<<

ITP–SB–93–23

hep-th/yymmnnn

Introduction to Supersymmetry

Martin Rocekˇ

Institute for Theoretical SUNY at Stony Brook Stony Brook, NY 11794-3840

ABSTRACT: This is a very lightly edited version of a set of lecture notes for five introductory lectures on supersymmetry given in June at the TASI summer school in Boulder CO. As such, it is often pretty telegraphic. My approach is basically quite old- fashioned, and mostly follows by Gates et al.

email: [email protected] Table of Contents:

LECTURE I.

I. Introduction to the introduction.

II. Introduction to supersymmetry.

III. Introduction to superspace.

IV. Supersymmetry in different dimensions.

V. Explicit supersymmetry algebras in D = 4, 3, 2.

LECTURE II.

I. Introduction and summary.

II. Models: Scalar multiplets

III. Super Yang-Mills .

LECTURE III.

I. Preview and summary.

II. D = 4, N = 1 models.

III. D=4 Super Yang-Mills theory.

LECTURE IV.

I. Preview and summary.

II. D = 4, N = 1 sigma Models.

III. D = 3, here we come!

IV. D = 2, nous sommes ici!

LECTURE V. Quotients and Duality

I. Preview and summary.

II. Quotients.

III. Duality

IV. Geometry of N = 2 Quotients.

V. CFT Quotients.

VI. The Surreal Pun.

2 LECTURE I.

I. Introduction to the introduction.

In the first lecture, I discuss schematically and show how to construct representations by introducing superspace. This is a coset of the Super Poincar´egroup

(SISO(D−1, 1)) mod the Lorentz group (SO(D−1, 1)), and is parametrized by coordinates x and anticommuting spinor coordinates θ. Then I construct a representation

of the SISO(D − 1, 1) algebra as superdifferential operators acting on superfields Φ(x, θ).

I also find a supercovariant derivative that (anti)commutes with (super)translations and

transforms as a spinor under the Lorentz group. I use this to define a superspace measure

that allows one to construct supersymmetric actions, as well as to define components of

superfields. Next, I describe how supersymmetry varies as one changes dimension. We will

see that starting from one (N = 1) in a higher dimension, we can get several (N > 1) in lower dimensions. Finally I’ll give the explicit supersymmetry algebras for N = 1, 2 in D = 4, 3, 2.

II. Introduction to supersymmetry.

What is supersymmetry? It is a between and with fermionic

charges and fermionic parameters. Consider a schematic action

Z S ∼ (B B + FDF/ ) ; (1)

Let’s look for an invariance of the form

δB = iεF

where F is commuting and ε is anticommuting. Dimensional analysis, or the form of the

3 action S implies

δF = −i(DB/ )ε

What is the symmetry algebra? We compute and find

µ [δε , δη]B = +2εγ ηDµB ⇒ {Q, Q} ∼P,/ (2) where Q is the of supersymmetry transformations:

δB = [iεQ, B] ,DµB = [iPµ,B] , etc.

Eq. (2) is what I mean by supersymmetry: A graded algebra with fermionic charges that anticommute to give a spacetime translation. Thus, for example, the phenomenologically generated graded symmetry that nuclear physicists call supersymmetry is not supersym- metry in my sense: it is an internal graded symmetry, and doesn’t have the form of (2).

Why is supersymmetry interesting? There a several reasons, some mathematical, and some physical. Historically, supersymmetry arose in two contexts: as a symmetry on the world sheet of the NSR formulation of the superstring, which is the only known consistent quantum theory of gravity, and as a symmetry of certain four dimensional ; in such supersymmetric theories, quadratic ultraviolet divergences are absent, and thus these theories are “natural” in a certain technical sense.

III. Introduction to superspace.

Still at a schematic level, let’s construct superspace. This will give us a more or less systematic way of studying representations of supersymmetry.

Recall how one finds representations of the Poincar´egroup ISO(D − 1, 1). One defines spacetime as the coset ISO(D −1, 1)/SO(D −1, 1). That is, for the ISO(D −1, 1) algebra

[J, J] ∼ iJ, [J, P ] ∼ iP, [P,P ] = 0 ,

4 we parametrize the quotient as

h(x) = eix·P mod SO(D − 1, 1) .

We can act by left multiplication:

h(x0) = eiw·J eix·P mod SO(D − 1, 1) = eiw·J eix·P e−iw·J mod SO(D − 1, 1)

iw·J −iw·J w = eix·(e P e )mod SO(D − 1, 1) = ei(e x)·P mod SO(D − 1, 1)

⇒ x0 = ewx for ew the vector representation of SO(D − 1, 1) induced by the

[J, P ]. Similarly,

h(x0) = eiξ·P eix·P mod SO(D − 1, 1) = ei(ξ+x)·P mod SO(D − 1, 1) ⇒ x0 = x + ξ

(Linear) representations of ISO(D − 1, 1) are found by introducing fields φ(x) that trans-

form as some matrix representation of SO(D − 1, 1):

[M, φ] ∼ φ.

We define the transformation by:

φ0(x0) = ew·M φ(x)e−w·M ,

which implies 0 δφ ≡ φ (x) − φ(x) = −δx · ∂xφ + wφ .

Identifying this as δφ = i[w · J + ξ · P, φ], we find

P = i∂x,J = ix ∧ ∂x − iM

For superspace, we just make ISO(D − 1, 1)→SISO(D − 1, 1) above; then

P → (P,Q), x → (x, θ), ξ → (ξ, ε) .

This leads to:

1 P = i∂ ,Q = i(∂ − θ∂/ ),J = i(x ∧ ∂ + θσ∂ ) − iM , x θ 2 x x θ 5 1 where σ is the 2 representation of the Lorentz group.

We find an important operator that (anti)commutes with Q, P by noting that left multiplication always commutes with right multiplication:

e−εDh(x, θ) ≡ h(x, θ)eiεQ ⇒ {D,Q} = 0, [D,P ] = 0 .

A B A+B+ 1 [A,B]+... Using the Baker-Campbell-Hausdorff theorem e e = e 2 we find

1 D = −iQ + θP/ = ∂ + θ∂/ θ 2 x

D is the covariant generalization of ∂θ. If Φ is a superfield, so is DΦ; neither QΦ nor ∂θΦ are (they do not transform simply under Q). D is a SO(D − 1, 1) spinor and obeys

{D,D} ∼ i∂/x . (3)

It has several important uses:

1) Components of a superfield are best defined by covariant expansion:

φ = Φ|, Ψ = DΦ|, . . . etc.

where | means the θ-independent projection.

2) The basic superspace invariant measure is Z dx(D)N L(Φ,DΦ,...)

where N is the number of (real) components of D. This is invariant because (D)N +1 ∼

∂x (because of (3)), and δL = iεQ(L) = −εD(L) + (∂xL-term).

IV. Supersymmetry in different dimensions.

Suppose we start with a basic in some dimension D:

{Q, Q} ∼P,/ [J, Q] ∼ iQ, [J, P ] ∼ iP, [J, J] ∼ iJ .

6 We can reduce to a lower D; that is we keep the following subalgebra:   JD−n 0 0 P → (P, 0) ,JD → , and all the Q s, 0 Tn where the T ’s are generators of some internal group G and the Q’s are now in (the spinor representation of SO(D − n, 1)) × (a representation of G). The remaining subalgebra is

{Q, Q} ∼P,/ [J, J] ∼ iJ, [J, Q] ∼ iQ, [J, P ] ∼ iP, [T,Q] ∼ iQ, [T,T ] ∼ iT .

Thus N = 1 in higher dimensions becomes, in general, higher N in lower dimensions; just how depends on the properties of the spinor representations of SO(D−1, 1). The following table summarizes the relations:

D 11 10 9 8 7 6 5 4 3 2 Spinor type M MW M W D W D W M MW Real spinor dim 32 16 16 16 16 8 8 4 2 1 Real/Complex R R R C C C C C R R N 1 2 2 2 2 4 4 8 16 (16,16) 1 1 1 1 2 2 4 8 (8,8) 1 1 2 4 (4,4) 1 2 (2,2) 1 (1,1)

Table I

Key: M–Majorana, W–Weyl, D–Dirac.

Comments: D = 4 spinors can be either Majorana or Weyl but not both; thus in D = 4, one can have either a real 4-dimensional spinor, or a complex 2-dimensional one. D = 2 is special, and allows (p, q) supersymmetry because it has real Weyl spinors; (p, q) counts the number of left and right handed spinors. (Actually, you can have this in any dimension where there is a real or pseudo real Weyl representation, e.g., D = 6, 10.)

V. Explicit supersymmetry algebras in D = 4, 3, 2.

We now examine some explicit examples of superalgebras. We begin in D = 4, where

7 ¯ † we have complex Weyl spinor supercharges Qα, Qα˙ = −(Qα) for α = (+, −). We can raise γβ γ α αβ and lower spinor indices by Cαβ = −Cβα,CαβC = δα; for example, Q = C Qβ,Qα = β 1 ¯ Q Cβα.* The spinors Q are in the (2, 0) representation of SL(2C), their conjugates Q 1 1 1 are in the (0, 2) representation, and vectors are in the (2, 2) representation. Thus vectors are naturally described as 2 × 2 matrices. Then the N = 1 has charges ¯ ¯ Qα, Qα˙ ,Pαα˙ ,Jαβ, Jα˙ β˙ which obey the following algebra: ¯ {Qα, Qα˙ } = Pαα˙ , 1 [J ,Q ] = i(C Q + C Q ) , αβ γ 2 γα β γβ α 1 [J ,P ] = i(C P + C P ) , αβ γγ˙ 2 γα βγ˙ γβ αγ˙ 1 [J ,J ] = i(C J + C J + C J + C J ) , αβ γδ 2 γα βδ γβ αδ δα βγ δβ αγ and all other commutation relations either vanish or follow by hermitian conjugation. The

spinor derivatives are:

i i D = ∂ + θ¯α˙ ∂ , D¯ = ∂¯ + θα∂ α α 2 αα˙ α˙ α˙ 2 αα˙ ¯ ¯ ¯ {Dα,Dβ} = {Dα˙ , Dβ˙ } = 0, {Dα, Dα˙ } = i∂αα˙

The algebra for N = 2 in D = 4 is essentially the same, except that Q and D get

doubled:

{D , D¯ b } = iδb∂ , (a = 1, 2). aα β˙ a αβ˙

One may also include an SU(2) internal symmetry under which the Q’s transform as

isospinors.

In D = 3, for N = 2, we just drop the “ ˙ ” from N = 1 above, impose Pαβ = Pβα, and

let Jαβ be real.

For D = 3, N = 1, there is only one real Qα, and of course Pαβ and Jαβ; the algebra

* Two frequently used identities for two component spinors are X[αβ] ≡ Xαβ − Xβα = γ −CαβX γ and X[αβγ] = 0. 8 is: {Qα,Qβ} = 2Pαβ , 1 [J ,Q ] = i(C Q + C Q ) , αβ γ 2 γα β γβ α 1 [J ,P ] = i(C P + C P + C P + C P ) , αβ γδ 2 γα βδ γβ αδ δα βγ δβ αγ 1 [J ,J ] = i(C J + C J + C J + C J ) , αβ γδ 2 γα βδ γβ αδ δα βγ δβ αγ

and the spinor derivatives are

α Dα = ∂α + iθ ∂αβ, {Dα,Dβ} = 2i∂αβ .

Finally, we consider D = 2. For N = (1, 1), we have the algebra of D = 3,N = 1 with

P+− = J++ = J−− = 0. This leaves Q±,P±±,J ≡ J+−; these obey

2 Q± = P±± , 1 [J, Q ] = ± Q , ± 2 ±

[J, P±±] = ±P±± .

The corresponding spinor derivatives are

± 2 D± = ∂± + iθ ∂±± ,D± = i∂±± .

Often, I’ll Wick rotate; then

± 2 D± = ∂± + θ ∂z,z¯ ,D± = ∂z,z¯ .

For N = (2, 2), we have

i i D = ∂ + θ¯±∂ , D¯ = ∂¯ + θ±∂ , ± ± 2 ±± ± ± 2 ±±

with 2 ¯ 2 ¯ ¯ ¯ D± = D± = 0 , {D±, D±} = i∂++ , {D+, D−} = {D−, D+} = 0.

The Wick rotated form is

1 1 D = ∂ + θ¯±∂ , D¯ = ∂¯ + θ±∂ , {D , D¯ } = ∂ . ± ± 2 z,z¯ ± ± 2 z,z¯ ± ± z,z¯ 9 ¯ a Finally, for N = (4, 4), we have four complex spinor derivatives D±a, D± obeying

¯ b b {D±a, D±} = iδa∂±± , (a = 1, 2) .

LECTURE II.

I. Introduction and summary.

In this lecture, I construct several N = 1 models in D = 3, 2. I write down actions in superspace, define components, and work out component actions. We will find actions that have the usual form for bosons and fermions, but with particular couplings dictated by supersymmetry. For example, a D = 3 or 2 scalar superfield has components:

 i  φ scalar i i Φ → ψα spinor  F i auxiliary which together form a scalar . The superspace action for a generalized

WZW-model, when expanded in components, will reveal an arbitrary bosonic WZW-model coupled to fermions in a way that is completely specified by supersymmetry. We also will explore the role of the auxiliary field F i, and will find that it is needed for two things: (1) Dynamics-independent transformations, and (2) off-shell closure of the algebra of transformations.

The next model that we’ll consider is D = 3,N = 1 super Yang-Mills theory. The basic superfield here is a spinor potential

 Aαβ - gauge potential Γα → λα - a spinor. 10 Γα has more components, but they can be gauged away by the gauge transformations

δΓα = ∇αK. In superspace, we’ll find that all other connections and all field strengths

can be constructed in terms of the basic object Γα. We’ll write down the superspace action for the usual Yang-Mills kinetic term as well as for the Chern-Simons term. The

resulting component action has just the usual bosonic terms with minimally coupled adjoint

representation fermions. The Chern-Simons term give an ordinary to the fermions.

Next, we’ll couple D = 3,N = 1 super Yang-Mills theory to , and find the usual

type of couplings with Yukawa couplings that have the same as the gauge

coupling. This is again typical of supersymmetry: supersymmetric theories are usual field

theories with particular restrictions on the types of fields and on their couplings. Finally,

I’ll briefly discuss coupling gauge fields to a nonlinear σ-model with some nonlinear field

dependent isometry generated by killing vectors.

II. Models: Scalar multiplets

Let’s start, for example, with a supersymmetric nonlinear σ-model for D = 3,N = 1:

1 R 3 2 α i j S = 8 d xD (D Φ DαΦ gij(Φ)). This describes maps from D = 3 superspace into a

target space manifold with metric g. We can easily reduce to D = 2:* recall that Dα is 2 best written in terms of its real chiral components D±,D± = ∂z,z¯. Then

1 Z S = − d2xD D (D ΦiD Φj − D ΦiD Φj)g (Φ) 4 + − + − − + ij 1 Z = − d2xD D (D ΦiD Φj + D ΦjD Φi)g (Φ) 4 + − + − + − ij 1 Z = − d2xD D (D ΦiD Φj)g (Φ) 2 + − + − ij

where the last equality follows because g is symmetric. A natural generalization of this

* We reduce the D = 3 action in the standard way: we assume that all fields are independent of one coordinate, and we drop the integration over that coordinate.

11 (special to D = 2) is

1 Z S = − d2xD D (D ΦiD Φj)(g (Φ) + b (Φ)) (4) gen 2 + − + − ij ij

where bij = −bji is an antisymmetric on the target space. Let’s find the component expansion; we define components

i i i i i i i i φ = Φ |, ψ = D+Φ |, ψ¯ = D−Φ |,F = iD+D−Φ | .

Then 1 Z  S = − d2xD (−D D ΦiD Φj − D Φi∂¯Φj)(g + b ) gen 2 + + − − + ij ij  i j k + D+Φ D−Φ (gij,k + bij,k)D−Φ 1 Z   = − d2x − (∂ψ¯i)ψ¯j + F iF j − ∂φi∂φ¯ j + ψi∂ψ¯ j (g + b ) 2 ij ij h i ¯j i ¯ j k i ¯j i j ¯k i ¯j ki + (−iF ψ + ψ ∂φ )ψ + (∂φ ψ + iψ F )ψ − iψ ψ F (gij,k + bij,k)  i ¯j l ¯k + ψ ψ ψ ψ (gij,kl + bij,kl) ,

where g = ∂ g , etc. We collect terms and integrate by parts to get: ij,k ∂φk ij 1 Z  S = d2x ∂φi∂φ¯ j(g + b ) − (ψi∇¯ ψj + ψ¯i∇ψ¯j)g 2 ij ij ij 1  − R− ψiψjψ¯kψ¯l − g (F i + iΓ−iψlψ¯k)(F j + iΓ−j ψmψ¯n) 2 ij,kl ij lk mn where ¯ i ¯ i +i ¯ j k ∇ψ = ∂ψ + Γjk ∂φ ψ . ¯i ¯i −i j ¯k ∇ψ = ∂ψ + Γjk ∂φ ψ − − Rijkl is the Riemann curvature of Γ , and  l  1 Γ±l = ± gli (b + b + b ) (5) jk jk 2 ij,k jk,i ki,j | {z } Tijk the torsion This is an N = 1 supersymmetric generalized WZW model. We will discuss this model later at great length, but let me make one comment here: if R− = 0, then this is precisely a Wess-Zumino-Witten (WZW) model.

12 We can also add a : Z Z 2 2 i i ¯j i d xD+D−V (Φ) = d xV,iF + iV,ijψ ψ

Finally, lets find the supersymmetry transformations of the components:

i i i i i δφ = −(ε+Q− − ε−Q+)Φ | = i(ε+D− − ε−D+)Φ | = i(ε+ψ¯ − ε−ψ )

(recall D = −iQ + 2θP )

i i i δψ = −(ε+Q− − ε−Q+)D+Φ | = i(ε+D− − ε−D+)D+Φ |

i i = −(ε+F + iε−∂zφ ) What is F i? Note that it enters the action only algebraically. It is called an auxiliary field.

It plays a very interesting role; its transformation is: i i i δF = −(ε+Q− − ε−Q+)(iD+D−Φ )| = (ε+D− − ε−D+)D+D−Φ |

i i = −(ε+∂ψ¯ + ε−∂ψ¯ ) . Note that δφi, δψi, and δF i are independent of the form of S; the transformations are valid

without the use of field equations, and the algebra closes without any field equations. If we i i −i l ¯k i − eliminate F by its equation of motion, F = −iΓlk ψ ψ , then δψ depends on Γ , which in turn depends on the form of S. Consider the simplest example: g + b = 1, Γ = R = 0. i i i Then on-shell F = 0, and δψ = −iε−∂zφ . Let’s check the algebra; on φ, we find i i i [δε, δη]φ = 2i(η+(iε+∂φ¯ ) − η−(−iε−∂φ ))

i i = +2(ε+η+∂φ¯ + ε−η−∂φ ) , which is fine. However, on ψi we have: i i i [δε, δη]ψ = −iη−∂(iε+ψ¯ − iε−ψ ) − η ↔ ε

i i = +2ε−η−∂ψ + (η−ε+ − ε−η+)∂ψ¯

i = 2(ε+η+∂¯ + ε−η−∂)ψ | {z } good i i − 2ε+η+∂ψ¯ + (η−ε+ − ε−ψ+)∂ψ¯ | {z } equations 13 Without auxiliary fields, in general, the algebra only close on-shell, that is, modulo equa-

tions of motion. Note that the action is still invariant without any equations of motions;

since these are defined by extremizing the action, they can never be used in verifying a

symmetry. Sadly, there are supersymmetric systems that we don’t really understand in

superspace, i.e., we only have on-shell representations of the algebra.

We will eventually run into this, but let me say now that this is typically the situation

for high D and/or N.

Another comment, referring back to off-shell supersymmetry transformations: how do

you break supersymmetry ? Well, you need hδ(something)i= 6 0. Note that φi 6= 0 does

not break supersymmetry (φ enters the transformation laws only through its derivatives),

but F i 6= 0 does. This is a very generic feature.

III. Super Yang-Mills theory.

Let’s look at some more models. Back in D = 3,N = 1, we can consider Super

Yang-Mills theory. We do this by covariantizing the spinor derivative Dα:

Dα → ∇α ≡ Dα − iΓα .

We can construct everything in terms of the basic object Γα. For example, we can define

the vector potential Γαβ by

1 {∇ , ∇ } = 2i∇ ≡ 2i(∂ − iΓ ) ⇒ Γ = − (iD Γ + {Γ , Γ }) α β αβ αβ αβ αβ 2 (α β) α β

(which means we absorb Fαβ into Γαβ). The components follow from the pieces of Γα that remain after we use the gauge transformation

δΓα = ∇αK 14 as follows: K| is the component gauge parameter

∇αK| gauges away Γα|

2 α ∇ K| gauges away D Γα| This leaves the component gauge field Aαβ and spinor λβ:

α D(αΓβ)| ∼ Aαβ D DβΓα| ∼ λβ .

A more careful definition of the components is given below.

The Bianchi identities

[∇(α{∇β, ∇γ)}] = 0

imply

[∇α, ∇βγ] ≡ −iFα,βγ = Cα(βFγ)

(that is F(α,βγ) = 0), which allows us to compute the superfield strength 1   F = i(DβΓ − ∂ Γβ) + [Γβ, Γ ] . α 3 αβ αβ αβ

We also can find the usual field strength as follows:

i i [∇ , ∇ ] = − [∇ , {∇ , ∇ }] = ({∇ , [∇ , ∇ ]} + {∇ , [∇ , ∇ ]}) , αβ γδ 2 αβ γ δ 2 γ δ αβ δ γ αβ

and hence i   [∇ , ∇ ] = C ∇ F + C ∇ F + C ∇ F + C ∇ F αβ γδ 2 δα γ β δβ γ α γα δ β γβ δ α . −i   = C (∇ F + ∇ F ) + α ↔ β + γ ↔ δ 4 αγ β δ δ β Thus we have found the spacetime Yang-Mills field strength in terms of the spinor field α β α strength Fα.* (Note that we also have:{∇ , [∇ , ∇αβ]} = 0 and hence ∇ Fα = 0.)

For an action, we have: Z 1 3 α 1 β S = d xD Dα T r(F F ) . 4 β * We use the two component spinor identities from the footnote of Section V, Lecture I.

15 We can also add a super Chern-Simons term: Z 1 3 α m β 1 γ S = d xD Dα T r(Γ (F − [Γ , Γ ])) . CS 2 g2 β 6 βγ

This is, of course, a special feature of D = 3. We can define components by projection as follows: The gauge field is

Γαβ| = Aαβ , which leads to a component field strength

∼ ∇(αFβ)| .

The physical spinor is

Fα| = λα|.

As described above, the rest of Γα can be gauge away.

In components, the sum of the two actions is simply a Yang-Mills field coupled to a massive adjoint representation spinor with mass equal to the topological mass of the

Yang-Mills field given by the Chern-Simons term.

Finally, let us consider matter couplings; that is, for some scalar fields transforming as i i A i δΦ = i(KΦ) ≡ iK (TAΦ) we couple to Yang-Mills fields by the usual minimal coupling i i i i prescription: DαΦ → (∇αΦ) = DαΦ − i(ΓαΦ) . This gives rise to a matter action 1 Z h i 1 Z h i S = − d3xDαD (∇βΦ) (∇ Φ)i = d3xDαD Φ (∇β∇ Φ)i 8 α i β 8 α i β 1 Z h  i = d3x ∇α∇ Φ (∇β∇ Φ)i 8 α i β 1 Z h i = d3x (∇α∇ Φ) (∇β∇ Φ)i + 2(∇αΦ) (∇ ∇β∇ Φ)i + Φ (∇α∇ ∇β∇ Φ)i 8 α i β i α β i α β 1 Z h1 1 i = d3x F F i − iψα(∇ ψβ)i − (∇αβφ) (∇ φ)i + i(ψα(λ φ)i − φ (λαψ )i) 2 4 i i αβ 2 i αβ i α i α with F i auxiliary again. More generally, we can consider an isometry that acts on some i A i i scalar fields by δΦ = K XA(Φ), where XA are killing vectors satisfying

Xi;j + Xj;i = 0 . 16 R 3 α β i j This leaves the action S = d xD DαgijD Φ DβΦ invariant, as can be seen as follows: Z   3 α β i A j β i j A k δS = d xD Dα 2gijD Φ Dβ(K XA) + gij,kD Φ DβΦ K XA Z  3 α A β i j β i j A k = d xD Dα 2K D Φ DβΦ XAi;j − D Φ DβΦ K XA(gij,k + gik,j − gjk,i)  β i j A k + gij,kD Φ DβΦ K XA

= 0 . i i A i Now minimal coupling takes the form (∇αΦ) = DαΦ − iΓα XA(Φ).

LECTURE III.

I. Preview and summary.

I describe D = 4, N = 1 supersymmetric models, in particular, I write down the most

general renormalizable D = 4 supersymmetric model. This will involve scalar multiplets

and Yang-Mills theories.* However, the superspace formulation of these theories is quite

different from the D = 3, N = 1 theories described in the previous lecture. We will see

that we will have to introduce superfields that are independent of some θ’s — constrained

superfields. These exist for D = 4,N = 1 because the superderivatives have an abelian

subalgebra (that is not real). These “chiral” superfields resemble complex D = 3, N = 1

scalar superfields, and describe matter with a complex scalar, a Weyl

spinor, and a complex auxiliary field. The Yang-Mills multiplets are described by an

unrestricted superfield. The structure of the theory is somewhat novel and is determined

by the structure of the matter multiplets: that is, by the principle that the restricted

(chiral) superfields that we use to describe matter should be allowed to carry Yang-Mills

. * Supersymmetric standard models are of this form; so is the only (proven) finite D = 4 theory, N = 4 Super Yang-Mills theory.

17 II. D = 4, N = 1 models.

By way of motivation, let’s consider some dimensional analysis. In D = 4, we have ¯ R 4 α ¯ α˙ ¯ 4 spinor derivatives (see table I): Dα, Dα˙ . Thus the measure is d xD DαD Dα˙ . Now ¯ 1 {Dα, Dα˙ } = i∂αα˙ , which implies the mass dimension of [D] = 2. Thus the measure has dimension 2. We expect, from our experience in D = 3, that φi = Φi|, and hence the physical scalar has [φ] = 1. Then a dimensionless action should have the form Z d4xD2D¯ 2|Φ|2.

αβ˙ But for an unconstrained superfield, this gives no dynamics; we want a ∂ φ∂αβ˙ φ term. We need a new idea. Let’s have a look at the D = 4 algebra of spinor derivatives:

¯ ¯ ¯ {Dα, Dα˙ } = i∂αα˙ , {Dα,Dβ} = 0, {Dα˙ , Dβ˙ } = 0.

Notice that by themselves, the D’s or the D¯’s are abelian. In particular, there is no inconsistency with constraining a superfield to be chiral:

¯ Dα˙ Φ = 0.

Loosely, this means Φ depends on θ but not θ¯; I say loosely, because the constraint is ¯ ¯ Dα˙ Φ = 0, not ∂α˙ Φ = 0. Φ had better be a complex superfield, because we have:

DαΦ¯ = 0

¯ ¯ and if Φ = Φ, then {Dα, Dα˙ }Φ = 0, which implies ∂αα˙ Φ = 0, i.e., Φ is a constant and that is not good for field theory. What are the components of Φ?

1 φ = Φ| , ψ = D Φ| ,F = DαD Φ| . α α 2 α

Does this solve our problem? Let’s see: 1 Z 1 Z S = d4xD2D¯ 2(ΦΦ)¯ = d4xD2(ΦD¯ 2Φ)¯ 0 4 4 1 Z = d4x(D2ΦD¯ 2Φ¯ + 2DαΦD D¯ 2Φ¯ + ΦD2D¯ 2Φ)¯ . 4 α 18 We need a few identities: ¯ 2 ¯ 2 ¯ α˙ DαD = D Dα − 2i∂αα˙ D

and 2 ¯ 2 ¯ αα˙ ¯ D D Φ = 2∂ ∂αα˙ Φ

which implies Z 1 S = d4x(F F¯ − iψα∂ ψ¯α˙ + φ∂αα˙ ∂ φ¯) , 0 αα˙ 2 αα˙ which is just what we want! Other constraints are possible; there is a systematic way to study all linear constraints, but it is far too technical to describe here (see Gates et al., section 3.11.a).

How about and interactions? Well, now we’re hoisted by our own petard:

R d4xD2D¯ 2|Φ|2 is a kinetic term, not a mass! How do we get no derivatives? We need a new superspace measure. We can get a clue as follows: The kinetic action can be written as R d4xD2(ΦD¯ 2Φ);¯ this almost looks like D = 3. Can we use D2 as a measure? Not in general, but for Lagrangians that are chiral, we can!

This leads to the following theorem:

¯ R 4 2 For L such that DL = 0,Schiral = d xD L is a superinvariant.

Proof: Z 4 2 α α˙ ¯ δSchiral = i d xD ( Qα + ¯ Qα˙ )L Z 4 2 α α˙ ¯ = − d xD ( Dα + ¯ Dα˙ )L + total derivatives Z 4 2 α = d xD Dα( L) = 0 ,

where the last equality follows from D3 = 0. R d4xD2 is a chiral measure good for chiral

Lagrangians. Now we’re in great shape: 1 Z S = d4xD2V (Φi) + c.c. V 2 Z 1 = d4x(V F i + V ψiαψj ) + c.c. ,i 2 ,ij α 19 1 2 1 3 For V = 2mΦ + 3!λΦ , we get Z  1 1  S = d4x m( ψαψ + φF ) + λ(φψαψ + F φ2) + c.c. m,λ 2 α 2 α (There is an old nomenclature that is still used: terms constructed with R d4xD2D¯ 2 are

called “D” terms, whereas terms constructed with R d4xD2 + c.c. are called “F ” terms.)

Eliminating F , we get Z h 1 1 d4x φ¯( ∂αα˙ ∂ − m2)φ − iψ¯α˙ ∂ ψα + m(ψαψ + ψ¯α˙ ψ¯ ) 2 αα˙ αα˙ 2 α α˙ 1 1 1 i − mλ(φφ¯2 + φφ¯ 2) − λ2φ2φ¯2 + λ(φψαψ + φ¯ψ¯α˙ ψ¯ ) 2 4 2 α α˙ This is the most general renormalizable supersymmetric model with one complex scalar φ

and one Weyl spinor ψ.

We can easily generalize to many Φ’s: Z Z Z 4 2 ¯ 2 i ¯ 4 2 i 4 ¯ 2 ¯ ¯ i S = d xD D Φ Φi + d xD V (Φ ) + d xD V (Φ )

for V (Φi) a cubic polynomial.

III. D=4 Super Yang-Mills theory.

As for the scalar multiplet, we immediately have a problem: where can we start? We

can’t use minimal coupling, as there are no derivatives in the Lagrangian! Let’s examine

the constraints on the matter superfields. Suppose we start with a chiral superfield; we

may ask that it remain chiral after a gauge transformation. This implies that the gauge

parameter must be chiral:

¯ i ¯ i i A i j Dα˙ Φ = 0 ; we ask Dα˙ (δΦ ) = 0 , but δΦ = iΛ (TA) j Φ , | {z } i Λj ¯ which implies that Λ better be chiral. This implies, however, that Dα˙ is gauge-covariant as it stands: ¯ i ¯ i j i ¯ j δ(Dα˙ Φ ) = Dα˙ (iΛjΦ ) = iΛjDα˙ Φ . 20 ¯ ¯ This implies: Fα˙ β˙ = i{∇α˙ , ∇β˙ } = i{Dα˙ , Dβ˙ } = 0. But what about Dα? We certainly i i j i j need Fαβ = i{∇α, ∇β} = 0, but Dα is not covariant: δ(DαΦ ) = i(DαΛj)Φ + iΛjDαΦ . −V V To satisfy Fαβ = 0, we must have ∇α = e Dαe ; how should V transform to fix things? Consider

δ(∇αΦ) = (δ∇α)Φ + i[∇α, Λ]Φ + iΛ∇αΦ which implies

δ∇α = −i[∇α, Λ] . this is accomplished by:

¯ (eV )0 = eiΛeV e−iΛ ,V = V † ⇒ δV = i(Λ¯ − Λ) + ...

Then the gauge covariant action is

Z d4xD2D¯ 2(Φ¯eV Φ)

¯ where Φ0 = eiΛΦ, Φ¯ 0 = Φ¯e−iΛ. This may seem like magic, but it is as deductive as I can make it. We see that V acts as a “bridge” or “converter” between chiral and antichiral representations of the gauge group. What does V contain?

¯ 2 ¯ 2 V |,DαV |, Dα˙ V |,D V |, D V | can be gauged away by ¯ ¯ 2 ¯ 2 ¯ ImΛ|,DαΛ|, Dα˙ Λ|,D Λ|, D Λ, respectively. What remains is the physical vector

¯ Aαα˙ ∼ [Dα, Dα˙ ]V | , the physical spinor ¯ 2 ¯ 2 ¯ λα ∼ D DαV | , λα˙ ∼ D Dα˙ V | , 21 and a hermitian auxiliary field: ¯ α˙ 2 ¯ D ∼ D D Dα˙ V | .

The full nonlinear expressions for these components can be found from the Bianchi iden- tities and the constraints:

¯ ¯ ¯ {∇α, ∇β} = {∇α˙ , ∇β˙ } = 0 , {∇α, ∇α˙ } = i∇αα˙ .

The first two constraints can be interpreted as integrability conditions for the existence of chiral and anti-chiral superfields, and the third is merely a matter of convenience: it defines the vector connection in terms of the spinor connections, just as in D = 3. The

Bianchi identities now imply:

i [∇ , ∇ ] = C W ,W = D¯ α˙ D¯ (e−V D eV ) α˙ ββ˙ α˙ β˙ β β 2 α˙ β

The covariant expressions for the components of the Yang-Mills multiplet are:

¯ ¯ Aαα˙ = Γαα˙ | , λα = Wα| , λα˙ = Wα˙ | ,

α α˙ α˙ D = ∇ Wα| , fαβ ≡ ∂α Aβα˙ + ∂β Aαα˙ = ∇(αWβ)| .

¯ We note that Dα˙ Wβ = 0. This allows us to write a consistent gauge invariant action for the Yang-Mills field: Z 1 4 2 α S = tr d xD (W Wα) . YM 4g2

This is real up to a topological F F˜ term. We can also add a term

Z d4xD2D¯ 2tr(νV )

for an abelian subgroup; I’ll leave it as an exercise to prove that this is gauge invariant

when ν commutes with all the TA’s. Such a term is called the Fayet-Illiopoulos term, and leads to spontaneous .

22 So finally, as promised, we have the most general D = 4 renormalizable supersymmetric action: Z h 1 S = d4x − (∇αα˙ φ)i(∇ φ¯) − iψαi(∇ ψ¯α˙ ) + i(φ¯ (λαψ )i − ψ¯α˙ (λ¯ φ)i) 2 αα˙ i αα˙ i i α i α˙ 1 + φ¯ (Dφ)i + F iF¯ + tr(−iλα[∇ , λ¯α˙ ] − fαβf + D2) i i αα˙ 2 αβ 1 1 i + tr(νD) + (V F i + V ψαiψj + V¯ iF¯ + V¯ ijψ¯α˙ ψ¯ ) . ,i 2 ,ij α , i 2 , i αj˙ Eliminating the auxiliary fields F, F¯, and D gives a scalar potential:

1 −[V V¯ i + (φ¯(T φ) + T r(νT ))2]. ,i , 4 A A

LECTURE IV.

I. Preview and summary.

In this lecture I discuss the relation between supersymmetry and the target space geometry of supersymmetric sigma-models. The main motivations for this are

1) theory — many conformal field theories (CFT’s) can be described as σ-models.

This is an important way to give a spacetime interpretation to conformal field theories.

(The other known way is Landau-Ginzburg models.)

2) Mathematics — supersymmetric sigma-models have led to important results in complex

manifold theory, and in particular, to the hyperk¨ahlerquotient.

For the applications, we want to study D = 2, N > 1 models, but for the mathematical applications, as well as for pedagogical reasons, I’ll start at D = 4 and work my way down. D = 4, N = 1 sigma models expressed in terms of chiral superfields have a unique and natural superspace Lagrangian: L ∼ K(Φi, Φ¯ i). Working out the relevant part of the component action, we’ll find that these describe manifolds that are

23 2 i ¯ j K¨ahler: ds = K,i¯dΦ dΦ . I’ll explain what K¨ahlergeometry is. It is possible to show that this is a general result for D = 4 σ-models, and doesn’t depend on using chiral

superfields. However, it is easier to do this after descending to D = 3. There we can

formulate the general question in N = 1 superspace; we’ll find that extra supersymmetries

correspond to covariantly constant hermitian complex structures on the target space. This

proves that we can have N = 2 supersymmetry if and only if the target space manifold is

K¨ahler, and N = 4 supersymmetry if and only if the target space manifold is hyperk¨ahler.

Descending further to D = 2, we have the option of adding torsion. One way to do this is

to take advantage of the existence of new kinds of constrained superfields: twisted chiral

superfields, which are left chiral and right antichiral, and can exist only in D = 2. However,

this does not give the most general model, and the general case is still not fully understood.

The final classification theorem is summarized in the table below:

D 6/5 4 3 2 Target Space (T=0) Dimension Geometry N 1 2 4 (4,4) 4k Hyperk¨ahler 1 2 (2,2) 2k K¨ahler 1 (1,1) k Riemannian

Table 2

Here k is an arbitrary integer.

For non-zero torsion, new geometries arise; these don’t have nice names, so I haven’t put them in the table. I will try to derive as much of this table as possible.

II. D = 4, N = 1 sigma Models.

The most general bosonic D = 4 σ-model can be described by an action:

1 Z S = − d4x(∂αα˙ φi∂ φj)g (φ) 0 2 αα˙ ij 24 and describes maps from 4 dimensions into a target space with a metric gij. On dimensional grounds (counting D’s), the superspace analog for chiral superfields must be:

1 Z S = d4xD2D¯ 2K(Φi, Φ¯ i) , D¯ Φi = D Φ¯ i = 0. 4 α˙ α

Let’s work out the terms in this corresponding to S0: Z 1 4  2 2 i α α˙ i j S = d x K¯D D¯ Φ¯ − 2K (D D¯ Φ¯ )(D D¯ Φ¯ ) 4 i ¯ı¯ α α˙  + F terms + ψ terms | 1 Z , = d4x(K ∂αα˙ ∂ φ¯i + K ∂αα˙ φ¯i∂ φ¯j) + ... 2 ¯ı αα˙ ¯ı¯ αα˙ 1 Z = − d4xK ∂αα˙ φ¯i∂ φj + ... 2 ¯ıj αα˙ where we have integrated by parts in the last step. Thus the metric gij is written as the matrix of second derivatives of the function K:

∂2K g ∼ , ds2 = K dφidφ¯j ij ∂φi∂φ¯j i¯ This is the form of an arbitrary K¨ahler metric. Notice that the geometry depends only on the Hessian of K; in particular, shifting K by the real part of a f(φ)+f¯(φ¯) does not change the metric. This is also true for the superspace action, because the full superspace measure R d4x D2D¯ 2 annihilates a chiral or antichiral integrand.

We have shown that for chiral superfields, the N = 1, D = 4 target space must be

K¨ahler; conversely, for any K¨ahler manifold, we can write an N = 1, D = 4 σ-model in terms of chiral superfields. This result is originally due to Zumino. To proceed further, we need to know more about K¨ahler geometry. Notice that the manifold is complex: the coordinates can be chosen to be complex (φ and φ¯). This gives an intuitive explanation of why supersymmetry knows about target space geometry: chiral superfields correspond to complex coordinates.

We can define a natural tensor:

j j ¯j j ¯ ¯ Ji = iδi ,Ji = J¯ı = 0,J¯ı = −iδ¯ı . 25 Then J has the following important properties:

0. The manifold on which it is defined is even (real) dimensional.

1. J2 = −1 (J is an almost complex structure).

J ¯ K 2. Let ∂±I = ∂I ± iJI ∂J , {I} = {i, j}. Then [∂±I, ∂±J ] = CIJ ∂±K (J is an integrable complex structure.)

J J 3. gIJ JK = −gKJ JI (The metric is hermitian.)

J 4. ∇IJK = 0 (J is covariantly constant).

K 5. This implies [∇I, ∇J ]JL = 0, which further implies

KM M M K 6. RIJ JL + RIJL JM = 0 (The of the Levi-Civita connection is not O(2K) but U(K).)

These properties characterize K¨ahlermanifolds.

III. D = 3, here we come!

As described in the first lecture, we can reduce to D = 3. We let:

¯ ¯ ¯ Dα → Dα, Dα˙ → Dα, ⇒ {Dα, Dβ} = i∂αβ = i∂βα .

The action remains

1 Z S = d3xD2D¯ 2K(Φi, Φ¯ i), D¯ Φi = D Φ¯ i = 0 4 α α

But, because we’re in D = 3, we can write N = 1 σ-models:

1 Z   S = d3xD2 DαΦiD Φjg (Φ) 8 α ij

(Recall lecture 2). Here, i, j are real indices. For future use we write down the superfield

α i i α j k equation of motion: D DαΦ + ΓjkD Φ DαΦ = 0.) Let’s try and see what happens 26 if we look for a second supersymmetry using N = 1 components. We want it to anti- commute with the first supersymmetry , so we should write it with Dα; the most general transformation that we can write down is:

i i α j δΦ = Jjε DαΦ ; (6) let’s study the consequences of

(A) imposing the supersymmetry algebra on Eq. (6); this will lead to conditions 0.+1.+2. above, and hence to the conclusion that the target space manifold is complex.

(B) imposing invariance of S on Eq. (6); this will lead to 3. + 4., and to the conclusion that the manifold is K¨ahler.

(A) Starting only from the supersymmetry algebra, we have

i i α j β k i k β l α j [δε, δη]Φ = Jjη Dα(Jkε DβΦ ) + Jj,kJl (ε DβΦ )(η DαΦ ) − η ↔ ε i j α β k i j α β l k = −JjJkη ε {Dα,Dβ}Φ − JjJ[k,l]η ε DαΦ DβΦ i k α β l j + J[j|,kJ|l]η ε DβΦ DαΦ which leads to

i j i −JjJk = δk ⇒ (conditions 0. + 1.) , and

i j j i JjJ[k,l] + J[kJl],j = 0 ⇒ (condition 2.) .

(B) Invariance of the action implies:

Z 1 3 2h β l i α i i α j k i 0 = δS = d xD − 2ε (DβΦ ) Jl gij(D DαΦ + ΓjkD Φ DαΦ ) . 8 | {z } Jjl

A nontrivial calculation reveals that this vanishes if and only if Jil = −Jli, which implies j j condition 3. (the metric is hermitian), and if Jil,k − JijΓlk − JjlΓik = 0, which implies condition 4. (the metric is K¨ahler).

27 So any N = 2,D = 3 model is K¨ahler, and hence can be written with chiral superfields, and hence the same is true for any N = 1,D = 4 model.

How about N = 3,D = 3? This means there are two complex structures J1,J2, and the metric is K¨ahlerwith respect to both of them. Imposing the supersymmetry algebra

2 implies [δ1, δ2] = 0 and hence J1J2 + J2J1 = 0. This implies that (J1J2) = −1, and hence

J3 ≡ J1J2

is another complex structure, and generates yet another symmetry (it is easy to show J3 is hermitian and covariantly constant). So N = 3 ⇒ N = 4, and the target space manifold

is hyperk¨ahler. Hyperk¨ahler geometry is characterized by the existence of 3 covariantly

constant complex structures Jx, obeying the algebra of the imaginary quaternions:

{Jx,Jy} = −2δxy .

IV. D = 2, nous sommes ici!

We now reduce once more to reach two dimensions. The resulting algebra of spinor

derivatives is:

2 2 ¯ ¯ ¯ ¯ D+ = D− = {D+,D−} = {D+, D−} = 0 , {D+, D+} = ∂, {D−, D−} = ∂ .

Again, we can define a chiral superfield that obeys: D¯±Φ = 0. But in D = 2, and only in

D = 2, {D¯+,D−} = 0, so we can also have twisted chiral superfields:

D¯+χ = D−χ = 0 ,

D+χ¯ = D¯−χ¯ = 0 .

Now let’s consider Z d2xD2D¯ 2K(Φi, χa) .

28 by the usual manipulations, this gives a bosonic part with a metric and in addition an

antisymmetric tensor b-field:

gi¯ = Ki¯, ga¯b = −Ka¯b, bia¯ = −Kia¯, ba¯ı = Ka¯ı.

Two obvious questions arise: is this the most general (2,2) model? What kind of geometry

describes its target space?

To get an answer, we go back to N = 1 superspace. Because we are in D = 2, we can

have separate left and right handed transformations with parameters ε+ and ε−:

I (+)I + J (−)I − J δΦ = JJ ε D+Φ + JJ ε D−Φ . (7)

Now we impose the supersymmetry algebra, and find

1. J2(±) = −1.

2. J(±) are integrable.

3. There is a term that involves the field equations and [J(+),J(−)]:

I [δ(+),δ(−)]Φ (−)I − J = δ(+)(JJ ε D−Φ ) − (+ ↔ −) (−)I − (+)J + K (−)I (+)K + L − J = JJ ε D−(JK ε D+Φ ) + JJ,K JL ε D+Φ ε D−Φ − (+ ↔ −)  (+)I (−)J (−)I (+)J  + − K = JJ JK − JJ JK ε ε D+D−Φ  (+)K (−)I (−)I (+)K + − ¯ L J − J[L J|J|,K] + J[K J|L|,J] ε ε D+Φ D−Φ

If [J(+),J(−)] = 0 and if this is integrable, then the algebra closes off shell. Then we can conclude that the model can be described in terms of chiral and twisted chiral superfields, and has the geometry descibed above. Otherwise, we need some auxiliary superfields.

For N = 2 some cases are known, but the general case is not solved. Note that when

J(+) = J(−), we’re fine: that’s the K¨ahler case.

29 Let’s now impose invariance of the action in Eq. (4) under the transformations in Eq.

(7): we find

4. g is bihermitian, that is, hermitian with respect to both J(±).

(±) (±)J ± 1 5. ∇I Jk = 0 with respect to a connection Γ with torsion T = 2db (see Eq. 5).

± The holonomy of the curvature with torsion RIJKL is U(k). This is a lot of inter- esting structure that the mathematicians haven’t analyzed yet, and that we don’t fully

understand.

For N = 4, we find left quaternionic structures and right quaternionic structures. This

structure is also very rich and interesting. For these models, the β-functions vanish. Some things are known about the off-shell structure, but a lot remains obscure.

LECTURE V. QUOTIENTS AND DUALITY

I. Preview and summary.

This time, I want to describe two closely related notions: quotients and duality. I won’t do fancy mathematics. The way physicists perform a quotient is very simple: gauge the symmetry that you want to quotient by, without a kinetic term for the gauge field, and then integrate out the gauge field. I’ll describe this for D = 2,N = 1, 2. A duality

transformation is just as simple: gauge the symmetry that you want to transform with

respect to, without a kinetic term, and add a Lagrange multiplier to constrain the gauge

field to be pure gauge; then integrate out the gauge field. In the N = 2 case, we’ll see that

duality swaps chiral superfields for twisted chiral ones, and vice-versa. Then I’ll briefly

discuss the geometry of N = 2 quotients, which is another very beautiful example of how

supersymmetry “knows” deep mathematics on the target space. Then, even more briefly,

30 I’ll discuss CFT quotients, and explain how duality transformations can be understood as

CFT quotients. I’ll end with a translation of a bad Swedish pun.

II. Quotients.

I will work with U(1) only; general groups are discussed in the literature (see Hull in particular) and U(1) is interesting enough.

Let’s start with a bosonic σ-model Z 2 i ¯ j S = d x(gij(φ) + bij(φ))∂φ ∂φ .

This is invariant under a symmetry δφi = Xi(φ) if

i 1. X is a killing vector Xi;j + Xj;i = 0, i.e., (LXg = 0) and

i i i 2. X bjk,i + X,jbik + X,kbji = w[k,j] for some wj, i.e., (LXb = dw).

We can always choose coordinates such that Xi is a constant, i.e., the symmetry acts only on one coordinate φ0 by δφ0 = α. Then condition 1. implies that g is independent

0 of φ , and contion 2. implies that we can find a 1-form Ωi such that shifting bij by a curl 0 Ω[i,j] makes b also independent of φ . Then the action takes the form (where now i, j do not run over φ0): Z 2 0 ¯ 0 0 ¯ i i ¯ 0 i ¯ i S = d x(g00∂φ ∂φ + (g0i + b0i)∂φ ∂φ + (gi0 + bi0)∂φ ∂φ + (gij + bij)∂φ ∂φ )

(i, j 6= 0) .

We can gauge this by minimal coupling: ∂φ0 → ∂φ0 +A, ∂φ¯ 0 → ∂φ¯ 0 +A¯; however, there is

an ambiguity, peculiar to D = 2, which arises as follows: The action S is invariant under

shifts bij → bij + Ω[i,j]. However, under such a shift, the Lagrangian changes by a total derivative. After gauging, the resulting A-dependent term is not a total derivative, and

contributes a gauge invariant term U(φi)(∂A¯ − ∂A¯). In CFT, this ambiguity is fixed by | {z } F (A)

31 the requirement of conformal invariance. Eliminating A + A¯ by their field equations, and substituting back into the action S, one finds: Z 2 −1 i ¯ j SQ = d x(gij + bij − (gi0 + bi0)g00 (g0j + b0j))∂φ ∂φ Q −1 Q −1 ⇒ gij = gij − g00 (gi0g0j + bi0b0j) , bij = bij + g00 (gi0b0j + bi0g0j .

Q Q Notice that gij and bij change non-trivially under b0a → b0a − U,a; this is precisely the ambiguity noted above.

This quotient can be used to construct σ-models, e.g., CP n models (S2n+1/U(1)).

The generalization to N = 1 superspace is straightforward: we just change the measure, and replace ∂, ∂,¯ A, A¯ by D+,D−, Γ+, Γ−.

N = 2 superspace is more interesting. Consider the case with chiral and twisted chiral multiplets; that is not the most general N = 2 case, but it is interesting and includes the general K¨ahler case. Recall the action

Z 2 2 2 i a d xD D¯ K(Φ , χ ), D¯+Φ = D¯−Φ = 0, D¯+χ = D−χ = 0

Note that this action is invariant (up to total derivatives) under the following shifts:

K → K + [f(Φ, χ) + g(Φ, χ¯) + c.c.]

This again leads to some ambiguities in gauging. The only that we can gauge are those that respect the constraints, i.e., δΦi = Xi(Φ) or δΛa = Y a(Λ). Without loss of generality, we can choose δΦi = Xi(Φ). Furthermore, we can choose coordinates and shifts such that there is some Φ0 for which the action becomes

K = K(Φ0 + Φ¯ 0, Φi, χa)

Then gauging is simple: Φ0 + Φ¯ 0 → Φ0 + Φ¯ 0 + V ; recall

δΦ0 = iΛ , δΦ¯ 0 = −iΛ¯ , δV = i(Λ¯ − Λ) .

32 We also include an explicit Fayet-Iliopoulos term −cV , although that is really just a gauging

of a “shift term” −(Φ0 + Φ¯ 0).* Thus we have: Z   d2xD2D¯ 2 K(φ0 + φ¯0 + V, Φi, χa) − cV

Eliminating V , we get:

i a i a i a KQ = K(V (Φ , χ ), Φ , χ ) − cV (Φ , χ )

i a ∂K i a with V (Φ , χ ) determined by ∂V (V, Φ , χ ) = c. Pn i ¯ i Φ0+Φ¯ 0 Pn i i For example: consider K = i=0 Φ Φ = e (1+ i=1 ϕ ϕ¯ ). Then KQ = c ln(1+ Pn i i n i=1 ϕ ϕ¯ ) is the K¨ahler potential for CP .

III. Duality

Again, let’s start with the bosonic case. We take the gauged action Z 2 h 0 ¯ 0 ¯ 0 ¯ i i ¯ 0 ¯ S = d x g00(∂φ + A)(∂Φ + A) + (g0i + b0i)(∂φ + A)∂φ + (gi0 + bi0)∂φ (∂φ + A) i ¯ ji + (gij + bij)∂φ ∂φ and add a term φ˜(∂A¯ − ∂A¯ ). Integrating out φ˜ implies ∂A¯ − ∂A¯ = 0 and hence, A = ∂λ;

after the shift φ0 + λ → φ0 we get back the original action. Actually there are subtleties:

F = 0 ⇒ A = ∂λ only locally. If φ˜ is periodic, then we get A = ∂λ with λ periodic. If

φ˜ and φ have the right periods, we get exact equivalence; otherwise, we get . (A

discussion of orbifolds and why the last claims I made are true is beyond the scope of these

lectures).

Let’s consider what happens if we integrate out the gauge field to get the dual model;

we find: Z ˜ 2 h 1 ˜¯˜ 1 ˜¯ i 1 i ¯˜ Q Q i ¯ ji S = d x ∂φ∂φ + (g0i + b0i)∂φ∂φ − (gi0 + bi0)∂φ ∂φ + (gij + bij)∂φ ∂φ g00 g00 g00 1 b0i ˜ g0i Q ˜ Q ⇒ g˜00 = , g˜0i = , b0i = , g˜ij = gij, bij = bij . g00 g00 g00 * Actually, c can be f(χ) + f¯(¯χ) for any f when there are twisted chiral fields χ in the theory.

33 Notice that b0i and g0i exchange values, and that the quotient metric and b-field enter.

Once more, the N = 1 generalization is tediously straightforward. The N = 2 story is very interesting. We need the analog of φF˜ (A). In D = 4, we had a superfield strength

2 Wα = D¯ DαV (in the abelian case). It turns out that in D = 2 we can take a complex scalar superfield strength

W = D¯+D−V ;

−V V this is invariant under δV = i(Λ + Λ).¯ (The non-abelian version of W is D¯+(e D−e ),

as follows from {∇¯ +, ∇−} = W .) So we take

0 0 i a K(V + Φ + Φ¯ , Φ , χ ) + iΨD¯+D−V + iΨ¯ D+D¯−V

where Ψ, Ψ¯ are unconstrained superfields. Integrating by parts and shifting V harmlessly,

we find

K(V, Φi, χa) − (˜χ + χ˜¯)V

whereχ ˜ = +iD¯+D−Ψ, χ˜¯ = +iD+D¯−Ψ¯ are twisted chiral and antichiral superfields re- spectively. Now eliminating V , we find

K˜ = K(V (Φi, χa, χ˜ + χ˜¯), Φi, χa) − (˜χ + χ˜¯)V where V (Φi, χa, χ˜ + χ˜¯) is found from

∂K =χ ˜ + χ˜¯; ∂V

K˜ is the Legendre transform of K. Notice that the chiral superfield Φ0 has been replace

by the twisted chiral superfieldχ ˜.

One final note here: For the reverse duality, that is if we want to replace a twisted

chiral gauge field by a chiral one, we need a twisted gauge multiplet. This is found by

switching D− with D¯− while leaving D+ alone. 34 IV. Geometry of N = 2 Quotients.

N = 2 supersymmetry requires an even-dimensional target space. A component quo- tient by U(1) removes one dimension, so you must lose N = 2 supersymmetry. How does the N = 2 quotient manage to preserve N = 2 supersymmetry? The answer is: symplectic

reduction. In addition to a quotient, one imposes a constraint. The N = 2 gauge multiplet

has a physical vector, a spinor, and an auxiliary scalar. The vector performs the quotient,

and the auxiliary field imposes the constraint. For simplicity, let’s consider the K¨ahler

i case. Then we have X is a holomorphic killing vector. That implies LXJ = 0, LXg = 0, and hence LX(Jg) = 0. This gives:

i i i X Jjk,i + X,jJik + X,kJji = 0 .

We also have that ∇J = 0, which in particular implies

Jjk,i + Jki,j + Jij,k = 0 .

Then we find i i i i −X Jki,j − X Jij,k + X,jJik + X,kJji = 0 , which finally implies i i (X Jik),j − (X Jij),k = 0 .

i Thus the curl of X Jij vanishes, and hence it is a gradient:

i X Jij = µ,j

µ is called the moment map.

If we now impose µ = 0 and take quotient on that submanifold, we preserve the ∂K K¨ahler structure. In our case, µ = ∂V |V =0 − c. The N = 2 quotient described earlier gives precisely the symplectic reduction of the target space as described here.

35 V. CFT Quotients.

The equations for invariance discussed so far only imply global invariance. The corre-

sponding Noether current is conserved: ∂J¯ + ∂J¯ = 0, but for conformal field theory, we

would like to have separate holomorphic currents: ∂J¯ = 0. This means that the symmetry

should survive for an arbitrary parameter α(¯z). A straightforward calculation shows that this implies

+ j ∇i X = 0.

k 0 k If TijkX = 0, φ is a distinct free field. If TijkX 6= 0, the action has the form (in special coordinates) Z 0 ¯ 0 i i ¯ 0 i (∂φ ∂φ + gi(φ )∂φ ∂φ ) + S[φ ]

If we also want an independent J¯, we are led to two coordinates Z h ¯ ¯ i ¯ R ¯ i L i ¯ i ∂φR∂φR + ∂φL∂φL + 2B(φ )∂φR∂φL + Gi ∂φR∂φ + Gi ∂φ ∂φL ¯ ¯ i + (∂φL∂φR − ∂φR∂φL) + S[φ ]

Then there are two natural quotients, vector:

δφR = α , δφL = α ,

or axial:

δφR = α , δφL = −α .

Since these are quotients of by Kaˇc-Moody symmetries, they make sense in CFT (for

example, they could be implemented by the GKO construction). If we simply perform

the quotients at the lagrangian level, we find that the two different quotients lead to

theories that are a dual pair! So dual backgrounds are different CFT quotients of a higher

dimensional theory.

36 VI. The Surreal Pun:

A man from Helsinki goes into a shop, and asks, in Swedish, “Do you have any candles?”

The salesperson answers, “Would you like apple juice or orange juice?” and the man replies

“I would like Christmas candles.”

ACKNOWLEDGEMENTS

It is a pleasure to thank the organizers of the school, Darla Sharp-Fitzpatrick for typing the first draft of these notes, and Fabian Essler, Amit Giveon, Kostas Skenderis, and Gene

Tyurin for their comments on the manuscript.

37 REFERENCES

Here is a very incomplete set of references.

The following are some standard texts on supersymmetry:

S.J. Gates, Jr., M.T. Grisaru, M. Roˇcek, and W. Siegel, Superspace, or One thousand

and one lessons in supersymmetry (Benjamin/Cummings, Reading, 1983)

J. Wess and J. Bagger, Supersymmetry and (Princeton University,

Princeton, 1983)

R.N. Mohapatra, Unification and supersymmetry: the frontiers of - physics (Springer-Verlag, New York, 1986)

P. West, Introduction to supersymmetry and supergravity (World Scientific, Sin- gapore, 1986).

A Physics Report on the subject that K.J. Schoutens recommends is:

M.F. Sohnius, Phys. Rep. 128 ( 1985) 39-204.

A nice set of lectures from an earlier TASI school is:

M.T. Grisaru, Introduction to Supersymmetry IN *ANN ARBOR 1984, PROCEED-

INGS, TASI LECTURES IN ELEMENTARY PHYSICS 1984*, 232-281.

The last two lectures make use of results in

S.J. Gates, C.M. Hull, and M. Roˇcek,Nucl. Phys. B248:157, 1984.

N.J. Hitchin, A. Karlhede, U. Lindstr¨om,and M. Roˇcek, Commun. Math. Phys. 108:535,

1987.

M. Roˇcek and E. Verlinde, Nucl. Phys. B373:630-646, 1992.

38 See also

C.M. Hull and B. Spence, Phys. Lett. B232:204,1989, Nucl. Phys. B353:379-426,1991,

Mod. Phys. Lett. A6:969-976, 1991.

39