Download as pdf or txt
Download as pdf or txt
You are on page 1of 114

Relaxing Symmetries in Field Theory:

from Noether theorem to noncommutativity and


the challenges to Lorentz invariance

Alfredo Iorio

Graduate Lecture Notes

IPNP Prague
25 October - 29 November, 2005
Contents

Introduction 4

1 Spacetime and Internal Symmetries 6


1.1 Emmy Noether’s Theorem . . . . . . . . . . . . . . . . . . . . 6
1.2 The issue of “improvements” . . . . . . . . . . . . . . . . . . 8
1.2.1 Maxwell’s − 14 F 2 . The canonical and symmetric en-
ergy momentum tensor . . . . . . . . . . . . . . . . . 8
1.2.2 The scale vs conformal symmetry case . . . . . . . . . 8
1.3 How to deal with broken symmetry . . . . . . . . . . . . . . . 9
1.4 The quantum case . . . . . . . . . . . . . . . . . . . . . . . . 11
1.5 Exercises . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11
1.6 Further Reading . . . . . . . . . . . . . . . . . . . . . . . . . 12

2 Combining Spacetime and Internal Symmetries 13


2.1 No-go Theorems . . . . . . . . . . . . . . . . . . . . . . . . . 13
2.2 O’Raifeartaigh Theorem(s) . . . . . . . . . . . . . . . . . . . 15
2.3 Exercises . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 16
2.4 Further Reading . . . . . . . . . . . . . . . . . . . . . . . . . 17

3 The (special) case of SUSY 18


3.1 The Haag-Lopuszansky-Sohnius Theorem . . . . . . . . . . . 18
3.1.1 Building up the Super-Poincarè Algebra. . . . . . . . . 18
3.1.2 Irreducible Representations and Supermultiplets. . . . 21
3.2 General features of SUSY theories . . . . . . . . . . . . . . . 24
3.3 The Wess-Zumino model and its (Noether) supercharges . . . 27
3.4 N=2 Susy Classical Georgi-Glashow model and the Mass For-
mula . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 30
3.5 Exercises . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 33

4 The infinite number of degrees of freedom of quantum fields,


the Haag’s theorem and the issue of locality 34

2
5 Spatiotemporal Noncommutativity 35
5.1 The fundamental length . . . . . . . . . . . . . . . . . . . . . 35
5.2 Examples in Nature . . . . . . . . . . . . . . . . . . . . . . . 35
5.3 Formalization: The gauge-covariant approach and the Seiberg-
Witten map . . . . . . . . . . . . . . . . . . . . . . . . . . . . 35
5.4 Problems? . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 35

6 From Lorentz symmetry violation to Lorentz symmetry en-


hancement? 36

7 The elusive effects of Lorentz violation: the case of − 14 F̂ 2 37

3
Introduction

These are days of a deep rethinking of fundamental issues in theoretical


physics, the most important of all being the symmetry principles. In these
lectures we want to introduce the challenges to Lorentz invariance that arise
from different areas of the theoretical investigation – especially the hypoth-
esis of noncommuting coordinates – and lead to everyday more precise ex-
perimental tests. This is a particularly fortunate case, as lately the most
advanced theoretical research in particle physics has been decoupled from
the experimental challenges.
We shall try to accomplish our goal by first presenting the most important
tools and ideas used in the application to physics of the symmetry principle.
In this way we shall have the chance to state and prove Emmy Noether’s
1918 theorem in field theory, discuss to the greatest possible extent the
subtleties involved with space-time symmetries, look closely at the difference
between internal and space-time symmetry, and introduce what ought to
be the highest possible symmetry of the S-matrix, namely Supersymmetry
(SUSY). This will close the first part of the course, where the old material
above described will be introduced in the most (to us) critical way.
Once we have solidly established our own technology on the fundamental
issues we shall be able to move to the new arena of what is known these days
as “Lorentz violation”. I shall first try to explain the theoretical motivations
behind the investigation of a departure from Lorentz invariance. Then I shall
touch upon as many as possible theoretical approaches to this fascinating
hypothesis. I will then focus on one of them, namely noncommutativity of
the spatiotemporal coordinates, for two reasons: a) it is what I understand
and like the most; b) there are very recent important results I want to
disclose to you. Eventually I will keep my promise of telling you about the
state-of-the-art with experiments and experimental proposals.
The lectures are two hours each. There are two (easy) exercises for each
lecture. I will propose a series of papers for further reading on the various
topics we will discuss, the most important papers will be listed separately.

4
The expected home-work is to solve the exercises and present a proposed
paper in the form of a journal club during the last class.

Further reading

Of course, it is purely a matter of taste to pick up one or two references on


the symmetry principle in physics! On the other hand, I do have my own
suggestions:
[1] K. Brading and E. Castellani Eds., “Symmetries in Physics: Philo-
sophical Reflections”, 2003, Cambridge University Press. This wonderful
book is a collection of novel contributions to the philosophical debate on
symmetry in physics. Do not get misled by the word “philosophy”, they
know a lot about mathematics and physics...
[2] L. O’Raifeartaigh, “The Dawning of Gauge Theory”, 1997, Princeton
University Press. This is a rare book on the history of physics wrote by
a first rank physicist. The early papers on gauge theory are reprinted and
commented upon in a deep scholarly way. Again, it is not just a book on
the history of the gauge principle, there is a lot to learn from the physics
side!

5
Chapter 1

Spacetime and Internal


Symmetries

1.1 Emmy Noether’s Theorem


In a theory with Lagrangian density L(Φi , ∂Φi ) (where the collective index
i takes care of the different fields, as well as their spin type) the Noether
current for space-time transformations has the form

Jfµ = Πµi δf Φi + Lδf xµ , (1.1)

where δf xµ = f µ , Πµi = δL/δ∂µ Φi , δf Φi = Φ0 i (x) − Φi (x), and the infinites-


imal quantities f µ take the following form

f µ = aµ or f µ = ωνµ xν or f µ = axµ or f µ = aµ x2 −2a·xxµ , (1.2)

for infinitesimal translations, rotations (and boosts), dilations, and special


conformal transformations, respectively, where, as usual, ω µν = −ω νµ .
The expression for the current (1.1) has been obtained by varying the action,
including the measure, under an arbitrary space-time transformation, and
only afterwards one tests the invariance. Let us sketch here a proof.
Let us consider the Action
Z
AΩ = d4 xL(Φi , ∂Φi ) (1.3)

where Ω is the space-time volume of integration. The infinitesimal transfor-


mations of the coordinates, of the fields and of the derivatives of the fields
are given respectively by

xµ → x0µ = xµ + δf xµ (1.4)

6
Φi (x) → Φ0i (x0 ) = Φi (x) + δf∗ Φi (x) (1.5)
∂µ Φi (x) → ∂µ0 Φ0i (x0 ) = ∂µ Φi (x) + δf∗ ∂µ Φi (x) (1.6)

note that δf∗ does not commute with the space-time derivatives.
When we act with this transformation the Action changes to
Z
A0Ω0 = d4 x0 L(Φ0i , ∂ 0 Φ0i ) . (1.7)
Ω0

If the transformation is a symmetry we have A0Ω0 − AΩ = 0, for any field


configuration, i.e. off-shell, therefore at the first order we obtain

0 = A0Ω0 − AΩ
Z h ³ ´ i
∂L ∗ ∂L
= d4 x (1 + ∂ρ δf xρ ) L(Φi , ∂Φi ) + δ f Φi + δf∗ ∂µ Φi − L
Ω ∂Φi ∂(∂µ Φi )
Z ³ ∂L ´
∂L
= d4 x δf∗ Φi + δf∗ ∂µ Φi + L∂ρ δf xρ (1.8)
Ω ∂Φi ∂(∂µ Φi )

where (1 + ∂ρ δf xρ ) is the Jacobian of the change of coordinates from x0 to


x at the first order.
Let us now introduce another variation δf that commutes with the deriva-
tives1 :
δf Φi = Φ0 i (x) − Φi (x) = Lf Φi , (1.9)
where Lf is the standard Lie derivative along the vector field f µ .
If we do so we can write

δf∗ Φi = ∂µ Φi (x)δf xµ + δf Φi
and δf∗ (∂µ Φi ) = ∂µ ∂ν Φi (x)δf xν + δf ∂µ Φi .
(1.10)
Substituting these back in (1.8) we obtain
Z h ³
´ ³ ∂L ´ i
∂L ∂L
d4 x ∂µ
δ f Φi + ∂µ Φi + ∂µ ∂ν Φi δf xµ + L∂µ δf xµ
Ω ∂(∂µ Φi ) ∂Φi ∂(∂ν Φi )
Z h ³ ∂L ´ ³ ∂L ´i
= d4 x ∂µ δ f Φi + δ f xµ
+ L∂µ δ f x µ
(1.11)
Ω ∂(∂µ Φi ) ∂xµ
Z
= d4 x(E.L.)i δf Φi ,

which finally gives the wanted conservation law on-shell ∂µ J µ = 0 where

J µ = Πµi δf Φi + Lδf xµ
1
Sometimes in literature, these transformations are referred to as “geometrical” trans-
formations.

7
Although this current is all we need to write our (conserved) quantities, to
make more transparent the appearance of known quantities we can write
back the space-time dependent variations δf∗ Φi we obtain

J µ = Πµi δf∗ Φi − (Πµi ∂ ν Φi − η µν L)δf xν


= Πµi δf∗ Φi − T µν δf xν (1.12)

that leads, for f µ = aµ , to the definition of the canonical (i.e. in general


non-symmetric, non-traceless, etc.) energy-momentum tensor T µν , and, for
f µ = ω µν xν , to the definition of the three rank tensor of angular momentum
M µνλ .
Noether theorem can also be stated in a different perspective, namely by
checking that δL = ∂µ V µ , which is only true for invariant actions, and then
writing the current as
Jµ = Nµ − V µ (1.13)
where, N µ ≡ Πµi δf∗ Φi is the rigid current, and we recall that in this case
δf∗ Φi = Φ0 i (x0 ) − Φi (x).
The choice δf∗ is the most useful in the case of gauge transformations and
Supersymmetry. In both cases it is not possible to write V µ = −Lδf xµ : for
SUSY that is because there is no δf xµ ; for gauge symmetry this is clearly
due to the fact that they are internal symmetries, hence do not involve
space-time transformations, and δf xµ = 0.

1.2 The issue of “improvements”


1.2.1 Maxwell’s − 41 F 2 . The canonical and symmetric energy
momentum tensor
See handwritten notes.

1.2.2 The scale vs conformal symmetry case


The virial current for the action A(Φ, ∂Φ), with Φ any field

Jµ = Πν Λµν Φ (1.14)

where Πµ = δA/δ∂µ Φ, Λµν = dΦ gµν + 2Σµν , and Σµν is the appropriate


spin-connection.
The general conditions for conformal invariance of a scale-invariant theory
are
for d ≥ 3 Jµ = ∂ ν Jµν for d = 2 Jµ = ∂µ J (1.15)

8
It is interesting to see what happens if we allow the kinetic term of the
Liouville field to be multiplied by scalar fields. Consider for example
Z µ ¶
1 ab h
2
i
A= d x η h(φ)∂a θ∂b θ + hαβ (φ)∂a φα ∂b φβ − eθ V (φ) (1.16)
2

where the φ-fields are conformal scalars and eθ has scale dimension deθ =
−2. In this case the virial current is Jµ = h(φ)∂µ θ and thus is not a total
derivative. It follows that although the action is rigid scale invariant it is
not conformally invariant, and if it is Weyl gauged the Weyl gauging cannot
be replaced by Ricci gauging. Thus, even for actions which are quadratic
in the derivatives of scalar fields, rigid scale invariance does not necessarily
imply conformal invariance.

1.3 How to deal with broken symmetry


With the current (1.1) one can: i) test whether the given transformation is
a symmetry by picking the correspondent f µ in (1.2), and checking whether
∂ J µ = 0, by using the equations of motion; ii) use the Noether charges Qf ≡
Rµ 3 0
d xJf , and the canonical equal-time Poisson brackets {Φi (x), Πj (y)} =
δij δ(~x − ~y ), to generate the transformations of an arbitrary function of the
canonical variables

{G(Φi , Πi ), Qf } ≡ ∆f G(Φi , Πi ) . (1.17)

Note also that, for f 0 = g 0 = 0 ,

{Qf , Qg } = Q[f,g] , (1.18)

and Eq.s (1.17) and (1.18) hold whether or not ∂0 Qf = 0. Of course, when
Qf acts on the fields it must reproduce the transformations one started with
∆f Φi = δf Φi .
Take the action
Z
1 1
Iˆ = − d4 x [F µν Fµν − θαβ Fαβ F µν Fµν + 2θαβ Fαµ Fβν F µν ] . (1.19)
4 2
We will learn more on this action later in the course. For the moment all
we need to know is that θαβ = −θβα and real.
The Noether currents for space-time transformations are

Jfµ = Πµν δf Aν + Lf µ , (1.20)

9
where the f µ s are given in (1.2), Πµν = δL/δ∂µ Aν , and, being Πµαβ =
δL/δ∂µ θαβ = 0, the transformations δf θαβ do not enter the Noether current.
Let us now analyze the symmetry properties by writing the divergence of
this current as
∂µ Jfµ = Πµν Fαν ∂µ f α − L∂µ f µ . (1.21)
We obtain
∂µ Jfµ = 0 translations (1.22)
for infinitesimal translations f µ = aµ , and

∂µ Jfµ = ωµα Πµν Fαν homogeneous Lorentz (1.23)

for infinitesimal rotations and boosts,

∂µ Jfµ = a(Πµν Fµν − 4L) dilations (1.24)

for infinitesimal dilations

∂µ Jfµ = 2Πµν Fαν (aα xµ − aµ xα − a · xδµα ) + 8a · xL special conformal (1.25)

infinitesimal special conformal transformations. For θµν = 0, Πµν = −F µν ,


and from (1.22)- (1.25) one immediately sees that ∂µ Jfµ = 0 for all f µ s
in (1.2), and the theory is invariant under the full conformal algebra (as
expected for classical electromagnetism!).
On the other hand, the action I, ˆ for θµν 6= 0, is only invariant under transla-
tions. Of course, this leaves room to special choices of the parameters and/or
of the field configurations, to obtain conserved currents. For instance, if one
performs two dependent infinitesimal boosts with parameters ω10 = ω20 = ω
, from the condition in (1.23) one obtains Π12 (E2 − E1 ) = E3 (Π13 + Π23 ).
Thus, for instance, one finds conservation if the electric field E ~ lives in the
(1, 2)-plane, and has equal components.
We now want to check the dynamical consistency of the transformations,
along the lines of what explained earlier. It is straightforward to see that
for Aµ and Fµν

∆f Aµ = δf Aµ and ∆f Fµν = δf Fµν , (1.26)

while for θµν


∆f θµν = 0 , (1.27)
for all f µ s in (1.2).

10
At this point one has to investigate whether θµν could be treated as a La-
grange multiplier (as e.g. in supersymmetric models). The answer in no. If
one uses
δ Iˆ
=0, (1.28)
δθµν
as a constraint this implies Fµν = 0, which is too trivial a theory (pure
gauge).
Note that if one uses (1.28) the expressions of ∂µ Jfµ in (1.22)-(1.25)are triv-
ially zero for all the space-time transformations.
We conclude that δf θµν cannot be fixed by any symmetry requirement (with
the only exception of translations), and that to have a physical meaningful
theory one should not make use of the “equations of motion” (1.28). There-
fore, among all possible δf θµν s that represent the conformal algebra, the
most natural choice is
δf θµν = ∆f θµν = 0
(which agrees with the translation invariance), and θµν does not transform
under dynamically consistent space-time transformations.

1.4 The quantum case


See handwritten notes.

1.5 Exercises
Exercise I.a Derive the angular momentum tensor Mµνρ for the Maxwell
theory − 14 F 2 and compare its energy momentum tensor with the one ob-
tained from translational symmetry.
Exercise I.b Prove that the virial current is a pure divergence for the
Maxwell theory (thus obtaining that scale invariance implies full conformal
invariance in this case).
Exercise I.c Suppose that the only nonzero components of θµν is θ3 (where
θ0i = 0 and θij = ²ijk θk with i, j, k = 1, 2, 3). In this case, which subgroup
of the Lorentz group SO(3, 1) survives in the theory analyzed in Section
1.3?
Exercise I.d Discuss the meaning of the quantum relation [H, Q] = 0, for
a theory with Hamiltonian H and a charge Q. Can you give an explicit
example?

11
1.6 Further Reading
E. Noether, Invarianten beliebiger Differentialausdrüke, Nachr. d. König.
Gesellsch. d. Wiss. zu Göttingen, Math-phys. Klasse (1918) 37;
J. Lopuszanski, “An introduction to Symmetry and Supersymmetry in quan-
tum field theory”, World Scientific, 1991;
A. Iorio, T. Sykora, On the Space-Time Symmetries of Noncommutative
Gauge Theories, Int.J. Mod. Phys. A 17 (2002) 2369;
A. Iorio, L. O’Raifeartaigh, I. Sachs and C. Wiesendanger Weyl-Gauging
and Conformal Invariance, Nucl. Phys. B 495 (1997) 433;
P. G. Federbush, K. A. Johnson, Phys. Rev. 120 (1960) 1926.

12
Chapter 2

Combining Spacetime and


Internal Symmetries

2.1 No-go Theorems


The ”no-go theorems” prove the impossibility of non-trivially combining
Lorentz invariance and internal symmetry for physical theories. The most
powerful is the Coleman-Mandula (CM) theorem [1]. The generalization of
this theorem lead to the discovery of supersymmetry.
The statement of the CM theorem is as follows: “Let E be a connected
symmetry group of the S matrix, and let the following five conditions hold:
(I) E contains a subgroup locally isomorphic to the inhomogeneous Lorentz
(Poincaré) group L; (II) all particle types correspond to positive-energy rep-
resentations of L, and, for any finite mass M , there are only a finite number
of particle types with mass less than M ; (III) elastic-scattering amplitudes
are analytic functions of the center of mass energy and of the momentum
transfer in some neighborhood of the physical region; (IV) at almost all ener-
gies, any two plane waves scatter; (V) the generators of E are representable
as integral operators in momentum space, with kernels that are distributions.
Then E is locally isomorphic to L × T , the direct product of the Poincaré
group and the internal symmetry group”.
In the 1960’s there were two kinds of motivations for investigating this prob-
lem in particle physics.
The first concerned the mass splitting occurring within the multiplets of
particles. The challenge was to find a group containing the internal symme-
try group, with non-trivial commutations among the generators of the latter
and those of space-time translations, Pµ . For example let us consider the

13
isospin group, whose SU(2) algebra we write in the Cartan-Weyl basis

[T+ , T− ] = 2T0 and [T0 , T± ] = ±T± , (2.1)

where T+ (T− )is the step-up (step-down) operator, and T0 is the generator
of the Cartan sub-algebra. The commutator [Pµ , T+ ] is zero for ordinary
isospin symmetry. However, if it is different from zero within the bigger
group, this would give account for the mass splitting as a higher symmetry
effect. Consider the doublet of nucleons |ni ≡ |+i and |pi ≡ |−i as the two
states of the fundamental irrep of SU(2)

T0 |±i = ±|±i , T± |±i = 0 , T∓ |±i = |∓i . (2.2)

By our hypothesis [Pµ , T+ ] = 0 ⇒ P 2 |±i = m2 |±i, which means m2p =


m2n . Hence the experimental results mp ∼ 938.3MeV and mn ∼ 939.6MeV,
cannot be explained. Suppose, instead, that

[Pµ , T+ ] = cµ T+ (2.3)

where cµ is orthogonal to Pµ (cµ P µ = 0) and commutes with the other


generators
³ ´
P 2 |+i = P 2 T+ |−i = T+ P 2 |−i + 2cµ P µ |−i + c2 |−i = T+ m2n |−i (2.4)

or
P 2 |−i = (m2n − c2 )|−i . (2.5)
i.e. the mass splitting
m2p = m2n − ∆m2 , (2.6)
where ∆m2 ≡ c2 . Nowadays the accepted explanation of the mass-splitting
phenomenon is the breakdown of flavor symmetry.
The second motivation to investigate non trivial combinations of space-time
and internal symmetries was the discovery of a model where the internal 3-
flavor symmetry group SU (3) and the non-relativistic spin group SU (2) were
non-trivially combined into SU (6), which contains but is not isomorphic to
SU (3) × SU (2). This gives the so-called static quark model. As a matter of
+
fact the barion octet J P = 12

n p
Σ− Σ0 /Λ Σ+
Ξ− Ξ0

14
3+
and decuplet J P = 2

∆− ∆0 ∆+ ∆++
Σ∗− Σ∗0 Σ∗+
,
Ξ∗− Ξ0
Ω−

which differ by spin, both fit into a 56-plet of SU(6). This is easily seen if
one considers that the dimension of the representation has to be d × (2J +
1), where d is the dimension of the SU(3) representation, and J the non-
relativistic spin. In this case: i) d = 10 and 2J + 1 = 4 gives 40, while ii)
d = 8 and 2J + 1 = 2 gives 16, which add up to 56. On the other hand the
tensorial representation of SU(6): 6 × 6 × 6 = 56 + 70 + 70 + 20. Similar
considerations hold for the meson nonets of J P = 0− and J P = 1− . The
natural task then was to extend this result to a fully relativistic theory.
These programs were brought to a negative end first by the O’Raifeartaigh
(LOR) theorem [2] which completes and generalizes the results of the work
started with the first “no-go” theorem of McGlinn (McG) [3], and later
by the CM theorem stated above. We shall later prove and discuss LOR
theorem in some detail. All these theorems hold if one considers Lie groups
as symmetry groups of the theory (LOR theorem holds for finite order Lie
algebras, while CM theorem holds also for the infinite case) and are of
local nature. Nevertheless if one relaxes the assumption of having only
standard Lie groups, for instance by allowing for graded structures, then the
negative-type conclusions no longer hold [4]. The most surprising feature of
these graded algebras is the occurrence of transformations among particles
differing by spin: this is the birth of supersymmetry. In Ref. [5] the most
general supersymmetric algebra of the S matrix was introduced and its
representations extensively studied, closing the era of the “no-go” theorems
with a “let’s go” theorem: the Haag-Lopuszánski-Sohnius (HLS) theorem.
(Note that the title of the paper in Ref. [5] is “All Possible Generators of
Supersymmetry of the S Matrix” as opposed to the title of the paper in Ref.
[1] “All Possible Symmetries of the S Matrix”). In the following we shall
state McG theorem, state and discuss LOR theorem, and finally prove and
discuss CM theorem.

2.2 O’Raifeartaigh Theorem(s)


LOR theorem. “Let L be the Lie algebra of the inhomogeneous Lorentz
group, consisting of the homogeneous part M and the translation part P .
Let E be any Lie algebra of finite order, with radical R and Levi factor G. If

15
L is a subalgebra of E, then only the following four cases occur: (i) R = P ;
(ii) R Abelian but larger than, and containing P ; (iii) R solvable but not
Abelian, and containing P ; (iv) R ∩ P = 0. In all cases, M ∩ R = 0”. The
main algebraic tools used in this theorem are the Levi decomposition and the
freedom of redefining the generators (the redefinitions have trivial physical
consequences). Levi’s radical-splitting theorem in Lie algebra theory states
that any Lie algebra E of finite order is the semidirect sum of a semisimple
algebra G (the Levi factor) and the radical R (an invariant solvable algebra,
where solvable means that for some integer k the k-derived algebra is zero).
LOR theorem enables one to classify the ways in which L can be embedded
in E. Case (i) is the only physical case and, up to a redefinition, reduces
to E = L ⊕ T , where T is a semisimple algebra (internal symmetry). Case
(ii) cannot be reduced to the previous direct sum but is unphysical since
it introduces a translation algebra of more than four dimensions. Case
(iii) is the most unphysical since, for non-trivial representations, hermitian
conjugation cannot be defined. Case (iv) amounts, up to a redefinition, to
embedding L as a subalgebra in a simple Lie algebra. It is again unphysical
due to the fact that the parameters corresponding to the Pµ have a non-
compact range and this lead to serious difficulties in defining multiplets,
even in the absence of mass-splitting. Thus, while it is possible to embed
L in a larger algebra E, the ways in which this may be done are restricted
and only the direct sum has a clear physical meaning. The McG result
can be obtained as a special case of LOR theorem by using the first McG
assumption alone and the redefinition freedom.
If one now considers the Hilbert space H on which any irreducible represen-
tation of the group generated by E operates, and if the mass operator P 2
has a discrete eigenvalue m2 and is self-adjoint on H, then the eigenspace
Hm belonging to m2 is closed and is invariant with respect to the elements
representing E on H. Hence the elements representing E cannot produce
the mass-splitting. Sometimes in literature this result (the mass-splitting
theorem) is referred to as the LOR theorem.

2.3 Exercises
Exercise II.a Show that the product of 3 fundamental (vector) represen-
tations of SU(6) boils down to the sum of 4 tensorial (symmetric, antisym-
metric and mixed) irreps given in the text 6 × 6 × 6 = 56sym + 70mix +
70mix + 20asym . (Hint: use the Young tableaux’ rules for SU(6).)
Exercise II.b Explain why for non-relativistic (Galilean invariant) theories
the O’Raifeartaigh theorem does not hold. (Hint. This has to do with the

16
compactness of the space-time group.)

2.4 Further Reading


[1] S. Coleman and J. Mandula Phys. Rev. 159 (1967) 1251;
[2] L. O’Raifeartaigh Phys. Rev. 139 (1965) 1052;
[3] W. D. McGlinn Phys. Rev. Lett. 12 (1964) 467;
[4] Yu. A. Gol’fand and E. P. Likhtman JETP Lett. 13 (1971) 323;
[5] R. Haag, J. Lopuszánski and M. Sohnius Nucl. Phys. B 88 (1975) 257;
[6] S. Weinberg, The Quantum Theory of Fields, Vol. III Supersymmetry,
(Chp. 24 ≡ first Chp.)

17
Chapter 3

The (special) case of SUSY

3.1 The Haag-Lopuszansky-Sohnius Theorem


3.1.1 Building up the Super-Poincarè Algebra.
Let us relax one of the conditions of the CM theorem: introduce anticom-
muting generators as possible symmetries of the S matrix. This corresponds
to consider a graded algebra with odd generators

Q, Q0 , Q00

besides the ordinary even ones

X, X 0 , X 00 ,

with the algebra among them being given by

[Q, Q0 ]+ = X [X, X 0 ]− = X 00 [Q, X 0 ]− = Q00

where the commutator (anti-commutator) is denoted by [, ]− ([, ]+ ). It is


important to realize that the CM theorem will not be proved false, it will
rather be generalized. Hence within the super-group we will indeed be able
to mix non-trivially internal and space-time symmetries, but in a more subtle
way than [Pµ , T+ ] 6= 0, and the CM results can still be used. For instance
the even generators have to be

L=M−
DP
X∈ or (3.1)
A = A1 ⊕ A2 ,

18
where L is the Poincarè algebra, semidirect product of Lorentz and transla-
tions, while the elements of the algebra of internal symmetry A are Lorentz
scalars, A1 is semi-simple, and A2 is Abelian.
From the point of view of Lorentz properties
Pµ transforms under the ( 12 , 12 ) vector irrep.

Mµν transforms under the (1, 0) + (0, 1) second-rank tensor irrep.

the elements of A transform under the (0, 0) scalar irrep.


The odd generators instead
X
Q= Qα1 ...αa ,α̇1 ...α̇b (3.2)

transform under the spin 12 (a + b) irrep of Lorentz, and, of course (a + b)


is odd. Suppose that Q̄ belongs to the algebra too, and consider the spin
(a + b) object
Q Q̄
| {z1̇} 1̇...
| {z } , 1̇...
1...1 | {z1̇} , |1...1
{z }
a times b times a times b times

If this product has to belong to the algebra hence (a + b) can only be either
0 or 1, because
[Q, Q̄]+ = X
and the spin (a+b) X can either be a scalar (in A) or a vector under Lorentz
(the second-rank tensor is ruled out by the hypothesis that Q is odd). Hence
the first anti-commutation relation of this algebra is easily written as
µ
[QL L
α , Q̄α̇M ]+ = 2σαα̇ Pµ δM , (3.3)

where L, M = 1, . . . , N , and for N > 1 we say that Susy is extended. Note


the space-time nature of the supercharges.
In order to construct the other relations one (cleverly) uses the following
super Jacobi identities

[A, [B, C]± ]± ± [B, [C, A]± ]± ± [C, [A, B]± ]± = 0 , (3.4)

where the use of the commutator [, ]− or the anti-commutator [, ]+ depends


on the even/odd nature of the generators involved, and the signs among the
terms depends on the number of commutations of the odd generators.
Let us look at the relation

[ QL M
α , Qβ ]+ = ²αβ Z
[LM ]
| {z } + M(αβ) Y
(LM )
(3.5)
|{z} |{z} | {z }
( 12 ,0) ( 1 ,0) (0,0)
(1,0)
2

19
where we indicate the properties under Lorentz. Since there are no spin 32
generators we have
[ QL
α , Pµ ]− = 0 (3.6)
|{z} |{z}
( 12 ,0) ( 12 , 12 )

and, by using the super Jacobi identities

0 = [Pµ , [QL M ρσ (LM )


α , Qβ ]+ ]− = [Pµ , Mρσ ]− σ Y ⇒ Y (LM ) = 0

since [Pµ , Mρσ ]− 6= 0. We can now write the N-extended Susy algebra

µ
[QL L
α , Q̄α̇M ]+ = 2σαα̇ Pµ δM (3.7)
[QL M
α , Qβ ]+ = ²αβ Z
[LM ]
(3.8)

[Q̄α̇L , Q̄β̇M ]+ = ²α̇β̇ Z[LM ] (3.9)
[QL
α , Pµ ]− = [Q̄α̇L , Pµ ]− = 0 (3.10)
[QLα , Bl ]− = (Sl )L
M Qα
M
(3.11)
l
[B , Q̄α̇L ]− = (S ∗l )M
L Q̄α̇M (3.12)
k
[Bl , Bm ]− = iClm Bk (3.13)

where the Bl s are the internal symmetry generators belonging to A1 , and


the Poincarè ”sector” of the algebra has been omitted. Note that, if one
wants to consider the largest (super) symmetry of the S-matrix, the full
conformal group has to be included1 .
By using the super-Jacobi for Bl , QL M
α , Qβ , it is easy to prove that the Z
[LM ] s

close and invariant sub-algebra of A = A1 + A2 . By using the super-


Jacobi for QL M
α , Qβ , Q̄γ̇K , one easily proves that [Q̄γ̇K , Z
[LM ] ]
− = 0 and
1 βα K N [LM ] [KN ] [LM ]
2 ² [[Qα , Qβ ]+ , Z ]− = [Z ,Z ]− = 0, hence, since from the hy-
pothesis A1 is semi-simple, the Z [LM ] s span A2 , the Abelian sub-algebra of
A, [Bl , Z [LM ] ]− = 0.
The generators Z [LM ] are central charges of the Susy algebra. They play a
major role in Susy gauge theories, where the Higgs mechanism to give mass
can actually take place within Susy, if the central charges are non-zero (see
next Section). It is important to note that, even when at the algebraic level
it is possible to have Z [LM ] 6= 0, as in the N=2 case, is the actual physical
realization of the algebra (i.e. the underlying model) that tells us if the
central charges are different from zero. For instance, in the phase where the
1
This means that besides the generators of the Poincarè group, Pµ and Mµν , also the
dilation generator D, and the special conformal generators Cµ have to be considered. See
also the Chapter on Noether currents.

20
gauge symmetry is not broken, even the N=2 SYM possesses trivial central
charges.
Finally, in the Exercises of this Chapter we propose to prove that the (Sl )L
Ms
form a representation of A.

3.1.2 Irreducible Representations and Supermultiplets.


We have now a fully relativistic (even conformal) way to combine space-time
and internal symmetries. We already saw that in the static non-relativistic
quark model the SU(6) (broken) symmetry combined together particles dif-
fering by spin (for instance the spin 1/2 octet and the spin 3/2 decuplet of
baryons). We have now a more powerful way of doing that within the Susy
multiplets.
Firstly we notice that for any finite dimensional representation of Susy (so
that the trace is well defined), we can prove that there are an equal number
of bosonic and fermionic degrees of freedom. Note here that we do not need
to have an equal number of bosonic and fermionic particles, but only of
degrees of freedom. To prove that it is just matter of defining a fermion
number operator NF , so that

(−)NF |BOSE >= +1|BOSE > and (−)NF |FERMI >= −1|FERMI >

We can now compute the following trace

Tr{(−)NF [QL
α , Q̄α̇M ]+ } = Tr{(−)
NF L
Qα Q̄α̇M + (−)NF Q̄α̇M QL
α}
= Tr{−QL
α (−)
NF
Q̄α̇M + QL
α (−)
NF
Q̄α̇M } = 0

where the minus sign in the first term of the second line is due to the fact
that Qα transforms bosons ↔ fermions, and the second term in the same
line is due to the cyclicity of the trace. By using the first relations of the
Susy algebra we then have

2σαµα̇ δM
L
Tr{(−)NF Pµ } = 0 ⇒ Tr{(−)NF } = 0 (3.14)

which proves our statement,


 
+1
 .. 
 . 
 
 +1 
NF  
(−) =  . (3.15)
 −1 
 
 .. 
 . 
−1

21
The irreducible representations (irreps) of extended Susy are easily found in
terms of suitable linear combinations of the supercharges QL α , L = 1, ..., N ,
to obtain creation and annihilation operators acting on a Clifford vacuum
Ω0 . There are three possible cases: I) M 6= 0 and Z = 0; II) M = 0 and
Z = 0; III) M 6= 0 and Z 6= 0.
I). Massive, Central-Charge-less
We move in the rest frame Pµ = (−M, 0, 0, 0), and write the Susy algebra
as
[QL M † L β
α , (Qβ ) ]+ = 2M δM δα (3.16)
[QL M L † M †
α , Qβ ]+ = [(Qα ) , (Qβ ) ]+ = 0 (3.17)

where L, M = 1, N . By defining aL
α ≡ (2M )
−1/2 QL , and (aL )† ≡ (2M )−1/2 Q̄ ,
α α α̇L
we have 2 (for α) × N (for L) = 2 N fermionic annihilation/creation opera-
tors, in terms of which the Susy algebra above becomes
†M †L †M
[aL β L
α , aβ ]+ = δα δM [aL M
α , aβ ]+ = [aα , aβ ]+ = 0 . (3.18)
The generic state is then given by
α1 1
Ω(n) A1 · · · αAnn = √ (aαA11 )† · · · (aαAnn )† Ω0 , (3.19)
n!
where
à a!A
α Ω0 = 0. Of course, the multiplicity (degeneracy) of this state is
2N
. Thus the dimension of this irrep is
n
2N
à !
X 2N
dI = = 22N . (3.20)
n=0
n

II). Mass-less, Central-Charge-less


To find the dimension of this irrep it suffices to notice that Pµ = (−E, 0, 0, E)),
hence à !
L 2E 0 L
[Qα , Q̄α̇M ]+ = 2 δM , (3.21)
0 0
αα̇
and zero the others. This means that only one element on the right hand
side is different from zero, while in the massive case there are two entries of
δαβ different from zero. Thus the number of annihilation/creation operators
is reduced to one half with respect to the massive case, and we can conclude
that the dimension of this irrep is
N
à !
X N
dII = = 2N , (3.22)
n=0
n

22

or dII = dI . The discrepancy of the dimension of the massive and mass-
less irreps causes problems in the Higgs mechanism. To cure it one has to
consider the next case.
III). Massive, Central-Charge
In the rest frame we can write

[QL M † L β
α , (Qβ ) ]+ = 2M δM δα (3.23)
[QL M
α , Qβ ]+ = ²αβ Z
[LM ]
(3.24)
[(QL
α)

, (QM †
β ) ]+ = ² αβ ∗
Z[LM ] (3.25)

where L, M = 1, ..., N . Let us consider N even, and take L ≡ (a, l), M ≡


(b, m), where a, b = 1, 2, and l, m = 1, ..., N/2.
By performing a unitary transformation on the QL α we can introduce new
charges Q̃L
α = U L QM so that in this basis Z [LM ] is mapped to ²ab 2|Z |,
M α n
where Zn = |Zn |eiζn , |Zn | ≥ 0, ζn = 0, n = 1, ..., N/2.
We can now define the following annihilation operators
1
alα = √ (Q̃1l 2l †
α + ²αγ (Q̃γ ) ) (3.26)
2
1
blα = √ (Q̃1l 2l †
α − ²αγ (Q̃γ ) ) (3.27)
2

and their conjugates a†α and b†α , in terms of which we can write the algebra
as

[alα , am l m l m
β ]+ = [bα , bβ ]+ = [aα , bβ ]+ = 0 (3.28)
[alα , am† lm
β ]+ = δ δαβ 2(M + |Zn |) (3.29)
[blα , bm† lm
β ]+ = δ δαβ 2(M − |Zn |) (3.30)

For α = β the anticommutators (3.29) and (3.30) are never less than zero
on any states. Therefore from (3.29) follows M + |Zn | ≥ 0 and from (3.30)
follows M − |Zn | ≥ 0. By multiplying these two inequalities together we
obtain
M ≥ |Zn | (3.31)
The states for which the inequality is saturated are called Bogomolnyi-
Prasad-Sommerfield (BPS) states, and we have the so-called short Susy
multiplet. The shortness is easily seen by considering a certain number
r ≤ n of BPS states, for which M = |Zi |, i = 1, ..., r. From Eq. (3.30) one
immediately sees that 2r operators blα must vanish. Thus, if r = n = N/2
all the N/2 operators blα must vanish. This reduces the number of creation

23
and annihilation operators of the Clifford algebra from 2N to N . Therefore
the dimension of the massive representation reduces to the dimension of the
massless one: from 22N to 2N , and we can implement the Higgs mechanism
without breaking Susy: dBPS
III = dII .

3.2 General features of SUSY theories


First let us discuss an example were Susy appears in the most simple way:
Susy Quantum Mechanics. These results are due to Witten. For a nice
account see Ref. [1].
Consider the motion in the direction x of an electron in a magnetic field
directed along z but function only of x: Bz (x). The hamiltonian

1 ~ 2 + i qe ∇ ~ + |qe | ~σ · B
~ ·A ~,
H= (~
p − qe A) (3.32)
2m 2m 2m
where qe and m are the charge and mass of the electron, respectively, and
the natural units are used h̄ = c = 1, with the choice

m
Ax = 0 = Az Ay = W (x) (3.33)
|qe |

becomes à !
1 p2 1 dW
H= + W 2 (x) + σ3 √ . (3.34)
2 m m dx
Now define
1 p 1 p
Q1 ≡ √ (σ1 √ + σ2 W ) Q2 ≡ √ (σ2 √ − σ1 W ) , (3.35)
2 m 2 m

and discover that Susy is on its way

[Qi , Qj ]+ = 2δij H [H, Qi ] = 0 i, j = 1, 2 . (3.36)

What we have done is to consider ”the square root” of the hamiltonian, in the
same spirit of how Dirac considered the ”square root” of the Klein-Gordon
equation (2 + µ2 )φ = 0 to obtain his (spinorial) equation (i 6 ∂ + µ)ψ = 0.
There are two important considerations to be done here: i) this toy model
Susy shares many of the properties of Susy in field theory; ii) here Susy
is implemented in a different spirit respect to the fundamental approach of
considering it as a symmetry of the S matrix. Here we list other areas of
physics where Susy is (or could be) implemented in this spirit:

24
* Nuclear Physics (Iachello)

* Chaotic Systems (Efetov)

* Superconductivity (Nambu)
On a different footing are the following areas, where the Susy scenario has
not been experimentally seen
0 Standard Model

0 Gravity (Supergravity)
After this brief overview we are ready to discuss the general features of
Susy models, based on the algebraic structure of the symmetry itself. The
hamiltonian of any Susy theory is given in terms of the Susy charges:

4H = σ̄ 0αα̇ [Qα , Q̄α̇ ]+ = (Qσ 0 Q̄ + Q̄σ̄ 0 Q) ≥ 0 (3.37)

on any physical state, where the N = 1 case is considered. In particular on


the vacuum |0i
< 0|H|0 >= 0 i.e. H =: H : . (3.38)
This can be seen by thinking of Susy as a space time symmetry implemented
by the group element of the super-Poincarè group

G(x, θ, θ̄) = exp{i(−xµ Pµ + θQ + θ̄Q̄)} . (3.39)

In gauge theory without SSB the invariance of a model under a given gauge
group G is implemented on the action (by construction) and on the vacuum
aT a
G|0i = eα |0i = |0i

hence, T a |0i = 0, where T a , with a = 1, ..., dimG, are the generators of the
group. Similarly in a Susy theory

Qα |0i = 0 Q̄α̇ |0i = 0 , (3.40)

which proves our statement about the automatic normal ordering of the
hamiltonian.
This is a first signal of the nice behavior of Susy theory under renormaliza-
tion. As a matter of fact, the normal ordering for a quantum field consists
in the subtraction of the infinite vacuum energy. For a one-dimensional
harmonic oscillator (H.O.)
1 h̄ω
H|ni = (n + )h̄ω|ni → H|0i = |0i
2 2

25
this subtraction is finite, and allowed. In (free) field theory the vacuum
energy one is discarding with the normal ordering is infinite
X h̄ωk
Evaccum = → ∞.
k
2

Already Pauli in the 1950s realized that in a theory


q with an equal number
of bosons and fermions with equal masses ωkbose = m2 + ~k 2 = ωkfermi ≡ ωk
the vacuum energy would have been automatically zero, just consider that
a fermionic oscillator has negative vacuum energy
X
HF = h̄ωkfermi (fk† fk − 1/2) (3.41)
k

and combine it with the bosonic partner (always there in Susy theories)
X
HB = h̄ωkbose (b†k bk + 1/2) (3.42)
k

to obtain X
H = HB + HF = h̄ωk (b†k bk + fk† fk ) =: H : (3.43)
k
The example above given holds for free fields. In fact, the important point
here is that this phenomenon holds no matter how complicated is the inter-
action, even for effective theories, as long as Susy is a symmetry. This is
particularly important if one considers that for interacting field theories the
vacuum-to-vacuum (or v2v) graphs < 0|0 > are, in principle, very nastily
divergent. I say in principle because in standard perturbative approaches
to QFT it is assumed that < 0|0 >= 1. We will see that this is a strong
requirement, not really satisfied by interacting quantum fields (Haag’s theo-
rem), especially in the relativistic regime (as was noticed by Dirac who put
it, more or less, this way “The relativistic interaction is to strong to keep the
incoming particles within the same Hilbert space as the outgoing particles.
This is sort of working for feeble interactions, while for fully relativistic ones
the state representing the incoming particle is kicked out of the Hilbert space
to a different one” [see the Introduction of Dirac’s lectures on quantum field
theories].
It is not surprising then that Susy theories have nice renormalization prop-
erties. These properties are exploited in full details in the so-called ”non-
renormalization theorems” mostly due to Seiberg, even if a better way of
calling these theorems would be ”no-need-to-renormalize theorems”. The
idea is again based on the fermi-bose symmetry, which one wants to im-
plement also at the level of the Feynman graphs. For instance, in Susy

26
¯ µ Aµ sψ, where
QED, the vertex qe ψ̄γ µ Aµ ψ has as Susy counterpart qe sψγ
sψ represents the super-partner of the electron ψ within the supermultiplet,
hence has the same mass, but bosonic statistics. The fermionic loop is then
cancelled by the bosonic one (same coupling qe and mass).
For instance the perturbative contributions to Quantum Super Yang-Mills
(SYM) theories are
N=1 Tree Level + 1-Loop + ... (well behaved)
N=2 Tree Level + 1-Loop
N=4 Tree Level
hence there are no quantum corrections at all to the N=4 classical SYM!
Note that the instanton, non-perturbative, contributions have not been con-
sidered.
Note also that in bosonic string theory (the first string theory after Veneziano’s
hadronic string) the number of space-time (target space) dimensions had to
depart from d = 4 to d = 26 to remove a quantum anomaly. This cancelation
takes place, instead, at d = 10 when Susy is implemented.
Let us conclude this overview with same nomenclature of the super-partners
of the particles within the Susy Standard Model and Gravity:
• fermions: electron, quarks, ... → sfermions: selectron, squarks, ...
• gauge bosons → gauginos: photino, gluino, ...
• graviton → gravitino
the exception is the neutrino which has as super-partner the neutralino.

3.3 The Wess-Zumino model and its (Noether) su-


percharges
The Lagrangian density and supersymmetric transformations of the fields
for the Wess-Zumino model are given by
i i ¯ m m
L = − ψ 6 ∂ ψ̄ − ψ̄ 6 ∂ψ − ∂µ φ∂ µ φ† + F F † + mφF + mφ† F † − ψ 2 − ψ̄ 2
2 2 2 2
(3.44)
and
√ √
δφ = 2²ψ δφ† = 2²̄ψ̄ (3.45)
√ µ √ √ µ α̇ √ α̇ †
α̇ †
δψα = i 2(σ ²̄)α ∂µ φ + 2²α F δ ψ̄ = i 2(σ̄ ²) ∂µ φ + 2²̄ F (3.46)
√ √
¯
δF = i 2²̄ 6 ∂ψ δF † = i 2² 6 ∂ ψ̄ (3.47)

27
where φ is a complex scalar field, ψ is its Susy fermionic partner in Weyl
notation and F is the complex bosonic dummy field.
As explained earlier, in this case one has to use the expression

Jµ = Nµ − V µ

to compute the Noether supercurrents. A note on partial integration in the


fermionic sector of (3.44) is in order. We see that ψ and ψ̄ play the double
role of fields and momenta at the same time. It is just a matter of taste to
choose Dirac brackets for this second class constrained system or to partially
integrate to fix a proper phase-space and implement the canonical Poisson
brackets.
If one chooses the canonical Poisson brackets, then it is only a matter of
convenience when to partially integrate the fermions. In fact, even if two
parts of the current N µ and V µ both change under partial integration, the
total current J µ is formally invariant, namely its expression in terms of fields
and their derivatives is invariant but the interpretation in terms of momenta
and variations of the fields is different. Of course both choices give the same
results, therefore one could either start by fixing the proper phase space
since the beginning or just do it at the end.
Let us keep (3.44) as it stands, define the following non canonical momenta

µ i
πψα = (σ µ ψ̄)α πψ̄µα̇ = 2i (σ̄ µ ψ)α̇ (3.48)
2
πφµ = −∂ µ φ† πφµ† = −∂ µ φ (3.49)

and use J µ = N µ − V µ to obtain the current.


We compute V µ by varying (3.44) off-shell, under the given transformations,
obtaining
† †
V µ = δφπφµ + δφ† πφµ† − δ φ ψπψµ − δ φ ψ̄πψ̄µ + δ F ψπψµ + δ F ψ̄πψ̄µ

−2δ Fon ψπψµ − 2δ Fon ψ̄πψ̄µ (3.50)

where δ X Y stands for the√part of the variation of Y which contains X (for


instance δ F ψ stands for 2²F ) and Fon , Fon † are the dummy fields given

by their expressions on-shell (F = −mφ , F † = −mφ). Note here that we


succeeded in finding an expression for V µ in terms of π µ ’s and variations of


the fields. Note also that the terms involving Fon and Fon † were obtained

without any request but they simply came out like that.
Then we write the rigid current

N µ = δφπφµ + δφ† πφµ† + δψπψµ + δ ψ̄πψ̄µ (3.51)

28
and the full current is given by
J µ = N µ − V µ = 2(δ on ψπψµ + δ on ψ̄πψ̄µ ) (3.52)

therefore J µ = 2(N µ )on fermi , with obvious notation. In the bosonic sector
N µ completely cancels out against the correspondent part of V µ . In the
fermionic sector δ F ψπψµ in N µ cancels out against the term coming from
V µ , δ φ ψπψµ in N µ and in V µ add up and combined with the 2δon F ψπ µ in V µ
ψ
gives 2δon ψπψµ in the full current J µ . Similarly for ψ̄.
Therefore we conclude that:
a the dummy fields are on-shell automatically2 ;
and, if we keep the fermionic non canonical momenta given in (3.48),
b the full current is given by twice the fermionic rigid current (N µ )on
fermi .
While the first result is general, the result b is only valid for simple
Lagrangians and it breaks down for less trivial cases. It is interesting to see
for which class of theories it holds.
Now we rewrite J µ in terms of fields and their derivatives

J µ = 2(ψ̄σ̄ µ σ ν ²̄∂ν φ + i²σ µ ψ̄Fon + h.c.) (3.53)
then choose one partial integration
√ √
J µ = δon ψπ µIψ + 2ψσ µ σ̄ ν ²∂ν φ† + i 2²̄σ̄ µ ψFon †
(3.54)
√ √
or = 2ψ̄σ̄ µ σ ν ²̄∂ν φ + i 2²σ µ ψ̄Fon + δon ψ̄π µII
ψ̄ (3.55)

where π µIψ = iσ µ ψ̄ (π µIψ̄ = 0) and π µII


ψ̄
= iσ̄ µ ψ (π µII
ψ = 0) are the canonical
momenta obtained by (3.44) conveniently integrated by parts, and perform
our computations using canonical Poisson brackets.
Choosing the setting I, for instance, what is left is to check that the charge
Z Z ³ √ √ ´
Q ≡ d3 xJ 0 (x) = d3 x δon ψπψI + 2ψσ 0 σ̄ ν ²∂ν φ† + i 2²̄σ̄ 0 ψFon †
(3.56)

correctly generates the transformations. This is a trivial task in this case


since the current and the expression of the dummy fields on-shell are very
simple and the transformations can be read off immediately from the charge
(3.56). It is worthwhile to notice at this point that to generate the transfor-
mations of the scalar field φ† one has to use

δ φ ψπ µIψ = δφ† πφ† + 2ψ̄σ̄ 0 σ i ²̄∂i φ (3.57)
2
As a matter of fact using Noether currents we are extracting dynamical information
out of a symmetry realized algebraically via the dummy fields, hence the procedure itself
forces supersymmetry to be realized at a dynamical level putting those non-dynamical
fields on shell.

29
Notice also that the transformation of ψ̄ is obtained by acting with the
charge on the conjugate momentum of ψ: {Q , πψI }− .

3.4 N=2 Susy Classical Georgi-Glashow model and


the Mass Formula
Edward Witten and David Olive (Phys. Lett. B 78 (1978) 97) considered
the classical N=2 supersymmetric Georgi-Glashow action with gauge group
O(3) spontaneously broken down to U(1) and its supercharges3 .
The action for this SU(2) N=2 Super-Yang-Mills (SYM) theory is
Z
1 1 a aµν
A = d4 x{− [ F F + Dµ φa Dµ φa† + iψ a 6 Dψ̄ a + iλa 6 Dλ̄a ,
g 2 4 µν
1 1 ϑ
+ √ (ψ a [λ, φ]a + h.c.) − ([φ, φ† ]a [φ, φ† ]a )] − F a F ∗aµν } .
2 µν
2 2 64π

where a = 1, 2, 3 is the gauge group index, SU(2) in our case; for each a, φ
is the scalar field, ψ, λ are two Weyl fermionic fields, Aµ is a vector field.
The Susy transformations of these fields are
first Susy, parameter ²1


δ1 φ~ = 2²1 ψ~

~α =
δ1 ψ ~
2²α1 E (3.58)
~ = 0
δ1 E

√ ~ + 2i[φ
δ1 E~ † = i 2²1 6 Dψ̄ ~ † , ²1~λ]
~ √ α
δ1 ψ̄ ~†
α̇ = −i 2²1 6 Dαα̇ φ (3.59)
~† = 0
δ1 φ

δ1~λα = −²β1 (σ µν α~ α~
β Fµν − iδβ D)
~ µ = i²1 σ µ~λ̄
δ1 A ~ = −²1 6 D~λ̄
δ1 D (3.60)
δ1~λ̄α̇ = 0

second Susy, parameter ²2


3
We shall consider the gauge group SU(2), since all the local considerations are the
same.

30

~ = − 2²2~λ
δ2 φ

δ2~λα = − 2²α2 E
~† (3.61)
~† = 0
δ2 E


δ2 E~ = −i 2²2 6 D~λ̄ + 2i[φ
~ † , ²2 ψ]
~

δ2~λ̄α̇ = i 2²α2 6 Dαα̇ φ
~† (3.62)
δ2 φ~† = 0

~ α = −²β (σ µν α F~µν + iδ α D)
δ2 ψ ~
2 β β

δ2 A ~
~ µ = i²2 σ µ ψ̄ ~
~ = ²2 6 Dψ̄
δ2 D (3.63)
~
δ2 ψ̄ α̇ = 0

We used dummy fields E, D explicitly in the transformations, and the


~ = 1 σ a X a with a = 1, 2, 3, we follow
generic SU(2) vector is defined as X 2
the summation convention, and the covariant derivative and the vector field
strength are given by

Dµ X a = ∂µ X a + ²abc Abµ X c (3.64)

and
a
Fµν = ∂µ Aaν − ∂ν Aaµ + ²abc Abµ Acν (3.65)
Note that the last term is the action above is the instanton term. This
is a pure divergence of the form ∂µ K µ , where

K µ ∼ ²µνρσ (Aaν ∂ρ Aaσ + ²abc Aaν Abρ Acσ )

and when integrated in the action gives the Pontryagin index. Note also,
that due to the pseudo-tensor (axial) nature of the dual Fµν ∗ , the time-

reversal symmetry is broken. That is why the ϑ-angle is related to CP


violation. On this point see, for instance, Ryder’s last chapter, and Wein-
berg’s first (discrete symmetries C, P, T) and second (topological objects in
QFT, in particular the QCD ϑ-vacua) volumes.
The central charge for this theory was obtained by Witten and Olive
√ Z 2
Z=i 2 d Σ ~ a φa + 1 B
~ · (Π ~ a φaD ) a = 1, 2, 3 (3.66)

31
~ is the measure on the sphere at spatial infinity S∞
where d2 Σ 2 , the φa ’s
~ a ’s are the magnetic fields, Π
are the scalar fields, the B ~ a is the conjugate
momentum of the vector field A ~ and φ ≡ τ φ , with
a a a
D

ϑ 4π
τ= +i 2 , (3.67)
2π g
the complex coupling constant whose real and imaginary part are related
to the CP violating ϑ-angle of the ϑ-vacua, and to the SU(2) coupling,
respectively.
In the classical case φaD is merely a formal quantity with no precise
physical meaning. On the contrary, in the low-energy sector of the quantum
theory, it becomes the e.m. dual of the scalar field.
In the unbroken phase Z = 0, but, as well known, in the broken phase
this theory admits ’t Hooft-Polyakov monopole solutions. In this phase the
a
scalar fields (and the vector potentials) tend to their vacuum value φa ∼ a rr
b
(Aai ∼ ²iab rr2 , Aa0 = 0), where a ∈ C, as r → ∞. This behavior gives rise
to a magnetic charge. By performing a SU(2) gauge transformation on this
radially symmetric (“hedgehog”) solution we can align < 0|φa |0 > along one
direction (the Coulomb branch), say < 0|φa |0 >= δ a3 a, and the ’t Hooft-
Polyakov monopole becomes a U(1) Dirac-type monopole.
In this spirit we can define the electric and magnetic charges as
Z
1 ~ ·Π
~ 3 φ3
qe ≡ d2 Σ (3.68)
a
Z Z
1 ~ · 1 B ~ 3 φ3 = 1 ~ · 1 B~ 3 φ3
qm ≡ d2 Σ d2 Σ D (3.69)
a 4π aD 4π
where aD = τ a and only the U(1) fields remaining massless after SSB appear.
These quantities are quantized, since4 qm ∈ π1 (U (1)) ∼ π2 (SU (2)/U (1)) ∼
Z and qe is quantized due to Dirac quantization of the electric charge in
presence of a magnetic charge.
Thus, after SSB, the central charge becomes

Z = i 2(ne a + nm aD ) (3.70)

The mass spectrum of the theory is then given by



M = 2|ne a + nm aD | (3.71)
4
We say that the ’t Hooft-Polyakov magnetic charge is the winding number of the map
SU (2)/U (1) ∼ S 2 → S∞ 2
, that identifies the homotopy class of the map. By considering
1 1 1 2
the maps U (1) ∼ S → S∞ , where S∞ is the equator of S∞ , it is clear that a similar com-
ment holds for the U(1) Dirac-type magnetic charge. It turns out that the two homotopy
groups, π2 (S 2 ) and π1 (S 1 ), are isomorphic to Z.

32
We shall call this formula the Montonen-Olive mass formula. It is now
crucial to notice that this formula holds for the whole spectrum consisting
of elementary particles, two W bosons and two fermions, and topological
excitations, monopoles and dyons. For instance the mass of the W bosons
and the two fermions can be √ obtained by setting ne = ±1 and nm = 0,
which gives mW = mfermi = 2|a|, whereas
√ the mass of a monopole (ne = 0
and nm = ±1) amounts to mmon. = 2|aD |. This establishes a democracy
between particles and topological excitations that becomes more clear when
e.m. duality is implemented.

3.5 Exercises
Exercise III.b Show that

[Sm , Sl ]K k K
L = iCml (Sk )L (3.72)

i.e. that the Sl s form a representation of the internal symmetry algebra of


k B .
the Bl s: [Bm , Bl ] = iCml k
Exercise III.a Using the supercharge given in the text, obtain the Wess-
Zumino transformations for all the fields, including the dummy fields.

33
Chapter 4

The infinite number of


degrees of freedom of
quantum fields, the Haag’s
theorem and the issue of
locality

See notes.

34
Chapter 5

Spatiotemporal
Noncommutativity

5.1 The fundamental length


The issue of locality: the Doplicher et al argument for a fundamental length
from the Heisenberg uncertainties in general relativity and space-time non-
commutatitivy.
See Notes

5.2 Examples in Nature


We closely follow the lecture by R.Jackiw, Physical instances of noncom-
muting coordinates, hep-th/0110057

5.3 Formalization: The gauge-covariant approach


and the Seiberg-Witten map
We closely follow the paper by J.Madore, S.Schraml, P.Schupp, J.Wess,
Gauge theory on Nocommutative Spaces, hep-th/0001203
On the Seiberg-Witten map for the Abelian case see notes.

5.4 Problems?
The problem of the infra-red/ultra-violet (IR/UV) connection ? The prob-
lem of unitarity with θoi 6= 0?

35
Chapter 6

From Lorentz symmetry


violation to Lorentz
symmetry enhancement?

See notes.
References:
M. Chaichian, P. P. Kulish, K. Nishijima and A. Tureanu, “On a Lorentz-
invariant interpretation of noncommutative space-time and its implications
on noncommutative QFT,” Phys. Lett. B 604, 98 (2004);
M. Chaichian, P. Presnajder and A. Tureanu, “New concept of rela-
tivistic invariance in NC space-time: Twisted Poincare symmetry and its
implications,” Phys. Rev. Lett. 94, 151602 (2005);
M. Dimitrijevic and J. Wess, “Deformed bialgebra of diffeomorphisms”,
arXiv:hep-th/0411224;
P. Aschieri, C. Blohmann, M. Dimitrijevic, F. Meyer, P. Schupp and
J. Wess, “A gravity theory on noncommutative spaces”, Class. Quant. Grav.
22, 3511 (2005).

36
Chapter 7

The elusive effects of Lorentz


violation: the case of − 14 F̂ 2

See notes.
References:
Z. Guralnik, R. Jackiw, S. Y. Pi and A. P. Polychronakos, “Testing non-
commutative QED, constructing non-commutative MHD”, Phys. Lett. B
517, 450 (2001);
P. Castorina, A. Iorio, D. Zappalà, Noncommutative Synchrotron, Phys.
Rev. D 69 (2004) 065008; On the Vacuum Čerenkov Radiation in Non-
commutative Electrodynamics and the Elusive Effects of Lorentz Violation,
Europhys. Lett. 71 (2005) 912; Violation of Lorentz Invariance and Dy-
namical Effects in High Energy Gamma Rays, Nucl. Phys. B (Proc. Suppl.)
136 (2004) 333.

37

You might also like