Download as pdf or txt
Download as pdf or txt
You are on page 1of 19

Special Topics Course Report: Spontaneous Symmetry

Breaking and the Higgs Mechanism


Karim Ibrahim, Omar Alsheikh, and Norhan Mahmoud
May 1, 2019

Abstract
In this report, we shall explore local gauge symmetries on the Standard Model Lagrangian
and show how interactions can be modeled as gauge fields. These interactions have to be
massless to keep the local symmetry. Furthermore, the local gauge symmetry in electroweak
fields demands that all fermions be massless, which is obviously not true for both cases. We
shall see that this can be resolved by introducing the idea of Spontaneous Symmetry Breaking
to the SM or what is called the Higgs Mechanism. Moreover, we shall discuss the detection
of the Higgs boson. Lastly, we shall look at an approach to the Higgs Mechanism without
Spontaneous Symmetry Breaking.

1 Symmetries
In general, when we mention a symmetry, we have to specify two things:
1. A corresponding transformation associated with this symmetry.
2. An object which is invariant under this transformation.
For example, a 2-sphere is symmetric under 3D rotations because it looks the same (invariant) after
this transformation. In classical field theory, a physical system enjoys some symmetry when its
equations of motion are invariant under the transformation associated with this symmetry. Since
these equations stem from the system’s action as Euler-Lagrange equations, the system enjoys the
symmetry when its action is invariant under the corresponding transformation.

1.1 Explicit vs. spontaneous symmetry breaking


According to the previous definition of symmetry, the symmetry is broken when the associated
object is not invariant under the corresponding transformation anymore. This can happen in two
different ways. An external effect (force) can actively break the symmetry in the physical sys-
tem. This is called explicit symmetry breaking. It occurs when we add an ’explicit’ term in the
Lagrangian that prevents it from being invariant. This term represents as an external force that
affects the original symmetric system. The second way is that the symmetry is broken without
any external effect. This is called spontaneous symmetry breaking. In this case, the Lagrangian is
still invariant under a specific transformation, but it could be written in new variables such that
the original symmetry is ’hidden’. This will be further illustrated in section 2.

Let’s take as an example a rod that is originally symmetric under rotations around its axis as
in figure 1(a). If we applied an external force perpendicular to its axis, the rod is bent and is
no longer rotationally symmetric as in figure 1(b). Here, the external force ’explicitly’ breaks the
symmetry. However, if two equal and opposite forces are applied in the longitudinal direction of
the rod, the rod will bend in an arbitrary direction as in figure 1(c). In other words, it bends
’spontaneously’. There is no external force here because the net force vanishes. In any case of
spontaneous symmetry breaking, we can notice two important features. First, there are several
equally favorable states (degenerate ground states) that the system can take. The rod can bend
in any direction perpendicular to its axis. Each direction represents a state. Second, these several
ground states are related by the original transformation. In the current case, the different direc-
tions of bending are linked through rotations about the rod axis. The symmetry is still there but

1
it is hidden in this link among the ground states. This implies that the Lagrangian of the system,
and in turn the equations of motion, still satisfy the symmetry, although the individual ground
states do not.

(a) Unbroken Symmetry: (b) Explicitly Broken Symme- (c) Spontaneously Broken Sym-
The rod is rotationally in- try: An external force bends the metry: The rod bends in an ar-
variant rod bitrary direction

Figure 1: Explicit vs spontaneous symmetry breaking in a rod

1.2 Physical vs. gauge symmetry


If a certain symmetry is physical, it relates states that are physically distinguishable. In other
words, the equally favorable ground states are not physically the same. Each bending direction is
observably different, we can tell which state the system chose, through some experiment. On the
other hand, a gauge symmetry relates states that are physically indistinguishable, but mathemat-
ically. different. Unlike the physical symmetry, a gauge symmetry is not a symmetry of nature
we observe, but rather a symmetry of the description of nature. Hence, in this case, we cannot
tell which ’mathematical’ state has been chosen by the system because all of them are physically
equivalent.

Therefore, we can anticipate that a spontaneous breaking of a gauge symmetry can not have
physical consequences. It does not matter which state the system picks, since they all have the
same physical content. The Higgs mechanism is generally presented in the literature as a case of
spontaneous gauge symmetry breaking, however, it plays a crucial physical role in the theory: it
generates the mass of the fundamental particles in the universe. This raises an important concep-
tual problem regarding the Higgs mechanism: if gauge symmetry breaking is not expected to have
physical consequences, how can the Higgs mechanism, have any? It turns out that a gauge in-
variant approach to the Higgs mechanism can be considered without resorting to any spontaneous
breaking of a gauge symmetry. This approach will be illustrated in section 4.

1.3 Global vs. local symmetry


In classical field theory, as previously mentioned, a symmetry is a transformations of the fields that
keeps the Lagrangian, and in turn the action, invariant. In this subsection the difference between
the global and local symmetries is explained, and their relation to gauge symmetry is emphasized.

A global symmetry depends on a set of parameters that are space-time independent. The transfor-
mation associated with this symmetry is carried out at every point in the space-time with the same
magnitude. For example, the rotational invariance of the rod is a global symmetry because every
point in space is rotated around the rod axis with the same angle which is constant in time. In
general, global rotations can be represented by constant special orthogonal matrix Sij (orthogonal
matrix with unit determinant). A local symmetry, on the other hand, depends on set of parameters
that are space-time dependent. In the case of rotation, the angle of rotation can be different at
every point in space and is allowed to vary through time.

2
The dependence on space-time is not the only difference between global and local symmetries.
A fundamental difference is concerned about the determinism of the theory at hand. A certain
physical theory is said to be deterministic if it does not have more than one solution for every
physical situation. In other words, given the initial conditions of the system and its action, we can
predict its future., without doubt.

In classical mechanics, the principle of least action states that the path the particle takes be-
tween two fixed points is the one extremizing the action. Extremizing the action of a certain
system of particles leads to Euler-Lagrange equations as its equations of motion:
d ∂L ∂L
− =0
dt ∂ q̇i ∂qi
This principle can be extended to classical field theory but instead of finding a path of a particle,
we find a ’history’ of a field. In other words, the history that the field will follow between two
fixed configurations is the one that makes the action extremum. In this case the corresponding
equations of motions are:  
∂L ∂L
∂µ − =0
∂(∂µ φi ) ∂φi
In the context of field theory, a theory is deterministic if the realised history (the solution of the
system) is unique, i.e. there is no more than one history that extremizes the action.

Now we are in a position to show the connection between the idea of determinism and the global/lo-
cal classification of symmetries. If we transformed both the action and the history that extremizes
it, the new history should extremize the new action. But, if this transformation is associated to
a symmetry of the action, the transformed action is the identical to the old one. Therefore, there
will be two histories extremizing the same action. Nevertheless, this does not necessarily violate
the determinism of the theory because the transformation might also changes the endpoints of the
history, giving a new history that minimizes the same action but for another situation (another
endpoints). Determinism is violated when the transformation changes the history, but keeps the
endpoints unchanged. There will be two realizable histories between the same endpoints leading
to an indeterministic theory.

A global symmetry can not ruin determinism because it changes the endpoints. The transfor-
mation is applied equally throughout time. Hence, the initial and final configurations can not
remain fixed when changing the realized history. This idea is explained in figure 2(a). The only
case in which the endpoints can remain fixed while the rest of the history changes is the case of
local symmetry. The time dependence of the parameters implies that a local transformation can
leave the endpoints invariant and changes the history between them (see figure 2(b)). Thus a local
symmetry leads, inevitably, to an indeterministic theory.

So, are deterministic theories only those with global symmetries? No, a theory can admit a
local symmetry and remain deterministic. This is possible only if this local symmetry is a gauge
symmetry, i.e. it relates states that are physically identical. Hence, the transformed history is
physically indistinguishable from the old one. In other words the two histories extremizing the
same action are actually the same and determinism is restored. Since any theory is desired to be
deterministic, the previous discussion leads to the following important result. In a deterministic
theory, a local symmetry must be a gauge symmetry.

1.4 Lagrangians
As previously stated, the equations of motions satisfied by a certain system of fields results from
plugging in the Lagrangian density of the system in Euler-Lagrange equations. For completeness
we list here the Lagrangian densities of the scalar, spinor, and vector fields and their associated
equations of motion1 .

1 In the following, we adopt the natural units (h̄ = c = 1)

3
(a) A global symmetry (b) A local symmetry can keep
must change the end points the endpoints unchanged

Figure 2: A local symmetry violates determinism

1. Spin-0 (scalar) field:


The Lagrangian of a complex scalar field:

L = (∂µ φ)(∂ µ φ∗ ) − m2 φφ∗ (1)

leads to the Klein-Gordon equation:

∂µ ∂ µ φ + m2 φ = 0 (2)

2. Spin-1/2 (spinor) field:


The Lagrangian of a spinor (Dirac) field:

L = iψ̄γ µ ∂µ ψ − mψ̄ψ (3)

leads to the Dirac equation:


iγ µ ∂µ ψ − mψ = 0 (4)

3. Spin-1 (vector) field:


The Lagrangian of a vector (Proca) field:
−1 µν 1 2 ν
L= F Fµν + m A Aν (5)
16π 8π
where:
F µν = ∂ µ Aν − ∂ ν Aµ (6)
leads to the Proca equation:
∂µ F µν + m2 Aν = 0 (7)
This equation is quite similar to Maxwell’s equations. In fact, the electromagnetic field is
exactly a massless vector field. If we put m = 0 in the Proca equation, we have Maxwell’s
equations in vacuum. However, the Lagrangian of a massless vector field Aµ interacting with
a source J µ :
−1 µν
L= F Fµν − J µ Aµ (8)
16π
leads to Maxwell’s equations in the presence of a source J µ :

∂µ F µν = 4πJ ν (9)

4
1.5 Imposing local U (1) symmetry
Let’s consider the Lagrangian of the Dirac field:

L = iψ̄γ µ ∂µ ψ − mψ̄ψ

This Lagrangian is manifestly invariant under the global U (1) phase transformation:

ψ → eiθ ψ

ψ̄ → ψ̄e−iθ
where θ is a constant scalar parameter. The partial does not act on the exponential and the
Lagrangian is unchanged. But what if the transformation is local? (i.e. θ is space-time dependent):
µ
ψ → eiθ(x ) ψ
µ
ψ̄ → ψ̄e−iθ(x )

Obviously, the previous Lagrangian is not invariant under this transformation, for now the partial
acts on θ and we have an extra term:

L → L − (ψ̄γ µ ψ)∂µ θ (10)

In the following, it is more convenient to work with the scaled parameter:


θ(xµ )
λ(xµ ) ≡ −
q
where q is the charge of the particle under study (i.e. the quanta of the involved field). Hence, in
terms of λ:
L → L + (q ψ̄γ µ ψ)∂µ λ
µ (11)
when: ψ → e−iqλ(x ) ψ

Now, let’s impose this local symmetry on the Lagrangian. In order to do that, we need to an extra
term to the original Lagrangian to absorb the offending term coming from the derivative of λ. We
can accomplish this by adding some new field to the Lagrangian that transforms, under local U (1),
in an opposite way to the transformation of the original Lagrangian. Let

L = iψ̄γ µ ∂µ ψ − m2 ψ̄ψ − (q ψ̄γ µ ψ)Aµ (12)

where Aµ is a some vector field that transforms, under local U (1), in the following way:

Aµ → Aµ + ∂ µ λ (13)

It is easily verifiable that this updated Lagrangian is now locally U (1) invariant. The extra term
in the transformation of Aµ exactly cancels the offending term in the transformation of the old
Lagrangian. However, (12) is not the end of the story. Since we introduced a new field Aµ , we
have to include, in the Lagrangian, its free Lagrangian (i.e. Proca Lagrangian)
−1 µν 1 2 ν
L= F Fµν + m A Aν
16π 8π
Unlike the first term, the second term (m2 Aν Aν ) is not invariant under (13). Hence, we have to
take the new field to be massless, if we want the local U (1) symmetry not to be spoiled.

Conclusion: If we insist that the Dirac Lagrangian be locally phase invariant, we are obliged
to introduce a massless vector field (gauge field) and the final Lagrangian becomes
1 µν
L = iψ̄γ µ ∂µ ψ − m2 ψ̄ψ − F Fµν − (q ψ̄γ µ ψ)Aµ (14)
16π
This Lagrangian is precisely the Lagrangian for quantum electrodynamics. It produces the equation
of a Dirac field ψ in an electromagnetic background Aµ and Maxwell’s equations with source:

J µ = q(ψ̄γ µ ψ) (15)

5
The Lagrangian (14) can be rewritten in the following way:
1 µν
L = iψ̄γ µ Dµ ψ − m2 ψ̄ψ − F Fµν (16)
16π
where
Dµ ≡ ∂µ + iqAµ (17)
We notice that if we replace the partial derivative ∂µ with the so-called ’covariant derivative’ Dµ in
the original non-locally invariant Lagrangian, the local symmetry is imposed. Hence, the covariant
derivative is a simple tool for converting a global symmetry into a local one.

Next, we apply the same concept to the Lagrangian of three quarks but with a different local
symmetry.

1.6 Imposing local SU (3) symmetry


Consider the following Lagrangian

L = [iψ̄r γ µ ∂µ ψr − m2 ψ̄r ψr ]
+[iψ̄b γ µ ∂µ ψb − m2 ψ̄b ψb ] (18)
µ 2
+[iψ̄g γ ∂µ ψg − m ψ̄g ψg ]

which is the Lagrangian of a particular quark flavor. It can be written in a more compact way if
we define:  
ψr 
ψ ≡  ψb  , ψ̄ = ψ̄r ψ̄b ψ̄g (19)
ψg
Now, the Lagrangian reads
L = iψ̄γ µ ∂µ ψ − mψ̄ψ
This is the same as the one-field Dirac Lagrangian. But, here ψ is composed of three elements,
each one is a four-component vector. This Lagrangian enjoys a more general symmetry than global
U (1)
ψ → Uψ (ψ̄ → ψ̄U † ) (20)
where U is a constant 3 × 3 unitary matrix. This is called global U (3) transformation. As in
the U (1) case, the current Lagrangian is invariant under U (3) because the partial derivative does
not act on the transformation parameters. Before imposing the local invariance, let’s write this
transformation in a more suggestive way. Any unitary matrix can be written in the form

U = eiH , H† = H (21)
In addition, we can expand H in the basis of 3 × 3 hermitian matrices as

H = θ11 + β · τ (22)

where 1 is the identity matrix and τ are the Gell-Mann matrices (the dot product denotes a linear
combination of the Gell-Mann matrices).

Therefore,
U = eiθ eiββ ·ττ (23)
β ·τ
iβ τ
We have already studied the U (1) part, now we focus on the second term. The matrix e belongs
to the group SU (3) (3 × 3 orthogonal matrices of unit determinant). Hence, we are now interested
in imposing the local SU (3) symmetry
µ
ψ → eiββ (x τ
)·τ
ψ (24)

on the three-field Lagrangian. It is also convenient here to work with the scaled parameters
β
λ≡−
q

6
Such that,
ψ → Sψ (25)
µ
S = e−iqλλ(x τ
)·τ
(26)
As before, the Lagrangian (18) is not invariant under local SU (3) transformation because the
derivative produces an extra term:

∂µ ψ → S∂µ ψ + (∂µ S)ψ (27)

To fix this problem, we try our previous trick. We replace the normal derivative with the covariant
derivative defined in this case as:
Dµ ≡ ∂µ + iqττ · A µ (28)
then we assign to the fields A µ (now, there are eight gauge fields, one for each Gell-Mann matrix)
a transformation rule, under SU (3), such that we have

Dµ ψ → S(Dµ ψ) (29)

However, this does not lead to the trivial transformation rule (13). It can be shown that the
condition (29) leads to the following rule

i
τ · A µ → S(ττ · A µ )S −1 + (∂µ S)S −1 (30)
q

It is now evident why the transformation rule does not reduce to (13). S does not necessarily
commute with neither τ · A µ nor τ · ∂µλ . In the case of infinitesimal (very small |λ
λ|), this rule
reduces to (after a couple of steps)

A µ → A µ + ∂µλ + 2q(λ
λ × Aµ) (31)

The new Lagrangian is


1 µν
L = iψ̄γ µ ∂µ ψ − mψ̄ψ − F · F µν − (q ψ̄γ µτ ψ) · A µ (32)
16π
where
F µν · F µν ≡ F1µν Fµν 1 + F2µν Fµν 2 + F3µν Fµν 3
is thus invariant under local SU (3). Again, we have excluded the mass terms (Aµ Aµ ) in the free
Lagrangians of the gauge fields because it spoils the symmetry, i.e. the eight gauge fields have to
massless. However, if we can not keep the same definition F µν ≡ ∂ µ Aν − ∂ ν Aµ because it is not
invariant under (30) or even under the infinitesimal transformation (31).

Instead, we take
F µν ≡ ∂ µA ν − ∂ ν A µ − 2q(A
Aµ × A ν ) (33)
This new F µν is still not invariant under (31). It transforms as

F µν → F µν + 2q(λ
λ × F µν ) (34)

but it can be shown that F µν · F µν is now invariant. There is a proof, beyond our scope, that
with the definition (33) the invariance of the Lagrangian (32) is extended to the case of finite local
SU (3) transformation.

Conclusion: We started with the Lagragnian (18) and imposed the local SU (3) symmetry. This
led us to the Lagrangian (32) describing the interaction between three Dirac fields (three colors
of a quark flavor) and eight massless gauge fields (the gluons, mediators of the strong force), i.e.
the Lagrangian of quantum chromodynamics. The quark fields act as eight source currents for the
gluon fields
J µ ≡ q(ψ̄γ µτ ψ) (35)

7
2 Spontaneous Symmetry Breaking and the Higgs Mecha-
nism
2.1 φ4 theory
To see how a gauge field can acquire mass without breaking the local gauge symmetry, we first
consider a simple Lagrangian of a self-interacting complex scalar field
1
L(φ, φ∗ ) = (∂µ φ)(∂ µ φ)∗ + µ2 φφ∗ − λ2 (φφ∗ )2 . (36)
2
We note that the mass term has the wrong sign, and one can naively conclude that this corresponds
to quanta of the φ (and φ∗ ) fields with imaginary masses that travel with superluminal velocities
i.e. tachyonic particles. However, we know from practice that such objects do not exist, and one
might be tempted to discard this Lagrangian all along. There is, however, a small catch: we always
assume that the minimum of the potential of the field is at the origin i.e. the field has a vanishing
vacuum expectation value, but this is not the case with this Lagrangian (see figure 3). The minima
of the potential U (φ, φ∗ ) = −µ2 φφ∗ + 21 λ2 (φφ∗ )2 is different from zero. This can be easily shown
by setting the first derivatives equal to zero:

∂U
= −µ2 φ + λ2 φ2 φ∗ = 0
∂φ∗
∂U
= −µ2 φ∗ + λ2 φφ∗2 = 0.
∂φ
One possible solution is φ = 0, but this is a local maximum, as evident from the potential profile.
The other solution is
−µ2 + λ2 |φ|2 = 0
or
µ2
|φ|2 = . (37)
λ2

Figure 3: Mexican Hat potential of the φ4 theory. Notice how the point φ = 0 is a local maximum,
and that there is a circle of minima centered around the origin.

But why are we interested in finding the minimum of the potential? The answer to that can be
summarized in the following three points:

8
1. In quantum field theory, particles are excitations (quanta) of the underlying quantum fields.

2. Feynman calculus, which is a fancy name for the calculational methods we use in quantum
field theory, is a perturbative approach. This means that we add small variations on the
background field configuration and produce the physics.
3. In this language, particles are just fluctuations on the ground state configuration (vacuum),
where expectation values of physical observables (energy, momentum, etc.) vanish. In other
words, crudely, the vacuum configuration is that of a static, constant field i.e. all derivatives
vanish. Thus, the Euler-Lagrange equations in this case are tantamount to minimizing the
potential.2
To make sense of the result from (37), we decompose φ into two real fields φ1 and φ2 as such:
1
φ = √ (φ1 + iφ2 ) (38)
2
and so (36) becomes
1 1 1 1 1
L(φ1 , φ2 ) = (∂µ φ1 )(∂ µ φ1 ) + µ2 φ21 + (∂µ φ2 )(∂ µ φ2 ) + µ2 φ22 − λ2 (φ21 + φ22 )2 . (39)
2 2 2 2 8

2.2 Breaking the global symmetry


We expand this Lagrangian around one possible solution to (37):
µ
φ1 = + η, φ2 = 0 + ξ (40)

λ
   
1 µ 2 2 1 µ
L(η, ξ) = (∂µ η)(∂ η) − µ η + (∂µ ξ)(∂ ξ) + · · · . (41)
2 2
We omitted interaction terms that are not of much relevance here. The first square brackets con-
tain a real massive scalar field η with a mass term with the correct sign, and the second square
brackets contain a real massless scalar field ξ.3

The original Lagrangian was invariant under the global phase transformation φ → eiθ φ, which
is the same thing as

φ1 → φ1 cos θ − φ2 sin θ
φ2 → φ1 sin θ + φ2 cos θ. (42)

However, it does not possess the same symmetry for η and φ i.e. L(η, ξ) is not invariant under

η → η cos θ − ξ sin θ
ξ → η sin θ + ξ cos θ. (43)

This is an example of a spontaneous global symmetry breaking. Please note that the Lagrangian
is still invariant under the original transformations (42). After all, all we did was redefine our
variables and substitute in. The symmetry is broken in the sense that we chose one minimum to
expand around out of the infinite number of possible minima from (37), and this procedure has
left us with the ‘physical’ fields η and ξ whose excitations are physically observable particles that
do not possess the original form of the symmetry anymore.

An intuitive way to understand why η is massive while ξ is not is by noticing that ’concavity
2 Actually, if we follow this logic only, then we have no right to eliminate solutions that are local maxima. This

is partially true, but suppose we really start off at a maximum: this would be an unstable configuration, and any
small perturbation would make the field ’roll away’ to a minimum.
3 The appearance of this field is no surprise, and is the result of the Goldstone theorem, which states that

spontaneously breaking a continuous symmetry is always accompanied by the appearance of one or more massless
scalar fields, or Goldstone bosons.

9
determines mass’. The mass term appears bilinear in the fields, and the relations
∂ 2 U (η, ξ) ∂ 2 U (φ1 , φ2 )

= = 2µ2 (44)
∂η 2 η=0,ξ=0 ∂φ21
φ1 =| µ |,φ2 =0
λ

∂ 2 U (η, ξ) ∂ 2 U (φ1 , φ2 )

= =0 (45)
∂ξ 2 η=0,ξ=0 ∂φ22
φ1 =| µ |,φ2 =0
λ

4
follow trivially. Thus, moving in the η (radial) direction would be moving uphill, and the system
experiences a restoring force which may be interpreted as mass. On the other hand, moving in
the ξ (tangential) direction is moving along the circle of minima, and the system experiences no
resistance at all.

While we are at it, evaluating the second derivative at φ = 0 (in any direction) would give negative
concavity. For example,
∂ 2 U (φ1 , φ2 ) ∂ 2 U (φ1 , φ2 )

= = −µ2 (46)
∂φ21
φ1 =0,φ2 =0 ∂φ 2
2

φ1 =0,φ2 =0

implying an unstable configuration. Any fluctuation (and in the quantum world, there are a lot of
fluctuations) and the field will evolve into a different solution, typically rolling down towards any
permissible minimum value of the potential.

There is nothing special about our choice in (40). Had we chosen another solution, we would
have obtained different masses for the fields. After all, this is what the spontaneity of the process
entitles; the field could have assumed any value on the circle of minima.

To summarize, we started with:


• two real (massive) scalar fields φ1 and φ2
• a simple self-interaction
and after breaking the symmetry we ended up with:

• one real massive scalar field η with mass 2|µ|
• one real massless scalar field ξ, a Goldstone boson
• a big bunch of interactions.

2.3 Breaking the local U (1) symmetry


A very interesting thing happens when we demand the original Lagrangian in (36) to be invariant
under local U (1) transformations. The covariant Lagrangian
1 1
L = (Dµ φ)(Dµ φ)∗ − Fµν F µν + µ2 φφ∗ − λ2 (φφ∗ )2 (47)
16π 2
becomes, after expanding using (40),
   
1 µ 2 2 1 µ
L = (∂µ η)(∂ η) − µ η + (∂µ ξ)(∂ ξ)
2 2
1 1 µ
  2
− Fµν F µν + q Aµ Aµ
16π
( 2 λ
µ
+ q η(∂µ ξ) − ξ(∂µ η) Aµ + q 2 ηAµ Aµ
 
λ
1
+ q 2 (ξ 2 + η 2 )Aµ Aµ − |λµ|(η 3 + ηξ 2 )
2 )
1 2 4  µ   µ2 2
− λ (η + 2η ξ + ξ ) + q (∂µ ξ)Aµ +
2 2 4
. (48)

4 λ 2λ
4 All interaction terms are of order 3 in the fields or higher, and thus their second derivatives vanish at η = 0.

10
The last term is a constant and can be dropped. Notice the second term in the second line; it
looks as if the gauge boson Aµ has acquired non-zero mass. Initially, we did not assume a non-zero
mass for Aµ , as per the local gauge invariance requirement. It just ‘popped up’, and this all has
to do with the fact the the field φ had a non-vanishing vacuum expectation value, which allowed
us to expand around a non-trivial minimum. This time, this is an example of a spontaneous local
symmetry breaking. Again, we could have equally chosen any minimum to expand around, and
hereby allowing the fields to acquire different masses. To summarize, we started with:

• two real (massive) scalar fields φ1 and φ2


• one massless gauge field Aµ
• simple interactions
and we ended up with:

• one real massive scalar field η with mass 2|µ|
• one real massless scalar field ξ, a Goldstone boson

• one massive gauge field Aµ with mass 2 πq µλ

• a big bunch of interactions.


This Lagrangian is called the Higgs Lagrangian, and η is the Higgs boson.5 The idea of sponta-
neous symmetry breaking, when applied to the Standard Model, is called the Higgs mechanism.
Everything seems to fall in place, except for this peculiar ξ field, or the Goldstone mode. To see
the problem, we note that we started with four degrees of freedom (two for the φ field and two
for the massless Aµ field), but we ended up with five degrees of freedom (two for the scalar fields
η and ξ and three for the massive Aµ field). Moreover, the second term in the last line in (48) is
suspicious. It corresponds to a vertex where a ξ boson changes into an Aµ boson.

This problem has a very simple solution, the degree of freedom associated with ξ is redundant!
To see this, we note that for the particular choice we made in (40), ξ = φ2 , and by choosing a
particular gauge, we can transform ξ entirely away. From (42):

φ1 sin θ + φ2 cos θ = 0

or
φ2
θ = − arctan (49)
φ1
gives ξ = 0.6 A degree of freedom associated with one of the original scalar fields, the Goldstone
mode, is lost, and is simultaneously gained by the gauge boson, acquiring a new mode (longitudinal
polarization). The gauge field is said to have ‘eaten’ the Goldstone boson, acquiring mass in the
process.

2.4 Electroweak symmetry breaking


The above discussion was just a toy model to illustrate how the Higgs mechanism works. We know
that the photon, which is the gauge boson for the U (1) group, is massless i.e. it does not interact
with the Higgs field. In reality, the symmetry that is broken is the SU (2)L × U (1)Y symmetry of
the electoweak sector of the Standard Model, giving the weakly interacting fermions as well as the
W and Z bosons masses. We will not go into details, for it is beyond the scope of this report and
is of much physical and mathematical sophistication, and we only give a very brief summary of
what happens.

Before symmetry breaking, local gauge invariance of the electroweak Lagrangian does not only
5 The Higgs Lagrangian is the only Klein-Gordon Lagrangian in the Standard Model i.e. the Higgs boson is the

only fundamental scalar particle in nature.


6 This is sometimes called the physical gauge, since only physical particles appear in the Lagrangian. However,

there is nothing special about this particular gauge. In fact, in some contexts, it is more convenient (and physically
equivalent) to work without fixing the gauge.

11
demand the gauge bosons to be massless, but it also prohibits fermions from having mass, which
we know is not true: the stuff that makes up the universe is massive. By adding the Higgs field
and allowing interactions with it, the problem is solved. After breaking the symmetry, fermions
interacting with the Higgs field can assume non-zero mass, as well as the gauge bosons. Namely,
The W + and W − , which are the two charged weak bosons and two of the gauge fields associated
with the SU (2)L symmetry group, acquire mass. Linear combinations of the third gauge boson
associated with SU (2)L , the W 0 boson, and the gauge boson for the U (1)Y group, the Y 0 boson,
give us the massive neutral Z boson and the massless A photon:

Aµ = Yµ0 cos θW + Wµ0 sin θW (50)


Zµ = −Yµ0 sin θW + Wµ0 cos θW (51)

where θW , known as the Weinberg angle or the mixing angle, is an experimentally determined
quantity.

12
3 Higgs Boson Detection at CERN
After we have discussed Spontaneous Symmetry Breaking and setting it as the framework of the
Higgs Mechanism in its role of mass generation, we move on to exploring the actual detection of
the Higgs Boson, the quantum of the Higgs field, in the LHC runs of 2011 and 2012.

3.1 Before the Higgs detection


The standard model (SM) of particle physics describes the fundamental particles and the electro-
magnetic, weak, and strong forces between them. These forces are all mediated by the so-called
gauge bosons. Photons mediate the electromagnetic force, Gluons mediate the strong force, and W
and Z bosons mediate the weak force. The two former bosons are massless while the two latter are
massive (masses are ∼ 100 times that of a proton). In the SM, they obtain their masses through
their interaction with a field of weak charge pervading the vacuum of space (assumingly the Higgs
field).

If enough energy is provided, excitation of this field will produce a massive particle with zero
spin, namingly the Higgs Boson. The mass of the Higgs Boson is not specified in the SM. However,
quantum mechanical effects link mH to properties of known particles. Experimental data was used
to determine a suitable mass region for the Higgs mass throughout the years.

3.2 Predictions of what was to be observed


According to the standard model, if the Higgs Boson were to have a mass of ∼ 126 GeV, it would
have 5 main decay channels (modes) as listed in figure 4 with their corresponding probabilities.

Figure 4: The Higg boson’s 5 main decay channels:


a. Higgs to b and b̄ b. Higgs to W and W∗
c. Higgs to Z and Z∗ d. Higgs to γγ
e. Higgs to τ and τ̄

The Higgs boson is a shortly lived particle of a lifetime of ∼ 10−22 s, meaning that it is practically
impossible for the LHC to directly detect the boson itself. The detectors at the LHC only record the
interactions of its decay products. Hence, evidence for Higgs boson production would be inferred
from statistically significant excesses of events of its decay products above background predictions
(events with similar products).

3.3 Signal predictions


According to the SM, the most important SM Higgs boson (with such predicted invariant mass)
production mode in the LHC energy range is expected to be gluon fusion (see figure 5). The two
most sensitive decay channels are the decay into two photons (H → γγ) and the decay into two
Z bosons which in turn decay into four leptons (H → ZZ∗ → llll). Both decay channels were

13
examined in the data from 2011 and 2012. Hence, with this information, it could be predicted
which events would be eligible for being considered for the search for the Higgs boson.

Figure 5: The production cross section and number of events for 105 pb−1 of the Higgs boson as
a function of the Higgs mass at the LHC.

3.4 Evidence of the Higgs Detection


After analyzing the data from the 2011 and 2012 runs of the LHC, it was found that there was
finally evidence for the detection of the Higgs boson. As previously mentioned, the two most sensi-
tive and accessible decay channels are H → γγ and H → ZZ∗ → llll, which both showed an excess
of events at an inverse mass of ∼ 125 GeV. We show this excess in figures 6 and 7.

It is needed to be mentioned that even with this excess of events, we cannot completely claim
that what is observed are the products of the Higgs boson decaying, as there may be some other
particle that we do not know of that is giving us these results or this may not be the Higgs boson
predicted by the standard model. Further studying of the properties and behaviour of the data
from the observed particle must be carried to make sure of the fact that this is indeed the Higgs
boson. Hence, we are only capable of claiming that this is the SM Higgs boson because of the fact
that what we have observed so far from the analysis of the data matches what the SM predicts.

3.4.1 H → γγ detected decay channel


The decay of the SM Higgs boson to a pair of photons occurs mainly via quantum loop processes
involving the W boson. This is not a highly probable decay channel, however, the signature of two
highly energized photons isolated from any other sizable activity in the detector is distinctive.

The ATLAS results are in the histogram in figure 6.

3.4.2 H → ZZ∗ detected decay channel


The decay channel in which the SM Higgs boson decays to two Z bosons, each decaying to either
e + e− or µ + µ−, offers the best signal purity. However, as only 7% of Z bosons decay this way,
the rate is low.

The ATLAS results are in the histogram in figure 7.

14
Figure 6: Distribution of the mass, mγγ , of weighted di-photon candidates. The selected events
are weighted by factors that reflect the signal-to-background ratio predicted for a SM Higgs boson.
The result of a fit to the data of the sum of a signal component fixed to mH = 126.5 GeV and a
background component described by a fourth-order polynomial are superimposed.

Figure 7: The distribution of the mass of the selected H→ZZ*→ llll candidate events, mllll . The
small peak at 90 GeV corresponds to a single Z boson decaying to four leptons, whereas the broad
structure around 200 GeV results from the direct production of Z boson pairs. An excess is seen
around 125 GeV; the expected signal from a SM Higgs boson at that mass (light blue) is added for
comparison.

15
4 The Higgs Mechanism Without Spontaneous Symmetry
Breaking
In this section, the option of explaining the Higgs mechanism without needing to resort to Spon-
taneous Symmetry Breaking will be explored. It may sound like an absurd feat, however, such an
option was already studied by Higgs himself in his 1966 paper.

Supposedly, the same results that we’ve obtained earlier in the presentation shall be obtained
by making convenient field transformations. Such a procedure will be worked out in detail for the
U (1) symmetry case.

4.1 Required field transformations


The starting point is once again the U (1) gauge invariant Lagrangian for φ4 theory.

Through certain field transformations, the set of fields (Aµ , φ) is transformed into the new set
of fields (Bµ , ρ, ξ) which are related in the following way:

1
Bµ = Aµ − ∂µ ξ(x) (52)
q

φ(x) = ρ(x)eiξ(x) 7 (53)

4.2 The gauge invariance of the newly obtained fields


Performing a gauge transformation on the new set of fields gives us the following:
0 0 1 0 1 1 0
Bµ (x) = Aµ (x) − ∂µ ξ (x) = Aµ + ∂µ α(x) − ∂µ ξ (x) (54)
q q q
0 0 0
eiα(x) φ(x) = φ = ρ (x)eiξ (x) (55)

And hence, we find that the transformed fields are related to the original fields in the following
manner:
0
Bµ (x) = Bµ (x)
0
ρ (x) = ρ(x) (56)
0
ξ (x) = ξ(x) + α(x)

Meaning that Bµ and ρ are both gauge invariant while ξ is a pure gauge variable.

4.3 The Lagrangian with the new fields


The original Lagrangian in (36) (written in its covariant form) now becomes as follows after sub-
stituting with the newly obtained fields:

L = −B µν Bµν + (∂µ + iqBµ )ρ(∂ µ − iqB µ )ρ − m2 ρ2 − λρ4 (57)


where
Bµν = ∂µ Aν − ∂ν Aµ (58)

We see that the pure gauge variable ξ(x) is not included in the Lagrangian. Hence, all fields in this
theory are now invariant under the U (1) gauge transformation; the U (1) symmetry has no grip on
this theory. Now that the U (1) symmetry has been factored out, the ground state of the system is
no longer degenerate. There is a unique ground state with Bµ,0 = 0 and ρ0 = v.

7 Both ρ and ξ are real.

16
To describe the perturbations around the ground state, ρ(x) is rewritten as

ρ(x) = v + η(x) (59)

and the Lagrangian now becomes

L = −B µν Bµν + ∂µ η∂ µ η + µ2 η 2 + q 2 v 2 Bµ B µ + · · · (60)

giving us the same result of the Spontaneous Symmetry Breaking approach.

4.4 The underlying feature of the Higgs mechanism


We can see that the differences between explaining the Higgs Mechanism in the framework of
Spontaneous Symmetry Breaking vs. without it are as so:

Spontaneous Symmetry Breaking No Spontaneous Symmetry Breaking


1. Breaking the symmetry first 1. Field transformations to gauge invariant fields first
2. Secondly, gauge fixing 2. Secondly, breaking the symmetry
3. ξ(x) is fixed 3. ξ(x) is a pure gauge variable

We can now come and ask ourselves if spontaneous symmetry breaking is the underlying feature
of the Higgs mechanism and if not, what is?

From our previous discussions of the two approaches, it is found that spontaneous breaking of
the gauge symmetry is not a necessary step in the Higgs mechanism. Is the possibility of describ-
ing the Higgs mechanism as gauge symmetry breaking then still problematic? No, spontaneous
gauge symmetry breaking is not connected to the description of physics, since the meaning of the
symmetry breaking disappears once the physics comes into play and a gauge is fixed. The possi-
bility of giving a gauge independent description of the Higgs mechanism underlines that.

The true underlying feature of the Higgs mechanism is the one crucial element in common be-
tween the two methods, which is the the non-zero vacuum expectation value of the Higgs field that
enforced the asymmetry in the ground state of the system.

17
5 Discussion and Conclusion
5.1 Was the theory ever symmetric?
A logical question one can ask is if there was a time the ground state respected the symmetry, or in
other words the Higgs field had a vanishing vacuum expectation value. In theory, this can be done
by promoting the parameters µ and λ to be functions of the cosmic time, or, which is physically
equivalent, the temperature of the universe. In fact, it is estimated that the electroweak scale, the
energy scale above which the electromagnetic and weak interactions merge into one, is above a
unification energy of ∼ 246 GeV, or around a temperature of ∼ 1015 -1016 K. According to the Big
Bang model, the universe maintained this temperature for only 10−12 s after the Big Bang, during
which | µλ |  1, followed by the quark epoch. After that, the symmetric potential relaxed into the
Mexican Hat potential of the φ4 theory, a process commonly referred to as condensation.

5.2 Domain walls


As we have seen, the vacuum expectation value assumed by the Higgs field has an effect on the
observed masses of the different particles. It is possible, although unlikely, that in different parts
of the universe the Higgs field assumed different vacua, implying that the mass of the electron, for
example, varies according to its location. Though this claim may sound absurd, it is not impossi-
ble, especially for causally disconnected regions within the universe.

This idea is very similar to the concept of domain walls in ferromagnetism, where the magnetic
dipoles align themselves parallel to each other within a domain, and two neighboring domains can
have different, even antiparallel, magnetic dipole moments.

5.3 Subsequent condensations and more Higgs modes


Another possible extension to the theory would be allowing for subsequent condensations, that is
allowing for further symmetry breaking. It may be the case that after the temperature drops to
a certain level, the current potential profile deforms once more, and hence changing the allowed
minima. Of course, this would affect any particle that interacts with the Higgs field, and may
hence change the mass of the known fundamental particles. It may also be the case that there is
more than one Higgs, or Higgs-like, field. However, current experimental data does not provide
any indication to the existence of such fields, and so the idea remains a matter of speculation.

18
References
[1] S. Van Dam, “Spontaneous symmetry breaking in the higgs mechanism,”
[2] L. H. Ryder, Quantum field theory. Cambridge university press, 1996.
[3] D. Griffiths, Introduction to elementary particles. John Wiley & Sons, 2008.

[4] A. Collaboration et al., “A particle consistent with the higgs boson observed with the atlas
detector at the large hadron collider,” Science, vol. 338, no. 6114, pp. 1576–1582, 2012.

19

You might also like