Tentative Title: Nonlinear Waves and Integrability I: Derivation of Basic Integrable and Non-Integrable Equations
Tentative Title: Nonlinear Waves and Integrability I: Derivation of Basic Integrable and Non-Integrable Equations
Contents
Part 1.
Hamiltonian Formalism
3
3
5
6
9
10
11
12
14
16
17
19
20
22
23
23
25
25
Bibliography
27
Index
29
iii
25
25
Part 1
Hamiltonian Formalism
CHAPTER 1
j = 1, . . . , N
are the system of the even number 2N (N is the positive integer) of ordinary
differential equations (ODE). That system is fully defined by the Hamiltonian H
which is the given smooth function H = H(p1 , . . . , pN , q1 , . . . qN , t) of independent
variables p := (p1 . . . , pN ), q := (q1 . . . , qN ) and time t. Here all components of p,
q and as well as H and t are real numbers R. The variables p and q are usually
called by the generalized momenta and coordinates, respectively. In most cases
below H does not depends on t explicitly (unless we explicitly specify the opposite)
and then the Hamiltonian is the constant of motion as follows from the dynamic
PN
H
H
equations (1.1): dH
j=1 p j pj + qj qj + H = 0, where here and below we use
dt =
the notation f := f
t for any function f (t). The set of all possible values of p and
q is called by a phase space P . The canonical Hamilton equations together with
the independent variables q and p and the phase space P form the equations of the
Hamiltonian mechanics.
For the following particular form of the Hamiltonian
(1.2)
H=
N
X
pi 2
+ U (q)
2mi
i=1
we recover from (1.1) the Newton equations for the motion of the particles with
the coordinates q, the momenta p and the masses mi in the potential U (q):
pi = mi qi , i = 1, . . . N, p = U
q , where here and below we use the notation
q :=
q1 , . . . , qN . For N/D particles in Ddimensional space with coordinates r1 , r2 , . . . , rN/D RD we set q = (r1 , r2 , . . . , rN/D ) RN (masses mi are
split in that case in N/D groups). Each component of q and p can take the arbitrary real values so the phase space is RN RN = R2N , the vector space of 2N real
N
P
pi 2
numbers. In the equation (1.2), the terms
2mi and U (q) are called by the kinetic
i=1
U (q) =
N
X
mj j2
j=1
qj 2 ,
pj = mj j aj cos (j t + j ),
qj = aj sin (j t + j ),
j = 1, . . . , N,
H=
p2
mgL cos ,
2mL2
.
(1.7)
qj pj
pj qj
j=1
The Hamiltonian equations (1.1) can be equivalently expressed through the
Poisson brackets between p, q and the Hamiltonian H as follows
dq
= {q, H},
dt
dp
= {p, H}.
dt
(1.8)
(1.9)
and satisfy a Jacobi identity
(1.10)
for any infinitely differentiable functions F (p, q), G(p, q) and H(p, q).
Time-derivative of any function F (p, q) with p(t) and q(t) being the solution
of (1.1) is conveniently represented through the Poisson bracket as follows
(1.11)
F
F
H F
H F
dF
= p
+ q
=
= {F, H}.
dt
p
q
p p
q q
{F, H} = 0.
H = H(I)
(1.15)
Thus the solution of Liouville integrable system is quasiperiodic with the rotation
of each angular coordinate j at the angular frequency j , j = 1, . . . , N .
For N = 1 any Hamiltonian system (1.1) is Lioville integrable because we use
the Hamiltonian as the integral of motion H = F1 . E.g., for the harmonic oscillator
2
p2
2
+ m
I = H, where H = 2m
2 q .
1.3. Canonical transformations and generating functions
The differential equations (1.13) in action-angle variables (I, ) is the particular
case of the Hamiltonian equations (1.1) with I being the generalized momentum
and being the generalized coordinate. Thus the change of variables from the
original variables (p, q) into action-angle variables save the general form (1.1) of
the Hamiltonian equations. In this Section we study the general transformation
of the Hamiltonian equations from the original variables (p, q) into new variables
variables (P, Q) such that
dQj
H
=
,
dt
Pj
(1.16)
H
dPj
=
, j = 1...N
dt
Qj
with P RN and Q RN being the new generalized momentum and generalized
S = [p(t) q(t)
(1.17)
H (p(t), q(t), t)] dt,
t1
between two fixed times t1 and t2 . The functional S here is understood as the line
integral along a curve in 2N + 1-dimensional real vector space (p, q, t) R2N +1 .
The principle of the least action states that the motion of the system between the
specified points q1 := q(t1 ) and q2 := q(t2 ) is determined by the the stationary
value (extremal) of the action S. Assume that the functions p(t) and q(t) realize
such stationary value. Then it means that the replacement of p(t) and q(t) in
(1.32) by their independent infinitesimal variations in the form p(t) + p(t) and
q(t) + q(t) with fixed generalized coordinates at initial and finite times,
(1.18)
q(t1 ) = q(t2 ) = 0
do not change the value of S in the linear order of p(t) and q(t). It is assumed
here that p(t), q(t), p(t) and q(t) are smooth functions at t1 t t2 to ensure
the differentiability of S. One can assume for example, that these functions are
infinitely differentiable. The variation of S is given by
Zt2
H (p + p, q + q, t)] dt
[(p + p) (q + q)
S =
t1
Zt2
Zt2
[p q H (q, p, t)] dt =
t1
(1.19)
p q + p q
H
H
q
p dt
q
p
t1
+ h.o.t.,
where h.o.t. means the higher order terms (quadratic and above) in q and p.
Below unless otherwise specified, we neglect h.o.t. in S, i.e. by S we assume the
first variation (only linear in q and p contributions). Also the scalar products
PN H
H
j=1 qj qj and similar expressions are implied here and below. The
q q =
assumption of the stationary value of S on p(t) and q(t) implies that S = 0. Then
integrating (1.34) by part for the term with q = dq
dt we obtain that
(1.20)
S = p
t=t
q|t=t21
Zt2
+
t1
H
q
p
Zt2
p dt +
H
q
q dt = 0.
t1
Taking into account the condition (1.33) we conclude that the integral in (1.35)
vanishes for any independent variations p and q. It means that the expressions
in brackets at each integrand in (1.35) also vanish resulting in the Hamiltonian
equations (1.1). Thus the Hamiltonian mechanics mechanics can be considered as
the reformulation of the classical mechanics using the principle of least action.
We note that the above derivation of the Hamiltonian equations (1.1) from the
variation of S needs that the motion of the mechanical system to be the stationary
value of S but does not actually require S to be the minimum. In most cases
however, for short pieces of the mechanical system trajectories, S is the minimum.
It justifies the historical use of the principle of least action terminology. Although
more accurately it should be called by the principle of stationary action. Later
in the book we additionally discuss the variational principle by introducing the
Lagrangian in Section???
Now we return to the transformation between variables (p, q) which satisfy
(1.1) and (P, Q) which satisfy (1.16). Because of (1.16), new variables must also
S =
Zt2 h
i
H
(P, Q, t) dt,
PQ
t1
H
+ F
p q H = P Q
2
because the difference in S and S is reduced to the constant term F |t=t
t=t1 which
vanishes under variation.
We consider the transformation
(1.23)
P = P(p, q, t)
and Q = Q(p, q, t)
from the old variables (p, q) to the new variables (P, Q), were p, q, P and Q RN .
The transformation (1.23) is called the canonical transformation if both (1.1) and
(1.16) are simultaneously satisfied and there exists a function F such that (1.22) is
valid. We define F in one of four possible forms as follows:
(1.24a)
F = F1 (q, Q, t),
(1.24b)
F = F2 (P, q, t) P Q,
(1.24c)
F = F3 (p, Q, t) + p q,
(1.24d)
F = F4 (p, P, t) + p q P Q.
H
H
+ F1 = P Q
+
p q H = P Q
F1
F1
F1
q +
Q+
.
q
Q
t
We assume the the old coordinate q and the new coordinate Q are separately independent. Then the equation (1.25) is identically valid if the expressions multiplying
are both identically zero as well as the sum of the remaining terms
both q and Q
is zero which give the expressions for the old and new generalized momenta as well
as the relation between the old and new Hamiltonians as follows
F1
F1
= H + F1 .
p=
, P=
, H
(1.26)
q
Q
t
We can define the transformation (p, q) (P, Q) from (1.26) as follows. The
first equation in (1.26) defines p as the function of q, Q and t. We assume that
this function can be inverted to give Q as the function of p, q and t, thus giving
the second part of the transformation (1.23). Then Q(p, q, t) can be substituted
into the second equation in (1.26) resulting in P(p, q, t), i.e. the first part of the
transformation (1.23). After that using the inverse of the transformation (1.23) in
in terms of
the third equation in (1.26) we immediately the new Hamiltonian H
the the new canonical variables (P, Q).
In a similar way, taking F from (1.24b),(1.24c) and (1.24d), respectively, and
substituting the total time derivative of F of into (1.22) result in
(1.27a)
(1.27b)
(1.27c)
F2
F2
= H + F2 ,
, Q=
, H
q
P
t
F3
F3
= H + F3 ,
q=
, P=
, H
p
Q
t
F4
F
F4
4
=H+
, Q=
, H
.
q=
p
P
t
p=
The last expression in each of the equations (1.26), (1.27a), (1.27b) and (1.27c)
in terms of the new canonical variables
give the formula for the new Hamiltonian H
(P, Q). We also note from these equations that if the generating function F does
In that case to obtain the new Hamiltonian
not depend on t explicitly then H = H.
where p(1) = (p1 , . . . pN1 ), Q(1) = (Q1 , . . . QN1 ), P(2) = (PN1 +1 , . . . PN ) and q(2) =
(qN1 +1 , . . . pN ), 1 N1 N 1, i.e. it is of the type 3 for first N1 components of
the canonical momenta and coordinates and it is of the type 2 for the other N N1
components with F in the equation (1.22) given by
(1.29)
The distinction between canonical coordinates and canonical momenta is essentially lost under the canonical transformation. This can be seen if we perform
a simple canonical transformation P = q and Q = p which interchanges the
canonical coordinates and canonical momenta. That canonical transformation is
of the type 1 with F = F1 = q Q. It suggest to call canonical coordinates and
canonical momenta as canonically conjugated quantities.
1.4. Canonical transformation through Poisson bracket
The Poisson bracket is invariant with respect to the canonical transformation,
i.e. the Poisson bracket is canonical invariant. To show that we define the Poisson
bracket with respect to the canonical variables Q and P as follows
N
X
F G
F G
{F, G}Q,P =
.
(1.30)
Qj Pj
Qj Qj
j=1
Then (1.7) defines {F, G}q,p . Substitution of (1.23) results in
Relation through the Poisson brackets.
?
10
=0
dt q
q
which is the system of N second order (in time) ODEs. That system is fully defined
t) of independent
by the Lagrangian L which is the given function L = L(q, q,
variables t, q RN and q RN (q is used as the second independent variable
in contrast to the Hamiltonian system (1.1), where p is the second independent
variable). The Lagrange equations together with the independent variables q and
q form the Lagrangian mechanics.
The Lagrange equations follows from the principle of least action (or more
accurately as the principle of stationary action as well as it is also called by the
Hamiltons principle) for the action S defined as
Zt2
(1.32)
L (q(t), q(t),
t) dt
S=
t1
between two fixed times t1 and t2 . The functional S here is understood as the
line integral along a curve in N + 1-dimensional real vector space (q, t) RN +1 .
The principle of the least action states that the motion of the system between the
specified points q1 := q(t1 ) and q2 := q(t2 ) is determined by the the stationary
value (extremal) of the action S. Assume that the function q(t) realizes such stationary value. Then it means that the replacement q(t) in (1.32) by its infinitesimal
variation in the form q(t) + q(t) with
(1.33)
q(t1 ) = q(t2 ) = 0
does not change the value of S in the linear order of q(t). It is assumed here
that both q(t) and q(t) are smooth functions at t1 t t2 to ensure the differentiability of S. One can assume for example, that these functions are infinitely
differentiable. The variation of S is given by
Zt2
Zt2
t) dt
L (q + q, q + q,
S =
t1
Zt2
t) dt =
L (q, q,
t1
L
L
q +
q dt
q
q
t1
(1.34)
+h.o.t.,
where h.o.t. means the higher order terms (quadratic and above) in q and q.
Below unless otherwise specified, we neglect h.o.t. in S, i.e. by S we assume the
first variation (only linear in q and q contribution). Also the scalar products
PN L
L
j=1 qj qj and similar expressions are implied here and below. The asq q =
sumption of the stationary value of S on q(t) implies that S = 0. Then integrating
(1.34) by part for the term with q = dq
dt we obtain that
(1.35)
t=t
Zt2
L 2
d L
L
S =
q
+
q dt = 0.
q t=t1
q dt q
t1
11
Taking into account the condition (1.33) we conclude that the integral in (1.35)
vanishes for any q. It means that the integrand in (1.35) also vanishes resulting in
the Lagrange equations (1.31). Thus the Lagrangian mechanics can be considered
as the reformulation of the classical mechanics using the principle of least action.
We note that the above derivation of the Lagrange equations (1.31) needs that
the motion of the mechanical system to be the stationary value of S but does not
actually require S to be the minimum. In most cases however, for short pieces of
the mechanical system trajectories, S is the minimum. It justifies the historical use
of the principle of least action terminology. Although more accurately it should be
called by the principle of stationary action.
1.6. Equivalence of the Lagrangian and Hamiltonian mechanics
We introduce the generalized momentum p in the Lagrangian mechanics as
follows
L
(1.36)
.
p=
q
Then the Lagrange equations (1.31) can be rewritten as the system of two sets of
equations
L
,
q
L
p=
.
q
p =
(1.37)
We define the relation between the Hamiltonian H(p, q, t) and the Lagrangian
t) describing the same mechanical system as follows
L(q, q,
(1.38)
t).
H(p, q, t) = pq L(q, q,
The equation (1.38) together with (1.38) define the Legendre transform from the
independent variables p, q and t to the independent variables q, q and t. The exact
differential dH of H is given by
(1.39)
dH =
H
H
H
dp +
dq +
dt
p
q
t
while the exact differential of the right-hand side (rhs) of (1.38) is given by
L
L
L
dq
dq
dt
q
q
t
L
(1.40)
dt.
= q dp pdq
t
Here we used (1.38). Because rhs of (1.39) equals to rhs of (1.40), we immediately
recover the Hamilton equations (1.1) from the common terms of dp and dq. Thus
we started from the Lagrange equations (1.37) together with (1.38) and derived
the canonical Hamilton equations (1.1). It is straightforward to show that (1.1)
with (1.1) results in the Lagrange equations (1.37). It proves the equivalence of the
Lagrangian and the Hamiltonian mechanics. Also the terms with dt results in the
following relation
d (pq L) = p dq + q dp
(1.41)
L
H
=
.
t
t
12
In most cases below both H and L do not explicitly depend on t meaning that
(1.41) is zero. Such systems are called autonomous.
For the motion of particles in the potential (1.2) we obtain from the Legendre
transform (1.38) expressing p through q that
(1.42)
L=
N
X
mi qi 2
i=1
U (q).
Notice that the potential energy U enters with the opposite sign in comparison with
the Hamiltonian (1.2).
1.7. Holonomic constraints and the Lagrangian mechanics on manifolds
The analysis of numerous mechanical systems can be greatly simplified if we
consider ideal objects like rigid body. The planar pendulum considered in Section
1.1 is the example of such simplification with the ideal rigid massless rod attached
to the point mass. That system can be considered as the limit of the Newton
equations of the potential motion where the mass of the rod approaches zero while
its stiffness (represented by the potentials for its stretching, bending and twisting)
goes to infinity. That limit can be taken rigorously (see e.g. [Arn89]). Such type of
constrains are called by holonomic constraints. These constraints can also depend
on time t.
Consider the general system of Np point masses in the D-dimensional real vector
space RD (usually D = 1, 2, 3) as in equation (1.2). The positions of these masses
are given by vectors r1 , r2 , . . . , rK RD , i.e. by a point in the real vector space
RNp D . Applying Nh independent holonomic constraints we restrict the allowed
values of these coordinates to Np D Nh -dimensional surface M in RNp D . From
point of view of equation (1.2) it means that we take the limit of the potential U
to be infinitely large if the holonomic constraints are not satisfied. Assume that
N = Np DNh and q := (q1 . . . , qN ) is the set of generalized coordinates on M . The
dynamics of Np point masses with Nh constraints is equivalent to the Lagrangian
mechanics for the generalized coordinates q on the surface M [Arn89].
A planar motion of a point mass (Np = 1 and D = 2) without constraint is
characterized by two coordinates (e.g. vertical and horizontal coordinates). The
planar pendulum considered in Section 1.1 has one holonomic constraint because
the rigid rod allows for the mass to move on the circle S 1 only. The position of the
point mass on that circle is fully characterized by a single generalized coordinate
q = .
The surface M is called the configuration space of the Lagrange equations
(1.31). M is the N -dimensional differential manifold. It means that the small
neighborhood of each point of M is homeomorphic to the N -dimensional Eucledian
space EN as well as that M has a globally defined differential structure (see Appendix ?? for the detailed definition). The Eucledian space EN is the vector space RN
PN
together with the standard scalar product x y = i=1 xi yi for any x, y RN . The
dimension N of the configuration space is called the number of degrees of freedom.
Below we often consider the dynamical systems with constraints which do have
any simple mechanical analogies and do not allow easy derivation of the constraints
from the unconstrained systems. However the notion of constraints remains extremely helpful for many Hamiltonian systems.
equations of Section 1.31 are valid assuming that q(t) M and (q(t), q(t))
TM.
defining the Lagrangian mechanics on manifolds.
Assume that x ENe are the coordinates for the Newtonian equations of the
potential motion of point masses given by the equations (1.42) and (1.31). Consider
the coordinate transform
(1.43)
x = g(q)
Ne
P
i=1
mi xi 2
2
L=
N
X
Aij (q)qi qj
U (q),
2
i,j=1
where
(1.46)
Aij =
Ne
X
k=1
mk
gk (q) gk (q)
qi
qj
are the elements of the symmetric N N matrix A = AT (AT means the transposed
matrix of A). It follows from (1.46) and positivity of masses mi > 0, i = 1, 2, . . . , Ne
that A positive-definite matrix (i.e. T A > 0 for any nonzero column vector
RN ). We also abused a notation by setting U (q) = U (x).
14
We use the example (1.45) to define the kinetic energy of in the Lagrangian
mechanics as the general symmetric positive-definite quadratic form in q:
N
X
Aij (q)qi qj
T =
.
2
i,j=1
(1.47)
N
X
i = 1, 2, , . . . , N,
j=1
where other terms stands for the terms which depend of q and q only but not
on higher derivatives of q over time. The symmetry the positive-definiteness of A
which provides the
ensures its invertibility. It means that (1.48) is solvable for q
existence and the uniqueness of the solutions of the Cauchy problem for the natural
Lagrange equations.
More general non-natural Lagrange systems have a Lagrangian L which is a
general general function of q and q (and t for nonautonomous systems). Then it
is generally impossible to separate the Lagrangian into the kinetic and potential
energy. The existence and the uniqueness of the solutions of the Cauchy problem
2
L
to be nonsingular on M .
for such systems require that the matrix qi q
j
1.8. Oscillations
Consider the equilibrium point q0 M of the Lagrangian system (1.31), i.e.
q=q0 = 0. For the natural Lagrangian system (1.45),(1.31) it implies that q0 is
q|
the stationary point of U (q),
U (q)
(1.49)
= 0.
q
q=q0
L2 =
N
N
X
X
Aij qi qj
Bij qi qj
,
2
2
i,j=1
i,j=1
and
(1.52)
2 U (q)
Bij :=
qi qj q=0
1.8. OSCILLATIONS
15
(1.53)
Q = Cq,
(1.55)
as follows
which results in the expression for L2 in new variables Q and Q
N
(1.56)
L2 =
1 X 2
Qi i Q2i
2 i=1
and the Lagrangian equations (1.53) turn into the the following decoupled system
of ODEs:
Qi = i Qi .
(1.57)
det(B A) = 0.
Notice that the positive-definiteness of A ensures its the invertibility and allows
to rewrite (1.58) as the usual characteristic equation det(A1 B I) = 0 for the
matrix A1 B, where I is the identity matrix.
The quadratic Lagrangian (1.56) is in the form (1.42) which implies that the
Hamiltonian of that system is
N
(1.59)
H2 =
1X
Pi 2 + i Q2i ,
2 i=1
Pi = Q i , i = 1, . . . , N.
16
Exercise 1.
Show that the transformation from q and p to Q and P defined by (1.50),
(1.55), (1.56) and (1.60) is the canonical transformation for the Hamiltonian equations (1.1) with the quadratic Hamiltonian (1.56).
The Hamiltonian (1.4) for the harmonic oscillator in physical variables is transformed into (1.71) by the following trivial change of variables
Qj = m1/2 qj , Pj = m1/2 pj , j = 1, . . . , N,
(1.61)
N1
X
aj cos (j t + j )qj +
j=1
NX
1 +N2
cj,1 ej t + c2,j ej t qj
j=N1 +1
(1.62)
N
X
(cj,1 t + cj,2 ) qj ,
j=N1 +N2 +1
q=
N1
X
aj cos (j t + j )qj .
j=1
17
k=1
the mass the kth atom. The potential energy has the general form U =
N
P
i,j=1
Bij qi qj
2
,
2
2
where M = A is the diagonal matrix with the main diagonal (m
1, . . . , m
N ). The
Hamiltonian is given by
L=
(1.64)
(1.65)
H=
N
N
X
1 X
pj 2
+
Bij qi qj ,
2m
j
2 i,j=1
j=1
(1.66)
Assume additionally that all atoms interacting only pairwise with each other by
the harmonic potentials with the positive interaction constants ij = ji > 0. The
Hamiltonian of that system takes the following form
H=
(1.67)
N
P
N
N
X
pj 2
1 X
+
ij (qi qj )2 .
2
m
4
j
j=1
i,j=1
j=1, j6=i
immediate follows from (1.67) that the matrix B is non-negative (i.e. x Bx 0 for
any x RN ) and respectively all j are nonnegative. Zero values of j have a simple
physical meaning representing the translation of the molecule in any direction.
There are three independent directions for such translation (i.e. for the motion of
the molecules center of mass). The general solution (1.62) is then reduced to
(1.68)
q=
N
3
X
aj cos (j t + j )qj +
N
X
(cj,1 t + cj,2 ) qj .
j=N 2
j=1
18
where we choose the positive sign j > 0 for all j. Solving (1.69) for Qj and Pj we
obtain that
1
Qj =
(aj + a
j ) ,
(2j )1/2
(1.70)
1/2
j
Pj = i
(aj a
j ) .
2
The Hamiltonian (1.59) simplifies by (1.69) into
H2 =
(1.71)
N
X
j |aj |2 ,
j=1
aj
The equation for
real function:
(1.75)
daj
dt
19
(1.69) is the classical analog of the quantum mechanical transformation from the
coordinate-momentum representation to the representation by the Bose creation
and annihilation operators.
1.10. Complex variables in weakly nonlinear case
Complex variables (1.70) are especially useful if the Hamiltonian can be expanded in a power series of the canonical variables Pj and Qj . E.g. it occurs if we
expand the Hamiltonian in the power series near the stationary point (1.49) and
take into account next terms beyond quadratic terms considered in Section 1.8.
The linearity of the transformation (1.70) implies a power series in variables aj and
a
j for the same Hamiltonian expanded in the power series.
Consider a nonlinear oscillator with the Hamiltonian
p2
(1.77)
H=
+ U (q),
2m
such that the potential U (q) has a minimum at q = 0. Without loss of generality
we set U (0) = 0 because the Hamiltonian equations (1.1) do not change if shift
U (q) by an arbitrary real constant.
Assume that q is small and expand U (x) in a Taylor series about q = 0:
m 2 2
q + q 3 + q 4 + . . . ,
2
where we define in the same way as in (1.3) so it would have a meaning of
frequency of small oscillations.
The Hamiltonian has now a form of power series in canonical variables p and q
as follows
p2
m 2 2
H2 =
+
q ,
2m
2
3
H3 = q ,
(1.79)
(1.78)
U (q) =
H4 = q 4 ,
...,
where subscripts in H represent the order of terms in powers of q an p.
Limiting to H = H2 we recover the harmonic oscillator considered in Section
1.9 with the dynamical equation (1.9). Substituting (1.70) into (1.79) we obtain
that
(1.80)
H3 =
1
2m
3/2
(a3 + a
3 + 3a
a2 + 3
aa2 )
U 3
a +a
3 + V |a|2 a
+ |a|2 a ,
3
da
+ ia = iU a
2 iV a2 2iV a
a,
dt
where
(1.82)
U = V = 3
1
2m
3/2
.
20
Similar, at the next order H = H2 + H3 + H4 we obtain from (1.70) and (1.79) that
2
1
4
H4 = q =
(1.83)
[a4 + a
4 + 6|a|4 + 4|a|2 (
a2 + a2 )]
2m
and the corresponding dynamic equation takes the following form
(1.84)
2
da
1
+ ia = iU a
2 iV a2 2iV a
a i
(4
a3 + 12|a|2 a + 12|a|2 a
+ 4a3 ).
dt
2m
1.11. Nonlinear oscillator and the generation of multiple harmonics
To analyze the equation (1.84) we note that its leading order solution in the
limit of small a is reduced to the harmonic oscillator equation da
dt + ia = 0 with
the exact solution a = C1 eit . We call that solution by a first (or fundamental)
harmonic. We assume that nonlinear terms in r.h.s of (1.84) result in slow time
dependence of C1 at times 1/ which motivates looking for the solution of (1.84)
in the following form
(1.85)
where C1 (t), C0 (t), C2 (t), C2 (t) are the slow functions of time. We refer to
C0 (t) and C2 (t) as the amplitudes of the zeroth and second harmonics, respectively.
C2 (t) can by called by the minus second harmonic but instead we also refer it as
the second harmonic keeping in mind that if we have a harmonic with a frequency
j then nonlinearity immediately results in the formation of j harmonic. All
these harmonics are generated by the quadratic nonlinear terms in r.h.s of (1.84).
It is seen if we substitute a in r.h.s. (1.84) by the leading order harmonic solution
a = C1 (t)eit and neglect for now cubic terms which gives
(1.86)
da
+ ia = iU C12 e2it iV C1 2 e2it 2iV |C1 |2
dt
C0 =
2V
V
U
|C1 |2 ; C2 = C12 ; C2 = C 2 .
These second and zero harmonics are of second order, |C1 |2 in amplitude of the
fundamental harmonic. Retaining slow time dependence of C1 (t), C0 (t), C2 (t), C2 (t)
2
results in the addition of small terms d|Cdt1 | 1 |C1 |2 into r.h.s, of each expression in (1.87). In a similar way, extending (1.85) to include third, C3 , fourth,
C4 etc harmonics and substitution into (1.86) allows to conclude that at leading
order |C3 | |C1 |3 , |C4 | |C1 |4 etc. Thus the nonlinearity in (1.84) results in the
generation of the harmonics (1.84) at multiple frequencies 0, 2, 3, 4, . . ..
These higher harmonics also modify the dynamics of the first harmonic C1 . To
understand that modification we return to the equation (1.81), collect all terms
which eit corresponding to the fundamental harmonic and use (1.87) to express C0 (t), C2 (t), C2 (t) through C1 (t). That procedure can be qualitatively
interpreted as the projection of a into the state eit and gives
dC1
= 2iU C2 C1 i2V C1 C0 2iV C1 C0 2iV C2 C1
dt
20
(1.88)
= i V 2 |C1 |2 C1 ,
3
where we used that U = V .
Returning back from C1 to the full form of the fundamental harmonic a1
C1 (t)eit results in the equation
(1.89)
da1
+ i( +
)a1 = 0
dt
where
20 V 2
|a1 |2 .
3
The expressions (1.89) and (1.90) are however not sufficient because we also
need to find a similar contribution to from H4 . For that we take into account the
term with in (1.84) and again project that term on the state eit which results
2
1
12|C1 |2 C1 in
in the equation similar to (1.88) but with the added term i 2m
r.h.s. Thus instead of equations (1.89) and (1.90) we obtain a full expression
(1.90)
da1
+ i( + )a1 = 0
dt
which includes all terms of the order O(|a1 |2 a1 ). Here is called by the nonlinear
frequency shift and is given by
"
#
2
1
20 V 2
(1.92)
= 3
|a1 |2 .
m
3
(1.91)
d|a1 |2
dt
= 0. It allows to
a1 (t) = a1 (0)ei(+)t ,
which has the same form as the solution a1 (t) = a1 (0)eit for the harmonic oscillator (1.76) except that the frequency is replaced by + .
Equation (1.91) has the complex Hamiltonian form (1.74) with the Hamiltonian
T
|a1 |4 := H2 + H4,ef f ,
2
where H2 = |a1 |2 is the quadratic part of the Hamiltonian and
2
1
20 V 2
(1.95)
T := 3
m
3
(1.94)
H = |a1 |2 +
is the interaction constant for the effective 4th-order Hamiltonian H4,ef f . There
1 2
are two contributions to T . The first one is given by 3 m
and can be immediately obtained from the averaging of the original Hamiltonian H4 : hH4 i =
3
1 2
|a1 |4 , where h. . .i means averaging over fast oscillations at time scale 1/.
2 m
V2
The second contribution, 10
3 , comes from taking into account of the zero and
22
second harmonics in the cubic terms H3 (1.80) as explained above. That contribution corresponds to the second order of the perturbation theory (at the first order
we calculated C0 , C2 , C2 in (1.87) and at the second order we found the influence
of these terms on the dynamics of a1 ). In other words, the interaction constant
1 2
3 m
of hH4 i is renormalized by the cubic terms H3 of the Hamiltonian in the
second order of the perturbation theory. That renormalization has the negative
sign which is consistent with the well-known fact of the quantum mechanics that
the shift of the ground state level from the second order perturbation theory is
negative one [LL76].
We conclude that there are two effects of nonlinearity of the oscillator:
1. Nonlinearity results in the formation of multiple harmonics.
2. Nonlinearity produces the nonlinear frequency of the fundamental harmonic.
The above technique of obtaining (1.93) represents the averaging method (averaging
of over fast oscillations eit assuming weak nonlinearity results in the calculation
of the nonlinear frequency shift (1.92)).
Exercise 2.
Find the nonlinear frequency shift for the pendulum (1.6).
1.11.0.1. Resonance in nonlinear oscillator. Consider a dynamics/exitation of
the nonlinear oscillator with an external periodic forcing. In the linear approximation the equation (1.91) with the added forcing if0 eit takes the following
form
a1
+ i( i)a1 = if0 eit ,
t
where f0 C and R are constants (i.e. the forcing is monochromatic). Also
we added a small damping coefficient > 0 ( ) into l.h.s. of equation (1.96).
A solution of (1.96) is the sum of the transient solution eitt (corresponds to the solution of the homogeneous part with forcing excluded of (1.96))
and the particular solution of full nonhomogeneous equation of (1.96). Assuming
t , the transient solution can be neglected and the remaining solution takes
the following form
(1.96)
f0
eit ,
+ i
which gives a time-independent resonance solution in the form of Lorentzian function for |a1 |2 as follows
a1 =
|a1 |2 =
(1.97)
|f0 |2
.
( )2 + 2
(1.98)
(1.99)
|a1 |2 =
|f0 |2
,
( T |a1 |2 )2 + 2
:= ,
EXERCISE 4.
23
where is the frequency detuning between the linear oscillator frequency and the
pumping frequency .
Equation (1.99) represents the cubic equation for the unknown amplitude |a1 |2
which remains bounded from all values of , and f0 . To see that we assume by
contradiction that |a1 |2 for any fixed , and f0 which is inconsistent with
(1.99) (l.h.s while r.h.s would 0 in that case). Thus the nonlinear frequency
shift regularizes the linear resonance. At that resonance 0 and 0 which
implies from (1.99) that
2/3
|f0 |
2
2
.
|a1 | |a1 |sat :=
T
Also |a1 |2 is close to |a1 |2sat for fixed > 0 and 0 provided T |a1 |2 ) .
Taking the limit f0 0 for each fixed values of > 0 and we recover the
linear solution (1.97). In that limit there is only one real solution of (1.99) for
|a1 |2 . If we increase |f0 | then for large enough amplitude |f0 | > |f0cr |, there are
three real solutions of (1.99) for |a1 |2 , where |f0 | = |f0cr | is the critical value of the
forcing |f0 | corresponding to the appearance of two additional real roots. +
Exercise 3.
Find |f0cr |.
Answer: |f0cr |2 =
2 3
T .
Exercise 4.
Determine the stability (in time) of all three roots |f0 | > |f0cr |. Find |f0cr |.
1.11.0.2. Parametric resonance. Another type or resonance occurs if the oscillator frequency has a periodic time dependence. Simplest example of the periodic
time dependence is given by
2 (t) = 02 (1 + cos t),
where 1 is the small parameter. The dynamics of oscillator without damping
is given by the Mathieus differential equation
d2 q
= 2 (t)q,
dt2
where q is the coordinate. To bring that equation to the complex form we introduce
the momentum as p = q (for simplicity we set the mass m = 1) an define a and a
similar to (1.72) as
(1.100)
1
(0 q + ip) ,
(20 )1/2
1
a
=
(0 q ip) ,
(20 )1/2
a=
(1.101)
except that now time-independent part 0 of the frequency (t) is used for that
transformation.
Consider the following Hamiltonian
0
(1.102)
H = 0 |a|2 +
(a + a
)2 cos(t),
4
24
20 =
a
4
(1.106)
d
a
hHi
0 it
+
a=i
= i0 a
+i
ae ,
dt
a
4
where we added a linear damping term a into l.h.s., similar to Section (1.11.0.1).
Comparison between first equations in (1.103) and (1.106) shows that (1.106) does
not include terms which oscillate with a different frequency than ei0 t . Taking
into account such non-resonant terms in (1.103) can be done by the expansion
in multiple harmonics, qualitatively similar to the expansion (1.85). These terms
however would modify (1.106) by the inclusion of O(2 ) terms which we neglect here.
25
t
2
, a
= c(t)e+i
t
2
0
+ i 0
+ c+i
c = 0,
t
2
4
(1.107)
0
+ i 0
+ c i
c = 0.
t
2
4
Assuming c, c et we obtain from (1.107) a homogeneous system of linear equations
0
c = i
c,
+ + i 0
2
4
(1.108)
0
+ i 0
c = i
c
2
4
for the unknowns c and c. That system is solvable provided the matrix of its
coefficients has a zero determinant which gives
s
2
(0 )2
0
.
(1.109)
=
16
2
It follows from (1.109) that the instability ( > 0) is possible for > cr =
4/0 provided 0 is close enough to /2 to ensure that the expression under the
square root in (1.109) is positive as well as to overcome dissipation rate . This
instability is called the parametric instability. The growth rate of the parametric
instability reaches maximum if the condition (1.104) is exactly satisfied.
A difference between the forced oscillator of Section (1.11.0.1) and the parametric amplification of Section (1.11.0.2) can be seen from the everyday experience
in playing on a childrens swing. Rocking back and forth pumps the swing as a
forced harmonic oscillator. Once moving, the swing can also be parametrically amplified by alternately standing and squatting at key points in the swing arc, i.e. by
periodically changing the moment of inertia of the swing and hence the resonance
frequency. The rate of that change is twice of the natural frequency of the swing
satisfying the parametric resonance condition (1.104).
Exercise 5.
Find a frequency width of the parametric resonance , i.e. the range of values
of at which a parametric excitation occurs assuming > cr .
1.12. Canonical transformations
1.13. Generalization of the canonical Hamilton equations. Symplectic
structure. Poisson mechanics.
1.14. Symplectic leaves. Example of free motion of rigid body
Bibliography
[Arn89] V. I. Arnold, Mathematical methods of classical mechanics, Springer, 1989.
[LL76] L. D. Landau and E. M. Lifshitz, Quantum mechanics (course of theoretical physics,
volume 3), Butterworth-Heinemann, New York, 1976.
[LL89]
, Mechanics (course of theoretical physics, volume 1), Pergamon Press, New York,
1989.
27
Index
averaging, 22
frequency detuning, 23
interaction constant, 21
kinetic energy, 4
nonlinear frequency shift, 21
parametric instability, 25
potential energy, 4
smooth manifold, 4
29