Classical Field Theory
Classical Field Theory
Takhtajan
i
MAT 560 Mathematical Physics I.
Classical Field Theory
Leon A. Takhtajan
v
vi CONTENTS
Classical Mechanics
LECTURE 1
Equations of motion
We assume that the reader is familiar with the basic notions from the the-
ory of smooth — C ∞ — manifolds, and recall here the standard notation.
Unless it is stated explicitly otherwise, all maps are assumed to be smooth,
and all functions are assumed to be smooth and real-valued. Local coordinates
q = (q 1 , . . . , q n ) on a smooth n-dimensional manifold M at point q ∈ M are
Cartesian coordinates on ϕ(U ) ⊂ Rn , where (U, ϕ) is a coordinate chart on M
centered at q ∈ U . For f : U → Rn we denote (f ◦ ϕ−1 )(q 1 , . . . , q n ) by f (q),
and we let
∂f ∂f ∂f
= ,..., n
∂q ∂q 1 ∂q
n
M
A• (M ) = Ak (M )
k=0
the graded algebra of smooth differential forms on M with respect to the wedge
product, and by d the de Rham differential — a graded derivation of A• (M ) of
degree 1, such that df is a differential of a function f ∈ A0 (M ) = C ∞ (M ). Let
Vect(M ) be the Lie algebra of smooth vector fields on M with the Lie bracket
[ , ], given by a commutator of vector fields. For X ∈ Vect(M ) we denote
by LX and iX , respectively, the Lie derivative along X and the inner product
with X. The Lie derivative is a degree 0 derivation of A• (M ) which commutes
with d and satisfies LX (f ) = X(f ) for f ∈ A0 (M ), and the inner product is
a degree −1 derivation of A• (M ) satisfying iX (f ) = 0 and iX (df ) = X(f ) for
f ∈ A0 (M ). They satisfy Cartan formulas
(1.1) LX = iX ◦ d + d ◦ iX = (d + iX )2 ,
(1.2) i[X,Y ] = LX ◦ iY − iY ◦ LX .
3
4 1. EQUATIONS OF MOTION
Γ : [t0 , t1 ] × [−ε0 , ε0 ] → M
1It follows from the Newton-Laplace principle that L could depend only on generalized
coordinates and velocities, and on time.
1.2. THE PRINCIPLE OF LEAST ACTION 5
∂Γ
δγ = ∈ Tγ P (M )
∂ε ε=0
d
S(γε ) = 0
dε ε=0
(q, v) = (q 1 , . . . , q n , v 1 , . . . , v n )
n
X ∂f ∂f
(1.3) v(f ) = vi i
=v .
i=1
∂q ∂q
where the dot does not stand for the time derivative. Since we only consider
paths in T M that are tangential lifts of paths in M , there will be no confusion4.
parts, gives
d
0= S(γε )
dε ε=0
Z t1
d
= L (q(t, ε), q̇(t, ε), t) dt
dε ε=0 t0
n Z t1
X ∂L i ∂L i
= δq + i δ q̇ dt
i=1 t0
∂q i ∂ q̇
n Z
X 1 ∂Lt n t1
d ∂L i
X ∂L i
= i
− i
δq dt + δq .
i=1 t0
∂q dt ∂ q̇ i=1
∂ q̇ i t0
The second sum in the last line vanishes due to the property δq i (t0 ) = δq i (t1 ) =
0, i = 1, . . . , n. The first sum is zero for arbitrary smooth functions δq i on the
interval [t0 , t1 ] which vanish at the endpoints. This implies that for each term
in the sum the integrand is identically zero,
∂L d ∂L
(q(t), q̇(t), t) − (q(t), q̇(t), t) = 0, i = 1, . . . , n.
∂q i dt ∂ q̇ i
P\(M ) at γ ∈ P\(M ) is the space of all smooth vector fields along the path γ in
M (no condition at the endpoints). The computation in the proof of Theorem
1.1 yields the following formula for the first variation of the action with free
ends:
Z t1 t1
∂L d ∂L ∂L
(1.6) δV S = − v dt + v .
t0 ∂q dt ∂ q̇ ∂ q̇ t0
In order for this system to be solvable for the highest derivatives for all initial
conditions in T U , the symmetric n × n matrix
2 n
∂ L
HL (q, q̇, t) = (q, q̇, t)
∂ q̇ i ∂ q̇ j i,j=1
should be invertible on T U .
Definition. A Lagrangian system (M, L) is called non-degenerate if for
every coordinate chart U on M the matrix HL (q, q̇, t) is invertible on T U .
Otherwise Lagrangian system is called singular.
Remark. Note that the n × n matrix HL is a Hessian of the Lagrangian
function L for vertical directions on T M . Under the change of standard coor-
dinates q 0 = F (q) and q̇ 0 = F∗ (q)q̇ it has the transformation law
where F∗ (q)τ is the transposed matrix, so that the condition det HL 6= 0 does
not depend on the choice of standard coordinates.
Inverting the matrix HL , we can write Euler-Lagrange equations for a non-
degenerate Lagrangian in the form
r 7→ g · r + r0 , t 7→ t + t0 ,
(1.8) r 7→ r + vt, t 7→ t,
so that
r g v r g · r + vt
7→ = .
t 0 1 t t
Galileo’s Relativity Principle. The laws of motion are invariant with
respect to the Galilean group.
These postulates impose restrictions on Lagrangians of mechanical systems.
In particular, Lagrangian L of a closed system 6 does not explicitly depend on
time.
Example 1.1 (Free particle). The configuration space for a free particle
is M = R3 , and it can be deduced from Galileo’s relativity principle that the
Lagrangian for a free particle is
L = 12 mṙ 2 .
Here m > 07 is the mass of a particle and ṙ 2 = |ṙ|2 is the length square of the
velocity vector ṙ ∈ Tr R3 ' R3 . Indeed, under the Galilean transformation (1.8)
d
(1.9) L = 12 mṙ 2 7→ L0 = L = 21 m(ṙ + v)2 = L + (mrv + 21 v 2 t),
dt
so that Lagrangians L and L0 have the same equations of motion (cf. Problem
1.2). Specifically, Euler-Lagrange equations give Newton’s law of inertia,
r̈ = 0.
M = R3N = R3 × · · · × R3
| {z }
N
where
N
X
1 2
T = 2 ma ṙa
a=1
is called kinetic energy of a system and V (r) is potential energy. The Euler-
Lagrange equations give Newton’s equations
ma r̈a = Fa ,
where
∂V
Fa = −
∂ra
is a force on the a-th particle, a = 1, . . . , N . Forces of this form are called
conservative. Thus the interaction of particles is through the action of potential
forces, and is an instantaneous action at a distance 8.
7Otherwise the action functional is not bounded from below.
8This means a phenomenon in which a change in intrinsic properties of one system induces
an instantaneous change in the intrinsic properties of a distant system without a process that
carries this influence contiguously in space and time.
1.4. EXAMPLES OF LAGRANGIAN SYSTEMS 11
It follows from the isotropy of space that V (r) depends only on relative distances
between the particles, so that the Lagrangian of a closed system of N particles
with pair-wise interaction has the form
N
X X
1 2
L= 2 ma ṙa − Vab (|ra − rb |).
a=1 1≤a<b≤N
L = 12 mq̇ 2 − V0 (q),
q̈ i + ωi2 q i = 0, i = 1, . . . , n,
where
1 σλ ∂gµλ ∂gνλ ∂gµν
Γσµν = g + −
2 ∂xν ∂xµ ∂xλ
are Christoffel’s symbols. The Euler-Lagrange equations of a free particle mov-
ing on a Riemannian manifold are geodesic equations.
Let ∇ be the Levi-Civita connection — the metric connection in the tangent
bundle T M — and let ∇ξ be a covariant derivative with respect to the vector
field ξ ∈ Vect(M ). Explicitly,
µ
µ ∂η µ λ ∂ ∂
(∇ξ η) = + Γνλ η ξ ν , where ξ = ξ µ (x) µ , η = η µ (x) µ .
∂xν ∂x ∂x
1.4. EXAMPLES OF LAGRANGIAN SYSTEMS 13
For a path γ(t) = (xµ (t)) denote by ∇γ̇ a covariant derivative along γ,
dη µ (t) ∂
(∇γ̇ η)µ (t) = + Γµνλ (γ(t))ẋν (t)η λ (t), where η = η µ (t)
dt ∂xµ
is a vector field along γ. Formula (1.5) can now be written in an invariant form
Z t1
δS = − h∇γ̇ γ̇, δγidt,
t0
which is known as the formula for the first variation of the action in Riemannian
geometry.
Problem 1.1. Show that the action functional is given by the evaluation of the
1-form Ldt on T M × R over the 1-chain γ̃ on T M × R,
Z
S(γ) = Ldt,
γ̃
where γε1 ,ε2 is a smooth two-parameter family of paths in M such that the paths γε1 ,0
and γ0,ε2 in P (M ) at the point γ0,0 = γ ∈ P (M ) have tangent vectors V1 and V2 ,
respectively. For a Lagrangian system (M, L) find the second variation of S and verify
that for given V1 and V2 it does not depend on the choice of γε1 ,ε2 .
Problem 1.5. Prove that the second variation of the action functional in Rie-
mannian geometry is given by
Z t1
δ2 S = hJ (δ1 γ), δ2 γidt.
t0
Here δ1 γ, δ2 γ ∈ Tγ P M , J = −∇2γ̇
− R(γ̇, · )γ̇ is the Jacobi operator, and R is a
curvature operator — a fibre-wise linear mapping R : T M ⊗ T M → End(T M ) of
vector bundles, defined by R(ξ, η) = ∇η ∇ξ − ∇ξ ∇η + ∇[ξ,η] : T M → T M , where
ξ, η ∈ Vect(M ).
LECTURE 2
To describe the motion of a mechanical system one needs to solve the Euler-
Lagrange equations — a system of second order ordinary differential equations
for the generalized coordinates. This could be a very difficult problem. There-
fore of particular interest are those functions of generalized coordinates and
velocities, which remain constant during the motion.
d
I(γ 0 (t)) = 0
dt
∂L
Lemma 2.1. The energy E = q̇ − L is a well-defined function on T M ×R.
∂ q̇
where
( n
)n
X ∂2F i k
G(q, q̇) = q̇ ,
∂q j ∂q k
k=1 i,j=1
15
16 2. INTEGRALS OF MOTION
so that
∂L 0 ∂L 0 ∂L
dL = dq + dq̇ + dt
∂q 0 ∂ q̇ 0 ∂t
∂L ∂L ∂L ∂L
= 0
F ∗ (q) + 0
G(q, q̇) dq + 0
F∗ (q)dq̇ + dt
∂q ∂ q̇ ∂ q̇ ∂t
∂L ∂L ∂L
= dq + dq̇ + dt.
∂q ∂ q̇ ∂t
Thus under a change of coordinates
∂L ∂L ∂L ∂L
0
F∗ (q) = and q̇ 0 0
= q̇ ,
∂ q̇ ∂ q̇ ∂ q̇ ∂ q̇
so that E is a well-defined function on T M .
Corollary 2.1. Under a change of local coordinates on M , components of
∂L ∂L ∂L
a vector (q, q̇, t) = , . . . , n transform like components of a 1-form
∂ q̇ ∂ q̇ 1 ∂ q̇
∂L
on M . In classical terminology, is a covariant vector.
∂ q̇
Let θL be a 1-form on T M , defined in standard coordinates associated with
a coordinate chart U ⊂ M by
n
X ∂L i ∂L
(2.1) θL = i
dq = dq.
i=1
∂ q̇ ∂ q̇
It follows from Corollary 2.1 that θL is a well-defined 1-form on T M .
Proposition 2.1 (Conservation of energy). The energy of a closed system
is an integral of motion.
Proof. For an extremal γ put E(t) = E(γ 0 (t)). We have, according to the
Euler-Lagrange equations,
dE d ∂L ∂L ∂L ∂L ∂L
= q̇ + q̈ − q̇ − q̈ −
dt dt ∂ q̇ ∂ q̇ ∂q ∂ q̇ ∂t
d ∂L ∂L ∂L ∂L
= − q̇ − =− .
dt ∂ q̇ ∂q ∂t ∂t
∂L
Since for a closed system = 0, the energy is conserved.
∂t
Conservation of energy for a closed mechanical system is a fundamental law
of physics, which follows from the homogeneity of time. For a general closed
system of N interacting particles considered in Example 1.2,
N
X N
X
E= ma ṙa2 − L = 1 2
2 ma ṙa + V (r).
a=1 a=1
In other words, the total energy E = T + V is a sum of the kinetic energy and
the potential energy.
2.2. NOETHER THEOREM 17
Thus
n n
X ∂ X ∂ai ∂
(2.2) X0 = ai (q) i
+ q̇ j j (q) i ,
i=1
∂q i,j=1
∂q ∂ q̇
∂L ∂L
dL(X 0 )(γ 0 (t)) = a+ ȧ.
∂q ∂ q̇
(2.3) I = θL (X 0 ).
where (q, τ ) are local coordinates on M1 and (q, τ, q̇, τ̇ ) are standard coordinates
on T M1 . The Noether integral I1 for a closed system (M1 , L1 ) defines an integral
of motion I for a system (M, L) by the formula
When the Lagrangian L does not depend on time, L1 is invariant with respect
to the one-parameter group of translations τ 7→ τ + s, and the Noether integral
∂L1
I1 = gives I = −E.
∂ τ̇
Noether’s theorem can be generalized further as follows.
Proposition 2.2. Suppose that for a given Lagrangian L : T M → R there
exist a vector field X on M and a function K on T M , such that for every path
γ in M ,
d
dL(X 0 )(γ 0 (t)) = K(γ 0 (t)).
dt
Then
n
X ∂L
I= ai (q) i (q, q̇) − K(q, q̇)
i=1
∂ q̇
is an integral of motion.
dI ∂L ∂L dK d
= a+ ȧ − = dL(X̃)(γ 0 (t)) − K(γ 0 (t)) = 0.
dt ∂q ∂ q̇ dt dt
2.3. Examples of conservation laws
Example 2.1 (Conservation of momentum). Let M = V be a vector space,
and suppose that a Lagrangian L is invariant with respect to a one-parameter
group gs (q) = q + sv, v ∈ V . According to Noether’s theorem,
n
X ∂L
I= vi
i=1
∂ q̇ i
N
X
1 ∂L ∂L ∂L
I= (u · ra ) 1
+ (u · ra )2 2 + (u · ra )3 3
a=1
∂ ṙa ∂ ṙa ∂ ṙa
0 0 0
is an integral of motion. Let u = u1 X1 +u2 X2 +u3 X3 , where X1 = 0 0 −1 , X2 =
0 0 1 0 −1 0 01 0
0 0 0 , X3 = 1 0 0 is the basis in so(3) ' R3 corresponding to the rota-
−1 0 0 0 0 0
tions about the vectors e1 , e2 , e3 of the standard orthonormal basis in R3 . Since
2.3. EXAMPLES OF CONSERVATION LAWS 21
I = u1 M1 + u2 M2 + u3 M3 ,
A complete general solution can be obtained for three very important exam-
ples: a motion on the real line, a system of two interacting particles, including
the Kepler problem, and the rotation of a rigid body.
L = 12 mẋ2 − V (x).
23
24 3. INTEGRATION OF EQUATIONS OF MOTION
V (x) = E
x1 x2 x3 x
Figure 1
Thus on Fig. 1 the motion between points x1 and x2 is periodic, and in the
region x3 ≤ x the motion is infinite; all other regions there are forbidden.
On the phase plane with coordinates (x, y) Newton’s equation reduces to the
first order system
dV
mẋ = y, ẏ = − .
dx
Trajectories correspond to the phase curves (x(t), y(t)), which lie on the level
sets
y2
+ V (x) = E
2m
of the energy function. The points (x0 , 0), where x0 is a critical point of the po-
tential energy V (x), correspond to the equilibrium solutions. The local minima
correspond to the stable solutions and local maxima correspond to the unstable
solutions. For the values of E which do not correspond to the equilibrium solu-
tions the level sets are smooth curves. These curves are closed if the motion is
finite.
The simplest non-trivial one-dimensional system, besides the free particle, is
the harmonic oscillator with V (x) = 21 kx2 (k > 0), considered in Example 1.4.
The general solution of the equation of motion is
q motion is infinite for E > V — the particle goes to ∞ with finite velocity
the
2
µ (E − V ).
Veff
r0
r
V0
Figure 2
This is the equation of a conic section with one focus at the origin. Quantity
2p is called the latus rectum of the orbit, and e is called the eccentricity. The
choice C = 0 is such that the point with ϕ = 0 is the point nearest to the origin
(called the perihelion). When V0 ≤ E < 0, the eccentricity e < 1 so that the
orbit is the ellipse2 with the major and minor semi-axes
p α p |M |
(3.6) a= 2
= , b= √ =p .
1−e 2|E| 1−e 2 2µ|E|
p p
Correspondingly, rmin = , rmax = , and the period T of elliptic orbit
1+e 1−e
is given by r
µ
T = πα .
2|E|3
The last formula is Kepler’s third law. When E > 0, the eccentricity e > 1
and the motion is infinite — the orbit is a hyperbola with the origin as internal
focus. When E = 0, the eccentricity e = 1 — the particle starts from rest at ∞
and the orbit is a parabola.
For the repulsive case α < 0 the effective potential energy Veff (r) is always
positive and decreases monotonically from ∞ to 0. The motion is always infinite
and the trajectories are hyperbolas (parabola if E = 0)
p
= −1 + e cos ϕ
r
with s
M2 2EM 2
p= and e = 1 + .
αµ µα2
Kepler’s problem is very special: for every α ∈ R the Lagrangian system on
R3 with
α
(3.7) L = 21 µṙ 2 +
r
has three extra integrals of motion W1 , W2 , W3 in addition to the components
of the angular momentum M . The corresponding vector W = (W1 , W2 , W3 ),
called the Laplace-Runge-Lenz vector, is given by
αr
(3.8) W = ṙ × M − .
r
αr
Indeed, using equations of motion µr̈ = − 3 and conservation of the angular
r
momentum M = µr × ṙ, we get
αṙ α(ṙ · r)r
Ẇ = µr̈ × (r × ṙ) − +
r r3
αṙ α(ṙ · r)r
= (µr̈ · ṙ)r − (µr̈ · r)ṙ − +
r r3
= 0.
2The statement that planets have elliptic orbits with a focus at the Sun is Kepler’s first
law.
3.4. THE MOTION OF A RIGID BODY 29
2M 2 E
(3.9) W 2 = α2 +
µ
where
p2 α
E= −
2µ r
is the energy corresponding to the Lagrangian (3.7). The fact that all orbits are
conic sections follows from this extra symmetry of the Kepler problem.
for all x, y, z ∈ g. Thus we have hΩ, Ωie = B(A·Ω, Ω) for some symmetric linear
operator A : g → g, which is positive-definite with respect to the Killing form.
Such linear operator A is called the inertia tensor of the body, and Lagrangian
(3.10) takes the form
Now we are ready to derive equations of motion for Lagrangian (3.11). Sim-
ilar to Sect. 1.2, for a path g : [t0 , t1 ] → G, consider the family
The corresponding angular velocity Ω(t, ε) = g −1 (t, ε)ġ(t, ε) ∈ g takes the form
Ω(t, ε) = Adhε (t) Ω(t) + εu̇(t), where hε (t) = exp{−εu(t)}, Ω(t) = Ω(t, 0),
and Adg stands for the adjoint action of G on g. Thus for the infinitesimal
variation
∂Ω(t, ε)
δΩ(t) = ∈g
∂ε ε=0
we readily obtain
(3.12) δΩ = u̇ + [Ω, u].
As in Sect. 1.2 in Lecture 1, consider the action functional
1 t1
Z
S(g, ġ) = B(A · Ω(t), Ω(t))dt.
2 t0
Using the symmetry of the operator A we obtain
1 t1
Z Z t1
δS = (B(A · δΩ(t), Ω(t)) + B(A · Ω(t), δΩ(t))) dt = B(A·Ω(t), δΩ(t))dt,
2 t0 t0
and using (3.12), ad g-invariance of the Killing form and integration by parts,
we get
Z t1
δS = B(A · Ω(t), u̇(t) + [Ω(t), u(t)])dt
t0
Z t1
= B(−A · Ω̇(t) + [A · Ω(t), Ω(t)], u(t))dt.
t0
A · Ω̇ = ad∗Ω (A · Ω),
where ad∗Ω is the adjoint of the operator adΩ on g with respect to the inner
product h , ie . These equations are called Euler-Arnold equations for the
geodesics of a left-invariant Riemannian metric on a Lie group G, finite or
infinite-dimensional.
3.4. THE MOTION OF A RIGID BODY 31
Returning to the case G = SO(3), the principal axes of inertia of the body
are orthonormal eigenvectors e1 , e2 , e3 of A; corresponding eigenvalues I1 , I2 , I3
are called the principal moments of inertia. Choosing the principal axes of
inertia as a basis in g and setting Ω = Ω1 e1 + Ω2 e2 + Ω3 e3 , we get the Lie
algebra isomorphism g ' R3 ,
0 −Ω3 Ω2
g 3 Ω = Ω3 0 −Ω1 7→ (Ω1 , Ω2 , Ω3 ) ∈ R3 ,
−Ω2 Ω1 0
where the Lie bracket in R3 is given by the cross-product (see Example 2.2).
Indeed, for the matrices
0 −a3 a2 0 −b3 b2
a = a3 0 −a1 and b = b3 0 −b1
−a2 a1 0 −b2 b1 0
[a, b] = c,
B(a, b) = a · b.
A · Ω = AΩ + ΩA,
I2 + I3 − I1 I1 + I3 − I2 I1 + I2 − I3
l1 = , l2 = , l3 = .
2 2 2
Thus
[A · Ω, Ω] = AΩ2 − Ω2 A
and (3.13) become celebrated Euler’s equations for rotation of a free rigid body
around a fixed point,
— the system of first order differential equations. Finally, the position g(t) of a
rigid body is determined from the first order linear matrix differential equation,
ġ = gΩ.
32 3. INTEGRATION OF EQUATIONS OF MOTION
then there are orbits with rmin = 0 — “fall” of the particle to the center.
Problem 3.3. Prove that all finite trajectories in the central field are closed only
when
α
V (r) = kr2 , k > 0, and V (r) = − , α > 0.
r
Problem 3.4 (Hamilton’s Theorem). Prove that the velocity vector v = ṙ(t)
of the Kepler problem moves along a circle C in the plane P from Sect. 3.2, not in
general centered at the origin. Any such “velocity circle” can occur, and a circle C,
together with its orientation, determines the orbit r = r(t) uniquely.
Problem 3.5. Derive Kepler’s third law from Kepler’s second law and equation
(3.6).
Problem 3.6. Find parametric equations for orbits in the Kepler’s problem.
Problem 3.7. For the Kepler problem, consider vector fields Y = (Y 1 , Y 2 , Y 3 )
on R6 , defined by (2.4) with aij (r, ṙ) = 2ṙi rj − ri ṙj − δ ij r · ṙ. Prove that they
2αr
satisfy (2.5) with K = = (K 1 , K 2 , K 3 ), and show that corresponding integrals
r
of motions are components of the Laplace-Runge-Lenz vector.
Problem 3.8. Prove that the Laplace-Runge-Lenz vector W points in the di-
rection of the major axis of the orbit and that |W | = αe, where e is the eccentricity
of the orbit.
Problem 3.9. Using the conservation of the Laplace-Runge-Lenz vector, prove
that trajectories in Kepler’s problem with E < 0 are ellipses. (Hint: Evaluate W · r
and use the previous problem.)
Problem 3.10. Derive Euler-Arnold equations.
Problem 3.11. In case g = so(3) prove that for every symmetric A ∈ End g
there is a symmetric 3 × 3 matrix A such that
A · Ω = AΩ + ΩA.
(p, q) = (p1 , . . . , pn , q 1 , . . . , q n )
∂f
pi (df ) = , i = 1, . . . , n.
∂q i
Equivalently, standard coordinates on T ∗ U are uniquely characterized by
the condition that p = (p1 , . . . , pn ) are coordinates in the fiber corresponding
∂ ∂
to the basis dq 1 , . . . , dq n for Tq∗ M , dual to the basis 1
, . . . , n for Tq M .
∂q ∂q
Definition. The 1-form θ on T ∗ M , defined in standard coordinates by
n
X
θ= pi dq i = pdq,
i=1
θL = τL∗ (θ).
1Following tradition, the first n coordinates parametrize the fiber of T ∗ U and the last n
coordinates parametrize the base.
33
34 4. LEGENDRE TRANSFORM AND HAMILTON’S EQUATIONS
∂L
where q̇ is a function of p and q defined by the equation p = (q, q̇) through
∂ q̇
∗
the implicit function theorem. The cotangent bundle T M is called the phase
space of the Lagrangian system (M, L). It turns out that on the phase space
the equations of motion take a very simple and symmetric form.
Theorem 4.1. Suppose that the Legendre transform τL : T M → T ∗ M is a
diffeomorphism. Then the Euler-Lagrange equations in standard coordinates on
TM,
d ∂L ∂L
i
− i = 0, i = 1, . . . , n,
dt ∂ q̇ ∂q
are equivalent to the following system of first order differential equations in
standard coordinates on T ∗ M :
∂H ∂H
ṗi = − , q̇ i = , i = 1, . . . , n.
∂q i ∂pi
Proof. We have
∂H ∂H
dH = dp + dq
∂p ∂q
∂L ∂L
= pdq̇ + q̇dp − dq − dq̇
∂q ∂ q̇ ∂L
p= ∂ q̇
∂L
= q̇dp − dq .
∂q p=
∂L
∂ q̇
J : T ∗ (T ∗ M ) → T (T ∗ M )
A1 (T ∗ M ) ' Vect(T ∗ M )
between the infinite-dimensional vector spaces, which is linear over the ring
C ∞ (T ∗ M ). Namely, if ϑ is a 1-form on T ∗ M , then the corresponding vector
field X = J(ϑ) on T ∗ M satisfies
and, correspondingly,
(4.4) df = −iXf ω.
d
(gt )∗ ω = LXH ω = 0,
dt t=0
where LXH is the Lie derivative along the Hamiltonian vector field XH . Since
for every vector field X,
LX (df ) = d(X(f )),
we compute
∂H i ∂H
LXH (dpi ) = −d and LXH (dq ) = d ,
∂q i ∂pi
38 4. LEGENDRE TRANSFORM AND HAMILTON’S EQUATIONS
so that
n
X
LXH (dpi ) ∧ dq i + dpi ∧ LXH (dq i )
LXH ω =
i=1
n
X ∂H ∂H
= −d ∧ dq i + dpi ∧ d = −d(dH) = 0.
i=1
∂q i ∂pi
Hamiltonian formalism
(cf. Section 1.2). The principle of least action in the phase space is the following
statement.
Theorem 5.1 (Poincaré). The admissible path σ in T ∗ M × R is an extremal
for the action functional
Z Z t1
S(σ) = (pdq − Hdt) = (pq̇ − H)dt
σ t0
if and only if it is a lift of a path γ(t) = (p(t), q(t)) in T ∗ M , where p(t) and
q(t) satisfy canonical Hamilton’s equations
∂H ∂H
ṗ = − , q̇ = .
∂q ∂p
Proof. As in the proof of Theorem 1.1, for an admissible family σε (t) =
(p(t, ε), q(t, ε), t) we compute using integration by parts,
n Z t1
d X ∂H ∂H
S(σε ) = q̇ i δpi − ṗi δq i − i δq i − δpi dt
dε ε=0 i=1 t0
∂q ∂pi
n
X t1
+ pi δq i t0
.
i=1
39
40 5. HAMILTONIAN FORMALISM
Since δq(t0 ) = δq(t1 ) = 0, the path σ is critical if and only if p(t) and q(t)
satisfy canonical Hamilton’s equations (4.1).
Remark. For a Lagrangian system (M, L), every path γ(t) = (q(t)) in the
configuration space M connecting points q0 and q1 defines an admissible path
∂L
γ̂(t) = (p(t), q(t), t) in the phase space T ∗ M by setting p = . If the Legendre
∂ q̇
transform τL : T M → T ∗ M is a diffeomorphism, then
Z t1 Z t1
S(γ̂) = (pq̇ − H)dt = L(γ 0 (t), t)dt.
t0 t0
on the space of all paths in the configuration space M connecting points q0 and
q1 and parametrized such that H( ∂L∂ q̇ (τ ), q(τ )) = E.
The functional Z
S0 (γ) = pdq
γ
where γ(τ ) is the extremal from the central field that connects q0Sand q. For
given q0 and t0 , the classical action is defined for t ∈ (t0 , t1 ) and q ∈ t0 <t<t1 Ut .
For a fixed energy E,
(5.2) S(q, t; q0 , t0 ) = S0 (q, t; q0 , t0 ) − E(t − t0 ),
where S0 is the abbreviated action from the previous section.
Theorem 5.4. The differential of the classical action S(q, t) with fixed initial
point is given by
dS = pdq − Hdt,
∂L
where p = (q, q̇) and H = pq̇ − L(q, q̇) are determined by the velocity q̇ of
∂ q̇
the extremal γ(τ ) at time t.
Proof. Let qε be a path in M passing through q at ε = 0 with the tangent
vector v ∈ Tq M ' Rn , and for ε small enough let γε (τ ) be the family of
extremals from the central field satisfying γε (t0 ) = q0 and γε (t) = qε . For the
infinitesimal variation δγ we have δγ(t0 ) = 0 and δγ(t) = v, and for fixed t we
get from the formula for variation with the free ends (1.6) that
∂L
dS(v) = v.
∂ q̇
∂S
This shows that = p. Setting q(t) = γ(t), we obtain
∂q
d ∂S ∂S
S(q(t), t) = q̇ + = L,
dt ∂q ∂t
42 5. HAMILTONIAN FORMALISM
∂S
so that = L − pq̇ = −H.
∂t
Corollary 5.5. The classical action satisfies the following nonlinear partial
differential equation
∂S ∂S
(5.3) +H , q = 0.
∂t ∂q
This equation is called the Hamilton-Jacobi equation. Hamilton’s equations
(4.1) can be used for solving the Cauchy problem
{ , } : C ∞ (T ∗ M ) × C ∞ (T ∗ M ) → C ∞ (T ∗ M )
(ii) (Skew-symmetry)
{f, g} = −{g, f }.
for all f, g, h ∈ C ∞ (T ∗ M ).
Proof. Property (i) immediately follows from the definitions of ω and J
in Section 4.3. Namely, it follows from (4.2) that
However, this expression does not contain second partial derivatives of h since
it is a commutator of two differential operators of the first order which is again
a differential operator of the first order!
44 5. HAMILTONIAN FORMALISM
The observable {f, g} is called the canonical Poisson bracket of the observ-
ables f and g. The Poisson bracket map { , } : C ∞ (T ∗ M ) × C ∞ (T ∗ M ) →
C ∞ (T ∗ M ) turns the algebra of classical observables C ∞ (T ∗ M ) into a Lie al-
gebra with a Lie bracket given by the Poisson bracket. It has an important
property that the Lie bracket is a bi-derivation with respect to the multiplica-
tion in C ∞ (T ∗ M ). The algebra of classical observables C ∞ (T ∗ M ) is an example
of the Poisson algebra — a commutative algebra over R carrying a structure of
a Lie algebra with the property that the Lie bracket is a derivation with respect
to the algebra product.
In Lagrangian mechanics, a function I on T M is an integral of motion for the
Lagrangian system (M, L) if it is constant along the trajectories. In Hamiltonian
mechanics, an observable I — a function on the phase space T ∗ M — is called an
integral of motion (first integral) for Hamilton’s equations (4.1) if it is constant
along the Hamiltonian phase flow. According to (5.6), this is equivalent to the
condition
{H, I} = 0.
It is said that the observables H and I are in involution (Poisson commute).
(5.7) g∗ ◦ J ◦ g ∗ = J.
∂K
P (t) = P (0), Q(t) = Q(0) + t (P (0)).
∂P
∂P
Assuming that the matrix is non-degenerate, the generating function S(P , q)
∂p
satisfies the differential equation
∂S
(5.10) H (P , q), q = K(P ),
∂q
where after the differentiation one should substitute q = q(P , Q), defined by
the canonical transformation g −1 . The differential equation (5.10) for fixed P ,
as it follows from (5.2), coincides with the Hamilton-Jacobi equation for the
abbreviated action S0 = S − Et where E = K(P ),
∂S
0
H (P , q), q = E.
∂q
Theorem 5.7 (Jacobi). Suppose that there is a function S(P , q) which de-
pends on n parameters P = (P1 , . . . , Pn ), satisfies the Hamilton-Jacobi equation
∂2S
(5.10) for some function K(P ), and has the property that the n×n matrix
∂P ∂q
is non-degenerate. Then Hamilton’s equations
∂H ∂H
ṗ = − , q̇ =
∂q ∂p
can be solved explicitly, and the functions P (p, q) = (P1 (p, q), . . . , Pn (p, q)),
∂S
defined by the equations p = (P , q), are integrals of motion in involution.
∂q
∂S ∂S
Proof. Set p = (P , q) and Q = (P , q). By the inverse function
∂q ∂P
theorem, g(p, q) = (P , Q) is a local canonical transformation with the gener-
ating function S. It follows from (5.10) that H(p(P , Q), q(P , Q)) = K(P ), so
that Hamilton’s equations take the form (5.9). Since ω = dP ∧ dQ, integrals of
motion P1 (p, q), . . . , Pn (p, q) are in involution.
Problem 5.5. Find the complete integral for the case of a particle in R3 moving
in a central field.
LECTURE 6
49
50 6. SYMPLECTIC AND POISSON MANIFOLDS
form
n
i X
ω = dz ∧ dz̄ = dxα ∧ dy α = dx ∧ dy.
2 α=1
This example naturally leads to the following definition.
Definition. A symplectic vector space is a pair (V, ω), where V is a vector
space over R and ω is a non-degenerate, skew-symmetric bilinear form on V .
It follows from basic linear algebra that every symplectic vector space V has
a symplectic basis — a basis e1 , . . . , en , f1 , . . . , fn of V , where 2n = dim V , such
that
ω(ei , ej ) = ω(fi , fj ) = 0 and ω(ei , fj ) = δji , i, j = 1, . . . , n.
2n
X ∂
J(dxi ) = − ω ij (x) , i = 1, . . . , 2n.
j=1
∂xj
∂H ∂H
ṗ = − , q̇ = .
∂q ∂p
Suppose now that the Hamiltonian vector field XH on M is complete. The
Hamiltonian phase flow on M associated with a Hamiltonian H is a one-
parameter group {gt }t∈R of diffeomorphisms of M generated by XH . The
following statement generalizes Theorem 4.3.
Theorem 6.2. The Hamiltonian phase flow preserves the symplectic form.
Proof. It is sufficient to show that LXH ω = 0. Using Cartan’s formula
(1.1) and dω = 0, we get for every X ∈ Vect(M ),
LX ω = (d ◦ iX )(ω),
LXH ω = −d(dH) = 0.
dft
= XH (ft )
dt
— Hamilton’s equation for classical observables. Hamilton’s equations for ob-
servables on M have the same form as Hamilton’s equations on M = T ∗ M ,
considered in Section 2.3. Since
{f, g} = ω(Xf , Xg ), f, g ∈ C ∞ (M ).
df
(6.2) = {H, f },
dt
understood as a differential equation for a family of functions ft on M with the
initial condition ft |t=0 = f . In local coordinates x = (x1 , . . . , x2n ) on M ,
2n
X ∂f (x) ∂g(x)
{f, g}(x) = − ω ij (x) .
i,j=1
∂xi ∂xj
Proof. The first two properties are obvious. It follows from the definition
of a Poisson bracket and the formula
[Xf , Xg ](h) = (Xg Xf − Xf Xg )(h) = {g, {f, h}} − {f, {g, h}}
The 2-tensor η ij (x), called a Poisson tensor, defines a global section η of the
vector bundle T M ∧ T M over M .
The evolution of classical observables on a Poisson manifold is given by
Hamilton’s equations, which have the same form as (6.2),
df
= XH (f ) = {H, f }.
dt
The phase flow gt for a complete Hamiltonian vector field XH = {H, · } defines
the evolution operator Ut : A → A by
so that ht and {ft , gt } satisfy the same differential equation (6.2). Since these
functions coincide at t = 0, (6.5) follows from the uniqueness theorem for the
ordinary differential equations.
Conversely, we get the Jacobi identity for the functions f, g, and H by dif-
ferentiating (6.5) with respect to t at t = 0.
1Here g is not the phase flow!
t
6.2. POISSON MANIFOLDS 55
corresponding vector field on M . The Jacobi identity for the Poisson bracket
{ , } is equivalent to LXf η = 0 for every f ∈ A, so that
LXf ω = 0.
Since Xf = Jdf , we have ω(X, Jdf ) = df (X) for every X ∈ Vect(M ), so that
ω(Xf , Xg ) = {f, g}.
By Cartan’s formula for the exterior differential,
1
dω(X, Y, Z) = 3 (LX ω(Y, Z) + LY ω(Z, X) + LZ ω(X, Y )
−ω([X, Y ], Z) − ω([Z, X], Y ) − ω([Y, Z], X)) ,
where X, Y, Z ∈ Vect(M ). Now setting X = Xf , Y = Xg , Z = Xh , we get
1
dω(Xf , Xg , Xh ) = 3 (ω(Xh , [Xf , Xg ]) + ω(Xf , [Xg , Xh ]) + ω(Xg , [Xh , Xf ]))
1
= 3 ω(Xh , X{f,g} ) + ω(Xf , X{g,h} ) + ω(Xg , X{h,f } )
1
= 3 ({h, {f, g}} + {f, {g, h}} + {g, {h, f }})
= 0.
The exact 1-forms df, f ∈ A, generate the vector space of 1-forms A1 (M )
as a module over A, so that Hamiltonian vector fields Xf = Jdf generate the
vector space Vect(M ) as a module over A. Thus dω = 0 and (M , ω) is a
symplectic manifold associated with the Poisson manifold (M , { , }).
Remark. One can also prove this theorem by a straightforward computation
in local coordinates x = (x1 , . . . , xN ) on M . Just observe that the condition
∂ηij (x) ∂ηjl (x) ∂ηli (x)
+ + = 0, i, j, l = 1, . . . , N,
∂xl ∂xi ∂xj
which is a coordinate form of dω = 0, follows from the condition
N
∂η kl (x) ∂η ik (x) ∂η li (x)
X
ij lj kj
η (x) + η (x) + η (x) = 0,
j=1
∂xj ∂xj ∂xj
so that
{Φξ , Φη } = Φ[ξ,η] + c(ξ, η)
for some constants c(ξ, η). The Hamiltonian action is called a Poisson action if
there is a choice of functions Φξ such that the linear mapping Φ : g → C ∞ (M )
is a homomorphism of Lie algebras,
H(g · x) = H(x), g ∈ G, x ∈ M .
Corollary 6.10. Let (M, L) be a Lagrangian system such that the Legendre
transform τL : T M → T ∗ M is a diffeomorphism. Then if a Lie group G
is a symmetry of (M, L), then G is a symmetry group of the corresponding
Hamiltonian system ((T ∗ M, ω), H = EL ◦ τL−1 ), and the corresponding G-action
on T ∗ M is Poisson. In particular, Φξ = −Iξ ◦ τL−1 , where Iξ are Noether
integrals of motion for the one-parameter subgroups of G generated by ξ ∈ g.
58 6. SYMPLECTIC AND POISSON MANIFOLDS
Proof. Let X be the vector field associated with the one-parameter sub-
group {esξ }s∈R of diffeomorphisms of M , used in Theorem 2.2, and let X 0 be
its lift to T M . We have2
(6.7) Xξ = −(τL )∗ (X 0 ),
and it follows from (2.3) that Φξ = iXξ (θ) = θ(Xξ ), where θ is the canonical
Liouville 1-form on T ∗ M . From Cartan’s formula (1.1) and formula LX 0 (θL ) = 0
(see Problem 2.4) we get
It follows from (6.1) that Xξ = J(dΦξ ) and the G-action is Hamiltonian. Using
again the formula LX 0 (θL ) = 0 and Cartan’s formula (1.2), we obtain
Φ[ξ,η] = i[Xξ ,Xη ] (θ) = LXξ (iXη (θ)) + iXη (LXξ (θ))
= Xξ (Φη ) = {Φξ , Φη }.
L = 21 mṙ 2 − V (r)
for a particle in R3 moving in a central field (see Section 3.2) is invariant with
respect to the action of the group SO(3) of orthogonal transformations of the
Euclidean space R3 . Let u1 , u2 , u3 be a basis for the Lie algebra so(3) corre-
sponding to the rotations with the axes given by the vectors of the standard
basis e1 , e2 , e3 for R3 (see Example 2.2 in Section 2.2). These generators satisfy
the commutation relations
[ui , uj ] = εijk uk ,
where i, j, k = 1, 2, 3, and εijk is a totally anti-symmetric tensor, ε123 = 1.
Corresponding Noether integrals of motion are given by Φui = −Mi , where
M1 = (r × p)1 = r2 p3 − r3 p2 ,
M2 = (r × p)2 = r3 p1 − r1 p3 ,
M3 = (r × p)3 = r1 p2 − r2 p1
According to Theorem 6.9 and Corollary 6.10, Poisson brackets of the compo-
nents of the angular momentum satisfy
{Mi , Mj } = −εijk Mk ,
(7.3) q̇ = v.
1To avoid confusion, here we do not denote standard coordinates on T M by (q, q̇).
61
62 7. HAMILTONIAN SYSTEMS WITH CONSTRAINTS
d ∂L ∂2L ∂L
0= − (q̇ − v) − .
dt ∂v ∂q∂v ∂q
ϑL = θL − Edt,
∂L
where θL = dq is the 1-form associated with the Lagrangian L : T M → R
∂v
∂L
(see formula (2.1) in Lecture 2), and E = v − L is the corresponding energy
∂v
(see Sect. 2.1 in Lecture 2).
Definition. Lagrangian L, given by (7.4), is called non-degenerate, if the
2-form !
N N
X
α
X ∂fβ
ω=d fα (ξ)dξ = (ξ)dξ α ∧ dξ β
α=1
∂x α
α,β=1
is non-degenerate on M.
7.2. SINGULAR LAGRANGIANS 63
It follows from the previous remark and Problem 2.1 in Lecture 2, that
for Lagrangians (7.1) this definition agrees with the one given in Lecture 1.
If the Lagrangian L is non-degenerate, it follows from the Darboux theorem
that N = 2n is even and there exist local canonical coordinates (p, q) =
(p1 , . . . , pn , q 1 , . . . , q n ) on M such that
L = pq̇ − H(p, q)
with the Hamiltonian function H(p, q). This repeats derivation of the Hamil-
ton’s equations given in Sect. 4.2 in Lecture 4, but without explicitly using
Legendre transform.
Remark. In this case we trivially have
∂L
p= and H = pq̇ − L.
∂ q̇
7.2. Singular Lagrangians
Here we consider important case when Lagrangian (7.4) is singular. Dar-
boux theorem is still applicable and guarantees existence of local coordinates
(p, q, λ) = (p1 , . . . , pn , q 1 , . . . , q n , λ1 , . . . , λm ) on M, where N = 2n + m, such
that
ϑL = pdq − H(p, q, λ)dt + dS
for some (local) function S(p, q, λ). Since addition of the exact form does not
change equations of motion (see Problem 1.2 in Lecture 1), the Euler-Lagrange
equations for the Lagrangian L have the following form
∂H ∂H ∂H
(7.6) ṗ = − , q̇ =
and = 0.
∂q ∂p ∂λ
2 m
∂ H
Now suppose that the m×m matrix has constant rank k on
∂λa ∂λb a,b=1
M. If k = m, it follows from the implicit function theorem that the equations
∂H
= 0 in (7.6) determine a submanifold M̃ in M of dimension N − m = 2n,
∂λ
given by the equations λa = λa (p, q), a = 1, . . . , m. In other words, in this case
it is possible to exclude coordinates λ = (λ1 , . . . , λm ). Putting
∂2H
= 0, a, b = k + 1, . . . , m,
∂λa ∂λb
so that H is linear function of λk+1 , . . . , λm . Thus from the very beginning we
can assume that M = M0 × Rm , where M0 has canonical coordinates (p, q)
and symplectic form ω = dp ∧ dq, and consider singular Lagrangians on T M of
the form
m
X
(7.7) L = pq̇ − H(p, q) − λa ϕa (p, q),
a=1
and
(7.10) ϕa (p, q) = 0, a = 1, . . . , m.
Lemma 7.3. For the first class constraints trajectories (p(t), q(t)) of the
Hamilton’s equations (7.8)–(7.9) lie on M0 if (p(0), q(0)) ∈ M0 .
and it follows from (7.11) that ϕ̇a = 0 on M0 . Thus ϕa (p(t), q(t)) = ϕa (p(0), q(0)) =
0, a = 1, . . . , m.
66 7. HAMILTONIAN SYSTEMS WITH CONSTRAINTS
and it follows from (7.11) that restriction of this equation to M0 does not depend
on the choice of a representative in f mod I and defines the evolution in the
algebra A/I. Still, this evolution depend on the choice of arbitrary parameters
λ1 , . . . , λm .
Definition. Admissible observables are functions f ∗ on M0 whose exten-
sions f to M satisfy
{f ∗ , g ∗ }0 = {f, g}|M0 ,
and it follows from (7.13) that admissible observables form a Poisson algebra
A∗ . For admissible observables equation (7.13) takes the form
(7.14) f˙∗ = {H ∗ , f ∗ }0
A0 /I ' A∗ .
Proof. Follows from Lemma 7.2, equations (7.11) and Jacobi identity.
We also have
ω(Xϕa , Xϕb ) = {ϕa , ϕb },
7.3. FIRST CLASS CONSTRAINTS AND REDUCED PHASE SPACE 67
so that
ωm (Xϕa , Xϕb ) = 0 for all m ∈ M0 .
Denote by Ya the vector vector fields Xϕa along M0 . It follows from (7.10) that
Ya are tangent to M0 and
ω|M0 (Ya , Yb ) = 0, a, b = 1, . . . , m.
(7.16) χa (p, q) = 0, a = 1, . . . , m,
(7.18) {χa , χb } = 0, a, b = 1, . . . , m,
such that
m
X
ω= dpa ∧ dq a + dp∗ ∧ dq ∗ .
a=1
68 7. HAMILTONIAN SYSTEMS WITH CONSTRAINTS
det {ϕa , ϕb } =
6 0,
are called the second class constraints. In this case m is necessarily even, m =
2k, and the submanifold M0 , determined by equations (7.10) is a symplectic
manifold with a symplectic form ω0 = ω|M0 . In this case Poisson bracket { , }0
corresponding to ω0 is obtained by the following construction. Let Cab be the
inverse matrix to ({ϕa , ϕb }), and let { , } be a Poisson bracket on M associated
with the symplectic form ω.
Definition. Dirac bracket { , }DB on M is given by the following formula
2k
X
(7.20) {f, g}DB = {f, g} − {f, ϕa }Cab {ϕb , g}.
a,b=1
Problem 7.1. Prove that formula (7.1) gives a well defined function L on T M.
Problem 7.2. Prove Lemma 7.2.
Problem 7.3. Prove that the symplectic quotient construction (see Problem 6.4
in Lecture 6) in case p = 0 is a particular case of the Dirac formalism, where ϕa are
the Hamiltonian functions of the Hamiltonian vector fields Xξa associated with a basis
ξ a of the Lie algebra g.
Problem 7.4. Prove (7.19) by computing Poisson bracket {f, g} on M in coor-
dinates η = (pa , p∗ , ϕa , q ∗ ).
Problem 7.5. Prove Lemma 7.5.
Problem 7.6. Prove Corollary 7.1.
Notes and references
The paper
71
72 NOTES AND REFERENCES
Maxwell equations
N
X
ρ(r) = ea δ(r − ra ).
a=1
dr0 (t)
j(r, t) = e0 v(t)δ(r − r0 (t)), where v(t) = .
dt
In general, the current density is
1We are using standard notations for the divergence and curl from the multivariable
calculus.
75
76 8. MAXWELL EQUATIONS
— there are no magnetic charges, the total magnetic flux through a closed
surface is zero;
∂B
(8.3) ∇×E =− (Faraday’s induction law)
∂t
— the voltage induced in a closed circuit is proportional to the rate of change
of the magnetic flux it encloses;
∂E
(8.4) ∇ × B = µ0 j + µ0 ε0 (Ampère’s circular law)
∂t
— the magnetic field induced around a closed loop is proportional to the electric
current plus displacement current (rate of change of electric field) it encloses.
Here the constant ε0 is called a permitivity of the free space and the constant
µ0 is called permeability of the free space or magnetic constant. They satisfy
1
µ0 ε0 = ,
c2
where c is the speed of light in the free space.2 Maxwell equations imply all
laws of the electromagnetism: Coulomb law, Bio-Laplace-Savart law, etc.
It follows from equation (8.2) that there is a vector-valued function A(r, t),
called vector potential, A = (Ax , Ay , Az ), such that
(8.5) B = ∇ × A.
Plugging (8.5) into (8.3) we get
∂A
∇× E+ = 0,
∂t
so that there is a function ϕ(r, t), called scalar potential, such that
∂A
(8.6) E = −∇ϕ − .
∂t
Formulas (8.5) and (8.6) solve the first pair of Maxwell equations — equations
(8.2)–(8.3).
and the formula for ∗F follows from the definition of the 2-form F .
To summarize, Maxwell equations in an empty space (without sources) can
be written succinctly as
(8.11) dF = 0 and d ∗ F = 0.
8.3. MAXWELL’S EQUATIONS WITH SOURCES 79
d∗F =
1 1
(∇ × B)x − ∂0 Ex dx0 ∧ dx2 ∧ dx3 − (∇ × B)y − ∂0 Ey dx0 ∧ dx1 ∧ dx3
c c
1 1
+ (∇ × B)z − ∂0 Ez dx0 ∧ dx1 ∧ dx2 − ∇ · E dx1 ∧ dx2 ∧ dx3 .
c c
Define the four-current
J = Jµ dxµ ,
where J0 = −cρ and J1 = jx , J2 = jy , J3 = jz . Using
∗d ∗ F = µ0 J.
d ∗ F = µ0 ∗ J,
be the total charge at time t. Then it follows from Stokes’s theorem for M =
[ct1 , ct2 ] × R3 that
Z Z
0= d∗J = ∗J = Q(t2 ) − Q(t1 ).
M ∂M
F µν = η µα η νβ Fαβ ,
which has the the same form as Fµν , where E is replaced by −E. It is related
to the dual strength field tensor by
1
(∗F )µν = εµναβ F αβ .
2
Then the second pair of Maxwell equations can be written in the following form
(8.12) ∂µ F µν = J ν , ν = 0, 1, 2, 3,
(8.13) dF = 0 and ∗ d ∗ F = J,
where the 4-current J satisfies the continuity equation. By Poincaré lemma, the
first equation has a solution
F = dA where A = Aµ dxµ .
where F = dA.
Proposition 8.1. The critical points of the action functional S(A) are given
by the Maxwell equations.
Proof. For given a ∈ A put
d
δS(A) = S(A + εa).
dε ε=0
where
1 c 1 2
(8.16) L (A) = − Fµν F µν = E − B2
16π 8π c2
P × G 3 (p, g) 7→ p · g ∈ P,
which preserves the Sfibers and is free and transitive. By definition, there is an
open covering M = α∈A Uα such that over each Uα there is a local trivializa-
tion, a diffeomorphism
ϕα : π −1 (Uα ) → Uα × G
such that
π(ϕ−1
α (x, g)) = x and ϕ−1 −1
α (x, g) = ϕα (x, e) · g for all x ∈ Uα , g ∈ G,
and
83
84 9. ELECTRODYNAMICS AS U(1) GAUGE THEORY
where (x, g) ∼ (y, h) if and only if x = y ∈ Uαβ and g = tαβ (x)h. Transition
functions tαβ and fα−1 tαβ fβ , where fα : U → G, are arbitrary smooth functions,
define the same bundle P . Sections of P over U ⊆ M are the maps S : U → P
satisfying π ◦ S = id|U . They are determined by the maps S α : Uα → G,
satisfying
Elements of the gauge group G(P ) are collections {fα }α∈A of arbitrary smooth
functions fα : Uα → G that map sections to sections by the formula S 0 = S ◦ f .
Explicitly,
0
S α = S α fα : Uα → G,
0
and Sα satisfy (9.4) with the transition functions
(9.5) E = (P × V )/G,
and
gαβ (d + Aβ )sβ = gαβ dsβ + gαβ Aβ sβ
are equal for all sβ |Uαβ , which is equation (9.8). Notation ∇A = d + A we will
used sometimes.
In local coordinates x1 , . . . , xn on a chart U ⊆ M ,
1Here we use a definition that does not use a notion of a connection on a principal
G-bundle.
2Such connections are obtained from connections on a principal G-bundle P .
86 9. ELECTRODYNAMICS AS U(1) GAUGE THEORY
∇2 : Ω0 (E) → Ω2 (E)
we obtain
∇2 (f ζ) = ∇(df ⊗ ζ + f ∇ζ)
= −df ∧ ∇ζ + df ∧ ∇ζ + f ∇2 ζ = f ∇2 ζ.
∇2 sα = Fα sα on Uα .
∇2 sα = (d + Aα )(dsα + Aα sα )
= dAα sα − Aα ∧ dsα + Aα ∧ dsα + Aα ∧ Aα sα
= (dAα + Aα ∧ Aα )sα ,
where Aα ∧Aα is understood as a product in End V together with the usual exte-
rior multiplication. Thus End E-valued 2-form F on M is a collection {Fα }α∈A
of End V -valued 2-forms on Uα ,
(9.12) Fα = dAα + Aα ∧ Aα ,
satisfying
−1
(9.13) Fα = gαβ Fβ gαβ on Uαβ .
9.1. BUNDLES, CONNECTIONS AND CURVATURE 87
Transformation law (9.13) follows from (9.8)–(9.12). We will often use notation
F = F (A) = dA + A ∧ A.
1 ∂Aν ∂Aµ
F = Fµν dxµ ∧ dxν , where Fµν = [∇µ , ∇ν ] = − + [Aµ , Aν ]
2 ∂xµ ∂xν
and [Aµ , Aν ] = Aµ ∧ Aν − Aν ∧ Aµ . It follows from (9.11) that curvature satisfies
the Bianchi identity,
(9.15) ∇EndE
A (F ) = dF + A ∧ F − F ∧ A = 0,
which we will simply write as ∇A F = 0. It can also be obtained from the Jacobi
identity
[∇µ , ∇ν ], ∇σ ] + [∇ν , ∇σ ], ∇µ ] + [∇σ , ∇µ ], ∇ν ] = 0.
Remark. In general, for B ∈ Ωk (M, End E) we have
∇EndE
A (B) = dB + A ∧ B − (−1)k B ∧ A.
dΦ(F ) = 0.
2. Cohomology class
[Φ(F )] ∈ H 2k (M )
3. A map
Φ 7→ Φ(F )
is a homomorphism of a commutative algebra of invariant polynomials
on End V into the commutative algebra H even (M ) of differential forms
of even degree on M .
The map Φ 7→ Φ(F ) is called Weil homomorphism, and cohomology classes
[Φ(F )] — characteristic classes of a bundle E, associated with the invariant
polynomial Φ. Let P i be elementary invariant polynomials of degree i =
1, . . . , n, defined by
Xn
det(B + tI) = P n−k (B)tk .
k=0
√
−1
Forms ci (F ) = P i F are called Chern forms, and corresponding co-
2π
homology classes — Chern classes. It is a fundamental fact in the theory of
characteristic classes, that
√
i −1
ci (E) = P F ∈ Ȟ 2i (M, Z), i = 1, . . . , n,
2π
where Ȟ 2i (M, Z) stands for the Čech cohomology with coefficients in the con-
stant sheaf Z.
(9.16) ∇ = d + Aα ,
where
√ Aα ∈ Ω1 (Uα ) are 1-forms on Uα with values in the Lie algebra u(1) '
−1 R of the Lie group U(1), satisfying
−1
Aα = Aβ − gαβ dgαβ on Uα ∩ Uβ .
F = dA
(9.18) dF = 0 and d ∗ F = 0.
for a complex U(1)-line bundle L there is a real rank 2 vector bundle L over M
with the symmetry group SO(2). Its first Pontryagin class p1 (L) ∈ Ȟ 4 (M, Z) is
given by
p1 (L) = −c2 (L ⊗ C),
where L ⊗ C is a complexification of the real bundle L — rank 2 complex vector
bundle over M , and c2 (L ⊗ C) is its second Chern class. It is easy to see
that L ⊗ C ' L ⊕ L̄, where L̄ is the line bundle with the transition functions
3Note that if A = A dxµ is a real-valued 1-form on M = R4 , used in Sect 8.2 in Lecture
µ √
8 in case L is a trivial line bundle, then in (9.16) we have ∇ = d + −1 A.
90 9. ELECTRODYNAMICS AS U(1) GAUGE THEORY
1
ḡαβ , so that c2 (L ⊗ C) is represented by the differential form F ∧ F . The
4π 2
corresponding first Pontryagin number is
Z
1
p1 = − 2 F ∧ F ∈ Z.
4π M
In case when M is a Riemannian manifold with the metric ds2 , then the
Maxwell’s equations on M have the form
dF = 0 and d ∗ F = 0,
√
where F ∈ Ω2 (M, −1 R) and ∗ is the Hodge star of the metric ds2 . They
characterize curvature forms F as harmonic 2-forms. Since in the Riemannian
case ∗2 = 1 on 2-forms, and we have a decomposition
√ √ √
(9.19) Ω2 (M, −1 R) = Ω2+ (M, −1 R) ⊕ Ω2− (M, −1 R)
according to the eigenspaces of the Hodge ∗-operator corresponding to the eigen-
values 1 and√−1. The 2-form F on √ M is called self-dual or anti-self-dual, if
F ∈ Ω2+ (M, −1 R) or F ∈ Ω2− (M, −1 R) respectively,
∗F = ±F.
Correspondingly, connection ∇ = d + A on a U(1)-line bundle L is called self-
dual or anti-self-dual, if its curvature 2-form F = dA is, respectively, self-dual or
anti-self-dual. Curvature forms of self-dual or anti-self-dual connections satisfy
Maxwell’s equations on a Riemannian 4-manifold M automatically!
From the inequality Z
− ω ∧ ∗ω ≥ 0
M
√
for all ω ∈ Ω2 (M, −1 R), we get for a curvature 2-form F of a line bundle
L → M,
Z Z
− F ∧ ∗F − 4π 2 p1 = − F ∧ ∗F + F ∧ F
M M
Z
1
=− (F − ∗F ) ∧ ∗(F − ∗F ) ≥ 0
2 M
and
Z Z
− F ∧ ∗F + 4π 2 p1 = − F ∧ ∗F − F ∧ F
M M
Z
1
=− (F + ∗F ) ∧ ∗(F + ∗F ) ≥ 0.
2 M
Yang-Mills theory
tr(ω1 ∧ ω2 ) = tr(ζ1 ζ2 ) ψ1 ∧ ψ2 .
by
?(ψ ⊗ ζ) = ∗ψ ⊗ ζ, ψ ∈ Ωp (M ), ζ ∈ Ω0 (M, End E).
Denote by AE the affine space of connections on E.
Definition. A Yang-Mills action functional S : AE → C is given by
Z
1
(10.2) S(A) = − tr(F ∧ ? F ), F = dA + A ∧ A, A ∈ AE .
4π M
If manifold M is non-compact, we assume that connections A are such that
the integral in (10.2) is convergent (e.g., F has compact support). It follows
from (9.14) that the functional S is invariant under the action of a gauge group
G with the symmetry group GL(r, C).
93
94 10. YANG-MILLS THEORY
(10.3) ∇A F = 0 and ∇A ? F = 0.
F (A + a) = F (A) + da + A ∧ a + a ∧ A + a ∧ a
= F (A) + ∇A a + a ∧ a.
Whence using the cyclic property of the trace, formula (8.15), Leibniz rule
d(a ∧ ? F ) = da ∧ ? F − a ∧ d ? F
d
δS(A) = S(A + εa)
dε ε=0
Z
1
=− tr ((da + A ∧ a + a ∧ A) ∧ ? F + F ∧ ?(da + A ∧ a + a ∧ A))
4π M
Z
1
=− tr ((da + A ∧ a + a ∧ A) ∧ ? F )
2π M
Z
1
=− tr (a ∧ (d ? F + A ∧ ? F − ? F ∧ A))
2π M
Z
1
=− tr(a ∧ ∇A ? F ).
2π M
Suppose that the vector bundle E is associated with a principal G-bundle
P over M through a representation R : G → GL(V ) of a compact Lie group
G. When representation R is unitary with respect to Hermitian inner product
in V , restriction of the Yang-Mills functional to the connections AEG with the
symmetry group G gives a functional taking non-negative values. Indeed, in
this case ρ(g) consists of skew-Hermitian endomorphisms V , and
− tr B 2 ≥ 0 for B = −B ∗ ∈ End V,
where adx ∈ End g is given by the adjoint action, adx (y) = [x, y]. The Killing
form defines positive-definite inner product if and only if a Lie group G is
10.1. YANG-MILLS EQUATIONS 95
are totally anti-symmetric. In case g = su(2) such basis in the defining two-
dimensional representation is given by the matrices
1 0 1 1 0 −i 1 1 0
x1 = , x2 = , x3 = ,
2i 1 0 2i i 0 2i 0 −1
and
hx, yi = −4 trC2 (xy).
The real vector bundle associated with a principal G-bundle P through the
adjoint representation of a Lie group G on its Lie algebra g is called an adjoint
bundle and is denoted by ad P . In case when (M, ds2 ) is a Riemannian manifold,
the Killing form defines on Ωp (M, ad P ) an inner product
Z
(10.5) (ω1 , ω2 ) = hω1 ∧ ? ω2 i
M
The symmetry
follows from (8.15) in Lecture 8 and the cyclic property of the trace. The Yang-
Mills functional is the L2 -norm of the curvature form F (A) ∈ Ω2 (M, ad P ),
1
S(A) = kF (A)k2 .
4π
96 10. YANG-MILLS THEORY
1 ∂Aν ∂Aµ
(10.7) F = Fµν dxµ ∧ dxν , Fµν = µ
− + [Aµ Aν ],
2 ∂x ∂xν
where1 Aµ = Aaµ Xa , Fµν = Fµν
a
Xa ∈ g and generators Xa satisfy (10.4). Corre-
sponding Yang-Mills functional (10.2) takes the form
Z Z
1 1
(10.8) S(A) = hFµν , F µν id4 x = − F a (F a )µν d4 x,
16π R4 8π R4 µν
∂F µν
(10.9) ∇µ F µν = + [Aµ , F µν ] = 0.
∂xµ
Yang-Mills equations (10.7) and (10.9) generalize U(1)-invariant Maxwell equa-
tions to the case of non-abelian symmetry group G. In terms of the components
equation (10.9) takes the form
∂(F a )µν
+ tabc Abµ (F c )µν = 0,
∂xµ
where tabc are totally anti-symmetric structure constants of g.
Remark. In physics one uses ∇µ = ∂µ − gAµ for the covariant derivative,
where g is a coupling constant of the theory. In Quantum Chromodynamics
(QCD) on considers G = SU(3) in the adjoint representation, and corresponding
components Aaµ (x), a = 1, . . . , 8, are the gluon fields; corresponding quark fields
are in the fundamental representation of SU(3). In our notation gluon part of
the QCD Lagrangian is
1 a
(10.10) L (A) = − F (F a )µν ,
4g 2 µν
a
where Fµν plays the role of gluon field strength tensor. Corresponding elemen-
tary particle — a gluon (or gauge boson) — is the exchange particle for the
strong force between quarks. In the Standard Model of electroweak and strong
interactions one uses the symmetry group G = SU(3) × SU(2) × U(1).
Recall that the first Pontryagin class of a real vector bundle ad P over M is
defined by
p1 (ad P ) = −c2 (adC P ) ∈ Ȟ 4 (M, Z),
where adC P = ad P ⊗R C is a complex vector bundle. Since c1 (adC P ) = 0, it is
1
easy to see that c2 (adC P ) is represented by the differential form tr(F ∧ F ),
8π 2
and the corresponding first Pontryagin number is
Z
1
p1 = − 2 tr(F ∧ F ) ∈ Z.
8π M
In case when M is a Riemannian four-manifold with the metric ds2 , we have
a decomposition
(10.11) Ω2 (M, ad P ) = Ω2+ (M, ad P ) ⊕ Ω2− (M, ad P )
according to the eigenspaces of the Hodge star operator ? corresponding to
the eigenvalues 1 and −1. Since operator ? is symmetric with respect to inner
product (10.5) in Ω2 (M, ad P ), these subspaces are orthogonal. Equivalently,
for F = F+ + F− , where F± ∈ Ω2± (M, ad P ),
Z Z
(F+ , F− ) = − hF+ ∧ F− i = − hF− ∧ ?F+ i = −(F− , F+ ),
M M
? F = ±F.
Correspondingly, connection ∇ = d + A on a real vector bundle ad P is called
self-dual or anti-self-dual, if its curvature is self-dual or anti-self-dual. Curvature
forms of self-dual or anti-self-dual connections satisfy Yang-Mills equations on
a Riemannian four-manifold M automatically!
Using the orthogonality of F+ and F− , we obtain
1 1
kF (A)k2 = kF+ k2 + kF− k2
S(A) =
4π 4π
and
Z
1
p1 = − tr((F+ + F− ) ∧ (F+ + F− ))
8π 2 M
1
= − 2 (F+ + F− , F+ − F− )
8π
1
= − 2 k(F+ k2 − kF− k2 .
8π
From here we obtain the inequalities
1 1
S(A) − 2πp1 ≥ kF+ k2 and S(A) + 2πp1 ≥ kF− k2 .
2π 2π
98 10. YANG-MILLS THEORY
Thus we see that the absolute minima of the Yang-Mills action on Aad G
P are
given by the self-dual connections in case p1 > 0, by the anti-self-dual connec-
tions in case p1 < 0 and by both these types in case p1 = 0. Number p1 in
called the instanton number. Solutions of the self-dual Yang-Mills equations for
M = S 4 and G = SU(2) in case p1 = k > 0 form he instanton moduli space
Mk , a smooth manifold of dimension 8k − 3.
Here
[Φ, Φ∗ ] = Φ ∧ Φ∗ + Φ∗ ∧ Φ
is a graded Lie bracket on adC P -valued 1-forms, and ∂¯A is a (0, 1)-component
of
∇A = ∂ + A1,0 dz + ∂¯ + A0,1 dz̄ = ∂A + ∂¯A .
It is remarkable that equations (10.13)–(10.14) make sense over a Riemann
surface M ! Namely, consider a principal G-bundle P over M , a connection A
in the adjoint bundle ad P and the Higgs field Φ ∈ Ω1,0 (M, adC P ). The pair
(A, Φ) satisfies self-duality equations over a Riemann surface M , if
∂Tαβ
(11.1) = 0, α = 0, 1, 2, 3.
∂xβ
The tensor Tαβ is traceless Tαα = 0 and symmetric, T αβ = T βα , where
1
(11.2) T αβ = η αγ Tγβ = −ηµν F αµ F βν + η αβ Fµν F µν .
4
The tensor T αβ is called the energy-momentum tensor. Its components contain
the energy density
00 1 1 2 2
T = E +B
2 c2
and the momentum density
1
T 0i = F 0k F ik = (E × B)i , i = 1, 2, 3.
c
The vector S = E × B is called the Poynting vector.
Remark. The conservation law (11.1)
∂T 00
= −∇ · S
∂t
101
102 11. ELECTROMAGNETIC WAVES IN A FREE SPACE
can be verified directly using Maxwell’s equations and the calculus formula
∇ · (a × b) = b · (∇ × a) − a · (∇ × b).
It also implies that implies that the total energy of the electromagnetic field
Z
1
E = T 00 d3 r
4π {ct}×R3
(11.3) dF = 0 and d ∗ F = 0
(11.4) ∗ d ∗ dA = 0.
(11.5) d ∗ A = 0,
A = 0,
where
=d∗d∗+∗d∗d
is the D’Alambertian — the Laplace operator of the Minkowski metric on R4 ,
acting on 1-forms. In terms of A = Aµ dxµ equation (11.5) becomes
∂
(11.6) ∂µ Aµ = 0, where ∂µ = ,
∂xµ
and is called the Lorenz1 gauge condition. Since equation (11.4) can be written
as
∂µ F µν = 0, where F µν = ∂ µ Aν − ∂ ν Aµ
(see (8.12) in Lecture 8), we readily obtain that in the Lorenz gauge
Aµ = 0, µ = 0, 1, 2, 3,
1Named after Danish physicist and mathematician Ludvig Lorenz, not to be confused
with Dutch physicist Hedrick Lorentz!
11.2. GAUGE FIXING 103
where
1 ∂2 ∂2 ∂2 ∂2
= ∂µ ∂ µ = 2 2
− 2
− 2 − 2.
c ∂t ∂x ∂y ∂z
For every A ∈ Ω1 (R4 ) there is a gauge equivalent 1-form Af = A + df
satisfying the Lorentz condition. Indeed, (11.5) gives
∗d ∗ df = − ∗ d ∗ A.
f = −∂µ Aµ .
Remark. Maxwell equations with sources in the Lorentz gauge have the
form
Aµ = J µ .
The Lorenz gauge is not unique: if A satisfies (11.6), so does Af , where
f = 0. In free and empty space one can make a unique choice by imposing
A0 = 0. In general, this gauge condition is called Hamilton gauge. Together
with Lorenz gauge it yields the Coulomb gauge,
(11.7) ∇ · A = 0.
(11.9) E = 0 and B = 0.
104 11. ELECTROMAGNETIC WAVES IN A FREE SPACE
where
E0 = iωA0 and B0 = ik × A0 .
Consider the vector E0 ∈ C and put b = E0 eiα , where E02 = E0 · E0 =
3
The electromagnetic waves describe photons, particles with 4-wave vector satis-
fying k02 = k2 .
where Cauchy data A0 (r) and A1 (r) satisfy Coulomb gauge condition
∇ · A0 = 0 and ∇ · A1 = 0
and rapidly decay as |r| → ∞.
Cauchy problem for the wave equation in R4 is solved by the Fourier trans-
form. Namely, let
Z
1
A0 (r) = 3 eik·r a0 (k)d3 k,
(2π) 2 R3
Z
1
A1 (r) = 3 eik·r a1 (k)d3 k,
(2π) 2 R3
where a0 (k) = ā0 (−k), a1 (k) = ā1 (−k) and k · a0 (k) = k · a1 (k) = 0. The
solution is given by
Z
1
(11.10) A(t, r) = 3 eik·r a(t, k)d3 k,
(2π) 2 R3
where
sin(c|k|t)
a(t, k) = cos(c|k|t)a0 (k) + a1 (k).
c|k|
Introducing
1 1
a(k) = a0 (k) + a1 (k),
2 2ic|k|
we can rewrite (11.10) as
Z
1
(11.11) A(t, r) = e−i(ωk t−k·r) a(k) + ei(ωk t−k·r) ā(k) d3 k,
(2π)3 R3
and
B =∇×A
Z
i
−i(ωk t−k·r) i(ωk t−k·r)
= 3 k × e a(k) − e ā(k) d3 k.
(2π) 2 R3
By Plancherel theorem we have for total energy of the electromagnetic field,
Z
1 1 2
E = E + B d3 r
2
8π R3 c2
Z
1
= (ω 2 a(k)ā(k) + (k × a(k)) · (k × ā(k))d3 k
4π R3 k
Z
1
= ω 2 a(k) · ā(k)d3 k,
2πc2 R3 k
11.4. THE GENERAL SOLUTION 107
where we have used the identity (k × a(k)) · (k × ā(k)) = |k|2 a(k) · ā(k), which
follows from k · a(k) = 0.
Similarly,
Z Z
1 3 1
Sd r = (E × B)d3 r
4π R3 4πc R3
Z
1
= ωk a(k) × (k × ā(k))d3 k
2πc R3
Z
1
= ωk (a(k) · ā(k))kd3 k.
2π R3
Finally, putting
ωk i
P (k) = √ (a(k) + ā(k)) Q(k) = √ (a(k) − ā(k)),
2c π 2c π
k · P (k) = k · Q(k) = 0.
LECTURE 12
where integration goes over the part of R4 between the slices t = t0 and t = t1
with fixed ϕ(t0 , x) = ϕ0 (x) and ϕ(t1 , x) = ϕ1 (x), or over R4 , where ϕ(x) is
assumed to be rapidly decaying as |x| → ∞. Corresponding Euler-Lagrange
equation δS = 0 takes the form
∂L ∂ ∂L
(12.1) − =0
∂ϕ ∂xµ ∂(∂µ ϕ)
and yields equation of motion of the massive real scalar field
0
(12.2) ( + m2 )ϕ + Vint (ϕ) = 0.
For the ϕ4 –model this equation takes the form
ϕ3
( + m2 )ϕ + g = 0,
3!
and is a nonlinear Klein-Gordon equation with cubic nonlinearity.
Remark. Let F be the space of scalar fields on R4 . The Lagrangian L is
map from F to the functions on R4 such that L(ϕ)(x) depends only on the 1-jet
of ϕ at x ∈ R4 , i.e., L(ϕ)(x) = L (ϕ(x), ∂µ ϕ(x)).
109
110 12. HAMILTONIAN FORMALISM. REAL SCALAR FIELD
∂L ∂L
∂ν L = ∂ν ϕ + ∂ν ∂µ ϕ
∂ϕ ∂(∂µ ϕ)
∂L ∂L ∂L
= − ∂µ ∂ν ϕ + ∂µ ∂ν ϕ .
∂ϕ ∂(∂µ ϕ) ∂(∂µ ϕ)
(12.3) ∂µ Tνµ = 0,
where
∂L
Tνµ = ∂ν ϕ − δνµ L
∂(∂µ ϕ)
is the energy-momentum tensor. The tensor T µν = η νλ Tλµ satisfies the conser-
vation law
∂µ T µν = 0,
and is defined up to the addition of ∂σ Ψµνσ , where Ψµνσ = −Ψµσν .
For the scalar field the tensor T µν = ∂ µ ϕ∂ ν ϕ − η µν L is symmetric and
1
T 00 = (∂0 ϕ)2 + (∇ϕ)2 + m2 ϕ2 + Vint (ϕ) ,
2
T 0k = ∂ 0 ϕ∂ k ϕ, T ij = ∂ i ϕ∂ j ϕ.
Conservation law for the energy-momentum vector (h, p), where h = T 00 and
p = (T 01 , T 02 , T 03 ) reads
∂h
+ ∇ · p = 0.
∂t
For the electromagnetic field L = − 16π
1
Fµν F µν , and the tensor
∂L
∂ ν Aσ − η µν L
∂(∂µ Aσ )
∂L
π(x) = = ∂0 ϕ(x)
∂(∂0 ϕ(x))
be canonically conjugated momentum to the field ϕ(x), and define the Hamil-
tonian functional density H (π, ϕ) by the Legendre transform
Remark. The Schwartz space S (R3 ) is a Fréchet space with the topology
defined by the system of the semi-norms
and
∞ X ∞ Z Z
δF X 1
= ··· cmn (x1 , . . . , xm ; x, y2 , . . . , yn )×
δϕ(x) m=0 n=1 m!(n − 1)! R3 R3
and
cmn (π1 ⊗ · · · ⊗ πm ⊗ ϕ2 ⊗ · · · ⊗ ϕn ) ∈ S (R3 )0
are represented by the Schwarz class functions.
Definition. The algebra A of classical observables on M is the algebra of
all admissible functionals on M .
The following result provides a rigorous foundation for the Hamiltonian me-
chanics with the infinite-dimensional phase space M .
Lemma 12.1. The symplectic form Ω endows A with the Poisson algebra
structure given by the Poisson bracket
Z
δF δG δF δG
(12.6) {F, G}(π, ϕ) = − d3 x,
R3 δπ(x) δϕ(x) δϕ(x) δπ(x)
where variational derivatives are evaluated at (π, ϕ) ∈ M .
Proof. It follows from the definition of real-analytic functionals and the
above remark that {F, G} ∈ A for F, G ∈ A . As in case of the canonical
Poisson bracket on R2n (see Sect. 5.3 in Lecture 5) the Jacobi identity for the
bracket given by (12.6) is proved by a direct computation.
The Darboux coordinates π(x), ϕ(x), considered as evaluation functionals of
(π, ϕ) at x ∈ R3 , do not belong to A . Nevertheless, we have in the distributional
sense,
δπ(x) δπ(x) δϕ(x) δϕ(x)
= δ(x − y), =0 and = 0, = δ(x − y),
δπ(y) δϕ(y) δπ(y) δϕ(y)
and it follows from (12.6) that
δF δF
{F, π(x)} = − and {F, ϕ(x)} = .
δϕ(x) δπ(x)
114 12. HAMILTONIAN FORMALISM. REAL SCALAR FIELD
Since for F ∈ A
Z
δF (π, ϕ) δF (π, ϕ)
∂0 F (π, ϕ) = ∂0 π(t, x) + ∂0 ϕ(t, x) d3 x,
R3 δπ(x) δϕ(x)
∂0 F = {H, F }
(12.7) {π(x), π(y)} = {ϕ(x), ϕ(y)} = 0 and {π(x), ϕ(y)} = δ(x − y),
(12.8) ( + m2 )ϕ(x) = 0
Here
ρ(k) = θ(k 0 )ρ1 (k) + θ(−k 0 )ρ2 (k),
where θ(k 0 ) is the Heavyside function and ρ1 , ρ2 are distributions supported
on R3√
. By definition of the distribution δ(k 2 − m2 ) = δ((k 0 )2 − ωk2 ), where
ωk = k2 + m2 > 0, for a test function u(k) ∈ S (R4 ) we have
where
u(ωk , k) u(−ωk , k)
u1 (k) = , u2 (k) = ∈ S (R3 ).
2ωk 2ωk
12.4. FOURIER MODES FOR THE KLEIN-GORDON MODEL 115
Whence
1 1
ϕ̂(k) = ρ1 (k)δ(k 0 − ωk ) + ρ2 (k)δ(k 0 + ωk ),
2ωk 2ωk
where reality condition ρ(k) = ρ(−k) gives ρ2 (k) = ρ1 (−k).
Substituting this ϕ̂(k) into the inverse Fourier transform
Z
1
ϕ(x) = e−ik·x ϕ̂(k)d4 k,
(2π)2 R4
√
introducing a(k) = 2πρ1 (k), ā(k) = a(k) and changing in the second integral
k by −k we obtain
d3 k
Z
1
(12.9) ϕ(x) = 3 a(k)e−ik·x + ā(k)eik·x , where k 0 = ωk .
(2π) 2 R3 2ωk
From this general distributional solution we can obtain a solution of the Cauchy
problem for the Klein-Gordon equation, which consists in finding a solution ϕ(x)
of (12.8) satisfying
Namely, from
d3 k
Z Z
1 1
ϕ(x) = 3 ϕ̂(k)eikx d3 k = 3 a(k)eikx + ā(k)e−ikx ,
(2π) R3
2 (2π) R32 2ωk
−i d3 k
Z Z
1
π(x) = 3 π̂(k)eikx d3 k = 3 ωk a(k)eikx − ā(k)e−ikx
(2π) 2 R3 (2π) 2 R3 2ωk
we get
a(k) = ωk ϕ̂(k) + iπ̂(k) ∈ S (R3 ),
so that (12.9) gives classical solution of the Cauchy problem.
It follows from Poisson brackets (12.7) that in the distributional sense
and
Z Z
1
{π̂(k), ϕ̂(l)} = {π(x), ϕ(y)}e−i(kx+ly) d3 xd3 y
(2π)3 R3 R3
Z
1
= e−i(k+l)x d3 x = δ(k + l),
(2π)3 R3
Z Z
1
{π̂(k), ϕ̂(l)} = {π(x), ϕ(y)}e−i(kx−ly) d3 xd3 y
(2π)3 R3 R3
Z
1
= e−i(k−l)x d3 x = δ(k − l).
(2π)3 R3
116 12. HAMILTONIAN FORMALISM. REAL SCALAR FIELD
Thus we obtain
(12.10) {a(k), a(l)} = {ā(k), ā(l)} = 0 and {a(k), ā(l)} = 2iωk δ(k − l).
d3 k
Z
= ā(k)a(k)k .
R3 2ωk
Thus we see that in terms of Fourier modes Hamilton’s equations (12.4)–
(12.5) decouple
117
118 13. HAMILTONIAN FORMALISM. GAUGE THEORIES.
so that the pairs (Ei (x), Ai (x)), are Darboux coordinates on M with the canon-
ical Poisson brackets
C(x) = ∇ · E(x) = 0, x ∈ R3 .
{C(x), C(y)} = 0 x, y ∈ R3
and also
∂2
{C(x), D(y)} = δ(x − y),
∂xi ∂y i
which is the integral kernel of the operator −∆, Laplace operator of the Eu-
clidean metric on R3 . Thus the reduced phase space M0 of classical electrody-
namics is a linear subspace in M defined by
Since
{D(x), D(y)} = 0,
Darboux coordinates for the symplectic form Ω0 = Ω|M0 can be found by the
general procedure described in Sect. 7.3 in Lecture 7.
Using Corollary 7.1 in Lecture 7, the Poisson bracket { , }0 on M0 , associated
with the symplectic form Ω0 , can be written as a restriction of the Dirac bracket
on M , associated with the second class constraints (C(x), D(x)). Namely, it
follows from (7.20) in Lecture 7 that
Z Z
{F, G}DB = {F, G} + {F, C(x)}G(y − x){D(y), G}−
R3 R3
(13.8) − {F, D(x)}G(x − y){C(y), G} d3 xd3 y,
or
eikx 3
Z
1
G(x) = d k.
(2π)3 R3 k2
Using (13.5) we readily compute that
and
⊥
(13.10) {Ei (x), Aj (y)}DB = 4πδij (x − y), x, y ∈ R3 ,
⊥
where the distribution δij (x) is the transverse δ-function,
Z
⊥ 1 ki kj
(13.11) δij (x) = δij − 2 eikx d3 k, i, j = 1, 2, 3.
(2π)3 R3 k
It satisfies
⊥
∂i δij (x) = 0, j = 1, 2, 3.
Thus Dirac bracket (13.8) yields a ‘transverse’ Poisson structure { , } on
M , determined by (13.9)–(13.10). It is degenerate and its center is generated
by C(x) and D(x) for x ∈ R3 . The Dirac bracket { , }DB restricts to M0 and
yields a non-degenerate Poisson bracket { , }0 associated with the symplectic
form Ω0 . Since Z
⊥
δij (x − y)fj (y)d3 y = fi (x)
R3
yield
∂E ∂A
= ∇ × B, where B = ∇ × A and = −E.
∂t ∂t
Together with the Gauss law, they give the full set of Maxwell equations in the
Coulomb gauge.
In terms of the normal modes P (k) and Q(k) (see Sect. 11.4 in Lecture 11),
satisfying
k · P (k) = k · Q(k) = 0,
the Poisson structure { , }0 is given by the transverse Poisson brackets
ki lj
{Pi (k), Qj (l)}0 = δij − δ(k − l).
k·l
tr(Xa Xb ) = −2δab , a, b = 1, . . . , n.
1
F = Fµν dxµ ∧ dxν , where Fµν = ∂µ Aν − ∂ν Aµ + [Aµ , Aν ],
2
be its curvature. Consider the Yang-Mills Lagrangian function (see Lecture 10),
1 1 a
L (A) = tr Fµν F µν = − Fµν (F a )µν ,
8 4
a
where Fµν = Fµν Xa , and we put g = 1 in formula (10.10). As in case of classical
electrodynamics, it can be written in the first order formalism
1 1
(13.12) L = tr ∂µ Aν − ∂ν Aµ + [Aµ , Aν ] − Fµν F µν .
4 2
Put
Ei = F0i and Bi = εijk F jk , i =, 1, 2, 3.
Using equations Fij = ∂i Aj − ∂j Ai + [Ai , Aj ] and the cyclic property of the
trace, we can rewrite (13.12) as follows
1 1 1
L = − tr ∂0 Ak − ∂k A0 + [A0 , Ak ] − Ek Ek + tr Fij F ij
2 2 8
1 1 2 1
= − tr Ek ∂0 Ak − (Ek + Bk ) + A0 (∂k Ek + [Ak , Ek ]) + ∂k (tr A0 Ek ) .
2
2 2 2
so that the pairs (Eka (x), Aak (x)), are Darboux coordinates on M with the canon-
ical Poisson brackets
(13.16) {Eka (x), Abl (y)} = δkl δ ab δ(x−y), i, j = 1, 2, 3 and a, b = 1, . . . , n.
The second term in (13.14) is the Hamiltonian of the Yang-Mills field,
Z Z
1
H (x)d x =
3
(Eka )2 (x) + (Bka )2 (x) d3 x.
(13.17) H=
R3 2 R 3
Comparing the last term in (13.3) with the corresponding term in (7.7) we
conclude that components Aa0 (x) of the gauge field are the Lagrange multipliers,
and the constraints C a (x) are given by the nonabelian Gauss law,
C a (x) = ∂k Eka (x) + tabc Abk (x)Ekc (x) = 0, x ∈ R3 .
As in case of classical electrodynamics, these are the first class constraints,
and we verify it by the following computation. Namely, it directly follows from
(13.16) that
{Eka (x), C b (y)} = −tab c
c Ek (x)δ(x − y)
and
{Aak (x), C b (y)} = {Aak (x), ∂l Elb (y) + tbcd Acl (y)Eld (y)}
∂
= −δ ab δ(x − y) − tbca Ack (x)δ(x − y)
∂yk
∂
= − δ ab + tab
c A c
(y) δ(x − y).
∂yk
Introducing
Z Z
1
C(f ) = − tr C(x)f (x)d3 x = C a (x)f a (x)d3 x
2 R3 R3
for a test function f : R3 → ad g, where f (x) = f a (x)Xa and f a (x) ∈ S (R3 , R),
we can succinctly rewrite these formulas as
def
(13.18) {Ek (x), C(f )} = {Eka (x), C(f )}Xa = [Ek (x), f (x)],
def
(13.19) {Ak (x), C(f )} = {Aak (x), C(f )}Xa = (∇k f )(x).
13.2. YANG-MILLS EQUATIONS 123
and
{[Ak (x), Ek (x)], C(g)} = Ak (x){Ek (x), C(g)} + {Ak (x), C(g)}Ek (x)
− Ek (x){Ak (x), C(g)} − {Ek (x), C(g)}Ak (x)
= [Ak (x), [Ek (x), g(x)]] + [(∇k g)(x), Ek (x)].
Whence
{C(x), C(g)} = [∂k Ek (x), g(x)] + [Ak (x), [Ek (x), g(x)]] + [[Ak (x), g(x)], Ek (x)]
= [∂k Ek (x), g(x)] + [[Ak (x), Ek (x)], g(x)]
= [C(x), g(x)]
and
Z
1
{C(f ), C(g)} = − tr ([C(x), g(x)]f (x)) d3 x
2 R3
Z
1
= tr (C(x)[f (x), g(x)]) d3 x.
2 R3
{Fij (x), C(f )} = {∂i Aj (x) − ∂j Ai (x) + [Ai (x), Aj (x)], C(f )}
= ∂i (∇j f )(x) − ∂j (∇i f )(x) + Ai (x)(∇j f )(x) + (∇i f )(x)Aj (x)
− (∇j f )(x)Ai (x) − Aj (x)(∇i f )(x)
= [∂i Aj (x) − ∂j Ai (x), f (x)] + [Ai (x), [Ak (x), f (x)] + [[Ai (x), f (x)], Ak (x)]
= [Fij (x), f (x)],
124 13. HAMILTONIAN FORMALISM. GAUGE THEORIES.
so that
{Bk2 (x), f (x)} = [B 2 (x), f (x)].
Whence
{Ek2 (x) + Bk2 , C(f )} = [Ek2 (x) + Bk2 , f (x)]
1
and for H (x) = − tr Ek2 (x) + Bk2 (x) we obtain
4
{H (x), C(f )} = 0.
Therefore
{H, C(f )} = 0
or equivalently,
{H, C a (x)} = 0.
This finishes the proof that Yang-Mills theory is a Hamiltonian theory with
first class constraints. As in the U(1)-case, for additional constraints one can
use non-abelian Coulomb gauge
D(x) = ∂k Ak (x) = 0.
∂2 ∂
(13.21) {C a (x), Db (y)} = δ ab k k
δ(x − y) + tab c
c Ak (x) δ(x − y).
∂x ∂y ∂y k
(13.22) M = −∆ + ad Ak (x)∂k ,
Classic text
• L.D. Landau and E.M. Lifschitz, The Classical Theory of Fields, 4th
edition, Butterworth-Heinemann, 1980
is the basic reference for classical electrodynamics, special relativity and theory
of gravity. For more detailed exposition, including applications, see
• D.J. Griffiths, Introduction to Electrodynamics, Prentice-Hall, NJ, 1999
• J.D. Jackson, Classical Electrodynamics, 3rd edition, Wiley, 1998.
The book
• B.A. Dubrovin, A.T. Fomenko and S.P. Novikov, Modern Geometry
— Methods and Applications: Part II: The Geometry and Topology of
Manifolds, 2nd edition, Springer, 1991
is a good introduction to theory of connections on principal and vector bundles,
see also
• T. Frankel, The Geometry of Physics: an Introduction, Cambridge
University Press, 1999.
For concise exposition for vector bundles, see the books
• P. Griffiths and J. Harris, Principles of algebraic geometry, Wiley-
Interscience, 1994.
• R.O. Wells, Differential Analysis on Complex Manifolds, Springer-
Verlag New York, 2008.
For global existence and uniqueness theorems for the Yang-Mills equations
on Minkowski spacetime see the papers
• D.M. Eardley and V. Moncrief, The global existence of Yang-Mills-
Higgs fields in 4-dimensional Minkowski space I. Local existence and
smoothness properties Comm. Math. Phys., 83 (1982), 171–191; II.
Completion of proof, ibid. 193–212
• M.V. Goganov and L.V. Kapitanskiĭ, Global solvability of the Cauchy
problem for Yang-Mills-Higgs equations, J. Sov. Math. 37 (1987),
802–822
125
126 NOTES AND REFERENCES
and
• P.T. Chruściel and J. Shatah, Global existence of solutions of the Yang-
Mills equations on globally hyperbolic Lorentzian manifolds, Asian J.
Math. 1:3 (1997), 530–548
The elegant proof in Lecture 13, that for the Yang-Mills theory C a (x) are the
first class constraints, is based on Faddeev’s lectures on Feynman path integral
and gauge fields (unpublished, 1974).
Note that for large Ak Faddeev-Popov operator M , defined by formula
(13.22), may have a zero eigenvalue, so that the Coulomb gauge condition in-
tersects the orbits of gauge group more that once. This is the so-called Gribov
ambiguity, which has been rigorously considered in
• I. M. Singer, Some remarks on the Gribov ambiguity, Commun. Math.
Phys. 60 (1978), 7–12.
However, this problem does not affect the perturbation theory (Feynman rules)
based on the Hamiltonian formulation of the Yang-Mills theory.
Part 3
Special relativity
Points in the spacetime are thought of as coordinates of events and the Minkowski
distance between two events P1 = (ct1 , x1 , y1 , z1 ) and P2 = (ct2 , x2 , y2 , z2 ) is
called the interval,
ds2 = a(v)ds0 2 ,
where the constant a(v) can depend only on the absolute value v = |v| of the
relative velocity v of the inertial frames K and K 0 . Applying this to three
reference frames K, K1 , K2 we get
a(v1 )
= a(v12 ),
a(v2 )
Om = {x ∈ M4 : xµ xµ = c2 t2 − x2 − y 2 − z 2 = m2 }
129
130 14. SPECIAL RELATIVITY
cone (see Fig. 1). Correspondingly, two events P1 , P2 ∈ M4 are called timelike
if s212 > 0, spacelike if s212 < 0 and lightlike if s12 = 0. It follows from the
transitivity of the L-action on orbits that for two timelike events there is a
Lorentz transformation such that they take place in the same point in space,
P2 − P1 = (t2 − t1 , 0, 0, 0), while for the two spacelike events there is a Lorentz
transformation such that they take place at the same time, P2 −P1 = (0, x2 −x1 ).
Clearly the spacelike events cannot be causally related. Correspondingly, the
points inside the light cone with t > 0 represent the absolute future of the event
at the origin O, while the points inside with t < 0 belong to the absolute past.
The points outside the light cone are not causally related to the origin O and are
absolutely remote relative to O. This means that the concepts “simultaneous”,
“earlier” and “later” are relative for these regions.
The Lorenz group L = O(1, 3) consists of 4 × 4 matrices Λ = {Λµα } satisfying
(14.1) Λt ηΛ = η,
so that Λ00 ≥ 1 or Λ00 ≤ −1. We also have det Λ = ±1, so that the Lorentz group
L has four connected components.
The component of the identity L↑+ preserves the future and past light cones
and is called the proper orthochronous Lorentz group or restricted Lorentz group.
14.1. THE RELATIVITY PRINCIPLE AND THE LORENTZ GROUP 131
Putting
v
1
cosh ψ = r , sinh ψ = r c ,
v2 v2
1− 2 1− 2
c c
v
x0 + vt0 t 0 + 2 x0
(14.2) x= r , y = y0 , z = z0 , t = r c .
v2 v2
1− 2 1− 2
c c
x = x0 + vt0 , y = y 0 , z = z 0 , t = t0 .
dr
Consider a particle in a reference frame K moving with velocity v = . In
0
dt
the reference frame K moving relative to K with velocity V in the x direction
dr 0
velocity of a particle is v 0 = 0 . Using
dt
V
dx0 + V dt0 dt0 + 2 dx0
dx = r 0 0
, dy = dy , dz = dz , dt = r c
V2 V2
1− 2 1− 2
c c
132 14. SPECIAL RELATIVITY
we obtain
dx vx0 + V
vx = = ,
dt v0 V
1 + x2
r c
V2
dy vy0 1 − 2
vy = = c ,
dt vx0 V
1+ 2
r c
V2
dz vz0 1 − 2
vz = = c .
dt vx0 V
1+ 2
c
When |V | c we get
x0 + vt0 x0 + vt0
x1 = r1 , x2 = r2
v2 v2
1− 2 1− 2
c c
and
∆x0
∆x = r .
v2
1− 2
c
Denoting by l0 = ∆x the proper length of the rod, the length in a reference
frame where it is at rest, and by l = ∆x0 its length in a moving reference frame
K 0 , we obtain the Lorentz contraction
r
v2
l = l0 1 − 2 ,
c
so that l < l0 .
Next consider the clock which is at rest in the moving reference frame K 0 .
Let (t01 , x0 , y 0 , z 0 ) and (t02 , x0 , y 0 , z 0 ) be two events occurring at the same point
(x0 , y 0 , z 0 ) in space in the reference frame K 0 , so that the time between these
events in K 0 is ∆t0 = t02 − t01 . It follows from (14.2) that in the fixed reference
14.3. LIE ALGEBRA OF THE LORENTZ GROUP 133
frame K
v v
t01 + 2 x0 t02 + 2 x0
t1 = r c , t2 = r c .
v2 v2
1− 2 1− 2
c c
Thus the time that elapses between these two events in the reference frame at
rest K is
∆t0
∆t = r ,
v2
1− 2
c
so that ∆t0 < ∆t. This is time dilation in special relativity: the time between
events occurring at the same place in a moving reference frame is always smaller
than the time between these events in a reference frame at rest. The time ∆t0
is called a proper time.
Remark. Note that notion of being on the same point in space depends
on the reference frame. Thus events (t01 , x0 , y 0 , z 0 ) and (t02 , x0 , y 0 , z 0 ) occur in the
same point in space in the reference frame K 0 , but in the reference frame K
x0 + vt01 x0 + vt02
x1 = r , x2 = r ,
v2 v2
1− 2 1− 2
c c
and x1 6= x2 .
[M λµ , M ρσ ] = −η λρ M µσ + η λσ M µρ − η µσ M λρ + η µρ M λσ .
Introducing
1
Ji = εikl M kl and Ki = M0i , i = 1, 2, 3,
2
134 14. SPECIAL RELATIVITY
[Ji , Jj ] = εijl Jl ,
[Ki , Kj ] = −εijl Jl ,
[Ji , Kj ] = εijl Kl , i, j = 1, 2, 3.
SO(3) × SO(3) ∼
= SO(4)/{I, −I}.
Remark. Replacing η = diag(1, −1, −1, −1) by ηc = diag(c, −1, −1, −1),
we get generators Ji and Kic , and since ηc−1 = diag(1/c, −1 − 1, −1) we obtain
1
[Kic , Kjc ] = − εijl Jl .
c2
[Ji , Jj ] = εijl Jl ,
[Ji , K̃j ] = εijl Kl ,
[K̃i , K̃j ] = 0,
which characterize the Lie algebra se(3) of the Euclidean group E(3) — the
homogenous Galilean group G0 — discussed in Sect. (1.3) in Lecture 1! Thus
we see that Euclidean Lie algebra se(3) is a contraction of the Lorentz Lie
algebra so(1, 3).
1Compare with formulas for X , X and X in Example 2.2 in Lecture 2.
1 2 3
14.4. LORENTZ GROUP AS DEFORMATION OF THE GALILEAN GROUP 135
Relativistic particle
137
138 15. RELATIVISTIC PARTICLE
dxµ
1 dxµ
δ(ds) = δdxµ + δdxµ
2 ds ds
= uµ dδxµ
duµ
= d(uµ δxµ ) − δxµ ds,
ds
and using δxµ (a) = δxµ (b) = 0, we obtain
Z b b
duµ
Z
δS = −mc δ(ds) = mc δxµ ds.
a a ds
15.2. ENERGY-MOMENTUM VECTOR 139
E0 = mc2 .
mv 2
E = E0 + + O(v 4 )
2
which, except for the rest energy, is the classical expression for the kinetic energy
of a free particle. We have
E2
= p 2 + m 2 c2 , p2 = p · p,
c2
∂H ∂H
ṗ = − , ṙ =
∂r ∂p
∂L
pµ = − .
∂ ẋµ
140 15. RELATIVISTIC PARTICLE
p
Remark. Since mcu0 = m2 c2 + p2 , equation (15.4) for ν = 0 follows
from (15.5).
Remark. In the non-relativistic limit |v| c equation (15.5) turns into
dv v
m =e E+ ×B
dt c
— Newton’s equation with the Lorentz force.
The Lagrangian of a charged particle in electromagnetic field is
r
2 v2 e
L = −mc 1 − 2 + A · v − eϕ.
c c
The canonically conjugated to r momentum of the charged particle, the gener-
alized momentum, is defined by
∂L mv e e
P = =r + A = p + A,
∂v v 2 c c
1− 2
c
and the corresponding energy is
∂L mc2
E =v −L= r + eϕ,
∂v v2
1− 2
c
p
2 2 2
= c m c + p + eϕ.
Hamiltonian formulation
[P µ , P ν ] = 0,
[M λµ , P σ ] = η λσ P µ − η µσ P λ ,
[M λµ , M ρσ ] = −η λρ M µσ + η λσ M µρ − η µσ M λρ + η µρ M λσ .
dxµ x0
ẋµ =
p
L = −mc ẋµ ẋµ , , t= ,
dt c
is invariant under the action (16.1) of the Poincaré group on M4 ,
q dx0µ x00
Ldt = L0 dt0 , where L0 = −mc ẋ0µ ẋ0µ , ẋ0µ = , t0 = .
dt c
According to Noether theorem in Sect. 2.2 in Lecture 2, there are ten integrals
of motion corresponding to the generators P µ and M λµ . The integrals of motion
for the abelian Lie algebra R4 are
∂L
pµ = − ,
∂ ẋµ
143
144 16. HAMILTONIAN FORMULATION
that is,
H p mv
p0 = = p2 + m2 c2 , p= r
c v2
1−
c2
(recall that pµ = (p0 , −p), see Sect. 15.2 in Lecture 15). The vector fields on R4
µν
which corresponds to the one-parameter subgroups esM of the Lorentz group
µν
generated by M are
∂ ∂
X µν = (M µν · x)σ σ
= (η σν xµ − η σµ xν ) σ .
∂x ∂x
The corresponding Noether integrals are given by
∂L
J µν = (η σν xµ − η σµ xν ) = xµ p ν − xν p µ .
∂ ẋσ
Thus we obtain components of the total angular momentum
Jx = J 23 = x2 p3 − x3 p2 , Jy = J 31 = x3 p1 − x1 p3 , Jz = J 12 = x1 p2 − p1 x2
Kx = J 01 = x0 p1 − x1 p0 , Ky = J 02 = x0 p2 − x2 p0 , Kz = J 01 = x0 p3 − x1 p3 .
maps B(0, c), the ball of radius c in R3 , onto R3 and the phase space of a free
relativistic particle of mass m is R6 . The inverse transform is
cp cp
(16.3) v=p = 0.
p2 + m2 c2 p
16.2. HAMILTONIAN ACTION OF THE POINCARÉ GROUP 145
J1 = x2 p3 − x2 p3 , J2 = x3 p1 − x1 p3 J3 = x1 p2 − x2 p1
(see Example 6.1 in Lecture 6) and Pi = −pi . Indeed, abelian group of transla-
tions of R3 acts on R6 by (p, r) 7→ (p, r + a) and the corresponding vector field
Xa is given by
d ∂f
Xa (f )(p, r) = f (p, r − a) = −ai (p, r).
du u=0 ∂xi
Thus the vector fields Xei are Hamiltonian vector fields with Hamiltonian func-
tions −pi , i.e.,
∂
Xei = − i = −J(dpi ), i = 1, 2, 3.
∂x
The one-parameter subgroup T of time translations acts on L by l 7→ l +
(x0 , 0, 0, 0) with the representative (r − x0 v/c, v). Thus T acts on R6 by
x0 p
r 7→ r − , p 7→ p
p0
pi ∂
and the corresponding vector field is X = . Using that
p0 ∂xi
∂ ∂
J(dp) = and J(dr) = − ,
∂r ∂p
0
(see Sect. 4.3 in Lecture 4) we obtain that X = J(dpp ), i.e., X is a Hamiltonian
vector with with the Hamiltonian function is p0 = p2 + m2 c2 , i.e., is 1/c times
the Hamiltonian of a free relativistic particle of mass m.
Next, consider the one-parameter subgroup K1 of P which consists on
Lorentz boosts in x0 x1 -planes,
Λ(ψ)x = (x0 cosh ψ + x1 sinh ψ, x0 sinh ψ + x1 cosh ψ, x2 , x3 ), ψ ∈ R.
To find the action of Λ(ψ) on R6 we need to determine how in acts on the
representative (r, v) of a straight line l. We have
so that
Λ(ψ)(p) = (p1 cosh ψ + p0 sinh ψ, p2 , p3 ).
The vector field corresponding to the K1 action on R6 is given by
d
X1 (f )(p, r) = f (Λ(−ψ)p, Λ(−ψ)r)
dψ ψ=0
x1
∂ ∂ ∂ ∂
= p1 + p2 2 + p3 3 − p0 .
p0 ∂x1 ∂x ∂x ∂p1
Thus we obtained thatpX is a Hamiltonian vector field with the Hamiltonian
function K1 (p, r) = x1 p2 + m2 c2 , i.e.,
X = J(dK1 ).
Similarly, we see that vector fields X2 and X3 for one-parameter subgroups K2
and pK3 are Hamiltonian vector field p with the Hamiltonian function K2 (p, r) =
2 3
x p + m c and K3 (p, r) = x p2 + m2 c2 .
2 2 2
where ra and pa are coordinates and momenta of the a-th particle, and with the
Hamiltonian function H . Suppose that (R6n , ω, H ) is a system of n relativistic
particles, that is, the principle of relativity holds in the following form:
a) There exists a set of ten generators of the Poincaré Lie algebra — ten
functions P0 = H /c, Pi , Ji and Ki on R6n with Poisson brackets
(16.4)–(16.6).
b) The coordinates of the particles transform correctly under the Poincaré
group — coordinates ra , a = 1, . . . , n, and the generators of the Poincaré
Lie algebra have Poisson brackets (16.7)–(16.9).
In addition, suppose that the system is non-degenerate,
( )
∂2H
det 6= 0.
∂pia ∂pjb
{H , {H , ra }} = 0, a = 1, . . . , n.
Equivalently, there are Darboux coordinates p̃a and ra (the coordinates of the
particles are unchanged) and the constants ma > 0 such that
n
X
P =− p̃a ,
a=1
n
X
H =
p
c p̃2a + m2a c2 ,
a=1
Xn
Ji = εijk xja p̃ka ,
a=1
Xn p
Ki = xia p̃2a + m2a c2 .
a=1
16.3. NO-INTERACTION THEOREM 149
General relativity
and F1 = −F2 . Obviously the Newton’s law is not a Lorentz invariant and one
needs to find a Lorentz invariant description of gravity.
The first attempt1 was to include the theory of gravity into the special
relativity by assuming that gravitation field is determined by the four potential
AGµ . The interaction of a relativistic particle of charge e and mass m would be
described by the action
Z Z Z
e
S = −mc ds − Aµ dxµ − m AG µ
µ dx .
c
Considering the case e = 0 and using AG µ = (ϕ, 0, 0, 0), one gets a Lorentz
invariant modification of Newton’s law of universal gravitation,
dp ∂ϕ mv
= −m , p= r .
dt ∂r v2
1− 2
c
However, this approach does not give a correct answer for the precession of the
perihelion of Mercury.
151
152 17. GENERAL RELATIVITY
To determine the metric dl2 = γij dxi dxj in space induced by ds2 we can-
not simply put dx0 = 0 since proper time at different points in space depend
differently on the coordinate x0 . However,
2
g0i i
ds2 = g00 (dx0 )2 + 2g0i dx0 dxi + gij dxi dxj = g00 dx0 + dx − γij dxi dxj ,
g00
where
g0i g0j
(17.1) γij = −gij + , i, j = 1, 2, 3
g00
is a three-dimensional metric tensor. Since g00 > 0 it is a Riemannian metric
tensor. It depends on x0 so that the distance in real space depends on time.
The relation
g0i i
dx0 + dx = 0
g00
can be integrated over any curve in space to define x0 along the curve. This
allows to synchronize the clocks in general relativity along any curve in space.
However, this synchronization depends on a curve connecting two points in
space. Proposition 17.1 asserts that for a globally hyperbolic spacetime one can
choose coordinates such that g0i vanish and one can synchronize clocks over all
space. The corresponding coordinates (reference system in physics terminology)
are called syncrhonous.
It is easy to see from (17.1) that
−γij g jk = δik .
The relations g00 > 0 and γij is positive-definite 3 × 3 matrix are equivalent to
the
g00 g01 g02
g g01
g00 > 0, det 00 < 0, det g10 g11 g12 > 0
g10 g11
g20 g21 g22
and
g00 g01 g02 g03
g10 g11 g12 g13
g = det
g20
< 0.
g21 g22 g23
g30 g31 g32 g33
Physically these conditions should hold for any choice of coordinates on M which
can be realized with the aid of “physical bodies”.
In other words, the action functional is −mc times the length functional in the
pseudo-Riemannian geometry. Correspondingly, the Euler-Lagrange equations
are the geodesic equations with respect to the natural parameter,
d2 xλ µ
λ dx dx
ν
+ Γ µν = 0,
ds2 ds ds
where
1 λσ ∂gµσ ∂gνσ ∂gµν
(17.2) Γλµν = g ν
+ µ
−
2 ∂x ∂x ∂xσ
are Christoffel’s symbols. The free particle in a gravitational field moves along
the geodesics.
∇ = d + A, where A = Aµ dxµ .
∂
Here Aµ (x) are linear operators in Tx M which in the basis are given by
∂xµ
the matrices
λ
(17.3) (Aµ )ν = Γλνµ .
∂
Thus a derivative of a (1, 0)-tensor, a vector field V = v µ µ , in the direction
∂x
∂
is given by
∂xµ
∂v λ
(∇µ V )λ = + Γλνµ v ν ,
∂xµ
while a derivative of a (0, 1)-tensor, a 1-form θ = aµ dxµ , is
∂aλ
(∇µ θ)λ = − Γνλµ aν .
∂xµ
Directional derivative ∇µ of an arbitrary (p, q)-tensor is defined similarly. We
have
where
∂Aν ∂Aµ
Fµν = µ
− + [Aµ , Aν ].
∂x ∂xν
On 2-forms B with values in End T M the connection ∇ acts by
∇B = dB + A ∧ B − B ∧ A,
∇F = 0
Using (17.3), we obtain the following formula for the Riemann curvature
λ
tensor Rλρµν = (Fµν )ρ ,
∂Γλρν ∂Γλρµ
(17.5) Rλρµν = − + Γλσµ Γσρν − Γλσν Γσρµ .
∂xµ ∂xν
The Bianci identity for the Riemann tensor has the form
∂Γλµν ∂Γλµλ
(17.7) Rµν = − + Γλµν Γσλσ − Γσµλ Γλσν .
∂xλ ∂xν
It follows from (17.2) that
1 λσ ∂gµσ ∂gλσ ∂gµλ
Γλµλ = g + −
2 ∂xλ ∂xµ ∂xσ
1 ∂gσλ
= g λσ
2 ∂xµ √
1 ∂g ∂ log −g
= = .
2g ∂xµ ∂xµ
Thus the Ricci tensor is symmetric, Rµν = Rνµ , and determines a symmetric
bilinear form Rµν dxµ dxν on the tangent space.
Finally, the scalar curvature R is the trace of Ricci curvature tensor,
R = g µν Rµν .
2∇µ Rσµ − ∇σ R = 0,
or
1
(17.8) ∇µ Rνµ − δνµ R = 0.
2
LECTURE 18
Einstein equations – I
Rµν = 0.
159
160 18. EINSTEIN EQUATIONS – I
where ηµν is Minkowski metric. It is also assumed that these asymptotics can
be differentiated with respect to xµ .
Timelike geodesic is slow if ẋi (t) c, where i = 1, 2, 3 and t = x0 /c. Since
1p µ ν
1
dτ = gµν ẋ ẋ dt = 1 + O 2 dt,
c c
the equation for slow geodesic takes the form
d2 xλ dxµ dxν
1
2
+ Γλµν =O .
dt dt dt c
we see that up to the order O(1/c) the geodesic equation becomes Newton’s
equation
∂ϕ
r̈ = − ,
∂r
∂ϕ
and the force acting on a particle is F = −m .
∂r
To find the potential ϕ we need to use Einstein equations. The energy-
momentum tensor of a macroscopic body which consists of slow moving particles
is given by
T µν = M (x)c2 uµ uν ,
where M (x) is the mass density of the body and uµ is a four-velocity vector.
If the macroscopic motion of the body is slow, we can put u0 = 1 and ui = 0,
i = 1, 2, 3. Thus the energy-momentum tensor takes the form
It follows from formula (17.7) in Lecture 17 that in the weak gravitational field
Rνµ = O(1/c2 ) and the only nontrivial contribution to Einstein equation (18.1)
is
4πG 4πM
R00 = 4 T = 2 .
c c
Since
∂Γi00
0 1 1 2 1
R0 = + O 3 = 2∇ ϕ + O 3 ,
∂xi c c c
Einstein equations for the weak gravitational field reduce to the Poisson equation
∇2 ϕ = 4πM
18.3. HILBERT ACTION 161
where R is the scalar curvature of the metric ds2 = gµν dxµ dxν ∈ M , and
√
−g d4 x is the corresponding volume form on M . Here integration goes over a
domain D in M (usually bounded by two spacelike Cauchy hypersurfaces) and
it is assumed that all metrics in M have the same boundary value on ∂D. In
addition, normal derivatives of gµν on ∂D are fixed.
Proposition 18.1. Let uµν = δgµν be a tangent vector to M at a point
gµν ∈ M and uµν = g µα g νβ uαβ . Then the Gato derivative of the Hilbert func-
tional S in the direction u is given by
√
Z
1
δu S = Rµν − gµν R uµν −g d4 x.
D 2
Proof. Putting
d
δS = SEH (gµν + εδgµν )
dε ε=0
we have
√ √
Z Z
µν µν 4
δS = (δg Rµν + g δRµν ) −g d x + Rδ( −g)d4 x.
D D
Here we used the Stokes theorem and the condition that δΓλµν = 0 on ∂D, which
follows from our assumptions on the space M of Lorentzian metrics on M .
Remark. ‘Tautologically’ computing variation of the Hilbert-Einstein ac-
tion we obtain the relation
√ √
1 1 ∂( −g R) ∂ ∂( −g R)
Rµν − gµν R = √ − .
2 −g
∂g
µν ∂xλ ∂g µν
∂xλ
18.3. HILBERT ACTION 163
Remark. If one fixes only the values of metric tensor gµν on ∂D then δS
will contain the boundary term. It is possible to add to the Hilbert-Einstein
functional S the so-called Gibbons-Hawking-York boundary term so that the δS
is still given by Hilbert’s formula. This boundary term is the integral over ∂D
of trace of the second fundamental form over the volume form of the induced
metric on ∂D.
Denote
c3
Sgravity = − S(g).
16πG
The total action of the gravitational field in the presence of a matter with the
density function Λ(x), depending only on gµν and its first derivatives, is given
by
S = Sgravity + Smatter ,
where
√
Z
1
Smatter = Λ −g d4 x.
c
Defining symmetric stress-energy tensor by
√ √
2c δSmatter 2 ∂( −g Λ) ∂ ∂( −g Λ)
Tµν = √ = √ −
−g δg µν −g ∂g
µν ∂xλ ∂g µν
∂xλ
from δS = 0 we obtain Einstein equations
1 8πG
Rµν − gµν R = 4 Tµν .
2 c
When Λ depends only on gµν , the formula for the stress-energy tensor sim-
plifies
∂Λ
Tµν = 2 µν − gµν Λ.
∂g
Thus for the electromagnetic field
1 1
Λ=− Fαβ F αβ = − Fαβ Fγδ g αγ g βδ
16π 16π
and we obtain
1 1
Tµν = −Fµλ Fνσ g λσ + gµν Fαβ F αβ .
4π 4
Up to the factor 1/4π this is formula (11.2) in Lecture 11. For a macroscopic
body the energy-momentum tensor is
Tµν = (p + ε)uµ uν − pgµν ,
where p is the pressure and ε is the energy density of the body.
For a complete determination of the distribution and motion of the matter
one must add to Einstein equations equation of the state of the matter, that is,
equation relating the pressure density and temperature. This equation must be
given along with the Einstein equations.
LECTURE 19
Einstein equations – II
165
166 19. EINSTEIN EQUATIONS – II
where
√
1 ∂( −gg µν )
Qµν
λ = − √ + g µν Γσλσ − g µσ Γνλσ − g νσ Γµλσ
−g ∂xλ
√
1 ∂( −gg µσ )
+δλν √ + g ρσ µ
Γ ρσ .
−g ∂xσ
whence
1 µν
∇ν g µν = g gσρ ∇ν g σρ .
2
Substituting this formula to (19.1) gives,
1 µν
(19.2) ∇λ g µν = g gσρ ∇λ g σρ .
2
Contracting (19.2) gµν using gµν g µν = 4 yields
gσρ ∇λ g σρ = 0,
∇λ g µν = 0.
This shows that ∇ is the Levi-Civita connection. Thus in the Palatini formalism
equations (17.2) for the Christoffel’s symbols appear from the principle of the
least action.
19.2. THE SCHWARZSCHILD SOLUTION 167
Kaluza-Klein theory
In the 1920s the only knows fundamental forces were electromagnetism and
the force of gravity, and the only known elementary particles were electron and
proton. Einstein’s idea of the unified field theory was to obtain electromagnetism
and general relativity from a single fundamental field. Toward this goal, T.
Kaluza (1921) and O. Klein (1926) proposed to consider the five-dimensional
space-time M = M × Sr1 , where the fifth dimension in the circle of small radius
r
~G
r= ∼ 10−35 m
c3
— the Planck’s length `P . The coordinates on M will be denoted by x̃a , a =
0, 1, 2, 3, 4, where x̃4 = θ, so that using xµ , µ = 0, 1, 2, 3, for coordinates on M
we have x̃µ = xµ . Consider the following pseudo-Riemannian metric on M of
signature (+, −, −, −, −),
g00 − A0 A0 g01 − A0 A1 g02 − A0 A2 g03 − A0 A3 A0
g10 − A1 A0 g11 − A1 A1 g12 − A1 A2 g13 − A1 A3 A1
g̃ab = g20 − A2 A0 g21 − A2 A1 g22 − A2 A2 g23 − A2 A3 A2
g30 − A3 A0 g31 − A3 A1 g32 − A3 A2 g33 − A3 A3 A3
A0 A1 A2 A3 −1
so that
ds̃2 = g̃ab dx̃a x̃b = gµν dxµ dxν − (Aµ dxµ − dθ)2 .
Also assume that the metric gµν dxµ dxν and the 1-form Aµ dxµ on M do not
depend on θ.
We have the following basic facts.
1) For g̃ = det g̃ab one has g̃ = −g, where g = det gµν .
2) The inverse matrix g̃ ab is given by
00
g 01 g 02 g 03 A0
g
g 10 g 11 g 12 g 13 A1
20 21
g
30 g 31 g 22 g 23 A 2
g g g 32 g 33 A 3
A0 A1 A2 A3 −1 + Aµ A µ
169
170 20. KALUZA-KLEIN THEORY
1
Γ̃µαβ = Γµαβ + g µσ (Aα Fσβ + Aβ Fσα ),
2
1
Γ̃µα4 = g µσ Fασ ,
2
4 µ 1 µ ∂Aα ∂Aβ
Γ̃αβ = Aµ Γαβ − A (Aα Fβµ + Aβ Fαµ ) − − ,
2 ∂xβ ∂xα
1
Γ̃4α4 = Aµ Fαµ ,
2
Γ̃a44 = 0.
As usual, here
∂Aβ ∂Aα
Fαβ = − .
∂xα ∂xβ
For the free particle of mass m on the five-dimensional space-time M we
have the action
Z r
dx̃a dx̃b
Z
S = −mc ds̃ = −mc g̃ab ds̃.
ds̃ ds̃
dx̃a
Using the formulas for Christoffel’s symbols Γ̃abc and putting ua = , we get
ds̃
the following equations
duµ
+ Γµαβ uα uβ = −g µσ Aα Fσβ uα uβ − g µσ Fασ uα u4 , µ = 0, 1, 2, 3,
ds̃
and
du4 ∂Aα α β
+ Aµ Γµαβ uα uβ = −Aσ Fασ uα u4 + Aσ Aα Fβσ uα uβ + u u .
ds̃ ∂xβ
Multiplying first equations by Aµ and adding them to the second equation yields
duµ
+ Γµαβ uα uβ = −ξg µν Fαν uα .
ds̃
20.2. HILBERT ACTION ON M 171
Putting
e
ξ=√
m2 c4 + e2
we see that the right hand side becomes
e µσ dxα
g F ασ .
mc2 ds
Thus we get the equation of a free charged particle moving in external gravita-
tional and magnetic fields, obtained from the action
Z Z
e
−mc ds − Aµ dxµ .
c
This is the so-called first Kaluza miracle.
c3 µν √
Z
1
SM = − R+ Fµν F −g d4 x.
16πG M 16πc
This is the desired unification of general relativity and electromagnetism. It
yields Einstein equations
1 8πG
Rµν − gµν R = 4 Tµν
2 c
172 20. KALUZA-KLEIN THEORY
∂2
4 − 2 Φ = 0,
∂θ
n=−∞
( 4 + m2n )ϕn = 0
with masses
n2
m2n = .
r2
However, these masses are very large! Thus assuming that n = 1 gives electron,
the obtained mass would me ∼ 3·1030 MeV, while the actual electron mass is
only 0.5 MeV.
Geometrically one can consider general Kaluza-Klein metrics
g − ΦAµ Aν ΦAµ
g̃ab (x, θ) = µν ,
ΦAν −Φ