Advanced SUSY 2020
Advanced SUSY 2020
Advanced Supersymmetry
Contents
1
2 Contents
Introduction
In this set of lectures, we will build on the formalism developed in the Supersym-
metry and Supergravity lectures. There, we learned how to write down classical
field-theory Lagrangians with supersymmetry, and we discussed important aspects
of the quantum theory in the simple case of theories with chiral multiplets only.
In the present lectures, we will explore aspects of the quantum dynamics of
supersymmetric gauge theories in 4d. These are close cousins to the gauge theories
that we use in Particle Physics. Supersymmetry leads to more technical countrol,
leading to many exact results that we could only dream off in ‘real-world’ QFT with-
out supersymmetry. It is thus of great theoretical interest to study these models in
details, for instance as ‘toy models’ for the much harder problem of understanding,
say, QCD at strong coupling from first principles.
Part I
Dynamics of 4d N = 1 gauge theories
1 Quantum aspects of 4d gauge theories
In this section, we discuss some general aspects of (gauge) anomalies in QFT, as
well as the role of the θ angle in gauge theories. (This is by no means self-contained,
but should amply suffice for our purpose in these lectures.)
Figure 1: The triangle one-loop diagram that determines the gauge anomaly in four
dimensions, with fermions running in the loop. The wiggly lines correspond to the
external gauge fields; equivalently, one inserts a current jaµ at each vertex.
(R )
schematically, with α(x) = αa (x)Ta ψ the gauge-transformation functions acting
on the fermions; the sum over the e g-indices a, b, c is understood. The anomaly
coefficients Aabc are given by:
(R ) (R ) (R )
Aabc = tr Ta ψ {Tb ψ , Tc ψ } . (1.8)
These are also called cubic anomalies, since they are cubic in the “charges” under
G.
e Physically, they can be extracted from the three-point function of the currents
jaµ in the original theory with Lagrangian L0 :
They are “mixed anomalies” between gauge invariance and diffeomorphism invari-
ance. They are often called “gravitational anomalies,” because they contribute to
the right-hand-side of (1.7) in the presence of a non-trivial metric gµν (with non-
zero curvature)—they arise at one-loop from the same triangle diagram as in Fig 1,
but with one gauge field and two gravitons for the external legs. We see from (1.10)
that linear anomalies can only be non-zero for abelian symmetries, the U (1) factors
inside G.
e
e = GF × G .
G (1.11)
while we have Tr(G) and Tr(GF ) for the linear anomalies. There are thus, in fact,
three types of anomalies that concern us when G is gauged, from the lethal to the
innocuous:
1.1 Anomalies for gauge and global symmetries 7
Tr(GGG) , (1.13)
for three currents of the gauge group G. If it were non-zero, the theory would be
inconsistent quantum-mechanically. Therefore, we have the anomaly-free condi-
tion:
Aabc = 0 , (1.14)
where a, b, c runs over the generators of G only. For a simple Lie algebra g, we
have:
1
Aabc = A R(Rψ dabc = 0 ,
(1.15)
2
which is a non-trivial condition only for SU (N ) with N > 2. (See below equation
(7.28) in the Susy&Sugra lectures.) Note also that, if the gauge group is a product:
Y Y
G= Gi × U (1)k ,
i j
with Gi some simple factors (as in the Standard Model, for instance), we also have
non-trivial constraints:
qk [ψ]T (Rψ
X X
i )=0 , qk [ψ]ql [ψ]qm [ψ] = 0 , (1.17)
ψ ψ
where T (Rψ ψ
i ) is the quadratic index of the representation Ri of Gi under which ψ
transforms, and qk [ψ] denotes the U (1)k charges of the fermions.
(ii) Anomalous global symmetries. We could also have so-called “mixed anoma-
lies:”
Tr(GGGF ) , Tr(GGF GF ) , (1.18)
between the gauge-symmetry and the global-symmetry currents. If G is semi-
simple, only Tr(GGGF ) can be non-trivial. Let a, b, · · · run over the generators Ta
of the gauge group G, and let α, β, · · · run over the generators Tα of GF . Then,
any anomaly:
Aabα 6= 0 , or Aaαβ 6= 0 , (1.19)
signals that the currents jαµ , or jαµ and jβµ , part of the naive symmetry GF , are
actually not conserved in the gauge theory. This is an anomalous global symmetry,
which is then not a symmetry of the quantum system.
We will mostly focus on G = SU (N ) and GF ⊃ U (1)A , with a particular
flavor symmetry U (1)A that can be anomalous. This non-conservation of U (1)A is
generally called a chiral anomaly.
8 1 Quantum aspects of 4d gauge theories
4 √
Z Z
θ µνρσ θ
Stop = − i d x g tr ( F F
µν ρσ ) = − i tr F ∧ F , (1.23)
64π 2 M4 16π 2 M4
where we used the form notation, with F = 21 Fµν dxµ ∧ dxν in local coordinates. It
is clear from the right-hand-side of (1.23) that Stop is independent of the metric on
M4 . In fact, the integrand is also a total derivative:
Z Z
2i
tr F ∧ F = d tr A ∧ dA − A ∧ A ∧ A , (1.24)
M4 M4 3
3 , at infinity:
with g(x) an gauge transformation on the three-sphere, S∞
g(x) 3
: S3 → G . (1.26)
S∞
Given a gauge field Aµ with field-strength Fµν , define the dual (or “magnetic”)
field-strength as:
1
Feµν ≡ µνρσ F ρσ . (1.29)
2
Definition: An (anti)-instanton is a gauge field configuration on M4 which is
(anti)-self dual:
Fµν = ±Feµν , (1.30)
with the + and − signs for the instanton and anti-instanton, respectively. In QFT,
we consider, in particular, M4 = R4 , which can be compactified to a sphere S 4 .
One physical significance of instantons is that they are the non-trivial classical
saddles of the Yang-Mills action. Indeed, we have:
Z Z Z
1 2 2
|F + F | = |F | ± F ∧ F ≥ 0 ,
e (1.31)
2
schematically, which implies that the YM action in always larger or equal to the
topological action, the sense that:
Z Z
|F |2 ≥ F ∧ F = 16π 2 |k| , (1.32)
with the inequality saturated for the (anti)-instanton configuration. We have k > 0
for an instanton and k < 0 for an anti-instanton. On any (anti)-instanton back-
ground, we have:
4 √ 4 √ µνρσ 8π 2
Z Z
1 µν 1
d x gFµν F = ± d x g Fµν Fρσ = |k| . (1.33)
4g 2 8g 2 g2
(k)
Thus, a k-instanton gauge field, Aµ , is weighted by a numerical factor:
8π 2 |k|
−SYM [A(k) ] −
e =e g2 , (1.34)
in the path integral. Note the similarlity to the (one-loop) strong-coupling scale:
8π 2
−
Λ = µe b0 g 2 (µ) , (1.35)
the IR, for an asymptotically-free gauge theory), the instantons will start giving
increasingly important contributions to the dynamics. At strong coupling, we would
even expect them to dominate.
While the trivial saddle Aµ = 0 (corresponding to k = 0) is the starting point
of ordinary perturbation theory, one can (and should) do a similar perturbative
expansion around all saddles. The computation of any given observable will a
priori receive contributions from all saddles (plus perturbative fluctuations). For a
simple, simply-connected gauge group, such as SU (N ) (for simplicity), we have:
Stop = −iθk , (1.36)
in Euclidean signature, in the presence of a k-instanton. Then, the Yang-Mills path
integral takes the schematic form:
8π 2 |k|
Z
iθk − g2
X
iSYM +iStop
Z = [DAµ ] e = e e Zk , (1.37)
k∈Z
where the sum is over all the chiral fermions ψ with non-zero U (1)A charges q[ψ].
Each such ψ sits in a representation Rψ of the gauge group G, with quadratic index
T (Rψ ). Let us define, then, the chiral anomaly coefficient:
X
AU (1)A ≡ qA [ψ] T (Rψ ) . (1.41)
ψ
θ → θ + α AU (1)A . (1.43)
Thus, even though the axial symmetry U (1)A does not exists in the quantum theory
if AU (1)A 6= 0, one can keep track of it (and extract physical consequences) by
assigning the U (1)A transformation (1.43) to θ, viewing θ itself as a background
field.
All the various interactions terms are then determined by the combination of gauge
invariance and supersymmetry, as well as by W .
1 2
L ⊃ D − φ̄Dφ , (2.2)
2g 2
2.1 Vacuum and beta functions 13
in the WZ gauge. The equations of motions for the auxiliary fields D = Da Ta give:
2 dim(G)
X ∂W g 2 X (R) 2
V0 = + φ̄Ta φ . (2.4)
∂φ 2
φ a=1
The first term is the contribution from the superpotential, which we discussed in
previous sections, while the second term can be viewed as a contribution from the
gauge interactions themselves. The real operators:
Any two solutions to (2.6) related by a (constant) gauge transformations are physi-
cally equivalent. So, we introduce the equivalence relation on the space of constant
field values:
a T (R)
φ0 ∼ φ if ∃ (αa ) ∈ Rdim(G) such that φ0 = eiα a
φ. (2.7)
The constant values of the scalar field φ ∈ Φ span the vector space:
VR ∼
= Cn , n ≡ dim(R) , (2.8)
on which the representation R acts. Then, the vacuum manifold of the gauge theory
takes the general form:
M = {φ ∈ VR | ∂φ W = 0 , µa = 0}/G , (2.9)
where the quotient by the gauge group corresponds to the equivalence relation
(2.7). In our discussion of theories with only chiral multiplets, we saw that the
vacuum moduli space was a purely algebraic object—in particular, everything was
holomorphic in φ. This is apparently not the case in a gauge theory, since the
formula (2.9) is non-holomorphic in two ways: the moment maps µa are real, and
the gauge equivalence (2.7) is in terms of real gauge parameters αa .
Nonetheless, there is a simple-looking (although by no mean obvious) way to
rewrite (2.9) more algebraically. It turns out that imposing the vanishing of the
14 2 General aspects of supersymmetric gauge theories
M = {φ ∈ VR | ∂φ W = 0}/GC . (2.10)
In this approach, we are considering the space of complexified gauge orbits (or,
more precisely, their closure), under the GC action:
a T (R)
φ0 ∼ φ if ∃ (ω a ) ∈ Cdim(G) such that φ0 = eiω a
φ. (2.11)
The fact that the two approaches (2.9) and (2.10) reproduce the same moduli space
was shown explicitly in [7]. 2
Conceptually, this was to be expected: the fact that we only divide by real gauge
transformations in (2.9) is an artefact of the WZ gauge. The supersymmetric gauge
transformations on chiral superfields,
δΩ Φ = eiΩ Φ , (2.12)
are really GC -valued gauge transformations. More generally, the F -term contribu-
tions to the Lagrangian of any supersymmetric gauge theory are invariant under
the complexified gauge group GC , while the total Lagrangian (in particular, the
D-term kinetic term for matter fields) is only G-invariant.
Finally, it is non-obvious but nonetheless true that the vacuum moduli space
M of a gauge-theory is also a Kähler manifold, just like in the case without gauge
fields.
where the parameters ri are the R-charges of the chiral multiplets, R[Φi ] = ri (so
the corresponding chiral fermions have R[ψi ] = ri − 1). One can then often find a
non-anomalous R-symmetry, by choosing the R-charges ri such that:
X
AU (1)R = T (adj) + (ri − 1)T (Ri ) = 0 . (2.20)
i
A digression: The reader might wonder about the fact that the same quadratic
indices of the gauge representations appear in the expression for the YM β function
(2.13) and for the chiral anomalies. In fact, one can easily define an R-symmetry:
2 2X
Rc ≡ R0 + A = R0 + qi , (2.21)
3 3
i
which is such that the chiral anomaly is exactly proportional to the β function:
2
AU (1)Rc = b0 . (2.22)
3
The R-charge Rc assigns rc = 23 to all chiral multiplets, which is compatible with
a classically-marginal superpotential, W = Φ3 (schematically). In fact, in the far
UV, such a super-Yang-Mills (SYM) theory is classically conformal (in particular,
scale invariant), and the R-charge combines with the supersymmetry current and
the energy-momentum tensor into a larger algebraic structure (known as N = 1
superconformal multiplet). Quantum corrections break both conformal invariance
(giving the running of g 2 , which is a “quantum anomaly” of scale invariance) and
the R-symmetry (through the chiral anomaly), but supersymmetry relates these
two “quantum anomalies” exactly as in (2.22).
16 2 General aspects of supersymmetric gauge theories
τ (µ0 )
Wµ0 (W, Φ, τ, λ) = − tr W α Wα + W (Φ, λ; µ0 ) . (2.23)
16πi
Here, λ denote the ordinary superpotential coupling. This F -term must depend
holomorphically on both τ and λ, which restricts the possible quantum corrections
to W, at least in some appropriate “holomorphic scheme.” We also introduced the
UV scale µ0 explicitly in (2.23).
Perturbatively in g 2 , the holomorphy in τ gives some very severe restriction
when combined with the fact that θ is an angle with period 2π, which means that
the theory is invariant under:
τ ∼τ +1 . (2.24)
This implies that τ can only appear linearly in Wµ , at any scale µ, and only in the
precise form:
τ (µ)
Wµ ⊃ − tr W α Wα . (2.25)
16πi
Indeed, in that case θ multiplies the topological term as in (1.23). Since it is a
topological invariant, it cannot depend on µ at all. The one-loop running of the
holomorphic coupling τ is given by:
b0 µ
τ (µ) = τ (µ0 ) − log . (2.26)
2πi µ0
since the theory should be regular in the limit Λ → 0. Note the expression:
b0 Λ
τ (µ) = log . (2.30)
2πi µ
One could still constraint more carefully the form of the possible quantum cor-
rections to W. This sort of analysis, however, is only reliable at weak coupling.
As we RG flow from an asymptotically-free theory in the UV toward µ ∼ |Λ|, the
theory become strongly coupled and we need new methods to explore the infrared
physics. As we will see in a particular example (SQCD), supersymmetry and gen-
eral symmetry arguments sometimes are exceptionally powerful in order to “guess”
what the infrared physics is.
where the ellipsis denotes contributions from higher-dimensional operators, and the
wave function renormalisation factor depends on all the coupling constants:
Aµ → gc Aµ , (2.33)
As we did for theories of chiral multiplets. It turns out, however, that these field
redefinitions are anomalous in the presence of the gauge interactions—that is, the
change of variable in the path integral gives a non-trivial Jacobian [8]. Heuristi-
cally, this can be understood as follows. The symmetry group of the F -terms is
complexified—we mentioned this fact before for the gauge group G, but that is true
of any global symmetries as well. The rescaling (2.35), in particular, can be under-
stood as a complexified “chiral rotation.” Since that chiral symmetry is anomalous,
the rotation shifts the θ angle by an imaginary amount. This, effectively, shift g12
by a real quantity.
V = gc Vc , (2.43)
to obtain the canonically normalised gauge field. This change of variables also has
a non-trivial Jacobian, which gives [8]:
Z
2 iθ 1 T (adj)
d θ − 2 + log(gc ) W α Wα . (2.44)
32π 2 4gR 32π 2
Then, equating the real coefficient inside the parenthesis with the canonically-
normalised coupling − 4g12 , we get:
c
1 1 T (adj) 1
= 2 + log , (2.45)
gc gR 16π 2 gc2
1 P
b0 + i T (Ri )γφi
1 2
β = 1 . (2.46)
gc2 8π 2 − 2 T (adj)gc
2
Note that the denominator is the same as in (2.41). This is the famous NSVZ
β-function, which was first derived by completely different methods [9]. It is some-
times called “the exact β-function,” in the sense that it depends only on the anoma-
lous dimensions of the fields. Note that the anomalous dimensions themselves de-
pend on gc2 (and on any other superpotential couplings), so it just tells us that the
β function is known is we know the exact anomalous dimensions. The latter depend
on the details of the Kähler potential, and thus we cannot have exact formulas for
them. This is thus similar to the “exact” β-function for the superpotential coupling
that we discussed previously, 4 namely:
eR ) = − 3 +
X 1 e
β(λ 1 + γ φi d i λR . (2.47)
2
i
One classic use that has been made of these “exact” results is to look for
perturbatively-exact fixed points of the RG flow. Combining the results for the
gauge-coupling and superpotential coupling constants, (2.46) and (2.47), we can
ask whether it is possible to find an solution to the equations:
1
β 2 =0, β(λ
eR ) = 0 . (2.48)
gR
Here we used gR , since its β-function has the same zeros as gc . The existence of such
fixed points should be completely scheme-independent. The equations (2.48) give
strong constraints on the anomalous dimensions γφ that can arise at any candidate
fixed point, even at strong coupling.
Qi , i = 1, · · · , Nf , ej ,
Q j = 1, · · · , Nf , (3.1)
respectively. For obvious reasons, the numbers Nc and Nf are called the number
of colors and flavors, respectively, and the scalar fields Q, Q e are called the squarks.
The supersymmetric Lagrangian takes the form:
Z Nf Nf
X X ¯e −2V e j
L = d2 θd2 θ̄ Q̄i e−2V Qi + Q je Q
i=1 j=1 (3.2)
Z Z
τ τ̄
− d2 θ tr(WW) + d2 θ̄ tr(W̄ W̄) .
16πi 16πi
More precisely, this is massless SQCD, with vanishing superpotential. We could
also consider adding Dirac masses for the quarks, through a superpotential:
W = µi j Q
e j Qi , (3.3)
with µi j the mass matrix. Note that all gauge indices are implicit. We denote by
a = 1, · · · , Nc the gauge indices in the fundamental representation. Then,
e j Qi ≡ Q
Q e ja Qai ,
which is obviously gauge invariant, and similarly for the contraction of the gauge
indices in (3.2).
3.1 Anomalies and anomaly-free R-symmetry 21
times an R-symmetry U (1)R . The U (1)A and U (1)B factors are called the axial
symmetry and the baryonic symmetry, respectively. The charges of the chiral su-
perfields under the gauge and global symmetries are summarised in Table 1. The
R-charge shown if for an R-charge:
R = R0 + rA , (3.5)
with A the generator of U (1)A and R0 the “reference” R-charge under which the
chiral multiplets are neutral.
However, there is a unique choice of r such that the U (1)R -SU (Nc )2 anomaly (3.7)
vanishes, namely:
Nc
r =1− . (3.8)
Nf
The other symmetries are non-anomalous. Thus, the global symmetry of SQCD,
at the quantum level, is:
Here, the “D-term constraints” can be written as the vanishing of the traceless
Nc × Nc matrices: 5
Nf Nf
δba δba
X
†i a †i †a e j † ej
X
µab ≡ Q b Qi − tr(Q Qi ) − Qj Qb −
e tr(Qj Q ) = 0 . (3.11)
e
Nc Nc
i=1 j=1
Note that the matrix µab is obviously SU (Nc ) × SU (Nc ) × U (1)B -invariant.
For any fixed number of “colors” Nc , the structure of the SU (Nc ) SQCD moduli
space changes as we vary the number of flavors, Nf . The basic physical reason is
the Higgs mechanism. A non-zero VEV for a single fundamental scalar Q = (Qa )
of SU (Nc ) breaks the gauge group as:
hQa i =
6 0 ⇒ SU (Nc ) → SU (Nc − 1) . (3.12)
By a gauge transformation, we can take the vector hQa i to be (q1 , 0, 0, · · · , 0), which
is obviously preserved by the SU (Nc − 1) subgroup of SU (Nc ). More generally, a
generic VEV for the Nf squarks breaks the gauge group according to:
hQai i =
6 0, i = 1, · · · , Nf ⇒ SU (Nc ) → SU (Nc − Nf ) . (3.13)
For Nf ≥ Nc , the SU (Nc ) gauge group is entirely broken at a generic point on the
vacuum moduli space. The VEVs also have to satisfy the D-term conditions (3.11)
in order to preserve supersymmetry.
Another approach to analysing the moduli space is to use the description (2.10).
In the absence of superpotential, this tells us that:
n o
M = (Qai , Qe ja ) ∈ C2Nc Nf /SL(Nc , C) . (3.14)
This space can be constructed algebraically by building all the possible gauge in-
variant chiral operators, X, and then imposing relations between the fields X that
follow from their definition—these are known as syzygies. This is a classic problem
in invariant theory. We will present explicit examples below.
5
Here we take:
1 a c
(T a )ab = (Tdc )ab = δbc δda −
δb δd ,
Nc
for the fundamental of SU (Nc ), and then (T̄ a )ab = −(T a )ab for the anti-fundamental.
3.2 The classical vacuum moduli space of SQCD and gauge-invariant operators 23
Plugging this into (3.11), we see that this solves the D-term condition if and only:
qe1 0 · · · 0 0 ··· 0
0 qe2 · · · 0 0 · · · 0
e ia ) =
(Q .. . . ,
with |qi | = |e
qi | , ∀i . (3.16)
. . 0 · · · 0
0 0 ··· qeNf 0 ··· 0
This is not the most general solution to (3.11), but any other allowed VEV can be
reached by considering the SU (Nf ) × SU (Nf ) orbit of (3.15)-(3.16). It turns out
that:
dimC (M) = Nf2 , (3.17)
as a complex space. This can be understood easily in terms of the Higgs mechanism.
At a generic point on M, the gauge symmetry is broken as in (3.12) and so:
gauge bosons get a mass, by each “eating” a complex scaler φ. Thus, out of the
2Nf Nc complex scalars in the UV, 2Nf Nc − NW = Nf2 scalars survive on the IR
moduli space, matching the counting (3.17).
We can also see this in the purely algebraic description (3.14). In this language,
we should construct all the gauge invariant scalars build out of the fundamental
e There are only N 2 of them, which we denote by:
squarks Q and Q. f
Mji ≡ Q
e ja Qai . (3.19)
They are usually called the SQCD mesons, since they are made of two fundamental
squarks, in analogy with the mesons of real-world QCD which are made of two
quarks. They are the natural coordinates on the moduli space for Nf < Nc , with:
2
M∼ N
=C f . (3.20)
Note also that the case Nf = Nc − 1 is special, since the gauge group is com-
pletely broken. (The “SU (1)” group is trivial.)
24 3 SQCD: Lagrangian, symmetries and classical moduli space
Let us see how this comes about in the language of gauge-invariant chiral fields.
Now, in addition to the mesons (3.19), we can build other gauge-invariants, by
making use of the fully-antisymmetric invariant tensor of SU (Nc ). These are the
so-called “baryons:”
a
Bi1 i2 ···iNc ≡ a1 a2 ···aNc Qai11 Qai22 · · · QiNNc , (3.22)
c
Again, the name comes from QCD, where a baryon is a gauge-invariant combination
of 3 fermions in the fundamental of the gauge group SU (3).
For future reference, we collect the symmetry charges of the mesons and baryons
in Table 2. In total, there are:
2 Nf
NX ≡ Nf + 2 , (3.24)
Nc
which follow directly from their definition in terms of fundamental fields. One can
easily check that:
j ···j [j
e j1 j2 ···jNc Bi i ···i − M 1 · · · M Nc = 0 , j ]
Pi11···iNNc (X) ≡ B 1 2 Nc i1 iN (3.26)
c c
i ···iNf j i ···iNf
Q Nc +2 (X) ≡ 1 Bi1 ···iNc MijN =0, (3.27)
c +1
j1 ···jNc j
Q
ej
Nc +2 ···jNf i (X) ≡ j1 ···jNf B Mi Nc +1 = 0 . (3.28)
The moduli space is then given explicitly in terms of generators and relations, as:
n o
M = X ∈ CNX P (X) = 0 , Q(X) = 0 , Q(X)
e =0 . (3.29)
Mathematically, this is known as a affine variety (the zero set of a some polynomials
in Cn ).
generators, while the moduli space is of dimension Nc2 + 1. Indeed, the generators
are:
Mij , B, Be , (3.31)
where we defined:
det(M ) − BB
e =0. (3.33)
On the other hand, the expected dimension of the moduli space is:
Note that there are NP = Nf2 + 2Nf relations amongst the NX = Nf2 + 2Nf
generators, but:
NP > NX − dim(M) . (3.38)
Thus, the relations cannot be all independent, but nonetheless there does not exist
a smaller set of relations. This is a common feature of algebraic varieties. (Variety
whose dimensions is given by the number of generators minus the number of rela-
tions are called complete intersections. The SQCD moduli space for Nf > Nc is
not a complete intersection.)
• Mass gap. The theory might be gapped. That is, there are no excitations
below a finite energy E0 , and therefore the far-infrared physics is trivial.
For instance, pure Yang-Mills theory in 4d is expected to have a mass gap.
(Proving that conjecture is a Millennium Prize Problem, literally and figura-
tively worth $1,000,000.)
3.5 Aspects of the quantum vacuum of SQCD 27
Note that the vacuum at a generic point of the classical SQCD moduli space is
an IR-free theory, consisting of n free chiral multiplets, with n = dim(M). On the
other hand, it is much more challenging to understand what happens at the origin
of the moduli space, where the gauge group SU (Nc ) is unbroken and we expect the
strongly-coupled gauge dynamics at scales µ ≤ Λ to be dominant.
It is expected that the theory confines and develops a mass gap in the IR, just
like pure YM theory. Moreover, the theory has Nc distinct vacua, in which the R-
symmetry Z2Nc is spontaneously broken to Z2 , due to the appearance of a gaugino
condensate in the supersymmetry-preserving vacuum:
2πin
hλα λα i = Λ3 e Nc , n = 1, · · · , Nc . (3.41)
The fact that the theory has (at least) Nc distinct vacua can also be inferred from
the Witten index of the theory, which is equal to Nc [10].
6
Free massless particles are CFTs too, but trivial ones.
28 3 SQCD: Lagrangian, symmetries and classical moduli space
with the dependence on det M only, to preserve SU (Nf ) × SU (Nf ), and the coef-
ficients:
Λ3Nc −Nf
WADS,Nf =Nc −1 = . (3.46)
det M
By consistency with various decoupling limits, this fixes:
1
Λ3Nc −Nf
N
c −Nf
WADS = (Nc − Nf ) , (3.47)
det M
hM i → ∞ . (3.49)
3.5 Aspects of the quantum vacuum of SQCD 29
hφi ≤ Λ . (3.51)
for the “meson” and “baryon” fields X in (3.25), which we now view as the “fun-
damental fields” in the low-energy description. The effective Lagrangian at scale
µ Λ is then given in terms of some unknown (and presumably complicated)
Kähler potential K, and some holomorphic superpotential W (X).
We can test this hypothesis using the ’t Hooft anomaly matching condition.
There are many non-trivial ’t Hooft anomalies for global symmetry group GF ×
U (1)R of the UV theory, massless SQCD. For instance:
and:
Nc4
tr(U (1)R ) = −Nc2 − 1 , tr(U (1)3 ) = Nc2 − 2 −1 , (3.55)
Nf2
where we used the non-anomalous R-charge (3.8). There are still more ’t Hooft
anomaly coefficients, whose computation is left as an exercice.
On the other hand, the naive theory (3.52) would have:
tr(SU (Nf )3+ ) = −Nf + A ⊗N A
c
Nf , (3.56)
where the first term is the contribution from the mesons M , and the second term is
the contribution from the baryons B to the cubic anomaly. In particular, we have:
det(M ) − BB
e =0. (3.60)
e = Λb0 ,
det(M ) − BB b0 = 2Nc . (3.61)
3.5 Aspects of the quantum vacuum of SQCD 31
3.5.5 Nf = Nc + 1: A σ-model
For Nf = Nc + 1, we saw above that we could saturate the ’t Hooft anomalies with
our naive σ-model of mesons and baryons. This cannot be the full description,
however, since the moduli space has a lower dimension than the number of fields,
NX = Nf2 +2Nf . Instead, we should have a superpotential to impose some relations
amongst the fields X. By symmetry, we can only have:
By various decoupling limits, one can fix α = −β = −1. We thus claim that the
correct superpotential is:
This is a rather strange result, since it does not seem to behave well in the classical
limit, Λ → 0. However, the numerator would also vanish in this limit, due to the
classical constraints (3.37)—in particular, det M = 0 classically since it is a matrix
of rank Nc < Nf .
In the low-energy description, the F -term equations that follow from (3.63),
when treating M and B, Be as fundamental fields, give us:
∂W
= B i Bej − (M −1 )ij det M = 0 ,
∂Mij
∂W
= Mij B i = 0 , (3.64)
∂B i
∂W
= Bej Mij = 0 .
∂ Bej
These are precisely the constraints (3.37) that define the classical moduli space.
Thus, for Nf = Nc + 1, the low-energy description seems to be in terms of a
σ-model whose vacuum moduli space is exactly the same as in the UV description.
7
Here, deformation is also a technical term; more precisely, we have a “complex structure
deformation” of an algebraic variety.
32 3 SQCD: Lagrangian, symmetries and classical moduli space
It is perhaps easiest to first discuss what happens near the upper bound Nf = 3Nc .
At precisely Nf = 3Nc , the Yang-Mills β-function vanishes at one-loop. We thus
have have a fixed-point of the RG flow, at first order—a four-dimensional CFT.
Consider the “exact” β function (2.46), which reads:
1 3Nc − Nf (1 − γφ )
β 2
= . (3.66)
gc 8π 2 − Nc gc2
for SQCD; here, we used the global symmetries to equate all the anomalous dimen-
sions,
γQi = γQej ≡ γφ . (3.67)
For Nf = 3Nc , there is a fixed point, at all order in perturbation theory, if γφ = 0.
That is, if the squarks retain their classical dimensions, ∆ = 12 . Therefore, it seems
that massless SQCD at Nf = 3Nc is an “almost free” conformal field theory.
Now, for Nf < 3Nc , we could try to obtain a zero of the β function:
1 Nf − 3Nc
β =0 ⇔ γφ = , (3.68)
gc2 Nf
where we used the global symmetries to equate all the anomalous dimensions. Con-
sider, in particular, the case of large number of colors, Nc 1 and 2Nc − Nf very
small; then, the anomalous dimensions are arbitrarily small and one can understand
the fixed point perturbatively. Such a fixed point is called a Banks-Zaks fixed point;
it also occurs in QCD-like theories without supersymmetry [12]. All we need is a
β-function of the form:
1 b0
β 2
= 2 − c0 Nf g 2 + O(g 4 ) , (3.69)
g 8π
at two-loop order, with c0 > 0 a positive numerical constant. Then, we have a
perturbative fixed point with a coupling constant:
1 b0
g∗2 = 1, (3.70)
8π 2 c0 Nf
if b0 is smaller than c0 Nf .
3.5 Aspects of the quantum vacuum of SQCD 33
SCFTs. The claim is that there is a non-trivial fixed point in the IR of SQCD in
the full range:
3Nc
≤ Nf ≤ Nc . (3.71)
2
This is called the SQCD conformal window. The gauge coupling gc2 at the fixed
point is small near the upper limit (for Nc and Nf sufficiently large, giving a Banks-
Zaks fixed point), but becomes strong (with g∗2 of order one) as we lower Nf , at fixed
Nc . The lower bound on the conformal window comes about as follows. Any fixed
point preserving supersymmetry necessarily enjoys a larger space-time symmetry
algebra, called the N = 1 superconformal algebra. It has generators: 8
Pµ ,
Qα , Q̄α̇ ,
∆, Mµν , R, (3.72)
Sα , S̄α̇ ,
Kµ ,
generalising the super-Poincaré algebra. Here, ∆ is the dilation operator, whose
eigenvalues are the quantum dimensions (or just “conformal dimensions”) of the
operators, by definition, and R is the U (1)R charge. The R-charge is now a non-
trivial part of the algebra. In 4d N = 1 superconformal field theories (SCFTs), the
scalar chiral operators Φ satisfy a BPS-type relation tying up their R-charges and
dimensions:
2
R[Φ] = ∆[Φ] . (3.73)
3
Note that this relation is compatible with a classically-marginal (that is, conformally-
invariant) superpotential. Another general fact about 4d CFTs (with or without
supersymmetry) is that the dimension of any well-defined (gauge-invariant) opera-
tor must satisfy:
∆(O) ≥ 1 , (3.74)
and the operator is free if and only if this so-called unitary bound is satisfied.
Consider the quantum dimension (3.68) for the squarks chiral superfields. Using
(3.73), that implies:
2 1 Nc
R[Q] = R[Q] =
e 1 + γφ = 1 − =r, (3.75)
3 2 Nf
precisely the anomaly-free R-charge of SQCD. Then, for the gauge-invariant meson
operators M = QQ e to satisfy the unitarity bound, we must have:
e =3−3 Nc 3
∆(QQ) ≥1 ⇔ Nf ≥ Nc . (3.76)
Nf 2
That explains the lower-bound on the conformal window.
8
Here the generators are organised according to their conformal dimensions, from ∆[Pµ ] = 1
to ∆[Kµ ] = −1. The Poincaré supercharges Qα , Q̄α̇ have dimension 21 , and the special conformal
supercharges Sα , S̄α̇ have dimension − 12 .
34 3 SQCD: Lagrangian, symmetries and classical moduli space
3
Nc + 1 < Nf < Nc . (3.77)
2
We saw that the naive mesons and baryons cannot give a good description of the
origin of the moduli space. One heuristic reason is that the only superpotential
term allowed for the mesons is of the form:
1
W ∼ (det M ) Nf −Nc , (3.78)
which is singular at the origin. Such singularities in the effective action typically
hint at the presence of light particles, which we forgot to take into account in the
Wilsonian action. The extraordinary claim, due to Seiberg, is that one should
describe the low energy physics in terms of some IR-free gauge theory with gauge
group:
SU (Nf − Nc ) (3.79)
and Nf flavors in chiral multiplets q i and qej , coupled to some additional Nf2 gauge-
singlets caled Mij , with a cubic superpotential:
This SQCD-like theory, which we call the “Seiberg-dual theory” of SQCD, has a
β-function coefficient:
bD
0 = 3(Nf − Nc ) − Nf = 2Nf − 3Nc , (3.81)
which is negative in the window (3.77). Thus, indeed, it becomes a free theory in
the infrared (and needs to be defined with a UV cut-off, at the scale Λ).
Let us repeat the claim: the low-energy theory for asymptotically-free SQCD
in the window (3.77) is given in terms of an IR-free gauge theory, with gauge group
SU (Nf − Nc ). The gauge bosons and matter fields of this “dual theory” have
nothing to do with the original fundamental fields of SQCD in the UV. Nonetheless,
one can check that all ’t Hooft anomalies match between SQCD and the proposed
IR description! The proposal, in fact, passes many other consistency checks, which
we will not discuss here.
Seiberg duality. Even more amazingly, this relation between two different gauge
theories, known as “Seiberg duality,” extends all the way into the conformal window,
where both the SU (Nc ) and the “dual” SU (Nf −Nc ) gauge group are asymptotically
free. In that case, we have two well-defined asymptotically free gauge theories in
the UV, written schematically as:
They are certainly two different theories, with different numbers of degrees of free-
dom in the UV. The claim is that, in the conformal window, they both flow to the
same SCFT in the infrared. Moreover, when one description is strongly coupled,
the other is weakly coupled—that fact, as you can imagine, can be very useful.
We should point out that the above intricate picture of the quantum vacuum struc-
ture of SQCD has no definite proof for Nf ≥ Nc , to this day, but it passes so many
highly non-trivial consistency checks that its correctness is beyond any reasonable
doubt.
Part II
Dynamics of 4d N = 2 gauge theories
We now turn our attention to gauge theories with minimally extended (N = 2)
supersymmetry. Our aim will be to introduce the most important elementary facts
about N = 2 supersymmetric gauge theories, and then to discuss the Seiberg-
Witten solution for their low-energy dynamics. For simplicity, we shall mainly
focus on the simplest case of a single SU (2) N = 2 vector multiplet.
1 1 1
L = tr − Fµν F µν − iλ̄σ̄ µ Dµ λ + D2 − Dµ φ̄Dµ φ
g2 4 2
√ √
− iψ̄σ̄ µ Dµ ψ + F̄ F − φ̄[D, φ] − i 2φ̄[λ, ψ] + i 2[λ̄, ψ̄]φ (4.2)
θ
µνρσ
− tr Fµν ρσ .
F
64π 2
36 4 Gauge theories with N = 2 supersymmetry
VN =2 = φ , φ̄ , Aµ , λI , λ̄I , DIJ ,
(4.3)
for a total of 8 + 8 off-shell degrees of freedom, with all fields transforming in the
adjoint representation of g = Lie(G). In the N = 1 notation above, we have:
Let us integrate out the auxiliary fields. Then, the Lagrangian (4.2) becomes:
1 1
LN =2 SYM = tr − Fµν F µν − Dµ φ̄Dµ φ − iλ̄I σ̄ µ Dµ λI
g2 4
(4.5)
1 i i
− [φ̄, φ]2 − √ φ̄IJ [λI , λJ ] + √ IJ [λ̄I , λ̄J ]φ ,
2 2 2
plus the topological term (with coupling constant θ), which is separately super-
symmetric. This Lagrangian is classically invariant under a SU (2)R × U (1)r R-
symmetry, with I = 1, 2 the SU (2)R index. We will explore this important point
momentarily.
Note also the form of the scalar potential that appears in (4.5):
1
tr [φ̄, φ]2 .
V0 (φ, φ̄) = 2
(4.6)
2g
with I, J = 1, 2 and 12 = 12 = 1 for the SU (2)R indices. The maximal R-
symmetry of the N = 2 supersymmetry algebra is:
U (2) ∼
= SU (2)R × U (1)r , (4.8)
4.1 N = 2 supersymmetry and R-symmetry 37
QIα (2)−1 ± 12 I
−δ 1
(4.9)
Q̄α̇J (2̄)1 ∓ 12 +δ I 1
Z (1̄)2 0
Aµ λI λ̄I φ φ̄ ,
(4.12)
SU (2)R × U (1)r (1)0 (2)1 (2̄)−1 (1)2 (1)−2
with all the other anticommutators vanishing. Let us define the operators:
1 1
aI = √ (QI1 + α(I) IJ Q̄2̇J ) , a†I = √ (Q̄1̇I + ᾱ(I) IJ Q̄J2 ) ,
2 2
(4.14)
1 1
bI = √ (QI1 − α(I) IJ Q̄2̇J ) , b†I = √ (Q̄1̇I − ᾱ(I) IJ Q̄J2 ) ,
2 2
with α1 , α2 some pure phases. One can check that the only non-zero commutators
amongst these operators are:
For any fixed Z, we can choose α(I) = ei arg(Z) , so that the RHS of the last two
lines in (4.15) vanish. We then find the interesting conditions:
(no summation on I here), since {aI , a†I } and {bI , b†I } are positive-definite. These
so-called BPS inequalities are very important in the study of 4d N = 2 quantum
field theories.
We can easily study the supermultiplets of one-particle states, as before. For
M 6= ±|Z|, we have an ordinary massive multiplet, also known as a long multiplet.
It has 24 = 16 states, 8 bosonic and 8 fermionic. When the so-called BPS condition:
M = |Z| , (4.17)
is satisfied, on the other hand, we have a short multiplet, with half the number of
components, since bI , b†I in (4.16) are then realised trivially on one-particle states.
MC = hC /WG . (4.19)
It is called the Coulomb branch, for reason that will be clear momentarily. Its
complex dimension is the rank of G.
We will focus on G = SU (2). Then, the scalar φ can be conjugated to:
a 0
φ= , a∈C. (4.20)
0 −a
It is clear that the VEV (4.20) with a 6= 0 breaks the gauge group according to the
pattern:
SU (2) → U (1) , (4.22)
by the Higgs mechanism. For a general G, we have:
at a generic point on the Coulomb branch. Thus, classically, the low-energy dynam-
ics is dictated by the Higgs mechanism. The strict IR limit contains only massless
U (1)
abelian gauge fields Aµ , and their N = 2 superpartners in abelian vector mul-
U (1)
tiplets VN =2 . Any vacuum moduli space of a supersymmetric theory with this
property is called a Coulomb branch, because of the presence of long-range inter-
actions between charged particle in the IR through the ‘Coulombic’ interactions
mediated by the abelian gauge fields. 10
On the Coulomb branch (4.20) of the SU (2) theory, we also have a perturbative
massive vector field, the W-boson, of mass:
MW = 2|a| . (4.24)
At large distance on the Coulomb branch, that is with |a| → ∞, the semi-classical
Higgs mechanism gives the full physics. On the other hand, we will see that, at
finite distance on the Coulomb branch, strong coupling effects will change the low
energy dynamics significantly.
10
In addition, we also have attractive forces from the exchange of the massless scalars. For a
BPS particle, the repulsive Coulombic forces and the scalar forces cancel out exactly.
40 4 Gauge theories with N = 2 supersymmetry
Here, we defined the electric and magnetic field in the usual way:
In the second line, we used the fact that the term |Bi ∓ Di ϕ|2 is positive definite,
and that the remaining term is a total derivative, so that the final integral is over
the sphere at spatial infinity. In this way, we learn that the energy of any field
configuration is bounded by the magnetic charge of the gauge field times the VEV
of the scalar ϕ at spatial infinity.
11
Like for our discussion of instantons in previous lectures, we cannot do justice to this beautiful
subject here. I encourage you to read more broadly about solitons in classical field theories.
4.3 ’t Hooft-Polyakov monopoles on the Coulomb branch 41
The lower bound (4.28) on the soliton energy is also called ‘the BPS bound.’
The bound is saturated by solutions that satisfy the Bogolmonyi equations:
Bi = Di ϕ . (4.29)
This is essentially the condition for the solution for the bosonic background (Aµ , φ)
of an N = 2 SU (2) vector multiplet to preserve half of the supersymmetry.
The ’t Hooft-Polyakov monopole is a specific solution for an adjoint scalar with
boundary conditions set at spatial infinity, and the solution can be chosen to satu-
rates the BPS bound. Let us pick the VEV (4.20) for the complex scalar φ. By a
U (1)r rotation, we can choose a = |a| ∈ R, for simplicity. This is thus equivalent
to:
|a| 0
ϕ= , (4.30)
0 −|a|
up to some unimportant constant factor. This breaks SU (2) to U (1), and the
profiles for the SU (2) gauge field compatible with the VEV are then of the form:
!
U (1)
Aµ 0
Aµ = U (1) , (4.31)
0 −Aµ
up to gauge transformations. The monopole solution is then indexed by the the
magnetic flux of this abelian gauge field at spatial infinity:
Z
1 U (1)
m= dni Bi . (4.32)
2π S∞ 2
and the VEV (4.30) for the scalar field. It has energy: 12
4π
E = Mmonopole = |a|m . (4.33)
g2
In conclusion, on the Coulomb branch of the SU (2) N = 2 SYM theory, there exists
monopoles in the classical theory. Note that their mass goes like g12 , and therefore
they are very massive in the perturbative regime, g 2 1 (much more massive than
the W-boson, whose mass is given by (4.24)).
In the quantum theory, we should view the monopoles as particle excitations,
on par with the W-boson. Due to the BPS bound, the monopole with the lowest
magnetic charge is a stable particle. Note that, while the W-boson is electrically
charged under the U (1) gauge symmetry in the IR, the monopole is magnetically
charged.
Simply comparing the semi-classical masses MW ∼ |a| and Mmonopole ∼ |a| g2
, we
may suspect that the W-bosons and the monopoles become much more similar in
the strong coupling regime. This is indeed the case. As we will see next, in the
full quantum field theory, the gauge coupling becomes strong near the origin of the
Coulomb branch, and the monopoles play a role very similar to the W-bosons in
the full quantum theory.
12
Up to some numerical factor we did not keep track of.
42 References
References
[1] K. A. Intriligator and N. Seiberg, “Lectures on supersymmetric gauge
theories and electric-magnetic duality,” Nucl. Phys. Proc. Suppl. 45BC
(1996) 1–28, arXiv:hep-th/9509066 [hep-th]. [,157(1995)].
[2] Y. Tachikawa, N=2 supersymmetric dynamics for pedestrians, vol. 890. 2014.
arXiv:1312.2684 [hep-th].
[12] T. Banks and A. Zaks, “On the Phase Structure of Vector-Like Gauge
Theories with Massless Fermions,” Nucl. Phys. B196 (1982) 189–204.