Izumiya - Differential Geometry From Singularity Theory Viewpoint PDF
Izumiya - Differential Geometry From Singularity Theory Viewpoint PDF
For photocopying of material in this volume, please pay a copying fee through the Copyright Clearance
Center, Inc., 222 Rosewood Drive, Danvers, MA 01923, USA. In this case permission to photocopy
is not required from the publisher.
Printed in Singapore
Preface
v
vi Di↵erential Geometry from a Singularity Theory Viewpoint
We are also very grateful to Masatomo Takahashi for reading the final draft
of the book and for his invaluable comments and corrections.
Preface v
1. The case for the singularity theory approach 1
1.1 Plane curves . . . . . . . . . . . . . . . . . . . . . . . . . 2
1.1.1 The evolute of a plane curve . . . . . . . . . . . . 3
1.1.2 Parallels of a plane curve . . . . . . . . . . . . . . 5
1.1.3 The evolute from the singularity theory viewpoint 7
1.1.4 Parallels from the singularity theory viewpoint . . 10
1.2 Surfaces in the Euclidean 3-space . . . . . . . . . . . . . . 11
1.2.1 The focal set . . . . . . . . . . . . . . . . . . . . . 13
1.3 Special surfaces in the Euclidean 3-space . . . . . . . . . . 14
1.3.1 Ruled surfaces . . . . . . . . . . . . . . . . . . . . 15
1.3.2 Developable surfaces . . . . . . . . . . . . . . . . 19
1.4 Notes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 21
ix
x Di↵erential Geometry from a Singularity Theory Viewpoint
Bibliography 347
Index 363
May 2, 2013 14:6 BC: 8831 - Probability and Statistical Theory PST˙ws
The study of curves and surfaces in the Euclidean space is a fascinating and
important subject in di↵erential geometry. We highlight in this chapter how
singularity theory can be used not only to recover classical results on curves
and surfaces in a simpler and more elegant way but also how it reveals the
rich and deep underlying concepts involved.
We start with the evolute and parallels of a plane curve. We first use
classical di↵erential geometry techniques to obtain the shape of the evolute
and parallels. We then define the family of distance squared functions on
the plane curve and recover from the singularities type of the members
of this family geometric information about the curve itself. We outline
how to use the Lagrangian and Legendrian singularity theory framework to
deduce properties of the evolute that are invariant under di↵eomorphisms.
We proceed similarly for surfaces in the Euclidean 3-space and consider
the singularities of their focal sets. We deal in the last section with the
singularities of ruled and developable surfaces.
We refer to [do Carmo (1976)] for a detailed study of the di↵erential
geometry of curves and surfaces.
Throughout this book, a given map is said to be smooth (or C 1 ) if its
partial derivatives of all order exist and are continuous.
The Euclidean n-space is the vector space Rn endowed with the scalar
product
hu, vi = u1 v1 + · · · + un vn
for any u = (u1 , . . . , un ) and v = (v1 , . . . , vn ) in Rn .
We also view the Euclidean n-space as a set of points. The vector
space Rn comes with a standard orthogonal basis e1 = (1, . . . , 0), . . . ,
en = (0, . . . , 1). We choose a point O = (0, . . . , 0) to be the origin and
denote by ⌃ = (O, e1 , . . . , en ) the standard orthonormal coordinates system
1
2 Di↵erential Geometry from a Singularity Theory Viewpoint
e1 · · · en
u11 · · · u1n
u1 ⇥ · · · ⇥ un 1 = .. . ,
. · · · ..
un1 1 · · · unn 1
n0 (s) = (s)t(s).
When the parameter t of the curve is not necessarily the arc length
parameter, the curvature is given by the formula
det( 0 (t), 00 (t))
(t) = .
|| 0 (t)||3
(One can re-parametrise by arc length and use the chain rule to get the
above formula, see for example [Bruce and Giblin (1992)].) If we write
(t) = (x(t), y(t)), then
x0 y 00 x00 y 0
= 3 ,
(x02 + y 02 ) 2
where all the functions are evaluated at t.
The curvature function determines completely the curve up to rigid
motions (i.e., up to translations and rotations about points in the plane).
Indeed,
where n(t) is the unit normal vector obtained by rotating the unit tangent
vector 0 (t)/|| 0 (t)|| anti-clockwise by ⇡/2.
The evolute is well defined and is a smooth curve away from the inflec-
tion points of . We can use classical di↵erential geometry techniques to
study its geometry.
An ellipse has four ordinary vertices and Figure 1.2, left, shows the
evolute of an ellipse with its four ordinary cusps. The vertices of a curve
are points where the curve is most or least curved and these can be detected
(approximately) by the naked eye for the ellipse in Figure 1.2, left. It is
not possible to do so for the ellipse in Figure 1.2, right, as its principal
axes have almost the same length. The ellipse in Figure 1.2, right, looks
like a circle but is not a circle as its evolute is not a point. We can find
the vertices of the ellipse in Figure 1.2, right, by considering the limiting
tangent lines to the evolute at its ordinary cusp. These lines intersect the
ellipse at its vertices.
Fig. 1.2 Evolutes of ellipses: the di↵erence between the lengths of the principal axes of
the ellipse on the left is noticeable, whereas that of the ellipse on the right is negligible.
The ellipse on the right looks like a circle but is not a circle.
trace the evolute of . Figure 1.3 shows the parallels of an ellipse with their
singular points tracing the evolute of the ellipse.
Fig. 1.3 The parallels of an ellipse. The ellipse and its evolute are drawn in thick.
⇢0d = (1 d)t,
⇢00d = d0 t + (1 d)n,
⇢000
d = (d00 + 2 (1 d))t + 0 (1 3d)n,
(4)
⇢d = (d000 + 30 (1 2d))t
+( (d00 + 2 (1 d)) + 00 (1 3d) 3d02 )n.
At a singularity s0 of the parallel d = 1/(s0 ), so
00 0 (s0 )
⇢d (s0 ) = t(s0 ),
(s0 )
00 (s0 )
000
⇢d (s0 ) = t(s0 ) 20 (s0 )n(s0 ).
(s0 )
The vectors ⇢00d (s0 ) and ⇢000
d (s0 ) are linearly independent if and only if
0
(s0 ) 6= 0, equivalently, if and only if (s0 ) is not a vertex of . If this is
The case for the singularity theory approach 7
the case, the parallel ⇢d with d = 1/(s0 ) has an ordinary cusp singularity
at s = s0 (see Definition 1.1).
Suppose that d = 1/(s0 ) and 0 (s0 ) = 0. Then ⇢0d (s0 ) = ⇢00d (s0 ) = 0
and
00 (s0 )
⇢000
d (s0 ) = t(s0 ),
(s0 )
(4) 000 (s0 )
⇢d (s0 ) = t(s0 ) 300 (s0 )n(s0 ).
(s0 )
(4)
The vectors ⇢000
d (s0 ) and ⇢d (s0 ) are linearly independent if and only if
00
(s0 ) 6= 0, equivalently, if and only if (s0 ) is an ordinary vertex of . If
this is the case, the parallel ⇢d with d = 1/(s0 ) has a (3, 4)-singularity at
s0 (see Definition 1.1). ⇤
1
A2 a = (s0 ) + (s0 )
n(s0 ), The centre of the circle C lies on the evolute of
0 (s0 ) 6= 0 but s0 is not a vertex of .
A3 a = (s0 ) + (s1 ) n(s0 ), The centre of the circle C lies on the evolute of
0
0 (s0 ) = 0, 00 (s0 ) 6= 0 and s0 is an ordinary vertex of .
D(p, a) = Da (p).
The case for the singularity theory approach 9
Proposition 1.4. (1) The local bifurcation set of the family of distance
squared functions on is the evolute of .
(2) The catastrophe set CD is a regular surface in I ⇥ R2 . The set of
critical values of the catastrophe map ⇡CD : CD ! R2 , with ⇡CD (s, a) = a,
is the local bifurcation set of the family of distance squared functions.
Proof. The proof of (1) follows from the definition of the bifurcation set
and from Proposition 1.3. As for (2), we prove in Chapter 5 a more general
result that shows that CD is a regular surface. ⇤
We can now outline the underlying singularity theory concepts involved
in the study of the evolute. These are developed in subsequent chapters.
T< ⇤ R2
L(D)
⇡
✏
CD / R2
⇡CD
The set
@ D̃r
⇤ (D̃r ) = {(s, a) 2 I ⇥ R : D̃r (s, a) = @s (s, a) = 0}
⌃± 2
P 9 T ⇤ R2
L (D̃r )
⇡
✏
⌃±
⇤ (D̃r )
/ R2
⇡2
L (D)(⌃± ⇤ (D̃)). The individual fronts are recovered by slicing the big
front by the planes r = constant.
8. We deduce that the singularities of the parallel ⇢r of the curve are
ordinary cusps at points ⇢r (s0 ) where r = 1/(s0 ) when s0 is not a
vertex of . The parallels undergo the transitions given by the generic
section of the swallowtail surface (big wavefront) at an ordinary vertex
of (Figure 1.4). See also Figure 1.3 which shows how the parallels to
an ellipse are stacked together at the vertices of the ellipse.
Fig. 1.4 A swallowtail surface (big wavefront) left and its generic sections (right).
Remark 1.1. There are non-isometric surfaces in R3 with the same Gaus-
sian curvature. A theorem of Bonnet states that if some compatibility
conditions (the Gauss and Mainardi-Codazi conditions) are imposed on the
coefficients of the first and second fundamental forms of the surface, then
these coefficients determine completely the surface up to isometries. See,
for example, [do Carmo (1976)] for a proof.
The case for the singularity theory approach 13
!! !$ !# "#% "#"
Fig. 1.6 Models of the focal set of a generic surface in R3 . The fourth and fifth figures
model the focal sets ε1 and ε2 joining at an umbilic point. The first three figures model
the focal sets ε1 or ε2 . One can have the following generic combinations for the pair
(ε1 , ε2 ): (A2 , A2 ), (A2 , A3 ), (A2 , A4 ), (A3 , A3 ) or vice-versa.
for some smooth function ⌧ (s). The scalar ⌧ (s) is called the torsion of
at s.
At each point on where (s) 6= 0, we have a positively oriented Serret-
Frenet frame {t(s), n(s), b(s)} which moves along the curve. This motion
is described by the Frenet-Serret formulae
8 0
< t (s) = (s)n(s)
n0 (s) = (s)t(s) + ⌧ (s)b(s)
: 0
b (s) = ⌧ (s)n(s).
The curvature and torsion functions determine completely the space
curve up to rigid motions. Indeed,
One can use singularity theory, as in the case of plane curves, to obtain
a great deal of geometric information about space curves (see for example
[Bruce and Giblin (1992)]). We consider below surfaces in R3 generated by
space curves.
and we get
0
xt ⇥ xu (t, u) = (t) ⇥ (t) + u 0 (t) ⇥ (t).
The surface M is singular at (t0 , u0 ) if and only if xt (t0 , u0 ) and
xu (t0 , u0 ) are linearly dependent, equivalently, if and only if xt ⇥ xu (t0 , u0 )
is the zero vector. This occurs when
0
(t0 ) ⇥ (t0 ) + u0 0 (t0 ) ⇥ (t0 ) = (0, 0, 0).
The surface M can be re-parametrised so as to make the singular points
easier to detect. We shall assume that 0 (t) 6= 0 for all t 2 I; such surfaces
are referred to as non-cylindrical ruled surfaces.
Given a non-cylindrical ruled surface M determined by and , there
is a curve : I ! R3 contained in M with
h 0 (t), 0
(t)i = 0 (1.3)
for all t 2 I. To show this, we observe that as the trace of lies on M , we
have
(t) = (t) + u(t) (t) (1.4)
for some smooth function u : I ! R, which we determine as follows. We
have, by di↵erentiating (1.4),
0 0
(t) = (t) + u0 (t) (t) + u(t) 0 (t),
so equation (1.3) becomes
h 0 (t) + u0 (t) (t) + u(t) 0 (t), 0
(t)i = 0. (1.5)
0
The assumption k (t)k = 1 implies h (t), (t)i = 0, so that equation
(1.5) becomes
h 0 (t), 0
(t)i + u(t)h 0 (t), 0
(t)i = 0.
Therefore,
h 0 (t), 0
(t)i
u(t) = . (1.6)
h 0 (t), 0 (t)i
The curve in (1.3) with u(t) as in (1.6) is called the striction curve
of the ruled surface M . It can be shown that the striction curve does not
depend on the choice of the base curve on M (see [do Carmo (1976)]).
Observe that the striction curve need not be a regular curve.
We can now re-parametrise M by taking the striction curve as the
base curve. The new parametrisation of M is given by
y(t, u) = (t) + u (t)
The case for the singularity theory approach 17
h 0 (t) ⇥ (t), 0
(t)i = det( 0 (t), (t), 0
(t)) = 0
for all t 2 I.
An example of a developable surface is the tangent developable of a
space curve, whose rulings are along the tangent directions of the space
curve ( (t) is parallel to 0 (t)).
Classically, non-singular developable surfaces are classified as follows
(see [Vaisman (1984)]).
Fig. 1.8 Generic singularities of developable surfaces: cuspidal edge (left), swallowtail
(centre) and cuspidal cross-cap (right).
Remark 1.3. The result in Theorem 1.6 shows that the generic singulari-
ties of developable surfaces are distinct from those of general ruled surfaces
(compare Theorem 1.4).
is still called the tangent developable surface of the singular space curve
(t) = (t2 , t3 , t4 ) ([Ishikawa (1995)]). This surface has a swallowtail singu-
larity at the origin (this was first observed by Arnol’d in [Arnol’d (1981)]).
We deform the curve within the family of curves ✏ (t) = (t2 , t3 ✏t, t4 ),
with 0 = . The tangent developable of the regular curve ✏ , with ✏ 6= 0,
is
1.4 Notes
We considered briefly the contact of plane curves with circles and showed
how this gives information about the evolute of the curve. This contact
captures also some infinitesimal symmetry of the curve. The locus of centres
of circles tangent to the curve at two or more points is called the Symmetry
Set (SS) of the curve. The SS is used extensively in computer vision and
shape recognition (see for example [Giblin and Brassett (1985); Siddiqi
and Pizer (2008)] and [Damon (2003, 2004, 2006)]).
The contact of a space curve with lines is measured by the singularities
of its projections to planes. The generic singularities of such projections
are obtained in [David (1983)], and in [Dias and Nuño (2008); Oset Sinha
22 Di↵erential Geometry from a Singularity Theory Viewpoint
23
24 Di↵erential Geometry from a Singularity Theory Viewpoint
The first fundamental form, also called the induced metric on M , ex-
presses the way the hypersurface M inherits the metric of the ambient space
Rn+1 . It is a tool for taking measurements on M , such as the length of
curves and angles between curves on M . For instance, the length of a curve
: [a, b] ! M , is defined as
Z b
l( ) = || 0 (t)||dt.
a
We can express the first fundamental form with respect to the basis
Pi=n
B(x) of Tp M at p = x(u) as follows. For w = i=1 wi xui (u) in Tp M , we
have
Xn
Ip (w) = wi wj gij (u)
i,j=1
with
gij (u) = hxui (u), xuj (u)i.
The functions gij are clearly smooth functions on U and satisfy gij (u) =
gji (u) for all i, j = 1, . . . , n and all u in U . These functions form an n ⇥ n-
symmetric matrix (gij ) and the first fundamental form can be written in
matrix form
Ip (w) = wT (gij (u))w.
Submanifolds of the Euclidean space 25
It is worth observing that the matrix (gij (u)) is not singular for any u
in U . Indeed, as it is a symmetric matrix, it is conjugate to a diagonal
matrix D. The scalar product is positive definite, so the diagonal entries
of D are all strictly positive numbers. The determinant of (gij (u)), being
equal to the product of the diagonal entries of D, is therefore distinct from
zero.
The tangent spaces Tp M and TN (p) S n are n-vector spaces in Rn+1 or-
thogonal to N (p) and can be identified. With this identification, the deriva-
tive map dNp of the Gauss map at p is considered as a linear transformation
Tp M ! Tp M.
with
hij (u) = hNui (u), xuj (u)i = hN (u), xui uj (u)i. (2.2)
Theorem 2.2 (The Weingarten formula). The matrix of the shape op-
erator Wp with respect to the basis B(x) of Tp M at p = x(u) is given by
(hji (u)) with
(hji (u)) = (hik (u))(g kj (u))
for (g ij (u)) = (gij (u)) 1 , where the matrices, (gij (u)) and (hij (u)) are,
respectively, those of the first and second fundamental forms of M at p.
That is,
Xn
Nui (u) = hji (u)xuj (u).
j=1
Proof. Since the vectors xu1 (u), . . . , xun (u), N (u) form a basis of Rn+1 ,
there exist scalar functions ↵ij (u) and i (u), i, j = 1, . . . , n such that
Xn
Nui (u) = ↵ij (u)xuj (u) + i (u)N (u).
j=1
Example 2.3. It follows from Example 2.1 and Example 2.2 that hyper-
planes and hyperspheres in Rn+1 are totally umbilic hypersurfaces.
Definition 2.11. The image of the map x⇤ is called the cylindrical pedal
of M and represents the dual hypersurface of M in S n ⇥ R ([Bruce (1981);
Romero Fuster (1983)]).
Proof. We have @hv /@ui (u) = hxui (u), vi. Thus, @hv /@ui (u) = 0 for
i = 1, . . . , n if and only if v is a unit normal vector of M at p = x(u).
Equivalently, if and only if v = ±N (u). ⇤
We fix now v = N (u) at a given point p = x(u) on M . The second
order partial derivatives of hv at u are, for i, j = 1, . . . , n,
@ 2 hv
(u) = hxui uj (u), vi = hxui uj (u), N (u)i = hij (u), (2.4)
@ui @uj
where hij are the coefficients of the second fundamental form as defined
in Definition 2.7. The second order derivatives of Hv at u form an n ⇥ n-
symmetric matrix, called the Hessian matrix of hv at u, which we denote
by
✓ 2 ◆
@ hv
H(hv )(u) = (u) .
@ui @uj
An immediate consequence of expression (2.4) and Corollary 2.1 is the
following.
Proof. Statement (i) follows from Corollary 2.2 as det H(hv )(u) = 0 if
and only if K(u) = 0. Statement (ii) follows from the expressions (2.4) and
Proposition 2.3. ⇤
Proof. From the expressions (2.5), rank H(da )(u) = 0 if and only if
6= 0 and hij (u) = gij (u)/ for all i = 1, . . . , n. Using matrix notation,
rank H(da )(u) = 0 if and only if
1
(hij (u)) = (gij (u)).
This is equivalent to the matrix (hji (u)) of the shape operator in the
basis B(x) being given by
1
(hji (u)) = (hij (u))(gij (u)) 1
= In ,
vN 2 Np M.
In what follows we denote by X (U ) (resp. N (U )) the set of di↵erentiable
vector fields tangent to M (resp. normal to M ).
Let r be the Riemannian connection on Rn+r . If !1 and !2 are local
vector fields on M, and ! 1 and ! 2 are local extensions to Rn+r , let
T
r! 1 ! 2 = r ! 1 ! 2
be the Riemannian connection relative to the metric induced on M.
Definition 2.14. The second fundamental form of the immersion x : U !
Rn+r is the mapping
B : X (U ) ⇥ X (U ) ! N (U ),
given by B(!1 , !2 ) = r!1 ! 2 r!1 !2 .
Submanifolds of the Euclidean space 37
All the concepts and results in section 2.1 derived from the Gauss map
and its derivative can be carried over to similar concepts along a given
normal vector field ⌫˜ on a higher codimension submanifold M of Rn+r .
Definition 2.18. The linear map Wp⌫ is called the shape operator of M at
p along the normal vector field ⌫, or simply, the ⌫-shape operator of M at
p.
with
(hji )⌫ = (h⌫ik )(g kj ),
where all the functions are evaluated at u. As a consequence, we have the
following Weingarten formula along ⌫
n
X
(⇡ T ⌫
G ) ui = (hji )⌫ xuj .
j=1
Submanifolds of the Euclidean space 39
Proof. Since {xu1 (u), . . . , xun (u), ⌫1 (u), . . . , ⌫r (u)} is a basis of Tp Rn+r
at p = x(u), there exist smooth functions ↵ij and ik on U such that
n
X r
X
G⌫ui = ↵ij xuj + k
i ⌫k .
j=1 k=1
Thus,
The expression for (⇡ T G⌫ )ui follows from the fact that the vectors
⌫i (u) are normal vectors. ⇤
⌫
det(h⌫ij (u))
K (u) = .
det(g↵ (u))
Remark 2.2. Any unit normal vector ⌫0 at a fixed point p = x(u0 ) can be
extended locally to a unit normal vector field ⌫ along M with ⌫(u0 ) = ⌫0 .
Di↵erentiating the identity h⌫(u), xuj (u)i = 0 and evaluating at u0 yields
Proof. The condition on ⌫ to be parallel gives ⌫ui (u) = ⇡ T ⌫ui (u), and
adding to it the condition on the submanifold to be totally ⌫-umbilic yields
The vectors xu1 (u), xu2 (u), . . . , xun (u) are lineally independent, so
⌫ui (u) = 0 for i = 1, . . . , n. This means that ⌫ (u) is a constant, say
⌫
.
Suppose that ⌫ 6= 0. Then the map x(u) + (1/⌫ )⌫(u) is a constant
map a, which implies that M is contained in the hypersphere of centre a
and radius 1/|⌫ |.
If ⌫ = 0, then ⌫ is a constant vector. It follows that hx(u), ⌫i is a
constant map c, so M is contained in the hyperplane H(⌫, c). ⇤
Submanifolds of the Euclidean space 41
Theorem 2.7. For " small enough, the canal hypersurface of M is locally
a smooth codimension 1 submanifold of Rn+r .
Proposition 2.12. For " small enough, the Gauss map G : CM (") ! S n
of the canal hypersurface of CM (") of M is given by
G(p, ⌫) = ⌫.
Proof. Following the proof of Theorem 2.7, we take y : U ⇥ S r 1 !
Rn+r , with y(u, µ) = x(u) + "N (u, µ), as a parametrisation of the canal
hypersurface. Then the map G is given by G(u, µ) = N (u, µ). (Observe
that N (u, µ) 2 (Np M )1 , with p = x(u), and (Np M )1 can be considered a
subset of S n .)
We have hN (u, µ), yui (u, µ)i = 0 as N (u, µ) 2 Np M at p = x(u), and
hN (u, µ), yµj (u, µ)i = 0 as the vectors ⌫i (u), i = 1, . . . , r are orthonormal.
It follows that N (u, µ) is a unit orthogonal vector to CM ("). ⇤
Proposition 2.13. Let K(p, ⌫) denote the Gauss-Kronecker curvature of
the canal hypersurface CM (") of M at (p, ⌫) and let K ⌫ (p) denote the
Lipschitz-Killing curvature of M at p along the normal vector ⌫. Then,
K(p, ⌫) = K ⌫ (p).
Proof. We take a parametrisation of M as before and choose an or-
thonormal frame along M in such a way that ⌫(u0 ) = ⌫1 (u0 ) at the
point p = x(u0 ). Then, with the notation of the proof of Theorem 2.7,
N (u0 , µ0 ) = ⌫(u0 ), with µ0 = (1, 0, . . . , 0) 2 U1+ .
By Theorem 2.5, we have, for i = 1, . . . , n,
X n
Nui (u0 , ⌫(u0 )) = ( ij )⌫ (u0 )xuj (u0 )
j=1
and from the proof of Theorem 2.7, for j = 2, . . . , r,
Nµj (u0 , µ0 ) = ⌫j (u0 ).
The matrix of the shape operator of CM (") at (u0 , µ0 ) with re-
spect to the basis {xu1 (u0 ), . . . , xun (u0 ), ⌫1 (u0 ), . . . , ⌫r (u0 )} has determi-
nant det( ( ij )⌫ (u0 )) which is precisely K ⌫ (u0 ). ⇤
Corollary 2.4. Let p be a point on M and let ⌫0 be a unit normal vector
at p. Then p is a ⌫0 -parabolic point of M if and only if (p, ⌫0 ) is a parabolic
point of the canal hypersurface CM (") of M , with " small enough.
where n(s) and b(s) are, respectively, the principal normal and binormal
vectors of at s.
The first order partial derivatives of y are
ys = (1 " cos ✓)t + "⌧ sin ✓n "⌧ cos ✓b
y✓ = "( sin ✓n + cos ✓b)
where , ⌧, t, n, b are evaluated at s. The matrix of the first fundamental
form is given by
✓ ◆
(1 " cos ✓)2 + "2 ⌧ 2 "2 ⌧
(gij ) = .
"2 ⌧ "2
The second order partial derivatives of y are given by
yss = "(0 cos ✓ + ⌧ 2 sin ✓)t + ("⌧ 0 sin ✓ + (1 " cos ✓) "⌧ 2 cos ✓)n
"(⌧ 0 cos ✓ + ⌧ 2 sin ✓)b,
ys✓ = "( sin ✓t + ⌧ cos ✓n + ⌧ sin ✓t),
y✓✓ = "(cos ✓n + sin ✓b).
The Gauss map of the canal surface is given by N (s, ✓) = cos ✓n(s) +
sin ✓b(s), so we can compute the matrix of the second fundamental form
(hij ) using the relations (2.2) and obtain
✓ ◆
(1 " cos ✓) cos ✓ "⌧ 2 "⌧
(hij ) = .
"2 ⌧ "
It follows by Corollary 2.1 that the Gaussian curvature of the canal
surface is given by
(s) cos ✓
K(s, ✓) = .
"(1 "(s) cos ✓)
In particular, the parabolic set of the canal surface consists of the two
curves (s) ± b(s).
Chapter 3
The birth of singularity theory can be traced back to the pioneering work
of Whitney [Whitney (1955)] where he showed that maps from the plane
to the plane have, in general, only folds and cusps singularities. Mather
introduced in his seminal papers [Mather (1968, 1969a,b,c, 1970)] groups
that act on the set of map-germs and set the foundations for the study of
finite determinacy of map-germs, a concept linked to the existence of versal
deformations and versal unfoldings. A versal deformation, is in some sense,
one that contains all possible deformations of the initial map-germ. This
has a wide range of applications in mathematics and other fields of science
(see for example [Koenderink (1990); Poston and Stewart (1996); Thom
(1983)]).
We give in this chapter some basic definitions and state the results that
we need in other chapters. For beginners in singularity theory, we rec-
ommend the books [Arnol’d, Guseı̆n-Zade and Varchenko (1985); Bröcker
(1975); Gibson (1979); Martinet (1982)] and [Bruce and Giblin (1992)] for
application to the geometry of curves. C. T. C. Wall’s survey article [Wall
(1981)] remains the first port of call for people embarking on the study of
finite determinacy of map-germs.
The definitions and results are presented for maps f : U ! Rm of class
C 1 from an open subset U of Rn to Rm , but they also hold for smooth
map f : M ! N , where M and N are any smooth dimensional manifolds.
45
46 Di↵erential Geometry from a Singularity Theory Viewpoint
map
(df )p : Rn ! Rm
is not maximal, that is, if rank(df )p < min(n, m). The point p is then said
to be a singular point of f . Otherwise, we say that f is non-singular at p
and p is a regular point of f . The critical set of f , denoted by ⌃(f ), is the
set of singular points of f , that is,
⌃(f ) = {p 2 U | rank(df )p < min(n, m)}.
The criminant of f , denoted by Cr(f ), is
Cr(f ) = {p 2 U | rank(df )p < m }.
When n m, Cr(f ) = ⌃(f ), and when n < m, Cr(f ) = U .
The discriminant of f , denoted by (f ), is the image of Cr(f ) by f :
(f ) = f (Cr(f )).
Observe that when n < m, (f ) = f (U ). The set f (⌃(f )) is called
the set of critical values of f. The above definitions can be localised at a
point p 2 Rn . A germ f : (Rn , p) ! Rm is said to be singular if one of
its representatives is singular at p. This definition does not depend on the
choice of the representative of f at p as any two of these are identical in
some neighbourhood of p.
The critical set (respectively, the criminant) of a map-germ f , still de-
noted by ⌃(f ) (respectively, Cr(f )), is the set germ (⌃(f˜), p) (respectively,
(Cr(f˜), p)), where f˜ is a representative of f in some neighbourhood U of p.
Again, this definition does not depend on the choice of the representative.
Likewise, the discriminant of f is (f ) = ( (f˜), f (p)).
Fig. 3.1 The fold (left) and cusp (right) singularities realised by the projection of a
surface to a plane. The singular sets and discriminants are the thick curves.
seen in Figure 3.1, right, as the image of the critical set of the projection
⇡|M .
(4) Consider the map-germ f : (R2 , 0) ! (R3 , 0) with f (x, y) =
(x2 , y, xy). The di↵erential map df at (x, y) is represented by the matrix
0 1
2x 0
@ 0 1A.
y x
This is singular if and only if x = y = 0, so the critical set ⌃(f ) is the origin
in the source. Its discriminant, which is its image, is as shown in Figure 3.2
and is called a cross-cap or a Whitney umbrella.
Remark 3.1. The group K is a natural one to use when one seeks to under-
stand the singularities of the zero level-sets of map-germs in Mn .E(n, m).
If two germs are K-equivalent, then their zero level-sets are di↵eomorphic
The action of the group A is finer than that of K. If two map-germs F and
G are A-equivalent, then G = k F h 1 for some (h, k) 2 A, so the level
sets G 1 (c) and F 1 (k 1 (c)) are di↵eomorphic for any c close to 0 2 Rm .
Therefore, the group A preserves also the smooth structure of nearby level
sets to the zero level set.
j k h.j k f = j k (h.f ).
#$!
!
"
tf : θn → θf
φ &→ df ◦ φ
54 Di↵erential Geometry from a Singularity Theory Viewpoint
Then f is k-G-determined.
Theorem 3.5. (i) For each map-germ f and a Mather group G, the fol-
lowing are equivalent:
(a) f is G-finite;
(b) f has a G-versal unfolding ;
(c) f has a Ge -versal unfolding.
(ii) The least number a0 of parameters for a G-versal (resp. Ge -versal)
unfolding is cod(f, G) (resp. cod(f, Ge )). A versal unfolding (a0 , F ) is called
miniversal.
(iii) Miniversal unfoldings are unique up to equivalence. Any versal
unfolding is equivalent to a suspension of a miniversal unfolding. Versal
unfoldings (a, F ) and (a, G) of f are equivalent.
58 Di↵erential Geometry from a Singularity Theory Viewpoint
g(x1 , . . . , xr ) + Q(xr+1 , . . . , xn ),
Definition 3.5. We say that the two families F and G are stably P -R+ -
equivalent if F + Q and G + Q0 are P -R+ -equivalent.
Definition 3.7. We say that the two families F and G are stably P -K-
equivalent if F + Q and G + Q0 are P -K-equivalent.
Proof. The result follows by applying Theorem 3.14 and Theorem 3.7 to
the normal forms in Table 3.1. ⇤
CF = {(x1 , u1 ) | u1 = −3x21 },
BF = {0}.
66 Di↵erential Geometry from a Singularity Theory Viewpoint
It follows that
CF = {(x1 , x2 , u1 , u2 , u3 ) | u1 = x21 + x22 2u3 x1 ,
u2 = 2x1 x2 2u3 x2 }
and
BF = {(u1 , u2 , u3 ) | u1 = x21 + x22 2u3 x1 ,
u2 = 2x1 x2 2u3 x2 ,
u23 = x21 + x22 , with (x1 , x2 ) 2 (R2 , 0) }.
The singularity of F (x, 0) is called elliptic umbilic and the bifurcation
set of F is as in Figure 3.4 (called pyramid).
(6) The D4+ -singularity
For F (x, u) = x31 + x23 + u3 x1 x2 + u2 x2 + u1 x1 , we have
@F
@x1 = 3x21 + u3 x2 + u1 ,
@F
@x1 = 3x22 + u3 x1 + u2 ,
and
@2F @2F
! ✓ ◆
@x1 2 @x1 @x2 6x1 u3
@2F @2F
= .
@x2 @x2 @x22
u3 6x2
Then,
CF = {(x1 , x2 , u1 , u2 , u3 ) | u1 = 3x21 u3 x2 ,
u2 = 3x22 u3 x1 }
and
BF = {(u1 , u2 , u3 ) | u1 = 3x21 u3 x2 ,
u2 = 3x22 u3 x1
u23 = 36x1 x2 , with (x1 , x2 ) 2 (R2 , 0) }.
The singularity of F (x, 0) is called hyperbolic umbilic and the bifurcation
set of F is as in Figure 3.4 (called purse).
(7) The D5 -singularity
Here too we take a di↵erent normal form F (x, u) = x21 x2 + 16 x42 + u4 x22 +
u3 x21 + u2 x2 + u1 x1 . We have
@F
@x1 = 2x1 x2 + 2u3 x1 + u1 ,
@F
@x2 = x21 + 23 x32 + 2u4 x2 + u2 ,
and
@2F @2F
! ✓ ◆
@x21 @x1 @x2 2x2 + 2u3 2x1
@2F @2F
= 2 .
@x2 @x1 @x22
2x1 2x2 + 2u4
68 Di↵erential Geometry from a Singularity Theory Viewpoint
We deduce that
CF = {(x1 , x2 , u1 , u2 , u3 , u4 ) | u1 = 2x1 x2 2u3 x1 ,
u2 = x21 23 x32 2u4 x2 }
and
BF = {(u1 , u2 , u3 , u4 ) | u1 = 2x1 x2 2u3 x1 ,
u2 = x21 23 x32 2u4 x2 ,
(x2 + u3 )(x22 + u4 ) x21 = 0, with (x1 , x2 ) 2 (R2 , 0) }.
Theorem 3.16. Let F and G be two P -K-equivalent families of germs of
functions. Then the discriminants DF and DG are di↵eomorphic.
(5) D4 -singularity
For F (x, u) = 13 x31 x1 x22 + u4 (x21 + x22 ) + u3 x2 + u2 x1 + u1 ,
Remark 3.2. It is not difficult to check that the Boardman symbol of the
catastrophe map-germ associated to an Ak -singularity is S1k 1 , where the
subindex k 1 refers to the number of entries of the subindex 1 in the sym-
bol, so S11 = S1,0 , S12 = S1,1,0 , S13 = S1,1,1,0 and so on. The Boardman
symbol of the catastrophe map-germ associated to a Dk -singularity is S2 .
3.11 Notes
Our aim in this chapter is to set the singularity theory notation and state
the results we need in this book. It is far short from being a survey of
results in the area of singularities of map-germs. We give below some of
the research directions in this area. The few mentioned references (mainly
books) are meant as appetisers.
We stated the results for C 1 (real) map-germs as these arise in applica-
tions to di↵erential geometry. The study of C r map-germs and of complex
holomorphic map-germs is very rich. The survey article of Wall ([Wall
(1981)]) touches on this.
J. Milnor ([Milnor (1968)]) considered the topology of the fibre of a
holomorphic function. If the germ of a map f at its singular point is R-
finite, then the intersection of the singular fibre with a small sphere has the
homotopy type of the wedge of µ(f )-spheres, where
µ(f ) = cod(f, R)
f /kf k : S"n 1
\ K ! S1
on any sufficiently small sphere S"n 1 , outside some small tubular neigh-
bourhood NK of the link K = f 1 (0) \ S"n 1 . This is known as Milnor
Fibration Theorem. (See [Brieskorn and Knörrer (1986); Wall (2004)] for
the case n = 2, and [Seade (2007)] for the real case.)
A classical example of a family of non-equivalent R-finite germs is
f (x, y) = xy(x y)(x + y), for 6= 0, 1. The parameter is a mod-
ulus and represents the cross-ratio of the four lines f 1 (0). If one relaxes
the equivalence relation and considers germs of homeomorphisms in the
source instead of di↵eomorphisms, then all the members of the family f
become topologically equivalent. We say that the family f is topologi-
cally trivial. The problem of determining whether a family of map-germs
is topologically trivial was and still remains a major problem in singularity
theory ([du Plessis and Wall (1995)]; see also [Damon (1988, 1992)] for
topological triviality for Damon’s geometric subgroups).
A G-invariant of a map-germ f is a number that depends only on the
G-orbit of f . For example, the Milnor number µ is an R-invariant of germs
of functions (it is also a topological invariant). It turns out that the con-
stancy of µ in a family of germs of functions implies that the family is
Singularities of germs of smooth mappings 71
73
74 Di↵erential Geometry from a Singularity Theory Viewpoint
(1986a)] (see also [Montaldi (1991)]) and some of them are reproduced in
section 4.1.
The maps defining the contact between a submanifold of Rn and model
submanifolds come naturally in a family of map-germs. One can ask if such
a family is versal and what possible singularities one can expect in the mem-
bers of the family. This leads to the notion of transversality and genericity
which are dealt with in section 4.2. The remaining sections give applica-
tions of Montaldi’s genericity theorems to the study of the families of height
functions, distance squared functions and projections of hypersurfaces to
Rn .
We start with the example of plane curves. Let ↵(t) = (x(t), y(t)) be a
regular plane curve and let be another plane curve given as the zero set
of a smooth function F : R2 ! R. We say that the curve ↵ has (k + 1)-
point contact at t0 with the curve if t0 is a zero of order k of the function
g(t) = F (↵(t)) = F (x(t), y(t)), that is,
where g (i) denotes the ith -derivative of the function g. Using the singularity
theory terminology in Chapter 3, the curve ↵ has (k + 1)-point contact at
t0 with if and only if the function g has an Ak -singularity at t0 .
A particular case of interest is when the curve is a circle or a line.
If for instance is a circle, then the singularities of g reveal geometric
information about the curve ↵ (see Proposition 1.3 in Chapter 1).
We extend in this chapter the notion of contact between plane curves
and define the contact between two submanifolds of an n-dimensional man-
ifold.
(M, x0 )
=f g
- (Rk1 , 0)
H *
HH
g H ⇡1
H
j
H =(f,f¯)
(Rn , 0) - (Rk1 , 0) ⇥ (Rk2 , 0)
* HH
ḡ ⇡2HH
¯=f¯ ḡ
H
j
(N, 0) - (Rk2 , 0)
where ⇡1 and ⇡2 are the canonical projections to the first and second com-
ponents respectively.
76 Di↵erential Geometry from a Singularity Theory Viewpoint
Montaldi proved the following result which shows that the K-equivalence
class of the contact map-germ f g determines the contact class of the pair
of submanifolds M and N.
The proof of Theorem 4.1 is done in two steps. The first step deals with
the equidimensional case, i.e., when dim Mi = dim Ni = m, and the second
deals with the general case.
Proof of Theorem 4.1 for the equidimensional case. Suppose
that the pairs (M1 , N1 ) and (M2 , N2 ) have the same contact type. Let
H be the di↵eomorphism of (Rn , 0) taking g1 (M1 ) to g2 (M2 ) and N1 to
N2 . As H|g1 (M1 ) : g1 (M1 ) ! g2 (M2 ) is a di↵eomorphism, there exists
a di↵eomorphism h : M1 ! M2 , such that H g1 = g2 h. We also
have (f2 H) 1 (0) = f1 1 (0). Then we can write each coordinate function
(f2 H)j of f2 H as
such that f 1 (0) is the linear subspace Rn k ⇥ {0}. Then, equation (4.1)
follows from Hadamard’s Lemma. Now, the map f2 H is also a submersion,
and as f1 (0) = 0, it follows that the k ⇥ k matrix [a1 (y), . . . , ak (y)] is
invertible in a neighbourhood of 0, where the vectors ai (y) have coordinates
(aij (y)).
We now define ✓ : Rn ⇥ Rk ! Rk by
and ✓0 : M1 ⇥ Rk ! Rk by
The proof of the general case of Theorem 4.1 requires the following two
lemmas which relate the contact between two submanifolds in Rn with that
of their suspensions.
Proof. The sufficient part of the statement follows from the fact that the
suspension of the di↵eomorphism taking M1 to M2 and N1 to N2 takes M10
to M20 and N10 to N20 .
For the converse, let H 0 : (Rn ⇥ Ra , y10 ) ! (Rn ⇥ Ra , y20 ) be the
di↵eomorphism-germ such that H 0 (M10 , y10 ) = (M20 , y20 ), H 0 (N10 , y10 ) =
(N20 , y20 ). We write H 0 = (H1 , H2 ), where H1 : (Rn+a , y10 ) ! (Rn , y1 ) and
H2 : (Rn+a , y10 ) ! (Ra , 0). As H 0 is a di↵eomorphism it follows that H1
and H2 are submersions.
Suppose that there exists a map-germ ⌘ : (Rn , y1 ) ! (Ra , 0) such that
Then the map H will be the required map. Indeed, for all x 2 (M1 , y1 )
we get (x, ⌘(x)) 2 M10 , then H 0 (x, ⌘(x)) 2 M20 and hence H(x) =
H1 (x, ⌘(x)) 2 M2 . Similarly, y 2 N1 gives (y, ⌘(y)) = (y, 0) 2 N10 , hence
H 0 (y, 0) 2 N20 and then H(y) 2 N2 .
We show now that the map ⌘ does exist. For (b) it is enough to show that
the derivative map dH is injective. We write U = d⌘ and dH1 = (A, B),
where A : Rn ! Rn and B : Ra ! Rn are linear mappings. Since (A, B)
has rank n, we require U so that A + BU has rank n.
For condition (a) we require U to be zero on T N1 . Since A restricted to
T N1 is injective, we can show that such U exists. ⇤
gi0 (x, u) !
7 (gi (x), u, 0),
fi0 (y, u, v) ! 7 (fi (y), u),
without changing the K class of the contact maps. Then fi0 gi0 (x, u) =
(fi gi (x), u), so fi0 gi0 is a suspension of fi gi , and the result follows by
Theorem 3.3 in Chapter 3. ⇤
Contact between submanifolds of Rn 79
Proof of Theorem 4.1 of the general case. The proof necessary part
is the same as that of the equidimensional case. For the sufficient part, if
dim(Mi ) 6= dim(Ni ), we can suspend the submanifold of lower dimension,
say Mi , with Ra , where a = dim(Ni ) dim(Mi ). This gives Mi0 = Mi ⇥ Ra
and Ni0 = Ni ⇥ {0} in Rn ⇥ Ra and the map-germs gi0 (x, a) = (gi (x), a) and
fi0 (y, a) = (fi (y), a).
We then have the following implications, which give the result:
(b)
f10 g10 ⇠K f20 g20 =) K(M10 , N10 ) = K(M20 , N20 )
*(a) +(c)
f1 g 1 ⇠ K f2 g 2 K(M1 , N1 ) = K(M2 , N2 )
where (a) follows by Lemma 4.4, (b) by the proof of the result in the
equidimensional case and (c) by Lemma 4.3. ⇤
(Y1 (c)) = Y2 (c), where Yi (c) = fi 1 (c) for all c 2 (R, 0). In this case we
write K(g1 (M1 ), Ff1 ; y1 ) = K(g2 (M2 ), Ff2 ; y2 ).
Here too we can replace the ambient space Rn by any manifold. The
following result characterises the contact of a hypersurface with a foliation
in terms of the R+ -singularities of functions.
4.2 Genericity
We shall apply the results in section 4.1 in the following way. We con-
sider the contact of a manifold M immersed in Rn with families of subman-
ifolds (more specifically, k-spheres and k-planes) and show that this contact
is generic for a dense subset of immersions of M in Rn . We also list the
generic contacts.
First we recall the notion of transversality and Thom’s transversality
theorem ([Golubitsky and Guillemin (1973)]).
TZ = {g 2 C 1 (X, Y ) | j r g t Z}
The key to the results in this section is the next proposition which is a
variant of Thom’s transversality theorem ([Thom (1956)], [Golubitsky and
Guillemin (1973)]).
82 Di↵erential Geometry from a Singularity Theory Viewpoint
TZ = {g 2 C 1 (X, Y ) | (j r g, id) t Z}
is smooth. By Theorem 4.1, the contact between g(M ) and a model sub-
manifold Nb at the point g(x) is given by the K-singularities of the com-
posite map fb g at x.
We denote by g,b : M ! Rk the map g,b (x) = fb g(x). We also
denote by g : M ⇥ B ! Rk the family of maps given by
Denote by Jyr (M, Rk ) the subset of the jet space J r (M, Rk ) of jets with
target y. In our application, 0 2 Rk is a preferred target as it is the target
of fb g. We consider all maps with non-zero target as being K-equivalent,
indeed if two map-germs have non-zero target their local algebras are iso-
morphic and equal to E(n). Moreover if one map has target 0, and another
does not, then they are not K-equivalent. Thus, any K-invariant subman-
ifold of J r (M, Rk ) is either all of the complement of J0r (M, Rk ) or is a
submanifold of J0r (M, Rk ).
RW = {g 2 Imm(M, Rn ) : j1r g t W}
is residual in Imm (M, Rn ), where j1r is the r-jet with respect to the first
variable, so j1r g maps M ⇥ B to J r (M, Rk ). Moreover, if B is compact and
W is closed then RW is open and dense.
Contact between submanifolds of Rn 83
j r g⇥id
- I r (M ⇥ Rn ) ⇥ B - J r (M, Rk )
r
M ⇥B
? jr g
?
M - I r (M, Rn )
is residual in Emb(M, Rn ).
If W is closed and M is compact, then in (i) and (ii) RW is open and
dense.
Proof. See [Montaldi (1986a)]. The arguments of the proof of (ii) hold
only for embeddings g : M ! Rn . ⇤
Remark 4.1. Theorem 4.4 still holds if one replaces the ambient space Rn
by a manifold Z and the target space by another manifold Q.
of jet spaces. Mather in a series of papers found the tools to build the
bridge between versality and transversality. His results give a method to
describe the generic singularities in terms of transversality of the r jet of
the mapping to a J r G invariant stratification of J r (M, Rk ). The idea is to
determine the pair of dimensions (m, k) for which the relevant strata of
this stratification are the r-jets of simple G-orbits. When G = K, Mather
computed in [Mather (1971)] the codimension (m, k) of the set of K non-
simple singularities in J0r (M, Rk ) - the fibre of J r (M, Rk ) over a point
(x, 0) in M ⇥ Rk - for sufficiently large r. As we will see, when (m, k)
is sufficiently large, transversality to the K-invariant stratification in jet
space means avoiding non-simple singularities. The pairs (m, k) satisfying
this condition form the nice dimensions.
The action of the group K does not move the origin in Rk , so the Ke -
tangent space to any singular jet is contained in J0r (M, Rk ). Hence, the
Ke -codimension of the set of all non-simple singularities in the jet space
J r (M, Rk ) is (m, k) + k.
For a given immersion g : M ! Rn , it also follows that the dimension of
the image of the associated jet-extension map j1r g : M ⇥ B ! J r (M, Rk )
is m + d, where d = dim B, and g is the composite F g.
Suppose that m + d < (m, k) + k (this will be the case for all the
applications we consider in this book). Let {W1 , . . . , Ws } be the finite set
of all the K-orbits in J r (M, Rk ) of Ke -codimension less than m + d and let
{Ws+1 , . . . , Wt } be a finite stratification of the complement of W1 [. . .[Ws .
Let R be the residual set of immersions given by the intersection of the sets
RWi , i = 1, . . . , t in Theorem 4.3. For g 2 R, it follows from the condition
m + d < (m, k) + k that the associated jet-extension map j1r g misses
the strata Wi for i > s, and is transverse to the strata Wi for i s. We
call the immersions g 2 R generic immersions.
When m+d (m, k)+k, a generic embedding does not avoid in general
the non-simple singularities. In this case, a stratification of J0r (M, Rk ) is
given by the strata which are the simple K-orbits together with strata
which are the union of the non-simple K-orbits parametrised by the moduli
(usually excluding some exceptional values of the moduli).
When the dimensions m, d, k are such that the non-simple singularities
are not encountered for a generic immersion, the strata of codimension
smaller than or equal to m + d are K-orbits, and the transversality of j1r g
to the stratification in jet space, for sufficiently high values of r gives that
g is Ke -versal. If, on the other hand, the non-simple singularities are
present, then all the singularities that arise for a generic immersion cannot
Contact between submanifolds of Rn 87
be presented transversely.
The formulae for the codimension r (m, k) of the algebraic variety con-
sisting of all non-simple singularities in J0r (M, Rk ), are given by Mather
in [Mather (1971)]. The number r (m, k) is a decreasing function of r,
and (m, k) is defined to be infr r (m, k). The formulae for (m, k) are as
follows:
Case I: m k
(m, k) = 6(k m) + 8 if k m 4 and m 4
= 6(k m) + 9 if 0 < k m < 3 and m > 4, or
if m = 3
= 7(k m) + 10 if m = 2
=1 if m = 1
Remark 4.2. In some applications in this book, for instance when con-
sidering the family of projections into linear spaces (section 4.6), we take
G = A or R+ if k = 1. The orbits of the extended actions of these groups
are invariant by translations y ! y + c, y and c in Rk . Then, for G = A, R+ ,
and f : (Rm , 0) ! (Rk , 0), the Ge -tangent space of f contains the vector
@ @
subspace of ✓f generated by { @y i
f }, i = 1, . . . , k where { @y i
, i = 1, . . . , k}
are the generators of ✓k .
When k = 1, the R+ -simple germs coincide with the K-simple germs
coincide. When k > 1 and G = A, a formula for the codimension A (m, k)
of the set of the Ae -non-simple singularities is not known in general (see
[Rieger and Ruas (2005)], [Oset Sinha, Ruas and Atique (2015)] for
some partial results). In this case, we proceed as follows. With similar
arguments as above, for each pair (m, k), if the A-classification of map-
germs (Rm , 0) ! (Rk , 0) of Ae -codimension d = dim(B) is finite, that
is, all orbits in this classification are A-simple, then we let {W1 , . . . , Ws }
to be the finite set of A-orbits in J r (M, Rk ) of Ae -codimension d,
and let {Ws+1 , . . . , Wt } to be a finite stratification of the complement of
W1 [ . . . [ Ws . In this case, we set W to be this stratification. Again, g is
a generic immersion if j r g t W.
88 Di↵erential Geometry from a Singularity Theory Viewpoint
Theorem 4.5. (i) For an open and dense set of immersions of a smooth
curve C in Rn , n e (resp. H) is locally Ke -versal (resp.
2, the family H
P-R+ -versal).
Contact between submanifolds of Rn 89
Proof. The argument is similar to that in the proof of Theorem 4.6. Here
m = 2 and k = 1 and one needs to stratify J0r (M, R). From Theorem 4.5
the relevant strata of the stratification in jet space are the orbits of the
simple singularities Ak , Dk , E6 , E7 , E8 .
Now, the generic singularities of the height function must have codi-
mension less than or equal to n. Hence, when n = 3, 4, 5 we get respectively
those in (i), (ii) and (iii). ⇤
Remark 4.3. Theorem 4.7 gives the possible singularities of the height
function on an immersed surface M in Rn , n = 3, 4, 5. The singularities
describe the contact of g(M ) with hyperplanes, where g is a generic immer-
sion M ! Rn . We extract in Chapters 6, 7, 8 extrinsic geometric properties
90 Di↵erential Geometry from a Singularity Theory Viewpoint
4.7 Notes
97
98 Di↵erential Geometry from a Singularity Theory Viewpoint
on R2n . The vector space R2n equipped with this symplectic form is the
standard model of a linear symplectic manifold.
!= d .
Example 5.4. Consider the linear symplectic space R2n and let ⇡ : R2n !
Rn be defined by ⇡(x1 , . . . , xn , p1 , . . . , pn ) = (x1 , . . . , xn ). Then ⇡ is a
Lagrangian fibration.
(Example 5.1).
Let F : (Rk ⇥Rn , 0) ! (R, 0) be an n-parameter family of germs of func-
tions from (Rk , 0) to (R, 0), and denote by (q, x) = (q1 , . . . , qk , x1 , . . . , xn )
the coordinates in Rk ⇥ Rn . In Chapter 3, we associated some set germs
the family of functions F . The germ of the catastrophe set of F is the set
germ
⇢
@F @F
(CF , 0) = (q, x) 2 (Rk ⇥ Rn , 0) (q, x) = · · · = (q, x) = 0
@q1 @qk
and its bifurcation set is the set germ
⇢ ⇣ @2F ⌘
n
(BF , 0) = x 2 (R , 0) 9(q, x) 2 CF such that rank (q, x) < k .
@qi @qj
Let ⇡2 : (Rk ⇥Rn , 0) ! (Rn , 0) denote the canonical projection and con-
sider the map-germ ⇡CF which is given by the restriction of the projection
⇡2 to (CF , 0). Thus,
⇡CF : (CF , 0) ! (Rn , 0)
102 Di↵erential Geometry from a Singularity Theory Viewpoint
with ⇡CF (q, x) = x for any (q, x) 2 (CF , 0). The map-germ ⇡CF is the
catastrophe map of F .
We say that F is a Morse family of functions if the map-germ F :
(R ⇥ Rn , 0) ! (Rk , 0) given by
k
✓ ◆
@F @F
F (q, x) = ,..., (q, x)
@q1 @qk
is not singular. When F is a Morse family of functions, (CF , 0) is a germ of
a smooth submanifold of (Rk ⇥ Rn , 0) of dimension n. We immerse (CF , 0)
in the cotangent bundle T ⇤ Rn by the map-germ L(F ) : (CF , 0) ! T ⇤ Rn
defined by
✓ ◆
@F @F
L(F )(q, x) = x, (q, x), . . . , (q, x) .
@x1 @xn
(One can check that this is indeed an immersion.) We have
Xn
⇤ @F
L(F ) = dxi |CF = dF |CF
i=1
@x i
so that
L(F )⇤ ! = L(F )⇤ d = dL(F )⇤ = d(dF |CF ) = (ddF )|CF = 0.
This proves that L(F ) is a germ of a Lagrangian immersion.
The family of map-germs F is called the generating family of the germ
of the Lagrangian submanifold L(F )(CF ) of T ⇤ Rn .
We have the following fundamental theorem.
Remark 5.1. The bifurcation set of F is the set of critical values of the
catastrophe map-germ ⇡CF . Since ⇡CF (q, x) = ⇡ L(F )(q, x), it coincides
with the caustic C(L(F )).
Proof. Statements (1) and (2) are equivalent by Theorem 5.4. By defi-
nition, statement (2) implies statement (3). Since ⇡ L(F ) and ⇡ L(G)
Lagrangian and Legendrian Singularities 105
Then
↵ ^ d↵n = ( 1)n(n+1)/2 dx1 ^ . . . ^ dxn ^ dy ^ dp1 ^ . . . ^ dpn ,
which is the standard volume form in R2n+1 . Thus ↵ is a contact form. It
defines the standard contact structure on M = R2n+1 .
Xn
1 ⇤ @f
j f ↵ = df (x) (x)dxi = df (x) df (x) = 0.
i=1
@x i
is not singular.
When F is a Morse family of hypersurfaces, the set-germ
⇢
@F @F
⌃F = (q, x) 2 (Rk ⇥ Rn , 0) | F (q, x) = (q, x) = · · · = (q, x) = 0
@q1 @qk
is a germ of a smooth (n 1)-dimensional submanifold of (Rk ⇥ Rn , 0).
Then we have the map-germ L (F ) : (⌃F , 0) ! P T ⇤ Rn defined by
✓ ◆
@F @F
L (F )(q, x) = x, (q, x) : · · · : (q, x) .
@x1 @xn
Pn
By definition the local contact form for K is given by ↵ = i=1 ⇠i dxi , where
((x1 , . . . , xn ), [⇠1 : · · · : ⇠n ]) are homogeneous coordinates of P T ⇤ (Rn ).
Therefore, we have
Xn
⇤ @F
L (F ) ↵ = dxi |⌃F = dF |⌃F = d(F |⌃F ) = 0.
i=1
@xi
We have P T ⇤ Rn ⌘ Rn ⇥ (RP n 1 ⇤
) , so we get
i1 (p` ) = (a, [⌫` ]) and i2 (qk ) = (a, [⇠k ]),
where ⌫` , ⇠k 2 (Rn )⇤ = HomR (Rn , R) and [⌫` ], [⇠k ] denote the homogeneous
coordinates of the dual projective space. Since ij , j = 1, 2 are embed-
dings, ⌫1 , . . . , ⌫m (respectively, ⇠1 , . . . , ⇠l ) are mutually distinct. Here, a is
a regular value of ij (j = 1, 2), so that ⌫` (respectively, ⇠k ) is considered
to be the tangent hyperplane of one of the components of the hypersurface
⇡ i1 (V1 ) (respectively, ⇡ i2 (V2 )) through a. Since ⇡ i1 (V1 ) = ⇡ i2 (V2 ),
we may conclude that m = l and i1 (p` ) = i2 (q` ) (` = 1, . . . , m). We set
Wj = (⇡ ij |Vj ) 1 (R), j = 1, 2. It follows that i1 (W1 ) = i2 (W2 ). By the
continuity of ij , j = 1, 2, we have i1 (W1 ) = i2 (W2 ). Therefore, it is enough
to show that Wj is dense in Vj . Suppose that (⇡ ij |Vj ) 1 (Z) has an in-
terior point. Since the set of regular points of ⇡ ij is dense in Vj , there
exists an open subset O` ⇢ Vj such that ⇡ ij (Oj ) ⇢ Z and ⇡ ij |Oj is
an immersion. For a point qj 2 Oj , let Tj be the tangent hyperplane of
the regular hypersurface ⇡ ij (Oj ) at qj . It follows that we have a local
di↵eomorphism
j : T` ! ⇡ Lj (Oj )
around qj . Since Zj is the critical value set of ⇡ ij , j (Zj ) is the critical
value set of j ⇡ ij , j = 1, 2. By Sard’s theorem, j (Z) is a measure
zero set. However, we have j ⇡ ij (Oj ) ⇢ j (Z). This is a contradiction,
therefore (⇡ ij |Vj ) 1 (Z) does not have interior points. Since
1 1
V j = (⇡ ij|Vj ) (S) = (⇡ ij |Vj ) (R [ Z)
1 1
= (⇡ ij |Vj ) (R) [ (⇡ ij |Vj ) (Z),
we get that Wj = (⇡ ij |Vj ) 1
(R) is dense in Vj . ⇤
We remark that in the original proof in [Zakalyukin (1984)] it is assumed
that ⇡ ij |U are proper mappings for j = 1, 2. This assumption is not
needed. The idea of the above proof for removing this assumption is given
in [Kokubu, Rossman, Saji, Umehara, and Yamada (2005)].
The assumption on the density of the regular sets of ⇡ i in Theorem
5.10 is a generic condition on i. If i is Legendrian stable, then there exist a
neighbourhood W of a representative i : V ! P T ⇤ Rn of i (in the Whitney
C 1 -topology) such that for any Legendrian immersion j 2 W , there exists
a point x0 2 V such that the germ j : (V, x0 ) ! P T ⇤ Rn is Legendrian
equivalent to i. Since the above property is generic, there exists a Legen-
drian immersion j : V ! P T ⇤ Rn 2 W and x0 2 V such that the germ
112 Di↵erential Geometry from a Singularity Theory Viewpoint
Proposition 5.2. Let F, G : (Rk ⇥ Rn , (0, 0)) ! (R, 0) be two Morse fami-
lies of hypersurfaces. Suppose that ⇡ L (F ) and ⇡ L (G) are Legendrian
stable. Then the following statements are equivalent.
The families F(q, x, y) and F (q, x) y define the same germ of a graph-like
Legendrian submanifold. Therefore F (q, x) y is a graph-like generating
family of L (F)(⌃F ). In this case,
⌃F = {(q, x, F (q, x)) 2 (Rk ⇥ (Rn ⇥ R), 0) | (q, x) 2 CF }
and L (F) : (⌃F , 0) ! J 1 (Rn , R) is given by
L (F)(q, x, F (q, x)) = (L(F )(q, x), F (q, x)) 2 J 1 (Rn , R) ⌘ T ⇤ Rn ⇥ R.
We summarise here the results in the previous sections which will be used
to describe generic properties of submanifolds in Euclidean and Minkowski
spaces. We start with the following observation.
Proposition 5.6. If F : (Rk ⇥ Rn , 0) ! (R, (0, 0)) is an R+ -versal defor-
mation of the germ f = F |Rk ⇥{0} , then F is a Morse family of functions.
Proof. By Thom’s splitting lemma (Lemma 3.1), we may assume that
f 2 M3k . By Theorem 3.12, we have
n o
LRe · f + R · 1, Ḟ1 , . . . , Ḟn = Ek ,
@F
where Ḟi (q) = @q i
(q, 0). By Theorem 3.1, f is R+ -finitely determined so
that there exists r 2 N such that Mrk ⇢ LRe · f. Then we have
n o
R · j r 1 Ḟ1 , . . . , j r 1 Ḟn = Mk /(LRe · f + Mrk ).
118 Di↵erential Geometry from a Singularity Theory Viewpoint
and
Fe : (Rk ⇥ Rn ⇥ R, 0) ! (R, 0)
(q, x, y) 7! F (q, x) y.
q1 y
CF WF
~
.(Id x F)
x1 x1
XF 1'
BF x1
Fig. 5.1 Catastrophe set, wavefront and bifurcation set of a 1-parameter versal unfold-
ing of the fold singularity.
y
q1 WF
CF
~
.(Id x F)
x2 x2
XF 1'
x1
x1
BF
x2
x1
Fig. 5.2 Catastrophe set, wavefront and bifurcation set of a 2-parameter versal unfold-
ing of the fold singularity.
q1 y
WF
CF
~
.(Id x F)
x2 x2
XF 1' x1
x1
BF x2
x1
Fig. 5.3 Catastrophe set, wavefront and bifurcation set of a 2-parameter versal unfold-
ing of the cusp singularity.
Proof. ) 2 Sn
Let v = (v1 , . . . , vnq 1
and assume, without loss of general-
ity, that vn > 0. Then vn = 1 v12 ··· vn2 1 and
q
H(u, v) = x1 (u)v1 + . . . + xn 1 (u)vn 1 + xn (u) 1 v12 ··· vn2 1.
with
0 v1 vn 1 1
(x1 )u1 (u) (xn )u1 (u) ... (xn 1 )u1 (u) (xn )u1 (u)
B vn vnC
B .. .. .. C
A=B . . . C
@ v1 vn 1 A
(x1 )un 1 (u) (xn )un 1 (u) . . . (xn 1 ) un 1
(u) (xn )un 1 (u)
vn vn
We show that the rank of the matrix A is n 1 at (u, v) 2 C(H). We
set
0 1
(xi )u1 (u)
B .. C
ai = @ . A for i = 0, . . . , n
(xi )un 1 (u)
CH = {(u, ±N (u)) |u 2 U }.
Lagrangian and Legendrian Singularities 123
Therefore,
v1 vn
detA = ( 1)n+1 det(a2 , . . . , an ) + · · · + ( 1)2n det(a1 , . . . , an 1)
vn vn
⌧✓ ◆
v1 vn
= ( 1)n 1
,..., , x u1 ⇥ · · · ⇥ x un 1
vn vn
( 1)n 1
= h±N, xu1 ⇥ · · · ⇥ xun 1 i
vn
( 1)n 1
=± kxu1 ⇥ · · · ⇥ xun 1 k
vn
and this is not zero for any (u, v) 2 CH . This proves that H is a Morse
family of functions. ⇤
Ui+ = {v = (v1 , . . . , vn ) 2 S n 1
| vi > 0 }, i = 1, . . . , n
and
Ui = {v = (v1 , . . . , vn ) 2 S n 1
| vi < 0 }, i = 1, . . . , n
of the (n 1)-sphere S n 1
. Since T ⇤ S n 1
|Ui± is a trivial bundle, we define
the maps
L± ⇤ n
i (H) : CH ! T S
1
|Ui± , i = 1, . . . , n
by
L±
i✓(H)(u, ±N (u)) ◆
= ±N (u), x1 (u) xi (u) vv1i , . . . , xi (u)\xi (u) vvii , . . . , xn (u) xi (u) vvni ,
q P
Proof. Since we have vi = ± 1 j6=i vj2 on Ui± ,
@H vj
(u, v) = xj (u) xi (u) , (j 6= i),
@vj vi
at (u, v) 2 U ⇥ Ui± . By the construction of the germ of the Lagrangian im-
mersion from the generating family in Section 5.1.2, L±
i (H) is a Lagrangian
immersion. ⇤
Remark 5.2. 1. The subset CH = {(u, ±N (u)) |u 2 U } is a double cover
of the hypersurface M . It is not difficult to see that CH is the catastrophe
set of H and the corresponding catastrophe map can be identified with
the normal Gauss map (up to a sign) on the hypersurface M . Thus, for
a generic embedding x, the Gauss map of M = x(U ) behaves as a stable
Lagrangian map.
2. We can apply the same construction to a submanifold M of codi-
mension higher than one. In this case, the catastrophe set CH is the unit
normal bundle and it can be identified with a canal hypersurface around
M in Rn . The catastrophe map coincides with the generalised Gauss map
on CH .
e on M is a
Proposition 5.8. The extended family of height functions H
graph-like Morse family of hypersurfaces.
Proof. Since the height function H is a Morse family of fucntions at each
e
point, the extended height function H(u, v, r) = H(u, v) r is a graph-like
Morse family of hypersurface at each point in U ⇥ (S n 1 ⇥ R) (see §5.3 and
the discussion before Proposition 5.5). ⇤
We can define now a graph-like Legendrian immersion whose generating
family at each point in the domain is the germ of the extended family of
e
height functions H.
We consider the contact manifold Rn ⇥ S n 1 whose contact structure is
given by the 1-form ✓ = hv, dxi|{kvk=1} , where (x, v) 2 Rn ⇥ S n 1 . Since
the tangent bundle T Rn is a trivial bundle, the above contact manifold
is identified canonically with the unit tangent bundle T1 Rn . The contact
structure on T1 Rn is given by the above 1-form ([Blair (1976)]).
We consider the projection : Rn ⇥ S n 1 ! S n 1 ⇥ R definded by
(x, v) = (v, hx, vi).
Lagrangian and Legendrian Singularities 125
It follows that
n
X n ✓
X ◆
⇤ vj
(dy pj dvj ) = dhx, vi|U + xj x1 dvj
j=2
1
j=1
v1
= hdx, vi|U + + hx, dvi|U + hx, dvi|U +
1 1 1
Let : (T ⇤ S n 1
|U + ) ⇥ R ! Rn ⇥ U1+ be defined by
1
n
X n
X
(v, p, y) = (v1 (y pj vj ), . . . , vn (y pj vj ), v).
j=2 j=2
LH,i : CH |U ⇥U ± ! T ⇤ S n 1
⇥ R|U ± ⇥R ⌘ J 1 (Ui± , R)
i i
Proposition 5.10. The germ defined by the family of the distance squared
functions D on M at each point (u0 , a0 ) 2 U ⇥ Rn is a Morse family of
functions.
Proof. We have
n
X
D(u, a) = (xi (u) ai ) 2 ,
i=1
where x(u) = (x1 (u), . . . , xn (u)), u = (u1 , . . . , un 1 ) 2 U and a =
(a1 . . . , an ) 2 Rn . We shall prove that the mapping
✓ ◆
@D @D
D(u, a) = ,..., (u, a)
@u1 @un 1
is not singular at any point (u, a). Its Jacobian matrix is given by
0 1
A11 · · · A1(n 1) 2(x1 )u1 (u) · · · 2(xn )u1 (u)
B .. .. .. .. .. .. C
@ . . . . . . A,
A(n 1)1 · · · A(n 1)(n 1) 2(x1 )un 1 (u) · · · 2(xn )un 1 (u)
where Aij = 2(hxui uj (u), x(u) ai + hxui (u), xuj (u)i). Since x : U ! Rn
is an embedding, the rank of the matrix
0 1
2(x1 )u1 (u) · · · 2(xn )u1 (u)
B .. .. .. C
x=@ . . . A
2(x1 )un 1 (u) · · · 2(xn )un 1 (u)
is n 1 at each point u0 in U. Therefore, the rank of the Jacobian matrix
of D is n 1, and this proves that the germ of D at (u0 , a0 ) is a Morse
family of functions. ⇤
As a consequence of Proposition 5.10, we can define a germ of a La-
grangian immersion whose generating family is the family distance squared
functions on the hypersurface patch M (Theorem 5.3).
Consider the set,
CD = {(u, x(u) + N (u)) | 2 R, u 2 U }
and define the smooth mapping L(D) : CD ! T ⇤ Rn by
L(D)(u, x(u) + N (u)) = (x(u) + N (u), 2 N (u)) .
Corollary 5.2. The map L(D) is a Lagrangian immersion and the family
of distance squared functions D is its generating family at each point in
U ⇥ Rn . The caustic of the Lagrangian map ⇡ L(D) is precisely the focal
set (or evolute) of M . (As a consequence, the focal set of a hypersurface in
Rn has Lagrangian singularities.)
128 Di↵erential Geometry from a Singularity Theory Viewpoint
Proof. By definition,
@D
(u, a) = 2(xi (u) ai ),
@ai
so that
✓ ◆
@D @D
(u, x(u) + N (u)), . . . , (u, x(u) + N (u)) = 2 N (u).
@a1 @an
By the construction of the Lagrangian immersion from the generating fam-
ily in §5.1, we have a Lagrangian immersion
L(D)(u, x(u) + N (u)) = (x(u) + N (u), 2 N (u)) .
Therefore the Lagrangian map is given by
⇡ L(D)(u, x(u) + N (u)) = x(u) + N (u). ⇤
Proof. Statements (4) and (5) are equivalent by Theorem 4.1. The
equivalence of the other statements follow from Proposition 5.2 and
Theorem 5.11. ⇤
130 Di↵erential Geometry from a Singularity Theory Viewpoint
Proof. By Proposition 5.4 and Theorem 5.14, the singular set of the
cylindrical pedal coincides with the singular set of the Gauss map, which
is the parabolic set. Thus, the corresponding Legendrian lifts LMi sat-
isfy the hypothesis of Theorem 5.10. If the germs of the cylindrical
pedals (M1⇤ , (v1 , r1 )) and M2⇤ , (v2 , r2 )) are di↵eomorphic, then LM1 and
LM2 are Legendrian equivalent, so H e1, H
e 2 are P -K-equivalent. There-
fore, e
h1,(v1 ,r1 ) , e
h1,(v2 ,r2 ) are K-equivalent, where ri = hxi (ui ), Ni (ui )i.
By Theorem 4.1, this condition is equivalent to K(M1 , T (M1 )p1 , p1 ) =
K(M2 , T (M2 )p2 , p2 ).
On the other hand, we have
i(x 1 (T (Mi )p ), ui ) = (e
i
h 1 (0), ui ).
i,(vi ,ri )
Since the K-equivalence preserves the zero level sets, the germs
(x1 1 (T (M1 )p1 ), u1 ), (x2 1 (T (M2 )p2 ), u2 ) are di↵eomorphic. ⇤
We call (x 1 (T (M )p0 ), u0 ) the germ of tangent indicatrix of M at u0
(or, p0 = x(u0 )), which is denoted by T I(M, p0 ). By Proposition 5.11, the
di↵eomorphism type of the germ of the tangent indicatrix is an invariant
of the di↵eomorphism type of the germ of the cylindrical pedal CPeM (U )
of M.
We can make use of some basic invariants of germs of functions. The
local ring of a germ of a function is a complete K-invariant for generic
germs, but it is not a numerical invariant. The Ke -codimension of the
germ is a numerical K-invariant ([Martinet (1982)]). We denote that
T-ord(x(U ), u0 ) = code (e
h(v0 ,r0 ) , K). By definition,
Cu10 (U )
T-ord(x(U ), u0 ) = dim ,
hhx(u), N (u0 )i r0 , hxui (u), N (u0 )iiCu10
where r0 = hx(u0 ), N (u0 )i. Usually T-ord(x(U ), u0 ) is called the Ke -
codimension of e
h(v0 ,r0 ) . We call it the order of contact of M with its tangent
hyperplane at p0 = x(u0 ).
We denote the corank of x at u0 by
T-corank(x(U ), u0 ) = (n 1) rank H(hv0 )(u0 ),
Lagrangian and Legendrian Singularities 131
where v0 = N (u0 ).
By Proposition 2.5, p0 is a parabolic point if and only if
T-corank(x(U ), u0 ) 1. Moreover p0 is a flat point if and only if
T-corank(x(U ), u0 ) = n 1.
By Thom’s splitting lemma (Lemma 3.1), if T-corank(x(U ), u0 ) = 1,
then generically the height function hv0 has the Ak -type singularity at u0 .
In this case we have T-ord(x(U ), u0 ) = k. For curves in the plane (i.e.,
n = 2), this number is equal to the order of contact of the curve with the
tangent line in the classical sense (see §4.1 in Chapter 4 and [Bruce and
Giblin (1992)]). This is the reason why we call T-ord(x(U ), u0 ) the order
of contact of M with its tangent hyperplane at p0 .
We now consider the contact of hypersurfaces with families of hyper-
planes. Let xi : (U, ui ) ! (Rn , pi ), i = 1, 2, be germs of parametrisations of
hypersurfaces Mi . We consider height functions Hi : (U ⇥ S n 1 , (ui , vi )) !
R on Mi , where vi = N (ui ). We denote that hi,vi (u) = Hi (u, vi ), then we
have hi,vi (u) = hvi xi (u). We also consider that the germ of the foliation
Fhvi defined in §4.1 for each i = 1, 2. We call Fhvi a tangent family of affine
hyperplanes of Mi at pi wich is denoted by T FH(Mi , pi ).
Moreover, if h1.v1 and h2,v2 are R+ -equivalent then the level set germs
of function germs h1.v1 and h2,v2 are di↵eomorphic. For a hypersurface
germ x : (U, u0 ) ! (Rn .p0 ), the set germ (x 1 (Fhv0 ), u0 ) is a singular
foliation germ at u0 2 U, where v0 = N (p0 ). We call it the tangential
Dupin foliation of M = x(U) at u0 , which is denoted by DF(x(U ), u0 ).
We have the following corollary.
Proof. Statements (1) and (2) are equivalent by Proposition 4.1. and
statements (2)–(5) are equivalent by Theorem 5.6. ⇤
Corollary 5.4. Under the hypotheses of Theorem 5.17 and if one of the
statement there is satisfied, then
Proof. Statement (1) follows from the fact that Lagrangian equivalences
among germs of Lagrangian immersions preserves their caustics. By def-
inition, the germ of the spherical Dupin foliation S-DF(xi (U ), ui ) is the
level-set foliation Fdi,xi of di,xi . Statement (2) follows from the fact that
R+ -equivalence sends the level sets of one germ to another. ⇤
134 Di↵erential Geometry from a Singularity Theory Viewpoint
i hxul , xui i = 0.
i=1
The matrix (hxul , xui i) is that of the first fundamental form of M , so
is not singular. Therefore, i = 0, i = 1, . . . s.
(ii) This follows from the fact that @ e
hv0 /@ui (u) = @hv0 /@ul (u). ⇤
Definition 5.7. Let M as above and let X ⇤ : U ⇥ S k 1
! Sn 1
⇥ R be
defined by
X ⇤ (u, µ) = (N (u, µ), hx(u), N (u, µ)i).
We call the map X ⇤ : the canal cylindrical pedal of M.
Lagrangian and Legendrian Singularities 135
Proof. We remark that the germs of the extended families of height func-
tions He i are the graph-like generating families of the Legendrian subman-
ifold covering (Xi⇤ (U ⇥ S k 1 ), (ui , µi )). By the relation b " y⇤i = Xi⇤ in
the proof of Proposition 5.15, such a Legendrian submanifold is Legendrian
equivalent to LCMi " .
Statements (1)–(4) are equivalent by Theorem 5.11 and Proposition
5.2. It also follows from the relation b " (CM (")⇤i = y⇤i (U ⇥ S k 1 ) that the
statements (1) and (6) are equivalent.
Lagrangian and Legendrian Singularities 137
Statements (6) and (7) are equivalent by Theorem 5.15 for hypersur-
faces.
Statements (4) and (5) are equivalent by Theorem 4.4 and the fact that
1
hvi (ri ) = T (Mi ; vi )pi . ⇤
May 2, 2013 14:6 BC: 8831 - Probability and Statistical Theory PST˙ws
139
140 Di↵erential Geometry from a Singularity Theory Viewpoint
Fig. 6.1 Visible parabolic curves (in thick) on the surface of a bell.
where
l = h Nu1 , xu1 i = hN, xu1 u1 i,
m = h Nu1 , xu2 i = hN, xu1 u2 i,
n = h Nu2 , xu2 i = hN, xu2 u2 i
are the coefficients of the second fundamental form of M . Following the
notation in Chapter 2, we have
l = h11 , m = h12 = h21 , n = h22 .
At each point on M , there are two principal curvatures 1 and 2 which
are the eigenvalues of Wp . The Gaussian curvature K and the mean cur-
vature H of M are given by
1
K = 1 2 , H = (1 + 2 ).
2
These can be expressed in terms of the coefficients of the first and second
fundamental forms as follows (see for example [do Carmo (1976)] and
Corollary 2.1)
ln m2
K=
EG F 2
and
lG 2mF + nE
H= .
2(EG F 2 )
The principal curvatures 1 and 2 are the solutions of the quadratic
equation
2 2H + K = 0,
p p
so, one is equal to H + H 2 K and the other to H H 2 K.
Recall from Chapter 1 that the Gaussian curvature K is an intrinsic
invariant of the surface M . However, the principal curvatures 1 and 2
and the mean curvature H are not intrinsic invariants of M . They provide
information about the geometry of M as a surface in the ambient space R3 ,
so they capture some extrinsic properties of M .
If p is not an umbilic point, that is, if 1 6= 2 , there are two orthogonal
principal directions at p denoted by v1 and v2 which are parallel to the
eigenvectors of Wp . At umbilic points , every tangent direction is consid-
ered a principal direction. A curve on M whose tangent direction at all
points is a principal direction is called a line of principal curvature. The
lines of principal curvature are the solution curves of the binary di↵erential
equation
(F n mG)du22 + (En lG)du1 du2 + (Em lF )du21 = 0, (6.1)
142 Di↵erential Geometry from a Singularity Theory Viewpoint
Fig. 6.2 Generic configurations of the lines of principal curvature at umbilic points:
Lemon (left), Star (centre), Monstar (right).
see [do Carmo (1976)] for a proof. Equation (6.1) can also be written in
the following determinant form
du22 du1 du2 du21
E F G = 0.
l m n
The lines of principal curvature form a pair of orthogonal foliations
Their generic topological configurations at umbilic points are determined in
[Sotomayor and Gutierrez (1982); Bruce and Fidal (1989)]; see Figure 6.2
and [Sotomayor (2004)] for historical notes. Porteous gave the colour blue
to one foliation of the lines of curvature and the colour red to the other.
In Figure 6.2, one foliation associated to the curvature i is drawn in solid
lines and the other associated to the curvature j , j 6= i, is drawn in dashed
lines.
The second fundamental form measures how the surface M bends in R3 .
Let ↵ : ( ✏, ✏) ! M be a smooth curve on M with ↵(0) = p. Suppose that
↵ is parametrised by arc length and denote by s the arc length parameter.
The acceleration vector ↵00 is orthogonal to the tangent vector t = ↵0 , so it
lies in the plane with orthonormal basis {t ⇥ N, N }, where N (s) = N (↵(s))
is the normal vector to M at ↵(s). We can write
↵00 = g t ⇥ N + n N,
for some scalars g (called the geodesic curvature) and n (called the normal
curvature). The component g t ⇥ N of ↵00 at s lies in the tangent plane to
M at ↵(s) and the component n N is parallel to the normal of the surface
at ↵(s).
If the geodesic curvature g = h↵00 , t ⇥ N i of the curve ↵ vanishes at
some point s, we say that the curve ↵ has a geodesic inflection at s. In
general, at a geodesic inflection, the curve ↵ is locally on both sides of
the plane at ↵(s) parallel to N (s) and ↵0 (s) (so the projection of ↵ to the
tangent plane T↵(s) M is a plane curve with an inflection).
Surfaces in the Euclidean 3-space 143
Proposition 6.1. All curves on a surface M with the same tangent line
at a point p on M have the same normal curvature at p.
If p is not an umbilic point, we can write w = cos ✓v1 + sin ✓v2 for any
unit tangent vector in Tp M , where v1 and v2 are the orthonormal principal
directions at p. Then
n (✓) = IIp (w) = hWp (cos ✓v1 + sin ✓v2 ), cos ✓v1 + sin ✓v2 i
= hcos ✓1 v1 + sin ✓2 v2 , cos ✓v1 + sin ✓v2 i
= 1 cos2 ✓ + 2 sin2 ✓.
Therefore, the principal curvatures 1 and 2 are the extrema of the normal
curvature n (✓) at p when ✓ varies in [0, ⇡].
A direction along which the normal curvature is zero is called an asymp-
totic direction. Thus, a direction w 2 Tp M is asymptotic if and only if
IIp (w) = 0. We observe that the normal section along an asymptotic di-
rection has an inflection and that there are two asymptotic directions at a
hyperbolic point, one at a parabolic point and none at an elliptic point (see
§2.1.4).
A curve on M whose tangent direction at each point is an asymptotic
direction is called an asymptotic curve. The asymptotic curves are the
solutions of the binary di↵erential equation
ndu22 + 2mdu1 du2 + ldu21 = 0. (6.3)
(ii) We have
IIp (v̄) = (am + bn)2 l 2(am + bn)(al + bm)m + (al + bm)2 n
= (nl m2 )(a2 l + 2abm + b2 n)
= (nl m2 )IIp (v).
(iii) It follows from (ii) that
1
n (v̄) = IIp (v̄)
||v̄||2
1
= (nl m2 )IIp (v)
||v̄||2
||v||2
= (nl m2 )n (v).
||v̄||2
(iv) We have
||v||2 ||v̄||2 hv̄, vi2 =
(a2 E + 2abF + b2 G)((am + bn)2 E 2(am + bn)(al + bm)F + (al + bm)2 G)
2
(a(am + bn)E + (a(al + bm) b(am + bn))F + b(al + bm)G)
= (EG F 2 )(a2 l + 2abm + b2 n)2
= (EG F 2 )IIp (v)2 . ⇤
the graph of some function z = f (x, y), with (x, y) in a open subset U of
R2 containing the origin. Then we have the Monge form parametrisation
(x, y) = (x, y, f (x, y)), (x, y) 2 U , of M at p. We shall say that M is
parametrised locally in Monge form z = f (x, y) at the origin p. Note that
the Taylor expansion of f at the origin has no constant or linear terms.
Proof. The result follows from Theorems 4.3, 4.5 and 4.7. ⇤
As (EG F 2 )(u) > 0, det H(hv )(u) and K(u) have the same sign, and
the result follows. ⇤
When the surface is taken in Monge form z = f (x, y), the conditions
for the height function hN0 (x, y) = f (x, y) along the normal direction N0 =
(0, 0, 1) to have a given singularity at the origin p and for the family of
height functions to be a R+ -versal unfolding of these singularities can be
expressed in terms of the coefficients of the Taylor expansion of f at p. We
write
f (x, y) = a20 x2 + a21 xy + a22 y 2 + a30 x3 + a31 x2 y + a32 xy 2 + a33 y 3 + O(4).
(6.4)
We can rotate the coordinate axes in the tangent plane Tp M and set a
chosen direction to be along the y-axis. The chosen direction could be, for
example, an asymptotic direction if p is not an elliptic point or a principal
direction.
Proof. For f as in (6.4), the 1-jets at the origin of the coefficients of the
second fundamental form for the Monge form setting in Proposition 6.3 are
given by
j 1 l = 2a20 + 6a30 x + 2a31 y,
j 1 m = a21 + 2a31 x + 2a32 y,
j 1 n = 2a22 + 2a32 x + 6a33 y.
It follows that
j 1 (ln m2 )(x, y) = 4a20 a22 a221 + 4(a20 a32 + 3a30 a22 a21 a31 )x+
4(3a20 a33 + a31 a22 a21 a32 )y.
(i) In particular, the origin is a parabolic point if and only if
(ln m2 )(0, 0) = 4a20 a22 a221 = 0.
(ii) At the origin p, l = 2a20 , m = a21 and n = 2a22 so a direction
a x (0, 0) + b y (0, 0) = (a, b, 0) is asymptotic at p if and only if
a20 a2 + a21 ab + a22 b2 = 0
Surfaces in the Euclidean 3-space 149
Theorem 6.2. The following hold for M in Monge form z = f (x, y) at the
origin p with f as in (6.4) and v = (0, 1, 0) 2 Tp M . Assume that p is not
a flat umbilic point.
(i) For p a parabolic point and v an asymptotic direction at p, the height
function HN0 has a singularity at p of type
A2 () 6 0,
a33 =
A3 () a33 = 0 and a232 4a20 a44 6= 0.
Proof. (i) The height function along the normal direction N0 = (0, 0, 1)
at the origin p is given by hN0 (x, y) = f (x, y). By Proposition 6.5, if p is a
parabolic point and v is an asymptotic direction at p, then a22 = a21 = 0
so that
f (x, y) = a20 x2 + a30 x3 + a31 x2 y + a32 xy 2 + a33 y 3 + ⌃4i=0 a4i x4 i y i + O(5).
(We chose the unique asymptotic direction at p to be along v in order
to have j 2 f = a20 x2 . This makes the task of recognizing the singularities
of hN0 much easier.) The hypothesis on p not being a flat umbilic point is
then equivalent to a20 6= 0. The height function f has an A 2 -singularity
at the origin if and only if a33 = 0. Then the singularity is of type A3
if and only if a20 x2 + a32 xy 2 + a44 y 4 is not a perfect square, equivalently,
a232 4a20 a44 6= 0.
(ii) Observe that the family of height functions H : R2 ⇥S 2 , (0, N0 ) ! R
is P -R+ -equivalent to the modified family, that we still denote by H :
R2 ⇥ R2 , (0, 0) ! R, which is given by
H(x, y, ↵, ) = ↵x + y + f (x, y)
150 Di↵erential Geometry from a Singularity Theory Viewpoint
(all we did here is to parametrise the sphere S 2 near (0, 0, 1) by (↵, , 1)).
Recall from Theorem 3.12 that H is an R+ -versal unfolding if and only
if
n o
LRe · f + R. 1, Ḣ1 , Ḣ2 = E2 . (6.5)
We have
Ḣ1 (x, y) = H↵ (x, y, 0, 0) = x,
Ḣ2 (x, y) = H (x, y, 0, 0) = y.
For f as in (6.4),
j 3 fx = 2a20 x + a21 y + 3a30 x2 + 2a31 xy + a32 y 2
+4a40 x3 + 3a41 x2 y + 2a42 xy 2 + a43 y 3
j 3 fy = a21 x + 2a22 y + a31 x2 + 2a32 xy + 3a33 y 2
+a41 x3 + 2a42 x2 y + 3a43 xy 2 + 4a44 y 3 .
The A1 -singularity is 2-R-determined, that is M32 ⇢ LR · f , so for H
to be an R+ -versal unfolding it is enough to show that equality (6.5) holds
modulo M32 , that is
n o
j (LRe · f + R. Ḣ1 , Ḣ2 + h1iR ) = J 2 (2, 1).
2
(6.6)
Fig. 6.3 Intersection of a surface with its tangent plane, top figures at an elliptic,
hyperbolic and parabolic point respectively, and bottom figures at an A+ 3 and A3 -
singularities of the height function.
Proof. If we take the surface in Monge form z = f (x, y), then the height
function HN0 along the normal N0 = (0, 0, 1) is just the function f . The
intersection of the surface with its tangent plane z = 0 is the zero set of the
function f , which is di↵eomorphic to the zero set of the R-normal form of
the singularity of hN0 = f . The zero sets of the normal forms (taken up to
a sign ±) are as follows
Surfaces in the Euclidean 3-space 153
!
Proposition 6.6 can be viewed as a local classification, up to diffeo-
morphisms, of the tangent Dupin indicatrices of a height function generic
surface. We can also consider the tangential Dupin foliation of M at p,
which is obtained by intersecting the surface M with planes parallel to
Tp M .
A+
1 A−
1 A2 A+
3 A−
3
Fig. 6.5 The dual surface of M at the Ak -singularities of the height function: A1 left,
A2 centre and A3 right.
Corollary 6.1. For a height function generic surface M, the Gauss map
at p 2 M is A-equivalent to:
It follows from Proposition 6.9 that the image of the parabolic set by
the Gauss map is a curve with a cusp singularity at an A3 -singularity of
the height function on M . For this reason, we have the following definition.
Definition 6.3. A point on the parabolic set where the height function
along the normal direction to the surface has an A3 -singularity is called a
cusp of Gauss. A cusp of Gauss p is elliptic if the height function along the
normal direction at p has an A+3 -singularity at p and hyperbolic if it has
an A3 -singularity.
"
!!
(M, p) ⇢ R3
6
x Pv
✏
(R2 , 0) / (Tv S 2 , Pv (p)) ' (R2 , 0)
Fig. 6.7 An apparent contour of a torus: the dashed parts represent the invisible part
of the contour generator and apparent contour when the surface is not transparent.
h k
! k◦P ◦h−1 !
v
(R2 , 0) " (R2 , 0)
It is thus appropriate to use the group A for studying the singularities
of the germs of orthogonal projections of M .
[v, xu1 u1 (u), xu2 (u)] + [v, xu1 (u), xu1 u2 (u)] = 0, (6.8)
[v, xu1 u2 (u), xu2 (u)] + [v, xu1 (u), xu2 u2 (u)] = 0. (6.9)
We drop the argument u and write v = axu1 + bxu2 . The vectors xui uj
can be written, with respect to the basis {xu1 , xu2 , N } of R3 , in the form
1 2
x u1 u1 = 11 xu1 + 11 xu2 + lN
1 2
x u1 u2 = 12 xu1 + 12 xu2 + mN
1 2
x u2 u2 = 22 xu1 + 22 xu2 + nN
where kij are the Christo↵el symbols (see for example [do Carmo (1976)]).
Then equation (6.8) becomes
[axu1 + bxu2 , 111 xu1 + 211 xu2 + lN, xu2 ]
+[axu1 + bxu2 , xu1 , 112 xu1 + 212 xu2 + mN ] = 0,
which is equivalent to
al + bm = 0
as [xu1 , xu2 , N ] 6= 0. Similarly, equation (6.9) is equivalent to am + bn = 0.
Therefore, the contour generator is singular if and only if
⇢
al + bm = 0
am + bn = 0
equivalently, nl m2 = 0 and v = nxu1 + mxu2 at u = (u1 , u2 ). These
are precisely the conditions for p to be a parabolic point and for v to be
the unique asymptotic direction at p.
(ii) The contour generator ⌃v must be a smooth curve at p otherwise
v is singular at Pv (p). We take, without loss of generality, am + bn 6= 0
at p so that ⌃v can be parametrised by u1 7! (u1 , u2 (u1 )) for some smooth
function u2 (u1 ) with
al + bm
u02 = . (6.10)
am + bn
The apparent contour v is then parametrised by
Pv (u1 ) = x(u1 , u2 (u1 )) hx(u1 , u2 (u1 )), viv.
162 Di↵erential Geometry from a Singularity Theory Viewpoint
We have
Pv0 = xu1 + u02 xu2 hxu1 + u02 xu2 , viv
= xu1 + u02 xu2 hxu1 + u02 xu2 , axu1 + bxu2 i(axu1 + bxu2 )
= (1 a(aE + (u02 a + b)F + u02 bG)) xu1
+ (u02 b(aE + (u02 a + b)F + u02 bG)) xu2
Therefore, v is singular if and only if
1 a(aE + (u02 a + b)F + u02 bG) = 0,
u02 b(aE + (u02 a + b)F + u02 bG) = 0.
The above two equations can be written, after substituting u02 by its
expression in (6.10), in the form
am + bn = a a2 (mE lF ) + ab(nE lG) + b2 (nF mG) , (6.11)
al + bm = b a2 (mE lF ) + ab(nE lG) + b2 (nF mG) . (6.12)
Observe that a2 (mE lF ) + ab(nE lG) + b2 (nF mG) = 0 means
that v is a solution of equation (6.1), that is, v is a principal direction. If
this is the case, the system of equations (6.11) and (6.12) becomes
⇢
al + bm = 0
am + bn = 0
equivalently, p is a parabolic point and v the unique asymptotic direction
at p. But this is excluded as it implies that ⌃v , and hence v , is singular.
Suppose then that a2 (mE lF ) + ab(nE lG) + b2 (nF mG) 6= 0.
Dividing side by side equation (6.11) by equation (6.12) yields
am + bn a
= .
al + bm b
Rearranging the above equality gives
a2 l + 2abm + b2 n = 0.
This implies that v is a solution of equation (6.3), that is, v is an asymptotic
direction.
Conversely, let v be a unit asymptotic direction (we project along v so
we require v 2 S 2 ). Then a2 l+2abm+b2 n = 0, so al+bm = (am+bn)b/a.
The right hand side of equation (6.11) can then be rearranged to get
a a2 (mE lF ) + ab(nE lG) + b2 (nF mG)
= a (a2 m + abn)E + (b2 n a2 l)F (abl + mb2 )G
✓ ◆
b2
= a a(am + bn)E + 2b(am + bn)F + (am + bn)G)
a
= (am + bn)(a2 E + 2abF + b2 G)
= am + bn.
Surfaces in the Euclidean 3-space 163
Proof. The curvature of the normal section is just the normal curvature
of the surface M at p along the tangent direction v and is given by
IIp (v)
n (v) = .
||v||2
Let ↵(s) be an arc length parametrisation of the contour generator ⌃v .
Then, by Proposition 6.12,
v̄
↵0 = t =
||v̄||
00
↵ = g t ⇥ N + n (v̄)N.
The apparent contour v is parametrised by (s) = Pv (↵(s)). The
projection Pv for v fixed is a linear map R3 ! Tv S 2 , so
0
= dPv (↵0 ) = ↵0 h↵0 , viv,
00
= dPv (↵00 ) = ↵00 h↵00 , viv.
164 Di↵erential Geometry from a Singularity Theory Viewpoint
The vector N (s) is a unit normal vector to the curve v at (s), and
0
(s), N (s) form a positively oriented frame. If we denote 0? the vector
obtained by rotating 0 anti-clockwise by ⇡/2, then 0? = || 0 ||N , so the
curvature ( v ) of the apparent contour is given by
0?
h , 00 i h|| 0 ||N, 00
i hN, ↵00 i n (v̄)
( v) = = = = .
|| 0 ||3 || 0 ||3 || 0 ||2 || 0 ||2
We have
0 v̄ v̄ v̄ v v 1
= dPv ( )= h , i = 2
(||v||2 v̄ hv̄, viv),
||v̄|| ||v̄|| ||v̄|| ||v|| ||v|| ||v|| ||v̄||
so that
1
|| 0 ||2 = 2 2
(||v||2 ||v̄||2 hv̄, vi2 )
||v|| ||v̄||
1
= (EG F 2 )IIp (v)2 (by Proposition 6.2(iv))
||v||2 ||v̄||2
✓ ◆
||v||2 2 IIp (v)2
= (EG F )
||v̄||2 ||v||4
||v||2
= (EG F 2 )n (v)2 .
||v̄||2
Therefore,
||v̄||2 n (v̄)
( v ) = .
||v||2 (EG F 2 )n (v)2
We obtain by substituting n (v̄) by its expression from Proposition 6.2 that
✓ ◆
||v̄||2 ||v||2 K(p)
( v ) = 2 2 2 2
(nl m2 )n (v) = .
||v|| (EG F )n (v) ||v̄|| n (v)
Consequently,
K(p) = n (v)( v ). ⇤
We have the following consequences of Koenderink’s theorem.
and has 4-point contact at (x, y) with the kernel of dP̃(↵, ) , that is,
@
@y (f (x + ↵y, y) y) = 0,
@2
@y 2 (f (x + ↵y, y) y) = 0,
@3
@y 3 (f (x + ↵y, y) y) = 0,
@4
@y 4 (f (x + ↵y, y) y) 6= 0.
(ii) The above three equations determine the flecnodal set. Evaluating
at ↵ = = 0, we get that the origin p is a flecnodal point if and only if fy =
fyy = fyyy = 0, at (0, 0), equivalently, a22 = a33 = 0, and a21 6= 0, a44 6= 0.
(As a22 = 0, the condition for ⌃↵, to be a smooth curve becomes a21 6= 0.)
We can use equation (6.14) to obtain as a function of (x, y, ↵). Sub-
stituting in equation (6.15), we can solve for ↵ as fxy (0, 0) = a21 6= 0.
We obtain ↵ as a function of (x, y) and substituting in equation (6.16),
we obtain an equation for the flecnodal set. The 1-jet of this equation is
a2
6(a43 a32 21
)x + 24a44 y. ⇤
and consider F1 (x, y + h(x, y)). The 3-jet of F1 (x, y + h(x, y)) is given by
(x, a21 xy + a21 h20 x3 + (a21 h21 + a31 )x2 y + (a21 h22 + a32 )xy 2 )
and equating the coefficients of (0, x3 ), (0, x2 y) and (0, xy 2 ) to zero gives
h20 = 0, h21 = a31 /a21 and h22 = a32 /a21 . This determines completely
the quadratic part h2 (x, y) of h(x, y).
We proceed inductively on the jet level of F1 (x, y +h(x, y)) to determine
completely h, and this gives F1 (x, y + h(x, y)), and hence j 7 Pv (x, y), A(7) -
equivalent to
with
1
↵1 = a21 (a21 a66 5a55 a32 ),
1
1 = a221
(a77 a221 6a66 a32 a21 5a21 a55 a43 + 20a55 a232 ).
5a55 h2 + (a21 a66 5a55 a32 )/a21 . Setting these to be zero and solving the
linear system in and h2 gives
a21 a66 5a55 a32 a21 a66 5a55 a32
= 2 , h2 = .
4a55 a21 4a55 a21
We then equate the coefficients of (0, Xy i ) in j 7 (k F2 H) to zero to get
h3 , . . . , h6 . As a result, j 7 (k F2 H), and hence j 7 Pv , is A(7) -equivalent
to (x, a21 xy + a55 y 5 + ⇤y 7 ) with
(8a55 a77 5a266 )a221 + 2a55 (a32 a66 20a43 a55 )a21 + 35a232 a255
⇤= .
8a221 a55
The singularity of Pv at p is a Butterfly if and only if ⇤ 6= 0.
• Lips/beaks
For the lips/beaks singularity which is 3-A-determined, we consider
j 3 Pv ⇠A(3) (x, a21 xy + a22 y 2 + a31 x2 y + a32 xy 2 + a33 y 3 ).
For Pv to have a lips/beaks singularity it is necessary that a21 = a22 = 0
and a33 6= 0. From Proposition 6.5 and Theorem 6.2, these conditions mean
that v is an asymptotic direction and p is a parabolic point but not a cusp
a32
of Gauss. Then the change of coordinates (x, y) 7! (x, y 3a 33
x) in the
source and a change of coordinate in the target yields
3a31 a33 a232 2
j 3 Pv ⇠A(3) (x, x y + a33 y 3 ).
3a33
This is equivalent to a lips/beaks singularity if and only if 3a31 a33 a232 6= 0.
By Proposition 6.10, 3a31 a33 a232 = 0 if and only if the geodesic curvature
of the image of the parabolic curve by the Gauss map vanishes at N (p).
• Goose
Suppose that a21 = a22 = 3a31 a33 a232 = 0 and a33 6= 0. Thus, p is a
parabolic point, v is an asymptotic direction and the geodesic curvature of
the image of the parabolic curve by the Gauss map is zero at N (p). The
coordinate changes for the lips/beaks give
j 4 Pv ⇠A(4) F3 = (x, a33 y 3 + ↵1 x3 y + ↵2 x2 y 2 + ↵3 xy 3 + ↵4 y 4 )
with
1
↵1 = (27a41 a333 18a42 a32 a233 + 9a43 a232 a33 4a44 a332 )
27a333
and ↵2 , ↵3 , ↵4 are constants depending on the coefficients aij . The terms
(0, ↵2 x2 y 2 +↵3 xy 3 +↵4 y 4 ) in F3 can be eliminated by a change of coordinate
of the form (x, y) 7! (x, y +g(x, y)) with g a polynomial with no constant or
172 Di↵erential Geometry from a Singularity Theory Viewpoint
Table 6.3 The conditions for the family of projections to be a versal unfolding.
Name Algebraic conditions Geometric interpretation
Fold – Always
Cusp – Always
Lips/beaks – Always
Swallowtail – Always
Goose a20 6= 0 p is not a flat umbilic
Gulls a232 4a20 a44 6= 0 The image of the parabolic set by the Gauss map
is a curve with a (2, 3)-cusp at N (p)
Butterfly a232 a21 a43 6= 0 The flecnodal curve is not singular
where
P̃˙ 1 (x, y) = P̃↵ (x, y, 0, 0) = (0, yfx (x, y)),
P̃˙ 2 (x, y) = P̃ (x, y, 0, 0) = (0, y).
The fold and cusp are stable singularities so are versally unfolded
by any family. We consider the remaining singularities in Table 6.1.
As these singularities are finitely k-A-determined for some k, we have
Mk+1
2 .E(2, 2) ⇢ LA · P̃0 , so to prove that P̃ is an Ae -versal unfolding of
the singularity of P̃0 we only need to show that
⇣ n o⌘
j k LAe · P̃0 + R. P̃˙ 1 , P̃˙ 2 = J k (2, 2). (6.17)
174 Di↵erential Geometry from a Singularity Theory Viewpoint
We shall work downwards on jet levels (as in the proof of Theorem 6.2)
and start by showing that all monomials (xi y j , 0) and (0, xi y j ) of degree
k are in the left hand side of (6.17). We have @@x P̃0
(x, y) = (1, fx (x, y)),
thus j k (Q @@x
P̃0
) = (Q(x, y), 0) is in the left hand side of (6.17) for any
monomial Q(x, y) of degree k. Therefore, we only need to consider the
monomials (0, xi y j ) of degree k. Once we got these, we can work modulo
Mk2 .E(2, 2) and consider the monomials of degree k 1. Again we only need
to consider monomials of the form (0, Q(x, y)). We observe that, using the
first component of P̃ , the monomials (xi , 0) and (0, xi ) are in LLe · P̃0 and
hence in the left hand side of (6.17). We also get the monomial (0, y) from
P̃˙ .
2
• Swallowtail
The swallowtail singularity is 4-A-determined, so we need to show that
(6.17) holds for k = 4. We have
j 3 ( @@x
P̃0
) = (1, 2a20 x + a21 y + 3a30 x2 + 2a31 xy + a32 y 2
+4a40 x3 + 3a41 x2 y + 2a42 xy 2 + a43 y 3 )
j 3 ( @@y
P̃0
) = (0, a21 x + a31 x2 + 2a32 xy
+a41 x3 + 2a42 x2 y + 3a43 xy 2 + 4a44 y 3 )
We get all the monomial of degree 4 of the form (0, xQ1 (x, y)) using
j (Q1 @@y
4 P̃0
), where Q1 is of degree 3. From this we also get the degree 3
monomial (0, x2 y) using j 4 (x2 @@y
P̃0
).
Working modulo these monomials, we have
⇠1 = j 3 (y @@y
P̃0
) ⌘ (0, a21 xy + a32 xy 2 )
⇠2 = j 3 (0, f (x, y)) ⌘ (0, a21 xy + a32 xy 2 + a44 y 4 )
so that ⇠2 ⇠1 ⌘ (0, a44 y 4 ). As a44 6= 0, we get (0, y 4 ).
Now j 3 (y 2 @@y
P̃0
) ⌘ (0, a21 xy 2 ) which gives (0, xy 2 ) as a21 6= 0. Similarly,
j 3 (x @@x
P̃0
) ⌘ (0, a21 xy) which gives (0, xy).
We get (0, y 3 ) from j 3 ( @@y
P̃0
) and (0, y 2 ) from P̃˙ 1 . Therefore (6.17) holds
without additional conditions on the coefficients of f . Thus P is always an
Ae -versal unfolding of the swallowtail singularity.
• Lips/beaks
The lips/beaks singularity is 3-A-determined, so we need to show that
(6.17) holds for k = 3. In this case we have
j 2 ( @@x
P̃0
(x, y)) = (1, 2a20 x + 3a30 x2 + 2a31 xy + a32 y 2 )
j 2 ( @@y
P̃0
(x, y)) = (0, a31 x2 + 2a32 xy + 3a33 y 2 ).
Surfaces in the Euclidean 3-space 175
• Gulls
Here we take k = 5 in (6.17) as the gulls singularity is 5-A-determined.
We have
j 4 ( @@x
P̃0
(x, y)) = (1, 2a20 x + 3a30 x2 + 2a31 xy + a32 y 2
+4a40 x3 + 3a41 x2 y + 2a42 xy 2 + a43 y 3
+5a50 x4 + 4a51 x3 y + 3a52 x2 y 2 + 2a43 xy 3 + a54 y 4 )
j 4 ( @@y
P̃0
(x, y)) = (0, a31 x2 + 2a32 xy
+a41 x3 + 2a42 x2 y + 3a43 xy 2 + 4a44 y 3
+a51 x4 + 2a52 x3 y + 3a43 x2 y 2 + 4a54 xy 3 + 5a55 y 4 )
When M is given in Monge form z = f (x, y), the conditions for da0 to
have one of the singularities in Proposition 6.9 and for the family D to be
a versal deformation of these singularities can be expressed in terms of the
coefficients of the Taylor expansion of f at the origin p. If p 2 M is not
an umbilic point, we can take the coordinate axes parallel to the principal
directions of M at p. We choose the x-axis in the direction of the principal
direction v1 and the y-axis in the direction of v2 . We can also rescale the
coordinates in the source if necessary, i.e., make a change of coordinates of
the form (x, y) 7! (↵x, y), so that the coefficients of the first fundamental
form at p are given by
E(0, 0) = 1, F (0, 0) = 0, G(0, 0) = 1.
A principal direction (a, b, 0) at p is a solution of the equation
b2 ab a2
E(0, 0) F (0, 0) G(0, 0) = 0,
l(0, 0) m(0, 0) n(0, 0)
that is,
m(0, 0)a2 + (n(0, 0) l(0, 0))ab + m(0, 0)b2 = 0.
As the principal directions at p are taken to be along (1, 0, 0) and
(0, 1, 0), it follows that m(0, 0) = 0, so fxy (0, 0) = 0. We have
l(0, 0) = fxx (0, 0), n(0, 0) = fyy (0, 0),
so the principal curvatures at the origin p are given by
1 (p) = fxx (0, 0), 2 (p) = fyy (0, 0).
Thus, at the origin p and with the above setting,
f (x, y) = a20 x2 + a22 y 2 + a30 x3 + a31 x2 y + a32 xy 2 + a33 y 3 + O(4) (6.18)
with 1 (p) = 2a20 and 2 (p) = 2a22 .
It follows from Proposition 2.6 that da0 is singular at the origin p if and
only if a0 = (0, 0, a3 ) with a3 some non-zero real number. That is, a is on
the normal line to M at p. The singularity is of type A1 if and only if a is
not on the focal set of M , that is, a3 6= 1/i (p), i = 1, 2.
180 Di↵erential Geometry from a Singularity Theory Viewpoint
A2 : always;
A3 : always;
A4 : 8a31 a420 8a31 a22 a320 4a51 a220 + 4(2a51 a22 a31 a42 )a20
3a231 a33 + 4a31 a42 a22 4a51 a222 6= 0;
D4 : 2a232 a231 + 3a31 a33 6= 0.
Proof. The proof follows the same steps as those of Theorem 6.2. We
consider only the singularities A4 and D4 .
We start with the A4 -singularity. We denote by dij the coefficient of
i j
x y in the Taylor expansion of da0 at the origin. Observe that
1 (p) 2 (p) a22 a20
d02 = = 6= 0
21 (p) 2a22
as p is not an umbilic point. The singularity is more degenerate than
an A3 , so a30 = 0 and a231 4(a22 a20 )(a40 a320 ) = 0. Then L =
d02 y 2 + d21 x2 y + d40 x4 is a perfect square.
We make the change of coordinates (x, Y ) = (x, y d21 /2(d02 )x2 ). Then
L(x, Y ) = d02 Y 2 and the singularity of da0 is of type A4 if and only if the
coefficient of x5 in the Taylor expansion of Da0 (x, Y ) is not zero. That
coefficient is not zero when the expression in the statement of the theorem
is not zero.
Surfaces in the Euclidean 3-space 181
with 0 1
1 0 1
M = @ 3a30 2a31 a32 A .
a31 2a32 3a33
The monomial of degree 2 are in the 3-jet of the left hand side of (6.19)
if and only if the determinant of the matrix M is not zero, that is, 2a232
a231 + 3a31 a33 6= 0. We then use Ḋ1 and Ḋ2 to get x, y in the 3-jet of the
left hand side of (6.19). Therefore, the family D is a versal deformation of
the D4 -singularity of da0 if and only if 2a232 a231 + 3a31 a33 6= 0. ⇤
Remark 6.3. The geometric interpretation of the algebraic conditions in
Theorem 6.10 are given in §6.6.3.
Geometric objects can be derived from the family of distance squared
functions on the surface M , such as the focal set of M and the spheri-
cal Dupin foliation defined below. For a distance squared function generic
surface, the local structure of these objects up to di↵omorphisms is deter-
mined by the R-singularity type of the distance squared function da0 . It
turns out that the di↵eomorphism models of these objects determine the
R-singularity type of da0 . We start with the focal set.
Theorem 6.11. Away from umbilic points, the focal set F1 and F2 of M
are disjoint surfaces, and for a distance squared function generic M they
are di↵eomorphic to
(i) a smooth surface if and only if da0 has an A2 -singularity at p;
(ii) a cuspidal edge surface if and only if da0 has an A3 -singularity at p;
(iii) a swallowtail surface if and only if da0 has an A4 -singularity at p,
where a0 = p + 1/i N (p), i = 1 or 2.
At an umbilic point the focal set F1 [ F2 is di↵eomorphic to a pyramid
(resp. purse) if and only if dc has a D4 (resp. D4+ )-singularity at p.
Proof. ([Arnol’d, Guseı̆n-Zade and Varchenko (1985)]) The focal set of
M is a caustic with D the generating family of the Lagrangian submanifold
L(D)(CD ) ⇢ T ⇤ R3 . Thus, it has Lagrangian singularities. For a distance
squared function generic surface M , the family D is R+ -versal, so the focal
set is di↵eomorphic to the bifurcation set of a model R+ -versal unfolding
(with two parameters) of the singularities that occur in a given distance
squared function. These are A1 , A2 , A3 , A4 and D4 -singularities (Theo-
rem 6.9). The bifurcation set of an R+ -versal 2-parameter family of these
singularities are as stated in the proposition (see also §3.9.2) and are as in
Figure 6.8. ⇤
Surfaces in the Euclidean 3-space 183
!! !$ !# "#% "#"
Fig. 6.8 Models of the focal set of a generic surface in R3 . The fourth and fifth figures
model the focal sets F1 and F2 joining at an umbilic point. The first three figures model
the focal sets F1 or F2 . One can have the following generic combinations for the pairs
(F1 , F2 ) or (F2 , F1 ): (A2 , A2 ), (A2 , A3 ), (A2 , A4 ), (A3 , A3 ).
A+
1 A−
1 A2
A+
3 A−
3 A4
D4+ D4−
pairs of foliations on the surface and one cannot trace in general an indi-
vidual curve when the surface is deformed. There may of course exist lines
of curvature or asymptotic curves which are homeomorphic to a circle, i.e.,
limit cycles. These are indeed robust features of the surface. Such curves
are studied in [Sotomayor and Gutierrez (1982); Garcia and Sotomayor
(1997)].
We consider below some special curves which are robust features on M ,
captured via the contact of the surface with planes, lines and spheres.
Remark 6.4. One can also use the arguments in the proof of
Proposition 6.13 to show that there is a smooth curve on M of pairs of
points where the surface admits a bi-tangent plane meeting tangentially
the parabolic curve at a cusp of Gauss (Figure 6.10). These pairs of points
are where the height function along their common normal direction has an
A1 -singularity. The curve of bi-tangent planes corresponds to the curve
of self-intersections of the dual surface of M (a swallowtail at the cusp of
Gauss). The set is denoted the A1 A1 -set or A21 -set. For a height function
generic surface, the A1 A1 -set is a smooth curve which has ordinary tan-
gency with the parabolic curve at a cusp of Gauss. As in Proposition 6.13,
to prove this it is enough to carry out the calculations for the A-model of a
swallowtail surface in Definition 1.4. The A1 A1 -set is modelled by the pair
of (distinct) points (s1 , t1 ) and (s2 , t2 ) which have the same image by the
map (s, t) = (3s4 + s2 t, 4s3 + 2st, t). Then t1 = t2 = t and
3u41 + s21 t = 3s42 + s22 t, 4s31 + 2s1 t = 4s32 + 2s2 t.
Equivalently,
(s21 s22 )(3s21 + 3s22 + t) = 0, 2(s1 s2 )(2(s21 + s1 s2 + s22 ) + t) = 0.
The solution s1 = s2 is excluded as it gives (s1 , t1 ) = (s2 , t2 ). The
system of equations 3s21 + 3s22 + t = 0 and 2(s21 + s1 s2 + s22 ) + t = 0 also
gives s1 = s2 . Therefore the solution of the system is s2 = s1 (from the
186 Di↵erential Geometry from a Singularity Theory Viewpoint
x0 y 0 x0 fx + y 0 fy
1
= x00 y 00 x00 fx + y 00 fy
fx fy 1
= (x0 y 00 x00 y 0 ) ,
Surfaces in the Euclidean 3-space 187
where = (1 + fx2 + fy2 )1/2 and all the partial derivatives of f are evaluated
at ↵(t). Thus [ 0 (t), 00 (t), N (↵(t))] = 0 if and only if (x00 y 0 x0 y 00 )(t) = 0,
which completes the proof. ⇤
fxx (g(t), t)g 0 (t)2 + fxy (g(t), t)g 0 (t) + fyy (g(t), t) = 0.
Theorem 6.6 gives the equations of the flecnodal curve for M in Monge
form. Using Proposition 6.15, we can obtain its equations in terms of the
coefficients of the second fundamental form and their derivatives.
Theorem 6.14. For a height function and projection generic surface, the
flecnodal set contains the cusps of Gauss, is a smooth curve at such points
and has 2-point contact with the parabolic curve there.
Flecnodal curve
A1A1curve
Parabolic curve
an equation in (x, y) and gives the flecnodal curve. This curve can be
parametrised locally in the form x = x2 (y) with
a44 2
j 2 x2 (y) = y . (6.21)
a32
It is clear from (6.20) and (6.21) that the flecnodal and parabolic curves
are tangential at p. We have j 2 (x1 x2 )(y)/y 2 = a32 /a20 6= 0, so the two
curves have 2-point contact at p. ⇤
The ridge inherits the colour of the principal directions. Following Por-
teous’ colouring, we have a red ridge associated, say, to the principal cur-
vature 1 and the blue ridge associated to the principal curvature 2 .
The ridge contains the points where the singularity of some distance
squared function is more degenerate than A3 . For a distance squared func-
tion generic surface, these are either A4 or D4 -singularities.
The image of the ridge associated to the principal curvature i by the
map "i (p) = p + 1/i (p)N (p), i = 1 or 2, is precisely the singular set of the
focal set, and captures the following geometric property of the surface.
Proof. Consider, for example, the focal set associated to the principal
curvature 1 which is parametrised by
1
"1 (u1 , u2 ) = (x + N )(u1 , u2 ).
1
The directions xu1 and xu2 are principal directions at all points in U so
Nui = i xui , i = 1, 2. Di↵erentiating "1 and dropping the argument, we
get
@"1 1 @k1
= N,
@u1 21 @u1
@"1 2 1 @k1
= (1 )xu2 N.
@u2 1 21 @u2
A point p = x(u1 , u2 ) is on the ridge if the focal set is singular at
"1 (u1 , u2 ). As p is not an umbilic point, the vectors @"1 /@u1 and @"1 /@u2
are linearly dependent at (u1 , u2 ) if and only if
@k1
(u1 , u2 ) = 0.
@u1 ⇤
Remark 6.6. It follows from the proof of Proposition 6.16 that, away from
umbilic points, the normal direction to the focal set F1 , which is given by
@"1 /@u1 ⇥ @"1 /@u2 , is the principal direction v1 associated to the principal
curvature 1 . Similarly, the normal direction to the focal set F2 is the
principal direction v2 associated to the principal curvature 2 .
Surfaces in the Euclidean 3-space 191
Re(z 3 + z 2 z̄)
= 2ei✓ + e 2i✓
: this is a hypocycloid in Figure 6.11 (right) and con-
sists of the umbilics which are more degenerate than
D4 , i.e., those cubic forms with repeated roots.
194 Di↵erential Geometry from a Singularity Theory Viewpoint
4
3
2
3 1 2
2
3
Regions 3 and 4 Region 1 Region 2
Fig. 6.11 Generic configurations of the ridge curves at umbilic points left and partition
of the -plane right. The ridge associated to one focal sheet is drawn in continuous line
and the one associated to the other focal sheet is drawn in dashed line.
The authors in [Bruce and Wilkinson (1991)] proved that the bifurca-
tion set of the family of folding maps is dual to the bifurcation set of the
family of distance squared function (which is the focal set). Thus, the fold-
ing maps capture the geometry of the focal set obtained via its contact with
planes. In particular, it captures its parabolic set. The set of points on the
surface M which correspond to the parabolic set of its focal set is defined in
[Bruce and Wilkinson (1991); Wilkinson (1991)] as the sub-parabolic curve
of M . Before giving other geometric characterisations of the sub-parabolic
curve, we require the following result.
@ 2 "1
mF1 = hNF1 , i=0
@u1 u2
196 Di↵erential Geometry from a Singularity Theory Viewpoint
@ 2 "1 2
nF1 = hNF1 , i = (1 p )hxu1 , xu2 u2 i
@u22 1 E
The coefficients of the first and second fundamental forms of F2 are
computed in a similar way. ⇤
@2
(2 1 )hxu1 , xu2 u2 i = G. (6.26)
@u1
As the surface patch is umbilic free, it follows from (6.26) that equation
(6.22) is satisfied if and only if
@2
= 0.
@u1 ⇤
L=Lemon
M=Monstar
S=Star M
M S
M
L
Lemon Star Monstar
Fig. 6.12 The partition of the -plane (right) and the generic configurations of the sub-
parabolic curves at an umbilic point (first three left). The sub-parabolic curve associated
to one focal sheet is drawn in continuous line and the one associated to the other focal
sheet is drawn in dashed line.
6.7 Notes
The contact of surfaces with planes and lines is an affine and projective
property of the surface, so does not depend on the metric in the ambient
space (see for example [Bruce, Giblin and Tari (1995); Shcherbak (1986)]).
The results in this chapter are, in particular, valid for surfaces embedded
in Minkowski 3-space. For such surfaces, additional geometric information
can be obtained when considering projections along the lightlike directions
in R31 , see [Izumiya and Tari (2013)].
Surfaces in the Euclidean 3-space 199
201
202 Di↵erential Geometry from a Singularity Theory Viewpoint
These curves are robust features of the surface and part of this chapter is
devoted to determining and characterising them. Certain robust curves are
related to the contact of the surface with two model objects. This suggest
a (duality) relationship between the mappings that define the contact of
the surface with the model objects. We explore briefly this relationship.
We recall in §7.1 some properties of the second fundamental form of
M and consider some of its local intrinsic and affine invariants. We also
define the asymptotic directions at a given point and give the expression
of the binary di↵erential equation of their integral curves. The coefficients
of this equation define at each point on the surface a point in the projec-
tive plane. It turns out that its polar line with respect to the conic of
degenerate quadratic forms represent the ⌫-shape operators. Exploring the
polarity further gives us a way of choosing one special pair of ⌫-lines of
curvature which we call the lines of curvature of M . We study the con-
tact of the surface with hyperplanes (§7.6), lines (§7.7) and planes (§7.8),
and derive robust features of the surface. The contact of a surface with
hyperplanes is intimately related to that of its canal hypersurface with hy-
perplanes (§7.6.1). The contact of the surface with lines is captured by the
singularities of orthogonal projections to 3-spaces. The image of the surface
by a projection can be singular and the geometry of the singular projected
surface gives geometric information about the surface itself. We discuss
in the last section (§7.9) the contact of the surface with hyperspheres and
give geometric characterisations of the generic singularities of the family of
distance squared functions on the surface.
by !i and !ij , i, j = 1, . . . , 4 by
!i = hdx, ei i and !ij = hdei , ej i,
where d is the exterior di↵erential. It is worth observing that !3 = !4 = 0
and !ij = !ji . The Maurer-Cartan structural equations (see [do Carmo
(1976)]) are
d!i = ⌃4j=1 !ij ^ !j and d!ij = ⌃4k=1 !ik ^ !kj . (7.1)
We have
0 = d!3 = !31 ^ !1 + !32 ^ !2 ,
0 = d!4 = !41 ^ !1 + !42 ^ !2 .
It follows by a lemma of Cartan (see [do Carmo (1976)]) that there
exist a, b, c, e, f, g, such that
!13 = a!1 + b!2 , !14 = e!1 + f !2 ,
!23 = b!1 + c!2 , !24 = f !1 + g!2 .
The second fundamental form of M is the vector valued quadratic form
associated to the normal component of the second derivative d2 x of x at p
(see Chapter 2), that is,
IIp = hd2 x, e3 ie3 + hd2 x, e4 ie4 .
We have
hd2 x, e3 i = hdx, de3 i
= h!1 e1 + !2 e2 , de3 i
= !1 he1 , de3 i !2 he2 , de3 i
= !1 !31 !2 !32
= !1 !13 + !2 !23
= a!12 + 2b!1 !2 + c!22 .
Similarly, hd2 x, e4 i = e!12 + 2f !1 !2 + g!22 , so that
IIp = (a!12 + 2b!1 !2 + c!22 )e3 + (e!12 + 2f !1 !2 + g!22 )e4 .
✓ ◆
ab c
The matrix ↵ = is called the matrix of the second fundamental
ef g
form with respect to the orthonormal frame {e1 , e2 , e3 , e4 }. Its entries are
called the coefficients of the second fundamental form with respect to that
frame.
204 Di↵erential Geometry from a Singularity Theory Viewpoint
The 1-form !12 is the connection form in the tangent bundle of M and
!34 is the connection form in the normal bundle of M , and d!12 and d!34
are the respective curvature forms in those bundles. Following Little, the
forms !1 , !2 and !12 can be regarded as forms in the tangent bundle and
depend only on the metric !12 + !22 on the surface. The Gaussian curvature
of the surface can be found using
d!12 = K!1 ^ !2 .
Little defines another scalar invariant N , called the normal curvature, by
d!34 = N !1 ^ !2 . The Gaussian and normal curvatures K and N can be
expressed as follows in terms of the coefficients of the second fundamental
form:
K = (ac b2 ) + (eg f 2) and N = (a c)f (e g)b.
Let ↵1 and ↵2 be the symmetric matrices associated to the quadratic
forms hd2 x, e3 i and hd2 x, e4 i respectively, so that
✓ ◆ ✓ ◆
ab ef
↵1 = and ↵2 = .
bc f g
Then
K = K1 + K 2 with K1 = det ↵1 , K2 = det ↵2 .
In fact, there is a more general result explaining the above relation (see [do
Carmo (1976); Basto-Gonçalves (2013)]).
with
H = 12 (a + c)e3 + 12 (e + g)e4 ,
B = 21 (a c)e3 + 12 (e g)e4 ,
C = be3 + f e4 .
The normal field H is called the mean curvature vector of M at p. The
map ⌘ can also be written in matrix form
✓ ◆
cos 2✓
⌘(✓) H = A , (7.3)
sin 2✓
with
✓1 ◆
A= 2 (a c) b
.
1
2 (e g) f
It follows from (7.3) that the image of the map ⌘ : S 1 ⇢ Tp M ! Np M
is an ellipse in the normal plane Np M with centre H and principal axes
along the vectors B and C.
23!#4
23!"4 23!#4
! ! 23!"4
"!# "!#
:(5+;59+'+).-,+%* 80/09,'+).-,+%*
23!#4
23!"4
!
"!# "!#
<1-(/9,'+).-,+%* 7''+-*+).-,+%*
23!#4
!
23!"4
!
" !# "!#
!# !"
! $!#
being inside or outside the curvature ellipse can also be determined by the
sign of ∆(p), positive for inside and negative for outside.
Points on the surface are classified in terms of the curvature ellipse.
When the curvature ellipse reduces to the point p, then p is said to be a flat
umbilic. A non inflection point p 2 M is called elliptic (resp. hyperbolic,
parabolic) when it lies inside (resp. outside, on) the curvature ellipse. See
Figure 7.1.
The notation hyperbolic, elliptic and parabolic point in Definition 7.2
has nothing to do with the sign of the Gaussian curvature. They are intro-
duced in [Mond (1982)] in an analogy to the behaviour of the asymptotic
directions and curves on surfaces in R3 (see §7.3).
f 3 = ↵ 1 e3 + 1 e4 ,
f 4 = ↵ 2 e3 + 2 e4 .
and denote by
a = hxu1 u1 , e3 i, b = hxu1 u2 , e3 i, c = hxu2 u2 , e3 i,
e = hxu1 u1 , e4 i, f = hxu1 u2 , e4 i, g = hxu2 u2 , e4 i
and
l1 = hxu1 u1 , f3 i, m1 = hxu1 u2 , f3 i, n1 = hxu2 u2 , f3 i,
l2 = hxu1 u1 , f4 i, m2 = hxu1 u2 , f4 i, n2 = hxu2 u2 , f4 i
the coefficients of the second fundamental form with respect to the basis
{e3 , e4 } and {f3 , f4 } respectively. Also denote by
En = hf3 , f3 i, Fn = hf3 , f4 i, Gn = hf4 , f4 i.
Remark 7.2. It is clear from (7.5) that some properties of the curvature
ellipse (Definition 7.2) remain invariant under the action of GL(2, R) on
Np M . For instance, the position of the point p with respect to the curvature
ellipse is an affine invariant property. The concept of an inflection point is
also affine invariant. However, semiumbilicity is not affine invariant.
Surfaces in the Euclidean 4-space 211
Theorem 7.2 ([Gibson (1979)]). The orbits of the action of GL(2, R)⇥
GL(2, R) on the set of pairs of quadratic forms are as in Table 7.1.
Proof. All the properties in the statement of the proposition are invariant
under the action of GL(2, R)⇥GL(2, R) on Tp M ⇥Np M . Thus, it is enough
to take (Q1 , Q2 ) as one of the normal forms in Table 7.1. After that, the
proof follows by straightforward calculations. ⇤
Corollary 7.1. The inflection points are the common solutions of the 2 ⇥ 2
minors of the matrix
✓ ◆
l1 m1 n1
↵(u) = (u),
l2 m2 n2
of the second fundamental form. That is, p = x(u) is an inflection point if
and only if
(l1 m2 m1 l2 )(u) = 0,
(l1 n2 n1 l2 )(u) = 0,
(m1 n2 n1 m2 )(u) = 0.
(i) If Q1 , Q2 are distinct, they determine a line in RP2 which meets the
conic in 0 (resp. 1, 2) points according to (p) < 0 (resp. = 0, > 0)
with
equivalently,
Eliminating gives
(af be) cos2 ✓ + (ag ce) cos ✓ sin ✓ + (bg cf ) sin2 ✓ = 0. (7.8)
Consider now a frame {xu1 , xu2 , e3 , e4 } of Tp M ⇥ Np M , write
✓ ◆ ✓ ◆ ✓ ◆✓ ◆
cos ✓ du1 ↵1 ↵2 du1
=⌦ =
sin ✓ du2 1 2 du2
and use the relations in Proposition 7.1. Then equation (7.8) becomes
1
(l1 m2 l2 m1 )du21 + (l1 n2 l2 n1 )du1 du2 + (m1 n2 m2 n1 )du22 = 0,
det ⌦
that is
(l1 m2 l2 m1 )du21 + (l1 n2 l2 n1 )du1 du2 + (m1 n2 m2 n1 )du22 = 0.
Similarly, using the relations in Proposition 7.2, equation (7.8) becomes
equation (7.7) with respect to a frame {e1 , e2 , f3 , f4 } in Tp M ⇥ Np M . ⇤
An immediate consequence of Theorem 7.3 is the following.
˜ 1 , u2 ). Consequently, there
(ii) The discriminant of equation (7.7) is −4∆(u
are 2 (resp. 1, 0) asymptotic directions at hyperbolic (resp. not inflec-
tion parabolic, elliptic) points.
Fig. 7.2 Topological configurations of the asymptotic curves at special parabolic points.
%
$
$&!"!#
!##
*+)
% !"# (
(
'
)
%
*+)
Fig. 7.5 Polar line of the asymptotic BDE tracing the pencil of ν-shape operators, and
the determination of a special shape operator.
The above system of three equations in λ1 and λ2 has a solution if and only
if
which is precisely the condition for the point Q to be on the polar line of
asymptotic BDE A. !
Remark 7.4. (1) The induced metric on M can also be represented at each
point by the point L = (E : 2F : G) in RP2 . The point L lies inside the
conic Γ as the metric is Riemannian. Its polar line L̂ represents BDE with
orthogonal solutions at p and intersect the polar line  of the asymptotic
BDE (7.7) at a unique point P (Figure 7.5). There is a unique point WpνP
on  such that A, C, WpνP form a self-polar triangle (see Figure 7.5). In fact
the solutions of P are the νP -principal directions at p. This construction
gives a way of choosing a unique pair of orthogonal foliation on M coming
from a ν-shape operator, which can be called the lines of principal curvature
of M .
(2) The above construction is in fact inspired from and is analogous to
that for surfaces in R3 . In that case too the asymptotic curves are given by
a BDE A and the lines of principal curvatures by a BDE P (see Chapter
6). There is a unique BDE C such that A, P, C form a self-polar triangle,
which is the BDE of characteristic curves ([Bruce and Tari (2005)]). In fact
using the metric L as above, the triple A, P, C is completely determined by
A. The construction runs into difficulties for surfaces immersed in higher
218 Di↵erential Geometry from a Singularity Theory Viewpoint
(i) E = 1 + f1 2x + f2 2x , F = f1 x f1 y + f2 x f2 y , G = 1 + f1 2y + f2 2y .
(ii) We can choose the frame F = {f1 , f2 , f3 , f4 } in U with
f1 = x = (1, 0, f1 x , f2 x )
f2 = y = (0, 1, f1 y , f2 y )
f3 = ( f1 x , f1 y , 1, 0)
f4 = ( f2 x + f1 y (f1 x f2 y f1 y f2 x ), f2 y f1 x (f1 x f2 y f1 y f2 x ),
f1 x f2 x f1 y f2 y , 1 + f1 x 2 + f2 x 2 )
The frame F at the origin is the standard basis of R4 .
Surfaces in the Euclidean 4-space 219
Hence,
✓ ◆2 ✓ ◆2 ! 2
˜ (w0 ) = @ 2 g1 @ 2 g1
+ (w0 ) 0.
@x2 @x2
fs (x, y) = (f1 (x, y, s), f2 (x, y), f3 (x, y), f4 (x, y)),
Surfaces in the Euclidean 4-space 221
with
1 1
f1 (x, y, s) = (1 cos(y)) cos(x) + sin(x) sin(y) + s cos(y),
10 10
1 1
f2 (x, y) = (1 cos(y)) sin(x) cos(x) sin(y),
10 10
2 4
f3 (x, y) = (1 cos(y)) cos(2x) + sin(2x) sin(y),
5 5
2 4
f4 (x, y) = (1 cos(y)) sin(2x) cos(2x) sin(y).
5 5
The inflection points on the image of fs can be computed numerically
for a given value of the parameter s. These are the solutions of any two of
the following three equations
↵ = l1 m2 m1 l2 = 0,
= l 1 n2 n1 l 2 = 0,
= m1 n2 n1 m2 = 0,
where li , mi , ni , i = 1, 2 are the coefficients of the second fundamental form
with respect to any frame.
Figure 7.6 left (s = 1/100) and Figure 7.6 right (s = 1/20) are computer
plots of the curves = 0, ↵ = 0, = 0 and = 0. The surface for
s = 1/100 has no inflection points. When s = 1/20 the surface has 4
inflection points which can be depicted as the triple points formed by the
intersection of the three curves ↵ = 0, = 0 and = 0. In fact, one can
see from Figure 7.6 right that the parabolic curve does not pass through
these points (it has a Morse singularity of type A+ 1 at such points), so the
inflection points are of imaginary type.
Fig. 7.6 No inflection points for s = 1/100 (left), and four inflection points for s = 1/20
(right) depicted as the points of intersection of the curves α = 0, β = 0 and γ = 0.
Then j 2 hv (x, y) is the pencil v3 Q1 (x, y) + v4 Q2 (x, y), and its R-singularity
type is completely determined by the GL(2, R) ⇥ GL(2, R)-class of the pair
(Q1 , Q2 ). In particular, we can take (Q1 , Q2 ) as in Table 7.1.
If p is a hyperbolic point, we take (Q1 , Q2 ) = (x2 , y 2 ) so that
j 2 hv (x, y) = v3 x2 + v4 y 2 . There are exactly two binormal directions at
p along (0, 0, 1, 0) and (0, 0, 0, 1).
If p is an elliptic point, we take (Q1 , Q2 ) = (xy, x2 y 2 ) so that
j 2 hv (x, y) = v3 xy + v4 (x2 y 2 ). The discriminant of this quadratic form
is 4v32 + v42 > 0, so the singularity of hv at the origin is always of type A1 .
Consequently, there are no binormal directions at p.
If p is a parabolic point, we take (Q1 , Q2 ) = (x2 , xy) so that j 2 hv (x, y) =
v3 x2 + v4 xy. There is one binormal direction at p along (0, 0, 1, 0).
Finally, if p is a non-degenerate inflection point and (Q1 , Q2 ) = (x2 ±
y 2 , 0), we get j 2 hv (x, y) = v3 (x2 ±y 2 ). The direction (0, 0, 0, 1) is the unique
binormal direction at p and the 2-jet of j 2 hv is identically zero along this
direction. We have thus the following proposition.
of H(hv )(p)) is an eigenvector of the shape operator Wpv along v and its
associated eigenvalue is zero. That is, u is v-principal direction.
Proposition 7.9. With notation as in §7.4, the conditions for hv0 to have
one of the generic singularities at the origin are as follows
A2 : b20 6= 0, b33 6= 0
A3 : b20 6= 0, b33 = 0, 4b20 b44 b232 6= 0
A4 : b20 6= 0, b33 = 0, 4b20 b44 b232 = 0, 4b20 2 b55 2b20 b32 b43 + b31 b232 6= 0
D4 : b20 = 0, b30 x3 + b31 x2 y + b32 xy 2 + b33 y 3 is non-degenerate.
The above singularities are R+ -versally unfolded by the family of height
functions if and only if
A2 : always
A3 : a22 6= 0 or b32 6= 0
A4 : a22 (b20 b43 b31 b32 ) b32 (b20 a33 21 a21 b32 ) 6= 0
D4 : 3b30 (a22 b32 32 a21 b33 ) b31 (a22 b31 12 a21 b32 )
+a20 (3b31 b33 b232 ) 6= 0.
Proof. The height function on M along the normal direction v0 =
(0, 0, 0, 1) is given by hv0 (x, y) = f2 (x, y).
We first assume that b20 6= 0. Then, the singularity of hv0 at the origin
is of type Ak . In this case, hv0 has an A2 -singularity at the origin if and
only if b33 6= 0. If b33 = 0, we can make changes of coordinates in the
source to reduce the 4-jet of f2 to j 4 f2 (x, y) = b20 x2 + b32 xy 2 + b44 y 4 .
Then, the singularity is of type A3 if and only if b33 = 0 and b20 x2 +
b32 xy 2 + b44 y 4 is not a perfect square, equivalently, b232 4b20 b44 6= 0.
If b33 = 0 and b20 x2 + b32 xy 2 + b44 y 4 is a perfect square, we can make
changes of coordinates in the source and show that j 5 f2 ⇠R(5) b20 x2 + Dy 5 ,
with D = a22 (b20 b43 b31 b32 ) b32 (b20 a33 12 a21 b32 ). Hence, hv0 has A4
singularity if and only if D 6= 0.
If b20 = 0, the result follows from the characterisation of a D4
singularity.
Surfaces in the Euclidean 4-space 225
LRe · f2 , hence
xk y d k
2 J d (LRe · f2 ). (7.11)
Then, when d = 2, if b32 6= 0 or a22 6= 0, we can solve (7.10) in J 2 (2, 1).
When d = 3 and 4, writing y i j 3 f2 x , and y i 1 j 3 f2 y i = 1, 2 and using
(7.11) to eliminate the monomials of degree 2 + i, i = 1, 2 divisible by x,
we can write the equations
2b20 xy i + b32 y 2+i 2 J 2+i (LRe · f2 ), i = 1, 2,
b31 x2 y i 1 + 2b32 xy i + 4b44 y 3+i 1 2 J 2+i (LRe · f2 ), i = 1, 2,
b31 x2 y i 1 2 J 2+i (LRe · f2 ).
Since 4b20 b44 b232 6= 0, it follows that all monomials of degree 3 and 4 are
in J 4 (LRe · f2 ). ⇤
Theorem 7.6. There is an open and dense set OH̄ in Imm(U, R4 ) such
that for any x 2 OH̄ , the surface M = x(U ) satisfies the conditions in
Theorem 7.5. Moreover, the following properties hold.
(i) For any v 2 S 3 , the height function h̄v on CM (") has only local sin-
gularities of R type A1 , A2 , A3 , A4 or D4 .
(ii) The singularities of h̄v are R+ -versally unfolded by the family H̄.
(iii) The generalised Gauss map G is stable as a Lagrangian map. The
singularities of G are given in Table 7.2.
2.13). It follows from the definition of the sets S1,0 (G), S1,1,0 (G), S1,1,1,0 (G)
that this direction is transverse to the surface S1,0 (G) and is tangent to the
surface Kc 1 (0) on the curve of points of type S1,1 (G). It is transverse to
this curve at general points and tangent to it at points of type S1,1,1 (G).
Let ⇠ : CM (") ! M be the projection of CM (") to M , given by
⇠(p, v) = p. Then, it follows from part (ii) in Proposition 7.11 that the
image of the parabolic set Kc 1 (0) by ⇠ is the set 0. More precisely,
let ⇠¯ be the restriction of ⇠ to the regular surface S1 (G) = Kc 1 (0) \ S2 (G).
We denote by M = {p 2 M : (p) < 0} the set of hyperbolic points
on M and by B = {(p, v) 2 Kc 1 (0) : p 2 M }.
Proposition 7.12. With notation as above,
(i) the restriction ⇠¯|B : B ! M is a local di↵eomorphism, more precisely,
it is a double cover;
(ii) a point p 2 is not an inflection point if and only if there exists v 2 S 3
such that (p, v) is a fold singularity of ⇠¯ .
Proof. We take M locally in Monge form with f1 and f2 as in Section
7.4 and choose the setting in the proof of Proposition 7.9. The modified
height function on M along v = (v1 , v2 , v3 , 1) near v0 = (0, 0, 0, 1) is given
by
hv (x, y) = v1 x + v2 y + v3 f1 (x, y) + f2 (x, y).
The function hv has a singularity at a point (x, y) if and only if
@f1 @f2 @f1 @f2
v1 = v3 (x, y) (x, y) and v2 = v3 (x, y) (x, y).
@x @x @y @y
The singularity p = (x, y) is degenerate if furthermore det H(hv )(x, y) =
0, equivalently Kc (x, y, v3 ) = 0 (see Proposition 7.11). We have
detH(hv )(x, y) = A0 (x, y)v32 + A1 (x, y)v3 + A2 (x, y) = 0,
where
2 2
@ 2 f1 2
A0 (x, y) = @@xf21 (x, y) @@yf21 (x, y) ( @x@y ) (x, y)
@ 2 f1 @ 2 f2 @ 2 f1 @ 2 f2
A1 (x, y) = @x 2 (x, y) @y 2 (x, y) + @y 2 (x, y) @x2 (x, y)
@ 2 f1 @ 2 f2
(7.12)
2 @x@y (x, y) @x@y (x, y)
@ 2 f2 @ 2 f2 @ 2 f2 2
A2 (x, y) = @x 2 (x, y) @y 2 (x, y) ( @x@y ) (x, y).
Write K̃c = det(H(hv )) so that Kc = K̃c for some function with
(x, y, v3 ) 6= 0. Then, denoting by z any of the variables x, y, v3 , one can
show inductively that
@Kc @ m Kc @ m+1 Kc
Kc = = ... = = 0 and 6= 0
@z @z m @z m+1
Surfaces in the Euclidean 4-space 229
and it follows that has non-zero torsion if and only if b33 6= 0. This is the
case if and only if p is an A2 singularity of hv .
Now, it follows from Proposition 7.10 and Theorem 7.6 that a hyperbolic
point p 2 M (✏) is a singularity of type A2 , A3 , A4 of hv if and only if
(p, v) 2 CM is respectively a fold, cusp or swallowtail singularity of G.
Hence, if b33 = 0, the following follows from Remark 7.6.
1 2
is A0 (x, y)A2 (x, y) 4 A1 (x, y), which is exactly the set . Then:
(a) p is an A2 singularity of hv if and only if @A
@y (0, 0) = 12b20 b33 6= 0 if
2
Corollary 7.3. For a height function generic immersion, the flat ridge set
1
is a regular curve in M which is tangent to the curve (0).
In §7.6, the second order geometry of a surface derived from its contact
with hyperplanes is determined by a pencil of quadratic forms. This pencil
also determines some of the second order geometry of the surface associated
to its contact with lines.
We take the surface in Monge form (x, y) = (x, y, f1 (x, y), f2 (x, y)),
at the origin p with f1 and f2 as in §7.4 and project along a unit tangent
direction v = (v1 , v2 , 0, 0) 2 Tp M (so that Pv is singular at p). Then
Pv (x, y) = (x, y) h (x, y), viv
= (x, y, f1 (x, y), f2 (x, y)) (xv1 + yv2 )(v1 , v2 , 0, 0)
= (x (xv1 + yv2 )v1 , y (xv1 + yv2 )v2 , f1 (x, y), f2 (x, y))
= (v2 (v2 x v1 y), v1 (v2 x v1 y), f1 (x, y), f2 (x, y)).
The image of Pv is in v? , the linear 3-space orthogonal to v. If we
denote by (X, Y, Z, W ) the coordinates in R4 , and if v2 6= 0, we can compose
Pv with the projection (X, Y, Z, W ) 7! (X, Z, W ) and obtain a map-germ
A-equivalent to Pv (the projection is a di↵eomorphism restricted to v? ).
If v2 = 0, we compose Pv with the projection (X, Y, Z, W ) 7! (Y, Z, W )
instead. In both cases we still denote by Pv the composite map. Rescaling
the first coordinate (or the second if v2 = 0) gives
Pv (x, y) = (v2 x v1 y, f1 (x, y), f2 (x, y)). (7.13)
We have j 2 Pv (x, y) = (v2 x v1 y, Q1 (x, y), Q2 (x, y)), with Q1 and Q2
representing j 2 f1 and j 2 f2 at the origin, respectively. We are interested in
the A-singularities of Pv . Then one of the conditions that determines the
234 Di↵erential Geometry from a Singularity Theory Viewpoint
Proof. (i) We prove the statement for the case p a hyperbolic or parabolic
point; the argument for the elliptic points is similar. We take (Q1 , Q2 ) =
(x2 , y 2 ) at a hyperbolic point so that j 2 Pv (x, y) = (v2 x v1 y, x2 , y 2 ). If v2 6=
0, we can make the change of coordinates X = v2 x v1 y, Y = y and obtain
a map-germ A(2) -equivalent to (X, 2v v2
1
XY, Y 2 ). We get a cross-cap if and
2
only if v1 6= 0. When v1 = 0, j 2 Pv is A(2) -equivalent to (X, 0, Y 2 ). If v2 = 0,
then j 2 Pv is A(2) -equivalent (Y, X 2 , 0). Therefore, the singularity is a cross-
cap unless v1 = 0 or v2 = 0. Observe that the degenerate singularities of
Pv are Sk , Bk or Ck .
At a parabolic point and with (Q1 , Q2 ) = (x2 , xy), we get j 2 Pv (x, y) =
(v2 x v1 y, x2 , xy). If v2 = 0, the singularity is a cross-cap. If v2 6= 0,
the same change of coordinates as above reduces the 2-jet to (X, 2v1 XY +
v12 Y 2 , XY +v1 Y 2 ) which is a cross-cap unless v1 = 0. If v1 = 0, j 2 Pv (x, y) is
A(2) -equivalent to (X, XY, 0), which leads to a projection P -generic surface
to H2 , H3 or P3 (c) singularity.
As before, the point (A : B : C) 2 RP2 represents the quadratic form
Ax2 +2Bxy+cy 2 . The conic has equation AC B 2 = 0 and any point on it
represents perfect squared (ax+by)2 , so has the form (a2 : ab : b2 ). It follows
that the tangent line to at (a2 : ab : b2 ) has equation Ca2 2Bab+Ab2 = 0.
Now the pencil determined by v consists of quadratic forms having
(v2 x v1 y) as a factor, so is the line tangent to at (v22 , v1 v2 , v12 ). There-
fore it has equation Cv22 + 2Bv1 v2 + Av12 = 0. The pencil determined by
(Q1 , Q2 ) intersects at (1 : 0 : 0) and (0 : 0 : 1). The point (1 : 0 : 0) (resp.
Surfaces in the Euclidean 4-space 235
Definition 7.8. We call S2 -curve (resp. B2 -curve and H2 -curve) the clo-
sure of the set of points p on M for which there exists a projection Pv
236 Di↵erential Geometry from a Singularity Theory Viewpoint
Remark 7.7. (1). The results in Proposition 7.17 follow in fact from
the duality in [Bruce and Nogueira (1998)] between certain strata of the
bifurcation set of the family of height functions with that of the family of
projections.
(2) The A4 -singularities hw are in general not related to the B3 -
singularities of Pv .
(3) The C3 -singularity of Pv occurs at a point of intersection of the
S2 -curve and B2 -curve. The S2 -curve is in general not related to the sin-
gularities of the height functions.
(4) For a projection generic surface, the robust curves captured by the
singularities of Pv are as in Figure 7.7. It is shown in [Bruce and Tari
(2002)] that the B2 -curve and the -set (i.e., the parabolic curve) meet
tangentially. The point of tangency is a P3 (c)-singularity of a Pv . The two
asymptotic direction fields in the hyperbolic region can be coloured, so we
have two di↵erent coloured B2 -curves. The B2 -curve changes colour at a
Surfaces in the Euclidean 4-space 237
!"#$%&'()*+%$,)'- !"#$%&'(
)"#$%&'(
)*+!*
/, !,
),
-,
.,
3-4($*/)'-+#')-/2 -"#$%&'(
.(()#/)*+%$,)'-
012$/
Fig. 7.7 Robust curves on the surface depicted by its contact with lines.
?
A2 : if p is a hyperbolic or parabolic point, w = vN and is a binor-
mal direction.
?
A3 : w = vN is a binormal direction, p is on the B2 -curve and v
is away from a circle of directions C in the sphere w? 2 D.
Then the v-P P S is a regular curve.
?
N V A3 : w = vN is a binormal direction, p is on the B2 -curve and
v 2 C. For generic v 2 C the singularity of the v-P P S is
an A1 . For isolated directions in C the singularity becomes
an A2 , and for special points on the B2 -curve it becomes an
A3 -singularity.
?
A4 w = vN is a binormal direction, p is an A4 -point on the B2 -
curve.
?
D4 : w = vN is a binormal direction, p is an inflection point.
Surfaces in the Euclidean 4-space 239
Proof. The proof can be found in [Oset Sinha and Tari (2010)]; see also
[Nuño-Ballesteros and Tari (2007)]. ⇤
Suppose now that v 2 S 3 is a tangent direction at p 2 M . Then Pv (M )
is a singular surface Pv (p). The map-germ Pv is of corank 1 and these can
be written, in some coordinate system, in the form
(x, y) = (x, p(x, y), q(x, y))
with p, q 2 M2 (x, y). We denote by Q1 (x, y) = j 2 p(x, y) = p20 x2 + p21 xy +
p22 y 2 and Q2 (x, y) = j 2 q(x, y) = q20 x2 + q21 xy + q22 y 2 . We can consider
the action of GL(2, R)⇥GL(2, R) on (Q1 , Q2 ) as in §7.2.1 to define an affine
property of the singular surface S, image of , at its singular point.
Definition 7.10. The singular point of the surface S is called hyperbolic
(resp. elliptic, parabolic) or (a generic) inflection point if the GL(2, R) ⇥
GL(2, R)-class of (Q1 , Q2 ) can be represented by (x2 , y 2 ) (resp. (x2 , x2 y 2 ),
(x2 , xy)) or (x2 ± y 2 , 0), as in Table 7.1.
The surface S = Pv (M ) has a cross-cap singularity if and only if v 2
Tp M is not an asymptotic direction at p. It is shown in [Bruce and West
(1998); West (1995)] that a parametrisation of a cross-cap can be taken,
by a suitable choice of a coordinate system in the source and affine changes
of coordinates in the target, in the form
(x, y) = (x, xy + f1 (y), y 2 + ax2 + f2 (x, y)), (7.14)
4 3
where f1 2 M (y) and f2 2 M (x, y). The following is also shown in [West
(1995)]. When a < 0, the height function along any normal direction at
the cross-cap point has an A1 -singularity. Such cross-caps are labelled hy-
perbolic cross-caps as the surface has negative Gaussian curvature at all
its regular points (Figure 7.8, left). When a > 0, there are two normal
p
directions (0, ±2 a, 1) at the cross-cap point along which the height func-
tion has a singularity more degenerate than A1 (i.e., of type A 2 ). Such a
cross-cap is labelled elliptic cross-cap (Figure 7.8, right). The singularity
of the height function along the degenerate normal direction is precisely of
type A2 if and only if j 3 f2 (⌥ p1a , 1) 6= 0. When a = 0, there is a unique
normal direction at the cross-cap point where the height function has a
singularity more degenerate than A1 . The singularity of its corresponding
3
height function is of type A2 if and only if @@xf32 (0, 0) 6= 0. Such a cross-cap
is labelled parabolic cross-cap (Figure 7.8, centre).
Theorem 7.10. A cross-cap is hyperbolic (resp. elliptic, parabolic) if
and only if its singular point is elliptic (resp. hyperbolic, parabolic) as in
Definition 7.10.
240 Di↵erential Geometry from a Singularity Theory Viewpoint
Fig. 7.8 Hyperbolic cross-cap (left) and elliptic cross-cap (right) separated by a
parabolic cross-cap (centre).
Table 7.5 The generic structure of the v-P P S and of its two components.
S B1 B2 B3
L1 A±
1 A2 A2
v-P P S D4± D5 D5
S S2 S3 C3
L1 A2 A±
3 D4±
l,m
I2,2 : (x2 + y 2l+1 , y 2 + x2m+1 ), l ≥ m ≥ 1
l
II2,2 : (x2 − y 2 + x2l+1 , xy), l ≥ 1.
244 Di↵erential Geometry from a Singularity Theory Viewpoint
with the second component tracing the pencil (Q1 , Q2 ) in RP2 . As we fixed
u, we cannot take (Q1 , Q2 ) as one of the normal forms in Table 7.1. We
have
and its bifurcation set BD is the focal set of M. It follows that the catastro-
phe map of the family D coincides with the map Ge : N M ! R4 defined
by Ge (u, a) = a, which we call the normal exponential map of M.
Proof. The proofs of (i) and (ii) follow from Theorem 4.8. Since the
catastrophe map of a Morse family of functions can be identified with the
Lagrangian map (cf. §5.4), Ge is a Lagrangian map. The proof of (iii)
follows from Theorem 5.4. ⇤
We prove next that the spherical contact directions are principal direc-
tions associated to normal fields pointing towards the focal points of the
surface.
Proof. Let
a 3 b3 c 3
↵(p) = ,
a 4 b4 c 4
be the matrix of the second fundamental form of x at p.
The curvature ellipse is given by
⌘(✓) = H + cos(2✓)B + sin(2✓)C.
Surfaces in the Euclidean 4-space 249
Proof. The result follows from Theorem 7.14 and Lemma 7.3. ⇤
7.10 Notes
With the aim of illustrating how the singularity techniques can be applied
to analyse the extrinsic geometry of surfaces in higher codimensions we
discuss in this chapter the geometrical properties associated to the contacts
of surfaces with hyperplanes and hyperspheres in R5 . In this, as well as in
higher codimensional cases, the curvature ellipse at each point of the surface
determines a proper subspace of the normal space and both, its dimension
and relative positions with respect to the considered point are relevant in
the description of the second order geometry of the surface at this point.
Moreover, di↵erently from the case of surfaces in R4 , the directions leading
to degenerate singularities of height functions at each point are not finite.
These directions determine a cone in the normal space at the considered
point. Attending to the behaviour of this cone we can distinguish among
di↵erent types of points. The corresponding properties are described in
§8.1. The generic singularities of the height functions on surfaces in R5 are
analysed in §8.2. Analogously to the case of surfaces in R4 , this setting
leads naturally to the introduction of binormal and asymptotic directions
at each point of the surface. Since the number of normal directions leading
to degenerate singularities of height functions at each point is not finite,
we use the concept of binormal direction for those leading to singularities
of height functions on the surface with codimension higher than one, which
only occur in a finite number of normal directions at every point. The
corresponding contact directions are the asymptotic directions at the point.
Alternative characterisations of asymptotic directions in terms of normal
sections and of the geometrical behaviour of orthogonal projections of the
surface into 3- and 4-spaces are also given. We see in Proposition 8.9 that
the number of asymptotic directions at each point is at least one and at
most five. The possible local configurations, described in Proposition 8.10,
251
252 Di↵erential Geometry from a Singularity Theory Viewpoint
where
H = 12 (a1 + c1 )e3 + 12 (a2 + c2 )e4 + 12 (a3 + c3 )e5 ,
B = 12 (a1 c1 )e3 + 12 (a2 c2 )e4 + 12 (a3 c3 )e5 ,
C = b 1 e3 + b 2 e4 + b 3 e5 .
Definition 8.2. The normal field
1 1 1
H = (a1 + c1 )e3 + (a2 + c2 )e4 + (a3 + c3 )e5
2 2 2
evaluated at a point p is the mean curvature vector of M at p.
We write
f 3 = ↵ 1 e3 + 1 e4 + 1 e5 ,
f 4 = ↵ 2 e3 + 2 e4 + 2 e5 ,
f 5 = ↵ 3 e3 + 3 e4 + 3 e5
and denote by
ai = hxu1 u1 , ei+2 i, bi = hxu1 u2 , ei+2 i, ci = hxu2 u2 , ei+3 , i , i = 1, 2, 3
and
li = hxu1 u1 , fi+2 i, mi = hxu1 u2 , fi+2 i, ni = hxu2 u2 , fi+2 i , i = 1, 2, 3
the coefficients of the second fundamental form with respect to the basis
{e3 , e4 } and {f3 , f4 , f5 } respectively. Also denote by
En = hf3 , f3 i, Fn = hf3 , f4 i Gn = hf3 , f5 i,
Hn = hf4 , f4 i, In = hf4 , f5 i, Jn = hf5 , f5 i.
Proposition 8.2. Denote by ai , bi , ci , i = 1, 2, 3 the coefficients of the sec-
ond fundamental form with respect to the basis {e1 , e2 , e3 , e4 , e5 } and by
li , mi , ni , i = 1, 2, 3 its coefficients with respect to the basis {e1 , e2 , f3 , f4 , f5 }.
Then
0 1 0 10 1 10 1
a 1 b1 c 1 ↵1 ↵2 ↵3 E n Fn G n l 1 m1 n 1
@ a 2 b 2 c 2 A = @ 1 2 3 A @ Fn H n I n A @ l 2 m 2 n 2 A .
a 3 b3 c 3 1 2 3 Gn I n J n l 3 m3 n 3
Definition 8.3. We define the subsets
Mi = {p 2 M | rank ↵p = i}, i 3.
a) M3 is an open subset of M .
b) M2 is a regularly embedded curve.
256 Di↵erential Geometry from a Singularity Theory Viewpoint
Proof. We show first that generically the rank of the second fundamental
form at any point is at least 2. Let x : U ! R5 a local parametrisation of M,
and {xu1 , xu1 , e3 , e4 , e5 } a frame in U. With respect to these coordinates,
for each p 2 U, the second fundamental form at p is represented by the 3⇥3
matrix
0 1
l1 m1 n1
↵p = @ l2 m2 n2 A
l3 m3 n3
where
j2x ⇧⇤
U ! J 2 (U, R5 ) ! M(3),
where ⇧⇤ (j 2 x(p)) = ↵p .
The mapping ⇧⇤ : J 2 (U, R5 ) ! M(3) is a submersion, as we can
take the variables li , mi and ni as coordinates in the jet space J 2 (U, R5 ).
Then, the stratification (Si ) in M(3) pulls back to a stratification (Ti ) in
J 2 (U, R5 ), such that cod Ti = cod Si .
Then, it follows from Thom’s Transversality Theorem that for a generic
embedding x : U ! R5 , the map j 2 x : U ! J 2 (U, R5 ) does not intersect
the strata Ti , i 1, since they have codimension greater than or equal to
4. Hence, for a generic embedding x, rank ↵p 2, 8p 2 U.
Let (p) = det(↵p ), with p = x(u) 2 M . We have that M M3 =
1
(0) and since is a continuous function on M , it follows that M3 must
be an open region in M .
We can consider a local representation of M in its Monge form at a
point p 2 M ,
: R2 , 0 ! R5
(x, y) 7 ! (x, y, f1 (x, y), f2 (x, y), f3 (x, y)).
In these coordinates
Surfaces in the Euclidean 5-space 257
(p) = f1xx f2xy f3yy f1xy f2xx f3yy f1xx f2yy f3xy + f1yy f2xx f3xy +
f1xy f2yy f3xx f1yy f2xy f3xx .
Now, it follows from this expression that, under appropriate transver-
sality conditions on the 3-jet of , the set = 0 represents a curve possibly
with isolated singular points determined by the vanishing of the deriva-
tives of the function . Since the orthogonality property of the frame
{e1 , e2 , e3 , e4 , e5 } is irrelevant for our study, we can take {e3 , e4 , e5 } such
that e5 generates Ker(Ap ).
If p 2 M2h , we choose {e3 , e4 } as the two degenerate directions in Np M .
Furthermore, we can also make linear changes of coordinates in source and
target, such that the two degenerate directions correspond to the quadratic
forms x2 and y 2 in C. Thus f can be locally written as
Cp
Fig. 8.1 The curvature ellipse and the cone of degenerate directions at a point on a
surface in R5 .
Proposition 8.4. The cone whose basis is the curvature ellipse is perpen-
dicular to the cone Cp of degenerate directions at p.
binormal direction has higher order contact with M at p and we call it os-
culating hyperplane at p. The corresponding contact directions are called
asymptotic directions on M . In the special case of the corank 2 singularities
D4± and D5 we refer to them as umbilic binormals.
Proof. This follows easily by taking the embedding in Monge form and
observing that the matrix of the second fundamental form in a normal
direction v at p coincides with the Hessian matrix of the height function
hv at p. ⇤
Taking into account the relation between the singularities of the gener-
alised Gauss map G, as a catastrophe map of the height functions family,
and the singularities of the height functions on the surface M (see Remark
3.2), we observe:
Then we can characterise the flat contact directions and the asymptotic
directions as follows:
We can study now the possible numbers and the distribution of asymp-
totic directions over the points of a generic surface M .
A1 = [ @n @m @n
@x , m, n] + 2[ @y , m, n] + [ @y , l, n]
A2 = [ @n @m @l @m @n
@x , l, n] + 2[ @x , m, n] + [ @y , m, n] + 2[ @y , l, n] + [ @y , l, m]
@l
A3 = [ @x , m, n] + 2[ @m @n @m @l
@x , l, n] + [ @x , l, m] + 2[ @y , l, m] + [ @y , l, n]
@l
A4 = [ @x , l, n] + 2[ @m @l
@x , l, m] + [ @y , l, m]
@l
A5 = [ @x , l, m]
where l = (l1 , l2 , l3 ), m = (m1 , m2 , m3 ), n = (n1 , n2 , n3 ).
Definition 8.8. The singular set of ⇠2 : S1,1 (G) ! M is known as the
criminant set and its image by ⇠2 , denoted by , is the discriminant set.
The discriminant set is a (non necessarily connected) curve separating
M into regions made of points with a constant number of asymptotic di-
rections. The possible local configurations of the asymptotic curves on a
surface generically immersed in R5 are described in the following proposi-
tion whose proof can be found in [Romero Fuster, Ruas and Tari (2008)].
Proposition 8.10.
(1) The local configurations of the asymptotic curves of a height function
generic surface in R5 are modeled by super-imposing in each quadrant
in Figure 8.2 1 and 2 one figure from the left column with one from
the right column.
(2) Let q 2 M3 be a point on the discriminant of the asymptotic IDE and
u the double asymptotic direction there. Then q is a folded-singularity
of the asymptotic IDE at (q, u) if and only if q is an A4 -singularity of
the height function along u⇤ (Figure 8.2 1 (a or b)+(g, h or i)).
Surfaces in the Euclidean 5-space 265
! %
&
+ "
' + $
( #
#
)
$
Definition 8.9. The flat ridge FR of M is the set of points for which there
exists a height function (in some binormal direction), having a singularity
of type Ak , k ≥ 4.
FR = ξ(S1,1,1 (G)).
1) There are at least one and at most five strong principal lines through
any point of M . The generic local qualitative behaviour of the strong
principal foliations of M corresponds to the one described above for the
asymptotic curves of surfaces generically immersed in R5 (see Figure
8.2).
Surfaces in the Euclidean 5-space 267
We deal now with the case when ⇡(w1 ,w2 ) |M is singular. This means
that the kernel of the projection ⇡(w1 ,w2 ) contains a tangent direction at
q. When {w1 , w2 } = Tq M , the map-germ ⇡(w1 ,w2 ) |M has rank zero at the
origin and does not identify the asymptotic directions. We shall assume
that {w1 , w2 } is distinct from Tq M . Then ⇡(w1 ,w2 ) |M has rank 1 at the
origin.
270 Di↵erential Geometry from a Singularity Theory Viewpoint
Fig. 8.3 Configurations of the asymptotic curves on surfaces in R3 at a flat umbilic point
(elliptic left and hyperbolic right). There are three separatrices at an elliptic umbilic
and one separatrix at a hyperbolic umbilic ([Bruce and Tari (1995)]).
The map ⇡(w1 ,w2 ) can be considered locally as a corank 1 smooth map-
germ R2 , 0 ! R3 , 0. Since ⇡(w1 ,w2 ) is a 6-parameter family, we expect the
map ⇡(w1 ,w2 ) to have only simple singularities of Ae -codimension 6 or
non-simple singularities with the Ae -codimension of the stratum 6 (as
dim G(2, 5) = 6). The A-simple singularities of map-germs R2 , 0 ! R3 , 0 of
Ae -codimension 6 are given in Table 4.2. The complete classification of
non-simple singularities of Ae -codimension of the stratum 6 is not known
so far.
⇧(w1 ,w2 ) |M : R2 , 0 ! R2 , 0.
⇧(u? ,v) |M (x, y) = (y, v1 (x2 +f1 (x, y))+v2 (xy+f2 (x, y))+v3 (cy 2 +f3 (x, y))),
(1) The projection ⇧(u? ,v) |M has a fold singularity for almost all v 2 S 2 .
272 Di↵erential Geometry from a Singularity Theory Viewpoint
(1) The 2-jet of the projection ⇧(w1 ,w2 ) |M is A-equivalent to (x2 , y 2 ), (x2
y 2 , xy) or (x2 , xy) if and only if q is, respectively, a hyperbolic, elliptic
or parabolic point of M(w1 ,w2 ) .
(2) The 2-jet of the projection ⇧(w1 ,w2 ) |M is A-equivalent to (x2 + y 2 , 0),
(x2 y 2 , 0), or (x2 , 0) if and only if q is, respectively, an inflection point
of real type, of imaginary type or of flat type of M(w1 ,w2 ) .
Moreover, if q 2 M3 then ⇧(w1 ,w2 ) |M satisfies (1) for every plane
{w1 , w2 } ⇢ Nq M . The point q 2 M2 if and only if there exists a
direction w2 2 Nq M such that q is an inflection point of M(w1 ,w2 ) , for
any w1 2 Nq M.
D : R2 ⇥ R 4 ! R
(u, a) 7 ! kx(u) ak2 = da (u)
Surfaces in the Euclidean 5-space 273
The catastrophe manifold and map are respectively given by the normal
bundle of M in R5 , N M = {(u, a) 2 U ⇥ R5 | a = x(u) + v, v 2 (Np M )1 }
and the normal exponential map,
Ge : N M ! R5
(u, a) 7 ! a.
(i) Given any point p 2 M , the distance squared function da from a point
a 2 R5 such that a p 2 Np M has only singularities of K- type
A1 , A2 , A3 , A4 , A5 , A6 , D4 , D5 and D6 .
(ii) The singularities of da are R-versally unfolded by the family D.
(iii) The normal exponential map Ge is stable as a Lagrangian map.
We observe that the proofs of Lemmas 7.2 and 7.3 can be easily adapted
to surfaces in R5 , so we can state the following:
The focal set of M clearly coincides with the bifurcation set of the family
D, given by
Definition 8.13. The rib set of M is the singular set of the focal set F.
where H, B and C are the vectors that determine the curvature ellipse of
M at p. We can express the focal set at p, in terms of these vectors:
n o
2 2
Fp = a = p + v 2 Np M ; (hH, vi 1)2 = hB, vi + hC, vi .
(2) For p 2 M2 and dim(Ep ) = 2 all focal points at p have corank 1 and
we may have:
(a) If p 2 M2e then Wp = ?.
(b) If p 2 M2h then Wp is a hyperbola in Ep .
(c) If p 2 M2p then Wp is a parabola in Ep .
Theorem 8.3.
(a) Given a point p 2 M3 , there exists a unique umbilical focus at p, given
by
1
ap = p + vp ,
d(p, Ep )
with vp = unit normal vector orthogonal to the plane Ep , such that
hvp , Hi > 0.
276 Di↵erential Geometry from a Singularity Theory Viewpoint
Corollary 8.3.
(a) If p 2 M3 , then u (p) is the curvature of the unique umbilic focal
hypersphere of M at p.
(b) If p 2 M2 is a non semiumbilic point, then u (p) = 0.
(c) If p 2 M2 is a semiumbilic point, u (p) is the maximum curvature
among those of all the tangent hyperspheres centred at umbilical foci of
M at p.
(d) If p is 2-regular then u (p) 6= 0.
Surfaces in the Euclidean 5-space 277
Proof. The proofs of all these assertions follow easily from the definition
of u . ⇤
8.5 Notes
Relative mean curvature foliations: The study of the mean directional con-
figurations carried out by Mello ([Mello (2003)]) on surfaces in R4 has
been generalised to surfaces in Rn , n 5 in [Gonçalves, Martı́nez Alfaro,
Montesinos Amilibia and Romero Fuster (2007)].
We point out first that surfaces immersed in codimension higher than
2 do not admit mean curvature directions in the way defined for surfaces
in R4 . Nevertheless it is possible to introduce certain foliations on surfaces
immersed in Rn , that in the particular case of surfaces in 4-space coincide
with the mean curvature foliations. The procedure is based on the fact
that, from a qualitative viewpoint, all the principal configurations on S
arise from normal vector fields parallel to the subspace determined by the
curvature ellipse at every point. In fact, any normal vector v 2 Np S can
be decomposed into a sum v > + v ? , with v > and v ? respectively parallel
and orthogonal to the plane determined by the curvature ellipse. Since the
shape operator associated to v ? is a multiple of the identity ([Moraes and
Romero Fuster (2005)]), then the eigenvectors of the shape operator Wv
and Wv> must coincide. This idea lead in [Gonçalves, Martı́nez Alfaro,
Montesinos Amilibia and Romero Fuster (2007)] to the definition of the
relative mean curvature directions at a point p of a surface immersed in Rn
with n > 4 as those inducing normal sections with curvature vector par-
allel to H(p)> . This gives rise to two orthogonal foliations whose critical
points are the semiumbilics and the pseudo-umbilics (with inflection points
and minimal points considered as non generic particular cases). Interest-
ing global consequences of this setting are the following facts ([Gonçalves,
Martı́nez Alfaro, Montesinos Amilibia and Romero Fuster (2007)]):
ing that they can be seen as the image of a stable map from a convenient
surface (non necessarily connected) into the 3-manifold. On the other hand,
Dreibelbis ([Dreibelbis (2012)]) generalises to n-manifolds in R2n the no-
tions of asymptotic directions and parabolic and inflection points of sur-
faces in R4 , analysing in detail the case of 3-manifolds in R6 . An interesting
fact, described in this work is that the parabolic subset of a 3-manifold M
generically immersed in R6 is a surface with normal crossings and possible
isolated cross-caps. Moreover, the cross-caps (resp. triple points) are the
points of M for which the origin of the normal space is a cross-cap (resp.a
triple) point of the curvature locus.
Surfaces in the Euclidean 5-space 279
281
282 Di↵erential Geometry from a Singularity Theory Viewpoint
Consider the orientation of R41 given by the volume form e⇤0 ^ e⇤1 ^ e⇤2 ^ e⇤3 ,
where {e⇤0 , e⇤1 , e⇤2 , e⇤3 } is the dual basis of the canonical basis of R41 . The
Minkowski space R41 has also a time-orientation given by the choice of e0
as a future timelike vector field.
As in the previous chapters, we are interested in the local geometric
properties of smooth surfaces in R41 . For this reason we consider a surface
patch M which is the image of an embedding x : U ! R41 , where U is an
open set of R2 (so M = x(U )).
We say that M is a spacelike surface if its tangent plane Tp M is spacelike
at all points p 2 M . Then the pseudo-normal space Np M of M at p is a
timelike plane. We denote by N (M ) the pseudo-normal bundle over M.
Since this is a trivial bundle, there are two well defined transverse lightlike
286 Di↵erential Geometry from a Singularity Theory Viewpoint
Lemma 9.1. For any given smooth lightlike vector field L on M , there is a
unique unit timelike normal vector field u on M and a unique unit spacelike
normal vector field v on M such that L = u + v. Observe that u and v are
necessarily pseudo-orthogonal on M .
Remark 9.1. The result in Lemma 9.1 shows that any normal lightlike
vector field on M can be obtained in the following way. Suppose that M is
parametrised by x : U ! R41 and choose a unit timelike normal vector field
nT on M (this can always be done as N (M ) is a trivial bundle). Then
nT (u) ^ xu1 (u) ^ xu2 (u)
nS (u) = (9.1)
knT (u) ^ xu1 (u) ^ xu2 (u)k1
is a unit spacelike normal vector field which is pseudo orthogonal to nT (u).
The vector fields nT + nS and nT nS are lightlike normal vector fields
on M and determine at each point p 2 M the two lightlike directions in
Np M . Any smooth lightlike vector field L on M coincides with nT + nS or
nT nS for some chosen nT (take nT to be u in Lemma 9.1 so v in Lemma
9.1 coincides with nS or nS ).
Since M is spacelike, e0 is a future directed timelike vector field along
M which is transverse to M at all points. We can write any v 2 Tp R41 |M in
the form v = v1 +v2 , where v1 2 Tp M and v2 2 Np M. If v is timelike, then
v2 is timelike. Let ⇡Np M : Tp R41 |M ! Np M be the canonical projection,
Spacelike surfaces in the Minkowski space-time 287
± T S
i = i (n ) ± i (n ).
Proof. This follows from the fact that W (L± )p = W (nT )p ± W (nS )p .⇤
ds2 = Edu21 + 2F du1 du2 + Gdu22 = g11 du21 + 2g12 du1 du2 + g22 du22 ,
where
E = g11 = hxu1 , xu1 i1 ,
F = g12 = g21 = hxu1 , xu2 i1 ,
G = g22 = hxu2 , xu2 i1 .
The second fundamental form associated to the normal vector field L,
which we refer to as the lightcone second fundamental form, is given by
Spacelike surfaces in the Minkowski space-time 289
IIp (u) = hW (L)p (u), ui, with u 2 Tp M . The coefficients of the lightcone
second fundamental form are given by
l = h11 = hLu1 , xu1 i1 = hL, xu1 u1 i1 ,
m = h12 = h21 = hLu1 , xu2 i1 = hLu2 , xu1 i1 = hL, xu1 u2 i1 ,
n = h22 = hLu2 , xu2 i1 = hL, xu2 u2 i1 .
Proof. The vectors xu1 , xu2 , nT , nS form a basis of R41 , so there exist real
numbers , µ such that
L u1 = (a11 xu1 + a21 xu2 ) + nT + µnS .
Since L = nT ± nS , we have hLu1 , nT i1 = . On the other hand
Lu1 = nu1 ±nu1 , so hLu1 , n i1 = hnu1 ±nu1 , n i1 . But hnT , nT i1 = 1 im-
T S T T S T
It follows that
The formula for Lu2 follows similarly. It follows now from the expres-
sions of Lu1 and Lu2 that
gives
✓ ◆ ✓ ◆✓ ◆ 1
a11 a21 h11 h12 g11 g12
Atp = =
a12 a22 h12 h22 g12 g22
so that
✓ ◆ 1 ✓ ◆ ✓ ◆✓ ◆
g11 g12 h11 h12 1 g22 g12 h11 h12
Ap = = 2 .
g12 g22 h12 h22 g11 g22 g12 g12 g11 h12 h22
⇤
(i) The shape operators W (L)p , W (nT )p and W (nS )p are self-adjoint op-
erators on Tp M .
(ii) The L-lightcone Gauss-Kronecker curvature is given by
L depends on the choice of the unit timelike normal vector field nT (see
Definition 9.1). However, the normal lightlike vector field L e does not depend
on the choice of nT . Indeed, given any two unit timelike normal vector fields
nT1 and nT2 , we have ⇡SL (nT1 ± nS1 ) = ⇡SL (nT2 ± nS2 ). We can thus introduce
the following map which is independent of the choice of nT .
Definition 9.3.
Let M be a spacelike surface patch parametrised by x : U ! R41 . The
normalised lightcone Gauss map of M is the map Le : U ! S 2 given by
+
e
L(u) = ⇡SL (L(u)).
K e p ) and H
e ` (p) = det(W (L) e ` (p) = 1 Trace(W (L)e p ).
2
A point p is said to be a lightcone parabolic point if K e ` (p) = 0.
Let Aep denote the matrix of the normalised lightcone shape operator
with respect to the basis {xu1 , xu2 } of Tp M at p = x(u). We write
L(u) = (`0 (u), `1 (u), `2 (u), `3 (u)).
Aep = 1 Ap .
`0 (u)
292 Di↵erential Geometry from a Singularity Theory Viewpoint
Proposition 9.4. For any future directed normal frame (nT , nS ) defining
L, Ke ` (p) = 0 if and only if K` (p) = 0, and H
e ` (p) = 0 if and only if
H` (p) = 0.
Remark 9.4. Observe that the eigenvectors of W (L)p coincide with those
e p . Therefore, the L-principal directions and L-umbilic points are
of W (L)
e
concepts that depend only on L.
Thus,
H(u, v) = hx(u), vi1 .
For v fixed, we write hv (u) = H(u, v) for the lightcone height function
on M along v.
We have the following characterisation of the lightlike parabolic points
and lightlike flat points in terms of the lightcone height function.
2
Proposition 9.6. Let p0 = x(u0 ) be a point on M and v0 a vector in S+ .
e 0 ).
(i) The function hv0 is singular at u0 if and only if v0 = L(u
e 0 ). Then,
(ii) Suppose that v0 = L(u
(a) The point p0 is a lightcone parabolic point if and only if
det Hess(hv0 )(u0 ) = 0.
(b) The point p0 is a lightcone flat point if and only if
rank Hess(hv0 )(u0 ) = 0.
Proof. (i) Since {xu1 (u), xu2 (u), nT (u), nS (u)} is a basis of Tp R41 for all
p = x(u) on M , there exist real numbers ⇠1 , ⇠2 , 1 , 2 such that
v0 = 1 xu1 (u0 ) + 2 xu2 (u0 ) + µ1 nT (u0 ) + µ2 nS (u0 ).
294 Di↵erential Geometry from a Singularity Theory Viewpoint
We have
@hv0
h , v0 i1 = hxu1 , v0 i1 = 1E + 2 F,
@u1
@hv0
h , v0 i1 = hxu2 , v0 i1 = 1F + 2 G.
@u2
Therefore @hv0 /@u1 (u0 ) = @hv0 /@u2 (u0 ) = 0 if and only if 1 = 2 = 0
(EG F 2 > 0 as M is spacelike). Thus, when hv0 is singular at u0 , we
have v0 = µ1 nT (u0 ) + µ2 nS (u0 ). Because v0 lightlike, µ21 + µ22 = 0. This
means that v0 = µ1 L(u0 ), and as v0 is in S+ 2 e 0 ).
, we have v0 = L(u
(ii)-(a) We have
!
e e
hxu1 u1 (u0 ), L (u0 )i1 hxu1 u2 (u0 ), L (u0 )i1
± ±
Hess(hv0 )(u0 ) = e ± (u0 )i1 hxu u (u0 ), L
e ± (u0 )i1
hxu1 u2 (u0 ), L
✓ ◆ 2 2
1 h11 (u0 ) h12 (u0 )
= `0 (u ) .
0 h12 (u0 ) h22 (u0 )
It follows from Propositions 9.2 and 9.3 that
✓ ◆
ep (u0 ) g 11 (u 0 ) g 12 (u 0 )
Hess(hv0 )(u0 ) = A ,
g12 (u0 ) g22 (u0 )
so
ep (u0 )) = det(Hess(hv0 )(u0 )) .
e ` (u0 ) = det(A
K
g11 g22 g122
Proof. Suppose that M is totally L-flat. This means that the matrix Ap
in Proposition 9.2 is identically zero for any future directed frame (nT , nS ).
Therefore, by Proposition 9.2, have
Lui (u) = ±hnS (u), nTui (u)i1 L(u)
at all points u 2 U and for i = 1, 2. Then
e ± e ± `±
2 (u) `± (u)
L(H)(u, L (u)) = (L (u), x2 (u) ⌥ x1 (u) ± , x3 (u) ⌥ x1 (u) 3± ),
`1 (u) `1 (u)
where L(u) = L± (u) = (`± ± ± ± ⇤
0 (u), `1 (u), `2 (u), `3 (u)) 2 LC . The following
result follows from Proposition 9.8.
Theorem 9.1. Let M and M̄ as above and suppose that h1,v and h2,v̄
satisfy the Milnor condition. Then the following assertions are equivalent.
Proof. The statements (i) and (ii) are equivalent by Proposition 4.1.
Since the germs L(H1 ) and L(H2 ) are Lagrangian stable, the germs H1
and H2 are R+ -versal unfoldings of h1,v and h2,v̄ , respectively. Then (ii)
is equivalent to (iii) by the uniqueness of R+ -versal unfoldings.
By Theorem 5.4, (iii) is equivalent to (iv) and it also follows from that
theorem that h1,v and h2,v̄ satisfy the Milnor condition. Then we can apply
Proposition 5.1 to show that (ii) is equivalent to (v). ⇤
Corollary 9.7. Let Embs (U, R41 ) be the set of spacelike embeddings U !
R41 endowed with the Whitney C 1 -topology. Then there exists a residual
subset O ⇢ Embs (U, R41 ) such that for any x 2 O, the following properties
hold.
Proof. We remark that Embs (U, R41 ) is an open subset of C 1 (U, R41 ).
The proof is similar to those of Theorems 4.7 and 6.1. We use the lightcone
300 Di↵erential Geometry from a Singularity Theory Viewpoint
height function H : U ⇥ S+ 2
! R here instead of the height function in
Theorem 4.7. For a generic spacelike embedding x : U ! R41 , the R+ -
singularities of hv are type Ak , k = 1, 2, 3 and H is an R+ -versal unfolding
2
of hv at any point (u, v) 2 U ⇥ S+ (Theorems 3.13 and 4.4). In this case
the catastrophe map-germ ⇡CH is A-equivalent to the fold singularity or
the cusp singularity by the calculation of §3.9.2. Moreover, the singular set
of the catastrophe map is a regular curve. Since the catastrophe map of H
is identified with the normalised lightcone Gauss map L, e the singular set of
the catastrophe map is the lightcone parabolic curve. This completes the
proof. ⇤
Remark 9.5. Let O be the set given in Corollary 9.7 endowed with the
Whitney C 1 -topology. For any x 2 O, we can show that the corresponding
Lagrangian submanifold L(H) is stable under the perturbation of x, and
it follows from this that the set O is open. The full details of the proof of
this observation requires the use of more tools from singularity theory and
is omitted.
@ H̄ @ H̄
(i) H̄(u0 , v) = (u0 , v) = (u0 , v) = 0 if and only if v =
@u1 @u2
e 0 )i1 L(u
hx(u0 ), L(u e 0 ).
(ii) Suppose that v0 = hx(u0 ), L(u e 0 )i1 L(u
e 0 ). Then
(v, r) = rv.
e = H̄
and denote by H 2
: U ⇥ (S+ ⇥ (R \ {0})) ! R the composite map
given by
e
H(u, v, r) = hx(u), vi1 r.
where x(u) = (x0 (u), x1 (u), x2 , x3 (u)). We show that the mapping
e e e e
⇤ e @ H , @ H = H,
e = H,
H e @H , @H
@u1 @u2 @u1 @u2
302 Di↵erential Geometry from a Singularity Theory Viewpoint
is not singular at any point in ⌃He = ( ⇤ H)e 1 (0). The Jacobian matrix of
⇤ e
H is given by
0 v2 v3 1
hxu1 , vi1 hxu2 , vi1 1 x1 + x2 x1 + x3
B v1 v1 C
B v2 v3 C
B hx C
B u1 u1 , vi1 hxu1 u2 , vi1 0 x1u1 + x2u1 x1u1 + x3u1 C .
B v1 v1 C
@ v2 v3 A
hxu2 u1 , vi1 hxu2 u2 , vi1 0 x1u2 + x2u2 x1u2 + x3u2
v1 v1
Consider the matrix
0 v2 v3 1
x1u1
+ x2u1 x1u1 + x3u1
B v1 v1 C
A=@ v2 v3 A
x1u2 + x2u2 x1u2 + x3u3
v1 v1
which is the same matrix in the proof of Proposition 9.8, so detA 6= 0 at
e 1 (0). Therefore rank ⇤ H
any point of ⌃He = ( ⇤ H) e = 3 on ⌃ e . ⇤
H
(i) The lightcone pedal germs LPM and LPM̄ are A-equivalent.
(ii) H̄1 and H̄2 are P -K-equivalent.
(iii) h̄1,v and h̄2,v̄ are K-equivalent.
(iv) K(M, T HP (x, u), v) = K(M̄ , T HP (x̄, ū), v̄).
(v) Q4 (x, u) and Q4 (x̄, ū) are isomorphic as R-algebras.
(i) The lightcone pedal surface is a cuspidal edge at each point of the light-
like parabolic curve except possibly at isolated points. At such points it
304 Di↵erential Geometry from a Singularity Theory Viewpoint
is a swallowtail surface.
(ii) A lightcone parabolic point is a fold of the normalised lightcone Gauss
map L e if and only if it is a cuspidal edge point of the lightcone pedal
map LPM .
(iii) A lightcone parabolic point is a cusp of the normalised lightcone Gauss
map L e if and only if it is a swallowtail point of the lightcone pedal map
LPM .
(A1 ) F (q, x1 , x2 ) = ±q 2 ,
(A2 ) F (q, x1 , x2 ) = q 3 + x1 q,
(A3 ) F (q, x1 , x2 ) = ±q 4 + x1 q + x2 q 2 .
The Euclidean space R30 and any spacelike hyperplanes in R41 are Rieman-
nian manifolds so any surface embedded in these spaces is a spacelike sur-
face. We illustrate below some properties concerning the lightcone Gauss
map of surfaces in these spaces. We also deal with the case of spacelike
surfaces contained in de Sitter 3-space and treat with more details the case
of surfaces in the hyperbolic 3-space.
Spacelike surfaces in the Minkowski space-time 305
L± (u) = nT (u) ± e3 ,
e ± )p =
W (L dnT (u).
Remark 9.6. The normalised lightcone Gauss map is the same as the
hyperbolic Gauss map defined [Epstein (1986); Bryant (1987); Kobayashi
(1986)] in the Poincaré disk model of the hyperbolic space.
3
Remark 9.7. The intersection of lightlike hyperplanes with H+ ( 1) are
the horospheres. Therefore, the family of lightcone height functions on
3
M ⇢ H+ ( 1) measures the contact of M with horospheres.
e + or L
Proof. It follows from Proposition 9.7 that L e is a constant light-
3
like vector v if and only if M ⇢ H(v, c) \ H+ ( 1), equivalently, M is a
subset of a horosphere. ⇤
For a general spacelike surface M in R41 , the image of the di↵erential
map dLp is not a subset of Tp M . In order to obtain a shape operator, we
308 Di↵erential Geometry from a Singularity Theory Viewpoint
Dv E = 1 x u1 + 2 x u2 + µ1 x + µ2 E
for some real numbers 1 , 2 , µ1 , µ2 .
It follows from the fact that hE, Ei1 = 1 that hDv E, Ei1 = 0, so µ2 = 0.
From the identity hE, xi1 = 0 we get hDv E, xi1 = hE, Dv xi. But E is
a normal vector and Dv x is a tangent vector to M , so hE, Dv xi = 0, which
implies hDv E, xi1 = 0. Therefore µ1 = 0. Consequently,
Dv E = 1 x u1 + 2 x u2
Remark 9.8. Let W (E)p = dEp denote the shape operator associated to
the de Sitter Gauss map and let Hd be the de Sitter mean curvature, so
Hd (u) = 12 Trace(W (E)p ), with p = x(u). Then
H`± (u) = ±Hd (u) 1.
It follows that surfaces in hyperbolic space with Hd ⌘ ±1 correspond
to surfaces with H`± ⌘ 0 which are marginally trapped in R41 .
Proof. We deal with the case for L+ , the case L follows similarly. By
definition, we have L+
ui = xui for i = 1, 2. Therefore, for i, j = 1, 2,
L+
ui uj = uj xui + xui uj .
Since L+
ui u j = L+
uj ui and xui uj = xuj ui , we have
u j x u i ui xuj = 0.
G(p, a) = hp a, p ai1 .
The fibres of the map G are the pseudo spheres in R41 of centre a.
310 Di↵erential Geometry from a Singularity Theory Viewpoint
so
✓ ◆ ✓ ◆
1 h11 h11 g11 g12
Hess(g)(u0 ) = 2 +2
`0 h12 h22 g12 g22
where all the entries of the right hand side of the above equality are evalu-
ated at u0 and `0 is the first coordinate of L. Therefore,
✓ ◆ 1
g11 g12 ep + I)
Hess(g)(u0 ) = 2( A 0
g12 g22
with I denoting the 2 ⇥ 2 identity matrix. Thus, det(Hess(ga0 ))(u0 ) = 0
if and only if 1/ is an eigenvalue of A ep , that is, = 1/e 1 (u0 ) or
0
= 1/e2 (u0 ). ⇤
Spacelike surfaces in the Minkowski space-time 311
for i = 1 or i = 2.
We define
2 ⇢
[ 1 e
FM = x(u) + L(u) | i (u) 6= 0, u 2 U ,
i=1
ei (u)
and call it the lightlike focal set of M.
From Theorem 9.3 and from the classification of stable Legendrian map-
pings (Theorem 5.12), we obtain the following classification of germs of
generic lightlike hypersurface.
Proof. By Theorems 5.11 and 9.3, the Lorentzian distance squared func-
tion G is a K-versal deformation of ga0 at each point (u, a0 ) 2 U ⇥ R.
Therefore, we can apply the classification of Legendrian stable map-germs
(Theorem 5.12). Then the generating family is P -K-equivalent to one for
the following normal forms:
(Ak ) F (q1 , q2 , x) = xk+1 ± q22 + x1 + x2 q1 + · · · + xk q1k 1
, 1 k 4,
(D4+ ) F (q1 , q2 , x) = q13 + q23 + x1 + x2 q1 + x3 q2 + x4 q1 q2 ,
(D4 ) F (q1 , q2 , x) = q13 q1 q22 + x1 + x2 q1 + x3 q2 + x4 (q12 + q22 )
with x = (x1 , x2 , x3 , x4 ). We consider the D4+ -case and take F as above
(the other cases follow by similar calculations). Then ⌃⇤ (F ) is given by
{(q1 , q2 , 2(q12 +q22 )+x4 q1 q2 , 3q12 x4 q2 , 3q22 x4 q1 , x4 ) | (q1 , q2 , x4 ) 2 R3 }.
If we change the parameters into (q1 , q2 , x4 ) = (x, y, z) and apply the
linear transformation (X, Y, Z, W ) = (X, Y, Z, W ), the corresponding
Legendrian map-germ is
f (x, y, z) = (2(x2 + y 2 ) + xyz, 3x2 + yz, 3y 2 + xz, z). ⇤
Spacelike surfaces in the Minkowski space-time 313
Proof. In the first place, we remark that K e ` (u) 6= 0 if and only if the
e
three vectors L(u), e u (u), L
L e u (u) are linearly independent.
1 2
E2
Qr (x, u) = ⇤ (M )E + Mr+1
.
g1,a 1 2 2
Proof. The discussion proceeding the proposition shows that (iii) and
(iv) are equivalent. The equivalence of the other statements follow by
Proposition 5.2. ⇤
Spacelike surfaces in the Minkowski space-time 315
Remark 9.9. Observe that ✓11 1 (0) and ✓12 1 (0) define the same field of
± 1 ± 1
tangent hyperplanes over 1 , denoted by K1 . Also ✓i1 (0) and ✓i2 (0)
±
define the same field of tangent hyperplanes over i , i = 2, 3, 4, denoted
by Ki± .
±
Consider, for example, the map 12 . We have
± ⇤ ±
( 12 ) ✓21 = hdv, ⌥v + wi1 | 1
F O X1
± ±
± 41 14
±
12 13
±
21
✏ ±
± 31
± 4 ±
42 : d 43
⌃ z ± ± ±
$ ⇠
24 34
± 23
/ ±
2 o 3
±
32
surfaces to surfaces in the lightcone. We construct here the basic tools for
the study of the extrinsic di↵erential geometry of spacelike surfaces in the
lightcone LC ⇤ as an application of the mandala of Legendrian dualities.
Let M be a spacelike surface patch in the lightcone parametrised by
x : U ! LC ⇤ . We shall show the existence and uniqueness of the lightcone
normal vector to M as a consequence of Theorem 9.5.
Consider the double Legendrian fibration ⇡41 : 4 ! LC ⇤ , ⇡42 : 4 !
LC ⇤ and let v 2 LC ⇤ , where we set 4 = 4 , ⇡41 = ⇡41 , ⇡42 = ⇡42 . The
fibre of ⇡411 (v) can be identified with
{w 2 LC ⇤ | hv, wi1 = 2},
which is the intersection of LC ⇤ with a lightlike hyperplane, so it is a two
dimensional spacelike submanifold. For any p = x(u) 2 M, the normal
space Np M is a timelike plane, so there exists two lightlike lines on Np M.
One of the lines is generated by p = x(u). We remark that a lightlike
plane consists of lightlike vectors and spacelike vectors only. Moreover,
all lightlike vectors are linearly dependent. Therefore, if another lightlike
line is generated by w, then we have hw, x(u)i1 = c 6= 0. If necessary,
we consider w e = cw/2. Then we have hw, e x(u)i1 = 2. Therefore, the
1
intersection of ⇡41 (v) with the normal plane of M (a timelike plane) in R41
consists of only one point at each point on M . Since ⇡41 : 4 ! LC ⇤ is
a Legendrian fibration, there is a Legendrian submanifold parametrised by
L4 : U ! 4 such that ⇡41 L4 (u) = x(u). It follows that we have a smooth
map x` : U ! LC ⇤ such that L4 (u) = (x(u), x` (u)), i.e., ⇡42 L4 = x` .
Since L4 is a Legendrian embedding, we have hdx(u), x` (u)i1 = 0, so x` (u)
belongs to the normal plane of M at x(u).
Given another Legendrian embedding L14 (u) = (x(u), x`1 (u)), we have
that x` (u) and x`1 (u) are parallel. However, the relation hx(u), x` (u)i1 =
hx(u), x`1 (u)i1 = 2 holds, so x` (u) = x`1 (u). This means that L4 is the
unique Legendrian lift of x.
Definition 9.6. We call the vector x` (u) = ⇡42 L4 (u) the lightcone normal
vector to M at p = x(u). The map x` : U ! LC ⇤ (or, its image) is called
the lightcone dual of M. The map x e ` : U ! S+2 e` (u) = ⇡SL (x` (u)),
, with x
is called the lightcone Gauss map of M .
ordinary Euclidean unit normal N (u) of r(U ) = ⇡(M ) and the Euclidean
Gauss map N : U ! S 2 ⇢ R30 . We can now define a transversal vector field
r` (u) to M along ⇡(M ) in R30 by
r(u) 2(r(u) · N (u))N (u)
r` (u) = ,
(r(u) · N (u))2
where “·” is the usual Euclidean scalar product. It follows by the uniqueness
of the vector x` that
kr(u)k
x` (u) = ( 2
, r` (u)).
(r(u) · N (u))
We can study the extrinsic di↵erential geometry of M using the normal
vector field x` . Here too, as in the case of surfaces in the hyperbolic space,
the di↵erential of x` at each point is a linear transformation of Tp M . Indeed,
One can obtain similar results to those in the previous sections about the
family H and derive accordingly information about the extrinsic geometry
of M in LC ⇤ .
L1 (u) = 41 L4 (u),
is a Legendrian embedding, where L4 is as in §9.11. If we write L1 (u) =
(xh (u), xd (u)), then
x(u) + x` (u) x(u) x` (u)
xh (u) = , xd (u) = .
2 2
Definition 9.7. We call xh (u) (resp. xd (u)) the hyperbolic normal vector
(resp. de Sitter normal vector) of M at p = x(u).
(E + l)(G + n) (F + m)2
Kd = .
4(EG F 2 )
Proof. We denote by
✓ ◆
a`11 a`12
A`p =
a`21 a`22
the matrix of the shape operator Wp` = dx` (u) with respect to the basis
{xu1 , xu2 } of Tp M . Following similar arguments in the proof of Proposition
9.2, we get a similar formula to (9.2), namely,
✓ ◆ ✓ ◆
EF ` l m
Ap = .
F G m n
Since xh = (x + x` )/2, we have
1
xhu1 = 2 (1 a`11 )xu1 a`21 xu2 ,
1
xhu2 = 2 a`12 xu1 + (1 a`22 )xu2
so the matrix of dxh with respect to the basis {xu1 , xu2 } of Tp M is
✓ ◆ ✓ ◆ ✓ ` ` ◆
1 1 a`11 a`12 1 10 a11 a12
` ` = 2( )
2 a21 1 a22 01 a`21 a`22
✓ ◆ 1 ✓ ◆ ✓ ◆✓ ` ` ◆
1 E F EF EF a11 a12
= 2 ( )
F G F G F G a`21 a`22
✓ ◆ 1 ✓ ◆ ✓ ◆
1 E F EF l m
= 2 ( )
F G F G m n
✓ ◆ 1✓ ◆
E F E l F m
= 12 .
F G F m G n
322 Di↵erential Geometry from a Singularity Theory Viewpoint
The curvature Kh is the determinant of the matrix in the left hand side
of the equality above, and the result follows by taking the determinant of
the right hand side of that equality.
The formula for Kd follows the same steps as above with an appropriate
change of signs as xd = (x x` )/2. ⇤
We denote, as usual, by kij the Christo↵el symbols of M (which is a
Riemannian manifold), where
X2 ⇢
k 1 @gjm @gim @gij
ij = g km + .
2 m=1 @ui @uj @um
Above, (gij ) is the matrix of the first fundamental form and (g km ) is its
inverse matrix. Using the notation E, F, G for the coefficients of the first
fundamental form, the Christo↵el symbols are given by the following six
functions:
1 GEu1 2F Fu1 + F Eu2 2 2EFu1 EEu2 F Eu1
11 = , 11 = ,
2(EG F 2 ) 2(EG F 2 )
1 1 GEu2 F Gu1 2 2 EGu1 F Eu2
12 = 21 = , 12 = 21 = ,
2(EG F 2 ) 2(EG F 2 )
1 2GFu2 GGu1 F Gu2 2 EGu2 2F Fu2 + F Gu1
22 = , 22 = .
2(EG F 2 ) 2(EG F 2 )
Proposition 9.18. Let M be a spacelike surface patch in the lightcone
parametrised by x : U ! LC ⇤ and let l, m, n denote the coefficients of
the second fundamental form associated to x` . Then the lightcone Gauss
equations are given by
1 2 1 `
x u1 u 1 = 11 xu1 + 11 xu2 2 (lx + Ex ),
1 2 1 `
x u1 u 2 = 12 xu1 + 12 xu2 2 (mx + F x ),
1 2 1 `
x u2 u 2 = 22 xu1 + 22 xu2 2 (nx + Gx ).
The expressions for xu1 u2 and xu2 u2 follow similarly. Moreover, we have
1
Eu1 = hxu1 u1 , xu1 i1 = ⇤111 E + ⇤211 F.
2
By the similar arguments to the above, we have
1 1
Eu2 = ⇤112 E + ⇤212 F, Gu1 = ⇤112 F + ⇤212 G,
2 2
1 1
Fu 2 Gu1 = ⇤122 E + ⇤222 F, Gu2 = ⇤122 F + ⇤222 G.
2 2
It follows that
✓ ◆✓ 1 1 1 ◆ ✓ ◆
EF ⇤11 ⇤12 ⇤22 1 Eu1 Eu2 2Fu2 Gu1
= ,
F G ⇤211 ⇤212 ⇤222 2 2Fu1 Eu2 Gu1 Gu 2
Kh Kd = Ks .
324 Di↵erential Geometry from a Singularity Theory Viewpoint
9.12 Notes
Global viewpoint
329
330 Di↵erential Geometry from a Singularity Theory Viewpoint
are of interest. (For example, the umbilic points on a surface in R3 are the
singular points of the integral curves of the principal directions, which are
the contact directions of the distance squared functions on the surface.)
One can use topological arguments, such as the Poincaré-Hopf formula, to
obtain lower bounds for the number of such points.
We outline several applications of the above approaches on surfaces M
in R3 and R4 (§10.1). We also consider the case of spacelike surfaces in
Minkowski space-time R41 (§10.2). We comment in §10.3 on some other
global results on submanifolds of Euclidean and Minkowski spaces.
We emphasise that our aim in this chapter is to give some applications of
singularity theory to the study of global properties of submanifolds, related
mainly to the work of the authors on the subject.
10.1.1 Surfaces in R3
We can view a finite plane graph as a stratified set. The vertices and
edges being respectively the 0- and 1-dimensional strata. Moreover, if this
graph lies on a surface, we can view the whole surface as a 2-dimensional
stratified set whose 2-dimensional strata are the connected components of
the complement of the graph.
Any embedding f of a surface M in R3 induces a stratification of the
2-sphere. This stratification is constructed according to the type of the
absolute minima of the family of height functions on the embedded sur-
face M . Provided that the embedding f is height functions generic, the
multi-transversality conditions on the height functions family imply that we
only have the following possibilities for the absolute minima of the height
functions on the surface M (a detailed definition for the general case of
hypersurfaces in Rn can be found in [Romero Fuster (1983)]):
map (i.e., cusps of the Gauss map lying on the boundary of the convex
hull of the surface). The other vertices of the graph (conflict strata of type
A1 A1 A1 ) are of degree 3 (i.e., they are the end points of exactly 3 edges)
and correspond to the (isolated) tri-tangent support planes of the surface.
The edges of the graph (conflict strata of type A1 A1 ) correspond to the
normal directions to the 1-parameter family of support bi-tangent planes
of the surface.
Applying equality (10.1) to surfaces in R3 leads to the following result.
10.1.2 Wavefronts
For a Legendrian fibration ⇡ : E ! N , where dim N = 3, we consider a
Legendrian immersion i : L ! E. The wavefront W (L) of a generic Legen-
drian immersion has singularitites of type cuspidal edges (A2 ), swallowtails
(A3 ) and points of transversal self-intersection (A1 A1 , A1 A2 , A1 A1 A1 ); see
[Arnol’d (1990)]. (In Theorem 10.2 and Corollary 10.3 bellow, generic
means that the wavefront has only the above local and multi-local singular-
ities.) If L is a closed surface, then the numbers of swallowtails and triple
points are finite and satisfy the following relation.
(1)
After this, we build up the one-skeleton KL of KL so that the im-
(1)
age of D2 (L) is a sub-complex of KL . We complete our procedure by
(2)
constructing the two-skeleton KL . Since ⇡|D2 (L) , ⇡|D2 (L,(2)) , ⇡|D3 (L) are
proper finite to one maps, we can pull back KL to obtain a triangulation
of L, D2 (L), D2 (L, (2)) and D3 (L). Let CjX be the number of j-cells in X,
where X = W (L), L, D2 (L), D2 (L, (2)) or D3 (L). Then equation (10.2) can
be written as
X W (L)
X X D 2 (L)
( 1)j Cj =↵ ( 1)j C Lj + ( 1)j Cj
j j j
X 2
D (L,(2))
X D 3 (L)
+ ( 1)j Cj + ( 1)j Cj ,
j j
where CjX = 0 for i > dim X. Therefore, if we can find real numbers ↵, ,
and such that
W (L) D 2 (L) D 2 (L,(2)) D 3 (L)
Cj = ↵CjL + Cj + Cj + Cj ,
for any j, then we have solutions of equation (10.2). We deal with the case
j = 0. We remark that ⇡|L is three-to-one over the points of ⇡(D3 (L)),
one-to-one over the points of ⇡(D2 (L, (2))), two-to-one over the points of
⇡(D2 (L) \ (D2 (L, (2)) [ D3 (L))), and one-to-one over the points of ⇡(N \
D2 (L)). It follows that equation
W (L) D 2 (L) D 2 (L,(2)) D 3 (L)
C0 = ↵C0L + C0 + C0 + C0
is equivalent to the system of linear equations
0 1 0 10 1
1 1000 ↵
B1C B2 2 0 0C B C
B C=B CB C
@1A @1 1 1 0A @ A .
1 3303
Solving the above linear system gives ↵ = 1, = 1/2, = 1/2 and
= 1/6. Thus, we have
1 1 1
(W (L)) = (L) (D2 (L)) + (D2 (L, (2))) (D3 (L)). (10.3)
2 2 6
We have, by definition, (D2 (L, (2)) = S(L) and (D3 (L)) = 3T (L).
Since D2 (L) is the union of closed curves on the surface L with 3T (L)
crossings, we can triangulate it with (3T (L) + n) 0-cells and (6T (L) + n)
1-cells, where n is the number of circles in D2 (L). We get the desired result
by substituting these in (10.3). ⇤
Global viewpoint 335
10.1.3 Surfaces in R4
Inflection points of a surface M in R4 are the singular points of the asymp-
totic curves on M (see Chapter 7). For a generic surface, the fields of
asymptotic directions have index ± 12 at inflection points of imaginary type.
The inverse of the stereographic projection from R3 to S 3 maps the lines
of principal curvature of a surface in R3 to the asymptotic curves of the
336 Di↵erential Geometry from a Singularity Theory Viewpoint
Theorem 10.3. For a height function generic closed and locally convex
surface M in R4 we have
2| (M )| ]{inflection points}.
Corollary 10.4. Any height function generic closed and locally convex sur-
face in R4 with non-vanishing Euler characteristic has at least 4 inflection
points.
The result in Corollary 10.5 is the generic version of the following con-
jecture.
Carathéodory conjecture: any closed, convex and sufficiently smooth
surface in three dimensional Euclidean space has at least two umbilic points.
A possible approach for solving the conjecture is by using Poincaré-Hopf
formula and investigating the possible values for the index of an umbilic
point. It is known, for instance, how to construct examples of local immer-
sions of surfaces with umbilics of any index 1. This leads to the following
conjecture which can be considered as a local version of the Carathéodory
conjecture.
Loewner conjecture: the index of the principal directions field at an
umbilic point of a sufficiently smooth surface in three dimensional Euclidean
space is at most 1.
A review on the state of Carathéodory’s conjecture can be found in
[Gutierrez and Sotomayor (1998)]. In view of Corollary 10.4, it is reason-
able to propose the following conjecture for surfaces in R4 .
Carathéodory conjecture for surfaces in R4 : every closed, convex
and sufficiently smooth surface in four dimensional Euclidean space which
is homeomorphic to a 2-sphere has at least two inflection points.
The above conjecture is shown to be true in some particular cases in
[Gutiérrez and Ruas (2003); Nuño Ballestero (2006)].
10.1.4 Semiumbilicity
Recall from Definition 7.2 that a point p on M ⇢ R4 is semiumbilic if the
curvature ellipse at p is a line segment that contains p. It can be shown
that the asymptotic directions of M are orthogonal at semiumbilic points.
Also, a point p is semiumbilic if and only if it is a ⌫-umbilic point of some
338 Di↵erential Geometry from a Singularity Theory Viewpoint
(i) p is semiumbilic.
(ii) N (p) = 0.
(iii) p is ⌫-umbilic for some normal field ⌫ on M .
(iv) There are two orthogonal asymptotic directions at p.
where k·k is the Euclidean norm in R30 . Suppose that there exists t0 2 (0, 1)
and p0 2 M such that
t0 N (p0 ) + (1 e 0 ) = 0.
t0 )⇡ L(p
e= 1
deg L (M ).
2
e ⇤ dvS 2 ,
e ` dvM = L
K +
eu , L
geij = hL e u i1
i j
1
= 2 hLui , Luj i1
`0
X2 2
X
1
= 2h h↵
i x u↵ , hi x u i 1
`0 ↵=1 =1
2
1 X ↵
= 2 hi hj hxu↵ , xu i1
`0
↵, =1
2
1 X ↵
= 2 hi hj g ↵ .
`0
↵, =1
e
for v = L(p). e hv has a Morse-type singular
Since v is a regular value of L,
point of index 0 or 2 at the minimum and maximum points. The light-
cone Gauss-Kronecker curvarture K e +
e ` is positive at such points, so L|M
is surjective. Since the area of the unit sphere is 4⇡, we have the desired
inequality by Proposition 10.1. ⇤
10.3 Notes
Curves. We have not touched on the subject of plane curves in this book.
However, there are some global properties of plane and space curves that
are worth mentioning here. We start with plane curves and the celebrated
4-Vertex Theorem which states that any smooth closed simple curve in
Euclidean plane has at least 4 vertices. An analogous result for curves in the
Minkowski plane R21 is proved in [Tabachnikov (1997)]. Using stereographic
projections H+ 2
( 1) ! R2 and S12 ! R2 , one can show that an analogous
2
result of the 4-vertex Theorem is true for curves in H+ ( 1) and in S12 .
Results of global nature on space curves obtained by similar methods to
those explored in this book concerning the number of tri-tangent planes, bi-
tangent osculating planes and torsion zero points can be found in [Bancho↵,
Ga↵ney and McCrory (1985); Freedman (1980); Nuño Ballesteros and
Romero Fuster (1992); Sedykh (1989); Ozawa (1985)].
It is worth observing that the case of spacelike embeddings of a circle
(i.e., spacelike knots) in R31 is di↵erent from ordinary knots R3 . There are,
for instance, many cases of closed spacelike curves which are un-knotted in
the ordinary sense ([Izumiya, Kikuchi and Takahashi (2006)]).
Umbilics on surfaces in R31 and H+ 3
( 1). It is shown in [Tari (2013)]
that any closed and convex surface in Minkowski 3-space R31 of class C 3
has at least two umbilic points. For ovaloids (i.e., strictly convex surfaces)
of class C 3 , the umbilic points all lie in the Riemannian part of the surface
and there are at least two of them.
Using the stereographic projection and Feldman result for generic sur-
faces in R3 ([Feldman (1967)]), it can be shown that the number of light-
344 Di↵erential Geometry from a Singularity Theory Viewpoint
(1996, 2004); Lippner and Szucs (2010); Szabó, Szucs and Terpai, (2010)].
Other approaches to global problems using Vassiliev type invariants and
h-principal can be found, for example, in [Ando (1985, 2007a,b); Eliashberg
and Mishachev (2002); du Plessis (1976a,b); Gromov (1986); Goryunov
(1998); Yamamoto (2006); Ohmoto and Aicardi (2006)].
Bibliography
347
348 Di↵erential Geometry from a Singularity Theory Viewpoint
Bancho↵, T., Ga↵ney, T. and McCrory, C. (1982), Cusps of Gauss mappings. Re-
search Notes in Mathematics 55. Pitman (Advanced Publishing Program),
Boston, Mass.-London.
Bancho↵, T., Ga↵ney, T. and McCrory, C. (1985), Counting tritangent planes of
space curves. Topology 24, pp. 15–23.
Bancho↵, T. and Thom, R. (1980), Erratum et compléments: “Sur les points
paraboliques des surfaces” by Y. L. Kergosien and Thom. C. R. Acad. Sci.
Paris Sér. A-B 291, A503–A505.
Basto-Gonçalves, J. (2013), Local geometry of surfaces in R4 . Preprint, arXiv :
1304.2242.
Binotto, R. (2008), Projetivos de curvatura. Doctoral Thesis, Instituto de
Matemática, Estatı́stica e Computação Cientı́fica, UNICAMP (Brasil).
Binotto, R, Costa, S. I. R. and Romero Fuster, M. C. (2015), Geometry of 3-
manifolds in Euclidean space. To appear in RIMS Kôkyûroku Bessatsu.
Birbrair, L. (2007), Metric theory of singularities. Lipschitz geometry of singular
spaces. Singularities in geometry and topology, pp. 223–233. (World Sci.
Publ., Hackensack, NJ).
Blair, D. E. (1976), Contact manifolds in Riemannian geometry, Lecture Notes
in Mathematics, 509, Springer-Verlag, Berlin.
Bleeker, D. and Wilson, L. (1978), Stablility of Gauss maps, Illinois J. Math. 22,
pp. 279–289.
Brieskorn, E. and Knörrer, H (1986), Plane algebraic curves. Birkhäuser.
Bröcker, Th. (1975), Di↵erentiable germs and catastrophes. London Mathematical
Societry Lecture Notes 17 (Cambridge University Press).
Bruce, J. W. (1981), The duals of generic hypersurfaces; Math. Scand. 49, pp.
36–60.
Bruce, J. W. (1984), Generic reflections and projections; Math. Scand. 54, pp.
262–278.
Bruce, J. W. (1986), On transversality. Proc. Edinburgh Math. Soc. 29, pp. 115-
123.
Bruce, J. W. (1994a), Generic geometry, transversality and projections. J. London
Math. Soc. 49, pp. 183-194
Bruce, J. W. (1994b), Lines, circles, focal and symmetry sets. Math. Proc. Cam-
bridge Philos. Soc. 118, pp. 411–436.
Bruce, J. W. and Fidal, D. (1989), On binary di↵erential equations and umbilics.
Proc. Royal Soc. Edinburgh,111A, pp. 147–168.
Bruce, J. W., Fletcher, G. J. and Tari, F. (2004), Zero curves of families of curve
congruences. Contemp. Math.,354, Amer. Math. Soc. Amer. Providence,
RI, pp. 1–18.
Bruce, J. W. and Giblin, P. J. (1990), Projections of surfaces with boundary.
Proc. London Math. Soc. 60, pp. 392–416.
Bruce, J. W. and Giblin, P. J. (1992), Curves and Singularities. Cambridge Uni-
versity Press.
Bruce, J. W., Giblin, P. J. and Gibson, C. G. (1985), Symmetry sets. Proc. Roy.
Soc. Edinburgh Sect. A 101, pp. 163–186.
Bruce, J. W., Giblin, P. J. and Tari, F. (1995), Families of surfaces: height func-
Bibliography 349
tions, Gauss maps and duals. Real and complex singularities (São Carlos,
1994). Pitman Res. Notes Math. Ser. (Longman, Harlow) 333, pp. 148–178.
Bruce, J. W., Giblin, P. J. and Tari, F. (1998), Families of surfaces: height
functions and projections to plane. Math. Scand. 82, pp. 165–185.
Bruce, J. W., Giblin, P. J. and Tari, F. (1999), Families of surfaces: focal sets,
ridges and umbilics. Math. Proc. Camb. Phil. Soc 125, pp. 243–268.
Bruce, J. W., Kirk N. P. and du Plessis A. A. (1997), Complete transversals and
the classification of singularities. Nonlinearity 10, pp. 253-275.
Bruce, J. W. and Nogueira, A. C. (1998), Surfaces in R4 and duality. Quart. J.
Math. Oxford Ser. 49, pp. 433–443.
Bruce, J. W., du Plessis, A. A. and Wall, C. T. C. (1987), Determinacy and
unipotency. Invent. Math. 88, pp. 521–554.
Bruce, J. W. and Romero Fuster, M. C. (1991), Duality and orthogonal projec-
tions of curves and surfaces in Euclidean 3-space. Quart. J. Math. 42, pp.
433–441.
Bruce, J. W. and Tari, F. (1995), On binary di↵erential equations. Nonlinearity
8, pp. 255–271.
Bruce, J. W. and Tari, F. (2000), Duality and implicit di↵erential equations.
Nonlinearity 13, pp. 791–811.
Bruce, J. W. and Tari, F. (2002), Families of surfaces in R4 . Proc. Edinb. Math.
Soc. 45, pp. 181–203.
Bruce, J. W. and Tari, F. (2005), Dupin indicatrices and families of curve con-
gruences. Trans. Amer. Math. Soc. 357, pp. 267–285.
Bruce, J. W. and West, J. M. (1998), Functions on a crosscap. Math. Proc.
Cambridge Philos. Soc. 123, pp. 19–39.
Bruce, J. W. and Wilkinson, T. C. (1991), Folding maps and focal sets. Pro-
ceedings of Warwick Symposium on Singularities, Springer Lecture Notes
in Math., Vol. 1462, pp. 63–72, Springer-Verlag, Berlin and New York.
Bryant, R. L. (1987), Surfaces of mean curvature one in hyperbolic space. in
Théorie des variétés minimales et applications (Palaiseau, 1983–1984),
Astérisque No. 154–155, 12, pp. 321–347, 353 (1988).
Buchner, M.A. (1974), Stability of the cut locus. Doctoral thesis, Harvard Uni-
versity.
Buosi, M., Izumiya, S. and Ruas, M. A. S. (2010), Total absolute horospherical
curvature of submanifolds in hyperbolic space. Adv. Geom. 10, pp. 603–620.
Buosi, M., Izumiya, S. and Ruas, M. A. S. (2011), Horo-tight spheres in hyperbolic
space. Geom. Dedicata 154, pp. 9–26.
do Carmo, M. P. (1976), Di↵erential geometry of curves and surfaces. Prentice-
Hall, 1976.
do Carmo, M. P. (1992), Riemannian geometry. Translated from the second Por-
tuguese edition by Francis Flaherty Mathematics: Theory & Applications.
Birkhuser Boston, Inc., Boston, MA.
Cecil, Th. E. (1974), A characterization of metric spheres in hyperbolic space by
Morse theory. Tohoku Math. J. 26, pp. 341–351.
Cecil, Th. E. (1980), Lie Sphere Geometry with Applications to Submanifolds.
Universitext, Springer-Verlag, New York Berlin.
350 Di↵erential Geometry from a Singularity Theory Viewpoint
Cecil, Th. E. and Ryan, P. J. (1979a), Distance functions and umbilic submani-
folds of hyperbolic space. Nagoya Math. J. 74, pp. 67–75.
Cecil, Th. E. and Ryan, P. J. (1979b), Tight ant taut immersions into hyperbolic
space. J. Lond. Math. Soc. 19, pp. 561–572.
Cecil, Th. E. and Ryan, P. J. (1985), Tight and taut immersions of manifolds.
Research Notes in Mathematics, 107. Pitman (Advanced Publishing Pro-
gram), Boston, MA.
Chen, L. and Izumiya, S. (2009), A mandala of Legendrian dualities for pseudo-
spheres in semi-Euclidean space Proc. Japan Acad. 85, pp. 49–54.
Cheng, S. -Y. and Yau, S.-T. (1973), Maximal space-like hypersurfaces in the
Lonrentz-Minkowski spaces. Ann. of Math. 104, pp. 407–419.
Chern, S. and Lashof, R. K. (1957), On the total curvature of immersed manifolds.
Amer. J. Math. 79, pp. 306–318.
Cleave, J. P. (1980), The form of the tangent developable at points of zero torsion
on space curves. Math. Proc. Camb. Phil. 88, pp. 403–407.
Costa, S. I. R. (1990), On closed twisted curves. Proc. Amer. Math. Soc. 109,
pp. 205–214.
Costa, S. I. R. and Romero Fuster, M. C. (1997), Nowhere vanishing torsion
closed curves always hide twice. Geom. Dedicata 66, pp. 1–17.
Costa, S. I. R., Moraes, S. M. and Romero Fuster, M. C. (2009), Geometric
contacts of surfaces immersed in Rn , n 5. Di↵erential Geom. Appl. 27,
pp. 442–454.
Damon J. N. (1984), The unfolding and determinacy theorems for subgroups of
A and K. Mem. Amer. Math. Soc. 50, No. 306.
Damon J. N. (1988), Topological triviality and versality for subgroups of A and
K. Mem. Amer. Math. Soc. 75, No. 389.
Damon J. N. (1992), Topological triviality and versality for subgroups of A and
K. II. Sufficient conditions and applications. Nonlinearity, 5, pp. 373–412.
Damon J. N. (2003), Smoothness and geometry of boundaries associated to skele-
tal structures. I. Sufficient conditions for smoothness. Ann. Inst. Fourier
(Grenoble) 53, pp. 1941–1985.
Damon J. N. (2004), Smoothness and geometry of boundaries associated to skele-
tal structures. II. Geometry in the Blum case. Compos. Math. 140, pp.
1657–1674.
Damon J. N. (2006), The global medial structure of regions in R3 . Geom. Topol.
10, pp. 2385–2429.
Dara, L. (1975), Singularités génériques des équations di↵erentielles multiformes.
Bol. Soc. Brasil Math. 6, pp. 95–128.
David, J. M. S. (1983), Projection-generic curves. J. London Math. Soc. 27, pp.
552–562.
Davydov, A. A. (1994), Qualitative control theory. Translations of Mathematical
Monographs, 142. AMS, Providence, RI.
Dias, F. S. and Nuño Ballesteros, J. J. (2008), Plane curve diagrams and geomet-
rical applications. Q. J. Math. 59, pp. 287–310.
Dreibelbis, D. (2001), Bitangencies on surfaces in four dimensions. Quart. J.
Math. 52, pp. 137–160
Bibliography 351
Publ., Polish Acad. Sci. Inst. Math., Warsaw 50, pp. 91–105.
Giblin, P. J. and Janeczko, S. (2012), Geometry of curves and surfaces through
the contact map. Topology Appl. 159, pp. 466–475.
Giblin, P. J. and Weiss, R. (1987), Reconstruction of surfaces from profiles. Proc.
First Internat. Conf. on Computer Vision; Computer Society of the IEEE,
pp. 136–144.
Giblin, P. J. and Zakalyukin, V. M. (2005), Singularities of centre symmetry sets.
Proc. London Math. Soc. 90, pp. 132–166.
Gibson, C. G. (1979), Singular points of smooth mappings. Research Notes in
Mathematics, 25. Pitman (Advanced Publishing Program), Boston, Mass.-
London.
Gibson, C. G., Hawes, W. and Hobbs, C. A. (1994), Local pictures for general
two-parameter planar motions. Advances in robot kinematics and compu-
tational geometry (Ljubljana), Kluwer Acad. Publ., Dordrecht, pp. 49–58.
Gibson, C. G., Wirthmüller, K., du Plessis, A. A. and Looijenga, E. J. N. (1976),
Topological stability of smooth mappings. Lecture Notes in Mathematics,
552. Springer-Verlag, Berlin-New York,
Golubitsky, M. and Guillemin, V. (1973), Stable mappings and their singularities
GTM, 14. Springer-Verlag, New York.
Golubitsky, M. and Guillemin, V. (1975), Contact equivalence for Lagrangian
manifold, Adv. Math. 15, pp. 375–387.
Golubitsky, M. and Schae↵er, D. G. (1985), Singularities and groups in bifurcation
theory. Vol. I. Applied Mathematical Sciences, 51. Springer-Verlag, New
York.
Gonçalves, R. A., Martı́nez Alfaro, J. A., Montesinos Amilibia, A. and Romero
Fuster, M. C. (2007), Relative mean curvature configurations for surfaces
in Rn , n 5. Bull. Braz. Math. Soc. New Series 38(2), pp. 1–22.
Goryunov, V. V. (1981a), Surface projection singularities. Ph. D. thesis, Moscow’s
Lomonosov State University.
Goryunov, V. V. (1981b), Geometry of bifurcation diagrams of simple projections
on a line. Funktsional. Anal. i Prilozhen 15, pp. 1–8, 96.
Goryunov, V. V. (1990), Projections of Generic Surfaces with Boundaries. Adv.
Soviet Math. 1, pp. 157–200.
Goryunov, V. V. (1998), Vassiliev type invariants in Arnold’s J + -theory of plane
curves without direct self-tangencies. Topology 37, pp. 603-620
Gromov, M. (1986), Partial di↵erential relations. Ergebnisse der Mathematik und
ihrer Grenzgebiete (3) [Results in Mathematics and Related Areas (3)], 9.
Springer-Verlag, Berlin.
Guillemin, V. and Pollack, A. (1974), Di↵erential topology. Prentice-Hall, Inc.,
Englewood Cli↵s, N.J.
Gutierrez, C., Guadalupe, I., Tribuzy, R. and Guı́ñez V. (1997), Lines of curvature
on surfaces immersed in R4 . Bol. Soc. Brasil. Mat. (N.S.) 28, pp. 233–251.
Gutierrez, C., Guadalupe, I., Tribuzy, R. and Guı́ñez, V. (2001), A di↵erential
equation for lines of curvature on surfaces immersed in R4 . Qual. Theory
Dyn. Syst., pp. 207–220.
Gutierrez, C. and Guı́ñez, V. (2003), Simple umbilic points on surfaces immersed
Bibliography 353
University of Southampton.
Romero Fuster, M. C. (1983), Sphere stratifications and the Gauss map. Proc.
Edinburgh Math. Soc. 95, pp. 115–136.
Romero Fuster, M. C. (1988), Convexly generic curves in R3 . Geom. Dedicata 28,
pp. 7–29.
Romero Fuster, M. C. (1997), Stereographic projections and geometric singulari-
ties. Workshop on Real and Complex Singularities (São Carlos, 1996). Mat.
Contemp. 12, pp. 167–182.
Romero Fuster, M.C. (2004), Semiumbilics and geometrical dynamics on surfaces
in 4-spaces. Real and complex singularities, pp. 259–276, Contemp. Math.,
354, Amer. Math. Soc., Providence, RI.
Romero Fuster, M. C. (2007), Geometric contacts and 2-regularity of surfaces in
Euclidean space. Singularity theory, pp. 307–325, World Sci. Publ., Hack-
ensack, NJ, 2007.
Romero Fuster, M. C., Ruas, M. A. S. and Tari, F. (2008), Asymptotic curves on
surfaces in R5 , Communications in Contemporary Maths. 10, pp. 1–27.
Romero Fuster, M. C. and Sanabria Codesal, E. (2002), On the flat ridges of
submanifolds of codimension 2 in Rn . Proc. Roy. Soc. Edinburgh Sect. A
132, pp. 975–984.
Romero Fuster, M. C. and Sanabria-Codesal, E. (2004), Lines of curvature, ridges
and conformal invariants of hypersurfaces. Biträge Algebra Geom. 45, pp.
615–635.
Romero Fuster, M. C. and Sanabria Codesal, E. (2008), Conformal invariants
interpreted in de Sitter space. Mat. Contemp. 35, pp. 205–220.
Romero Fuster, M. C. and Sanabria Codesal, E. (2013), Conformal invariants and
spherical contacts of surfaces in R4 . Rev. Mat. Complut. 26, pp. 215–240.
Romero Fuster, M. C. and Sánchez-Bringas F. (2002), Umbilicity of surfaces
with orthogonal asymptotic lines in R4 . Di↵erential Geom. and Appl. 16,
pp. 213–224.
Romero Fuster, M. C. and Sedykh, V. D.(1995), On the number of singularities,
zero curvature points and vertices of a simple convex space curve. J. Geom.
52, pp. 168–172.
Romero Fuster, M. C. and Sedykh, V. D. (1997), A lower estimate for the number
of zero-torsion points of a space curve. Biträge Algebra Geom. 38, pp. 183–
192.
Ruas, M. A. S. and Tari, F. (2012), A note on binary quintic forms and lines of
principal curvature on surfaces in R5 . Topology Appl. 159, pp. 562–567.
Saeki, O. (1996), Simple stable maps of 3-manifolds into surfaces. Topology 35,
pp. 671–698.
Saeki, O. (2004), Topology of singular fibers of di↵erentiable maps. Lecture Notes
in Mathematics, 1854. (Springer-Verlag, Berlin).
Saji, K. (2010), Criteria for singularities of smooth maps from the plane into the
plane and their applications. Hiroshima Math. J. 40, pp. 229–239.
Saji, K., Umehara, M. and Yamada, K. (2009), The geometry of fronts. Ann. of
Math. 169, pp. 491–529.
Saloom, A. and Tari, F. (2012), Curves in the Minkowski plane and their contact
Bibliography 361
363
364 Di↵erential Geometry from a Singularity Theory Viewpoint
lightcone height functions, 293, 319 surfaces in R3 , 147, 155, 156, 184
Lorentzian distance squared surfaces in R4 , 226, 227, 229, 231,
functions, 309 232
orthogonal projections, 93, 159 surfaces in R5 , 259
finite determinacy, 55 Hessian matrix, 34, 41
first fundamental form, 24, 25, 36, horosphere, 307
140 horospherical
flat rib, 230 Chern-Lashof Type Theorem, 344
flat ridge, 230, 232, 265 Gauss-Bonnet Theorem, 344
flecnodal Horospherical Geometry, 281, 325
curve, 187 hyperbolic 3-space, 284
point, 165, 166, 187 hyperbolic Gauss map, 307
set, 166, 188 hyperbolic point
focal set, 13, 30, 182, 195 surfaces in R3 , 12, 33
fold, 49 surfaces in R4 , 207, 212, 215, 223,
future 229, 238, 239
directed, 284 hypocycloid, 193, 198
direction, 284
inflection, 3, 164, 186
Gauss map, 12, 25–27, 32, 43, 44, 140 inflection point
Gauss map with respect to a normal surfaces in R4 , 206, 210, 212, 214,
vector field, 37 215, 220, 223, 228, 231, 234,
Gauss-Bolyai-Lobachevski, 284 238, 335, 336
Gauss-Bonnet Theorem, 335 internal points, 330
Gaussian curvature, 12, 141, 163, 204,
239 Koenderink Theorem, 163
generalised Gauss map, 226, 227
generating family, 10, 11, 102, 109, Lagrangian
116 di↵eomorphism, 100
generic equivalent, 101
embedding, 81 fibration, 100
immersion, 81, 85–87 immersion, 9
property, 81 map, 9, 101
geodesic inflection, 142, 186, 187, 197 stable, 102
geometric subgroups, 69 submanifolds, 100
germ, 46 surface, 9
graph-like left group, 51
Legendrian left-right group, 51
immersion, 113 Legendrian
submanifold, 113 di↵eomorphism, 108
Morse family of hypersurfaces, 115 equivalent, 108, 109
wavefront, 116 fibration, 107
gravitational collapse, 292 immersion, 10
lift, 108
height function generic map, 10, 108
stable, 109
366 Di↵erential Geometry from a Singularity Theory Viewpoint