Weak interactions and modern particle theory Howard Georgi pdf download
Weak interactions and modern particle theory Howard Georgi pdf download
https://ebookgate.com/product/weak-interactions-and-modern-
particle-theory-howard-georgi/
https://ebookgate.com/product/particle-lung-interactions-second-
edition-peter-gehr/
ebookgate.com
https://ebookgate.com/product/gauge-theories-of-the-strong-weak-and-
electromagnetic-interactions-second-edition-chris-quigg/
ebookgate.com
https://ebookgate.com/product/constructing-reality-quantum-theory-and-
particle-physics-5th-edition-john-marburger/
ebookgate.com
https://ebookgate.com/product/advanced-concepts-in-particle-and-field-
theory-revised-edition-tristan-hubsch/
ebookgate.com
Non Covalent Interactions Theory and Experiment 1st
Edition. Edition Hobza Pavel
https://ebookgate.com/product/non-covalent-interactions-theory-and-
experiment-1st-edition-edition-hobza-pavel/
ebookgate.com
https://ebookgate.com/product/elementary-particle-theory-
volume-1-quantum-mechanics-1st-edition-eugene-stefanovich/
ebookgate.com
https://ebookgate.com/product/charged-particle-optics-theory-an-
introduction-1st-edition-timothy-r-groves/
ebookgate.com
https://ebookgate.com/product/small-boats-weak-states-dirty-money-
piracy-and-maritime-terrorism-in-the-modern-world-1st-edition-martin-
n-murphy/
ebookgate.com
1 — Classical Symmetries 2
1.1 Noether’s Theorem – Classical . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2
1.2 Examples . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4
1b — Gauge Symmetries 21
1b.1 Noether’s Theorem – Field Theory . . . . . . . . . . . . . . . . . . . . . . . . . . . . 21
1b.2 Gauge Theories . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 24
1b.3 Global Symmetries of Gauge Theories . . . . . . . . . . . . . . . . . . . . . . . . . . 26
i
Weak Interactions — Howard Georgi — draft - February 10, 2009 — ii
6a - Anomalies 105
6a.1 Electromagnetic Interactions and π 0 → 2γ . . . . . . . . . . . . . . . . . . . . . . . . 105
6a.2 The Steinberger Calculation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 110
6a.3 Spectators, gauge invariance and the anomaly . . . . . . . . . . . . . . . . . . . . . . 111
— Classical Symmetries
The concept of symmetry will play a crucial role in nearly all aspects of our discussion of weak
interactions. At the level of the dynamics, the fundamental interactions (or at least that subset of
the fundamental interactions that we understand) are associated with “gauge symmetries”. But
more than that, the underlying mathematical language of relativistic quantum mechanics — quan-
tum field theory — is much easier to understand if you make use of all the symmetry information
that is available. In this course, we will make extensive use of symmetry as a mathematical tool to
help us understand the physics. In particular, we make use of the language of representations of
Lie algebras.
where L is the local Lagrangian density. The index, j, is what particle physicists call a “flavor”
index. Different values of j label different types, or “flavors”, of the field φ. Think of the field,
φ, without any explicit index, as a column vector in flavor space. Assume, for simplicity, that the
Lagrangian depends only on the fields, φ, and their first derivatives, ∂µ φ. The equations of motion
are
δL δL
∂µ = . (1.1.2)
δ(∂µ φ) δφ
Note that (1.1.2) is a vector equation in flavor space. Each side is a row vector, carrying the flavor
index, j.
A symmetry of the action is some infinitesimal change in the fields, δφ, such that
S[φ + δφ] = S[φ] , (1.1.3)
or
L(φ + δφ, ∂µ φ + δ∂µ φ) = L(φ, ∂µ φ) + ∂µ V µ (φ, ∂µ φ, δφ) , (1.1.4)
where V µ is some vector function of the order of the infinitesimal, δφ. We assume here that we can
throw away surface terms in the d4 x integral so that the V µ terms makes no contribution to the
action. But
δL δL
L(φ + δφ, ∂µ φ + δ∂µ φ) − L(φ, ∂µ φ) = δφ + ∂µ δφ , (1.1.5)
δφ δ(∂µ φ)
2
Weak Interactions — Howard Georgi — draft - February 10, 2009 — 3
because δ∂µ φ = ∂µ δφ. Note that (1.1.5) is a single equation with no j index. The terms on the
right hand side involve a matrix multiplication in flavor space of a row vector on the left with a
column vector on the right. From (1.1.2), (1.1.4) and (1.1.5), we have
∂µ N µ = 0 , (1.1.6)
where
δL
Nµ = δφ − V µ . (1.1.7)
δ(∂µ φ)
Often, we will be interested in symmetries that are symmetries of the Lagrangian, not just
the action, in which case V µ = 0. In particular, our favorite symmetry will be linear unitary
transformations on the fields, for which
δφ = ia T a φ , (1.1.8)
where the T a for a = 1 to m are a set of N × N hermitian matrices acting on the flavor space, and
the a are a set of infinitesimal parameters. We can (and sometimes will) exponentiate (1.1.8) to
get a finite transformation:
a
φ → φ0 = eia T φ , (1.1.9)
which reduces to (1.1.8) for small a . The T a ’s are then said to be the “generators” of the trans-
formations, (1.1.9)
We will be using the T a ’s so much that it is worth pausing to consider their properties system-
atically. The fundamental property of the generators is their commutation relation,
[T a , T b ] = i fabc T c , (1.1.10)
where fabc are the structure constants of the Lie algebra, defined in any nontrivial representation (a
trivial representation is one in which T a = 0, for which (1.1.10) is trivially satisfied). The generators
can then be classified into sets, called simple subalgebras, that have nonzero commutators among
themselves, but that commute with everything else. For example, there may be three generators,
T a for a = 1 to 3 with the commutation relations of SU (2),
[T a , T b ] = i abc T c , (1.1.11)
and which commute with all the other generators. Then this is an SU (2) factor of the algebra. The
algebra can always be decomposed into factors like this, called “simple” subalgebras, and a set of
generators which commute with everything, called U (1)’s.
The normalization of the U (1) generators must be set by some arbitrary convention. However,
the normalization of the generators of each simple subgroup is related to the normalization of the
structure constants. It is important to normalize them so that in each simple subalgebra,
X
facd fbcd = k δab . (1.1.12)
c,d
Because the infinitesimal parameters are arbitrary, (1.1.15) actually defines m conserved currents,
δL
Jaµ = −i T aφ for a = 1 to m. (1.1.16)
δ(∂µ φ)
I stress again that this discussion is all at the level of the classical action and Lagrangian. We
will discuss later what happens in quantum field theory.
1.2 Examples
Example 1
Let Φj , for j = 1 to N , be a set of real scalar boson fields. The most general possible real quadratic
term in the derivatives of Φ in the Lagrangian is
1
LKE (Φ) = ∂µ ΦT S ∂ µ Φ . (1.2.1)
2
where S is a real symmetric matrix. Note the matrix notation, in which, the Φj are arranged
into an N -component column vector. In physical applications, we want S to be a strictly positive
matrix. Negative eigenvalues would give rise to a Hamiltonian that is not bounded below and zero
eigenvalues to scalars that do not propagate at all. If S is positive, then we can define a new set of
fields by a linear transformation
Lφ ≡ Φ, (1.2.2)
such that
LT S L = I . (1.2.3)
In terms of φ, the Lagrangian becomes
1
LKE (φ) = ∂µ φT ∂ µ φ . (1.2.4)
2
Weak Interactions — Howard Georgi — draft - February 10, 2009 — 5
This is the canonical form for the Lagrangian for a set of N massless free scalar fields. Under an
infinitesimal linear transformation,
δφ = G φ , (1.2.5)
where G is an N × N matrix, the change in LKE is
1
δLKE (φ) = ∂µ φT (G + GT ) ∂ µ φ . (1.2.6)
2
If G is antisymmetric, the Lagrangian is unchanged! We must also choose G real to preserve the
reality of the fields. Thus the Lagrangian is invariant under a group of SO(N ) transformations,
δφ = ia T a φ , (1.2.7)
where the T a , for a = 1 to N (N −1)/2 are the N (N −1)/2 independent, antisymmmetric imaginary
matrices. These matrices are a representation of the SO(N ) algebra.1 When exponentiated, (1.2.7)
produces an orthogonal transformation, a representation of the group SO(N )
φ → φ0 = O φ , (1.2.8)
where
OT = O−1 . (1.2.9)
This is a rotation in a real N -dimensional space.
Notice that we have not had to do anything to impose this SO(N ) symmetry ex-
cept to put the Lagrangian into canonical form. It was an automatic consequence of
the physical starting point, the existence of N free massless scalar fields. The canon-
ical form of the derivative term in the kinetic energy automatically has the SO(N )
symmetry.
The corresponding Noether currents are
The symmetry, (1.2.8), is the largest internal rotation symmetry that a set of N real spinless
bosons can have, because the kinetic energy term, (1.2.4), must always be there. However, the
symmetry may be broken down to some subgroup of SO(N ). This can happen trivially because of
the mass term if the scalars are not all degenerate. The mass term has the general form:
1
Lmass (φ) = − φT M 2 φ , (1.2.11)
2
where M 2 is a real symmetric matrix called the mass matrix. Its eigenvalues (if all are positive)
are the squared masses of the scalar particles. The mass term is invariant under an orthogonal
transformation, O, if
If M 2 is proportional to the identity matrix, then the entire SO(N ) is unbroken. In general,
the transformations satisfying (1.2.12) form a representation of some subgroup of SO(N ). The
subgroup is generated by the subset of the generators, T a , which commute with M 2 .
1
Usually, I will not be careful to distinguish an algebra or group from its representation, because it is almost
always the explicit representation that we care about.
Weak Interactions — Howard Georgi — draft - February 10, 2009 — 6
The mass matrix can be diagonalized by an orthogonal transformation that leaves the deriva-
tive term, (1.2.4), unchanged. Then the remaining symmetry is an SO(`) for each ` degenerate
eigenvalues. For example, if K of the fields have mass m1 and the other N − K have mass m2 , the
diagonal mass matrix can be taken to have to form
m21
..
. 0
m21
M2 = . (1.2.13)
m22
..
.
0
m22
The symmetry is an SO(K) × SO(N − K) which rotates the degenerate subsets, of the form
O1 0
O= , (1.2.14)
0 O2
where O1 and O2 act on the fields with mass m1 and m2 respectively. The finite group elements,
(1.2.14), are generated by exponentiation of the K(K − 1)/2 + (N − K)(N − K − 1)/2 generators
S1 0 0 0
and , (1.2.15)
0 0 0 S2
where S1 and S2 are antisymmetric imaginary matrices.
If the mass matrix is non-degenerate, that is with no pair of eigenvalues equal, the SO(N )
symmetry is completely broken and there is no continuous symmetry that remains, although the
Lagrangian is still invariant under the discrete symmetry under which any component of φ changes
sign.
The symmetry can also be broken down by interaction terms in the Lagrangian. With cubic and
quartic terms in φ, the symmetry can be broken in more interesting ways. Consider, as a simple
example that will be of use later, a Lagrangian with N = 8 scalars, with K = 4 with mass m1 (the
φj for j = 1 to 4) and the rest with mass m2 (the φj for j = 5 to 8). The kinetic energy and mass
terms then have the symmetry, SO(4) × SO(4). Arrange the 8 real fields into two complex doublet
fields as follows: √ √
(φ1 + iφ2 )/√2 (φ5 + iφ6 )/√2
ξ1 ≡ , ξ2 ≡ . (1.2.16)
(φ3 + iφ4 )/ 2 (φ7 + iφ8 )/ 2
Now consider an interaction term of the form
Lint = λ ξ1† ξ2 ξ2† ξ1 . (1.2.17)
This interaction term is not invariant under the SO(4) × SO(4), but it is invariant under an
SU (2) × U (1) symmetry under which
ξj → U ξj for j = 1 to 2, (1.2.18)
where the 2×2 matrix, U , is unitary. The U can written as a phase (the U (1) part) times a special
unitary matrix V (det V = 1), so it is a representation of SU (2) × U (1). This is the subgroup of
SO(4) × SO(4) left invariant by the interaction term, (1.2.17).
The symmetry structure of a field theory is a descending hierarchy of symmetry,
from the kinetic energy term down through the interaction terms. The kinetic energy
term has the largest possible symmetry. This is broken down to some subgroup by the
mass term and the interaction terms. This way of looking at the symmetry structure is
particularly useful when some of the interactions terms are weak, so that their effects
can be treated in perturbation theory.
Weak Interactions — Howard Georgi — draft - February 10, 2009 — 7
Example 2
Let ψj for j = 1 to N be free, massless, spin- 21 , four-component Dirac fermion fields with Lagrangian
The Dirac fermion fields are complex, thus (1.2.19) has an obvious SU (N ) × U (1) symmetry under
which, for infinitesimal changes,
δψ = ia T a ψ , (1.2.20)
where the T a are the N 2 hermetian N × N matrices, which generate the defining (or N ) represen-
tation of SU (N ) × U (1). When exponentiated this gives
ψ → U ψ, (1.2.21)
P± ≡ (1 ± γ5 )/2 . (1.2.25)
ψL ≡ P+ ψ , ψ R ≡ P− ψ . (1.2.26)
Then
ψL = ψP− , ψR = ψP+ . (1.2.27)
But then because
P+ γ µ = γ µ P− , (1.2.28)
we can write
L(ψ) = i ψ ∂/ψ = i ψL ∂/ψL + i ψR ∂/ψR . (1.2.29)
Evidently, we have the freedom to make separate SU (N ) × U (1) transformations on the L and R
fermions:
δψL = iL a
a T ψL , δψR = iR a
a T ψR , (1.2.30)
Weak Interactions — Howard Georgi — draft - February 10, 2009 — 8
or
ψL → UL ψL , ψR → UR ψR . (1.2.31)
The physical interpretation of the L and R states is that they are helicity eigenstates for on-mass-
shell fermions. Consider a plane wave moving in the 3 direction. The fermions are massless, thus
p0 = p3 while p1 = p2 = 0. The Dirac equation in momentum space is p/ψ = 0, or p0 (γ 0 − γ 3 )ψ = 0,
or
γ0ψ = γ3ψ . (1.2.32)
The spin angular momentum in the 3 direction is
σ 12 iγ 1 γ 2
J3 = = . (1.2.33)
2 2
Then
i i
J 3 ψL = 2 γ 1 γ 2 ψL = 2 γ 0 γ 0 γ 1 γ 2 ψL
i i (1.2.34)
= 2 γ 0 γ 1 γ 2 γ 0 ψL = 2 γ 0 γ 1 γ 2 γ 3 ψL
= − 21 γ5 ψL = − 12 ψL .
Thus ψL describes a particle with helicity − 12 , that is a left handed particle, while ψR describes a
particle with helicity 21 , that is a right handed particle.
The symmetry, (1.2.30), under which the L and R fields transform differently, is called a chiral
symmetry. Note that the generators of the chiral symmetry all commute with the fermion number.
The chiral symmetry is still not the largest internal symmetry of (1.2.19). To identify the largest
symmetry, we must introduce the idea of charge conjugate fields. Define
ψc = Cψ ∗ , , (1.2.35)
and
ψcL = P+ ψc , ψcR = P− ψc , (1.2.36)
where C is the charge conjugation matrix acting on the Dirac indices, satisfying
C2 = 1 , C† = C , C γ µ∗ C = −γ µ . (1.2.37)
We will sometimes take C to be nontrivial in flavor space as well, but for now, we will assume that
it acts in the space of the Dirac indices. The form of C depends on the representation of the γ
matrices. For example, in the so-called Majorana representation in which all the γ’s are imaginary,
C = 1, while in the standard representation of Bjorken and Drell, C = iγ 2 .
The name “charge conjugation” comes originally from the effect of the transformation (1.2.35)
on the Dirac equation for a particle in an electromagnetic field. Such a particle, with mass m and
charge e, satisfies
i ∂/ψ − eAψ
/ − mψ = 0 , (1.2.38)
where Aµ is the (hermetian) electromagnetic field. Taking the complex conjugate, multiplying by
C and using (1.2.35) gives
i ∂/ψc + eAψ
/ c − mψc = 0 . (1.2.39)
The C operation changes the particle of charge e to its antiparticle, with the same mass but charge
−e.
Weak Interactions — Howard Georgi — draft - February 10, 2009 — 9
The interesting thing about charge conjugation from our present point of view is that it changes
L to R,
1−γ5∗
∗ = C (P )∗ ψ ∗ = C
CψR ψ∗
− 2
(1.2.40)
1−γ5∗
=C 2 CC ψ∗ = P+ ψc = ψcL ,
where we have used Cγ5∗ C = −γ5 . This is what we need to find a larger symmetry of (1.2.19).
The point is that with a purely internal symmetry, we cannot mix up the ψL with the ψR . A
transformation such as
δψL = ia T a ψR (1.2.41)
makes no sense at all. You can see this trivially by acting on both sides with P− .
But if we first change the ψR into ψcL with charge conjugation, then we can mix them with the
ψL . Such a transformation will not commute with the U (1) of fermion number, because ψ and ψc
have opposite fermion number, but at this point, we do not care.
We can use (1.2.40) to show that
We drop the total derivative, which does not effect the action, and then combine the L fields into
a 2N component column vector:
ψL
ΨL = . (1.2.43)
ψcL
In terms of Ψ, the Lagrangian is
L(Ψ) = i Ψ ∂/Ψ . (1.2.44)
Clearly, the Lagrangian, (1.2.44), in invariant under SU (2N ) × U (1) transformations, generated by
hermetian 2N × 2N matrices acting on the extended flavor space of (1.2.43). Note that if fermion
number is not conserved, there is not even any reason that there should be an even number of L
fields!. For example, in the standard model, we seem to have 3 approximately massless left-handed
neutrinos, each described by a two-component Weyl field.
Most of the Lagrangians that we actually deal with will have interactions that distinguish
between particles and their antiparticles, thus breaking the symmetries that mix fermions with
antifermions. For example, if the fermions have a charge, as in QED, the symmetry is broken
down to a chiral symmetry, (1.2.30), because the fact that the charge changes sign under charge
conjugation breaks the larger symmetry.
Fermion Masses
The most general fermion mass term that preserves fermion number has the form,
where
ψL = (ψL )† γ 0 , ψR = (ψR )† γ 0 , (1.2.46)
and the mass matrix, M , is an arbitrary complex matrix. Note that the two terms in (1.2.45)
are hermetian conjugates of one another. Because the ψL and ψR terms are coupled together, the
Lagrangian is not invariant under the chiral symmetry, (1.2.30). A chiral transformation changes
M
M → UL† M UR . (1.2.47)
Weak Interactions — Howard Georgi — draft - February 10, 2009 — 10
However, that means that we can use the chiral symmetry to put the mass matrix, M , into a
canonical form. We can diagonalize it and make each of the eigenvalues real and positive!2
The symmetry of the mass term that commutes with fermion number is then determined by
the degeneracy of the diagonal mass matrix. If all the masses are different, then the only symmetry
that remains is a U (1)N symmetry, separate conservation of each of the fermion flavors. If ` flavors
have the same mass, there is a non-Abelian SU (`) symmetry. For example, if all N fermions are
degenerate, the mass matrix is proportional to
Lmass = ψψ , (1.2.48)
This breaks the chiral symmetry, (1.2.30), but it is still invariant under the SU (N )×U (1) symmetry,
(1.2.20), in which the L and R fields rotate together.
If fermion number is not conserved, so that the fermions are most appropriately described by
N left-handed Weyl fields, ΨL , the most general mass term has the form
ΨcR M ΨL , (1.2.49)
where M is a complex, symmetric matrix. The kinetic energy term is invariant under an SU (N ) ×
U (1) symmetry:
ΨL → U ΨL . (1.2.50)
Under (1.2.50), the mass matrix, M changes to
M → M0 = UT M U , (1.2.51)
This can be used to make M diagonal and positive.
Problems
1.1. The Lagrangian for a set of N massless free scalar fields, (1.2.4), has a rather peculiar
symmetry under which
φ → φ + a,
for any constant a. What is the Noether current? Note that this “symmetry” really is peculiar.
While it is a symmetry of the classical Lagrangian, it is not a symmetry of the corresponding quan-
tum theory. This is one of the simplest examples of a continuous symmetry that is spontaneously
broken. Spontaneous symmetry breaking is a subject to which we will return many times.
1-2. Find a basis for the γ µ matrices in which the charge conjugation matrix, C, is the identity.
This is called a Majorana representation. In this representation, we can define real 4-component
fields
√ √
ψ1 = (ψ + ψc )/ 2 = (ψ + ψ ∗ )/ 2 ,
√
ψ2 = i (ψ − ψ ∗ )/ 2 .
Rewrite the mass term for N degenerate Dirac fermions, (1.2.48), in terms of these real fields,
and identify the symmetry of this term. What’s the symmetry group? Show that this is also a
symmetry of the kinetic energy term for free Dirac fermions.
1-3. Show that C in (1.2.37) is essentially unique in any given representation of the γ matrices.
2
This is all at the classical level — we will come back to the question of real masses later on when we discuss the
strong CP puzzle.
Chapter 1a
Many of the results of the previous chapter can be carried over immediately to quantum field
theory. However, the necessity of regularization and renormalization sometimes makes things more
interesting. In this chapter, we discuss an approach to quantum field theory that is particularly
useful for the discussion of weak interactions and the Standard model in general. We will assume
that the reader is familiar with quantum field theory, and has available an encyclopedic text such as
[Itzykson]. Here we will not pretend to anything like completeness, but will concentrate instead on
explaining what is really going on. The detail of our regularization and renormalization scheme will
seldom matter, but when it does, we will use dimensional regularization and minimal subtraction.
A brief review will be included in an appendix to this book.
11
Weak Interactions — Howard Georgi — draft - February 10, 2009 — 12
small distances is modified in some way to make the theory well-defined. Then the dependence on
the short distance physics is incorporated into a set of parameters that can be related to physical
quantities at measurable distances. A renormalizable theory is one in which only a finite number
of parameters are required to absorb all the dependence on short distance physics.
We will have much more to say about these issues as we develop the idea of effective field
theory. Eventually, we will make the idea of hiding unknown physics at small distances into one of
our principle tools. For now, we simply discuss renormalizable, local quantum field theory as if it
were a God-given truth.
Let us begin by discussing a simple massless scalar field theory, defined by
1 µ λ
L(φ) = ∂ φ∂µ φ − φ4 + sφ φ . (1a.1.1)
2 4!
In addition to the quantum field, φ, with mass dimension 1, we have included a c-number source,
sφ , with mass dimension 3 for the scalar field φ. This is convenient, because it gives us a handle
with which to tweak the theory to see how it responds. All the Green functions are implicitly
present in the structure of the “vacuum” as a function of the source,
W (s) is the generating function for the connected Green functions. Actually, as we will discuss
later, it is not so easy to define this so-called λφ4 theory beyond perturbation theory, because the
theory is sick at short distances, but for now, we will be happy with a perturbative definition.
Without the source term, the action derived from (1a.1.1) is classically scale invariant. This
means that it is invariant under the following transformation:
where
x0 = e λ x φ0 (x0 ) = e−λ φ(x) (1a.1.4)
The classical scale invariance would be broken if we put in a mass term. We will leave the mass
term out so that we have only one parameter to talk about — the coupling λ.
The first interesting feature that we encounter when we look at a quantum field theory defined
by dimensional regularization and minimal subtraction (DRMS) is that scale invariance is broken.
All the parameters in the theory depend on the renormalization scale, µ, used to define their
dimensional extensions. The appearance of the parameter µ has an important physical consequence,
which is easy to understand in perturbation theory. Quantities calculated in DRMS depend on
ln µ. If all the other relevant dimensional parameters in the calculation are of the same order as
µ, these logs cause no problems. However, if some dimensional parameters are very different from
µ, perturbation theory will break down before it should. Thus calculations should be organized so
that, as much as possible, all relevant dimensional parameters are of order µ. Fortunately, µ can
be chosen at our convenience, and given the parameters for one value of µ, we can calculate the set
of parameters for a different value of µ that lead to the same physics.
translation can be built up by putting small ones together! Consider the parameter λ in (1a.1.1).
If we know λ at µ1 , then we can compute it at µ2 using perturbation theory:
for some function L. If both λ(µ1 ) and ln(µ2 /µ1 ) are small, we can reliably compute L in per-
turbation theory. However, so long as both λ(µ1 ) and λ(µ2 ) are small, we can dispense with the
restriction that ln(µ2 /µ1 ) is small. Instead of using (1a.1.5) directly, we can get from µ1 to µ2 in
a series of small steps. The lowest order perturbative result is
µ2
2
λ(µ2 ) = λ(µ1 ) + B λ(µ1 ) ln + O(λ3 ) , (1a.1.6)
µ1
for some constant B.
Let
1 µ2
∆ = ∆(ln(µ)) ≡ ln . (1a.1.7)
N µ1
Then
λ(µ1 e∆ ) = λ(µ1 ) + B λ(µ1 )2 ∆ + O(λ3 ) ,
···
or in the limit as N → ∞,
Z µ2
dµ
λ(µ2 ) = λ(µ1 ) + B λ(µ)2 + O(λ3 ) , (1a.1.10)
µ1 µ
We have arrived, by a rather backwards route, at the lowest order integral equation for the “running”
coupling, λ(µ). Note that we have not done anything at all except to build up a large change in
µ out of infinitesimal changes. If we differentiate with respect to µ2 and set µ2 = µ, we get the
standard differential equation for the running coupling:
∂
µ λ(µ) = βλ (λ(µ)) , (1a.1.11)
∂µ
where
βλ (λ) = B λ2 + O(λ3 ) (1a.1.12)
is the β-function. We will see below that B > 0, so that λ is an increasing function of µ for small
λ.
Ignoring the terms of order λ3 , we can solve (1a.1.12) to obtain
λ(µ0 ) 1
λ(µ) = = . (1a.1.13)
1 + B λ(µ0 ) ln(µ0 /µ) B ln(Λ/µ)
Weak Interactions — Howard Georgi — draft - February 10, 2009 — 14
This solution exhibits the most interesting feature of the running coupling, dimensional transmuta-
tion[Coleman 73]. Because the dimensionless quantity, λ, is a calculable function of the dimensional
quantity, µ, the parameter that actually characterizes the physics is a dimensional one, Λ. (1a.1.13)
also shows the difficulty with this theory. The scale µ cannot be taken arbitrarily large. Above
µ ≈ Λ, the coupling is large and the theory is not defined perturbatively. There are good reasons
to believe that this difficulty persists beyond perturbation theory and that the λφ4 theory does not
make sense at short distances.
This is sometimes discussed in terms of the rather ridiculous word, “triviality”. The word is a
holdover from an old fashioned way of looking at field theory. All it means is that the λφ4 theory is
not a useful description of the physics at arbitrarily short distances. If you try to define the theory
in isolation, by cutting the physics off at some small distance, renormalizing, and then letting the
distance go to zero, the theory is “trivial” because the renormalized coupling is driven to zero. This
is simply the flip side of (1a.1.13) in which a finite renormalized coupling leads to a disaster at a
finite small distance. In practice, this means that the physics has to change at some momentum
scale smaller than the Λ in (1a.1.13). However, as we will see in more detail later on, this does not
prevent us from making sense of the theory at scales at which λ(µ) is small.
+
(1a.1.15)
This vanishes in DRMS for a massless scalar (and in any sensible scheme, it does not depend on
the external momentum) and thus the wave function renormalization is 1 in one loop and γφ = 0
to order λ.
The Feynman graphs that contribute the Γ4 to one loop are
@
+
@
@ + crossed
@ graphs
@
@ @ (1a.1.16)
Weak Interactions — Howard Georgi — draft - February 10, 2009 — 15
3λ2
βλ − + O(λ3 ) = 0 , (1a.1.18)
16π 2
or
3λ2
βλ = + O(λ3 ) . (1a.1.19)
16π 2
It is worth commenting on some minus signs that frequently cause confusion. Note that the µ
dependence of λ(µ) is opposite to that of Γ4 . Because the wave function renormalization vanishes,
Γ4 is a directly physical scattering amplitude, to this order. Thus when µ changes, Γ4 remains
constant. That is a direct translation of (1a.1.14), with γφ = 0.
On the other hand, the “solution”, to the renormalization group equation, is
for λ(µ) satisfying (1a.1.13). Since λ(µ0 ) can be regarded as a function of λ(µ) and µ, (1a.1.20)
says that Γ4 is not a function of λ(µ) and µ separately, but only of the combination, λ(µ0 ). The
µ dependence of λ is then related, in the absence of wave function renormalization, not to the µ
dependence of Γ4 , but to its dependence on the scale of p, because it is when p ≈ µ, that Γ4 ≈ λ(µ).
Secondly, because W (s) is the generating function for the connected Green functions, we can obtain
(1a.1.21) for all n from the single equation
∂ ∂ ∂
Z
µ + βλ + βs(x) d4 x W (λ, s) = 0 . (1a.1.22)
∂µ ∂λ ∂s(x)
In words, (1a.1.24) means that a change in µ can always be compensated by suitable changes in all
the parameters and the fields.
Of course this assumes, as usual, that we have included all the parameters required to absorb
the dependence on the physics at very short distances. Once all the relevant parameters have been
included, then µ is arbitrary, and (1a.1.24) is almost a tautology.
In the simple example, (1a.1.1), the field, φ, is multiplicatively renormalized, because it is the
only field around. In our renormalization group language, that statement is equivalent to (1a.1.23).
Note, in fact, that (1a.1.12) is the most general thing we can write down consistent with dimensional
analysis. However, in general, the situation is more complicated.
Suppose, for example, that your theory depends on several real scalar fields, φα , for α = 1 to
K, with Lagrangian
1 X µ
L(φ) = (∂ φα ∂µ φα + sα φα )
2 α
(1a.1.25)
X λ α1 α2 α3 α4
− φα1 φα2 φα3 φα4 ,
α1 α2 α3 α4 4!
(K+3)!
where λα1 α2 α3 α4 is some symmetric tensor in the α’s. If all 4!(K−1)! of the independent α’s are
nonzero, then there are no symmetries that distinguish one φ from another. In this case, each of
the βλ -functions depends on all of the λ’s, and the most general form for βs is
X
βsα (x) = sα0 (x) γα0 α (λ) . (1a.1.26)
α0
This exhibits the phenomenon of operator mixing. Of course, all of the γ’s are zero in one loop in
the scalar field theory, but in higher order, a change in µ requires not just a rescaling of each of the
φα ’s, but a rotation in the flavor space, to get the theory back into a canonical form. The fields
are not multiplicatively renormalized.
The failure of multiplicative renormalization in (1a.1.26) is an example of the general principle
that whatever can happen will happen. This is particularly true in quantum field theory, where if
you fail to write down an appropriately general expression, the theory is likely to blow up in your
face, in the sense that you will not be able to absorb the dependence on short distance physics.
However, as long as you are careful to include everything (consistent, as we will emphasize time
Weak Interactions — Howard Georgi — draft - February 10, 2009 — 17
and time again, with all the symmetries of the system), the situation is really no more difficult
to understand than in a simple theory with multiplicative renormalization. You just have to keep
track of some indices.
As an example of operator mixing in action, we will show how to go back and forth from 1PI
2-point functions to connected 2-point functions. Consider the connected 2-point function
δ2
Gα1 α2 (x1 , x2 ) = − W (s)
δsα1 (x1 )δsα2 (x2 ) (1a.1.27)
= h0 | T φα1 (x1 )φα2 (x2 ) | 0i .
Because W (s) satisfies, (1a.1.24), we can differentiate twice and use (1a.1.26) to obtain with
∂ ∂
µ + βλ ≡ D, (1a.1.28)
∂µ ∂λ
X h i
D δα1 α01 δα2 α02 + γα1 α01 + γα2 α02 Gα01 α02 (x1 , x2 ) = 0 . (1a.1.29)
α01 α02
Let G̃(p) be the Fourier transform of G(x, 0). To avoid getting lost in indices, it will be useful
to write (1a.1.29) for G̃ (Fourier transforming commutes with applying D) in an obvious matrix
notation,
D G̃ + γ G̃ + G̃ γ T = 0 . (1a.1.30)
The 1PI two point function is the inverse of G̃ — symbolically
Γ G̃ = G̃ Γ = I . (1a.1.31)
= (D Γ) G̃ − Γ γ G̃ − Γ G̃ γ T . (1a.1.32)
= (D Γ) G̃ − Γ γ G̃ − γ T
Multiplying by Γ on the right and using (1a.1.31) gives the renormalization group equation for the
1PI 2-point function:
D Γ − Γ γ − γT Γ = 0 . (1a.1.33)
Note not only the sign change, compared to (1a.1.30), but the change from γ to γ T as well.
where
δW
Φα (x) = , (1a.1.35)
δsα (x)
Weak Interactions — Howard Georgi — draft - February 10, 2009 — 18
But
δΓ
s(x) = −
, (1a.1.38)
δΦ
which follows from (1a.1.34). Putting this into (1a.1.23) and using (1a.1.35), we find
∂ ∂ X ∂
µ + βλ − Φ α γ α0 α Γ(Φ)
∂µ ∂λ α,α0 ∂Φα0
(1a.1.39)
∂ ∂ T
X ∂
= µ + βλ − Φα γαα 0 Γ(Φ) = 0 .
∂µ ∂λ α,α0 ∂Φα0
The quantum fermion fields, ψ, have dimension 3/2. Their sources, η, have dimension 5/2. Both
are anticommuting.
The Lagrangian, (1a.1.40), is the most general Lagrangian we can write down for massless fields
with these dimensions, consistent with a parity symmetry under which the φα fields are odd. Thus,
(1a.1.24) will be satisfied, where the couplings now include the gα ’s as well as the λ’s, and the
sources include the η and η, as well as the sα ’s. A feature of this theory that makes it a nice
pedagogical example, is that all β-functions are nonzero at the one loop level. You will find all of
these for yourselves in problem (1a-3).
1. Couple a source, t, to each composite operator in the Lagrangian. We will find that if we
include an operator of a given dimension, we will be forced to include sources for all the other
operators with the same dimension and the same quantum numbers under all the symmetries
of the theory. When masses are present, we may have to include others as well, but the theory
will complain if we do not have the right sources.
2. Renormalize the theory as usual. To do this, you will have to include extra parameters for
the new “interactions” that you can build with your new sources, one for each new dimension
4 object you can build. These will be required to absorb dependence on the unknown short
distance physics.
3. Once this is done, the generating functional, W , depends on the new sources, and we can
find the effect of an insertion of a composite operator in any Green function simply by
differentiating with respect to the appropriate source.
As a fairly simple example, but one which exhibits many of the interesting features of composite
operators, consider the fermion example, (1a.1.40), with K = 1 (dropping the unneeded subscripts),
and include a source for the operator ψ ψ. This is the only dimension 3 operator that is a scalar
under parity, so we do not need any other sources. The Lagrangian, constructed according to the
above rules, is
1 µ
L(φ, ψ) = i ψ ∂/ ψ + ηψ + ψη +∂ φ∂µ φ + sφ φ
2 (1a.2.1)
λ κ1 µ κ2 4 κ3 2 2
− φ4 − g ψ iγ5 φ ψ + tψ ψ − ∂ t∂µ t − t − t φ ,
4! 2 4! 4
The point of the extra terms in (1a.2.1) is that they are require for the calculation of graphs
involving more than one insertion of the new operator. For example, consider the graph
× ×
(1a.2.2)
where the × indicates the insertion of the operator ψ ψ. When computed using DRMS, the graph
is µ-dependent. This means that it depends on the details of how we constructed the ψ ψ operator
at short distances, so we have no right to calculate it. Thus we need a new, a priori unknown
parameter in the theory. This is κ1 .
The κ parameters have β-functions just like the other coupling constants. The µ-dependence
of (1a.2.2) produces µ-dependence in κ1 , which is just its β-function in the renormalization group.
If, somehow, we manage to discover the values of this and the other κ parameters at some µ (this
requires somehow specifying the physics — we will see some examples later on), then we can use
the renormalization group to find its value at any other µ.
The κ parameters are the translation into our language of so-called “subtractions” which must be
made to define the Green functions of composite operators. We now see that they are something that
we already understand. The sources for the composite fields differ from the sources for elementary
fields in that their dimensions are smaller. This is what allows us to build the nontrivial κ terms.
We will see many examples of composite operators as we explore the standard model. The thing
to remember is that in this way of looking at things, they are not so different from other fields and
sources.
Weak Interactions — Howard Georgi — draft - February 10, 2009 — 20
Problems
1a-1. Find the one-loop β-functions for (1a.1.25). If you use (1a.1.19), no calculation should
be required, beyond drawing the graphs and keeping track of the indices.
1a-3. Find all the β-functions, both for the couplings and for the sources, for the theory
described by (1a.1.40), for K = 2.
Chapter 1b
— Gauge Symmetries
In this section, we will discuss theories with local or gauge symmetry. First, we will use gauge
symmetry as a device to define the Noether currents in a quantum field theory. Then we will go
on the discuss dynamical gauge symmetry.
Given a symmetry of a classical Lagrangian without sources, we can always extend it to a
symmetry of the Lagrangian including sources, by requiring the sources to transform appropriately.
The interesting question is, can we always define a corresponding quantum theory that exhibits
the symmetry.
The answer is — sometimes! When it is possible to find a regularization and renormalization
scheme that preserves the symmetry, then the symmetry can be extended into the quantum theory,
but sometimes this is not possible. We will discuss important examples of the “anomalies” that can
occur in quantum field theory to prevent the realization of a classical symmetry later. For now we
will ignore this subtlety, and discuss symmetry in the language we have developed to discuss field
theory, without asking whether any particular symmetry is consistent with our DRMS scheme.
L0 (φ, ∂ µ φ, ψ, ∂ µ ψ) + sT φ + η ψ + ψ η , (1b.1.1)
φ → Dφ (g −1 ) φ ,
ψL → DL (g −1 ) ψL , (1b.1.2)
ψR → DR (g −1 ) ψR ,
where g is and element of G and the D’s are unitary representations of the symmetry group. Note
that because the φ fields are real, the representation Dφ must be a “real” representation. That is,
the matrices, Dφ , must be real. Such a symmetry can then be extended to include the sources as
1
We use L0 for the Lagrangian without any sources.
21
Weak Interactions — Howard Georgi — draft - February 10, 2009 — 22
well, as follows:
s → Dφ (g −1 ) s ,
ηR → DL (g −1 ) ηR , (1b.1.3)
ηL → DR (g −1 ) ηL .
Note the switching of L ↔ R for the fermion sources, because the fermions source terms are like
mass terms that couple fields of opposite chirality. Now if the regularization and renormalization
respect the symmetry, the vacuum amplitude, Z(s, η, η) will also be invariant under (1b.1.3).
Generators — T a
It is often convenient to use the infinitesimal version of (1b.1.2) and (1b.1.3), in terms of the
generators, T a ,
δφ = ia Tφa φ ,
The Tja for j = φ, L and R, are generators of the representation of the Lie algebra of G, corre-
sponding to the representations,
a
Dj (g −1 ) = eia Tj . (1b.1.5)
A sum over repeated a indices is assumed. Often we will drop the subscript, and let the reader
figure out from the context which representation we are discussing.
Gauge Symmetry
In order to discuss the conserved Noether currents associated with a symmetry of the quantum
field theory as composite operators, it is useful to convert the global symmetry, (1b.1.4), into a
gauge symmetry, that is a symmetry in which the parameters can depend on space-time,
a → a (x) . (1b.1.6)
To do this, we must introduce a set of “gauge fields”, which in this case will be classical fields, like
the other sources. Call these fields, hµa . Under the symmetry, (1b.1.4), they transform as follows:
Now we modify the Lagrangian, (1b.1.1), by replacing the derivatives by “covariant” derivatives,
Dµ ≡ ∂ µ − i hµa T a , (1b.1.8)
Weak Interactions — Howard Georgi — draft - February 10, 2009 — 23
where the T a are the generators of the representation of the object on which the derivative acts.
Thus,
Dµ φ ≡ ∂ µ − i hµa Tφa φ ,
(1b.1.9)
Dµ ψ ≡ (∂ µ − iP+ hµa TLa − iP− hµa TRa ) ψ.
The resulting Lagrangian is now invariant under (1b.1.4) and (1b.1.7), for space-time dependent
parameters, a .
Furthermore, the classical gauge fields, hµa , act as sources for composite fields in the quan-
tum field theory. These are the classical Noether currents corresponding to the global symmetry,
(1b.1.2). To see this, note that
δ Dµ φ
= −i δνµ Tφa φ (1b.1.10)
δhνa
Suppose that the Lagrangian depended only on the φs. Then using (1b.1.10), the chain rule for
functional differentiation and the fact that φ doesn’t depend on hµa , we could write
δL δL δ Dν φ δL
µ = = −i Ta φ (1b.1.11)
δ ha δ Dν φ δ hµa δ Dµ φ φ
which is the classical Noether current.
This works in general, because for any field ξ,
δ Dµ ξ
= −i δνµ Tξa ξ (1b.1.12)
δhνa
where Tξa are the generators of the symmetry acting on the field ξ. Thus
δL X δL δ Dν ξ X δL
µ = νξ δ hµ = −i Ta ξ
µξ ξ
(1b.1.13)
δ ha ξ
δ D a ξ
δ D
Thus we expect such terms to be required by renormalization. With our convenient normalization,
(1.1.12), the parameter κab is
κab = δab κa (1b.1.16)
where κa is independent of a within each simple subalgebra.2
2
For U (1) factors of the symmetry algebra, there is no normalization picked out by the algebra, and no constraint
of the form (1b.1.16).
Weak Interactions — Howard Georgi — draft - February 10, 2009 — 24
hµ ≡ T a hµa . (1b.1.17)
where = a Ta . In this form, the gauge transformation can be integrated to give the finite form
where
Ω = ei . (1b.1.23)
= ∂ µ Gν − ∂ ν Gµ − i [Gµ , Gν ] .
The κ term of (1b.1.14) becomes the kinetic energy term for the gauge fields,
X 1
− Gaµν Gµν
a . (1b.2.2)
a 4ga2
Again, the normalization, that we have here called 1/4ga2 , is equal for a in each simple component
of the gauge group. The constant, ga , is the gauge coupling constant. We can rescale the fields
to make the coefficient of the kinetic energy term the canonical 1/4 if we choose. Then the gauge
transformation law has ga dependence. Often, we will keep the form of the gauge transformations
fixed and let the gauge coupling constant appear as the normalization of the kinetic energy term.
This is particularly convenient for the study of the symmetry structure of the theory.
Weak Interactions — Howard Georgi — draft - February 10, 2009 — 25
Gauge Fixing
The field theory treatment of gauge theories is more complicated than that of theories without
dynamical gauge symmetry because the gauge symmetry must be broken in order to define the
theory, at least perturbatively. The kinetic energy term, (1b.2.2), does not involve the longitudinal
components of the gauge fields, and is not invertible in momentum space. Thus the propagator is
not well defined.
In the functional integral formulation of field theory, one tries to write the Green functions as
functional integrals:
hg̃i
h0 | T g(G, φ) | 0i = , (1b.2.3)
h1̃i
where Z
hg̃i = g(G, φ) eiS(G,φ) [dG][dφ] . (1b.2.4)
In the functional integral description, the difficulty with gauge invariance shows up in the indepen-
dence of the action on gauge transformation,
This defines an “orbit” in the space of gauge field configurations on which the action is constant. The
part of the gauge field functional integral over the gauge transformations is completely unregulated,
and gives rise to infinities that are the translation into the functional integral language of the
problem with the gauge field propagator. Thus hg̃i has spurious infinities, and is not well defined.
Having formulated the problem in this language, Fadeev and Popov suggested a general proce-
dure for solving it. They introduced 1 into (1b.2.3), in the following form:
Z
1 = ∆(G, φ) f (GΩ , φΩ ) [dΩ] , (1b.2.6)
where f is a functional that is not gauge invariant and [dΩ] is the invariant measure over the gauge
transformations, satisfying
[dΩ] = [d(Ω0 Ω)] = [d(ΩΩ0 )] , (1b.2.7)
for fixed Ω0 . The functional, ∆, called the Fadeev-Popov determinant, is gauge invariant. This can
be seen as follows (using (1b.2.7)):
1
∆(GΩ , φΩ ) = R [dΩ0 ] (1b.2.8)
f (GΩΩ0 , φΩΩ0 )
1 1
R
0
=R = ∆(G, φ) . (1b.2.9)
f (GΩΩ0 , φΩΩ0 ) [d(ΩΩ )] f (GΩ , φΩ ) [d(Ω)]
Inserting (1b.2.6), we can write
Z
hg̃i = eiS(G,φ) g(G, φ) ∆(G, φ) f (GΩ , φΩ ) [dG][dφ][dΩ] . (1b.2.10)
If g(G, φ) is gauge invariant, then (1b.2.10) can be written with the integral over the gauge trans-
formations factored out, as
Z Z
hg̃i = eiS(G,φ) g(G, φ) ∆(G, φ) f (GΩ , φΩ ) [dG][dφ] [dΩ] . (1b.2.11)
Weak Interactions — Howard Georgi — draft - February 10, 2009 — 26
R
The Fadeev-Popov suggestion is then to drop the factor of [dΩ], and define
Z
hgi = eiS(G,φ) g(G, φ) ∆(G, φ) f (GΩ , φΩ ) [dG][dφ] . (1b.2.12)
This process is called “choosing a gauge”. The gauge is defined by the choice of the fixing function,
f . The important point to notice is that hgi should be completely independent of the gauge fixing
function, f (G, φ), so long as the Green function, g(G, φ), is gauge invariant. Because it is the gauge
invariant physical matrix elements that we actually want, this is a perfectly reasonable definition.
Of course, this means that in general, Green functions are not invariant for a given gauge fixing
function, f . That means that the consequences of gauge symmetry are not as obvious as those of
a global symmetry. However, it is possible to organize the calculations in gauge theory in a way
that makes gauge invariance manifest, by using what is called “background field gauge”. This is
discussed in Appendix B.
and study its global symmetries. Of course, the gauge symmetry itself defines a global symmetry,
but the theory may have additional symmetry. We will reserve the term “flavor symmetry” to refer
only to those internal symmetries of the theory that are not gauged. The first important comment
is that any such symmetry must map the gauge fields into themselves. Because the gauge fields are
associated with the generators of the gauge symmetry, this implies that the gauge symmetry is a
normal subgroup of the full symmetry group. For continuous flavor symmetries, the constraint is
even stronger. The generators of continuous flavor symmetries must commute with all the gauge
generators.3 This, in turn, implies that the flavor generators are constant within each irreducible
representation of the gauge group, and also that if two fields are rotated into one another by the
flavor symmetry, they must belong to the same irreducible representation of the gauge group. If
we choose a canonical form for the gauge generators in each irreducible representation of the gauge
symmetry,4 then we can regard the ψ and φ fields as having two sets of indices: gauge indices, on
which the gauge symmetries act; and flavor indices, on which the flavor symmetries act.
First consider the fermion kinetic energy term. In the basis in which all the fermions are left
handed, this has the general form X
i / ψrL ,
ψ rLD (1b.3.2)
r
where the sums run over the different irreducible representations of the gauge group, and where,
for each r, the fermions field is a vector in a flavor space with dimension dr . Each of these flavor
vectors can be rotated without changing (1b.3.2). Thus, at least classically, (1b.3.2) has a separate
SU (dr ) × U (1) flavor symmetry for each irreducible representation that appears in the sum (with
non-zero dr ).5
If we charge conjugate some of the left handed fermion fields to write them as right handed
fields, we must remember that the charged conjugate fields may transform differently under the
3
Discrete symmetries, like parity, my induce nontrivial automorphisms on the gauge generators.
4
See problem (3-2) for an example of what happens when you do not do this.
5
As we will see later, some of the U (1)s are broken by quantum mechanical effect, “anomalies”.
Weak Interactions — Howard Georgi — draft - February 10, 2009 — 27
gauge transformations. If
δψL = ia T a ψL , (1b.3.3)
then
δψcR = ia (−T a∗ )ψcR , (1b.3.4)
In a “real” representation, the generators can be taken to be imaginary, antisymmetric matrices, so
that −T a∗ = T a , and the right handed fields transform the same way. However, if the representation
is “complex”, then −T a∗ is a different representation. Finally, the representation may be “pseudo-
real”, in which case the representation generated by −T a∗ is equivalent to that generated by T a ,
− T a∗ = S T a S −1 , (1b.3.5)
but the matrix S is antisymmetric, so the generators can not be made entirely real. In this case, one
can include the matrix S into the definition of charge conjugation to make the right handed fields
transform like the left handed fields. The most familiar example of a pseudo-real representation is
the Pauli matrices, generating the spin 1/2 representation of SU (2).
The situation with scalar fields is somewhat more subtle, because the global symmetry asso-
ciated with a given representation of the gauge group depends on whether it is real, complex, or
pseudo-real.
For d sets of scalars in a real irreducible representation, rR , of the gauge group, the flavor
symmetry is SO(d). Here the kinetic energy term looks like
Dµ ΦT Dµ Φ , (1b.3.6)
where Φ is a vector in flavor space (as well as the space of the gauge group representation).
If, instead, the scalar field representation, rC , is complex, then because the elementary scalar
fields are real, both the rC representation and its complex conjugate, rC , appear in the theory.
Suppose that there are d copies of the representation, rC , which we can organize into a vector in
flavor space, as usual. We can write the kinetic energy term as
D µ φ† Dµ φ . (1b.3.7)
However, (1b.3.7) describes the rC representation as well, because we could just as well have used
the field φ∗ , which transforms under the rC representation. If the representation, rC (and rC ) is
n dimensional, then there are actually 2nd real fields described by (1b.3.7), because each of the
components of φ is complex. The flavor symmetry of (1b.3.7) is an SU (d) acting on the flavor
indices of φ.
The most bizarre situation occurs for pseudo-real representations. It turns out that a pseudo-real
representation has an SU (2) structure built into it. The generators of a 2n dimensional irreducible
pseudo-real representation can be taken to have the form6
~a ,
Ta = Aa + ~τ · S (1b.3.8)
~ matrices
where ~τ are the Pauli matrices, and the n × n matrix A is antisymmetric while the n × n S
are symmetric. Then
τ2 Ta∗ τ2 = −Ta . (1b.3.9)
6
Note that all pseudo-real representations have even dimensionality.
Weak Interactions — Howard Georgi — draft - February 10, 2009 — 28
The flavor structure of the theory involves the ~τ matrices in a non-trivial way. If there are d
identical representations, we can describe them by a 2n × 2d matrix field of the following form:
···
Σ11 Σ1d
.. .. ..
. . . , (1b.3.10)
Σn1 ··· Σnd
for real σij and ~πij . The gauge generators act on Σ on the left. Now the kinetic energy term has
the form
tr Dµ Σ† Dµ Σ . (1b.3.12)
This is invariant under an Sp(2d) flavor symmetry, generated by acting on Σ on the right with
matrices of the form
~x ,
Tx = Ax + ~τ · S (1b.3.13)
where the A and S~ are d × d matrices, antisymmetric and symmetric, respectively. Like the gauge
transformation, (1b.3.13) preserves the special form of Σ.
Thus the largest possible flavor symmetry of a set of d pseudo-real representations of scalars is
Sp(2d). This apparently arcane fact will be important when we discuss the standard model with a
fundamental Higgs boson.
Problems
1b-1. In the theory described by (1b.1.1) modified by the replacement of ∂ µ with the covariant
derivative (1b.1.8), find the µ dependence of κab defined in (1b.1.14) in the one loop approximation.
This does not involve any of the unspecified interactions, because it arises from one loops diagrams
like (1a.2.2). It will depend on the representation matrices, Tfa .
/ 1 ψ1 + iψ 2D
iψ 1D / 2 ψ2
where
Djµ = ∂ µ − iGa Tja
with
1
T1a = τa
2
where τa are the Pauli matrices, but
1 1 1
T21 = τ2 , T22 = τ3 , T23 = τ1 .
2 2 2
This theory has an SU(2) gauge symmetry, and also an SU(2) global flavor symmetry, because ψ1
and ψ2 transform under equivalent representations of the gauge symmetry which we have chosen,
perversely, to write in different forms. Find the generators of the SU(2) flavor symmetry.
Chapter 2
2. it does not carry color and therefore does not participate directly in the strong interactions.
This second property is a defining characteristic of the class of spin 1/2 particles called “leptons”
comprising the electron e− and its neutrino νe ; the muon µ− and its neutrino νµ and the tau
t− and its neutrino ντ — and by their antiparticles. So far as we know today, the µ− and the
τ − seem to be heavier copies of the electron, distinguished only by their larger masses. As far
as we can tell, the weak and electromagnetic interactions act on the µ and τ exactly as they act
on the electron. Because they are heavier, they decay, by weak interactions. The electron, the
lightest charged particle, must be absolutely stable unless electromagnetic gauge invariance and
global charge conservation are violated.
The SU (2) × U (1) theory of the weak and electromagnetic interactions was first written down
as a model of leptons, simply because at the time the strong interactions the weak interactions of
the hadrons were not completely understood.
νeL , e−
L, e−
R (2.2.1)
νµL , µ−
L, µ−
R (2.2.2)
ντ L , τL− , τR− (2.2.3)
There is no compelling evidence for right-handed ν’s or left-handed ν’s, so we do not need a νR
field.
29
Weak Interactions — Howard Georgi — draft - February 10, 2009 — 30
The neutrinos are known to be very light. They may be massless, but there are some (still
confusing) indications that they may have small masses. If the neutrinos are massless, the weak
interactions conserve electron number, µ number and τ number separately. Formally, this means
that the theory has global symmetries:
We will come back in later chapters to the lepton number violating effects that can occur if the
neutrino masses are non-zero. For now, we will ignore neutrino masses. Then the symmetries,
(2.2.4), simplify the construction of the model because the different families do not mix with one
another.
Further, the interactions of the µ and τ are exact copies of those of the electron, so we can
discuss only the fields in the electron family (2.2.1).
The gauge group is SU (2) × U (1), which means that there are four vector fields, three of which
are associated with the SU (2) group that we will call Waµ , where a = 1, 2 or 3, and one Xµ
associated with the U (1).
The structure of the gauge theory is determined by the form of the covariant derivative:
Dµ = ∂ µ + igWaµ Ta + ig 0 X µ S (2.2.5)
where Ta and S are matrices acting on the fields, called the generators of SU (2) and U (1), respec-
tively. Notice that the coupling constants fall into two groups: there is one g for all of the SU (2)
couplings but a different one for the U (1) couplings. The SU (2) couplings must all be the same
(if the Ta ’s are normalized in the same way κ tr(Ta Tb ) = δab ) because they mix with one another
under global SU (2) rotations. But the U (1) coupling g 0 can be different because the generator S
never appears as a commutator of SU (2) generators. Even if we did start with equal g and g 0 ,
we would be unable to maintain the equality in the quantum theory in any natural way. The two
couplings are renormalized differently and so require different infinite redefinitions in each order of
perturbation theory. Since we need different counterterms, it would be rather silly to relate the
couplings.
To specify the gauge structure completely, we must define the action of Ta and S on the fermion
fields. Define the doublet
νeL
ψL ≡ (2.2.6)
e−
L
Then the Ta ’s are defined as
τa
Ta ψL =ψL , Ta e− R =0 (2.2.7)
2
where τa are the Pauli matrices. Both of these sets of T ’s satisfy the SU (2) commutation relation
although for the e−R field, which is called an SU (2) singlet, (2.2.8) is satisfied in a rather trivial way.
In order to incorporate QED into the theory we are building, we must make certain that some
linear combination of the generators is the electric-charge matrix Q. The matrix T3 is clearly related
Weak Interactions — Howard Georgi — draft - February 10, 2009 — 31
to the charge because the difference between the T3 values of each multiplet (the doublet ψL and,
trivially, the singlet e−
R ) us the same as the charge difference. Thus, we define
Q = T3 + S (2.2.9)
which defines S. We have done this very carefully so that S will be proportional to the unit matrix
on each multiplet.
1
QνeL = 0 ; Qe− −
L = −eL ; SψL = − ψL ; Qe− − −
R = SeR = −eR . (2.2.10)
2
We defined S this way so it satisfies the SU (2) × U (1) commutation relations
[Ta , S] = 0 . (2.2.11)
(2.2.5)-(2.2.11) completely define the gauge couplings to the leptons. To see just what these
gauge couplings do, we will look first at the interactions that change particle identity, the couplings
of W1µ and W2µ . If we write out the coupling of W1µ and W2µ to the fermions just by inserting the
standard forms of the Pauli matrices into (2.2.5)-(2.2.7), we find the interaction terms
gn o
− νeL (W / 2 )e−
/ 1 − iW L + e − /
L (W 1 + iW/ 2 )νeL (2.2.12)
2
plus an analogous term for muons. The “charged” fields defined by
W1 ∓ iW2
W±µ = √ (2.2.13)
2
create and annihilate charged intermediate vector bosons. W±µ annihilates (creates) W ± (W ∓ )
particles. (2.2.10) can give rise to µ decay as shown in Figure 2-1.
ν −
.. e
µ
..
..
.. ..... .........
. .
..... .....
...
....... ..........
. . . . .
µ− ..........................................................................................
W− ......
......
......
....
νe
Figure 2-1:
There are two things wrong with Figure 2-1 as a picture of the weak interactions. The µ− and
e− are massless, and the W ± are massless. The SU (2) × U (1) gauge symmetry does not allow a
lepton mass term e− −
L eR or a W
± mass term. The leptons must obviously get mass somehow for
the theory to be sensible. The W ± must also be very heavy in order for the theory to agree with
data. A massless W ± would give rise to a long-range weak force. In fact, the force has a very short
range.
Despite these shortcomings, we will press on and consider the neutral sector. If the theory is to
incorporate QED, one linear combination of the W3µ and X µ fields must be the photon field Aµ .
Thus, we write
Aµ = sin θ W3µ + cos θ X µ (2.2.14)
Then the orthogonal linear combination is another field called Z µ :
The two independent fields are orthogonal linear combinations of W3 and X because I have taken
the kinetic-energy terms for the gauge fields to be
1 1
− Waµν Waµν − X µν Xµν (2.2.16)
4 4
where
W µν = ∂ µ Waν − ∂ ν Waµ − gabc Wbµ Wcν
. (2.2.17)
X µν = ∂ µ X ν − ∂ ν X µ
Thus, only the orthogonal combinations of W3 and X have independent kinetic energy terms.
At this point, sin θ is an arbitrary parameter. But if we insert (2.2.14-2.2.15) into the covariant
derivative, we must obtain the photon coupling of QED. This determines the couplings of g and g 0
in terms of sin θ and e.
Schematically, the couplings of the neutral gauge particles are
/ 3 T3 + g 0XS
gW / (2.2.18)
Figure 2-2:
The SU (2) × U (1) gauge structure was first written down by Glashow in 1960. At the time, the
nature of the weak interactions was by no means obvious. That he got the right form at all was
a great achievement that had to wait over 12 years for experimental confirmation. Of course, he
did not know how to give mass to the W ± and Z without breaking the gauge symmetry explicitly.
The Z, like the W ± , must be very heavy. That much was known immediately because a massless
Z would give rise to a peculiar parity-violating long-range force (see Figure 2-2.).
Weak Interactions — Howard Georgi — draft - February 10, 2009 — 33
But all is not lost. If we somehow preserve the gauge-invariance structure and give mass to
the W ± and Z, we may be able to preserve renormalizability. This is what Weinberg and Salam
did by making use of spontaneous symmetry breaking. We will return to the general discussion of
why (and whether) the theory should be renormalizable when we discuss effective field theories in
Chapter 8.
1 m2 2 λ 4
L(φ) = ∂ µ φ∂µ φ − φ − φ (2.4.1)
2 2 4
With only a single field, there can be no continuous internal symmetry, but this L(φ) is invariant
under the reflection
φ → −φ (2.4.2)
This has allowed us to omit a φ3 (and φ) interaction term so that the theory has only two parameters,
m and λ. If λ is small, we can hope to treat the λφ4 term as a perturbation.
But suppose the sign of the mass term is changed so that the L(φ) becomes (discarding a
constant)
L(φ) = 21 ∂ µ φ∂µ φ − V (φ)
2 . (2.4.3)
λ m2
V (φ) = 4 φ2 − λ
In this case, we cannot perturb around the φ = 0 vacuum because the free theory contains a
tachyon, a particle with imaginary mass. Nor would we want to do so because the potential looks
like what we see in Figure 2-3. It is clear that the φ = 0 vacuum is unstable
p and that the theory
much prefers to spend its time near one of two degenerate minima, φ = ± m2 /λ. It doesn’t care
which.
... .
... ..
.
...
... ..
.
... .
... ..
... ...
...
... ..
... ...
... ............ .
...
..
. ....... ........... ..
.
... . ....
... .... ..
...
... .
.... .... .
.
... ... ... .
... .
.... ...
. ...
... ... ... .
... ... ... ..
.
. ... .
... . ....
.... ...... ..... ....
............. .............
− m2 /λ
p 0 p
m2 /λ
Figure 2-3:
It is easier to see what is going on if we rewrite the theory in terms of a field with zero VEV.
q
φ0 = φ − m2 /λ (2.4.4)
The result is
L(φ0 ) = 21 ∂ µ φ0 ∂µ φ0 − V (φ0 )
2
V (φ0 ) = λ
φ0 2 + 2 m2 /λ φ0 . (2.4.5)
p
4
= λ4 φ0 4 + m2 /λ φ0 3 + m2 φ0 2
p
Several points should stand out. The symmetry is hidden. It has been spontaneously broken by the
choice of vacuum. However, the theory is still described by only two parameters, for this relation
between the φ0 4 , φ0 3 and φ0 2 terms will be preserved by quantum renormalization effects. This
enhanced renormalizability, the fact that we need only two independent counterterms, is the legacy
of the spontaneously broken symmetry. The φ0 field describes a massive scalar field with mass
√
2m, with the original “imaginary” mass setting the scale, but not trivially related to the eventual
physical particle masses.
The existence of the other possible vacuum, φ = − m2 /λ or φ0 = −2 m2 /λ, does not show
p p
up in perturbation theory. It is infinitely far away because the field must be changed everywhere
in space-time to get there. Note that spontaneous symmetry breaking occurs only in p infinite
space-time. In a finite space, the ground state would be a linear combination of the φ = m2 /λ
vacua, invariant under the discrete symmetry. In the infinite space, such states are forbidden by
superselection rules. The existence of the other vacuum does give rise to a variety of interesting,
and sometimes even more important, nonperturbative effects in the infinite volume theory, but we
will ignore them for now and press on.
∂n
Vj1 ···jn (φ) = V (φ) (2.5.3)
∂φj1 . . . ∂φjn
Vj (λ) = 0 (2.5.4)
Weak Interactions — Howard Georgi — draft - February 10, 2009 — 36
The second derivative matrix Vjk (λ) is the meson mass-squared matrix. We can see this by
expanding V (φ) in a Taylor series in the shifted fields φ0 = φ − λ and noting that the mass term
is 12 Vjk (λ)φ0j φ0k . Thus, (2.5.5) assures us that there are no tachyons in the free theory about which
we are perturbing.
Now comes the interesting part, the behavior of the VEV λ under the transformations (2.5.2).
There are two cases. If
Ta λ = 0 (2.5.6)
for all a, the symmetry is not broken. This is certainly what happens if λ = 0. But (2.5.6) is
the more general statement that the vacuum doesn’t carry the charge Ta , so the charge cannot
disappear into the vacuum. But it is also possible that
Then the charge Ta can disappear into the vacuum even though the associated current is conserved.
This is spontaneous symmetry breaking.
Often there are some generators of the original symmetry that are spontaneously broken while
others are not. The set of generators satisfying (2.5.6) is closed under commutation (because
Ta λ = 0 and Tb λ ⇒ [Ta , Tb ] λ = 0) and generates the unbroken subgroup of the original symmetry
group.
Now let us return to the mass matrix. Because V is invariant under (2.5.2), we can write
Setting φ = λ in (2.5.9), we find that the second term drops out because of (2.5.4), and we obtain
region and then go smoothly back to our vacuum outside. The point is that the energy of such a
state can be made arbitrarily close to the energy of our vacuum state by making the region larger
and the transition smoother. But if there are states in the theory with energy arbitrarily close
to the energy of the vacuum state, then there must be massless particles in the theory. These
are Goldstone bosons. Their masslessness is the translation into local field theory of the global
degeneracy of the vacuum.
δψ = ia Ta ψ (2.6.2)
However, as we have seen, the kinetic-energy term for massless fermions is automatically invariant
under the larger group of symmetries, SU (2) × SU (2):
δψL = iaL Ta ψL
. (2.6.3)
δψR = iaR Ta ψR
These are the chiral symmetries. (2.6.3) can be rewritten in terms of the infinitesimal parameters
The invariance of the Yukawa couplings may be more transparent in terms of finite transformations.
The infinitesimal transformations (2.6.3) and (2.6.6) can be integrated to obtain
Σ → LΣR† (2.6.9)
where L and R are independent 2 × 2 unitary matrices with determinant 1,
Σ = σ + iτa πa (2.6.11)
It is not obvious (at least to me) that this form is preserved by the transformations (2.6.6).1 But
it is true, and you can work out by explicit calculation the transformations of the σ and πa fields:
δσ = a5 πa
. (2.6.12)
δπa = −abc b π c − a5 σ
† † 2 2 2 2
Another
√ way to see this is to note that Σ Σ = ΣΣ = σ + ~π I and det Σ = σ + ~π . Thus
2 2
Σ is σ + ~π times a unitary unimodular matrix. Obviously, if we multiply Σ on the left or on
√
the right by a unitary unimodular matrix, the result is still of the form σ + ~π 2 times a unitary
2
unimodular matrix.
Inserting (2.6.11) into L, we can write the Yukawa couplings as
From (2.6.13), you can see that the πa fields have the right form to describe the π’s. The coupling
g is the πN N coupling gπN N .
We still don’t have a nucleon mass term, but it is clear from (2.6.13) that if we can give σ
a VEV, we will be in good shape. To this end, we must ask how to build L(Σ) invariant under
SU (2) × SU (2). It is clear from (2.6.9) that
1 †
tr Σ Σ = σ 2 + πa2 (2.6.14)
2
is invariant. In fact, the most general invariant (without derivatives) is just a function of σ 2 + πa2 .
This can be seen by noting that more complicated traces just give powers of σ 2 + πa2 because
Σ† Σ = σ 2 + πa2 (2.6.15)
and is proportional to the identity in the 2 × 2 space. Alternatively, we can recognize (2.6.12) as
the transformation law of a 4-vector in four-dimensional Euclidean space. The invariant (2.6.14) is
just the length of the vector, the only independent variant.
1
These objects are actually related to interesting things called quaternions.
Weak Interactions — Howard Georgi — draft - February 10, 2009 — 39
Now we can use our freedom to make SU (2) × SU (2) transformations to rotate any VEV into the
σ direction so that without any loss of generality we can assume
σ 0 = σ − Fπ , hσ 0 i = 0 (2.6.20)
in terms of which L is
L = iψ ∂/ψ − gFπ ψψ − gσ 0 ψψ
This describes nucleons with mass gFπ coupled to the scalar σ 0 field and massless pseudoscalar πa ’s.
Why did Gell-Mann and Levy think (2.6.21) had anything to do with the world? For one thing,
the physical pion is very light compared to other hadrons. For example, m2π /m2N ' 1/50. Perhaps
a theory in which it is massless is not such a bad approximation. But there was another reason.
The parameter g can clearly be measured in π-nucleon interactions. It turns out that Fπ can also
be measured. The reason is that, as we shall discuss in enormous detail later, it determines the
rate at which π ± decay through the weak interactions. The crucial fact is that the axial vector
current, the current associated with a5 transformations, has the form
µ
j5a = − (∂ µ πa ) σ + (∂ µ σ) πa − ψγ µ γ5 τa ψ (2.6.22)
The other terms are all bilinear in the fields. The point is that this current has a nonzero matrix
element between the vacuum and a one-pion state
µ
h0|j5a |πb i = iFπ pµ δab (2.6.24)
Weak Interactions — Howard Georgi — draft - February 10, 2009 — 40
where pµ is the pion momentum. This is odd. A normal current, like the charge with which it is
associated, just moves you around within multiplets. (2.6.24) is a sign of spontaneous symmetry
breaking. At any rate, the decay π + → µ+ νµ is proportional to Fπ2 , and so Fπ can be measured.
Then the nucleon mass can be predicted according to
mN = gπN N Fπ (2.6.25)
This relation is called the Goldberger-Treiman relation, and it works fairly well (actually (2.6.25)
us a special case of the general Goldberger-Treiman relation that works even better).
The pion is a Goldstone boson in this model. This is probably obvious, but we can use the
formal machinery of Section 2.5 to see it directly. The Goldstone-boson directions are defined by
Ta λ where λ is the VEV. In this theory, the isospin generators annihilate the vacuum, so isospin is
not spontaneously broken and there is no scalar Goldstone boson. But the chiral transformations
rotate the VEV into the π directions (σπa = · · ·+a5 Fπ from (2.6.12)), and thus the chiral symmetry
is broken and the π’s are Goldstone bosons.
We will discuss explicit chiral symmetry breaking later, but now notice that we can incorporate
a pion mass by adding to V the term
m2π 2
σ + π 2 − 2Fπ σ + Fπ2 (2.6.26)
2
which is not invariant because of the linear term m2π Fπ σ.
− f e− †
R φ ψL + h.c. (2.7.1)
~τ 1
T~ φ = φ, Sφ = φ , (2.7.2)
2 2
then (2.2.6) and (2.2.10) with (2.7.2) imply that (2.7.1) is invariant under the SU (2) transformation
δ = ia T a + iS. Evidently, φ must be a doublet field
+
φ
=φ (2.7.3)
φ0
We can do everything in this notation, but it is slightly more convenient to go over to a notation
in which the scalar fields are self-adjoint. We can easily do this by rewriting the complex φ+ and
φ0 fields in terms of their real and imaginary parts:
+ √
φ (φ3 + iφ4 )/√2
= (2.7.7)
φ0 (φ1 + iφ2 )/ 2
√
The conventional 2 is intended to ensure that the fields are normalized in the same way. At
any rate, we can arrange the real fields φj in a real 4-vector and ask how the generators T~ and S
translate into this notation. Thus, we write
φ3
φ4
Φ=
φ1 (2.7.8)
φ2
The space of the Φ field has an obvious tensor-product structure. It is a tensor product of the two-
dimensional space of the original φ, on which the τ ’s act, and a two-dimensional space corresponding
to the real and imaginary parts of the components of φ, on which we can define an independent set
of Pauli matrices ~σ . The 15 traceless Hermitian matrices acting on φ can be written as
In this notation, it is easy to write down the generators T~ and S. The procedure is simple and quite
general: Leave antisymmetric matrices in φ space unchanged, but multiply symmetric matrices by
−σ2 to make them antisymmetric. Thus
T1 Φ = − 21 τ1 σ2 Φ
T2 Φ = 21 τ2 Φ
(2.7.10)
T3 Φ = − 21 τ3 σ2 Φ
SΦ = − 21 σ2 Φ
You can easily check that (2.7.10) gives the same transformation laws as (2.7.2) and (2.7.7). The
advantage of this notation is really marginal. For example, it makes it slightly easier to identify
the Goldstone bosons. But since it is so easy, we might as well use it.
In terms of Φ, the kinetic-energy term becomes
LKE (Φ) = 21 Dµ ΦT Dµ Φ
h i
= 1
2 ∂ µ ΦT − iΦT e
sin θ T~ · W
~µ+ e
cos θ SXµ (2.7.11)
h i
∂µ + i e
sin θ T~ · W
~µ+ e
cos θ SXµ Φ
Weak Interactions — Howard Georgi — draft - February 10, 2009 — 42
The point of all this is that (2.7.11) contains terms like ΦT W1µ W1µ Φ. If Φ has a VEV, this looks
like a W1 mass. Perhaps we will be able to give mass to the W ± and Z. Of course, we hope to do
it without giving mass to the photon, and we still have the Goldstone bosons to worry about. But
let’s go on.
The most general SU (2) × U (1) invariant potential depends only on the combination φ† φ. In
particular, if it has the form
λ † 2
V (φ) = φ φ − v 2 /2 (2.7.12)
2
Then φ will develop a VEV.
We want the VEV of φ0 to be real and positive to that (2.7.1) gives an electron mass term
that is real and positive. This is purely conventional, of course. We can make an SU (2) × U (1)
transformation to make any VEV of φ+ and φ0 have the form
√
hφ+ i = 0, hφ0 i = v/ 2 (2.7.13)
φ → exp(2iθ3 T3 )θ
we get
a → eiθ3 a, b → e−1θ3 b (2.7.15)
By a suitable choice of θ3 we can make a and b have the same phase. Then
φ → exp(2iφ2 T2 )φ (2.7.16)
is an orthogonal transformation by which we can rotate the (a, b) vector into the φ0 direction.
Then by another (2.7.15) transformation we can make the VEV real and positive. Thus, all we
have done is to choose a convenient form for the VEV so that our original labeling of the fermion
fields is consistent.
In any case, (2.7.13) is equivalent to
0
0
hΦi =
v (2.7.17)
0
Now we want to rewrite (2.7.13) in terms of the VEV, λ, and the shifted scalar field:
Φ0 = Φ − λ (2.7.18)
Of course, LKE (Φ) contains the kinetic energy term for the Φ0 fields. But for the moment we will
concentrate on the terms that are quadratic in the fields. There are two types:
1 T e ~ ~µ e e ~ e
λ T ·W + SX µ · T · Wµ + SXµ λ (2.7.19)
2 sin θ cos θ sin θ cos θ
and
e ~ ~ e
µ 0T
i∂ Φ T · Wµ + SXµ λ (2.7.20)
sin θ cos θ
Weak Interactions — Howard Georgi — draft - February 10, 2009 — 43
The (2.7.19) are the W and Z mass terms we want, as we will soon see. But (2.7.20) looks
dangerous. It describes some sort of mixing between the Goldstone bosons and the gauge fields.
The Goldstone-boson fields are
Φ0 T T~ λ = ΦT T~ λ (2.7.21)
The field Φ0 T Sλ is not independent, since T3 λ = −Sλ.
But we have not yet used all our freedom to make local SU (2) × U (1) transformations. In
fact, we can choose a gauge, called the unitary gauge, in which the Goldstone-boson fields just
vanish, so we can throw (2.7.20) away. To see this, we need to return to the unbroken theory. We
saw in (2.7.14-17) that we use the global SU (2) × U (1) symmetry to take an arbitrary VEV and
rotate it into the φ3 direction. But since we have the freedom to make different SU (2) × U (1)
transformations at each point in space-time, we can take an arbitrary field Φ(x) and rotate it into
the φ1 direction! So that after these rotations
Of course, we now have no further freedom to rotate Φ without disturbing (2.7.22). In other
words, we have chosen a gauge. Comparing (2.7.21) and (2.7.22), we can see explicitly that the
Goldstone-boson fields vanish in this gauge. The VEV is a VEV of the single remaining field φ3 .
Having disposed of (2.7.20), we can now return to (2.7.19) and show that it is a mass term for
the W ± and the Z. Note first that we can separate the problem into a neutral-gauge boson mass
matrix and a separate-charged boson matrix. Electromagnetic charge conservation prevents mixing
between the two. The neutral sector is more complicated than the charged sector, so we will do it
first. We saw in Section 2.2 that we could rewrite the covariant derivative in terms of Aµ and Z
fields as follows:
e e e
T3 W3µ + SX µ = eQAµ + T3 − sin2 θ Q Z µ (2.7.23)
sin θ cos θ sin θ cos θ
Now we can see why the photon field doesn’t get a mass — because
Qλ = 0 (2.7.24)
so that Aµ doesn’t appear in (2.7.19) and (2.7.20). This, of course, is the way it had to work
out. (2.7.24) is the statement that the electromagnetic gauge invariance is not broken by the
vacuum. Thus, it must imply that the photon remains massless, and it does. (2.7.24) also shows
that the sin2 θQ term in the Z µ coupling is irrelevant to the mass. Putting (2.7.23)-(2.7.24) into
(2.7.19)-(2.7.20), we get
2
1 e2 1 ev
2 2
Z µ Zµ λT T32 λ = Z µ Zµ (2.7.25)
2 sin θ cos θ 2 2 sin θ cos θ
Thus
ev
MZ = (2.7.26)
2 sin θ cos θ
An even simpler analysis for the charged fields gives
2
1 ev
(W1µ W1µ + W2µ W2µ ) (2.7.27)
2 2 sin θ
Thus
ev
MW = = MZ cos θ (2.7.28)
2 sin θ
Other documents randomly have
different content
Sie wurde nachdenklich – vergaß die begonnene Arbeit, riß den
Hut vom Haken und drückte ihn auf das Haar. – Wenn sie hier fort
wollte, mußte ein neuer Unterschlupf gefunden werden. – Und fort
wollte sie. Je früher, desto besser. – Im Laufschritt eilte sie die
breite, stille Straße hinunter. – Wollte zu der Zweigniederlassung der
von der Präsidentin bisher gelesenen Zeitung, um ein Gesuch nach
einer Stellung aufzugeben – vergaß dann aber sofort wieder diesen
Vorsatz und eilte gedankenlos weiter, den wundervollen, schattigen
Plätzen entgegen, an denen die prunkvollen Häuser der glücklichen
Besitzer lagen.
Die Welt war klar, satt und durstlos. An stillen Seitenstraßen
schienen die jungen Buchen zu bluten, als verschenkten sie freudig
ihren Lebenssaft. Unbeschreibliche Sehnsucht nach einem
Menschen, der sie in dieser Stunde haltloser Verzweiflung voll
verstehen könnte, überkam Eva von Ostried. Sie wußte sich
Niemand!
Ihre Schönheit hatte zu allen Zeiten glühende Bewunderer
gefunden. Aber sie kannte sich selbst noch zu wenig, um schon zu
wissen, daß sich lediglich ihre stark entwickelte Eitelkeit durch die
unverhüllten Blicke der Leidenschaft befriedigt gefühlt.
Wäre es anders gewesen, hätte sie damals unmöglich Paul
Karlsens gestohlene Zärtlichkeit als eine unerhörte Beleidigung
empfinden können. Ihr Herz war bisher völlig unberührt geblieben.
Ihre Frauensehnsucht suchte indessen unbewußt – an den lauten
Huldigungen vorbei – nach den stillen Gassen, die zu dem Tempel
reiner Liebe führen.
Und dennoch sträubte sie sich heftig gegen die Zumutung, die
Krone des Frauendaseins einzig in der Ehe mit einem Manne zu
suchen.
Plötzlich verlangsamten sich ihre Schritte. Lauschend neigte sich
der Kopf. Rächten sich die Stunden der Aufregung und gaukelten ihr
Töne vor aus jener Welt, die ihr von heute an verschlossen war.
Oder gehörte die jauchzende Stimme hinter ihrem Rücken der
Wirklichkeit an?
P
aul Karlsen ging mit gemächlichen Schritten über den
rostfarbenen Kies. Zu beiden Seiten des schmalen Weges
blühte der Vorgarten. Ueber dem weinumzogenen Haus lag die
Mittagssonne. Augenscheinlich hatte er es nicht eilig. Auch die
wenigen bequemen Marmorstufen der Treppe nahm er fast zögernd.
In dem Vorraum, der zur eigentlichen Diele führte, erwartete ihn die
steife Gestalt eines alten Dieners, der etwas eigentümlich Lebloses
hatte. Paul Karlsens Augen waren noch von der Fülle der Sonne
geblendet. Er erschrack, als sich eine Hand nach seinem Hut
ausstreckte, trotzdem er dies Bild nun doch nachgerade kennen
mußte.
„Na – bin ich heute pünktlich, alter Hagen,“ fragte er lässig.
Das Gesicht veränderte sich nicht. Nur die leise Stimme klang
vorwurfsvoll.
„Die gnädige Frau wartet seit einer Stunde mit dem Essen!“
Er lachte kurz auf, warf den Kopf in den Nacken und murmelte
etwas.
„Verdammter Zwang,“ hieß es. –
In dem großen, sehr kühlen Eßzimmer harrten auf köstlichem
Leinen zwei Gedecke. – Dieser Raum wirkte pomphaft und
erdrückend. Die Bespannung der Wände mit schwarzem Rupfen allzu
feierlich. Die wuchtigen Möbel spreizten sich in ihrer Kostbarkeit. Die
Sonne, welche durch stilvoll bemalte Scheiben ohnehin ihren Weg
niemals finden konnte, war vollends von schweren Vorhängen
abgesperrt. Nur die Tafel mit dem blendend weißen Leinen trug eine
Fülle blutroter Rosen und dunkelblauem Kristall.
Plötzlich löste sich aus der halbdunklen Schwermut die
überschlanke Gestalt einer weißgekleideten Frau und schritt auf Paul
Karlsen zu. Das längliche Gesicht war auffallend bleich. Die Nase trat
scharf hervor, als habe ein kürzlich überstandenes Krankenlager den
Wangen die natürliche Rundung genommen.
Karlsen führte ihre Hand an die Lippen und ließ den Wortlaut
seiner Stimme in gut gespielter Ueberraschung klingen:
„Du hast ja diese Leichenkammer heute so herrlich geschmückt,
kleine Frau. Wer soll denn beigesetzt werden? Und ein neues
Gewand hast du ebenfalls angelegt.“
Ihr stiegen die Tränen auf. Nicht weil er sie warten ließ. O nein –
daran hatte sie sich längst gewöhnt. Aber – daß er nicht – daran
dachte.
„Das Kleid,“ sagte sie hastig, um nicht laut aufweinen zu müssen,
„kennst du es wirklich nicht, Paul?“
Er zog sie nach einem der hohen Fenster herüber und zerrte den
Vorhang zurück. In dieser Bewegung lag ein Aufbäumen auch gegen
vieles andere.
„Nee, mein Kind. Keine Ahnung habe ich.“
„Ich trug es an dem Tage unserer heimlichen Verlobung in
Oeynhausen.“
Er lachte verlegen auf.
„Richtig! – Natürlich! – Jetzt sehe ich es. Das sind aber doch
höchstens vier Monate her und noch längst kein Jahr. Wo ist also der
geschätzte Anlaß zu einer besonderen Feier?“
„Heute sind wir einen Monat Mann und Frau,“ sagte sie leise und
konnte nun doch nicht hindern, daß ein runder Tropfen auf das
kostbare Gewand fiel. – Er zog ungeduldig die Stirn empor.
„Schön – also einen Monat! Was ist das im Vergleich zu all den
Jahren, die hoffentlich noch vor uns liegen. – Also, ich habe dieses
hohe Fest verschwitzt. Nimm’s nicht übel. Mir brummt der Kopf. Es
gibt doch mehr Arbeit und Schwierigkeiten zu überwinden, als ich
anfänglich annahm.“
„Ich störe dich doch nicht etwa bei deinen Studien, Paulchen?“
Er hatte seinen Rufnamen überhaupt niemals gemocht. Dies
„Paulchen“, das er ihr nicht abgewöhnen konnte, reizte ihn zuweilen
bis zur Tollheit. Jetzt überhörte er es, weil er etwas erreichen wollte.
„Du im Besonderen bist das bescheidenste und leiseste Wesen,
das es geben kann. Im allgemeinen freilich wäre ich gerade jetzt für
eine kurze Zeit nicht eben ungern solo.“ Sie sah entsetzt zu ihm auf.
„Soll das heißen!“ Sie konnte nicht vollenden. Ihre Stimme
erstickte in Tränen. Er schüttelte sich, als fröre er.
„Tu mir den einzigen Gefallen und höre mit dem Weinen auf,
Elfriede. Ich komme mir ja andauernd wie ein Barbar vor. Nein, nicht
du sollst für wenige Tage deine zur Zeit kränkelnde Mutter, eine
Straße weiter, besuchen und sie dadurch halb unsinnig vor Freude
machen – welchen Wunsch sie mir schon vor einer Woche, allerdings
mit der Bitte, ihn dir vorläufig zu verheimlichen, verraten hat –
sondern ich werde zu meinem Lehrer unter den blendenden
Dachgarten ziehen. Denn, weißt du, kleine Frau, ich muß üben und
immer nur üben – kann mich nicht mehr an eine feste Tischzeit
binden – vertrage überhaupt zu solchen Zeiten vorübergehend keine
andere Gesellschaft als eine männliche.“
Sie legte die Hand auf seinen Arm.
„Paulchen, schenk mirs zum heutigen Tag, daß ich in mein altes
Mädchenstübchen zur Mutter darf. Du mußt deine Bequemlichkeit
gerade jetzt haben.“
„Das würde eine schöne Geschichte geben, mein liebes Kind!
Deine Mutter würde plötzlich vergessen, wie sehr sie sich nach dir
gesehnt und felsenfest glauben, ich behandele dich schlecht und
lieblos. Denn sieh mal, immerhin bleibt es etwas wunderbar, wenn
eine junge, liebliche Frau nach einmonatlicher Ehe ihren Ehemann –
wenn auch nur vorübergehend – verläßt.“ Der letzte Satz gab ihr
eine ungeheure Kraft.
„Glaubst du wirklich, Paulchen, daß ich der Mutter meinen
Besuch in diesem Lichte hinstellen würde?“
„Na, na, Kleines – wer kennt sich mit euch Frauen aus? In
gewissem Sinne ähnelt ihr euch alle verteufelt.“
Sie widersprach mit jähaufflackerndem Rot.
„Hast du schon vergessen, was ich dir in der grünen Einsamkeit
des Siels am Karpfenteich gelobt habe?“ Natürlich hatte er nicht die
geringste Ahnung. Aber er hütete sich es einzugestehen.
„Frauengelöbnisse sind unberechenbar, wie eure Eifersucht,
Schatz.“
„Hältst du mich für eifersüchtig?“
„Es käme auf die Probe an. Glatt verneinen möchte ich das
nicht!“
„Ich würde sie bestehen. Verlaß dich drauf.“
„Lieber nicht. Deine Mutter wohnt ein bißchen zu nahe, Kleines.“
„Wie tief mußt du mich einschätzen, Paul!“
„Bewahre. Riesig hoch sogar. Hätte ich dich denn sonst
geehelicht?“
Sie legte mit einer rührenden Gebärde der Demut ihr Gesicht auf
seine schlanke Hand.
„Sage so etwas niemals wieder, Paulchen. Wir wollen uns doch
fest, ganz fest vertrauen.“ Ihm wollte ein Lachen aufsteigen. Es
wurde aber zuletzt ein Hüsteln daraus.
„Wollen wir auch. Natürlich. Aber jetzt komm gefälligst. Ich
verspüre einen Bärenhunger.“ Erschrocken drängte sie ihn zur Tafel
hinüber.
„Verzeih – ich vergesse das so oft neben dir!“
Er musterte ihre magere, noch kindlich unentwickelte Gestalt und
seufzte leicht auf.
„Leider, mein guter Schatz! Eß und trink, lieb und sing. Ja – so
stand es an einem alten Bauernhaus in Sachsen. Und recht hat der
Spruch! – Wie ich sehe, hast du zur Feier des hohen Tages auch
herrlich für Stoff gesorgt. Hoffentlich ist er gut.“
Sie ließ es sich nicht nehmen, ihm aus der schweren
Kristallkaraffe die funkelnde Schale zu füllen.
„Probiere ihn, Paulchen.“ Er hob das kostbare Glas und ließ es
hell an das ihre klingen.
„Herrlich! – Ueberhaupt – das muß ich immer wieder
anerkennen, du bist eine ganz prachtvolle, kleine Hausfrau.“
Strahlend sah sie zu ihm auf.
„Darum habe ich auch einen Wunsch frei, ja?“
Der Diener trug die Suppe auf. Die Unterhaltung verstummte.
Sobald er unhörbar entschwunden war, sagte Paul Karlsen spöttisch:
„Er liebt mich nicht, Elfchen. Weißt du das eigentlich?“
„Er liebt jeden, der mir gut ist,“ sagte sie ruhig, fast streng.
„So? Na, weißt du, das bezweifle ich stark. Oder willst du etwa
andeuten, daß ich –“
Sie ließ ihn nicht zu Ende kommen. Sanft legte sie ihre Hand auf
seinen Mund.
„Ich bin dir unaussprechlich dankbar dafür. Trotzdem wünsche ich
mir noch eine Kleinigkeit.“
„Was denn, Kleines?“
„Den Besuch bei meiner Mutter.“
„Ausgeschlossen! Die Gründe für meine Härte habe ich dir
genannt.“
„Sie sind sämtlich hinfällig. Ich fange es eben so geschickt an,
daß Mama zum Schluß sich heimlich bei dir bedanken wird.“
„Wie wolltest du das anstellen?“
„Sehr einfach. Heute nachmittag zur üblichen Whistpartie, wäre
ich doch herübergegangen. Da werde ich also ausnehmend blaß
aussehen müssen. – Lache nicht – ein wenig Weiß genügt schon. Sie
wird mich wieder zur Schonung quälen, in ihrer Ueberängstlichkeit
meinen längeren Besuch verlangen, damit sie sich selbst von
meinem Gesundheitszustand überzeugen kann und zwar dies alles in
deiner Gegenwart.“
„Um Gottes willen, ich soll dich doch nicht etwa begleiten. Das
hast du bisher doch klug zu vermeiden gewußt.“
„Bringe mir dies Opfer, Liebster.“
„Also gut! Ich will sogar gern mitkommen. Das heißt höchstens
für ein bis zwei Stunden.“
„Solange wird es gar nicht nötig sein,“ meinte sie froh. „Aber nun
höre weiter. Du sperrst dich gegen das von ihr Geforderte und
verweigerst schließlich in aller Form deine Erlaubnis. – Dann wird sie
hitzig werden und unter allen Umständen darauf bestehen. – Ich
kenne sie doch.“
„Du bist ja eine ganz gefährliche, kleine Heuchlerin, Schatz.“
Er zog sie leicht in die Arme. In tiefem Glücksgefühl schloß sie
die Augen, die das einzig Schöne in ihrem Gesicht waren.
„Ist das nicht ein feiner Plan, Paulchen?“
„Ausgezeichnet sogar, wenn mir inzwischen die Sache nicht
wieder leid geworden wäre. Du hast als Ernst aufgefaßt, was bei mir
nur eine Art Gefühlsausbruch war.“
„Daß du es, wenn auch nur einen Augenblick gewünscht hast,
zeigt mir die Notwendigkeit und nachher – wird es um so schöner
sein.“
„Gelt, das hätten wir vor einem Vierteljahr auch noch nicht
gedacht?“
„Was denn,“ schnurrte er mit erwachender Behaglichkeit.
„Daß wir so schnell unser Glück erzwingen würden.“
Er nickte mit vollem Mund, denn inzwischen war der Braten
gekommen, der, zart und saftig, selbst den größten Feinschmecker
befriedigt hätte.
„Wärst du nicht plötzlich nach der schroffen Ablehnung meines
Werbens durch die Frau Kommerzienrat, wollte natürlich sagen,
deiner lieben Mama, kränker geworden und dadurch jegliche
Wirkung der Kur auf dein rebellisches Herzlein in Frage gestellt – wer
weiß, wer dann heute an meiner Stelle neben dir säße –“
„Wie wenig du mich im Grunde doch kennst, Paulchen. Fühlst du
nicht, daß ich niemals einem andern als dir gehört hätte?“
Er nickte ihr zu.
„Kleines Treues – du!“ Dann begann er zu scherzen und von
jener Zeit zu plaudern, weil er genau wußte, daß ihr dies die liebste
Unterhaltung war. Seine feurigen Augen strahlten tief in die ihren.
Das schmeichlerische weiche Organ machte auch das
unbedeutendste Wort zu einer Zärtlichkeit. Seine Laune war plötzlich
glänzend.
Ueber den blutroten Rosen und dem blauen Kristall schien die
Krone des Glückes, die allein die Liebe gibt, in warmen Glanz zu
schweben! – –
„Ja,“ sagte einige Stunden später Frau Kommerzienrat Eßling zu
ihrer alten Freundin und Vertrauten, die – wie seit Jahren – als Erste
zur Whistpartie gekommen war, „in der Nähe hätte ich sie nun ja.
Aber, was will das sagen. So viel man auch aufpaßt – allwissend ist
doch Niemand. Wer sagt mir, ob Elfriede unter seiner Anleitung nicht
ebenfalls Komödie zu spielen gelernt hat?“
Frau Generalkonsul Enck war keine mißtrauische Natur. Aber
dieser überstürzt geschlossenen Verbindung zwischen dem
überzarten, beständig kränkelnden Mädchen und diesem
bildhübschen Leichtfuß, dem Karlsen, brachte sie doch ihre schärfste
Mißbilligung entgegen. Hätte man sie, wie das sonst bei jeder
wichtigen Entscheidung der Fall gewesen, nur um Rat gefragt. Man
hatte jedoch, einfach über ihren Kopf fort, in aller Stille dem
durchaus nicht von ihr ernstgenommenen Verlöbnis, die eheliche
Verbindung auf dem Fuße folgen lassen.
Nun kamen natürlich Reue und Gewissensbisse über die
besorgte, selbst leidende Mutter. Anderseits kannte sie die
bewundernswerte Energie der Kommerzienrätin zu genau, um dieses
Bündnis von vornherein als dauerndes anzusehen.
„Sie hätten es sich gründlicher überlegen sollen,“ konnte sie sich
nicht versagen, zu erwidern. Die andere sah starr auf das feine
Porzellan der kostbaren Teeschalen herab.
„Sie haben niemals Kinder besessen. Da können Sie so etwas
wohl sagen. Stehen Sie nur an zwei Krankenbetten, in denen
scheinbar bisher kerngesunde, bildhübsche, lebenslustige Mädchen –
– Auch die andern Aerzte haben zuerst keine Ahnung davon gehabt.
Denn daß mein Mann an den Folgen einer hartnäckigen
Lungenentzündung in jungen Jahren starb, gab noch allein keinen
Grund zur Beängstigung für seine Kinder ab. Erleben Sie mal erst,
was ich ertragen habe. – Wie habe ich damals gegen das furchtbare
Gespenst gerungen. Hart bin ich gewesen – so hart.“
In ihrem energischen Gesicht, aus dem die scharfe Nase, wie sie
auch ihre jetzt noch einzige Tochter hatte, auffallend hervorsprang,
zuckte es.
„Regen Sie sich nicht mit den alten Geschichten auf, Frau Eßling.“
„Die Aussprache mit Ihnen tut mir wohl. Zu wem sollte ich wohl
davon reden, wenn nicht zu Ihnen, vor der ich kein Geheimnis habe.
– Seitdem ich meinen alten Franz, den Diener, meiner Elfriede
gegeben habe, weiß niemand im Haus um diese Sachen.“
„Malen Sie sich nicht zu schwarz, Beste,“ verteidigte die Konsulin.
„Sie mögen damals streng gewesen sein. Wer wäre es in der
gleichen Lage nicht gewesen. An eine Härte glaube ich nicht.“
„Sie sollen selbst urteilen. In St. Blasien war’s, wohin ich nach
den erfolglosen Kuren in Hohenhonnef und Davos aus eigenem
Entschluß noch mal mit den beiden ältesten Töchtern ging. Denn Sie
wissen, ich konnte und wollte nicht daran glauben, daß alles
vergeblich sein sollte. In der Liegehalle war ein vergnügliches Leben
unter dem jungen Volke, und keines war da, das an ein frühzeitiges
Sterben gedacht hätte. Als Gesunder läßt man die sonst im Verkehr
der verschiedenen Geschlechter streng beobachteten Richtlinien
außer Acht, weil die armen totgeweihten Geschöpfe doch keine
Vollmenschen mehr sind. Nicht wahr, wenn unsereins so ein
schmalschultriges Kerlchen mit fieberroten Flecken auf den
herausstehenden Backenknochen sieht, dann fragt man nicht erst
lange danach, was er sonst ist, hat und will, selbst wenn er
augenscheinliches Wohlgefallen an dem eigenen Fleisch und Blut
zeigt. Im Gegenteil, man freut sich noch gar darüber, und kommt
sich wer weiß wie großmütig und gar edel vor, weil man die leibliche
Mutter von seinem Glückserreger ist. Darum bin ich auch nicht einen
Augenblick besorgt gewesen, als der junge Bildhauer meiner kranken
Aeltesten über alle Gebühr hinaus den Hof machte. Erst, als der ihn
behandelnde Arzt, dem ich mein Bedauern über diesen
hoffnungslosen Fall aussprach, mir rund heraus und lachend erklärte,
er wäre froh, wenn jeder seiner Kranken so gesund wäre, wie dieser
Künstler, der sicher im nächsten Jahr wieder völlig obenauf sein
würde, wurde ich nachdenklich, vorsichtig und streng. – Mein Mädel
nahm ich ins Gebet. Den Bildhauer behandelte ich so schlecht, wie
es nur irgend ging. – Es war für alles zu spät. – Eines Tages erklärte
mir meine Tochter, daß sie sich mit dem Jüngling von Habenichts
verlobt habe. Sie hat vor mir auf den Knien gelegen und mich um
meine Einwilligung angefleht. Ich blieb hart. Daß der offensichtlich
seinem Aussehen nach Totgeweihte lediglich an den Folgen einer
schweren Rippenfellentzündung schonungsbedürftig sei, hatte meine
Hoffnung bezüglich der eigenen Kinder wunderbar gekräftigt. –
Einen Tag nach dem vergeblichen Flehen meiner Aeltesten reisten
wir, die noch nicht zur Hälfte vollendete Kur abbrechend, nach
Hause. Briefe kamen, wurden von mir abgefangen und prompt
vernichtet. Jede Nacht hörte ich das bitterliche Schluchzen meiner
Aeltesten – merkte, wie sie bleicher und hinfälliger wurde und
glaubte plötzlich doch nicht mehr an den Ernst des Verhängnisses.
Es war so nahe. Meine kleine Elfriede, die wenigst anmutigste der
Drei, hatte ich indessen aufs Land in Pension gegeben, weil der Arzt
von der Möglichkeit einer Ansteckung, selbst bei größester Vorsicht,
gesprochen. Nun konnte ich ganz der Pflege und Sorge für die
beiden andern leben. – Einmal hat der Bildhauer gewagt, bis in mein
Haus vorzudringen. Ich habe ihn auch empfangen. – Seitdem hat er
keine Zeile mehr geschrieben. Denn ich war deutlich gewesen. – Vier
Wochen nachher hat meine Tochter, unterstützt von ihrer Schwester,
noch einen letzten Sturm auf mein Mutterherz gemacht. Weiß Gott,
es hat sich in dieser Stunde nicht geregt. Ich habe es als Laune und
Eigensinn empfunden, was doch mehr gewesen ist.“
Die Andere legte begütigend die Hand auf die zuckende Schulter
der Kommerzienrätin.
„Wir wissen alle, was Sie die langen Jahre für eine aufopfernde,
prachtvolle Mutter gewesen sind.“
„So prachtvoll, daß ich mich hinterher noch meines gefestigten
Charakters gefreut und ein paar Tage ernsthaft mit dem armen Kind
geschmollt habe. Auch meine Zweite hat begonnen für sie und den
Bildhauer unentwegt zu betteln. – Als sie einsah, daß ich nicht
nachgab, verstummte sie zwar, aber es war seltsam, auch mit ihr
wurde es seitdem schlechter. Sie schienen sich beide in das
Unabänderliche meines Willens gefügt zu haben, bis zu jenem
schrecklichen Augenblick, an dem mich die Pflegerin in der Nacht
rief. Da hat meine Aelteste, die stets ein sanftes, scheues Ding war,
mir gesagt, wie unerträglich ihr Dasein ohne den Geliebten gewesen
und wie wenig sie sich freue, daß es nun endlich aufhören dürfe. –
Welcome to Our Bookstore - The Ultimate Destination for Book Lovers
Are you passionate about books and eager to explore new worlds of
knowledge? At our website, we offer a vast collection of books that
cater to every interest and age group. From classic literature to
specialized publications, self-help books, and children’s stories, we
have it all! Each book is a gateway to new adventures, helping you
expand your knowledge and nourish your soul
Experience Convenient and Enjoyable Book Shopping Our website is more
than just an online bookstore—it’s a bridge connecting readers to the
timeless values of culture and wisdom. With a sleek and user-friendly
interface and a smart search system, you can find your favorite books
quickly and easily. Enjoy special promotions, fast home delivery, and
a seamless shopping experience that saves you time and enhances your
love for reading.
Let us accompany you on the journey of exploring knowledge and
personal growth!
ebookgate.com