Geometrical Methods in The Theory of ODE
Geometrical Methods in The Theory of ODE
ARNOLD
Volume 250
Grundlehren
der mathematischen
Wissenschaften
A Series of
Comprehensive Studies GEOMETRICAL METHODS
in Mathematics
IN THE THEORY
OF ORDINARY
DIFFERENTIAL EQUATIONS
Biblioteca rCEyN * UBA
SECOND EDITION
Springer
Grundlehren der
mathematischen Wissenschaften 250
A Series of Comprehensive Studies in Mathematics
Editors
M. Artin S.S. Chem J.L. Doob A. Grothendieck
E. Heinz F. Hirzebruch L. Hormander S. Mac Lane
W. Magnus C.C. Moore J.K. Moser M. Nagata
W. Schmidt D.S. Scott J. Tits B.L. van der Waerden
Managing Editors
M. Berger B. Eckmann S.R.S. Varadhan
Grundlehren der mathematischen Wissenschaften
A Series of Comprehensive Studies in Mathematics
A Selection
Mark Levi
Department of Mathematics
Boston University
Boston, MA 02215
U.S.A.
98765432
Since 1978, when the first Russian edition of this book appeared, geometrical
methods in the theory of ordinary differential equations have become very
popular. A lot of computer experiments have been performed and some
theorems have been proved. In this edition, this progress is (partially) repre¬
sented by some additions to the first English text. I mention here some of
these recent discoveries.
Moscow V. Arnold
September 10, 1987
Preface to the First Edition
*In the exposition of some special questions, we have also used or recalled elementary
information on differential forms. Lie groups, and functions of a complex variable. This
information is not necessary for the understanding of most of the book.
X Preface to the First Edition
years. The author is grateful to the many readers and students of these
courses for a series of valuable remarks used in the preparation of the
book. The author is grateful to referees D. V. Anosov and V. A. Pliss for a
careful and helpful review of the manuscript.
Chapter 1
Special Equations. 1
§ 1. Differential Equations Invariant under Groups of Symmetries. 1
§ 2. Resolution of Singularities of Differential Equations. 9
§ 3. Implicit Equations. 14
§ 4. Normal Form of an Implicit Differential Equation in the Neighborhood of
a Regular Singular Point. 25
§ 5. The Stationary Schrodinger Equation. 31
§ 6. Geometry of a Second-Order Differential Equation and Geometry of a
Pair of Direction Fields in Three-Dimensional Space. 43
Chapter 2
First-Order Partial Differential Equations . 59
§ 7. Linear and Quasilinear First-Order Partial Differential Equations. 59
§ 8. The Nonlinear First-Order Partial Differential Equation. 68
§9. A Theorem of Frobenius. 86
Chapter 3
Structural Stability. 89
§10. The Notion of Structural Stability. 89
§11. Differential Equations on the Torus . 97
§ 12. Analytic Reduction of Analytic Circle Diffeomorphisms to a Rotation . . 114
§ 13. Introduction to the Hyperbolic Theory. 121
§14. Anosov Systems. 128
§15. Structurally Stable Systems Are Not Everywhere Dense . 141
Chapter 4
Perturbation Theory. 144
§ 16. The Averaging Method. 144
§17. Averaging in Single-Frequency Systems. 148
§ 18. Averaging in Systems with Several Frequencies. 153
xii Contents
Chapter 5
Normal Forms. 180
§ 22. Formal Reduction to Linear Normal Forms. 180
§23. The Case of Resonance. 184
§ 24. Poincare and Siegel Domains. 187
§ 25. Normal Form of a Mapping in the Neighborhood of a Fixed Point. 192
§ 26. Normal Form of an Equation with Periodic Coefficients. 195
§ 27. Normal Form of the Neighborhood of an Elliptic Curve. 202
§ 28. Proof of Siegel’s Theorem. 214
Chapter 6
Local Bifurcation Theory. 222
§ 29. Families and Deformations. 222
§ 30. Matrices Depending on Parameters and Singularities of the Decrement
Diagram. 238
§31. Bifurcations of Singular Points of a Vector Field . 260
§32. Versal Deformations of Phase Portraits. 265
§ 33. Loss of Stability of an Equilibrium Position. 270
§ 34. Loss of Stability of Self-Sustained Oscillations . 286
§35. Versal Deformations of Equivariant Vector Fields on the Plane. 302
§36. Metamorphoses of the Topology at Resonances . 323
§37. Classification of Singular Points. 337
Index 345
Notation
Special Equations
v(gx) = g+xv(x).
Examples
^x
Figure 1.
2. The vector field xdx + ydy on the Euclidean (x, y)-plane is invariant under the
dilations g(x, y) = (Ax, Ay) and rotations.
Exercise. Determine the symmetry group of the field xdx + ydy in the coordinate
plane (x, y).
Examples
1. The direction field of the equation x = v(x) is invariant under translations along
the /-axis (Fig. 2a).
2. The direction field of the equation x = v(t) is invariant under translations along
the x-axis (Fig. 2b).
xA
y / f
x / l
t
(a) (b)
Figure 2.
§ 1. Differential Equations Invariant under Groups of Symmetries 3
Example
The family of integral curves of the equation x = v(t) is transformed into itself by
translations along the x-axis, and that of the equation x = v(x) by translations along
the /-axis.
B. Homogeneous Equations
Definition. The direction field in the plane without the origin 0 is said to be
homogeneous if it is invariant under all the dilations
Example
The function f is said to be homogeneous of degree d if j\exx, exy) = eXdj\x, y). Any
form (homogeneous polynomial) of degree d can serve as an example. Let P and Q be
two forms of degree ^depending on the variables x and y. The differential equation
x = P, y = Q
Figure 3.
4 1. Special Equations
is given by the vector field (P, Q) in the plane. The corresponding direction field in
the domain P # 0 is the direction field of the homogeneous equation
dy = 9, L dy = ax + by dy = *2 - y2 \
dx P \ ' dx cx + dy' dx x2 + y2 ’ /
Exercise. Let P and Q be forms of degree d. Prove that the phase curves of the system
x = P, y = Q are obtained from each other by dilations (Fig. 4).
If one of these curves is closed and has period T, then the dilation gx transforms it
into another closed phase curve having period T/eMd~l).
Let us fix the real numbers a and /? and consider the family of transformations
Example
Exercise. Choose weights in such a way that the function x2 + xy3 is quasi-homo¬
geneous.
Exercise. Prove that the function v(x, y) is the right-hand side of a quasi-homogeneous
differential equation (with weights (a, /?)) if and only if it is quasi-homogeneous of
degree d = ft — a.
6 ]. Special Equations
Remark. The above definitions and theorems can be easily extended to the
case of more than two variables and to differential equations of order greater
than 1. In particular, it is easy to prove the following.
Theorem. Let y: y = ^(x) be a curve in the (x, _y)- plane and let dky/dxk = F
at the point (x0, y0). Then for the curve gsy we have
— e(P-ka)sF
dxk~
Exercise. Prove that if a particle in a homogeneous force field of degree moves along
the trajectory Y in time T, then the same particle moves along the dilated trajectory
XV in time
Exercise. Prove Kepler’s third law: The squares of the times needed for similar trajec¬
tories in the gravitational field are proportional to the cubes of the linear measurements
of these trajectories.
Solution. From the solution of the preceding exercise with d = — 2 (the law of universal
gravity), we obtain T' = X3/2T.
Exercise. Determine how the period of oscillation depends on the amplitude in the
case of a restoring force proportional to the elongation (linear oscillator) and to the
cube of the elongation (weak force).
Answer. In the case of a harmonic oscillator, the period does not depend on the am¬
plitude; in the case of a weak oscillator, it is inversely proportional to the amplitude.
Exercise. It is known that a top with a vertical axis has a critical angular velocity; if
the angular velocity is greater than the critical velocity, then the top stands up firmly
vertically, and if it is less, it falls.
How does the critical angular velocity change if we take the top to the moon, where
the gravitational force is six times less than on the earth?
Figure 7.
More precisely, consider a point jc0 e IR” and assume that the orbit of {<f} going
through x0 is a curve a. Through x0 let us draw an (n — l)-dimensional local trans¬
versal Z to a. In the neighborhood of x0, introduce the local coordinate system (j, u),
where the point gsu of the initial space corresponds to the pair s e IR, u e Z. Then in
the neighborhood of x0, the projection p onto the orbit space and the action of the
group gs of symmetries are given by the formulas
Moreover, {gs} consists of translations along the j-axis; therefore, v does not depend
on s. The vector field v(u) on X defines a direction field on this (n — l)-dimensional
surface. If we know its integral curves, we can find the solutions to the equation dufds
= v(u) (by quadrature) and, consequently, the integral curves of the initial equation.
In the special case n = 2, the passage to the coordinates (s, u) immediately leads
to the integrable equation du/ds = v(u).
~~ = v(u)f(z)
Example
be integrated explicitly, where the weight of jc is a and that of y is ft (so that v is a quasi-
homogeneous function of degree /J — a).
Solution. We may take u = y^/x^, z = x (in a domain where x ^ 0) (cf., Fig. 8b.)
Exercise. Give the equation with separated variables explicitly to which the equation
of the preceding exercise reduces in the coordinates (u, z).
Figure 8.
§ 2. Resolution of Singularities of Differential Equations 9
du _ xy*~lu — flux0~'
dx x0
du at/“-1/°r)w(«) — flu
dx x
A. cr-Process
In the neighborhood of a nonsingular point, all vector fields are simple and
have identical structures.
For the study of fine details of all mathematical objects near a singular
point a special apparatus is devised, having fine resolution similar to a
microscope. It is called the resolution of singularities. From the analytical
point of view, we speak of the choice of a coordinate system near the singular
point in which, to a small displacement near the singularity, there corres¬
ponds a great change in coordinates.
The polar coordinate system has this property; however, the passage to
polar coordinates requires transcendental (trigonometric) functions; there¬
fore, algebraically, it is often more convenient to use another procedure:
the so-called <r-process or resolution of singularities.
We begin with an auxiliary construction. Let p : R2\<9 —> [RP1 be the
standard fibration defining the projective line. (The projective line is a
manifold whose points are the straight lines in the plane going through the
origin of the coordinate system. To a point in the plane, the mapping p
assigns the straight line connecting the point with the origin.)
Let us consider the graph T of the mapping p. This graph represents a
smooth surface in the Cartesian product (1R2\(9) x [RP1 (Fig. 9). Embedding
the punctured plane in the plane, we may consider the graph as a smooth
surface T in the Cartesian product [R2 x [RP1. The natural projection
7ij : (R2 x (RP1 -> IR2 maps T onto the punctured plane 1R2\0 diffeomor-
phically. (In order to picture this more clearly, it is useful to note that T locally
has the form of a spiral ladder; globally, the projective line is diffeormorphic
to the circle, and the product (R2 x [RP1 to the interior of the torus.)
10 1. Special Equations
Figure 10.
Theorem. The closure of the graph T of the mapping p in IR2 x RP1 is a smooth
surface Tj = F U (O x RP1). The surface F, is diffeomorphic to the Mobius
strip {Fig. 10).
Let (x, y) be the coordinates in the plane, and u = yjx the affine local
coordinate in RP1. Then (x, y, u) is a local coordinate system in R2 x RP1.
In this coordinate system, T is given by y = ux, x ^ 0, and T, by the local
equation y = ux. This surface is smooth; it is obtained by adding to the part
of T covered by our coordinate system the part of the projective line O x RP1
falling there.
The proof of the smoothness of Tj can be completed by considering a
second coordinate system (x, y, v), where a: = vy.
The projection n2 : R2 x RP1 —► RP1 foliates Tj into straight lines. When
the circle RP1 is traversed once, the corresponding straight line in R2 turns
by the angle n. It follows from this that Tj is a Mobius strip. ^
Example
Consider three straight lines passing through the point O. On T,, this corresponds to
three straight lines intersecting IRP1 at distinct points (Fig. 11).
Exercise. Consider two curves which are tangent to each other of order n (for example,
y = 0 and y = x2, n = 2). Prove that on T, two corresponding curves have tangency
of orders — 1 at the corresponding point 0, (Fig. 12).
If after the er-process singularities do not reduce to transversal intersections, then
§ 2. Resolution of Singularities of Differential Equations 11
D?P'
Figure 13.
another a-process can be performed at the remaining singular points, etc. until all
singularities reduce to transversal intersections. It can be proved that every algebraic
curve can be resolved (reduced to transversal intersections) in a Finite number of steps.
B. Resolution Formulas
In practice, a cr-process means the passage from the coordinates (x, y) to the
coordinates (x, u = yjx), where x ^ 0, and to the coordinates (o = x/y, y),
where y ^ 0 (Fig. 14). Let us see what this does to the differential equation
given by a vector field in the plane (x, y). We shall assume that the point O
is a singular point of our vector field.
Theorem. A smooth vector field w with singular point at O after the o-process
turns into a vector field on Y extendible to a smooth field on Tj.
Let w be the field given by the system x = P(x, y), y = Q(x, y). Using the
coordinates (x, u = yjx), we find
Figure 14.
12 1. Special Equations
The right-hand sides are smooth, since P(0, 0) = Q{0, 0) = 0. In the other
coordinate system (v = x/y, y), we also obtain a smooth field. ►
Remark. It can happen that the vector field obtained in the er-process
vanishes on the whole straight line pasted in the cr-process. Then, one can
divide the field by x in the domain of the first coordinate system, and by y in
the domain of the second. Division does not change the direction of the
vectors of the field. One obtains a direction field on ^ with singular points
lying on the pasted line but not filling it out completely. In the neighborhood
of every singular point, the direction field is given by a smooth vector field.
To every “entering direction” of phase curves of the initial field at O there
corresponds a singular point of such a field on Tj lying on the straight line
[RP1 pasted in the cr-process.
If these singular points Ot are not sufficiently simple, then er-processes
can be performed at them. Continuing in this way, we finally arrive at the
case where at least one of the eigenvalues of the linearization of the field is
different from 0 at every singular point.
In many cases, the first cr-process enables us to analyze the behavior of
phase or integral curves near a singular point. For example, the integral
curves of a homogeneous equation turn into the integral curves of an
equation with separated variables after our change of variables (x, y) h-►
(x, u = y/x).
We illustrate the method by the trivial example of a linear equation. The equation of
a pendulum with a friction coefficient k has the form x + kx + x = 0. In the phase
plane, this equation is equivalent to the system
x = y, y = -ky - x.
& = _ *
dx y
According to the general theory, the variables should separate after the cr-process,
i.e., in the coordinate system (x, u = yjx). Indeed, dujdx = —(u2+ku+ 1 )/ux. In¬
troducing log |x| = z, we obtain
du
-k u +
dz
We study the integral curves of this equation for different values of the coefficient
k > 0. The graph of the function /' = u + 1/w is a hyperbola (Fig. 15). Consequently,
the graph of the function —k— /(«) has the form shown in Fig. 16. Correspondingly,
§ 2. Resolution of Singularities of Differential Equations 13
U \
J ul \
\ \ -
2
/ z
0<k<2 A ^2
Figure 17.
Figure 18.
the integral curves of the equation du/dz = — k — J\u) have the form shown in Fig.
17. Returning to the phase plane (x, y), we obtain Fig. 18.
Hence, for small values of the coefficient of friction (0 < k < 2), the pendulum
performs an infinite number of oscillations, and for k > 2 the direction of the motion
of the pendulum changes no more than once.
Exercise. Determine the phase curves of the equations z = az" and z = az", re C.
Theorem. Assume that all phase curves passing through points close to the equilibrium
point O are closed. Then the period of oscillation in the neighborhood of O tends to the
period of oscillation in the linearized system as the amplitude of the oscillation tends to 0.
14 1. Special Equations
^ In the er-process, the closed phase curves going around 0 turn into curves on the
Mobius strip closing after two revolutions; that the amplitude of the oscillation con¬
verges to 0 corresponds to the fact that the phase curve on the Mobius strip converges
to the projective line pasted in the a-process (the middle line of the Mobius strip).
Since a solution depends continuously on the initial condition, the limit of the
period of oscillation, as the amplitude tends to 0, is equal to the doubled period of
revolution on the pasted straight line IRP1 in the system obtained in the <r-process. The
velocities of motion on the pasted straight line for the field under consideration and
for its linearization are identical (cf., the equation for u in § IB), It is easy to verify
that all phase curves of the linearized equation are closed. These closed curves in the
linear system are traversed in the same time, since the linear vector field is transformed
into itself by dilations of the phase plane. Consequently, the limit of the period of
oscillation in the initial system is equal to the limit of the period of oscillation in the
linearized system, and, consequently, to the period of oscillation in the linearized
system.
Exercise. Calculate the period of small oscillations of the pendulum x = — sinx near
the equilibrium point x = 0.
§ 3. Implicit Equations
A. Basic Definitions
F(x, y, p) = 0, (1)
where p = dy/dx.
Examples
denoted by dx(£), dy(£), dp(£). Hence, dx, dy, and dp are not mysterious,
infinitely small quantities, but completely defined linear functions of £>.
At the point (x, y, p) of the space of jets, consider the plane consisting of
the vectors £ for which dy — pdx. In other words, the vector ^ at the point
(x, y,p) lies in the indicated plane (Fig. 19) if its projection onto the Eucli¬
dean (x, >>)-plane forms an angle with the x-axis whose tangent is equal to p.
The plane thus constructed is called a contact plane. Hence, a contact plane
passes through every point of the space of 1-jets; all of these planes form the
contact field ofplanes (or, in other words, a contact structure) in the space of
1-jets.
Exercise.* Are there any surfaces in the space of 1-jets whose tangent at every point
is the contact plane drawn at that point?
Answer. No.
Assume that the surface given by Eq. (1) in the space of 1-jets is smooth.
[This is not a strong restriction, since for a generic smooth (infinitely
differentiable) function F, the value 0 is not critical and the set of zeros is
smooth. If this is not so for the function under consideration, then for almost
every arbitrarily small perturbation of Fthe set of zeros becomes smooth: for
example, it is sufficient to add a small constant to F (cf., Sard’s theorem,
§ 10E)].
We consider a point on the smooth surface M given by Eq. (1) and assume
that at this point the tangent of the surface does not coincide with the contact
plane. Then these two planes intersect in a straight line. Moreover, the tangent
and contact planes intersect in straight lines at all nearby points of M, so a
direction field arises on M in the neighborhood of the point being considered.
The integral curves of Eq. (1) are, by definition, the integral curves of
the direction field thus obtained on M. To solve Eq. (1) it is necessary to
find these curves. The connection between the integral curves on M and
the graphs of solutions of Eq. (1) on the (x, >»)-plane is discussed below.
We emphasize that the integral curves on M are defined in terms of contact
planes and not solutions of Eq. (1).
16 1. Special Equations
The direction of the /7-axis in the space of jets will be called vertical direction.
Let M be the smooth surface given by Eq. (1) in the space of jets. Consider
the projection
Example
The discriminant curve of the equation p2 = x is the j/-axis, and that of the equation
p2 = y is the x-axis (Fig. 20).
Theorem. Projection onto the (x, y)-plane turns the integral curves of Eq. (1)
on M in the neighborhood of a regular point exactly into the integral curves
oj the equation
dy
v(x, y) (2)
dx
Remark. Globally, the projections of the integral curves of Eq. (1) onto the
(x, _y)-plane are not, in general, integral curves of some direction field. The
projections of integral curves of Eq. (1) onto the (x, y)-plane have cusps on
the discriminant curve in the general case; however, for some cases of
Eq. (1), these projections remain smooth at points of the discriminant curve.
C. Examples
Consequently, in coordinates (/?, y), the integral curves are determined from the
equation dy = 2p2 dp.
Hence, the integral curves on M are given by the relations y + C = fp3, x = p2.
Their projections onto the (x, y)-plane are semicubic parabolas.
p2
Ip dp = dy,
dy = pdx.
p = 0, v = 0 or x = 2p + C, y = p2.
The projections of these curves onto the (x, _y)-plane are parabolas tangent to the
discriminant curve y = 0.
y = px + f{p),
dy = pdx.
Then (jc + f')dp = 0. The points where x + f' = 0 are critical, and
the remaining points are regular. On the (jc, p)-plane the integral curves
are the straight lines p = const = C; generally these straight lines inter¬
sect the line x + f' = 0 of critical points.
The projections of the integral curves onto the (jc, y)-plane are the
straight lines y = Cx + /(C) tangent to the discriminant curve. (Strictly
speaking, the points of intersection with the critical line do not belong to
the integral curves on M, since for the equation under consideration,
the direction field is not defined at these points: the contact plane is
tangent to M.)
The discriminant curve can be determined from the conditions
y = px + f(p), x + f = o.
Figure 24.
§ 3. Implicit Equations 19
D. Legendre Transformation
Examples
1. Let f(x) = x2j2. Calculating the Legendre transform, we obtain g(p) = p2 /2.
2. Let J\x) = x“/a.. Then g{p) = pfiIP, where a-1 + P~l = 1. (Here x, p, a, and ft
are nonnegative.)
Let / be strictly convex (f" > 0) and its derivative define a diffeomor-
phism of the real line onto itself. Then g is also strictly convex and the
supremum is attained at the uniquely determined point where f'(x) = p.
Points x and p are said to correspond to each other under the Legendre
transformation.
*In the literature several different objects, also associated with the names Minkowski and
Young, are called Legendre transforms as well. Nevertheless, we shall not try to be completely
pedantic concerning terminology.
Figure 25.
20 1. Special Equations
Example
In the corollaries below, we assume that f and g are strictly convex and
their derivatives are diffeomorphisms of the real line onto itself.
This follows from the fact that the inequality in the preceding theorem
is symmetric with respect to f and g. ►
E. Projective Duality
RP" = (Rn+1\0)/(R\O).
tive space for which the corresponding points of the linear space belong to
a hyperplane passing through the origin.
Let us consider the set of all hyperplanes in ^-dimensional projective
space. This set itself is an ^-dimensional projective space in a natural way.
Indeed, a hyperplane in the projective space is given by a homogeneous
equation
where [Rn+1* is the space of linear functions on [R"+1 (this space is linear of
dimension n -1-1 and called the conjugate or the dual space of the initial
space IR"+1).
Hence, to a hyperplane in the projective space, there corresponds a non¬
zero vector in 1R"+1*, determined up to multiplication by a nonzero number.
Consequently, the set of all hyperplanes in IRP" has the natural structure
of a projective space of dimension n:
IRP"* = (Or+1*\0)/(R\0).
Examples
All straight lines passing through one point of the projective plane form a
straight line in the dual plane, as is easily seen. All straight lines passing
through a given point of the projective plane inside an angle with vertex at
this point form an interval in the dual plane.
Theorem. The graphs of a strictly convex function and its Legendre transform
are projectively dual to each other.
Consider all straight lines on the affine (x, y)-plane not parallel to the
y-axis. These straight lines themselves form a plane: a straight line can be
given by the equation y = px — z, and we can consider (p, z) as affine
coordinates in the new plane. In this case, the Legendre transformation
22 1. Special Equations
reduces to the passage from the graph of a function f to the family of tangents
of this graph; when the point on the (x, j>)-plane runs over the graph of /,
the tangent of the graph of / runs over a curve in the (p, z)-plane which is
the graph of the Legendre transform, z = g(p). ►
Therefore, the Legendre transformation is the passage from a curve to
the projectively dual curve given in affine coordinates.
Example
Let the graph of/be a convex polygon. A supporting line is, by definition, a straight
line which is below the graph and which has a point in common with the graph.*
Consider all supporting lines of the convex polygon.
It is easy to see that they form a convex polygon in the dual plane. Indeed, the
supporting lines form an angle at every vertex of the initial polygon and, consequently,
form a segment or closed interval in the dual plane. Similarly, the vertices in the dual
plane are obtained from the sides of the initial polygon.
Projective duality allows us to consider more general cases than that of the Legendre
transformation.
Exercise. Determine the curve projectively dual to the curve in Fig. 26.
Hint. To the double tangents of the initial curve, there correspond points of self-
intersection of the dual curve. Consequently, the dual curve has four points of self¬
intersection.
To points of inflection of the initial curve there correspond cusps of the dual curve.
Indeed, iff = x3, then the tangent of the graph at the point x = t is given by the coor¬
dinates p = 312, z = 213. These relations define a curve with cusp in the (p, z)-plane.
Hence, the dual curve has eight cusps, two between any pair of consecutive self¬
intersection.
Moreover, consider the initial curve as a pair of intersecting ellipses somewhat
smoothed near the points of intersection (the part of each ellipse inside the other is
eliminated).
* In general, a supporting hyperplane of a convex body is a plane which has a common point
with the body and is such that the body lies in one of the half-spaces into which the plane divides
the space.
§ 3. Implicit Equations 23
Figure 27.
The dual curve is also connected with a pair of ellipses. To the points of intersection
of the initial ellipses there correspond double tangents of the dual ellipses. It is easy
to see how to construct the dual curve from the pair of ellipses with their double tangents
(Fig. 27).
Definition. A norm in IR" is, by definition, a real nonnegative convex positively homo¬
geneous even function of degree 1. It is equal to zero only at the origin:
Let/be a norm in IR". The function /is determined by the set where it is equal to
1. This set is- a convex hypersurface centrally symmetric with respect to zero. Con¬
versely, every convex compact body in IR” centrally symmetric with respect to zero
and containing zero inside it defines a unique norm which is equal to 1 on its boundary.
This hypersurface/ = 1 is called the unit sphere of the norm f.
Theorem. The unit sphere of the dual norm is the set of supporting hyperplanes of the
unit sphere of the initial norm.
24 1. Special Equations
The condition g(p) = 1 means that (p, x) has maximum equal to 1 on the unit
sphere, i.e.. the plane p = 1 is a plane of support for the initial sphere. ^
The set of all hyperplanes of support of a given convex hypersurface is called the
dual convex hypersurface. Hence, the unit spheres of dual norms are dual.
Exercise 2. Determine the hypersurfaces in [R3 dual to the (a) sphere, (b) tetrahedron,
(c) cube, (d) octahedron.
From what was said above it is clear that the passage from a convex hypersurface
to its dual is locally given by the Legendre transformation.
x = x(s, /), y - t)
of two variables define a family of curves in the plane parametrized by the parameter
t (indicating the point on the curve) and indexed by the parameter s (indicating the
index of the curve).
(a) x = (s + /2), v = t;
(b) x = s + st + t3, y = t2, 5 and t small;
(c) x = (5 + t2)2, y = t.
It can be shown that for a generic family the envelope is a curve whose only singular¬
ities are cusps (as in the case of a semicubic parabola) and points of self-intersection;
in the neighborhood of every point where the envelope is smooth, the family reduces
to one of the normal forms (a), (b), or (c) by means of a smooth transformation ATv, y),
K(.v, y) of the coordinates and a smooth transformation S(s), T(s, t) of the parameters.
This is not true in the holomorphic case. See J. P. Dufour, Families de courbes planes
differentiables, Topology 22, 4 (1983), 449-474; and V. I. Arnold, Wave fronts
evolution and the equivariant Morse lemma. Comm. Pure Appl. Math. 29, 6 (1976),
557-582.
(c)
Figure 28.
§ 4. Normal Form of an Implicit Differential Equation 25
2
dy
X
dx
by means of a smooth transformation X(x, y), Y(x, y) of the coordinates (cf., § 4).
The projections of the integral curves onto the (A\ T)-plane are semicubic parabolas
in these coordinates. Hence, the discriminant curve is the envelope of the projections
of the integral curves only for exceptional equations (for example, Clairaut’s equation).
In particular, for a small generic perturbation of Clairaut’s equation, the discriminant
curve turns from the envelope into the locus of cusps of the projections of the integral
curves.
A. Singular Points
Definition. A point of the surface F = 0 is said to be singular for Eq. (1) if the projec¬
tion (x, y, p) i—► (x, y) of the surface to the plane is not a local diffeomorphism in the
neighborhood of this point.
By the implicit function theorem, the singular points are those points of the surface
F = 0 at which dF/dp - 0.
B. Criminant
Consider the set of all singular points of Eq. (1). In the three-dimensional space of
jets, this set is given by the two equations F = 0 and dFjdp = 0. Therefore, “generally
speaking”, the singular points form a curve.
By the implicit function theorem, the criminant is a smooth curve in the three-
dimensional space of jets in the neighborhood of each of its points where the rank of
the derivative of the mapping (x, y, p) i—> (F, dF/dp) of the three-dimensional space to
the plane is maximal (equal to 2).
C. Discriminant Curve
Definition. The projection of the criminant to the (x, jy)-plane parallel to the p-direction
is called the discriminant curve*.
Remark. Under the above conditions, the discriminant curve can well have singularities.
These singularities arise because, in general, several points of the criminant curve
can be mapped onto the same point of the discriminant curve by the projection. Ge-
nerically, these singularities will be points of self-intersection of the discriminant curve.
For the generic equation in the neighborhood of such a point, the discriminant curve
consists of two branches intersecting at a nonzero angle.
On the other hand, to the points where the tangent to the criminant is parallel to
the p-direction, generically there correspond cusps on the discriminant curve.
All more complicated singularities of the discriminant curve besides points of self¬
intersection and cusps can be eliminated by a small perturbation of the equation.
These two types of singularities are preserved, only a little displaced, for small deforma¬
tions of the equation.
At every point (x, y, p) of the space of jets, there is a contact plane dy = pdx. In parti¬
cular, such a plane exists at the points x of the criminant. The tangent of the criminant
at a given point may lie in the contact plane or it may intersect it.
Definition. A point of the criminant is called a point of tangency with the contact plane
if the tangent of the criminant at this point lies in the contact plane.
We note that points of tangency of the criminant with the p-direction are points of
tangency with the contact plane. Indeed, the contact plane contains the p-direction at
every point.
Definition. A singular point of Eq. (1) is said to be regular if at this point the condition
of smoothness of the criminant*
Example
Remark. For a generic equation, almost all singular points are regular: the nonregular
points lie on the criminant discretely. If this is not so for a given equation, then it can
always be achieved by a small perturbation. This and the previous "genericity argu¬
ments” can be verified by Sard’s theorem in § 10.
Corollary. In the neighborhood of a regular singular point, the family of integral curves
of Eq. (1) is diffeomorphic to the family of semicubic parabolas y = x312 + C.
The diffeomorphism mentioned in the theorem maps integral curves of Eq. (1) in
the (a, >’)-plane to integral curves of the equation P2 = X in the (X, T)-plane. These
latter integral curves are semicubic parabolas with cusp on the discriminant curve:
dY/dX = ^/X, Y = \X312 + C. ►
1. Reduction to the Case Where the Criminant Is the y-Ax is. Let(.v0, v0,/>0)
be a regular singular point of the equation F(x, y, p) = 0. Then the discriminant curve
is smooth in the neighborhood of (x0, y0). Consider the projections of the contact
planes onto the (x, y)-plane at the points of the criminant. In the neighborhood of
(*0’ To)> we obtain a family of straight lines not tangent to the discriminant curve.
Now choose a local coordinate system near (x0,y0) in the (x, y)-plane so that (1)
* M. Cibrario, Sulla reduzione a forma canonica delle equationi lineari alle derivate parziali di
secondo ordine di tipo mislo\ Accademia di scienze e lettere, Instituto Lombardo, Rendiconti 65
(1932), 889-906.
28 1. Special Equations
the equation of the discriminant curve is x = 0; (2) the straight lines y = const inter¬
sect the discriminant curve in the directions just constructed.
As before, we shall denote the coordinates by (x, y) and the derivative dyfdx by p.
The singular point has the coordinates (0, 0, 0).
det
D(F, Fp)
* 0, i.e..
FXI, * 0
D(x,p)
FJ0, y, 0) A 0, Fpp(0, y, 0) * 0.
It follows from the above relations that A(0,y) = 0, 5(0, y) = 0. Therefore, we may
write A(x, y) = xa(x, y), B(x, y) = xp(x, y), where a and P are smooth functions.
The regularity conditions of the criminant curve have the form Ax(0, y) ¥= 0,
C(0, y, 0) # 0. In what follows we may even assume that C > 0, Ax < 0 (if this is
not so, we change the signs of F and/or x). Thus a(0, 0) < 0, C(0, 0, 0) > 0.
2C 2C
where y = —4otC -I- xP2 is a function of (x, y, p), and y(0, 0, 0) = — 4a(0, 0, 0) x
C(0, 0, 0) > 0.
Finally, let x = £2. Keeping only the sign “ + ” in “ ± ”, we obtain
-epg\y) + ZJv(e,y,p)
2C(£2, y, p)
We apply the implicit function theorem to this equation in p(£, y). We obtain a solution
p = £cu(<f, y), where (o is a smooth function and a>(0, 0) # 0.
4. A Differential Equation for y{f). We note that p = dyfdx — dy/2£ d£,. There¬
fore, we obtain the following differential equation fory(^):
In the (<!;, y)-plane the integral curves intersect the axis ^ = 0 and have points of tangency
of order 2 with the straight lines y = const. Therefore, the equation has a first integral
of the form /(<{;, rj) = y — £3K(£, y), where K is a smooth function, and A^(0, 0) ¥= 0.
(/ is the coordinate of the intersection point of the integral curve through (<!;, .y) with
the axis £ = 0; K # 0 since oj # 0.)
Here L and M are smooth functions of x and y, and L(0, 0) ^ 0. With this notation
we have /(£, y) = y — CMif2, y) — £3 L(£2, y). We introduce new variables Y and E
by the formulas
Then 1 = Y - E3.
Consider also X - E2. Then
These formulas give a local diffeomorphism of the plane in the neighborhood of (0, 0),
since L(0, 0) # 0. The first integral takes the form
I = Y — X3'2.
Now (dY/dX)2 = \X. The normal form can be obtained by stretching one of the
coordinate axes. ^
H. Remarks
The basic step in the above proof is the substitution x = f2, i.e., the passage to a two-
sheeted covering of the (jc, jy)-plane with branching along the discriminant curve. From
topological arguments (although, in a complex domain), it is clear a priori that the
two-valuedness of p(x, y) disappears on this two-sheeted covering, and the equation
splits into two. To trouble with the quadratic equation is necessary only for the proof
of this fact in a real domain. It remains to reduce Eq. (2) obtained on the covering
to its normal form by a diffeomorphism of the covering plane into itself. This can
easily be achieved by decomposing the first integral into even and odd parts with
respect to
Remark. Our proof used the representation of an even function by a function of the
square of its argument. For analytic functions (or formal series) such a representation
is obvious. In the case of a smooth function, it needs a proof.
Indeed, an infinitely differentiable even function can be considered as a function
of the square of the argument with values on the positive semiaxis. It is infinitely
differentiable at all points of this semiaxis, including zero. We have to represent it as
30 1. Special Equations
y = (P + kx)2
by a diffeomorphism of the (jc, j/)-plane; for the proof, see A. A. Davydov, The normal
form of the differential equation implicit in the derivative, Funct. Anal. Appl. 19, 2
(1985), 1-10. The topological case was settled in A. V. Phakadse, A. A. Shestakov,
On the classification of singular points of a first-order differential equation implicit in
the derivative. Mat. Sbornik 49, 1 (1959), 3-12.
The direction field on the surface defined by the equation has, at a generic point
of tangency of a contact plane with the surface, a singularity of the same type as a
direction field of a plane generic vector field at its singular points—that is, a saddle
node or a focus. Hence these singular points of the implicit differential equations are
called folded saddle (nodes, foci): they are obtained from ordinary saddles (nodes, foci)
by a folding mapping.
It is interesting to note that the foldings generate no new moduli: the parameter k
in the normal form is defined by the ratio of the eigenvalues of the linearization of the
vector field, whose phase portrait generates the folded singularity under the folding
mapping (Fig. 29).
In contrast with these singularities, those corresponding to the cusps of the dis¬
criminant curve have functions-moduli not only with respect to diffeomorphisms, but
even with respect to homeomorphisms of the plane (x, y) (see the Davydov paper
quoted above). These cusps are the projections of Whitney pleats on the (jc, j>)-plane.
The projections of the integral curves on this plane may be described as the family
of projections of the swallowtail surface’s generic plane sections under a generic (rank
Figure 29.
§ 5. The Stationary Schrodinger Equation 31
2) mapping of the ambient space on the plane. The swallowtail surface is given by
^ + (E - t/(*))T - 0 (1)
ties of Eq. (1) and equations and systems of partial differential equations
generalizing it.
Example
Let U = 0. Then the particle is said to be free. The Schrodinger equation for a free
particle with energy E = k2 has the form
These two solutions describe the particle moving to the right (with momentum k > 0)
and the particle moving to the left, respectively. Hence, the space of solutions of a free
particle with energy £ is a two- dimensional complex space.
Physicists call the square of the absolute value of the wave function the
probability density of the event that the particle is at a given place. Hence,
a free particle with impulse k “can be found at any point with the same
probability.” (This terminology can be used regardless of the meaning of
these words and regardless of how all this is connected with probability
theory.)
B. Potential Barriers
We assume that the potential has compact support (is different from zero only
in some domain). If U ^ 0, then we say that we are given a potential barrier,
if U ^ 0, then we say that we are given a potential well. The domain where
the potential is different from zero is called the support of the potential
(Fig. 31).
We assume that the energy E = k2 of the particle is positive. Left of the
support, Eq. (1) coincides with Eq. (2) of a free particle. Consequently,
Schrodinger’s equation has two solutions which coincide with etkx and
e~ikx left of the support. These two solutions are called the particle incoming
from the left and the particle outgoing to the left, respectively. We note that
U\
Figure 31.
§ 5. The Stationary Schrodinger Equation 33
these solutions are defined for all x, but coincide with e'kx and e lkx only
left of the support.
Exactly in the same way, there exist two solutions which coincide with
e'kx and e~lkx right of the support. These solutions are called the particle
outgoing to the right and the particle incoming from the right, respectively.
Exercise. Can a particle arriving from the left be entirely reflected to the left (i.e., can
a wave function be zero right of the barrier and not zero left of it)? Can the particle
depart to the right entirely?
C. Monodromy Operator
Notation. Denote by [R2 the space of real solutions of the Schrodinger equation
(1). The space of states of the particle (i.e., the space of complex solutions of
the equation) is the complexification of IR2 ; we denote it by C2 = C!R2. All
four states of the particle arriving and departing to the right and left belong to
this space.
The space of real solutions of eq. (2) of a free particle is denoted by [Rq
(since U = 0 for a free particle). In this space, we have the following natural
basis:
/. = 3ikx
A =
- ikx
Note that el and e2 determine a basis in the state space, too. These two bases
are connected via the relations
A = ei + iei. A = ei
- ie2.
34 !. Special Equations
Theorem. The matrix of the monodromy operator in the basis (/, ,f2) belongs to
the group SU(1, 1).
The reason why the monodromy operator belongs to SU(1, 1) is that the
phase flow of Eq. (1) preserves area. For the proof, recall some information
on the group SU(1, 1).
Consider the real linear space R2 and its complexification C2. In R2, choose
an area element and denote by [£, >7] the oriented area of the parallelogram
generated by the vectors £ and rj. The skew-symmetric inner product [ , ] is
called a symplectic structure. If a basis (ex,e2) is fixed in R2 for which
[ex,ef\ = f then [£, >7] is equal to the determinant consisting of the
components of the vectors £ and rj in the basis ex, e2.
The complexification of the bilinear form [ , ] defines a symplectic
structure in C2, which we shall denote by the same square brackets.
We note that the form [ , ] is nondegenerate: if [£, rf\ = 0 for all then
>7 = 0.
Consider the Hermitian form >/> = rf\ in C2. This is indeed a
Hermitian form: g)> = >7), rj) = <>7, £>. For the following, it is
useful to calculate the Hermitian products of the vectors fx = ex + ie2 and
f2 = ex — ie2. It is easy to prove the following lemma.
^ For example.
1
</,,/,> = j [/.,/,] 2 [-A ’-^2] — 2 1. ^
—i
Hence, the Hermitian form < , > is of “type (1, 1)” (one positive and one
negative square in the canonical form <z, z> = |zj|2 — \z2\2)-
Now we consider the linear transformations of the plane C2 preserving
the Hermitian, symplectic, and real structures.
§ 5. The Stationary Schrodinger Equation 35
Figure 32.
Theorem. The intersection of any two of these groups coincides with the inter¬
section of all three groups {Fig. 32).
This intersection is called the special (1, 1 )-unitary group* and is denoted
bySU(l, 1). [It is also called real unimodular group and is denoted by SL(2, R)
or the real symplectic group of the second order and denoted by Sp(l, R).]
If the transformation A is real and unimodular, then [A£, Ag] = [<!;, rf\
and A£ = A£. Therefore, (Ag, Ag> = i/2\_Ag, Arf\ = i/2\_Ag, Atf] =
1/2[f, g~\ = <£, *?>• _ _
If A is real and (1, l)-unitary, then <Ag, Ag} = (g, g} and Ag = A£.
Therefore, [AAg] = —2i(Al;,Agy= -2i<A£,Agy = -2i(£,rjy = [£,g].
If A is (1, l)-unitary and unimodular, then [Alj, Ag] = [<^, g~] and
[A£, Ai{] = [<^, rj\. Therefore, [A£, Ag] = [A^, Arf\ for all ^ and g. Con-
* We emphasize that we are dealing with operators and not matrices. The matrices of these
operators belong to the matrix group SU(1, 1) for the special choice of basis indicated above.
36 1. Special Equations
sequently, [£, At] — Atf\ = 0 for all £, and hence At] = At] for all t], i.e., A is
real. ^
Corollary. If the matrix of an operator in the real basis (el, e2) is real uni-
modular, then the matrix of this operator in the complex conjugate basis
(Ji = el + ie2,f2 — e\ ~ ief) is special {1, l)-unitary, and conversely.
M The Hermitian scalar square of the vector z1fl + z2f2 can be expressed
by the coordinates (zl,z2) in the basis (f\,f2) by the formula <z, z) =
1-zJ2 — |z2|2 (cf., the lemma above). Therefore, the following statements
are equivalent:
(i) the matrix of A in the basis (ex, e2) is real and unimodular;
(ii) A e GL(2, R) H SL(2, C).
(iii) A e SU(1, 1); and
(iv) the matrix of A in the basis (fx, f2) is (1, 1 )-unitary and unimodular.
The matrix groups SL(2, IR) and SU(1, 1) are connected with the Lobachevsky
geometry in the following way (Fig. 33).
A real unimodular matrix of the second order defines a linear-fractional
transformation z —► (az + b)/{cz 4- d) mapping the upper half-plane onto
itself. This transformation is a motion of the Lobachevsky plane represented
in the form of the upper half-plane. All motions of the Lobachevsky plane can
be obtained in this way. The group of motions of the Lobachevsky plane is
isomorphic to SL(2, IR)/±E.
A 2 x 2 unimodular (1, l)-unitary matrix defines a linear-fractional trans¬
formation of the unit disk onto itself. Indeed, the cone \zl |2 < |z2|2 is mapped
onto itself by any (1, l)-unitary unimodular transformation. Under the
natural mapping
this cone turns into the unit disk |w| < 1 and the linear transformations of
C2 into linear-fractional transformations of CP1 (Fig. 34).
The linear-fractional transformations of the unit disk onto itself expressed
by matrices belonging to SU(1, 1) are the motions of the Lobachevsky plane
represented as the interior of the unit disk. All motions of the Lobachevsky
plane can be obtained in this way. The group of motions of the Lobachevsky
plane is isomorphic to SU(1, 1 )/±E.
The matrix groups SL(2, R) and SU(1, 1) are isomorphic: they are
obtained from the same group of operators. The matrices of these operators
in the real basis (el5 e2) belong to SL(2, R), and in the complex conjugate
basis, belongs to SU(1, 1). The passage from SL(2, R) to SU(1, 1)
corresponds to the passage from the real basis to the complex conjugate
basis, and from the model of the Lobachevsky plane in the upper half-plane to
its model in the unit disk.
Problem. Prove that SL(2, R) is homeomorphic to the solid torus S 1 x D2 (the interior
of a doughnut).
assigning the initial condition at the point x to each (real) solution of Eq. (1).
This operator is an isomorphism. The isomorphism g*2 = BXl{Bx')~l is
called the phase transformation from xy to x2.
For Eq. (2) of a free particle, the operator (solution i—► phase point)
Bx : R2 - R2
M ,2
a
38 1. Special Equations
Here / indicates a point left of the support, and r a point to the right of it. The
operator M does not depend on the choice of these points.
M In the space IRg of real states of a free particle, we choose the basis
c, = coskx, e2 — sin/:*. In the real phase space [R^, the coordinates T, 4^
are chosen, thereby determining a basis as well. In this basis, the matrix of
Bq has the form
f cos kx sin/or \
0 \— A: sin/:* k coskx)
Problem. Prove that the Schrodinger equation (1) does not have a nonzero solution
coinciding with ae'kx to the left of the support and with be~,kx to the right of it (no
particle can arrive without departing).
Solution. The monodromy operator preserves the (1, 1 )-Hermitian square \zl |2 — \z2\2.
On the other hand, <ae‘kx, ae'kxy = \a\2, <be~'kx, be~lkxy = —\b\2 (cf., the Lemma in
§ 5D). Consequently, \a\2 = —16|2, i.e., a = b = 0.
Definition. We say that a particle arriving from the left with impulse k > 0
passes the barrier with transmission coefficient \A\2 and reflection coefficient
\B\2 if the Schrodinger equation (1) in which E — k2 has a solution equal to
§ 5. The Stationary Schrodinger Equation 39
Lemma. The solution and the complex constants A and B satisfying the above
conditions exist and are unique for every k > 0.
Problem. Prove that the transmission coefficient is always different from zero.
Solution. If *P = 0 to the right of the barrier, then 4* = 0 to the left of it, as well.
-4 Lemma. In the basis (/t = e,kx,f2 = e~‘kx), the matrix of the monodromy
operator can be expressed in terms of the complex coefficients A and B by the
formula
\/A — B/A \
(M)
-B/A 1/A )'
\/A B/A \
(AT1)
B/A 1/A )
Problem. Calculate the transmission and reflection coefficients for the potential equal
to a constant U0 for 0 < x ^ a and zero at the remaining points (Fig. 36).
Solution
Uq sin2 ak t \ 1
4E(E - U0) ) ’
where E = k2, E — U0 = k\. (Passing over the barrier, the particle slows down;
therefore, the probability density of finding it within the barriers is larger than finding
it outside.) For large E, the coefficient of reflection converges to 0,
Ul
B\2 sin2 aki.
4E(E - U0)
If the energy of the particle is less than the height of the barrier, then the transmission
coefficient is exponentially small:
4k2x2
(k2 + *2)sh2 aye + Ak2*2'
Figure 38.
I. Scattering Matrix
Along with the passage through a barrier from left to right we may consider
passage from right to left. The corresponding solution yY2 is equal to
A B
B2 a2
Consequently,
42 1. Special Equations
Comparing this with the matrix calculated in the Lemma of § 5H, we obtain
A2 = A, B2 = -BA/A.
Since \A\2 + \B\2 = \A2\2 + |i?2|2 = 1 and AB2 + BA2 = 0, the matrix S is
unitary. ►
Remark. We have considered the Schrddinger equation (1) with the real spectral
parameter E = k1. It proves to be very useful to consider complex vaules of k as well.
The scattering matrix remains unitary and symmetric for complex k. Besides, S is
''real’', S( — k) = S(A:) and "analytic”, A(k) is the boundary value of a function,
analytic in the upper half-plane Im/r > 0, and having a finite number of poles on the
imaginary axis.
Since the transmission and reflection coefficients can be measured, there arises the
so-called inverse problem of scattering theory of determining the potential U from the
scattering matrix S(k).
The potential U is given by a real function on the real line or two real functions on
the half-line. The coefficients A and B are two complex functions on the semi-axis
k > 0, i.e., four real functions on this semi-axis. The unitarity condition\A |2 + \B\2 = 1
decreases the number of real functions on the semi-axis from four to three.
Since3 > 2, it can be expected that not every pair A, B satisfying the condition\A |2 +
\B\2 = I corresponds to a potential: in order to reconstruct the potential from A and
B, another condition has to be imposed on these coefficients. Analyticity turns out to
be such a condition.
It is surprising that these heuristic arguments based on the calculation of the number
of arbitrary functions can be turned into theorems which can be formulated and
proved in an exact manner (however, these theorems go beyond the scope of this course).
J. Bound States
Figure 39.
§ 6. Geometry of a Second-Order Differential Equation 43
the left of the well and the coefficient of the exponential function increasing
to the right vanishes to the right of the well. A solution with these properties
does not exist for every negative value of the energy.
It turns out that if the well is sufficiently deep and wide, then there exist a
finite number of negative values of E for which the particle can be stationary
in the well; the deeper and wider the well, the larger the number of these
values.
These values of E are called stationary levels, and the wave functions 4*
decaying as x —> ± oo are called bound states (in case the well is not compact,
J|T|2 dx < oo is required).
Problem. Determine the stationary energy levels in the rectangular well of depth U{t
located between x = 0 and x = a (Fig. 40).
(tan e > 0 for the first equation, and tan d; < 0 for the second).
Problem. Prove that the wave functions T. corresponding to bound states with distinct
energy levels are orthogonal:
T, T2 dx = 0.
Problem.* What is the connection between stationary energy levels of bound states
and the poles of the S-matrix on the imaginary axis?
The graphs of solutions of the equation d2yjdx2 = 0 (straight lines) satisfy the con¬
figuration theorems (Papp, Desargues, etc.) of projective geometry.
d2y dv
-A + «(*)-/ + b^y = 0
dx dx
is locally (in the neighborhood of any point x = x0) diffeomorphic* to the family of graphs
of solutions of the simplest equation d2yjdx2 = 0.
Corollary. The configuration theorems of projective geometry hold (locally) for the graphs
of solutions of any second-order linear equation, for example, for the family of curves
y = /4sinx -I- Bcos x or y = Aex — Be~x.
y, yi
Remark 2. Two locally projective manifolds are said to be equivalent if there exists a
diffeomorphism mapping one locally projective structure into the other.
♦That is, there exists a diffeomorphism of the neighborhood of the straight line x = x0
onto the (x, y)-plane turning the graphs of solutions into straight lines.
§ 6. Geometry of a Second-Order Differential Equation 45
Hint. All locally projective structures on a straight line are induced by a mapping into
the projective line (i.e., the circle) with nonvanishing derivative; the number of inverse
images of points of the circle under this mapping is an invariant of the structure.
On the circle the two-sheeted covering of the projective line defines a locally projec¬
tive manifold structure, not equivalent to the structure of the projective line. However,
not every locally projective structure on the circle is induced from the structure of the
projective line. The classification of locally projective structures on the circle is con¬
nected with the classification of Hill’s equations (second-order linear equations with
periodic coefficients). Even equations with constant coefficients introduce structures
not induced from the projective line.
dzy = <t> x, y
dy
dx2 dx
We will study the geometry of the two-parameter family of curves given by this
equation in the (x, >)-plane. In particular, we are interested in finding out whether the
configuration theorems hold for this family and whether this family can be rectified
(turned into a family of straight lines) by an appropriate diffeomorphism of the plane.
We shall see that such a rectification is not always possible, and find invariants measur¬
ing “infinitesimal nondesarguesness” (violation of the configuration theorems).
Theorem. In the neighborhood of every linear element (x, y, p) of the plane of the depen¬
dent and independent variables, every second-order differentia! equation can be reduced to
* It is useful to observe that if the right-hand side <J> is a polynomial of degree w ^ 1 in the
argument p, then it is also such a polynomial in the new coordinates constructed in § 6A.
46 1. Special Equations
sequently, the Taylor series of O in y and p begins with terms of not less than second
order:
.x = F(X, Y), y = Y
(converting the point with coordinates (x, y) into the point with coordinates {X, T)).
dfy dy
#(*, y, P), P
dx2 dx
d2 Y dY
= ®{X, Y, P), P
dX2 dX'
where
here A = F' + PFy ; the prime denotes the partial derivative with respect to X, and the
arguments of F and of its derivatives X and Y.
Let >’ = u(x) be a solution of the initial equation and Y = U(X) its image. Then
dU _ du P
P = (F' + PFY), ^
dX dx dx A'
FiX.lHX))
Consequently,
d2U A d du
dX2 dX dx FiX.lHX))
+ i(F"+ 2F'r + F”p2 +
Moreover,
d du d2u
A = U(X)), U(X), 0A.
dXdx dP
FiX.UiX))
Hence,
§ 6. Geometry of a Second-Order Differential Equation 47
PFV\ d2U
= <D 2 pf; + p2fyy)
dX2
Corollary 2. Let <S>bea polynomial in p of degree not greater than 3. Then <t> is a polynomial
in P of degree not greater than 3.
Corollary 3. A differential equation of the second order def ines the structure of a local
projective line on the graph of every solution, and the structure of a local projective plane
on the normal bundle of the graph.
d2 Y dy
4>(jc, y, p). P
d2X dx
d2Z
o(z, z, n), n = —,
dX2 dX
where
here the prime indicates partial derivation with respect to X, the arguments of G and its
derivatives are always X and Z.
^ Let y = v(x) be a solution of the initial equation, and Z = V(X) its image. Then
dv _
s-c +
(Here and in the following the arguments of G and its derivatives are X and V(X).)
Moreover,
Determining V" from this equation, we obtain the formula of the lemma. ^
The lemma just proved implies the following corollaries.
Corollary 1. Let <D = 0. Then <I> is a polynomial in FI of degree not greater than 1.
Corollary 2. Let <t> be a polynomial in p of degree not greater than n, n ^ 2. Then <t> is
a polynomial in FI of degree not greater than n.
dy
* P)’ P <D = Q(\y\2 + |p|2)
dx dx’
§ 6. Geometry of a Second-Order Differential Equation 49
into
d2Z
4^2, 2, n). n = —,
dX2 dX
where
This can be proved by applying the lemmas concerning the transformations of the
independent and then the dependent variable. Expanding the right-hand sides of the
formulas obtained there in a series in (Z, 11) and retaining only the quadratic terms,
we obtain the expression Q indicated above as an addendum to the quadratic terms
of O. h
5. Reduction of the Quadratic Terms. Denote the quadratic terms of the initial
right-hand side by
I = 6A - 2B' + C"
C. Infinitesimal Nondesargueness
The coordinate system in which the second-order differential equation has the form
50 1. Special Equations
near the graph of a fixed solution is not determined uniquely. We study the extent to
which the coefficient A is invariant, i.e., independent of the method of reduction to
normal form. A obstructs the rectifiability of the family of solutions and measures the
infinitesimal nondesargueness at the point and in the direction in question.
oj = A{x)\dx\512
In other words, if (X, Y) is another coordinate system in which the equation is also
in normal form and y = 0 corresponds to the solution Y = 0 and the coefficient A(x)
is changed to A(X), then
dx 5/2
A(X) = CA(x)
dX
The vector of the normal bundle at the point x with >’-component £ is converted to a
vector at the point f0(x) with the _y-component gfx)£,.
The projective structure of the normal bundle is defined in an invariant manner
(cf., §6B); therefore, the transformation (x, i—» (/0(x), gi(xK), constructed by means
of the diffeomorphism (x, y) i—*■ (X, Y), converting an equation in normal form into
another equation in normal form must be projective. Thus, we find
. _ ax + b _ C
0 cx + d' cx + d
Every diffeomorphism leaving the axis y = 0 invariant and preserving the projective
structure of its normal bundle can therefore be represented as the product of a special
transformation
X = f0(x), Y = ygA(x),
invariant I consisting of those terms of the right-hand side of the equation which are
quadratic in y and p (cf., § 6B5) does not change under this diffeomorphism.
We study the behavior of / under a special projective transformation.
Every projective transformation of the line splits into the product of translations,
expansions, and the transformation _y i—> l/.v.
The expression I is invariant with respect to translations, and expansions of x and y
only multiply I by a constant. Therefore, it is sufficient to consider the behavior of
I under the substitution x = 1/A\ v = Y/X.
Calculating the derivatives P = dY/dX and dP/dX = d1 YjdX2, we find that
P = y — px, dPIdX = X~3dp/dx (where p = dy/dx). Consequently,
The differential form to introduced above yields scalar functions connected with the
equation in an invariant manner.
First of all, we note that with any differential form (of arbitrary order) on a one¬
dimensional manifold, one can associate a vector field in an invariant manner. The
value of the form on a vector of this field is equal to 1 at every point.
For example, the vector field v(x)d/dx, where v = A~2/5, is invariantly connected
with the form A(x)(dx)s/2.
Theorem. Let v(x)d]dx be a vector field on the line. Then the following scalar functions
are connected with this field invariantly with respect to projective transformations of the
line:
Remark l. The invariant I2 is constructed by the following procedure. The Lie algebra
of the projective group of the line is generated by the fields d/dx, xdjdx, and x23jdx*.
/I A /exp(//2) 0 \ / I °\
VO 1/ \ 0 exp( —1/2))' \-t 1/
52 1. Special Equations
Therefore, every vector field can be approximated by a projective field (a field from
the Lie algebra of the projective group) up to quadratic terms at every point.
Under projective transformations, the projective field approximating the initial
field at the initial point turns into a new projective field approximating the transformed
field at the image point. The action of the projective transformations of the line on the
three-dimensional space of projective fields is the adjoint action of the projective group
on its Lie algebra. This action preserves the quadratic form on the algebra. Indeed,
if we express projective transformations by second-order matrices, and the projective
fields by matrices of the infinitesimal generators of the one-parameter groups corre¬
sponding to them, then the action of the transformation g on the field v is described
as the matrix product gv~lg. On the other hand, det gvg~l = det r. Therefore, the deter¬
minant of v is a quadratic form on the Lie algebra of the projective fields invariant
with respect to the adjoint representation. Consequently, this determinant, calculated
for the projective field approximating the vector field under consideration is a scalar
connected with the field in an invariant manner with respect to projective transforma¬
tions.
In the above basis of the Lie algebra, the approximating projective field has the
components (u, v', v"/2). Therefore, the matrix corresponding to the field has the form
( v'!2 v ).
V- v"l2 - v'/2) ’
Remark 2. It seems clear that every function (polynomial, series, etc.) of the values of r
and a finite number of derivatives of v which is invariantly connected with v with
respect to projective transformations can be expressed in terms of the invariants Ik.
Remark 3. The projective invariants of a function on the projective line can be con¬
structed in the following way: let its differential (1-form) correspond to the function,
let its field correspond to its form, and let the invariants Ik correspond to the field. In
particular, the simplest invariant of / with respect to projective transformations of the
independent variable is
(This differs from Schwartz’ derivative which is invariant with respect to projective
transformations of the axis of the values of the function by the factor J'2 in the de¬
nominator).
Remark 4. The invariants I2, /3, • ■ ■ are multiplied by a2, A3, ... if the vector field
is multiplied by A. It is easy to construct combinations of them which are not sensitive
to multiplication of the field by a number; for example, J = /f//|.
0 IN (h 0 \ / 0 0\
0 o/ Vo -h)' V-1 0/
§ 6. Geometry of a Second-Order Differential Equation 53
The vanishing of the form co of nondesargueness along any solution is necessary for
the rectifiability of the equation (its reduction to the form d2yjdx2 = 0), but, as we
shall immediately see, it is not sufficient.
d2v dy
y< p). p
'dx2 dx ’
In other words, a differential equation of the family of all lines on the plane described
in an arbitrary system of curvilinear coordinates has, for its right-hand side, a poly¬
nomial in the first derivative of degree not greater than 3.
Theorem 1 follows from the (curious) fact below.
d2y
<D(x, y, p)
dx2
is a polynomial in p of degree not greater than 3. Then every diffeomorphism of the plane
transforms the equation into an equation of the same kind, i.e., the right-hand side remains
a polynomial in the derivative of degree not greater than 3.
Theorem 2 follows from the lemmas in § 6B on the effect of the change of the indepen¬
dent and dependent variables on the right-hand side of the equation (cf., the corollaries in
§ 6B), since every local diffeormorphism of the plane can be obtained by successive
applications of these changes of variables. ^
•4 Theorem 1 follows from Theorem 2 and the fact that zero is a polynomial in p of
degree not greater than 3. ►
d^y
a0y oty bxy' - b0
dx2
(where a, and b{ are functions of x and y), make the change of variables (x, y) i—»(y, x).
Solution
d2x
b0x'3 b^x1 + axx' — a0.
~d?
54 I. Special Equations
Remark. It can be shown that the conditions oj = 0 and d4d>jdp4 are independent.
Therefore, the condition to = 0 is not sufficient for the reducibility of the equation
to the form
The two conditions [to = 0, d4<J?/dp4 = 0} together are sufficient for the reducibility
of the equation to the form d2y/dx2 = 0. This can be seen from the formulas of § 6B
(after some calculation).
Exercise. Prove that every second-order equation can be reduced to the form d2v/dx2 =
p2 B(x, y, p) locally (in the neighborhood of the point x = 0 of the solution y = 0).
We consider a pair of direction fields in three-dimensional space. It turns out that the
local classification (up to a diffeomorphism of the space) of such pairs in general
position is equivalent to the local classification of second-order differential equations
(up to diffeomorphisms of the plane of the dependent and independent variables:
local, meaning near a given point with direction).
First of all, we associate a two-parameter family of curves in the plane with the pair
of direction fields in three-dimensional space.
To do this, we rectify the first field by a local diffeomorphism of the space, converting
the family of integral curves of the first field into a family of parallel vertical lines.
After this, we project the integral curves of the second field onto the horizontal plane
in the direction of the vertical lines. On the horizontal plane (on the quotient plane of
the space modulo the integral curves of the first field), we obtain a two-parameter family
of curves.
From this two-parameter family of curves, we construct a second-order differential
equation for which these curves are graphs of solutions.
To this end, we note that in a generic local two-parameter family of curves in the
plane, a unique curve of the family passes through every point of the plane in every
direction near every linear element (a point and a direction) of every curve of the family.
(For this, only a certain Jacobian has to be different from zero).
If the family is generic in the indicated sense and (x, >•) are the coordinates in the
plane, then d2yldx2 is a smooth function of the element under consideration i.e.,
(x, y, dy/dx). In this way we obtain a second-order differential equation d2y/dx2 =
<D(x, y, dy/dx). The graphs of the solutions of this equation are curves of the family
(by the uniqueness theorem).
Consequently, to a pair of direction fields in three-dimensional space we assigned
a second-order differential equation under the nondegeneracy condition that the
corresponding Jacobian is different from zero.
Remark. It is easy to see that the definition is unambiguous: if the second field becomes
nondegenerate with respect to the first after some diffeomorphism rectifying the
integral curves of the first field, then it becomes so after any other diffeomorphism
rectifying the first field. It is also not difficult to see that the nondegeneracy of a pair
is preserved if the order of the fields is changed (i.e., it is immaterial which one of the
two fields is rectified).
It follows from the theorem just proved that all our results concerning the geometry
of the family of solutions of a second-order equation can be reformulated in terms of
56 1. Special Equations
G. Duality
Exercise. Prove that the nondesargueness form io of a second-order equation (cf., § 6B) is
equal to zero if and only if the right-hand side of the dual equation is a polynomial in
the derivative of degree not greater than 3.
Hence, the rectifiability condition for a second-order equation can be formulated
as follows:
An equation d2yjdx2 = <I>(x, y, dy/dx) can be reduced to the form d2yjdx2 = 0 if
and only if the right-hand side is a polynomial in the derivative of order not greater than
3 both for the equation and for its dual.
H. Survey
side of the equation and its derivatives of order not greatei than k at point (x, y, p).
If it vanishes for the equation at the given point of this three-dimensional space of
elements, then it will vanish for the equation transformed by a diffeomorphism of the
plane at the transformed element.
It turns out that there are exactly two functionally independent semi-invariants of
order 4. One of them is d*<&ldpA. The other is the scalar invariant I2 constructed from
the form of infinitesimal nondesargueness (cf., § 6D). For an equation in the normal
form d2y/dx2 = A(x)y2 + 0(|jy|3-.+ |/?|3), we have
In this theory, the problem of orbit space has apparently remained unresolved, in
particular, the question of determining how many parameters are needed to enumerate
the orbits of a A--jet of a second-order equation (under the action of the group of dif-
feomorphisms of the plane on the space of A--jets of the equation), in the neighborhood
of a generic point in the space of A-jets.
It seems that the Poincare series counting these parameters are rational (in this
problem, as in many other analytical classification problems, where functional moduli
occur) and that these series provide some rigorous meaning to the expression “depends
on p arbitrary functions in q variables”.
Chapter 2
Partial differential equations have been studied much less than ordinary
differential equations. Part of the theory—that of a single partial differential
equation of the first order—can be reduced to the study of special ordinary
differential equations, the so-called characteristic equations. This is possible
because, physically speaking, non-interacting particles can be described by
either the partial differential equation of the field or ordinary equations for
the particles.
In this chapter we develop this theory (which mathematically reduces to
the geometry of the so-called contact structures). We also consider the
problem of integrability of a field of hyperplanes (the theorem of Frobenius).
The assertion of the theorem is local and invariant with respect to diffeo-
morphisms. Therefore, it is sufficient to prove it for a standard field of
parallel directions in a linear space. In this case, the assertion of the theorem
is obvious. [It reduces to the fact that a function given on an interval is
constant if and only if its derivative is equal to zero (Fig. 42)].
Let T be a ^-dimensional submanifold in an ^-dimensional manifold X
(Fig. 43). r is called a hypersurface if k = n — 1.
Definition. The Cauchy problem for the direction field V with initial manifold
T is the problem of finding a (k + l)-dimensional integral submanifold of
V containing the initial submanifold T.
We note that the integral curves passing through T do not always form
a submanifold even locally in the neighborhood of T, cf., Fig. 44.
Lau = 0, (1)
Let the field a have the components (a1, ••,«„) in the coordinates
(jtj, - - - , x„); every component is a function of the coordinates. Equation
(1) becomes
du du
a + + a = 0. (2)
1 dxx Vx.
Definition. The field a is called the characteristic vector field of Eq. (1) and
its phase curves are called characteristics. The equation x = a(x) is called
the characteristic equation of the partial differential equation (1).
Definition. The Cauchy problem for Eq. (1) is the problem of finding a
solution u satisfying the condition u\ = cp, where y is some smooth hyper¬
surface in M and q> is a prescribed smooth function on this hypersurface.
The hypersurface y is called the initial hypersurface and the function (p
the initial condition.
We note that such a problem does not always have a solution. Indeed,
on each characteristic, the solution u is constant. However, a characteristic
may intersect the initial surface y several times. If the values of the prescribed
function <p are different at these points, then the corresponding Cauchy
problem does not have a solution in any domain containing the characteristic
in question (Fig. 45).
Figure 46.
cu
= 0, MU,=0 = <P- (3)
<x i
The unique solution (in a convex domain) is u(xx, x') = <p(x'), where
x' = (x2, ■ ■ ■ , x„). ►
Laii = b, (4)
cu cu u
a + +
,
°n?- = *. (5)
cx,
Figure 47.
The Cauchy problem for Eq. (4) can be formulated in the same way as
for the homogeneous equation (1).
~~ = b, u\x o = <P(x').
rx, 1
In other words, Eq. (4) means that the derivative of the solution along
the characteristic is the known function b. The increment of the solution
on a segment of the characteristic has to be equal to the integral of b over this
segment (Fig. 47).
Lau = p, (6)
This differs from a linear equation in that the coefficients ak and b may
depend on the unknown function.
Example
cu . cu cu ru
(p = —<p + — = u— 4- —.
(X (t (X ct
uux + ut — 0. (8)
Problem. Construct the graph of the function u{ ■, t) if the graph of the function w( , 0)
has the form shown in Fig. 48.
Answer. Cf., Fig. 49. For t $= there exists no smooth solution. Beginning at t = tx,
the particles collide in the medium. [The physical hypothesis of inertial motion, i.e.,
the lack of interaction between the particles, becomes unrealistic and has to be replaced
by other physical hypotheses, the description of the nature of the collision. Thus arise
so-called shock waves, functions of the form shown in Fig. 50, satisfying Eq. (8) outside
the discontinuity and some additional conditions of physical origin at the discontinuity.]
We have just seen how useful it is to pass from the field of velocities to the
motion of the particles for the special quasilinear equation (8). Something
similar can be done in the case of the general equation (6). In this case, the
role of the motion of the particles is played by some curves in the direct
product of the domain and range of the unknown function; these curves
are called the characteristics of the quasilinear equation.
The quasilinear equation (6) for an unknown function u\ M —* IR, i.e.,
means that if the point x leaves x0 and begins to move on M with velocity
a(x0, u0), then the value of the solution u = u0 begins to change with
velocity b(x0, u0).
In other words, the vector A(x0, u0) applied at the point (x0, u0) of the
space M x IR, which has components a(x0, u0) along M and b(x0, u0) along
IR, is tangent to the graph of the solution (Fig. 51).
Definition. The vector A(x0, u0) is called the characteristic vector of the
quasilinear equation (6) at the point (x0, u0). The characteristic vectors at
all points of the space M x IR form a vector field A. This field is called the
characteristic vector field of the quasilienear equation (6). The phase curves
of the characteristic vector field are called the characteristics of the quasi¬
linear equation.
The differential equation given by the field A of phase velocities is called
the characteristic equation.
Example
Example
dx j _ dx2 _ _ dxn _ du
ai a2 an b
◄I This is obvious, since Eq. (6) says that a characteristic vector is tangent
to the graph. ^
§ 7. Linear and Quasilinear First-Order Partial Differential Equations 67
Definition. The Cauchy problem for the quasilinear equation (6) with initial
condition cp on y consists of determining a solution u which is equal to cp
on y.
The solution of this problem reduces to the solution of the Cauchy
problem for the characteristic direction field.
We consider the graph of the function cp: y —>■ IR. This graph is a hyper¬
surface in the direct product y x R. Since y is imbedded in M, we may
regard the graph T of <p as a submanifold (of codimension 2) of M x IR
(Fig. 52).
Remark. If the equation is linear,.then the vector a(x0, u0) does not depend
on u0. Therefore, noncharacteristic points of the surface y can be defined.
Figure 52.
68 2. First-Order Partial Differential Equations
On the other hand, for a quasilinear equation, only a point (x0, u0) e T can
be characteristic or noncharacteristic, but we cannot speak of a point
x0 e y being characteristic.
M If the initial condition is not characteristic at the point x0, then we have:
(1) The characteristic field A does not vanish in the neighborhood of the
point (x0, u0). Therefore, a smooth characteristic direction field is
defined in the neighborhood of this point.
(2) The characteristic direction is not tangent to the initial manifold T
at the point under consideration, and consequently, in its neighbor¬
hood. Therefore, the local integral surface, containing the initial
manifold T, of the characteristic direction field exists and is unique
(cf.,§7A).
(3) The tangent plane to the integral surface at the point (x0, u0) is
nonvertical (does not contain the w-axis). Therefore, the integral
surface is the graph of a function. This function is the desired solution
(cf.,§7E).
Remark. The proof also contains a procedure for constructing the solution
of the Cauchy problem for a quasilinear equation.
A. Contact Manifolds
Figure 53.
may not have integral surfaces whose dimension is equal to the dimension
of the planes. In order to measure the obstruction to the existence of integral
surfaces of a field of hyperplanes, we carry out the following construction.
In the neighborhood of the point O of the manifold under consideration,
we introduce coordinates so that the plane of the field at O becomes a
coordinate hyperplane; we shall call the corresponding coordinates horizontal
and the remaining coordinate vertical.
For any path in the horizontal plane, we construct a vertical cylinder
over the path. The trace of our field of planes on the lateral surface of the
cylinder gives a direction field. The integral surfaces (if they exist) intersect
the cylinder in integral curves of the direction field (Fig. 53).
Consequently, we may lift any path from the horizontal plane to the
desired surface.
Now let (£, rj) be two vectors in the horizontal coordinate plane at the
point under consideration. Take the parallelogram spanned by (£, rj). There
are two paths along the sides of the parallelogram from the point under
consideration to the opposite vertex of the parallelogram. By lifting these
paths, we obtain two points above the opposite vertex which are different
in general. Their difference is the obstruction to the construction of an
integral surface, or, as is said, to the “integrability” of the field of hyper-
planes.
We consider the difference between the vertical coordinates of the two
points obtained above. The principal bilinear part (with respect to t, and r\)
of this difference measures the degree of nonintegrability of the field. In
order to give a formal definition, we carry out the following construction.
The field of tangent hyperplanes to the manifold can be given locally by
a differential 1-form a which is nowhere equal to the zero form. It is defined
up to multiplication by a function which vanishes nowhere: the planes of
the field are the null spaces of the form (those subspaces of the tangent space
on which the form vanishes).
Example
In the space IR2n+1 with coordinates (xj, . . . , xn ; w; /?,, . . . , pn), consider the 1-form
cl = du — pdx (where pdx = p, dxx 4- - • ■ + p„dxn). This 1-form is not equal to the
70 2. First-Order Partial Differential Equations
zero form at any point of R2n+1 and, consequently, defines the field of 2«-dimensional
planes a = 0 in [R2n+1 (Fig. 54).
Example
The form constructed in the preceding example is a contact form. Indeed, the exterior
derivative of the form a is equal to
In the plane cl = 0, (.x.. . ... , pn) may serve as coordinates. The matrix of
, f° ~E\
the form co = rfa ,_n has the form in these coordinates, where E is the
11-0 \E 0 J
identity matrix of order n. The determinant of this matrix is equal to 1. Consequently,
the 2-form co is nondegenerate.
Figure 55.
d(fa) — df a a + f da.
Remark I. From the preceding theorem, it follows that it does not depend
on the choice of the form whether the 1-form which defines the field of
planes is a contact form. This is determined by the field of contact planes
itself. Indeed, if f is another form defining the same field, then (locally) (3
differs from a by a nowhere-vanishing factor; consequently, the forms /?
and a are either both contact forms or neither of them is.
Definition. The standard contact form on the manifold JfV, IR) of 1-jets of
functions is the 1-form
a = du — p dx (p dx = p l dxt + • • • + pndxn).
The first part follows from the definition of the total differential du
[) dx: of a function. The second part follows from the existence of a function
with prescribed partial derivatives at a given point. ►
From the theorem just proved, it follows that the field of planes defined
by the standard contact form in the space of 1-jets does not depend on the
choice of the coordinate system in the definition of the standard contact
form.
Problem. The groups of diffeomorphisms of the spaces V and R act naturally on the
manifold JX{V, (Rt) of 1-jets of functions. Prove that the standard contact structure of
the space of 1 -jets is invariant under this action.
Moreover, the standard contact 1-form a is invariant under the action of the group
of diffeomorphisms of V and is multiplied by a nonvanishing function under the
action of diffeomorphisms of the real line IR.
Px = TXE n n,.
* A generic hypersurface in a contact space may have some isolated points, where the contact
plane is tangent to the hypersurface and the transversality does not hold, that is, there may be
singular points of the equation for partial derivatives defined by the hypersurface.
For a generic surface at some neighborhood of a point we can reduce both the surface and
the contact structure to the normal form
where Q is a nondegenerate quadratic form inp and q (which may be chosen in the form lkpkqk
in the complex case).
This theorem for smooth (C°°) functions is proved in V. V. Lychagin, The local classification
of first-order partial differential equations, Uspekhi Mat. Nauk (Russian Math. Surveys) 30, 1
(1975), 101-171 (see also: V. I. Arnold, A. B. Givental, Symplectic Geometry, Itogi Nauki i
Techniki Sovremennye Problemy Mathematiki, Fundamentalnye Napravleniia Dynamicheskie
Systemy, Vol. 4, pp. 5-140, Mosc. VINITI, 1985 (Springer translation, 1988). The analytical
version of the Lychagin theorem has been proved by M. Ya. Zhitomirskii, Funct. Anal. Appl.
20,2(1986), pp. 65-66.
Figure 56.
74 2. First-Order Partial Differential Equations
D. Skew-Orthogonal Complement
Examples
1. Let (L, <o) be a Euclidean space, i.e., let L be a linear space and co a
scalar product. To every vector £ there corresponds a 1-form to(£, ■), the
scalar product with the vector The value of this 1-form at the vector
rj is equal to cu(^, rj).
For example, grad / is the vector corresponding to the 1-form dj
To a straight line in L there corresponds the orthogonal complement
of this line (the plane of zeros of the 1-form corresponding to the vector
of the line). Every plane of codimension 1 is the orthogonal complement
of a line in L (Fig. 57). We note that multiplication of to by a nonzero
number preserves orthogonality of vectors. Therefore, the correspon¬
dence between straight lines and their orthogonal complements does
not change if we multiply the scalar product by a nonzero number.
2. Let (L, oj) be a symplectic space, i.e., let L be a linear space and to a
skew-scalar product (a bilinear skew-symmetric nondegenerate form).
To every vector £ there corresponds a 1-form cu(£, • ), the scalar
product with £. The value of this 1-form at the vector rj is equal to <u(£, rj)
(Fig. 58).
For example, a Hamiltonian field with Hamiltonian function H is a
field corresponding to the 1-form dH.
To a straight line in L there corresponds its skew-orthogonal comple¬
ment (the plane of zeros of the 1-form corresponding to the vector of the
line). Every plane of codimension 1 is the skew-orthogonal complement
of a unique line in L.
Problem. Prove that every straight line lies in its skew-orthogonal complement.
Solution. <o(c. c) = f) = 0.
Problem. Prove that if all vectors in a suhspace of a 2n-dimensional sympletic space are
skew-orthogonal to each other, then the dimension of this subspace does not exceed n.
in nf1.
Figure 59.
76 2. First-Order Partial Differential Equations
a(v) - 0.
Lemma. In order that the field of zeros of a be invariant with respect to the
phase flow of the field v, it is necessary and sufficient that da(v, f) = 0 for
all £ in the plane a(^) = 0 of the field.
The assertion of the lemma is local and invariant with respect to diffeo-
morphisms. Therefore, it is sufficient to prove it for the standard field
v = f/f.v, in a Euclidean space with coordinates (jc,, . . . , xj (in view of
the theorem on the rectification of a vector field). Let a = a 1 dx: -F • • • +
andxn. By assumption, we have at = 0 (since a(v) = 0).
Problem. Prove that the value of the exterior derivative of the form a at the
pair (v, O (where v = f/Cr,, a{v) = 0, and c, is an arbitrary vector) coin¬
cides with the value of the partial derivative ca/cxl at the vector £.
Solution
< a-
da = X dxj a dxl 'da(v. () = Zfi, - Zf-tj
( -V; v 3xt dx.
However, a, = 0.
Lemma A. The form a vanishes on the planes tangent to Y at the points of the
initial manifold.
Lemma B. The diffeomorphisms g' convert null planes of a into null planes.
Lemmas A and B imply that the form a vanishes at all vectors tangent
to Yn. Consequently, Yn is an integral manifold.
Hence, we have constructed an integral submanifold Yn a E2n of the
field of contact hyperplanes, passing through the initial manifold N"-1.
H. Proof of Uniqueness
<X>U, u, p) = 0. (1)
Theorem. The solutions of Eq. (1) are the functions whose l-graphs consist of
characteristics on E2n.
Definition. The Cauchy problem for Eq. (1) is the problem of determining a
solution u : V —► IR which is equal to q> on y.
Figure 64.
§ 8. The Nonlinear First-Order Partial Differential Equation 81
(3) The value of the total differential of the function at this point is such
that its restriction to the plane tangent to yn~l is equal to the total
differential of the initial condition q>.
(4) The jet belongs to E2n.
Remark. We say that “the derivatives of the unknown function are deter¬
mined by the initial condition in n — 1 directions on y, and the derivative
in the last direction (transversal to y) is determined by Eq. (1).”
Example
Let y be given by the equation xx = 0 in the space with coordinates (jcj , x').
Then N is given by the conditions
dtp
xt = 0, u = <p(x'). P' =
Theorem. Assume that the point (x0, u0,p0) of the space of 1 -jets is a non¬
characteristic point of the initial manifold N. Then the solution of Eq. (1) with
initial condition N exists in some neighborhood U of the initial point x0 and
is locally unique (in the sense that every two solutions of the equation satisfying
the initial condition u\vrty — cp\uny,u(x0) = u0,du(x0) = p0 coincide in some
neighborhood of x0).
This follows from the theorem of § 8G, which also provides a means of
constructing the solution. ►
K. Explicit Formulas
Problem. Calculate explicitly the differential equation of the characteristics for the
equation <l>(x, u, p) = 0.
Answer.
* = <t>P,
P = -4>x - ;
ti =
82 2. First-Order Partial Differential Equations
Solution. Consider a vector tangent to the manifold 0 = 0. For such a vector with
components (X, U, P) we have
®xX + Ou£/ + O pP = 0.
This vector lies in the contact plane if the form du- pdx vanishes for it, i.e., JJ = pX.
Hence, a condition necessary and sufficient for the vector (X, pX, P) to lie in the
intersection of the contact plane with the tangent to the manifold O = 0 is that
The characteristic vector (x, u = px, p) is defined by the condition that the skew-
scalar product with all vectors (X, pX, P) satisfying the preceding equality vanishes.
This skew-scalar product is equal to the value of da. = dx a dp at the pair of the
vectors (x, u = px, p), (X, U = pX, P), i.e., equal to xP — pX.
Hence, Eq. (2) for X and P must be equivalent to the equation
xP - pX = 0. (3)
Consequently, the coefficients of X and P in Eqs. (2) and (3) are proportional. This
provides the above answer in view of the equality u = px.
Problem. Give explicit conditions for y, <p, and O under which the point (x0, u0, p0) is
noncharacteristic for the equation ®(x, «, p) = 0 with the initial condition (p on y.
Solution. We project the plane tangent to the surface formed by the characteristics
passing through the initial manifold onto the x-space. If the surface is the 1-graph of
the solution and the point (x0, u0,p0) is noncharacteristic, then the tangent plane is
generated by the tangent to N and the characteristic direction. It is projected isomor-
phically. Consequently, the x-component of the characteristic vector must be trans¬
versal to y at x0. But this component is (cf. § 8K).
Conversely, let Op be not tangent to y at x0. Then:
(1) In the neighborhood o/'(x0, u0,p0), the hypersurface 0 = 0 is smooth. Indeed,
Op ¥= 0, and consequently, */0|<JCq „o , ^ 0.
(2) At (x0, u0,p0), the equation 0 = 0 is noncharacteristic. Indeed, the vector
(0, 0, Op) lies in the contact plane and is not tangent to the surface O = 0 at (x0, m0, p0),
since Op(x0, u0,pQ) =£ 0.
(3) Near (x0, u0, p0) the initial manifold N is smooth.
Indeed, we choose coordinates (Xj, . . . , xj = (x,, x') so that the local equation of
y takes the form x, = 0. Then the condition of solvability of the equation
for pfx') takes the form dO/Sp, |(.V(rl<i],P(i) ^ 0; the condition that the vector <t>p is not
tangent to y has the same form.
(4) The point (,x0, uQ, p0) of the initial manifold is noncharacteristic.
Indeed, if the characteristic vector were tangent to the initial manifold N, then its
projection <l>p would be tangent to the projection y of N onto the .x-space.
(5) The characteristics intersecting the initial manifold in the neighborhood of (,x0,
u0, po) form a smooth manifold in this neighborhood which is projected dijfeomorphically
onto the x-space (and thus is a \-graph of a function).
Indeed, the image of the plane tangent to this manifold at (,x0, «0, pQ) under projec¬
tion onto the .x-space contains the tangent plane of y and a vector transversal to it.
Consequently, at (,x0, u0, p0), the derivative of the mapping under consideration is an
isomorphism, and the projection mapping itself is a local diffeomorphism (by the
inverse function theorem).
Hence, all five conditions for being noncharacteristic are satisfied at (.x„, u0,p0) if
<5 is not tangent to y at this point.
H{x,ux) = 0. (1)
Example
Let y be a smooth hypersurface in the Euclidean space IR" and let u(x) be
the distance of the point x from y (Fig. 65). The function u satisfies the
Hamilton-Jacobi equation
2
tu
| + • • • + (2)
.ex.
Figure 65.
84 2. First-Order Partial Differential Equations
Problem. Prove that every solution of a Hamilton-Jacobi equation (2) is locally the
sum of the distance from a hypersurface and a constant.
Remark. This system of differential equations is called the system of Hamilton's canonical
equations. The corresponding vector field is defined on not only the surface H = 0
but on the whole phase space.
Figure 67.
86 2. First-Order Partial Differential Equations
§ 9. A Theorem of Frobenius
A direction field in the plane always defines a family of integral curves and
is locally rectifiable (it reduces to a field of parallel lines by a diffeomor-
phism). In a three-dimensional space this is not true: a field of planes in
R3 may in general not have integral surfaces.
In this section we discuss conditions for the local rectifiability of a field
of hyperplanes, i.e., conditions under which the field is a field of tangent
hyperplanes to a family of smooth hypersurfaces.
Remark. The property of local integrability does not depend on the choice
of the form a which defines it locally, since under the multiplication of a
by a nonvanishing function, the form da|a = 0 is multiplied by this function,
(cf.. ij 8A).
a. a da. = 0.
Figure 68.
§ 9. A Theorem of Frobenius 87
Lemma. The phase curves of the field v passing through points of the integral
manifold Fk form a smooth integral manifold Tk + 1 of a completely integrable
field of planes a = 0.
Denote by {g‘} the local phase flow of the field v. Then (a) the diffeomor-
phisms g‘ convert the planes of our field a — 0 into planes of the field.
Indeed, da(£,) = 0 for every vector £ of a plane of the field. Therefore,
the field of planes is invariant under the diffeomorphisms g' according to
the lemma of § 8F.
Moreover, (b) the tangent space to Yk + l lies in the plane of the field at
the points of the initial manifold Tf
Indeed, both the tangent plane of the integral manifold rk and the vector
v belong to a plane of the field, and the tangent space to Yk+1 at x is spanned
by v(x) and Txrk.
Figure 69.
88 2. First-Order Partial Differential Equations
From (a) and (b) it follows that the manifold Yk+l is an integral manifold
for the field of planes a = 0. ►
Structural Stability
x - p(x), x e M,
given by the vector field v on the manifold M. We will say that the field v
defines a dynamical system (or briefly, a system). We shall assume (as a rule)
that the solutions of the equation can be extended infinitely; this is always
so if M is compact.
Example
If k = 0,then all phase curves are closed. If k > 0, they spiral towards
the singular point O, which is of focal type. Consequently, a small change
in the friction coefficient changes the behavior of the phase curves qual¬
itatively if the coefficient had been zero before the change. It does not change
the qualitative picture if the coefficient had been positive.
The definition of structural stability given below formalizes this differ¬
ence: a pendulum without friction turns out to be a structurally unstable
system, and a pendulum with friction is a structurally stable one.
B. Topological Equivalence
Example
the second, which converts the phase flow of the first system into the phase
flow of the second:
hg\x = g2hx.
C. Orbital Equivalence
Example
Consider a vector field having a closed phase curve, say, a limit cycle. Then
every topologically equivalent system also has a limit cycle with the same
period. A small change in the field may cause a small change in the period.
Consequently, the period of motion on the cycle is a continuously changing
invariant (modulus) with respect to topological equivalence as well. In order
to get rid of such moduli, a classification of systems is introduced which is
still coarser than the classification up to a homeomorphism.
Let M be a compact manifold (of class Cr_1, r ^ 1). Let v be a vector field
of class Cr (if M has a boundary, then it is assumed that v is not tangent
to it).
§ 10. The Notion of Structural Stability 93
Let all singular points of the field be nondegenerate. There are a finite
number of them and they are alternately stable and unstable. Every non¬
constant solution of the equation x = v(x) tends to a stable equilibrium
position as t —> + oo and to an unstable one as t —> — oo. This easily implies
all assertions of the theorem except one: it remains to be proven that all
singular points can be made nondegenerate by means of an arbitrarily small
perturbation of the system.
It is convenient to prove the last assertion by means of Sard’s theorem.
Lemma. The measure of the set of critical values of a smooth function defined
on the interval [0, 1] is equal to zero.
We divide the interval into N equal parts and single out those which
contain critical points. If N is sufficiently large, then the derivative of the
function does not exceed C/N on each of the parts singled out (C is a constant
independent of N). Therefore, the length of the image of each of the parts
singled out does not exceed C/N2. Cover this image by an interval of length
2C/N2. We have obtained a cover of the set of critical values by intervals of
total length not greater than 2C/N. ►
We consider a family of vector fields with parameter e on the circle given
by the formula u(xr, e) = v(x) — e. A point x is a degenerate singular point
of the field corresponding to the value e of the parameter if and only if e is
a critical value of the function v at the point x.
94 3. Structural Stability
All critical points form a set of measure zero. Therefore, there exist
arbitrarily small noncritical values. Fix a noncritical value e. All singular
points of the Field corresponding to this value of the parameter are non¬
degenerate. P>
Let /: A/ -*■ ,/V be a smooth mapping of manifolds of any dimension. A point of the
source space is said to be critical if the dimension of the image of the differential of
the mapping at the point is smaller than the dimension of the image space. The value of
the mapping at a critical point is said to be a critical value.
Theorem. The measure of the set of critical values of every sufficiently smooth mapping
is equal to zero.
/.If the dimension of the source space is equal to zero, then the theorem is obvious;
if the dimension is equal to 1, then it has already been proved above. We assume that
the theorem is proved in all cases where the dimension of the source space is equal to
m — 1, and prove it in the case where it is equal to m.
2. We divide the set K of critical points of the mapping into several parts. A point
of the source space is called a point of Jlattening of order r if all partial derivatives of
orders 1, . . . , r are equal to zero at this point. We denote by Kr the set of all points of
flattening of order r.
3. First we consider the set of critical points of K\KX. We prove that the measure of
the corresponding set of critical values (i.e., the measure of the set f{K\Kx)) is equal to
zero.
At every point of K\KX, one of the partial derivatives is different from zero, say,
cj/j/cbr,. In the neighborhood of such a point,/, can be chosen as a local coordinate
instead of x,, retaining the other coordinates x2, . . ., xm in the source space. In these
coordinates,/can be written as a one-parameter family of smooth mappings of (m — 1 )-
dimensional spaces into (n — 1 /dimensional ones:
We fix a value c of the parameter of J\. The mapping /induces a mapping/ of the
(m — 1/dimensional plane j\ = c in the source space into the (n — 1/dimensional
plane/ = c in the image.
The set of critical values of the mapping/ has (n — 1/dimensional measure zero
in the plane / = c by the induction hypothesis (the theorem is proved for (m — 1/
dimensional source spaces). By Fubini’s theorem, the n-dimensional measure of the
union of the sets of critical values of the mappings/ is equal to zero.
On the other hand, the image of the set of critical points in K\KX lying in the neigh¬
borhood of the point under consideration is contained in this union. This implies that
the measure of j\K\Kf is equal to zero.
4. We consider the set Kr\Kr+l of points of r-flattening. We prove that the measure
of the corresponding set j\Kr\Kr+l) of critical values is equal to zero.
§ 10. The Notion of Structural Stability 95
At every point of Kr\Kr+l, one of the partial derivatives of order r 4- 1, say dgjdxl,
is different from zero, where g is one of the partial derivatives of /, of order r (in an
appropriate local coordinate system).
In the neighborhood of such a point, the set Kr\Kr+1 is contained in the smooth
{m — 1/dimensional hypersurface g = 0. The points of Kr\Kr+i are critical for the
restriction of /to this'hypersurface, since df = 0 on Kr. By assumption, the measure
of the image of the set of critical values of the restriction of f to this hypersurface is
equal to zero. Consequently, the measure of J\Kr\Kr+l) is equal to zero.
5. Finally, we consider the set Kr of r-flat critical points for sufficiently large r. We
prove that the measure of the corresponding set f(Kr) of critical values is equal to zero if
r is sufficiently large.
To this end, we divide each side of the m-dimensional cube in the source space
(choosing local coordinates) into N equal parts, divide the cube into Nm equal small
cubicles, and single out those which contain points from Kr. The diameter of the
image of any cubicle singled out does not exceed c(l/7V)r+1 (where the constant c does
not depend on TV). Therefore, the images of all cubes singled out are covered by open
cubes with total measure not greater than
,n(r+ 1)
c, TVm
TV
Remark. A nondegenerate singular point of the field does not disappear for
a small perturbation of the field, and moves only slightly (by the implicit
function theorem). In contrast to this, a degenerate singular point, in general,
bifurcates under a small perturbation of the field (splits into several non¬
degenerate ones) or disappears. Therefore, all singular points of a structurally
stable system are nondegenerate.
Figure 75.
Remark. The proof of structural instability if at least one of conditions (I )-(4) is violated
is not complicated (cf., Fig. 76). The proof of the fact that conditions (l)-(4) imply
Figure 76.
§11. Differential Equations on the Torus 97
Theorem. The structurally stable vector fields form an everywhere dense open
set in the space of vector fields on the two-dimensional sphere.
Remark. Analogous results hold for a vector Field on the disk not tangent
to the bounding circle.
In this section we discuss the theory of vector fields without singular points
on the two-dimensional torus, due to Poincare and Denjoy. In particular,
all structurally stable fields are described.
The ^-dimensional torus is the direct product of n copies of the circle. The
two-dimensional torus T2 — S1 x S1 can be represented as the square
with opposite sides pasted together [the points (0, y) and (27t, y) as well as
the points (x, 0) and (x, 2n) are identified (Fig. 77)].
The torus can also be viewed as the set of cosets of the group R2 modulo,
the subgroup 2nJ.2 of vectors with integral components multiplied by 2n:
(2n
Zrt
Figure 77.
98 3. Structural Stability
‘ o o o
o o o
Figure 78.
Every vector field on the torus defines a field on the plane periodic of period
2n in both coordinates. Conversely, to every field 2n periodic in both
coordinates in the plane, there corresponds a vector field on the torus.
Example
The equations
x = a, y = P,
where a and P are constant, define a vector field on the torus without singular
points.
Theorem. If the ratio X = P/a is rational, then all phase curves are closed
on the torus. If it is irrational, then they are everywhere dense.
•4 (1) Let X = p/q. The phase curve passing through the point (x0, >»0) has
the equation — y0 = (x — x0)p/q. Ifx — x0 = 2nq, then>> — y0 —
2np and, consequently, (x, y) — (x0, y0) mod 27r, i.e., the phase curve
is closed.
§11. Differential Equations on the Torus 99
(2) Below we prove that (in the case where X is irrational) any phase curve
is uniformly distributed on the torus, i.e., in any part of the torus*,
it spends a time proportional to the area of that part. This implies
that, in particular, a sufficiently long segment of the phase curve is
as close to any point of the torus as we wish, i.e., the phase curve is
everywhere dense.
C. Uniform Distribution
t (A T, z)
lim
T-co T MAO'
is called the time average of the function / (here gf is the phase flow).
Remark. Of course, this limit does not always exist, and even if it does,
it may, in general, depend on the initial point.
The theorem on the uniform distribution on the torus follows from the
theorem below.
*By “a part of the torus” we mean a Jordan measurable domain, for example, a domain
with piecewise smooth boundary.
100 3. Structural Stability
fdx dy.
We denote the vector with components (a, (I) by co. Then the solution with
initial condition z has the form z + cot. The assertion of the theorem reads
as follows:
/(z + oot) dt = ^
We note that among the functions on the torus, there are the harmonics
of the form e'(fc,z), where A: is a vector with integral components. For har¬
monics, the theorem can be verified by direct calculation of the integral. Let
k ^ 0. We have
pH"-.-*;
i(k,z + a>t) ,i(k,o)t Jt _
dt = el * = TTTTTTn [el(k,a>)r - 1],
i{k, oo)
\p - P\ < e, \q - Q\ < e.
1 CT
pT(z) = — p(z + cot) dt, etc.
Jo
Then pT(z) — e < fT(z) < qT(z) + e for any T. For T sufficiently large, we
have
E. Some Consequences
1. The fact that the torus was two-dimensional did not play any role in
the preceding. We consider the equation z = co,z e T", on the «-dimensional
torus. The frequency vector co is said to be a resonant vector if there exists
a nonzero vector k with integral components such that (co, k) = 0.
If the vector co is nonresonant, then the time and space averages of con¬
tinuous (or even Riemann integrable) functions coincide and the solutions
are uniformly distributed.
2. From the uniform distribution theorem, it follows that the first digit
of the number 2" is equal to 7 more often than to 8. More precisely, denote
by Nk(n) the number of positive integers m ^ n for which 2m begins with
the digit k. Then
The same distribution governs the first digits of the population figures
of the world’s countries, but not those of the lengths of rivers (N. N.
Konstantinov). One possible explanation is the exponential growth of popu¬
lations with time. Such growth implies the desired distribution for the popula¬
tion figures of every fixed country in different years. To obtain the required
explanation we have only to substitute temporal averaging for “spatial”
averaging (averaging over the set of countries).
/«) = £ ake“°-'.
k= 1
We give the answer for the nonresonant vectors co = (<ol, - - - , con). Let
n = 3. If a triangle can be constructed from the three line segments (aj,
a2, a3), then co = ^ockcok/n, where ak is the angle opposite ak.
For an arbitrary n, the solution has the form of a weighted average of
the frequencies tok: co = £ Wkcok. The weights Wk can be calculated in the
following way. Denote by fV(at, ■ ■ ■, as; b) the probability of the event
that the distance from the origin to the endpoint of the planar polygon
with .v sides of lengths ax, ■ ■ ■, as with random directions is smaller than b.
We then have Wk = fV(ak, ak) (ak is the collection of all numbers at except
ak).
For a proof, cf., for example, H. Weyl, Mean motion, Am. J. Math.
60, (1938), 889-896; 61 (1939), 143-148.
Lagrange encountered the above problem (the so-called problem of
mean motion) in the following way. Consider the vector connecting the
Sun with the center of the ellipse on which a planet moves (it is called the
Laplace vector). In the first approximation of the perturbation theory, the
evolution of the Laplace vector under the influence of mutual attraction of
planets has the form of a motion of a sum of uniformly rotating vectors
(their number equals the number of planets).
If for the planets of the solar system we calculate the frequencies cok
and amplitudes ak, it turns out that for all planets except the Earth and
Venus one of the amplitudes ak is greater than the sum of the others. There¬
fore, Lagrange managed to determine the mean motion of the perihelia
of all planets except the Earth and Venus. In the case of the Earth and
Venus, however, several terms have approximately the same amplitude. The
problem was solved only in the Twentieth Century by Bohl, Sierpinski, and
H. Weyl.
on the torus. Assume that the field to has no singular points and oq # 0.
[If there are no singular points or cycles, then in an appropriate coordinate
system, the first component of the field is everywhere different from zero
(cf., Siegel, Note on differential equations on the torus, Ann. Math. 46,
3 (1945), 423-428); it is easy to construct a field without singular points
but with cycles which do not admit such a coordinate system].
Now we pass to the study of integral curves of a nonautonomous equation
with doubly periodic right-hand side:
dx w1
All solutions of this equation can be infinitely extended, since the right-
hand side is bounded.
G. Rotation Number
Definition. The rotation number of the equation dy/dx = X{x, y) on the torus
is the number
(p(x)
^ = lim
X-*OQ X
Theorem. The limit in the definition of the rotation number exists and does
not depend on the initial point; it is rational if and only if some power of the
diffeomorphism has a fixed point (i.e., the differential equation has a closed
phase curve).
Indeed, the inequality holds for \yt — y2\ < 2n, since the mappings A
and Ak of the line convert intervals of length 2n into intervals of length 2n.
§11. Differential Equations on the Torus 105
au(y) _ "h
2nkl k
Indeed, |ak(y) — 2nmk\ < 4n for any y according to § 11G1, and therefore,
ak(y) _
2nk k
3. Denote by <jk the interval [(mk — 2)/k, (mk + 2)/&]. We have proved
that akl{y)/2nkl belongs to <Jk for all l. We prove that the intervals ak with
distinct k intersect each other.
Indeed, akl{y)j2nkl belongs to both ak and cr,.
5. Assume that Aq has a fixed pointy on the circle; then the corresponding
point on the line translates by an integral multiple of2n under g-fold applica¬
tion of the corresponding mapping, i.e., aq(y) — 2np. In this case for any /
we have aql(y) = 2npl, and, therefore, the rotation number p = p/q is
rational.
6. Let p = pjq. If for all y we have aq(y) > 2np, then for some £ > 0
we have aq(y) > 2np + e for all y.
Then p > pfq. If we had aq(y) < 2np for all y, then we would have p <
pjq. Hence, aa — 2np changes sign. Consequently, y exists such that aAy) =
2np. ►>
Remark. If the rotation number p is irrational, then the order of the points
(y. Ay, A2y, .. . , ANy) on the circle is the same for any y as in the case of a
rotation by angle 2np. Indeed, aq(y) > 2np if and only if p > pjq.
106 3. Structural Stability
^ This theorem follows from the analogous assertion proved below for
orientation-preserving diffeomorphisms of the circle. ^
Remark. The derivatives of Aq are similar at the points of the same cycle
and, therefore, all points of a cycle are degenerate or nondegenerate at the
same time.
*A. G. Maier, Robust transformations of the circle into the circle, Ucenye Zapiski GGU,
12 (1939), 215-229; V. A. Pliss, On the robustness of differential equations given on the torus,
Vestnik LGU, Ser. Mat. 13, 3 (1960), 15-23.
§11. Differential Equations on the Torus 107
M I- The ordering of the points . . ., A~xy, y. Ay, ... of the orbit of the
mapping A on the circle is the same as the ordering of the points of the orbit
of rotation by the angle 2np (cf., § 11G). Therefore, to prove the theorem it
is sufficient to establish that the orbit of A is everywhere dense on the circle.
Indeed, we then would obtain a homeomorphism of the circle converting A
to a rotation by continuously extending the mapping which takes the points
of the orbit . . . , A~xy, y. Ay, . . . into the corresponding points of the orbit
of the rotation.
2. If on the circle there is an arc free from points of the orbit of A, then
the images of this arc under the powers of A are mutually disjoint. Indeed,
consider the maximal arc containing the given arc and free from points of
the orbit. Its images are also maximal arcs. The endpoints of a maximal arc
belong to the closure of the orbit. Therefore, the endpoints of a maximal
arc cannot lie inside maximal arcs. Hence, any two intersecting maximal
arcs must coincide. On the other hand, if a maximal arc coincides with its
image, then its boundary point belongs to a cycle, in contridiction with the
irrationality of p.
un = in
=o
-r; W’
ay vn = n -z—0*
1=0ay ‘y)>
we have
uN dy -► 0, vNdy -» 0
Indeed, consider the arc (as, aq+s), s < q, of length <5, the distance of
a0 from aq. Assume that ar lies on this arc. If r < s, then a0 lies on the arc
(as_r, <x5_r+q). Therefore, the distance from as_r to a0 is less than S, contrary
to the choice of aq. If r > s, then ar_s lies on the arc (a0, aq), and therefore,
r — s > q. Then the distance from a0 to car_s_q is less than <5. Hence, the
arc (as, <xq+s) contains no points ar, r < 2q, which was to be proved.
5. Consider the points (y. Ay, . . . , Aq~ly) and (A~ly, . . . , A~qy). These
two sets of points alternate (§ 1114). Therefore, for any function/of bounded
variation defined on the circle, for any point y, and for any q defined in
§ 1114, the quantity
H n
is bounded from below and above by positive constants not depending on y
or q (if q is chosen as in § 1114).
s[uo»qdy uq dy vqdy. ►
and so on for all cycles (the order of points of a cycle on the circle is the
same as for the rotation). This correspondence can then be extended to
adjacent intervals using the following lemma, which can be proved easily:
Any two homeomorphisms of an interval onto itself without fixed points
are topologically conjugate.
2. If the rotation number is rational and all cycles are nondegenerate, the
rotation number, the number of cycles, and the nondegeneracy of cycles are
preserved under a small perturbation (according to the implicit function
theorem). Consequently, a diffeomorphism with rational rotation number and
nondegenerate cycles is structurally stable (cf., § 11J1).
K. Discussion
tions. If the rotation number is rational, then the invariant measure is con¬
centrated on separate points. If the rotation number can be approximated
very rapidly by rational numbers with not too large denominators, then the
invariant measure can be approximated so rapidly by measures concentrated
on separate points that it cannot be even absolutely continuous with respect
to Lebesgue measure. Therefore, the homeomorphisms in Denjoy’s theorem
cannot be replaced by diffeomorphisms.
C
>
2+e
q
for any integers p, q > 0. This gives rise to the conjecture that the phenom¬
enon in § 11K2 occurs with probability 0. We formulate two results in this
direction.
Theorem. For almost every rotation number p, every sufficiently smooth (of
class C3 or smoother) dijfeomorphism of the circle with rotation number p is
smoothly equivalent to the rotation by the angle 2np (Herman, 1976).
Here “almost every” means that the Lebesgue measure of the exceptional
set of rotation numbers is equal to zero.
Herman’s theorem was preceded by an analogous theorem for mappings
close to a rotation and by the following result (proved in the analytic case
in 1959* and in the smooth case by J. Moser in 1962).
Figure 81.
Theorem. For any irrational number p, there exist arbitrarily accurate rational
approximations whose error is less than the reciprocal value of the square of
the denominator:
The vectors ek are the vertices of our two convex hulls (the upper one
for even k and the lower one for odd k).
A For the initial parallelogram (e0, e_x), this is evident. Every following
parallelogram has a common side with the preceding one and equal altitude,
and gives an opposite orientation of the plane. ►
Corollary. Denote by qk andpk the coordinates of the point ek. The difference
of two subsequent convergent fractions is equal to
Pk _ Pk±i = (-Pft
Qk Qk +1 QkQk+1
Proof of Theorem. The vectors ek lie alternately on one or the other side of
the line y = px.
Therefore, the convergent fractions are alternately larger or smaller than
p. Consequently, the difference between p and a convergent fraction is
smaller than the modulus of the difference between the convergent fraction
and the subsequent convergent fraction. By the corollary, the absolute value
of this difference is equal to \/qkqk+i, which is not larger than l/qk, since
Qk+1 > Qk for A: ^ 0. ►
114 3. Structural Stability
Remark. The numbers ak are called partial quotients. The convergent frac¬
tions can be expressed in terms of the partial quotients in the following way:
— = a0 +
ai +
1
H-
ak-1
Denote by Ylp the strip |Imy| < p. For a holomorphic function a bounded in this
strip, we shall write
Let p be an irrational number, and let K > 0, a > 0. We say that p is a number of
type (AT, a) if for any integers p and q ^ 0,
Theorem. There exists e > 0 depending only on K, p, and a such that if a is a 2n-periodic
analytic function, real on the real axis with ||«||p < e and such that the transformation
def ines a diffeomorphism of the circle with rotation number p of type (K, <r), then this
diffeomorphism is analytically equivalent to the rotation by the angle 2np.
B. Homological Equation
Denote by 91 the rotation by the angle 2np and by H the desired diffeomorphism
converting the rotation into A. The following diagram is commutative:
§ 12. Analytic Reduction of Analytic Circle Diffeomorphisms to a Rotation 115
S1 A S'
«T T H, i.e., //■ 91 = A H.
S' 4 s'
We write H in the form Hz = z + h(z), h(z + 2n) = /j(z). For h, we obtain the
functional equation
h(z + 2npi) — /z(z) = a(z + h(z)).
We expand the given function a and the unknown function h in the Fourier series:
hk = - —-
g2nikfi _ J '
For the solvability of the equation, it is necessary that the denominators vanish only
together with the numerators. In particular, the homological equation is not solvable
if a0 # 0. If a0 = 0 and the rotation number // is irrational, then the preceding formulas
give a solution of the homological equation in the class of formal Fourier series. In
order to obtain the actual solution, it is necessary to study the convergence of the
series.
geometric progression:
.41 Me~wp.
As is known, fk = ~~ (j) f(z)e ,kz dz. Let k > 0. Let us shift the path of integration
down (by —ip). The integral does not change, since the integrals on the vertical sides
of the rectangle thus obtained are equal. Hence,
2n
1
fk f(x - ip)e~ikx-kp dx, \fk\ 5= Me~kp.
,k ~ 2n
For k < 0 the path of integration has to be shifted upward (by ip). ►
Lemma 2. If\fk\ Me~Wp, then the function f = ~Zfke,kz is analytic in the strip FIP and
4M
/ llp-a ^
for S < p, S < 1.
Remark. In the case of functions of n variables. Lemma 1 still holds. In Lemma 2, the
estimate 4M/d has to be replaced by CM/S", where C = C(n) is a constant independent
of 6 andf
E. Small Denominators
In solving the homological equation, the Fourier coefficients of the right-hand side
have to be divided by the numbers e2n‘kp — 1. If p is irrational, then for k # 0, these
numbers are different from 0. Nevertheless, some of them are very close to 0. Indeed,
every number p admits rational approximations pjq with error |p — (p/q)\ < \/q2
with arbitrarily large q. For k = q, the denominator e2"'ktl — 1 will be very small.
It turns out that, with probability 1, all these small denominators admit lower
estimates in terms of powers of k.
Lemma 3. Let a > 0. For almost every real p there exists K = K(p, a) > 0 such that
P K
P :>
2 + <r
q q
Consider those numbers p in the interval [0, 1], for which the above inequality
(with fixed p, q, K, and a) is violated. These numbers form an interval of length not
greater than 2K/q1+a. The union of these intervals for all p (for fixed q > 0, K, and cr)
has total length not greater than 2K/q1+a. By summing over q, we obtain a set of
measure not greater than CK, where C = 2Lq~{X+a) < x. Consequently, the set of
numbers pe [0, 1] for which the desired AT does not exist is covered by sets of arbitrarily
§ 12. Analytic Reduction of Analytic Circle Diffeomorphisms to a Rotation 117
small measure. Consequently, this set is of measure zero (on the interval [0, 1] and,
therefore, on the entire line).
Remark. The numbers p satisfying the above inequality are called numbers of type
(AT, a) in § 12A.
For a number p of type (K, a), a small denominator admits the following lower estimate:
Indeed, the distance of kp from the closest integer can be estimated from below by
the number K/\k\i+a \ a chord of the unit circle is not shorter than the length of the
small arc subtended by it divided by n. ►
Lemma 4. For almost all p, the homological equation has a 2n-periodic analytic solution
(which is real if a is a real function). There exists a constant v = v(K, a) > 0 such that
if p is of type (K, a), then for any 6 > 0 smaller than p and any p < f we have
We solve the homological equation with right-hand side a = a — a0 (a0 is the mean
value of the function a). Denote by h° the solution. We define a mapping H0 by the
formula H0(z) = z + h°(z). We construct the mapping Ax = Hq 1 o A o H0 and
define a function a1 by the relation A^z = z + 2np + al(z).
118 3. Structural Stability
A 2 = Hf1 o A1 o Hl.
Lemma 5. There exist constants x, A > 0 depending only on K and a such that for every
6 in the interval (0, p), where p < \, we have
^ /. Let be a convex domain in C" (or IR") and let h: Q -* C" (or IR") be a smooth
mapping with ||/j#|| = suPxea||^*Cx)|| < 1. Then the mapping H taking x into x + h{x)
is a diffeomorphism ofQ onto H O..
The eigenvalues of //*(*) are different from 0; therefore, H is a local diffeomor-
Figure 82.
§12. Analytic Reduction of Analytic Circle Diffeomorphisms to a Rotation 119
phism. In view of the condition |/j J < q < 1 and the convexity of Q, h is a contraction.
Consequently, the difference between the displacements of any two distinct points under
the mapping H is smaller than the distance between these points and, therefore, their
images are distinct, i.e., H is one-to-one. &■
The expression within the first pair of square brackets can be estimated using the mean
value theorem and the Cauchy inequality. By § 12H2, we obtain
Hence,
4. Now estimate the quantity |a0| using the fact that the rotation number of A and,
therefore, of Ax is equal to 2np.
◄ From this it follows that a1 vanishes at some real point z0. We substitute the value
z0 into the formula for al(z). We obtain aQ = a(z0) — a(z0 + /z°(z0)), and, conse¬
quently, |a0| ^ M23~u (cf., § 12H3). ►
120 3. Structural Stability
1. We shall consider the mapping An constructed in the «th step in a strip of radius
P„ decreasing with every step: p0 = p, p„ = p„_, -
The sequence of the numbers is chosen decreasing in the following way:
f’n = <2-
M„ = <
A sufficiently large number TV (depending only on K and a) will finally be chosen below.
Note that Mn = M*L\ -
provided that TV > 2a. We choose TV larger than 2A and x. Then we obtain
IK l, =
c = no + <**).<• = no -
This implies that is a diffeomorphism of flp and that the sequence #„ converges
in np/2. Indeed,
Denote by H the limit of the sequence Passing to the limit in A o J^n = o A„,
we obtain A o = .jf o 'it, where 91 is the rotation by 2np. The theorem is proved.
J. Remarks
1. Moser observed that an analogous theorem can be proved in the case of finite
smoothness by combining the above approximation with a smoothing process of Nash
(cf., J. Moser, A rapidly converging iteration method and nonlinear partial differentia!
equations, 1, Ann. Scuola Norm. Sup. Pisa (3), 20 (1966), 265-316, II, 20 (1966), 499).
In the first publications of Moser, hundreds of derivatives were required. Subsequent
efforts of Moser and Riissmann reduced the number of derivatives (H. Riissmann,
Kleine Nenner II: Bemerkungen zur Newtonischen Methode, Nachr. Acad. Wiss.
Gottingen, Math. Phys. Klasse 1 (1972), 1-10). See also the recent works by Sinai,
Khanin and others (1987).
Figure 83.
1. Fix an integer TV. Then the points of the torus whose coordinates are
rational multiples of 2n with denominator TV form a finite set. The mapping
A maps this set onto itself. Consequently, all points of this set belong to
cycles.
mes(T"F) D G mes G
lim
n—*co mes F mes T2
§ 13. Introduction to the Hyperbolic Theory 123
lim (An*fi g) =
n-co (1, 1)
Theorem 3. On the torus there exist two direction fields invariant with respect
to the automorphism A. The integral curves of each of these direction fields
are everywhere dense on the torus. The automorphism A converts the integral
curves of each field into integral curves of the same field, expanding by X > 1
for the first field and contracting by X for the second (Fig. 84).
R2 4 (R2
H | |H
[R2 4 [R2
§13. Introduction to the Hyperbolic Theory 125
commutative, where B{x) - A(x) + f(x) and the functions / and h are
27t-periodic in x.
From the diagram, we obtain the following nonlinear functional equation
for h:
We assume that f is small and that h turns out to be small of the same
order. Therefore, we replace the right-hand side by /(x), omitting a
“quadratically small quantity”. We obtain the following linearized equation:
Lemma 1. The space of vector fields on the torus splits into the direct sum of
two subspaces invariant with respect to L.
The spaces of vector fields parallel to the first and second eigendirections
of the operator A are invariant under the transformation, and every vector
field can be represented uniquely as the sum of two fields in the eigen¬
directions.
Let / = fel + f2e 2, h = h1el + h2e 2, be the decompositions of the
fields f and h. Then the homological equation takes the form
Let / = 1. Then
||(S - A,£)->|| «
Analogously,
Consequently, ZC1 exists and ||L 1/(1 — X). The homological equation
is solved, fe-
The nonlinear functional equation from § 13D can now be solved by the
simple contraction mapping principle. We set
Lemma 2. If the C1 -norm off is sufficiently small, then the operator L-10 is a
contraction in the space C°.
<4 It is sufficient to verify that the nonlinear operator <I> satisfies a Lipschitz
condition with a small constant. Indeed, according to E we have
(1) The tangent space of M is decomposed into the direct sum of two
subspaces at every point of M:
TXM = Xx 0 Yx.
(2) The fields of planes X = {Xx} and Y = { Yx} are continuous and
invariant with respect to A.
(3) For some Riemannian metric, A contracts the planes of the first
field and expands the planes of the second field : there exists a number
X < 1 such that for any point x of M,
Example
AJVl2 1
1
If this condition is satisfied for one metric, then it is satisfied for any other
(possibly with a different c). This condition also implies the above inequalities
(possibly in a modified metric).
Remark 2. The definition does not require smoothness of the fields X and Y
of planes. A torus diffeomorphism close to the automorphism
2 1
l 1
The proof can be carried out following the method of § 13 for auto¬
morphisms of the torus; the details can be found, for example, in J. Mather,
Anosov diffeomorphisms. Bull. Am. Math Soc. 73 (1967), 747-817 (ap¬
pendix).
The first proof was related to the following property of Anosov diffeo-
rriorphisms.
that the Frobenius theorem cannot be used, since our Fields of planes are
not smooth.
The proof is based on the observation that under the diffeomorphism
the angle between planes not too different from the planes of the expanding
Field decreases: the expanding field is an attracting fixed point in the space
of fields of planes under the action of the Anosov diffeomorphism on this
space.
In order to construct the expanding foliation, we can partition the mani¬
fold into sufFiciently small domains and take, in every domain, an arbitrary
foliation whose leaves have dimension equal to that of the planes of the
expanding Field and form a not too large an angle with these planes. We apply
the Anosov diffeomorphism and its iterates to these foliations. It turns
out that the sequence of partial foliations thus obtained converges to the
true expanding foliation.
Remark. A special case of the above construction is the construction of stable and
unstable invariant manifolds of a Fixed point of a diffeomorphism in the case where the
absolute values of the eigenvalues of the linear part of the diffeomorphism are all
different from 1. For the construction of the unstable manifold, we may apply the
iterates of the diffeomorphism to any manifold tangent to the unstable invariant
subspace of the linear part of the diffeomorphism.
bo = 0 aia2 -anBn.
n~> + go
In the same way, there exists a unique horizontal perturbed disk whose
images under the negative powers of our diffeomorphism do not leave the
neighborhoods with negative indices.
The intersection of the perturbed horizontal and vertical disks thus
constructed defines the point which is assigned by the conjugating homeo-
morphism to the initial phase point.
It is not difficult to verify the fact that this construction actually defines
a homeomorphism conjugating the unperturbed Anosov diffeomorphism
with the perturbed one.
Anosov diffeomorphisms with invariant measure given by a positive
density have an everywhere dense set of periodic points (cycles). Quite
complete study of the ergodic (mixing, etc.) properties of Anosov diffeo¬
morphisms with invariant measure was carried out by Anosov and J. G.
Sinai (cf., D. V., Anosov, Geodesic flows on closed Riemannian manifolds
of negative curvature, Trudy Steklov, 90 (1967), 3-209: Proceedings of the
Steklov Institute of Mathematics, Petrovskii and Mikolskii, ed., American
Mathematical Society, 1969, pp. 1-235.
C. Anosov Flows
Figure 85.
132 3. Structural Stability
TXM = Xx © Yx © Zx.
(2) The fields X, Y, and Z of planes are continuous and invariant with
respect to the phase flow.
(3) The field Z is generated by the field of phase velocity.
(4) For some positive constants c, A, and for some Riemannian metric on
M we have
Then the phase flow is called an Anosov jlow and the equation x = v(x)
an Anosov system.
Example
Consider a vector field directed along the factor [0, 1] in the direct
product T2 x [0,1]. Upon gluing boundaries in the direct product, this
field turns into a smooth (why?) field v on M.
The field v thus obtained defines an Anosov flow on M.
Some Anosov flows have infinite sets of closed phase curves. Conse¬
quently, even confined to structurally stable vector fields, in the multi¬
dimensional case we cannot expect to obtain as simple a picture with a
finite number of equilibrium positions and cycles as in the case of systems
on the two-dimensional sphere.
In 1961, Smale constructed the first examples of structurally stable systems
with an infinite number of cycles. In these examples, an exponential diver¬
gence takes place not on the entire phase space, but on a closed subset.
Such sets are now called hyperbolic. The general theory of hyperbolic sets
was created later, under the influence of the theory of Anosov systems.
The appearance of similar examples led to a sharp change in the inter¬
pretation of the behavior of phase curves of multidimensional systems.
Some specialists hastened to announce these results as not having any real
meaning, since such systems, however structurally stable they are, “cannot
describe any real physical processes” in view of the instability of the individ¬
ual trajectories.
Nevertheless, there are very important real cases where systems with
exponential divergence of trajectories apparently describe reality best. We
speak of the mathematical description of phenomena of the type of tur¬
bulence and of the motion of colliding particles (say, in models of a gas
consisting of rigid spheres). Simpler, but quite real, is the problem of motion
along geodesics of manifolds of negative curvature. We are now going to
analyze the simplest version of this problem, the problem of geodesics on
surfaces of constant negative curvature. To do this, we need some informa¬
tion about Lobachevsky geometry.
The Lobachevsky plane is the upper half-plane Imz > 0 with the metric*
2 dx2 + dy2
ds = --2->
y
where z = x -\- iy. The straight line y — 0 is called the absolute. We note
that angles in this metric coincide with Euclidean angles, and the distance
from the absolute is infinite.
Theorem. The geodesics of the Lobachevsky plane are all circles and straight
lines orthogonal to the absolute {Fig. 86).
symmetries z i—> — z (this is obvious). It is easy to verify that (4) the metric
is also invariant under the inversion z —> 1/z.
It follows from (l)-(4) that the metric is invariant under all fractional
linear transformations of the upper half-plane into itself. Besides, it follows
from (3) that the y-axis is a geodesic. On the other hand, the 7-axis can be
converted into any circle or straight line orthogonal to the absolute line by
a real fractional linear transformation. Consequently, they all are geodesics.
Conversely, a circle or a straight line orthogonal to the absolute passes
through every point in every direction. Consequently, there are no other
geodesics. ^
Remark. At the same time, we have proved that the motions (preserving
the metric and orientation) of the Lobachevsky plane are the fractional
linear transformations of the upper half-plane into itself.
Theorem. The circles of the Lobachevsky plane are all Euclidean circles not
intersecting the absolute.
M Consider the unit disk. The upper half-plane can be transformed into the
unit disk by a fractional linear transformation (cf., Chap. 1, § 5E). Therefore,
the interior of the unit disk can be regarded as a model of the Lobachevsky
plane (Fig. 34).
Then a fractional linear transformation preserving the upper half-plane
turns into a fractional linear transformation leaving the unit disk fixed.
Therefore, the metric of the Lobachevsky plane in the model on the disk is
invariant with respect to all fractional linear transformations of the disk
onto itself.
On the other hand, among these transformations are the rotations around
the center. Consequently, all points of a Euclidean circle with the same center
as the unit disk are at a constant distance from the center in the sense of the
Lobachevsky metric. Hence, a Euclidean circle is a Lobachevsky circle
provided that its center is at the center of the disk.
On the other hand, any Euclidean circle not intersecting the absolute can
be converted into a Euclidean circle with center at the origin by a motion
of the Lobachevsky plane. Consequently, every Euclidean circle not inter¬
secting the absolute is a circle in the sense of the Lobachevsky metric (in
§ 14. Anosov Systems 135
Figure 87.
both the model on the disk and the model in the half-plane). From this, it
follows that, conversely, every Lobachevsky circle is a Euclidean circle.
Theorem. The horocycles of the Lobachevsky plane are exactly the Euclidean
circles and straight lines tangent to the absolute.
Remark 1. By the same passage to the limit t —* oo, we can construct horo¬
cycles on surfaces of negative curvature and horospheres on manifolds of
negative curvature.
Remark 2. There are two horocycles with common tangent passing through
every point of the Lobachevsky plane; they are obtained from the above
construction as t —*■ +oo and as t -* — oo.
2. Conditions (2) and (3) express the invariance of geodesics and horo-
cycles with respect to the geodesic flow and can be verified directly. Indeed,
the family of geodesics perpendicular to a given horocycle intersects the
absolute at the point of tangency of the absolute with the horocycle, and
every horocycle tangent to the absolute at this point is orthogonal to all
geodesics of the family (Fig. 89).
Therefore, the geodesic flow converts every horocycle (equipped with the
field of normal vectors) into the horocycle tangent to the absolute at the
same point (and also equipped with the field of normal vectors).
Figure 90.
F. Billiard Systems
Presently, Anosov systems and related objects are in the same situation
as limit cycles were in the time of Poincare. The whole mathematical ap¬
paratus for the study of limit cycles had been created, but serious application
of limit cycles in engineering began only several decades later, when the
progress of radio technology turned the theory of nonlinear oscillations
into a field of applied mathematics.
Since the beginning of the 1960’s, it has been conjectured that a natural
area of application of Anosov flows is in the theory of turbulent motions
of a fluid. Let us imagine a closed vessel filled with an incompressible viscous
fluid brought into motion by some exterior force (stirring). Stirring is
necessary; otherwise viscosity would damp motion over time.
The Navier-Stokes equations of hydrodynamics define a dynamical
system* in a function space (the points of this infinite-dimensional phase
space are divergence free vector fields, the fields of the velocities of the
fluid).
The equilibrium positions of this dynamical system are stationary fields
of velocities, i.e., those motions of the fluid for which the velocity does not
change over time at each point of the space. To the cycles of the system
there correspond periodic motions of the fluid, in which the velocity changes
periodically at every point of the space. Such a motion may be observed
by turning on a water faucet.
The conjecture concerning the mathematical description of turbulence
is based on the fact that the phenomenon essentially reduces to a finite¬
dimensional dynamical system, since viscosity extinguishes high harmonics
rapidly. In other words, it is assumed that in the infinite-dimensional
phase space, a finite-dimensional manifold or set exists to which all phase
curves are attracted; on this very set, the phase flow is an Anosov system
or has the similar properties of exponential instability of trajectories and
mixing.
In this case, the observable properties of the motion of the fluid have to
be as follows: under any initial condition, the motion rapidly assumes a
limiting behavior; however, this behavior is neither stationary nor periodic;
although the limiting motion is in fact determined by a finite number of
parameters (“phase” of the limiting behavior), the parameters themselves
are highly unstable (limiting flows with close initial phases diverge ex¬
ponentially) ; actually, the statistical characteristics of the flow do not depend
on these unstable phases.
The following results have so far been obtained in this direction. If the
viscosity is large enough, then the Navier-Stokes system has a unique
* Actually, the theory of partial differential equations has not been able to solve the problem
of existence and uniqueness of solutions of the three-dimensional Navier-Stokes equations.
However, we will ignore this circumstance for now.
§ 15. Structurally Stable Systems Are Not Everywhere Dense 141
fixed point which attracts all phase curves. This is the so-called laminar
motion. Every other flow tends to turn into a laminar flow under the effect
of viscosity. With the decrease of viscosity, the laminar flow may loose
stability. Then a stable limit cycle appears (cf., chap. 6). With further
decrease of viscosity, the cycle may lose its stability and give rise to a more
complicated aperiodic motion which attracts its neighbors. It is expected
that this motion will, in general, have the properties of exponential insta¬
bility of phase curves on the attracting set. Although much theoretical and
experimental work has been devoted to this question in the past (cf., for
example, the survey of J. B. McLaughlin, P. C. Martin, Transition to tur¬
bulence of a statically stressedfluid system, Phys. Rev. A 12 (1975), 186—203),
the above conjecture is far from being a theorem.
It should be noted, however, that the appearance of the attracting set
with exponentially unstable trajectories is not necessarily connected with
the loss of stability of the laminar flow; this set may occur far from the
equilibrium position and even for the values of viscosity for which the
laminar flow is still stable.
A. Smale’s Example
A(x,y; z) 2x + y, x + y; -
In the neighborhood of the point O with coordinates (0, 0, 7r), the mapping
A is given by the formula
A(x, y; 7i -f m) = ^ ; 7t + 2uj.
2. From what has already been proved, it follows that A has the invariant
manifold T2 close to T2 and on it a countable everywhere dense set of
periodic points. Through every point of the neighborhood U of the torus
T2 there passes a uniquely determined smooth leaf of the two-dimensional
contracting foliation of A, which depends continuously on the point (it
consists of the points which approach each other under the iterates of the
diffeomorphism).
D. Structural Instability
Perturbation Theory
In this section, we describe the recipe of the averaging method in its simplest
variant. The problems of the justification of this method are discussed in the
following subsections.
f ir+e u.
TL
B
Figure 93.
The functions on the base B of the fibering n determine first integrals of the
equation x = v(x) on M. The vertical vector field v is said to be unperturbed.
A perturbed field is, by definition, a field v 4- £t>, close to v. Consider the
following perturbed differential equation:
JC = v(x) +
Example
Consider the planetary system. The unperturbed equations only take into
account the interaction of the Sun and the planets. In the unperturbed
motion, the planets move along Keplerian orbits. The role of perturbation is
played by the mutual attraction of planets. The role of e is played by the ratio
of the mass of the planets to that of the Sun; this is a quantity of the order
1/1000. The characteristic unit of time is the period of revolution around the
Sun, i.e., a quantity of the order of a year or decade, and the characteristic
unit of length is the radius of a planetary orbit.
In this example, M is the phase space, the base B is the space of collections
of Keplerian ellipses, and the fibers are tori of dimension equal to the number
of planets (every collection of Keplerian ellipses determines a torus whose
point is given by indicating the positions of the planets on the ellipses). Then
a displacement on the base by a quantity of order 1 corresponds to a change
in the orbit, say, a doubling of the radius. Time on the order of 1 /e is time on
the order of thousands or tens of thousands of years.
Consequently, in this example, a systematic slow motion (drift) on the base
at velocity e could double the radius of the orbit of the Earth over a time
period on the order of thousands of years. This would be fatal to our civiliza¬
tion, which owes its existence to the fact that this drift does not occur (at least
146 4. Perturbation Theory
not in the direction of the change of radii of orbits, a change in the eccen¬
tricities takes place and apparently has effects on ice ages).
f <p = "CO,
l / = o,
where to is a vertical vector field given by a frequency vector (tol(I), . . . ,
toff)) depending on the point I of the base.
where f and g are 27i-periodic in <p and £ <^ 1 is a small parameter. The
angular coordinates tpi are called fast variables and the coordinates Ij on the
base are called slow variables.
J = eG(J),
where G(J) = §g(J, (p, 0)d(pj§dtp is the mean value of the function g on a
fiber.
The solutions of the averaged equation are called averaged motions.
§ 16. The Averaging Method 147
Example
j = ea.
F/ . r eb sin cot
I(t) = I0 + eat + -
co
which differs from the solution of the averaged equation J{t) = I0 + eat by
an oscillating small term only. ^
h
Figure 95.
Time T > 1 is large; therefore, the expression within the square brackets
is close to the time average of g.
We introduce the slow time x — et. The variation of x from 0 to 1 corre¬
sponds to the variation of t from 0 to 1 /e. We will denote by a prime the
velocity of motion with respect to the slow time. Then the preceding equation
takes the form
We replace the time average by the space average and obtain the averaged
equation
D. Discussion
The use of the averaging method consists of replacing the perturbed equation
by the much simpler averaged equation. The solutions of the averaged
equation are studied on time intervals on the order of 1/e (i.e., on intervals
of the order of 1 of the slow time). Then conclusions are drawn on the
behavior of the perturbed motion over time on the order of 1/e. (The con¬
clusion is usually that the / component of the solution of the perturbed
equation is close to the solution of the averaged equation over time 1/e.)
This conclusion does not follow from the preceding reasoning and
requires a proof. Indeed, in deducing the averaged equation, we replaced
time averages by space averages. The replacement is reasonable if the
trajectory of the unperturbed motion is uniformly distributed over the
^-dimensional torus, i.e., when the frequencies are incommensurable. In
cases of resonance, the trajectory of the unperturbed motion is everywhere
dense on a torus of dimension smaller than n. Therefore, the replacement of
the time average by space average on the ^-dimensional torus is not legitimate
near resonances.
Indeed, there are examples which show that the difference between the
projection of the perturbed trajectory on the base and the solution of the
averaged equation may be on the order of 1 over time 1/e: the averaged
drift and the projection of the true motion point in different directions.
Practically, the only case studied completely is that of single-frequency
systems where the fibers are one-dimensional tori, i.e., circles.
§ 17. Averaging in Single-Frequency Systems 149
Here we formulate and prove a theorem which justifies the averaging method
for single-frequency systems.
/ = eg(I, (p, e)
Theorem. Assume that the frequency co does not vanish in the domain B. Then
the distance between the value of the solution J(t) of the averaged equation
and the I component I(t) of the solution of the perturbed equation with 7(0) =
7(0) remains small for t e [0, T/e] if e is sufficiently small:
The basic idea of the proof of the theorem consists of trying to annihilate
the perturbation by means of an appropriate change of variables. This
idea has many applications (cf., for example; the preceding and following
chapters) and is the basis of the whole formal apparatus of perturbation
theory.
In place of 7 we choose a new coordinate P = I + eh(I, cp) such that
the P component of the solution cease to oscillate. For this we want to
150 4. Perturbation Theory
ch
P = £ g + co —
C(p
+ r.
where the argument e of the function g is replaced by zero; the remainder r
(as we shall verify below) is of second order in e.
We try to choose h so that the terms of first order in s are annihilated,
i.e., the square brackets vanish. Formally, we obtain
<p
1
h(I, <p) = g(i, <A, 0) dip.
co(7)
<p 0
(Here the condition co # 0 of the theorem has been used.) Actually, such a
method of solving the equation g + codh/cq) = 0 is not legitimate, since the
function h has to be 27r-periodic in q> for the mapping (/, cp) i—> {P, cp) to be
defined on M.
The preceding formula defines the function h on the circle (and not on a
covering line of it) only if the average value of g on the circle is equal to zero.
Consequently, the choice of h does not allow us to annihilate the whole
perturbation g, but only its oscillating part
The average of g over the period is equal to zero, and we can define a periodic
function h by the formula
1
h(I, cp) = g(7, if/, 0)dif/. (1)
co(7)
o
P = eG(P) + eR.
j = eG(J)
by the small quantity eR of order e2. Therefore, the solutions diverge with
§ 17. Averaging in Single-Frequency Systems 151
velocity of order e2. Consequently, over time 1/e, they diverge to a distance
of order e. The difference between P and / is also of order e. Therefore, the
distance between I(t) and J(t) remains of order e over a time period of order
l/e.
To prove this assertion, one needs (simple) estimates of the terms that
were omitted above.
C. Estimates
2. We prove that the mapping A: (/, ip) i—► (P, ip) is a diffeomorphism of K x S1
for suff iciently small e.
It follows from the definition of h (Eq. 1) that he C]. Consequently, |e/r[, < 1
for sufficiently small e. If two points were mapped onto one by A, then the difference
of the values of eh at these points would be equal to the difference of the values of /;
this contradicts the inequality \eh\, < 1, since K is convex. It also follows from |e/t |, < 1
that A is a local diffeomorphism. Hence A is a diffeomorphism. $>■
3. Estimation of R. We have
/?, = g(I, tp, 0) - g(P{l, ip, e), cp, 0), R2 = g(l, (p, e) — g(J, <p, 0),
P = I + eh(I, ip),
we obtain
The norms of f g, and h appearing here can be estimated by r,. Finally, if / and
P(I, cp, e) belong to K, then
r = GiJ).
\Z \ ^ a\Z\ + b,
for et ^ T as long as lit). Pit) = Pfit), (pit), £), and ./(/) remain in K.
We denote by p the distance from the trajectory of the averaged motion {-/(/),
et ^ T} to the boundary of K. If (c3 + c4)e < p, then, according to the preceding
estimates, lit). Pit), and Jit) cannot go out to the boundary of K for Et T. Over
all this time interval we obtain
D. Example
x = —x + e(l — x2)x.
Figure 96.
The case of several frequencies is studied much less than that of a single
frequency. In this section, we give a survey of the basic results in this area.
A. Resonance Surfaces
Tm = co(/)) - 0}
Example
01 = 7j , (f>2 = I2 ’ / = 0
Example
with three frequencies. Here B is the plane with coordinates f and /2, and
the resonance surfaces are all straight lines given by rational equations.
In this case, during the motion of the point I in the plane, the point /,
even if it intersects all resonance curves transversally, will always intersect
many of them at small angles, because for any linear element there exists
a linear element of a resonance curve arbitrarily close to it (Fig. 98).
§ 18. Averaging in Systems with Several Frequencies 155
Figure 98.
Remark. What has been said above may become clearer if we consider the
following mapping of the base into the projective (n — l)-space:
Example
as is easily seen. In the averaged system, J2 does not change over time.
156 4. Perturbation Theory
Example
7, = e, J2 = 0.
7,(0) = 1, 72(0) = 1,
gives
tion of the perturbed motion under consideration lies in the fact that
the perturbed trajectory remains on the resonance surface all the time,
where averaging is evidently not applicable, since the time average is
not close to the space average over the whole ^-dimensional torus.
Example
For the study of this system, we consider the equation (p = e{a 4- sin q> —
(p) of a pendulum with torque and friction to which it can be reduced easily.
We introduce the slow time t = yfst (t ~ 1/^/e corresponds to the interval
t ~ 1/e). Denoting by a prime the derivation with respect to slow time, we
obtain the equation
*This system is obtained from a single-frequency system by adding the trivial equation
<p2 = 1; the resonance in the system thus obtained corresponds to the vanishing of the frequency
in the single-frequency system.
158 4. Perturbation Theory
If a < 1, then the pendulum may oscillate (the loop inside the separatrix
in the phase picture). This oscillating behavior corresponds to trajectories
constantly remaining near resonance.
The effect of the small friction consists, in essence, of the destruction
of the loop of the separatrix. In its place, a narrow (of width on the order of
*J~e) strip appears in the plane (<p, (p') along the unbounded part of the
separatrix consisting of the phase points captured by an attracting equili¬
brium position; the entire domain inside the separatrix is also captured
(Fig. 100).
Returning to the initial system, we see that for a < 1 there is capture
into resonance. The resonance captures a small portion of all the tra¬
jectories. [The measure of the set of initial conditions (/, <p) captured over
time 1/e is of order >/e.] For these captured initial conditions, the difference
between the variation of the slow variable / and the variation of the solution
J of the averaged equation reaches order 1 over time 1/e.
For the remaining initial conditions (i.e., for all initial conditions outside
a set of measure of order ^/e), the difference between / and J remains small
over time 1/e (it is of order yfe In e, as can be calculated).
On the other hand, if a > 1, then capture into resonance does not occur
at all.
We consider the following system with two frequencies cof!) and co2(/):
4 = £01, 4 = £02-
Definition. The system satisfies condition A if the rate of change of the ratio aJl/(o2
of the frequencies along the trajectories of the perturbed system is everywhere different
from zero:
rco2
co,-l-q ^ co
2 cl y ~JT 0 -
§ 18. Averaging in Systems with Several Frequencies 159
The system satisfies condition A if the rate of change of the ratio oiJa>2 of the fre¬
quencies is different from zero everywhere along trajectories of the averaged system:
rcoi coj 2
co,—^ -G.
cl cl
Theorem. If the system satisfies condition A then the difference between the slow motion
I(t) in the perturbed system and J(t) in the averaged one remains small over time 1 je:
^ The proof is based on singling out a finite number of resonances with small indices
(the number of resonances is large for small £). Outside small neighborhoods of the
chosen resonance surfaces, the usual changes of variables are used (cf., § 17).
Passage through the neighborhoods of the resonances chosen leads to a divergence
of order y/e (as in the examples above).
Combining results for the divergence near the resonances chosen and the drift in
the segments between them, we obtain the estimate above. &
For details, cf., V. I. Arnold, Conditions for the applicability and estimation of the
error of the averaging method for systems which go through resonances in the process of
,
evolution, Sov. Math. Dokl. 161 1 (1965); A. I. Neistadt, On passage through resonances
,
in the problem with two frequencies, Sov. Math. Dokl. 22 2, (1975); A. I. Neistadt’s
dissertation “On some resonance problems in nonlinear systems”, Moscow University,
1975 contains a proof of the above estimate with yfe replacing the original estimate
yfe In2 s in the first reference.
Theorem (A. I. Neistadt). If the system satisfies condition A and some other condition
B (satisfied almost always), then for all initial points (70, (p0) outside a set of measure
not exceeding cI%/e, the difference between the slow motion l(t) in the perturbed system
and the motion J{t) in the averaged system remains small over time 1/e:
◄ The proof is based on the choice of a finite number of resonances with small indices.
Outside small neighborhoods of the resonance surfaces chosen, the usual changes of
variables are used.
In the study of the resonances chosen, one averages over circles which are trajec¬
tories of the unperturbed motion at the resonance.
To do this, we fix the index (m,, m2) of the resonance, where mx and m2 are relatively
prime, and choose new angular coordinates (y, (5) on the torus in place of the angular
coordinates (<pl, tp2), where y = ml<pl + m2q>2. The rate of change of the angular co¬
ordinate y in the unperturbed motion vanishes at resonance, since mxa>l + m2(o2 — 0.
For the base, we also introduce the special coordinate p = m1 cu1 + m2co2. Now
the equation of the resonance surface has the form p = 0, i.e., p characterizes the
deviation from resonance. A point of the resonance surface will be denoted by a. In
160 4. Perturbation Theory
the neighborhood of this surface, a point of the base can be characterized by the distance
p from resonance and the projection a on the resonance surface.
In these coordinates, the perturbed system takes the form
The functions Gk have period 2n with respect to y. They also depend on p and e.
We introduce the slow time t = yfet and the normalized distance r — p/y/l from
resonance. Denoting by a prime derivatives with respect to t, we write the averaged
equation in the form
This means that we obtain as a first approximation the equation of a pendulum with
torque depending on the parameter o. The Hamiltonian character of this equation is
a surprising fact, discovered through calculation and by no means obvious beforehand.
We consider the phase portrait of the equation of the first approximation in the
plane (y, r). It looks similar to the portrait in Example 3 of § 18B for a < 1 or a > 1,
depending on whether or not the function u changes sign.
It turns out that loops occur in the separatrix only for a small number of resonances
with indices which are not too large (here condition A is used). Indeed, condition A im¬
plies that the average value of the function u with respect to y is different from zero.
For resonances with large indices in the equation of the first approximation, we obtain
a function u which differs little from its average value (since in this case the average with
respect to (5 is close to the average on the torus). Therefore, it is everywhere different
from zero. This corresponds to a pendulum with external torque that is large compared
to the gravitational torque. In this case, the equation of the first approximation has
neither an equilibrium position nor a domain of oscillation.
When passing from the equation of the first approximation to the complete equa¬
tion, a loop of the separatrix changes to a zone of capture to resonance, as in Example
3 of § 18B*. The measure of the set of captured points of the phase space can be estimated
by a quantity of order yfs if all equilibrium positions of the equation of the first approx¬
imation are simple (i.e., if the zeros of u are simple: u = 0 => du/dy ^ 0). This restric¬
tion concerning simplicity is in fact condition B in Neistadt’s theorem. We note that
the condition is imposed on equations of the first approximation corresponding to a
* In contrast with Example 3 of§ 18B, in the general case, the “captured” trajectories are not
bound to remain close to resonance forever.
§ 18. Averaging in Systems with Several Frequencies 161
finite number of resonances (since under condition A the equations of the first approx¬
imation have equilibrium positions only for a finite number of resonances).
The proof of the theorem is completed by combining the estimates of the variation
of 1 in the nonresonance regions and near resonances—in the uncapturable portion
of the phase space. For details, cf., Neistadt’s dissertation cited above.
Remark. For systems with two frequencies, the case where condition A is violated has
not been studied yet. In this case the ratio of the frequencies of the fast motion does
not vary monotonically in the averaged motion. Such a behavior is not possible in the
case of a one-dimensional base: however, if the number of slow variables 1 is two or
larger, then the reversal of evolution of the ratio of the frequencies is a generic phenom¬
enon not removable by a perturbation of the system.
The case where the number of frequencies is larger than two has been
studied much less than the case of two frequencies. For generic systems, the
frequencies of the fast motion are incommensurable for almost all values
of the slow variables. Therefore, it is natural to expect that for the majority
of initial conditions, the averaging method describes the evolution of slow
variables on time intervals of order 1/e accurately.
The first general theorems in this direction are due to Anosov [D. V.
Anosov, Averaging in systems of ordinary differential equations with rapidly
oscillating solutions, Izv. A. N. USSR, Ser. Math. 24, 5 (1960), 721-742] and
Casuga [T. Casuga, On the adiabatic theorem for the hamiltonian system of
differential equations in the classical mechanics, I, II, III, Proc. Japan Acad.
37, 7 (1961)]. Anosov’s theorem asserts that for any positive number p, the
measure of the set of initial conditions (in a compactum in the phase space)
for which
passing from one to another. This is also possible if the number of frequencies
is larger than two.
It is of interest to estimate the measure of this set accurately. For example,
for systems with two frequencies Neistadt’s results (cf., § 18C) imply the
estimate |/(7) — J{t)\ ^ c2N/e|lne| outside a set of measure not greater than
clsfi (under insignificant restrictions on the system).
We assume that the frequencies are independent, i.e., the rank of doo/dl
is equal to the number of frequencies.
mesE(e, p) ^ .
P
The idea of the independence of the increments of the deviation of I from J can
apparently be justified much more completely when the fast motion is not quasi-
periodic but is an Anosov system. This is indicated, in particular, by the central limit
theorem for functions on the phase space [Ja. G. Sinai, The central limit theorem for
geodesic flows on manifolds of negative constant curvature, Sov. Math. Dokl. 133, 6
(1960), 1303-1306; M. E. Ratner, The centra! limit theorem for Anosov flows on three-
dimensional manifolds, Sov. Math. Dokl. 186, 3 (1969), 519-521]. This theorem jus¬
tifies the probabilistic approach when both the slow and fast motions are independent
of the slow variables:
However, in the case where the number of slow variables is small (smaller than the
number of frequencies minus one), even this condition cannot be satisfied.
An extension of Neistadt’s theorem to the case where the number of slow variables
is essentially smaller than that of frequencies requires a study of Diophantine approx¬
imations on submanifolds of Euclidean spaces.
For mappings
for almost all / in (Rfc, which holds for almost all points of IR".
Results of this kind have been obtained for special curves (cos = Is); cf., V. G.
Sprindzuk, Mahler’s Problem in Metric Number Theory, Minsk, 1967; concerning
the general case, cf., A. S. Pjartli, Diophantine approximations on submanifolds of a
Euclidean space, Funct. Anal. Appl. 3, 4 (1969), 59-62.
164 4. Perturbation Theory
We note that these publications do not solve either the above problem of generalizing
Neistadt’s theorem or the arithmetic problem of giving a sharp estimate for v (which,
by the way, does not have a great significance for our problem, in which the variation
of v will only change the necessary smoothness of the equations).
In generic systems with arbitrary numbers of fast and slow variables depending
on sufficiently many parameters for almost every (in the sense of Lebesgue measure)
value of the parameter one has:
(i) the measure of the set of the initial conditions, for which the deviation is greater
than p for some t in (0, 1/e), does not exceed Cy/s/p;
(ii) the integral of the deviation over the set of other initial conditions (the phase
space is supposed to be compact) does not exceed CN/e (the deviation equals
max |I(t) — /(r)| for t e [0, 1/e].)
For mappings a> outside some exceptional set of codimension N in the space of
mappings we may substitute e1/(p+1) for ^/e in the previous estimate, where
n ^ C£+p — k - N + 1
(here as above n is the number of fast variables and k is the number of slow variables).
The constant C in the estimate depends on the mapping, but this time the estimate
holds for all systems, not for almost all, and we do not need to consider parametrized
families—the result holds for individual systems.
Both theorems formulated above are proved by V. I. Bachtin, On the averaging in
systems with several frequencies, Uspekhi Math. Nauk 40, 5 (1985), 304-305, Funct.
Anal. Appl. 20, 2 (1986), 1-7; see also V. I. Arnold, V. V. Kozlov, A. I. Neistadt,
The Mathematical Aspects of the Classical and Celestial Mechanics, Itogi Nauki i
Techniki, Sovremennye Problemy Mathematiki, Fundamentalnye Napravleniia. Dy-
namicheskie Systemy, Vol. 3. pp. 3-304, see p. 181, Mosc. VINITI, 1985 (Springer
translation, 1988).
The proofs depend on the estimate
holding generically for almost all values of /. This estimate implies that the deviation,
averaged over the initial conditions, does not exceed (generically) a value of order
gi/iP+D if n ^ Cp+r
Let us note that these estimates provide new information, even for those values of
the dimensions (n and k) where the Neistadt theorem is applicable, in the case where
the Jacobian vanishes (but not identically) on some surface. Indeed, to apply the
Neistadt theorem, we have to consider time intervals of lengths Cje smaller than 1 /e,
where C depends on the initial point (so as to prevent the averaged orbit from the
intersection with the surface of zero Jacobian). In the theorem of Bachtin the set swept
by the orbits intersecting the zero surface of the Jacobian, is included in the exceptional
set of small measure controlled by the estimate.
cH_ cH
<p = / = —< ^
cl ’ ccp
(p = oj(J), 1=0,
dHx cHx
<P = CO(7) + £ i
dl ccp
B. Kolmogorov’s Theorem
We assume that the frequencies are independent in the sense that the deri¬
vative cco/cl of the frequencies with respect to the action variables is non¬
degenerate. In this case, as has been established by Kolmogorov [A. N. Kol¬
mogorov, On the preservation of quasi-periodic motions under a small variation
of Hamilton'sfunction, Sov. Math. Dokl. 98,4(1954), 527-530], the majority
*The coordinates (/, tp) are said to be canonically conjugate if the symplectic structure of
the phase space can be written in the form io = 'fdlk a dtpk.
166 4. Perturbation Theory
C. Nehoroshev’s Theorem
It turns out that the average velocity of the deviation of the action variables
from their initial values is so small in every generic Hamiltonian system
that it cannot be detected by any approximation method of perturbation
theory (i.e., it does not appear in the form of noticeable deviation in time
of order l/eN for any N, where e is the parameter of the perturbation).
More precisely, Nehoroshev [N. N. Nehoroshev, On the behavior of
Hamiltonian systems close to integrable ones, Funct. Anal. Appl., 5, 4
(1971), 82-83 ; N. N. Nehorosh cv. Exponential estimate of the time of stability
of Hamiltonian systems close to integrable ones, /, UMN, 32, 6, (1977),
5-66; II Trudy seminara im. I. G. Petrovskogo, V. 5 (1979), 5-50; N. N.
Nehoroshev, Stable estimates from above for smooth mappings and smooth
function gradients, Math. Sbornik 90, 3 (1973), 432—472. Cf., also his dis¬
sertation, MGU, 1973] proved that for almost every unperturbed Hamil¬
tonian function H0(I), there exist positive numbers a and b such that the
average rate of change of the action variables / does not exceed eb in the
perturbed system over time T = ellz\
We note that T increases more rapidly then any power of 1/e as e —*■ 0,
so that the variation of / over time l/eN is small for any N.
The constants a and b depend on geometric properties of the unperturbed
§ 19. Averaging in Hamiltonian Systems 167
dcp
*The nonsteepness condition, equivalent to that of Nekhoroshev for all analytic systems
(with the exception of codimension infinity), admits a very simple formulation: The critical points
of all the restrictions of the unperturbed Hamiltonian function to the affine subspaces of all
dimensions of the space of action variables should be of finite multiplicity (that is, isolated over
complex numbers)—this remark is due to Ju. S. Iliashenko, 1985; it is assumed that the unper¬
turbed Hamiltonian function has no critical points on any part of the action variable spaces
which we consider.
+ We note that in the space of action variables, the affine structure is uniquely determined
and so is the identification of vectors of the space dual to the space of frequencies with vectors
of the space of action variables.
168 4. Perturbation Theory
Here we give a review of the basic results of the theory of adiabatic invariants
in Hamiltonian systems with slowly varying parameters.
Example
co
The proof of this theorem is based on the averaging method. Let cp be the
angular coordinate on closed phase curves chosen in such a way that it
varies in proportion to the time of the motion on the curve and increases by
2n during every revolution (of course, the angular coordinate cp and the
action variable I both depend on not only the phase coordinates (p, q), but
also on the parameter A).
For every value of A, the equation of our system can be written in the
form of the standard unperturbed system of the averaging method:
<P = w + ef 7 = eg, i = e,
Lemma. The action variable is a first integral of the averaged system (that is,
the average of g with respect to (p is equal to zero).
Problem. A little ball moves horizontally between two vertical absolutely elastic walls
whose distance apart varies slowly. Prove that the product of the velocity of the ball
with the distance between the walls is an adiabatic invariant.
Problem. A charged particle moves in a magnetic field which varies slowly in the
course of a Larmour loop of the particle around a magnetic line of force. Prove that
an adiabatic invariant is the ratio v2JH of the square of the component of the velocity
of the particle normal to the line of force to the magnetic field strength (cf., for example,
L. A. Arcimovich, Controllable Thermonuclear Reactions, Moscow, Fizmatgiz, 1961).
det
CO)
det (?2Ho # o.
If V ?j2,
We assume, as above, that the parameter X begins to vary slowly. The
variation of p and q can be described by Hamilton’s equations with a varying
§ 20. Adiabatic Invariants 171
Corollary. The action variables 1 are first integrals of the averaged system.
Example
For arbitrarily small s (i.e., for an arbitrarily slow variation of the para¬
meter), parametric resonance is possible in which the equilibrium position
-v = 0 becomes unstable. It is clear that in parametric resonance an adiabatic
invariant of a linear pendulum changes unboundedly (over an infinite time
interval).
It turns out that this behavior of the adiabatic invariant in a system with
a periodic slowly varying parameter is connected with the linearity of the
system, more precisely, with the independence of the period of oscillations
from the amplitude. If, in a Hamiltonian system with periodic slowly varying
parameter, the derivative of the frequency of the fast motion with respect to
the action variable is different from zero, then the action variable varies
little over an infinite time interval [cf., V. I. Arnold, On the behavior of an
adiabatic invariant in a periodic slow variation of the Hamiltonian function,
Sov. Math. Dokl., 142, 4(1962), 758-761].
The proof is based on the fact that invariant tori exist in this situation
(cf., Kolmogorov’s theorem, § 19B).
Another interesting case is the one in which the parameter varies in such a
way that it has definite limits as t —»■ — x or t —> -hoc. In this case, it makes
sense to speak about the values of an adiabatic invariant at minus infinity
and plus infinity and its increment
A/ = /( + 30) - /(-oo)
one can prove that the increment of an adiabatic invariant over infinite
time is exponentially small in e (under the assumption that co is an analytic
function which must not change sign and must behave reasonably at infin¬
ity). Moreover, the principal term of the asymptotics of the increment of an
adiabatic invariant can be given explicitly as e —* 0 [cf., A. M. Dyhne,
Quantum passages in adiabatic approach, JETP 38, 2 (1960), 570-578 and
A. A. Sludskin, JETP 45, 4 (1963), 978-988]. Analogous results have been
obtained for multidimensional systems as well. The accurate formulations
and proofs can be found in M. V. Fedorjuk, Adiabatic invariant of a system
of linear oscillators and scattering theory, Differential Equations 12, 6 (1976),
1012-1018 (in which references to earlier work by physicists have been
omitted, however). See also M. Levi, Adiabatic invariants of linear Hamiltonian
systems with periodic coefficients, J. Diff. Equat. 42, 1 (1981), 47—71.
The problem of the increment of an adiabatic invariant of a one-dimen¬
sional nonlinear system has also been studied by physicists: they have
proved the smallness of the increment compared to eN, i.e., the absence of
variations of an adiabatic invariant in all orders of perturbation theory
[A. Lenard, Ann. Phys. 6 (1959), 261-276]. Neistadt obtained an exponential
§21. Averaging in Seifert’s Foliation 173
A. Seifert’s Foliation
Seifert’s foliation is a partition of the direct product IR2 x S’1 into circles.
It is constructed in the following way. In Euclidean three-dimensional space,
we consider a cylinder with horizontal bases and vertical axis. We foliate
174 4. Perturbation Theory
the interior of the cylinder by vertical segments and identify the upper and
lower bases of the cylinder after having rotated the upper base by an angle
of Inpjq [we glue the point (z, 0) of the lower base to the point (Az, 1) of
the upper base, where A is the rotation by angle 2npfq and p and q are
relatively prime numbers].
Let a vector field be given in the space R 2 x S' of Seifert’s foliation. Then
the covering foliation contains a vector field, too. Every vector of the field
can be projected on the base R2 of the covering foliation. We average the
vector thus obtained on the base along a leaf of the covering foliation. At
every point of the base we obtain a well-defined vector. Consequently, we
have defined a vector field on the base. The above operation of constructing
(from a field in the Seifert foliation space) a field on the plane is called
averaging the original f ield along Seifert's foliation.
In other words, averaging on Seifert’s foliation of type (/?, q) is defined
as the usual averaging in its <y-sheeted covering foliation.
§ 21. Averaging in Seifert’s Foliation 175
We realize the base as one of the bases of the initial cylinder. Then the
averaging along Seifert’s foliation turns into averaging on q intervals paral¬
lel to the axis of a cylinder. Upon rotation by an angle 2njq, these q intervals
turn into each other. It is easy to see that averaging commutes with rotation
by angle 2nlq. (After the rotation, we have to average along the same
intervals but in a different order.) ►
D. An Example
z - eF(z),
where the vector field F turns into itself under rotation of the z-plane by
an angle 2njq.
^ Asa result of the uniqueness of the Taylor series, every term of the series
defines a vector field invariant under rotation. Under rotation of 2 by
the angle 2njq, the vector zkz‘ turns by the angle (k — l)2nlq. This rotation
is a rotation by the angle 2 n/q if and only if A: — /is congruent with 1 modulo
q. ►
We consider the quadrant of the lattice of integral nonnegative points
(k, /). Among them, we distinguish those for which /c — / is congruent with 1
modulo q. The point (1,0) and all integral points on the line issued from
this point and parallel to the bisector of the quadrant are always distin¬
guished. These points correspond to fields z<I>(|z|2) invariant under rotation
by any angle.
The point (0, q — 1) is always among the distinguished points. This
point corresponds to the field zq~x invariant with respect to rotation by
the angle 2nfq. All distinguished points form a series of rays parallel to the
bisector issued from the points (0, mq — 1) and {mq + 1, 0) on the sides
of the quadrant.
Omitting the last term, we obtain the following simplest differential equation
with symmetry of order 3 :
z = az + bz 2.
Here the coefficients a and b and the phase coordinate z are complex.
We assume that a ^ 0 and b A 0. Multiplying z by a suitable number
and changing the unit of time, one obtains b — 1 and |a| = 1. The variation
of the phase portrait is shown for a = eltp, ft = 1 in Fig. 102. For any a,
there are four equilibrium positions at the vertices of an equilateral triangle
and at its center. For purely imaginary a, the system is Hamiltonian. In
order to study the system for arbitrary a, it is sufficient to notice that it
§ 21. Averaging in Seifert’s Foliation 177
Now we try to take into account the omitted terms 0(|z|3). We assume that
\a\ is small (this corresponds to the assumption that in the original system
of differential equations a resonance of third order almost takes place).
Then the radius of the triangle of singular points is also small (of order
\a\). We consider our symmetric vector field in the neighborhood of z = 0,
which is large compared to \a\, but small compared to 1.
In the same neighborhood, the omitted terms 6>(|z|3) are small compared
to the terms we kept. It is easy to show that their effect does not essentially
change the form of the phase portrait if it is structurally stable. In our case,
the phase portrait is structurally unstable only for purely imaginary a,
when the system is Hamiltonian. The Hamiltonian character is not preserved
if we take into account the omitted terms.
For every ray in the complex a-plane not coinciding with the imaginary
axis and for sufficiently small \a\ ^ 0, the shape of the phase portrait of the
complete system (under the condition b ^ 0) in a neighborhood of the origin
small compared to 1 and large compared to \a\ is indicated in Fig. 102,
<P ^ ± tt/2.
A study of the change of the phase portrait when a intersects the imaginary
axis constitutes a special problem, to which we shall return in Chap. 6.
In the generic case, the change is determined by just one term of the Taylor
series: all happens in the same way as for the equation
z — az + z2 czlzl2,
where Re c # 0.
For resonances of order q higher than 3, we obtain for the averaged system
of the first nontrivial approximation
z = az + Az\z\2 + z 3.
z = az + Az\z\2 + z4
Normal Forms
A. Resonances
K = (rn, A),
§ 22. Formal Reduction to Linear Normal Forms 181
Example
B. Poincare’s Theorem
Theorem. If the eigenvalues of the matrix A are nonresonant, then the equation
x = Ax + • ■ •
y = Ay
x = Ax + v(x) + • • •
*That is, let h be a vector field whose components are polynomials. A vector-valued poly¬
nomial is the sum of vector-valued monomials: the latter are fields, one component of which
is a monomial and the remaining components zeros. The order of a polynomial is the degree
of the lowest term.
182 5. Normal Forms
<vA ,
x = \ E + f1 \ Ay =
8h
E + — )A(x — h(x) + ■ ■ -)
'
WAy~
dh
= Tx + — Ax — Ah(x) +
cx
Remark. The square brackets contain the Poisson bracket of the vector fields
Ax and h{x).
We shall denote by LA the operator converting every field into the Poisson
bracket of the linear field Ax with the given field:
Ph
LAh = ffAx - Ah(x).
LAh = v,
The linear operator LA acts in the space of formal vector fields. It leaves
the spaces of homogeneous vector-valued polynomials of any degree
invariant.
We calculate the eigenvalues and eigenvectors of LA. Denote by ei an
eigenvector of A with eigenvalue Xt. We denote by (xl5 . . . , xn) coordinates
with respect to the basis (e1, . .. , en). As usual, xm will stand for x™1 • • • x“-.
Let h - xmes. Then the only nonzero component of the vector (8hjdx)Ax
is the jth one, which is equal to
Remark. If A is not diagonal (has Jordan blocks), then LA also has Jordan
blocks. Nevertheless, as is easily seen, the eigenvalues are given by the same
formula as in the diagonal case. Therefore, for nonresonant (possibly
multiple) eigenvalues, LA is invertible on the space of homogeneous vector¬
valued polynomials. Hence, the corollary above holds in the case of multiple
eigenvalues as well.
Remark 4. If the original equation is real but the eigenvalues are not, then an
eigenbasis can be chosen from complex conjugate vectors. In this case all
substitutions in Poincare’s theorem can be chosen real, i.e., converting
complex conjugate vectors into complex conjugate vectors.
A. Resonant Monomials
Example
For the resonance Xx = 2X2, the unique resonant monomial is x\ex. For the
resonance X1 + X2 = 0, all monomials (x 1x2)kxses are resonant.
y = Ay + w{y)
^ We begin killing the nonlinear terms of the series v. After several steps,
we may encounter an unsolvable homological equation
§ 23. The Case of Resonance 185
v^Tvm.sxmes, h = YJhm_sxmes
and set
h = --
m'5 (m. A) - Xs
for those m and 5 for which the denominator is different from zero. In this
way we define the field h.
We perform the usual substitution x = y + h(y) of Poincare’s theorem.
In the original equation, all terms of degree r will vanish except the resonant
ones, which will remain unchanged. The equation will take the form
y = Ay + wr(y) + • • • ,
y = Ay + w2(y) + - - - + vt’^y) + • • ■
is converted into
C. Examples
In practice, the Poincare-Dulac theorem is usually used to single out resonance terms
of small degree and remove perturbation up to terms of some finite order, i.e., reduce
the equation to the form
x = Ax + H’O) + o( | jc|
186 5. Normal Forms
Examples
1. We consider a vector field in the plane with a node-type singular point with resonance
A, = 2A2. The Poincare-Dulac theorem enables us to reduce (formally) the equa¬
tion to the normal form
f X, = A,X, + CX2
\ X2 = ^2X2-
In this case, the normal form is polynomial, since the number of resonant terms
is finite (just one).
2. We consider a vector field in the plane IR2 with a singular point with purely imaginary
eigenvalues a , 2 = +/co (center in linear approximation).
We pass to an eigenbasis. The eigenvectors can be chosen complex conjugate.
It is customary to denote by z and z the coordinates in C2 with respect to a basis
of complex conjugate vectors (these numbers are actually conjugate only in the
real plane [R2 cz <C2).
Our differential equation on [R2 gives an equation on C2 which can be written
in the form
Z — AZ + • • ■ , Z = Az 4-
(The dots denote a series in terms of the powers of z and z.) Since the second equa¬
tion is obtained from the first by conjugation, we may omit it.
We have the resonance A, 4- A2 = 0. By the Poincare-Dulac theorem, our
equation can be reduced to the form
£ = K + cC|C|2 + 0(|c|5)
If the real part of c is negative (positive), then the equilibrium position is stable
(unstable).
Consequently, the first few steps of the Poincare method provide a method for
solving the problem of the stability of a singular point neutral in linear approxima¬
tion. Moreover, it is immaterial whether the construction can be continued and
whether the procedure converges. It is only important that the “nonlinear decrement”
Rec be different from zero.
u = S + N, SN = NS.
A. Resonant Planes
Varying the integral vector m and the index 5, we obtain countably many
resonant planes. We shall study how the set of resonant planes is located
in the space C" of eigenvalues. It turns out that the resonance planes lie
discretely in one part of Cn, but are everywhere dense in another part.
188 5. Normal Forms
Remark. For n > 2, the Poincare and Siegel domains are open and separated
by a cone. For n = 2, the Siegel domain has real codimension 1 in C2.
Theorem 1. Every point of the Poincare domain satisfies not more than a
finite number of resonance relations Xs = (m, X), \m\ ^ 2, m{ ^ 0, and has a
neighborhood not intersecting the other resonant planes.
In other words, the resonance planes lie discretely in the Poincare domain.
By definition, in the plane of complex numbers there exists a real straight
line separating the collection of eigenvalues from zero. We consider the
orthogonal projections of the eigenvalues onto the line normal to this
straight line pointing away from zero. All these projections are not smaller
than the distance of the separating line from zero.
On the other hand, the coefficients mi of the resonance relation are
nonnegative. Consequently, for sufficiently large m, the projection of (m, X)
onto the normal will be larger than the largest projection of the eigenvalues
onto the normal to the separating line. p
Now we assume that the collection X of eigenvalues lies in the Siegel domain.
Theorem 3. The resonance planes are everywhere dense in the Siegel domain.
§ 24. Poincare and Siegel Domains 189
^ The point 0 lies either inside some triangle with vertices (Ax, X2, X3) or on
the interval (Al5 X2). In the first case, we consider the angle with vertex 0
formed by the linear combinations of the numbers Ax and X2 with real
nonnegative coefficients.
The negative multiples of the number X3 lie in this angle. We divide the
angle into parallelograms with vertices at the integral linear combinations
of Aj and X2. Let d be the diameter of such a parallelogram. For any natural
number N, the number — NX3 lies within one of our parallelograms. Con¬
sequently, it lies not farther than d from one of the vertices, so that
This inequality implies that the distance of our point from the resonance
plane X3 = mlXl + m2X2 + {N + 1)A3 does not exceed d/N. Hence, the
theorem is proved if zero lies in the triangle.
If 0 lies on the line segment between Xx and X2, arbitrarily large integers
px andp2 exist such that |plXl + p2X2\ ^ d.
This yields a resonance plane at a distance from X smaller than djp. ►
K- (.*>. -Dl» A
H
for all vectors m with nonnegative integral components mf, YJmi = \m\ ^ 2.
Theorem 4. The measures of the set of points which are not of type (C, v)
for any C > 0 is equal to zero provided that v >{n- 2)12.
We fix a ball in C" and estimate the measure of the set of points in it
which are not of type (C, v). The inequality in the definition determines a
neighborhood of width not greater than C, C/|m|v+1 of the resonance plane.
Therefore, the measure of the part of this neighborhood which is contained
in the ball does not exceed C2C2j\m\2v+2. Summing over m with fixed \m\,
we obtain not more than |m|”_1 C3C2/|m|2v+2. Summing over \m\, we obtain
C4(v)C2 < oo if v > (n — 2)/2. Consequently, the set of non-(C, v) points
in the ball is covered by sets of arbitrarily small measure. ^
In the real case, v > n — 1 is needed in Theorem 4.
Now we assume that the vector field is given by a convergent rather than
formal series, i.e., we consider a differential equation with a holomorphic
right-hand side.
190 5. Normal Forms
In other words, Poincare’s series are convergent for almost all (in the
sense of measure theory) linear parts of a field at a singular point.
Remark. All nonresonant vectors in the Poincare domain are vectors of type
(C, v) for some C > 0. On the other hand, in the Siegel domain, the sets
of the following vectors are everywhere dense: vectors of type (C, v), resonant
vectors, and nonresonant vectors which are not of type (C, v) for any C
and v.
For n-tuples of eigenvalues of the last type, although incommensurable
but very close to commensurability, Poincare’s series may be divergent, so
that the Field may be formally equivalent to its linear part but biholomor¬
phically inequivalent to it.
Remark. In contrast to that, if the eigenvalues lie in the Siegel domain, then
the series leading to the formal normal forms are often divergent in the case
§ 24. Poincare and Siegel Domains 191
f x = x2,
1 y = y - x,
the origin is a saddle-node type singular point. Despite the analyticity of the
right-hand side, the separatrix separating both sides of the half-plane x < 0
is not analytic, but only infinitely differentiable: y = ^{k — 1)! xk.
Many examples of the divergence of Poincare’s series have been con¬
structed by Brjuno (A. D. Brjuno, Analytic form of differential equations,
Trudy MMO 25 (1971), 119—262; in this work, the convergence of series is
also proved in some cases going beyond the scope of the Siegel theorem).
The Poincare and Poincare-Dulac theorems can be carried over to the real-analytic
case and the case of infinitely differentiable vector fields, or even the case of fields with
finite (sufficiently large) smoothness.
In the Siegel case, such a generalization is also possible [cf., for example, S. Stern¬
berg. On the structure of local homeomorphisms of Euclidean »-space, Amer. J. Math.
,
80, 3 (1958), 623-631, 81 3 (1959), 578-604],
However, one should mention that the situations in which these theorems are
applicable are topologically trivial. Indeed, the Poincare case for a real field can be
encountered only if the eigenvalues all lie either in the left half-plane or the right half¬
plane. In this case (independently of resonances), the system is topologically equivalent
to the standard system x = — x (or x = +x) in the neighborhood of a stable point in
a real space. All phase curves enter an asymptotically stable equilibrium position as
t —*■ xd (or emanate from an equilibrium position close to — oo).
In the situation of the Siegel theorem in a real domain, the Grobman-Hartman
theorem is applicable (the system is topologically equivalent to a standard saddle).
Indeed, if at least one of the nonzero eigenvalues of the linear part lies on the imaginary
axis, then the complex conjugate eigenvalue lies on the imaginary axis, too; the pair
Xl 2 = +ia> leads to the resonance Xi + X2 = 0. A vanishing eigenvalue is always
resonant. Consequently, in the real case, Siegel’s theorem is applicable only to systems
without eigenvalues on the imaginary axis, and such systems are locally topologically
equivalent to their linear part (the Grobman-Hartman theorem, § 13).
In contrast with Poincare’s and Siegel’s theorems, Poincare’s method is applicable
to the study of topologically complicated cases where there are eigenvalues on the
imaginary axis, for the method can be applied to the normalization of a finite number
of terms in the Taylor series. After this, one shows that the terms of higher order do
not change the qualitative picture.
The simplest example of this kind has been analyzed in § 23C. This method is
especially useful in bifurcation theory (cf., Chap. 6).
192 5. Normal Forms
We consider a formal mapping F: C" —> C" given by a formal power series
Tfx) = Ax + ■ ■ ■ . Let (ij, . . . , Xn) be the eigenvalues of the linear operator
A. A resonance is, by definition, a relation
Example
For n = 1, resonant eigenvalues are 0 and all roots of unity. All other
numbers X are nonresonant.
B. Formal Linearization
First of all, we consider the problem of the formal normal form of a map¬
ping at a fixed point.
F = JP-A.
where the dots indicate terms of order greater than r. The expression in
square brackets is a homogeneous vector-valued polynomial of degree r.
This polynomial depends on h linearly. The linear operator
MAh = v.
C. Convergence Problems
Poincare’s and Siegel’s theorems can be carried over to the case of discrete
time in the following way.
Siegel’s Theorem. For almost all {in the sense of Lebesgue measure) collec¬
tions of eigenvalues of the linear part of a holomorphie diffeomorphism at a
fixed point, the diffeomorphism is biholomorphically equivalent to its linear part
at the fixed point.
Example
We consider a mapping of C1 into itself with fixed point O and eigenvalue A which is
an «th root of 1. Such a mapping can be reduced to
This formula enables us to study the stability of the fixed point of a real mapping.
Indeed, the square of the mapping assumes the form
Hence, the first few steps of the Poincare method enable us to study the stability
of a fixed point in a case when it is neutral in linear approximation.
x = A(t)x
in a complex phase space, where the complex linear operator A{t): C" —»■ C”
depends 27t-periodically on t.
The monodromy operator is, by definition, the linear operator M: C" —»■ C"
converting the initial condition for t — 0 into the value of the solution with
this initial condition at t = 2n. (The monodromy mapping is defined not
only for linear equations, but also for any equation with periodic coefficients;
in this more general case, the monodromy mapping is usually called the
Poincare return map, or simply the Poincare map.)
y =
Remark. The proof of Floquet’s theorem only uses the representation of the
monodromy operator in the form M = e2KK. Therefore, a periodic change
of variables reduces to an equation with constant coefficients not only a
complex equation with a diagonal monodromy operator, but also every
equation for which the monodromy operator has a logarithm.
196 5. Normal Forms
A real linear operator does not always have a real logarithm, even if its
determinant is positive (the determinant of the monodromy operator is
always positive). Indeed, consider, for example, a linear operator on the
plane with eigenvalues (—1, — 2). If this operator is the exponent of another
linear operator, then the eigenvalues of the latter operator are complex but
not complex conjugate. Therefore, our operator on the real plane does not
have a real logarithm.
On the other hand, it is easy to see that the square of a real operator
always has a real logarithm. This leads to the following.
x = y + h(y, t)
Lemma. Ifh — 0(yr) (or the series h begins with terms of degree not smaller
than r) and r ^ 2, then
dh dh
x = Ax + — Ax - Ah + — +
.ox ot.
where the dots indicate terms of order greater than r with respect to x.
LAh + ht = v
We shall also consider the case where h and v are formal series with co¬
efficients 27i-periodic in t.
v(x, o = I h =
h _m,k,s —_
m’k’s ik + {m,X)-X;
As = (m. A) + ik,
y = Ay + iv(>-, /),
Example
r = i co¬
in an appropriate coordinate system. (As usual, the equation for z is omitted, since it
is conjugate to the one above.) The eigenvalues are A, 2 = ±ko. The resonant terms
in the equation for i are determined from the condition
ik + (/77, — m? — 1)/oj = 0.
By a true (not formal) change of variable, the equation can be reduced, for example,
to the form
z = iioz + c,z|z|2 + - ■ • ,
where the (27t-periodic) dependence on / has been preserved only in terms of order of
smallness 5 in z (denoted by dots).
We note that in this case every step of the Poincare method reduces to averaging
with respect to t and arg z and that the equation thus obtained is invariant under trans¬
lations of / and rotations of z.
§ 26. Normal Form of an Equation with Periodic Coefficients 199
Let us now assume that in the preceding example that co is rational, co = pjq. In this
case, we obtain from the equation for the resonant terms
k = pr, mx = m2 + 1 — qr.
For the study of the normal form, it is convenient to consider a g-sheeted covering on
the time axis. We note that the integral curves of the linear part of our equation form
a Seifert foliation of type (p, q) (cf., § 21). On the space of the ^-sheeted covering, the
integral curves form a trivial foliation and we can introduce direct product coordinates.
The coordinate on a fiber will be denoted by t{mo62nq). The coordinate £ on the
base is determined from the condition
z = eioj,£-
In this notation, the linear part of our equation takes the form £ = 0. The normal
form is a formal series
£ = iC*c'
In this case, every step of the Poincare method reduces to averaging along Seifert’s
foliation. Therefore, the equation thus obtained is invariant under translations of /
and rotations of £ by angles which are multiples of 2tz/q.
A study of the equations thus obtained will be carried out in Chap. 6.
Nevertheless, for almost all {in the sense of Lebesgue measure) collections
of eigenvalues X of the operator A, in the neighborhood of the null solution,
a holomorphic differential equation x = Ax + • • • 2n-periodic in t can be
reduced to the autonomous normal form x = Ax by a biholomorphic
transformation 2n-periodic in t (Siegel’s theorem for the case of periodic
coefficients).
The proof is as usual, cf., § 28. ^
/
G. The Neighborhood of a Closed Phase Curve
Example
The phase space is the Mobius band and the phase curve is the circular axis.
In general, the neighborhood of a circle on a manifold is not a direct
product if and only if the manifold is not orientable and the circle is a
disorienting path. In this case, one has to resort to a two-sheeted covering
of the original circle for the passage to an equation with periodic coefficients.
The theory of normal forms of equations with periodic coefficients could have been
deduced from the theory of normal forms of their Poincare mappings, i.e., normal
forms of diffeomorphisms in the neighborhood of a fixed point. Conversely, the study
of a diffeomorphism in the neighborhood of a fixed point can be reduced to the study
of an equation with periodic coefficients for which the diffeomorphism is the Poincare
mapping.
In the case of finite or even infinite real smoothness, the construction of a differential
equation with periodic coefficients* from a given Poincare mapping does not present
*It is usually not possible to imbed a given mapping in the phase flow of an autonomous
equation (an example is a diffeomorphism of the circle with irrational rotation number, which
is not equivalent diffeomorphically to a rotation, cf., § 11).
§ 26. Normal Form of an Equation with Periodic Coefficients 201
major difficulties. In the analytic or holomorphic case, the situation is more complicated.
This problem is equivalent to the problem of analytic (holomorphic) triviality of
analytic (holomorphic) foliations over a circular ring under the assumption of topolog¬
ical triviality. Although a positive answer follows essentially from the theory of sheaves
and Stein manifolds, as far as this author knows, no proof has been published. (The
author is indebted to V. P. Palamodov and Ju. S. Il’jasenko for an explanation of this
matter.) We shall not go into this theory, especially because the results necessary for
the study of differential equations and diffeomorphisms need not be deduced from
each other, but can be obtained independently, using the same method of proof.
x = Ax + v(x, cp),
(p = <x>,
where k runs through the lattice of integral points of the r-dimensional space,
and m satisfies the usual conditions mp Ss 0, X>p ^ 2.
Let v be analytic (holomorphic) in x and cp of period 2n in cp = {cpx, . . . ,
cpr). Then one can prove the reducibility of the system to the form
y = Ay, cp = co
A. Elliptic Curves
Example
We consider the plane C of a complex variable and two complex numbers (co l5 co2)
whose ratio is not real. We identify every point <p of C with any point obtained from it
by translation by col and co2 (and consequently, with all points cp + k1co1 + k2co2,
where kl and k2 are integers). After this identification, C becomes the elliptic curve
M. - -.
ccqZ + oj2Z
Consequently, the elliptic curve T can be viewed as a parallelogram with sides (ccq, a>2),
in which corresponding points are identified on the opposite sides.
It can be proved that all elliptic curves can be obtained by the construction above
(up to biholomorphic equivalence). This fact is by no means an obvious theorem.
For example, consider the strip 0 ^ Im cp ^ x and glue all points (p, <p -I- 2n and
also points of the boundaries of the strip, identifying a point <p with cp + h + a +
0.5 sin <p for real numbers <p. The manifold thus obtained can be mapped biholo-
morphically onto the quotient manifold C/cOjZ + cn2Z; nevertheless, it is not easy
to prove this. Probably, under usual Diophantine conditions, to1/co2 converges to the
rotation number as t —> 0.
The numbers col and co2 are called the periods of the curve. If we multiply both
periods by the same complex number, we obtain new periods which give an elliptic
curve biholomorphically equivalent to the initial one. Therefore, the periods can always
be chosen so that coj = 2n.
In this case, we can denote the second period by co. We may always assume that
Im co > 0. To different co there correspond generally biholomorphically inequivalent
elliptic curves (more precisely, the curves are inequivalent biholomorphically if the
corresponding lattices cOjZ -I- co2Z do not turn into each other upon multiplication
by a complex number).
Problem*. Prove that the phase curves of the one-dimensional Newton equation with
*The solution of this and the following problems is based on elementary knowledge of the
topology of Riemann surfaces contained in any course on the theory of functions of a complex
variable.
§ 27. Normal Form of the Neighborhood of an Elliptic Curve 203
potential energy of degree 3 or 4 are elliptic curves (if they are considered in a complex
domain).
Hint. The role of the coordinate <p in the plane covering the elliptic curve is played
by the time t of motion on the phase curve, the time being defined by the relation
dt = dxjy (the time is also called an elliptic integral of the jirst kind).
Problem. Let the potential energy be a polynomial of degree 4 with two minima. Prove
that the periods of (not necessarily small) oscillations with the same total energy
coincide in the two wells.
Hint. The integrals of the first kind on any two meridians of the torus coincide.
Problem. Let the potential energy be a polynomial of degree 3 with a local maximum
and minimum. Prove that the period of oscillations in the well is equal to the period
of the motion from infinity to infinity on a noncompact phase curve with the same
total energy.
Remark. Choosing potential energy of degree 3 or 4, we may obtain any elliptic curve.
It follows from results of the preceding problems that an elliptic curve is an algebraic
manifold.
The simplest surface containing an elliptic curve is the direct product of an elliptic
curve with the complex line. Besides the direct product of a circle with the line, there
is a nontrivial bundle over the circle with fiber equal to the line (the Mobius band);
similarly, besides the direct product, there are other fiber bundles over an elliptic curve
with fiber C.
We consider a fibering of the plane of two complex variables into complex lines.
We shall call the fibers of this bundle vertical lines.
We shall denote by (r, <p) the coordinates in C2, where r is the vertical and <p the
horizontal coordinate. Our fibering C2 —» C assigns the point q> of the horizontal
complex line to the point (r, q>).
Let T be an elliptic curve covered by the horizontal line. The curve T is obtained
from the horizontal <p-axis by identifying points differing by integral multiples of the
periods (oo1, co2).
In the plane C2 we identify those vertical lines whose projections on the horizontal
line differ by integral multiples of the periods. This identification turns C2 into a
fiber bundle over the elliptic curve T. The identification of vertical lines itself can be
carried out in different ways. (Similarly, upon pasting fibers over a circle, out of a
rectangle we may obtain either a cylinder or a Mobius band, depending on how the
vertical lines are glued together.)
The simplest method of pasting is to identify (r, <p) with the points (r, q> + oq)
and (r, <p + a>2). Then we obtain a direct product. The next simplest method of pasting
includes a twisting of the vertical lines being glued together.
204 5. Normal Forms
Example
Let A be a complex number different from zero and let T be the elliptic curve with
periods (27i, to). In the plane C2 we identify the points with coordinates
After this identification, C2 turns into a smooth complex surface Z and the fibering
C2 —> C, (r, tp) i ► tp turns into a fiber bundle Z —► T whose base is the elliptic curve T
and whose fiber is C. The equation r = 0 yields an imbedding of T in Z.
The surface Z can be imagined in the following way (in the case of a real X). We
consider the real three-dimensional space foliated by vertical lines with horizontal
plane {tp eC). Consider the strip 0 ^ Im tp < Imco. We glue the vertical planes which
are the boundaries of this strip, by identifying the point (r, tp) in the vertical plane
Im tp = 0 (r is the vertical coordinate) with the point (Xr, tp + to) in the plane Im tp =
Imcn. Moreover, we glue the points differring by 2n in the coordinate tp. We obtain a
fibering over an elliptic curve whose fiber is a line.
To imagine the complex surface Z, we only have to replace real vertical lines by
complex ones.
Topologically, the fiber bundle thus constructed is a direct product. Nevertheless,
from the holomorphic point of view, this bundle is generally nontrivial.
In the direct product, T can be deformed: for any e the equation r = e defines an
elliptic curve in the direct product. Let Tj be an elliptic curve in the total space Z near
T, which is the zeroth section of the fiber bundle (the equation of T has the form r = 0).
The curve Tj is given by an equation r = f(tp), where f(tp + 27i) = f(tp),f(tp + to) =
Xf(tp). Expanding/in a Fourier series,/ = Xfke'k<p> we find that jke,kw = Xfk. Conse¬
quently, fk = 0 and Tt coincides with T. Hence, our elliptic curve is not deformable
in a fiber bundle with X ^ e‘ka>. gfc-
Problem. Prove that for X = e'ko3 the fiber bundle with the projection Z -> T is a direct
product (trivial holomorphically).
Problem. Prove that the fiber bundles Zx —> T, Z2 -* T given by the complex numbers
Aj, X2 are equivalent biholomorphically if and only if Xx = X2e'km for some integer k.
Remark. The classes of biholomorphically equivalent fiber bundles of this form (called
one-dimensional bundles of degree 0) over a fixed elliptic curve T constitute a group
(where multiplication is multiplication of the numbers A).
It follows from the results of the preceding problems that this group can be identified
naturally with the quotient group of the multiplicative group of complex numbers
modulo the subgroup of numbers of the form e,k<a. The quotient group C*/{e,k<°} itself
§ 27. Normal Form of the Neighborhood of an Elliptic Curve 205
can be mapped biholomorphically onto the original elliptic curve. This group is also
called the Picard group or the Jacobi manifold of F. (These notions are defined for not
only elliptic curves, but also for arbitrary algebraic manifolds and, in the general
situation, they do not coincide with the original manifold.)
Remark. It can be proved that all topologically trivial one-dimensional vector bundles
over an elliptic curve are biholomorphically equivalent to the bundles Z -> T described
above.
All fibrations described above are topologically trivial (homeomorphic to the direct
product). The index of self-intersection of the zeroth section is an invariant enabling us
to distinguish topologically inequivalent bundles.
Let Mv, M2 be smooth oriented compact submanifolds of an oriented smooth real
manifold M (considering manifolds without boundary). We assume that the dimension
of M equals the sum of the dimensions of Mx and M2 ■ We also assume that Ml and
M2 intersect transversally (i.e., at every point of intersection, the sum of the tangent
spaces of the two submanifolds equals the tangent space of M).
The index of intersection of M1 and M2 in M is, by definition, the number of points
of intersection counting orientation. (A point of intersection is counted with positive
sign if the frame orienting My positively followed by the frame orienting M2 positively
determines a frame orienting M positively.)
Let Mx be an oriented compact smooth submanifold of dimension half that of M.
The index of self-intersection of Ml in M is defined as the index of intersection of Mx
with a manifold M2 obtained from Ml by a small deformation and intersecting M,
transversally. For example, the index of self-intersection of the meridian of a torus is
equal to zero, since the neighboring meridians do not intersect.
It can be proved that the index of self-intersection of Mx in M does not depend on
the choice of M2 if it is obtained from Mx by a small deformation.
Problem. Find the index of self-intersection of the sphere S2 in the space of its tangent
bundle.
Answer. +2. In general, the index of self-intersection of a manifold with its tangent
bundle is equal to the Euler characteristic of the manifold.
Now, over an elliptic curve, we consider the one-dimensional vector bundle Z -*• T
obtained from the fibering C2 -» C by pasting vertical lines (the vertical coordinate is
denoted by r) according to the rule
206 5. Normal Forms
Remark. All one-dimensional vector bundles over an elliptic curve are exhausted by
the bundles above up to biholomorphic equivalence.
Example
We consider a family of elliptic curves such that neighboring curves of the family are
not equivalent to each other biholomorphically. Such a family may be obtained, for
example, by identifying the points (cp, to), (cp + 2n, co) (cp -I- co, co) in the plane of the
two complex variables (cp, co). Upon identification, the domain Im co > 0 turns into
the union of the elliptic curves co = const. No neighborhood of any of these curves
can be mapped holomorphically onto this curve so that the curve itself remains fixed.
Indeed, if such a mapping existed, we would obtain a biholomorphic mapping
close to the identity mapping of elliptic curves with close distinct co onto each other,
which is impossible.
It turns out that in a certain sense this example is exceptional: the neighborhood
of an elliptic curve imbedded in a complex surface with vanishing index of self-inter¬
section is, genetically, biholomorphically equivalent to a neighborhood of the curve
in the normal bundle (in the same sense as a differential equation is generically equiv-
§ 27. Normal Form of the Neighborhood of an Elliptic Curve 207
/ r\ / r \ / rA(r, cp) \
Theorem. The function C defining the linear change of the vertical coordinate can be
chosen so that in the identification relation in the new coordinates the function A (0, cp)
takes the form Xe‘p<p (where p is the integer equal to the negative of the index of self¬
intersection of the elliptic curve r = 0 in the surface under consideration).
The function A(0, cp) defines the normal bundle of the curve r = 0 on our surface.
This bundle is biholomorphically equivalent to the bundle obtained by the pasting
208 5. Normal Forms
(cf., the Remark in § 27E). The linear change of the variable r reducing the pasting of
the normal bundle to this canonical form leads to the form of A(0, cp) given above,
In the following, we shall not always indicate identification of points differring only
by 2n in the coordinate cp, because the functions we encounter are 27r-periodic and cp
can be assumed to belong to the cylinder C mod 2n.
Definition. A pair of numbers (A, co) is said to be resonant if Xn = e‘kaj for some integers
n and k not vanishing simultaneously.
Theorem. The resonant pairs form an everywhere dense set in the space of all pairs of
complex numbers.
This follows from the fact that the set of numbers of the form i{{kjn)co + (m/n)2n)
(where k and m are integers and n is a natural number) is everywhere dense in the
complex plane. ►
Theorem. A pair (A, oj) is resonant ij and only if the bundle corresponding to the pasting
(r, cp) ~ (r, + 2n) ~ (Xr, cp + to) is trivial over some cyclic n-sheeted covering of the
original elliptic curve.
^ If A" = e,k<p, then (r, cp) ~ (e‘kcor, cp + noo) and, consequently, the bundle over
C/27tZ + nooZ is trivial (cf., § 27C). The converse can be proved analogously. ►
rA{r,cp) \
W \cp + co + rB(r, cp))'
where A and B are formal power series in r with coefficients analytic on the real <p-axis
and 27t-periodic in cp and A(0, cp) =£ 0.
A formal change of variables is a “mapping”
A / rC(r,<p) \
cp) \cp + rD(r, cp))’
§ 27. Normal Form of the Neighborhood of an Elliptic Curve 209
where C and D are formal power series in r with coefficients analytic in the strip 0 C
Im cp < Im o) of the complex <p-axis and 27r-periodic in cp and C(0, <p) ^ 0.
A formal change of variables g acts on a formal pasting/according to the formula
/^jo/o/1- (The right-hand side is defined by a natural substitution of power series
and is itself a formal pasting.)
+ ra^' ^ \
\<p) \<P + oi + rb(r, <p)j
fW
\<pj )
\<p + <*>/
We shall successively annihilate the terms of degree 1,2, ... with respect to r in
ra and rb. For this, as usual, we need to solve a linear homological equation. We write
the equation for the normalization of terms of degree n.
Let
A direct substitution shows that after the change of variables the coefficients of r" in
ra and rb take the form
210 5. Normal Forms
Theorem. For every fixed to, the nonnormal pairs form an everywhere dense set of Lebesgue
measure 0.
reduces to the linear normal form (r, tp) i—► (r2, (p + to) by a holomorphic change of
variables.
Corollary. In the sense of Lebesgue measure, almost all one-dimensional vector bundles
of degree 0 over an elliptic curve are rigid.
Remark. For certain nonresonant bundles in which the pairs (2, to) are nonnormal, the
formal series reducing the pasting to normal form may diverge. These nonnormal pairs
(2, to) constitute an everywhere dense set of measure zero. This problem will be dis¬
cussed in more detail in § 36.
§ 27. Normal Form of the Neighborhood of an Elliptic Curve 211
I. Negative Neighborhoods
Let us consider the case where the index of self-intersection of the elliptic curve on the
surface is different from zero. If the index is negative, then the curve cannot be deformed
in the class of holomorphic curves. Indeed, otherwise the deformed curve would
intersect the original curve with a positive index of intersection (since both curves are
complex).
Consequently, a curve with negative index of self-intersection lies isolated on the
surface. Such a curve is called an exceptional curve and its neighborhood a negative
neighborhood.
Theorem (Grauert). The normal bundle of an exceptional curve is always rigid, i.e., the
neighborhood of a curve with negative index of self-intersection on a complex surface is
defined by the normal bundle of the curve {up to biholomorphic equivalence').
We outline a simple proof of this theorem for the case of an elliptic curve.
We begin with the preliminary normal form of the pasting
We assume that the terms of degree in r smaller than n are already annihilated in
ra and rb, i.e., that
ra = r"an{(p) + - - • , rb = r"bn((p) + • • •.
After this change of variables (i.e., for the pasting g o fo g~l), the coefficients of r"
in ra and rb have the form
We make b„ and <5„ vanish. To this end, we first determine Dn from the second and
then C„ from the first equation. In both cases we have to solve a homological equation
of the form
with respect to the unknown 27r-periodic function u with the known 27r-periodic function
v.
212 5. Normal Forms
We consider the Fourier series expansions of the unkown and known functions:
« = v = Y.vkeikr
ocs2 + fis + y,
K. Positive Neighborhoods
Let us assume that the index of self-intersection of the elliptic curve on the surface is
positive. In this case, the homological equation studied in the preceding section is, in
general, unsolvable, since |e«2+/Js+y| increases as |j| -» oo. This means that not only
can the neighborhood of an elliptic curve with positive index of self-intersection on a
general complex surface not be mapped biholomorphically onto a neighborhood of
this curve in the normal bundle, but also such a mapping is impossible on the level of
2-jets (i.e., if we neglect terms of order 3 with respect to the distance from the curve).
The neighborhood of an elliptic curve with positive index of self-intersection is said to
be positive.
According to what has been said above, a positive neighborhood of an elliptic curve
has to have moduli and even functional moduli: the “normal form” of the neighborhood
has to contain arbitrary functions (and even, apparently, functions of two variables or
germs of functions of two variables at several points).
While a curve with negative index of self-intersection lies isolated on the surface,
an elliptic curve with positive index of self-intersection can always be deformed.
Much of what has been said about the neighborhood of an elliptic curve on a surface
can be carried over to the case of an elliptic curve in a multidimensional space. To
that end, the variable r has to be considered multidimensional in the formulas above.
Vector bundles of an arbitrary dimension over an elliptic curve can be described by
pastings
where A(cp) is a linear operator in Jordan normal form with eigenvalues of the form
Xeip<p.
A bundle is said to be negative (nonpositive, zero) if all numbers p are positive
(nonnegative, zero).
We assume that the normal bundle of our elliptic curve is negative. Then the neigh¬
borhood of the curve in a manifold can be mapped biholomorphically onto its neigh¬
borhood in the normal bundle (Grauert’s theorem), i.e., the normal bundle is rigid.
In the class of zero normal bundles, rigidity is violated only with probability zero.*
The condition of resonance takes the form As = X"e'ka>, where k is an integer, A" =
't'11 • • • K” m = dim nj > > 2.
A nonresonant bundle is formally rigid. For true holomorphic rigidity, the following
inequality (of (C, v)-normality) is sufficient:
for all n ^ 0, £".■ > 2 and the integral k. The measure of the vectors X which are not
(C, v)-normal for any (C, v) > 0 is equal to zero.
Rigidity (apparently) takes place with probability 1 for nonpositive bundles as well.
Structures of neighborhoods of curves of genus greater than 1 have not been studied
very much, except the case where the normal bundle is negative and, consequently,
rigid according to Grauert’s theorem.
In this section, we prove the theorem on the holomorphic local equivalence of a mapping
and its linear part at a fixed point.
for any .s and any vector k with nonnegative integral components with \k \ larger than 1
(C > 0, v > 0).
Theorem. We assume that in the neighborhood of the fixed point O in Cn, the collection
of eigenvalues of the linear part of a holomorphic mapping has multiplicative type (C, v).
Then in some neighborhood of O, the mapping is biholomorphically equivalent to its
linear part.
* For a proof, see Yu. S. Ilyashenko, A. S. Pyartly, Neighborhoods of zero degree of imbedded
complex tori, Trudy Seminara im. I. G. Petrouskogo 5 (1979), 85-95.
§ 28. Proof of Siegel’s Theorem 215
where the Taylor series of a and h at zero do not contain terms of degree 0 or 1. We
write the mapping H o A o H~l calculating the terms of zeroth and first degree with
respect to h and a and obtain
where the remainder R has second order of smallness with respect to a and h in a sense
which will be defined precisely later. We have enclosed the arguments of R in square
brackets in order to emphasize that the operator acts on functions rather than their
values at the point z.
We consider the following homological equation with respect to h :
Here the Taylor series of the known vector-valued function a and the unknown
function h do not have constant or linear terms. In the class of such series, the equation
can be solved uniquely, since the collection of eigenvalues is nonresonant.
In § 28C we shall prove that the series thus obtained are convergent if the collection
of eigenvalues has multiplicative type (C, v) for some positive C and v. We denote by
U the operator converting the right-hand side a of the homological equation into its
solution h = £/([a]).
By induction we define functions as, hs by the formulas
K = as+i = [AJ)
beginning with a0 = a.
We construct the mappings H0, Hx, ... defined by the formulas
Hs(z) = z + hs(z).
H = lim Hs o • • • o //j o Hq
We assume that the Taylor series of the right-hand side and the solution of the homo-
logical equation
Lemma 1. Let the collection of eigenvalues of the diagonal linear operator A have multi¬
plicative type (C, v). Assume that the right-hand side a of the homological equation is
continuous is the polydisk \zj\ ^ r and holomorphic inside this polydisk.
Then the homological equation has a solution h holomorphic inside the polydisk. For
every 5 such that 0 < <5 < j, the inequality
^ We expand a and h in Taylor series and denote the coefficients of zkes by ask and
hsk. Then hsk = asJ{Xs — Xk). We estimate the numerators using the Cauchy inequality
for the coefficients of a Taylor series and the denominators using the fact that {As} is
of type (C, v).
Let maX|z|^r|a(z)| = M. According to the Cauchy inequality, we have |a£| ^ M/rw.
Consequently, |A£| MC~l |/c|v/r|k|. Let us estimate the Taylor sum £hskzk. We con¬
sider the terms of degree |A:| = p. Their number does not exceed cx{n)p"~l and, there-
fore, \Y,\k\=PKzk\ ^ Mc2pm\(z/r)k\, where c2 = cJC and m = v + n — 1.
The function xme~x has a maximum (m/e)m and, therefore, pme~3P12 ^ c3d~m, where
c3 — (2rn/e)m. Consequently, for |z| ^ re~d, we have
(h o A)(z) = h(\z).
\*, - ^| s* Colkl-'l^l
for all s = 1, . . . , n and for all vectors k with nonnegative integral components whose
sum |k| is not smaller than 2.
We denote max |AS| by p. For |2.k| sj 2p we may take C0 = C/2p. For |2k| > 2u we
may take C0 =
Using this remark we obtain the following estimate for the Taylor series:
D. Order of Operators
For the following estimates, it is convenient to introduce some new notation. Let f be
a function continuous in the polydisk |z| ^ r, holomorphic at the interior points, and
vanishing at the center of the polydisk.
For such functions, we introduce the norm
\m
sup
Osg |z|< r
Example
The function /(z) = ez has the norm |e| independent of the radius of the polydisk.
ll*/°*-1IL = ll/ll,-
The values of the function f may not only be numbers but elements of a normed
space, for example, vectors, matrices, etc.
Let <1> be an operator acting on functions of the class described above*. Let d, a,
and p be positive numbers and let 0 < r < 1.
Definition. The operator <t> has order (d\ a\P) if for every 5 in the interval (0, 1/2) and
every r in (0, 1),
ll<i>([/])IL-'
provided that ||/||r ^ <5/
We shall write this relation in the form 0([/]) -3 fd(oc\P) or more briefly, <£([/]) -3
fd.
* We shall denote by the same letter operators acting on functions from classes with distinct
r under the condition that they coincide as operators on the space of germs of functions. This is
the same as in ordinary calculus when the notation of sine is not changed when its domain is
changed.
218 5. Normal Forms
An operator has order d if constants a and /? exist such that it has order (rf; oc|/J).
[It is essential that a and /? are independent of/, r e (0, 1), and 5 e (0, 1/2)].
Examples
First, we note that the Cauchy estimate implies the following inequality:
The convergence for |z| ^ re~s and the estimate g -3 h follow easily from theorems on
contraction mappings. ►
Example
h — g -3 h2.
for IzI ^ re d.
§28. Proof of Siegel’s Theorem 219
We note that in our notation 2/-9 f,f2 -3 /; if j\ -3 f2 and/2 -3 f3, then/, -3 f3.
We extend our notation to operators of several functions. Let the operator E
convert a pair of functions t], £ into the function £. Let <p be a polynomial. We write
£ -3 (pin, £)
if positive constants (a; /?,, /32) exist such that for any <5 in the interval (0, and any
r in (0, 1), we have the inequality
Example
We have £ -3 rjC-
This can be proved by means of inequality (1) which we have used in Examples 2
and 3. ►
We explicitly write out the remainder R defined in section B. We introduce the notation
By definition,
-^j The estimate Rx -3 /z2 has been proved in Example 3 of § 28D and the inequality
E([a], [gr]) -3 ag in Example 4. According to Example 2, we have G([/?]) -3 h, and,
therefore, /?2([a], [/z]) -3 ah.
From the estimates g -3 h, we obtain v -3 h + a. Consequently, according to the
estimate of the operator E from Example 4, Ri -3 u{h + a).
Estimate 2. Let U be the operator solving the homological equation. The operator
given by the formula = 2?([a], C/([a])) has order 2.
The conclusion of the proof of Siegel’s theorem is exactly the same as the estimates of
§ 12.
We choose the following number sequences:
Let as+l = i?([as], £/([as])). We assume that ^ Ms = <5SN. Thenfor N > ft,
the preceding inequality can be applied with <5 = 6S and we obtain
ik+1il, ^ = 3™~a-
If N > 2a, then the right-hand side does not exceed Ms+l = S2N/2. Therefore, we fix
N > (P, 2a). If ||a0||r0 ^ Mo = then < Ms = <5* for all s.
Finally, we choose r0. By assumption, the initial function a0 = a has a root of order
at least 2 at the origin. Consequently,
\a{z)\ ^ K\z\2
Hence the condition ||«0||ro <5q *s satisfied for sufficiently small r0. We fix such an
r0. Now all numbers ds, Ms, and rs are defined. The inequalities ||as||r < Ms are satisfied
for all 5. They imply estimates for hs. Therefore, the products Hs o ■ ■ ■ o Hl are deter¬
mined for |z| ^ rj2 and converge to a limit H ass -» oo. It is easy to verify that for the
limit diffeomorphism, we have H o A o H~l = A. ►
Chapter 6
The word bifurcation means division into two and is used in a wide sense
to indicate every qualitative topological metamorphosis of a picture under
the variation of parameters on which the object being studied depends. The
objects can be diverse: for example, real or complex curves or surfaces,
functions or mappings, manifolds or fiber bundles, vector fields or equations,
differential or integral.
If an object depends on parameters, we speak of a family. If we are
interested in a family locally, i.e., in small variations of the parameters in the
neighborhood of fixed values, then we speak of deformations of the object
corresponding to these values of the parameters.
It turns out that in many cases the study of all possible deformations
reduces to that of a single one from which all others can be obtained. In
some sense, such a deformation is the richest; it has to yield all possible
bifurcations of the object. It is called a versal deformation.
In this chapter, we mainly study bifurcations and versal deformations of
phase portraits of dynamical systems in the neighborhood of equilibrium
positions and closed trajectories.
For more details on nonlocal bifurcations, see “Bifurcation theory” (V. I.
Arnold, V. S. Afraimovich, Ju. S. Iljashenko, L. P. Shilnikov, eds.), in:
Modern Problems in Mathematics, Fundamental Directions, Dynamical Systems,
Vol. 5, Moscow, VINITI, 1986 (Springer translation. Encyclopaedia of
Mathematical Sciences, 1988). See also the survey “Ordinary differential
equations” (Arnold, Iljashenko) in the same series of surveys (Vol. I,
Moscow, VINITI, 1985), and the survey “Ergodic theory of smooth dy¬
namical systems” (Ja G. Sinai, Ja. B. Pesin, L. A. Bunimovich, M. V.
Jakobson, eds.), Vol. 2, Moscow, VINITI, 1985.
In the study of any kind of analytic objects (for example, differential equa¬
tions, boundary value problems or optimazation problems), one can usually
identify the generic cases. For example, the nodal, focal, and saddle points
are generic among the singular points of a vector field in the plane while, say,
the centers are destroyed by an arbitrarily small perturbation of the field.
The study of generic cases is always a primary problem in the analysis
of the phenomena and processes described by a given mathematical model.
Indeed, by an arbitrarily small change in the model, a nongeneric case may
turn into a generic case; on the other hand, the parameters of the model
are usually determined only approximately.
Nevertheless, there are instances in which it is necessary to study non¬
generic cases. Let us assume we do not study an individual object (say, a
vector field), but a whole family whose objects depend on a number of
parameters.
In order to visualize the situation more clearly, we consider a function
space whose points are the objects themselves (say, the space of all vector
fields). The nongeneric cases correspond to some hypersurface of codimen¬
sion 1 in the space. A point can be moved by an arbitrarily small shift off the
hypersurface to the domain of generic cases. The hypersurfaces of singular
cases form the boundaries of the domains of the generic cases (Fig. 105).
A family with k parameters is represented by a k-dimensional manifold in
our function space. For example, a one-parameter family is a curve in the
function space (the bold line in Fig. 105).
A curve in our function space may intersect a hypersurface of singular
cases. If this intersection occurs at “nonzero angle” (transversally), then it
is preserved under a small perturbation of the family: every nearby curve
intersects the hypersurface of singular points at some nearby point (the
nonbold line in Fig. 105).
Consequently, although every individual member of the family can be
made generic by an arbitrarily small perturbation, it is impossible to achieve
the condition that all members of the family be generic at the same time:
by a deformation of the family, the nongeneric case can be avoided for
every fixed value of the parameter, but for some nearby value nongenericity
occurs just as well.
224 6. Local Bifurcation Theory
where dk — djdxk and ,x, = 0 on the boundary [cf., S. M. Visik, On the oblique derivative
problem, Vestnik MGU, Ser. Mathem. 1 (1972), 21-28],
The oblique derivative problem apparently has to be stated according to the
following scheme. The manifold of tangency of the field with the boundary, the manifold
of tangency of the field with the tangency manifolds, etc. divide the boundary into parts
of various dimensions. On some of these parts of the boundary, conditions have to be
given : on some other parts, the boundary function itself has to satisfy certain conditions
for the existence of the classical solution of the problem.
In spite of the abundance of research concerning the oblique derivative problem (this
problem is studied especially thoroughly in the publications of Maz’ya, whose work was
preceded by that of Maljutov and Egorov with Kondrat’ev), the program described
above has been carried out only in the two-dimensional case (where the boundary is a
circle).
*Thal is. for every Held in an open everywhere dense set in the function space of smooth
fields.
226 6. Local Bifurcation Theory
Here p denotes some Riemannian metric; it is easy to see that the property
of A-tangency does not depend on the choice of the metrics pM and pN.
Two mappings are 0-tangent at a point x if their values coincide at x.
Tangency of order k is an equivalence relation
Notation
The point x is called the source and the point/(x) the target of the jet.
Example
The 0-jet of a mapping/of the x-axis into the y-axis at point x is given by
a pair (x, y) of numbers where y = /(x). A 1-jet is given by a triple (x, y, p),
where p = df/dx.
Besides tangency of order k, there is another equivalence relation leading
to germs of mappings instead of jets.
Definition. Two mappings in two neighborhoods of one and the same point
have a common germ at that point if they coincide in a third neighborhood
" of this point. (The third neighborhood can be smaller than the intersection
of the first two neighborhoods.)
The germ of a mapping at a point is an equivalence class with respect to
the equivalence relation introduced above.
As for a mapping, for a germ we can define its 0-jet at a point, 1-jet at
a point, etc.
We consider the set of all /c-jets of germs* of smooth mappings from M
into N at all possible points of M.
Example
* In the real smooth case, it is immaterial whether we consider jets of germs or jets of map¬
pings on the whole of M, since every germ is the germ of a global mapping. On the other hand,
in the complex situation, a global mapping with a given jet may not even exist.
228 6. Local Bifurcation Theory
Example
J*x(M) = GL(Rm).
It is easy to see that these mappings (the mappings ignoring the terms of degree k in
§ 29. Families and Deformations 229
the Taylor polynomial) are homomorphisms and their kernels are commutative groups.
For example, let m = 1.
^ Then: if f(x) — x + axk(mod(xk+l)) and g(x) = x + bxk mod(x’l+l), (fog)(x)
x + axk + bxk(modxk + l).
M ^ TM
is commutative.
The definition of germs, jets, and spaces of jets of vector fields mimic the
above definitions.
The group of diffeomorphisms of a manifold M acts on the space of all
vector fields on M and also on the spaces of fc-jets of vector fields on M.
The group of &-jets of local diffeomorphisms of a manifold Mata point
acts on the space of (k — l)-jets of vector fields on M at the point; this
action is linear.
Example
Lety = atx + a2x2 + •-• be the 2-jet of a local diffeomorphism at zero. The image of
the I-jet r(x) = r0 + r, x + ■ • • of a field is given by the formula h (.y) = iv0 + iiyc +
■ ■ • , where w0 = at v0, ny — atvtaJl + 2a2afv0.
This formula can be obtained by writing the equation x = v(x) in y coordinates. ^
L = X + Y.
(Here and in the following the word manifold means a manifold without
boundary.)
Example
Let C be the union of two planes intersecting in a straight line in the three-
dimensional space. The stratification is the partition into the line of intersec¬
tion and four half-planes. Transversality to C means transversality to each of
the planes and transversality to the line of intersection. For example, a curve
§ 29. Families and Deformations 231
Figure 108.
transversal to the stratified variety C does not intersect the straight line of
singularities of C.
This theorem is called the weak transversality theorem. Its assertion means
that a mapping not transversal to a fixed submanifold can be turned into a
transversal mapping by a small perturbation (Fig. 108). If, on the other hand,
transversality is present, then it is preserved under small perturbations.
We consider the special case where B is a linear space, B = Rn and C is its
subspace
We represent B in the form of the sum B = C + D of two subspaces of
complementary dimensions, C = IR"-* and D = (Rk. We project B onto D in
the direction of C; we denote this projection by n. Let us consider the
mapping n ofA -* D.
The point 0 is a critical value for this mapping if and only if the mapping
/: A —» B is not transversal to the submanifold C c B. By Sard’s lemma
(§ 10), almost all points of D are not critical values for n of Let £ be a point in
D which is not a critical value for n of We construct a mapping fz: A —► B
by setting^(n) = f{a) — e. The mapping/^ is transversal to C. Since £ can be
chosen arbitrarily small, we have proved that the set of transversal mappings
is everywhere dense in our special case. That this set is open follows from the
implicit function theorem. The general case can easily be reduced to the one
just considered. ^
Remark. If C is not compact, then the term “open” has to be replaced, in general, the
term “intersection of countably many open”.
Example
2. B is the plane, A is a circle imbedded in the plane, and C is a tangent line to A (without
the point of tangency). The imbedding is transversal to C but there exist nontrans¬
versal mappings to C arbitrarily close.
232 6. Local Bifurcation Theory
Example
1. Let C be a finite union of planes in a linear space stratified in a natural way (for
example, a pair of intersecting planes in F53). Transversality to IR* implies transver¬
sality to the ambient R'. Therefore, our condition is satisfied.
2. Let C be the cone x2 = y2 + z2 in R3 and let the stratification be partition into the
point 0 and the two skirts. As is easily seen, our condition is satisfied.
*The umbrella includes a handle indicated by the bold line in Fig. 109.
§ 29. Families and Deformations 233
Figure 109.
: (R2 —> [R3 given by the formulas x = u, z = v2, y = uv. Whitney has proved that
(1) the type of singularity (up to diffeomorphisms of (R2 and (R3) is preserved under
a small perturbation of q>\ (2) this is the only singularity of mappings of two-
dimensional manifolds into three-dimensional manifolds, which is preserved under
small perturbations (except for lines of self-intersection); all other singularities split
into singularities of this type under a small perturbation.] Transversality to the
singular line x = y — 0 does not imply transversality to the manifold of regular
points of the surface close to this line. (The plane z = 0 is transversal to the line
but not to the surface.)
If the condition
Example
Let B be the space of linear operators b: IR" —> IRn and let C be the set of
operators of nonmaximum rank. The operators of rank r form a smooth
manifold whose codimension in the space B is equal to (m — r)(n — r). The
partition into manifolds of operators of distinct ranks defines a stratification
on C.
A mapping f : A -* B is a family of linear operators from IRm into Rn
smoothly depending on a point of A as a parameter. The manifold A is called
the base of the family. The transversality theorem immediately implies the
following corollary.
Remark. The weak transversality theorem can be obtained from the above theorem for
k = 0. On the other hand, the strong theorem cannot be obtained from the weak
theorem. It is possible to apply the weak theorem to a mapping/: M —► Jk and obtain a
mapping close to/and transversal to C. However, this nearby mapping will not generally
be a A-jet extension of any smooth mapping of M into N.
-^j The essence of the proof consists of the same kind of reduction to Sard’s
lemma as in the case of the weak transversality theorem. The main difference
lies in the fact that the transversalizing deformation is not sought for the class
of mappings/ = / — e, but for the larger class of polynomial deformations
/ = f -t- eyel + • • • + eses, where ei are all possible vector monomials of
degree not greater than k.
M The singular points are preimages of a smooth manifold (the zeroth sec¬
tion) in the space of O-jets of vector fields. Nondegeneracy of a singular point
is transversality to this manifold of the 0-jet extension of the field. ►
Hence, a degenerate singular point disintegrates into nondegenerate points
under an arbitrarily small perturbation of the field.
Example
x = x2, y = -y.
Remark. Strictly speaking, multiplicity is defined in the following way: (1) a sufficiently
small neighborhood of the singular point in a complex space is fixed; (2) this neigh¬
borhood determines the smallness of a perturbation; (3) for the perturbed field, the
number of singular points is calculated in the neighborhood of the given point.
Definition. The support of the series/is the set of points m in the octant Z+ for
which/, A 0. Notation:
Definition. The Newton polyhedron of the series f is the convex hull of the
union of octants parallel to Z” with vertices at the points of the support in
the octant R" of the real linear space. Notation :
{a + b)2 - a2 - b2
a2 = ab |a=b, ab
2
Example
In the planar case, n = 2 and the mixed volume of a pair (Tj, F2) is F(r,, T2)
= [W, + r2) - viT,) - F(r2)]/2.
Example
/i = min(a1h2, a2bx).
The reduction of a matrix to the Jordan normal form is not a stable operation
if there are multiple eigenvalues. Indeed, in the presence of multiple eigen¬
values, an arbitrarily small change in the matrix may change the Jordan form
completely. Therefore, if the matrix is known only approximately, then its
reduction to Jordan normal form is practically impossible in the case of
multiple eigenvalues. It is not even necessary, since a generic matrix does not
have multiple eigenvalues.
Multiple eigenvalues are unremovable under small perturbations when we
are interested in a family of matrices depending on parameters rather than in
an individual matrix. In this case, although we can reduce every individual
matrix of the family to a Jordan normal form, both this normal form and
the transformation leading to it generally depend discontinuously on the
parameter.
Consequently, a problem arises: What is the simplest form to which a
family of matrices depending smoothly (for the sake of definiteness, holomor-
phically) on the parameters can be reduced by a change of coordinates
depending smoothly (holomorphically) on the parameters?
Let us consider the set of all complex square matrices of order n as a linear
space of dimension n2. The relation of similarity of matrices partitions the
whole space C"2 into manifolds (orbits of a linear group): two matrices lie in
the same orbit if their eigenvalues and the dimensions of the Jordan blocks
coincide. Because of the eigenvalues, this partition is continuous. As a rough
model, one can imagine the division of the three-dimensional space into the
strata of the manifolds x2 + y2 — z2 = C (Fig. 110).
A family of matrices is given by a mapping of the space of the parameters of
the family into the space C" of matrices. It turns out that from all families of
matrices we can select such families that a reduction to them can be effected
by a change of basis which now depends on the parameters smoothly (and by
a smooth change of parameters). Such families are called versal deformations
(the precise definition is given below). Versal deformations with the minimum
possible number of parameters are said to be miniversal.
Consequently, miniversal deformations are normal forms with the smallest
possible number of parameters in the reduction to which the smooth depen¬
dence on the parameters can be preserved.
240 6. Local Bifurcation Theory
Example
If all eigenvalues of a diagonal matrix are distinct, then we may take the
family of all diagonal matrices as its miniversal deformation (the parameters
are the eigenvalues).
B. Versal Deformations*
A'(X) = C{X)A{X)(C{X)yf
Definition. The family induced from A by the mapping <p is the family cp* A :
(cp*A)(p) = A(<p(p)).
The induced deformation <p*A of the matrix yl(0) is defined by the same
formula.
Example
The family of diagonal matrices with diagonal entries (oq + Xt), where all oq
are distinct and the Xt are the parameters of the deformation, is a versal,
universal, and miniversal deformation of the matrix (a,).
* Versal deformations were introduced by Poincare (Lemma IV of his Thesis), he has studied
the versal deformations of equilibria and cycles in dynamical systems.
§ 30. Matrices Depending on Parameters and Singularities of Decrement Diagrams 241
n i (a,) $5 «2(a,.) ^ •
A0 + B(X)
in this basis, where A0 is the Jordan upper triangular matrix of'A(A0) and B(X)
a block-diagonal matrix whose blocks correspond to the eigenvalues of A0.
The block Bt corresponding to the eigenvalue a, has all zeros except at the
places indicated in Fig. Ill; these places are taken by holomorphic functions of
X.
In Fig. Ill, three normal forms are depicted. In the first two, the number
of nonzero entries of B{ is equal to nf a,.) + 3n2(ai) + • • • ; in the third, all
entries are equal on every slanted line. We obtain miniversal deformations of
A 0 if the indicated entries of the matrices B{ are considered independent vari¬
ables ; in all three cases their number is equal to £ [nx(a,-) -I- 3n2(ai) + • • • ].
The advantage of the first two normal forms is that the number of nonzero
entries is the smallest possible. The advantage of the third form is the ortho-
n, ”2 n3
If—
[ _
Figure 111.
242 6. Local Bifurcation Theory
gonality of the versal deformation to the corresponding orbit (in the sense of
the entry-wise scalar product of matrices).
Lemma 2. In the neighborhood of\e, 0), the mapping <D is a local dijfeomorphism
on (C"\ A0).
For the proof of Lemma 2 we consider the mapping i// of the group of
nonsingular matrices into the space C"2 of all matrices given by the formula
i//(b) = bA0b~l.
I. The derivative of the mapping tjj at the identity is the operator of com¬
mutation with A0: i]/*: C"2 —» C"2, = [*L Z0].
In the space Cn\ we introduce the Hermitian scalar product (A, S') =
tr(AB*), where B* is the matrix obtained from B by transposition and
complex conjugation. The corresponding scalar square is simply the sum
of the squares of the absolute values of all the entries of the matrix.
-4 The vectors tangent to the orbit are the matrices representable in the
form [T, Aq\. Orthogonality of B to the orbit means that <[T, A0], By = 0
for any X. In other words,
Lemma 4. The matrices in Fig. 112 and only they commute with the matrix A0.
ss 5
Figure 112.
244 6. Local Bifurcation Theory
Proof of Lemma 2. ^ The derivative of <I> with respect to p at (e, 0) is if/*, and
the derivative with respect to X is A . According to what has been proved
above, these operators isomorphically map the space tangent to P at e and
the space tangent to A at 0 onto transversal spaces of the same dimensions
(the tangent space to the orbit C at A0 for P and a space transversal to it
for A). Consequently, the derivative of at (e, 0) is an isomorphism between
linear spaces of dimension n2. By the inverse function theorem, is a local
diffeomorphism. ►
D. Examples
fa i\ /0 0
(1)
V 0 cj+U ^2
fa 0\ / Aj X2
\° aJ VA3 ^4
fa 1 0\ / 0 0 0\
I 0 a 1 )+( 0 0 0 j.
Vo 0 a) \Xx X2 xj
fa 1 0\ /0 0 0\
I 0 a 0 J+f Xl X2 X3 J.
Vo 0 a) Va4 0 Xj
form of (1) for nearby values of the parameter, where k1 , A2 are holomorphic
functions of the parameters.
In the study of many problems about operators depending on parameters,
the normal forms constructed above enable us to restrict ourselves to special
families: miniversal deformations. One of these problems is that of the
structure of bifurcation diagrams.
E. Bifurcation Diagrams
Figure 113.
Figure 114.
248 6. Local Bifurcation Theory
way. First, let the real operator in !R2" for which the versal deformation is
sought have a unique pair of complex-conjugate eigenvalues x + iy (y # 0)
with Jordan blocks of dimensions nx ^ n2 ^ ,so that n1 + n2 + ■ ■ ■ =
n. In some real basis in IR2", the matrix of the operator has the same form
as the matrix of the decomplexification of the complex Jordan operator
A0 : C" —>■ C" with the only eigenvalue x + iy and Jordan blocks of dimen¬
sions «] ^ n2 ^ • • • , i.e., the form
X -yE
Ao (2)
yE x
(; :K a-
i.e., the following deformation with the parameters px, p2, r,, t2 :
1 —y 0 0
°\ 1° 0 \
X 0 -y ! Pi P2 -T1 -T2 ' z = x + iy
—L_
1
0 X 0 0 h = Pk + nk
y 0 il U
1 °
T2 Pi
0 1
Pi 1
By means of a similarity over the Field of real numbers, every real matrix
can be reduced to a block-diagonal form where a real Jordan matrix corre¬
sponds to every real eigenvalue and a block of the form (2) to every pair of
complex-conjugate eigenvalues.
We obtain a real versal deformation of a matrix reduced to this form
(with the smallest possible number of parameters) if we replace every block
by its minimal versal deformation. The minimum number of parameters of
a real versal deformation is thus given by the formula
where the summation is extended to all v eigenvalues, both real and complex.
Explicit formulas of versal deformations and tables of bifurcaction
§ 30. Matrices Depending on Parameters and Singularities of Decrement Diagrams 249
These results may also be applied to nonlinear systems having stationary points
depending smoothly on the parameters: at these points, the decrement of the lineariza¬
tion of the nonlinear system will have only the simplest singularities as a function of the
parameters (in the case of a generic family.)
When we apply these results to nonlinear systems, however, we have to exclude the
part of the boundary of the domain of stability corresponding to vanishing roots, since
the smooth dependence of the stationary point on the parameters is not preserved.
Consequently, the desription of singularities of the boundary of the domain of stability
of a generic nonlinear system (and the description of decrement diagrams in the neigh¬
borhood of points of this boundary) requires additional analysis. We return to this
question in the following subsections.
In the study of iterations of mappings, and also of equations with periodic coefficients
or motions in the neighborhood of a periodic trajectory, the role of the decrement is
played by the largest absolute value of the eigenvalues. If this absolute value is different
from 1, then its singularities (as a function of the parameters in a generic family) are the
same as those of the decrement of a generic family. Therefore, in the following, we
consider only the decrement.
In the study of the absolute values of eigenvalues in noninear problems of the types
just indicated, the results of this section are applicable outside the boundary of stability
and at those points of the boundary for which 1 is not an eigenvalue.
G. Decrement Diagrams
Definition. The increment* of the family is the function J of the parameter whose value at
/. is equal to the largest real part of the eigenvalues of A(/.) :
The function / is continuous but not necessarily differentiable. Our problem is the study
of singularities of / for a generic two-parameter family. Consequently, the parameter
space A may be assumed to be the plane IR2 or a domain in the plane.
A family of level curves of / in the plane A will be called a decrement diagram. A dash
across a level curve indicates the direction of the slope, i.e., the direction in which /
decreases; in other words, the dash points in the direction of increasing stability.
Example
z = xz + yz
*ln engineering, the quantity [ / | is called the decrement for / < 0 and the increment for
/ > 0. ^
§ 30. Matrices Depending on Parameters and Singularities of Decrement Diagrams 251
depending on two parameters (x, y). The matrix of the corresponding system has the
form
/0 1
A{x,y) =
\x y
The decrement diagram is given in Fig. 115. The parabola 4x + y2 = 0 divides the
(x, y)-plane into two parts. In each of them, the increment is a smooth function. To
the left of the parabola the eigenvalues are complex and f = y/2. To the right, the
eigenvalues are real and f = (y ± -J4x + y2)/2. The level curves of the increment are
tangent rays to the parabola.
All points of the parabola are singular points of the decrement diagram. To them
there correspond matrices A with Jordan blocks of order 2. To the left of the parabola
the increment changes linearly, and to the right it changes as the root.
It is clear that the singularity indicated here is unremovable by a small perturbation
of the family. There exist other unremovable singularities as well, our purpose is to
give their complete list.
If one real eigenvalue or one pair of complex-conjugate eigenvalues* has the maximal
real part, then the increment is a smooth function in the neighborhood of the value
20 of the parameter.
The smoothness is lost only in the case where the eigenvalue with maximal real
part is not unique. The matrices for which several eigenvalues have the maximum real
part form a closed semi-algebraic subvariety Fin the space 1R” of all matrices of order
n. The codimension of this variety is equal to 1. Its complement consists of two open
components:
. Stratum (a). Exactly one real eigenvalue has the maximal real part.
D2. Stratum (a ± ioj). Exactly one complex-conjugate pair has the maximum real
part.
* Here and in the following we assume that the members of a complex-conjugate pair are
not real.
252 6. Local Bifurcation Theory
It is easy to stratify the manifold F. The strata of maximal dimension (of codimension
1) are exhausted by the following list:
FY. Stratum (a2). Exactly two coinciding eigenvalues have the maximal real part:
these eigenvalues are real, and a Jordan block of order 2 corresponds to them.
F2. Stratum (a, a + ico). Exactly three eigenvalues have the maximal real part: one
is real and the other two form a complex conjugate pair.
F3. Stratum {a. + z'co,, a ± z'a>2). Exactly two distinct complex-conjugate pairs have
the maximal real part.
It is clear that the strata /j, F2, and F3 are regular smooth nonclosed disjoint
submanifolds of codimension 1 in the space R" of matrices. The remainder TV/7, U
F2 U F3) of F (the variety of matrices with several eigenvalues with maximal real part)
is a closed semi-algebraic* subvariety of codimension 2 in the space R" of all matrices.
The strata of maximum dimension of F\(Fy Uf2U F3) have codimension 2 in R" . It is
easy to list them:
G}. Stratum (a3). Exactly three eigenvalues have the maximal real part; they are
real and a Jordan block of order 3 corresponds to them.
G2. Stratum ((a ± ico)2). Exactly two coinciding pairs of complex-conjugate num¬
bers have the maximal real part; Jordan blocks of order 2 correspond to them.
G3. Stratum (a2, a ± ico). Exactly four eigenvalues have the maximal real part; a
Jordan block of order 2 corresponds to two real ones and the two complex ones form
a complex-conjugate pair.
G4. Stratum (a, a + iojy, a ± ioj2). Exactly five eigenvalues have the maximal real
part: one real and two distinct complex conjugate pairs.
G5. Stratum {a ± zoz,, a + z'a>2, a + ico3). Exactly three distinct complex conjugate
pairs have the maximal real part.
*A semi-algebraic subvariety of a linear space is, by definition, a finite union of sets given
by finite systems of polynomial equations and inequalities.
tAll varieties D-„ Zj, G, are connected for sufficiently large n. Exceptions: for n = 2: D2
and Fy, for n = 4: F3, G2, and G3, and for n = 6: Gs have two components.
§ 30. Matrices Depending on Parameters and Singularities of Decrement Diagrams 253
In the complement of the set F of singularities, the increment /is a smooth function
of the parameters. Nevertheless, at some points of this complement, the decrement
diagrams may have singularities: they are the critical points of the function /
Outside F, the increment of a generic family has only simple critical points, i.e.,
points of the following three types (or six types, if we distinguish between the cases
(£>,) of real eigenvalues and (D2) of complex eigenvalues):
D°. Minimum. In the neighborhood of the critical point of the parameter plane
under consideration, we can choose smooth coordinates (x, y) such that the increment
will have the form f = const + x2 + v2.
£>/. Saddle Point. In appropriate coordinates, we have/= const + x2 — y2.
D2. Maximum, f = const — x2 — y2.
Let us now study the behavior of the function /near nonsingular points of F, i.e.,
near interior points of the curves / in the parameter plane. We distinguish between
two cases: a point of a curve F, is either noncritical for the increment viewed as a smooth
function on the curve, or it is critical.
The transversality theorem implies that in generic families, critical points of the
restrictions of the increment to the curves / can only be nondegenerate maxima or
minima.
Combining this information with the explicit formulas of versal families of matrices
from § 30B, we obtain, without any difficulty, the following normal forms of the
increment near points of strata of codimension 1.
Figure 117.
if x 0,
/ const + y +
if * ^ 0.
f= const + x + |>’|.
The curves F° and F° divide the domains Z), and D2 of real and complex roots.
The level curves of the increment meet F: tangentially from the side of the real roots
and transversally from the side of the complex roots. The level curves of the decrement
meet F2 and F3 transversally from both sides at the points F2 and F°. The angle of level
curves (smaller than 180°) on the segments f contains the direction of the decrease of
/along the curve in all cases.
The four values of k are obtained by combining the two signs of e and the two
possibilities for a:
k 1, 2 3, 4
1
a | (0, 1) (1, +30)
Figure 118.
this case, the level curves of / consist of segments of two parabolas translated along
the y-axis.
7 C/p if y 0,
./ = const + ex- + cp(y) + <
(_ 0 if y 0.
k 1 2 3 4
+
sign of £, sign of a
1
++
1
1
1
Combining the transversality theorem and the explicit formulas for versal families
of matrices from § 30B, we obtain the following normal forms of the increment near
points of codimension 2 strata.
Theorem. In the neighborhood of a point of every codimension 2 stratum (G, in the notation
of § H) in the plane of parameters of a generic family, smooth coordinates (x. y) can be
chosen so that the increment f assumes one of the 18 forms listed below {Fig. 119).
where /. is the largest among the real parts of the roots of the cubic equation /.3 = x/. + y
and tp is a smooth function such that (f <p/Cx)( 0, 0) = a ^ 0.
The form of the decrement diagram is determined by the sign of the number a.
The signs “ + ” and “ —” in Gf correspond to a > 0 and a < 0. To clearly see
the form of the decrement diagram, it suffices to consider the cases <p — ±x. Two
singular curves meet tangentially at the point x = y = 0: a ray F2 {y — 0, .v < 0) and
one-half of a semicubic parabola Fj (4x-3 = 21y2,y < 0). These two curves separate
the (convex) domain D2 of complex-conjugate roots from the domain Z), of real roots.
In moving along the boundary of the domains £), and Dz, the increment / changes
monotonically if a > 0 and has a minimum at the point G,- if a < 0. From the side
of D,, the level curves of / are tangent to the semicubic parabola Fi.
Here Re is the real part and <p is a smooth function such that {d<pjdx){0, 0) = a ^ 0.
§ 30. Matrices Depending on Parameters and Singularities of Decrement Diagrams 257
The form of the decrement diagram is determined by the sign of the number a.
The signs + and — in G 2 correspond to a > 0 and a < 0. To visualize the form of
the decrement diagram clearly, it is enough to consider the cases (p = ±x. The ray F3
(y = 0, x < 0) arrives (and ends) at the point x = y = 0. In the case a < 0, the
function /has a minimum at the point GJ(x = y = 0). In the case a > 0, the point
GJ {x = y = 0) is topologically nonsingular for / The level curve of /going through
this point has a singularity of semicubic type.
\Jx, <p(x, y) if x ^ 0,
/ const + y + max
0, <p(x, y) if x 5= 0.
The six values of k are obtained by combining the two possibilities for the sign of
a and the three intervals of variation of b:
k I 2 3 4 5 6
sing of a + + - +
interval of b (0, +oo) (0, +30) (-1,0) (-1,0) (-00,-1) (-30,-1)
To visualize the form of the decrement diagram clearly, it is enough to consider a linear
function q>. Three smooth rays , F2, and F3 enter the origin. Fl and F2 come from
opposite directions (with tangency of the first order), and F3 comes transversally from
the side D2 of complex roots. In the case G3 (i.e., where a < 0, b < — 1), the increment
has a minimum at the point x = y = 0; in the remaining cases, the point G3(k =F 5)
is a topologically nonsingular point of/
k 1 2 3
To see the form of the decrement diagram clearly, it suffices to consider a linear
function <p.
In each of the three cases (k = 1,2, 3), three smooth branches of the curve F3
converge transversally at the point Gk. In the last case, this point is a minimum of the
increment; in the first and second cases, it is a topologically nonsingular point. In
approaching the point Gk along k of the three rays, the increment decreases. It increases
on the remaining ones.
258 6. Local Bifurcation Theory
Cases Gj(k = ±1, ±2, 3) (Double Passing with the Participation of a Real Root).
The increment is given by the same formula as in cases G5\ but we have to distinguish
between more cases depending on the question to which of the sectors the real root
corresponds.
If k is negative, one obtains cases in which, in approaching the point G4, the incre¬
ment increases on the curve F3 (on which the complex pairs collide). The two other rays
are branches of F2.
K. Discussion
Considering the normal forms listed above, we can arrive at a series of conclusions of
general character concerning the structure of the decrement diagram both locally and
globally. First of all, our theorems imply the following corollary.
These points are minimum points of the types Df Ff, G,_2, G|, G4 5.
The points D} and F,3 are topologically equivalent to a saddle point. In the neigh¬
borhood of the maximum points (Df), the increment is a smooth function. The points of
the remaining types are topologically nonsingular.
This corollary obviously implies inequalities for the numbers of the singular points of
various types. In particular, if a closed level curve of the increment encloses a simply
connected domain, then the total number of points of types Df1, Ff G^ 2, G 3, G3 5 inside
this domain exceeds the number of points Df, F3 by l. It is not known whether the
corollary can be carried over to /-parameter families with / > 2*.
The fact that the segments F, and F2 together form closed curves and the description
of the singularities at the ends of the segments F3 imply the following.
Corollary. If the parameter space A is a closed two-dimensional manifold, then (i) the
numbers oj points of types G, and G3 have the same parity and (ii) the total number of
points of type G2, G3, G4, or G5 is even.
If, for the parameter space A, we take a compact domain with boundary transversally
intersecting F, and not going through the points G,, then the result is changed in the
following way: the total number of points of type G{ or G3 has the same parity as the
total number of points of intersection of the boundary with F, or F2, and the total
number of points of type G2, G3, G4, or G5 has the same parity as the number of points of
intersection of the boundary with F3.
In particular, this study of the increment enables us to study the singularities of the
boundary of stability (i.e., curves of increment zero) in the parameter plane of generic
two-parameter systems. Our theorems imply the following corollary.
* We note that in the case 1=2, the singularities of the increment of a generic family are
of the same type as the singularities of the largest real part of the roots of an algebraic equation
whose coefficients are generic functions depending on / parameters. For / ^ 3, this is not so:
the increment can have more complicated singularities.
§ 30. Matrices Depending on Parameters and Singularities of Decrement Diagrams 259
We note that the corner points of the boundary of stability can be of types F,°
(“Jordan 2-block”) or F°, F° (“simple passing”), according to the classification of § 301
and § 30J. Therefore, each of the arcs of the boundary of stability can be continued
through its endpoints without losing smoothness. Moreover, the total number of corner
points of types F° or F° on every closed component of the boundary of stability is
always even.
We also note that our analysis of the singularities of the increment of two-parameter
families is sufficient for the study of the boundary of stability in three-parameter
families.
Indeed, according to the transversality theorem, the singular points of strata of
codimension 3 and the critical points of the restrictions of the increment to strata of
codimension 0, 1, or 2 can be removed from the boundary of stability by a small per¬
turbation of the family. Consequently, the boundary of stability of a general family
consists of smooth surfaces, and its singularities lie on curves in which the boundary of
stability intersects the surfaces of types Ft and at points of intersection of the boundary of
stability with the strata G, (the latter appear in the form of curves in generic three-
parameter families).
Moving along such a curve G,, we may consider our three-parameter family as a
one-parameter family of two-parameter families (two of the parameters are coordinates
in a small area transversal to G, and one is the coordinate t along G,). Considering the
normal forms of § 30J and § 30K, we now have to assume that all constants and the
arbitrary functions <p depend smoothly on the parameter t. Moreover, generically, we
may use the functions <p{x, y, t) as the parameter z. This leads to the following conclu¬
sion.
The angles of the boundary of stability are always directed outside, driving a wedge
into the domain of instability. This is apparently the consequence of a very general
principle, according to which everything good is fragile1.
What has been stated previously also implies some global properties of the boundary
*In generic ^-parameter families of matrices, only a finite number of singularities of the
stability boundary occur for k arbitrarily large (up to a local diffeomorphism of the ambient
space (Levantovski)).
260 6. Local Bifurcation Theory
of stability. For example, if the boundary is closed, then the total number of vertices of
types (G,, i > 1) is even, just as the total number of vertices of type G, or G3 are even.
The proofs of the theorems above may be found in the article V. I. Arnol’d, Lectures
on bifurcations and versa! systems, Uspekhi Mat. Nauk 27, 5 (1972), 119-184 [Russ.
Maths. Surveys, pp. 54-123]. See also, L. V. Levantovski, On the singularities of the
stability boundary, Vestn. Mosk. Univ. Ser. 1 Mat. Mekh., 1980, 6, 20-22; and L. V.
Levantovski, On the boundary of the set of stable matrices, Uspekhi Mat. Nauk, 35, 2
(1980), 213-214. L. V. Levantovski, Singularities of the stability boundary, Funct.
Anal. Appl. 16, 1 (1982), 44-48.
<4 In the neighborhood of the point and the value of the parameter studied, a
family of fields in an ^-dimensional phase space is given by n functions of
n + 1 variables (n phase coordinates and the parameter e). The singular
points are given by a system v(x, e) = 0 of n equations for the n + 1 variables.
By the implicit function theorem, these equations locally determine a
smooth curve x - y(£) if at the original point the determinant of dv/dx is
different from zero. On the other hand, this determinant is equal to the
product of the eigenvalues of the linearization of the field at the singular
point. It is different from zero by assumption. ►
Remark. The singular points at which all eigenvalues of the linearized field
are different from zero are called' nondegenerate. Consequently, if a field
depends smoothly on a parameter, then its singular points depend smoothly
on the parameter, provided that they remain nondegenerate.
The above proof remains valid for any dimension of the parameter space.
All singular points of a generic field are nondegenerate. Nevertheless, if we
consider a family of vector fields, then degeneracies, which are unremovable
by a small perturbation of the family may occur for some values of the
parameter.
We study degeneracies in generic one-parameter families for vector fields
in an rc-dimensional space.
Let us consider the (n + l)-dimensional space which is the direct product
of the phase space with the axis of the parameter e. We shall denote a point of
the phase space by x. Our family determines a family
x = v(x, e)
of differential equations.
In our n + 1-space, we consider the set formed by the singular points of
the equations of the family for all values of the parameter (Fig. 121):
Figure 121.
262 6. Local Bifurcation Theory
Theorem. For a generic family, the set of singular points is a smooth curve.
Here and in the following, the words “generic families” mean “families
from an everywhere dense set in the space of all families”; the everywhere
dense set in question is open if the domain of definition of the family is
compact or if the families are considered in the fine topology (cf., § 29); in
any case, this everywhere dense set is the intersection of a countable number
of open sets.
The theorem follows from the transversality theorem (§ 29) or from
Sard’s lemma {§ 10).
Indeed, by the implicit function theorem, T is a locally smooth curve if 0 is
not a critical value of the local mapping (Rn+1 —► IR", (x, e) t—> v(x, e). On the
other hand, for a generic mapping, the value 0 is not critical.
><
Figure 122.
§31. Bifurcations of Singular Points of a Vector Field 263
We assume that the set of singular points of the family is a smooth curve
(rank(ru/c(x, e)) = n). We consider the mapping of projection of this
smooth curve onto the axis of the values of the parameter. The points at
which the curve is projected badly onto the g-axis are exactly the degenerate
singular points. Indeed, by the implicit function theorem, the curve of
singular points is the graph of a smooth function of the parameter in the
neighborhood of a nondegenerate singular point.
Remark. The theorem states that, as the parameter varies, the singular points
of a generic family may only annihilate each other pairwise or be born in
pairs, when the parameter passes through bifurcation values (Fig. 121). The
bifurcations of this type are stable (are preserved under a small perturbation
of the family). All more complicated bifurcations disintegrate into several
bifurcations of the type described (Fig. 123) under a small generic pertur¬
bation.
x = ±xz + e, x e IR, e e R.
For £ = 0, this vector field has the simplest nondegenerate singular point
(x = 0). If the parameter passes through the regular bifurcation value £ = 0,
then, depending on the sign of x2, either the two singular points (stable and
unstable) annihilate each other or a pair of singular points is born whose
members run apart immediately (with asymptotics VRI)-
It is easy to verify that the bifurcation in this example is the only bifurca¬
tion that cannot be removed in generic one-parameter families of vector
fields on a line.
The function v(x, e) determining the field changes its sign on the curve T.
We choose the origin of the coordinates (x, e) at the bifurcation point. In
view of the nondegeneracy of this point, the equation of T has the form
£ = Cx2 + 0(|x|3), C ^ 0. This immediately implies our state. ^
§ 32. Versal Deformations of Phase Portraits 265
The theorem just proved, together with the theorem of § 31B, furnishes a
complete topological description of bifurcations of singular points of vector
fields on a line in generic one-parameter families.
Remark. The search for the periodic solution <D in the form of a series in e is
called the Poincare method of a small parameter. The solution cp is called the
generating solution. An analogous method can be used in the nonautonomous
case where v has period T(e) in t and we look for TXQ-periodic solutions.
Problem. With an error on the order of e2, find a 27t-periodic solution of the equation
x = sin x + e cos t with x = 0 for e = 0.
Example
Figure 126.
y = —y y e R"~
z = z, z e R" +
where n_ and n+ are the numbers of the roots of the characteristic equation in
the left and right half-planes, respectively. For example, for n — 2, this
system describes the coalescence of a nodal and saddle points (Fig. 126). For
£ = 0 we obtain a so-called saddle-node.
In § 31 we called bifurcation points those points in the direct product of the
phase space with the space of the values of the parameter for which the
characteristic equation has a vanishing root.
*As usual, the set of generic families is the intersection of a countable number of open
sets. It is open if the domain of definition of the family is compact or if we use the fine topology.
§ 32. Versal Deformations of Phase Portraits 269
x0 = 0 e R", e0 = 0 e Rfc.
y - -y, y e R"-,
z = z, z e R”+.
The existence of the five foliations was proved by Tikhonova, independently of the
needs of bifurcation theory [E. A. Tikhonova, Analogy and homeomorphism of perturbed
and unperturbed systems with a block-triangular matrix. Differential Equations 6, 7
(1970), 1221-1229], and by Hirsch, Pugh, and Shub [M. W. Hirsch, C. C. Pugh, M.
Shub, Invariant manifolds. Bull. Am. Math. Soc. 76, 5 (1970), 1015-1019], The case
n+ — 0 was considered earlier by Pliss [V. A. Pliss, A reduction principle in stability
theory of motions, Izv. AN. USSR, Ser. Mathem. 28, 6 (1964), 1297-1324],
Example
From what has been said above, it specifically follows that, when a pair
of singular points in a generic one-parameter family of vector fields is
born, one (and only one) phase curve leads from one of the new singular
points to the other (for values of the parameter close to the bifurcation
values).
In the analysis of singular points performed here locally in z, we observe that the
singular point loses stability at the moment e = 0, but we omit an important phenomenon
§ 33. Loss of Stability of an Equilibrium Position 271
connected with the loss of stability: the birth of a limit cycle (cf., Fig. 129). To avoid
such a mistake, one must consider a neighborhood of zero in the (z, e)-space and not
in the z-space for fixed e.
p = 2p(e + cp), p ^ 0.
In the case c > 0 (Fig. 130), a limit cycle exists for e < 0 and is unstable.
As e converges to 0, the cycle settles onto an equilibrium position, which was a
stable focus for e < 0. For s = 0, the focus becomes unstable (the instability
is weak, nonexponential). For positive £, the focus is unstable even in linear
approximation.
This loss of stability is called hard for the following reason.
Imagine that the system is near a stable equilibrium position and that
under the variation of the parameter this equilibrium position loses stability.
In the case c > 0, as £ approaches 0 from the negative side (or even somewhat
sooner) the ever present perturbations pull the system out of the neighbor¬
hood of the equilibrium position and it jumps into some other regime
(for example, a faraway equilibrium position, limit cycle, or a more com¬
plicated attracting set). Consequently, under the continuous variation of
the parameter, the behavior of the motion changes stiffly, in a jump.
In the case c < 0, although the amplitude of the self-sustained oscillations
born from the equilibrium does not depend on the parameter smoothly
(radical singularity) it does so continuously; in this sense, the behavior of
the motion changes softly.
In the study of Eq. (1), we have essentially used the “versal” point of
view: if we had considered a neighborhood in the z-space for fixed £ instead
of a neighborhood in the (z, £)-space, we would have missed limit cycles.
This agrees with the fact that a degeneracy of codimension k has to be
studied in a ^-parameter family: our case of codimension 1 is imbedded
in a one-parameter family.
In fact, the example just considered exhausts the bifurcations of the phase
portrait in generic one-parameter families occurring under the loss of stability
of an equilibrium position in the plane and, more generally, under the passage
of a pair of roots of the characteristic equation through the imaginary axis.
Theorem. Generic families with the indicated properties are locally topolog¬
ically equivalent to the family of the preceding example.
§ 33. Loss of Stability of an Equilibrium Position 273
Z; = Al (e)z ! + al(e)zfz2 + • -
i2 = a2(e)z2 + b,(e)zlz2 + • • •.
We note that since the original equation is real, the eigenbasis can be chosen
from complex-conjugate vectors, and the normalizing changes can be chosen
real. In such a case, the second equation follows from the first by conjugation.
Moreover, z2 = z, in the real plane, and we may write only the first equation,
denoting z, by z and z2 by z. This equation can be considered as another
version of the original system in the real plane IR2 written as an (nonholo-
morphic) equation on the complex line C1 with coordinate z :
z = 2j(£)z + al(e)z2z + • • •.
z = Xi(e)z + al(e)z2z.
This can be studied in the same way we studied the special example in § 33A.
The correspondence between the two kinds of notation is as follows:
274 6. Local Bifurcation Theory
It is easy to see from this formula that 0(p2) does not affect the bifurcations
of the phase portrait at some neighborhood (independent of e) of the
origin. ►
The theorem above was essentially known to Poincare; the explicit formulation and
proof were given by Andronov [A. A. Andronov, “Mathematical problems of the
theory of self-sustained oscillations.” In; Pervaya vsesojuznaya konferencija po kole-
banijam (I all—Union conference on oscillations), M. L. GTTI, 1933 (reproduced in
A. A. Andronov, Collected Works, AN SSSR, Moscow, 1956, pp. 85-124). A. A.
Andronov, Application of Poincare's theorem on “bifurcation points” and “change in
stability” to simple autooscillatory systems, C. R. Acad. Sci. (Paris) 189, 15 (1929),
559—561; A. A. Andronov, E. A. Leontovic-Andronova, Some cases of the dependence
of periodic motions on a parameter, Ucenye Zapiski Gorki Gosudarstvenny Univ.,
1939, Vol. 6, p. 3 (see also A. A. Andronov, Collected Papers, pp. 186—216)]. R. Thom
to whom I taught this theory in 1965, began to promote it under the name “Hopf
bifurcation” (cf., for example, the work ofSmaleand Hirsch*). Curiously, the 20 pages
of bibliography in The Hopf Bifurcation and Its Applications by J. E. Marsden and
M. McCracken (as well as the 30 pages of bibliography in the second edition of
Foundations of Mechanics by R. Abraham and J. Marsden) does not contain the basic
Andronov-Leontovich-Andronova paper. The theorem is also presented in the well-
known Andronov-Vitt-Chaikin book Theory of oscillations, Moscow (1937) (English
translation by University Press, Princeton (1949)).
u = —u, u e R"-,
v = v, v e Rn+, n = n_ + n+ + 2.
Example
Let n = 3, n+ =0, and let the sign in front of zz be negative. In this case,
the theorem asserts that in the neighborhood of the origin of coordinates
not depending on e, the birth of an invariant cylinder of radius yfe. attracting
neighboring trajectories takes place as the pair of eigenvalues passes through
the imaginary axis. On the cylinder itself, there is a stable limit cycle onto
which all trajectories wind. Consequently, this case corresponds to a soft
loss of stability with the appearance of self-sustained oscillations.
This degeneracy has been studied by many authors; in particular, Hopf studied the
birth of a cycle in the multidimensional case [E. Hopf, Abzweigung einer periodischen
Losung von einer stationaren Losung, Bereich Sachs. Acad. Wiss. Leipzig, Math. Phys.
Kl. 94, 19 (1942), 15-25], Further results have been obtained by Neimark [J. I. Nelmark,
On some cases ofperiodic motions depending on parameters, Dokl. Acad. Nauk SSSR 129
(1959, 736-739] and Bruslinskaja [N. N. Bruslinskaja, Qualitative integration of a system
of n differential equations in a region containing a singular point and a limit cycle, Dokl.
Acad. Nauk SSSR 139 (1961), 9-12],
Nevertheless, the general theorem formulated above, covering a complete study of
the metamorphoses of the phase portrait (and not only bifurcations of a cycle) has only
been proved in Sositaisvilli’s work on reduction (cited above) using two-dimensional
results of Andronov and Poincare.
Figure 131.
where the coefficient v denotes viscosity; /is the field of nonpotential mass
forces; the pressure p is determined from the condition of incompressibility.
On the boundary of D, we may have, say, conditions of adhesion (v\dD = 0).
It is assumed that the initial field of velocities determines the entire motion, so that
the equation defines a dynamical system in the infinite-dimensional space of divergence-
free vector fields equal to 0 on the boundary of D. [In fact this has only been proven in
the two-dimensional case. Extensive literature is devoted to problems of existence,
uniqueness, and properties of solutions of the Navier-Stokes equations; nevertheless,
the basic problems have remained unsolved.]
It seems that all models in which hyperbolic attracting sets have so far been found
contain terms of the type of a pump or negative viscosity, which are absent in the
Navier-Stokes equation. At any rate when, in 1964, this author attempted to find a
hyperbolic attracting set in the six-dimensional phase space of the Galerkin approxima¬
tion of the Navier-Stokes equation on the two-dimensional torus with sinusoidal
external force (using a computer programmed by Vvedenskaya), the attracting set turned
out to be apparently the three-dimensional torus (maybe because of the too small
Reynolds’ number). As far as this author knows, hyperbolic attracting sets for the
Navier-Stokes equations or their Galerkin approximations have not yet been found.*
On the other hand, the numerical experiment described above has served as a starting
point for a series of publications on the application of geodesic flows on groups of
diffeomorphisms to hydrodynamics. [V. I. Arnold, Sur la geometrie differentielle des
groupes de Lie de dimension infine et ses applications a I’hydrodynamique des fluides
parfaits, Ann Inst. Fourier, Grenoble 16, 1 (1966), 319-361; D. G. Ebin, J. Marsden,
Groups of diffeomorphisms and the motion of an incompressible fluid, Ann. Math. 92
(1970), 102—163; A. M. Lukatskii, On the curvature of the group of measure-preserving
diffeomorphisms of a 2-sphere, Functional Anal. Appl. 13, 3 (1979), 23—27; A. M.
Lukatskii, On the curvature of the group of measure preserving diffeomorphisms of an
n-torus, Uspekhi Mat. Nauk 36, 2 (1981), 187-188. That the curvature is negative
indicates exponential instability of the corresponding attractors. Lukatskii found
negativeness for flows on S2 and T", for most sections, and negativeness of the Ricci
curvature (per unity of dimension). See further applications of these ideas in D. D.
Holm, J. E. Marsden, T. Ratiu, A. Weinstein, Nonlinear stability of fluid and plasma
equilibria, Phys. Rep. 123, 1&2 (1985), 1-116; A. I. Shnirelman, On the geometry of
the diffeomorphism groups and on ideal fluid dynamics. Mat. Sbornik 128 (170), 1(9),
(1985), 82-109; McIntyre and T. G. Sheperd, An exact conservation theorem for
finite-amplitude disturbances to nonparallel shear flows, with remarks on Hamiltonian
structure and on Arnold's stability theorem, J. Fluid Mech. 181 (1987), 527-565.]
A very simple model with unstable trajectories on an attracting set has been proposed
by Henon [M. Henon, A Two-dimensional mapping with a strange attractor. Comm.
Math. Phys. 50 (1976), 69-77]. Henon considers a quadratic “Cremona transformation”
on the plane having the form T — 7] 7] 7], where
Interestingly, the experimentally observed (for a = 1.4, b = 0.3) attraction to the set
having locally the form of the product of the Cantor set with an interval, could not be
* We may think that when the viscosity becomes small (the Reynolds’ number becomes large)
the minimum of the dimensions of the minimal attractors grows indefinitely. But this is not
proved even for the maximum of these dimensions (it has not been proved rigorously that this
maximum exceeds one).
§ 33. Loss of Stability of an Equilibrium Position 279
Figure 133.
The determination of the regime occurring after the Poiseuille flow loses
stability is, in the opinion of the experts, on the borderline of the capabilities
of modern computers.
In this situation one must probably not disregard the qualitative pre¬
dictions which can be made completely without calculations, relying on
the general bifurcation theory expounded above.
In the problem under consideration, there are two parameters, a and R.
Consequently, besides singularities of codimension 1, we may also encounter
singularities of codimension 2. We direct our attention to the one that is
associated with the change of the sign of c.- Calculations show that for a
sufficiently large Reynolds’ number R, the rough generation on the lower
side of the tongue of the loss of stability becomes soft. In order to understand
what happens at that moment, we have to construct a two-parameter versal
family for such a two-fold degeneracy. Such a family can easily be con¬
structed. It has the form
1. c2 < 0, e2 < 0. When r,t passes from negative to positive values, the
systems exits softly to a periodic stable regime of self-sustained oscillations
(Fig. 134).
quencies; undoubtedly, this can be explained by the fact that other dynamical systems
were not known to him.
In 1965, this author lectured on the theory discussed above in the seminar of Thom at
IHES at Bures-sur-Yvette. Six years later, Ruelle and Takens constructed [in On the
nature of turbulence. Comm. Math. Phys. 20 (1971), 167-192; 23 (1971)] examples of
the loss of stability of a cycle with the occurrence of a behavior more complicated than
a quasiperiodic one; however, their example has an exotic character, since it corresponds
to a metrically very skinny (although open) part of the space of the parameters of the
deformation.
We would like to mention that for the applicability of the results of the indicated
works, it is necessary that the loss of stability take place in a soft regime, whereas the
behavior of the loss of stability of the Poiseuille flow turned out to be rough.
For a review of the subsequent experimental work, cf., J. B. McLaughlin, P. C.
Martin, Transition to turbulence of a statically stressed fluid system, Phys. Rev. Lett.
33 (1974); Phys. Rev. A 12 (1975), 186-203.
E. Degeneracies of Codimension 2
x = ±x3 + + e2, x e IR
(Fig. 135). It is easy to verify that the deformation thus described is (topo¬
logically) versal; in the multidimensional case, a versal deformation can be
obtained by a suspension of a saddle.
The bifurcation diagram (for the case +x3) is illustrated in the left-hand
part of Fig. 135. The semicubic parabola divides the parameter plane into
Figure 135.
§ 33. Loss of Stability of an Equilibrium Position 283
I
Figure 136.
two parts. In the smaller part, the system has three equilibrium positions
near x = 0 and in the larger one there is one equilibrium. The metamorphoses
of the phase portrait, as the parameter goes around the point e = 0 on a
small circle, are shown on the right-hand side of Fig. 135. The direct product
of this circle with the one-dimensional phase space is an annulus, and the
equilibrium positions form a closed curve in this ring; the behavior of vectors
of the field is clear from Fig. 135.
4. An Imaginary Pair and Still Another Vanishing Root. These cases have
not yet been studied sufficiently to describe versal families; moreover,
it is not clear whether in the case of two imaginary pairs a two-parameter
(or, at least, a finite-parameter) topologically versal family exists (even
under the assumption of normal incommensurability of the ratio of fre¬
quencies in the case of their simultaneous passage from one half-plane to
the other). See, however, E. I. Horozov, Bifurcations of a vector field near a
singular point in the case of two pairs of imaginary eigenvalues, I and II,
Comptes Rendus de l’Academie Bulgare des Sciences, 34 (1981) and 35 (1982),
149-152; and also a paper by H. Zol^dek on case 4.
These problems then lead to the problem of bifurcations of phase portraits
of planar vector fields when the equations are averaged along the fast rotation
284 6. Local Bifurcation Theory
(or when finite-order Poincare normal forms are used). These special vector
fields have an invariant line (in the case of one imaginary pair and one zero
eigenvalue) or two intersecting invariant lines (in the case of two imaginary
pairs) for all values of the parameters, and have zero eigenvalues for zero
value of both parameters (at the origin which is a singular point of the vector
field). For the case of two pairs the equations are
x i = x2,
*2 = £1 + e2Xl + *1 ± *1*2
* H. Zondek, Bifurcations of certain family of planar vector fields tangent to axes, Journal of
Differential Equations 67, 1 (1987), 1-55.
§ 33. Loss of Stability of an Equilibrium Position 285
of equations in the plane with parameters (el9 e2). The bifurcation diagram
divides the £-plane into four parts denoted by A, B, C, and D in Fig. 137,
corresponding to the choice + x±x2 in the formula.
The phase portraits corresponding to each of the four parts of the £-plane
are shown in Fig. 137. To the branches of the bifurcation diagram there
correspond the systems (P, Q, R, S) with degeneracies of codimension 1
depicted in Fig. 137.
We note that the bifurcation on the branch S—the birth of a cycle from
a loop of the separatrix—is not included in our classification of singularities
of codimension 1, since it is not a local (near a singular point) but a global
phenomenon. We see, therefore, that in the local study of bifurcations of
singular points with the increase of the number of parameters of the family,
global bifurcations of small codimensions begin to play a role. It follows that
in a local problem for a sufficiently large number of parameters, we will
encounter the same difficulty of structurally stable systems being not every¬
where dense, which was discovered by Smale in the global problem of vector
fields on a manifold (cf., § 15).
The bifurcations in the case corresponding to the choice “ — ” of the sign
in the formula can be reduced to the preceding ones by changes in the signs
of t and x2 ■
Theorem. Vector fields generic among those with two vanishing roots of the
characteristic equation at a singular point in the phase plane have a topo-
286 6. Local Bifurcation Theory
This theorem, proved by Bogdanov in 1971, was first published in the survey of
V. I. Arnold, Lectures on bifurcations and versal systems, Uspekhi Math. Nauk 27,
5 (1972), 119-184 [Russ. Math. Surveys 27, (1972), 54-123], Takens announced an
analogous result in 1973. The proof of versality is not simple: the main difficulty is the
study of uniqueness of the limit cycle. Bogdanov overcomes this by nontrivial arguments
on the behavior of elliptic integrals depending on a parameter [cf., R. I. Bogdanov,
Bif urcations of a limit cycle of a family of vector f ields in the plane, Trudy Seminara
Imeni I. G. Petrovskogo, 1976, vol. 2, 23-36: A versa! deformation of a singular point of
a vector field in the plane in the case of vanishing eigenvalues, Ibid., pp. 37-65. English
translation: Selecta Math. Sovietica, I, 4 (1981), 373-388, 389-421].
Figure 138.
B. Simple Degeneracies
*One should better call it “contracting manifold”, because the neighboring phase curves are
repelled by this manifold, hence the motion along this manifold is highly unstable. The misleading
terms “stable”, “unstable” and notations “s”, “u” were introduced by Smale.
288 6. Local Bifurcation Theory
Examples
1. We consider the mapping of the x-axis into itself given by the formula
x i—»■ x + x2. The point x = 0 is fixed and its multiplier is equal to 1.
We consider the following one-parameter deformation with parameter
s close to zero:
I = x + x2 + e.
2. Consider the mapping of a linear space into itself given by the formula
where y, z, «, and v are points of four spaces whose direct product is our
space. We shall call such a mapping a standard saddle. (The dimensions
of the spaces to which y and u belong are arbitrary, and the dimensions
of the spaces to which z and v belong are equal to 0 or 1.)
We consider any smooth mapping with a fixed point. Let us assume
that none of the multipliers lie on the unit circle. Then in the neighborhood
of the fixed point, the mapping is topologically equivalent to a standard
saddle (which follows easily from the Grobman-Hartman theorem, § 13).
If the multiplier — 1 appears, then the closed phase curve depends smoothly
on the parameter and does not bifurcate itself. However, another closed
phase curve branches out, winding twice. In order to understand how this
occurs, we again appeal to the Poincare map.
Example
/oW = —X ± X3.
Figure 140.
292 6. Local Bifurcation Theory
case where the multiplier becomes equal to 1, when the multiplier generally
does not pass through the circle). At the moment of the passage of the multi¬
plier through — 1 from inside to outside, the fixed point loses stability. Then
there are two possibilities depending on the sign of the coefficient of x3:
(i) Along with the point which has lost stability (at a distance on the order of
the square root of the difference of the parameter from the critical value), a
stable cycle of period 2 arises (two fixed points of the square of the mapping).
This is a case of the soft loss of stability, (ii) The domain of attraction shrinks
to 0 because of the approach of a cycle of order 2 before the loss of stability
(rough loss of stability).
The multi-dimensional picture can be obtained by a suspension of a
saddle, as described above.
Applying all of what has been said about mappings to the Poincare map
of a closed phase curve, we obtain the picture in Fig. 141 in the case of the
soft loss of stability: the original cycle loses stability, but a stable cycle
apperars with a period which is approximately twice the original one.
The phenomena described here can well be observed in experiments. The following
example is borrowed from a lecture by Barenblat delivered at the seminar of I. G.
Petrovskil. We consider a polymer film stretched slowly by a weight. For small expan¬
sions, the process is quasistationary. (Time can be considered the parameter; the phase
point is in a stable equilibrium position; and all observable quantities are constant for
every value of the parameter, i.e., they actually change slowly with the variation of time).
However, for some value of the parameter (i.e., for a sufficient expansion of the film),
the picture changes and the form of the various physical parameters (say, the tension
x of the film) as functions of time becomes such as depicted in Fig. 142. (Every
oscillation in this picture can be considered as taking place for a fixed value of the
parameter, and in the next oscillation the parameter changes only slightly.)
The interpretation of this evolution of phase variables over time is as follows: Point 1
corresponds to a soft loss of stability of the equilibrium with the onset of oscillations;
it is clear that their amplitude grows in proportion the square root of supercriticality.
Point 2 corresponds to a soft loss of stability of a cycle with a multiplier passing through
- 1.
Indeed, assume that the metamorphoses indicated in Fig. 141 take place in the
phase space.
§ 34. Loss of Stability of Self-Sustained Oscillations 293
x i—» Axe~x
(here the multiple e~x describes the effect of the competition decreasing the Malthusian
growth coefficient A). (Fig. 143). For small values of A the fixed point 0 of the mapping
is stable (extinction). For larger values of A a positive, stable fixed point appears (Fig.
143), which later, at some value Al of the parameter, loses the stability while its eigen¬
value goes through — 1 (our mapping, of course, is not a diffeomorphism) generating
a stable cycle of period 2,* thus starting a sequence of doublings (Fig. 143).
These doublings were found immediately by ecologists when they started computer
experiments with the above model and other similar models (see A. P. Shapiro,
“Mathematical models of competition,” in: Upraylenie i Informaciya/Control and
Information, Vladivostok: DVNC AN SSSR, 10 (1974), 5-75; R. M. May, Biological
*This cycle explains, it seems, the well-known observation that the gorbusha (humpbacked
salmon) catch oscillates with a two-year period: years with a large gorbusha population are
followed by years with a small population, and vice versa.
294 6. Local Bifurcation Theory
Figure 143.
populations obeying difference equations', stable points, stable cycles, and chaos, J.
Theor. Biol. 51 (1975), 511—524; Simple mathematical models with very complicated
dynamics. Nature 261 (1976), 459-466). It was the analysis of these experiments that
led M. Feigenbaum to his remarkable discovery of the stability of infinite cascades of
doublings and to the even more remarkable universality law.
Let us consider a mapping of a line into inself, x i—► f(x), in a neighborhood of the
maximum point of function f The squared (iterated twice) mapping will have a graph
with two maxima and one minimum point between them, as is easily seen (Fig. 145).
In some neighborhood of the minimum point the graph, after a coordinates dilatation
and a sign change, looks very similar to the graph of the original function. A computer
experiment shows that this similarity becomes closer and closer under the sequential
doublings (squarings) of the mapping.
§ 34. Loss of Stability of Self-Sustained Oscillations 295
Figure 145.
This leads to an attempt at finding a function, which repeats itself exactly after the
squaring of the mappings (followed by a rescaling); that is, to find a fixed point for
the doubling operator J, defined on functions as is described below. Such a fixed point
happens to exist, namely, the even analytic function
(Jf){x) = -I/(/(-ax)).
a
The normalization constant a for <J> is a = 0.3995... = — <I>(1) (all the numerical
coefficients are given approximately).
The operator J, having fixed point <J>, is defined on the even C1-mappings /:
[—1, 1] —> [—1, 1] with one maximum point 0, satisfying the following conditions (see
Fig. 146):
for some a and b (depending on/). The operator 7is a hyperbolic (at point <F) mapping
of the functions space. It has a one-dimensional dilatating invariant manifold (a curve)
and a codimension 1 contracting submanifold. The eigenvalue corresponding to the
dilatation is the Feigenbaum constant 4.6692... .
The universality of the period-doubling cascades is explained by this picture as we
shall now see. A one-parameter family of mappings (even for the simplicity of the
arguments) is represented by a curve in the functions space (y at Fig. 147). If this curve
§ 34. Loss of Stability of Self-Sustained Oscillations 297
is not too far from O its images become closer and closer to the dilatating invariant
curve T+ under consecutive iterations by /.
On the other hand, let us consider the bifurcation surface E in the functions space,
consisting of the mappings/of a line having a fixed point with eigenvalue — 1. It turns
out that the dilatating invariant curve T+ intersects this hypersurface Z transversally.
Hence the preimages of the bifurcation surface E under / and under its iterations form
a sequence of hypersurfaces of codimension 1 in the functions space, converging to
the contracting manifold F_ of the fixed point <D when the number of iterations grows
(Fig. 147).
The points of intersection of these surfaces, with the curve y representing our
family, define the parameter values corresponding to the period doublings. This
explains the origin of the geometrical progression and the universality of its ratio: it
depends on the sequence of surfaces, and not on the special choice of the transversal
curve. For the (computer-assisted) proof and for a bibliography see P. Collet, J.-P.
Eckmann, Iterated Maps of the Interval as a Dynamical System. Boston: Birkhauser,
1980, 248 p.
The situation remains more or less the same for the doubling of cycles of differential
equations (for instance, the universal constant 4.6... remains unchanged), while some
of the details of the cascades of period doublings are different for differential equations
and for line mappings. The multiplier (the eigenvalue of the mapping) must change
its value from 1 to — 1 between two doublings. The reason is that the doubling cycle
has (at the moment of doubling) a multiplicator equal to — 1. Hence the doubled cycle
has (at the moment of its birth) a multiplier (— l)2 = 1. But the path from 1 to — 1
along the real line is forbidden for an eigenvalue of a diffeomorphism (and hence for
a multiplier of a cycle of a differential equation), since the eigenvalue of a diffeo¬
morphism may never become zero.
Thus the cascades of period-doubling bifurcations in differential equations and
diffeomorphisms involve a second multiplier, besides that coming from 1 at some point
on the real axis. After a collision both become complex (conjugate) and turn around
zero in the upper and lower half-planes, collide once more on the negative part of the
real axis, and finally diverge—one going to — 1, the other in the opposite direction.
The curve, followed by the multiplier, is extremely close to a circle, and displays some
universality features for repeated doublings (M. V. Jakobson, 1985).
The universal period-doubling cascades have their close analogs at other reso¬
nances. Let us consider, for instance, the 1 : 3 resonance when a multiplier leaves the
unit circle through the cubic root of unity. To observe such a phenomenon in a generic
family we need at least two parameters. In the parameter plane the moments of tripling
are represented by isolated points. These points occur, as for the doubling, in the form
of infinite series. The differences of consecutive tripling values of the parameters form
asymptotically geometric progressions whose ratio is a universal constant (the same
one for all generic families). But in contrast to the case of doublings the universal
constant this time is not a real, but a complex number, hence the tripling values of the
parameters on the parameters plane are organized in spiraling sequences.
There exist cascades of doublings in Hamiltonian systems too. This time the — 1
multiplier is always of even multiplicity, generically of multiplicity 2. In a doubling
the doubled cycle is born, having (at the moment of birth) a double multiplier equal
to 1. This pair of multipliers moves (until the next doubling) along the unit circle from
1 to — 1 in the upper and lower half-plane. The corresponding Feigenbaum constant
is much larger than for ordinary doublings (about 8). On the way from 1 to — 1 the
298 6. Local Bifurcation Theory
multipliers visit all the roots of unity. Thus between two doublings a lot of other events
occur with the same universality (family independence) as for the doublings: the birth
of cycles, whose periods are all kinds of multiples of the original one (at the moment
of birth). A newborn cycle of longer period has (at the moment of birth) a pair of
multipliers, equal to 1. Not infrequently these multipliers will travel to — 1 along the
unit circle. Thus, in principle, a very complicated system of bifurcation values occurs.
The bifurcation values of the parameter may be enumerated by all the sequences of
rational numbers.
The usual cascades of doublings correspond to the series 1, 2, 3, ...; the cascade
of doublings of a tripled cycle corresponds to 1/3, 1, 2, 3, ...; we may consider more
sophisticated cases, like 1/3, 1/3,... (triplings cascade) or 1/3, 1/2, 1/3, 1/2,...—every
such case may lead to new universal constants.
The consequences of these bifurcations were observed in the numerical experiences
of the early 1960s. For instance, the hat-shaped invariant curves (“islands”), see
Fig. 148) appear as a result of cycle quadrupling and were observed in the first
experiences of Henon, Heiles, and Chirikov* (with no explanation, it seems).
This case has been understood much less than either of the preceding two.
Topologically versal deformations are not known and may not exist. Never¬
theless, Poincare’s method allows us to obtain significant information. We
begin with the case where the argument of the multiplier falling on the unit
circle is incommensurable with 27i. (This case can be considered typical,
since the measure of the set of rational numbers is equal to zero.)
We shall assume that the dimension of the space being mapped is equal to
2. In this case, after an appropriate smooth change of coordinates depending
smoothly on the parameter, our family of mappings can be reduced to the
form
*M. Henon, C. Heiles, The applicability of the third integral of motion: some numerical
experiments, Astron. J. 69 (1964), 73; B. V. Chirikov, Studies of Nonlinear Resonances and
Stochasticity, Novosibirsk, Izv. Sib. Otd. Akad. Nauk. SSSR, N267, 1969.
§ 34. Loss of Stability of Self-Sustained Oscillations 299
where the real number e is the parameter of the family and 2(0) = e,cc, a #
2npjq. For a generic family, we have c/|/.|/<i£|0 ^ 0, which means we may
choose |A| — 1 for the parameter.
We assume that the term 0(|z|4) is absent. In this case, it is easy to study
the mapping. Indeed, the modulus of the image of a point is determined by
the modulus of the preimage, which gives rise to the real mapping:
| X | 11 + ar21 a; 1 + e Re ar2 + • - • .
Figure 149.
(1) Up to terms of degree q + 1 (and even up to terms of arbitrarily high degree) the
Poincare map coincides with a transformation of the phase flow of a vector field
on the plane.
(2) The indicated vector field is invariant with respect to a cyclic group of diffeomor-
phisms of the plane (of order q).
(3) Conclusions 1 and 2 hold not only for an individual Poincare map, but also for a
family depending smoothly on the parameters; moreover, both the group and the
field invariant under it depend smoothly on the parameters.
Remark 2. In general, the exact Poincare map is not a transformation of the phase flow
of any vector field and does not commute with any finite group of diffeomorphisms.
x = y, y = 0.
^ The Taylor series is unique, and, therefore, each of its terms has to turn
by the angle 2nfq when z turns by the angle Injq. The point zkzl of the
complex plane rotates by an angle 2n(k — l)lq. This rotation coincides
with rotation by an angle 2n!q exactly when the above condition is
satisfied. ^
§ 35. Versal Deformations of Equivariant Vector Fields on the Plane 303
Figure 150.
In the (k, /)-plane, we consider the integer points satisfying the congruence
k — / = 1 (mod q). These points are located on rays parallel to the bisector
of the positive quadrant and emanating from points corresponding to the
monomials z, zq+1, z2q+1, . . . ; zq~l, z2q~l, .... Among these monomials,
we shall seek monomials of smallest degree (Fig. 150).
First, let us successively obtain several monomials on the ray emanating
from the point z (i.e., monomials of the form z\z\2k) and then the monomial
z~q~l; all other monomials have degree not smaller than q + 1 (Fig. 150). ►
Definition. The preceding equation, without the term O, is called the principal
equation and is invariant with respect to rotation by an angle 2njq. The
right-hand side of a principal equation is called a principal ^-equivariant
field.
Example
z = ez 4- Tz|z|2 + Bz 2 ; z = ez + Az\z\2 + Bz 3.
Analogous definitions can be given for germs of vector fields and for
deformations in a class of fields with special properties (for example, with
a fixed linear part at a singular point). Now we pass to the construction of
versal deformations.
C. Principal Deformations
Definition. A field is said to be singular if the linear part of the field at zero
is equal to zero.
Example
In the cases q = 3,4, the principal deformations are given by the equations
Remark. The objects defined above arise in the study of the loss of stability
of a cycle as a pair of complex-conjugate multipliers crosses the unit circle.
In the function space of all systems, the systems with such a passage form a
hypersurface. This hypersurface is one of the three hypersurfaces bounding
the domain of stability. The other two hypersurfaces correspond to the
passages of one multiplier through the unit circle at the point 1 and at the
point — 1.
The boundary of the hypersurface corresponding to the passage of a
complex pair of multipliers consists of two parts (of two hypersurfaces of
codimension 2 in the function space of all systems). One of these surfaces
of codimension 2 corresponds to a pair of multipliers equal to 1 forming a
Jordan block of order 2 and the other to such a block with two eigenvalues
equal to — 1.
The study of the loss of stability in the neighborhood of these codimension
2 surfaces leads to the study of bifurcations in two-parameter families of
vector fields on the plane having a nilpotent Jordan block of order 2 as the
linear part and invariant under rotation by an angle 2nq, q = 2 (for multi¬
pliers equal to — 1) or q = 1 (for multipliers equal to 1). To include these
cases in the general scheme, it is convenient to give the following definitions.
D. Cases q = 1 and q = 2
^ The linear part of our field has the form ydjcx. We now construct the
homological equation corresponding to this linear field. To do this, we
306 6. Local Bifurcation Theory
calculate the Poisson bracket of our linear field, A = yc/Px, with an arbitrary
vector field h — PP/Px + Qc/cy.
We find
r r 11 r i ( n <
<
y — ,Px- y — ,Q — yQx^~ - Q-^,
II
=
L cx (XJ1 cx | cx cy (y cx
yPx - Q + u = 0, yQx + v = 0.
Here u and v are known functions, namely the components of the vector
field vr = uP/Px + vc/ry, which we wish to annihilate by a change of
variables.
We study this system. Expressing Q by the first equation and substituting
it in the second one, we obtain
y2Pxx = —yux — v.
[A, h\ + u- = (u0(*) +
cy
x = a(x) + yb(x).
also be chosen odd (commuting with the rotation), since in the preceding
formulas the degrees of (P, Q) and (u, v) are the same. Then, in the formal
normal form, the series of (a, b) will consist solely of terms of odd degree.
Restricting ourselves to a few first approximations in Poincare’s method,
we obtain the proposition formulated above.
and the vector fields defining them in the phase plane (x, y = x).
z = ez + Az\z\2 + Bz3, II
z = ez + Az\z\2 + Bz2, 4 = 3 ,
K>
II
Here the variables z, e, A, and B are complex, x, y, a, (3, a, and b are real, the
parameters of the deformation are denoted by Greek letters, and y = x.
“Theorem”. For every q, all principal singular fields can be divided into de¬
generate and nondegenerate fields in such a way that
(1) the degenerate fields form the union of a finite number of submanifolds
in the space of all principal singular fields ;
(2) the nondegenerate fields form the union of a finite number of open
connected domains ;
(3) the principal deformations of the germs of nondegenerate fields at zero
are versal;
(4) the principal deformations of the germs of nondegenerate f ields are
topologically equivalent in every connected component.
F. Description of Bifurcations
1m e ^ /'(Ree) + c|Ree\(q~2)/2, q ^ 5.
The proofs of the above theorem and statements are simple for q ^ 5.
They are contained in the publications of Bogdanov cited in § 33 in the case
q = 1, are unknown in the case q = 4, and are outlined below in the cases
q = 2, 3. Some versions of metamorphoses are illustrated below for <7 = 4
in Figs. 157, 158, and 160. [Cf., also, Additional Problems (1)—(4)].
1. Let A = 0. In this case, the system can be obtained from a certain Hamiltonian
system by a rotation of the field. Namely, consider an equilateral triangle formed by
singular points (saddle points). The sides of this triangle determine three linear non-
homogeneous functions. The product of these three functions is the desired Hamiltonian.
The sign of the derivative of this function along the solution of our system
z = ez T- Bz2 is determined by the sign of the real part of the parameter e. This enables
us to use the indicated function as the Liapunov function. Therefore, for A = 0, the
principal deformation can be studied without difficulty.
— = EZ + Z2 + \e\AZ\Z\2, E = —
dT ill] |£|
dZ
- + iZ + Z ■“
dT
4. The remaining proof of the versality of a principal family and of the form of
bifurcation diagrams and phase portraits can be carried out as in the case <7=1.
312 6. Local Bifurcation Theory
/. By expansions of x and changes of time and the parameters, the family can be
reduced to the form
y- x *
H = y - sgn a -- - sgn a —.
s/\a\b
2 0> + b'x2y. 0' = b'
>/M>
3. Integrating the rate of change of H along the level line H = 6, we obtain the
following condition of the birth of a cycle from exactly that line as a and 0 = «a pass
through zero: u = r2, where
r2 _ ff x2 dx dy
II dx dy
is the square of the radius of inertia with respect to the y-axis of the domain bounded
by the line H — h.
4. If |or| 10J, we make the substitution x: = kx\ t = xt', converting |u| into 1 and
|6|into 10|:
Here both parameters 0' ~ V|0| and a' ~ «/0 are small. For 0' = 0, we obtain the
following Hamiltonian:
y2 ax2 ax4
H ~ y - ~2 - 4~
The remainder of the study can be carried out in the usual way.
§ 35. Versal Deformations of Equivariant Vector Fields on the Plane 313
It has been shown above that the study of the behavior of cycles in our families can be
reduced to the solution of special cases of the following “weakened Hilbert’s 16th
problem”:
Let H be a real polynomial of degree n and let P be a real polynomial oj degree m in
the variables (x, y). How many real zeros can the function
"If
1(h) = | Pdxdy
have ?
For the study of symmetries of order 2, we need the case
2 2 4
y ax ax
P — p + Ax2, ti — — ■— - — -
2 2 4
For P = p + Ax2, the problem reduces to the study of the monotonicity of the function
r(h), where r is the radius of inertia with respect to the j^-axis of a domain bounded by
a cycle.
Lemma. On intervals between critical values of H, the function r behaves in the following
way :
property is not clear, but its investigation has led to the above-mentioned papers as
well as to the discovery (by B. and M. Shapiro) of the relation of the Chebyschev
systems to the natural stratification of the Schubert cells of the flag manifolds and to
their natural Bruhat ordering, which we will not discuss here.
z = £Z + Azjzj3 + z3.
—% = A + TV, where TV =
r
Let us now consider the circle with radius 1 and center at A (Fig. 154). The value of
s for which the point z is singular lies on the ray opposite the ray connecting zero with
the point A + TV on our circle. Moreover, the closer the point of the circle to zero,
the larger the modulus of e.
From what has been already stated, it is clear that the cases where \A \ is smaller or
larger than 1 are very different. If \A\ < 1, zero lies inside the circle. In this case, for
any e (except zero), the equation has four singular points at the vertices of a square.
If e goes around zero, rotating by 360°, the square of the singular points turns by angle
90° in the opposite direction.
If, on the other hand, \A \ > 1, then in the plane of the variable e, there is an angle
bounded by the extensions of the tangents to our circle. For e inside this angle, there
are eight singular points and none outisde. When e turns from one side of the angle to
the other, four singular points are born at the vertices of the square. This square is
immediately duplicated. Then nearby singular points begin to diverge. When £
Figure 154.
§ 35. Versal Deformations of Equivariant Vector Fields on the Plane 315
approaches the other side of the angle, every singular point of the first square dies,
colliding with a point of the second square, which initially stood at 90° angle from it
(so that one of the squares of singular points turns by 90° with respect to the other).
£ = PS + QL
Lemma. The type oj the singular point 0 does not depend on the argument oj Q. This point
is a saddle jor |F] < \Q\, a focus jor |Im P\ > \Q\, and a node jor |Im P\ < \Q\ <
the focus is stable jor Re P < 0 and unstable jor Re P > 0 {Fig. 147).
<3 By multiplying £ by a complex number A, the coefficient P does not change, and the
coefficient Q is multiplied by I/A. By changing A, we can make the argument of Q
arbitrary; this proves the first assertion of the lemma. For the proof of the second
assertion, we consider the case Q = 1. Let P = a + i/3. We write the matrix of the
equation in the basis (1, i). This matrix has the form
fa. + 1 — (3 \
M = ( ), tr M = 2a, det M = a2 + (32 — 1.
V P a - 1/
A1-2 = a ± V1 - p2.
they are real if\/i\ < 1. The roots have opposite signs if a2 + ft2 < 1. The lemma is
proved for £9=1. The conditions of a saddle, node, or focus for any |Q\ now follow
from similarity arguments: for dilations of time, P and Q are multipled by the same
real number. &■
Figure 155 shows clearly the relative location of nodes, foci and saddles in the
function space. This is useful to keep in mind in any study of bifurcations of singular
point in the plane.
ImP
Figure 155.
316 6. Local Bifurcation Theory
3. Study of the Saddle Points.We return to the original nonlinear equation. Let
z() = re‘v be a singular point. We linearize the equation at this point. Let z = z0 + f
Lemma. If\A\ < 1, then all singular points are saddles. lf\A\ > 1, then for every £ the
singular point with the smaller modulus is a saddle and that with the larger modulus is
not.
By the lemma of § 35J2, the condition for a saddle has the form \ A — N\ < \ A + 37V|.
We consider the points A — N and A + 3TV (Fig. 154). These points are symmetric
with respect to A + N, and the straight line connecting them goes through A. Which
of these points is closer to zero depends on which side of the tangent to our circle
at the point A + N lies the point zero. If \A\ < 1, then zero always lies on the same
side of the tangent (where A — N is located). If, on the other hand, \A\ > 1, then the
answer depends on which of the two arcs bounded by the tangents to the circle from
the origin lies A + N. The arc farther from zero corresponds to saddles and singular
points near 0 (cf., § 35J1).
1 + Re2 4
Im A
yjl - Re2 A
The corresponding line in the plane of the variable A is tangent to the circle \A\ — 1
at the points A = ±i and has the straight lines |Re^4| = 1 as asymptotes (Fig. 156).
5. Behavior at Infinity. For large z, we can “ignore” the term e.z. Setting u = r2,
we obtain the linear equation
vi- = 2 Aw + 2w.
To study this equation, we apply the lemma of § 35J2. Consequently, for \ A\ < 1, the
singular point in the n-plane is a saddle point, and for \A \ > 1, all trajectories from
infinity are attracted to a finite domain if Re A < 0.
6. Bifurcations of the Phase Portrait. If \A\ < 1, then the diagram (Fig. 157)
is apparently the same as for the third-order resonances (cf., § 35G). If [Re ^ | > 1,
then apparently the same occurs as for resonances of order 5 and above (Figs. 158
§ 35. Versal Deformations of Equivariant Vector Fields on the Plane 317
B■BBm^Pg
Figure 156.
lw!iS
Figure 157.
w* X X Figure 158.
and 153), although, in this case, singular points may be born not on cycles, as well,
[cf.. Additional Problems (1), (2), and (3)].
The main difficulties are represented by the case | Re A | < 1,|^| > 1.
7. New Normalizations and Notation. For the study of the case |Re/l| < 1 it is
useful to consider the asymptotics as |Im,4| -> x. Instead of letting A go to infinity,
we may let B converge to zero in the original equation. In order to study this case,
we introduce the notation
e = a + it, A = — ia — y, B = fi
and assume that fi, a = itfi and y = vfi (fi -* 0, u ~ u ~ 1) are small parameters of
the same order.
We choose the multipliers normalizing the expansion of coordinates and time so
that ot = 1 and r = 1 and introduce the symplectic polar coordinates p = |r|2/2,
(P = argz.
This transforms the original equation into the system
P — — + n p, (p — Hp.
Lemma. Let SH he the increment of H under one revolution along the dosed phase curve
H = h. Then
= — nqdp = f. An dpdq. ►
u cr
This method enables us to study our system outside the annulus in which p is close
to 5. Therefore, the case of a a close to 7 has to be considered separately.
— = »• + cos4<p, = -IP
dT dT
sin 4 <p
H
11nc\ =
— *P2 -p W(p
H = p — p2 — /?p2sin4<p — fiw<p,
4*',o^
n = ap2--(a - y)p.
where S is the area inside the well in the (p, <p)-plane and p0 is the coordinate of the
center of gravity of the well.
The condition of the birth of a cycle is a = 4yp0. Consequently, we need to calculate
p0. We know that p0 = \ + ~JPpx + - ■ ■ . Hence, the condition of the birth of a
cycle has the form a = 2y + 4y ^fppx + ■••,»■ = (y//?) + 4yp, 4- • ■ • .
We calculate p,. In the coordinates P, <p, the exact equation of the closed phase
curve H = const has the form
p + p = ypsin4cp
1 2 1 + /?sin4<p
p, = — 2 sin 4<p
(where the bar means “averaging” along the unperturbed phase curve for ft = 0).
From the location of the well with respect to the maximum and minimum of sin4cp
and its variation as w changes (Fig. 159), we obtain the information on the behavior
of Pi, on which the diagrams of metamorphoses are based (Fig. 160).
The system of metamorphoses depicted in Fig. 160 is realized if \lm A\ is large
compared to |Z?| and 0 < |Re/4| < B.
Of course, the above arguments cannot replace proofs and are only first steps in
the study of bifurcations in a principal family of 4-symmetric equations.
Results for the cases of symmetry of order q ^ 4 have apparently been known to
specialists for a long time; see, e.g., V. K. Melnikov, Qualitative description oj resonance
phenomena in nonlinear systems, Dubna, OIJaF, P-1013 (1962), 1-17. Takens announced
them in a preprint in 1973 but his proofs have not yet appeared. Our exposition is based
320 6. Local Bifurcation Theory
)
?c
M ip
||§!i 13
—
Figure 160.
IS
on this author’s article in Funct. Anal. Appl. 11 No. 2 (1977) 1 —10; detailed proofs have
been given by E. I. Horozov, Versal deformations ofequivariant vector fields for the cases
of symmetries of order 2 and 3, Trudy Sem. I. G. Petrovskogo 5 (1979), 163-192.
The proof relies on the fact that the reduction to normal form of terms
of degree not greater than N with retention of the resonant terms is imple¬
mented by diffeomorphisms depending smoothly on the parameters.
Combining Lemma 2 with the description of bifurcations of phase flows
from the preceding subsections, we obtain information on the loss of stability
of the fixed point 0 of the mapping f (or of the periodic motion for which
/ is the Poincare map).
L. Discussion
L For the translation of the results obtained above into the language of
bifurcations of periodic solutions, the fixed points in the plane have to be
replaced by closed trajectories in space, the separatrices of points by invariant
stable and unstable manifolds of these closed trajectories, and the limit
cycles in the plane by invariant tori. The situation will be significantly
different only for the metamorphoses of separatrices: while, in the plane.
322 6. Local Bifurcation Theory
*A homoclinic (heteroclinic) picture is, by definition, the net formed in a section plane by
the intersecting traces of the stable and unstable invariant manifolds of one (two) closed
trajectory (trajectories).
§ 36. Metamorphoses of the Topology at Resonances 323
of a torus, which soon becomes pinched along a parallel, so that the form
of the meridian of the torus approaches the shape of a figure eight, in
approaching the center (which represents an unstable cycle), the attracting
set, remaining close to the torus with the meridian almost contracted into
the eight, is destroyed near a homoclinic separatrix (Neimark).
In this case, the phase trajectory loops around one and then the other side
of the collapsed torus, jumping over from one side to the other seemingly
at random.
One obtains transversality with the sphere if the 1-form dr2 does not vanish
on the tangent plane of a leaf. On the other hand, the form A dt A- A dt
is a zero form only for A = 0. The relation A = 0 is not satisfied in the
Poincare case (and only in the Poincare case) for any z # 0. Hence, in the
linear case, the theorem is proved : the leaves intersect the sphere at a nonzero
angle a(z).
We consider the minimum a0 of the angle a(z) on the sphere \z\ — r. The
qunatity a0 does not depend on r (since a(cz) = a(z)). Hence, a(z) $: a0 > 0
for all z # 0.
Now we turn to the nonlinear system. The angle between the direction
fields of the nonlinear system and its linear part is small with |z|. Therefore,
in a sufficiently small neighborhood of zero, it is smaller than a0, and the
phase curves of the nonlinear system intersect the sphere transversally. ►
We find ourselves in the Poincare domain if the ratio A = Aj/A2 is not a negative
real number. First, we consider the foliation on S3 corresponding to the linear part of
the system.
The separatrices zt = 0, z2 = 0 intersect the sphere in large circles which are cycles
of the system on S3. Their linkage coefficient is equal to 1.
If A is not a real number (the case of a “focus”), then all remaining curves of the
foliation on the sphere wind away from one cycle and wind onto the other. We study
the Poincare map of the cycles.
We note that these maps can be assumed to be holomorphic. Indeed, they are
equivalent, with respect to real diffeomorphisms, to complex Poincare maps mapping
a holomorphic transversal to the separatrix into itself. Consequently, they become
holomorphic by an appropriate choice of the complex structure on the real two-
dimensional transversal to a cycle in S3. It also follows that the multipliers of our
cycles are equal to e±2ntX and e±2m,/l~‘
The foliations on S3 corresponding to all foci are homeomorphic but not diffeo-
morphic to each other: A2 + A-2 is an invariant under diffeomorphisms.
If A is real and positive (the case of a node), we also find ourselves in the Poincare
domain. In this case, the part of S3 between two linked cycles is foliated into two-
dimensional tori filled out by windings with the rotation number A the same on all
tori.
Now we consider a nonlinear system. In the case of a focus, resonance is impossible.
Therefore, in the nonlinear case, the foliation on a sphere is diffeomorphic to the
foliation described above, constructed for a linear system. The same is true for a
nonresonant node, i.e., for all A > 0, excluding the cases where A or 1/A is an integer.
We consider, for example, the resonance A = 2. In this case, the Poincare normal
form is
This system has only one separatrix for c ^ 0, and the foliation on S’3 has only one
cycle. We replace A by a nonreal value close to 2. The resulting system obtained on S3
is, on the one hand, close to a resonant system; on the other hand, it is diffeomorphic
to the system constructed from a linear focus studied earlier and has two cycles with
326 6. Local Bifurcation Theory
linkage coefficient 1. It can be shown that one of these cycles, Cx, is close to the unique
cycle C of the resonant system. The other cycle, C2, lies on a thin torus with axis C{
and closes after two trips along C1, making one revolution on the meridian (so that
the linkage coefficient of C2 and Cl is equal to 1). Hence, close to the resonance X — 2,
the metamorphosis of the system on S3 consists of a bifurcation of a two-fold periodic
trajectory from a periodic trajectory with eigenvalues (—1, — 1).
Remark. The preceding exposition follows the author’s article Remarks on singularities
of finite codimension in complex dynamical systems, Funct. Anal. Appl. 3, 1 (1969),
1-6. The results of this paper were later generalized by J. Guckenheimer, N. Kuiper,
N. N. Ladis, Ju. S. Il’jasenko, et al. Most complete are the results on the topological
type of the foliation given by a linear system in the neighborhood of a singular point
in a complex space.
We consider, in particular, the case where the phase space is three-dimensional and
the triangle of the eigenvalues contains zero in its interior. It turns out that the topolog¬
ical type of the foliation in a complex space is determined by the triple of the reciprocals
of the eigenvalues considered as a triple of vectors in a real plane (i.e., considered up
to linear transformations of the decomplexified plane of one complex variable). In
the multi-dimensional case, the real type of the collection of the reciprocals of the
eigenvalues determines the topological type of the complex foliation in a one-to-one
manner in the neighborhood of a singular point of a linear system, if zero belongs to the
convex hull of the eigenvalues and their pairwise ratios are not real [cf., C. Camacho,
N. Kuiper, J. Palis, C. R. Acad. Sci. Paris, 282, (1976), 959-961; N. N. Ladis Topological
invariants of complex linear flows. Differential Equations 12, 12 (1976), 2159-2169;
Ju. S. Il’jasenko, Remarks on the topology of singular points of differential equations in
a complex domain and Ladis’ theorem, Funct. Anal. Appl. 11, 2 (1977), 28-38].
In other words:
In the neighborhood of the point O a local family of analytic (,holomorphic,
smooth) vector fields with singular point O of Poincare type is analytically
{holomorphically, smoothly) equivalent to a family consisting of sufficiently
long segments of the Taylor series of these fields at O.
By assumption, the singular point is nondegenerate and, consequently,
depends smoothly on the parameter. Therefore, the singular point can be
moved to the origin of coordinates by a smooth change of variables depend¬
ing smoothly on the parameters. We assume that the eigenvalues are simple.
Then the eigenvalues can be chosen to depend smoothly on the parameters.
§ 36. Metamorphoses of the Topology at Resonances 327
*k = K(e)xk + ■ ■ • , k = 1, . . . , n.
Applying the Poincare method (Chap. 5), we shall annihilate only those
terms which remain nonresonant for £ = 0. Then the substitutions depend
smoothly on the parameter. Since the eigenvalues belong to the Poincare
domain, there are finitely many resonances, and the convergence of the
whole procedure can be proved without difficulty.
We obtain a coordinate system in which the right-hand sides of all
equations of the family are polynomials. ►
The case of finitely (or infinitely) differentiable right-hand sides can be studied
without difficulty, too. For details, cf., N. N. Bruslinskaja A finiteness theorem for
families of vector fields in the neighborhood of a singular point oj Poincare type, Funct.
Anal. Appl. 5, 3 (1971), 10-15, where the case of multiple eigenvalues is also considered.
D. Materialization of Resonances
It can be assumed that in the local problem treated above, the effect of
resonances on divergence has an analogous nature, but is connected with
changes in the topology of the foliation formed by the phase curves in the
complex rather than in the real domain. Such a change, even if it does not
manifest itself at all on the real part of the phase space, it necessarily obstructs
the analytic reduction and may interfere with the Cr-smooth reduction.
We note that the system xk = lkxk + ■ • - can be reduced to the form (*)
by the substitution x = el° (real <x> correspond to purely imaginary 2). The
usual methods of investigating the limit cycles of the system (*) lead to the
consideration of the first integral p = ei(e,k) of the unperturbed system; in
the notation of the original system, one obtains p = xk. An equation of
first approximation for the invariant manifold corresponding to resonance
can be obtained from the relation
(k, 2(e))
~ (k, c(e))‘
Our series are divergent in general, and the inference needs a verification.
For n — 2, our arguments can be confirmed by a rigorous proof of the
existence of a complex limit cycle, which has the indicated asymptotics near
resonance [A. S. Pjartli, Birth of complex invariant manifolds near a singular
point of a vector field depending on parameters, Funct. Anal. Appl. 6, 4 (1972),
95-96; for details of the proofs, see: Cycles of a system of two complex
differential equations in a singular point neighborhood, Trudy Mosc. Ob. 37
(1978), 95-106].
At the moment of resonance, when (k, 2) = 0, a cycle (a complex non-
simply connected phase curve) approaches the complex seperatrices of the
singular point. The noncontractible path which exists on this cycle disap¬
pears at resonance, merging with an equilibrium position. A particular case
is the birth (or destruction) of a cycle from an equilibrium position with
loss of stability (cf., § 33). In this case, k = (1, 1, 0, • • •), kt + 22 = 0.
The phenomenon can be observed in the set of real numbers (cf.. Fig. 129).
In other cases (even at the same resonance; for example, in the case of a
saddle), the topology of real phase curves can persist at resonance.
The difference between the topologies of complex phase curves of an
equation (or family) and of its normal form is an obstruction to the analytic
reduction to the normal form. Moreover, if this difference is determined
by a jet of finite order (as is usually the case), then it prevents not only
the analytic but also a finitely smooth reduction to normal form. For
example, in the case where the ratio of the eigenvalues can be approximated
§ 36. Metamorphoses of the Topology at Resonances 329
The study of the question of divergence of Poincare series is far from complete.
The divergence proofs preceding Pjartli’s work (Poincare, Siegel, and Brjuno) are based
on a calculation of the growth of coefficients and do not illuminate the causes of
divergence in the same sense as the calculation of the coefficients of the series of arctan z
proves divergence for|z| > 1 but does not show the reason, i.e., singularities at z = ± i.
The work of Pjartli is based on a method of E. Hopf. Other proofs and generaliza¬
tions of the first two results of Pjartli have been proposed by Brjuno [A. D. Brjuno,
Normal form of differential equations with a small parameter, Matem. Zametki 16, 3
(1974), 407—414; Analytic invariant manifolds, Dokl. Akad. Nauk USSR 216, 2 (1974),
253-256; Integral analytic sets, Dokl. Akad. Nauk USSR 220, 6 (1975), 1255-1258],
where the triangle with vertices Xx, X2, X3 contains 0. This resonance also has
been studied by Pjartli and Brjuno. Here a branching from the separatrices
of a singular point of an invariant manifold has been proved for n = 3.
Let (z1, z2, z3) be the phase coordinates and let e be the parameter of the
deformation (the resonance corresponds to e = 0). Then in the space with
coordinates (z, e), the invariant manifolds fill a holomophic hypersurface
whose equation has the form zkx'z22z3 = e in first approximation in an
appropriate coordinate system.
330 6. Local Bifurcation Theory
Pjartli has proved that in the generic case for “abnormally commensur¬
able” (Al A ,A3) forming a Siegel triangle in any neighborhood of the
3 2
series in this neighborhood. It follows from Pjartli’s result cited above that
in the generic case for “abnormally commensurable” (Aj, A2, A3), the indi¬
cated series diverge in any neighborhood of the origin.
Remark. To every linear vector field in C”, there corresponds a linear trans¬
formation in C (the “Poincare map”). Namely, let the field define the
"-1
differential equation
ls = e2"ia'/a- (j = , 1 ). 1
k1rxl + k2a 2 + k3 a3 = 0
X™'X™* = 1.
H. Local Shifts
2nZ. + co(e)Z’
where 2n corresponds to the distinguished circle. The function co(a) has the
limit co0 for a —> 0. From the formulas of § 361, it follows that Xt = e,co°m2,
X2 = e~iw°m‘.
The index of self-intersection of the curve T with the surface Z is equal
to zero. Therefore, Z is topologically the direct product of T with a disk.
Nevertheless, Z is not necessarily a direct product analytically.
334 6. Local Bifurcation Theory
Moreover:
Let (zL, z2) be the eigencoordinates. The differential forms dzjzk on the
cylinder are holomorphic and invariant with respect to the mapping and
determine holomorphic forms on the elliptic curve.
We calculate the integrals of these forms on generators of the group of
homologies of the torus. One of the generators corresponds to the directrix
circle y on the cylinder. For this, we have
— 2nim,,
co _ In A i _ In X2 ^
2n 2nim2 2nim1
where a, /?, a, and b are constants, ir = rne'k<p, and A and B are power series in £ and
h' beginning with terms of degree 2. This substitution transforms w into w(l + ye +
ch- + C), where y = na + ikfl, c = na + ikfi; C can be written in terms of second and
higher degree in e and w.
The equation ye 4- cw + C = 0 determines a branching curve. For a generic family,
we have y ^ 0, c ^ 0. After an appropriate change of the coordinates e and r, this
equation has the form = e.
The convergence can be studied in the same way as in the works of Pjartli and
Brjuno cited previously.
Remark. It is easy to verify that the condition }!'e'k<0 = 1 means exactly that the normal
bundle is analytically trivial over some finitely sheeted cyclic covering of the elliptic
curve.
All fibrations considered so far are topologically trivial. In particular, the elliptic
curve branching off at resonance is projected (nonholomorphically) onto the curve
T(e) “along the r-direction”. This projection is a topological finitely sheeted cyclic
covering of the torus. This is the same covering over which the normal bundle becomes
trivial at the moment of passage through resonance.
If (n, k) = d > 1 (but there are no resonances of smaller orders, ^ 1 for
0 < m < n), the branching curve is not connected. In this case, it consists of d com¬
ponents, each of which is an («/c/)-sheeted topological covering of the original torus.
For some nonresonant bundles (i.e., pairs X, co), the series reducing the pasting to
normal form are divergent.
§ 37. Classification of Singular Points 337
The branching off of the curves in the case of resonance allows us to “explain” the
divergence of the series linearizing the pasting. Let us assume that a pair (A, to) is
nonresonant but very close to a resonance. Then, in a small neighborhood of the
initial elliptic curve, another elliptic curve will generically exist, namely the curve
materializing the resonance. If the pair (A, oj) is sufficiently close to an infinite number
of resonances, then in an arbitrarily small neighborhood of the initial elliptic curve, an
infinite number of curves exist, materializing distinct resonances and cyclicly covering
the initial curve.
The normal bundle of the initial curve is nonresonant. Nonresonant normal bundles
of degree 0 do not have sections over any cyclic finitely sheeted covering of an elliptic
curve. Therefore, in the normal bundle of the initial elliptic curve, there are no elliptic
curves cyclicly covering the initial curve. This means that no neighborhood of the
initial curve on the surface can be mapped biholomorphically on a neighborhood of
the zeroth section of a normal bundle. Therefore, the series diverge for generic pastings
if the pair (A, co) can be approximated too well by resonant pairs. This account follows
the author's article in Funct. Anal. Appl. 10 (1976), 1-12. Details of the proofs can be
found in a sequence of papers by Ju. S. Il’jasenko and A. S. Pjartly Zero-type neighbor¬
hoods of imbedded complex tori, Trudy Sem. I. G. Petrovskogo 5 (1979), 85-95 ; 7 (1981),
3-49; 8 (1982), 111 -127. Ju. S. Il’jasenko, Embeddings of elliptic curves of positive type
into complex surfaces, Trudy Moskov. Mat. Obschestva 45 (1982), 37-67.
Jk = I U II U III;
The jets of types I or II are said to be sufficient and those of type III are
said to be susceptible.
In our case the sets I, II, and III have the following two properties.
1. Semialgebraic Property. Each of the sets I, II, and III is a semialgabraic
submanifold in the jet space Jk.
A semialgebraic set in (RN is defined as a finite union of subsets, each of
which is given by a finite system of polynomial equations and inequalities.
If inequalities are not needed, then the set is said to be algebraic. A useful
property of semialgebraic sets is expressed by the following theorem. [For
the proof, cfi, A. Seidenberg, A new decision method for elementary algebra,
Ann. Math., Ser. 2 60 (1954), 356-374; E. A. Gorin, On asymptotic prop¬
erties of polynomials and algebraic functions, Russ. Math. Surveys 16, 1
(1961), 91-118.]
We note that the projection of an algebraic set may not be algebraic but
only semialgebraic (for example, the projection of a sphere onto a plane).
2. Almost Finite Determinacy. As k —> oo, the codimension of the set
III Jk of susceptible jets tends to infinity.
In other words, the susceptible jets in Jk are determined by a number of
conditions increasing with A:. As a result, it turns out that the set of functions
for which it is undecidable whether 0 is a point of local minimum from any
number of terms in the Taylor series is very thin: it has infinite codimension
in the function space.
B. Other Examples
The analogous problem for functions of several variables does not admit
such a simple algorithm: if the second differential degenerates, then we have
to appeal to higher derivatives and we arrive at the problem of the classifica¬
tion of algebraic curves, surfaces, etc. Nevertheless, in this case as well, the
decomposition Jk = I U II U III of the space of A:-jets of functions on R"
are semialgebraic and almost finitely determined, although it is hopeless to
explicitly write out the equations and inequalities on the Taylor coefficients
for arbitrarily large n and k. The existence of these equations and inequalities
can be derived from the Tarsky-Seidenberg theorem, the proof of which
also contains an algorithm for obtaining these equations and inequalities
(generalized Sturm’s theory).
The initial segment of the classification has been calculated explicitly and turns out
to be connected (quite mysteriously) with the classification of regular polyhedra, the
Coxeter, Weyl, and Lie groups of the series Ak, Dk, and Ek, automorphic functions,
triangles in the Lobachevsky plane, singularities of caustics, wave fronts, and oscillating
integrals of the method of stationary phase [cf., V. I. Arnold, Singular points of smooth
functions and their normal forms, Russ. Math. Surveys 30, (1975), 1-75 and references
therein; Vasil’ev, V. A., The asymptotics of exponential integrals, Newton diagram, and
classification of points of minimum, Funct. Anal. Appl. 11, 3 (1977), 1-11],
to the number of eigenvalues in the left half-plane. The space of 1-jets can
be divided into a finite number of parts corresponding to the number of
roots in the left half-plane. Each of these parts is a semialgebraic set in the
space of jets; the polynomial inequalities defining it can even be given
explicitly (Routh-Hurwitz condition; cf., for example, F. R. Gantmaher,
Theory of Matrices, Moscow, Nauka, 1967).
The susceptible 1-jets form a semialgebraic submanifold of codimension
1, which separates the domains corresponding to distinct numbers of roots
in the left half-plane. In the preceding sections, we considered a series of
examples of investigations of what happens in these degenerate cases in the
passage to 2-jets, etc. Thus, this gives the impression that here, too, we may
go arbitrarily far, and only the complexity of calculations and the abundance
of cases do not permit us to give an algebraic classification in the cases of an
arbitrarily large codimension. It turns out that this is not so [V. I. Arnold,
Algebraic unsolvability of the problem of Lyapunov stability and the problem
of topological classification of singular points of an analytic system of differ¬
ential equations, Funct. Anal. Appl. 4, 3 (1970), 1-9].
The semialgebraic property is lost even in such a simple case as in the
problem of distinguishing between a center and a focus for vanishing roots
of the characteristic equation (Brjuno and Il’jasenko, cf., Ju. S. Il’jasenko,
Algebraic unsolvability and almost algebraic solvability of the center-Jocus
problem, Funct. Anal. Appl. 6, 3 (1972), 30-37]. Consequently, for the
problems of stability and topological classification no algebraic algorithm
can exist*).
There remains some hope for the existence of a nonalgebraic algorithm,
nevertheless, i.e., that the property of almost finite determinacy holds: the
set of germs whose topological type (or stability) is not determined by any
finite segment of the Taylor series may have infinite codimension. The
question of whether this is so presents serious difficulties; its formulation
has to be made precise, by indicating the exact sense of the word codimension.
The sets in the space of A:-jets whose codimension has to be defined are not
algebraic and set-theoretical difficulties may arise. Thom conjectured that the
answer to this question is negative. See F. Takens, A nonstabilisable jet of a
singularity of a vector field, Dynamical Systems (ed. M. M. Peixoto), Academic
Press, New York (1973), 583-597.
We also mention the problem of algorithmical decidability of the stability
of a stationary point for a vector field with polynomial components over
the ring of integers.
* Most recently, L. Hazin and E. Snol' established the algebraic unsolvability of the problem
of stability in the case of two pairs of purely imaginary eigenvalues with resonance 3:1. This
case corresponds to a submanifold of codimension 3 in the function space. See L. G. Hazin,
E. E. Snol’, Simplest cases of algebraic unsolvability in the problems of asymptotic stability,
Dokl. Akad Nauk SSR 240 (1978), 1309-1 311; Stability conditions at resonance I ; 3, Prikl.
Mat. Mekh. 44 (1980), 229-244. See, also, P. M. Elisarov, Nonalgebraicity of some manifolds
of differential equations, Vestn. Mosk. Univ. Mat. Mekh. 2, (1978), 57-64.
§ 37. Classification of Singular Points 341
Variant 1
I. Let £ = 0.
II. Let Eq. (2) be the linearized equation along the solution indicated
in problem (5).
Variant 2
I. Let a vector field in three-dimensions have the origin as its singular
point and let one of the eigenvalues of this singular point be equal to zero
and the other two purely imaginary.
(1) Reduce to normal form the terms of degree 1 of the Taylor series
expansion at zero of the components of the field [1],
(2) Do the same for terms of degree 2 [3].
(3) Do the same for terms of any degree [8].
(4) Average the system with respect to the fast rotation given by the
linear part of the field [12].
(5) Reduce a beginning segment of the Taylor series at zero of the fields
of the family to the simplest possible form by a diffeomorphism
depending smoothly on parameters varying in the neighborhood of
zero [10].
(6) Average the same system with respect to the fast rotation given by
the linear part of the initial field [20],
IV. Let a straight line going through zero be distinguished in the plane.
A diffeomorphism of the plane is said to be distinguished if it transforms the
distinguished line into itself. A vector field is said to be distinguished if it is
tangent to the distinguished line at all of its points. Let a distinguished field
be given which has a singular point at zero with two vanishing eigenvalues.
(12) Reduce a segment of the Taylor series of the field at zero to the sim¬
plest possible form by means of a distinguished diffeomorphism [12],
344 Samples of Examination Problems
Additional Problems
(1) Let z = sz + Tz|z|2 + z3. Prove that the number of limit cycles is
not greater than 1 if |Re/l| > 1. Hint: Divide the field by zz and
use the formula div P(z, z) = 2 Re(dP/dz).
(2) Let A = (3 + i)/y/2. Then, for argg = 5n/4, the singular separatrix
of every saddle-node coincides with the nonsingular separatrix of the
next saddle-node. Hint: At the moment of saddle-node coalescence,
the equation can be reduced to the form w = el6 [_/?w(| w|2 — 1) +
i(w2 — w2)], where A = (R — i)e,e. If R = 2 and 9 = n/4, the
separatrices are straight lines.
(3) Analyze the curves in the complex A -plane which divide domains
where, when arg e varies, singular points coalesce on the cycle, inside
the cycle, and outside the cycle. Hint: If one varies 6, the field vectors
rotate. The curves are located approximately like the four parabolas
a2 = 2 ( + b ± 1), A = a -I- ib.
(4) For small | Re A | and 1 < | Im ^41 < c « 4.11, the equation of problem
(1) has (for appropriate e) two limit cycles, with nine singular point
inside the inner cycle. For | Im A | > c, only one cycle exists (Neistadt).
The boundary between domains of existence of one or two cycles
looks like an ellipse with the major axis 1 ^ |Im/l| <4.11 and the
minor axis of length 2.
(5) Prove that there are no limit cycles in the generalized system of
Lotka-Volterra, x = x (a + ax + by), y = y (/? + cx + dy). Hint:
At stability loss, the system has a first integral: a product of degrees
of three linear functions (Bautin).
Index
258. Smoller: Shock Waves and 286. Andrianov: Quadratic Forms and Hecke
Reaction-Diffusion Equations, 2nd Operators
Edition 287. Maskit: Kleinian Groups
259. Duren: Univalent Functions 288. Jacod/Shiryaev: Limit Theorems for
260. Freidlin/Wentzell: Random Perturbations Stochastic Processes
of Dynamical Systems 289. Manin: Gauge Field Theory and Complex
261. Bosch/Guntzer/Remmert: Non Geometry
Archimedian Analysis—A System 290. Conway/Sloane: Sphere Packings,
Approach to Rigid Analytic Geometry Lattices and Groups
262. Doob: Classical Potential Theory and Its 291. Hahn/O’Meara: The Classical Groups and
Probabilistic Counterpart K-Theory
263. Krasnosel’skil/Zabrelko: Geometrical 292. Kashiwara/Schapira: Sheaves on
Methods of Nonlinear Analysis Manifolds
264. Aubin/Cellina: Differential Inclusions 293. Revuz/Yor: Continuous Martingales and
265. Grauert/Remmert: Coherent Analytic Brownian Motion
Sheaves 294. Knus: Quadratic and Hermitian Forms
266. de Rham: Differentiable Manifolds over Rings
267. Arbarello/Comalba/Griffiths/Harris: 295. Dierkes/Hildebrandt/Kuster/Wohlrab:
Geometry of Algebraic Curves, Vol. I Minimal Surfaces I
268. Arbarello/Comalba/Griffiths/Harris: 296. Dierkes/Hildebrandt/Kiister/Wohlrab:
Geometry of Algebraic Curves, Vol. II Minimal Surfaces II
269. Schapira: Microdifferential Systems in 297. Pastur/Figotin: Spectra of Random and
the Complex Domain Almost-Periodic Operators
270. Scharlau: Quadratic and Hermitian Forms 298. Berline/Getzler/Vergne: Heat Kernels and
271. Ellis: Entropy, Large Deviations, and Dirac Operators
Statistical Mechanics 299. Pommerenke: Boundary Behaviour of
272. Elliott: Arithmetic Functions and Integer Conformal Maps
Products 300. Orlik/Terao: Arrangements of
273. Nikol’skil: Treatise on the Shift Operator Hyperplanes
274. Hormander: The Analysis of Linear 301. Loday: Cyclic Homology
Partial Differential Operators III 302. Lange/Birkenhake: Complex Abelian
275. Hormander: The Analysis of Linear Varieties
Partial Differential Operators IV 303. DeVore/Lorentz: Constructive
276. Liggett: Interacting Particle Systems Approximation
277. Fulton/Lang: Riemann-Roch Algebra 304. Lorentz/v. Golitschek/Makovoz:
278. Barr/Wells: Toposes, Triples and Constructive Approximation. Advanced
Theories Problems
279. Bishop/Bridges: Constructive Analysis 305. Hiriart-Urruty/Lemarechal: Convex
280. Neukirch: Class Field Theory Analysis and Minimization Algorithms I.
281. Chandrasekharan: Elliptic Functions Fundamentals
282. Lelong/Gruman: Entire Functions of 306. Hiriart-Urruty/Lemarechal: Convex
Several Complex Variables Analysis and Minimization Algorithms II.
283. Kodaira: Complex Manifolds and Advanced Theory and Bundle Methods
Deformation of Complex Structures 307. Schwarz: Quantum Field Theory and
284. Finn: Equilibrium Capillary Surfaces Topology
285. Burago/Zalgaller: Geometric Inequalities
9 783540'966494
http://www.springer, de