Mathematical Foundations of Quantum Field Theory
Mathematical Foundations of Quantum Field Theory
Albert Schwarz
University of California at Davis, USA
World Scientific
NEW JERSEY • LONDON • SINGAPORE • BEIJING • SHANGHAI • HONG KONG • TAIPEI • CHENNAI • TOKYO
Printed in Singapore
v
b2530 International Strategic Relations and China’s National Security: World at the Crossroads
Preface
vii
March 30, 2020 11:10 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-00-fm page viii
Preface ix
Contents
Preface vii
Introduction xvii
xi
March 30, 2020 11:10 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-00-fm page xii
Contents xiii
Contents xv
Appendix 375
A.1 Hilbert spaces . . . . . . . . . . . . . . . . . . . . . 375
A.2 Systems of vectors in a pre-Hilbert
vector space . . . . . . . . . . . . . . . . . . . . . . 376
A.3 Examples of function spaces . . . . . . . . . . . . . 377
A.4 Operations with Hilbert spaces . . . . . . . . . . . . 381
A.5 Operators on Hilbert spaces . . . . . . . . . . . . . 383
A.6 Locally convex linear spaces . . . . . . . . . . . . . 392
A.7 Generalized functions (distributions) . . . . . . . . 393
A.8 Eigenvectors and generalized eigenvectors . . . . . . 403
A.9 Group representations . . . . . . . . . . . . . . . . . 404
Bibliography 409
Index 413
b2530 International Strategic Relations and China’s National Security: World at the Crossroads
Introduction
xvii
March 30, 2020 11:10 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-00-fm page xviii
Introduction xix
this review, while others could read it after finishing the book or in
parallel with reading the book. An advanced reader can jump to the
last chapter after reading the review.
Hamiltonian formalism
The equations of motion of a three-dimensional non-relativistic
particle in a potential field U (~x) have the form
d~
p
= −∇U,
dt
March 30, 2020 11:10 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-00-fm page xxi
Introduction xxi
~
where p~ = m dx
dt stands for the momentum of the particle. To solve
these equations (i.e. to find the trajectory of the particle), we should
know the initial data: the coordinates and the momenta at some
moment of time. One says that the coordinates and the momenta
specify the state of our particle at the given moment and that the
equations of motion allow us to find the state of the particle at any
moment if we know it at one of the moments. The equations of motion
can be written in the form
d~
p ∂H
=− ,
dt ∂~x
~
dx ∂H
= ,
dt ∂~
p
2
where H = pm + U (~x) is called Hamiltonian function (one can
say the Hamiltonian function is the energy expressed in terms
of momenta and coordinates). Similar equations are valid for any
mechanical system, but the number of degrees of freedom (the
number of coordinates and momenta) and the Hamiltonian function
can be arbitrary. This gives the so-called Hamiltonian formalism
of mechanics. (In Lagrangian formalism, the state is specified by
coordinates and velocities.)
In Hamiltonian formalism, the (pure) state of a classical mechan-
ical system (at the time t) is characterized by 2n numbers: p =
(p1 , . . . , pn ) (generalized momenta) and q = (q 1 , . . . , q n ) (gener-
alized coordinates). (Together, these numbers specify a point of
2n-dimensional space called the phase space of the system).
More generally, we can define a state as a probability distribution
on the phase space. The set D of probability distributions is a convex
set, the pure states can be identified with extreme points of this set.
Every state can be considered as a mixture of pure states. (The
mixture of states ω1 , . . . , ωn with probabilities p1 , . . . , pn is the state
p1 ω1 +· · ·+pn ωn . If states are labeled by continuous parameter R λ ∈ Λ,
one defines the mixture of the states as an integral ω(λ)ρ(λ)dλ
where ρ(λ) stands for the density of the probability distribution
on Λ. Note that the definition of mixture can be used for any
convex set.)
March 30, 2020 11:10 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-00-fm page xxii
1
The multiplication and the Poisson bracket specify the structure of Poisson
algebra on the space of functions on the phase space (see the definition of Poisson
algebra in Section 1.3). This means that the phase space is a Poisson manifold.
Moreover, it is a symplectic manifold, i.e. the Poisson structure is non-degenerate.
We have considered the Hamiltonian formalism on a flat symplectic manifold, but
it can be considered on any symplectic manifold.
March 30, 2020 11:10 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-00-fm page xxiii
Introduction xxiii
d
f (p(t), q(t)) = {f, H}. (I.5)
dt
The extreme points of this set are called pure states. Every state is
a mixture of pure states.
Let us denote by Aut the group of automorphisms of the algebra
A commuting with involution. This group naturally acts on states.
In quantum system, the state depends on time and this dependence
can be described by the evolution operator U (t) transforming ω(0)
into ω(t). This is the so-called Schrödinger picture; it is equivalent
to Heisenberg picture where the elements of A depend on time, but
the states do not: ω(t)(A) = ω(A(t)).
The evolution operator satisfies the equation
dU
i = H(t)U (t), (I.6)
dt
Introduction xxv
where A1 , . . . , An ∈ A.
The Green functions in the state ω are defined by the formula
Introduction xxvii
1 ∂A ∂B
{A, B} = σ kl (u) k l . (I.7)
2 ∂u ∂u
Introduction xxix
Stationary states
A state that does not depend on time is called stationary state. In
what follows, we work in the formalism where the states are described
by density matrices in Hilbert space E. The state represented by
a density matrix K is stationary if K obeys HK = 0, i.e. if
K commutes with the Hamiltonian Ĥ. (Recall that H acts as a
commutator with Ĥ.) An important particular case of a stationary
state is the Gibbs state K = Z −1 e−β Ĥ where Z = T re−β Ĥ . This state
corresponds to the equilibrium state with the temperature T = β −1 .
It tends to the ground state as T → 0.
Let us assume that the operator Ĥ has discrete spectrum with
orthonormal basis φn of eigenvectors with eigenvalues En . Then it
is convenient to work in representation where Ĥ is represented by
a diagonal matrix with entries En (it is called Ĥ-representation).
In this representation, the eigenvectors of the operator H in the
space L are matrices ψmn having only one non-zero entry equal
March 30, 2020 11:10 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-00-fm page xxx
Adiabatic approximation
Let us consider the case of slowly varying Hamiltonian Ĥ(t). We
will assume that the energy levels En (t) are distinct and vary
continuously with t, corresponding eigenvectors will be denoted by
φn (t). We assume that these eigenvectors constitute an orthonormal
system. Then it is easy to prove that in the first approxima-
tion the evolution operator Û (t) transforms eigenvector to eigen-
vector
dαn (t)
Û (t)φn (0) = e−iαn (t) φn (t), = En (t). (I.8)
dt
To verify (I.8), we check that the RHS satisfies the equation of motion
up to terms that are small for slow varying Ĥ(t) (see Section 4.3 for
more details).
Let us introduce the operators ψmn (t) by the formula ψmn (t)x =
hx, φn (t)iφm (t). (These operators are eigenvectors of the operator
H(t) in the space L.) Applying (I.8) or analyzing directly the
evolution operator U (t) in L, we obtain that U (t) transforms
eigenvector into eigenvector
dβmn (t)
U (t)ψmn (0) = e−iβmn (t) ψmn (t), = Em (t) − En (t) (I.9)
dt
(this equation is true up to terms that can be neglected for slowly
varying Hamiltonian). Note that βmm does not depend on t.
March 30, 2020 11:10 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-00-fm page xxxi
Introduction xxxi
Decoherence
Let us consider a quantum system (atom, molecule, etc.) described
by a Hamiltonian Ĥ with simple discrete spectrum. We assume that
the system is “almost closed” in the following sense: the interaction
with the environment can be described as an adiabatic change of
the Hamiltonian Ĥ. Let us consider the evolution operator U (T )
assuming that Ĥ(0) = Ĥ(T ). If the Hamiltonian is time-independent,
then for the eigenvector φn with the eigenvalue En , we can say that
Û (T )φn = Cn (T )φn where Cn (T ) = e−iEn T . If the Hamiltonian is
slowly changing, we have the same formula with Cn (T ) calculated
from (I.8). Hence, if we started with pure stationary state, we remain
in the same state. Similarly, U (T )ψmn = Cmn (T )ψmn and the phase
factor is constant for m = n.
It is natural to assume that the environment is random (the
time-dependent Hamiltonian depends on some parameters λ ∈ Λ
with some probability distribution on Λ), then for m 6= n, we have
a random phase factor Cmn (λ, T ). If we start with density matrix
P
K= kmn ψmn (with matrix entries kmn in Ĥ-representation), then
P
the density matrix Kλ (T ) is equal to Cmn (λ, T )kmn ψmn , i.e. the
matrix entries acquire phase factor Cmn (λ, T ). Now, we should take
the mixture K̄(T ) of states Kλ (T ) (this means that we should take
the average of phase factors). It is obvious that non-diagonal entries
of K̄(T ) are smaller by absolute value than corresponding entries of
K. Imposing some mild conditions on the probability distribution on
Λ, one can prove that the non-diagonal entries of K̄(T ) tend to zero
as T → ∞. In other words, the matrix K̄(T ) tends to a diagonal
matrix K̄ having the same diagonal entries as K. (See Schwarz and
Tyupkin (1987) and Schwarz (2019a) for more details.)
The matrix K̄ can be considered as a mixture of pure states
corresponding to the vectors φn with probabilities knn .
This phenomenon is known as decoherence.
Introduction xxxiii
where
R f stands for any piecewise continuous function and f (A) =
f (a)ρ(a)da denotes the mean value of f (A) with respect to the
probability distribution ρ(a)da. (It is also called the expectation
value of f (A).) If the operator  has simple discrete spectrum, (I.10)
is equivalent to the formulas for probabilities we gave in this case.
If Â1 , . . . , Ân are commuting self-adjoint operators, one can define
the joint probability distribution of corresponding physical quantities
A1 , . . . , An using the formula
Integrals of motion
One can work either in the Schrödinger picture where states are
time-dependent, but observables do not depend on time, or in the
Heisenberg picture, where states do not depend on time, but the
observables do. These pictures are equivalent:
dÂ
i = [Ĥ, Â].
dt
March 30, 2020 11:10 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-00-fm page xxxiv
Introduction xxxv
p̂2 q̂ 2 1
Ĥ = + = a∗ a + .
2 2 2
Noting that Ĥa∗ = a∗ (Ĥ + 1), Ĥa = a(Ĥ − 1), we obtain that the
operator a∗ transforms an eigenfunction φ with eigenvalue E in an
eigenfunction with eigenvalue E + 1. Similarly, the operator a either
sends φ to zero (if φ is the ground state) or to an eigenfunction with
eigenvalue E − 1. Using this fact, we can check that the ground state
θ has energy 21 and the states √1n! (a∗ )n θ constitute an orthonormal
basis consisting of eigenfunctions of Ĥ with eigenvalues n + 21 .
In appropriate coordinates, the Hamiltonian of multidimensional
quantum oscillator can be considered as a sum of non-interacting
one-dimensional oscillators:
X ωk
Ĥ = ωk a∗k ak + ;
2
again applying many times operators a∗k to the ground state θ, we
obtain a basis of eigenfunctions.
We see that it is convenient to use the operators a∗k , ak in the
analysis of excitations of the ground state. In quantum field theory,
we are interested first of all in the excitations of ground state; this
is one of the many reasons why these operators are so useful.
Let us consider now the Weyl algebra with infinite number of
generators. This notion can be made precise in different ways. We
March 30, 2020 11:10 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-00-fm page xxxvi
Introduction xxxvii
Here, Γk1 ,...,km ,l1 ,...,ln = Γ̄ln ,...,l1 ,km ,...,k1 (this condition guarantees that
h is self-adjoint).
In the case of infinite number of generators, one can modify this
construction to obtain other infinitesimal automorphisms. Namely,
we can consider (I.18) as a formal expression; if the expressions
[ak , h], [a∗k , h] can be regarded as elements of A, these formulas specify
a derivation of algebra. This happens if for every k, there exists
only a finite number of summands in (I.18) where one of the indices
is equal to k. (Then only a finite number of terms survives in the
commutator.) To check that the derivation specifies a one-parameter
family of automorphism, we should verify that the equations of
dak a∗k
motion i dt = [ak , h], i dt = [a∗k , h] have a solution. If h is a quadratic
hamiltonian, the equations of motion are linear and the solution is
the same as in classical theory. If h is a sum of quadratic Hamiltonian
and a summand multiplied by a parameter g, then it is easy to
prove that the equations of motion can be solved in the framework
of perturbation theory with respect to g.
March 30, 2020 11:10 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-00-fm page xxxviii
Introduction xxxix
Introduction xli
Introduction xliii
Concluding remarks
In conclusion, a couple of general words about quantum theory.
A quantum mechanical system is specified by a Hamiltonian
(infinitesimal generator of time evolution) and a QFT is specified
by a Hamiltonian and a momentum operator (generator of spatial
translations). Knowing these data, we can define a notion of particle
in QFT; a particle can be interpreted as an elementary excitation of
ground state. Under certain conditions, we can define the scattering
of particles.
Note that string field theory can be regarded as a quantum field
theory in the sense of this definition.
Quantum mechanics is a deterministic theory. This means that
knowing the state at some moment and the Hamiltonian governing
the evolution of the state, we can in principle predict the state at
any moment — precisely as in classical mechanics. The probabilities
in quantum mechanics can be explained in the same way as in
classical statistical physics — they come from a random environment
(see Sections “Decoherence” and “Observables and probabilities” on
page xxxi for an explanation of how probabilities can be obtained
from decoherence).2
The notion of particle in quantum field theory is a secondary
notion. There exists no physical difference between elementary and
composite particles. (There are theories that can be represented in
two different ways: elementary particles of one approach or com-
posite particles of another approach.) Probably, the most revealing
illustration of properties of quantum particles is given by analogy
with nonlinear scattering in classical field theory. Such a theory can
have particle-like solutions, for example, solitons. (A soliton is a finite
energy solution of the form s(x − vt). It can be visualized as a bump
moving with constant speed that does not change its form. If we have
a bump moving with constant average speed, but the form of the
bump changes with time, we can talk about generalized soliton or, in
different terminology, about particle-like solution.) It seems that for
2
We do not consider subtle questions of measurement theory. The above
statement means only that the standard formulas for probabilities can be obtained
from decoherence.
March 30, 2020 11:10 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-00-fm page xliv
Introduction xlv
Notation
A∪B the union of sets A and B;
A∩B the intersection of sets A and B;
A×B the cross product of sets A and B (the set of all (a, b)
with a ∈ A and b ∈ B);
En the n-dimensional Euclidean space;
L2 (X) the Hilbert space of square-integrable complex
functions on the measure space X;
hx, yi the scalar multiplication of elements of (pre-)Hilbert
space (for Euclidean space elements x, y ∈ E n , this
quantity will sometimes be denoted xy);
S(E )n the space of smooth functions of n variables that
rapidly decay; by smooth, we mean infinitely
differentiable and by rapidly decaying, we mean
faster than any power;
Ω the cube with length L edges in E 3 , satisfying the
equations 0 ≤ x ≤ L, 0 ≤ y ≤ L, 0 ≤ z ≤ L;
TΩ a set of vectors with k = 2πnL , where n is an integer
vector;
φk the orthonormal basis in the space L2 (Ω), formed by
the functions φk (x) = eixk ;
H1 + H2 a direct sum of Hilbert spaces;
H1 + · · · + a direct sum of an infinite sequence of Hilbert spaces;
Hn + · · ·
F (H) the Fock space built with the Hilbert space H (if
H = L2 (X), then F (L2 (X)) = F0 + F1 + · · · +
Fn + · · · , where Fn consists of square-integrable
symmetric functions in n variables in the case of
bosons; in the case of fermions, the functions are
antisymmetric);
March 30, 2020 11:10 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-00-fm page xlvi
Chapter 1
1
March 26, 2020 11:53 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch01 page 2
equation:
∂U (t, t0 )
i = H(t)U (t, t0 ), (1.1)
∂t
with the initial condition U (t0 , t0 ) = 1. Here, H(t) is a self-adjoint
operator called the Hamiltonian operator of the quantum system.
The Schrödinger equation can also be written in the form
dψt
i = H(t)ψt , (1.2)
dt
though less precisely, since the operator H(t) may not be defined on
the full Hilbert space R, while U (t, t0 ) is defined everywhere.
It is easy to check that U ∗ (t, t0 ) = U (t0 , t); using the group
property, we can see that the operator is unitary (to be rigorous, we
should assume that the evolution operator is unitary and prove that
the Hamiltonian is self-adjoint using the unitarity of the evolution
operator).
If the Hamiltonian does not depend on time explicitly, we can
write U (t, t0 ) in the form U (t, t0 ) = exp(−iH(t − t0 )).
For a non-relativistic one-dimensional particle, the state space R
can be taken to be L2 (E 1 ), the space of square-integrable functions
ψ(x) of one variable x, where −∞ < x < ∞. The Hamiltonian
1 d2
operator in this case can be written as H(t) = − 2m dx2
+ V̂(x, t)
where V̂(x, t) is the multiplication operator by the function V(x, t)
(t plays the role of a parameter here).1 This Hamiltonian describes
the non-relativistic particle with mass m, moving in the field with
time-dependent potential V(x, t).
1
It is easy to check that V̂(x, t) = V(x̂, t), where x̂ is the operator of
multiplication by x.
March 26, 2020 11:53 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch01 page 3
Ari = ari .
To each state vector ψ, we can assign the sequence (c1 , . . . , ci , . . .)
of coefficients that arise from the decomposition of the vector ψ with
respect to the basis ri , consisting of the eigenvectors of operator A.
This construction specifies an isomorphism of the space R and the
space l2 . We call this isomorphism an A-representation and the
sequence (c1 , . . . , ci , . . .), we call the A-representation of the vector ψ.
The A-representation of the vector ψ allows us to easily calculate the
probabilities of values of the observable a in the state ψ.
The notion of an A-representation can also be defined when the
operator A has continuous spectrum. Moreover, it can be generalized
to the case when we are dealing with a family of commutative self-
adjoint operators. Namely, if A1 , . . . , Ak are self-adjoint, pairwise
commuting operators, then an (A1 , . . . , Ak )-representation is an
isomorphism of Hilbert spaces R and L2 (M ) in which the operators
A1 , . . . , Ak are transformed to the multiplication operators by the
functions a1 (m), . . . , ak (m) (here M is a measure space, with m ∈ M ).
In the (A1 , . . . , Ak )-representation of the vector ψ ∈ R, one can easily
March 26, 2020 11:53 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch01 page 5
Chapter 2
9
March 26, 2020 12:16 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch02 page 10
For brevity, we will often use the vector operators: r̂ = (x̂, ŷ, ẑ),
p̂ = (p̂x , p̂y , p̂z ), M̂ = r̂ × p̂ = (M̂x , M̂y , M̂z ).
For the Heisenberg operators r̂t and p̂t , the equations of motion
are
dr̂t p̂t dp̂t dV
= ; = − (r̂t ).
dt m dt dr
Up to this point, the states for one particle were written as ψ(r)
with r ∈ E 3 . If α is a unitary operator, transforming the space
L2 (E 3 ) to the space L2 (M ), where M is some measure space (an
isomorphism of the space L2 (E 3 ) and the space L2 (M )), then we
can equivalently represent the state vectors ψ ∈ L2 (E 3 ) as vectors
ψ̃ = αψ ∈ L(M ). For every operator A acting on the space L2 (E 3 ),
there will be a corresponding operator à = αAα−1 acting on L2 (M ),
and from hf (Ã)ψ̃, ψ̃i = hf (A)ψ, ψi, it follows that the probabilities
obtained from the operator à and the state vector ψ̃ will be the same
as those obtained from A and ψ.
Changing the representation of a state vector is useful in many
cases. For example, let α be the Fourier transform
Z
−3/2
ψ̃(p) = (2π) exp(−ipr)ψ(r)dr
ψ1 (r)
ψ (r)
2
ψ= .. ,
.
ψk (r)
p̂2 x̂2 1 d 1
H= + =− + x̂2 .
2 2 2 dx2 2
March 26, 2020 12:16 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch02 page 18
First, let us find the Heisenberg operators p̂t and x̂t from the
Heisenberg equations
dp̂t dx̂t
= −x̂t ; = p̂t .
dt dt
This system of operator equations is linear and therefore it can be
solved precisely as the corresponding system of numerical equations.
One of the possible ways to solve it is based on the introduction of
auxiliary operators
1 1
â = √ (x̂ + ip̂); â+ = √ (x̂ − ip̂).
2 2
The equations for the Heisenberg operators ât and â+
t have a very
simple form
dât dâ+
= −iât ; t = iâ+
t ,
dt dt
hence â+ +
t = â exp(it); ât = â exp(−it). Using these relations, we
can immediately obtain the formulae x̂t = √12 (ât + â+
t ) and p̂t =
√1 (â+ − ât ).
2 t
The operators â+ and â are very convenient for solving problems
related to the harmonic oscillator.
Note, first of all, that the Hamiltonian can be expressed in terms
of these operators by the formula H = â+ â + 1/2. Second, note
that the commutation relations with the Hamiltonian have the form
[H, â+ ] = â+ , [H, â] = −â, and their commutator is [â, â+ ] = 1.
Let us find the stationary states of the Hamiltonian H. We use
the following statement: If Hφ = Eφ, then H(âφ) = (E−1)âφ (hence
if φ is a stationary state and âφ 6= 0, then âφ is also a stationary
state). This statement follows from the relations H(âφ) = â(H −
1)φ = (E − 1)âφ.
Similarly, H(â+ φ) = â+ (H + 1)φ = (E + 1)â+ φ.
Note that the ground state φ0 should satisfy the relation âφ0 = 0
(otherwise, âφ0 is a stationary state with lower energy). Solving
the equation âφ0 = √12 (x + dx d
)φ0 = 0, we obtain that φ0 =
π −1/4 exp(− 21 x2 ) (the constant π −1/4 comes from the normalization
March 26, 2020 12:16 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch02 page 19
1 = hφn , φn i
2 1
= γn H+ φn−1 , φn−1
2
= nγn2 ,
√
hence, γn = 1/ n and cn = (n!)−1/2 .
The orthonormal system of stationary states φn = √1n! â+n φ0 ,
with energy levels En = n + 1/2, exhausts all stationary states. One
can check this by proving the completeness of functions φn in the
space L2 (E 1 ). However, one can give a more direct proof. Let φ
denote a stationary state of H. Let us denote by n then, the minimal
number satisfying ân φ0 = 0 (such numbers necessarily exist because
the eigenvalues of the Hamiltonian H are bounded from below).
Then, ân−1 φ = λφ0 , where λ 6= 0. Applying the operator (â+ )n−1
to this equation, after some easy calculations, we obtain that φ is
proportional to φn−1 .
p̂2 2 2
The more general Hamiltonian H = 2m + mω2 x̂ can be reduced
to the Hamiltonian we have considered by means of changing the
system of units. One can, however, use the same consideration in
March 26, 2020 12:16 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch02 page 20
√ 1 √
1 ip̂ + ip̂
â = √ mωx̂ + √ ; â = √ mωx̂ − √ .
2 mω 2 mω
These operators are also related by the formula [â, â+ ] = 1, however,
the expression of the Hamiltonian in terms of these operators has
the form H = ω(â+ â + 1/2), hence [H, â] = −ωâ, [H, â+ ] = ωâ+ .
The dependence of operators ât and â+ t on time has the form ât =
+ +
â exp(−iωt), ât = â exp(iωt). The stationary states can be written
in terms of the ground state φ0 by the formula φn = (n!)−1/2 (â+ )n φ0 ,
with the energies equal to En = (n + 1/2)ω.
Adding to the quadratic potential energy some higher order
terms with respect to x̂, we obtain the Hamiltonian of anharmonic
oscillators. It is convenient to express the Hamiltonian of anharmonic
oscillators in terms of the operators â+ and â. For example, if
p̂2 2 2
H = 2m + mω2 x̂ + αx̂3 + β x̂4 , then its expression in terms of â+
and â looks as follows:
ω
H= + ωâ+ â + γ(â+ + â)3 + δ(â+ + â)4 ,
2
n
X ∂2 X
H=− αj + ki,j x̂i x̂j
j=1
∂x2j i,j
P
(here, αj > 0 and the quadratic form kij xi xj is positively definite).
In this way, we can write down, for example, the Hamiltonian for a
system of oscillators coupled by means of elastic forces. By means
P
of a linear change of variables ξi = j aij xj (the coordinates ξi are
called normal coordinates), the Hamiltonian can be written in the
March 26, 2020 12:16 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch02 page 21
following form:
n n
1 X ∂2 1 X 2 ˆ2
H=− 2 + ωi ξi .
2 ∂ξi 2
i=1 i=1
Chapter 3
23
March 26, 2020 13:43 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch03 page 24
formulas
1 X
Ps ψ(ξ) = ψ(π(ξ)),
n! π
1 X
Pa ψ(ξ) = (−1)γπ ψ(π(ξ))
n! π
1
More precisely, the definition of the operator Hn can be given as follows: Hn is
a self-adjoint operator on Bn that transforms functions φ1 (ξ1 ) . . . φn (ξn ), where
the
Pn functions φi belong to the domain of the operator H1 , into the function
i=1 φ1 (ξ1 ) . . . φi−1 (ξi−1 )(H1 φi )(ξi )φi+1 (ξi+1 ) . . . φn (ξn ). Such an operator is
unique.
2
In other words, in the system of non-interacting identical fermions, two fermions
cannot be found in the same state. This statement is called the Pauli exclusion
principle.
March 26, 2020 13:43 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch03 page 27
Here, F0s and F0a are spaces of constant functions (i.e. one-
dimensional spaces).
The spaces F s and F a are called Fock spaces. The elements of
Fock spaces represent the states of a system of identical bosons
or fermions in the case when the number of particles is not fixed.
Speaking of bosons and fermions simultaneously, we will use the
notation F for one of the spaces F s or F a . Vectors in the space
F s or F a can be considered as sequences (f0 , f1 , . . . , fn , . . . ), where
fn ∈ Fns (or correspondingly fn ∈ Fna ), satisfying the condition
P∞ 2
n=0 kfn k < ∞. In other words, the elements of the space F can
be considered as column vectors
f0
f1 (ξ1 )
..
f = .
fn (ξ1 , . . . , ξn )
..
.
of symmetric (antisymmetric) functions fn (ξ1 , . . . , ξn ) as entries; they
should satisfy the condition3
∞ Z
X
|fn (ξ1 , . . . , ξn )|2 dξ1 . . . dξn < ∞.
n=0
The spaces Fn are naturally embedded in the Fock space F (to the
vector f ∈ Fn , there corresponds a sequence (f0 , . . . , fn , . . . ) ∈ F ,
where fn = f, fk = 0 for k 6= n).
Let us introduce the operator of the number of particles
N = N0 + N1 + · · · + Nn + · · · ,
3
Such column vectors are called Fock states.
March 26, 2020 13:43 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch03 page 29
The symbols a(x), a+ (x) are not operators. The integral f (x)
R
It follows from the relations (3.10) and (3.12) that the Hamiltonian H
of the system of non-relativistic identical particles can be represented
as a sum of operators of the form (3.9) and (3.11). Namely, in the
coordinate representation,
XZ
+ ∆
H= a (x, s) − a(x, s)dx
s
2m
X X1Z
+
+ V(x)a (x, s)a(x, s)dx + W(|x1 − x2 |)a+ (x1 , s)
s s
2
X Z k2
H= a+ (k, s)a(k, s)dk
s
2m
X
+ Ṽ(k1 − k2 )a+ (k1 , s)a(k2 , s)dk1 dk2
s
X1Z
+ W̃(k1 − k4 )δ(k1 + k2 − k3 − k4 )
s
2
can define an operator on Fock space only in the case when the
function α is square integrable. Namely, by the general definition,
the operator defined by the expression (3.8) should transform the
sequence (φ0 , . . . , φn , . . . ) ∈ D into the sequence (ψ0 , . . . , ψn , . . . ),
March 26, 2020 13:43 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch03 page 36
where
s
n!
ψn (ξ1 , . . . , ξn ) = P φn−m (ξ1 , . . . , ξn−m )α(ξn−m+1 , . . . , ξn ).
(n − m)!
If the function α is square integrable, then the functions ψn are also
square integrable, hence the expression (3.18) specifies an operator
A on the whole space D and this operator transforms D into
itself. However, if the function α is not square integrable, then
the function ψn may be square integrable only under the condition
φn−m ≡ 0. This means that the domain of the operator specified by
the expression (3.18) contains only the zero vector; in other words,
the expression (3.18) does not specify any operator because the
domain of the operator should be dense in Fock space.
Let us consider the Fock space F (L2 (E 3 )), constructed starting
with the measure space E 3 , and operators defined by expressions of
the form
m+n≤s
X Z
A= Am,n (k1 , . . . , km |l1 , . . . , ln )
m,n
× δ(k1 + · · · + km − l1 − · · · − ln )
× a+ (k1 ) . . . a+ (km )a(l1 ) . . . a(ln )dm kdn l (3.19)
(operators of this form commute with the momentum operator). Let
us suppose that the functions Am,n belong to the space S(E 3(m+n) )
of smooth, rapidly decreasing functions (faster than any power
function). Let us single out the subspace S∞ ⊂ F that consists
of sequences (φ0 , . . . , φk , . . . ) ∈ D, obeying φk ∈ S(E 3k ). In what
follows, it will be convenient to consider operators on the space
F (L2 (E 3 )) only on the set S∞ . By means of the relation (3.6), it is
easy to check that in the case when Am,0 ≡ 0, the operator specified
by the expression (3.19) is well defined on all elements of the set
S∞ . This operator transforms every sequence in S∞ into a sequence
belonging to the same set. If one of the functions Am,0 is non-zero,
then the expression (3.19) cannot define an operator on Fock space
because the function Am,0 (k1 , . . . , km )δ(k1 + · · · + km ) is not square
integrable.
March 26, 2020 13:43 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch03 page 37
Chapter 4
∂S(t, t0 )
i = g Ṽ (t)S(t, t0 ), (4.1)
∂t
where Ṽ (t) = exp(iH0 t)V (t) exp(−iH0 t); the initial condition is
specified in the form S(t0 , t0 ) = 1.
In some physical situations, the operator H0 can be considered as
a free Hamiltonian (i.e. it describes non-interacting particles), and
the operator H − H0 = gV (t) is the interaction Hamiltonian. This
terminology is also commonly used in quantum field theory. However,
39
March 26, 2020 14:5 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch04 page 40
Substituting this series into the equation for the operator S(t, t0 ), we
obtain the recurrence relation
∂Sn (t, t0 )
i = Ṽ (t)Sn−1 (t, t0 ).
∂t
Using the initial condition S(t0 , t0 ) = 1, we see that S0 (t0 , t0 ) = 1
and Sn (t0 , t0 ) = 0 for n ≥ 1, hence
1 t
Z
Sn (t, t0 ) = Ṽ (τ )Sn−1 (τ, t0 )dτ. (4.2)
i t0
one can solve the integral equation using the method of iterations.
From (4.2) or from the integral equation, we can conclude that
1 t
Z
S1 (t, t0 ) = Ṽ (τ )dτ,
i t0
2 Z t Z τ1
1
S2 (t, t0 ) = dτ1 dτ2 Ṽ (τ1 )Ṽ (τ2 ),
i t0 t0
..
.
n Z t Z τ1 Z τn−1
1
Sn (t, t0 ) = dτ1 dτ2 · · · dτn Ṽ (τ1 ) . . . Ṽ (τn ).
i t0 t0 t0
March 26, 2020 14:5 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch04 page 41
1 1 n t
Z Z t
Sn (t, t0 ) = ··· T (Ṽ (τ1 ) . . . Ṽ (τn ))dτ1 . . . dτn .
n! i t0 t0
∂Sα (t, t0 )
i = exp(−α|t|)Ṽ (t)Sα (t, t0 )
∂t
Indeed, if this condition is not satisfied, one can always find a phase
factor exp(iα(g)), such that the vector φ̃g = exp(−iα(g))φ gE satisfies
R D
1 g dφλ
the condition: namely, we can choose α(g) = i 0 φλ , dλ dλ (the
function α(g) is real, since from the relation hφλ , φλ i = 1, one can
March 26, 2020 14:5 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch04 page 44
obtain that
dφλ dφλ dφλ dφλ
φλ , + , φλ = φλ , + φλ , = 0).
dλ dλ dλ dλ
By differentiating in g the relation
H(g)φg = E(g)φg ,
we obtain
dφg dH(g) dE(g)
(H(g) − E(g)) =− φg + φg . (4.5)
dg dg dg
Taking the scalar product of this relation with φg , we obtain
dE(g) dH(g)
= φg , φg . (4.6)
dg dg
Let us consider the commonly occurring case of H(g) = H0 + gV ;
then, it is clear that
dE(g)
= hV φg , φg i , (4.7)
dg
dφg
(H0 + gV − E(g)) = (hV φg , φg i − V )φg . (4.8)
dg
We will sometimes assume that the vector φg is analytic with respect
to g in the neighborhood around g = 0, and we will search for an
expansion of the vector φg and the eigenvalue E(g) in a series in
g (stationary perturbation theory). Formulas (4.7) and (4.8), when
g = 0, provide the linear terms in g (first-order terms), namely,
E(g) = E(0) + g hV φ0 , φ0 i + · · · ,
φg = φ0 + gψ + · · · ,
where ψ satisfies the relation
(H0 − E(0))ψ = (hV φ0 , φ0 i − V )φ0 . (4.9)
D E
dφ
Given that the vector φg satisfies the equation φg , dgg = 0, we
obtain the following condition on the vector ψ:
hψ, φ0 i = 0. (4.10)
April 6, 2020 15:51 Mathematical Foundations of Quantum Field Theory 9in x 6in 11222-ch04 page 45
Rg
where C(g) = 0E(λ)dλ, sg is defined by the relation
dφg dsg
i = (Hg − E(g))sg , , φg = 0, sg0 = 0,
dg dg
and the norm of the vector r(g, α) is bounded above by a constant that
does not depend on g or α (here, g lies in the interval g0 ≤ g ≤ g1 ).
To prove this lemma, let us first use the change of variables ψ(g) =
exp( αi C(g))σ(g) to transform equation (4.12) to the form
dσ(g)
iα = (Hg − E(g))σ(g). (4.13)
dg
Let us assume that the solution of equation (4.13) has the form
∞
X
σ(g) = αn σn (g). (4.14)
n=0
dV (g)
iα = (Hg − E(g))V (g),
dg
V (g0 ) = 1.
R0
where C = aE(h(λ))dλ. We obtain
i a
φh(0) = φ̃0 = lim exp C Uα 0, φ . (4.24)
a→0 α α h(a)
Introducing the operator
Sα (t, T ) = exp(iHg0 t)Uα (t, T ) exp(−iHg0 T ),
we can rewrite the equation (4.14) in the form
Z 0
i a
φh(0) = lim exp [E(h(λ)) − E(h(a))]dλ Sα 0, φ .
a→0 α a α h(a)
(4.25)
Since Sα ( αa , T ) = 1 for T < αa and h(λ) = h(a) for λ < a, using the
relation (4.25), we can show that
Z 0
i
φh(0) = lim exp [E(h(λ)) − E(h(−∞))]dλ
a→0 α −∞
× Sα (0, −∞) φh(−∞) . (4.26)
To obtain (4.11) from (4.26), we must choose for the family of
Hamiltonians Hg the family H0 + gV with 0 ≤ g ≤ g1 , and for
h(τ ) choose the function g exp(−α|τ |). The function exp(−α|τ |) does
not vanish for τ 0 as required of the function h(τ ), therefore,
strictly speaking, we cannot use equation (4.26). However, a slight
modification of the above proof, based on Remark 4.2, allows us to
verify (4.11).
Remark 4.3. If the family of Hamiltonians Hg also depends on
another parameter Ω, then it is not difficult to outline the conditions
under which the limit in (4.26) is uniform in Ω (for this, it is
necessary to give uniform in Ω estimates in the proof of the lemma).
In particular, if HgΩ = H0Ω + gV Ω (0 ≤ g ≤ g1 , Ω ∈ O), then the limit
in (4.26) will be uniform in g and in Ω if the norm of the operators
V Ω is bounded by a constant not depending on Ω and it is possible
to find a δ in such a way that the interval (E Ω (g) − δ, E Ω (g) + δ) for
any Ω and g contains no eigenvalues of the operator HgΩ except for
E Ω (g) (see Tyupkin and Shvarts, 1972).
March 26, 2020 14:41 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch05 page 51
Chapter 5
51
March 26, 2020 14:41 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch05 page 52
where
This means that the limits of Φx (t) for t → ±∞ exist for x ∈ T and
therefore, the limits exist for all x ∈ R (see Appendix A.5).
Let us consider the relation of Møller matrices S± and the
S-matrices S to the adiabatic Møller matrices Sα± and the adiabatic
March 26, 2020 14:41 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch05 page 53
If the operator S+ is unitary, then the relations (5.4) and (5.5) imply
slim Sα = S. (5.6)
α→0
It follows from this inequality that the limits Φαx (±∞) exist, hence
the operators Sα± also exist. Furthermore, it follows from this
March 26, 2020 14:41 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch05 page 54
lim Σ± = S± .
→+0
which holds if the vector function f (t) is bounded and has the limit
f (±∞) = limt→±∞ f (t).1 It is clear that ψλ± = lim→+0 ψλ± , where
Z ±∞
±
ψλ = Σ± φλ = ± exp(−|t|) exp(iHt) exp(−iEλ t)φλ dt
0
±i
= φλ .
H − Eλ ± i
1
If kf (t)k ≤ A and for t ≥ T , we have that kf (t) − f (+∞)k ≤ δ, then
Z ∞
Z ∞
exp(−t)f (t)dt − f (+∞)
=
exp(−t)(f (t) − f (+∞))dt
0 0
≤ 2T A + δ.
March 26, 2020 14:41 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch05 page 56
1
( ψλ− − φλ , V ψµ+ − V ψλ− , ψµ+ − φµ )
−
Eλ − Eµ
1 1
−
= δ(λ − µ) + − V ψλ , φµ
Eλ − Eµ + i0 Eλ − Eµ
1 1
φλ , V ψµ+ .
+ + (5.11)
Eµ − Eλ + i0 Eλ − Eµ
2
We will not go further into the delicate question of specifying the precise
meaning of the Lippman–Schwinger equation. The calculations at the end of this
section are also informal.
March 26, 2020 14:41 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch05 page 57
3
We consider only the scattering on non-zero angles.
March 26, 2020 14:41 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch05 page 59
p2
(exp(−iH0 t)Sφ)(p) = exp −i t ψ(p),
2m
hence the probability of having the momentum of outgoing particles
directed in the solid angle Ω is equal to
Z Z Z
2
|ψ(p)| dp = dp dqdq0 S(p, q)S(p, q0 )φ(q)φ(q0 ).
Ω Ω
δ(p − q) δ(p − q0 )
× · α(q)α(q0 ) exp(i(q − q0 )a),
2p 2p
Integrating over a, we obtain
δ(p − q)
Z Z
σΩ = (2π) 2
dp dqdq0 S1 (p, q)S1 (p, q0 )
Ω 4p2
×α(q)α(q0 )δ(q − q0 )δ(qT − q0T );
here qT , q0T are projections of the vectors q and q0 onto the plane
orthogonal to the vector p0 ; we used the fact that
Z
exp(i(q − q0 )a)da = (2π)2 δ(qT − q0T )].
a⊥p0
It is easy to check that
δ(q − q 0 )δ(qT − qT 0 ) = 2qδ(q 2 − q 02 )δ(qT − qT 0 )
= 2qδ(qn2 − qn02 )δ(qT − q0T )
q
= δ(qn − qn0 )δ(qT − q0T )
qn
q
+ δ(qn + qn0 )δ(qT − q0T )
qn
q
= [δ(q − q0 ) + δ(Iq − q0 )],
qn
where qn , qn are the projections of the vectors q, q0 on the vector p0
0
If p0 6= 0, then
Z
f (q)α(q)α(Iq)dq ≈ 0.4
4
It is not difficult to convert the above considerations into a rigorous proof.
To be precise, one needs to consider a sequence of normalized functions αn with
support tending towards the point p0 6= 0 and use the fact that for such a sequence
and a continuous function f (q), we have
Z
f (q)|αn (q)|2 dq → f (p0 ),
Z
f (q)αn (q)αn (Iq)dq → 0.
March 26, 2020 14:41 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch05 page 62
Φs (k) = a+
s (k)θ. (5.24)
H = H0 + W, (5.25)
where
X Z
W = Wm,n (k1 , . . . , km |l1 , . . . , ln )a+ (k1 )
m≥2,n≥2
× φn (k1 , i1 , . . . , kn , in )a+ +
i1 (k1 ) . . . ain (kn )θdk1 . . . dkn .
Let us now define the in- and out-operators Ain (k, τ ) and
Aout (k, τ ), corresponding to the particle Φ(k), as the limits
Ain (k, τ ) = lim exp(iω(k)(t − τ ))A(k, t), (5.29)
t→−∞
Ψout (k1 , i1 , . . . , kn , in ) = A+ +
out (k1 , i1 ) . . . Aout (kn , in )θ
are called in- and out-states.
March 26, 2020 14:41 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch05 page 70
Sometimes it is useful to note that the vectors Ψin and Ψout can
also be represented in the form
∗
Ain (k, s) = ain (k, s) = S− as (k)S− , (5.32)
∗
Aout (k, s) = aout (k, s) = S+ as (k)S+ . (5.33)
March 26, 2020 14:41 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch05 page 71
To check these equalities, we will use the relations (5.1) and (5.2). It
follows from these relations that
+
S∓ as (k)S∓ = slim exp(iHt) exp(−iH0 t)as (k)
t→∓∞
× exp(iH0 t) exp(−iHt)
= slim exp(is (k)t)as (k, t) = a in (k, s).
t→∓∞ out
The relations (5.32) and (5.33) imply that
+
ain (l1 , σ1 ) . . . a+ + +
in (ln , σn )θ, aout (k1 , s1 ) . . . aout (km , sm )θ
= S− a+ + + +
σ1 (l1 ) . . . aσn (ln )θ, S+ as1 (k1 ) . . . asm (km )θ
= Sa+ + + +
σ1 (l1 ) . . . aσn (ln )θ, as1 (k1 ) . . . asm (km )θ
× a+ +
s1 (k1 ) . . . asm (km )dk1 . . . dkm
where
exp(−iPx)A+ +
i (k) exp(iPx) = exp(−ikx)Ai (k),
March 26, 2020 14:41 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch05 page 74
we see that
Z
Bα (fα , t) = f˜α (x, t) exp(−ikx)φα (k)A+
iα (k, t)dk
Z
= exp(−iωiα (k)t)fα (k)φα (k)A+
iα (k, t)dk,
hence
Z
slim Bα (fα , t) = fα (k)φα (k)A+in (k, iα )dk. (5.36)
t→∓∞ out
. . . b+ (f n φn , in )θ.
March 26, 2020 16:10 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch06 page 75
Chapter 6
1
In the definition of the representation of CCR and CAR, one should assume
that the operators a(f ), a+ (f ) are defined on the same linear subspace D that is
dense in the space H. They should transform D into itself. The operators a(f )
and a+ (f ) are conjugate (i.e. ha(f )x, yi = x, a+ (f )y for all x, y ∈ D).
75
March 26, 2020 16:10 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch06 page 76
2
Using CR, it is easy to check that every vector of the form a(f1 , 1 ) . . . a(fn , n )θ
can be represented as a linear combination of vectors of the form
a+ (φ1 ) . . . a+ (φm )θ. Hence, θ is also a cyclic vector with respect to the family
of operators a+ (f ).
March 26, 2020 16:10 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch06 page 78
3
If the first representation of CR is a Fock representation, but the second
representation has a vector θ2 ∈ R2 satisfying the condition a2 (f )θ2 = 0 (not
necessarily cyclic), then precisely in the same way, we can construct the operator
α satisfying the relation αa1 (f ) = a2 (f )α, αθ1 = θ2 . However, in general, α will be
an isometry that is not necessarily unitary (this statement is used in Section 5.3
to construct the operators S− and S+ ).
March 26, 2020 16:10 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch06 page 79
For every vector ξ ∈ F (L2 (E n )), the function hξ, a+ (x)θi is square
integrable because this function enters the Fock column correspond-
ing to the vector ξ.4 This means that in the case when φ ∈ / L2 (E n ),
equation (6.9) does not have a solution in the Fock space.
A broader class of representations of CCR is described by the
following formulas:
Z
+
b(f ) = a(Φf ) + a (Ψf ) + f (x)φ(x)dx,
Z (6.10)
+ +
b (f ) = a(Ψf ) + a (Φf ) + f (x)φ(x)dx
4 + +
We have used the fact
that hξ, θi 6= 0. If hξ, θi = 0, then hξ, a (x1 )a (x2 ) . . .
+ + +
a (x+n )θi = −φ(x 1 ) ξ, a (x2 ), . . . , a (xn )θ . By induction on n, we derive
ξ, a (x1 ) . . . a+ (xn )θ = 0, hence ξ = 0.
March 26, 2020 16:10 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch06 page 82
[Nk , a+ +
l ] = δk,l al , [Nk , al ] = −δk,l al .
If φ is a common eigenvector of the operators Nk , in other words
Nk φ = nk φ, then Nk a+ + + +
l φ = al Nk φ = nk al φ for l 6= k, Nk ak φ =
a+ +
k (Nk + 1)φ = (nk + 1)ak φ. Similarly, Nk al φ = nk al φ for l 6=
k, Nk ak φ = (nk − 1)ak φ. We conclude that vectors of the form
March 26, 2020 16:10 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch06 page 83
a+n +nr
k1 . . . akr θ, where ni are non-negative integers and ki 6= kj ,
1
5
More precisely, two vectors of this kind are either proportional or orthogonal.
The completeness of this system of eigenvectors follows from the cyclicity of the
vector θ with respect to the family of operators a+ (f ).
6
More precisely, two vectors of this kind are either orthogonal or proportional.
March 26, 2020 16:10 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch06 page 84
7
The operator H0 is an essentially self-adjoint operator for any choice of real
numbers ωk (this follows from the fact that the operator H0 has a complete
system of eigenvectors). In particular, the operator of the number of particles N
is essentially self-adjoint. As usual, we identify the essentially self-adjoint operator
with its self-adjoint extension.
March 26, 2020 16:10 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch06 page 85
define an operator on the space F (B) = F (L2 (X)). Note that only
for the case X = E r some results in this direction are proven in
Section 3.2.
In the case of CAR an operator corresponding to a physical quantity
(in particular, a Hamiltonian) should always contain an even number
of creation and annihilation operators (i.e. Γm = 0 for odd m).
For example,
A1 A2 B1 B2 = B1 B2 A1 A2 + B1 B2 A1 A2 + B1 B2 A1 A2 + B1 B2 A1 A2
+B1 B2 A1 A2 + B1 B2 A1 A2 + B1 B2 A1 A2
+ B1 B2 A1 A2 .
+ · · · + BA1 . . . Am . (6.21)
to the edges. (Here, l is the index of the end of the edge and p is the
index of the beginning of the edge.) The direction of the edge will
always be fixed as the direction from the out-vertex of the diagram of
the operator K to the in-vertex of the diagram of the operator L. For
every diagram, we construct a product in which the internal vertices
of the diagram contribute the factors K(k1 , . . . , km |l1 , . . . , ln ) and
L(p1 , . . . , pr |q1 , . . . , qs ), where k1 , . . . , km , p1 , . . . , pr are the indices
of the ends of the lines entering the vertex and l1 , . . . , ln , q1 , . . . , qs
are the indices of the ends of the lines exiting the in-vertex. The
factor a+ p al = δlp is assigned to every connecting edge. Finally, the
8
It is useful to note that an orthonormal basis φk , with k ∈ M , for the space
B, specifies an isomorphism between the spaces B and L2 (M ). (The set M is
considered as a measure space, equipped with the counting measure, where the
measure of a finite subset is equal to the
P number of its elements; then the integral
of f (k) over the set M is equal to k∈M f (k).) This remark allows us to say
that the representation of an operator in the form (6.20) is a particular case of
representation in the form (6.22); correspondingly, the diagram technique for the
operators of the form (6.22) is a generalization of the techniques for operators of
the form (6.20).
March 26, 2020 16:10 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch06 page 96
(where the sum is taken over all permutations P = (i1 , . . . , im )). The
formula (6.23) can also be expressed in the form
+
a (x1 ) . . . a+ (xm )θ, a+ (y1 ) . . . a+ (yn )θ
X
n
= δm δ(x1 , yi1 ) . . . δ(xm , yim ), (6.24)
P
where
X X
V (t) = vm,n (k1 , . . . , km |p1 , . . . , pn |t)a+ +
k1 . . . akm
m,n k1 ,...,km
×ap1 . . . apn
(here, a+ +
k = a (φk ), ak = a(φk ) are operators in the Fock space F (B)
corresponding to the orthonormal basis φk ∈ B and the index k runs
over the set M ). For concreteness, we consider the case of CCR;
however, the consideration of this section can also be performed
in the case of CAR (in the case of CAR, the Hamiltonian H is
assumed to be Fermi-even, i.e. it is assumed that every summand
in the operator H contains an even number of operators a+ and a
(m + n is even)).
Using the representation of the operator S(t, t0 ) in the form of
T -exponent (4.4) and Wick’s theorem (Section 6.3), we can construct
diagram techniques for the calculation of the operator S(t, t0 ). More
precisely, we obtain the decomposition of the normal form of S(t, t0 )
as a power series in g.
Let us recall that a star is a diagram consisting of a point with m
incoming lines and n outgoing lines. The lines belonging to a star are
depicted as dotted lines (Fig. 6.1 depicts a star with three incoming
and two outgoing lines).
A diagram of the operator S(t, t0 ) is a collection of several stars
and several edges that are depicted as directed lines. Every edge
starts with an out-vertex of some star and ends in an in-vertex of
another star. Let us assume that two lines belonging to a diagram
cannot have common internal points and can have only one common
March 26, 2020 16:10 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch06 page 99
X k2
H0 = a+ ak ,
2m k
1 2π 3 X
V = W̃(k1 − p2 )δpk11+p
+k2
2
× a+ +
k1 ak2 ap1 ap2 ,
2 L
k1 ,k2 ,p1 ,p2
× ã+ + + +
k1 (τ )ãk2 (τ )ãp1 (τ )ãp2 (τ )dτ
1 2π 3 X
= W̃(k1 − p2 )δpk11+p
+k2
2
2 L
k1 ,k2 ,p1 ,p2
2α
× a+ +
k1 ak2 ap1 ap2 .
k12 k22 p21 p22 2
2m + 2m + 2m + 2m + α2
The operator corresponding to the second diagram has the form
1
Z X X 1 2π 6
dτ1 dτ2 W̃(k1 − p2 )
2l 0 0 0 0
4 L
k1 ,k2 ,p1 ,p2 k1 ,k2 ,p1 ,p2
p21
× δpk11+p
+k2
exp(−α|τ1 |) exp −i
2
(τ1 − τ2 ) δkp01 θ(τ1 − τ2 )
2m 1
p2
k0 +k0
× exp −i 2 (τ1 − τ2 ) θ(τ1 − τ2 )δkp22 W̃(k10 − p02 )δp01+p02
2m 1 2
× exp(−α|τ2 |)ã+ +
k1 (τ1 )ãk2 (τ1 )ãp01 (τ2 )ãp02 (τ2 )
(where ki , pi and ki0 , p0i are summed over a lattice).
March 26, 2020 16:10 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch06 page 103
Ṽ (τ ) = exp(iH0 τ )V (τ ) exp(−iH0 τ )
X X
= vm,n (k1 , . . . , km |p1 , . . . , pn |τ )ã+
k1 (τ )
m,n k1 ,...,km
. . . ã+
km (τ )ãp1 (τ ) . . . ãpn (τ )
X X Xm n
X
= exp i (kj ) − (pj ) τ
m,n k1 ,...,km j=1 j=1
XZ
V (t) = vm,n (k1 , . . . , km |p1 , . . . , pn |t)a+ (k1 )
m,n
Chapter 7
107
March 26, 2020 16:33 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch07 page 108
wn (λ1 , t1 + τ, . . . , λn , tn + τ ) = wn (λ1 , t1 , . . . , λn , tn ).
(7.1)
(2) Hermiticity:
+ a, . . . , λn , tn + a)da = 0 (7.4)
wn (k1 , 1 , t1 , . . . , kn , n , tn )
= ha(k1 , 1 , t1 ) . . . a(kn , n , tn )Φ, Φi .
Gn (k1 , 1 , t1 , . . . , kn , n , tn )
= lim C1 (α)C2 (α)
α→0
× hSα (0, ti1 )ã(ki1 , i1 , ti1 )Sα (ti1 , 0)Sα (0, ti2 ) . . .
× ã(kin , in , tin )Sα (tin , 0)Sα (0, −∞)Φ0 , Sα (0, +∞)Φ0 i
hAα Φ0 , Φ0 i
= lim ,
α→0 hSα (∞, −∞)Φ0 , Φ0 i
where
Aα = T (ã(k1 , 1 , t1 ) . . . ã(kn , n , tn ))
Z ∞
1
× exp g exp(−α|τ |)Ṽ (τ )dτ
i −∞
∞
X 1 g n
=
n! i
n=0
Z
× T (ã(k1 , 1 , t1 ) . . . ã(kn , n , tn )
hAα Φ0 , Φ0 i
. (7.10)
hSα (∞, −∞)Φ0 , Φ0 i
Then
w2 (λ1 , t1 , λ2 , t2 ) = ha(λ1 , t1 )a(λ2 , t2 )Φ, Φi
D E
= a(λ2 , t2 )Φ, a(λ̃1 , t1 )Φ
X D E
= ha(λ2 , t2 )Φ, Φn i Φn , a(λ̃1 , t1 )Φ
n
X
= hexp(iHt2 )a(λ2 ) exp(−iHt2 )Φ, Φn i
n
D E
× Φn , exp(iHt1 )a(λ̃1 ) exp(−iHt1 )Φ
X
= exp(i(En − E0 )t2 ) ha(λ2 )Φ, Φn i
n
D E
× exp(i(E0 − En )t1 ) Φn , a(λ̃1 )Φ .
Introducing the notation ρλn = hΦn , a(λ)Φi, we obtain the
representation
X
w2 (λ1 , t1 , λ2 , t2 ) = exp[−i(En − E0 )(t1 − t2 )]ρλ̃n1 ρ̄λn2 .
n
If the operator H additionally has continuous spectrum, then the
complete system of eigenvectors consists of normalized eigenvectors
Φn and generalized eigenvectors Φγ (we assume that the vectors
Φn are orthonormal and Φγ are δ-normalized; the corresponding
eigenvalues are denoted by En and Eγ ). Repeating the above
calculations and using the relation
X Z
hx, yi = hx, Φn i hΦn , yi + hx, Φγ i hΦγ , yi dγ,
n
we see that
X
w2 (λ1 , t1 , λ2 , t2 ) = exp[−i(En − E0 )(t1 − t2 )]ρλ̃n1 ρ̄λn2
n
Z
+ exp[−(Eγ − E0 )(t1 − t2 )]ρλ̃γ 1 ρ̄−λ
γ dγ,
2
(7.11)
where
ρλn = hΦn , a(λ)Φi , ρλγ = hΦγ , a(λ)Φi .
March 26, 2020 16:33 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch07 page 118
Indeed,
1
A rigorous proof of this statement can be given under the assumption of
absolute continuity of the continuous spectrum.
March 26, 2020 16:33 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch07 page 119
G2 (k1 , 1 , t1 , k2 , 2 , t2 )
X
= θ(t1 − t2 ) exp[−i(En − E0 )(t1 − t2 )]ρkn1 ,−1 ρ̄kn2 ,2
n
X
± θ(t2 − t1 ) exp[−i(En − E0 )(t2 − t1 )]ρkn2 ,−2 ρ̄kn1 ,1
n
(here, ρk,
n = hΦn , a(k, )Φi and Φn are the stationary states of the
Hamiltonian H and En are the corresponding eigenvalues).
Let us also consider the functions
w̃2 (k1 , 1 , ω1 , k2 , 2 , ω2 )
Z
−1
= (2π) exp[i(1 ω1 t1 + 2 ω2 t2 )]
where
Z
G(k1 , k2 , ω) = exp(iωτ )G2 (k1 , 1, τ, k2 , −1, 0)dτ
where a+
= a+ (φk ), ak = a(φk ), φk is an orthonormal basis in
k
B, the functions Γm,n are symmetric with respect to variables
ki and lj in the case of CCR and antisymmetric in the case of
CAR. Then the Heisenberg equations for the operators ak (t) =
exp(iHt)ak exp(−iHt), a+ +
k (t) = exp(iHt)ak exp(−iHt) can be writ-
ten in the form
1 dak (t)
= [H, ak (t)]
i dt
X
=− mΓm,n (k, k1 , . . . , km−1 |l1 , . . . , ln )
m,n,k1 ,...,km−1 ,l1 ,...,ln
× a+ +
k1 (t) . . . akm−1 (t)al1 (t) . . . aln (t);
1 da+
k (t)
= [H, a+
k (t)]
i dt
X
= nΓm,n (k, k1 , . . . , km |l1 , . . . , ln−1 , k)
m,n,k1 ,...,km−1 ,l1 ,...,ln
× a+ +
k1 (t) . . . akm (t)al1 (t) . . . aln−1 (t).
These equations immediately imply equations for Wightman
functions
wn (k1 , 1 , t1 , . . . , kn , n , tn ).
Let us calculate, for example, the expression of the derivative of
the function wn with respect to the variable t1 in terms of Wightman
March 26, 2020 16:33 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch07 page 121
The first two summands in the formula for ∂G ∂t1 can be expressed
2
Chapter 8
Translation-Invariant Hamiltonians
123
March 27, 2020 8:20 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch08 page 124
2
Indeed, in the case of vacuum polarization, the expression (8.1) does not
determine an operator on Fock space at all (see Section 3.2 for details).
3
To construct an operator on Fock space corresponding to a translation-invariant
Hamiltonian with vacuum polarization, we should make the volume or, in other
words, the infrared cutoff (see below).
March 27, 2020 8:20 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch08 page 125
XZ
H= Λm,n (k1 , . . . , km |l1 , . . . , ln )
m,n
× δ(k1 + · · · + km − l1 − · · · − ln )
× a+ (k1 ) . . . a+ (km )a(l1 ) . . . a(ln )dm kdn l. (8.2)
(in other words, we obtain the expression for HΩ from the expression
L 3/2 + L 3/2
for H by replacing a+ (k), a(k) with ( 2π ) ak , ( 2π ) ak , replacing
the integration with summation over the lattice TΩ and multiply-
ing by ( 2π 3
L ) , and replacing the function δ(k1 +· · ·+km −l1 −· · ·−ln )
L 3
by ( 2π ) δk1 +···+km ,l1 +···+ln ).
March 27, 2020 8:20 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch08 page 126
4
In the case of CCR, with the given assumptions on the coefficient functions,
one can prove only that the formula (8.3) defines a Hermitian operator.
March 27, 2020 8:20 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch08 page 127
∂a(k, −1, t) X
Z
i = m Λm,n (k, k1 , . . . , km−1 |l1 , . . . , ln )
∂t m,n
× δ(k + k1 + · · · + km−1 − l1 − · · · − ln )
× a(k1 , 1, t) . . . a(km−1 , 1, t)a(l1 , −1, t)
. . . a(ln , −1, t)dm−1 kdn l; (8.4)
∂a(k, 1, t) X Z
i =− n Λm,n (k1 , . . . , km |l1 , . . . , ln−1 , k)
∂t m,n
× δ(k1 + · · · + km − l1 − · · · − ln−1 − k)
× a(k1 , 1, t) . . . a(km , 1, t)a(l1 , −1, t)
. . . a(ln−1 , 1, t)dm kdn−1 l. (8.5)
It would be more precise to say that the functions w̌n are Wightman
functions in (k, t)-representations. Their Fourier transforms with
respect to the variables k1 , . . . , kn are the functions
Z X
− 23 n
wn (x1 , 1 , t1 , . . . , xn , n , tn ) = (2π) exp i j xj kj
|ρ(−1, λ)|2
Z
=i δ(k + k(λ))dλ
ω − E(λ) + i0
|ρ(+1, λ)|2
Z
∓i δ(−k + k(λ))dλ. (8.10)
ω + E(λ) − i0
March 27, 2020 8:20 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch08 page 131
w̌n (k1 , 1 , t1 , . . . , kn , , tn )
3n
L 2
= lim wnΩ (k1 , 1 , t1 , . . . , kn , n , tn ) (8.11)
Ω→∞ 2π
We assume that the limit (8.11) exists (the existence of this limit
can be checked in the framework of perturbation theory).
The functions w̌n (k1 , 1 , t1 , . . . , kn , n , tn ) will be considered as
generalized functions with respect to the variables k1 , . . . , kn and
conventional functions with respect to the variables t1 , . . . , tn .
Green functions Ǧn (k1 , 1 , t1 , . . . , kn , n , tn ) of the translation-
invariant Hamiltonian H are defined in the same way, as limits
of finite-volume Green functions. Namely, we use the following
definition:
Ǧn (k1 , 1 , t1 , . . . , kn , n , tn )
3n
L 2 Ω
= lim Gn (k1 , 1 , t1 , . . . , kn , n , tn ),
Ω→∞ 2π
March 27, 2020 8:20 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch08 page 132
1 n
where GΩ n (k1 , 1 , t1 , . . . , kn , n , tn ) = hT (ak1 (t1 ) . . . akn (tn ))ΦΩ , ΦΩ i
are Green functions of the Hamiltonian HΩ constructed with the
basis φk , where k ∈ TΩ . The limit is understood in the same way as
the limit in the definition of Wightman functions.
Green functions Ǧn can be expressed in terms of Wightman
functions w̌n . Namely, taking the limit of Ω → ∞ in the relation (7.7)
applied to the functions GΩ Ω
n and wn , we obtain
Ǧn (k1 , 1 , t1 , . . . , kn , n , tn )
X
= (−1)γ(π) θπ (t)w̌nπ (k1 , 1 , t1 , . . . , kn , n , tn ) (8.13)
π
(2) Hermiticity:
. . . , kn , n , tn + a)da = 0, (8.14)
if ω < 0. The relation (8.14) is also valid under a weaker
condition — it is sufficient to assume that ω does not belong
to the spectrum of the Hamiltonian H.
(5) Symmetry with respect to the permutation of arguments: Let us
introduce the notation
w̌n(i) (k1 , 1 , t1 , . . . , ki , i , ti , ki+1 , i+1 , ti+1 , . . . , kn , n , tn )
= ±w̌n (k1 , 1 , t1 , . . . , ki+1 , i+1 , ti+1 , ki , i , ti ,
. . . , kn , n , tn )
(here, the plus sign corresponds to the case of CCR and the
(i)
minus sign corresponds to the case of CAR). In other words, w̌n
is obtained from w̌n by permuting ki , i , ti with ki+1 , i+1 , ti+1
(with a change of sign in the case of CAR).
If ti = ti+1 , then
w̌n(i) (k1 , 1 , t1 , . . . , kn , n , tn )
= w̌n (k1 , 1 , t1 , . . . , kn , n , tn ) + A0 δ(ki − k0i )
× w̌n−2 (k1 , 1 , t1 , . . . , ki−1 , i−1 , ti−1 , ki+2 , i+2 , ti+2 ,
. . . , kn , n , tn ) (8.15)
0
(the definition of the matrix A can be found in Section 6.1).
(6) Translation-invariance:
w̌n (k1 , 1 , t1 , . . . , kn , n , tn )
X
= vn (k1 , 1 , t1 , . . . , kn , n , tn )δ j kj
All the above listed properties (with the exception of (6)) can
be easily proven if we recall that w̌n are obtained as limits of the
functions wnΩ , which have similar properties proven in Section 7.1.
Let us start the proof from the second part of the theo-
rem. Let us consider the set N of sequences of functions f =
5
We will consider the space of test functions to be S(E 3 ) (i.e. all functions f in
the formulation of the theorem satisfy f ∈ S(E 3 )).
March 27, 2020 8:20 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch08 page 135
In the space
R D, we can define the operator generalized functions
a(f, , t) = f (k)a(k, , t)dk, assuming that a(f, , t)Ψg = Ψφg ,
where φ(k, σ, τ ) = f (k)δσ, δ(t − τ ), and we can define the family
of operators Ṽτ , W̃α , satisfying Ṽτ Ψg = ΨVτ g , W̃α Ψg = ΨWα g .
Let us now define the Hilbert space H as the completion of the
pre-Hilbert space D.
The operators Ṽτ , W̃α map the set D into itself and preserve inner
products. Therefore, they can be extended by continuity to unitary
operators on the space H. In this way, we obtain a one-parameter and
a three-parameter group of unitary operators on H; the generators
of these groups will be denoted by Ĥ and P̂ (in other words, Ṽτ =
exp(−iHτ ), W̃α = exp(−iαP̂)). The symbol Φ will denote the vector
Ψθ , where θ is a function sequence with f0 = 1, fn = 0, for n > 0.
The operator generalized functions a(f, , t) are defined on the dense
subset D ⊂ H.
Hence, starting with the Wightman functions, we have con-
structed the objects that were described in the reconstruction theo-
rem. It is easy to check that they have all the necessary properties.
The only point we will consider in detail is the proof that the vector
Φ is the ground state of the Hamiltonian Ĥ. We derive this fact from
the following lemma.
Lemma 8.1. The number ω does not belong to the spectrum of the
operator Ĥ if and only if for all Wightman functions we have
Z
exp(−iωτ )w̌n (k1 , 1 , t1 , . . . , ki , i , ti , ki+1 , i+1 , ti+1 + τ,
. . . , kn , n , tn + τ )dτ = 0. (8.19)
Then, by property (4) of Wightman functions and the lemma, it
follows that the operator Ĥ is non-negative. Taking into account that
ĤΦ = 0, we see that Φ is the ground state.
To prove (2), we first note that the functions w̌n are Wightman
functions of the operator Ĥ with respect to the operator generalized
function a(k, , t) (in the sense of the definition in Section 7.1).
Therefore, it follows from property (4) (Section 7.1) that for every
ω that does not belong to the spectrum of the operator Ĥ, we
have (8.19). To prove the inverse statement, it is sufficient to check
March 27, 2020 8:20 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch08 page 138
easy to give this proof, using the remark that the Wightman functions
of the Hamiltonian H satisfy
1 ∂
w̌r (k1 , 1 , t1 , . . . , kr , r , tr )
i ∂t1
X Z
1
= δ1 n Λm,n (p1 , . . . , pm |q1 , . . . , qn−1 , k1 )δ(p1 + · · · + pm
m,n
w̌n (k1 , 1 , t1 , . . . , kn , n , tn )
= v̌n (k1 , 1 , . . . , kn , n , t2 − t1 , . . . , tn − tn−1 ).
and note that by the spectrum condition the support of the function
vn is contained in the set ω1 ≥ 0, . . . , ωn−1 ≥ 0). The analytic
continuation of the function ṽn will be denoted by the same symbol.
March 27, 2020 8:20 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch08 page 141
Sn (k1 , 1 , t1 , . . . , kn , n , tn )
= v̌n (k1 , 1 , . . . , kn , n , i(t1 − t2 ), . . . , i(tn−1 − tn ))
where p̂k , q̂k are self-adjoint operators satisfying the canonical com-
mutation relations (CCR)
~
[p̂k , p̂l ] = [q̂k , q̂l ] = 0; [p̂k , q̂l ] = δkl
i
March 27, 2020 8:20 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch08 page 142
where π(x) are the generalized momentum variables and φ(x) are the
generalized coordinates. For definiteness, we assume that x runs over
three-dimensional Euclidean space. We consider only translation-
invariant functionals (i.e. we assume that the function Vn (x1 , . . . , xn )
has the form vn (x1 − xn , . . . , xn−1 − xn )). It is natural to conjecture
that by quantizing such a system we will obtain a quantum system
described by the Hamiltonian
Z XZ
1
H= π̂ 2 (x)dx + Vn (x1 , . . . , xn )φ̂(x1 ) . . . φ̂(xn )dx1 . . . dxn ,
2 n
(8.23)
where π̂(x), φ̂(x) are Hermitian operators (more precisely, operator
generalized functions) that obey the commutational relations
(1)
∂ 2 φ̂(x, t) X Z
= − n Vn (x, x1 , . . . , xn−1 )
∂t2 n
(2)
i −i
exp τ Ĥ φ̂(x, t) exp τ Ĥ = φ̂(x, t + τ ),
~ ~
−i i
exp ap̂ φ̂(x, t) exp ap̂ = φ̂(x̂ + a, t).
~ ~
d
R
(3) RThe operators φ̂(f, t) = f (x)φ̂(x, t)dx and π̂(f, t) = dt φ(f, t) =
∂ 3
f (x) ∂t φ(x, t)dx, where f ∈ S(E ), are defined on a dense
subset D of the space H and transform the subset into itself;
if the function f is real, then these operators are Hermitian. The
expressions hφ̂(f, t)Ψ1 , Ψ2 i and hπ̂(f, t)Ψ1 , Ψ2 i should depend
March 27, 2020 8:20 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch08 page 144
As the space H, we should take the Fock space F (L2 (E 3 )), and the
energy operator H and the momentum operator P̂ should be defined
by the formulas
Z
Ĥ = ~ ω(k)a+ (k)a(k)dk,
Z
P̂ = ~ ka+ (k)a(k)dk,
dk
+ a(k) exp(−iω(k)t + ikx)) p .
2ω(k)
The ground state Φ coincides with Fock vacuum θ. It is easy to
check that the objects we have constructed satisfy the conditions for
an operator realization and that any other operator realization is
unitarily equivalent to the realization we have described.
Let us now consider an arbitrary Hamiltonian H of the
form (8.23). Let us express it in terms of the symbols a+ (k), a(k)
satisfying CCR, assuming that
Z
dk
φ̂(x) = (2π)−3/2 ~1/2 (a+ (k) exp(−ikx) + a(k) exp(ikx)) p ,
2ω(k)
Z p
−3/2 1/2 i ω(k) +
π̂(x) = (2π) ~ √ (a (k) exp(−ikx)
2
−a(k) exp(ikx))dk, (8.27)
where ω(k) is an almost everywhere positive function. This expres-
sion can be written in normal form by means of CCR; we obtain
a quadratic expression plus a (possibly infinite) constant. We will
discard this constant and as a result we will obtain a formal
expression for H̃ of the form (8.2).
It is easy to see that the problem of constructing an operator
realization of the Hamiltonian H is equivalent to the same problem
March 27, 2020 8:20 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch08 page 146
dk
+ a(k, −1, t) exp(ikx)) p .
2ω(k)
This remark allows us to transfer to Hamiltonians of the form (8.23)
everything we know for Hamiltonians of the form (8.2).
March 27, 2020 9:30 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch09 page 147
Chapter 9
147
March 27, 2020 9:30 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch09 page 148
X Z
V = Vm,n (k1 , . . . , km |p1 , . . . , pn )a+ (k1 )
m,n≥1
1 dξ(t)
exp(itH)V exp(−itH0 )x = ,
i dt
where ξ(t) = exp(iHt) exp(iH0 t)x and hence
Z t2
1
exp(iHt)V exp(−iH0 t)xdt = (ξ(t2 ) − ξ(t1 )).
t1 i
March 27, 2020 9:30 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch09 page 149
XZ
V exp(−iH0 t)x = fn (k1 , . . . , kn |t)a+ (k1 ) . . . a+ (kn )dn kθ,
n
where
Z
fn (k1 , . . . , kn |t) = vn,1 (k1 , . . . , kn |k1 + · · · + kn ) (k1 + · · · + kn )
× exp(−it(k1 + · · · + kn )).
Noting that
X Z
kV exp(−iH0 t)xk = n! |fn (k1 , . . . , kn |t)|2 dn k,
n
we see that k dξ
dt k does not depend on t. Similar considerations show
d2 ξ d
that k dt2 k = k dt (exp(iHt) × V exp(−iH0 t)x)k = k exp(−iHt)(HV −
V H0 ) × exp(−iH0 t)xk = k(HV − V H0 ) exp(−iH0 t)xk does not
depend on t. Now, to finish the proof, we should apply the following
lemma to the vector η(t) = dξ dt : Rif kη(t)k does not depend on t
dη t
and k dt k is bounded above, then t12 η(t)dt cannot tend to zero as
t1 , t2 → ∞.
This mathematical statement — the fact that it is impossible
to give the definition of a Møller matrix in the same way as in the
theory of potential scattering — has a clear physical background. The
problem is that the states a+ (k)θ (one particle states) are eigenstates
of the Hamiltonian H0 but are not eigenstates of the full Hamiltonian
H. In Section 5.3, we introduced the notion of a particle (single-
particle state) for the Hamiltonian H as a generalized vector function
Φ(k) satisfying the conditions (5.20)–(5.22).
March 27, 2020 9:30 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch09 page 150
If we know the in- and out-operators, then the Møller matrices can
be defined by the relations
(the equivalence of this definition with the first one was discussed in
Section 5.3).
Let us start by generalizing the third definition. The limit in (9.4)
does not exist for vm,1 6≡ 0; this is clear from the fact that
the expression
Z
exp(iHt) exp(−iH0 t) f (k)a+ (k)θdk
Z
= f (k) exp(−i(k)t)a+ (k, t)θdk
a+ +
in (k) = wlim Λ(k) exp(−iω(k)t)a (k, t),
t→−∞
a+ +
out (k) = wlim Λ(k) exp(−iω(k)t)a (k, t),
t→+∞
1
The definition of Møller matrices by means of (9.10) was suggested by
I. Ya. Arefieva.
March 27, 2020 9:30 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch09 page 154
the limit in (9.11) exists and satisfies CR) will be discussed in Chap-
ters 10 and 11. To denote the in- and out-operators simultaneously,
we will use the notation aex (k, , t).
Let us establish a few simple properties of in- and out-operators:
Then,
× a(k, , τ ) exp(−iĤt)
= exp(iĤt)( wlim Λ∓ (k, ) exp(iω(k)τ )a(k, , τ ))
τ →∓∞
hence,
From (9.12), it follows that aex (k, −1)Φ = 0 (if this condition is
not satisfied for k in the set K having non-zero measure, then the
generalized vector function aex (k, −1)Φ is a generalized eigenfunction
of the operator Ĥ, hence the number −ω(k), where k ∈ K, belongs
to the spectrum of the operator Ĥ; this is impossible because the
operator Ĥ is positive).
Applying CR and (3), we see that the generalized function
Φ∓ (k) = aex (k, 1)Φ is δ-normalized (hΦ∓ (k) , Φ∓ (k0 )i =
haex (k0 , −1)aex (k, 1)Φ, Φi = h[aex (k0 ), a+ 0
ex (k)]∓ Φ, Φi = δ(k − k )).
Formula (9.12) implies that Φ∓ (k) is a generalized eigenfunction
of the operator Ĥ. In order to prove that Φ∓ (k) is a generalized
eigenfunction of the operator P̂, we should recall that by (2), we have
(it follows from the results of Section 6.1 that such operators exist
and are defined by conditions (9.13) and (9.14) uniquely).
The scattering matrix of a translation-invariant Hamiltonian H
is defined as the operator S = S+ ∗S .
−
It is easy to verify that the scattering matrix S will be unitary if
and only if the spaces Hin = S− Has and Hout = S+ Has coincide.
We will assume that H = Hin = Hout (in other words, not only is
the S-matrix unitary, but so are the Møller matrices S− and S+ ).
Some of the relations proven later, in particular (9.22), are correct
without this assumption.
March 27, 2020 9:30 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch09 page 158
(as usual, b+ (k) = b(k, 1), b(k) = b(k, −1)). It is easy to see that
b(k, , t) = exp(iĤas t)b(k, ) exp(−iĤas t) = exp(iω(k)t)b(k, ),
exp(iP̂as α)b(k, , t) exp(−iP̂as α) = exp(iαk)b(k, , t).
Using this relation and the properties (1) and (2) of the operators
aex , we can show that
ĤS± = S± Ĥas , P̂S± = S± P̂as (9.15)
(i.e. the operators S+ and S− specify unitary equivalences between
the operators Ĥ and Ĥas , P̂ and P̂as ). From (9.15), it follows that
the scattering matrix commutes with the operators Ĥas and P̂as :
S Ĥas = Ĥas S,
S P̂as = P̂as S.
Let us now prove that the scattering matrix defined above has the
following properties:
(1) Sθ = θ (the vacuum is stable),
(2) Sb+ (k)θ = c(k)b+ (k)θ,
where |c(k)| = 1 (single-particle states are stable). The second of
these statements will be proved only under certain restrictions on the
function ω(k); it is sufficient to assume that it is strongly convex.
The stability of the vacuum follows immediately from the relations
S− θ = Φ, S+ θ = Φ. To prove that single-particle states are stable,
we note that
Ĥas Sb+ (k)θ = S Ĥas b+ (k)θ = ω(k)Sb+ (k)θ,
P̂as Sb+ (k)θ = S P̂as b+ (k)θ = kSb+ (k)θ.
March 27, 2020 9:30 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch09 page 159
Hence, the generalized vector function Ψ(k) = Sb+ (k)θ satisfies the
condition Ĥas Ψ(k) = ω(k)Ψ(k), P̂as Ψ(k) = kΨ(k); it is easy to see
that this generalized vector function is δ-normalized. It follows that
Ψ(k) = c(k)b+ (k)θ, where |c(k)| = 1 (see Section 5.3).
We will now consider an ambiguity in our definition of in- and out-
operators. It is not difficult to check that the functions Λ− (k) and
Λ+ (k) are not specified uniquely by the requirement the operators
aex (k, t) obey CR. Namely, we can replace the functions Λ− (k) and
Λ+ (k) by functions Λ0− (k) and Λ0+ (k) that have the same absolute
value as the functions Λ− (k) and Λ+ (k) and obtain the new in- and
out-operators a0in (k, t) and a0out (k, t):
(here, exp(iφ∓ (k)) = Λ0∓ (k)Λ−1 ± (k) and φ∓ (k) are real-valued
functions).
One can check that under the above assumptions, all possible in-
and out-operators can be represented as operators a0ex (k, t).
It is easy to check that the Møller matrices S−0 and S 0 constructed
+
0
by means of the operators aex (k, t) are related to the Møller matrices
S− and S+ corresponding to the operators aex (k, t) by the formula
0
S∓ = S∓ U∓ ,
In what follows, we always assume that the in- and out-operators are
chosen in such a way that condition (9.17) is satisfied.
March 27, 2020 9:30 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch09 page 160
Hence, we have proven the equality Λ− (k) = Λ+ (k) and also related
these quantities with the function ρ(1, k). Using this relation and
the Källén–Lehmann representation, we will prove that the functions
ω(k) and |Λ(k)| can be expressed in terms of the Green function
We obtain
i|ρ(−1, −p)|2 i|ρ(+1, p)|2
G(p, ω) = ∓
ω − ω(−p) + i0 ω + ω(p) − i0
XZ |ρ(−1, k1 , . . . , kn )|2
+i
ω − (ω(k1 ) + · · · + ω(kn )) + i0
n≥2
= a+ + + +
in (q1 ) . . . ain (qn )Φ, aout (p1 ) . . . aout (pn )Φ .
We will now show how to express the functions σm,n in terms of the
Green functions of a translation-invariant Hamiltonian H.
We will prove the following Lehmann–Symanzik–Zimmermann
formula (LSZ):
ω1 , . . . , pm , −1, ωm ), (9.22)
× T (a(k2 , 2 , t2 ) . . . a(kn , n , tn ))
= wlim exp(−i2 w(k1 )t1 )
t1 →+∞
Indeed, this follows from the remark that if t1 > t2 , . . . , > tn , then
Similarly,
T (a(k2 , 2 , t2 ) . . . a(kn , n , tn ))ain (k1 , 1 )
= lim Λ(k1 , 1 ) exp(−i1 ω(k1 )t1 )
t1 →−∞
. . . a(kn , n , tn )))dt1
Z ∞
= L1 T (a(k1 , 1 , t1 ) . . . a(kn , n , tn ))dt1 .
−∞
March 27, 2020 9:30 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch09 page 165
Applying the relation (9.25) several times, we can express the matrix
entries of the scattering matrix in terms of Green functions.
We will give another, shorter, proof. Let us introduce the oper-
ators c(k, , t) = S−∗ a(k, , t)S . We can easily express the Green
−
function of the Hamiltonian H in terms of these operators:
Ǧn (k1 , 1 , t1 , . . . , kn , n , tn )
= hT (a(k1 , 1 , t1 ) . . . a(kn , n , tn ))S− θ, S− θi
= hT (c(k1 , 1 , t1 ) . . . c(kn , n , tn ))θ, θi .
∗ = SS ∗ and on the right by S
Multiplying (9.25) on the left by S+ − −
and using the relations (9.13) and (9.14), we obtain
[S · T (c(k2 , 2 , t2 ) . . . c(kn , n , tn )), b(k1 , 1 )]
Z ∞
=− L1 S · T (c(k1 , 1 , t1 ) . . . c(kn , n , tn ))dt1 . (9.26)
−∞
Applying this formula several times, we see that
[. . . [S, b(k1 , 1 )] . . . b(kn , n )]
Z
n
= (−1) Ln . . . L1 (S · T (c(k1 , 1 , t1 ) . . . c(kn , n , tn )))dt1 . . . dtn .
(9.27)
The expression for the functions σm,n can be obtained from the
remark that
σm,n (p1 , . . . , pm |q1 , . . . , qn )
= (−1)n [. . . [[. . . [S, b+ (q1 )] . . . b+ (qn )]b(p1 )] . . . b(pm )]θ, θ .
(9.28)
Namely, from (9.27) and (9.28) and the relation Sθ = θ, it follows
that
σm,n (p1 , . . . , pm |q1 , . . . , qn )
Z
m
= (−1) Lm+n . . . L1 Ǧm+n (q1 , 1, t1 , . . . , qn , 1, tn , p1 , −1,
In order to obtain (9.22) from the above formula, we should take the
Fourier transform over the variables t1 , . . . , tm+n , using that
Z Z ∞
d
Lf (k, , t)dt = Λ(k, ) (exp(−iω(k)t)f (k, , t))dt
−∞ dt
= iΛ(k, ) lim (ω − ω(k))
ω→ω(k)
Z ∞
× exp(−iωt)f (k, , t)dt.
−∞
We will consider the adiabatic S-matrix SαΩ = SαΩ (∞, −∞) corre-
sponding to the pair of operators (HΩ , H0Ω ) (the Hamiltonian HΩ is
defined as the Hamiltonian H with volume cutoff, see Section 8.1).
× qQ (9.29)
m Ω
Qn Ω
i=1 hpi |Sα |pi i j=1 hqj |Sα |qj i
. . . b+ (pm )θ ,
D +
E
p1 , . . . , pm |SαΩ |q1 , . . . , qn = SαΩ a+ + +
(recall that the operator SαΩ , like the operator HΩ , acts on the Fock
space FΩ and the symbols a+ +
p , ap denote the operators a (φp ), a(φp ),
where φp (x) = L−3/2 exp(ipx) and p runs over the lattice TΩ ,
consisting of the vectors 2π
L n).
m+n −1
p1 , . . . , pm |SαΩ |q1 , . . . , qm θ|SαΩ |θ 2
× qQ .
m Ω |p i
Qn Ω |q i
i=1 hp |S
i α i j=1 hq |S
j α j
qQ
m Ω
Qn Ω
i=1 hpi |Sα |pi i j=1 hqj |Sα |qj i
a+ Ω +
k θ = lim exp(2iρΩ (k))Sα ak θ.
α→0
We conclude that for α → 0,
k|SαΩ |k0 ≈ exp(2iρΩ (k))δkk0 ,
(9.31)
Ω
θ|Sα |θ ≈ exp(2iρΩ ), (9.32)
and therefore,
!
X
UαΩ ≈ ŨαΩ = exp −iρα − i (ρΩ (k) − ρΩ )a+
k ak . (9.33)
k
March 27, 2020 9:30 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch09 page 169
where S̃αΩ are operators that are N -equivalent to the operators SαΩ .
March 27, 2020 9:30 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch09 page 170
Let us recall that the limit in the expression of the form (9.37) is
understood in the sense of the convergence of the matrix entries of
S̃αΩ in the basis a+ +
k1 . . . akn θ to the matrix entries of the operator S
in the generalized basis b+ (k1 ) . . . b+ (kn )θ.
It is clear from the relation (9.36) that a scattering matrix
specified by Definition 7.1 is also a scattering matrix in the sense of
Definition 9.1. However, unlike Definition 7.1, Definition 9.1 does not
specify a scattering matrix unambiguously: using the ambiguity in
the choice of the operators SαΩ , it is easy to see that an operator that
is S-equivalent to the scattering matrix is also a scattering matrix in
the sense of Definition 9.1. The converse of this observation holds as
well: two scattering matrices of a Hamiltonian H are S-equivalent.
Note that in Section 9.2 the scattering matrix was also defined up to
S-equivalence.
In the framework of perturbation theory, one can prove that
Definition 9.1 is equivalent to the definition of the scattering matrix
given in Section 9.2.2 However, to show this equivalence, we need to
make a few assumptions about the Hamiltonian H (it is sufficient
to require that the Hamiltonian belongs to the class M defined in
Section 11.1 and the function (k), entering the definition of the
Hamiltonian H0 , satisfies the condition (k1 + k2 ) < (k1 ) + (k2 )).
We consider the connection between adiabatic S-matrices and
scattering matrices in Section 10.6 (in the framework of axiomatic
scattering theory) and in Section 11.5 (for Hamiltonians that do
not generate vacuum polarization). The proof of the equivalence of
Definition 9.1 with other definitions of scattering matrices in the
framework of perturbation theory with partial summation can be
obtained by modifying the considerations in Section 10.6. Another
less rigorous proof can be found in Likhachev et al. (1972).
2
More precisely, to show this equivalence, we should perform a partial summa-
tion of the series in perturbation theory for SαΩ .
March 27, 2020 9:30 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch09 page 171
. . . a+
km ap1 . . . apn ; (9.42)
XX
r= ρm,n (k1 , . . . , km |p1 , . . . , pn )a+
k1
m,n ki ,pj
. . . a+
km ap1 . . . apn
[h0 , x], we obtain the relation between the unknown functions ξm,n
and the known functions ρm,n :
3
It is easy to see that the operators C(t) = exp(−itα)b exp(itα) and D(t) =
P∞ (it)n dC(t)
n=0 n! [. . . [b, α] . . . , α] satisfy identical equations i dt = [α, C(t)], i dD(t)
dt
=
[α, D(t)] with the same initial condition C(0) = D(0) = b; hence
C(t) = D(t).
March 27, 2020 9:30 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch09 page 174
h0n = i[h0 , αn ] + qn ,
× δkp11+···+k
+···+pn +
a . . . a+
m k1 km ap1 . . . apn ,
March 27, 2020 9:30 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch09 page 176
0Ω
the functions ωkΩ , vm,n converge to a limit as Ω → ∞.4 Let us
introduce the notation
lim ωkΩ = ω(k);
0Ω
lim vm,n (k1 , . . . , km |p1 , . . . , pn )
0
= vm,n (k1 , . . . , km |p1 , . . . , pn ), (9.45)
Z
H00 = ω(k)a+ (k)a(k)dk;
XZ
0 0
V = vm,n (k1 , . . . , km |p1 , . . . , pn )
m,n
H 0 = lim (H 0Ω − C Ω ) (9.47)
Ω→∞
× δkp11+···+k
+···+pn +
a . . . a+
m k1 km ap1 . . . apn ,
4 0Ω
The functions ωkΩ , vm,n are expressed in the form of a series in powers of g;
we understand the convergence as Ω → ∞ in the sense of convergence for every
r
power of g. It is useful to note that we can find such functions νm,n ∈ S and σ r (k)
that do not grow faster than a power function and that satisfy the conditions
0Ω
|(vm,n )r | ≤ νm,n
r
, |(ωkΩ )r | ≤ σ r (k) (by the symbol f r here, we denote a term of
order r in g in the expansion of the function f ).
March 27, 2020 9:30 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch09 page 177
Ω (k , . . . , k |p , . . . , p ) have a limit as
where the functions αm,n 1 m 1 n
Ω → ∞,
Ω
αm,n (k1 , . . . , km |p1 , . . . , pn ) = lim αm,n (k1 , . . . , km |p1 , . . . , pn )
Ω→∞
(9.48)
in each term in the series with respect to g. Using the formal
translation-invariant expression
XZ
A= αm,n (k1 , . . . , km |p1 , . . . , pn )
m,n
It is easy to check that the S-matrix (9.38) can be written in the form
× exp(−iH00Ω t0 )
S= lim Ω
lim exp(iHkb t)(DΩ )−1
t→∞,t0 →−∞ Ω→∞
Formula (9.49) shows that the family of operators (WΩ )−1 is a family
of dressing operators in the sense of the above definition.
In Section 11.4, we will describe a broad class of families of
dressing operators that includes the operators (W Ω )−1 .
The theorem establishing the equivalence of Faddeev’s definition
with other definitions of scattering matrices is a particular case
of a theorem of invariance of scattering matrices under canonical
March 27, 2020 9:30 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch09 page 179
where σm,n are smooth functions whose derivatives do not grow faster
than a power.
The transformation from the symbols a+ (k), a(k) into the sym-
bols b+ (k), b(k) is called a canonical transformation if the sym-
bols b+ (k), b(k) obey the same commutation (or anticommutation)
relations as the symbols a+ (k), a(k) (i.e. obey CR). In the case of
CCR, we can calculate the commutators for the symbols b+ (k), b(k)
by using CCR for the operators a+ (k), a(k), distributivity, and
the relation [A, B, C] = [A, C]B + A[B, C] (we can calculate the
anticommutator in the case of CAR analogously).
The above equivalence theorem can be reformulated (not quite
rigorously) in the following way.
Theorem 9.1. The canonical transformation (9.50) transforms the
translation-invariant Hamitlonian H of the form (9.1) into the
translation-invariant Hamiltonian H̃ having the same scattering
matrix.
(The Hamiltonian H̃ can be constructed in the following way:
in the expression of the Hamiltonian H with respect to the symbols
March 27, 2020 9:30 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch09 page 180
1 X
+ cijk xi xj xk + · · · .
3!
i,j,k
H = H0 + V ;
1X 2 1X
H0 = p̂i + cij x̂i x̂j + U (q10 , . . . , qn0 );
2 2
1 X
V = cijk x̂i x̂j x̂k + · · · .
3!
The operators p̂i , x̂i satisfy the same commutation relations as p̂i , q̂i ,
i.e. the transformation between these operators can be considered to
be a canonical transformation.
The eigenvalues of the operator H0 are specified by the expres-
sion (9.52) (see Section 2.6). The eigenvalues of the operator H =
H0 + V can be calculated in the framework of perturbation theory
by considering the operator V as a perturbation. It is easy to see
that the corrections to the energies of a weakly excited state are
small. This becomes evident if we express H0 and V in terms of the
operators a+ i , ai , satisfying CCR (we rely on the considerations in
March 27, 2020 9:30 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch09 page 182
Section 2.6):
1 X X
H0 = U (q10 , . . . , qn0 ) + ~ ωi + ~ ωi a+i ai ;
2
1
Γm,n
X X
+
V = ~( 2 )(m+n) i1 ,...,im ,j1 ,...,jn ai1 . . . ajn
m+n≥3 i1 ,...,im ,j1 ,...,jn
where
where
Z
νi1 ,...,im = vi1 ,...,im (x1 , . . . , xm−1 )dm−1 (x)
√
Z Z
V ± = a4 ξˆ4 (x)dx ± 2 2a2 a4 ξˆ3 (x)dx.
0 , we get
Starting with the operator realization of the Hamiltonian H+
an equivalent operator realization of the Hamiltonian H that gives
D E ra
2 ˆ t)Φ, Φi
lim φ̂(x, t)Φ, Φ = + lim hξ(x,
~→0 2a4 ~→0
r
a2
= > 0.
2a4
For an operator realization of the Hamiltonian H, constructed in a
similar way by means of H− 0 , we get that
r
a2
lim hφ̂(x, t)Φ, Φi = − < 0,
~→0 2a4
and hence, we have constructed two essentially different operator
realizations.
Note that we have encountered an important phenomenon called
symmetry breaking. The initial Hamiltonian function is invariant
with respect to the transformation φ(x) → −φ(x); however, the
classical vacuum is not invariant with respect to this transformation
(it transforms one classical vacuum into another). A similar situation
arises in the quantum case: the Hamiltonian H is invariant with
respect to the transformation φ̂(x) → −φ̂(x) and therefore, from
one operator realization of the Hamiltonian H, we can get another
by replacing the operators φ̂(x, t) by the operators −φ̂(x, t). If
hφ̂(x, t)Φ, Φi =
6 0, then this replacement gives an operator realization
that is not equivalent to the original.
Let us consider one more interesting example, the Hamiltonian
Z Z X
1X 2
2
H= π̂i (x)dx + φ̂2i (x) − a2 dx. (9.57)
2
Here,
we can write down the Hamiltonian (9.35) in the form (8.2). This
allows us to apply the considerations of this section to the analysis
of the Hamiltonian (8.2).
March 27, 2020 15:20 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch10 page 191
Chapter 10
191
March 27, 2020 15:20 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch10 page 192
(10.1)
Z
dp
f˜j (x|t) = exp(−iω(p)t + ipx)fj (p) , (10.2)
(2π)3
and the relation (10.4) implies the stability of the vacuum and single-
particle states:
Sθ = θ,
Sa+ (k)θ = a+ (k)θ.
Lemma 10.1. The set of vectors of the form BΦ, where B runs over
all good operators in the algebra A, is dense in the single-particle
space H1 .
1
The algebras A and A0 asymptotically commute if every operator in A
asymptotically commutes with every operator in A0 .
March 27, 2020 15:20 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch10 page 197
of the product Aλ1 (x1 , t1 ) · · · Aλn (xn , tn ) in the ground state Φ will
be called the (n-point) Wightman function (see Section 7.1). Let us
define the truncated Wightman function wnT (λ1 , x1 , t1 , . . . , λn , xn , tn )
as the truncated vacuum expectation value of this product:
In other words,
n X
X
wn (λ1 , x1 , t1 , . . . , λn , xn , tn ) = wαT1 (π1 ) · · · wαTk (πk ).
k=1 ρ∈Rk
(10.11)
w3 (λ1 , x1 , t1 , λ2 , x2 , t2 , λ3 , x3 , t3 )
= w3T (λ1 , x1 , t1 , λ2 , x2 , t2 , λ3 , x3 , t3 )
+ w1T (λ1 , x1 , t1 )w2T (λ2 , x2 , t2 , λ3 , x3 , t3 )
+ w1T (λ2 , x2 , t2 )w2T (λ1 , x1 , t1 , λ3 , x3 , t3 )
+ w1T (λ3 , x3 , t3 )w2T (λ1 , x1 , t1 , λ2 , x2 , t2 )
+ w1T (λ1 , x1 , t1)w1T (λ2 , x2 , t2)w1T (λ3 , x3 , t3).
Lemma 10.3. If all the operators Aλi are smooth, then the truncated
Wightman function
in the expression for hA0 (f0 , t) . . . An (fn , t)iT some of the operators
Aj (fj , t) are replaced by their time derivatives Ȧj (fj , t) and the
operators (Aj (fj , t))∗ are replaced by (Ȧj (fj , t))∗ .
Z
= f˜(x|t) exp(itω(p) − ipx)φ(p)Φ(p)dxdp
Z
= f (p)φ(p)Φ(p)dp, (10.13)
d d
Ψ(B|f |t) = B(f, t)Φ = 0. (10.14)
dt dt
i.e.
Z Z
S∓ ( f (p)φ(p)dpθ) = f (p)φ(p)Φ(p)dp. (10.15)
h(Ḃi (fi , t))∗ Ḃj (fj , t)iT = h(Ḃi (fi , t))∗ Ḃj (fj , t)Φ, Φi
or the form
h(Ḃi (fi , t))∗ Ḃj (fj , t)iT = h(Bi (fi , t))∗ Ḃj (fj , t)Φ, Φi
and both expressions are zero by the relation (10.14). Hence, all
non-zero terms in the sum above contain either a factor Ik,l with
March 27, 2020 15:20 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch10 page 202
for k + l = 3, and
where the functions fi are smooth and finite and the functions φi
correspond to good operators.
Therefore, let us now prove that the relation (10.1) defines S∓
uniquely and, furthermore, this operator can be uniquely extended
to a linear isometric operator defined on the whole space Has . To
show this, we will use the following general statement.
Let us suppose that on the total2 subset L of the Hilbert space
H, we have a multi-valued isometric mapping α taking values in H0
(i.e. to every point ξ ∈ L corresponds a set of points α(ξ) ⊂ H0 in
such a way that for any selection of points ξ1 , ξ2 ∈ L, x1 , x2 ∈ H0
satisfying x1 ∈ α(ξ1 ), x2 ∈ α(ξ2 ) we have hx1 , x2 i = hξ1 , ξ2 i). Then
2
A subset is called total if the set of its linear combinations is dense in H.
March 27, 2020 15:20 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch10 page 203
The totality of the set L follows from the fact that functions of
the form φB (k)f (k), where f (k) is a smooth function with compact
support and B is a good operator, are dense in L2 (E 3 ) (as we have
noted above, this fact follows from Lemma 10.1).
To prove relation (10.17), let us represent the left side of this
relation in the form
and then expand the expression under the limit sign in terms
of truncated Wightman functions. Every term of the resulting
expansion consists of a product of factors of the form
T
Rk,l (t) = (Bi01 (fi1 , t))∗ · · · (Bi0k (fik , t))∗ Bj1 (fj1 , t) · · · Bjl (fjl , t) .
It follows from Lemma 10.4 that for k + l ≥ 3 the factors Rk,l (t)
converge to zero as t → ±∞. On the other hand, the factors Rk,l are
equal to zero if k = 0 or l = 0, by the relation BΦ = 0. Therefore,
the terms that differ from zero in the limit t → ∓∞ are the ones
where all the factors have the form R1,1 (t). It is easy to check that
T
R1,1 (t) = (Bi0 (fi0 , t))∗ Bj (fj , t) = (Bi0 (fi0 , t)∗ Bj (fj , t))
and that R1,1 in fact does not depend on t [in the calculations we
have used the formula (10.13)].
The statements proven above imply that the expression (10.18)
n
P Qn 0 0
is equal to δm s=1 hφs fs , φis fis i [the sum is taken over all
permutations (i1 , . . . , in )]. This justifies equation (10.17).
March 27, 2020 15:20 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch10 page 205
we obtain
where
Applying (10.21) and noting that the left part of (10.1) does not
depend on the choice of the function Φ(k) representing the particle,
we obtain that the Møller matrices S± do not depend on the choice
of the function Φ(k).
has support that does not intersect the multi-particle spectrum and
the half-space ω ≤ 0. Then the operator
Z
B = α(t, x)A(x, t)dxdt,
where A ∈ A, is good.
March 27, 2020 15:20 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch10 page 208
= α̂(−H, −P)A∗ Φ = 0,
since the support of the function α̂(−ω, −p) does not intersect the
spectrum of the operators H, P.
Let us now show that vectors BΦ, where B is a good operator,
constructed by the method above, are dense in the single-particle
subspace. R
To prove this, let us consider the vector Φ(λ) = λ(k)Φ(k)dk,
where λ(k) is a smooth function with compact support, in H1 .
We will construct a sequence of good operators Bn such that
Bn Φ → Φ(λ) (this is enough to prove the statement above since
smooth functions with compact support λ(k) areR dense in L2 (E 3 )
and therefore the corresponding vectors Φ(λ) = λ(k)Φ(k)dk are
dense in H1 ).
By the cyclicity of the vacuum vector there exists a sequence of
operators An ∈ A so that An Φ → Φ(λ). The necessary sequence
R of
good operators can now be constructed by setting Bn = α(t, x)
An (x, t)dxdt, where
Z
dωdp
α(t, x) = α̂(ω, p) exp[−i(ωt − px)] ,
(2π)4
the properties of the supports of the functions α̂ ensure that the
operators Bn are good, and the function α̂ satisfies the condition
March 27, 2020 15:20 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch10 page 209
= α̂(H, P)Φ(λ)
Z
= λ(k)α̂(H, P)Φ(k)dk
Z
= λ(k)α̂(ω(k), k)Φ(k)dk
Z
= λ(k)Φ(k)dk = Φ(λ).
3
In the proof, we have assumed that the function χ(p) is smooth. One can relax
this condition assuming that this function is m times continuously differentiable.
Then we have the inequality (10.23) for n = m.
March 27, 2020 15:20 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch10 page 211
When r ≥ 1.5
−2ρ+1 , we can show that |I1 | ≤ C|t|−3/2 .
March 27, 2020 15:20 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch10 page 212
where
Z
dp
f˜(x) = f (p) exp(ipx) ,
(2π)3
3/2
2π i X
G(x, t) = | det γjk |−1/2 exp λjk xj xk ,
it 2
j,k
∂ 2 ω(p0 )
γjk = ,
∂pj ∂pk
λjk is the inverse matrix of the matrix γjk .
The inequality (10.8) follows from the inequality (10.7) in the case
x ∈ U t and from the inequality (10.9) in the case x 6∈ U t.
Remark 10.1. With some extra conditions on the behavior of the
function ω(p) at infinity, we can prove Lemma 10.2 for any smooth,
rapidly decaying function f (p).
Proof of Lemma 10.3. Let us begin by proving the simple but
important identity that connects the truncated vacuum expectation
value
hAB(t)iT = hAB(t)Φ, Φi − hAΦ, Φi hBΦ, Φi
with the vacuum expectation value of the commutator [A, B(t)]. To
prove this, let us consider the smooth function h(ω), satisfying the
conditions h(ω) = 0 for ω ≤ 0 and h(ω) = 1 for ω ≥ δ [the number
δ > 0 is chosen such that all points of the spectrum of the operator
H, with the exception of 0, lie on the ray (δ, +∞)]. Let us note that
h(H) = 1 − P0 ,
(10.26)
h(−H) = 0,
where P0 is a projection operator on the vacuum vector Φ (these
equations can be checked by using the isomorphism between the
spaces H and L2 (M ), which sends the operator H to an operator
March 27, 2020 15:20 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch10 page 214
Indeed, we have
Z Z
fh (t) hAB(t)i dt = fh (t) hA exp(iHt)Bi dt
= hBh(−H)f˜(−H)Ai = 0. (10.29)
Here, we have used the relation (10.26) and the equations
Z
fh (t) exp(iωt)dt = h(ω)f˜(ω),
Z
f (t) exp(iωt)dt = f˜(ω).
wn (x1 , t1 , . . . , xn , tn )
(1) (2)
≈ wk (xi1 , ti1 , . . . , xik , tik )wl (xj1 , tj1 , . . . , xjl , tjl ).
March 27, 2020 15:20 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch10 page 216
where Cij , Dij are operators whose norms are bounded by a constant
not depending on xi , ti , ξi , σi . Using this fact, asymptotic commu-
tativity, and formula (10.30), we obtain the asymptotic factorization
property for the case K = {1, . . . , k}, L = {k+1, . . . , n}. We can show
that the general case can be reduced to this special case. Suppose the
set K consists of the points (i1 , . . . , ik ) and the set L consists of the
points (j1 , . . . , jl ), where i1 < · · · < ik , j1 < · · · < jl . If the quantity
d = mini∈K,j∈L |xi − xj | is large, we can use the approximation
that the lemma holds for functions wn−1 T and then show it for
T
functions wn . However, we will first prove the following geometric
lemma.
Let N be a set consisting of the points x1 , . . . , xn in Euclidean
space and let D be the diameter of N (i.e. D = maxxi ,xj ∈N |xi − xj |).
Then, we can partition N into two subsets P and Q in such a way
so that ρ(P, Q) = minxi ∈P,xj ∈Q |xi − xj | ≥ 2Dn .
We will prove this fact by induction on n. Let D = |xα − xβ |. Let
us delete from N the point xγ which does not coincide with xα or
with xβ , obtaining as a result the set N 0 , consisting of n − 1 points
and having diameter D. By the inductive hypothesis, we can split the
set N 0 into two sets P 0 and Q0 so that ρ(P 0 , Q0 ) = minxi ∈P,xj ∈Q |xi −
D
xj | ≥ 2n−1 . By the triangle inequality, the point xγ cannot satisfy
the inequalities ρ(P 0 , xγ ) = minxi ∈P 0 |xi − xγ | < 2Dn and ρ(Q0 , xγ ) =
minxi ∈Q0 |xi − xγ | < 2Dn simultaneously; for concreteness, suppose
that ρ(P 0 , xγ ) ≥ 2Dn . Then, it is clear that the necessary partition
can be constructed by taking P to be P 0 and taking Q to be
Q0 ∪ {xγ }.
Let us now suppose that Lemma 10.3 holds for functions wkT with
k < n. To prove the lemma for functions wnT , we need to estimate
wnT (x1 , t1 , . . . , xn , tn ) under the condition max1≤i≤j≤n |xi − xj | = D
(with the times t1 , . . . , tn fixed). Using the geometric lemma we just
proved, let us partition the set {1, . . . , n} into the subsets K and L
such that mini∈K,j∈L |xi − xj | ≥ 2Dn . To estimate the functions wnT ,
let us note that in the right-hand side of the relation (10.11) (the
recurrence relation defining the functions wnT ) all factors have the
form wαr (πr ) with the set πr containing both elements of K and L.
By induction, we can obtain an estimate of the form CD−m , where
m is arbitrary and C depends on m. It therefore follows that every
term on the right-hand side of (10.11) contains at least one factor of
the form just described and therefore admits an estimate of the form
CD−m , since the factors wαTi (πi ) do not exceed a constant, depending
only on kA1 k, . . . , kAn k and the number n (this statement becomes
clear if we note that
where the sum is taken over partitions ρ of the set {1, . . . , n} such
that every set πj is completely contained in either K or L.
It is easy to check that the sum in (10.32) equals the product
wk (xi1 , ti1 , . . . , xik , tik )wl (xj1 , tj1 , . . . , xjl , tjl ), where (i1 , . . . , ik ) are
elements of K and (j1 , . . . , jl ) are elements of L.
Hence, we see that
wnT (x1 , t1 , . . . , xn , tn )
≈ wn (x1 , t1 , . . . , xn , tn )
− wk (xi1 , ti1 , . . . , xik , tik )wl (xj1 , tj1 , . . . , xjl , tjl ).
To finish the proof of Lemma 10.3, we apply the asymptotic
factorization property of Wightman functions.
Remark 10.2. Because the operators Ai are smooth, the Wightman
functions wn (x1 , t1 , . . . , xn , tn ) and the truncated Wightman func-
tions wnT (x1 , t1 , . . . , xn , tn ) are smooth, and hence the derivatives of
these functions can be viewed as Wightman and truncated Wightman
functions constructed with different smooth operators. R (This is clear
if we note that for a smooth operator A = f (τ, x)B(x, τ )dξdτ ,
where f ∈ S(E 4 ), the following relation holds:
Z
D(α) A(x, t) = D(α) f (τ − t, ξ − x)B(ξ, τ )dξdτ
Z
= g(τ − t, ξ − x)B(ξ, τ )dξdτ = C(x, t),
∂ |α|
where D(α) = α α α
∂tα0 ∂x1 1 ∂x2 2 ∂x3 3
is a differential operator; g(t, x) =
(−1)|α| D(α) f (t, x);
R C is a smooth operator defined by C =
g(τ, ξ)B(ξ, τ )dξdτ ; operator derivatives are understood as norm
derivatives.) Using this fact and the statement of Lemma 10.3,
March 27, 2020 15:20 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch10 page 220
with the operators (Aj (fj , t))∗ or (Ȧj (fj , t))∗ requires no new
ideas.
S− a+ (k) = a+
in (k)S− ,
4
Under the assumption of strong convexity on the functions ω(k), the family
f1 , . . . , fn will be non-overlapping if and only if the supports of these functions
suppfj are pairwise non-intersecting. Therefore, in the case at hand, the definition
of non-overlapping family can be simplified. However, in the case when we have
multiple particles, the definition in the main text is necessary.
March 27, 2020 15:20 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch10 page 223
where
It follows from Lemma 10.2 that the function f˜j (x|t), for large t, is
small outside the set tU (fj ), where the symbol U (fj ) denotes an
-neighborhood of the set U (fj ). Under the assumed conditions, for
small enough and large enough t, the sets tU (fj ) are far apart.
Let us consider vectors of the form Ψ− (B1 , . . . , Bn |f1 , . . . , fn )
[correspondingly, of the form Ψ+ (B1 , . . . , Bn |f1 , . . . , fn )], where
B1 , . . . , Bn are good operators and f1 , . . . , fn are functions with
compact support. We will call them non-overlapping in-vectors (out-
vectors) if f1 , . . . , fn is a non-overlapping family of functions. The
set of linear combinations of non-overlapping in-vectors (out-vectors)
will be denoted by Din (correspondingly, Dout ); Din and Dout are
linear subspaces that are dense in Hin and Hout , respectively.
Let us now prove a theorem that describes the asymptotic
behavior of operators A(x, t), where A ∈ A, in terms of in- and
out-operators. For concreteness, we will formulate the theorem in
the case t → −∞.
March 27, 2020 15:20 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch10 page 224
in the form
Z
f1 (p1 , σ1 ) exp(−iσ1 ω(p1 )t) . . . fn (pn , σn )
w̃nT (p1 , σ1 , . . . , pn , σn )
Z X
= exp i pj xj hA1 (x1 , t, σ1 ) . . . An (xn , t, σn )iT dn x,
Ω(p2 , . . . , pn ) = σ1 ω(−p2 − · · · − pn )
+ σ2 ω(p2 ) + σ3 ω(p3 ) + · · · + σn ω(pn ).
March 27, 2020 15:20 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch10 page 226
1 ∂Ω ∂Ω −2
u(p) = ,
t ∂p2 ∂p2
and therefore
The proof in the case when the operators Ai (fi , t, σi ) are replaced
by their derivatives is similar to the proof of Lemma 10.4.
where
Ψ2 (t) = Ψ(B1 , . . . , Bm |g1 , . . . , gm |t).
It follows from the boundedness of kξ(t)k and kΨ2 (t)k that
lim hξ(t), Ψ2 i = lim hξ(t), Ψ2 (t)i.
t→−∞ t→−∞
Dθ = Φ, (10.43)
+
Da (k)θ = Φ(k). (10.44)
Theorem 10.1. Let us suppose that (a) the operator D has norm 1,
(b) for all non-overlapping families of functions with compact support
f1 (k), . . . , fn (k) and for all smooth operators A1 , . . . , An ∈ A, we
have, for t < 0,
t t
hDa+ (f 1 ) . . . a+ (f n )θ, A1 (g1 ) . . . An (gn )Φi
X
with error not exceeding cν n (g)|t|−N (here, fit (k) = fi (k) exp
(−iω(k)t), P is the set of permutations R π = (i1 , . . . , in ) of3 indices
(1, . . . , n), ν(g) = maxi (sup |gi (x)| + |gi (x)|dx); gi ∈ S(E ); N is
an arbitrary number; C is a constant depending on N but not on the
functions gi ). Then the operator D is in-dressing.
The condition for an operator to be out-dressing can be obtained
from the theorem above by replacing t < 0 with t > 0.
To prove the above theorem, let us consider the vector x = S− y,
where x = Ψ− (B1 , . . . , Bn |f1 , . . . , fn ); y = a+ (φ1 f 1 ) . . . a+ (φn f n )θ;
B1 , . . . , Bn are good operators and f1 , . . . , fn are non-overlapping
functions, and
x(t) = Ψ− (B1 , . . . , Bn |f1t , . . . , fnt )
t t
= S− a+ (φ1 f 1 ) . . . a+ (φn f n )θ.
It is clear that
t t
exp(−iHas t)y = a+ (φ1 f 1 ) . . . a+ (φn f n )θ (10.46)
and therefore, by the relation (10.5), we have
x(t) = exp(−iHt)x. (10.47)
For large t, the vector x(t) negligibly differs from the vector ξ(t) =
Ψ(B1 , . . . , Bn |f1t , . . . , fnt |0) = B1 (f1t , 0) . . . Bn (fnt , 0)Φ (to prove this
statement, note that for t → ±∞ we have
kx(t) − ξ(t)k = k exp(iHt)(x(t) − ξ(t))k
= kΨ− (B1 , . . . , Bn |f1 , . . . , fn )
− Ψ(B1 , . . . , Bn |f1 , . . . , fn |t)k → 0). (10.48)
On the other hand, by relation (10.46), we have that
lim hD exp(−iHas t)y, ξ(t)i
t→−∞
t t
= lim hDa+ (φ1 f 1 ) . . . a+ (φn f n )θ, B1 (f1t , 0) . . . Bn (fnt , 0)Φi
t→−∞
X
Φ(φ1 f1t ), Bi1 (fit1 , 0)Φ . . . Φ(φn fnt ), Bin (fitn , 0)Φ
= lim
t→−∞
π∈P
= hy, yi . (10.49)
March 27, 2020 15:20 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch10 page 231
constitute a total set in the space Has , it is clear that the relation
holds for a dense set of vectors and hence for any vector y ∈ Has (see
Appendix A.5).
We have thus obtained sufficient conditions for an operator D to
be in-dressing and out-dressing. These conditions can be relaxed by
changing the requirement that kDk = 1 with
t t t t
lim kDa+ (f 1 ) . . . a+ (f n )θk = ka+ (f 1 ) . . . a+ (f n )θk (10.51)
t→−∞
(t→+∞)
holds. (In (10.53), we assume that the functions f˜1 (x), . . . , f˜n (x) ∈ S
have distant supports; the error should not exceed
Cν n (f )d−N ,
(1) T Φ = Φ̃,
(2) T Φ(k) = Φ̃(k),
(3) if A ∈ A, then T −1 AT ∈ A (here, Φ and Φ̃ are ground states
and Φ(k) and Φ̃(k) are single-particle states corresponding to the
energy operators H and H̃).
If D is a dressing operator for the operator H, satisfying the
conditions of Theorem 10.2, then the operator D̃ = T D is a dressing
operator for the energy operator H̃.
The proof of this theorem consists of checking that the conditions
of Theorem 10.2 apply to the operator D̃. This verification can be
March 27, 2020 15:20 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch10 page 234
depicts the collision process where we begin (at t → −∞) in the state
Here, sn,k are functions that are related to the matrix elements Sn,k
of the scattering matrix by the relation
f (k) and g(k) are normalized wave functions differing from zero only
in a small neighborhood of the points q and 0, respectively; α is a
vector orthogonal to the vector v(q) (as explained in Section 5.2, this
vector is analogous to the impact parameter in classical mechanics.)
The collision process is described by the vector
xα (t) = exp(−iHt)S− ξα ,
and at the end of the process (with t → +∞) we obtain the state
D exp(−iH0 t)(Sξα ).
The probability that for initial state (10.57) we obtain the final
state Sξα of n particles with momenta (p1 , . . . , pn ), belonging to
March 27, 2020 15:20 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch10 page 238
X 1 Z
Sξα = Sm,2 (p1 , . . . , pm |k1 , k2 )f (k1 )
m
m!
+ +
× exp(ik1 α)g(k2 )a (p1 ) . . . a (pm )θdp1 . . . dpm dk1 dk2 .
5
Strictly speaking, this formula contains a limit: the probabilities wα (G) should
be defined with the functions fν (k), gν (k), whose supports contract to the points
q, 0, respectively, as ν → ∞.
March 27, 2020 15:20 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch10 page 239
Z Z
1
+ dp1 . . . dpn dk1 dk2 (sn,2 (p1 ,
n! G
. . . , pn |k1 , k2 )s̄n,2 (p1 , . . . , pn |β(k1 , k2 ), γ(k1 , k2 ))
× f (k1 )f¯(β(k1 , k2 )g(k2 )ḡ(γ(k1 , k2 ))α(k1 , k2 )
× δ(p1 + · · · + pn − k1 − k2 )δ(ω(p1 )
+ · · · + ω(pn ) − ω(k1 ) − ω(k2 ))).
To obtain the final expression for σG , we should use the fact that
for normalized functions f (k) with support in a small neighborhood
of the point q and for functions λ(k), continuous at the point q, we
have the relations
Z
λ(k)|f (k)|2 dk = λ(q),
Z
λ(k)f (k)f¯(ρ(k))dk = 0 (10.61)
10.5 Generalizations
In this section, we describe some generalizations of the results of
previous sections of this chapter.
1. The condition of strong convexity imposed on the dispersion
law ω(p) can be substantially relaxed. Namely, it is enough to require
that for any open set G ⊂ E 3 the function ω(p) is not linear (i.e. for
any neighborhood of a point p0 ∈ E 3 there is another point p1 ∈ E 3
such that ∂ω(p) ∂ω(p)
∂p |p=p1 6= ∂p |p=p0 ).
In this case, the definition of the Møller matrix should be modi-
fied: in the formula (10.1), we should consider only non-overlapping
families of functions f1 , . . . , fn . The main modification required in the
proofs is to use Lemma 10.5 of Section 10.3 in place of Lemma 10.2
of Section 10.1.
These modifications in the definitions and proofs allow us to
construct a scattering theory in the one- and two-dimensional cases,
March 27, 2020 15:20 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch10 page 241
i.e. in the case when the momentum operator P has only one or
two components instead of three. (In higher dimensions, there are no
essential modifications.)
2. Up to this point, we have only considered situations with a
single one-particle state. The generalization to an arbitrary number
of one-particle states is straightforward.
A good operator B, in this case, can be defined as a smooth
operator with (1) B ∗ Φ = 0; (2) we can findR a particle Φi (k) and a
function φB (k) for which BΦ = Φi (φB ) = φB (k)Φi (k)dk. [Recall
that in Section 10.1 we have agreed to fix a complete orthogonal
system of particles Φ1 (k), . . . , Φs (k).]
The modifications of the statements and their proofs consist
mostly of adding indices; perhaps the most important change comes
from the need to use the word “total” instead of “dense” in the
formulation of Lemma 10.1.
It is easy to check that the Møller matrices and the scattering
matrix do not depend on the choice of complete system of particles
(i.e. they depend only on the operators H, P, and the algebra A).
3. All the constructions in this chapter can be adapted, with
the appropriate modifications, to the case when we replace the
condition of asymptotic commutatitivty by the condition of asymp-
totic anticommutatitivity of the family of operators A (asymptotic
anticommutativity means that for any two operators A, B ∈ A and
any n, we can find C and r such that
1 + |t|r
k [A, B(x, t)]+ k ≤ C .
1 + |x|n
If we have an asymptotically anticommutative family A, we
impose an additional condition that any vector A1 . . . A2k+1 Φ, where
Ai ∈ A, is orthogonal to the vacuum Φ (this condition becomes
necessary to construct the scattering matrix). The space H can then
be expanded into a direct sum of two subspaces Hg and Hu in such a
way that the vector A1 . . . An Φ belongs to the space Hg , if n is even,
and to the space Hu , if n is odd. We will call vectors in the spaces
Hg and Hu even and odd, respectively.
Let us use the symbol à to denote the smallest algebra of
operators that contains the family A. The algebra Ã, like the linear
March 27, 2020 15:20 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch10 page 242
where N1 is the set of even particles and N2 is the set of odd particles.
In the case considered in this subsection, the construction of the
scattering matrices is performed in the same way as when the family
A is asymptotically commutative.
In the situation at hand, the elementary particles are fermions
while bosons arise only as composite particles. If the elementary
particles can be either fermions or bosons, then the family A must
March 27, 2020 15:20 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch10 page 243
(10.63)
6
The second part of the condition C2 implies that the function hA1 (x1 , t1 ) . . .
Ar (xr , tr )Φ, Φi is a locally summable function of moderate growth.
March 27, 2020 15:20 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch10 page 245
where
fit (k) = fi (k) exp(−iω(k)t),
Z
B(fi |t) = fit (k)B(k|t)dk,
t
B(k|0)Φ = φ(k)Φ(k).
The proof of this statement follows the reasoning in Section 10.1.
Let us now define the notion of asymptotically commutative
families of operator generalized functions.
March 27, 2020 15:20 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch10 page 246
D4. For every function A(x, t) in family B, the family also contains
the adjoint operator generalized function A+ (x, t).
D5. For any operator generalized functions A1 (x, t), . . . , Ar (x, t),
A(x, t), B(x, t) in B, we can find a δ > 0 such that the real-valued
March 27, 2020 15:20 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch10 page 247
generalized function
T (x, t, ξ1 , τ1 , . . . , ξr , τr )
= h[A(x + a, t + α), B(a, α)]A1 (ξ1 , τ1 )
. . . Aj (ξj , τj )Φ, Aj+1 (ξj+1 , τj+1 ) . . . Ar (ξr , τr )Φi
T (x, t, ξ1 , τ1 , . . . , ξr , τr )
!k
X X X
= D(α) 1 + |ξi |2 + |τi |2
|α|≤N (n) i i
where
a+ (k, t) = exp(iH0 t)a+ (k) exp(−iH0 t)
= exp(iω(k)t)a+ (k),
a(k, t) = exp(iH0 t)a(k) exp(−iH0 t) exp(−iω(k)t)a(k),
constitute an asymptotically commutative family B0 of generalized
functions (as always, we assume that these operator generalized
March 27, 2020 15:20 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch10 page 249
7
Conditions T1–T3 mean that the set A is a complete locally convex topological
algebra with involution.
March 27, 2020 15:20 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch10 page 251
respectively].
0 0
Lemma 10.9. For any norm kf kα,β on the space S(E 3(m+m +n+n −2r) )
and any number N , we can find norms kf kγ,δ and kf kγ 0 ,δ0 on the
0 0
spaces S(E 3(m+n) ) and S(E 3(m +n ) ) and a number C so that
C
kλr (Vx f, g)kα,β ≤ kf kγ,δ kgkγ 0 ,δ0 .
1 + |x|N
[Here, r > 0 and λr (f, g) denotes the function
Z
f (k1 , . . . , km |p1 , . . . , pn−r , q1 , . . . , qr )
It follows from Lemma 10.8 that for any seminorm p(A) on the
space A(m+n) we can find a seminorm q(A) on A(m+n) and a number
k so that for any operator A ∈ Am,n we have
q(A)q 0 (B)
p([A(x), B]) ≤ .
1 + |x|N
the equation
dSα (t, t0 )
i = exp(iH(0)t)(H(h(αt)) − H(0)) exp(−iH(0)t)Sα (t, t0 )
dt
with the initial condition Sα (t0 , t0 ) = 1.
We will similarly use U (t, t0 |g(τ )) to denote the evolution operator
corresponding to the time-dependent Hamiltonian H(g(τ )), where
g(τ ) is a function taking values in [0, 1]; using this notation, we can
write Uα (t, t0 ) = U (t, t0 |h(ατ )).
We will call the operators Sα (0, ±∞) = slimt→±∞ Sα (0, t) the
adiabatic Møller matrices and the the operator Sα = Sα (∞, −∞) =
slim t→∞ Sα (t, t0 ) will be called the adiabatic S-matrix. The con-
t0 →−∞
struction of the adiabatic S-matrix and the adiabatic Møller matrices
provided in Section 4.1 for the pair of operators H, H0 corresponds
to H(g) = H0 + g(H − H0 ) and h(τ ) = exp(−|τ |).
In what follows, for the sake of simplifying the proofs, we will
assume that the function h(τ ) satisfies a few somewhat stronger
conditions than we have assumed above. Namely, we will assume
that this function is smooth, even, has compact support, and satisfies
h(0) = 1 in a neighborhood of the point τ = 0. The radius of this
neighborhood will be denoted by the symbol δ and the radius of the
support of the function h(τ ) will be denoted by ∆ (i.e. h(τ ) = 1 for
|τ | < δ and h(τ ) = 0 for |τ | > ∆).
Let us now consider the situation when the energy operator in
an axiomatic scattering theory depends on a parameter g. More
precisely, suppose that we have an energy operator H(g) that acts on
the Hilbert space H, where the parameter g is in the interval [0, 1], a
momentum operator P = (P1 , P2 , P3 ), and a family of operators A,
satisfying the following conditions:
(1) H(g)Φ = PΦ = 0;
(2) H(g)Φ(k|g) = ω(k|g)Φ(k|g),
where ω(k|g) is a positive function, infinitely differentiable in k, g
and strongly convex in k;
(3) PΦ(k|g) = kΦ(k|g);
(4) hΦ(k|g), Φ(k0 |g)i = δ(k − k0 );
(5) for any k0 ∈ E 3 , 0 ≤ g0 ≤ 1, we can find an operator A ∈ A,
such that hΦ(k0 |g0 ), AΦi =
6 0;
(6) for any k0 ∈ E 3 , 0 ≤ g0 ≤ 1, there exist > 0, δ > 0, so that for
any point (ω, k) belonging to the multi-particle spectrum of the
operators (H(g), P) and satisfying |k − k0 | < δ, |g − g0 | < δ, we
have the inequality ω > ω(k0 |g0 ) + .
(b) Let us denote by D the set of vectors of the form AΦ where
A ∈ A. We assume that D is dense in H and invariant with respect
to the operators exp(iPx) and U (t, t0 |g(τ )), where g(τ ) is a smooth
function taking values in [0, 1]. For any two operators A, B ∈ A,
the linear combination λA + µB, the product AB, and the adjoint
operators A+ , B + all belong to A (in other words, the family A is
an operator algebra with an involution). We will assume that A is
equipped with a locally convex topology, in which A is complete and
the product and involution operations are continuous.
The topology on A should satisfy the following conditions:
(1) hAΦ, Φi continuously depends on A ∈ A;
(2) if A ∈ A, g(τ ) is a smooth function with values in [0, 1], x ∈ E 2 ,
and −∞ < t0 , t1 < ∞, then the operator
∂m
p A(x, t|g) ≤ (1 + |x|k + |t|k )q(A), (10.69)
∂g m
q(A)q(B)
p([A(x), B]) ≤ . (10.70)
1 + |x|n
where
Z 0
ρ(k) = (ω(k|h(σ)) − ω(k|0))dσ,
−∞
−1
ain (k) = S− (0)a(k)S− (0),
−1
aout (k) = S+ (0)a(k)S+ (0)
are in- and out-operators, constructed for the energy operator H(0).
The proof of these relations (the adiabatic theorem) is the main result
of this section.
Of course, one can easily state the analog of relations (10.71) and
(10.72) for the operators S± (g), where 0 ≤ g ≤ 1; we will not stop
to do this.
The relations (10.71) and (10.72) are clearly equivalent to the
relations
1
ρ(k) = lim (rα (k|t) − tω(k|0))).
α t→−∞
k, such that
dk
Z
φ(x, t)A(x, t|g)dxdt
dg k
continuously depends on φ ∈ T , A ∈ A, g ∈ [0, 1] in the topology
on A.
Let us suppose that Bg1 , Bg2 are two families of good operators that
are infinitely differentiable in g in the topology of A; the functions
r1 (k|g) and r2 (k|g) will be defined by the relation
Z
i
Bg Φ = ri (k|g)Φ(k|g)dk.
It follows from relation (10.75) and Lemma 10.12 that the function
r1 (k|g)r2 (k|g) is infinitely differentiable in k and g.
Let us now show that for all points (k0 , g0 ) we can find an
infinitely differentiable family of good operators Bg , such that
hBg0 Φ, Φ(k0 |g0 )i =
6 0.
We will use the fact that for every operator A ∈ A and every
smooth function with compact support σ(k, ω|g), which is equal to
0 for ω ≤ 0 and for any point (ω, k) belonging to the multi-particle
spectrum of the operators H(g), P, we can construct a family of good
operators Bgσ,A by means of the formula
Z
Bgσ,A = σ̃(x, t|g)A(x, t|g)dxdt,
where
Z
1
σ̃(x, t|g) = exp(ikx − iωt)σ(k, ω|g)dkdω
(2π)4
(see Lemma 10.1 in Section 10.1). By Lemma 10.11, this family is
infinitely differentiable in g. It is easy to check that
σ,A
Bg Φ, Φ(k|g) = σ(k, ω(k|g)|g) hAΦ, Φ(k|g)i . (10.76)
March 27, 2020 15:20 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch10 page 261
is infinitely differentiable.
It follows from relation (10.76) that the same choice of the
function λ(k|g) implies the differentiability of all functions of the
form exp(iλ(k|g)) hAΦ, Φ(k|g)i, where A ∈ A; this finishes the proof
of the lemma.
Remark 10.4. Using the same arguments, one can prove, without
requiring condition a(5), that for any operator A ∈ A, the function
|hΦ(k|g), AΦi|2 is infinitely differentiable. This allows us to consider
the value of the function |hΦ(k|g), AΦi| at individual points. There-
fore, condition a(5) has a precise meaning.
Let us note that for every point (k0 , g0 ) there exists a neighbor-
hood U , an operator A ∈ A, and a function σ, such that the function
hBgσ,A , Φ, Φ(k|g)i is non-zero in the neighborhood U (here, Bgσ,A is a
family of good operators, constructed in the proof of Lemma 10.12).
If the support of the function f (k|g) is contained in the neighborhood
U , then the necessary family Bgf can be obtained with the help of
relation (10.77), using Bgψ = Bgσ,A , Bgψφ = Bgf in the relation. In other
words, we should choose function (10.76) as the function ψ(k|g) and
March 27, 2020 15:20 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch10 page 263
the function
f (k|g)
hBgσ,A Φ, Φ(k|g)i
as the function φ(k|g) (clearly, the latter function is smooth and has
compact support if we define it to be zero at the values where both
the numerator and the denominator are zero).
An arbitrary function f can be represented as a finite sum of
functions fi , for which the family Bgfi can be obtained by the
construction above. It is clear that the operators
X
Bgf = Bgfi
(Bgf )∗ Φ = 0.
where
Z
1
µ̃(x, t|g) = µ(k, ω|g) exp(ikx − iωt)dkdω,
(2π)4
µ(k, ω|g) is a smooth function with compact support, equal to 1 if
(k, g) ∈ supp f , ω = ω(k|g).
Let us first note that, by Lemma 10.14, the vector Φ(f |g) can
be represented in the form Bgf Φ, where the operator Bgf is infinitely
|g)
differentiable in g; hence, the derivative dΦ(f
dg exists. Furthermore,
the bilinear form
0 dΦ(f |g) 0
A(f, f ) = , Φ(f |g)
dg
is translation-invariant [i.e. A(f, f 0 ) = A(fx , fx0 ), where fx (k) =
exp(−ikx)f (k)] and, therefore, can be written in the form (10.78).
Let us consider the function
dΦ(f |g) 0
ζ(x|g) = , Φ(fx |g)
dg
dΦ(f |g) 0
= , exp(−iPx)Φ(f |g)
dg
* +
dBgf f0
= Φ, Bg (x)Φ .
dg
It is clear that
Z
ζ(x|g) = exp(ikx)ν(k|g)f (k)f 0 (k)dk. (10.79)
It follows from relation (10.79) and Lemma 10.12 that the function
ν(k|g)f (k)f 0 (k) is smooth. Since f (k), f 0 (k) are arbitrary smooth
functions, the proof of the lemma follows.
It follows from Lemma 10.15 that the single-particle state Φ(k|g)
can be chosen in such a way that for every f, f 0 we have
dΦ(f |g)
, Φ(f 0 |g) = 0. (10.80)
dg
[If equality (10.80) is not satisfied, then we can satisfy it if we replace
Φ(k|g) with the generalized vector function
Φ̃(k|g) = exp(iτ (k|g))Φ(k|g),
where τ (k|g) is a solution of the equation
∂τ (k|g)
i = −ν(k|g).
∂g
March 27, 2020 15:20 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch10 page 265
Lemma 10.15 implies that τ (k|g) is a smooth function and hence the
function hAΦ, Φ̃(k|g)i is smooth for all A ∈ A.
In what follows, we will assume that the single-particle state
Φ(k|g) is chosen such that it satisfies condition (10.80).
Let us denote by t0 a moment in time, satisfying the relation
t0 ≤ − ∆α (recall that the symbol ∆ denotes the support radius of the
fixed function h(τ ), so that h(τ ) = 0 for τ ≤ αt0 ).
Lemma 10.16. For any smooth function with compact support f (k),
we have
f,σ
We will prove this lemma by induction. Suppose the operator Di−1
is given. Let us consider the vector ζi (f |σ), defined by the relations
dξi−1 (f |σ)
i = (H(h(σ)) − ω(P|h(σ)))ζi (f |σ),
dσ
ζi (f |σ), Φ(f 0 |h(σ)) = 0,
where γ̃(t, x|g) is the Fourier transform of γ(ω, p|g) in the variables
f,σ f,σ
ω, p and Ei−1 = (d/dσ)Di−1 . Equation (10.88) implies that ζi (f |σ) =
f,σ f,σ R f,σ
Fi Φ, where Fi = γ̃(t, x|h(σ))Ei−1 (x, t|σ)dxdt (this operator is
infinitely differentiable in σ, by Lemma 10.11). The vector ξi (f |g)
can be expressed in terms of the vector ζi (f |σ) in the following way:
R R
relations ξi (f |σ) = f (k)ξi (k|σ)dk, ζi (f |σ) = f (k)ζi (k|σ)dk, the
formulas (10.89) and (10.90) can be written in the form
where µσi (k) = β(k)λσi (k) and the symbol β(k) denotes a smooth
function with compact support, equal to 1 when k ∈ suppf ).
Remark 10.5. For all σ in the interval [−δ, δ], where the function
h equals 1, we can show the following equations by induction:
dξi (f |σ)
= 0; ζi (f |σ) = 0; Fif,σ = 0;
dσ
∂λσi (k) λσ f
= 0; Dif,σ = B1 i .
∂σ
Let us define the functions φ̃t,α (x), φt,α (k) by the formulas
Z
t,α −3
φ̃ (x) = (2π) φt,α (k) exp(ikx)dk,
implies that
k1 − k2
e= , ν(ζ) = rα (k2 + ζe|t)
|k1 − k2 |
C|τ |r
p([Uα (t + τ, t)Q1,t 2,t
s Uα (t, t + τ ), Qs ]) ≤
1 + |t|n
Qi,t t
s = Qs (fi , φi , α),
Ψn (t) = Uα (t, t0 )S− (0)Wα (t0 )b+ (f¯1 φ̄1 ) . . . b+ (f¯n φ̄n )θ
where
Qi,t t
s = Qs (fi , φi , α),
Ψn (t + 1) = Uα (t + 1, t)Ψn (t)
≈ Uα (t + 1, t)Q1,t n,t
s . . . Qs Φ
= Uα (t + 1, t)Q1,t n−1,t
s . . . Qs Uα (t, t + 1)Uα (t + 1, t)Qn,t
s Φ
≈ Uα (t + 1, t)Q1,t n−2,t
s . . . Qs Uα (t, t + 1)Uα (t + 1, t)
× Qn−1,t
s Uα (t, t + 1)Qn,t+1
s Φ (10.99)
Uα (t + 1, t)Q1,t n−1,t
s . . . Qs Uα (t, t + 1)Rs (t, t + 1),
March 27, 2020 15:20 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch10 page 273
last statement follows from (10.91)]. Using these remarks and the
inequalities |t| ≤ |t0 |, s ≥ 32 n + 1, we can show that as we go
from (10.98) to (10.99) the error does not increase by more than
Cα2 .
By Lemma 10.21, we can see that
Ψn (t + 1) ≈ Uα (t + 1, t)Q1,t n−2,t
s . . . Qs
× Uα (t, t + 1)Qn,t+1
s Uα (t + 1, t)Qn−1,t
s Φ. (10.100)
Similarly, the additional error obtained from transferring to (10.100)
from (10.99) does not exceed Cα2 (moreover, this error does not
exceed Cαr , where r is arbitrary). To prove this, we should use, in
addition to Lemma 10.21, the inequality t < −aα− and the estimate
for Qi,t
s shown above.
Using equation (10.95) again, we obtain
Ψn (t + 1) ≈ Uα (t + 1, t)Q1,t
s
. . . Qn−2,t
s Uα (t, t + 1)Qn,t+1
s Qn−1,t+1
s Φ.
Applying equation (10.95) several times and then applying
Lemma 10.21, we obtain
Ψn (t + 1) ≈ Qn,t+1
s . . . Q1,t+1
s Φ,
≈ Q1,t+1
s . . . Qn,t+1
s Φ. (10.101)
The error arising from each transformation does not exceed Cα2 .
Hence, we have derived (10.101) from (10.98) and have proven that
the error in (10.101) also does not exceed Cα2 . The inductive step
is shown. Since the case t = t0 for equation (10.97) is obvious,
Lemma 10.22 is proven.
We can now finally easily show the main result of this section.
In the formulation of Lemma 10.22, let us take t = − αδ , where δ
is such that h(τ ) ≡ 1 for −δ ≤ τ ≤ δ. Then we can show that
φ̃t,α
i (x) = φ̃i (x, t)
Z
−3
= (2π) φi (k) exp(ikx − iω(k|1)t)dk, (10.102)
March 27, 2020 15:20 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch10 page 274
n Z s
αi Djfi ,−δ (x) dx Φ,
Y X
Ψn (t) ≈ φ̃i (x, t)
i=1 j=0
Chapter 11
Translation-Invariant Hamiltonians
(Further Investigations)
275
March 27, 2020 16:3 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch11 page 276
(11.1)
where a+ in (k,
R τ ), ain (k, τ ) are in-operators, defined in Section 9.2;
µ̃(ω(k)) = µ(τ ) exp(iω(k)τ )dτ .
Relations (9.18) and (9.21) imply that
Z
BΦ, a+ λ(k)µ(τ ) a+ (k, τ )Φ, a+
in (p)Φ = in (p)Φ dkdτ
Z
λ(k)µ̃(ω(k)) a+ (k)Φ, a+
= in (p)Φ dk
−1
= λ(p)µ̃(ω(p))ρ(1, p) = λ(p)µ̃(ω(p))Λ (p).
(11.2)
H = H0 + V
Z X XZ
+
= (k)a (k)a(k)dk + gr (r)
vm,n (k1 , . . . , km |p1 , . . . , pn )
r≥1 m,n
1 P n
In other words, we will consider formal expressions n An g , where An are
numbers, vectors, or operators; if An are operators, we will assume that they
are all defined on the same P domain. ThePsums and products of two (number
n
or
P operator) formal series n An g and n Bn g n are understood as the series
n P P n
n (An + Bn )g and n( k+l=n Ak Bl )g . The rest of this section will be
considered in the framework of perturbation theory (without always stating so).
In order to give precise mathematical meaning to the results in this section and
their proofs, all the relations should be understood in terms of formal power series.
March 27, 2020 16:3 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch11 page 278
2
g r Ur (t) constitute a group of unitary operators if
P
The operators
X
Uk (t)Ul (t0 ) = Ur (t + t0 ),
k+l=r
3
The same modification of the Møller matrix definition is used for a different
purpose in Section 10.5, part 5.
March 27, 2020 16:3 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch11 page 280
A similar situation arises for all diagrams without outer vertices (for
vacuum loops): in the expression corresponding to the diagram, the
integrand contains a product of δ-functions whose arguments are
linearly dependent, rendering the expression meaningless.
Therefore, we cannot write a solution to the Schrödinger equation
i dψ
dt = Hψ even in perturbation theory.
However, using the expression (11.3), we can formally write
the Heisenberg equations which remain meaningful in perturbation
theory.
Indeed, the Heisenberg equations,
∂a(k, t)
= i[H, a(k, t)], (11.4)
∂t
∂a+ (k, t)
= i[H, a+ (k, t)], (11.5)
∂t
can be written in the form
∂a(k, t)
i − (k)a(k, t) = I(k, t), (11.6)
∂t
∂a+ (k, t)
−i − (k)a+ (k, t) = I + (k, t), (11.7)
∂t
March 27, 2020 16:3 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch11 page 281
where
X X Z
r (r)
I(k, t) = g m vm,n (k, k1 , . . . , km−1 |p1 , . . . , pn ) (11.8)
r≥1 m,n
m−1
X n
X
× δ k + ki − pj a+ (k1 , t) (11.9)
i=1 j=1
a(k, t) = exp(−i(k)t)a(k)
Z t
+ exp(−i(k)(t − τ ))I(k, τ )dτ, (11.11)
0
a(k, t) = exp(−i(k)t)a(k)
X XZ
+ gr r,t
fm,n (k1 , . . . , km |p1 , . . . , pn )
r≥1 m,n
m
X n
X
× δ k + ki − pj a+ (k1 ) · · · a+ (km )a(p1 )
i=1 j=1
a+ +
k (t) = exp(iHΩ t)ak exp(−iHΩ t).
The operators a+k (t), ak (t), for fixed t, clearly satisfy CCR; the trans-
formation from the operators a+ +
k , ak to the operators ak (t), ak (t)
specifies a canonical transformation in the volume Ω. Let us write
March 27, 2020 16:3 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch11 page 285
the operators a+
k (t), ak (t) in normal form:
ak (t) = exp(−i(k)t)ak
X X
+ gr Ω r,t
fm,n (k1 , . . . , km |p1 , . . . , pn )
r≥1 ki ∈TΩ
pj ∈TΩ
3 (m+n−2)
2π 2
× δpk11 +···+p
+···+km
n
a+ +
k1 · · · akm ap1 · · · apn ,
L
a+ +
k (t) = exp(i(k)t)ak
Ω r,t
X X
+ gr f m,n (k1 , . . . , km |p1 , . . . , pn )
r≥1 ki ∈TΩ
pj ∈TΩ
3 (m+n−2)
2π 2
× δpk11 +···+p
+···+km
n
a+ +
p1 · · · apn ak1 · · · akm .
L
r,t r,t
For the functions Ω fm,n , just as for the functions fm,n , we can derive
recurrent formulas from the Heisenberg equations. It is easy to check
that the recurrent formulas admit the limit Ω → ∞, and therefore,
it follows that
Ω r,t r,t
lim fm,n = fm,n .
Ω→∞
where
H̃ = H0 + Ṽ , (11.23)
where
X XZ
r (r)
Ṽ = g ṽm,n (k1 , . . . , km |p1 , . . . , pn )
r m,n
m
X n
X
×δ ki − pj a+ (k1 ) · · · a+ (km )a(p1 )
i=1 j=1
Consider the equations for a0r (k, −1, t) following from the Heisenberg
equations. It is easy to check that they have the form
∂a0r (k, −1, t)
i − (k)a0r (k, −1, t) = Fr , (11.24)
∂t
where Fr is an expression composed of a0s (k, 1, t) and a0s (k, −1, t)
with s < r. In order to prove the necessary relation, we should show
that a0r (k, −1, t)Φ0 = 0; we will perform the proof by induction in r.
4
The existence of the operator U satisfying these conditions follows from the
results of Section 6.1 (more precisely, from their analogs concerning the unitary
equivalence of the Fock representations of CCR, written in the form of a formal
expression in powers of g). Using the Heisenberg equations, we can show that
U a(k, , t) = a0 (k, , t)U . From the condition (b) in the definition of operator
realization, we obtain that U H = H 0 U, U P = P0 U .
March 27, 2020 16:3 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch11 page 291
Suppose that a0s (k, −1, t)Φ0 = 0 for s < r. By the absence of vacuum
polarization, every summand in the expression Fr will contain at
least one operator of the form as (k, −1, t) and therefore Fr Φ0 = 0.
Using equation (11.24) on the vector Φ, we can see that
∂ 0
i a (k, −1, t)Φ0 = (k)a0r (k, −1, t)Φ0 ,
∂t r
from which it follows that
Let
R us now calculate the expectation value of the energy in the state
f (k)a (k, −1, t)dkΦ0 :
0
Z Z
0 0 0 0 0
H f (k)a (k, −1, t)dkΦ , f (k)a (k, −1, t)dkΦ
Z Z
1 ∂ 0 0 0 0
= f (k) a (k, −1, t)dkΦ , f (k)a (k, −1, t)dkΦ
i ∂t
X1 Z
∂ 0
= 0
f (k)f (k ) a (k, −1, t)Φ , as0 (k , −1, t)Φ dkdk0
0 0 0 0
i ∂t s
s≥r
where
Z
1 ∂ 0
C2r = f (k)f (k0 ) a (k, 1, t)Φ , ar (k , −1, t)Φ dkdk0
0 0 0 0
i ∂t r
Z
=− (k)f (k)f (k0 ) exp(−i((k) − (k0 ))t)
Since
0
ar (k, −1, 0)Φ0 , a0r (k0 , −1, 0)Φ0 = σ(k)δ(k − k0 ),
If a0r (k, −1, t)Φ0 6≡ 0, then σ(k) 6≡ 0, and therefore, we can find a
function f (k) such that C2n < 0. This contradicts the non-negativity
assumption of the operator H 0 .
Let us now check the applicability conditions of axiomatic field
theory in each power of g. Let us first consider the spectrum of the
operators Ĥ, P̂ in the operator realization (H, Ĥ, P̂, ã(k, , t), Φ) of
a Hamiltonian of the form (11.23). For this Hamiltonian, the (bare)
Fock vacuum θ and the bare single-particle state a+ (k)θ satisfy the
conditions
Ĥθ = P̂θ = 0,
(r)
X
Ĥa+ (k)θ = (k) + g r ṽ11 (k) a+ (k)θ,
R
the form f (x)ã(x, , t)dx that we have considered belong to the
algebra A, consisting of operators of the form
X XZ
gr (r)
fm,n (k1 , . . . , km |p1 , . . . , pn )a+ (k1 )
m,n
First, let us note that the discussions imply that the scattering
matrix of the Hamiltonian H coincides with the scattering matrix of
the Hamiltonian H̃, obtained from H by means of Faddeev canonical
transformation.5 Indeed, these two scattering matrices are built on
the asymptotically commutative families B and B̃, which are both
contained in the same asymptotically commutative algebra A; by
the results of Chapter 10, the Møller and scattering matrices based
on these families are identical.
The statement we have just proved implies the following equiva-
lence theorem.
5
It is important to note that in this statement in place of the Faddeev
transformation, we can select any canonical transformation transforming the
Hamiltonian H into a Hamiltonian of the form (11.23), not necessarily the
concrete transformation constructed in Section 9.4.
March 27, 2020 16:3 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch11 page 295
P (r)
We will write A in the form A = Am,n , where Am,n = r Am,n g r ,
P
Z
Am,n = A(r)
(r)
m,n (k1 , . . . , km |p1 , . . . , pn )
(r )
where at least one of the operators Amii ni satisfies the condition
ni ≥ 2 (i.e. contains at least two annihilation operators), and the
functions f1 (k), . . . , fs (k) constitute a non-overlapping family of
smooth functions with compact support. The above completes the
proof of the theorem in the case when the operator D takes the
March 27, 2020 16:3 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch11 page 299
form N (exp A). The proof of the theorem in the general case can
be reduced to the cases already considered using the following two
lemmas.
Lemma 11.1. Any operator of the form exp(A1 + iA2 ), where
A1 , A2 ∈ M, can be written in the form N (exp(A01 + iA02 )), where
A01 , A02 ∈ M.
Lemma 11.2. Any operator of the form exp(A1 + iA2 ) × exp(A01 +
iA02 ), where A1 , A2 , A01 , A02 ∈ M, can be written in the form
N (exp(A001 + iA002 )), where A001 , A002 ∈ M.
We will not provide proofs of these lemmas.
Let us now consider the case of an arbitrary Hamiltonian H ∈ M.
The symbol HΩ , as usual, denotes the operator H with finite volume
cutoff (see Section 8.1); the operator HΩ acts on the Fock space
FΩ = F (L2 (Ω)). In perturbation theory, we can calculate the
operator exp(iHΩ t) (in other words, HΩ can be viewed as a self-
adjoint operator in the framework of perturbation theory).
The family of operators DΩ will be called a family of dressing
operators for the Hamiltonian H if the scattering matrix S of the
Hamiltonian H can be written in the form
S = lim lim exp(i(cΩ + HasΩ )t)(DΩ )−1
t→∞ Ω→∞
t0 →−∞
× D exp(−iHas t0 ),
which follows from the already-proven relations (11.25) and the
formula S = S+ ∗ S . Thus, the necessary statement is proven for
−
March 27, 2020 16:3 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch11 page 301
× D0Ω exp(−iHasΩ
0
t0 ), (11.34)
where
D0Ω = exp(iWΩ )DΩ
(to derive equation (11.34), one needs to use the coincidence of
the asymptotic Hamiltonians Has and Has 0 , corresponding to the
0
Hamiltonians H and H ). Applying the statement of the theorem
to a Hamiltonian H 0 not generating vacuum polarization, we see
that the operators D0Ω constitute a family of dressing operators
for the Hamiltonian H 0 . This implies that the right-hand side of
equation (11.34), in the limit as t → ∞, t0 → −∞, equals
the scattering matrix of the Hamiltonian H 0 . Since the scattering
matrices of the Hamiltonians H and H 0 coincide, equation (11.34)
proves the theorem in the general case.
6
This statement, just as the rest of the statements inPthis section, applies to
the more general class of Hamiltonians of the form H0 + r≥1 g r Wr (x)dx, and
R
· · · W (xr , τr ),
Z
V (τ ) = exp(iH0 τ )V exp(−iH0 τ ) = W (x, τ )dx,
R0 (A) = A,
Rr (x1 , t1 , . . . , xr , tr |A)
= [W (x1 , t1 ), Rr−1 (x2 , t2 , . . . , xr , tr |A)].
Ur (t)AΦ = Br (t)Φ,
where
Z t Z τ1 Z τr−1 Z
1
Br (t) = dτ1 dτ2 · · · dτr dr x exp(−iH0 t)
r! 0 0 0
× Rr (x1 , τ1 , . . . , xr , τr |A) exp(iH0 t).
7
In order to prove the existence of Møller matrices and scattering matrices
constructed with the operators H(g), P, and the algebra A, it is enough to assume
that the joint spectrum of the operators H0 , P satisfies conditions 4 and 5 from
Section 10.5.
March 27, 2020 16:3 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch11 page 307
X X
P g r Φr (k) = k g r Φr (k) , (11.42)
∂ X r X
r 0
g Φr (k), g Φr (k ) = 0. (11.43)
∂g
By the relations (11.41) and (11.42), we obtain that
Z
ωr (k)f (k)f 0 (k)dk
− Φ1 (ωr−1 f ), Φ0 (f 0 ) ,
(11.44)
where
R f (k) is a smooth function with compact support; Φr (f ) =
f (k)Φr (k); γ(ω, k) is a smooth function with compact support
equal to (ω − ω0 (k))−1 , if k ∈ supp f and (ω, k) belongs to the multi-
particle spectrum of the operators H, P and also if k = 0, ω = 0. It
should be equal to zero if ω = ω0 (k). Using relations (11.43)–(11.45),
we can find functions λr (k), ωr (k), and vectors Φr (k); simultaneously
by induction on r, we can prove that ωr (k) and λr (k) are smooth
functions and the vector Φr (f ) can be written in the form
(r)
Φr (f ) = Bf Φ, (11.46)
(r)
where Bf ∈ A. (The representation (11.46) follows from equa-
tion (11.45) with the help of the reasoning employed in the proof
of Lemma 10.17 in Section 10.6; the infinite differentiability of ωr (k)
and λr (k) can be shown by a slight modification of the proof of
Lemma 10.15 in Section 10.6.)
Let us now show that the Møller matrices S± (g), corresponding
to the operators H(g), P, and the algebra A, can be constructed at
least in the framework of perturbation theory. More precisely, we will
construct the Møller matrix as a formal series
(r)
X
S± (g) = g r S± ,
March 27, 2020 16:3 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch11 page 308
(r)
where S± are operators defined on the set D ⊂ Has , consisting of
linear combinations of vectors of the form a+ (f1 ) · · · a+ (fn )θ (here,
f1 (k), . . . , fn (k) is a non-overlapping family of smooth functions with
compact support. (Recall that this condition on a family of functions
means that for any k ∈ supp fi , k0 ∈ supp fj , i 6= j, the gradients of
the functions ω0 (k) at the points k and k0 do not coincide.)
To construct Møller matrices, we repeat the construction in
Section 10.1, considering all objects, entering into the formula (10.1),
as formal series in powers of g. The proof of the correctness of
this definition is nearly identical to the proof in Chapter 10. It is
important to note that the estimate (10.7) cannot be proven in every
order in g, in other words, if f (k) is a smooth function with compact
support with
Z X
f˜(x|t) = (2π)−3 exp −it g r ωr (k) + ikx f (k)dk
r≥0
X
= g r f˜r (x|t),
r≥0
× δ(k1 + · · · + km − p1 − · · · − pn )
× a+ (k1 ) · · · a+ (km )a(p1 ) · · · a(pn )dm kdn p,
March 27, 2020 16:3 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch11 page 311
Chapter 12
Axiomatic Lorentz-Invariant
Quantum Field Theory
the definitions of the space Has differ. However, in light of the fact
that the single-particle space B defined in this section is contained in
the single-particle space in the sense of Section 10.1, the discussions
in Chapter 10 do not need to be modified in this case.) The Lorentz
invariance of the Møller matrices and the scattering matrix can be
proved by the discussions in Section 6, Chapter 6 in Jost (1965); we
will sketch another proof in Chapter 13.
Let us note that the results of this section hold if the asymp-
totically commutative family of operators is replaced by an asymp-
totically commutative family of operator generalized functions (in
the sense of Section 10.5). The domain of definition of these
operator generalized functions should be invariant with respect to
the operators U (g), where g ∈ P. Conditions A2 and A3 require
corresponding modifications (we ought to require that the vector Φ
is cyclic vector for this family and that for every generalized operator
function A(x, t), the corresponding generalized operator functions
Ag (x, t) = Ug A(x, t)Ug−1 , where g is an arbitrary element of the
group P, should also belong to the family).
Let us assume that the algebra R(O) has the following properties:
R1. If O1 ⊂ O2 , then R(O1 ) ⊂ R(O2 ).
R2. Covariance with respect to the Poincaré group:
Ug R(O)Ug−1 = R(gO).
where x = (x, t). Since the sets O1 and O2 are bounded, there exists
an a such that for |x|2 − |t|2 ≥ a2 , any point of the set O1 − x is
separated from any point of the set O2 by a space-like interval. Using
condition R3, we obtain that for |x|2 − |t|2 ≥ a2 ,
[A(x, t), B] = 0.
2
For more details, see Streater and Wightman (2016), Jost (1965), and
Bogolyubov et al. (1969).
March 27, 2020 16:33 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch12 page 323
Thus, we can see that all the statements which can be obtained
from the axioms in Section 12.1 can be derived from Wightman’s
axioms. However, there are theorems which can be proved in
Wightman’s axiomatics that have not been proven (and, apparently,
cannot be proven) in the axiomatic formulation in Section 12.1.
(This is not surprising; Wightman axiomatized local quantum field
theory; the axioms of Section 12.1 are based on weaker assumption
of asymptotic commutativity).
The most important statement of this type are dispersion rela-
tions obtained first in Bogolyubov et al. (1956) and derived from
Wightman axioms in Hepp and Epstein (1971).
Let us also mention the CPT theorem, stating that the Wightman
axioms imply CPT invariance (i.e. the existence of additional
symmetries).
Let us formulate Borchers’ theorem:
dk
+ a(k) exp(−iω(k)t + ikx)) p . (12.12)
2ω(k)
The ground state Φ coincides with the Fock vacuum θ.
March 27, 2020 16:33 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch12 page 326
3
We should note that in the one-dimensional case (i.e. in the case when in the
Hamiltonian (12.14) integration is performed over a single variable) calculations of
Green functions of the Hamiltonian (12.14) do not give rise to diverging integrals.
March 27, 2020 16:33 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch12 page 328
Let us consider for the sake of concreteness the case when ν(φ) =
bφ4 . We consider the family of Hamiltonians of the form
H = H0 (w) + bVΛ
Z Z
1 2 1
= π̂ (x)dx + w(x − y)φ̂(x)φ̂(y)dxdy
2 2
Z
+ b fΛ (x1 − x4 , x2 − x4 , x3 − x4 )
Chapter 13
331
March 30, 2020 11:40 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch13 page 332
− log K − 1 − β Ĥ − ζ = 0.
S = βE + log Z.
F = −T log Z.
Here, · · · stands for higher order terms with respect to λ. Note that
Tr Ae−β Ĥ
Ā =
Z
is the mean value (expectation value) of the observable A in the
equilibrium state. We will also use the notation hAi for the expecta-
tion value. The expectation values hA1 · · · An i or, more generally,
hA1 (t1 ) · · · An (tn )i are called correlation functions (here, Ai are
observables and Ai (t) are the corresponding Heisenberg operators).
If we want to emphasize that the correlation functions are calculated
for the equilibrium state with the temperature T = β1 , we will use
the notation hA1 (t1 ) · · · An (tn )iβ .
If A and B are two observables, then
hA(t)Biβ = hBA(t + iβ)iβ . (13.3)
This equation, called the Kubo–Martin–Schwinger (KMS) condition,
can be proved by simple formal manipulations if the Hilbert space
H is finite-dimensional. In the infinite-dimensional case, one should
check that the correlation function hBA(t)iβ can be extended
analytically to the strip 0 ≤ =t ≤ β (here, = stands for imaginary
part).
It is easy to see that
∂ log Z ∂F
Ā = −T = (13.4)
∂λ ∂λ
(the derivatives are calculated at the point λ = 0).
If the Hamiltonian depends linearly on a set of parameters
λ1 , . . . , λk , we can calculate the correlation functions by differenti-
ating the free energy F . For example, if Ĥ = Ĥ0 + λ1 A1 + · · · + λk Ak ,
we have
∂2F
= hAi Aj i − hAi ihAj i. (13.5)
∂λi ∂λj
(The derivatives are calculated at the point λi = 0.) The RHS
of (13.5) is called a truncated correlation function. Higher truncated
correlation functions can be defined as higher derivatives of F. (See
Section 10.1 for the definition of truncated correlation functions in
terms of correlation functions.)
March 30, 2020 11:40 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch13 page 334
13.1.1 Examples
Let us consider the multidimensional harmonic oscillator as an
example. Introducing the creation and annihilation operators
obeying CCR, we can represent the Hamiltonian in the form
Ĥ = i ωi â+
P P
i âi + C, where C = ωi /2 is the energy of the ground
state. We shift the energy scale assuming that the ground state has
P
zero energy: C = 0. Then the energy levels are ni ωi where ni is a
non-negative integer. It is easy to calculate the partition function
1
Z=Π ,
1 − e−βωi
the free energy
1X
F = log(1 − e−βωi ),
β
and the mean value of energy
X
H̄ = ωi n̄i ,
Z = Π(1 + e−βωi ),
1X
F =− log(1 + e−βωi ),
β
X
H̄ = ωi n̄i ,
1
where n̄i = eβωi +1
.
formula for the equilibrium state makes no sense (the integral for
the partition function diverges). Moreover, as we have seen, the
formula (13.7) in general does not define an operator on Fock
space. However, the Hamiltonian with volume cutoff ĤΩ specifies an
operator in Fock space under some mild conditions and we can define
an equilibrium state and correlation functions. In particular, we
can consider the correlation functions β wnΩ (k1 , 1 , t1 , . . . , kn , n , tn ) =
hak11 (t1 ) · · · aknn (tn )iβ generalizing Wightman functions. (Wightman
functions are correlation functions for T = 0.)
We can define correlation functions in infinite volume β wn =
limβ wnΩ by taking the limit Ω → ∞ (removing the volume cutoff).
As in Section 8.2, the limit is understood in the sense of generalized
functions. Under some mild conditions, one can prove that the
correlation functions in infinite volume obey KMS condition (13.3)
March 30, 2020 11:40 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch13 page 336
function of x and t.
We assume that Φ(f ) is normalized (i.e. hΦ(f ), Φ(f 0 )i = hf, f 0 i).
Note that it is possible that there exist several types of (quasi-)
particles Φr (f ) with different dispersion laws. We say that the space
spanned by Φr (f ) is one-particle space and denote it H1 . If the theory
is invariant with respect to spatial rotations, then the infinitesimal
rotations play the role of components of angular momentum. If d = 3,
a particle with spin s can be described as a collection of 2s + 1
functions Φr (p) obeying PΦr (p) = pΦr (p), HΦr (p) = ε(p)Φr (p).
The group of spatial rotations acts in the space spanned by these
functions; this action is a tensor product of the standard action
of rotations of the argument and irreducible (2s + 1)-dimensional
representation. (Here, s is half-integer, the representation is two-
valued if s is not an integer.) In relativistic theory, an irreducible
subrepresentation of the representation of the Poincaré group in H
specifies a particle with spin.
13.3.3 Scattering
Let us assume that we have several types of (quasi-)particles defined
as generalized functions Φk (p) obeying PΦk = pΦk (p), HΦk (p) =
εk (p)Φk (p), where the functions εk (p) are smooth and strictly
convex. Take some good operators Bk ∈ A, Robeying B̂k Φ = Φ(φk ).
Define B̂k (f, t), where f is a function of p as f˜(x, t)B̂k (x, t)dx and,
March 30, 2020 11:40 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch13 page 342
(13.13)
This formula specifies S± on a dense subset of the Hilbert space Has .
These operators are isometric, hence they can be extended to Has
by continuity.
One can say that the vector
e−iHt Ψ(k1 , f1 , . . . , kn , fn | ± ∞)
= Ψ(k1 , f1 e−iεk1 t , . . . , kn , fn e−iεkn t | ± ∞) (13.14)
describes the evolution of a state corresponding to a collection
of n particles with wave functions f1 φ1 e−iεk1 t , . . . , fn φn e−iεkn t as
t → ±∞.
March 30, 2020 11:40 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch13 page 343
a+ ¯ a+ ¯
in (f φ̄) = lim B̂(f, t),
t→−∞ out (f φ̄) = lim B̂(f, t).
t→∞
(13.15)
in the vectors (13.12), we can first take the limit for i > 1 and then
the limit t1 → ∞.
The formula (13.15) can be written in the following way:
a+ ¯
in (f φ̄)Ψ(f1 , . . . , fn | − ∞) = Ψ(f, f1 , . . . , fn | − ∞),
a+ ¯
out (f φ̄)Ψ(f1 , . . . , fn | ∞) = Ψ(f, f1 , . . . , fn | ∞).
Similarly,
If the operators S+ and S− are unitary, we say that the theory has
a particle interpretation. In this case (and also in the more general
case when the image of S− coincides with the image of S+ ), we can
define the scattering matrix
∗
S = S+ S− .
Smn (p1 , . . . , pm | q1 , . . . , qn )
= ha+ + + +
in (q1 ) · · · ain (qn )Φ, aout (p1 ) · · · aout (pm )Φi. (13.18)
Smn (f1 , . . . , fm | g1 , . . . , gn )
Z Y Y
= dm pdn q fi (pi ) gj (qj )Smn (p1 , . . . , pm | q1 , . . . , qn )
= ha+ + + ¯ + ¯
in (ḡ1 ) · · · ain (ḡn )Φ, aout (f1 ) · · · aout (fm )Φi.
Smn (f1 , . . . , fm | g1 , . . . , gn )
Note that in the same way, we can obtain a more general formula
Smn (f1 , . . . , fm | g1 , . . . , gn )
where R is negligible,
First of all, we consider the integral (13.23) over the domain Γ(t)
defined in the following way. Take non-overlapping compact sets G ⊂
Rd , G0 ⊂ Rd . We say that the point (x, x0 ) belongs to Γ(t) if there
0
exist points v ∈ G, v0 ∈ G0 such that |v − xt | < C|t|−ρ , |v0 − xt | <
C|t|−ρ where ρ = 12 − . (We will apply our estimate in the case when
G is the set of points of the form ∇εki (p) where p belongs to the
support of fki and G0 is defined in similar way using the function
fkj .) For large t, the volume of this domain is less than t2d(1−ρ) and
the norm of the integrand is less than t−d × t−a . (We have used the
fact that the distance between x and x0 grows linearly with t in the
integration domain, hence the norm of the commutator in (13.23) is
less than t−a .) This allows us to say that the norm of the integral
does not exceed t−a+ .
To estimate the integral (13.21) over the complement to Γ(t), we
consider the integral
Z x
exp(−iε(p)t + ipx)µ tρ v(p) − f (p)dp, (13.24)
t
where µ(x) is a smooth function equal to zero for |x| ≤ ν1 and equal
to one for |x| ≥ ν2 . One can prove that for every r, the absolute value
of this integral does not exceed C|t|−(1−2ρ)r and for |x| > bt, it does
not exceed C|x|−r |t|−(1−2ρ)r .
The integral (10.24) is equal to f˜(x, t) if tρ (v(p)− xt ) ≥ ν2 for p ∈
supp f. This means that the estimate of this integral can be applied
to at least one of the factors f˜ki (x, t), f˜kj (x0 , t) in the integrand of
the integral (13.23) on the complement to the domain Γ(t). Using
this fact and the inequality
Z
d
dx|f˜k (x, t)| ≤ C|t| 2 ,
we obtain that the norm of the integral (13.23) over the complement
to Γ(t) tends to zero faster than any power of |t| as t → ±∞.
Combining this estimate with the estimate for the integral over Γ(t)
March 30, 2020 11:40 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch13 page 350
and taking < a − 1, we obtain the estimate for the norm of (13.23):
for large |t|, it is less than |t|−b where b > 1. The same arguments
work if the dot (time derivative) in (13.23) is removed.
˙
To prove (13.21), we note that calculating B̂(f, t) we should take
into account that f depends on t. It is easy to check that
Z Z
˙ ˙
B̂(f, t) = dxg̃(x, t)B̂(x, t) + dxf˜(x, t)B̂(x, t),
where g(p) = −iε(p)f (p). Using this expression, we obtain that the
commutator in (13.21) can be written as a sum of (13.23) and similar
expression with removed dots. Both summands do not exceed |t|−b
with b > 1 for large |t|. This implies (13.21).
Note that the proof based on asymptotic commutativity also
allows us to estimate the speed of convergence to the limit. In the
case of strong asymptotic commutativity, the difference between Ψ(t)
and Ψ(±∞) tends to zero faster than any power of t.
and taking the limit i → −εi (pi ) for 1 ≤ i ≤ m and the limit j →
εj (pj ) for m < j ≤ m+n, we obtain on-shell Green function denoted
by σmn . We prove that it coincides with scattering amplitudes:
σm,n (p1 , . . . , pm+n ) = Smn (p1 , . . . , pm | pm+1 , . . . , pm+n ). (13.25)
First of all, we note that the on-shell Green function can be expressed
in terms of the asymptotic behavior of the Green function in (p, t)-
representation: if for t → ±∞ and fixed p, this behavior is described
March 30, 2020 11:40 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch13 page 352
1
The physical meaning of this condition: the energy conservation law forbids the
decay of a particle. This condition is not always satisfied, however, stability of a
particle is always guaranteed by some conservation laws. Our considerations can
be applied in this more general situation.
March 30, 2020 11:40 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch13 page 353
Gn = ω(M N ),
March 30, 2020 11:40 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch13 page 354
where
N = T (B1 (x1 , t1 ) · · · Bn (xn , tn ))
stands for chronological product (times decreasing) and
M = T opp (B1∗ (x01 , t01 ) · · · Bn∗ (x0n , t0n ))
stands for antichronological product (times increasing).
One can give another definition of GGreen functions introducing
the operator
Q = T (B1 (x1 , t1 ) · · · Bn (xn , tn )B̃1 (x01 , t01 ) · · · B̃n (x0n , t0n )),
where the operators Bi , B̃i act on the space of linear functionals on
A. (Recall (13.8) and (13.9) that operators B and B̃ act on linear
functionals defined on A; they transform ω(A) into ω(BA) and in
ω(AB ∗ ) correspondingly.) It is easy to check that
Gn = (Qω)(1).
Let us define inclusive S-matrix as on-shell GGreen function. We
will show that in the case when the theory has particle interpretation,
inclusive cross-section can be expressed in terms of inclusive S-
matrix.
Recall that the inclusive cross-section of the process (M, N ) →
(Q1 , . . . , Qm ) is defined as a sum (more precisely, a sum of integrals)
of effective cross-sections of the processes (M, N ) → (Q1 , . . . , Qm ,
R1 , . . . , Rn ) over all possible R1 , . . . , Rn . If the theory does not have
particle interpretation, this formal definition of inclusive cross-section
does not work, but still the inclusive cross-section can be defined in
terms of probability of the process (M, N → (Q1 , . . . , Qn + something
else)) and expressed in terms of inclusive S-matrix.
Let us consider the expectation value
ν(a+ +
out,k1 (p1 )aout,k1 (p1 ) · · · aout,km (pm )aout,km (pm )), (13.27)
where ν is an arbitrary state. This quantity is the probability density
in momentum space for finding m outgoing particles of the types
k1 , . . . , kn with momenta p1 , . . . , pm plus other unspecified outgoing
particles. It gives inclusive cross-section if ν = ν(t) describes the
evolution of a state represented as a collection of incoming particles.
March 30, 2020 11:40 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch13 page 355
ν(a+ +
out,k1 (f1 )aout,k1 (g1 ) · · · aout,km (fm )aout,km (gm )), (13.28)
ν(a+ +
out,k1 (p1 )aout,k1 (q1 ) · · · aout,km (pm )aout,km (qm ));
13.4 L-functionals
In the approach of Section 13.2, for every equilibrium state, we should
construct a Hilbert space depending on the temperature, CCR or
CAR are represented in this space, the equilibrium state is described
by a vector in the space. In this section, we will describe a formalism
of L-functionals (positive functionals on Weyl algebra) that can be
used to describe the states corresponding to vectors and density
March 30, 2020 11:40 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch13 page 356
prompted by relations
ÃALK = LAKA+ .
(13.33)
(13.34)
March 30, 2020 11:40 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch13 page 359
(13.35)
The equation of motion for the L-functional L(α∗ , α) has the form
dL
i~ = ĤL = HL − H̃L. (13.36)
dt
(We introduced the notation Ĥ = H − H̃.) It corresponds to the
equation of motion for density matrices; this follows from the formula
ĤLK = LHK−KH + .
Note that often the equations of motion for L-functionals make
sense even in the situation when the equations of motion in the
Fock space are ill defined. This is related to the fact that vectors
and density matrices from all representations of CCR are described
by L-functionals. This means that by applying the formalism of
L-functionals, we can avoid the problems related to the existence of
inequivalent representations of CCR. It is well known, in particular,
that these problems arise for translation-invariant Hamiltonians; in
perturbation theory, these problems appear as divergences related to
infinite volume. Therefore, in the standard formalism, it is necessary
to consider at first a Hamiltonian in finite volume Ω (to make volume
cutoff or, in another terminology, infrared cutoff) and to take the
limit Ω → ∞ in physical quantities (see Sections 8.2 and 13.2 for
more details). In the formalism of L-functionals, we can work directly
in infinite volume.
In general, the ground state of formal Hamiltonian (13.7) does
not belong to Fock space, but the corresponding L-functional is well
defined. (Note that we always assume that ultraviolet divergences are
absent.) Similarly, the equilibrium states for different temperatures
belong to different Hilbert spaces, but all of them are represented by
well-defined L-functionals.
In what follows, we consider (13.33) as a formal expression;
we assume that it is formally Hermitian. There is no necessity to
March 30, 2020 11:40 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch13 page 360
Ĥ(g)L(g) = 0.
~ω(k)
c1 (k)LT = e− T (−~c+
2 (k) + c1 (k))LT , (13.39)
hence
c1 (k)LT = −n(k)c+ +
2 (k)LT , c2 (k)LT = −n(k)c1 (k)LT , (13.40)
where
~
n(k) = ~ω(k)
. (13.41)
e T −1
We obtain
α∗ (k)n(k)α(k)dk
R
LT = e− . (13.42)
If we are interested in equilibrium state for given density, we should
replace ω(k) with ω(k) − µ in (13.41) (here, µ stands for chemical
potential).
The Hamiltonian Ĥ governing the evolution of L-functionals can
be written in the form
Z Z
Ĥ = ω(k)b+ (k)b(k)dk − ω(k)b̃+ (k)b̃(k)dk
Z
= ~(ω(k)c+ + +
1 (k)c1 (k)dk − ω(k)c2 (k)c2 (k))dk.
March 30, 2020 11:40 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch13 page 362
G2 (k1 , t1 , σ1 , k2 , t2 , σ2 )Λ
= G2 (k1 , t1 , σ1 , k2 , t2 , σ2 )Λ
Z
+ dk20 dt02 dσ20 dk200 dt002 dσ200 G2 (k1 , t1 , σ1 , k20 , t02 , σ20 )Λ
GΛ Λ Λ Λ Λ
2 = G2 + G 2 M G2
or
−1 −1
(GΛ
2) = (GΛ
2) + M Λ. (13.50)
where Ŝa = Ŝa (∞, −∞) stands for the adiabatic S-matrix,
Z
Va = exp i ra (k)(c+ +
1 (k)c1 (k) − c2 (k)c2 (k))dk
we obtain
σ̂m,m0 (k1 , . . . , km |k10 , . . . , km
0
0 |Ψ)
p p
(m + n)! (m0 + n)!
X Z
= dp1 · · · dpn
n
n!
× hSΨ|k1 , . . . , km , p1 , . . . , pn ihSΨ|k10 , . . . , km
0 , p , . . . , p i.
0 1 n
× δ(k10 + · · · + km
0
− (k1 + · · · + km )),
where
σ̃m,m0 (k1 , . . . , km |k10 , . . . , km
0
0 |Ψ)
p p
(m + n)! (m0 + n)!
X Z
= dp1 · · · dpn
n
n!
Let us now consider the case when Φ is the ground state. Then one
can obtain an expression of Ŝ in terms of GGreen functions that is
analogous to the LSZ formula. It will be derived from some identities
that were used in the proof of LSZ, namely, we can use the identity
[· · · [S, ain (k1 , σ1 )] · · · ain (kn , σn )]
Z
n ∗
= (−1) S− S+ dt1 · · · dtn Ln · · · L1 T (a(k1 , t1 , σ1 ) · · · a(kn , tn , σn )),
(13.53)
where S− , S+ are Møller matrices, the scattering matrix S is
represented in terms of in-operators ain (k, 1) = a+ in (k), ain (k, −1) =
ain (k), operators +
R a(k, t, σ) are Heisenberg operators a (k, t), a(k, t),
the operators dti Li in (k, )-representation canR be interpreted
as “on-shell operators.” (To apply the operator dti Li in (k, )-
representation, we should multiply by iΛ(ki , σi )(i − ω(ki )) and take
the limit i → ω(ki ). Here, ω(k) stands for the location of the pole
of the two-point Green function and Λ can be expressed in terms
of the residue in this pole. Note that in the transition to (k, )-
representation, we are using direct Fourier transform for σ = −1
and inverse Fourier transform for σ = 1.)
The identity (13.53) can be obtained, for example, from (32.17)
of Section 9.2 (by means of conjugation with S− ).
It follows from (13.53) that Green function defined as vac-
uum expectation value of chronological product T (a(k1 , t1 , σ1 ) · · ·
a(kn , tn , σn )) is related to scattering amplitude: one should take the
Fourier transform with respect to time variables and “go on-shell”
in the sense explained above. This gives the LSZ formula (see the
preceding section for more general approach). We remark that these
considerations also go through in the case when instead of vacuum
expectation value h0|A|0i where |0i obeys ain (k)|0i = 0 (represents
physical vacuum), we can take matrix elements h0|A|p1 , . . . , pn i
where |p1 , . . . , pn i = √1n! a+ +
in (p1 ) · · · ain (pn )|0i. This remark allows us
to express “on-shell” GGreen functions as sesquilinear combinations
of scattering amplitudes; comparing this expression with the formula
for ŜLK , we obtain the expression of Ŝ in terms of GGreen functions
“on-shell” (an analog of LSZ formula). Indeed, we can consider
March 30, 2020 11:40 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch13 page 372
· · · a+
in (pr )ain (q1 ) · · · ain (qs ).
We assume that all pi ’s are distinct and all qj ’s are distinct, then the
coefficient functions in normal form coincide with scattering ampli-
tudes hSa+ + + +
in (q1 ) · · · ain (qs )θ | ain (p1 ) · · · ain (pr )θi. Let us take σi = 1
for i ≤ m, σi = −1 for i > m in (13.53). Introducing the notation
qi = ki−m , we obtain that
1
hp1 , . . . , pn |LHS|0i = √ σm+n,l (p1 , . . . , pn , k1 , . . . , km | q1 , . . . , ql ).
n!
Here, LHS stands for the LHS of (13.53). Now, we can apply (13.53)
and (13.54) to identify on-shell GGreen functions with matrix entries
of the inclusive scattering matrix Ŝ.
Our considerations used LSZ relations that are based on the
conjecture that the theory has particle interpretation. Moreover, the
very definition of Ŝ that we have applied requires the existence of
conventional S-matrix. However, using (13.51) as the definition of
the inclusive scattering matrix Ŝ, one can prove the relation between
Ŝ and on-shell GGreen functions analyzing diagram techniques for
these objects. This proof can also be applied in the case when the
theory does not have particle interpretation. Moreover, the same
March 30, 2020 11:40 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch13 page 373
Appendix
Appendix 377
(1) A = ∪∞ i=1 Ai ,
(2) the sets A1 , . . . , An , . . . are pairwise disjoint. (In other words, we
assume that the measure is countably additive.)
The set M , equipped with the family B and the countably additive
measure µ, specifies a measure space.
A set R ⊂ M (not necessarily belonging to B) is called a set of
measure zero if for every > 0, there exists a set A ∈ B, such that
R ⊂ A and µ(A) < . One says that a sequence of functions fn (x)
converges to a function f (x) almost everywhere on a set M if the
set of points x ∈ M , such that the sequence fn (x) does not converge
to f (x), is a set of measure zero. In general, if a relation is satisfied
everywhere except a measure zero set, then we say that it is satisfied
almost everywhere.
March 30, 2020 10:29 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-x99a-app page 379
Appendix 379
One can prove that this limit always exists and does not depend on
the choice of the sequence fn (x).
If the function f is unbounded and non-negative, then one can
define its Lebesgue integral by the relation
Z Z
f (x)dµ = lim fn (x)dµ,
A n→∞ A
If the
R function f (x) is Rcomplex, then the Lebesgue integral is defined
by A Re(f (x))dµ + i A Im(f (x))dµ, where Re(f (x)), Im(f (x)) are
real and imaginary parts of f (x).
We note some important properties of the Lebesgue integral:
(Fubini’s theorem).
Appendix 381
1
Two equivalent measurable functions specify the same element of the space of
L2 (M ).
2
R
Integrating over the Rwhole measure space M , we use the notation φ(x)dµ
instead of the notation M φ(x)dµ.
March 30, 2020 10:29 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-x99a-app page 382
hi , h0i .
=
i
Appendix 383
Appendix 385
Appendix 387
Appendix 389
2. Let us suppose that for every smooth finite function χ(ω) with
support on (a, b) we have
Z
χ̃(t) exp(iHt)dt = 0,
where
Z
χ̃(t) = exp(−iωt)χ(ω)dω.
Then the spectrum of the operator H lies outside the interval (a, b).
Two operators A and B, defined on the whole space H, are called
commuting if AB = BA. The same definition can be used in the
case where operators A and B are defined on the same set D and
transform it into itself. However, for self-adjoint operators, one should
use another definition of commuting operators. One says that two
self-adjoint operators A and B in the Hilbert space H commute if
φ(A)ψ(B) = ψ(B)φ(A) for all bounded functions φ(λ), ψ(λ) (if this is
the case, we say that AB = BA). Note that the operators φ1 (A) and
φ2 (A) commute (here, φ1 (λ) and φ2 (λ) are arbitrary real functions
and A is a self-adjoint operator).
One can prove the following theorem.
For any family of self-adjoint commuting operators A1 , . . . , An , . . .
in a Hilbert space H, one can find a measure space M and an isomor-
phism α between the space H and the space L2 (M ) transforming each
of these operators into the operator of multiplication by a function
(i.e. αAi α−1 = âi (x)).
As in the case of one operator, this theorem allows us to
define functions of families of operators. Namely, if A1 , . . . , An are
March 30, 2020 10:29 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-x99a-app page 390
Appendix 391
(a) if for a dense set of vectors x the limit slim An x exists and the
sequence kAn k is bounded, then the sequence of operators An is
strongly convergent;
(b) if the limit lim hAn x, yi exists for all x from a dense subset X
and for all y from a dense subset Y and the sequence kAn k is
bounded, then the sequence An is weakly convergent;
(c) if An → A, Bn → B and the sequence kAn k is bounded, then
An Bn → AB (the analog of this statement for weak convergence
does not hold); if An → A and the sequence kA−1 n k is bounded,
then A−1n → A −1 ;
Appendix 393
3
R
Let us emphasize that while f (x)φ(x)dx is understood as another way of
writing the number f (φ) and is defined if φ is a test
R function and f is a generalized
function, in this section we use the notation f (x)φ(x)dµ for the Lebesgue
integral of the function f (x)φ(x). Clearly, the Lebesgue integral is defined when
r
the function f (x)φ(x) is summable. ForR integrals in the Euclidean space E in
both cases, we use the same notation f (x)φ(x)dx.
March 30, 2020 10:29 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-x99a-app page 395
Appendix 395
Appendix 397
4
The integral in this formula can be understood in any sense; it is important
only that it is linear with respect to φ.
March 30, 2020 10:29 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-x99a-app page 398
Appendix 399
Let us fix the space of test functions R and a dense linear subspace D
in the Hilbert space H. We will consider only the generalized operator
functions that transform D into itself (i.e. in other words, we assume
March 30, 2020 10:29 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-x99a-app page 400
C(x, y) = A(x)B(y).
Appendix 401
p (xp )
. . . A+
p p
1 (x1 )A1 (y1 ) . . . Ap (yp )a, a d xd y,
(A.6)
where
ρm,n (x1 , . . . , xp |y1 , . . . , yp ) = ψ m (x1 , . . . , xp )ψm (y1 , . . . , yp )
+ψ n (x1 , . . . , xp )ψn (y1 , . . . , yp )
−ψ n (x1 , . . . , xp )ψ( y1 , . . . , yp )
where
Z
0 0
C (ψ ) = ψ 0 (x1 , . . . , xq )A01 (x1 ) . . . A0q (xq )dq x,
Z
0 0
C(ψ)C (ψ ) = ψ(x1 , . . . , xp )ψ 0 (x01 , . . . , x0q )A1 (x1 )
Appendix 403
(i)
is convergent. Then ηn → 0. Here, ai ∈ D, φn ∈ S, ψn ∈ S, ψn → ψ
(i)
in the topology of S, Ak ∈ RD ,
Z
(i) (i) (i) (i) (i) p(i)
Cn (φn ) = φn (y1 , . . . , yp(i) )A1 (y1 ) . . . Ap(i) (yp(i) )d y.
i.e. Af (φ) = f (aφ). The last equation must be satisfied in the cases
when it makes sense (i.e. when φ ∈ R and aφ ∈ R).
The generalized vector function φa = (2π)−r/2 exp(iax), defined
in (A.6), turns out to be a generalized eigenfunction for the operators
1 ∂ 1 ∂
i ∂x1 , . . . , i ∂xr (and more broadly for differential operators with
constant coefficients):
1 ∂
φa = ak φa .
i ∂xk
Note the following important statement.
For any commuting self-adjoint operators A1 , . . . , An , acting on
the Hilbert space H, we can find a generalized basis of functions that
are eigenfunctions for each of the operators (more precisely, we can
find a generalized basis f that is δ-normalized (hf (x), f (y)i = δ(x, y))
and satisfies the equation Ai f (x) = ai (x)f (x) for every i = 1, . . . , n).
The proof of this statement is easily obtained if we recall that
there exists an isomorphism α of the space H and the space
L2 (M ), by which the operators Ai are transformed into operators of
multiplication by the function ai (x). Indeed, the generalized vector
function f can be represented by the formula
Z
f (φ) = f (x)φ(x)dx = α−1 φ,
where φ ∈ L2 (M ).
Te = 1, Tgh = Tg Th . (A.7)
Appendix 405
T(1,a) = exp[−i(a0 H − a1 P1 − a2 P2 − a3 P3 )]
= exp[−i(a0 H − aP)],
Appendix 407
on L2 (E 3 ):
(V (Λ, a)fˇ) = U (Λ, a)fˇ.
Let us describe (up to equivalence) all the irreducible representa-
tions of the group P in which the energy operator H is non-negative
and not equal to zero.
For every non-negative number m, there exists an infinite series
of unitary irreducible representations, where the joint spectrum of
the operators H, P in these representations is the set Um . The
representation of this series will be denoted by the symbols (m, σ),
where σ is a non-negative integer for positive m and any integer for
m = 0.
A representation of the type (m, σ) can be realized via unitary
operators on the space L2 (E 3 × S), where S is a set consisting of
2σ + 1 elements for positive m and of one element for m = 0. Here,
the energy and momentum operators on L2 (E 3 × S) are given by the
formulas
p
Hf (p, s) = p2 + m2 f (p, s), (A.10)
Pf (p, s) = pf (p, s). (A.11)
For example, the representation described above is a representation
of type (m, 1).
Let us note that in addition to the single-valued unitary rep-
resentations of P described here, there exist two-valued unitary
representations. Irreducible two-valued representations, where the
energy operator is non-negative and is not identically zero, can be
described in the same way as single-valued (with the difference that
σ must be a half-integer).
b2530 International Strategic Relations and China’s National Security: World at the Crossroads
Bibliography
409
March 30, 2020 10:34 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-x99b-bib page 410
Bibliography 411
Index
adiabatic, viii, xxxi, 43, 45, 52, 101, generalized function, xxii, xxxviii, 30,
150, 166, 170, 253, 254, 309, 393–400
362–365, 367, 368 generalized Green function, xix, 353,
angular momentum, xxxiv, 9–12, 341 363
asymptotically commutative, 194, GNS (Gelfand–Naimark–Segal)
233, 242, 292–294, 313 construction, xxv, xxvi, 337, 341,
357, 360, 362
CAR, xxxvi, xxxviii, xxxix, 30, 75, Green function, xxvi, xxvii, 111–114,
76, 79, 82–86, 88, 93, 97, 98, 116, 118–122, 128, 161, 327,
109–113, 120, 123, 133, 156, 163, 350–353, 366, 370
179, 334, 355, 356
CCR, xxviii, xxix, xxxiv, xxxvi, Heisenberg equation, xxxiii, 6, 7, 18,
xxxviii, xxxix, 30, 69, 75–77, 87, 84, 121, 138, 139, 143, 278–280,
90, 93, 95, 97, 98, 109–113, 120, 285, 290, 305
126, 127, 133, 141, 156, 163, 179,
181, 193, 277, 281–285, 290, 327 inclusive, xix, xlii, 345, 353, 368–370,
correlation function, xxvi, xxvii, xl, 372, 373
333, 335, 336, 343
cross section, xix, xlii, 58–61, 162, KMS (Kubo–Martin–Schwinger)
236–238, 345, 354, 368–370 condition, xli, 333, 335, 336, 340
413
April 1, 2020 14:21 Mathematical Foundations of Quantum Field Theory 9in x 6in 11222-x99c-index page 414
momentum, xxi, xxxiv, xli, xliii, xlvi, Schrödinger equation, 1, 2, 7, 46, 63,
3, 4, 9–11, 14–17, 27, 29, 34–36, 280
58–60, 62–64, 123, 124, 126, 138, soliton, xliii–xlv, 188, 189
142, 172, 184, 233, 275, 301, 335, spin, 11–16, 23, 57, 123, 316, 326, 341
407 stationary, xxvi, xxix–xxxi, 6, 7, 15,
Møller matrix, 73, 149, 196, 203, 240, 16, 18–22, 27, 39, 43, 44, 46, 63,
249, 253, 274, 279, 307–309, 311 119, 168, 178, 209, 306, 332, 360,
373
operator realization, viii, xxxix, symmetry breaking, 186, 187
126–129, 138, 139, 143–146, 153,
154, 166, 184–188, 275, 277–279, translation-invariant Hamiltonian,
286–290, 292, 293, 325, 336 viii, xxxix, 64, 65, 68, 123–126, 128,
oscillator, xxxv, 17, 18, 20, 21, 77, 147, 275, 311,
334 truncated, xl, 197–199, 333
partition function, 331, 332, 334, 335 Wightman function, xxvii, xxxix, xl,
perturbation theory, viii, xvii, xviii, 107–113, 322
xxxvii, xxxix, 39, 44, 45, 65, 86,
114, 129, 131, 170, 277, 301, 303,
305–309, 311, 327, 328, 338, 359