Добавил:
Опубликованный материал нарушает ваши авторские права? Сообщите нам.
Вуз: Предмет: Файл:

Supersymmetry. Theory, Experiment, and Cosmology

.pdf
Скачиваний:
79
Добавлен:
01.05.2014
Размер:
12.6 Mб
Скачать

 

 

 

 

 

 

 

 

BPS states 69

Maxwell equations in the vacuum

 

 

 

 

 

 

µF

µν

˜µν

= 0,

˜

1

µνρσF

ρσ

(4.63)

 

= 0, ∂µF

Fµν

2

 

are invariant under the duality transformation

F

µν

˜µν

.

 

→ F

This may be rephrased in terms of the electric F 0i = −Ei and magnetic F ij fields9:

(4.64)

= ijkBk

E B, B → −E,

(4.65)

which may be generalized to the continuous transformation

E + iB = e(E + iB) .

(4.66)

However, introducing charged matter through the current jEµ

µF

µν

ν

˜µν

= 0

(4.67)

 

= jE ,

µF

introduces an imbalance between electric and magnetic. At the classical level, the obvious solution is to introduce magnetic sources through a magnetic current jMν in the right-hand side of the second equation. This is not as straightforward in the quantum theory because the wave equation for a charge particle involves the vector potential Aµ which is ill-defined once one introduces magnetic charges; indeed, if the vector potential A is uniquely defined, we can surround any magnetic charge Qm by a closed surface S and write

Qm =

B.dS = A.dS = 0.

(4.68)

S

S

 

However, the description of the electromagnetic field by the potential Aµ is redundant: it is the famous gauge invariance Aµ = Aµ − ∂µα. Since Aµ and Aµ describe the same physical reality, we may use either of them in di erent parts of space, under the condition that one makes sure that there is full compatibility in overlapping regions. This is precisely what one does in the presence of magnetic charges since A cannot be uniquely defined.

Let us consider a magnetic monopole Qm. As shown on Fig. 4.1, we draw a closed surface around the monopole and consider a plane going through the monopole. The vector potential is A(1) above the plane and A(2) below: in order that the electromag-

netic field be uniquely defined, we must

have A(1)

A(2) =

 

α. The plane divides

 

(1)

 

 

 

 

(2)

: S

(1)

∩ S

(2)

is

the surface between an upper hemisphere S

 

and a lower sphere S

 

 

 

a curve C parametrized by an angle φ (0 ≤ φ ≤ 2π). Then

 

 

 

 

 

 

 

 

Qm = S(1) S(2)

B.dS = S(1)

B.dS + S(2)

B.dS

 

 

 

 

 

 

= C A(1) A(2)

.dl = α(2π) − α(0).

 

 

 

 

(4.69)

9Note that the square of this transformation corresponds to charge conjugation.

70 The supersymmetry algebra and its representations

S(1)

φ

Qm

C

S(2)

Fig. 4.1 Closed surface around a magnetic monopole

Under the gauge transformation, the wave function corresponding to a field of electric charge Qe transforms as

ψ (x) = eiQeα/ ψ(x).

(4.70)

Requiring such wave functions to be single-valued (as φ → φ + 2π) implies that

QeQm = 2nπ , n Z,

(4.71)

which is the famous Dirac quantization condition [118, 381]: the presence of a single monopole imposes that all electric charges are multiples of a given unit of charge (2π /Qm). This gives a rationale for the problem of the quantization of charge mentioned in Chapter 1. We will encounter a seemingly di erent reason when we discuss grand unified theories in Chapter 9 but we will see that, at a deeper level, the two solutions coincide.

Until now, we have only considered electric charges and magnetic monopoles. One may envisage the presence of dyons [241], i.e. objects with both electric charge Qe and magnetic charge Qm. If we have two dyons, the Dirac quantization condition becomes [332, 390]:

Qe1Qm2 − Qe2Qm1 = 2nπ , n Z.

(4.72)

We deduce that, if the theory has electrons of charges (−e, 0), then the quantization reads, for any dyon field (Qe, Qm): Qm = 2nπ /e. Let us consider for example the dyon (Qe, 2π /e) and assume that we have CP invariance. Under CP, the electric charge is odd whereas the magnetic charge is even. We thus also have a dyon (−Qe, 2π /e). Applying the quantization condition (4.72) to this pair, we obtain Qe = ne/2: such dyons must have integer or half-integer charge.

Moreover, applying again the quantization condition to (Qe, 2π /e) and (Qe, 2π /e), we obtain Qe − Qe = ne. Hence all of them must have integer charges, or have half-integer charges.

However, if there is a source of CP violation, we may expect that there will be a departure from these quantized values. Witten [370] has in particular stressed the importance of adding the CP violating term

δL = −θ

e2

µν ˜

(4.73)

32π2

F

Fµν ,

BPS states 71

which gives a contribution of order θ to the dyon mass. We will return in more details to this issue in Section 4.5.7.

4.5.3Monopoles in the Georgi–Glashow model

In order to be able to consider the magnetic charge as a topological charge, we need to unify the electromagnetic U (1) symmetry as part of a larger nonabelian symmetry. The simplest case is the model of [180] (see also Exercise 9 of Appendix Appendix A) with a SU (2) gauge symmetry. We note that, in such cases, the monopole is a solution of the field equations and therefore must exist, in the context of these theories. From now on, we set = 1.

The Lagrangian of the Georgi–Glashow model reads

 

 

 

1 a

aµν

 

 

θ

2

 

a ˜aµν

1

 

µ

a

 

a

 

 

 

L

=

 

Fµν F

 

 

g

 

Fµν F

 

+

 

D

 

φ

Dµφ

 

− V (φ),

(4.74)

 

4

 

32π2

 

 

2

 

 

where (a = 1, 2, 3)

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

Fµνa

= µAνa − ∂ν Aµa + g abcAµb Aνc

 

 

(4.75)

 

 

 

 

Dµφa = µφa + g abcAµb φc,

 

 

 

 

 

 

(4.76)

and

 

 

 

 

 

 

 

 

 

λ

 

 

2

2

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

a

a

 

 

 

 

 

 

 

 

 

 

 

 

 

V (φ) =

 

 

φ φ

− a .

 

 

 

 

 

(4.77)

 

 

 

 

 

 

 

 

 

 

 

 

 

 

The corresponding Euler equations are4

 

 

 

 

 

 

 

 

 

 

 

 

 

DµF aµν = −g abcφbDν φc

 

 

 

 

where DµF

aµν

 

 

(DµDµφ)a = −λφa φbφb − a2

,

 

 

 

(4.78)

 

is defined in (A.57) of Appendix Appendix A (as is well known, the

θ term does not contribute to the equation of motion thanks to the Bianchi identity

˜aµν

= 0).

 

 

 

DµF

 

 

 

The symmetric energy–momentum tensor simply reads:

 

 

 

 

Θµν = −F aµρF ρ + DµφaDν φa − gµν L.

(4.79)

If we want to find the ground state, we must minimize the energy density

 

 

Θ00 =

1

(EiaEia + BiaBia + D0φaD0φa + DiφaDiφa) + V (φ),

(4.80)

 

2

 

 

 

 

with standard definitions for the electric and magnetic fields (see the preceding section):

a

a0i

a

1

ajk

. Hence the vacuum corresponds to a

vanishing gauge

Ei = −F

 

, Bi

= 2 ijkF

 

 

φ

a

φ

a

 

a

2

,

field and a constant scalar

field value that minimizes the potential:

 

 

 

2

 

3

 

 

1

 

 

2

 

 

 

 

 

 

 

which defines a two-sphere Svac. For example, φ

 

= a = 0 and φ

= φ

= 0 breaks

 

 

 

 

 

 

 

µ

µ

µ

 

 

 

 

 

 

3

 

 

 

 

SU (2) to U (1) and gives a mass ag to the fields W ±

A1

iA2

 

/2, where the

 

 

 

 

 

 

 

 

The field A

 

A

 

 

remains

index ± refers to the charge under the U (1) symmetry.

 

µ

 

 

µ

 

 

 

 

massless.

We are interested in finite energy configurations. Obviously, for the energy to remain finite, each term in (4.80) must fall o faster than 1/r2 at infinity. On the surface at infinity, the two-sphere S2 , we must recover the vacuum. Hence a finite

72 The supersymmetry algebra and its representations

energy configuration defines a mapping φa: S2 → Svac2 . Such a map is characterized by an integer, the winding number n, which counts the number of times one wraps one sphere around the other. This can be expressed as

 

 

 

 

1

S2 dSi ijk

abc ˆa

ˆb

ˆc

(4.81)

 

 

 

 

n =

 

φ

j φ

kφ

 

 

 

 

8π

ˆa

≡ φ

a

 

 

 

 

 

 

 

ˆa

of

where φ

 

/a. Note that this expression is invariant under a small change δφ

 

 

 

ˆa

 

 

 

10

 

 

 

 

the unit vector φ

: it is a topological invariant .

 

 

 

 

ˆa We may express this in a more mathematical way. We first note that the vector φ

which describes the vacuum parametrizes the coset group G/H = SU (2)/U (1) in the breaking of G = SU (2) to H = U (1) (see Section A.2.3 of Appendix Appendix A). We have thus defined a mapping from S2 to SU (2)/U (1). The group of inequivalent mappings (i.e. mappings which cannot be deformed continuously into one another) from the k-sphere Sk to a manifold M (such as G/H) is called its kth homotopy group and noted πk(M). We have just established that π2(SU (2)/U (1)) is isomorphic to Z.

We must have in particular

ˆa

= i

ˆa

+ g

abc

 

b ˆc

0

(4.82)

Diφ

 

φ

 

Ai φ

on the surface at infinity. This is solved by

 

 

 

 

 

 

a

 

1

 

abc ˆb

 

ˆc

ˆa

 

 

Ai

 

 

 

φ

iφ

 

+ φ

Ai

(4.83)

g

 

 

where Ai is a SU (2) invariant potential. We then deduce from (4.82) that, at large distances, the field strength is aligned along the scalar field11, that is along the direction of unbroken U (1) symmetry:

a ˆa

Fij

 

 

 

(4.84)

F ij = φ

 

 

 

with

 

 

 

 

 

 

 

1

 

abc ˆa

ˆb

ˆc

 

 

Fij = iAj − ∂j Ai

 

 

φ

iφ

j φ

.

(4.85)

g

The magnetic charge corresponding to this field configuration is expressed in terms of the winding number n given in (4.81) as:

Qm = S2

BidSi =

1

S2

ijkdSiFjk =

1

S2

 

abc ˆa

ˆb

ˆc

=

4πn

2

 

2g

dSi ijk

φ

j φ

kφ

g

.

(4.86) which corresponds to the Dirac quantization condition (4.71) since, in this theory, fields in the fundamental of SU (2) (not introduced here) would have an electric charge Qe = g/2. We conclude that the finite energy configurations of the Georgi–Glashow model are the magnetic monopoles.

10

ˆa

ˆa

 

 

ˆa

:

 

Since φ

is a unit vector, δφ

, and its spatial derivatives all lie in the plane perpendicular to φ

thus δn = 0.

 

 

= −i abc in the adjoint representation of

11Using (A.51) of Appendix Appendix A and (T a)bc

SU (2), we have

 

= −g φˆ Fij

a

 

 

 

0 = [Di, Dj ]ab φˆb = −g abcφˆbFijc

 

 

 

.

 

BPS states 73

More generally, if the symmetry group of a theory is G and the unbroken subgroup H, then monopoles exist if the second homotopy group π2(G/H) is nontrivial, i.e. not reduced to the identity. If G is simply connected (i.e. any loop can be continuously deformed into a point: π1(G) is trivial) such as SU (n) or any simple compact connected Lie group besides SO(n), then

π2(G/H) = π1(H).

(4.87)

For example, if H = U (1), π1(U (1)) = Z, then monopoles exist and are characterized by a integer n. This is what happens with the ’t Hooft–Polyakov model just discussed or the breaking of the grand unified group SU (5) (see Chapter 9). If H = SO(n) (n ≥ 3), π1(SO(n)) = Z2, there is a single type of monopole corresponding to the nontrivial element 1 of Z2. If G is not simply connected, one may find useful the following theorem: π2(G/H) is isomorphic to the kernel of the homomorphism of π1(H) into π1(G) (a loop in H is naturally a loop in G).

4.5.4BPS monopoles

We note that, according to (4.84), we may write the electric and magnetic charges with respect to U (1) as

Qe =

1

 

dSi Eiaφa

 

a

 

 

Qm =

1

 

dSi Biaφa.

(4.88)

 

a

Following [47], we may derive a lower bound for the energy of the stationary configurations that we have just identified. For simplicity, we work in the temporal gauge12 Aa0 = 0. We obtain for the energy density (4.80)

E = T + V,

 

T =

1

d3x [0Aia0Aia + 0φa0φa] ,

(4.90)

 

2

V =

1

d3x [(Bia Diφa)(Bia Diφa) + 2V (φ)] ±

d3x BiaDiφa,

 

2

where we have separated the contribution which is quadratic in time derivatives

(“kinetic energy” T ) from the one which does not involve time derivatives (“potential”

V). Since, according to (4.88), aQm = d3x BiaDiφa, one obtains (choosing the upper, respectively lower, sign for positive, respectively negative, magnetic charge):

E ≥ a |Qm| .

(4.91)

12Accordingly, we must impose the Gauss law constraint, which corresponds to ensuring the A0

equation of motion (4.78):

 

Di (0Aia) + g abcφb0φc = 0.

(4.89)

74 The supersymmetry algebra and its representations

 

The bound is saturated for states satisfying the Bogomol’nyi conditions

 

Eia = 0 , D0φa = 0

(4.92)

Bia = ± Diφa

(4.93)

in the limit of vanishing potential V (φ). This limit is known as the Prasad–Sommerfield limit [315] and the states which saturate the bound are thus known as BPS states. The upper (lower) sign corresponds to the monopole (antimonopole) solution.

As we will emphasize later, the Prasad–Sommerfield condition of vanishing potential is unstable under quantum fluctuations unless we are in a supersymmetric context.

The Bogomol’nyi equations can be solved using the following static and spherically symmetric ansatz:

φa(x) = aH(ξ)

xa

, A0a(x) = 0,

 

 

r

 

 

 

 

 

 

 

xj

 

 

Aia(x) = aij

 

(1 − K(ξ))

(4.94)

gr2

where r = |x| and H and K are functions of the rescaled radial variable ξ = agr. Plugging this ansatz into the Bogomol’nyi equation (4.93) for the (anti)monopole yields

H (ξ) =

1

(K2 1)

 

ξ2

 

K (ξ) =

 

KH,

(4.95)

 

the boundary conditions being H → 0, K → 1 as ξ → 0 and H → 1, K → 0 as ξ → ∞. The solution corresponding to the monopole (upper sign in the preceding equation) reads:

H(ξ) = coth ξ −

1

 

K(ξ) =

ξ

(4.96)

 

,

 

.

ξ

sinh ξ

4.5.5Moduli space

Before quantizing the monopole, we need to identify its collective coordinates. The monopole solution which has been constructed in the previous subsection sits at the origin but the translation invariance of the equations implies that a solution with the same energy can be constructed at any point in space. Thus the spatial coordinates of the monopole center of mass provide three collective coordinates.

Since the inclusion of these degrees of freedom does not change the (potential) energy of the monopole, they are zero modes, or moduli, of the system and their quantization requires a special treatment. Just as for the kink, the standard way [316] is to start with the classical solution (4.94) which we denote φacl(x), Aai cl(x) and to

translate this solution to a time-dependent position X(t) to form the ansatz:

 

φa(x, t) = φcla (x X(t)) , Aia(x, t) = Aia cl (x X(t)) .

(4.97)

BPS states 75

This represents, as long as the position X(t) is a slowly moving function, the low energy ansatz obtained by integrating out the massive modes (i.e. ignoring in the semiclassical approach contributions of order ωi where ωi are the nonvanishing frequencies associated with the eigenvalues of the double derivatives of the potential energy, evaluated around the monopole solution). Substituting this ansatz into the action yields

S =

dtL =

 

πa

 

dt (T − V) =

4

 

g

dt

1

 

X(t)

 

d

2

dt

2

1 .

(4.98)

This is the moduli space approximation: the motion in moduli space is made at constant potential energy since moduli are associated with the zero modes of the system. One remains with a kinetic energy associated with the slow motion in moduli space (here in position space).

In order to search for all the moduli fields, we adopt a more general strategy. We start with the BPS monopole solution satisfying Bia = Diφa and deform it (Aai + δAai , φa+δφa) while keeping the potential energy fixed. Then, to first order, the Bogomol’nyi equation yields:

ijkDj δAak = Diδφa + g abcδAibφc,

(4.99)

where Di is the covariant derivative in the adjoint representation (DiXa = iXa +

g abcAbi Xc).

This has to be complemented by the Gauss law constraint (4.89): writing for a

general modulus Z, as in (4.97),

 

φa(x, t) = φcla (x, Z(t)) , Aia(x, t) = Aia cl (x, Z(t)) ,

(4.100)

the constraint reads

 

Di (δAia) + g abcφbδφc = 0.

(4.101)

A general deformation is a combination of spacetime translation and gauge transformation (δAai = (1/g)Diαa, δφa = Cabcαbφc). We have already considered moduli associated with spacetime translations. The only remaining modulus corresponds to the gauge transformation exp [(t)T aφa(x)/a]. We note that, since φa is nontrivial at spatial infinity, this gauge transformation does not approach the identity: in the standard language of gauge theories, this is a large gauge transformation which relates two distinct physical points (by comparison, a small gauge transformation, i.e. one that approaches the identity at spatial infinity, corresponds to a mere redundancy in the description of physical states). Writing thus αa(x) = −χ(t)φa(x)/a, the new modulus corresponds to the deformation13

δAia =

1

 

Di [χ(t)φa(x)] , δφa = 0.

(4.102)

ga

 

 

 

It is straightforward to check that this satisfies the linearized Bogomol’nyi constraint (4.99) (using the original equation Bia = Diφa) and the Gauss law constraint (4.101)

1

 

13In order to keep A0a = 0 we must complement the ansatz (4.100) with A0a =

 

χ˙ (t)φa(x).

ga

76 The supersymmetry algebra and its representations

(using the scalar field equation of motion (4.78) with vanishing scalar potential). We see that the new coordinate is conjugate to the electric charge (just as the space coordinates are conjugate to momentum): motion in moduli space generates momentum and charge for the monopoles, which thus become dyons. We note that there are no moduli corresponding to angular momentum: classical motion in moduli space does not generate intrinsic angular momentum for the monopoles.

Since the unbroken U (1) gauge group is compact, χ has a compact range (typically

0

χ < 2π). Thus the corresponding space is the circle S1 and the full moduli space

 

 

 

14

is R

3

× S

1

. In the moduli space approximation, we now have

for a single monopole

 

 

 

 

 

4πa

 

 

1

 

 

 

dX(t) 2

1

 

(t)

2

 

 

 

S =

 

 

 

dt

 

 

 

 

 

+

 

 

 

1 .

(4.103)

 

 

 

g

 

2

dt

2a2g2

dt

We note that the corresponding plane waves are simply eiP·Xeineχ, with ne integer. Charge is simply Qe = −ig∂χ: such plane waves correspond to dyons with charge Qe = neg. We may also infer their masses in this approximation: since the corresponding Hamiltonian is

 

1

 

g

1 ag3

4πa

 

 

H =

 

 

 

P2 +

 

 

 

Pχ2 +

 

,

(4.104)

2

4πa

2

4π

g

where P and Pχ are the canonical momenta associated with X and χ, respectively, the mass of the dyons is simply

M =

4πa

+

1 ne2ag3

=

4πa

1 +

1 ne2g4

.

(4.105)

g

 

2

 

4π

g

 

2

 

16π2

It can be showed [81] that the classical solutions of the equations of motion satisfy the

 

 

 

 

 

 

 

 

general Bogomol’nyi-type bound M ≥ a

 

Qe2 + Qm2 . We see that the dyons saturate

this bound: indeed, (4.105) is the

expansion for g

 

1 of the formula

 

 

"

 

 

 

M = a"

 

= a |Qe + iQm| ,

(4.106)

Qe2 + Qm2

since Qm = 4π/g and Qe = neg.

4.5.6Supersymmetry

The picture that emerged in the last section is likely to su er some undesirable modifications once one includes quantum corrections. First, as already noted, radiative corrections should generate a nonvanishing potential, thus contradicting the Prasad– Sommerfield condition of vanishing potential. Similarly, the mass bounds should be corrected by quantum fluctuations. Obviously, this makes such bounds less interesting.

We now know enough about supersymmetry to understand that it might provide the extra ingredient required to control quantum corrections. Indeed, both vanishing potentials and masses are protected by supersymmetry. Since we already have vector (Wµ± and Aµ) and scalar (monopole) degrees of freedom, the natural set up, once one includes fermionic degrees of freedom, is N = 2 supersymmetry.

14More generally, the moduli space of n monopoles is 4n-dimensional [357].

BPS states 77

We thus consider a N = 2 vector supermultiplet, whose action is written in (4.36) (we will set P r = 0 and thus consider the on-shell action). We note that the condition of vanishing potential requires [A, B] = 0, which may be satisfied for nonzero vacuum expectation values of A and B. Supersymmetry ensures that this holds at the quantum level.

Choosing the gauge symmetry group SU (2), as in the Georgi–Glashow model, we may set the vacuum expectation value for B to zero. A nonzero value for A (sayA3 = a = 0) breaks SU (2) to U (1). The supersymmetry current for the model considered reads (see (3.37) in Chapter 3 and (C.73) in Appendix C):

1

1

 

 

 

Jµi = σρσ Fρσγµγ5λi +

 

Dρργµλi +

 

Dρργµγ5

λi + µγ5

[A, B] λi. (4.107)

2

2

One may then deduce the supersymmetry charge anticommutation relations, with special attention to the boundary terms. Introducing

U = d3x ∂i AaF0ai + Ba

1

 

 

ijkFjka

 

2

 

1

 

 

 

 

V = d3x ∂i Aa

 

ijkFjka

+ BaF0ai ,

(4.108)

2

one obtains [369]

 

 

 

 

Qri, Q¯sj = δij γrsµ Pµ + ij δrsU + γrs5 V .

(4.109)

Since we set B = 0, we have, using (4.88),

 

 

 

 

U = aQe,

V = aQm.

(4.110)

We see that, in (4.108), U and V play the rˆole of central charges. We thus expect mass bounds similar to the ones derived in subsection 4.3.3. Indeed, if we work in the frame where Pµ = (M, 0, 0, 0), then (4.109) reads

$Qri, Qsj%

 

 

 

 

 

= M δij δrs + a ij

Qe + 5Qm

γ0

rs .

(4.111)

Since the left-hand side is positive definite, the eigenvalues of the right-hand side are

 

 

Qe

m

 

 

positive. But

Qe + 5Qm

γ0

2 =

 

Qe + 5Qm

Qe − iγ5Qm = Qe2 + Qm2 . The

eigenvalues are then M

± "

2 + Q2

and one obtains the constraint

 

 

 

M ≥ a"

 

.

(4.112)

 

 

 

Qe2 + Qm2

Obviously states which saturate the bound are annihilated by at least one of the supersymmetry generators.

We note that all states in the theory saturate the bound: for example, the photon Aµ is massless whereas Wµ± has mass ga; and dyon masses are given by (4.106). This explains what could have been seen as a puzzle otherwise. Indeed, we started with the eight states of the (non-CPT self-conjugate) N = 2 vector supermultiplet: Aµ, λi, A, B. Symmetry breaking gives mass to some of these states but does not provide the

78 The supersymmetry algebra and its representations

degrees of freedom necessary to form a (long) massive N = 2 vector supermultiplet, which has 16 degrees of freedom. But these eight states form a short massive N = 2 supermultiplet, in presence of the central charges.

This also explains why the Bogomol’nyi bound remains una ected by radiative corrections or nonperturbative e ects in this context. If it did, this would mean that the mass spectrum departs from the one fixed by the central charges. The massive states would then fall into long supermultiplets but we are lacking the necessary extra degrees of freedom.

4.5.7Montonen–Olive duality

Since magnetic monopoles carry magnetic charge, they generate a magnetic field which should be described by a gauge theory. The coupling associated with this theory should satisfy, according to (4.86),

gm =

4π

.

(4.113)

 

 

g

 

A key property of this relation is that it exchanges strong and weak coupling regimes: if the “electric” gauge symmetry is weakly coupled, then the “magnetic” gauge symmetry is strongly coupled, and vice versa. This implies that the fundamental degrees of freedom of both theories are not simultaneously accessible: in the electric description, the degrees of freedom of magnetic charge qm ≡ Qm/gm = ±1 (i.e. the magnetic monopoles) arise as solitons; conversely, the gauge fields of charge qe ≡ Qe/g = ±1 should arise as solitons of the magnetic theory. It might seem surprising to find here topological solitons with spin. It is well-known [223, 236] that, in the context of the quantum theory, monopoles convert isospin into spin.

Making full use of the mass spectrum derived above, i.e. M = a |Qe + iQm|, we may represent the di erent fields on the lattice (qe, qm). The fundamental fields of the electric gauge theory are located at points (0, 0) and (±1, 0) whereas the fundamental fields of the magnetic gauge theory are found at (0, 0) and (0, ±1). The sites o the axis correspond to dyon states. The electric–magnetic duality encapsuled in (4.113) appears in this diagram as the symmetry associated with a rotation of π/2. Such a duality was conjectured by [288].

Besides the mass spectrum, a nontrivial check is provided by monopole interactions. The force between two (positively charged) monopoles is velocity-dependent [281]: it vanishes at rest. Duality would imply that the static W +−W + force vanishes as well. It turns out that the standard Coulomb repulsion is cancelled by the force resulting from Higgs exchange (note that, in the Prasad–Sommerfield limit, the Higgs is massless).

Obviously, it is crucial to the Montonen–Olive duality conjecture that the mass formula does not receive contributions from radiative corrections or nonperturbative e ects. This is why it has to be considered in the supersymmetric context. We have already discussed what N = 2 supersymmetry brings forth. It is only when one goes to N = 4 supersymmetry that one finds monopole of spin 1 and thus a full realization of Montonen–Olive duality [301]15. This goes beyond the scope of this chapter and we refer the reader to the existing reviews on the subject (see for example [222]).

15Moreover the beta function of the N = 4 symmetric Yang–Mills theory vanishes, which settles the question of whether the couplings that we have been using are running or not.