QFT Lectures
QFT Lectures
Lasciate ogni speranza, voi ch’ intrate. (Dante Aligheri, Commedia Divina)
Contents
1 Minkowski space 5
1.1 Coordinates in Minkowski space . . . . . . . . . . . . . . . . . . . 5
1.2 Causal structure . . . . . . . . . . . . . . . . . . . . . . . . . . . 6
1.3 Lorentz and Poincaré groups . . . . . . . . . . . . . . . . . . . . 6
1.4 Euclidean space . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8
1.5 Joint spectrum . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8
1.6 Quantum mechanics . . . . . . . . . . . . . . . . . . . . . . . . . 9
1.7 Relativistic quantum mechanics . . . . . . . . . . . . . . . . . . . 10
1
3 The Laplace and Helmholtz equation 20
3.1 Tempered distributions . . . . . . . . . . . . . . . . . . . . . . . . 20
3.2 Fourier transformation . . . . . . . . . . . . . . . . . . . . . . . . 21
3.3 Green’s functions on the Euclidean space . . . . . . . . . . . . . 21
3.4 Bessel equations . . . . . . . . . . . . . . . . . . . . . . . . . . . 23
3.5 Macdonald function . . . . . . . . . . . . . . . . . . . . . . . . . 23
3.6 Bessel and Hankel functions . . . . . . . . . . . . . . . . . . . . . 25
5 Second quantization 32
5.1 Vector and Hilbert spaces . . . . . . . . . . . . . . . . . . . . . . 32
5.2 Direct sum . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 33
5.3 Tensor product . . . . . . . . . . . . . . . . . . . . . . . . . . . . 34
5.4 Fock spaces . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 35
5.5 Creation/annihilation operators . . . . . . . . . . . . . . . . . . . 36
5.6 Integral kernel of an operator . . . . . . . . . . . . . . . . . . . . 37
5.7 Second quantization of operators . . . . . . . . . . . . . . . . . . 38
5.8 Symmetric/antisymmetric tensor product . . . . . . . . . . . . . 39
5.9 Exponential law . . . . . . . . . . . . . . . . . . . . . . . . . . . . 40
5.10 Wick symbol . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 40
5.11 Wick symbol and coherent states . . . . . . . . . . . . . . . . . . 42
5.12 Particle number preserving operators . . . . . . . . . . . . . . . . 42
5.13 Examples . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 42
2
7.6 Representations of the CCR . . . . . . . . . . . . . . . . . . . . . 55
7.7 Fock representations of the CCR . . . . . . . . . . . . . . . . . . 56
7.8 Equivalence of representations of CCR . . . . . . . . . . . . . . . 57
7.9 Two steps of quantization with an infinite number of degrees of
freedom . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 59
7.10 Positive energy Fock quantization . . . . . . . . . . . . . . . . . . 60
7.11 Positive energy quantization for charged systems . . . . . . . . . 63
3
12 Metaplectic group in the Schrödinger representation 96
12.1 Generating functions of symplectic transformations . . . . . . . . 96
12.2 Action integral . . . . . . . . . . . . . . . . . . . . . . . . . . . . 97
12.3 Composition of generating functions . . . . . . . . . . . . . . . . 98
12.4 Linear symplectic transformations . . . . . . . . . . . . . . . . . 99
12.5 The metaplectic group . . . . . . . . . . . . . . . . . . . . . . . . 100
12.6 The Weyl-Wigner quantization in the Schrödinger representation 102
4
17 Time-dependent Hamiltonians 129
17.1 Schrödinger and Heisenberg picture . . . . . . . . . . . . . . . . . 129
17.2 Time-ordered exponential . . . . . . . . . . . . . . . . . . . . . . 129
17.3 Schrödinger and Heisenberg picture for time-dependent Hamilto-
nians . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 130
17.4 Classical dynamics . . . . . . . . . . . . . . . . . . . . . . . . . . 131
17.5 Time-dependent perturbations . . . . . . . . . . . . . . . . . . . 133
1 Minkowski space
1.1 Coordinates in Minkowski space
By definition, the Minkowski space, denoted R1,n , is the vector space R1+n
equipped with the canonical pseudo-Euclidean form of signature (− + · · · +).
Its coordinates will be typically denoted by xµ , µ = 0, 1, . . . , n. The pseudo-
Euclidean form is then given by
n
X
gµν xµ xν = −(x0 )2 + (xi )2 . (1.1)
i=1
(Throughout these notes the velocity of light has the value 1 and we use the
Einstein summation convention). We use the metric tensor [gµν ] to lower indices
and its inverse [g µν ] to raise indices:
xµ = gµν xν , xµ = g µν xν .
5
x onto Rn . Latin letters i, j, k will sometimes denote the spatial indices of a
vector. Note that xi = xi .
On R1,n we have the standard Lebesgue measure denoted dx. The notation
d~x will be used for the Lebesgue measure on Rn ⊂ R1,n .
We will often write t for x0 = −x0 . The time derivative will be often denoted
by a dot:
∂f (t) ∂f (x0 )
f˙(t) = = ∂t f (t) = = ∂0 f (x0 ) = f,0 (x0 ).
∂t ∂x0
6
It consists of four connected components
SO0 (1, n), T·SO0 (1, n), P·SO0 (1, n), X·SO0 (1, n).
(1.5) coincides with SO(1, n) for even 1+n and (1.6) coincides with SO(1, n)
for odd 1 + n.
The affine extension of the full Lorentz group R1,n o O(1, n) is called the full
Poincaré group. Its elements will be typically written as (y, Λ). On x ∈ R1,n it
acts by
(y, Λ)x := y + Λx.
Here is the multiplication:
We will often write y instead of (y, 1l) and Λ instead of (0, Λ). It is the full
symmetry group of the Minkowski space.
Example 1.2. Let us determine O(1, 1). We set
1 1
x+ := x + t, x− := x − t; x= (x+ + x− ), t= (x+ − x− ).
2 2
a b
Now, let A = .
c d
is solved by
ad + bc = 1, ac = 0, bd = 0.
This has 4 types of solutions:
Finally, we set
1 1 −1 a b 1 1
A= .
2 1 1 c d −1 1
7
1.4 Euclidean space
We will sometimes consider the Euclidean space Rd equipped with the form
The orthogonal group O(d) has only two connected components, one of them is
the group SO(d). We also have the Euclidean group Rd o O(d).
Note that if d = 1 + n and we set x0 = ±ixd , then (1.13) becomes the
Minkowski form (1.1). This trick is called the Wick rotation.
where 1la (A) is the orthogonal projection onto eigenvectors of A with eigenvalue
a. More generally, if Ω ⊂ sp(A), then we set
X
1lΩ (A) := 1la (A).
a∈Ω
8
Let A1 , . . . , An be self-adjoint operators that strongly commute, that is their
spectral projections commute. We say that (a1 , . . . , an ) ∈ sp(A1 , . . . , An ) if for
any > 0
1l[a1 −,a1 +] (A1 ) · · · 1l[an −,an +] (An ) 6= 0. (1.16)
Clearly
Example 1.3. Consider the commuting self-adjoint operators L2 = L2x +L2y +L2z
and Lz .
sp(pj ) = R; (1.22)
sp(−∆) = [0, ∞[ (1.23)
sp(p1 , p2 , p3 , −∆) = {(k1 , k2 , k3 , k12 + k22 + k32 | k1 , k2 , k3 ∈ R}. (1.24)
The Stone Theorem says that U (t) := e−itH for a uniquely defined self-adjoint
operator H, called a Hamiltonian.
In typical situations the Hamiltonian is bounded from below, which means
that there exists E ∈ R such that
Equivalently, sp(H) ⊂ [E, ∞[. It does not affect any physical predictions if we
subtract from the Hamiltonian the infimum of its spectrum.
The Hamiltonian has often a ground state, that means inf sp(H) is an eigen-
value. The ground state is often nondegenerate.
It will be convenient to formalize these properties.
9
Definition 1.5. We will say that H, H, Ω satisfy the standard requirements of
quantum mechanics (QM) if
(1) H is a Hilbert space;
(2) H is a positive self-adjoint operator on H (called the Hamiltonian);
(3) Ω is a normalized eigenvector of H with eigenvalue 0;
(4) Ω is nondegenerate as an eigenvector of H.
10
(1) Existence of a Poincaré invariant vacuum: There exists a (normalized)
vector Ω invariant with respect to R1,3 o SO0 (1, 3).
(2) Spectral condition: The joint spectrum of the energy-momentum operator
is contained in the forward light cone, that is, sp(P ) ⊂ J + .
(3) Uniqueness of the vacuum: The vector Ω is unique up to a phase factor.
11
By measuring thePobservable A we mean measuring the partition of unity (2.1).
Clearly, A = a1la (A). Hence, the average eigenvalue of A in such an
a∈sp(A)
experiment equals X
TrρA = aTrρ1la (A). (2.2)
a∈sp(A)
such that only self-adjoint operators that preserve each subspace Hi are mea-
surable. We say then that Hi , i = 1, . . . , n, are superselection sectors.
Let Qi denote the orthogonal projection onto Hi . Then linear combinations
of Qi can be measured simultaneously with all other observables. We say that
they are classical observables.
If we choose an o.n. basis of H compatible with (2.3), then only block
diagonal self-adjoint matrices are observable. States are also described by block
diagonal matrices.
Superselection sectors arise typically when we have a strictly conserved quan-
tity, this means a self-adjoint operator Q that commutes with all possible dy-
namics. For instance, the total charge of the system usually determines a su-
perselection sector. Another example of a superselection sector is the fermionic
parity: states of an even and odd number of fermions form two superselection
sectors.
12
Let H1 = Cp with an o.n. basis e1 , . . . , ep and H2 = Cq with an o.n. basis
f1 , . . . , fq . Then ei ⊗ fj i = 1, . . . , p, j = 1, . . . , q is an o.n. basis of H1 ⊗ H2 .
Matrices in B(Cp ) ⊗ 1lCq have the form
A 0
0 A
, A ∈ B(Cp ),
A
and matrices in 1lH1 ⊗ B(H2 ) have the form
b11 1l b12 1l
b21 1l b22 1l
[bij ] ∈ B(Cq ),
,
bqq 1l
2.4 ∗-algebras
n
Consider the Hilbert space H = CN , N =
P
pi qi ,
i=1
n
H = ⊕ Cpi ⊗ Cqi ,
i=1
13
(1) φ(λA) = λφ(A);
(2) φ(A + B) = φ(A) + φ(B);
(3) φ(AB) = φ(A)φ(B).
We say that an algebra A is a ∗-algebra if it is equipped with an antilinear
map A 3 A 7→ A∗ ∈ A such that (AB)∗ = B ∗ A∗ , A∗∗ = A and A 6= 0 implies
A∗ A 6= 0.
If H is a Hilbert space, then B(H) equipped with the hermitian conjugation
is a ∗-algebra
If A, B are ∗-algebras, then a homomorphism π : A → B satisfying π(A∗ ) =
π(A)∗ is called a ∗-homomorphism.
Theorem 2.1. (1) Every finite dimensional ∗-algebra A is ∗-isomorphic to
n
⊕ B(Cpi ),
i=1
for some p1 , . . . , pn
(2) If in addition A is a subalgebra of B(CN ) and contains the identity on CN ,
n
pi qi , and a basis of CN such that
P
then there exist q1 , . . . , qn with N =
i=1
n
A = ⊕ B(Cpi ) ⊗ 1lqi . (2.4)
i=1
2.5 Commutant
If B ⊂ B(H), then the commutant of B is defined as
2
We say that A ⊂ B(H) is a von Neumann algebra if A = A00 . Clearly, von
Neumann algebras are ∗-algebras.
14
It is easy to see that all ∗-subalgebras of B(CN ) containing 1lN are von
Neumann algebras. Indeed, if A is given by (2.4), then A is obviously ∗-invariant
and n
A0 = ⊕ 1lpi ⊗ B(Cqi ).
i=1
00
So, A = A.
A = {A1 , . . . , An }00 .
15
1. Consider the operators φ̂i , i = 1, 2, 3 on L2 (R3 ). They are self-adjoint and
commute. They have simple joint spectrum. The von Neumann algebra
generated by φ̂i , i = 1, 2, 3 is equal to the operators of multiplication by
functions in L∞ (R3 ).
2. Consider in addition the operators π̂i := i−1 ∂xi , i = 1, 2, 3 on L2 (R3 ).
The von Neumann algebra generated by φ̂i , π̂i , i = 1, 2, 3, coincides with
B(L2 (R3 )).
16
Remark 2.4. One can ask why von Neumann algebras are used in the Haag-
Kastler axioms to describe sets of observables. We would like to argue that it is
a natural choice.
Suppose we weaken the Haag-Kastler axioms as follows: We replace the
family of von Neumann algebras A(O) by arbitrary sets B(O) of self-adjoint
elements of B(H), and otherwise we keep the axioms unchanged. Then, if we
set A(O) := B(O)00 (which obviously contain B(O)), we obtain a family of von
Neumann algebras satisfying the usual Haag-Kastler axioms. In particular, to
see that the Einstein causality still holds, we use the following easy fact:
Let B1 , B2 , be two ∗-invariant subsets of B(H) such that
A1 ∈ B1 , A2 ∈ B2 implies A1 A2 = A2 A1 .
A1 ∈ A1 , A2 ∈ A2 implies A1 A2 = A2 A1 .
One can try to interpret quantum fields as “operator valued tempered dis-
tributions”, which become (possibly unbounded) self-adjoint operators when
smeared out with real Schwartz test functions. We can organize the internal de-
grees of freedom of neutral fields into a finite dimensional vector space V = Rn .
Thus for any f = (fa ) ∈ S(R1,3 , Rn ) we obtain a smeared-out quantum field,
which is the operator
XZ
φ̂[f ] := fa (x)φ̂a (x)dx. (2.8)
a
17
Definition 2.5. We assume that the basic requirements of Relativistic Quantum
Mechanics are satisfied and V is a finite dimensional real vector space equipped
with a representation
is continuous.
(2) Poincaré covariance: for (y, Λ) ∈ R1,3 o SO0 (1, 3) we have
∗
= φ̂ σ(Λ)f ◦ (y, Λ)−1 .
U(y,Λ) φ̂[f ]U(y,Λ)
(4) Cyclicity of the vacuum: Let Falg denote the algebra of polynomials gener-
ated by φ̂[f ]. Then Falg Ω is dense in H.
(5) Hermiticity: For any Φ, Ψ ∈ D,
In what follows a map (2.10) satisfying Axiom (1) will be called an operator
valued distribution. By saying that it is cyclic we will mean that it satisfies
Axiom (4).
Note that free bosons satisfy both the Haag-Kastler and Wightman axioms.
In particular, the joint spectrum of the Hamiltonian and momentum is contained
in the forward light cone:
18
2.10 Relationship between Haag-Kastler and Wightman
axioms
“Morally”, Wightman axioms are stronger than the Haag-Kastler axioms. In
fact, let Aalg (O) be the algebra of polynomials in φ̂[f ] with suppf ⊂ O, which
can be treated as a ∗-subalgebra of L(D). Then the family O 7→ Aalg (O) is
almost a net of field algebras. Unfortunately, elements of Aalg (O) are defined
only on D and not on the whole H, and often do not extend to bounded operators
on H.
We know that the fields φ̂[f ] are Hermitian (symmetric) on D. Suppose
they are essentially self-adjoint. Then their closures are self-adjoint operators
on H. We could consider the von Neumann algebra A(O) generated by bounded
functions of φ̂[f ], suppf ⊂ O. Then there is still no guarantee that the net
O 7→ A0 (O) satisfies the Haag-Kastler axioms: we are not sure whether the
Einstein causality holds.
To understand this, we recall that there are serious problems with commu-
tation of unbounded operators [Reed-Simon-I]. One says that two self-adjoint
operators commute (or strongly commute) if all their spectral projections com-
mute. There exist however examples of pairs of two self-adjoint operators A, B
and a subspace D ⊂ DomA ∩ DomB with the following property:
(1) A and B preserve D and are essentially self-adjoint on D.
(2) A and B commute on D.
(3) A and B do not commute strongly.
(4) D is dense.
One of the most important topics in QFT is that of gauge invariance. In the
older literature one distinguishes between gauge invariance of the first kind–wrt
a global symmetry–and of the second kind–wrt a local symmetry. In modern
physics literature the first meaning seems to have disappeared, although it is
still used in some parts of mathematical literature. Thus in the modern physics
usage by gauge invariance one one means local gauge invariance.
Global symmetries are well understood in the framework of Haag-Kastler
axioms, thanks to the work of Doplicher-Haag-Roberts. Unfortunately, to my
understanding, we do not know how to accommodate (local) gauge invariance
in axioms of QFT.
The Haag-Kastler axioms are so abstract, general and have so little structure
that we do not know how to see the gauge invariance. The Wightman axioms do
not apply to gauge fields, because apparently for them one needs an indefinite
product Hilbert space or nonlocal fields like Wilson loops. I am not aware of a
successful adaptation of Wightman axioms that accommodates gauge fields.
19
3 The Laplace and Helmholtz equation
3.1 Tempered distributions
The space of Schwartz functions on Rn is defined as
S(Rn ) := Ψ ∈ C ∞ (Rn ) : |xα ∇βx Ψ(x)|2 dx < ∞,
α, β ∈ Nn .
R
(3.1)
Remark 3.1. (3.1) is equivalent to the definition
S(Rn ) = Ψ ∈ C ∞ (Rn ) : |xα ∇βx Ψ(x)| ≤ cα,β ,
α, β ∈ Nn . (3.2)
more common in the literature.
S 0 (Rn ) denotes the space of continuous linear functionals on S(Rn ). This
means. a linear functional S(Rn ) 3 Ψ 7→ hT |Ψi ∈ C belongs to S 0 iff there
exists N such that
X Z 21
|hT |Ψi| ≤ C |xα ∇βx Ψ(x)|2 dx .
|α|+|β|<N
(3.6) is simply given by the locally integrable function tλ , however for λ < −1
it is not.
Here some examples of functions that do not correspond to elements of S 0 (R):
t 1 1
e , |t| , t . The first blows up at infinity too fast. The last two can be regularized
to make a distribution in S 0 (R). Note the Sochocki formula and its consequence:
1 1
= P ± iπδ(t), (3.7)
t ∓ i0 t
1 1
− = 2πiδ(t). (3.8)
t − i0 t + i0
Note that for λ > −1
20
3.2 Fourier transformation
The definition of the Fourier transform of Rd 3 ~x 7→ f (~x) on a Euclidean space
will be standard: Z
~
fˆ(~k) := e−ik·~x f (~x)d~x.
F is unitary on L2 (Rd ).
Often, we will drop the hat – the name of the variable will indicate whether
we use the position or momentum representation:
Z Z
~ 1 ~
f (~k) = e−ik·~x f (~x)d~x, f (~x) = eik·~x f (~k)d~k.
(2π)d
On a Minkowski space, for the time variable (typically t) we reverse the sign
in the Fourier transform:
Z Z
1
f (ε) = eiεt f (t)dt, f (t) = e−iεt f (ε)dε.
2π
21
solves (3.12)
Assume that ζ, f ∈ S 0 (Rd ) and apply the Fourier transformation
Z Z
ζ(x) = (2π)−d eipx ζ(p)dp, f (x) = (2π)−d eipx f (p)dp (3.15)
to (3.12). Then
[
Using G ? f = Ĝfˆ we obtain
eixp
Z Z
ζ(x) = Gd,m (x − y)f (y)dy, Gd,m (x) = (2π)−d dp. (3.17)
(p2 + m2 )
The Euclidean invariance suggests to look for Green’s functions that depend
only on r = |x|, so that one can write G(x) = G(r). Such Green’s functions
away from r = 0 satisfy the radial part of the Helmholtz equation:
d−1
− ∂r2 − ∂r ± m2 G(r) = 0 (3.19)
r
The massless Green’s function can be expressed in terms of elementary func-
tions.
Γ( d
2 −1) 2−d
Theorem 3.2. For d ≥ 3, Gd,0 (r) = d r .
4π 2
1 1
E.g. G3,0 (r) = 4πr , G4,0 (r) = 4π 2 r 2 .
22
But the lhs of (3.21) equals
Z ∞
d−1
Z
G(x)∆Φ(x)dx = G(r)rd−1 |Sd−1 | ∂r2 + ∂r Φ(r)dr (3.22)
0 r
Z ∞ d − 1
= cd |Sd−1 | Φ(r) ∂r2 − ∂r rdr (3.23)
0 r
!
0
∞
+ rΦ (0) − Φ(0) + (d − 1)Φ(0) (3.24)
0
23
Theorem 3.3. For | arg z| < π − ,
Kµ (z)
lim
−z
√ = 1.
|z|→∞ e √ π
2z
1 ∞ −µ−1
Z
Kµ (z) = t exp(zφ(t))dt
2 0
1 ∞ φ00 (t0 )
Z
' exp zφ(t0 ) + z (t − t0 )2 dt
2 −∞ 2
Z ∞ √
1 −z z 2
1 −z 2π
= e exp (t − 1) dt = e √ .
2 −∞ 2 2 z
2
mµ ∞
Z mr
µ
m Kµ (mr) = exp − (s + s−1 ) sµ−1 ds (3.32)
2 0 2
Z ∞
m2 r2 µ−1
1 2
µ
= exp −t − t dt (3.33)
2 r 0 4t
1 r −µ
→ Γ(µ). (3.34)
2 2
2
For half-integer µ the Macdonald function can be expressed in terms of
elementary functions, e.g.
π 12
K± 12 (r) = e−r . (3.35)
2r
Theorem 3.5.
1 d
Gd (r) = d r1− 2 K d −1 (r). (3.36)
(2π) 2 2
e−r e−r
E.g. G1 (r) = 2 , G3 (r) = 4πr .
24
Proof. We will use the following identities:
Z ∞
1
= e−sA ds, (3.37)
A 0
Z
sp2
2π d2 x2
dpe− 2 eipx = e− 2s . (3.38)
s
Now
eipx dp
Z
(2π)−d
(1 + p2 )
|x| ∞
Z Z
2 |x|s
= (2π)−d ds dpe−(1+p ) 2 eipx
2 0
|x| 1− d2 d Z ∞ d 1 |x|
= (2π)−d π2 dss− 2 e−(s+ s ) 2
2 0
d
d
|x| 1− 2
= (2π)−d 2π 2 K1− d (|x|).
2 2
2
Using
Gd,0 (x) = lim md−2 Gd (mx), (3.39)
m&0
Γ( d
2 −1) 2−d
we obtain an alternative proof of the zero mass formulas Gd,0 (r) = d r ,
4π 2
d ≥ 3.
(−∆ − m2 )ζ = f. (3.42)
25
The Green’s function Gd,m is well defined not only for m ≥ 0, but also for
Re(m) > 0, which guarantees m2 ∈ C\] − ∞, 0]. Taking the limit at imaginary
line, that is setting Gd,±im , we obtain two Green’s functions of (3.42):
e−ixp
Z
Gd,∓im (r) = (2π)−d dp (3.43)
(p2 − m2 ∓ i0)
i m d2 −1 ±
=± H d −1 (mr). (3.44)
4 2πr 2
Thus in the case −m2 we have many Green’s functions that belong to S 0 .
We also have many solutions in S 0 of the (homogeneous) Helmholtz equation,
e.g.
1 m d2 −1
i Gd,im (r) − Gd,−im (r) = J d −1 (mr). (3.45)
2 2πr 2
satisfies (4.5).
We say that G• is a solution the Klein-Gordon equation if
26
The ansatz analogous to (3.17)
eixp
Z
• −d
G (x) = (2π) dp (4.9)
(p2 + m2 )
is incomplete and needs to be precised to make it well defined.
Every bisolution of G• can be written as
Z
dp
G• (x) = eipx g(p)δ(p2 + m2 )
(2π)3
where g is a function on the two-sheeted hyperboloid p2 + m2 = 0 (see below
for the meaning of δ(p2 + m2 )).
Then Z X φ(si )
lim δ (f (s))φ(s)ds = . (4.10)
&0 |f 0 (si )|
f (si )=0
27
4.3 Propagators for the Klein-Gordon and wave equation
Introduce
eix·p
Z
1
G∨/∧ (x) := d
dp, (4.14a)
(2π) p + m ∓ i0sgn(p0 )
2 2
eix·p
Z
1
GF/F (x) := d
dp, (4.14b)
(2π) p + m2 ∓ i0
2
(4.14a), (4.14b) are distinguished Green’s functions (inverses) and (4.14c), (4.14e)
are distinguished solutions of the Klein–Gordon equation −2 + m2 . We will call
them jointly “propagators”.
Note the identities satisfied by the propagators:
G∨ − G∧ = GPJ (4.15a)
(+) (−)
= iG − iG , (4.15b)
GF − GF = iG(+) + iG(−) , (4.15c)
F F ∨ ∧
G +G =G +G , (4.15d)
GF = iG(+) + G∧ = iG(−) + G∨ , (4.15e)
F (+) ∨ (−) ∧
G = −iG + G = −iG +G . (4.15f)
28
4.4 Einstein causality of propagators
Proposition 4.2. We have suppG∨/∧ ⊂ J ∨/∧ and suppGPJ ⊂ J.
Proof. Let us prove that suppG∨ ⊂ J ∨ . By the Lorentz invariance it
suffices to prove that G∨ is zero on the lower half-plane. We write
eipx
Z
dp
G∨ (x) =
(p + m − i0sgnp ) (2π)4
2 2 0
0 0
e−ip x +i~p~x dp0 d~
Z
p
= .
p~2 + m2 − (p0 + i0)2 (2π)4
Hence
1 mθ(−x2 )θ(±x0 ) p
G∨/∧ (x) = θ(±x0 )δ(x2 ) − √ J1 (m −x2 ).
2π 4π −x2
1 mθ(−x2 ) ∓ p 2
GF/F (x) = δ(x2 ) − √ H1 (m −x )
4π 8π −x2
miθ(x2 ) √
± √ K1 (m x2 ). (4.21)
4π 2 x2
29
• The Pauli-Jordan
1 msgnx0 θ(−x2 ) p
GPJ (x) = sgnx0 δ(x2 ) − √ J1 (m −x2 ).
2π 4π −x2
We obtain (4.21). This is obvious for spacelike x0 , ~x, where the Wick rotation
does not change the sign of x2 and therefore we still obtain an expression in-
volving the Macdonald function. Inside the light cones the sign of x2 changes.
More precisely, we can interpret the Wick rotation as e±iφ x0 with φ ∈ [0, π2 ]
and we obtain expressions involving Hankel functions. For φ ∈]0, π2 [ we have
±iφ 0 2
±Im (e x ) + (~x) > 0. Therefore, the Eclidean square |x|2 has to be re-
2
placed with the Lorentzian square x2 ± i0. On the surface of the light cone the
terms involving u, v are sufficiently regular and cause no problem. However the
function |x|1 2 becomes
1 1 1 iπ
= 2 ∓ iπδ(x2 ) = 2 ∓ δ(x0 + |~x|) + δ(x0 − |~x|) ,
(4.25)
x2 ± i0 x x 2|~x|
1
where we used the Sochocki formula and x2 is meant in terms of the principal
value.
30
Then we compute
Finally, we can read off the formulas for positive/negative frequency solutions
from (4.17).
Below we give the formulas for the massless propagators in the position
representation in dimension 1 + 3.
1 δ(x0 ∓ |~x|)
G∨/∧ (x) = θ(±x0 )δ(x2 ) = .
2π 4π|~x|
31
GPJ (x − y)f (y)dy is a solution of
R
(2) ζ(x) :=
Proposition 4.4. Let α, β ∈ Cc∞ (Rn ). Then there exists a unique ζ ∈ Csc
∞
(R1,n )
that solves
(−2 + m2 )ζ = 0 (4.32)
with initial conditions ζ(0, ~x) = α(~x), ζ̇(0, ~x) = β(~x). It satisfies suppζ ⊂
J(suppα ∪ suppβ) and is given by
Z Z
ζ(t, ~x) = ĠPJ (t, ~x − ~y )α(~y )d~y + GPJ (t, ~x − ~y )β(~y )d~y . (4.33)
Rn Rn
Proof. Clearly, (4.33) satisfies (4.32). Using (4.31) we check that ζ(0, ~x) =
α(~x). We have
G̈PJ (0, ~x) = (∆ − m2 )GPJ (0, ~x) = 0. (4.34)
Now we can verify that ζ̇(0, ~x) = β(~x). 2
5 Second quantization
In this chapter we describe the terminology and notation of multilinear algebra.
We will concentrate on the infinite dimensional case, where it is often natural to
use the structure of Hilbert spaces. We will introduce Fock spaces and various
classes of operators acting on them. In quantum physics the passage from a
dynamics on one-particle spaces to a dynamics on Fock spaces is often called
second quantization – hence the name of the chapter.
We will consider two setups: that of vector spaces and that of Hilbert spaces.
If X , Y are vector spaces, then L(X , Y) will denote the set of linear operators
from X to Y. If X , Y are Hilbert spaces, then B(X , Y) will denote the set of
bounded operators fro X to Y.
32
v ∈ V to {ei : i ∈ I} ⊂ V then it is not linearly independent anyP more. Note
that every v ∈ V can be written as a finite linear combination v = i∈I λi ei in
a unique way.
Let V be a vector space over C or R equipped with a scalar product (v|w)
(positive, nondegenerate, sesquilinear form). It defines a metric on V by
p
kv − wk := (v − w|v − w). (5.2)
It consists of sequences (vi )i∈I , which are zero for all but a finite number of
elements.
al
If (Vi )i∈I is a family of Hilbert spaces, then ⊕ Vi has a natural scalar prod-
i∈I
uct. X
(yi )i∈I (vi )i∈I = (yi |vi ). (5.4)
i∈I
al
If I is finite, then ⊕ Vi = ⊕ Vi
i∈I i∈I
Let (Vi ), (Wi ), i ∈ I, be families of vector spaces. If
ai ∈ L(Vi , Wi ), i ∈ I,
al al
then their direct sum is denoted ⊕ ai and belongs to L ⊕ Vi , ⊕ Wi . It is
i∈I i∈I i∈I
defined as
⊕ ai (vi )i∈I = (ai vi )i∈I (5.5)
i∈I
33
Let Vi , Wi , i ∈ I be families of Hilbert spaces, and ai ∈ B(Vi , W
i ) with
supi∈I kai k < ∞. Then the operator ⊕ ai is bounded. Its extension in B ⊕ Vi , ⊕ Wi
i∈I i∈I i∈I
will be denoted by the same symbol.
In particular, in general
(λv) ⊗ w = λv ⊗ w, v ⊗ (λw) = λv ⊗ w,
(v1 + v2 ) ⊗ w = v1 ⊗ w + v2 ⊗ w, v ⊗ (w1 + w2 ) = v ⊗ w1 + v ⊗ w2 .
Vectors of the form v ⊗ w are called simple tensors. Not all elements of V ⊗ W
al
are simple tensors, but they span V ⊗ W.
If {ei }i∈I and {fj }j∈J are bases of V, resp. W, then {ei ⊗ fj }(i,j)∈I×J is a
al
basis of V ⊗ W,
al
If V, W are Hilbert spaces, then V ⊗ W has a unique scalar product such
that
To see this it is enough to choose o.n.b’s {ei }i∈I and {fj }j∈J in V, resp. W.
al
Then every element of V ⊗ W can be written as an (infinite) linear combination
of ei ⊗fj and we can use them as an orthonormal set defining this scalar product.
We set
al
V ⊗ W := (V ⊗ W)cpl ,
34
and call it the tensor product of V and W in the sense of Hilbert spaces. If
{ei }i∈I and {fj }j∈J are o.n.b’s of V, resp. W, then {ei ⊗ fj }(i,j)∈I×J is an
o.n.b. of V ⊗ W,
al
If one of the spaces V or W is finite dimensional, then V ⊗ W = V ⊗ W.
Let V1 , V2 , W1 , W2 be vector spaces. If a ∈ L(V1 , V2 ) and b ∈ L(W1 , W2 ),
al al
then there exists a unique operator a ⊗ b ∈ L(V1 ⊗ W1 , V2 ⊗ W2 ) such that on
simple tensors we have
To see this it is enough to choose bases (ei )i∈I in V1 and (fj )j∈J in W1 and to
define a ⊗ b on the basis (ei ⊗ fj )(i,j)∈I×J by
Then we check that thus defined operator satisfies (5.9) and is unique. It is
called the tensor product of a and b.
If V1 , V2 , W1 , W2 are Hilbert spaces and a ∈ B(V1 , V2 ), b ∈ B(W1 , W2 ), then
a ⊗ b is bounded. It extends uniquely to an operator in B(V1 ⊗ W1 , V2 ⊗ W2 ),
denoted by the same symbol.
To prove the boundedness of a ⊗ b = a ⊗ 1l 1l ⊗ b, it is sufficient to consider
al al
the operator a ⊗ 1l from V1 ⊗ W to V2 ⊗ W. Let e1P , e2 , . . . and f1 , f2 . . . be
orthonormal bases in V1 , W resp. Consider a vector cij ei ⊗ fj .
X 2 X X 2
a ⊗ 1l cij ei ⊗ fj = cij aei
i j i
X X 2 X X
≤ kak2 cij ei ≤ kak2 |cij |2
j i j i
X 2
= kak2 cij ei ⊗ fj .
ij
To see that Θ(σ) is well defined we first choose a basis {ei }i∈I of Y. Then
al n
we define Θ(σ) on the corresponding basis of ⊗ Y:
al n
Then we extend by linearity Θ(σ) to the whole ⊗ Y. It is easy to see that the
operator defined in this way satisfies (5.11).
35
We can check that
al n
Sn 3 σ 7→ Θ(σ) ∈ L(⊗ Y) (5.12)
is a group representation.
al n
We say that a tensor Ψ ∈ ⊗ Y is symmetric, resp. antisymmetric if
Θ(σ)Ψ = Ψ, (5.13)
resp. Θ(σ)Ψ = sgn(σ)Ψ. (5.14)
36
Then on ⊗ns/a Y we have
√
a∗ (z) = Θn+1
s/a n + 1|z) ⊗ 1ln⊗ , (5.17)
√
a(z) = n(z| ⊗ 1l(n−1)⊗ . (5.18)
The space ⊗ns/a Z can then be identified with the space of symmetric/antisymmetric
square integrable functions L2 (Rnd ), and then
√
a(ξ)Φ (ξ10 , . . . , ξn−1
0
) = nΦ(ξ, ξ10 , . . . , ξn−1
0
). (5.22)
Note that strictly speaking A(·, ·) does not have to be a function. E.g. in the
case X = Rd it could be a distribution, hence one often says the distributional
kernel instead of the integral kernel. Sometimes A(·, ·) is ill-defined anyway. At
least formally, we have
Z
AB(x, y) = A(x, z)B(z, y)dz,
37
A∗ (x, y) = A(y, x).
Here is a situation where there is a good mathematical theory of inte-
gral/distributional kernels:
Theorem 5.2 (The Schwartz kernel theorem). B is a continuous linear trans-
formation from S(Rd ) to S 0 (Rd ) iff there exists a distribution B(·, ·) ∈ S 0 (Rd ⊕
Rd ) such that
Z
(Ψ|BΦ) = Ψ(x)B(x, y)Φ(y)dxdy, Ψ, Φ ∈ S(Rd ).
Γ(q) = q ⊗ ··· ⊗ q .
⊗n
s/a
Z ⊗n
s/a
Z
38
Let us prove it in the bosonic case. Let Φ ∈ Γns (Z).
â∗i âj Φ = nΘns |ei ) ⊗ 1l(n−1)⊗ (ej | ⊗ 1l(n−1)⊗ Φ (5.25)
(n−1)⊗
= nΘns |ei )(ej |
⊗ 1l Φ (5.26)
1 X
= Θ(σ)|ei )(ej | ⊗ 1l(n−1)⊗ Θ(σ)−1 Φ (5.27)
(n − 1)!
σ∈Sn
n
X
= 1l(k−1)⊗ |ei )(ej | ⊗ 1l(n−k)⊗ Φ. (5.28)
k=1
39
5.9 Exponential law
Let Z, W be Hilbert spaces. We can treat them as subspaces of Z ⊕ W. Let
Φ ∈ ⊗ns/a Z, Ψ ∈ ⊗m
s/a W. We can identify Φ ⊗ Ψ with
r
(n + m)!
U Φ ⊗ Ψ := Φ ⊗s/a Ψ ∈ ⊗n+m
s/a (Z ⊕ W). (5.38)
n!m!
Theorem 5.3. The map (5.38) extends to a unitary map
It satisfies
U Ω ⊗ Ω = Ω, (5.40)
dΓ(h ⊕ g)U = U dΓ(h) ⊗ 1l + 1l ⊗ dΓ(g) , (5.41)
Γ(p ⊕ q)U = U Γ(p) ⊗ U Γ(q), (5.42)
∗ ∗ ∗
a (z ⊕ w)U = U a (z) ⊗ 1l + 1l ⊗ a (w) , (5.43)
a(z ⊕ w)U = U a(z) ⊗ 1l + 1l ⊗ a(w) , in the bosonic case, (5.44)
∗
= U a∗ (z) ⊗ 1l + (−1)N ⊗ a∗ (z) ,
a (z ⊕ w)U (5.45)
= U a(z) ⊗ 1l + (−1)N ⊗ a(z) , in the fermionic case.
a(z ⊕ w)U (5.46)
Proof. Let us prove the unitarity of this map in the symmetric case:
1 X
Φ ⊗s Ψ = Θ(σ)Φ ⊗ Ψ (5.47)
(n + m)!
σ∈Sn+m
n!m! X
= Θ(σ)Φ ⊗ Ψ. (5.48)
(n + m)!
[σ]∈Sn+m /Sn ×Sm
The terms on the right are mutually orthogonal. The maps Θ(σ) are unitary.
The number of cosets in Sn+m /Sn × Sm is (n+m)!
n!m! . Therefore the square norm
of (5.47) is
n!m!
kΦ ⊗ Ψk2 . (5.49)
(n + m)!
2
40
be a complex function. Note that (5.50) can be also interpreted as the integral
kernel of an operator b from ⊗k Z to ⊗m Z:
Z Z
(Φ|bΨ) = · · · Φ(ξ1 , · · · ξm )b(ξ1 , · · · ξm , ξk0 , · · · , ξ10 )
â∗ (ξ1 ) · · · â∗ (ξm )â(ξk0 ) · · · â(ξ10 )dξ1 , · · · dξk dξ10 · · · dξm
0
.
(Actually, by (5.52), in (5.53) and (5.54) we can consider b which is not
symmetric/antisymmetric.)
Here is an equivalent definition of b(â, â): Its only nonzero matrix elements
are between Φ ∈ ⊗p+m p+k
s/a Z, Ψ ∈ ⊗s/a Z, and equal
p
(m + p)!(k + p)!
∗
(Φ|b(â , â)Ψ) = (Φ|b ⊗ 1⊗p
Z Ψ). (5.55)
p!
To see this it is enough to use the formal identity (5.22) several times:
Φ|â∗ (ξ1 ) · · · â∗ (ξm )â(ξk0 ) · · · â(ξ10 )Ψ
(5.56)
= â(ξm ) · · · â(ξ1 )Φ|â(ξk0 ) · · · â(ξ10 )Ψ
(5.57)
p
= (m + p) · · · (p + 1)(k + p) · · · (p + 1) (5.58)
Z
0
× Φ(ξm , . . . , ξ1 , ηp , . . . , η1 )Ψ(ξm , . . . , ξ10 , ηp , . . . , η1 )dηp · · · dη1 . (5.59)
41
Essentially every operator on a Fock space can be written as a linear com-
bination of (5.54).
5.13 Examples
Consider the Schrödinger Hamiltonian of n identical particles on L2 (RdN )
n
X X
Hn = − ∆i + V (xi − xj ), (5.68)
i=1 1≤i<j≤n
n
X 1
Pn = ∂xi , (5.69)
i=1
i
42
In the momentum representation
n
X
Hn = p2i
i=1
X
+(2π)−d δ(p0i + p0j − pj − pi )V̂ (p0i − pi ).
1≤i<j≤N
n
X
Pn = pi .
i=1
2π d
Consider L2 ([0, L]d ) ' L2 L Z and its 2nd quantization. Again we use
x, y in the position representation with periodic boundary conditions and k, k 0
in the momentum representation. We can pass from one representation to the
other by
Z
d d
X
a∗ (k) = L− 2 a(x)e−ikx dx, a∗ (x) = L− 2 a(k)eikx , (5.76)
k
Z
−d −d
X
a(k) = L 2 a(x)eikx dx, a(x) = L 2 a(k)e−ikx . (5.77)
k
43
Here are the analogs of (5.73) and (5.75):
X
H= p2 a∗p ap
p
XXX
+ L−d V̂ (k)a∗p+k a∗q−k aq ap ,
p q k
X
P = pa∗p ap .
p
d
F (ζ) = hdF (ζ)|V (ζ)i. (6.2)
dt
If we fix coordinates xi on Y, then for each y ∈ Y, the tangent space Ty Y
∂
is spanned by the vectors ∂x i and its dual, called the cotangent space by the
d i
ζ (t) = V i ζ(t) ,
(6.4)
dt
d ∂F (ζ)
F (ζ) = · V i (ζ). (6.5)
dt ∂xi
44
Another form:
d φ 0 1 ∂φ H
= . (6.7)
dt π −1 0 ∂π H
A vector field of the form
∂ ∂
∂πi H(φ, π) i
− ∂φi H(φ, π) (6.8)
∂φ ∂πi
for some function H is called a Hamiltonian vector field. Thus the Hamilton
equations are given by a Hamiltonian vector field.
Given two functions F, G on Y one introduces the Poisson bracket which is
a bilinear antisymmetric map C ∞ (Y) × C ∞ (Y) → C ∞ (Y), as
45
Clearly,
D ∂ ∂ E D ∂ ∂ E
ω| , = ω| , = 0, (6.18)
∂φi ∂φj ∂πi ∂πj
D ∂ ∂ E D ∂ ∂ E
ω| , j = − ω| i , = δji . (6.19)
∂πi ∂φ ∂φ ∂πj
One can introduce a linear map ω : Ty R2n → T∗y R2n such that
Thus
∂
ω =dπi , (6.21)
∂φi
∂
ω = − dφi . (6.22)
∂πi
0 −1 −1 0 1
Thus ω is given by the matrix ω = . Clearly, ω = . A
1 0 −1 0
vector field on a symplectic manifold is called Hamiltonian if it has the form
In the previous subsection the phase space was Y = R2n and φi , πi were
functions on Y satisfying the Poisson bracket relations (6.12).
As a side remark, let us note the following properties of the symplectic form:
1. ω(y) is nondegenerate at every point y ∈ Y;
2. dω = 0.
Nondegenerate means: Let z ∈ Ty Y be a tangent vector. If for any vector
v ∈ Ty Y we have ω(z, v) = 0, then z = 0.
One can be more general: Let Y be a manifold equipped with a 2-form
satisfying 1. and 2. Then we say that Y, ω is a symplectic manifold. The
Darboux Theorem says that on any symplectic manifold locally we can always
choose coordinates, say φi , πj , i = 1, . . . , n, such that (6.17) holds.
Specifying a symplectic form is equivalent to specifying a Poisson bracket.
Indeed, dF = (∂φ F, ∂π F ), dG = (∂φ G, ∂π G), we can write the Poisson bracket
as
D
{F, G} = hdF |ω −1 dGi = − ω|ω −1 (dF ), ω −1 (dG) , F, G ∈ C ∞ (Y). (6.24)
In most our applications, the phase space wil have the structure of a vector
space and φi , πj , i = 1, . . . , n, can be chosen to be the coordinates in a basis.
Then the tangent space to Ty Y at any point y ∈ Y can be identified with Y
itself and the form ω is simply a nondegenerate antisymmetric bilinear form on
46
Y. The Darboux Theorem says that we can identify a symplectic manifold with
a symplectic vector space at least locally.
The space Y T of linear functionals on Y obviously is contained in C ∞ (Y).
For a linear functional on Y, its derivative is the original functional itself. There-
fore, (6.17) can be simplified and written as
ω = π i ∧ φi , (6.25)
If dim Y is finite, then Y T = ωY (because ω is nondegenerate) and (6.24)
determines the Poisson bracket on the whole C ∞ (Y), consistently with (6.17).
47
6.6 Classical field theory
Consider the space Rd (the “spacetime”, where however the metric or Lorentz
structure is for the moment irrelevant) and a space Rn (the “internal degrees
of freedom”, whose indices will be as a rule omitted). A field configuration is a
function Rd 3 x 7→ ζ(x) = [ζ i (x)] ∈ Rn . The classical field φi (x) is the “value
of the ith coordinate at x of the field configuration”, that is
is called the action. Thus the action is a (typically nonlinear) functional on field
configurations.
Let us compute the the derivative in the direction of ε ∈ Cc∞ (Rd ) of the
action at the configuration ζ:
hI|ζ + εi − hI|ζi
Z
= L x, ζ(x) + ε(x), ζ,µ (x) + ε,µ (x) − L x, ζ(x), ζ,µ (x) dx
Z Z
∂L ∂L
≈ (x, ζ(x), ζ,µ (x))ε(x)dx + (x, ζ(x), ζ,µ (x))ε,µ (x)dx
∂φ(x) ∂φ,µ (x)
Z
∂L ∂L
= (x, ζ(x), ζ,µ (x)) − ∂µ (x, ζ(x), ζ,µ (x)) ε(x)dx (6.35)
∂φ(x) ∂φ,µ (x)
Z ∂L
+ ∂µ (x, ζ(x), ζ,µ (x))ε(x) dx
∂φ,µ (x)
The last term vanishes by the Stokes Theorem. Hence the derivative in the
direction of ε is given by (6.35). If we require that ζ is stationary, that is this
derivative vanishes, we obtain the Euler-Lagrange equations:
∂L(x)
∂φ(x) L(x) − ∂µ =0 (6.36)
∂φ,µ (x)
Let π µ (x) denote the canonical momentum conjugate to φ(x) in the direction
of xµ :
∂L(x)
π µ (x) = . (6.37)
∂φ,µ (x)
Consider the space of solutions of the Euler-Lagrange equations. The opera-
tion d will denote the exterior derivative on this (infinite dimensional) space.
Similarly, on this space we use ∧. If something holds on solutions of the Euler-
Lagrange equation, we will say that it is true on shell.
48
We introduce
Indeed,
∂L(x) ∂L(x)
∂µ θµ (x) = ∂µ ∧ dφ(x) + ∧ dφ,µ (x) (6.42)
∂φ,µ (x) ∂φ,µ (x)
∂L(x) ∂L(x)
= ∧ dφ(x) + ∧ dφ,µ (x) = dL. (6.43)
∂φ(x) ∂φ,µ (x)
∂L(x)
Tνµ (x) := − φ,ν (x) + δνµ L(x). (6.44)
∂φ,µ (x)
Here is the Noether Theorem in the context of classical field theory: on solutions
of the Euler-Lagrange equation we have
49
which leads to the Klein-Gordon equation with nonlinear terms:
The momentum conjugate to the field φ(t, ~x) coincides with the temporal coor-
dinate of (6.37):
∂L(t)
π(t, ~x) := = π 0 (t, ~x). (6.51)
∂ φ̇(t, ~x)
The Hamiltonian obtained from the Lagrangian by the Legendre transformation
coincides with the integral of the 00-component of the stress-energy tensor over
the corresponding constant time surface:
Z Z
H(t) := φ̇(t, ~x)π(t, ~x)d~x − L(t) = T00 (t, ~x)d~x. (6.52)
Using (6.41) we can show that (6.53) does not depend on t. Actually, instead
of integrating over {t} × Σ, we can integrate the symplectic current over any
“Cauchy surface” obtaining the same ω.
The symplectic form ω corresponds to the equal-time Poisson brackets
such that ω(y, z) = −ω(z, y) and for every y 6= 0 we can find z such that
ω(y, z) 6= 0.
50
Every finite dimensional symplectic vector space has an even dimension. If
we choose a basis, we can write
X
ω(y, z) = ωij y i z j , (7.2)
ij
51
Remark 7.1. There exists a useful generalization of a symplectic vector space:
a symplectic manifold. More precisely, let Y be a manifold equipped with a
2-form ω such that dω = 0 is ω is nondegenerate on the whole Y. Then we
say that (Y, ω) is a symplectic manifold. The Darboux Theorem says that on
any symplectic manifold locally we can always choose coordinates, say xi , pj ,
i = 1, . . . , n, such that (7.8) holds. One can also define the Poisson bracket on
Y. However, we will avoid using this concept.
1̂ = 1l; (7.12)
1
(F̂ Ĝ + ĜF̂ ) ≈ Fd G; (7.13)
2
\
[F̂ , Ĝ] ≈ i~{F, G}. (7.14)
Here ~ is a small positive parameter and ≈ means some kind of equality modulo
terms small for small ~.
One can prove that one cannot replace ≈ with =.
Almost always we will assume that Y is a symplectic vector space. As we
discussed above, if Y is finite dimensional, we can always find coordinates xi , pi ,
i = 1, . . . , n satisfying (7.8). In other words, Y = R2n is a symplectic vector
space equipped with a symplectic basis.
With the above classical system we associate a quantum system as follows.
Let ~ be a real parameter. We consider the Hilbert space H := L2 (Rn ) equipped
with the operators
∂
x̂i Ψ(x) = xi Ψ(x), p̂i Ψ(x) := ~ Ψ(x). (7.15)
i∂xi
They satisfy the Heisenberg commutation relations:
V = ξi xi + η i pi + c, (7.17)
i i
V̂ = ξi x̂ + η p̂i + c1l. (7.18)
[V̂ , V̂ 0 ] = i~{V,
\ V 0} (7.19)
One would like to extend the quantization to more general functions, not
only 1st order polynomials. There are many possibilities in the literature (e.g.
52
the Weyl quantization, Wick quantization, etc.). We will not discuss them here.
Actually, for our present purposes we will only need quantization of second order
polynomials, which we define as follows:
1 1
H= Aij xi xj + B ij pi pj + Cij xi pj , (7.20)
2 2
1 1 1
Ĥ = Aij x̂i x̂j + B ij p̂i p̂j + Cij (x̂i p̂j + p̂j x̂i ). (7.21)
2 2 2
Note that with this definition we have exact versions of (7.13) and (7.14)
[Ĥ, Ĥ 0 ] = i~{H,
\ H 0 }, (7.22)
1
(V̂ V̂ 0 + V̂ 0 V̂ ) = Vd
V0 (7.23)
2
for polynomials V, V 0 of degree ≤ 1 and for polynomials H, H 0 of degree ≤ 2.
Proof. Clearly, (7.27) is satisfied for t = 0. We check that φ̂(t) satisfies the
Heisenberg equations:
d i
i~ \
φ̂ (t) = i~{φ(t), H} = [φ̂i (t), Ĥ], (7.28)
dt
This implies (7.27) for all t. 2
Thus we can first solve the classical Hamilton equations obtaining φ(t) and
then put the hat, or first put the hat and then solve the quantum Heisenberg
equation—we obtain the same φ̂(t).
53
7.4 Weyl operators
Proposition 7.3 (Baker-Campbell-Hausdorff formula). Suppose that
[A, B], A = [A, B], B = 0.
Then
1
eA+B = eA eB e− 2 [A,B] .
Proof. We will show that for any t ∈ R
1 2
et(A+B) = etA etB e− 2 t [A,B]
. (7.29)
First, using the Lie formula, we obtain
∞ n
X t
etA Be−tA = adnA (B)
n=0
n!
= B + t[A, B].
Now
d tA tB − 1 t2 [A,B] 1 2
e e e 2 = AetA etB e− 2 t [A,B]
dt
1 2
+etA BetB e− 2 t [A,B]
1 2
−etA etB t[A, B]e− 2 t [A,B]
1 2
= (A + B)etA etB e− 2 t [A,B]
.
Besides, (7.29) is true for t = 1. 2
Let ξ = (ξ1 , . . . , ξd ), η = (η 1 , . . . , η d ) ∈ Rd . We will write
x̂(ξ) := ξi x̂i , p̂(η) := η j p̂j .
Clearly,
[x̂(ξ), p̂(η)] = i~ξ · η.
Therefore,
i~
eix̂(ξ) eip̂(η) = e− 2 ξη ei(x̂(ξ)+p̂(η)) (7.30)
= e−i~ξ eip̂(η) eix̂(ξ) . (7.31)
The operators ei(x̂(ξ)+p̂(η)) are sometimes called Weyl operators. They satisfy
the Weyl commutation relations:
0 0 i~ 0 0 0 0
ei(x̂(ξ)+p̂(η)) ei(x̂(ξ )+p̂(η )) = e− 2 (ξη −ηξ ) ei x̂(ξ+ξ )+p̂(η+η ) . (7.32)
The Weyl commutation relations, at least formally, imply the Heisenberg com-
mutation relations.
Weyl operators translate the position and momentum:
i i
e ~ (−p̂(y)+x̂(w)) x̂e ~ (p̂(y)−x̂(w)) = x̂ − y,
i i
e ~ (−p̂(y)+x̂(w)) p̂e ~ (p̂(y)−x̂(w)) = p̂ − w.
54
7.5 Stone-von Neumann Theorem
Operators x̂i , p̂i are unbounded. Hence the Heisenberg commutation relations
(7.16) are problematic–without specifying the domain it is not clear what they
precisely mean. The Weyl commutation relations (7.32) involve only bounded
operators, hence their meaning is clear.
The following theorem is one of mathematical foundations of Quantum Me-
chanics:
Theorem 7.4. Suppose that
If we drop the irreducibility condition, then there exists a Hilbert space K and a
unitary operator U : H → L2 (Rn ) ⊗ K such that
55
In this case Y = Rn ⊕ Rn , y = (ξ, η), y 0 = (ξ 0 , η 0 ),
ω(y, y 0 ) = ξη 0 − ηξ 0 . (7.39)
Set ∗
Ŵ (z) := eâ (z)−â(z)
. (7.48)
Then, using the Baker-Campbell-Hausdorff formula we see that
Z 3 z 7→ Ŵ (z) ∈ U Γs (Z) (7.49)
56
is a regular irreducible CCR representation over the symplectic space Z. Of
course, if Z is finite dimensional, it is unitarily equivalent to the Schrödinger
representation, which follows from the Stone-von Neumann Theorem, but is also
an obvious consequence of the well known theory of the harmonic oscillator.
We have
(z|z) ∗
Ŵ (z) = e− 2 eâ (z) −â(z)
e , (7.50)
â(z)Ω = 0, z ∈ Z. (7.51)
Hence,
(z|z)
(Ω|Ŵ (z)Ω) = e− 2 . (7.52)
Therefore, if we know the vacuum state we can recover the real part of the scalar
product on Z.
The symplectic form ω fixes the imaginary part of the scalar product, see
(7.47). Then there are many ways you can complete it to a full scalar product.
Each of them leads to a Fock representation of CCR satisfying (7.52). If Z
has an infinite dimension, the resulting representations in general will not be
unitarily equivalent, as we will illustrate in the next section.
The Stone-von Neumann Theorem says that all CCR representations over a
finite dimensional symplectic space are unitarily equivalent up to a multiplicity.
This is not the case for an infinite number of degrees of freedom.
In fact, for instance, there are many inequivalent Fock representations over
the same infinite dimensional symplectic space.
Let us do a computation that illustrates this. Let us start with the symplectic
space R2 with the Schrödinger representation on L2 (R) given by x̂, p̂. For any
ω > 0 we introduce the creation/annihilation operators
1 √ ip̂ 1 √ ip̂
a∗ω := √ ωx̂ − √ , aω := √ ωx̂ + √ . (7.54)
2 ω 2 ω
The vectors
1
eω,n := √ a∗n
ω Ωω (7.56)
n!
57
form an orthonormal basis of eigenvectors of the harmonic ascillator a∗ a. Thus
∞
X
Uω Φ := |n)(eω,n |Φ) (7.57)
n=0
we obtain
√ 1
Vω x̂Vω−1 = ωx̂, Vω p̂Vω−1 = √ p̂, (7.60)
ω
hence
Vω â∗1 Vω−1 = â∗ω , Vω â1 Vω−1 = âω , Vω e1,n = eω,n , (7.61)
and Vω = Uω−1 U1 . We can also compute
s
2
(Ωω |Ω1 ) = 1 1 . (7.62)
ω + ω− 2
2
So far we treated L2 (R) as the underlying space. Let us now treat the
Fock space Γs (C) as the main space. Then for any ω > 0 we have a Fock
representation
iξ √
2 ∗ iη ω ∗
R 3 (ξ, η) 7→ Wω (ξ, η) := exp √ (a + a) + √ (a − a) . (7.63)
2ω 2
They are different, but unitarily equivalent.
It is straightforward to generalize this construction to a finite number of
degrees of freedom. Thus, for any finite sequence of positive numbers we obtain
the equivalence of two CCR representations of the symplectic space R2n : the
Schrödinger representation on L2 (Rn ) and the Fock representation on Γs (Cn ).
The vacua corresponding to ω = (ω1 , . . . , ωn ) and 1 = (1, . . . , 1) have a positive
scalar product:
n s
Y 2
(Ωω |Ω1 ) = 1
−1
> 0. (7.64)
j=1 ωj2 + ωj 2
The Schrödinger representation does not work for n = ∞, since there is
no generalization of the Lebesgue measure to R∞ . However, the Fock space
Γs (l2 ) is well defined. Thus we can define creation/annihilation operators a∗i , aj ,
i = 1, 2, . . . .
58
For any infinite sequence ω = (ω1 , ω2 , . . . ) we can represent the commutation
relations
[x̂i , p̂j ] = iδij (7.65)
by setting √
1 ωi
x̂ω,i = √ (a∗ + ai ), p̂ω,i = i √ (a∗i − ai ). (7.66)
2ωi i 2
However, if 1 = (1, 1 . . . , . . . ), the infinite product
∞ s
Y 2
(Ωω |Ω1 ) = 1
− 12
. (7.67)
j=1 ωj + ωj
2
is usually zero. One can show that if (7.67) is zero then the representations
given by ω and 1 are inequivalent.
2. Then one chooses a Hilbert space where the quantum fields are repre-
sented. (Typically, this is a Fock space)
In the mathematically oriented literature there exist several equvalent ways
of presenting CCR relations. Probably, the most economical way involves the
notion of a CCR representation, which we introduced above. An alternative
way is to use a ∗ algebra of CCR. Then the two steps described above can be
described as foillows:
1. Introduce an abstract ∗-algebra of CCR over the space (Y, ω).
2. Find a representation of this algebra.
This approach has a minor problem: there are several, essentially equivalent
but mathematically different ∗-algebras that can be used to describe Canonical
Commutation Relations. Let us describe two of them (there are others).
The “Weyl CCR C ∗ -algebra” is the C ∗ algebra generated by {Ŵ (y) | y ∈ Y}
satisfying
59
The “field CCR ∗-algebra” is the ∗-algebra generated by {φ̂(y) | y ∈ Y}
satisfying
The GNS representation wrt σ acts naturally in the bosonic Fock space, as
described in Subsection 7.7. The state σ corresponds to the Fock vacuum:
σ = (Ω| · Ω).
Because of a finite number of degrees of freedom this is well defined and essen-
tially unique, since all irreducible representations are equivalent, so that this
quantization is essentially unique.
If the number of degrees of freedom is infinite, usually only the phase space
with its symplectic structure is given beforehand. In the quantum theory one
also needs a CCR representation on a Hilbert space, which is less canonical and
more tricky to choose. How to select a physically motivated CCR representation,
if we are given a symplectic space?
60
Suppose that the symplectic form is given by the matrix [ωij ], with its in-
verse denoted [ω ij ]. This is expressed by two equivalent identities, one for the
symplectic form, the other for the Poisson bracket:
X 1
ω= ωij φi ⊗ φj = ωij φi ∧ φj ; (7.75)
ij
2
{φi , φj } = ω ij . (7.76)
For ζ ∈ R2n , note that kζ = ω −1 dH(ζ) is the Hamiltonian vector field generated
by H. Thus the dynamics generated by H is rt = etk . Thus on the classical
level we have the dynamics t 7→ φi (t) with φi (0) = φi is given by the linear
transformation
φi (t) = rt,j
i
φj . (7.78)
We would like to have a quantization, such that the analogous identity is true
on the quantum level. In other words, we would like to have a Hilbert space H
equipped with operators φ̂i and Ĥ such that
Note that −ik is self-adjoint in the scalar product (7.80). Therefore, we can
diagonalize −ik in a basis orthonormal in (7.80). Note that all eigenvalues of −ik
are nonzero and real. Besides, if v i is an eigenvector of −ik with eigenvalue εi ,
then vi is an eigenvector with eigenvalue −εi , because −ik is purely imaginary.
Thus
X
hζ|hξi = hζ|hv i ihvi |hξi + hζ|hvi ihv i |hξi , (7.81)
i
X
−1
ε−1
ihζ|ωξi = hζ|h(−ik) ξi = i hζ i |hv i ihvi |hξi − hζ i |hvi ihv i |hξi ,
i
61
where we can assume that all εi are positive. Introduce the following functionals
acting on real vectors ζ:
1
hai |ζi := √ hvi |hζi, (7.82)
εi
1 1
ha∗i |ζi := √ hv i |hζi = √ hvi |hζi. (7.83)
εi εi
Now we quantize the fields on a Hilbert space with the vector Ω, so that
Clearly, what we obtain is the bosonic space with the 1-particle space spanned
by â∗i Ω. We choose the Hamiltonian
X
Ĥ = εi â∗i âi . (7.89)
sym
Note that Ĥ is positive, ĤΩ = 0 and
P (7.79) holds. If Ĥ is well defined, then
1
it differs from Ĥ by a constant 2 λi .
Let us now describe the above procedure in a basis independent way. Sup-
pose that H is a positive quadratic form on a real vector space Y equipped with
a symplectic form ω. We first extend H to a bilinear form on Y by polarization
identity. This form is generated by an operator from Y to the dual of Y, which
we denote by h
62
as a scalar product on CY called the energy scalar product. We easily check
that −ik is a self-adjoint operator on CY. We can diagonalize −ik. Because of
nondegeneracy of ω, this operator has a zero nullspace. Hence we have also a
nondegenerate sesquilinear form
ihζ|ωζi = hζ|h(−ik)−1 ζi
Let W (+) and W (−) be the positive and negative subspace of CY of the
self-adjoint operator −ik. We have
We have
eitĤ W (z)e−itĤ = W (etk z), z ∈ Z. (7.95)
The scalar product (7.93) will be treated as the basic one on Z and called
the dynamical scalar product. We have
for some symplectic matrix ω on YR ' YI . The element eiθ ∈ U (1) acts on the
coordinates as
φiR 7→ cos θφiR − sin θφiI , φiR 7→ sin θφiR + cos θφiI . (7.98)
63
In such a case it is customary to treat the space Y as a complex space
Y = CYR , setting
1 1
ψ i = √ (φiR + iφiI ), ψ i∗ = √ (φiR − iφiI ), (7.99)
2 2
so that
{ψ i , ψ j } = {ψ i∗ , ψ j∗ } = 0, {ψ i , ψ j∗ } = −ω ij , (7.100)
and the action of the group is ψ i 7→ eiθ ψ i , ψ i∗ 7→ e−iθ ψ i∗ .
Note that the space Y is equipped with complex conjugation c. It satisfies
cψ i c = ψ i∗ , c2 = 1l. It can be interpreted as the charge conjugation.
Suppose that H is a (real) quadratic Hamiltonian that is invariant wrt U (1).
Then it can be written as X
H= hij ψ i∗ ψ j (7.101)
ij
for some Hermitian matrix [hij ]. The symplectic form leads to the following
sesquilinear form, which will be called the charge:
X
Q=i ωij ψ i∗ ψ j (7.102)
ij
(Note that (7.104) and (7.105) are equivalent). We rename ai = b∗i , a∗i = bi for
i ∈ I− . Thus
X X
H= εi a∗i ai + εi b∗i bi , (7.106)
i∈I+ i∈I−
X X
Q= a∗i ai − b∗i bi , (7.107)
i∈I+ i∈I−
64
If in addition the matrix [hij ] is real then we have
1X
hij φiR φjR + φiI φjI
H= (7.109)
2 ij
(The choice of smooth functions does not matter much, since the Lagrangian
formalism serves mainly to obtain formal identities). One also chooses a local
Lagrangian density L(x), which is a function of the field φ(x), ∂µ φ(x) =: φ,µ (x)
and of x ∈ R1,3 . The Euler-Lagrange equation reads then
∂L(x)
∂φ(x) L(x) − ∂µ =0 (8.2)
∂φ,µ (x)
Recall that a function on R1,3 is called space compact if there exists a com-
pact K ⊂ R1,n such that suppf ⊂ J(K). The set of space compact smooth
∞
functions will be denoted Csc (R1,3 ). Therefore all solutions of (8.5) with com-
pactly supported Cauchy data are space compact.
65
Let YKG denote the space of real, resp. space-compact solutions of the Klein-
Gordon equation We can endow the space YKG with the standard topology of
Cc∞ (R3 ) ⊕ Cc∞ (R3 ) given by the initial conditions. The space of real continuous
T T
functionals on YKG will be denoted by YKG . The action of T ∈ CYKG on
ζ ∈ YKG will be denoted by hT |ζi, and sometimes simply by T ζ.
The field φ(x) understood as (8.1) will be called an off-shell field. It is used
in the Lagrangian formalism. When we go from the Lagrangian to Hamiltonian
formalism, we enforce the on-shell condition, that is, we restrict ourselves to
solutions of the E-L equation. We are also more careful in the choice of the
space on which the fields act. We restrict ourselves to YKG . Thus, for x ∈ R1,3 ,
the on-shell field φ(x) acting on ζ ∈ YKG gives
hφ(x)|ζi := ζ(x),
We will not distinguish the notation for on-shell and off-shell fields.
Clearly, for any ζ ∈ YKG we have
(−2 + m2 )hφ(x)|ζi = 0.
∂L(x)
π(x) := = φ,0 (x) = φ̇(x). (8.7)
∂φ,0 (x)
Equivalently,
T−1
r(y,Λ) φ(x) = φ(Λx + y). (8.9)
66
This implies Green’s identity
Z Z
j (t+ , ~x)d~x − j 0 (t− , ~x)d~x
0
(8.11)
Z
= (−2 + m2 )ζ1 (x)ζ2 (x) − ζ1 (x)(−2 + m2 )ζ2 (x) dx. (8.12)
t− <x0 <t+
or more simply, Z
ω= φ(t, ~x) ∧ π(t, ~x)d~x. (8.14)
By (8.14), the symplectic structure on the space YKG leads to the Poisson
bracket
67
The relations (8.15) can be viewed as mnemotechnic identities that yield the
correct Poisson bracket for more regular functions, eg. the smeared out fields
Z
φ[f ] := f (x)φ(x)dx. (8.17)
We have Z Z
{φ[f ], φ[g]} = − f (x)g(x)GPJ (x − y)dxdy. (8.18)
Note that formally φ(t, ~x) and π(t, ~x) generate the algebra of all functions
on YKG .
∂L(x) ,ν
T µν (x) := − φ (x) + g µν L(x) (8.19)
∂φ,µ (x)
1
∂ µ φ(x)∂ ν φ(x) − g µν ∂α φ(x)∂ α φ(x) + m2 φ(x)2 .
=
2
It is easy to check that the stress-energy tensor is conserved on solutions of
the Klein-Gordon equation (on shell):
∂µ T µν (x) = 0.
We express the stress-energy tensor in terms of φ(x) and π(x). Its compo-
nents with the first temporal coordinate are called the Hamiltonian density and
momentum density:
1 ~
2
H(x) := T 00 (x) = π(x)2 + ∂φ(x) + m2 φ(x)2 ,
2
P i (x) := T 0i (x) = −π(x)∂ i φ(x).
68
where S is any Cauchy subspace. They are examples of quadratic functionals:
Z
1 ~
2
hH|ζi = ζ̇(t, ~x)2 + ∂ζ(t, ~x) + m2 ζ(t, ~x)2 ,
2
Z
hP i |ζi = − ζ̇(t, ~x)∂ i ζ(t, ~x).
69
where we smear out ζ (±) (k) and with “positive”, resp. “negative frequency
plane waves”
1 ~ 0 ~
|k) = p q ei(−ε(k)x +k~x) , (8.32)
(2π)3 2ε(~k)
1 ~ 0 ~
| − k) = |k) = p q e−i(−ε(k)x +k~x) . (8.33)
(2π)3 2ε(~k)
We obtain
Z Z
(+) (+) (−) (−)
iω(ζ1 , ζ2 ) = ζ1 (k)ζ2 (k)d~k − ζ1 (k)ζ2 (k)d~k (8.34)
Z
1
ε(~k) ζ (+) (k)ζ (+) (k) + ζ (−) (k)ζ (−) (k) d~k
hH|ζi = (8.35)
2
Z
1
i
k i ζ (+) (k)ζ (+) (k) + ζ (−) (k)ζ (−) (k) d~k.
hP |ζi = (8.36)
2
70
(8.34) can be rewritten with ζ1 , ζ2 ∈ YKG as
Z
iω(ζ1 , ζ2 ) = ha(k)|ζ1 iha(k)|ζ2 i − ha(k)|ζ1 iha(k)|ζ2 i d~k. (8.40)
Rewriting it in a shorter form we see that a(k), a∗ (k) diagonalize the symplectic
form:
Z
iω = d~ka∗ (k) ∧ a(k). (8.41)
(8.41) is equivalent to
Hence,
71
We can use f ∈ ZKG to smear out the functionals a(k) and a∗ (k):
Z Z
ha(f )|ζi = f (k)ζ (+) (k)d~k, a(f ) = f (k)a(k)d~k; (8.51)
Z Z
∗ (+)
ha (f )|ζi = f (k)ζ (k)dk, ~ a (f ) = f (k)a∗ (k)d~k;
∗
(8.52)
(−)
and ζ (−) (k) = (−k|ζ ).
(−) (+)
Note that WKG = WKG , where we use the usual (internal) complex con-
(−) (+)
jugation in WKG . Therefore in principle we could identify ZKG and ZKG . In
particular, with this identification
| − k) = |k). (8.56)
This identification is consistently applied in this section, however in the next
(−) (+)
section we treat ZKG and ZKG as two separate Hilbert spaces.
(+) (−)
R1,3 o O↑ (1, 3) acts on ZKG and ZKG in a natural way.
(+)
We have a natural identification of YKG with WKG . Indeed, ζ ∈ YKG can be
(+)
projected onto ζ (+) ∈ WKG , as in (8.30).This identification allows us to define
a real scalar product on YKG :
(+) (+)
hζ1 |ζ2 iY := Re(ζ1 |ζ2 ).
We can compute explicitly this scalar product:
Z Z
hζ1 |ζ2 iY = ζ̇1 (0, ~x)G(+) (0, ~x − ~y )ζ̇2 (0, ~y )d~xd~y (8.57)
Z Z
+ ζ1 (0, ~x)(−∆~x + m2 )G(+) (0, ~x − ~y )ζ2 (0, ~y )d~xd~y .
72
8.7 Quantization of scalar fields
There are several equivalent presentations of free scalar quantum fields.
The description in typical physics textbooks can be described more or less
as follows. We want to construct H, Ĥ, Ω such that H is a positive self-adjoint
operator on H, Ω is a normalized eigenvector of H with eigenvalue 0 and a
self-adjoint operator valued distribution
R1,3 3 x 7→ φ̂(x), (8.58)
˙
such that, with π̂(x) := φ̂(x),
(1) (−2 + m2 )φ̂(x) = 0,
(2) [φ̂(0, ~x), φ̂(0, ~y )] = [π̂(0, ~x), π̂(0, ~y )] = 0,
[φ̂(0, ~x), π̂(0, ~y )] = iδ(~x − ~y ).
(3) eitĤ φ̂(x0 , ~x)e−itĤ = φ̂(x0 + t, ~x).
(4) Ω is cyclic for φ̂(x).
Let us describe quantum scalar fields following the above strategy, as an
(essentially unique) solution of the above problem. Let R1,3 3 x 7→ φ̂(x), π̂(x)
satisfy (1). Then the Fourier transform of φ̂ has to be supported on the mass
hyperboloid. Therefore, by the same argument as in the classical case we can
intorduce â∗ (k) and â(k) such that
d~k
Z
eikx â(k) + e−ikx â∗ (k) ,
φ̂(x) = q (8.59)
(2π)3 2ε(~k)
p
q
Z d~k ε(~k)
√ eikx â(k) − e−ikx â∗ (k) ,
π̂(x) = p (8.60)
i (2π)3 2
with the inverse transformation
s !
ε(~k)
Z
d~x −i~
k~x i
â(k) = p e φ(0, ~x) + q π(0, ~x) , (8.61)
(2π)3 2 ~
2ε(k)
s !
ε(~k)
Z
d~x ~ i
â∗ (k) = p eik~x φ̂(0, ~x) − q π̂(0, ~x) . (8.62)
(2π)3 2
2ε(~k)
(These are identities (8.38) and (8.39) decorated with hats). Again, repeating
the classical arguments, (2) implies
[â(k), â(k 0 )] =[â∗ (k), â∗ (k 0 )] = 0, (8.63)
[â(k), â∗ (k 0 )] =δ(~k − ~k 0 ). (8.64)
We still need the Hilbert space and the Hamiltonian. Since we have an
infinite number of degrees of freedom we cannot use the symmetric quantization
73
to define Ĥ, because this would produce an infinite constant. Differentiating
(3) wrt time we obtain
i[Ĥ, φ̂(x)] = π̂(x). (8.65)
This is equivalent to
(+)
Recall that L2 (R3 ) coincides with ZKG , the
completion of WKG . Thus the
Hilbert space H can be identified with Γs ZKG , Ω with the Fock vacuum, â∗ (k)
with the creation operators in the “physicist’s notation”.
As usual, we can also introduce the smeared versions of (8.61), (8.62) for
f ∈ ZKG :
Z
â(f ) = f (k)â(k)d~k; (8.69)
Z
â∗ (f ) = f (k)â∗ (k)d~k; (8.70)
74
Using the “smeared notation” on the right we can write
This is true even though we only required that time translations are imple-
mented.
One of possible alternative presentations of quantization of the free scalar
field in the mathematical style goes as follows. We have the symplectic space
(YKG , ω). This symplectic space is equipped with a symplectic dynamics rt
generated by a positive classical Hamiltonian H. We would like to find a CCR
representation
YKG 3 ζ 7→ W (ζ) ∈ U (H) (8.73)
We want the quantum Hamiltonian to be compatible with the classical Hamil-
tonian, so that
W (rt (ζ)) = eitĤ W (ζ)e−itĤ . (8.74)
We assume that the representation is Fock. Then we apply the method of a
positive energy representations, as described in Subsection 7.10.
75
In fact,
d~kd~k 0
Z Z
0
(Ω|φ̂(x)φ̂(y)Ω) = √ √ eikx−ik y (Ω|â(k)â∗ (k 0 )Ω)
(2π)3 2ε 2ε0
d~k
Z
= eik(x−y)
(2π)3 2ε(~k)
= G(+) (x − y);
(Ω|T(φ̂(x)φ̂(y))Ω) = θ(x0 − y 0 )(Ω|φ̂(x)φ̂(y)Ω) + θ(y 0 − x0 )(Ω|φ̂(y)φ̂(x)Ω)
= θ(x0 − y 0 )G(+) (x − y) + θ(y 0 − x0 )G(−) (x − y)
= −iGF (x − y),
The operators φ̂[f ] are essentially self-adjoint for f ∈ Cc∞ (O, R) on, say, smooth
vectors in the Fock space with compact supports. Therefore, we can define
W (f ) := eiφ̂[f ] .
We have
Z Z
[φ̂[f ], φ̂[g]] = −i f (x)g(x)GPJ (x − y)dydy. (8.82)
76
For an open set O ⊂ Rd we set
A(O) := {exp(iφ̂[f ]) : f ∈ Cc∞ (O, R)}00 .
The algebras A(O) satisfy the Haag-Kastler axioms.
where φ̂(x) are the free fields. Suppose we measure the vaccum–vacuum ampli-
tude. The resulting quantity is
Z
Ω|Texp − i ĤInt (t)dt Ω (8.89)
∞
X Z Z
= (−i)n ··· (Ω|ĤInt (tn ) · · · ĤInt (t1 )Ω)dtn · · · dt1 (8.90)
n=0 tn >···>t1
λ2 Z Z
= exp − f (x)f (y)(Ω|T{φ̂(x)φ̂(y)}Ω)dxdy , (8.91)
2
which is expressed in terms of the Feynman propagator.
Note that the second scenario is probably more realistic. Thus one can argue
that the Feynman propagator is more physical than the 2-point function.
77
9 Free charged scalar bosons
The formalism used in physics to describe complex fields, and especially to
quantize them, is different from the real case, therefore we devote to it a separate
section.
78
WKG will denote the space of smooth space-compact complex solutions of the
Klein-Gordon equation
(−2 + m2 )ζ = 0. (9.6)
(In the context of neutral fields, it was denoted CYKG , because it was an aux-
iliary object, the complexification of the phase space YKG . Now it is the basic
object, the phase space itself). In the on-shell formalism we will consider ψ(x),
ψ ∗ (x), η ∗ (x) and η(x) as functionals on WKG .
(8.6) implies
Z Z
ψ(t, ~x) = ĠPJ (t, ~x − ~y )ψ(0, ~y )d~y + GPJ (t, ~x − ~y )η(0, ~y )d~y . (9.7)
or equivalently
1 1
ψ ∗ (x) = √ φR (x) − iφI (x) .
ψ(x) = √ φR (x) + iφI (x) ,
2 2
Clearly, the Lagrangian density can be rewritten as
1 1
L(x) = − ∂µ φR (x)∂ µ φR (x) − m2 φR (x)2 (9.9)
2 2
1 1
− ∂µ φI (x)∂ µ φI (x) − m2 φI (x)2 . (9.10)
2 2
The usual real formalism yields a pair of neutral fields with the usual equal time
Poisson brackets (we write only the non-vanishing ones):
{φR (t, ~x), πR (t, ~y )} = {φI (t, ~x), πI (t, ~y )} = δ(~x − ~y ). (9.11)
79
9.3 Classical 4-current
The Lagrangian is invariant w.r.t. the U (1) symmetry ψ 7→ e−iθ ψ. The Noether
4-current associated to this symmetry is called simply the 4-current. It is
∂L(x) ∂L(x)
J µ (x) := i ψ ∗ (x) ∗
− ψ(x)
∂ψ,µ ∂ψ,µ
= i ∂ ψ (x)ψ(x) − ψ ∗ (x)∂ µ ψ(x) .
µ ∗
∂µ J µ (x) = 0,
µ ∗
J (x) = J µ (x).
hJ µ (x)|ζi = ij µ (ζ, ζ, x)
i ∂ µ ζ(x)ζ(x) − ζ(x)∂ µ ζ(x) .
=
is conserved (does not depend on time) and coincides with the quadratic form
obtained from (8.27):
hQ|ζi = iζωζ. (9.16)
It is conserved on shell
∂µ T µν (x) = 0.
80
The components of the stress-energy tensor with the first temporal coordinate
are called the Hamiltonian density and momentum density. We express them
on-shell in terms of ψ(x), ψ ∗ (x), η(x) and η ∗ (x):
We easily check
Note that (8.28) and (8.29) from the neutral case differ from (9.24) and (9.25)
only by the prefactor 21 .
81
We use the Fourier transformation and (8.30):
Z
ζ (+) (k)ζ (+) (k)d~k − ζ (−) (k)ζ (−) (k) d~k
hQ|ζi = (9.26)
Z
hH|ζi = ε(~k) ζ (+) (k)ζ (+) (k) + ζ (−) (k)ζ (−) (k) d~k
(9.27)
Z
hP i |ζi = k i ζ (+) (k)ζ (+) (k) + ζ (−) (k)ζ (−) (k) d~k.
(9.28)
(+) (−)
R1,3 o O↑ (1, 3) acts on ZKG and ZKG in a natural way.
82
Thus
Z r
ε(~
p) i
~ d~x
a(k) = ψ(0, ~x) + q η(0, ~x) e−ik~x p ,
2 (2π)3
2ε(~k)
s
ε(~k) ∗
Z
i
~ d~x
a∗ (k) = ψ (0, ~x) − q η ∗ (0, ~x) eik~x p ,
2 (2π)3
2ε(~k)
s
ε(~k) ∗
Z
i
~ d~x
b(k) = ψ (0, ~x) + q η ∗ (0, ~x) e−ik~x p ,
2 (2π)3
2ε(~k)
s
ε(~k)
Z
i
~ d~x
b∗ (k) = ψ(0, ~x) − q η(0, ~x) eik~x p ,
2 (2π)3
2ε(~k)
d~k
Z
eikx a(k) + e−ikx b∗ (k) ,
ψ(x) = q
(2π)3 2ε(~k)
p
q
Z d~k ε(~k)
√ eikx a(k) − e−ikx b∗ (k) .
η(x) = p
3
i (2π) 2
Z
d~k a∗ (k)a(k) − b∗ (k)b(k) .
Q =
(+) (−)
As usual, for f ∈ KKG , g ∈ KKG , we have smeared versions of the above
83
functionals:
Z Z
ha(f )|ζi = f (k)ζ (+) (k)d~k, a(f ) = f (k)a(k)d~k; (9.35)
Z Z
ha∗ (f )|ζi = f (k)ζ (+) (k)d~k, a (f ) = f (k)a∗ (k)d~k;
∗
(9.36)
Z Z
hb(g)|ζi = b(k)ζ (−) (k)d~k, b(g) = g(k)b(k)d~k; (9.37)
Z Z
hb∗ (g)|ζi = g(k)ζ (−) (k)d~k, b∗ (g) = g(k)b∗ (k)d~k; (9.38)
Thus, for k on the mass shell, using physicist’s notation on the left and
mathematician’s on the right, we can write
a∗ (k) = a∗ |k) ,
(9.40)
∗ ∗
b (k) = b c| − k) . (9.41)
9.8 Quantization
In principle, we could quantize the complex Klein-Gordon equation as a pair of
real Klein-Gordon fields. However, we will use the formalism of quantization of
charged bosonic systems, see Subsect. 7.11.
We want to construct (H, Ĥ, Ω) satisfying the usual requirements of QM
(1)-(3) and an operator valued distribution
˙
satisfying, with η̂(x) := ψ̂(x),
(1) (−2 + m2 )ψ̂(x) = 0;
(2) the only non-vanishing 0-time commutators are
[ψ̂(0, ~x), η̂ ∗ (0, ~y )] = iδ(~x − ~y ), [ψ̂ ∗ (0, ~x), η̂(0, ~y )] = iδ(~x − ~y ); (9.43)
84
letter b. Thus we put hats and do all the obvious modifications to the classical
formulas. Ω is the Fock vacuum. The quantum field is
d~k
Z
ψ̂(x) := q eipx â(k) + e−ikx b̂∗ (k) ,
(2π)3 2ε(~k)
p
q
Z d~k ε(~k)
η̂(x) := p √ eikx â(k) − e−ikx b̂∗ (k) .
i (2π)3 2
Thus all these operators are expressed in terms of the Wick quantization of their
classical expressions.
Note thatthe whole
R1,3 o O↑ (1, 3) acts unitarily on H by U (y, Λ) :=
group
Γ r(y,Λ) (+)
⊗ Γ r(y,Λ) (−)
, with
ZKG ZKG
Moreover,
[ψ̂(x), ψ̂ ∗ (y)] = −iGPJ (x − y), [ψ̂(x), ψ̂(y)] = 0.
Note the identities for the 2-point functions:
85
9.9 Smeared fields
For f ∈ Cc∞ (R1,3 , C) we set
Z Z
ψ̂[f ] := f (x)ψ̂(x)dx, ψ̂ ∗ [f ] := f (x)ψ̂ ∗ (x)dx. (9.46)
The observable algebra A(O) is the subalgebra of F(O) fixed by the automor-
phism
B 7→ eiθQ̂ Be−iθQ̂ .
The algebras F(O) and A(O) satisfy the Haag-Kastler axioms.
1 m2 2
L(x) = − ∂µ φ(x)∂ µ φ(x) − φ (x) for bosons, (10.1)
2 2
L(x) =iψ(x)γµ ∂ µ ψ(x) + mψ(x)ψ(x), for fermions. (10.2)
86
3. Compute formally the perturbation expansion of the scattering operator
between in states Φ− and out states Φ+ :
10.2 Renormalizability
Consider the Lagrangian of the form (10.1) or (10.2) in d spacetime dimensions.
We want the action to be scalar and the kinetic term to have no dimensionful
coefficient. This implies that in the units of length the boson field has the
dimension [φ] = 1 − d2 and the fermion field the dimension [ψ] = 21 − d2 . Clearly
deg ∂ = −1.
The full Lagrangian is typically the sum of monomials in fields. The action
integral should be dimensionless. The integral includes the Lebesgue measure
dd x, which has dimension d. Therefore, if the dimension of a monomial is c,
then one has to put a coupling constant in front of dimension −d − c.
We will say that a monomial is
1. super-renormalizable if the corresponding coupling constant has a negative
dimension.
87
validity is limited to low energies. Therefore, one should not treat them as
ultimate theories.
Nevertheless, renormalizable theories are distinguished. In fact, they have
clearly a better predictive power than non-renormalizable theories. There are
also well-known arguments attributed to Kenneth Wilson, involving the change
of scale, often called the renormalization group, which explain a different role
of renormalizable and non-renormalizable terms in the Lagrangian.
Let us review the renormalizability of various monomials in the Lagrangian
in various dimensions.
super-renormalizable φ3 ;
marginally renormalizable φ4 , φ2 ∂φ, φψψ.
marginally renormalizable φ3 .
10.3 Counterterms
Let us consider for instance the λφ44 theory. Naive computations based on the
Lagrangian
1 m2 2
L(x) = − ∂µ φ(x)∂ µ φ(x) − φ (x) − λφ(x)4 (10.5)
2 2
lead to ill-defined quantities. In order to obtain sensible predictions one has to
consider the Lagrangian involving counterterms, which is traditionally written
88
as
1 m2 2
L(x) = − Z ∂µ φ(x)∂ µ φ(x) + φ (x) − Z 2 gφ(x)4 (10.6)
2 2
∞
X
Z= Zn λ n , Z0 = 1; (10.7)
n=0
X∞
m= mn λn , m0 = m; (10.8)
n=0
X∞
g= gn λn , g1 = 1. (10.9)
n=1
89
this approximation, there is a pole for large energies and it is clear that (10.11)
becomes worthless. Its discovery is attributed to Landau, and the corresponding
pole is called the Landau pole. The same phenomenon can be seen for most other
QFT’s, such as λφ4 .
There exists a version of the above argument based on the renormalization
group equation that also predicts the existence of a pole in these theories and
is even more convincing.
Landau, after discovering the pole named after him, announced the death of
Quantum Field Theory. In reality, if the coupling constant is very small, the pole
is very far in high energies and one can ignore it in perturbative calculations—
this is the case of QED. However, when the coupling constant is not so small,
problems related to the Landau pole may appear close to physical energies. This
is the case of the Standard Model, where there is a φ4 term in the Higgs field.
Quantization of the Yang-Mills theory is much more difficult than that of
fields with an Abelian gauge or no local gauge at all. Nevertheless, by the early
70’s it was well understood and it was proven by t’Hooft and Veltman that
the gauge invariance survives renormalization. In the Yang-Mills theory instead
of photons we have gluons. Gluons interact with themeselves, therefore in the
gluon self-energy beside fermion loops we have (bosonic) gluon loops. They
change the sign of the self-energy, and therefore (if there are not too many
fermions) there is no Landau pole in the ultraviolet and the gluon propagator
becomes suppressed for large energies. This property of the Yang-Mills theory
is called asymptotic freedom. It was discovered in the early 70’s. It implies that
the Yang-Mills theory can be applied in large energies. This lead in the 70’s to
an enthusiastic revival of QFT.
Apparently, t’Hooft did first the computation proving the asymptotic free-
dom of the Yang-Mills theory, but did not recognize its physical importance.
The asymptotic freedom of Yang-Mills was shown (simultaneously?) by Gross-
Wilczek and Politzer. For many years they were on the list of candidates for
the Nobel Prize, until eventually they got it.
90
Another current of mathematical research has been devoted to perturbative
Quantum Field Theory. There is good understanding of renormalization of
quantum fields, and also on a curved background.
g2
ψ a (iγ µ ∂µ − m)ψ a + (ψ ψ a )2 . (10.12)
2N a
91
(N is the number of fermion species. For N = 1 the model reduces to the com-
pletely integrable Thirring model). It was constructed by Gawedzki–Kupiainen,
and by Rivasseau et al.
In the early 90’s the interest in constructive field theory waned and essen-
tially this topic was abandoned by researchers. One reason for the collapse of
the topic was that it became prohibitively complicated. Another reason was
the philosophy proclaimed by Wilson saying that quantum field theories that
we know are probably only low-energy effective approximations and one should
not be surprised if they cannot be expressed in a mathematically satisfactory
way.
Nowadays it is even very difficult to determine what has been proven and
what are the proofs—the old literature is mostly unreadable. I know only one
recent result in some kind of constructive field theory. Unfotunately, it is nega-
tive: Aizenman and Duminil-Copin proved that the λφ44 theory is trivial. More
precisely, if one tries to approximate it on a lattice, then in the limit one obtains
a trivial theory.
At the turn of millenium the Clay Institute funded prizes of 1 milion dollars
each for proving 7 important mathematical conjectures. One of them was proven
(The Poincare Conjecturé by Grigorii Perelman, who declined the prize). Other
are still open, including two conjectures in mathematical physics.
One of them is the existence of solutions to the Navier–Stokes equation. The
problem is clearly formulated (by Charles Fefferman).
The other is the construction of the Yang-Mills Theory and the proof of
the existence of a positive mass gap, (formulated by Arthur Jaffe and Edward
Witten).
To my understanding, the formulation of the problem is vague. The descrip-
tion of the problem mentions Wightman axioms, however they seem not suitable
for gauge theories. What is worse, even if we construct something that the Prize
Committee will accept as the quantized Yang-Mills Theory, we have to prove
the positivity of the mass gap (which involves controlling not only ultraviolet
divergencies, but also the large volume limit). In other words, we need to show
that the lightest glueball is massive—which is supposed to be the expression of
the confinement.
.
11 Time-dependent Hamiltonians
11.1 Schrödinger and Heisenberg picture
Suppose that H is a (time-independent) Hamiltonian. It generates the dynamics
e−itH on the Hilbert space H. If we prepare a state ρ at time 0 and measure an
observable A at time t > 0, then the expectation value of the measurement is
92
(1) The Schrödinger picture: We let the state evolve ρ(t) := e−itH ρeitH and
keep the observable constant. Then (17.1) equals Trρ(t)A.
(2) The Heisenberg picture: We let the observable evolve A(t) := eitH Ae−itH
and keep the state constant. Then (17.1) equals TrρA(t).
(By the Schrödinger picture one also means the unitary evolution Ψ(t) :=
e−itH Ψ on H.)
t 7→ H(t). (11.2)
For brevity, we will write U (t+ , t− ) for (17.3) and call it the dynamics generated
by t 7→ H(t). Note that U (t+ , t− ) are unitary. (The above constructions can
be easily made rigorous if H(t) are bounded. If they are unbounded, the above
definition should be viewed only as a heuristic indication how to define the
family of unitary operators U (t+ , t− ). In most of this subsection we are not
very precise about the boundedness of operators, types of limits, etc.)
We also set U (t− , t+ ) := U (t+ , t− )−1 . Thus U (t+ , t− ) is the solution of the
following two equivalent equations:
d
U (t+ , t− ) = −iH(t+ )U (t+ , t− ), U (t, t) = 1l; (11.4)
dt+
d
equivalently, U (t+ , t− ) = U (t+ , t− )iH(t− ), U (t, t) = 1l. (11.5)
dt−
Clearly, if H(t) = H does not depend on time, then U (t+ , t− ) = e−i(t+ −t− )H .
93
We also have
n i(t − t ) jt + (n − j)t
+ − + −
Y
U (t+ , t− ) = lim exp − H , (11.6)
n→∞
j=1
n n
where in the product the indices increase from the right to the left:
n
Y
Aj := An · · · A1 . (11.7)
j=1
Thus the family of self-adjoint operators (17.2) can be called the Hamiltonian
in the Schrödinger picture. Hence the evolution of a density matrix ρ in the
Schrödinger picture from time 0 to time t is
where we treat t = 0 as the reference time. We then have two ways to express
the time evolution of the expectation value:
d
A(t) = i H Hp (t), A(t) ,
(11.14)
dt
A(0) = A,
94
where the Hamiltonian in the Heisenberg picture is defined as
The equation
d
x(t) = X t, x(t) . (11.21)
dt
for any initial condition x(0) = x0 ∈ Rd , we obtain a solution R 3 t 7→ x(t, x0 ).
This defines a flow R(t, 0)on Rd such that R(t, 0)x0 = x(t, x0 ) and (17.22) can
be rewritten as
d
R(t, 0)x0 = X t, R(t, 0)x0 . (11.22)
dt
In classical mechanics the phase space is described by coordinates (φ, π) ∈
Rm × Rm with the Poisson bracket
{φi , φj } = {πi , πj } = 0,
i
{φ , πj } = δji .
95
and the Hamilton equations
Note that the classical evolution equations (17.22) and (17.24) are analogs
of the quantum equations in the Heisenberg picture (17.14). In particular, the
classical Hamiltonian (17.23) is the analog of the quantum Hamiltonian in the
Heisenberg picture (17.15).
There exists also a dual picture, which is the analog of the Schrödinger
picture of Quantum Mechanics. Consider the evolution given by the backward
flow. The analog of the equation (17.22) is
d
R(0, t)x0 = −X Sp t, R(0, t)x0 ,
(11.25)
dt
where
X Sp (t, y) = R0 (0, t)X t, R(t, 0)y .
(11.26)
Thus the vector field is minus the backward transport of the original field.
R0 (0, t) is the derivative of the flow.
In the Hamiltonian case we have the dynamics
Thus the classical and the quantum cases are analogous. However, whereas
in the quantum case the Schrödinger picture seems preferred, in the classical
case the analog of the Heisenberg picture seems to be more common. Therefore,
in the quantum case we put the superscript Hp but not Sp, and the other way
around in the classical case.
96
We can introduce the so-called interaction picture or the Furry picture. The
evolution in the interaction picture is
d
UInt (t+ , t− ) = −iHInt (t+ )UInt (t+ , t− ); (11.30)
dt+
d
UInt (t+ , t− ) = UInt (t+ , t− )iHInt (t− ), (11.31)
dt−
Thus if
97
Clearly, S = S +(−1) S − .
Note that both Møller operators and the scattering operator trivially exist
if V (t) decays sufficiently fast as |t| → ∞. In fact, this is a typical situation in
QFT, where we usually impose a temporal “adiabatic cutoff”.
In quantum mechanics one often applies this formalism to time independent
potentials, but this is a different story.
The interaction picture described above has also its “Heisenberg picture
version”:
Hp
UInt (t+ , t− ) := U (0, t+ )e−i(t+ −t− )Hfr U (t− , 0) (11.39)
Z t+ !
Hp
= Texp − HInt (t)dt , (11.40)
t−
Hp
HInt (t) := UInt (0, t)HInt (t)UInt (t, 0) = U (0, t)V (t)U (t, 0). (11.41)
does not depend on the choice of a basis and defines a number ∈ [0, ∞] called
the trace of A. We say that A is trace class if TrA < ∞.
If A is any operator, then it is trace class if it can be written as a linear
combination of positive trace class operators
n
X
Ai = c i Ai . (12.2)
i=1
One can show that (18.3) does not depend on the decomposition (18.2). Note
that for any unitary U
TrA = TrU AU ∗ . (12.4)
If A is an operator on L2 (Rd ), and A(x, y) is its distributional kernel, then
under some conditions one can show that
Z
TrA = A(x, x)dx. (12.5)
98
For instance, if A = f (x̂)g(p̂), then
Z Z
dp dxdp
A(x, y) = f (x)g(p)ei(x−y)p , TrA = f (x)g(p) . (12.6)
(2π)d (2π)d
Suppose that A is an operator such that ]−∞, 0] is disjoint from its spectrum.
Then we can define ln(A). Suppose in addition that A − 1l is trace class. Then
it is easy to show that ln(A) is trace class. We then can define the so-called
Fredholm determinant of A:
1
is called the partition function. The parameter β has the interpretation of kT ,
where T is the temperature. Anyway, we will assume β = 1.
In many situations, e.g. the thermodynamic limit, the partition function is
infinite and instead it is more natural to consider the ratio of partition functions
for two Hamiltonians, I and I0 . We can then introduce the parameter E equal
to the logarithm of its relative partition function:
R −I(φ)
−E e Dφ
e := R −I (φ) . (12.9)
e 0 Dφ
99
Unfortunately, (18.13) are divergent. We will show how to renormalize E. This
means, we will show how to modify the Hamiltonian by adding local countert-
erms so that we obtain a finite expression.
Let us try to analyze E, in spite of the fact that it is ill defined. We have
∞ ∞
X (−1)n+1 n X
E = Tr κ(x̂)(p̂2 + m2 )−1 =: En .
n=1
2n n=1
Note that Z Z
1 dp
E1 = E1cl = κ(x)dx (12.16)
2 (2π)4 (p2 + m2 )
is divergent. But the classical and quantum quantities coincide. We will simply
drop them.
E2 and E2cl are also divergent, but in general different:
1
−E2 = Tr κ(x̂)(p̂2 + m2 )−1 κ(x̂)(p̂2 + m2 )−1
4Z
1 dp2 dp1
= κ(p1 − p2 ) κ(p2 − p1 ) (12.17)
4 (2π)4 (p22 + m2 ) (2π)4 (p21 + m2 )
|κ(k)|2
Z
1 dk dq
= (12.18)
4 (2π) (2π) ((q + 2 k) + m2 )((q − 12 k)2 + m2 )
4 4 1 2
Z
dk
= |κ(k)|2 π(k 2 ) , (12.19)
(2π)4
Z Z
1 dp
−E2cl = κ(x)2 dx (12.20)
4 (2π) (p2 + m2 )2
4
Z Z
1 dk dp
= |κ(k)|2 , (12.21)
4 (2π) (2π) (p2 + m2 )2
4 4
Z
dk
= |κ(k)|2 π(0) , (12.22)
(2π)4
where in (18.18) and (18.19) we used κ(−k) = κ(k) and we set p1 = q − 12 k,
p2 = q + 12 k, and we introduced
d4 q
Z
1 1
π(k 2 ) = , (12.23)
4 (2π) ((q + 2 k) + m )((q − 21 k)2 + m2 )
4 1 2 2
100
All the terms En and Encl with n ≥ 3 are convergent.
To rigorously define the above expressions we need first to perform some
kind of regularization. For instance, we can use the the Pauli-Villars method
with m0 = m, C0 = 1, m1 = M , and C1 = −1. (See Subsection 18.4 for more
about this method). Let us set π(m, k 2 ) to be the formal expression (18.23).
Set
1
X
π reg (k 2 ) := π(m, k 2 ) − π(M, k 2 ) = Ci π(mi , k 2 ). (12.26)
i=0
ren 2
We will compute π (k ).
In order not to clutter the formulas, in the following computations we use the
ill defined π(k 2 ). The following formulas become well defined when we replace
1
Ci , so that instead of π(k 2 ) we obtain π reg (k 2 ):
P
m with mi and insert
i=0
Z 4
d q 1
4π(k 2 ) = 4 1
(2π) ((q + 2 k) + m )((q − 21 k)2 + m2 )
2 2
∞ Z ∞
d4 q
Z Z
2 1 2 2
= dα 1 dα2 exp −(α 1 + α 2 ) q + k + m − (α1 − α2 )qk
(2π)4 0 0 4
Z ∞ Z ∞
1 1 α1 α2 2
= dα1 dα2 exp −(α1 + α2 )m2 − k
(4π)2 0 0 (α1 + α2 )2 α1 + α2
Z 1 Z ∞
(1 − v 2 )k 2
1 dρ 2
= dv exp −ρ m +
(4π)2 2 −1 0 ρ 4
Z 1 2 2
1 k (1 − v )
= 2
dv log m2 +
(4π) 2 −1 4
Z 1
(1 − v 2 )k 2
1 2
= dv log 1 + + log m .
(4π)2 2 −1 4m2
We used the identities (18.28) and (18.29):
Z ∞
1
= dα exp(−αA), (12.28)
A 0
Z
dp 1
exp −(ap2 + bp) = exp b2 /4a .
(12.29)
(2π)4 (4π)2 a2
Then we changed the variables to
(1 − v) (1 + v)
α1 = ρ , α2 = ρ , (12.30)
2 2
1
so that α1 + α2 = ρ, − α1 + α2 = ρv, dα1 dα2 = ρdvdρ.
2
101
At the end we use the identity (18.31):
Z ∞X
dρ X
Ci e−ρAi = − Ci log(Ai ). (12.31)
0 i
ρ i
P
valid if Ci = 0.
Now the renormalized π is given by
Note that
π ren (0) = 0, (12.33)
which is our renormalization condition. Finally, we evaluate π ren using
Z
(1 + w)
log(1 − w2 )dw = w log(1 − w2 ) − 2w + log , 0 < w < 1. (12.34)
(1 − w)
We obtain
r
ren 2 1 1 1+θ k2
π (k ) = 2
log −2 , θ= . (12.35)
4(4π) θ 1−θ k2 + 4m2
102
where
K0 := p̂2 + m2 , K := − ∂ µ + ieAµ (x) ∂µ + ieAµ (x) + m2 .
(12.38)
are operators on L2 (R4 ).
Our goal is to compute the renormalized value of
E := Tr log(K) − log(K0 ) . (12.39)
We have
E =Tr log 1l + − ieAµ ∂µ − ie∂µ Aµ + e2 Aµ Aµ (p̂2 + m2 )−1
(12.40)
∞
X (−1)n+1 n
Tr − ieAµ ∂µ − ie∂µ Aµ + e2 Aµ Aµ (p̂2 + m2 )−1 .
=
n=1
n
103
Using
2
X 2
X
Ci = Ci m2i = 0 (12.44)
i=0 i=0
we can check that with this choice the sums used in the following computations
are integrable.
d4 q
Z
2 4qµ qν
2Πµν (k) = e
(2π)4 (q + 12 k)2 + m2i − i0 (q − 21 k)2 + m2
gµν gµν
− −
(q + 12 k)2 + m2 (q − 12 k)2 + m2
d4 q 4qµ qν − 2gµν (q 2 + 14 k 2 + m2 )
Z
2
= e
(2π)4 (q + 12 k)2 + m2 (q − 12 k)2 + m2
Z ∞ Z ∞
d4 q
Z
1 2
= e2 dα 1 dα 2 4q q
µ ν − 2gµν q 2
+ k + m 2
(2π)4 0 0 4
1
× exp −(α1 + α2 ) q 2 + k 2 + m2 − (α1 − α2 )qk
4
Z 4 Z ∞ Z ∞
d q 1 2
= e2 dα 1 dα 2 4∂ ∂
zµ zν − 2g µν ∂z
2
+ k + m 2
(2π)4 0 0 4
1
× exp −(α1 + α2 ) q 2 + k 2 + m2 − (α1 − α2 )qk + zq
4 z=0
2 Z ∞ Z ∞
e 1
2 1 2 2
= dα1 dα 2 4∂ ∂
zµ zν − 2g µν ∂z + k + m
(4π)2 0 0 (α1 + α2 )2 4
1 2
1
× exp −(α1 + α2 ) k 2 + m2 + (α1 − α2 )k − z
4 4(α1 + α2 ) z=0
∞ ∞
e2 (α1 − α2 )2
Z Z
= dα1 dα2 − (gµν k 2 − kµ kν )
(4π)2 0 0 (α1 + α2 )4
!
m2
α1 α2 2 1
−2gµν k + +
(α1 + α2 )4 (α1 + α2 )3 (α1 + α2 )2
α1 α2 2
× exp −(α1 + α2 )m2 − k
α1 + α2
=: (−gµν k 2 + kµ kν )2Πgi (k 2 ) + 2Πgd 2
µν (k ).
104
Let us compute the gauge dependent part of the vacuum energy tensor:
Πgd 2
µν (k )
Z ∞ Z ∞
e2
2 α1 α2 2
= dα 1 dα 2 exp −(α 1 + α 2 )m − k
(4π)2 0 0 α1 + α2
2 2
α1 α2 k 1 m
× gµν + +
(α1 + α2 )4 (α1 + α2 )3 (α1 + α2 )2
1 ∞
e2 (1 − v 2 )ρ 2 (1 − v 2 ) 2 m2
Z Z
2 1
= dv dρ exp −ρm − k gµν k + 2+
(4π)2 2 −1 0 4 4ρ ρ ρ
1 ∞
e2 (1 − v 2 )ρ 2
Z Z
d 1
=− 2
dv dρ exp −ρm2 − k gµν
(4π) 2 −1 0 dρ ρ 4
Z 1
e2 (1 − v 2 )ρ 2
1 ρ=∞
2
=− 2
dv exp −ρm − k gµν .
(4π) 2 −1 ρ 4 ρ=0
105
Note that the Fourier transform of the electromagnetic field is
Hence
1
− Fµν (k)F µν (k) = −k 2 |A(k)|2 + |kA(k)|2 . (12.48)
2
Thus the renormalized 1st order contribution to the vacuum energy is
Z
dk
E1ren = − Πren (k 2 )Fµν (k)F µν (k). (12.49)
2(2π)4
with the renormalization condition
106
Clearly, if the spectral shift function is well defined, then
Z
D(z) = ξ(s)(s − z)−1 ds, (12.54)
Then there exists an improved method of computing the spectral shift, which
is essentially an adaptation of the Pauli-Villars method from QFT. Choose
c0 , , . . . cn and Λ0 , . . . , Λn such that
c0 + · · · + cn = 0,
c0 Λ0 + · · · + cn Λn = 0,
...
c0 Λn0 + · · · + cn Λnn = 0.
Hence
n n
Dreg (z) :=Tr log Π (H + Λi − z)ci − log Π (H0 + Λi − z)ci
i=0 i=0
Z Xn
= ξ(s) ci (s + Λi − z)−1 ds. (12.58)
i=0
107
13 Scalar field with a masslike perturbation
13.1 Lagrangian and Hamiltonian formalism
Consider the Lagrangian density
1 1
L(x) = − ∂µ φ(x)∂ µ φ(x) − (m2 + κ(x))φ(x)2 , (13.1)
2 2
where R1,3 3 x 7→ κ(x) is a given function. In most of this subsection we will
assume that κ is Schwartz and m > 0. (19.1) leads to the equations
The theory of the Klein-Gordon equation with a variable mass is very similar
to the one with a constant mass. We can define the corresponding retarded and
advanced propagators as the unique distributional solutions of
satisfying
suppG∨/∧ ⊂ {x, y : x ∈ J ∨/∧ y)}.
The Pauli-Jordan function:
satisfies
suppGPJ ⊂ {x, y : x ∈ J(y)}.
and can be used to solve the initial value problem of (19.2):
Z
φ(t, ~x) = − ∂s GPJ (t, ~x, s, ~y ) φ(0, ~y )d~y
s=0
Z
+ GPJ (t, ~x, 0, ~y )π(0, ~y )d~y . (13.5)
The free (constant mass) field will be denoted by φfr , πfr . We can use the
same phase space YKG for the free and perturbed scalar particle, identifying
them for the initial condition at t = 0:
φfr (0, ~x) = φ(0, ~x), πfr (0, ~x) = π(0, ~x). (13.6)
108
Thus we can introduce the plane wave functionals a(k), a∗ (k) as in the free case,
since they depend only on the Cauchy data at time 0.
We easily obtain the Hamiltonian density from the Lagrangian:
1 2 1 ~ 2 1 2
H(x) = π (x) + ∂φ(x) + (m + κ(x))φ2 (x),
2 2 2
so that the full Hamiltonian generating the dynamics, first in the “Heisenberg
picture” and then in the “Schrödinger picture”, are
Z
H(t) = H(t, ~x)d~x
Z
1 2 1 ~ 2 1
= π (t, ~x) + ∂φ(t, ~x) + (m2 + κ(t, ~x))φ2 (t, ~x) d~x, (13.7)
2 2 2
Z
Sp 1 2 1 ~ 2 1 2
H (t) = π (~x) + ∂φ(~x) + (m + κ(t, ~x))φ2 (~x) d~x, (13.8)
2 2 2
109
We obtain
Hp
ȧ∗t (k)
∗
= at (k), HInt (t)
dk1 κ(t, −~k + ~k1 )
~
Z
= i q q
(2π)3 2ε(~k) 2ε(~k1 )
~ ~ ~ ~
× e−itε(k)−itε(k1 ) at (−k1 ) + e−itε(k)+itε(k1 ) a∗t (k1 ) ,
a∗0 (k) = a∗ (k).
more precisely,
" ∗
a∗t− (k1 )
# " #" #
at+ (k) Z pt+ ,t− (k, k1 ) qt+ ,t− (k, k1 )
= d~k1 .
at+ (k) qt+ ,t− (k, k1 ) pt+ ,t− (k, k1 ) at− (k1 )
(19.12) has a limit as t+ , −t− → ∞, which can be called the classical scattering
operator.
One can try to solve the equations of motion by iterations. The first iteration
is often (at least in the quantum context) called the Born approximation, and
it gives the following formula for the elements of (19.12):
t+
κ(s, −~k + ~k1 )
Z
~ ~
pBorn
t+ ,t− (k, k1 ) = δ(~k − ~k1 ) + i ds q q e−isε(k)+isε(k1 ) ,
t− (2π)3 2ε(~k) 2ε(~k1 )
t+
κ(s, −~k + ~k1 )
Z
~ ~
qtBorn
+ ,t−
(k, k1 ) = i ds q q e−isε(k)−isε(k1 ) .
t− (2π)3 2ε(~k) 2ε(~k1 )
13.3 Quantization
We are looking for quantum fields R1,3 7→ φ̂(x) satisfying
˙
with the conjugate field π̂(x) := φ̂(x) having the equal time commutators
110
and coinciding with the free field at time 0. The solution is given by putting
“hats” onto (19.5):
Z
φ̂(t, ~x) = − ∂s GPJ (t, ~x, s, ~y ) φ̂(0, ~y )d~y
s=0
Z
+ GPJ (t, ~x, 0, ~y )π̂(0, ~y )d~y . (13.15)
Clearly,
[φ̂(x), φ̂(y)} = −iGPJ (x − y).
We would like to check whether the classical scattering operator and the
classical dynamics are implementable in the Fock space for nonzero κ. By Thm
14.5, we need to check the Shale condition, that is, whether the off-diagonal
elements of (19.12) are square integrable. For simplicity, we will restrict our-
selves to the Born approximation; the higher order terms do not change the
conclusion.
The verification of the Shale condition is easier for the scattering operator.
Consider
Z ∞
κ(s, −~k + ~k1 ) ~ ~
Born
q∞,−∞ (k, k1 ) = i ds q q e−isε(k)−isε(k1 ) . (13.16)
−∞ (2π)3 2ε(~k) 2ε(~k1 )
This decays in ~k and ~k1 as any inverse power, and hence is square integrable on
R3 × R3 . Therefore the classical scattering operator is implementable.
Next let us check the implementability of the dynamics, believing again that
it is sufficient to check the Born approximation. We integrate by parts once:
qtBorn
+ ,t−
(k, k1 )
~ ~ ~ ~
−κ(t+ , −~k + ~k1 )e−it+ ε(k)−it+ ε(k1 ) + κ(t− , −~k + ~k1 )e−it− ε(k)−it− ε(k1 )
= q q
(2π)3 2ε(~k) 2ε(~k1 ) ε(~k) + ε(~k1 )
~ ~
∂s κ(s, −~k + ~k1 )e−isε(k)−isε(k1 )
Z t+
+ ds q q . (13.18)
t− (2π)3 2ε(~k) 2ε(~k1 ) ε(~k) + ε(~k1 )
Using that κ(s, ~k + ~k1 ) decays fast in the second variable, we see that (19.18)
can be estimated by
C
,
(ε(~k) + ε(~k1 ))2
111
which is square integrable. Therefore, the dynamics is implementable for any
t− , t+ .
By a similar computation we check that if we freeze t0 ∈ R, the dynamics
generated by the momentary Hamiltonian HInt (t0 ) is implementable.
Formally, the quantum Hamiltonians, first in the Heisenberg, then in the Schrödinger
picture, are given by
Z
1 2 1 2 1
Ĥ Hp (t) := π̂ (t, ~x) + ∂~ φ̂(t, ~x) + (m2 + κ(t, ~x))φ̂2 (t, ~x) d~x, (13.20)
2 2 2
Z
1 2 1 ~ 2 1 2
Ĥ(t) := π̂ (~x) + ∂ φ̂(~x) + (m + κ(t, ~x))φ̂2 (~x) d~x. (13.21)
2 2 2
We will treat the Schrödinger picture Hamiltonian, that is (19.21), as the stan-
dard one. It is expressed in terms of zero time fields. It is clear that it is
ill-defined. One could improve it by putting : · · · :, that is by the Wick order-
ing. We will see later on that even the Wick-ordered expression (19.21) does
not define an operator.
Formally (19.47) remains true if we add a time dependent constant C(t)
to (19.21). We will see that in order to define correct Hamiltonians Ĥ(t) this
constant has to be infinite, even after Wick ordering. We will obtain bounded
from below Hamiltonians Ĥren (t), however the vacuum will not be contained in
their form domain. Therefore, the condition (Ω|Ĥren (t)Ω) = 0 for all t, which
is equivalent to the Wick ordering, cannot be imposed.
The interaction Hamiltonian is formally given by
Z
1
ĤInt (t) = κ(t, ~x)φ̂2fr (t, ~x)d~x, (13.22)
2
d~k
Z
~ ~ ~ ~
φ̂fr (x) = q e−itε(k)+ik~x â(k) + eitε(k)−ik~x â∗ (k) , (13.23)
(2π)3 2ε(~k)
p
where (19.23) is taken from (8.59). Here is the Wick ordered interaction Hamil-
tonian:
d~k1 d~k2 κ(t, ~k1 + ~k2 ) −itε(~k1 )−itε(~k2 )
Z
1
:ĤInt (t): = q q e â(−k1 )â(−k2 )
2
(2π)3 2ε(~k1 ) 2ε(~k2 )
~ ~ ~ ~
+2eitε(k1 )−itε(k2 ) â∗ (k1 )â(−k2 ) + eitε(k1 )+itε(k2 ) â∗ (k1 )â∗ (k2 ) .
112
13.5 The vacuum energy
We would like to compute
!
Z
e−iE := (Ω|ŜΩ) = Ω|Texp − iĤInt (t)dt Ω . (13.24)
113
where p̂2 is the 4-dimensional Laplacian.
As we computed in Subsection 18.2, E E needs to be renormalized and equals
(see (18.35)):
∞
X
E E,ren = E2E,ren + EnE , (13.31)
n=3
Z
dk E,ren 2
E2E,ren = − |κ(k)|2 π (k ), (13.32)
(2π)4
r
1 1 1+θ k2
π E,ren (k 2 ) = log − 2 , θ= . (13.33)
4(4π)2 θ 1−θ k 2 + 4m2
The Wick rotation consists in replacing x4 with ix0 everywhere except for
κ(x), where we do not change anything. More precisely, we replace x4 with zx0 ,
where z is a complex parameter which is continued as eiα with α ∈ [0, π2 ]. Thus
the Euclidean quantities are treated as functions F (z), where F (z) z=1 is the
Euclidean value and F (z) z=i the Wick rotated value.
This implies replacing p4 with −ip0 . Moreover, the Euclidean x2 and p2
are replaced with Minkowskian x2 + i0 and p2 − i0. The Euclidean Lebesgue
measures dx and dk are replaced in Rthe Minkowski space by idx, resp. −idk.
Consequently, the Euclidean action LE (x, φ, φ,µ )dx becomes after the Wick
E
rotation −i L(x, φ, φ,µ )dx. Thus e−E (z) z=i = e−iE , and consequently
R
114
Taking into account (19.33), we obtain
√
ren 2 1 1 1+θ k2
π (k ) = 2
log −2 , θ= √ , 0 < k 2 ; (13.39)
4(4π) θ 1−θ k + 4m2
2
√
1 2 −k 2
= 2
arctan θ − 2 , θ = √ , −4m2 < k 2 < 0; (13.40)
4(4π) θ k 2 + 4m2
√
1 1 θ+1 −k 2
= log − iπ − 2 , θ = √ , k 2 < −4m2 . (13.41)
4(4π)2 θ θ−1 −k 2 − 4m2
Here is the calculation. First we assume that k 2 > 0. Then we just take the
Euclidean value. Then we use analytic continuation, remembering
√
that k 2 may
2 k 2
have negative imaginary part. As k decreases, θ := √k2 +4m2 first varies from
1+θ
1 to 0, then from 0 to −i∞, finally, from ∞ to 1. Therefore, 1−θ first varies
from ∞ to 1, then goes over the lower semicircle, finally, from −1 to −∞. Next
1+iy
we use log 1−iy = 2i arctan y, and for y < 0, log(y − i0) = log |y| − iπ.
(19.39) corresponds to a spatial transfer of energy-momentum. In (19.40)
the transfer is time-like, but the energy is below the 2-particle threshold. In
(19.41) it is above this threshold, and π ren acquires a nonzero imaginary part
responsible for the decay of the vacuum.
Here are operator-theoretic formulas for the renormalized vacuum energies:
i 2
E2ren = − Tr κ(x)(−2 + m2 − i0)−1 − κ(x)2 (−2 + m2 − i0)−2 ,
4
i
ren
= − Tr log 1l + κ(x)(−2 + m2 − i0)−1 − κ(x)(−2 + m2 − i0)−1
E
2
1
+ κ2 (x)(−2 + m2 − i0)−2 . (13.42)
2
where C is the constant responsible for the Wick ordering and π(0) is the 2nd
order renormalization. Note that both infinite quantities are quite well behaved
– they depends locally on the interaction, and therefore the renormalization
preserves the Einstein causality. This manifests itself in the identity
115
have
Ŝ ren (t3 , t2 )Ŝ ren (t2 , t1 ) = Ŝ ren (t3 , t1 ). (13.45)
Thus we get a unitary evolution given by
116