhausen and Thurston
has attracted much attention:
Conjecture 1.1. Let M be a closed, irreducible 3-manifold with infinite π
1
.
Then M has a finite cover which is Haken.
The focus of this paper is the following, stronger, conjecture:
Conjecture 1.2. Let M be as above. Then M has a finite cover
˜
M with
H
1
(
˜
M, Z) infinite.
If M is a compact 3-manifold, we say that M is Z-representable if
H
1
(M, Z) is infinite. If M satisfies the conclusion of Conjecture 1.2, we
say that M is virtually Z-representable.
We shall give what appear to be the first examples of 3-manifolds with
torus boundary for which all but finitely many fillings are virtually Z-
representable, but not Z-representable (in fact non-Haken). Boye r and
Zhang have independently given examples of knot complements for which
all but finitely many fillings are virtually Haken, but non-Haken [BZ].
Before we can state our results, we must establish some notation. Let F
be a once-punctured torus with π
1
(F ) = h[x], [y]i, and basepoint x
0
F
(see Fig. 1).
Any orientation-preserving homeomorphism h : F F is isotopic to one
of the form h = D
r
1
x
D
s
1
y
· · · D
r
k
x
D
s
k
y
. Here D
x
and D
y
are Dehn twists along
simple closed curves homologous to x and y, respectively. The twists D
x
205
206 JOSEPH D. MASTERS
h
x
0
x
y
β
α
Figure 1. Notation for the once-punctured torus bundle M.
and D
y
induce the following actions on π
1
(F ):
D
x]
(x) = x
D
x]
(y) = yx
D
y]
(x) = yx
D
y]
(y) = y.
We may assume h fixes F . Let M
h
= (F × I)/h be the once-punctured
torus bundle with monodromy h. We specify a framing for H
1
(M
h
, Z) by
setting the longitude β = F oriented counter-clockwise, and the meridian
α = (x
0
×I)/h, where x
0
is some point on F , and α is oriented as in Fig. 1.
Then, for coprime integers (µ, λ), M
h
(µ, λ) denotes the manifold obtained
by gluing a solid torus to M
h
in s uch a way that the curve α
µ
β
λ
becomes
homotopically trivial.
We shall prove:
Theorem 1.3. Let M
h
be a once-punctured torus bundle over S
1
, with
monodromy h = D
r
1
x
D
s
1
y
· · · D
r
k
x
D
s
k
y
, and let n = g.c.d{s
1
, . . . , s
k
}, R =
r
1
+ · · · + r
k
.
(i) If n is divisible by some m such that m 6 and m is even or m = 7,
and if |λ| > 1, then all but finitely many Dehn-fillings M
h
(µ, λ) are
virtually Z-representable.
(ii) If n is divisible by some m such that m 5, m is odd, and m 6= 7,
and if 1/|Rµ λ| + 1/| 2λ| + 1/|λ| < 1, then M
h
(µ, λ) is virtually
Z-representable.
(iii) If n is divisible by 4, and if 2/| 2λ| + 1/|λ| < 1, then M
h
(µ, λ) is
virtually Z-representable.
VIRTU AL HOMOLOGY OF SURGERED TORUS BUNDLES 207
Remarks. 1. Analogous results hold if we replace n by gcd{r
1
, . . . , r
k
} and
R by s
1
+ · · · + s
k
.
2. It was shown in [B1] that if m 2, n 2 and mn 8 but mn 6= 9,
then all non-integral surgeries are virtually Z-representable. In [B2] it was
shown that if 4|n, then for each µ, M
h
(µ, λ) is virtually Z-representable for
all but finitely many λ coprime to µ.
3. From [CJR] and [FH], all but finitely many surgeries on a once-
punctured torus bundle over S
1
yield non-Haken manifolds.
Theorem 1.3 may b e used to show that, for certain choices of f , all but
finitely many surgeries on M
f
are virtually Z-representable. For example:
Theorem 1.4. Let f = (D
x
D
y
)
18
. Then every surgery on M
f
is virtually
Z-representable.
The proof of Theorem 1.4 appears in Section 3.
In order to state the next theorem, we require some notation. Let 1 =
(D
x
D
1
y
D
x
)
2
, the central involution on the punctured torus. If h is a home-
omorphism of the punctured torus, h stands for (1)h.
Theorem 1.5. Let N = M
D
x
D
y
(also known as “the figure eight knot’s
sister”). Then if 1/|µ λ| + 1/|µ 2λ| + 1/|λ| < 1, N(µ, λ) is virtually
Z-representable.
Theorem 1.6. Let K denote the figure-eight knot and let M denote S
3
K.
Then, with respect to the canonical framing of knots in S
3
, any surgery of
the form M(2µ, λ) is virtually Z-representable.
Other results on virtually Z-representable figure-eight knot surgeries may
be found in [M], [KL], [H], [N] and [B3]. In particular, it was shown in [KL]
and [B3] that surgeries of the form M (4µ, λ) are virtually Z-repres entable.
It was also shown in [B3] that surgeries of the form M (2µ, λ) are virtually
Z-representable if λ = ±7µ (mod 15). Finally, it was shown in [Bart] that
every non-trivial surgery of M contains an immersed incompressible surface.
Our techniques are extensions of Baker’s. The main new ingredient is the
use of group theory to encode the combinatorics of cutting and pasting.
I would like to thank Professor Alan Reid for his help and patience.
2. Construction of covers.
We begin by recalling Baker’s construction of covering spaces of M
h
(µ, λ)
(see [B1], [B2]). Let n be as in the statement of Theorem 1.3, and let
ˆ
F
be the kn-fold cover of F associated to the kernel of the map φ : π
1
(F )
Z
k
× Z
n
, with φ([x]) = (1, 0) and φ([y]) = (0, 1) (see Fig. 2).
Now create a new cover,
˜
F , of F by making vertical cuts in each row of
ˆ
F , and gluing the left side of each cut to the right side of another cut in the
208 JOSEPH D. MASTERS
x
y
n
k
Figure 2. The cover
ˆ
F of F .
same row. An example is pictured in Figure 3, where the numbers in each
row indicate how the edges are glued.
If h lifts to a map
˜
h :
˜
F
˜
F , then the mapping cylinder
˜
M
h
=
˜
F /
˜
h is
a cover of M
h
. Furthermore, if the loop α
µ
β
λ
lifts to loops in
˜
M
h
, then the
cover extends to a cover
˜
M
h
(µ, λ) of M
h
(µ, λ).
If the cover
˜
M
h
exists, then we m ay compute its first Betti number with
the formula b
1
(
˜
M
h
) = rank (fix(
˜
h
)), where
˜
h
is the map on H
1
(
˜
M, Z)
induced by
˜
h, and f ix(
˜
h
) is the subgroup of H
1
(
˜
M, Z) fixed by
˜
h
(see [H]
for a proof). We shall use this formula to prove that, in some cases, b
1
(
˜
M)
is greate r than the number of boundary components of
˜
M, which ensures
that b
1
(
˜
M(µ, λ)) > 0.
We now introduce some notation to describe the cuts of
˜
F (see Fig. 3).
˜
F
is naturally divided into rows, which we label 1, . . . , n. The cuts divide each
row into pieces, each of which is a square minus two half-disks; we number
them 1, . . . , k. If we slide a point in the top half of the i
th
row through the
cut to its right, we induce a permutation on {1, . . . , k}, which we denote
VIRTU AL HOMOLOGY OF SURGERED TORUS BUNDLES 209
Piece 3, Row 1
1 3 4
65
1
2
3
4
5 6
1 2 3 4
5 6
6 2
1 5
2 3
2
3
4 5 6
1
3 6 2 1
4 5
1
3
2 6 3 2
4 1
5 4
56
σ = (15)(2463)
σ = (146235)
σ =
1
2
3
Row 1
Row 2
Row 3
2
1
Figure 3. The permutations encode the combinatorics of
the gluing.
σ
i
. Thus the cuts on
˜
F may be encoded by elements σ
1
, . . . , σ
n
S
k
, the
permutation group on k letters .
Next, we find algebraic conditions on the σ
i
’s which will guarantee that
the cover of F extends to a cover of M(µ, λ). We first must pick k, n, and
{σ
1
, . . . , σ
n
} so that h lifts to
˜
F .
Lemma 2.1. If
I. [σ
i
, σ
1
σ
2
· · · σ
i1
] = 1 for all i and
II. σ
1
σ
2
· · · σ
n
= 1
then h lifts to
˜
F .
Proof. Note that D
n
y
lifts to Dehn twists on
˜
F . Therefore, we need only
ensure that D
x
lifts. We shall attempt to lift D
x
to a sequence of “fractional
Dehn twists” along the rows of
˜
F . Let ˜x
i
denote the disjoint union of the
lifts of x to the i
th
row of
˜
F . We first attempt to lift D
x
to row 1, twisting
1/k
th
of the way along ˜x
1
. Considering the action on the bottom half of row
1, we find that the cuts are now matched up according to the permutation
σ
1
1
σ
2
σ
1
. Thus, for D
x
to lift to row 1 we assume σ
1
and σ
2
commute.
We now twist along ˜x
2
. The top halves of the squares in row 2 are moved
according to the permutation σ
1
σ
2
, and the lift will extend to all of row 2 if
and only if σ
3
commutes with σ
1
σ
2
. We continue in this manner, obtaining
the conditions in I. After we twist through ˜x
n
, we need to be back where we
started in row 1, so we require the additional condition σ
1
σ
2
· · · σ
n
= 1.
Note that the loop α
µ
lifts homeomorphicly to loops in
˜
M
h
if
˜
h
µ
= id, and
that the loop β
λ
lifts to loops in
˜
M
h
if (σ
i+1
σ
1
i
)
λ
= id for all i = 1, . . . , n.
Then, by considering the action of
˜
h on
˜
M
h
, the following condition for a
loop in M
h
to lift to
˜
M
h
is easily verified:
210 JOSEPH D. MASTERS
Lemma 2.2. The loop α
µ
β
λ
M
h
lifts homeomorphicly to loops in
˜
M
h
if and only if
III. (σ
1
· · · σ
i
)
(σ
i+1
σ
1
i
)
λ
= 1, for i = 1, . . . , n.
Therefore we may construct covers of M
h
(µ, λ) simply by finding permu-
tations satisfying conditions I-III.
Proof of Theorem 1.3.
Case 1. m = 4.
Construction of the cover of M
h
(µ, λ).
To construct a cover of M
h
(µ, λ), we must first construct a cover of F .
It was shown in the discussion prior to Lemma 2.1 that there is a unique
such cove r associated to any four permutations σ
1
, σ
2
, σ
3
and σ
4
in any
permutation group S
k
.
To ensure that the cover of F extends to a cover of M
h
, we shall set
σ
2
= σ
1
1
and σ
4
= σ
1
3
(see Fig. 4a). Then conditions I and II of Lemma
2.1 are satisfied automatically, so that any choice of σ
1
and σ
3
will determine
a cover of M
h
.
To ensure that the cover extends to M
h
(µ, λ), we must arrange for the
surgery curve α
µ
β
λ
to lift to
˜
M
h
. By Lemma 2.2, α
µ
β
λ
will lift provided
that σ
1
, . . . , σ
4
satisfy condition III, which reduces to:
σ
2λ
1
= 1(1)
(σ
3
σ
1
)
λ
= 1(2)
σ
2λ
3
= 1(3)
(σ
1
σ
3
)
λ
= 1.(4)
Any pair of permutations σ
1
and σ
3
satisfying Equations (1)-(4) deter-
mines a unique cover of M
h
(µ, λ). We now turn our attention to the con-
struction of such permutations.
Consider the abstract group G generated by the symbols ¯σ
1
and ¯σ
3
, satis-
fying relations (1)-(4). G is a (| 2λ|, | 2λ|, |λ|)-triangle group. It is
well-known that if 1/|2λ|+1/|Rµ2λ|+1/|λ| < 1, then G is residually
finite, and hence surjects a finite group H such that the images of ¯σ
1
, ¯σ
3
,
and ¯σ
3
¯σ
1
have order | 2λ|. By taking the permutation representation
of H, we then obtain permutations σ
1
and σ
3
satisfying conditions (1)-(4).
Note that the permutations act on |H| letters, so
˜
M is a 4|H|-fold cover of
M
h
.
Associated with the permutations σ
1
and σ
3
we have covers
˜
M
h
and
˜
F of
M
h
and F , and a cover
˜
M
h
(µ, λ) of M
h
(µ, λ);
Claim. b
1
(
˜
M
h
(µ, λ)) > 0.
VIRTU AL HOMOLOGY OF SURGERED TORUS BUNDLES 211
σ
σ
σ
σ
1
−1
1
3
−1
3
σ
−1
σ
−1
1
σ
1
σ
id
4
4
ba
Figure 4
a. The cover when n = 4. b. The cover when n = 5.
Proof of claim. It suffices to show that
˜
h
has a non-peripheral class [δ]
H
1
(
˜
F ) with
˜
h
([δ]) = [δ]. To construct this element, we shall first find a
non-peripheral class [δ
2
] in row 2, as follows.
σ
1
−1
σ
3
Figure 5. The surface
˜
F
2
(with |H| = 4).
Consider the sub-surface
˜
F
2
obtained by deleting rows 1, 3 and 4 from
˜
F
(see Fig. 5). The punctures of
˜
F
2
are in 1-1 correspondence with the cycles
212 JOSEPH D. MASTERS
of σ
1
, σ
3
and σ
3
σ
1
. Any permutation τ coming from the permutation repre-
sentation of H decomposes as a product of |H|/order(τ) disjoint order(τ )-
cycles. Therefore
˜
F
2
has |H|(1/order(σ
1
) + 1/order(σ
3
) + 1/order(σ
3
σ
1
)) <
|H| punctures. Since χ(
˜
F
2
) = −|H|, we deduce that
˜
F
2
contains a non-
peripheral class [δ
2
]. The class δ
2
also represents a non-peripheral class in
˜
F , since it has non-zero intersection number with a class of
˜
F in row 2.
We may find a corresponding non-peripheral loop δ
4
in row 4, such that
I([δ
2
+ δ
4
], [˜y
i
]) = 0 for all i (see Fig. 6 for the notation and the idea of
the proof). Let [δ] = [δ
2
+ δ
4
]. Then, since [δ] has non-zero intersection
number with classes in row 2 and row 4, it is a non-peripheral class. We
have I[δ, ˜y
i
] = 0 for all i (where I(., .) denotes oriented intersection number);
therefore [δ] is fixed by
˜
D
4
y
, and since
˜
D
x
fixes rows 2 and 4, it is fixed by
˜
D
x
. Therefore it is fixed by
˜
h
, concluding the proof of the claim, and of
Case 1.
Case 2. m 5 and m is odd.
Case 2a. m = 5.
The construction proceeds analogously to the case m = 4. We require per-
mutations σ
1
, . . . , σ
5
satisfying conditions I-III. Again, to simplify matters,
we shall impose some extra conditions: σ
2
= id, σ
3
= σ
1
1
, and σ
5
= σ
1
4
(see Fig. 4). Then conditions I-III reduce to:
σ
λ
1
= 1
(σ
1
σ
4
)
λ
= (σ
4
σ
1
)
λ
= 1
σ
2λ
4
= 1.
Again, these relations determine a triangle group, which, under the hypothe-
ses on µ and λ, is hyperbolic. The rest of the proof is identical to Case 1,
except that now the fixed class is in rows 3 and 5.
Case 2b. m 9 and m is odd.
Consider the cover obtained by setting σ
2
= σ
1
1
, σ
4
= id, σ
5
= σ
1
3
,
σ
6
= σ
1
, σ
7
= σ
1
1
, and for i = 4, . . . , k, σ
2i+1
= σ
1
2i
(see Fig. 7a).
The corresponding relations are:
σ
2λ
1
= 1(5)
σ
λ
3
= 1(6)
(σ
3
σ
1
)
λ
= (σ
1
σ
3
)
λ
= 1(7)
(σ
8
σ
1
)
λ
= 1(8)
σ
2λ
2i
= 1 for i = 4, . . . , k(9)
(σ
2i+2
σ
2i
)
λ
= 1 for i = 4, . . . , k 1(10)
(σ
1
σ
2k
)
λ
= 1.(11)
VIRTU AL HOMOLOGY OF SURGERED TORUS BUNDLES 213
2
δ
2
δ
4
δ
4
i
y
i
y
i
y
i
y
δ
σ
σ
1
σ
σ
1
−1
σ
σ
−1
1
σ
σ
1
−1
3
3
3
3
−1
row 4
row 2
row 4
row 2
2
δ
2
δ
δ
4
δ
4
i
y
i
y
i
y
i
y
σ
σ
1
σ
σ
1
−1
σ
σ
1
σ
σ
1
−1
3
3
3
3
2
δ
2
δ
δ
4
δ
4
i
y
i
y
i
y
i
y
σ
σ
1
σ
σ
1
−1
σ
σ
1
−1
σ
σ
1
3
3
3
3
−1
−1
−1
−1
Figure 6. How to find cancelling loops in rows 2 and 4.
These relations again determine a Coxeter group. It is well-known (see
[V]) that any such group surjects a finite group “without collapsing”– i.e.,
such that the orders of the images of the σ
i
’s and σ
i
σ
j
’s are as given in
(5)-(11). Then, arguing as in Case 1, we may find a non-peripheral fixed
class in rows 2 and 5.
214 JOSEPH D. MASTERS
1
−1
σ
σ
1
σ
σ
σ
−1
−1
1
1
σ
σ
−1
2k
2k
id
σ
3
3
σ
σ
σ
σ
1
−1
1
−1
σ
σ
σ
σ
1
−1
1
3
−1
3
7
7
σ
σ
−1
2k-1
2k-1
a b
Figure 7. a. The cover for n = 2k + 1 9. b. The cover
for n = 2k 8.
Case 3. n = 6
Case 3a. 2/| λ| + 1/|λ| < 1.
Again, we need permutations σ
1
, . . . , σ
6
satisfying I-III. In this case we
impose the additional conditions σ
2
= id, σ
3
= σ
1
1
, σ
5
= id, and σ
6
= σ
1
4
.
VIRTU AL HOMOLOGY OF SURGERED TORUS BUNDLES 215
Then conditions I-III reduce to:
σ
λ
1
= 1
(σ
1
σ
4
)
λ
= (σ
4
σ
1
)
λ
= 1
σ
λ
4
= 1.
These relations dete rmine a triangle group, and we find a fixed class in
rows 3 and 6.
Case 3b. |λ| > 2 and |3λ| |λ|, or λ is even (non-zero) and |3λ|
4.
When n = 3, conditions I-I II may be abelianized to obtain a cyclic
group of order | 3λ|. Specifically, they are satisfied by setting σ
1
=
(1, 2, . . . , Rµ 3λ), σ
2
= σ
2
1
, and σ
3
= σ
1
. For n = 6, we may “double”
this cover: That is take σ
1
, σ
2
, σ
3
as above, and then set σ
4
= σ
1
, σ
5
= σ
2
,
and σ
6
= σ
3
. Then we modify the corresponding cover
˜
M(µ, λ) of M(µ, λ)
by making horizontal cuts in adjacent squares of row 3 and gluing the flaps
back together as indicated by Fig. 8. If λ is even, we make two non-adjacent
cuts and glue the top of one to the bottom of the other. If λ is odd, we
make (|λ|1)/2 pairs of adjacent cuts and glue the top of the one cut to the
bottom of the other cut in its pair. Now make the same cuts in row 6, with
the same identifications. Since rows 3 and 6 are fixed by
˜
D
x
, D
x
still lifts
to the modified
˜
M
h
(µ, λ), and since the ˜y’s still project 6 to 1 onto y, D
y
lifts also; so h lifts. Also, one may check that α
µ
β
λ
still lifts, so
˜
M
h
(µ, λ)
remains a cover of M
h
(µ, λ).
To see that b
1
(
˜
M
h
(µ, λ)) > 0, note that
˜
D
x
fixes rows 3 and 6, so it is
enough to find a non-peripheral loop in row 3 and add it to the corresponding
loop in row 6 with opposite orientation. As in Case 1, the existence of such
a non-peripheral loop follows from an Euler characteristic argument (or see
Fig. 8).
Note that Case 3a or 3b applies to all but finitely many (µ, λ) with |λ| > 1.
Case 4. n = 2k 8.
Case 4a. 2/| 2λ| + 1/|λ| < 1. Set σ
2
= σ
1
1
, σ
4
= σ
1
3
, σ
5
= σ
1
,
σ
6
= σ
1
1
, and σ
2i
= σ
1
2i1
for i = 4, . . . , k (see Fig. 7b). Then, as in Case
2, these relations determine a Coxeter group. We may find a non-peripheral
fixed class in rows 2 and 4.
Case 4b. | λ| 2 We cannot guarantee, in this case , that there will
always be a cover with b
1
> 0, but we shall show that there are at most
finitely many exceptions.
We argue as in Case 3b. Take permutations σ
1
, . . . , σ
k
, and consider the
relations obtained by abelianizing conditions I-III. We claim that they can
be satisfied by setting σ
1
= (1, 2, 3, . . . , N), for some N, and setting each
216 JOSEPH D. MASTERS
1
4
1
3
4
32
1
4
1
3
4
2
b
b
2
a
32
2
3
1 2 3 4
a
b
c
d
1
4
a
d
c
2
3
1 2 3 4
a
b
c
c
d
1
4
b
a
d
Figure 8. a. The cover and fixed class for n = 6, 3λ =
4, λ = 3. b. The cover and fixed class for n = 6, Rµ 3λ =
4, λ = 2.
σ
i
to an appropriate power of σ
1
. We have already seen that this may be
done when k = 3.
The σ
i
’s must s atisfy the following c onditions:
σ
λ
1
σ
λ
2
= 1(12)
σ
1
σ
λ
2
σ
λ
3
= 1(13)
.
.
.
σ
1
σ
2
· · · σ
k2
(14)
σ
λ
k1
σ
λ
k
= 1(15)
σ
+λ
1
σ
2
· · · σ
k1
(16)
σ
λ
k
= 1(17)
σ
1
σ
2
· · · σ
k
= 1.(18)
VIRTU AL HOMOLOGY OF SURGERED TORUS BUNDLES 217
We shall assume that this system has a cyclic solution, so we may sub-
stitute σ
i
= σ
e
i
1
. Then, Equations (12)-(18) are e quivalent to the following
conditions on the exponents (all of the following equations in this case are
taken mod N):
λ + λe
2
= 0(19)
+ ( λ)e
2
+ λe
3
= 0(20)
.
.
.
+ Rµe
2
+ · · · + Rµe
k2
+ ( λ)e
k1
+ λe
k
= 0(21)
+ λ + Rµe
2
+ · · · + Rµe
k1
+ ( λ)e
k
= 0(22)
1 + e
2
+ · · · + e
k
= 0(23)
(22) and (23) imply that λ = λe
k
. Let us set e
k
= 1, eliminating Equation
(22). Then, using (23), we may pair off (19) and (21) to deduce that λe
2
=
λe
k1
, and we set e
2
= e
k1
to eliminate (21). Similarly, we set e
3
= e
k2
,
and so on. If k is even, we are left with equations:
λ + λe
2
= 0(24)
+ ( λ)e
2
+ λe
3
= 0(25)
.
.
.
+ Rµe
2
+ · · · + (Rµ λ)e
k/21
+ λe
k/2
= 0(26)
+ Rµe
2
+ · · · + (Rµ λ)e
k/2
+ λe
k/2
= 0(27)
2 + 2e
2
+ · · · + 2e
k/2
= 0.(28)
If we replace (28) by
1 + e
2
+ · · · + e
k/2
= 0(29)
then we may eliminate (27). Then solve for λe
2
, λ
2
e
3
, . . . , λ
k/21
e
k/2
. By
(29), we have:
λ
k/21
+ λ
k/22
(λe
2
) + λ
k/23
(λ
2
e
3
) + · · · + λ
k/21
e
k/2
= 0.(30)
Substituting our solutions for λe
2
, λ
2
e
3
and so on, we get the equation N = 0
for some integer N; the system has a solution in Z/NZ.
If k is odd, then our reduced system looks like:
λ + λe
2
= 0(31)
+ ( λ)e
2
+ λe
3
= 0(32)
218 JOSEPH D. MASTERS
.
.
.
+ Rµe
2
+ · · · + (Rµ λ)e
(k1)/2
+ λe
(k+1)/2
= 0(33)
+ Rµe
2
+ · · · + (Rµ λ)e
(k+1)/2
+ λe
(k1)/2
= 0(34)
2 + 2e
2
+ · · · + 2e
(k1)/2
+ e
(k+1)/2
= 0.(35)
Adding (33) and (34) gives a multiple of (35), so we may eliminate (34).
Then we s olve for λe
2
, λ
2
e
3
, . . . , λ
(k1)/2
e
(k+1)/2
. By (35), we have:
2λ
(k1)/2
+ 2λ
(k3)/2
(λe
2
) + 2λ
k5/2
(λ
2
e
3
) + · · ·
+2λ(λ
(k3)/2
e
(k1)/2
) + λ
(k1)/2
e
(k+1)/2
= 0.
And again we get a solution in Z/NZ for some N.
Then, as in Case 3b, M(µ, λ) will have a cover with b
1
> 0, provided that
|N| |λ| and |λ| > 2. Solving for N, if k is even, gives:
N = λ
k/21
+ λ
k/22
(λ ) + λ
k/23
[(λ )
2
Rµλ](36)
+λ
k/24
[(λ )((λ Rµ)
2
Rµλ) Rµλ(λ ) Rµλ
2
] + · · ·
and if k is odd:
N = 2λ
(k1)/2
+ 2λ
(k3)/2
(λ ) + 2λ
(k5)/2
[(λ )
2
Rµλ](37)
+2λ
(k7)/2
[(λ )((λ Rµ)
2
Rµλ) Rµλ(λ ) Rµλ
2
]
+ · · · + 1[..].
We are supposing that |Rµ λ| 2. So for large µ or λ, Rµ/λ 1, and
for k even,
N = o[λ
k/21
+ λ
k/23
(λ
2
) + λ
k/24
(λ
3
) + · · · ]
= o[λ
k/21
λ
k/21
λ
k/21
· · · ].
So if k is even and k 8, then for all but finitely many µ and λ, |N | > |λ|,
and we are done. Similarly, if k is odd and k 7, then we are done. In the
remaining cases, |N| is given by:
k = 4, |N| = | 2λ|
k = 5, |N| = |()
2
5Rµλ + 5λ
2
|
k = 6, |N| = |()
2
4Rµλ + 3λ
2
|.
One may check that each condition is satisfied by only finitely many rela-
tively prime pairs (µ, λ) with | λ| 2. This concludes the proof in Case
4b.
Case 5. n = 7, and |λ| > 1.
Case 5a. 1/| λ| + 1/|λ| < 2/3 and |( 2λ)
2
2λ
2
| > 2|R|.
VIRTU AL HOMOLOGY OF SURGERED TORUS BUNDLES 219
We shall consider covers with σ
2
= id, σ
3
= σ
1
1
, σ
6
= σ
1
5
, and σ
7
= σ
1
4
(see Fig. 9a).
σ
1
σ
σ
σ
id
σ
2
−1
σ
1
−1
6
6
−1
2
b
σ
1
σ
σ
σ
σ
−1
σ
id
1
−1
4
5
4
5
−1
a
Figure 9. Two covers for n = 7.
We obtain conditions:
[σ
4
, σ
5
] = 1(38)
σ
λ
1
= 1(39)
(σ
4
σ
1
)
λ
= 1(40)
220 JOSEPH D. MASTERS
σ
4
(σ
5
σ
1
4
)
λ
= 1(41)
(σ
4
σ
5
)
σ
2λ
5
= 1(42)
(σ
1
σ
4
)
λ
= 1.(43)
Let us also assume for simplicity that σ
5
commutes with σ
1
. Equations (38),
(41) and (42) determine an abelian group A of order |(Rµ 2λ)
2
2λ
2
|; we
must show that σ
2
4
is non-trivial in A. The elements σ
2
4
and σ
5
generate a
subgroup H of A of index at mos t 2. If σ
2
4
= id, then H is cyclic of order
gcd(|λ|, | 2λ|). Then |( 2λ)
2
2λ
2
| = |A| 2|H| = 2gcd(|λ|, |
2λ|) = 2gcd(|λ|, |R|) 2|R|. So if
|( 2λ)
2
2λ
2
| > 2|R|(44)
then σ
2
4
6= id. Therefore, under our hypotheses, the relations generate a
group which is isomorphic to the direct sum of a cyclic group with a hy-
perbolic triangle group. As in the previous cases, we may then find a non-
peripheral fixed class (in rows 3 and 7), and we are done.
However, note that if R = 1, then Equation (44) is false for all (µ, λ)
satisfying
(µ + 2λ)
2
2λ
2
= 1.
This is an example of Pell’s equation, which has infinitely many solutions,
and hence (44) may be false infinitely often.
Case 5b. 1/| 2λ| + 1/|λ| < 2/3 and |( λ)
2
2λ
2
| > 2|R|.
Let σ
3
= id, σ
4
= σ
1
2
, σ
5
= σ
1
1
, and σ
7
= σ
1
6
(see Fig. 9b). The
conditions for a cover are:
[σ
1
, σ
2
] = 1(45)
σ
1
(σ
2
σ
1
1
)
λ
= 1(46)
(σ
1
σ
2
)
σ
λ
2
= 1(47)
(σ
6
σ
1
)
λ
= 1(48)
σ
2λ
6
= 1(49)
(σ
1
σ
6
)
λ
= 1.(50)
For simplicity, suppose also that σ
2
commutes with σ
6
. Then (45), (46), (47)
determine an abelian group B of order |(Rµ λ)
2
Rµλ|. If σ
2
1
= 1, then
|B| 2gcd(|λ|, | λ|) = 2gcd(|λ|, |R|) 2|R|. Therefore, in this case,
the group determined by conditions (45)-(50) is again the direct pro duct
VIRTU AL HOMOLOGY OF SURGERED TORUS BUNDLES 221
of an abelian group with a hyperbolic triangle group, and we may find a
non-peripheral fixed class in rows 5 and 7.
Note that Case 5a or 5b applies to all but finitely many surgeries where
|λ| > 1.
This concludes the proof of Theorem 1.3.
3. Examples.
We begin with the proof of Theorems 1.5 and 1.6 (see Section 1 for notation).
Lemma 3.1. Let g = D
5
y
D
1
x
and h = D
x
D
y
. Then M
h
(µ, λ)
=
M
g
(µ, λ
µ), and M
h
2
(µ, λ)
=
M
(h)
2
(µ, λ + µ)
=
M
g
2
(µ, λ µ).
Proof. Re call that the mapping class group of the once-punctured torus
is isomorphic to SL
2
(Z), under the identifications D
x
R = [
1
0
1
1
] and
D
y
L = [
1
1
0
1
]. Under these identifications, we compute that h has mon-
odromy matrix [
2
1
1
1
], (1) has monodromy matrix [
1
0
0
1
], and g has mon-
odromy matrix [
1
5
1
4
]. The homeomorphisms h
2
and (h)
2
have the same
monodromy matrix, and hence are isotopic. Therefore M
h
2
=
M
(h)
2
. Also,
[
1
2
1
1
](RL)[
1
2
1
1
]
1
= L
5
R
1
, so g and h have conjugate monodromy
matrices. It follows that M
h
=
M
g
, and M
(h)
2
=
M
g
2
.
It remains to determine the effect of these homeomorphisms on the fram-
ings. Computing the maps on π
1
(F ) gives:
(h)
2
]
= (x
1
yxy
1
)(h
2
]
)(x
1
yxy
1
)
1
.
Therefore the isotopy which takes h
2
to (h)
2
twists F once in a counter-
clockwise manner, so the induced bundle homeomorphism sends M
h
2
(µ, λ)
to M
(h)
2
(µ, λ + µ).
Let f = D
2
y
D
1
x
. The bundle homeomorphism induced by conjugation
preserves the framing, so M
h
(µ, λ)
=
M
f(h)f
1
(µ, λ). The homeomor-
phisms f (h)f
1
and g have identical monodromy matrices, and hence
are isotopic. We compute g
]
= (yx
1
y
1
x)f(h)f
1
]
(yx
1
y
1
x)
1
so the
isotopy from f (h)f
1
to g twists F once in a clockwise manner. The
induced bundle homeomorphism sends M
f(h)f
1
(µ, λ) to M
g
(µ, λ µ). So
M
h
(µ, λ)
=
M
g
(µ, λ µ).
Likewise, M
f(h)
2
f
1
(µ, λ)
=
M
g
2
(µ, λ 2µ). Thus
M
h
2
(µ, λ)
=
M
(h)
2
(µ, λ + µ)
=
M
f(h)
2
f
1
(µ, λ + µ)
=
M
g
2
(µ, λ µ).
Proof of Theorem 1.5. This is an immediate consequence of Lemma 3.1 and
Theorem 1.3.
222 JOSEPH D. MASTERS
Proof of Theorem 1.6. We have M(2µ, λ)
=
M
h
(2µ, λ), which is double cov-
ered by M
h
2
(µ, λ)
=
M
g
2
(µ, λµ). So it is enough to show that M
g
2
(µ, λµ)
is virtually Z-representable. By Theorem 1.3, we are done unless
1/| 2µ (λ µ)| + 1/| 2µ 2(λ µ)| + 1/|λ µ| 1
or, simplifying:
1/|µ + λ| + 1/|2λ| + 1/|µ λ| 1.(51)
By [B3], M(2µ, λ) is virtually Z-representable if 2µ is divisible by 4; hence
we may assume µ is odd. Also, since gcd(2µ, λ) = 1, we may assume λ is
odd, and, assuming (µ, λ) 6= (±1, 1), |λ| 6= |µ|. It follows that
|µ λ| 2(52)
|µ + λ| 2.(53)
The only simultaneous solutions to inequalities (51), (52) and (53) with
µ and λ odd are: (µ, λ) = ±(3, 1) and ±(3, 1). So the only possible
exceptions to Theorem 1.6 are M(6, 1)
=
M(6, 1) and M(2, 1)
=
M(2, 1).
The virtual Z-representability of these manifolds may be verified with either
of the computer programs GAP or Snappea.
We now turn to the proof of Theorem 1.4.
Proof of Theorem 1.4. Let g and h be as in the statement of Lemma 3.1,
let f = h
18
, and let i = D
2
x
D
4
y
D
x
D
4
y
D
x
. Both h
3
and i have monodromy
matrix [
13
8
8
5
]; hence h
3
and i are isotopic. By arguments similar to those
used in the proof of Lemma 3.1, we compute that M
h
3
(µ, λ)
=
M
i
(µ, µ + λ).
Therefore M
f
(µ, λ)
=
M
i
6
(µ, λ + 6µ). Hence by Theorem 1.3 iii, M
f
(µ, λ)
is virtually Z-representable if
1/|6µ λ| + 1/|6µ + λ| < 1.(54)
By Lemma 3.1 we have M
f
(µ, λ)
=
M
g
18
(µ, λ 9µ). Hence by Theo-
rem 1.3 ii, M
f
(µ, λ) is virtually Z-representable if
1/|9µ + λ| + 1/|2λ| + 1/|9µ λ| < 1.(55)
The only simultaneous solutions to the inequalities 54 and 55 have µ =
0. The proof is completed by noting that M(0, 1) has positive first Betti
number, as it is a torus bundle over S
1
.
We remark that the same methods may be applied to many other ex-
amples of once-punctured torus bundles, to show that all but finitely many
surgeries are virtually Z-representable. The idea is to start with a mon-
odromy f to which Theorem 1.3 i or ii applies. Since L
4
and R generate
a finite-index subgroup of SL
2
(Z), there exists an integer ` such that f
`
is
VIRTU AL HOMOLOGY OF SURGERED TORUS BUNDLES 223
isotopic to a g satisfying the hypotheses of Theorem 1.3 iii. Usually The-
orem 1.3 will then imply that all but finitely many surgeries on M
f
`
are
virtually Z-representable.
References
[B1] M. Baker, Covers of Dehn fillings on once-punctured torus bundles, Proc. Amer.
Math. Soc., 105 (1989), 747-754.
[B2]
, Covers of Dehn fillings on once-punctured torus bundles II, Proc. Amer.
Math. Soc., 110 (1990), 1099-1108.
[B3] , On coverings of figure eight-knot surgeries, Pacific J. Math., 150 (1991),
215-228.
[Bart] A. Bart, Surface groups in surgered manifolds, to appear in Topology.
[BZ] S. Boyer and X. Zhang, Virtual Haken 3-manifolds and Dehn filling, Topology, 39
(2000), 103-114.
[CJR] M. Culler, W. Jaco and H. Rubinstein, Incompressible surfaces in once-punctured
torus bundles, Proc. London Math. Soc., 45(3) (1982), 385-419.
[FH] W. Floyd and A. Hatcher, Incompressible surfaces in punctured torus bundles,
Topology and it Applications, 13 (1982), 263-282.
[H] J. Hempel, Coverings of Dehn fillings of surface bundles, Topology and its Appli-
cations, 24 (1986), 157-170.
[KL] S. Kojima and D. Long, Virtual Betti numbers of some hyperbolic 3-manifolds, A
Fete of Topology, Academic Press, 1988.
[M] S. Morita, Finite coverings of punctured torus bundles and the first Betti number,
Sci. Papers College Arts Sci., Univ Tokyo, 35 (1986), 109-121.
[N] A. Nicas, An infinite family of hyperbolic non-Haken 3-manifolds with vanishing
Whitehead groups, Math. Proc. Camb. Phil. Soc., 99 (1986), 239-246.
[V] E.B. Vinberg, Groups defined by periodic paired relations, Sbornik: Mathematics,
188 (1997), 1269-1278.
Received June 15, 1998.
University of Tex as at Austin
Austin, Texas 78712
Rice University
Houston, Texas 77005-1892
E-mail address: mastersj@rice.edu
PACIFIC JOURNAL OF MATHEMATICS
Vol. 195, No. 1, 2000
GROUP ACTIONS ON POLYNOMIAL AND POWER
SERIES RINGS
Peter Symonds
When a finite group G acts faithfully on a graded integral
domain S which is an algebra over a field k, such as a poly-
nomial ring, we consider S as a kG-module. We show that S
is asymptotically mostly projective in each degree, and also
that it is in fact mostly free in an appropriate sense. Simi-
lar results also hold for filtered algebras, such as power series
rings.
1. Introduction.
Let S =
L
n=0
S
n
be a graded algebra over a field k. We suppose that
S is finitely generated over k as a k-algebra and that the homogeneous
components S
n
are finite dimensional vector spaces over k. Let G be a finite
group of grading preserving automorphisms of S (so G acts faithfully). We
are concerned with the structure of S as a kG-module.
The classical theory of Hilbert and Serre asserts that for large n, dim
k
S
n
is given by a function
φ
S
(n) = c
d1
(n)n
d1
+ c
d2
(n)n
d2
+ · · · + c
0
(n),
where the c
i
(n) are rational valued functions periodic in n, i.e., φ
i
(n +
p) = φ
i
(n) for some integer p (see Section 2). If c
d1
is assumed not to be
identically zero then d is equal to the dimension of the ring in various senses.
If S is a polynomial ring then d is equal to the number of variables.
From now on, we assume that S is an integral domain. Let P
n
denote the
maximal projective summand of S
n
(defined up to isomorphism).
Theorem 1.1. dim
k
(S
n
/P
n
) is bounded by a polynomial in n of degree d2.
Thus S
n
is mostly projective, and if S is a polynomial ring then the
non-projective part grows like a polynomial ring in one fewer variables.
In fact S is mostly free, although the individual S
n
do not have to contain
a free module at all; the different projectives can occur in different degrees.
To explain this let R = S
G
, the ring of invariants.
225
226 PETER SYMONDS
Theorem 1.2. S contains a free kG-submodule F of rank 1, a sum of ho-
mogeneous pieces, such that the product map R
k
F S is injective. De-
note its image by RF =
L
n
(RF )
n
. Then RF is a free summand of S and
dim
k
(S
n
/(RF )
n
) is bounded by a polynomial of degree d 2.
Of course, the first theorem is a corollary to the second. Versions of these
theorems were proven by Howe [4] in characteristic 0 and by Bryant [2, 3]
for polynomial rings.
Section 2 contains the main proof, except for some technical details which
appear in Section 3. Section 4 proves similar results for filtered algebras.
2. Main Proof.
Proof. We can assume that k is a splitting field for G, since if a kG-module
contains a free or projective summand after extension of scalars then it did
so before. Let Q
S
(resp. Q
R
) denote the fields of fractions of S (resp. R),
so Q
S
=
Q
R
R
S. By the Normal Basis Theorem, Q
S
is a free module of
rank 1 over Q
R
G; let e be a generator. Then, over kG, e generates a free
submodule E of rank 1 such that Q
S
=
Q
R
k
E. Now there is an r R
such that re S. Let F be the kG-module generated by re, so F S and
F = rE
=
kG. Also the product map R
k
F RF S is injective.
We claim that F can be assumed to be a sum of homogeneous pieces,
F =
L
i
F
n
i
. The proof of this plausible statement is somewhat delicate,
and we postpone it to the next section.
Let x
1
, . . . , x
s
be homogeneous generators for S as a k-algebra. Then
x
i
=
P
j
a
ij
b
ij
e
j
, where a
i,j
, b
i,j
R and the e
j
form a homogeneous k-basis
for F . By writing b
i,j
x
i
=
P
j
a
i,j
e
j
and taking the homogeneous component
of this equation in some degree where b
i,j
x
i
is non-zero, we see that we may
assume that the b
i,j
are homogeneous. Let α R
a
be the product of all the
b
i,j
. Then each x
i
α
1
RF , so S α
1
RF .
Thus
(RF )
n
S
n
α
1
(RF )
n+a
,
and so, identifying RF with R
k
F , we have
M
i
R
nn
i
F
n
i
S
n
α
1
M
i
R
n+an
i
F
n
i
.
In particular, the dimension of S
n
/(RF )
n
is bounded by the difference in
the dimensions of the two sides, i.e., by
X
i
(φ
R
(n + a n
i
) φ
R
(n n
i
)) dim
k
F
n
i
.
GROUP ACTIONS ON POLYNOMIAL RINGS 227
But
φ
R
(n + a n
i
) φ
R
(n n
i
) = c
d1
(n + a n
i
)(n + a n
i
)
d1
c
d1
(n n
i
)(n n
i
)
d1
+ lower degree terms,
and c
d1
is periodic, with period dividing a (see 3.1), so the n
d1
term
cancels and we are done.
3. Technical Details.
The form of φ
S
(n) given above is not quite the standard one, although it is
quoted in [4]. The usual references deal with a module over a polynomial
ring which has all the variables in degree 1, and then all the coefficients of
φ are constants. To deduce the version given in the introduction, note that
if S is generated by x
1
, . . . , x
s
then it is a finitely generated module over
k[x
1
, . . . , x
s
]. By taking suitable powers y
i
of the x
i
we can get all the y
i
in
the same degree b, and S will still be finitely generated over k[y
1
, . . . , y
s
].
For 0 j b 1, set T
j
=
L
l=0
S
j+lb
. Then R
=
L
j
T
j
as a k[y
1
, . . . , y
s
]-
module, and after regrading each T
j
so that each y
i
can have degree 1, we
can apply the usual theory ([1] 11.2, [5] VII Theorem 41) to each T
j
and
sum the results. It is the summation that leads to the periodic coefficients.
Lemma 3.1 ([4]). If R is an integral domain (as it always is for us), c =
gcd{r Z|R
r
6= 0} and φ
R
(n) = c
d1
(n)n
d1
+ · · · + c
0
(n), then there is a
constant b such that
c
d1
(n) =
b, if n|c,
0, otherwise.
Proof. If 0 6= α R
a
, then multiplication by α embeds R
n
in R
n+a
, so
for large n, φ
R
(n) φ
R
(n + a). Now consider the limit of φ
R
(n)/n
d1
as
n through elements of the same residue class modulo the period of
c
d1
to see that c
d1
(n) c
d1
(n + a). This, together w ith the periodicity,
implies the result.
Now we prove the claim made in Section 2.
Proposition 3.2. The free module F S can be assumed to be a sum of
homogeneous pieces in such a way that the product map Q
R
k
F Q
S
is
still an isomorphism.
Proof. For each simple kG-module V , let T
V
= Hom
kG
(V, S), a graded R-
module. Now soc F is a direct sum of simples. Let soc
V
(F ) denote the sum
of those isomorphic to V , so soc
V
(F ) = V
1
· · ·V
s
, where V
i
=
V , and let
P
V
i
be a projective summand of F with soc(P
V
i
) = V
i
. The inclusions of the
V
i
in S give us s homomorphisms f
i
T
V
, which are linearly independent
over R.
228 PETER SYMONDS
Lemma 3.3. Let f
1
, . . . , f
s
be elements of a graded R-module T which are
linearly independent over R. Write each f
j
as a sum of its homogeneous
components; f
j
=
P
k
f
j
k
, f
j
k
T
k
. Then for each j there is an integer k
j
such that f
1
k
1
, . . . , f
s
k
s
are linearly independent over R.
Proof. For each 0 t s, let P
t
be the claim that there exist integers
k
1
, . . . , k
t
such that f
1
k
1
, . . . , f
t
k
t
, f
t+1
, . . . , f
s
are linearly independent over
R. P
0
is true by hypothesis and we want P
s
. We give a proof by induction
on t, so assume P
t
.
If P
t+1
is false, then for each k Z we can find u
k
, r
i
k
R, u
k
6= 0, such
that
u
k
f
t+1
k
= r
1
k
f
1
k
1
+ · · · + r
t
k
f
t
k
t
+ r
t+2
k
f
t+2
+ · · · + r
s
k
f
s
.
Let u be the product of the u
k
for which f
t+1
k
6= 0. Then uf
t+1
=
P
k
(
u
u
k
)u
k
f
t+1
k
, contradicting P
t
.
Applying this to the {f
i
} T
V
we obtain homogeneous {
¯
f
i
} T
V
,
¯
f
i
T
a
i
, say, linearly independent over R.
Lemma 3.4. The evaluation map ev : T
V
k
V S is injective.
Proof. In fact ev : Hom
kG
(V, M)
k
V M is injective for any kG-module
M. This is because it factors through soc
V
(M), which is a direct sum
of V ’s, so we are reduced to proving the case M = V . But then ev is an
isomorphism, since Hom
kG
(V, V )
=
k, by the assumption that k is a splitting
field.
Corollary. The product map R
k
(
L
i
¯
f
i
(V )) S is injective.
Now let
¯
P
V
i
be the image of the projection of P
V
i
to S
a
i
. The projection
map is injective on soc(P
i
V
), by the construction of a
i
, so
¯
P
V
i
=
P
V
i
and
soc(
¯
P
V
i
) =
¯
f
i
(V ). Let
¯
P
V
=
L
i
¯
P
V
i
and consider the product map R
k
¯
P
V
S. Since soc(R
k
¯
P
V
) = R
L
i
¯
f
i
(V ), it is injective on the socle,
so is injective.
Finally, we sum the
¯
P
V
over the simples V to obtain
¯
F , a free kG-module
of rank 1, which is a sum of homogeneous pieces, as required.
Remark. If G is a p-group, where p is the characteristic of k, then the proof
is much simpler because soc(F )
=
k. Under at least one of the projections
of F onto its homoge neous components the image of soc(F ) must be non-
zero. Let
¯
F be the image of F under this projec tion. Then
¯
F
=
kG and
Q
R
k
F Q
S
is an isomorphism because it is injective on the socle, and
both sides have the same dimension over Q
R
.
This is enough to prove 1.1 for general G. For if P = Syl
p
(G) then S is a
direct summand of Ind
G
P
Res
G
P
S.
GROUP ACTIONS ON POLYNOMIAL RINGS 229
Remark. It is not hard to see that, given any degree m, the summ ands of
¯
F can be moved by multiplication by a scalar to lie in T
m+lc
= S
m+lc
· · ·
S
m+(l+1)c1
, for some l. The argument of the proof of 1.2 now shows that
the non-free part of T
n
has dimension bounded by a polynomial of degree
d 2 (cf. [2, 3]).
4. Filtered Rings.
The case of filtered rings is slightly different. Consider the power series ring
k[[x]] in characteristic 2 and let the group of order 2 act by x 7→ x/(x + 1) =
x + x
2
+ x
3
+ · · · . The action on the associated graded ring is trivial, yet
the action on k[[x]] certainly c ontains free summands (the only alternative
is trivial).
We consider finitely generated k-algebras S which are integral domains
and have a filtration S = I
0
I
1
I
2
· · · . Each S/I
n
is assumed to
be finite dimensional over k, and I
n
= {0}. There is a finite group G
of automorphisms of S, which preserves the filtration. The invariants are
R = S
G
with the induced filtration J
n
= R I
n
. Again there is a function
χ
S
(n) = c
d
(n)n
d
+ c
d1
n
d1
+ · · · c
0
(n),
where the c
i
are periodic, such that dim
k
(S/I
n
) = χ
S
(n) for large n. If S is
a power series ring, then d is equal to the number of variables.
As before there is a free kG-module of rank 1 in S, and the product map
R
k
F S is injective. Since F is finite dimensional there is s ome integer
f such that F I
f
= 0, so F injects into S/I
f
.
For each n, let K
n
be a vector space complement to J
n
in R. Then the
product map K
n
F S/I
f+n
is injective, so its image, K
n
F is a free
summand of S/I
f+n
.
Proceeding in the same way as before we can prove:
Theorem 4.1. dim
k
((S/I
f+n
)/K
n
F ) is bounded by a polynomial of degree
d 1.
So S is mos tly free. Again, for a power series ring, the non-free part grows
like a power series ring in one fewer variables.
References
[1] M.F. Atiyah and I.G. Macdonald, Introduction to Commutative Algebra, Addison-
Wesley, Reading, MA, 1969.
[2] R.M. Bryant, Symmetric powers of representation of finite groups, Jour. Algebra, 154
(1993), 416-436.
[3] , Groups acting on polynomial algebras, Finite and locally finite groups. NATO
Advanced Science Institutes Series C: Mathematical and Physical Sciences, 471, 327-
346, Kluwer Acad. Publ. Dordrecht, 1995.
230 PETER SYMONDS
[4] R. Howe, Asymptotics of dimensions of invariants for finite groups, Jour. Algebra,
122 (1989), 374-379.
[5] O. Zariski and P. Samuel, Commutative Algebra, Vol. II, Van Nostrand, Princeton,
1960.
Received October 13, 1998.
Department of Mathematics
U.M.I.S.T.
Manchester M60 1QD
England
E-mail address: psymonds@umist.ac.uk
PACIFIC JOURNAL OF MATHEMATICS
Vol. 195, No. 1, 2000
THE MODULI OF FLAT PU(2,1) STRUCTURES ON
RIEMANN SURFACES
Eugene Z. Xia
For a compact Riemann surface X of genus g > 1,
Hom(π
1
(X) , PU(p, q))/PU(p, q) is the moduli space of flat
PU(p, q)-connections on X. There are two integer invari-
ants, d
P
, d
Q
, associated with each σ Hom(π
1
(X) , PU(p, q))/
PU(p, q). These invariants are related to the Toledo invari-
ant τ by τ = 2
qd
P
pd
Q
p+q
. This paper shows, via the theory of
Higgs bundles, that if q = 1, t hen 2(g 1) τ 2(g 1).
Moreover, Hom(π
1
(X) , PU(2, 1))/PU(2, 1) has one connected
component corresponding to each τ
2
3
Z with 2(g 1)
τ 2(g 1). Therefore the total number of connected com-
ponents is 6(g 1) + 1.
1. Introduction.
Let X be a smooth projective curve over C with genus g > 1. The deforma-
tion space
CN
B
= Hom
+
(π
1
(X), PGL(n, C))/ PGL(n, C)
is the space of equivalence classes of semi-simple PGL(n, C)-represe ntations
of the fundamental group π
1
(X). This is the PGL(n, C)-Betti moduli space
on X [22, 23, 24]. A theorem of Corlette, Donaldson, Hitchin and Simpson
relates CN
B
to two other moduli spaces, CN
DR
and CN
Dol
—the PGL(n, C)-
de Rham and the PGL(n, C)-Dolbeault moduli spaces, respectively [3, 5, 11,
21]. The Dolbeault moduli space consists of holomorphic objects (Higgs
bundles) over X; therefore, the classical results of analytic and algebraic
geometry can be applied to the study of the Dolbeault mo duli space.
Since PU(p, q) PGL(n, C), CN
B
contains the space
N
B
= Hom
+
(π
1
(X), PU(p, q))/ PU(p, q).
The space N
B
will be referred to as the PU(p, q)-Betti moduli space which
similarly corresponds to some subspaces N
DR
and N
Dol
of CN
DR
and CN
Dol
,
respectively. We shall refer to N
DR
and N
Dol
as the PU(p, q)-de Rham and
the PU(p, q)-Dolbe ault moduli spaces.
The Betti moduli spaces are of gre at interest in the field of geometric
topology and uniformization. In the case of p = q = 1, Goldman analyzed
231
232 EUGENE Z. XIA
N
B
and determined the number of its connected c omponents to be 4g 3
[6]. Hitchin subsequently considered N
Dol
in the case of p = q = 1 and
determined its topology [11].
In this paper, we analyze N
Dol
for the case of p = 2, q = 1 and deter-
mine its number of connected components. In addition, we produce a new
algebraic proof, via the Higgs-bundle theory, of a theorem by Toledo on the
bounds of the Toledo invariant [26, 27].
An element σ Hom
+
(π
1
(X), PU(p, q)) defines a flat principal PU(p, q)-
bundle P over X. Such a flat bundle may be lifted to a principal U(p, q)-
bundle
ˆ
P with a Yang-Mills connection D [2, 3, 5, 11, 21]. Let E be the
rank-(p + q) vector bundle associated with (
ˆ
P , D). The second c ohomology
H
2
(X, Z) is isomorphic to Z, so one may identify the Chern class c
1
(E)
H
2
(X, Z) with an integer, the degree of E. Suppose we impose the additional
condition
0 deg(E) < n.
Then the above construction give s rise to a unique obstruction class o
2
(E)
H
2
(X, π
1
(U(p, q))) [25]. The obstruction class is invariant under the conju-
gation action of PU(p, q); therefore, one obtains the obstruction map:
o
2
: Hom
+
(π
1
(X), PU(p, q))/ PU(p, q) H
2
(X, π
1
(U(p, q)))
=
Z × Z.
The maximum compact subgroup of U(p, q) is U(p) × U(q). Hence topo-
logically E is a direct sum E
P
E
Q
with
deg(E) = deg(E
P
) + deg(E
Q
).
The obstruction class o
2
(E) is then (deg(E
P
), deg(E
Q
)) Z ×Z. Associated
with σ is the Toledo invariant τ which relates to d
P
= deg(E
P
) and d
Q
=
deg(E
Q
) by the formula [7, 26, 27]
τ = 2
deg(E
P
E
Q
)
p + q
= 2
qd
P
pd
Q
p + q
.
This explains why the Toledo invariant of a PU(2, 1) representation cannot
be an odd integer [7]. The main result presented here is the following:
Theorem 1.1. Hom
+
(π
1
(X), PU(2, 1))/ PU(2, 1) has one connected com-
ponent for each τ
2
3
Z with 2(g 1) τ 2(g 1). Therefore the total
number of connected components is 6(g 1) + 1.
We shall also provide a new proof en route to the following theorem:
Theorem 1.2 (Toledo). Suppose σ Hom
+
(π
1
(X), PU(p, 1)) and τ is the
Toledo class of σ. Then
2(g 1) τ 2(g 1).
Moreover τ = ±2(g 1) implies σ is reducible.
FLAT PU(2,1) STRUCTURES ON RIEMANN SURFACES 233
These results are related to the results of Domic and Toledo [4, 26, 27]
and, as being pointed out to the author recently, are also related to the work
of Gothen [8] which computed the Poincar´e polynomials for the components
of Hom(π
1
(X), PSL(3, C))/ PSL(3, C), where deg(E) is coprime to 3.
Acknowledgments.
Most of this research was c arried out while the author was at the Univer-
sity of Maryland at College Park. I thank J. Adams, K. Coombes, P. Green,
K. Joshi, S. Kudla, P. Newstead, J. Poritz and especially W. Goldman and
C. Simpson for insightful discussions over the course of the research. I thank
the referee for helpful suggestions.
2. Backgrounds and Preliminaries.
In this section, we briefly outline the constructions of the Betti, de Rham
and Dolbeault moduli spaces. For details, see [2, 3, 5, 11, 12, 18, 21, 22,
23, 24].
2.1. The Betti Moduli Space. The fundamental group π
1
(X) is gener-
ated by S = {A
i
, B
i
}
g
i=1
, subject to the relation
g
Y
i=1
A
i
B
i
A
1
i
B
1
i
= e.
Denote by I and [I] the identities of GL(n, C) and PGL(n, C), respectively.
Define
R : PGL(n, C)
2g
PGL(n, C)
R : GL(n, C)
2g
GL(n, C)
to be the commutator maps:
(X
1
, Y
1
, . . . , X
g
, Y
g
)
R,R
g
Y
i=1
X
i
Y
i
X
1
i
Y
1
i
.
The group
{ζI : ζ C, ζ
n
= 1}
is isomorphic to Z
n
. The space R
1
(Z
n
) is identified with the representation
space Hom(Γ, GL(n, C)), where Γ is the central extension [2, 11]:
0 Z
n
Γ π
1
(X) 0.
Each element ρ R
1
(Z
n
) acts on C
n
via the standard representation of
GL(n, C). The representation ρ is called reducible (irreducible) if its action
on C
n
is reducible (irreducible). A representation ρ is called semi-simple if
it is a direct sum of irreducible representations. Let ζ
1
= e
2πi/n
and define
CM
B
(c) = {σ R
1
(ζ
c
1
I) : σ is semi-simple}/ GL(n, C),
234 EUGENE Z. XIA
CM
B
=
n1
[
c=0
CM
B
(c),
CN
B
(c) = CM
B
(c)/ Hom(π
1
(X), C
)
= Hom
+
(π
1
(X), PGL(n, C))/ PGL(n, C).
Fix p, q such that p + q = n. Denote by R
U
the restriction of R to the
subgroup U(p, q)
2g
. Define
M
B
(c) = {σ R
1
U
(ζ
c
1
I) : σ is semi-simple}/ U(p, q),
M
B
=
n1
[
c=0
M
B
(c).
Note the center of U(p, q) is U(1) and is contained in the center of GL(n, C).
It follows that M
B
(c) CM
B
(c). Define
N
B
(c) = M
B
(c)/ Hom(π
1
(X), U(1))
N
B
= M
B
/ Hom( π
1
(X), U(1)) = Hom
+
(π
1
(X), U(p, q))/ U(p, q).
All the spaces constructed here that contain the symbols M
B
or N
B
will
be loosely referred to as Betti moduli spaces. The subspace of irreducible
elements of a Betti moduli space will be denoted by an s superscript. For
example, CM
s
B
denotes the subspace of irreducible elements of CM
B
.
2.2. The de Rham Moduli Space. Suppose P is a principal GL(n, C)-
bundle on X, E its associated vector bundle of rank n and G
C
(E) the group
of GL(n, C)-gauge transformations on E. A connection is called Yang-Mills
(or central) if its curvature is c entral [2]. The gauge group G
C
(E) acts
on the space of GL(n, C)-connections on E and preserves the subspace of
Yang-Mills connections. Fix E with deg(E) = c. The de Rham moduli space
CM
DR
(c) on E is defined to be the G
C
(E)-equivalence classes of Yang-Mills
connections.
Let M
DR
(c) denote the space of U(p, q)-gauge equivalence classes of
U(p, q)-central connections on E. In other words, M
DR
(c) is constructed as
CM
DR
(c), but with U(p, q) replacing GL(n, C). Since the center of U(p, q)
is contained in the center of GL(n, C), M
DR
(c) CM
DR
(c).
The space of C
-gauge equivalence classes of C
-connections on X is
H
1
(X, C
) which acts on CM
DR
(c) [2]. Denote the quotient CN
DR
(c).
This action corresponds to the action of Hom(π
1
(X), C
) on CM
B
(c) and
the quotient CN
DR
(c) corresponds to CN
B
(c). Similarly, the space of U(1)-
gauge equivalence classes of U(1)-connections on X is H
1
(X, U(1)) which
acts on M
DR
(c) and the quotient is denoted by N
DR
(c). De fine
CM
DR
=
[
c=−∞
CM
DR
(c), CN
DR
=
[
c=−∞
CN
DR
(c)
FLAT PU(2,1) STRUCTURES ON RIEMANN SURFACES 235
M
DR
=
[
c=−∞
M
DR
(c), N
DR
=
[
c=−∞
N
DR
(c).
All the spaces constructed here that contain the symbols M
DR
or N
DR
will be loosely referred to as de Rham mo duli spaces. A central connection
is irreducible if (E, D) = (E
1
E
2
, D
1
D
2
) implies rank(E
1
) = 0 or
rank(E
2
) = 0. The subspace of irreducible elements of a de Rham moduli
space will be denoted by an s superscript.
Theorem 2.1. The moduli space CM
B
(c) is homeomorphic to CM
DR
(c).
Proof. See [3, 5, 11].
Consider all the objects we have defined so far with subscripts B or DR.
With Theorem 2.1, one can verify the following: Suppose two objects have
subscripts B or DR. Then the two objects are homeomorphic if they only
differ in subscripts. For example, N
B
(c) is homeomorphic to N
DR
(c).
Since the maximum compact subgroup of U(p, q) is U(p) × U(q), (E, D)
M
DR
implies E is a direct sum of a U(p) and a U(q)-bundle:
E = E
p
E
q
,
where the ranks of E
p
and E
q
are p and q, respectively. Therefore, associated
to each (E, D) are the invariants
d
P
= deg(E
P
) and d
Q
= deg(E
Q
),
with
d
P
+ d
Q
= deg(E) = c.
The Toledo invariant τ is [7, 26, 27]
τ = 2
deg(E
P
E
Q
)
n
= 2
qd
P
pd
Q
n
.
The subspace of M
DR
(c) with a fixed Toledo invariant τ is denoted by M
τ
DR
.
By the equivalence of Betti and de Rham moduli spaces, one may define the
Toledo invariant on M
B
(c). Denote by M
τ
B
the subspace of M
B
(c) with
a fixed Toledo invariant τ. The H
1
(X, U(1)) action on M
DR
(c) preserves
M
τ
DR
and the quotient is denoted by N
τ
DR
. In the Betti moduli space, the
Hom(π
1
(X), U(1)) action on M
B
preserves M
τ
B
, and the quotient is denoted
by N
τ
B
.
2.3. The Dolbeault Moduli Space. Let E be a rank n complex vector
bundle over X with deg(E) = c. Denote by the canonical bundle on
X. A holomorphic structure on E induces holomorphic structures on the
bundles End(E) and End(E) Ω. A Higgs bundle is a pair (E
, Φ), where
is a holomorphic structure on E and Φ H
0
(X, End(E
) Ω). Such a Φ
is called a Higgs field. We denote the holomorphic bundle E
by V .
236 EUGENE Z. XIA
Define the slope of a Higgs bundle (V, Φ) to be
s(V ) = deg(V )/ rank(V ).
For a fixed Φ, a holomorphic subbundle W V is said to be Φ-invariant if
Φ(W ) W Ω. A pair (V, Φ) is stable (semi-stable) if W V is Φ-invariant
implies
s(W ) < ()s(V ).
A Higgs bundle is called poly-stable if it is a direct sum of stable Higgs
bundles of the same slope [11, 22].
The gauge group G
C
(E) acts on holomorphic structures by pull-back and
on Higgs fields by conjugation. Moreover the G
C
(E) action preserves stabil-
ity, poly-stability and semi-stability. The Dolbeault moduli space CM
Dol
(c)
on E (with deg(E) = c), is the G
C
(E)-equivalence classes of poly-stable
(or S-equivalence classes of semi-stable [18]) Higgs bundles (V, Φ) on X
[11, 12, 18, 22]. A Higgs bundle is called reducible if it is poly-stable but
not stable. Let
CM
Dol
=
[
c=−∞
CM
Dol
(c).
If D CM
DR
(c), then for any Hermitian metric h on E, there is a
decomposition,
D = D
A
+ Ψ,
where D
A
is compatible with h and Ψ is a 1-form with coefficients in p. The
(0, 1) part of D
A
determines a holomorphic structure
A
on E while the
(1, 0) part of Ψ is a section of the bundle End(E) Ω. There exists a metric
h such that the pair
(V, Φ) = (E
A
, Ψ
1,0
)
so constructed is a poly-stable Higgs bundle [11, 21, 22]. Therefore this
construction gives a map
f : CM
DR
(c) C M
Dol
(c).
Theorem 2.2 (Corlette, Donaldson, Hitchin, Simpson). The map f is a
homeomorphism.
Proof. See [3, 5, 11, 21].
3. The U(p, q)-Yang-Mills Connections.
Assume p q and p + q = n. From the previous section, we know that
M
DR
CM
DR
. Let D CM
DR
(c) be a GL(n, C)-Yang-Mills connection
on a rank n vector bundle
E X.
FLAT PU(2,1) STRUCTURES ON RIEMANN SURFACES 237
Proposition 3.1. D is a U(p, q)-Yang-Mills connection if and only if its
corresponding Higgs bundle (V, Φ) CM
Dol
(c) satisfies the following two
conditions:
1) V is decomposable into a direct sum:
V = V
P
V
Q
,
where V
P
, V
Q
are of rank p, q, respectively.
2) The Higgs field decomposes into two maps:
Φ
1
: V
P
V
Q
,
Φ
2
: V
Q
V
P
.
Proof. Suppose D is a U(p, q)-Yang-Mills connection. Denote by h the
Hermitian-Yang-Mills metric on (E, D). Then D decomposes as
D = D
A
+ Ψ,
where D
A
is the part compatible with h. The Cartan decomposition (g =
k p) for u(p, q) is
u(p, q) = (u(p) u(q)) p.
If we take the standard representation of u(p, q), then elements in k are of
the form
a 0
0 d
where a u(p), b u(q), respectively. The elements in p are then of the
form
0 b
c 0
,
where b Hom(C
q
, C
p
), c Hom(C
p
, C
q
), respectively. Hence on local
charts, D
A
and Ψ have coefficients in k and p, respectively. In particular,
the connection D
A
is reducible.
The Higgs bundle corresponding to D is (E
A
, Φ) where
A
is the (0, 1)-
part of D
A
and Φ, the (1, 0)-part of Ψ, is considered as a holomorphic bundle
map:
Φ : V V .
Since D
A
has coefficient in k, the holomorphic structure on V defined by
D
0,1
A
is a direct sum:
V = V
P
V
Q
.
Since Ψ is block off-diagonal, Φ is also block off-diagonal implying Φ can be
decomposed into two maps:
Φ
1
: V
P
V
Q
,
Φ
2
: V
Q
V
P
.
This proves the only if part of the proposition.
238 EUGENE Z. XIA
Suppose (V, Φ) is a Higgs bundle that satisfies the two conditions of Propo-
sition 3.1. Let α be the constant gauge
α =
I
p
0
0 I
q
,
where I
p
, I
q
are p × p, q × q identity matrices, respectively. Then α acts on
the space of holomorphic structures on E and fixes V . Moreover,
αΦα
1
= Φ
since Φ is of the form
Φ =
0 Φ
1
Φ
2
0
.
Hence by a theorem of Simpson, the corresponding Hermitian-Yang-Mills
metric h is invariant under the action of α [21]. In other words, on local
charts, h is a Hermitian matrix of the form
h =
a 0
0 d
,
where a, d are Hermitian matrices of dimension p × p, q × q, respectively.
Hence the corresponding Yang-Mills connection is
D = D
A
+ Φ + Φ
,
where Φ
is the adjoint of Φ with respect to h. In local coordinates, D
A
has
coefficient of the form
a 0
0 d
and Φ + Φ
is of the form
0 b
b
0
.
Hence D
A
and Φ + Φ
have coefficients in u(p) u(q) and p, respectively.
This implies D is a U(p, q)-Yang-Mills connection.
Denote by M
Dol
(c) the subspace of CM
Dol
(c) satisfying the hypothesis of
Proposition 3.1. Then M
Dol
(c) is homeomorphic to M
DR
(c).
The invariants d
P
, d
Q
and τ on (E, D) translate to invariants on the
corresponding U(p, q)-Higgs bundles (V
P
V
Q
, Φ):
d
P
= deg(V
P
), d
Q
= deg(V
Q
), τ = 2
qd
P
pd
Q
n
.
The subspace of M
Dol
(c) with a fixed Toledo invariant τ is denoted by M
τ
Dol
.
FLAT PU(2,1) STRUCTURES ON RIEMANN SURFACES 239
4. Group Actions and ahler Structures on CM
Dol
.
4.1. The Action of line bundles. The space of holomorphic line bundles,
H
1
(X, O
), acts freely on CM
Dol
as follows:
H
1
(X, O
) × CM
Dol
7− CM
Dol
,
(L, (V, Φ)) 7− (V L, Φ 1),
where 1 is the identity map on L. An immediate consequence is:
Proposition 4.1. If c
1
c
2
mod n, then CM
Dol
(c
1
) is homeomorphic to
CM
Dol
(c
2
).
4.2. The Action of H
0
(X, Ω). The vector space H
0
(X, Ω) acts freely on
CM
Dol
as follows:
H
0
(X, Ω) × CM
Dol
7− CM
Dol
,
(φ, (V, Φ)) 7− (V, Φ + φI).
The actions of H
1
(X, O
) and H
0
(X, Ω) commute and the quotient is
defined to be
CN
Dol
= CM
Dol
/(H
1
(X, O
) × H
0
(X, Ω)).
The H
1
(X, O
) action preserves the subspaces M
Dol
(c) and M
τ
Dol
. The
quotients are defined to be
N
Dol
(c) = M
Dol
(c)/H
1
(X, O
),
N
τ
Dol
= M
τ
Dol
/H
1
(X, O
).
All the spaces constructed so far that contain the symbols M
Dol
or N
Dol
will be loosely referred to as the Dolbeault moduli spaces. The subspace
of stable Higgs bundles of a Dolbeault moduli space will be denoted by an
s superscript. For example, CM
s
Dol
will denote the subspace of irreducible
elements of CM
Dol
.
Remark 1. The Betti, de Rham and Dolbeault moduli spaces CM
B
,
CM
Dol
and CM
Dol
constructed here are variations of those of Simpson’s
[22, 23, 24].
With Theorems 2.1 and 2.2, one can obtain the following equivalence
relations between the various Betti, de Rham and Dolbeault moduli spaces.
Corollary 4.2. Suppose M
τ
DR
M
DR
(c). Then one obtains the following
commutative diagram:
M
τ
B
M
B
(c) CM
B
(c)
y
y
y
M
τ
DR
M
DR
(c) CM
DR
(c)
y
y
y
M
τ
Dol
M
Dol
(c) CM
Dol
(c).
240 EUGENE Z. XIA
Moreover the horizontal maps are continuous injections and vertical maps
are homeomorphisms. One obtains three additional commutative diagrams
by respectively replacing the symbol M by M
s
, N and N
s
in the above
diagram. In the case of M
s
, the maps in the commutative diagram are
smooth.
4.3. The Dual Higgs Bundles. There is a Z
2
action on CM
Dol
. Let
(V, Φ) CM
Dol
where Φ is a holomorphic map:
Φ : V V .
This induces a map on the dual bundles
Φ
: V
V
.
Tensoring with Ω,
Φ
1 : V
V
,
where 1 denotes the identity map on Ω. This produces the dual Higgs bundle
(V
, Φ
1). We shall abbreviate it as (V
, Φ
).
Proposition 4.3. If (V, Φ) CM
Dol
(c), then (V
, Φ
) CM
Dol
(c).
Proof. One must show that (V, Φ) is stable (semi-stable) implies (V
, Φ
) is
stable (semi-stable). Suppose W
1
V
is Φ
-invariant. Then we have the
following commutative diagram
0 W
1
V
W
2
0
y
Φ
y
Φ
y
Φ
0 W
1
V
W
2
0
where W
2
= V
/W
1
. The proposition follows by dualizing the diagram.
In light of Propositions 4.1 and 4.3 we have:
Corollary 4.4. If c
2
= ±c
1
mod n, then CM
Dol
(c
1
) is homeomorphic to
CM
Dol
(c
2
).
4.4. The U(1) and C
-Actions on the Complex Moduli Spaces. If
(V, Φ) CM
Dol
(c), then for t C
, (V, tΦ) CM
Dol
(c). This defines an
analytic action [11, 12, 22]
C
× CM
Dol
(c) 7− C M
Dol
(c).
Since U(1) C
, this also induces a U(1)-action on CM
Dol
(c).
FLAT PU(2,1) STRUCTURES ON RIEMANN SURFACES 241
4.5. The Moment Map. The moduli space CM
Dol
(c)
s
is ahler [11, 12].
Denote by i, ω the corresponding complex and symplectic structures, respec-
tively. Define the Morse function [11, 12]
m : CM
Dol
(c)
s
R,
m(V, Φ) = 2i
Z
X
tr(ΦΦ
),
where Φ
is the adjoint of Φ with respect to the Hermitian-Yang-Mills metric
on (E, D). Denote by X the vector field on CM
Dol
(c)
s
such that [12]
grad m = iX.
Theorem 4.5.
1) The map m is proper.
2) The U(1)-action generates X.
3) The C
action is analytic with respect to i; therefore, the orbit of C
is locally an analytic subvariety with respect to i.
Proof. See [11, 12, 22].
Corollary 4.6. Each component of CM
Dol
(c) contains a point that is a
local minimum of m.
Corollary 4.7. If the C
action preserves M CM
Dol
(c)
s
, then the gra-
dient flow grad m preserves M.
Let m
r
be the restriction of m to the subspace M
τ
Dol
CM
Dol
(c).
Corollary 4.8. Every component of M
τ
Dol
contains a point that is a local
minimum of m
r
. If (V, Φ) is stable and is a local minimum of m
r
, then
(V, Φ) is a critical point of m.
Proof. Consider
M
τ
B
M
B
(c) CM
B
(c).
Since U(p, q) is closed in GL(n, C), M
B
(c) is a closed subspace of CM
B
(c).
Since the obstruction map o
2
is continuous, M
τ
B
is a closed subspace of
M
B
(c). Hence M
τ
B
is closed in CM
B
(c). Hence by Theorem 4.5, m
r
is
proper. Thus each c omponent of M
τ
Dol
contains a local minimum of m
r
.
The points in (M
τ
Dol
)
s
are smooth. Suppose (V, Φ) (M
τ
Dol
)
s
. Then
(V, Φ) is of the form described in Propos ition 3.1. Hence the C
action
preserves the subspace (M
τ
Dol
)
s
CM
s
Dol
. By Corollary 4.7, the gradient
flow of m preserves (M
τ
Dol
)
s
. Hence
grad m
r
= grad m = iX.
If m
r
is a local minimum at (V, Φ), then
grad m(V, Φ) = grad m
r
(V, Φ) = 0.
Hence (V, Φ) is a c ritical point of m.
242 EUGENE Z. XIA
5. Bounds on Invariants.
In this section, we assume q = 1 and let n = p + q = p + 1. In light of
Proposition 4.3 and Corollary 4.4, one may further as sume that τ 0 and
0 c < n, or equivalently,
s(V
Q
) s(V ) s(V
P
), 0 c < n.
Proposition 5.1. If (V, Φ) = (V
P
V
Q
,
1
, Φ
2
)) M
Dol
(c)
s
(M
Dol
(c)),
then
d
P
< ()
c(n 1)
n
+ (g 1)
d
Q
> ()
c
n
(g 1).
Proof. Suppose (V
P
V
Q
, Φ) M
Dol
(c)
s
with Φ =
1
, Φ
2
) in the notation
of Proposition 3.1. Since s(V
P
) s(V ),
Φ
1
: V
P
V
Q
is non-zero.
Construct the canonical factorization for Φ
1
[20]: There exist holomorphic
bundles V
1
, V
2
and W
1
, W
2
such that the following diagram
0 V
1
f
1
V
P
f
2
V
2
0
Φ
1
y
ϕ
y
0 W
2
g
2
V
Q
g
1
W
1
0
commutes, and the rows are exact, rank(V
2
) = rank(W
1
) and ϕ has full rank
at a generic point of X. This implies
(
deg(V
1
) + deg(V
2
) = d
P
deg(W
1
) + deg(W
2
) = d
Q
+ 2(g 1).
Since Φ
1
6≡ 0, we have ϕ 6≡ 0, rank(W
2
) = 0 and W
1
= V
Q
Ω.
The case of p = 1 has been dealt with by Hitchin [11], so we assume p > 1.
Then V
1
is a Φ-invariant subbundle of positive rank. Stability implies
s(V
1
) < s(V ) = (d
P
+ d
Q
)/n = c/n.
Since the map
V
2
ϕ
W
1
= (V
Q
Ω)
is not trivial,
deg(V
2
) deg(W
1
) = deg(V
Q
Ω).
FLAT PU(2,1) STRUCTURES ON RIEMANN SURFACES 243
So one has
s(V
1
) < s(V )
d
P
= deg(V
1
) + deg(V
2
)
deg(V
2
) d
Q
+ 2(g 1).
This implies
d
P
<
(n 2)c
n
+ d
Q
+ 2(g 1).
Since d
P
+ d
Q
= c,
d
P
<
c(n 1)
n
+ (g 1)
and
d
Q
>
c
n
(g 1).
When (V, Φ) is semi-stable, one has either Φ 6≡ 0 or Φ 0. In the former
case, one has s(V
1
) s(V ) implying
d
P
c(n 1)
n
+ (g 1)
d
Q
c
n
(g 1).
In the latter case , V
p
is Φ-invariant. By the assumption s(V
Q
) s(V
P
),
d
P
= d
Q
= 0 and τ = 0.
By definition,
τ = 2
d
P
pd
Q
n
2
n
c(n 1)
n
+ (g 1) (n 1)
c
n
+ (n 1)(g 1)
= 2(g 1).
Equality holds only when (V, Φ) is semi-stable but not stable, in which case,
the associated flat connection is reducible. This proves Theorem 1.2.
6. Reducible Higgs Bundles.
Let p = 2 and q = 1 and assume τ 0 and 0 c < 3. By definition, a
reducible poly-stable Higgs bundle is a direct sum of stable Higgs bundles
of the same slope. These Higgs bundles correspond to the reducible repre-
sentations in M
B
. A direct computation shows that if (V, Φ) is reducible,
then
deg(V ) = d
P
+ d
Q
= 0
and the associated Toledo invariant τ is an even integer. Hence one has:
244 EUGENE Z. XIA
Proposition 6.1. If c = deg(V ) 6= 0 and (V, Φ) M
Dol
(c), then (V, Φ) is
stable. In particular, M
Dol
(c) is smooth.
An example of a reducible Higgs bundle is (O
1
2
1
2
, Φ), where
Φ :
1
2
1
2
is a holomorphic bundle isomorphism. That is, Φ is of the form
0 0 0
0 0 0
0 1 0
.
The Toledo invariant in this case is 2(g 1). All the flat U(2, 1)-connections
with τ = 2(g 1) are reducible by Proposition 5.1. The fact that there
is no irreducible deformation for the U(2, 1)-connections with τ = 2(g 1)
was first demonstrated by Toledo [26]. In particular, this component is
connected [6, 11].
7. Hodge Bundles and Deformation.
Let p = 2 and q = 1 and assume τ 0 and 0 c < 3. A Hodge bundle on
X is a direct sum of holomorphic bundles [22]
V =
M
s,t
V
s,t
together with holomorphic maps (Higgs field)
Φ
i
: V
s,t
V
s1,t+1
.
An immediate consequence of Proposition 3.1 is:
Corollary 7.1. Suppose (V
P
V
Q
,
1
, Φ
2
)) M
Dol
(c) (in the notations
of Proposition 3.1). Then (V
P
V
Q
,
1
, Φ
2
)) is a Hodge bundle if and only
if (V
P
V
Q
,
1
, Φ
2
)) is either binary or t ernary in the following sense:
1) Binary: Φ
2
0.
2) Ternary: V
P
= V
1
V
2
and the Higgs field consists of two maps:
Φ
1
: V
2
V
Q
,
Φ
2
: V
Q
V
1
.
Denote by B(d
P
, d
Q
) the space of all poly-stable (or S-equivalence classes
of semi-stable) binary Hodge bundles (V
P
V
Q
,
1
, 0)) with deg(V
P
) = d
P
and deg(V
Q
) = d
Q
. Denote by T (d
1
, d
2
, d
Q
) the space of all poly-stable
(or S-equivalence classes of semi-stable) ternary Hodge bundles (V
1
V
2
V
Q
,
1
, Φ
2
)) with deg(V
1
) = d
1
, deg(V
2
) = d
2
and deg(V
Q
) = d
Q
. Denote
the subspaces of stable Hodge bundles by B(d
P
, d
Q
)
s
, T (d
1
, d
2
, d
Q
)
s
. When
τ is not an integer, these are the type (2,1) and (1,1,1) spaces in [8]. Note
the (1,2) typ e s give τ < 0 and therefore need not be considered here.
FLAT PU(2,1) STRUCTURES ON RIEMANN SURFACES 245
Proposition 7.2. Every stable binary Hodge bundle in (M
τ
Dol
)
s
may be
deformed to a stable ternary Hodge bundle within M
τ
Dol
.
A family (or flat family) of Higgs pairs (V
Y
, Φ
Y
) is a variety Y such
that there is a vector bundle V
Y
on X × Y together with a section Φ
Y
Γ(Y, (π
Y
)
(π
X
End(V
Y
))) [18]. CM
Dol
being a moduli space implies that
if Y is a family of stable (poly-stable or S-equivalence classes of semi-stable)
Higgs bundles, then there is a natural morphism [15, 17]
t : Y CM
Dol
.
Moreover t takes every point y Y to the point of CM
Dol
that corresponds
to the Higgs bundle in the family over y [15, 17, 18].
The s pace M
Dol
(c) is a subvariety of CM
Dol
(c); hence, to show that two
stable (poly-stable or S-equivalence classes of semi-stable) Higgs bundles
(V
1
, Φ
1
) and (V
2
, Φ
2
) belong to the same component of M
Dol
(c), it suffices
to exhibit a connected family Y (within M
Dol
(c)) of stable (poly-stable or
S-equivalence classes of semi-stable) Higgs bundles containing both (V
1
, Φ
1
)
and (V
2
, Φ
2
).
Proof. Suppose (V, Φ) = (V
P
V
Q
,
1
, 0)) B(d
P
, d
Q
)
s
(M
τ
Dol
)
s
. Since
s(V
P
) s(V ) (This is due to the assumption τ 0, and 0 c < 3), Φ
1
6≡ 0.
Construct the canonical factorization for Φ
1
:
0 V
1
f
1
V
P
f
2
V
2
0
Φ
1
y
ϕ
y
0 W
2
g
2
V
Q
g
1
W
1
0 .
V
1
being Φ
1
invariant implies
deg(V
1
) = s(V
1
) < s(V ) s(V
P
) s(V
2
) = deg(V
2
).
The space P ic
0
(X) of line bundles of degree 0 over X is identified with
the Jacobi variety J
0
(X). Construct the universal bundle [2, 19]
U X × J
0
(X)
such that U restricts to the bundle L V
1
V
1
2
on (X, L). Let π be the
projection
π : X × J
0
(X) J
0
(X).
Applying the right derived functor R
1
to π gives the sheaf F = R
1
π
(U)
[10] such that
F|
L
= H
1
(X, L V
1
V
1
2
).
Since
deg(L V
1
V
1
2
) = deg(V
1
) deg(V
2
) < 0,
246 EUGENE Z. XIA
by Riemann-Roch,
h
1
(L V
1
V
1
2
) = h
0
(L V
1
V
1
2
) deg(L V
1
V
1
2
) + (g 1)
= deg(V
2
) deg(V
1
) + (g 1)
is a constant. By Grauert’s theorem, F is locally free, hence, is associated
with a vector bundle
F 7− J
0
(X)
of rank deg(V
2
)deg(V
1
)+(g1). In particular the total space F is smooth
and parameterizes extensions [9, 10]:
0 L V
1
f
3
W
P
f
4
V
2
0
for fixed V
1
, V
2
. Tensoring the above sequence with gives:
0 L V
1
g
3
W
P
g
4
V
2
0.
Fix ϕ. Then F also parameterizes a family of Higgs bundles (W
P
, Φ
0
1
) that
fit into the factorization
0 L V
1
f
3
W
P
f
4
V
2
0
Φ
0
1
y
ϕ
y
0 W
2
g
2
V
Q
g
1
W
1
0 .
Let V F be the subset of stable extensions (i.e., W
P
V implies W
P
is
a stable holomorphic bundle [19]).
Lemma 7.3. V H
1
(L V
1
V
1
2
) and V are open and dense in H
1
(L
V
1
V
1
2
) and F , respectively. Moreover if W
P
V, then (W
P
V
Q
,
0
1
, 0))
is stable.
Proof. Since deg(L V
1
) < deg(V
2
) for each L J
0
(X), by a theorem of
Lange and Narasimhan [13], there always exists a stable extension W
P
H
1
(L V
1
V
1
2
). In addition, a theorem of Maruyama states that being
stable is an open prope rty [14]. The open dense property follows from the
smoothness of F and H
1
(L V
1
V
1
2
).
Let p
P
, p
Q
be the projections of W
P
V
Q
onto its W
P
and V
Q
factors,
respectively. Suppose W is
0
1
, 0)-invariant. Suppose W has rank 1. If
P
Q
(W ) = 0, then W = L V
1
; otherwise, deg(W ) deg(V
Q
). In either
case, s(W ) < s(V ). Suppose W has rank 2. If p
Q
(W ) = 0, then W = W
P
and s(W ) < s(V ). Suppose P
Q
(W ) 6= 0. Then there exists a line bundle L
1
such that
0 L
1
W
p
Q
p
Q
(W ) 0.
Now let L
P
= p
P
(L
1
) W
P
. Then
deg(W ) = deg(L
1
) + deg(p
Q
(W )) deg(L
P
) + deg(V
Q
).
FLAT PU(2,1) STRUCTURES ON RIEMANN SURFACES 247
Since W
P
is stable, s(L
P
) < s(W
P
). By the assumptions τ 0 and 0 c <
3, one has s(V
Q
) 0 and s(W
P
) 0. Therefore,
s(W ) s(L
P
V
Q
) =
s(L
P
) + s(V
Q
)
2
<
s(W
P
) + s(V
Q
)
2
=
deg(W
P
)
4
+
deg(V
Q
)
2
deg(W
P
) + deg(V
Q
)
3
= s(V ).
Thus (W
P
V
Q
,
0
1
, 0)) is stable.
Since Φ
1
6≡ 0, deg(V
2
) d
Q
+ 2(g 1) and
deg(V
1
) = d
P
deg(V
2
) d
P
d
Q
2(g 1).
Hence
deg(V
1
Q
V
1
Ω) = d
Q
+ deg(V
1
) + 2(g 1) d
P
2d
Q
> 0.
Hence there exists L
0
J(X) such that
h
0
(V
1
Q
L
0
V
1
Ω) > 0
implying there exists a non-trivial holomorphic map
φ : V
Q
L
0
V
1
.
Fix φ 6≡ 0. By Lemma 7.3, the family parameterized by V contains both
(V
P
V
Q
,
1
, 0)) and (W
P
V
Q
,
0
1
, 0)) implying there is deformation
between the two.
Set L = L
0
and Φ
0
2
= g
3
φ. Then the family of stable Higgs bundles
parameterized by H
0
(X, V
1
Q
L
0
V
1
Ω) contains (W
P
V
Q
,
0
1
, 0)) and
(W
P
V
Q
,
0
1
, Φ
0
2
)).
Now the family of bundle extensions of V
2
by L
0
V
1
is H
1
(L
0
V
1
V
1
2
).
With a fixed φ and the canonical factorization with ϕ fixed, H
1
(L
0
V
1
V
1
2
) parameterizes a family of Higgs bundles. This family contains (W
P
V
Q
,
0
1
, Φ
0
2
)). The zero element in H
1
(L
0
V
1
V
1
2
) corresponds to the
bundle extension
0 L
0
V
1
f
5
(L
0
V
1
) V
2
f
6
V
2
0.
Tensoring with gives
0 L
0
V
1
g
5
((L
0
V
1
) V
2
)
g
6
V
2
0.
Lemma 7.4. If (W
P
V
Q
,
0
1
, Φ
0
2
)) is stable (semi-stable), then H
1
(L
0
V
1
V
1
2
) parameterizes a stable (semi-stable) family.
Proof. Suppose (U
p
V
Q
,
1
, Ψ
2
)) H
1
(L
0
V
1
V
1
2
) and W U
P
V
Q
is
1
, Ψ
2
)-invariant. Since ϕ, φ 6≡ 0, one has W = V
1
or W = V
Q
V
1
. A
direct computation shows s(W ) < s(U
P
V
Q
) (s(W ) s(U
P
V
Q
)).
248 EUGENE Z. XIA
Proposition 7.2 follows from Lemma 7.4 because the family of Higgs bun-
dles parameterized by H
1
(L
0
V
1
V
1
2
) contains (W
P
V
Q
,
0
1
, Φ
0
2
)) and
((L
0
V
1
) V
2
V
Q
, (g
1
ϕ f
6
, g
5
φ)).
To summarize, a stable binary Hodge bundle (V
P
V
Q
,
1
, 0)) is first
deformed to (W
P
V
Q
,
0
1
, 0)) such that non-trivial holomorphic maps exist
between V
Q
and (L
0
V
1
) W
P
Ω. Such a non-trivial map Φ
0
2
is then
chosen and attached to the existing Higgs field Φ
0
1
. Finally W
P
is deformed
to a direct sum making the resulting stable Higgs bundle a ternary Hodge
bundle.
Let B = B(0, 0)\(B(0, 0)
s
T (0, 0, 0)).
Proposition 7.5. B is connected and can be deformed to a stable ternary
Hodge bundle in M
0
Dol
.
Proof. Consider the space U × J
0
(X), where J
0
(X) is the Jacobi variety
identified with the s et of holomorphic line bundles of degree zero on X and
U is the moduli space of rank-2 poly-stable holomorphic bundles of degree 0
on X. The space U is connected [2, 19]. Hence U ×J
0
(X) is connected. Each
poly-stable Higgs bundle in B is contained in the family of Higgs bundles
parameterized by U × J
0
(X). Hence the natural morphism
t : U × J
0
(X) B
is surjective. This proves that the set B is connected.
Choose holomorphic line bundles V
1
, V
2
, V
Q
of degrees 1, 1, 0, respec-
tively such that
h
0
(X, V
1
2
V
Q
Ω) > 0,
h
0
(X, V
1
Q
V
1
Ω) > 0.
Choose
0 6≡ ψ
1
H
0
(X, V
1
2
V
Q
Ω)
0 6≡ ψ
2
H
0
(X, V
1
Q
V
1
Ω).
The space of extension of V
2
by V
1
,
0 V
1
f
1
V
P
f
2
V
2
0,
is H
1
(X, V
1
V
1
2
). Tensoring the exac t sequence with gives
0 V
1
g
1
V
P
g
2
V
2
0.
Since deg(V
1
) < deg(V
2
), by the theorem of Lange and Narasimhan [13],
stable extensions always exist. Fix a stable extension V
P
and set
Φ
1
= ψ
1
f
2
,
Φ
2
= g
1
ψ
2
.
Note (V
P
V
Q
, 0) B. The connected family
F C = H
0
(X, V
1
2
V
Q
Ω) × H
0
(X, V
1
Q
V
1
Ω)
FLAT PU(2,1) STRUCTURES ON RIEMANN SURFACES 249
of Higgs bundles contains (V
P
V
Q
, 0) and (V
P
V
Q
,
1
, Φ
2
)). Note the
family F C contains s emi- stable Higgs bundles. This is allowed since the
points in the moduli space M
Dol
are also interpreted as S-equivalence classes
of sem i-stable Higgs bundles. However one may choose F C to be a strictly
poly-stable family:
F C = (H
0
(X, V
1
2
V
Q
Ω) × H
0
(X, V
1
Q
V
1
Ω)) \
(({0} × H
0
(X, V
1
Q
V
1
Ω)) (H
0
(X, V
1
2
V
Q
Ω) × {0})).
Since V
P
is stable, by Lemma 7.3, any element in F C is semi-stable. Hence
the family F C provides a deformation between (V
P
V
Q
, 0) and (V
P
V
Q
,
1
, Φ
2
)). The cohomology H
1
(X, V
1
V
1
2
) parameterizes bundle ex-
tensions of V
2
by V
1
and also parameterizes a family of Higgs bundles
with fixed ψ
1
, ψ
2
. By Lemma 7.4, this is a stable family which contains
(V
P
V
Q
,
1
, Φ
2
)) and (V
1
V
2
V
Q
, (ψ
1
f
4
, g
3
ψ
2
)) where f
3
, f
4
, g
3
, g
4
come from the trivial extensions
0 V
1
f
3
V
1
V
2
f
4
V
2
0,
0 V
1
g
3
(V
1
V
2
)
g
4
V
2
0.
Hence H
1
(X, V
1
V
1
2
) provides a deformation between (V
P
V
Q
,
1
, Φ
2
))
and (V
1
V
2
V
Q
, (ψ
1
f
4
, g
3
ψ
2
)) T (1, 1, 0).
To summarize, one first shows that the space B is connected. Then choose
a specific element (V
P
V
Q
, 0) B with V
P
a stable extension of V
2
by V
1
and that there exists non-trivial holomorphic maps
ψ
1
: V
2
V
Q
ψ
2
: V
Q
V
1
.
This provides a deformation from (V
P
V
Q
, 0) to (V
P
V
Q
,
1
, Φ
2
)). Finally,
since V
P
is an extension of V
2
by V
1
, (V
P
V
Q
,
1
, Φ
2
)) is deformed to
(V
1
V
2
V
Q
, (ψ
1
f
4
, g
3
ψ
2
)) in H
1
(X, V
1
V
1
2
).
Corollary 7.6. Every Binary Hodge bundle can be deformed to a ternary
Hodge bundle.
Proof. Every poly-stable reducible Hodge bundle is either ternary or in B.
The result then follows from Proposition 7.2 and 7.5.
Lemma 7.7. For fixed integers d
1
, d
2
, d
3
, T (d
1
, d
2
, d
3
) is connected.
Proof. We first consider the stable bundles. Stability implies the Higgs fields
Φ
1
, Φ
2
are not identically zero. Denote by J
d
(X) the Jacobi variety identified
with the set of holomorphic line bundles of degree d. For each L
1
J
d
1
(X),
the set of all (L
3
, Φ
2
) such that L
3
J
d
3
(X) and
0 6≡ Φ
2
H
0
(X, L
1
3
L
1
Ω)
250 EUGENE Z. XIA
is C
× Sym
d
1
+2(g1)d
3
X, where Sym
d
X is the d-th symmetric product of
X. Hence the set of all triples (L
3
, L
1
, Φ
2
) such that
L
3
Φ
2
7− L
1
with Φ
2
6≡ 0 is the space (C
× Sym
d
1
+2(g1)d
3
X) × J
d
1
(X).
Similarly, for each L
3
J
d
3
(X), the space of all triples (L
2
, L
3
, Φ
1
) such
that
L
2
Φ
1
7− L
3
with Φ
1
6≡ 0 is C
× Sym
d
3
+2(g1)d
2
X. The set of Higgs bundles parame-
terized by the total space
S = (C
× Sym
d
3
+2(g1)d
2
X) × (C
× Sym
d
1
+2(g1)d
3
X) × J
d
3
(X)
contains every Higgs bundle in T (d
1
, d
2
, d
3
). Hence the natural morphism
t : S T (d
1
, d
2
, d
3
)
is surjective. Since S is connected, T (d
1
, d
2
, d
3
) is connected.
The reducible bundles consist of T (0, 0, 0) and T (0, d
2
, d
2
). All poly-
stable Higgs bundles associated with the points in T (0, 0, 0) and T (0, d
2
, d
2
)
are contained in the families parameterized by
S
1
= J
0
(X) × J
0
(X) × J
0
(X)
and
S
2
= (C
× Sym
2(g1)2d
2
X) × J
d
2
(X) × J
0
(X),
respectively. Both S
1
, S
2
are connected. Since the natural morphisms
t
1
: S
1
T (0, 0, 0)
t
2
: S
2
T (0, d
2
, d
2
)
are surjective, both T (0, 0, 0) and T (0, d
2
, d
2
) are connected.
Proposition 7.8. Every component of M
τ
Dol
contains a Hodge bundle.
Proof. By Corollary 4.8, every component of M
τ
Dol
contains a local minimum
(V, Φ) of m
r
. If (V, Φ) is a smooth point, then (V, Φ) is a critical point of m.
A theorem of Hitchin and Simpson implies that (V, Φ) is a Hodge bundle
[12, 22]. Singular points of M
τ
Dol
correspond to reducible Higgs bundles.
The space of all reducible Higgs bundles correspond to either the space of
U(2) × U(1) representations or the space of U(1) × U(1, 1) representations.
Each component of U(2) × U(1) and U(1) × U(1, 1) representations contains
points that correspond to Hodge bundles [11]. In fact, these points are
exactly the ones corresponding to the points in B and T (0, d
2
, d
2
).
Let K be a divisor of and let
w : X |K|
=
CP
g1
be the canonical map [10].
FLAT PU(2,1) STRUCTURES ON RIEMANN SURFACES 251
Lemma 7.9. has a section with simple zeros.
Proof. The linear system |K| is base point free [10]. If X is hyperelliptic,
then the map w is a 2-1 branch map into CP
g1
and an embedding otherwise .
In both cases, by Bertini’s theorem, there exists a hyperplane H CP
g1
such that H X is regular. Then w
1
(H) is an effective divisor equivalent
to K and with simple zeros.
Choose
K = {x
1
, x
2
, . . . , x
2(g1)
},
such that the x
i
’s are all distinct.
Proposition 7.10. Let 0 τ < 2(g 1). Suppose
T (d
1
1, d
2
+ 1, d
Q
), T (d
1
, d
2
, d
Q
) M
τ
Dol
.
Then there is deformation between T (d
1
, d
2
, d
Q
) and T (d
1
1, d
2
+ 1, d
Q
)
within M
τ
Dol
.
Proof. Suppose
(V
1
V
2
V
Q
,
1
, Φ
2
)) T (d
1
1, d
2
+ 1, d
Q
),
(U
1
U
2
U
Q
,
1
, Ψ
2
)) T (d
1
, d
2
, d
Q
).
By the semi-stability of (U
1
U
2
U
Q
,
1
, Ψ
2
)) and the assumptions τ
0, 0 c < 3, one has d
Q
0 and
d
1
1 < d
1
d
P
+ d
Q
3
< 1;
hence,
d
1
1 < d
1
0 and d
2
+ 1 > 0.
This implies (V
1
V
2
V
Q
,
1
, Φ
2
)) is stable. Hence Φ
1
6≡ 0 and
deg(V
2
) + d
Q
+ 2(g 1) 0.
On the other hand, deg(V
1
) + deg(V
2
) = d
P
, so
d
P
deg(V
1
) d
Q
2(g 1),
d
1
< 1 d
1
= deg(V
1
) d
P
+ d
Q
+ 2(g 1) 2(g 1).
In light of Lemma 7.7, it suffices to demonstrate the existence of (U
1
U
2
U
Q
,
1
, Ψ
2
)) T (d
1
, d
2
, d
Q
) and (V
1
V
2
U
Q
,
1
, Φ
2
)) T (d
1
1, d
2
+ 1, d
Q
) and a deformation between the two.
Since |K| is base point free, there exists K
0
|K| such that
K
0
= {y
1
, y
2
, . . . , y
2(g1)
}
252 EUGENE Z. XIA
with y
i
6= x
2(g1)
for all 1 i 2g. The b ounds on the degrees of the
various bundles allow us to construct the following divisors:
D
1
= {−x
1
, . . . , x
deg(U
1
)
}
D
2
= {y
1
, . . . , y
d
P
deg(V
1
)
, x
2(g1)
}
D
Q
= {−y
d
P
deg(V
1
)+1
, . . . , y
d
P
deg(V
1
)d
Q
}.
Let u be the basic epimorphism [1]
u : Div (X) H
1
(X, O
)
and set
U
1
= u(D
1
)
U
2
= u(D
2
)
U
Q
= u(D
Q
)
U
P
= U
1
U
2
.
Let ψ
1
, ψ
2
be meromorphic sections associated with the divisors D
1
, D
2
.
Then the meromorphic section ψ
1
ψ
2
of U
P
is associated with the divisor
D
0
1
= {−x
1
, . . . , x
deg(U
1
)
, x
2(g1)
}.
Hence there exists V
1
U
P
[9] such that
V
1
= u(D
0
1
).
Let
V
2
= U
P
/V
1
.
Since
V
1
V
2
= det(U
P
) = U
1
U
2
,
V
2
= u(D
0
2
),
where
D
0
2
= {y
1
, . . . , y
d
P
deg(V
1
)
}.
In short, the bundle U
P
is constructed in such a way that it is the trivial
extension of U
2
by U
1
, and is also an extension of V
2
by V
1
:
0 U
1
f
1
U
P
f
2
U
2
0
0 V
1
f
3
U
P
f
4
V
2
0.
Tensoring with gives
0 U
1
g
1
U
P
g
2
U
2
0
0 V
1
g
3
U
P
g
4
V
2
0.
FLAT PU(2,1) STRUCTURES ON RIEMANN SURFACES 253
Since
D
2
+ D
Q
+ K
0
=
n
x
2(g1)
, y
d
P
deg(V
1
)d
Q
+1
, . . . , y
2(g1)
o
D
Q
+ D
1
+ K =
n
y
d
P
deg(V
1
)+1
, . . . , y
d
P
deg(V
1
)d
Q
,
x
deg(U
1
)+1
, . . . , x
2(g1)
o
are effective divisors, there exists
0 6≡ ψ
1
H
0
(X, U
1
2
U
Q
Ω)
0 6≡ ψ
2
H
0
(X, U
1
Q
U
1
Ω).
Set
Ψ
1
= ψ
1
f
2
and Ψ
2
= g
1
ψ
2
.
Then (U
1
U
2
U
Q
,
1
, Ψ
2
)) is a semi-stable ternary Hodge bundle.
The divisors
D
0
2
+ D
Q
+ K
0
=
n
y
d
P
deg(V
1
)d
Q
+1
, . . . , y
2(g1)
o
D
Q
+ D
0
1
+ K =
n
x
deg(U
1
)+1
, . . . , x
2(g1)1
,
y
d
P
deg(V
1
)+1
, . . . , y
d
P
deg(V
1
)d
Q
o
are effective. Hence there exist
0 6≡ φ
1
H
0
(X, V
1
2
U
Q
Ω)
0 6≡ φ
2
H
0
(X, U
1
Q
V
1
Ω).
Remark 2. This is the critical step where the assumption τ < 2(g 1) is
needed. In the case of τ = 2(g 1), the degree of V
1
2
U
Q
equals
1 thus rendering it impossible to find a non-zero global section φ
1
. This
reflects the fact that every representation with τ = 2(g 1) is reducible.
(See Section 6.)
Set
Ψ
0
1
= φ
1
f
4
and Ψ
0
2
= g
3
φ
2
.
Then (U
P
U
Q
,
0
1
, Ψ
0
2
)) is a semi-stable Higgs bundle. Since
h
0
(X, U
1
2
U
Q
Ω) > 0
h
0
(X, U
1
Q
U
1
Ω) > 0,
H
0
(X, U
1
1
U
Q
Ω) and H
0
(X, U
1
Q
U
2
Ω) are proper subspaces of
H
0
(X, U
1
P
U
Q
Ω) and H
0
(X, U
1
Q
U
P
Ω), respectively. Hence
F C = (H
0
(X, U
1
P
U
Q
Ω) \ H
0
(X, U
1
1
U
Q
Ω)) ×
(H
0
(X, U
1
Q
U
P
Ω) \ H
0
(X, U
1
Q
U
2
Ω))
254 EUGENE Z. XIA
is connected and parameterizes a family of sem i-stable Higgs bundles that
contains both (U
P
U
Q
,
1
, Ψ
2
)) and (U
P
U
Q
,
0
1
, Ψ
0
2
)). Hence there is
deformation between the two.
The space of bundle extensions of V
2
by V
1
,
0 V
1
f
5
V
f
6
V
2
0,
is parameterized by the vector space H
1
(V
1
V
1
2
) containing both U
P
and
V
1
V
2
(the zero element in H
1
(V
1
V
1
2
)). Again tensoring with gives
0 V
1
g
5
V
g
6
V
2
0.
Let
Φ
1
= φ
1
f
0
6
and Φ
2
= g
0
5
φ
2
,
where
0 V
1
f
0
5
V
1
V
2
f
0
6
V
2
0
0 V
1
g
0
5
(V
1
V
2
)
g
0
6
V
2
0
correspond to the trivial extensions. By Lemma 7.4, H
1
(V
1
V
1
2
) param-
eterizes a family of semi-stable Higgs bundles that contains both (U
P
U
Q
,
0
1
, Ψ
0
2
)) and (V
1
V
2
U
Q
,
1
, Φ
2
)).
To summarize, the first step consists of fixing U
P
= U
1
U
2
and deform
the Higgs field
1
, Ψ
2
) to
0
1
, Ψ
0
2
). In the second step, fix φ
1
, φ
2
and deform
U
P
to V
1
V
2
.
Consider the space T (0, d
2
, d
2
). By Proposition 7.5, one may assume d
2
>
0. To deform points in T (0, d
2
, d
2
), the family F C constructed in the
above proof contains semi-stable Higgs bundles. However, one may also opt
to construct the deformation family of poly-stable Higgs bundles by setting:
F C = (H
0
(X, U
1
P
U
Q
Ω) \
(H
0
(X, U
1
1
U
Q
Ω) H
0
(X, U
1
2
U
Q
Ω))) ×
(H
0
(X, U
1
Q
U
P
Ω) \
(H
0
(X, U
1
Q
U
2
Ω) H
0
(X, U
1
Q
U
1
Ω)))
(H
0
(X, U
1
2
U
Q
Ω) × {0}).
The case with τ = 2(g 1) has been covered in Section 6 and M
2(g1)
Dol
is connected. Suppose τ < 2(g 1). By Proposition 7.8, every component
of M
τ
Dol
contains a Hodge bundle. By Corollary 7.6, every component of
M
τ
Dol
contains a ternary Ho dge bundle. It follows from Proposition 7.10
and induction that M
τ
Dol
is connected. Since
N
τ
Dol
= M
τ
Dol
/H
1
(X, O
),
Theorem 1.1 then follows from Corollary 4.2.
FLAT PU(2,1) STRUCTURES ON RIEMANN SURFACES 255
References
[1] E. Arbarello, M. Cornalba, P. Griffiths and J. Harris, Geometry of Algebraic Curves,
Vol. I, Springer-Verlag, 1985.
[2] M. Atiyah and R. Bott, The Yang-Mills equations over Riemann surfaces, Philos.
Trans. Roy. Soc. London, Ser., A 308 (1982), 523-615.
[3] K. Corlette, Flat G-bundles with canonical metrics, J. Diff. Geom., 28 (1988), 361-
382.
[4] A. Domic and D. Toledo, The Gromov norm of the ahler class of symmetric do-
mains, Math. Ann., 276 (1987), 425-432.
[5] S. Donaldson, Twisted harmonic maps and the self-duality equations, Proc. London
Math. Soc., 55 (1987), 127-131.
[6] W. Goldman, Topological components of spaces of representations, Invent. Math., 93
(1988), 557-607.
[7] W. Goldman, Michael E. Kapovich and Bernhard Leeb, Complex Hyperbolic Mani-
folds Homotopy-Equivalent to a Riemann Surface, Preprint.
[8] P. Gothen, The Betti numbers of the moduli space of stable rank 3 Higgs bundles on
a Riemann surface, Internat. J. Math., 5(6) (1994), 861-875.
[9] R. Gunning, Lectures on Vector Bundles Over Riemann Surfaces, Princeton Univer-
sity Press, 1967.
[10] R. Hartshorne, Algebraic Geometry, Springer-Verlag, 1977.
[11] N. Hitchin, The self-duality equations on a Riemann surface, Pro c. London Math.
So c., 55 (1987), 59-126.
[12] , Lie groups and Teichm¨uller space, Topology, 31 (1992), 449-473.
[13] H. Lange and M. Narasimhan, Maximal subbundles of rank two vector bundles on
curves, Math. Ann., 266 (1983), 55-72.
[14] M. Maruyama, Openness of a family of torsion free sheaves, J. Math. Kyoto Univ.,
16-3 (1976), 627-637.
[15] D. Mumford and J. Fogarty, Geometric Invariant Theory, 2nd edn., Springer, Berlin,
1982.
[16] M. Narasimhan and C. Seshadri, Stable and unitary vector bundles on a compact
Riemann surface, Annals of Math., 82 (1965), 540-567.
[17] P.E. Newstead, Introductions to Moduli Problems and Orbit Spaces, Springer-Verlag,
1978.
[18] N. Nitsure, Moduli space of semistable pairs on a curve, Pro c. London Math. Soc.,
62 (1991), 275-300.
[19] C. Seshadri, Fibr´es vectoriels sur les courbes algebriques, Ast´erisque, 96 (1982).
[20] S. Shatz, The decomposition and specialization of algebraic families of vector bundles,
Comp os itio Mathematica, 35 (1977), 163-187.
[21] C. Simpson, Constructing variations of Hodge structures using Yang-Mills Theory
and applications to uniformization, Journal of the A.M.S., 1 (1988), 867-918.
[22] , Higgs bundles and local systems, Publ. Math. I.H.E.S., 75 (1992), 6-95.
[23] , Moduli of representations of the fundamental group of a smooth projective
variety, I, Publ. Math. I.H.E.S., 79 (1994), 47-129.
256 EUGENE Z. XIA
[24] , Moduli of representations of the fundamental group of a smooth projective
variety, II, Publ. Math. I.H.E.S., 80 (1994), 5-79.
[25] N. Steenrod, The Topology of Fiber Bundles, Princeton University Press, 1951.
[26] D. Toledo, Harmonic maps from surfaces to certain ahler manifolds, Math. Scand.,
45 (1979), 13-26.
[27] , Representations of surface groups in complex hyperbolic space, J. Differential
Geometry, 29 (1989), 125-133.
Received September 2, 1998 and revised March 16, 1999.
University of M assachusetts
Amherst, MA 010 03 -45 15
E-mail address: xia@math.umass.edu
Guidelines for Authors
Authors may submit manuscripts at pjm.math.berkeley.edu/about/journal/submissions.html
and choose an editor at that time. Exceptionally, a paper may be submitted in hard copy to
one of the editors; authors should keep a copy.
By submitting a manuscript you assert that it is original and is not under consideration
for publication elsewhere. Instructions on manuscript preparation are provided below. For
further information, visit the web address above or write to pacific@math.berkeley.edu or
to Pacific Journal of Mathematics, University of California, Los Angeles, CA 90095–1555.
Correspondence by email is requested for convenience and speed.
Manuscripts must be in English, French or German. A brief abstract of about 150 words or
less in English must be included. The abstract should be self-contained and not make any
reference to the bibliography. Also required are keywords and subject classification for the
article, and, for each author, postal address, affiliation (if appropriate) and email address if
available. A home-page URL is optional.
Authors are encouraged to use L
A
T
E
X, but papers in other varieties of T
E
X, and exceptionally
in other formats, are acceptable. At submission time only a PDF file is required; follow
the instructions at the web address above. Carefully preserve all relevant files, such as
L
A
T
E
X sources and individual files for each figure; you will be asked to submit them upon
acceptance of the paper.
Bibliographical references should be listed alphabetically at the end of the paper. All
references in the bibliography should be cited in the text. Use of BibT
E
X is preferred but
not required. Any bibliographical citation style may be used but tags will be converted to
the house format (see a current issue for examples).
Figures, whether prepared electronically or hand-drawn, must be of publication quality.
Figures prepared electronically should be submitted in Encapsulated PostScript (EPS) or
in a form that can be converted to EPS, such as GnuPlot, Maple or Mathematica. Many
drawing tools such as Adobe Illustrator and Aldus FreeHand can produce EPS output.
Figures containing bitmaps should be generated at the highest possible resolution. If there
is doubt whether a particular figure is in an acceptable format, the authors should check
with production by sending an email to pacific@math.berkeley.edu.
Each figure should be captioned and numbered, so that it can float. Small figures occupying
no more than three lines of vertical space can be kept in the text (“the curve looks like
this:”). It is acceptable to submit a manuscript will all figures at the end, if their placement
is specified in the text by means of comments such as “Place Figure 1 here”. The same
considerations apply to tables, which should be used sparingly.
Forced line breaks or page breaks should not be inserted in the document. There is no point
in your trying to optimize line and page breaks in the original manuscript. The manuscript
will be reformatted to use the journal’s preferred fonts and layout.
Page proofs will be made available to authors (or to the designated corresponding author)
at a Web site in PDF format. Failure to acknowledge the receipt of proofs or to return
corrections within the requested deadline may cause publication to be postponed.
PACIFIC JOURNAL OF MATHEMATICS
Volume 195 No. 1 September 2000
Motion of hypersurfaces by Gauss curvature 1
BEN ANDREWS
Summation of formal solutions of a class of linear difference equations 35
B.L.J. BRAAKSMA, B.F. FABER AND G.K. IMMINK
Approximation of recurrence in negatively curved metric spaces 67
CHARALAMBOS CHARITOS AND GEORGIOS TSAPOGAS
Cohomology of singular hypersurfaces 81
BERNARD M. DWORK
Riemannian manifolds admitting isometric immersions by their first
eigenfunctions 91
AHMAD EL SOUFI AND SAÏD ILIAS
The Rubinstein–Scharlemann graphic of a 3-manifold as the discriminant set
of a stable map 101
TSUYOSHI KOBAYASHI AND OSAMU SAEKI
Geometric realizations of Fordy–Kulish nonlinear Schr
¨
odinger systems 157
JOEL LANGER AND RON PERLINE
Seiberg–Witten invariants for 3-manifolds in the case b
1
= 0 or 1 179
YUHAN LIM
Virtual homology of surgered torus bundles 205
JOSEPH D. MASTERS
Group actions on polynomial and power series rings 225
PETER SYMONDS
The moduli of flat PU(2,1) structures on Riemann surfaces 231
EUGENE Z. XIA
0030-8730(200009)195:1;1-#
Pacific Journal of Mathematics 2000 Vol. 195, No. 1
Pacific
Journal of
Mathematics
Volume 195 No. 1 September 2000
... The main contribution of this paper is to provide a very natural argument based on anisotropic flows developed by Andrews [5] to handle the case −n − 1 < p < 1, or equivalently, the case 1 n+2 < α < ∞. Entropy functional. ...
... When α = 1 and f (x) ≡ 1, then the above quantity agrees with the entropy in [28], first introduced by Firey [25] for the centrally symmetric Ω. General integral quantities studied by Andrews in [3,5]. Here we shall assume that ...
... where ν(x, τ ) is the unit exterior normal atX(x, τ ) ofM τ =X(M, τ ), andK(x, τ ) is the Gauss curvature ofM τ atX(x, τ ). Andrews [5] proved that flow (1.7) contracts to a point under finite time if the initial hypersurface M 0 is strictly convex. Under a proper normalization, the normalized anisotropy flow of (1.7) is ...
Article
Full-text available
We provide a natural simple argument using anistropic flows to prove the existence of weak solutions to Lutwak’s $L^p$ -Minkowski problem on $S^n$ which were obtained by other methods.
... The main contribution of this paper is to provide a very natural argument based on anisotropic flows developed by Andrews [4] to handle the case −n − 1 < p < 1, or equivalently, the case 1 n+2 < α < ∞. Entropy functional. ...
... When α = 1 and f (x) ≡ 1, then the above quantity agrees with the entropy in [27], first introduced by Firey [24] for the centrally symmetric Ω. General integral quantities studied by Andrews in [2,4]. Here we shall assume that S n f (x) dθ(x) = 1; namely, 1 ωn f (x)dθ(x) is a probability measure. ...
... where ν(x, τ ) is the unit exterior normal atX(x, τ ) ofM τ =X(M, τ ), andK(x, τ ) is the Gauss curvature ofM τ atX(x, τ ). Andrews [4] proved that flow (1.7) contracts to a point under finite time if the initial hypersurface M 0 is strictly convex. Under a proper normalization, the normalized anisotropy flow of (1.7) is ...
Preprint
We provide a natural simple argument using anistropic flows to prove the existence of weak solutions to Lutwak's $L^p$-Minkowski problem on $S^n$ which were obtained by other methods.
... In particular, solutions to (1.2) are smooth if f is smooth and positive. For the case p ∈ (−∞, −n − 1], the smoothness of the solutions can be deduced from the result of Andrews [3]. We note that, for the affine critical case where p = −n − 1 (in which the equation is invariant under affine transformations), solutions might not exist in general, as described in [28]. ...
... On the other hand, in the case −n − 1 < p < 0, there is no uniform diameter estimate depending on f by the result in Jian-Lu-Wang [40]. (iii) Although the smoothness of solutions and curvature estimates are already known for p ∈ (−∞, −n + 1] in Andrews [3] and Chou-Wang [28], we included this range of p in Theorem 1.1 because our curvature estimates also work in this range and the proofs are new. (iv) It was previously known in Chou-Wang [28] that when p ∈ (−n + 1, 1), the associated convex hypersurface is of class C 1 (or C 1,γ , respectively) provided that f ∈ L ∞ (S n ) (or C 0,1 (S n ), respectively). ...
... To prove Theorem 1.1, we utilize the anisotropic α-Gauss curvature flow, defined in [3], with α = 1 1−p . We say that a one-parameter family of complete convex hypersurfaces {M t } t∈I is an anisotropic α-Gauss curvature flow if ...
Preprint
Full-text available
We establish curvature estimates for anisotropic Gauss curvature flows. By using this, we show that given a measure $\mu$ with a positive smooth density $f$, any solution to the $L_p$ Minkowski problem in $\mathbb{R}^{n+1}$ with $p \le -n+2$ is a hypersurface of class $C^{1,1}$. This is a sharp result because for each $p\in [-n+2,1)$ there exists a convex hypersurface of class $C^{1,\frac{1}{n+p-1}}$ which is a solution to the $L_p$ Minkowski problem for a positive smooth density $f$. In particular, the $C^{1,1}$ regularity is optimal in the case $p=-n+2$ which includes the logarithmic Minkowski problem in $\mathbb{R}^3$.
... wherẽis a positive smooth function on n . This type of flow was treated in [12] when f is homogeneous, in particular for power of Gauss curvature f ( ) = K . ...
Article
Full-text available
We consider a general class of non-homogeneous contracting flows of convex hypersurfaces in R n + 1 ${\mathbb{R}}^{n+1}$ , and prove the existence and regularity of the flow before extincting to a point in finite time.
... Being a self-similar solution to the isotropic Gauss curvature flow, the case p = 0 and µ = m of (1.1) describes the ultimate fate of a worn stone in a model proposed by Firey [30]. This model was extended and further studied in [1,2,3,5,15,23,25,74,75]. In the general anisotropic α-power-of-Gausscurvature flow, x : S n−1 × [0, T ) → R n evolves according to ...
Preprint
Let $K$ be a smooth, origin-symmetric, strictly convex body in $\mathbb{R}^n$. If for some $\ell\in GL(n,\mathbb{R})$, the metric $\frac{1}{2}D^2 \Vert\cdot\Vert_{\ell K}^2$ is comparable to the standard Euclidean metric of $\mathbb{R}^{n}$ up-to a factor of $\gamma > 1$, we show that $K$ satisfies the even $L^p$-Minkowski inequality and uniqueness in the even $L^p$-Minkowski problem for all $p \geq p_\gamma := 1 - \frac{n+1}{\gamma}$. This result is sharp as $\gamma \searrow 1$ (characterizing centered ellipsoids in the limit) and improves upon the classical Minkowski inequality for all $\gamma < \infty$. In particular, whenever $\gamma \leq n+1$, the even log-Minkowski inequality and uniqueness in the even log-Minkowski problem hold.
... Among other reasons, the interest in K α -flows comes from the fact that they naturally relate to a wide range of research fields, varying from image processing to affine geometry and geometric analysis (cf. [1]). ...
Article
Full-text available
We consider translating solitons to flows by positive powers α of the Gaussian curvature, called Kα-flows, in Riemannian products M×R. We prove that, when M is the Euclidean space Rn, the sphere Sn, or one of the hyperbolic spaces HFm, there exist complete rotational translating solitons to Kα-flow in M×R for certain values of α.
Article
We consider an anisotropic area-preserving nonlocal flow for closed convex plane curves, which is a generalization of the model introduced by Pan and Yang (J. Differential Equations 266 (2019), 3764–3786) when τ = 1. Under this flow, the evolving curve maintains its convexity and converges to a homothety of a smooth symmetric strictly convex plane curve in the C ∞ sense. The analysis of the asymptotic behavior of this flow implies the possibility of deforming one curve into another within the framework of Minkowski geometry.
Article
Let $K$ be a smooth, origin-symmetric, strictly convex body in ${\mathbb{R}}^{n}$. If for some $\ell \in \textrm{GL}(n,{\mathbb{R}})$, the anisotropic Riemannian metric $\frac{1}{2}D^{2} \left \Vert \cdot \right \Vert_{\ell K}^{2}$, encapsulating the curvature of $\ell K$, is comparable to the standard Euclidean metric of ${\mathbb{R}}^{n}$ up-to a factor of $\gamma> 1$, we show that $K$ satisfies the even $L^{p}$-Minkowski inequality and uniqueness in the even $L^{p}$-Minkowski problem for all $p \geq p_{\gamma }:= 1 - \frac{n+1}{\gamma }$. This result is sharp as $\gamma \searrow 1$ (characterizing centered ellipsoids in the limit) and improves upon the classical Minkowski inequality for all $\gamma < \infty $. In particular, whenever $\gamma \leq n+1$, the even log-Minkowski inequality and uniqueness in the even log-Minkowski problem hold.
Article
Existence of symmetric solutions to the Gaussian Minkowski problem was established by Huang et al. In this paper, we show the existence of non-symmetric solutions to this problem by studying the related Monge–Ampère type equation on the sphere.
Article
Full-text available
We consider parabolic curvature flows for hypersurfaces in Euclidean space, defined in terms of the mixed discriminants of Aleksandrov. Dilation-invariant curvature integral estimates are proved for convex hypersurfaces evolving by these equations, generalising the so-called entropy estimates which have been proved for special cases by Richard Hamilton, Ben Chow, and the author. The earlier proof by contradiction is replaced here by a new proof which is more direct and applies more generally, and which demonstrates the intimate connection between entropy estimates and the Aleksandrov-Fenchel inequalities.
Article
Soit C(•,0):S 1 →R 2 une courbe lisse plongee dans le plan. Alors C:S 1 ×[0,T)→R 2 existe en satisfaisant δC/δt=K•N, ou K est la courbure de C, et N est son vecteur unite normal entrant. C(•,t) est lisse pour tout t, il converge vers un point quand t\T et sa forme limite quand t→T est un cercle rond, avec convergence dans norme C ∞
Article
We give a new proof of multisummability of formal power series solutions of a nonlinear meromorphic differential equation. We use the recent Malgrange-Ramis definition of multisummability. The first proof of the main result is due to B. Braaksma. Our method of proof is very different: Braaksma used Écalle definition of multisummability and Laplace transform. Starting from a preliminary normal form of the differential equation xdy → dx=G → 0 (x)+λ(x)+A 0 y →+x μ G →(x,y →), the idea of our proof is to interpret a formal power series solution as a holomorphic cochain, whose coboundary is exponentially small of some order. Then we increase this order in a finite number of steps. (In this process we use the knowledge of the slopes of a Newton polygon.) The key lemma is based on reductions to some resonant normal forms and on a precise description of some nonlinear Stokes phenomena.