Sei sulla pagina 1di 346

Nuclear Physics B 757 (2006) 1–18

Finite Heisenberg groups and Seiberg dualities


in quiver gauge theories
Benjamin A. Burrington, James T. Liu, Manavendra Mahato,
Leopoldo A. Pando Zayas ∗
Michigan Center for Theoretical Physics, Randall Laboratory of Physics, The University of Michigan,
Ann Arbor, MI 48109-1040, USA
Received 18 April 2006; accepted 9 June 2006
Available online 12 July 2006

Abstract
A large class of quiver gauge theories admits the action of finite Heisenberg groups of the form Heis(Zq ×
Zq ). This Heisenberg group is generated by a manifest Zq shift symmetry acting on the quiver along with
a second Zq rephasing (clock) generator acting on the links of the quiver. Under Seiberg duality, however,
the action of the shift generator is no longer manifest, as the dualized node has a different structure from
before. Nevertheless, we demonstrate that the Zq shift generator acts naturally on the space of all Seiberg
dual phases of a given quiver. We then prove that the space of Seiberg dual theories inherits the action of
the original finite Heisenberg group, where now the shift generator Zq is a map among fields belonging to
different Seiberg phases. As examples, we explicitly consider the action of the Heisenberg group on Seiberg
phases for C3 /Z3 , Y 4,2 and Y 6,3 quivers.
© 2006 Elsevier B.V. All rights reserved.

1. Introduction

Generalizing an observation of [1], it was shown in [2] that for a class of Zq orbifold quiver
gauge theories with gauge group SU(N )p , there is a set of discrete transformations A, B and C
satisfying

Aq = B q = C q = 1, AB = BAC. (1.1)

* Corresponding author.
E-mail address: lpandoz@umich.edu (L.A. Pando Zayas).

0550-3213/$ – see front matter © 2006 Elsevier B.V. All rights reserved.
doi:10.1016/j.nuclphysb.2006.06.030
2 B.A. Burrington et al. / Nuclear Physics B 757 (2006) 1–18

Here the A generator is inherited from the Zq orbifold action, and corresponds to the manifest
Zq shift symmetry acting on the quiver. The B generator may be thought of as a clock generator,
which acts by rephasing the links of the quiver. This combination of clock and shift then generates
the finite Heisenberg group as indicated above, where C is essentially a uniform rephasing of all
the links. These transformations satisfy three important properties: (i) they leave the superpoten-
tial invariant; (ii) they satisfy anomaly cancellation constraints for all SU(N ) gauge groups; and
(iii) the above group relations are true up to elements in the center of the gauge group SU(N )p ,
that is, up to gauge transformations.
The generators A, B and C can be interpreted, in the dual string theory, as operators counting
the number of wrapped F-strings, D-strings and D3 branes respectively. Thus, the above Heisen-
berg group implies that the charges of D branes in the presence of Ramand–Ramond flux do
not commute. In fact, an important motivation for our study is provided by recent investigations
along these lines due to Belov, Moore and others [4,5].
An extension of this structure was considered in [3] were nonconformal quiver gauge theo-
ries were considered. In this case, the Heisenberg group gets centrally extended as a result of
having gauge groups with different ranks. In other words, the addition of fractional branes to the
background induces a central extension of the Heisenberg group.
We believe the study of discrete symmetries of quiver gauge theories is interesting in its own
right. As field theories, a natural question that arises in this context is the interplay between
the existence of a finite Heisenberg group and Seiberg duality. Seiberg duality is an equivalence
between two gauge theories [6], and has been extensively studied in the context of quiver gauge
theories [7–10].
In this paper we study the interplay between the existence of a finite Heisenberg group acting
on orbifold quiver gauge theories and Seiberg duality. The generator A, realized as a shift sym-
metry, acts manifestly only on the symmetric phase of the theory. After Seiberg duality, most
quivers lose this manifest shift symmetry associated with A. However, we demonstrate that this
symmetry is naturally restored in the space of all Seiberg dual quivers.
Our main result is as follows: for a quiver gauge theory admitting the action of a finite Heisen-
berg group Heis(Zq × Zq ) as above (1.1), there exists a similar finite Heisenberg group acting
on the space of Seiberg dual theories to the quiver. Moreover, the action of the shift generator in
the Heisenberg group maps fields among different Seiberg phases.
The paper is organized as follows. In Section 2 we review some known aspects of quiver
gauge theories including the algorithm for performing Seiberg duality in quiver gauge theories
and some of its properties. Also in Section 2 we provide a constructive proof of the existence of a
Heisenberg group acting on the spaces of Seiberg dual quivers. Section 3 contains several explicit
examples of our construction; we consider C3 /Z3 , Y 4,2 and Y 6,3 quivers. We then conclude in
Section 4.

2. Seiberg duality and discrete symmetries

In this section we begin with a discussion of generalities of Seiberg duality acting in quiver
gauge theories. We then examine the implication of Seiberg duality on the B and C genera-
tors, which act by rephasings on the links. As discussed in [1–3], these Abelian generators may
be constructed by assigning discrete U (1) charges to each node in the quiver consistent with
anomaly cancellation. The anomaly cancellation requirement has a natural description in terms
of the adjacency matrix, and we demonstrate how this carries over into the Seiberg dual phases
as well. As a result, the B and C generators have a natural extension when acting on the space of
B.A. Burrington et al. / Nuclear Physics B 757 (2006) 1–18 3

Seiberg duals. By including the A generator (now acting on the space of Seiberg duals), we are
then able to prove that the space of Seiberg dual quivers naturally inherits the action of the finite
Heisenberg group of the original (symmetric phase) quiver.

2.1. Seiberg duality on quivers

Seiberg duality has been discussed extensively in the context of quiver gauge theories. In this
subsection we review its implementation and discuss some of its implications for quiver gauge
theories. We consider only quiver gauge theories with gauge group SU(Ni ). In this case, a quiver
gauge theory is completely defined by giving the rank of the gauge groups, the matter content and
the superpotential. The matter content is conveniently encoded in the (antisymmetric) adjacency
matrix denoted by aMN .
A given Seiberg duality acts on a single node of the quiver. Thus, to state the rules of Seiberg
duality, it is convenient to split the indices (labels for the nodes) according to: 0 to indicate
the specific node under investigation; i, j, . . . to indicate neighboring nodes that have arrows
pointing from the 0th to the ith node; ĩ, j˜, . . . to denote neighboring nodes that have arrows
pointing from the j th to the 0th node; and a, b, . . . to denote the remaining nodes, which we
emphasize are not directly connected to the 0th node. Seiberg duality of the 0th node is captured
by the transformation
â0M = −a0M , âij = aij ,
âĩ j˜ = aĩ j˜ , âaM = aaM ,
âi j˜ = ai j˜ − a0i aj˜0 , âj˜i = aj˜i + a0i aj˜0 , (2.1)
where the caret denotes the new quantities after Seiberg duality. All other components follow
from the antisymmetry of a and â. If we further make the assumption that any two nodes are only
connected by edges with the same directionality, the adjacency matrix completely determines the
field content and charges of the new theory. In addition, the rank of the 0th gauge group is now
changed to be
 
N̂0 = (a0i Ni ) − N0 = (aj˜0 Nj˜ ) − N0 . (2.2)
i j˜

Here we use the notation NM to denote the rank of the gauge group at the Mth node. Note that
the second equality follows as a result of the original anomaly cancellation condition, and that
the sum term above is easily understood as the effective number of flavors.

2.2. Discrete transformations and Seiberg duality

The B and C operators both act by rephasings of the links of the quiver. A convenient manner
to construct these operators is to assign discrete U (1) charges to the nodes, say charge nK for the
Kth node. These charges are not arbitrary, but must satisfy the constraint of anomaly cancellation.
(The superpotential constraint is automatically satisfied, since it corresponds to closed loops
along the links.) Here we show that this anomaly cancellation constraint has a natural formulation
in terms of the adjacency matrix a. However, once we do this, it then becomes clear how to
extend B and C to act on the Seiberg dual quivers. The result essentially follows by replacing
the adjacency matrix a with the corresponding dual one â. This results in a map of the charges
nK to n̂K in a similar fashion as (2.2).
4 B.A. Burrington et al. / Nuclear Physics B 757 (2006) 1–18

Before turning to the B and C rephasing operators, however, we first display the general
feature of cubic anomaly cancellation for SU(N )3 and explicitly demonstrate that it is still met
in the Seiberg dual phase. This calculation will then be extended to the discrete rephasing case
in a straightforward manner. To begin, we note that the anomaly cancellation condition in the
original Seiberg phase is given by
  
aM0 N0 + (aMi Ni ) + (aM j˜ Nj˜ ) + (aMa Na ) = 0. (2.3)
i j˜ a

Given this condition, we wish to show that


  
âM0 N̂0 + (âMi N̂i ) + (âM j˜ N̂j˜ ) + (âMa N̂a ) = 0, (2.4)
i j˜ a

which is the equivalent statement in the Seiberg dual phase. The last summation factor in either
of the above
 is unaffected by Seiberg duality of the 0th node. Hence we will denote this term
simply as a (âMa Na ) = a (aMa Na ) = Σ, because âMa = aMa . In fact, as a result of this,
the anomaly condition is trivially met by all nodes except for the 0th and those connected to it.
Therefore, we will only consider the M = k and M = k̃ terms. For M = k, the left-hand side
of (2.4) becomes
    
−ak0 (a0i Ni ) − N0 + (aki Ni ) + (ak j˜ − a0k aj˜0 )Nj˜ + Σ
i i j˜
   
= ak0 N0 + (aki Ni ) + (ak j˜ Nj˜ ) + Σ − ak0 (a0i Ni ) − (a0k aj˜0 Nj˜ )
i j˜ i j˜
   
= − ak0 (a0i Ni ) + (ak0 a0j˜ Nj˜ ) = 0, (2.5)
i j˜

where the terms are 0 as a result of the kth and 0th component of the original anomaly condition.
This condition is likewise met by the M = 0 node because the only terms appearing simply
flip sign. The j˜ nodes follow in exactly the same manner as the above calculation because of
relation (2.2). This shows that anomaly cancellation holds in the Seiberg dual phase, so long as
it holds in the original phase.
We now turn to the discrete U (1) rephasings used to construct the B and C operators. Because
these are Abelian rephasings, we need to consider mixed anomalies, where the anomaly comes
from the j μ SU(N )2 triangle diagram where j μ is the conserved current of the U (1) that the
rephasing is associated with. To accomplish the rephasing, we associate the phase ωK and charge
nK with the Kth node. A field represented by an arrow going from the Kth node to the Lth node
nK −nL
gets rephased by ωK ωL , and we mean this component by component in the superfield.
We now consider an instanton number 1 at the Mth node, and so all fields represented by an
arrow with either end on that node have a fermion zero mode (again, counting the other end of
the arrow as an effective flavor symmetry). We follow the previous work [2,3] and find that the
general expression for the anomaly with this instanton number is
 n0 −nM N0 a0M  nj −nM Nj aj M  nj˜ −nM N ˜ a ˜  n −nM Na aaM
ω0 ω M ωj ωM ω ˜ ωM j jM ωa a ωM = 1,
j
j j˜ a
(2.6)
B.A. Burrington et al. / Nuclear Physics B 757 (2006) 1–18 5

where the equality is to be read as a requirement of the set of numbers nK and phases ωK .
One can easily see that the factors of ωM cancel in this expression because the NM are a zero
eigenvector of the adjacency matrix a, and follows because the effective number of arrows in and
out are the same. The expression simplifies to
 n0 N0 a0M   nj Nj aj M   nj˜ Nj˜ aj˜M   n N a 
ω0 ωj ω˜ ωa a a aM = 1. (2.7)
j
j j˜ a

The above expression is general, and valid for any assignment of numbers and phases. As a
further simplification, we wish to express all phases in terms of a single phase, ω. We note also
that the phase ωK and the number nK overparameterize exactly how each field is charged. We
take that the phases are all eiφK where φK ∈ (0, . . . , 2π], so that 1 is parameterized only by e2πi .
nK
One can vary ωK and nK while leaving ωK fixed. Therefore, we find it useful to tune all of the
ωK such that

ωK = ω1/NK , (2.8)
for some fixed ω that is independent of K, leaving all parametrization of the phases represented
by the charges nK . Again, if ω = 1 we parameterize this as ω = e2πi such that the roots above
defining ωK are nontrivial (we will return to this case in a moment). This greatly simplifies our
expression, and we find
 n a   n a   n ˜ a ˜   n a 
ω 0 0M ω j jM ω j jM ω a aM
j j˜ a
  
n0 a0M + j (nj aj M )+ j˜ (nj˜ aj˜M )+ a (na aaM )
=ω = 1. (2.9)
This, then, is the general expression for the anomaly when the Mth node has an instanton number
1. We will abbreviate the exponential above as ← n− · a for obvious reasons.
We therefore require that this condition is met by the numbers nK for all values of M above.
One may worry that this does not capture all possible instanton numbers. However, because the
Pontryagin number for composite connections is additive, the basis of taking instanton number
1 for each node spans the space of all possible instanton numbers, and satisfying the anomaly
is most restrictive for these individual basis vectors. One may see that the Pontryagin number
is additive because the two triangle diagrams j μ SU 1 (N1 )2 and j μ SU 2 (N2 )2 exist (for a field
charged under both of these gauge groups), but no cross term exists. Therefore, a fermionic field
charged under both SU 1 (N1 ) and SU 2 (N2 ) will have J1 × N2 + J2 × N1 fermion zero modes,
where Ji are the instanton numbers of the SU i (Ni ) gauge groups. The extra factors of Ni show
up as a result of the trace over the gauge indices not “coupled to” in the triangle diagram.
A few words are now in order to discuss possible solutions to the above equations. If one
picks the ←n− to be a zero eigenvector of the adjacency matrix, there is no more condition on ω
and hence ω is arbitrary. This is a full U (1) symmetry that is nonanomalous, and explains why
the D and E operations of [2,3] were found to be simply Zp subgroups of these continuous U (1)
factors. This also suggests
 how to take the rephasings of this kind through Seiberg duality: one
N / a N −N 
reassigns ω̂0 = ω0 0 i 0i i 0 , and then reassigns nˆ0 = i a0i ni − n0 (all other nM remain
the same) and one again gets a U (1) in the new Seiberg phase. One should note that the nodal
charge of the fields remains unchanged in this case because the reassignment of the value of n0
n n̂
exactly cancels the change in the value of ω0 so that ω0 0 = ω̂0 0 .
6 B.A. Burrington et al. / Nuclear Physics B 757 (2006) 1–18

Let us now consider when the nM are integers, a case that matches the B and C operations
of [2,3]. One can see that if one takes the nM such that
 −

GCD (← n · a)M = λ, (2.10)
then we may simply require that

ωλ = 1, (2.11)
and we again have a symmetry. This time, however, the symmetry is not continuous, but is a λth
root of the center of the gauge group. If we label this rephasing as Q, we have that Qλ = 1 up to
the center of the gauge group. Certainly a class of such vectors is possible if a has any (nonzero)
integer eigenvalues. Also note that the case λ = 1 corresponds to the ω = 1 case previously
mentioned. As stated, we parameterize this by ω = e2πi such that the roots (2.8) are nontrivial.
This is easily seen to be a rephasing using the center of the gauge group, because the additional
root of (2.8) makes this an Ni root for a node with rank Ni . These rephasings are therefore gauge
equivalent to 1. This also means that the integers nK in this case are only understood modulo λ,
as one may shift the center of each gauge group independently.
Let us show how such symmetries map through Seiberg duality. First, in the original Seiberg
phase, we assume that there is a solution to

a · n ≡ 0 (mod λ). (2.12)


We therefore wish to find the new vector nˆ after Seiberg duality that satisfies

â · nˆ ≡ 0 (mod λ). (2.13)


We again suppose that we only wish to change n̂0 = n0 , leaving all other integers alone
n̂M = nM , M = 0. We now note that again the M = a components of the above equation are
automatically satisfied: only the anomalies away from the node being dualized are affected. We
now make an educated guess as how one transforms n0 . We guess that
  
n̂0 = (a0i ni ) − n0 = (a0i ni ) + (a0a na ) − n0 (2.14)
i i a

≡ (aj˜0 nj˜ ) − n0 (mod λ), (2.15)

where we have used a0a = 0 and the original anomaly cancellation conditions. The calculation
now goes through exactly as it did for the zero eigenvectors (2.5): all = signs are simply replaced
with ≡(mod λ) . Actually, one may expect this structure from the discussions of [11].
We now make one final comment. Seiberg duality of a given node is it’s own inverse, and in
fact the rephasings discussed above transform into themselves up to the center of the gauge group.
This is noted simply by the fact that after two Seiberg duals of one node, the phase associated
with the node has gone from

a N −N0
N0 N0 j˜ j˜0 j˜
 
i a0i Ni −N0 i a0i Ni −N0 N0
ω0 → ω0 → ω0 = ω0 , (2.16)
where in the second Seiberg duality we used that what we were calling “in” arrows are now called
“out” arrows, hence the switch in the sum from indices with no tilde to ones with tilde. This is
of course the same because the effective number of in and out arrows are the same, but this will
B.A. Burrington et al. / Nuclear Physics B 757 (2006) 1–18 7

be more important for the mapping of the numbers nM . For this, we note that the number n0 gets
mapped as
   
n0 → a0i ni − n0 → aj˜0 Nj˜ − a0i ni − n0
i j˜ i
 
= n0 + aj˜0 Nj˜ − a0i ni ≡ n0 (mod λ). (2.17)
j˜ i

In the last line, we have again used the fact that we are modding out by the center of the gauge
group, which corresponds to taking the number nM as only being defined mod λ.

2.3. General proof of the existence of the Heisenberg group

Given the natural mapping of discrete U (1) charges under Seiberg duality found above, we
now combine this with the results of [2,3] to show that there is generically an action of the
Heisenberg group on the space of Seiberg dual quivers. Here, we are assuming that there is a
symmetric Seiberg phase with a natural shift symmetry, which we will call A (see Section 3 for
some examples). A large class of these was examined in [2,3] and the action of a finite Heisenberg
group was found. Assuming that there is a symmetric phase, and an action of a Heisenberg group
on this phase, we will show the existence of an action of a Heisenberg group on the space of all
Seiberg phases.
First, note that if there is an A symmetry of the symmetric phase, this descends to the entire
tree of possible Seiberg dual quivers, however mapping from one phase to another. This space
was discussed in [9] where it was given the name of duality tree; each point represents a quiver.
In particular the three-node tree was shown to be related to Markov’s equation. We will be careful
to call the points on the duality tree points, which represent entire quivers, to distinguish them
from the nodes of the quivers themselves which represent gauge groups. The action on the duality
tree can be seen easily. Take that we have a symmetric phase with x × y nodes, and there is a
manifest Zx symmetry that permutes the Mth node to the (M + y)th node. The entire tree is
defined by taking an arbitrary number of Sieberg dualities on any number of the nodes in any
given order. We therefore label the point on the Markov tree that is  Seiberg dualities away from
the symmetric phase by the Seiberg dualities it takes to get there: (s1 , s2 , s3 , . . . , s ). Here, the
integers si label which node of the quiver diagram is to be Seiberg dualized, i.e. si ∈ 1, . . . , x × y.
There is an analogous quiver given by (s1 + y, s2 + x, s3 + y, . . . , s + y) ≡ s + y with the
same matter content, same gauge groups and couplings, simply with the labels changed. This,
therefore, defines the action of the A operator on this quiver: it maps the two differently labeled
field theoretic degrees of freedom into each other. So, if one wants to go from one branch to the
other, one must apply a set of inverse Seiberg dualities to get back to the symmetric phase, apply
the A operation (as many times as needed) to reorder the fields, and then apply the same set of
Seiberg dualities. However, because the ordering of the fields have been switched, the Seiberg
dualities are in fact being performed along a different branch of the duality tree. We emphasize
here that this is much the same as the operation B where the rephasings are really on the order
one gives the fields in, not their subscripts: A may have already rearranged them, or may not
have. B’s matrix representation does not depend on this! So, when we say that “B is rephasing
the field U1 as u1 × U1 ”, we really mean that it rephases this way only when the fields are
given in the canonical order (U1 , U2 , . . .); it really rephases the first field in the list. A of course
changes exactly which field this is. We think of the Seiberg dualities in the same way. We give the
8 B.A. Burrington et al. / Nuclear Physics B 757 (2006) 1–18

gauge groups and fields a canonical ordering which A shuffles. The Ss simply Seiberg dualize
according to s in the order that the gauge groups are listed, not what their labels are, and so may
or may not move out along a different branch, depending on whether an A is present or not.
We will make these comments more precise here. We have labeled the series of Seiberg dual-
ities with a vector s, and we will call the series of Seiberg dualities Ss . The inverse is of course
given by the same vector, simply with its entries reversed Ss−1 = Sreverse−1
s ) . A is simply
order (
defined using the symmetric phase

Âs ,s = δs ,s Ss ASs−1 (2.18)

with all other entries for s and s equal zero. One must be careful here to note that something
new has actually happened. The S −1 that maps back to the symmetric phase and the S that maps
one out are (as matrices) identical. However, because the fields on which they act have been
shifted by A one is actually going out on a different branch of the duality tree. It is equivalent to
going out on the one shifted by the y vector. The rephasing operations B and C may be similarly
defined as:

B̂s,s ≡ Ss BSs−1 δs,s , Ĉs,s ≡ Ss CSs−1 δs,s , (2.19)

where in the last section we have shown how one maps these rephasings through a Seiberg
duality explicitly (they are diagonal on the space of Seiberg phases because the B operation
simply rephases the fields, and does not mix different fields). Hence one should think of B̂ as
working on all points of the duality tree simultaneously, and one thinks of  as mapping the
entire tree into itself.
The above operators satisfy the Heisenberg group structure on the duality tree. Let us consider
a general quiver in the tree Qs1 and the action of Â, B̂ and Ĉ on this quiver

Âs,s B̂s ,s1 Qs1 = δ(s ,s1 ) Ss ABSs−1


1
Qs1 = δ(s ,s1 ) Ss BACSs−1
1
Qs1

= B̂s,s Âs ,s Ĉs ,s1 Qs1 , (2.20)

where we make a special note that there is only one nonzero entry for each s and s such that
the only implicit summation is over those s in the δ symbols: the indices in the Seiberg duals are
fixed. All of them are in fact exactly equal to s1 because they each contain a Kronecker δ. It is
again the action of A which is nontrivial and mixes the fields, so exactly which gauge group the
entries of s are referring to depend only on the order in which the gauge groups are listed, which
is changed by the presence of an A. In fact all of the relations found in the symmetric phase
descend to the entire tree, and so we find

ÂB̂ = B̂ ÂĈ, Ĉ commutes with everything, Âx = B̂ x = Ĉ x = 1. (2.21)

This, then, is an action on the whole space of possible Seiberg dual theories, and the equals sign
are read only up to the center of the gauge group. The ranks and possible gauge couplings are
different in each Seiberg phase, and so we take that the gauge redundancies are modded out in
each phase individually. There may be some understanding of this as some duality among the
relevant Faddeev–Popov ghosts, however we satisfy ourselves here by simply noting that the
extra factors are purely gauge.
B.A. Burrington et al. / Nuclear Physics B 757 (2006) 1–18 9

3. Finite Heisenberg groups acting on the space of Seiberg dual quivers

3.1. C3 /Z3

Let us first consider a simple example discussed originally in [1]. We take the gauge theory
dual to the Maldacena limit of string theory on C3 /Z3 where the orbifold action is given by
(z1 , z2 , z3 ) → (ξ z1 , ξ z2 , ξ −2 z3 ) where ξ is a cubic root of unity. The quiver diagram is repre-
sented in the center of Fig. 1. Let the rank of each of the gauge group be Nc . Performing Seiberg
duality at the node doubles its rank and the node is represented by a filled circle. Three different
quivers can be obtained by performing duality at the 3 different nodes (see Fig. 1). We will show
in this section that there exists a Heisenberg group acting on the three Seiberg dual quivers.
We label the dual quivers on the top, left and right as T , L and R respectively. The A trans-
formation permutes the nodes between different Seiberg dual quivers. Let it act as
1L → 2T → 3R, 2L → 3T → 1R, 3L → 1T → 2R, (3.1)
where the number in the front denotes the node corresponding to the quiver represented by the
alphabet next to it. B and C transformations are phase transformations of chiral fields. Let us
denote the phase of field Ui by ui . These phases respect invariance of the superpotential and
anomaly cancellation. These relations for the phases in the left quiver are
u1 u2 u3 = 1, (3.2)
 3 3 N  6 6 N  N
u1 u2 = 1, u2 u3 = 1, u63 u61 = 1. (3.3)
They are further simplified to obtain
u3 = (u1 u2 )−1 , (u1 u2 )3N = 1, 1 = u2 = 1.
u6N 6N
(3.4)

Fig. 1. C3 /Z3 quivers. Each Seiberg dual does not possess a Z3 symmetry. However, the space of Seiberg duals has a
manifest shift symmetry generated by Z3 .
10 B.A. Burrington et al. / Nuclear Physics B 757 (2006) 1–18

Below we write a particular solution for the phase assignments. We write it in terms of a 3N th
root of unity denoted as ω i.e. ω3N = 1. The topmost row denote the subscript corresponding to
the fields whose phases are written below in that column. The numbers in the rows corresponding
to B and C denote the powers to which ω is raised.

1 2 3
B L 0 0 0
T 0 1 −1
R 1 0 −1 (3.5)
C L −1 1 0
T 0 −1 1
R 1 0 −1

Let us make a general remark about a technical point. The generators of the Heisenberg group
satisfy
AB = BAC, AC = CA, BC = CB, Aq = B q = C q = 1, (3.6)
up to an element in center of the gauge group. These can be rewritten as
C −1 A−1 B −1 AB = Z1c , A−1 C −1 AC = Z2c , B −1 C −1 BC = 1,
Aq = 1, B q = Z3c , C q = Z4c , (3.7)
where Zic s are elements in center of the gauge group.
Thus, after explicitly constructing the elements A, B and C, we need to present the central
elements involved in the construction of the group. The particular elements in the center of the
gauge group needed for the Heisenberg group can be written in terms of a set of 3 integers
(a1 , a2 , a3 ). These three elements will stand for a rephasing at the nodes (1, 2, 3) by ω3a1 /2 ,
ω3a2 , ω3a3 respectively for the case of the left quiver diagram. The factor of 1/2 in the exponent
of the phase associated with node 1 is due to the fact that the rank of the gauge group at node 1
is twice that of the others for the left quiver diagram. The needed elements of the center of the
gauge group as denoted in Eqs. (3.7) can then be written as
c
Z4L : a2 = a3 = 1, a1 = 0,
Z1T : a1 = a3 = −1,
c
a2 = 0,
Z3R : a1 = a2 = 1,
c
a3 = 0,
Z3T = Z1T ,
c c
Z4T = −Z1T
c c
, c
Z4R = Z3R
c
,
Z1L = Z1R = Z2L = Z2T = Z2R = Z3L = 1.
c c c c c c

The center of the gauge group corresponds to the left, top and the right quiver diagrams depend-
ing on the extra subscript (L, T or R) respectively.

3.2. Heisenberg group in Y 4,2

Consider the quiver diagram for Y 4,2 as shown in Fig. 2 with chiral fields Ui , Yi and Zi It
has a cyclic Z2 -symmetry interchanging the nodes as (15) (26) (37) (48). It is convenient to
introduce the following notation, U = [U2 , U3 , U4 , U6 , U7 , U8 ]T , Y = [Y1 , Y3 , Y4 , Y5 , Y7 , Y8 ]T
and Z = [Z1 , Z5 ]T to denote the sets of chiral fields. The A transformation acts on chiral fields
B.A. Burrington et al. / Nuclear Physics B 757 (2006) 1–18 11

Fig. 2. Y 4,2 quiver diagram.

as

0 I3 0 1 0 I3
AU = U, AZ = Z, AY = Y, (3.8)
I3 0 1 0 I3 0
where I3 is a 3 × 3 matrix. We will try to find other symmetries which are related to phase
changes of the chiral fields. Let the phases be represented by small case letters. The relations
among the phases arising due to invariance of the superpotential are
z1 u2 y3 u8 = 1, u2 u3 y4 = 1, u3 u4 y5 = 1,
u4 z5 u6 y7 = 1, u6 u7 y8 = 1, u7 u8 y1 = 1. (3.9)
These equations determine the action on Y. Y → [(u7 u8 )−1 , (z1 u2 u8 )−1 , (u2 u3 )−1 , (u3 u4 )−1 ,
(u4 u6 z5 )−1 , (u6 u7 )−1 ]T Y. There are further constraints on the phases due to anomalies. They
are
 2 N  2 N  2 N  2 N
u8 y1 z1 = 1, z1 u2 y4 = 1, u2 y3 y5 u23 = 1, u3 y4 y7 u24 = 1,
 2 N  N  2 N  2 N
u4 y5 z5 = 1, z5 y8 u26 = 1, u6 y7 y1 u27 = 1, u7 y8 y3 u28 = 1.
(3.10)
These can be simplified to obtain

2 = u4 = u6 = u8 ,
u2N 2 = u6 , 4 = u8 ,
2N 2N 2N
uN N
uN N
 N  N  N  N
u3 u2 u3 u3
= , z1N = , z5N = . (3.11)
u7 u8 u2 u4
Both the other generators of the Heisenberg group, namely B and C, are phase changes which
will act on U, Z and Y diagonally. They should also satisfy relations (3.11) and (3.9). A particular
solution is
B: z1 = u4 = u7 = u8 = 1, u2 = u3 = ω, z5 = u6 = ω−1 ,
C: u2 = u4 = u6 = u8 = 1, z1 = z5 = ω, u3 = u7 = ω−1 . (3.12)
Here, ω2N = 1. Thus, we have the action of Heis(Z2 × Z2 ).
12 B.A. Burrington et al. / Nuclear Physics B 757 (2006) 1–18

Center of the gauge group


The Heisenberg group is closed up to the center of the gauge group. It will be useful to
first give a quick look at the center. The center has 8 generators with 8 parameters (ai say), one
acting at each node. It changes the chiral field in its fundamental (antifundamental) representation
by ω±ai . It is convenient to work with a new set of generators αi , such that αi = ai − ai+1
(α8 = a8 − a1 ). However, there exists a relation between them.
The center acts on the chiral fields as

U → diag ω2α2 , ω2α3 , ω2α4 , ω2α6 , ω2α7 , ω2α8 U,

Z → diag ω2α1 , ω2α5 Z,

Y → diag ω−2α7 −2α8 , ω−2α2 −2α8 −2α1 , ω−2α2 −2α3 , ω−2α3 −2α4 , ω−2α4 −2α6 −2α5 ,

ω−2α6 −2α7 Y. (3.13)
The generators αi satisfy a relation

α1 + α2 + α3 + α4 + α5 + α6 + α7 + α8 = 0. (3.14)
For the transformations given in (3.12), the particular elements in the center are

Z1c : α2 = α3 = 1, α5 = α6 = −1, α1 = α4 = α7 = α8 = 0,
Z2 : αi = 0,
c

Z3c = Z1c ,
Z4c : α1 = α5 = 1, α3 = α7 = −1, α2 = α4 = α6 = α8 = 0. (3.15)

3.2.1. Seiberg duals of Y 4,2


Toric duality in the case of Y p,q spaces shifts one of the singlet fields in the outside of the
quiver diagram [12]. Therefore, we concentrate on two Seiberg duals phases of Y 4,2 in Fig. 3.
They are obtained by Seiberg dualizing at node 2 and at node 6. We can think of the A symmetry
here as the one which takes node i of the left quiver to a node i + 4 (i − 4 for i > 4) of the right
quiver, and vice versa. B and C transformations are changes of phases of chiral fields, which are

Fig. 3. Two toric phases of Y 4,2 quiver.


B.A. Burrington et al. / Nuclear Physics B 757 (2006) 1–18 13

constrained. For the left quiver, the superpotential conditions are

y1 u7 u8 = 1, y3 u8 u1 = 1, y2 u1 u3 = 1,
y5 u3 z2 u4 = 1, y7 u4 z5 u6 = 1, y8 u6 u7 = 1. (3.16)

The anomaly cancellation conditions are


 2 2 N  N  2 2 N  N
u8 u1 y1 y2 = 1, u23 y2 z2 = 1, u1 u3 y3 y5 = 1, u24 z2 y7 = 1,
 2 N  2 N  2 2 N  N
u4 z5 y5 = 1, u6 y8 z5 = 1, u6 u7 y1 y7 = 1, u27 u28 y8 y3 = 1.
(3.17)
The former set helps to write the phases yi in terms of the other phases. The second set can be
reduced to obtain
 N
u7
u2N
1 = u 2N
7 , uN
4 = uN
8 = ,
z2
 N  
u1 u1 z2 N
uN3 = u N
6 = , z5
N
= . (3.18)
z2 u7
A particular solution is to consider the phases

u1 = u7 = 1, u3 = u4 = u6 = ω, z2 = z5 = u8 = ω−1 . (3.19)

Next, we construct the phases for the transformations B and C acting on all the chiral fields in the
two quivers. We present them in the table below, where the topmost row are the subscripts of ui
or zi . The numbers in other rows are the powers to which ω is raised for the given transformation
indicated on the left. The letters L and R in the second column refer to left and right quivers
drawn in Fig. 3. We assume ω2N = 1.

1 2 3 4 5 6 7 8
B L 0 −1 1 1 −1 1 0 −1
R 0 0 0 0 0 0 0 0 (3.20)
C L 0 −1 1 1 −1 1 0 −1
R −1 1 0 −1 0 −1 1 1

The particular elements of the center of the gauge group needed to satisfy the Heisenberg algebra
in this case can be written in the notation used in Section 3.2. We will use notation Zic to denote
various elements in the center and αi as its generator. We will also put a subscript L or R to
denote the action of the center on the two different quivers. The necessary Zic are
c
Z1L = Z1R
c
= Z2L
c
= Z2R
c
= Z3R
c
=I (identity element),
c
Z3L : α3 = α4 = α6 = 1, α2 = α5 = α8 = −1, α1 = α7 = 0,
c
Z4L = Z4R
c
,
Z4R : α2 = α7 = α8 = 1,
c
α1 = α4 = α6 = −1, α3 = α5 = 0. (3.21)
14 B.A. Burrington et al. / Nuclear Physics B 757 (2006) 1–18

Fig. 4. Quiver gauge theory for Y6,3 with a manifest Z3 shift symmetry.

3.3. Y 6,3 quiver with manifest shift symmetry A

The symmetric phase of the Y 6,3 quiver is given in Fig. 4. It possesses an A symmetry which
act on nodes as (1, 5, 9) (2, 6, 10) (3, 7, 11) (4, 8, 12). The Heisenberg group for this case has
been worked out in [2]. The superpotential and anomaly conditions are

u1 u2 y3 = 1, u2 u3 y4 = 1, u3 z4 u5 y6 = 1,
u5 u6 y7 = 1, u6 u7 y8 = 1, u7 z8 u9 y10 = 1,
u9 u10 y11 = 1, u10 u11 y12 = 1, u11 z12 u1 y2 = 1, (3.22)

 2 N  N  2 2 N  2 N
u1 y3 z12 = 1, u21 u22 y2 y4 = 1, u2 u3 y3 y6 = 1, u3 z4 y4 = 1,
 2 N  2 2 N  2 2 N  2 N
u5 z4 y7 = 1, u5 u6 y6 y8 = 1, u6 u7 y7 y10 = 1, u7 z8 y8 = 1,
 2 N  2 2 N  2 2 N
z8 u9 y11 = 1, u9 u10 y10 y12 = 1, u10 u11 y11 y2 = 1,
 2 N
u11 z12 y12 = 1. (3.23)
One particular solution is

B: u1 = u6 = u7 = z8 = 1, u2 = u5 = u11 = z12 = ω,
−1
u3 = z4 = u9 = u10 = ω ,
C: u1 = u3 = z4 = u5 = u7 = z8 = u9 = u11 = ω−1 ,
u2 = u6 = u10 = ω, z12 = ω5 . (3.24)
In order to write the center of the gauge group, we will use a similar notation as in Section 3.2
for the Y 4,2 case where ai for i = 1, . . . , 12 denote the generator at each node. These genera-
tors are commutative and each of them changes the phase of an incoming (outgoing) quiver by
ω−3ai (ω+3ai ). We take linear combinations of the generators as αi = ai − ai+1 (α12 = a12 − a1 ).
B.A. Burrington et al. / Nuclear Physics B 757 (2006) 1–18 15

Then the elements in the center of the gauge group are

Z1c : α5 = α11 = 1, α10 = α12 = −1 zero otherwise,


Z2c : α12 = 2, α8 = −2 zero otherwise,
Z3c : α2 = α5 = α11 = α12 = 1, α3 = α4 = α9 = α10 = −1,
α1 = α6 = α7 = α8 = 0,
Z4c : α2 = α6 = α10 = 1, α1 = α3 = α4 = α5 = α7 = α8 = α9 = α11 = −1,
α12 = 5. (3.25)

3.3.1. Seiberg duals of Y 6,3


Now we look at certain Seiberg duals of the Y 6,3 quiver. Let us label the figures in Fig. 5 on
the top, left and right as T , L and R. The A transformation acts by taking a node i of the top

Fig. 5. Three Seiberg phases of Y6,3 . Each phase lacks a Z3 shift symmetry but the symmetry is present when considering
the space of three Seiberg phases.
16 B.A. Burrington et al. / Nuclear Physics B 757 (2006) 1–18

(left) quiver to a node i + 4 (i − 8 if i > 8) of the right (top) quiver as well as a node i of the
right quiver to a node i + 4 (i − 8 if i > 8) of the left quiver. Let us consider the quiver diagram
on the right and find the conditions for the phases associated with the B and C transformations.
Clearly, similar conditions can be written for the top and the left.

y4 u2 z1 u3 = 1, y6 u3 z4 u5 = 1, y7 u5 u6 = 1, y8 u6 u7 = 1, y10 u7 z8 u9 = 1,
y11 u9 u10 = 1, y12 u10 u11 = 1, y2 u11 u12 = 1, y1 u12 u2 = 1, (3.26)

 N N  2 N
z1 u22 y1 = 1,
u22 u212 y4 y2 = 1, z1 u3 y6 = 1,
 2 N  2 N  2 2 N
u3 z4 y4 = 1, z4 u5 y7 = 1, u5 u6 y6 y8 = 1,
 2 N  2 N  2 N
u6 y7 y10 u27 = 1, u7 z8 y8 = 1, z8 u9 y11 = 1,
 N  N  2 2 N
y10 u29 u210 y12 = 1, u210 u211 y11 y2 = 1, u11 u12 y12 y1 = 1. (3.27)
Here the set (3.26) comes from superpotential invariance and the set (3.27) are the anomaly
constraints. The first set allows the yi ’s to be solved in terms of other phases. The second set can
be simplified to obtain
   N
z1 u2 N u23
u3N
3 = u3N
2 , z4N= , uN
5 = , 6 = (z1 u3 ) ,
uN N
u3 u2
   N
z1 u3 N z1 u22
uN
7 = uN
2 , z8 =
N
, uN
9 = uN
3 , uN
10 = ,
u2 u3
 N
u22
uN
11 = , 12 = (z1 u2 ) .
uN N
(3.28)
u3
For a particular solution, we will write B and C acting on ui and zi only. The numbers in the top
row of the table below denote the subscripts of ui and zi . The numbers in the rows corresponding
to B and C denote the powers to which ω is raised.

1 2 3 4 5 6 7 8 9 10 11 12
B R 1 0 1 0 −1 −1 0 −1 1 0 −1 1
L 0 0 0 0 0 0 0 0 0 0 0 0
T 1 1 0 1 −1 0 1 −1 −1 0 −1 0 (3.29)
C R 1 0 1 0 −1 −1 0 −1 1 0 −1 1
L 1 0 −1 1 1 0 1 0 −1 −1 0 −1
T −1 −1 0 −1 1 0 −1 1 1 0 1 0

We will write the center of the gauge group in terms of generators αi as defined in the previous
subsection. We reinterpret few of the generators as

for L quiver α4 = a4 − a6 , α5 = a5 − a7 , α6 = a6 − a5 ,
for T quiver α8 = a8 − a10 , α9 = a9 − a11 , α10 = a10 − a9 ,
for R quiver α1 = a1 − a3 , α2 = a2 − a1 , α12 = a12 − a2 .
B.A. Burrington et al. / Nuclear Physics B 757 (2006) 1–18 17

The elements in the center of the gauge group needed for the algebra are
c
Z1R = Z1L
c
= Z2R
c
= Z2L
c
= Z2T
c
= Z3L
c
=I (identity element),
c
Z3R : α1 = α3 = α9 = α12 = 1, α5 = α6 = α8 = α11 = −1,
α2 = α4 = α7 = α10 = 0,
c
Z4L : α5 = α8 = α9 = α11 = 1, α1 = α2 = α4 = α7 = −1,
α3 = α6 = α10 = α12 = 0,
c
Z1T : α3 = α9 = α10 = α12 = 1, α1 = α4 = α9 = α7 = −1,
α2 = α6 = α8 = α11 = 0,
c
Z4R = Z3R
c
, c
Z3T = −Z4T
c
= Z1T
c
. (3.30)

4. Conclusion

We have demonstrated that the space of Seiberg duals to a given quiver gauge theory inherits
the action of a finite Heisenberg group. An interesting new feature is that to consider the action
of the Heisenberg group we are forced to enlarged the space from one quiver to the set of quiver
gauge theories arising from performing Seiberg duality at various nodes.
It has been shown that different Seiberg phases are related to different toric phases [7,8,10]. It
would, therefore, be nice to study the action of the Heisenberg group purely as a symmetry in the
space of toric phases. This is particularly interesting in the space of quiver gauge theories which
are Seiberg dual with different ranks of the gauge groups.
Assuming that there is a well-defined string theory dual for each Seiberg phase, our findings
imply that charges of branes of one string theory are related to the charges of branes of a different
theory. In a sense this provide a sort of stringy toric duality. A perhaps more speculative way of
describing this situation is that string theory views different Seiberg duals as twisted sectors of a
given theory. It would be interesting to explore these suggestions and we hope to return to some
of these questions in the future.

Acknowledgements

We are grateful to Ami Hanany for a very enlightening discussion. B.B. wishes to thank David
Morrissey and Paul de Medeiros for useful conversations and clarifications. This work is partially
supported by US Department of Energy under grant DE-FG02-95ER40899 to the University of
Michigan.

References

[1] S. Gukov, M. Rangamani, E. Witten, Dibaryons, strings, and branes in AdS orbifold models, JHEP 9812 (1998)
025, hep-th/9811048.
[2] B.A. Burrington, J.T. Liu, L.A. Pando Zayas, Finite Heisenberg groups in quiver gauge theories, hep-th/0602094.
[3] B.A. Burrington, J.T. Liu, L.A. Pando Zayas, Central extensions of finite Heisenberg groups in cascading quiver
gauge theories, hep-th/0603114.
[4] G. Moore, Mathematical aspects of fluxes, talk at KITP: http://online.kitp.ucsb.edu/online/strings05/moore/;
D. Belov, Holographic approach to the action of self-dual fields and RR fields, talk at KITP: http://online.kitp.
ucsb.edu/online/strings05/belov/.
[5] D. Belov, G.W. Moore, Conformal blocks for AdS5 singletons, hep-th/0412167.
18 B.A. Burrington et al. / Nuclear Physics B 757 (2006) 1–18

[6] N. Seiberg, Electric–magnetic duality in supersymmetric non-Abelian gauge theories, Nucl. Phys. B 435 (1995)
129, hep-th/9411149.
[7] B. Feng, A. Hanany, Y.H. He, A.M. Uranga, Toric duality as Seiberg duality and brane diamonds, JHEP 0112 (2001)
035, hep-th/0109063.
[8] B. Feng, A. Hanany, Y.H. He, D-brane gauge theories from toric singularities and toric duality, Nucl. Phys. B 595
(2001) 165, hep-th/0003085.
[9] S. Franco, A. Hanany, Y.H. He, P. Kazakopoulos, Duality walls, duality trees and fractional branes, hep-th/0306092.
[10] C.E. Beasley, M.R. Plesser, Toric duality is Seiberg duality, JHEP 0112 (2001) 001, hep-th/0109053.
[11] C. Csaki, H. Murayama, ’t Hooft anomaly matching for discrete symmetries, hep-th/9805053.
[12] S. Benvenuti, A. Hanany, P. Kazakopoulos, The toric phases of the Y (p, q) quivers, JHEP 0507 (2005) 021, hep-th/
0412279.
Nuclear Physics B 757 (2006) 19–46

Living dangerously with low-energy supersymmetry


Gian F. Giudice ∗ , Riccardo Rattazzi
CERN, Theory Division, CERN, CH-1211 Geneva 23, Switzerland
Received 9 June 2006; accepted 14 July 2006
Available online 25 September 2006

Abstract
We stress that the lack of direct evidence for supersymmetry forces the soft mass parameters to lie very
close to the critical line separating the broken and unbroken phases of the electroweak gauge symmetry.
We argue that the level of criticality, or fine-tuning, that is needed to escape the present collider bounds
can be quantitatively accounted for by assuming that the overall scale of the soft terms is an environmental
quantity. Under fairly general assumptions, vacuum-selection considerations force a little hierarchy in the
ratio between m2Z and the supersymmetric particle square masses, with a most probable value equal to a
one-loop factor.
© 2006 Elsevier B.V. All rights reserved.

1. Introduction

For almost three decades, the gauge hierarchy problem has been the only reason to think that
the Standard Model (SM) should be overthrown right around the weak scale. It has inspired the
construction of a huge stack of new models and is arguably one of the main motivation to build
the Large Hadron Collider. As it is normally formulated, the problem lies in the difficulty to
understand the relatively low value of the Higgs mass parameter |m2H | ∼ (100 GeV)2 in a frame-
work in which the SM is valid up to some ultra-high scale Λ, for instance for Λ of the order of
the Planck scale MP . We can equivalently picture the problem as one of criticality. Imagine the
fundamental theory at the Planck scale has a few free parameters. In string theory, to be perhaps
more concrete, these parameters may correspond to the (discrete) set of vacuum expectation val-
ues of the moduli fields. Let us consider the phase diagram for electroweak symmetry breaking,
in the space of these parameters. Over the bulk of the parameter space, |m2H | is expected to be of

* Corresponding author.
E-mail address: gian.giudice@cern.ch (G.F. Giudice).

0550-3213/$ – see front matter © 2006 Elsevier B.V. All rights reserved.
doi:10.1016/j.nuclphysb.2006.07.031
20 G.F. Giudice, R. Rattazzi / Nuclear Physics B 757 (2006) 19–46

order MP2 , and therefore either H  ∼ MP or H  = 0 depending on the sign of m2H . The hierar-
chy problem is now simply stated as: if the critical line separating the two phases is not special
from the point of view of the fundamental theory, why are the parameters in the real world so
chosen as to lie practically atop the critical line?
Supersymmetry is relevant to this puzzle for two reasons. First, because it selects the crit-
ical line as a locus of enhanced symmetry.1 More precisely, in the minimal supersymmetric
SM with both supersymmetry and Peccei–Quinn (PQ) symmetry unbroken, the Higgs poten-
tial V ∝ (|H1 |2 − |H2 |2 )2 is indeed “critical”, in the sense that the symmetric H1 = H2 = 0
point is a minimum of the potential, but it can be destabilized by arbitrarily small mass per-
turbations. Second, supersymmetry is, under rather general circumstances, broken only by tiny
non-perturbative effects [1]. These will unavoidably move the theory slightly off the critical
line. Effectively this corresponds to the generation of tiny mass terms ∼ MP e−1/α  MP which
generically lead to electroweak symmetry breakdown (while stabilizing at the same time the flat
direction |H1 | = |H2 |) at a correspondingly low scale.
In hidden sector models, at energies below the Planck mass MP , supersymmetry breaking is
accurately parametrized by soft supersymmetry-breaking terms of order MS . The electroweak
vacuum dynamics is then controlled by renormalization-group (RG) evolution of the soft terms
from MP down to MS . One nice feature of this evolution is that, over a wide region of the soft
parameter space, one of the eigenvalues of the Higgs squared-mass matrix flows to a negative
value somewhere between MP and MS [2]. This makes electroweak symmetry breaking a rather
natural phenomenon within supersymmetric extensions of the SM.
For the sake of this discussion we should, however, be slightly more precise. Notice that, since
the RG evolution is homogeneous in the soft terms, the RG scale Qc at which the Higgs mass
eigenvalue crosses zero depends on MP and on dimensionless ratios of soft parameters, but it is
parametrically unrelated to MS . Furthermore, as long as the Higgs mass matrix is positive def-
inite at MP , since the evolution is logarithmic in the RG scale, Qc is exponentially suppressed
with respect to MP (see Fig. 1, top frame). Therefore, the supersymmetric parameter space is
essentially divided into two regions (phases) characterized respectively by MS  Qc  MP and
Qc  MS  MP . In the first region, at the scale MS where RG evolution of the soft terms is
frozen, the Higgs mass matrix has a negative eigenvalue of magnitude ∼ MS2 , due to the hierar-
chical separation between MS and Qc . Given the structure and size of the Higgs quartic potential,
this implies a weak scale H 2 ∼ MS2 /g 2 and m2Z ∼ MS2 . In the second region, RG evolution is
frozen with a positive definite Higgs mass matrix so that the Higgs field does not break elec-
troweak symmetry. We call this region the unbroken phase, although electroweak symmetry is
still spontaneously broken, but only by fermion condensation in QCD. The resulting spectrum is
therefore vastly different than in the H  ∼ MS /g phase. All elementary SM particles, including
W and Z, weigh less than about 100 MeV, while the superpartners are still at MS , so that the
pattern is mZ < ΛQCD  MS . While the unbroken phase region does not resemble even approx-
imately the world we live in, the broken phase region makes instead supersymmetry relevant to
phenomenology, as it solves the gauge hierarchy problem and explains electroweak symmetry
breaking in a unitary conceptual framework.
The generic spectrum MS ∼ mZ of the broken phase also held up great expectations for a
discovery of supersymmetry at LEP [3]. As those expectations were then frustrated by the exper-

1 Conformal symmetry provides in principle an alternative symmetry principle. The reason why it is not viable is very
simple: the presence of a fundamental physics scale, say MP , both defines the hierarchy problem and explicitly breaks
conformal invariance.
G.F. Giudice, R. Rattazzi / Nuclear Physics B 757 (2006) 19–46 21

Fig. 1. The running of the Higgs mass parameter m22 as a function of the RG scale Q. The top frame shows the case of a
generic supersymmetric setup, leading to |m22 (MS )| = O(MS2 ) and MS  Qc  MP . The bottom frame corresponds to
a fine-tuned choice of soft terms, such that |m22 (MS )|  MS2 and MS  Qc .

imental data, in the post-LEP era also the broken phase does not seem to qualitatively describe
our world. The direct and indirect limits placed by LEP point instead to a spectrum where MS
is at least an order of magnitude larger than mZ ( ΛQCD ), corresponding to the boundary be-
tween the two phases. The strongest, but not unique, constraint is given by the experimental lower
limit on the mass of the lightest CP-even Higgs mh . Given the tree-level theoretical upper bound
mh < mZ | cos 2β|, the experimental constraint can be satisfied only by pumping up the top-stop
quantum corrections to the Higgs quartic coupling [4]. Over most of the parameter space, this
implies stop masses mt˜ that range closer to a TeV than to 100 GeV. We can then work out where
Qc should be, by expanding the RG evolution of the negative mass eigenvalue in the Higgs po-
tential between Qc and MS . For the sake of the argument we can focus on the case tan β → ∞,
where electroweak breaking is driven by the Higgs mass parameter m22 and where we find

m2Z dm22  Qc
= −m2 
2
 ln . (1)
2 d ln Q Qc MS
For typical choices of supersymmetric parameters, the stop masses dominate the RG evolution
and dm22 /d ln Q  0.1m2t˜ . For mt˜ ∼ 1 TeV, we find ln(Qc /MS ) < 1, which is so small that
there is not even a meaningful scale separation between the overall supersymmetric scale MS
and Qc (see Fig. 1, bottom frame). The coincidence of these two conceptually unrelated mass
scales is one way of viewing the fine-tuning problem of supersymmetric models. Why should the
fundamental theory prefer such critical choice of parameters? A more quantitative illustration of
this question will be given in the next section and the rest of the paper is an attempt to provide
an answer.
22 G.F. Giudice, R. Rattazzi / Nuclear Physics B 757 (2006) 19–46

2. Fine-tuning and criticality in low-energy supersymmetry

In this section, we want to explain, in a more quantitative fashion, the connection between fine
tuning and criticality. Let us consider the phase diagram in the parameter space of the minimal
supersymmetric model, spanned by all independent dimensionless ratios of the coefficients of
soft supersymmetry-breaking terms. For illustrative purposes, we reduce this multi-dimensional
space into a plane, by taking unified gaugino masses (M) and universal scalar masses (m2 ) at the
GUT scale. For the moment, we also set to zero all trilinear soft terms at the GUT scale (A = 0),
and choose a small bilinear term B at the scale MS , corresponding to a fixed and moderately large
value of tan β, in the region where radiative electroweak breaking occurs. These hypotheses are
just meant to simplify the visualization, but the discussion we present here remains valid also for
general soft terms. In the case under consideration, the phase diagram can be described in terms
of only two variables, which we take to be m2 /μ2 and M 2 /μ2 , the square ratios of the common
scalar and gaugino masses to the μ parameter, with all quantities defined at the GUT scale.
The SM presents two phases, with broken (H  = 0) or unbroken (H  = 0) electroweak sym-
metry. The situation is more complicated in the supersymmetric version, because of the extended
structure of the Higgs sector and of the properties of supersymmetry. We recall that the Higgs
potential, along the neutral field components is
g2 + g 2  2
V= |H1 |2 − |H2 |2 + m21 |H1 |2 + m22 |H2 |2 − m23 (H1 H2 + h.c.), (2)
8
and that we are working in the limit of small m23 . The boundary condition at the GUT scale is
m21,2 = m2 + μ2 .
The phase diagram of the minimal supersymmetric SM is shown in Fig. 2. A first peculiarity
of supersymmetry is the existence of phases where color and electric charge are broken. This
happens, for instance, at negative m2 and small M, where the third-generation squark Q̃3 gets
a vacuum expectation value. Actually, assuming strict universality, there is an even larger re-
gion where the selectron gets a vacuum expectation value, which is not shown in Fig. 2 since,
for the sake of argument, we take a common scalar mass only for the particles involved in the
conventional SU(2) × U (1) breaking pattern (third-generation squarks and the two higgses).
More interesting is a special multi-critical point, separating the various Higgs phases, that cor-
responds to vanishing Higgs bilinear terms (m21 = m22 = m23 = 0).2 This point, which is actually
a surface in the case of general soft terms, occurs at negative m2 , in the example we are consider-
ing. Moving away from the multi-critical point, different phases emerge, depending on the signs
and the values of m21 and m22 at the scale MS . For positive m21,2 , the potential is stabilized at the
origin; the scale MS is larger than the critical scale Qc , and electroweak symmetry is unbroken.
Notice that this phase (marked as H1,2  = 0 in Fig. 2) extends, for M = 0, to rather large val-
ues of m2 /μ2 . This is a peculiarity of the assumption of strict universality which, together with
the known value of the top mass, leads to a certain cancellation of the contribution to m2Z pro-
portional to m2 . Varying the top Yukawa coupling (or, ultimately, tan β), one can obtain higher
degrees of cancellation, approaching what is known as “focus point” [5]. To compensate for
this reduced dependence on m2 (a characteristic of universality, not shared by generic soft-term
structures) we have expanded in Fig. 2 the scale of the vertical axis, with respect to the horizontal

2 These three conditions cannot be in general satisfied in the case of only two free parameters. However, Fig. 2 corre-

sponds to fixed tan β, and thus m23 automatically vanishes, whenever m21 = m22 = 0.
G.F. Giudice, R. Rattazzi / Nuclear Physics B 757 (2006) 19–46 23

Fig. 2. The phase diagram of the minimal supersymmetric SM, assuming a universal scalar mass m2 , a gaugino unified
mass M, a higgsino mass μ, and trilinear term A = 0, with all parameters defined at the GUT scale. The top Yukawa
coupling is fixed such that mt = 172.7 GeV and tan β = 10 in the usual phase with electroweak breaking. Some contours
are shown for masses of the lightest stop (Mt˜ ), the gluino (Mg̃ ), and the lightest chargino (Mχ + ). The green (gray) area
1
shows the region of parameters allowed after LEP Higgs searches. (For interpretation of the references to colour in this
figure legend, the reader is referred to the web version of this article.)

axis. For the same reason, the precise location of the boundary between the broken and unbroken
phases at small M sensitively depends on the values of the coupling constants and on the degree
of accuracy of the calculation. In our figures, we have chosen αs (mZ ) = 0.1176, and fixed the
top Yukawa coupling corresponding to mt = 172.7 GeV and tan β = 10 in the broken phase. We
have also limited our RG evolution to one-loop approximation.
In the limit of exact supersymmetry and PQ symmetry, all quadratic Higgs terms in Eq. (2)
vanish. Actually, since in supergravity scenarios the PQ breaking can easily arise only from su-
persymmetry breaking [6], we will refer to this case (m21,2,3 = 0) as the supersymmetric limit. In
this limit, the Higgs potential has a flat direction H1  = H2 , characteristic of supersymmetric
D-terms. Supersymmetry breaking stabilizes this direction as long as m21 + m22 > 2|m23 |. If this is
not the case, the Higgs field slides up to the renormalization scale where the previous inequality
is satisfied, as in the Coleman–Weinberg mechanism. If m2 /μ2 < −1, this scale is actually larger
than the GUT scale cutoff. At any rate, the important point is that the Higgs vacuum expecta-
tion value is unrelated to the supersymmetry scale MS and in particular mZ  MS . This region,
which is of course experimentally ruled out is marked in Fig. 2 by m2A < 0. Indeed, its boundary
is characterized, in our analysis with fixed tan β, by the condition mA ≡ m21 + m22 < 0 since, in
the region of conventional electroweak breaking, 2m23 = sin 2β(m21 + m22 ).
In the rest of the phase diagram in Fig. 2 the Higgs vacuum expectation value is proportional
to supersymmetry-breaking terms, and therefore MS controls the size of electroweak breaking,
thus providing a potentially realistic solution to the hierarchy problem. Depending on whether
m21 or m22 is driven negative, we obtain two possible regions marked in Fig. 2 as H1  > H2 
and H2  > H1 , respectively. The first region, which occurs only at negative m2 , has phenom-
24 G.F. Giudice, R. Rattazzi / Nuclear Physics B 757 (2006) 19–46

enological difficulties in maintaining a perturbative top Yukawa coupling to large scales and in
making the Higgs mass sufficiently heavy. Therefore, we will concentrate on the region with
H2  > H1 .
In this region, the inequality Qc > MS is satisfied, and we can determine the overall mass scale
of supersymmetric particles from the condition that radiative electroweak breaking reproduces
the known value of mZ . The complete mass spectrum can then by computed at a given point of
the phase diagram, and in Fig. 2 we show some characteristic values of supersymmetric particle
masses. In the bulk of the region, we find that supersymmetric colored particles weigh typically
less than 2–3 times mZ , while some electroweak particles are lighter than mZ . The values of the
supersymmetric masses have only mild variations in the bulk of the region, but they precipitously
increase in the proximity of the critical line separating the broken and unbroken phases, where
the critical scale Qc rapidly approaches MS . Only near the boundary we can find supersymmetric
masses compatible with the present bounds from collider experiments. For instance, the chargino-
mass LEP bound Mχ + > 103.5 GeV at 95% CL [7] rules out all the region to the right of the
corresponding blue line in Fig. 2, allowing only the narrow strip between the blue and critical
lines. Actually the negative Higgs searches impose even stronger constraints on the allowed
region. Taking into account the limits on Higgs production at LEP [8] in the channels Zh, ZH ,
hA, H A (where h, H , A are the three neutral supersymmetric higgses), we find that the only
allowed points in Fig. 2 are those inside the green (gray) region, clustering along the critical line.
A first conclusion that we can draw from these results is that the most natural prediction of
supersymmetry on the spectrum of new particles has already been ruled out, and only small
corners of parameter space are still allowed. This conclusion is of course well known and it has
been already quantified in different ways [3,9]. Fig. 2 presents an alternative way to illustrate the
problem.
However, Fig. 2 also leads us to a new way of characterizing the allowed region, in terms of
criticality condition. The problem of understanding why supersymmetry may have chosen highly
untypical values of soft parameters, which appear to have the only effect to hide it from collider
searches, is now turned into the question of why supersymmetry wants to lie in a near-critical
condition. In the following, we will discuss possible statistical (or dynamical) attempts to explain
this puzzle. But before ending this section, we want to address the question of how general is our
conclusion that the only allowed parameter region of low-energy supersymmetry lies close to the
critical line, and we investigate if other regions, albeit tuned, can arise far from it.
It is well known that the experimental SM Higgs mass bound mh > 114.4 GeV at 95% CL
[10] does not directly apply to the supersymmetric case since, for a pseudoscalar mass mA near
mZ , the coupling of the lightest Higgs boson to the Z boson is reduced. In Fig. 3 we zoom
into the region of parameter space where this happens. The thin sliver extending away from the
multi-critical point corresponds to the region allowed by LEP searches where mA  mZ and
mh < 114 GeV. Besides the consideration that this region appears as a very special tuning of
the underlying parameters, we observe that it does not allow to depart significantly from the
critical condition. As explained in the appendix, the reason for this limited effect is that, to have
a suppressed hZZ coupling without conflict with the LEP searches in the hA channel, one needs
large corrections to the Higgs quartic coupling. This requires again to be near criticality.
It is also known that the stringent lower limit on the stop mass derived from the Higgs-mass
bound can be significantly relaxed for large values of the trilinear term At . This limit on Mt˜1 is
more than 1 TeV for vanishing stop mixing, but it is reduced to 200–300 GeV when the mixing
reaches the condition (At − μ/ tan β)2 /(Mt˜1 Mt˜2 ) = 6. This allows lighter stops and, apparently,
less fine tuning. Indeed, if we increase the value of A, the region allowed by Higgs searches
G.F. Giudice, R. Rattazzi / Nuclear Physics B 757 (2006) 19–46 25

Fig. 3. Same as Fig. 2, zooming in the allowed region where the Higgs pseudoscalar mass is close to mZ .

becomes slightly larger than what shown in Fig. 2. However, at the same time, large A terms
contribute to m2Z and reduce the overall value of MS . This has the effect of predicting a lighter
supersymmetric spectrum and push the mass contour lines of Fig. 2 closer to the critical line. In
this case, the chargino mass limit plays the dominant role, and the allowed region is still clustered
along the critical line. A certain relaxation could be achieved if gaugino-mass unification does
not hold, and if M2 > M3 at the GUT scale. In this case, the chargino mass limit plays a more
limited role, and we can increase further the value of A and make the Higgs boson heavier.
However, this possibility is limited by the bound on the stop mass. In conclusion, we find that the
connection between experimentally-allowed supersymmetric parameters and criticality is robust
under variations of the soft-term structure.

3. Statistical criticality

There have been various attempts to explain the tuning of low-energy supersymmetry by dy-
namical mechanisms or through extra symmetries [11–14]. Ref. [11] marries the little Higgs idea
to supersymmetry, suitably extending the minimal model in order to make one combination of
the two higgses a pseudo-Goldstone boson. The papers in Ref. [12], by providing extra contri-
butions to the Higgs quartic coupling, focus just on the tuning produced by the mh bound. The
papers in Refs. [11,12] represent departures from the minimal supersymmetric SM right at the
superparticle mass scale, which are in principle testable at future colliders. However these mod-
els are rather complicated and it is hard to believe that nature would choose such complication
just to hide supersymmetry at LEP. It is also fair to say that they do not fully solve the fine-tuning
problem of supersymmetry. Notice in passing that this last problem is also shared by the exten-
sion of the minimal model involving an extra Higgs singlet superfield, unlike what is commonly
stated but as it has been recently emphasized in a detailed study [15]. In the models in Ref. [13,
14] the theory retains the minimal field content up to some ultra high scale, and the apparent
26 G.F. Giudice, R. Rattazzi / Nuclear Physics B 757 (2006) 19–46

tuning is supposedly explained by the supersymmetry-breaking dynamics. Ref. [13] represents


a remarkable supergravity scenario where Qc is parametrically tied to MS , but it seems that the
lifting potential, upon which this results is fully based, does not have any sensible microscopic
motivation [16]. In Section 5 we will illustrate in more detail why the dynamical explanation in
Ref. [14] has difficulties.
Here we take a different approach and try instead to provide a statistical explanation of criti-
cality. We will be working under the multiverse or landscape hypothesis [17]. According to this
hypothesis, the fundamental description of nature features a tremendous multiplicity, a landscape,
of physically inequivalent vacua and our local universe represents but one domain of a multiverse.
With the parameters of the low-energy effective field theory changing from domain to domain,
statistical considerations can be applied to deduce, under some assumptions, the likelihood of
parameter configurations. In particular, observed properties of our domain, through their physi-
cal relations to the parameters, some of which known and some of which unknown, can imply
conditional probabilities on the unknown parameters. Weinberg’s prediction of the anthropically
favoured size of the cosmological constant [18] is an example of that, with the existence of galax-
ies and the size of primordial density perturbations playing the role of the measured data of our
domain.
In analogy with Weinberg’s approach to the cosmological constant problem, we will assume
that the soft-supersymmetry breaking mass parameters are environmental quantities varying
across the multiverse. Working in the context of hidden-sector models, this means that each
different vacuum in the landscape gives rise to a different set of soft mass parameters {mi } up
at a fixed scale, say at MP . As the simplest possibility, let first us assume that only the overall
supersymmetric mass scale MS is environmental. More precisely let us assume that at the Planck
scale the soft masses including μ are given by
mi = ci MS , (3)
with the dimensionless coefficients ci fixed everywhere throughout the landscape, while MS
varies. Let us also assume that all the other dimensionless couplings (gauge and Yukawa) are
fixed at the Planck scale. It is possible to think of field theoretic landscapes that realize this
condition, as we will discuss in Section 6. Let us consider the normal situation in which the
Higgs mass matrix is positive definite at the Planck scale. Under the above conditions also Qc
is fixed. Indeed the RG equations are homogeneous in the soft terms, so that the RG evolution
is written as the evolution of the ci with MS constant. Then Qc , corresponding to the RG scale
where
 
det M2 (Q) = m21 (Q)m22 (Q) − m43 (Q) ≡ MS4 c12 (Q)c22 (Q) − c34 (Q) (4)
turns negative, depends on the high-energy scale MP and on the dimensionless couplings
Qc = MP × F (ci , αa ) , (5)
but not on MS . Here by αa we collectively denote the gauge and Yukawa couplings at the Planck
scale. The physical values of the Higgs mass parameters are, in leading log approximation, equal
to the running masses computed at the RG scale Q ≡ MS . Two possibilities for the value of
MS in the multiverse domain comprising our universe are then given: (i) MS > Qc , for which
M2 is positive definite and thus H  = 0; (ii) MS < Qc , for which M2 as at least one negative
eigenvalue, implying H  = 0.
It is pretty clear we do not live in region (i), and in fact it is not even sure if in region (i)
there can exist anyone to ask this question [19,20]. Although a rich atomic structure may exist
G.F. Giudice, R. Rattazzi / Nuclear Physics B 757 (2006) 19–46 27

in this region [19], it would look so different from our world that it seems rather unlikely it
would be hospitable to life. Moreover, and more simply, it has been shown [20] that for H  = 0
any primordial baryon density is very efficiently converted into leptons (mostly neutrinos) by
electroweak sphalerons. These effects are now active down to temperatures of the order of ΛQCD ,
at which conversion of baryons into leptons is energetically favored. This feature of the H  = 0
universe seems rather solid as it does not depend very much on the Yukawa couplings of quarks
and leptons (as long as they remain weak). Therefore region (ii) is also strongly favored over
region (i) for anthropic reasons.
Compatibly with the prior that we must live in region (ii), we can ask what is the most likely
value we expect MS to have. The problem is phrased in complete analogy with Weinberg’s ap-
proach to the cosmological constant, with H  = 0 replacing the datum that galaxies exist. Then,
under the assumption that the distribution of MS is reasonably flat and featureless, and not peaked
at MS = 0, we expect MS ∼ Qc . For instance, as we will show in Section 6, in simple field theo-
retic modelling of the landscape, a typical expectation is that the number of vacua with MS < m
grows like a positive power mn . Then, treating all vacua as equally probable, the prior H  = 0
leads to a conditional probability
  M n dM
n QSc MSS for MS < Qc ,
dP = (6)
0 for MS > Qc .
By this equation the average logarithmic separation of scales is given by ln Qc /MS  = 1/n,
which agrees with the rough expectation MS ∼ Qc for a smooth distribution with n = O(1).
As the RG evolution slowly proceeds by 1-loop effects, the Higgs mass matrix M2 typically
develops a small O(α) negative eigenvalue at the scale MS ∼ Qc , where the running is frozen.
The weak scale will correspondingly be parametrically smaller than MS . By minimizing the
scalar potential at leading order in LS = ln Qc /MS , we have
  2
  2 d(det M2 )  α
m2Z cos2 2β + δ sin4 β = 2 L + O L 2
, (7)
m1 + m22 d ln Q Q=Qc
S
16π 2 S
where δ, defined in the appendix, represents the top-stop quantum correction to the Higgs quartic
coupling and where sin 2β = 2m23 /(m21 + m22 ). For tan β  5–10, the above equation is well
approximated by its tan β → ∞ limit
dm22
m2Z (1 + δ)  2 LS (8)
d ln Q


 2  g12  2   2 
2 3LS
= λt mt˜ + mt˜ + |At | −
2 2 2
M1 + μ − g2 M2 + μ
2 2
. (9)
L R 5 4π 2
By assuming the stop parameters to dominate the above equation with m2t˜ ∼ m2t˜ ∼ |At |2 ∼ MS2
L R
and by using ln Qc /MS  = 1/n, we find the following average little hierarchy
2 
mZ 9λ2t 1 114 GeV 2 1
 ×  0.15 × . (10)
MS2 4π 2 (1 + δ) n mh n
Notice that while the loop factor in Eq. (10) helps to explain the little hierarchy problem in
supersymmetry, it still falls short to explain it completely. Indeed, the Higgs mass limit requires
stop masses close to 1 TeV, a therefore a little hierarchy m2Z /MS2  0.01–0.1. However, as we
will argue in Section 6, it is rather reasonable to imagine field theoretic landscapes where n
28 G.F. Giudice, R. Rattazzi / Nuclear Physics B 757 (2006) 19–46

is somewhat bigger than 1, say O(a few), though not much bigger. For instance if there are
O(10500 ) vacua, as perhaps suggested by string theory [21], and if MS can range up to MP , then
n < 30. So it is reasonable for the ratio in Eq. (10) to be between 0.01 and 0.1 but not much
smaller, thus providing an argument why supersymmetry should be elusive at LEP but not at the
LHC. Of course there has been a price to pay. Supersymmetry looks tuned because throughout
the landscape it is much more likely to be in the region with H  = 0 than in the region H  = 0
: the most likely points with H  = 0 are then close to the boundary of the two regions, where a
little hierarchy is present.
We should stress that these conclusions are based on statistical arguments and that the average
in Eq. (10) has a variance of the same order, i.e., (X 2  − X2 )/X2 = 1 with X ≡ m2Z /MS2 .
In Section 7 we will discuss in a little more detail the natural ranges of particle masses in this
scenario. For the remaining part of this section we want instead to address some technical and
conceptual questions.
Let us address a technical question first. We have so far been working with 1-loop RG evolu-
tion. Accordingly we should be controlling only leading log effects ∼ (α ln)n and neglect finite
threshold effects. On the other hand, we have used as part of our logic the RG evolution between
two scales Qc and MS that practically coincide, ln Qc /MS ∼ 1/n  1, in apparent contradic-
tion to our leading-log approximation. Our results are nonetheless correct. Consider indeed the
physical supersymmetric mass parameters as computed using RG evolution and thresholds to
all orders (our boundary conditions at MP should also be given within some renormalization
scheme)

m2i phys = MS2 Fi (MS /MP , αa , cj ). (11)
Now we can define Qc as the critical value of MS below which the Higgs mass matrix develops
a negative eigenvalue. With this definition we can go back and follow the same logic from above
Eq. (6). By varying MS the Fi vary, at leading order in αa , according to the 1-loop RG equation
and all our results follow, to that accuracy. Notice that Qc shifts by O(1) when going from
leading to next-to-leading order, but by the above definition ln Qc /MS is a scheme-independent
physical quantity.
Another issue concerns the dependence of our conclusions on the “choice” of priors. We have
used a weak prior corresponding to the rather weak request that SU(2)L × U (1) be dominantly
broken by an elementary scalar field. In anthropic considerations the interesting implications of
a stronger prior, normally called the atomic principle, have also been studied [19]. The atomic
principle is based on the remark that the existence of a rich spectrum of nuclei and atoms, crucial
for the existence of the complex world we live in, severely constrains the ratio of the electroweak
and strong scales vF /ΛQCD . In particular it is rather safe to argue that if vF , keeping every
other coupling fixed, were just a factor 5 bigger than its actual value, then neutrons decay inside
nuclei and no complex chemistry is possible. For vF significatively smaller than its actual value,
it is harder to control exactly what happens but, as we already said above, atomic physics is
drastically modified, and moreover for vF  ΛQCD it seems very difficult to have enough baryons
[20]. While this last bound seems rather robust even when the other parameters are varied, the
upper bound on vF can in principle be lifted by allowing suitable variations of both Yukawa
couplings and cosmological parameters. For instance it has been recently pointed out [22] that
the upper bound on vF can be fully relaxed, thus allowing a so-called weakless universe, once
some cosmological parameters are modified. As we shall see in a moment, under very reasonable
assumptions, our results do not crucially depend on the upper bound on vF , so that the existence
or the absence of this bound do not really affect our conclusions. In what follows we will simply
G.F. Giudice, R. Rattazzi / Nuclear Physics B 757 (2006) 19–46 29

indicate by atomic principle the request that vF  103 ΛQCD while by weak principle we will
indicate the basic request H  = 0.
If the weak principle is taken, like we did so far, as the only relevant prior, then the value of
the Z mass in our patch is expected to be typical of the set of patches where the electroweak
symmetry is broken, and therefore mZ  ∼ 90 GeV. Since we have found that, up to the little-
hierarchy factor, Qc ∼ mZ , and since ΛQCD is fixed to 1 GeV throughout the multiverse, we
conclude that the presence of atoms is a fundamental and generic feature of practically all vacua
breaking electroweak symmetry. Parametrically this corresponds to the fact that the fundamental
ratio Qc /ΛQCD does not scan and it is fixed to the right value in our theory. However, if also
the Atomic Principle is taken as a prior, then, by definition, our local patch is not expected to
be typical among those where electroweak symmetry is broken and the critical scale Qc will no
longer be around the TeV.
Consider the case Qc 1 TeV and write Eq. (7) as m2Z = αMS2 ln(Qc /MS ), where α de-
scribes a one-loop factor depending on coupling constants and dimensionless soft-term ratios.
The weak and atomic principles correspond to the condition 0 < mZ < 103 ΛQCD ≡ ΛW , and
therefore restrict the distribution of vacua
dN ∝ dMSn (12)
only to values of MS in the two domains
D1 : 0 < MS < x1 Qc , D2 : x2 Qc < MS < Qc . (13)
Here x1,2 are the solutions of αx 2 ln x = −Λ2W /Q2c with x1 < x2 . We are considering the case of
small Λ2W /Q2c , where x1 = O(ΛW /Qc ) and x2 is close to unity, x2  1 − Λ2W /(αQ2c ).
The relative number of vacua in the two domains D1 and D2 is
 
dN x1n α ΛW n−2
 D1 =  . (14)
D2 dN 1 − x2n n Qc
Therefore, for n < 2 vacua in D1 dominate, while the region D2 is favored for n > 2. The average
value of the ratio between the weak and supersymmetric scales in region D2 (for n > 2) is
2 
mZ D2 dN α ln Qc /MS Λ2W
≡   . (15)
MS2 D2 D1 ∪D2 dN 2Q2c
In these vacua there is a huge hierarchy between mZ and MS , and the low energy theory is either
the SM or split supersymmetry [23], depending on the masses of higgsinos and gauginos. In
region D1 (for n < 2), we find
2
mZ Qc
2
 α ln , (16)
MS D1 ΛW
and the low-energy theory is just ordinary untuned supersymmetry. We never encounter the situ-
ation where a little hierarchy m2Z /MS2 ∼ α is favored, unless we artificially tune ΛW  Qc .
The strong dependence of our conclusions of the number of priors is not surprising, given that
we only have one aleatory variable MS at hand. So in order to reach a more robust conclusion we
should consider the general situation where also some other parameter is scanned. The atomic
principle depends crucially on ΛQCD and so we will assume this parameter to be scanned in some
range. The scanning of ΛQCD is given by a corresponding scanning of αs up at the Planck scale.
Since ln ΛQCD /MP ∝ −1/αs and since it is reasonable to expect 1/αs to be scanned roughly
30 G.F. Giudice, R. Rattazzi / Nuclear Physics B 757 (2006) 19–46

linearly, we will assume the measure to be ∝ d ln ΛQCD . Of course it is hard to imagine Qc not
to scan when ΛQCD does. Actually, as also Qc is related to dimensional transmutation, we will
assume a measure for the distribution of vacua

dN ∝ dMSn d ln Qc d ln Λ, (17)

where, for convenience, we defined Λ ≡ ΛW / α. The weak and atomic principles restrict the
acceptable vacua in the region

D: 0 < MS < Qc , Qc < Q̄c , MS ln Qc /MS < Λ < Λ̄, (18)
where Q̄c and Λ̄ are the maximum values of Qc and Λ over the landscape. We take Λ̄  Q̄c , to
allow a sufficient scanning range for ΛW .
Since we are mainly interested in the distribution of r ≡ MS /Qc , it is convenient to integrate
Eq. (17) over the other variables in the region D admitted by the weak and atomic principles. We
find


Q̄nc Λ̄ 1 1 1
dN ∝ ln − ln ln + dr n . (19)
n Q̄c r 2 r n
Aside from a mild ln(1/r) dependence, the above measure describes a probability function es-
sentially identical to Eq. (6). Indeed, we find an average little hierarchy
2 
mZ dN α ln Qc /MS αC 2−n
≡ D  = , C =1+ , (20)
MS 2
D dN n 4 + n(γ + ln nΛ̄2 /Q̄2c )
where γ is the Euler constant. Therefore, up to an O(1)-coefficient C, we obtain the same result
as in Eq. (10), corresponding to the case where only MS scans and where only the weak principle
is used. In practice by integrating over a logarithmically distributed ΛQCD we have “integrated
out” the atomic principle. Finally notice that scanning ΛQCD is crucial. If we scanned only Qc ,
we would not reach this conclusion. In that case the distribution of r would end up being

d ln 1r
dN ∝ (21)
(ln 1r )n/2
showing once again that, depending on n large or smaller than 2, either split or untuned super-
symmetry are favored.
This discussion also partially illustrates the result that will be obtained with an independent
scan of all soft terms. If the scan is restricted to a parameter region such that the squared masses
for all scalars are positive at the high-energy scale, then Qc has a maximum, which acts as an
upper bound on MS under the weak principle. Again, our considerations will lead to a little
hierarchy in m2Z /MS2 . The case of an independent scan of the higgsino mass parameter μ is
particularly interesting, and it will be the subject of Section 4.
Our results so far have been based on the standard soft term scenario, where the Higgs squared
mass matrix starts out positive at the high-energy scale and develops a negative eigenvalue while
flowing to lower energy scales. In this scenario we have argued that, when large values of MS are
statistically favored, then the weak principle implies criticality of the physical Higgs mass. Large
values of MS are generically favored when supersymmetry is broken at tree level, as discussed
in Section 6. On the other hand, dynamical supersymmetry breaking is an interesting possibility
which suggest that MS could be logarithmically distributed or even peaked at low values. It is
G.F. Giudice, R. Rattazzi / Nuclear Physics B 757 (2006) 19–46 31

easy to see that also in this second case, where


 
dN = d 1/MSn (22)
with n > 0, the weak principle can lead to criticality. However, this will require the remarkably
unusual situation where M2 has a negative eigenvalue in the ultraviolet and flows to become
positive definite in the infrared. We can again define as Qc the RG scale where this happens.
Then the sign of ln Qc /MS and of dm22 /d ln Q in Eq. (9) will both be reversed with respect to
the previous case. The weak principle and Eq. (22) imply ln Qc /MS  = −1/n meaning that RG
evolution gets frozen just before M2 turns positive. The phenomenology of this scenario has
been recently discussed in Ref. [24].

4. Scanning μ

Among the soft parameters the role of μ is special, as it does not break supersymmetry while
it breaks a global PQ-symmetry which is respected by the other soft parameters. This properties
of μ make its origin often problematic from the model building viewpoint. This is generically the
case in models with gauge or anomaly mediated supersymmetry breaking. On the other hand, in
ordinary tree-level gravity mediation, the size of μ (and of m23 ≡ Bμ) can be naturally associated
to MS [6]. In this section we will present an alternative viewpoint on the μ problem, proposing
a solution based on statistical considerations, obtained by exploring the implications of scanning
μ independently of MS over the landscape. This will on one side illustrate some features of
the general case in which all soft terms are independently scanned, but on the other side it will
remarkably predict a favored range for the size of μ and of tan β.
We consider a slight modification of our previous ansatz at the Planck scale with
μ = μ0 , m23 = bMS μ0 (23)
and all other soft terms given by Eq. (3), taking μ0 and MS to scan independently, while b is
fixed. Writing the running Higgs mass matrix as
 2
m̃1 + μ2 Bμ
M =
2
, (24)
B ∗μ m̃22 + μ2
the condition of criticality is
 
det M2 = m̃21 m̃22 + m̃21 + m̃22 − |B|2 μ2 + μ4 = 0. (25)
The critical line in the μ–MS plane, corresponding to Eq. (25), is shown in Fig. 4, in the
case of scalar√ universality and gaugino unification with m = M = MS , A = 0 and B = 0 (solid
line), B = 2MS (dashed line) at the GUT scale. To understand the shape of these curves,
let us indicate by Q̄c the critical RG scale for μ = 0 and let us start with the case in which
m̃21 + m̃22 − |B|2 > 0 for Q > Q̄c (as for the solid line of Fig. 4). Then, in ordinary RG flows with
d m̃22 /d ln Q > 0, an increase in μ will lower the value of Qc . Let us study the critical line close
to Ms = Q̄c . The small eigenvalue of M2 is
 
λsmall  m̃22 + (1 − rB )μ2 + O μ4 , (26)
where rB = |B|2 /m̃21 . By using the qualitative behaviour 2m̃22 = αMS2 ln MS /Q̄c , the critical
curve is then given by
α
λsmall = 0 ⇒ μ2  M 2 ln Q̄c /MS ≡ f 2 (MS ). (27)
2(1 − rB ) S
32 G.F. Giudice, R. Rattazzi / Nuclear Physics B 757 (2006) 19–46

Fig. 4. The critical line separating the broken and unbroken phases. Here μ is defined at the scale MS and we have fixed
the top Yukawa coupling corresponding to mt = 172.7 GeV for large tan β. Scalar universality and gaugino √ unification is
assumed with boundary conditions at the GUT scale m2 = M 2 = MS2 , A = 0 and B = 0 (solid line), B = 2MS (dashed
line).

This equation implies that in the region MS ∼ Q̄c , which is favored when the number of vacua
grows with MS , there exist a moderate O(α) hierarchy also between μ2 and MS2 . To see how this
works explicitly, let us make the simple assumption on the distribution of vacua

dN ∝ dMSn dμm (28)


with m, n > 0. The weak principle restricts the acceptable vacua to lie in the region

Dμ : 0 < MS < Q̄c , 0 < μ < f (MS ). (29)


Since m2Z  −2λsmall , the average ratio m2Z /MS2 is given by
  2 
Qc
α ln M − 2(1 − rB ) μ 2
m2Z Dμ dN S MS α
  = . (30)
MS2 Dμ dN n + m

This shows that an independent scanning of μ changes Eq. (10) simply by the replacement n →
n + m. The extra suppression is easy to understand, since a positive m pushes μ towards the
critical line, as a positive n pushes MS . The average ratio μ2 /MS2 is instead

μ2 m
=α . (31)
2
MS 4(1 − rB )(n + m)
This shows that higgsinos are lighter than the typical supersymmetric particles by the square root
of a loop factor.
G.F. Giudice, R. Rattazzi / Nuclear Physics B 757 (2006) 19–46 33

Using sin 2β = 2Bμ/(m21 + m22 ), we can express the average value of tan β as

1 MS
tan β  √ . (32)
rB μ

Therefore, tan β ∼ 1/ α ∼ 5–10 is the most natural expectation in this framework. This is wel-
come, because | cos 2β|  1 plays a non-negligible role in pushing the mass of the lightest Higgs
above the LEP bound. The generic prediction with soft parameters of the same order of magni-
tude is tan β = O(1), and it is well known that large tan β can be obtained only with a fine tuning
of order 1/ tan β in Bμ. In our scenario, tan β 1 is just an added bonus of statistical criticality.
So far we have considered the case in which m̃21 + m̃22 − |B|2 is always positive for Q > Q̄c .
If this is not the case, the coefficient of μ2 in Eq. (25) becomes negative before m̃22 , in the RG
evolution starting from MGUT . Therefore we expect that the most probable values of μ/MS are
of order unity, with no loop-factor suppression. This is confirmed by the result shown by the
dashed line in Fig. 4, which corresponds to the case of a large B.
In conclusion, the weak principle and statistical considerations based on an independent scan
of MS and μ offer a solution to the μ problem, since the most probable values of μ turn out to
be close to MS . Moreover, in the case of positive m̃21 + m̃22 − |B|2 for Q > Q̄c (which is actually
the most likely situation for typical soft terms of comparable sizes), our anthropic assumption
gives the testable prediction that both μ/MS and 1/ tan β are of the order of the square of a loop
factor.

5. Dynamical criticality

Our environmental argument to explain (if not post-dict) the little hierarchy of the minimal
supersymmetric SM followed very closely Weinberg’s approach to the cosmological constant
problem. Like for the cosmological constant, we think it is instructive to see why a dynamical
mechanism, where MS is a dynamical rather than aleatory variable, is difficult to be realized.
It is well known that directions which are flat in the supersymmetric limit can dynamically
generate 1-loop hierarchies via the Coleman–Weinberg mechanism. Indeed, in the presence of
soft supersymmetry breaking, the effective potential along a flat direction φ can be written in gen-
eral as V = m2 (φ)φ 2 , where m2 is a running effective mass squared. If, starting from a positive
√ smaller φc , then m (φ) = αMS ln φ/φc
value at some high-energy scale, m2 crosses zero at some 2 2

and the minimum of the potential will be at φ = φc / e, where e is Napier’s constant. At this
minimum, m2 is one-loop suppressed with respect to the typical soft mass scale thus dynamically
realizing a little hierarchy. Barbieri and Strumia [14] have proposed a set up where a Coleman–
Weinberg potential explains the little hierarchy. Following Ref. [25] (see also Ref. [26]), these
authors have considered the situation in which the overall supersymmetric scale MS in Eq. (3)
is itself a scalar field, a modulus, with respect to which the potential should be minimized. The
latter can generally be written as

V (MS , H1 , H2 ) = VMSSM (MS , H1 , H2 ) + VS (MS ), (33)


where VMSSM represents the ordinary potential of the supersymmetric SM, with terms quadratic
and quartic in the Higgs fields and with the running soft mass matrix M2 . If, for some reason,
VS could be neglected, then the minimization of VMSSM would dynamically realize an O(α)
hierarchy between H 2 and MS2 . Indeed, under the same assumptions of Section 3, VMSSM is
strictly positive for MS > Qc . For MS < Qc we can minimize VMSSM with respect to H1,2 thus
34 G.F. Giudice, R. Rattazzi / Nuclear Physics B 757 (2006) 19–46

yielding an effective negative definite potential for MS


2MS4 [c12 (MS )c22 (MS ) − c34 (MS )]2
eff
VMSSM (MS ) = − , (34)
g22 + g12 [c12 (MS ) + c22 (MS )]2 − 4c34 (MS )
where, as before, the Higgs mass parameters are expressed as m2i = ci2 MS2 (i = 1, 2, 3). Here to
simplify the notation we have neglected the top-stop corrections to the Higgs quartic coupling.
Expanding the ci in ln Qc /MS , we find (indicating schematically the powers of loop factors for
later use)


α 4 2 Qc α Qc
eff
VMSSM (MS ) = − M ln 1 + O ln , (35)
(4π)3 S MS 4π MS
which, at leading order in α, is minimized at ln Qc /MS = 1/2. Compared to our Eq. (10), this
result correspond to dynamically predicting n = 2 which, in principle, is a testable relation among
soft parameters at the weak scale. It falls short, as already explained, to fully account for the little
hierarchy, but nonetheless it is a remarkable result.
Unfortunately this result totally rests on our assumption of negligible VS (MS ). Now, VS con-
(1) (2)
sists of two pieces: VS = VS +VS . The first is truly incalculable, as it is quadratically divergent
(1)
with the cutoff VS = Λ2 MS2 . There is no symmetry reason to really control this contribution,
which for Λ ∼ MP becomes, understandably, of the order of the supersymmetry breaking scale
(1)
in the hidden sector MI4 . The presence of VS does not only disrupt our dynamically critical min-
imum, but also implies a minimum for MS which is either 0 or O(Λ). In other words MS is no
longer a flat direction. Without any solid physical motivation one must then assume that by some
(1)
clever short-distance conspiracy VS ≡ 0. Yet this not sufficient, due to the second contribution
  
VS(2) = c0 (MS )MS4 1 + O α(MS ) , (36)
where the O(α) term indicates threshold correction effects at the supersymmetric scale and where
c0 satisfies an RG equation


dc0 1 STr M 4 α
= +O . (37)
d ln Q 64π 2 MS4 4π
Here M is the mass matrix of all particles that become massive through their couplings to
MS including, in particular, the supersymmetric partners of SM fields. The natural size of c0
(2)
is ∼ 1/(4π)2 ln MP /MS , which makes VS parametrically bigger than VMSSM eff in the region
MS ∼ Qc . Then, in order to preserve the minimum of the full potential VMSSM + VS(2) near the
eff
(2)
critical point MS ∼ Qc , also VS should independently have a minimum in this region. As the
(2)
stationary points of VS are determined by dimensional transmutation through the logarithmic
evolution of c0 , this coincidence represents a tuning, which we can roughly estimate to be of
order 1/(ln MP /Qc ). Unfortunately, this is precisely what a dynamical-relaxation model was
designed to avoid. Moreover the presence of VS(2) would destroy the prediction ln Qc /MS = 12 .
Finally one could argue that, although there is no solid field-theoretic reason to neglect VS , per-
haps this could follow from whatever mechanism solves the cosmological constant problem. We
believe it is difficult to imagine how this could work. Indeed from a strict field-theoretic point
eff (2)
of view the only distinction between VMSSM and VS is diagrammatic: the latter is determined
by 1PI diagrams of supersymmetric SM fields, while the former involves also the one-particle
G.F. Giudice, R. Rattazzi / Nuclear Physics B 757 (2006) 19–46 35

reducible diagrams with tree-level Higgs exchange. How can the solution of the cosmological
constant problem distinguish among different contributions to the potential of the same field
MS ? Perhaps the only way to proceed is to see if there are consequences following from the
(1)
vanishing of the potential at the minimum. Neglecting VS , again without any explanation, the
potential V = ceff (MS )MS4 consists of the addition of Eqs. (35) and (36). The coefficient ceff
varies logarithmically with MS , and it is in principle possible to fine tune the parameters so that
vanish simultaneously at some point. This point would correspond to
ceff and its derivative ceff
a minimum with vanishing vacuum energy. It is easy to see that even this criterion in no way
singles out the minimum of Eq. (35).

6. Distribution of supersymmetry-breaking scales

In this section we shall produce an argument on the possible distribution of the supersymme-
try-breaking scale based on simple effective field theory. Our considerations and results are in
line with what was previously found in type IIB Calabi–Yau orientifolds [21] or, in the same
spirit of this section, in effective supergravity [27].
Suppose we have a general supersymmetric theory with N chiral superfields Ψi , and a general
superpotential W (Ψ ) and Kähler potential K(Ψ, Ψ † ). We will assume that both of them include
higher-dimension operators suppressed by some fundamental scale M∗ , that we set to unity in this
discussion. We will also ignore supergravity corrections by assuming that M∗ is parametrically
smaller than MP ; all the vacua that we will find in our analysis below are then smoothly deformed
into vacua of the full theory with supergravity effects included.
Of course, there will be a large number (exponential in N ) of supersymmetric minima as-
sociated with the stationary points of W . However, we also expect to have a large number
of metastable non-supersymmetric minima. It is easy to see that this is only possible due to
higher-order terms in the Kähler potential. Indeed, with a canonical Kähler potential, any non-
supersymmetric extremum is either a saddle point or it is associated with an exactly flat direction.
Consider, for instance, the theory of a single chiral superfield with
a3 3 a 5 5
W = a1 X + X + X + ···. (38)
3 5
For simplicity, to begin with we assume a discrete R symmetry that makes the superpotential
odd in X and makes K a function only of X † X. For generic coefficients ai , supersymmetric
minima do exist. There is also a local maximum of the potential at X = 0 with a quadratic in-
stability along the direction where a1∗ a3 X 2 is real and negative. If we tune a3,5,... = 0, then we
have supersymmetry breaking but also an exactly flat direction for X. This is why an arbitrar-
ily small a3,5,... can restore supersymmetry; it lifts the flat direction and drives the field to the
supersymmetric minimum.
However, the story changes if we have corrections to the Kähler potential
c2  † 2 c4  † 3
K = X† X − X X + X X + ···, (39)
4 9
because the higher-order terms in K can also lift the flat direction and, if c2 is large enough
relative to a3 , stabilize a non-supersymmetric vacuum at X = 0. Indeed, the potential is

|a1 + a3 X 2 + a5 X 4 + · · · |2
V= . (40)
1 − c2 X † X + c4 (X † X)2 + · · ·
36 G.F. Giudice, R. Rattazzi / Nuclear Physics B 757 (2006) 19–46

As long as
2|a3 | < |a1 |c2 , (41)
there is a non-supersymmetric local minimum at X = 0. Note that the condition to find a non-
supersymmetric local minimum does not depend on the values of a5,... and c4,... ; the reason is
that these terms do not contribute to the quadratic curvature around the extremum at X = 0.
Suppose we mediate the supersymmetry breaking to the SM sector via higher-dimensional
operators, so that the overall scale of the soft terms is MS ∼ |FX | ∼ |a1 | (working in units with
M∗ = 1). For c2 ∼ O(1), to get a small MS , we need not only |a1 | ∼ MS but also |a3 |  MS to
be small. If the X sector coupled to a landscape of vacua, so that the complex parameters a1 , a3
scan, it is natural that, when they are small, they scan with a uniform distribution. So, since these
are two complex parameters, the number of vacua with supersymmetry breaking scale smaller
than MS is
N(MS ) ∝ MS4 . (42)
Let us now consider the most general case where X is not charged under any symmetries, so
that W and K are general functions of X. By shifting X, we can always assume without loss of
generality that there is an extremum for X located at X = 0, and we can expand W, K around
this point
 an  cpq
W= Xn , K= X (p+1) X †(q+1) , (43)
n
n p,q
(p + 1)(q + 1)
∗ = c . The potential, expanded at quadratic order in X, is
with c00 = 1 and cpq qp

|∂X W |2  
V= = |a1 |2 + k1 X + k2 X 2 + h.c. + k3 |X|2 + · · · , (44)
∂X ∂X† K
k1 ≡ a1∗ (a2 − a1 c10 ), (45)
k2 ≡ a1∗ (a3 − a1 c20 ) − c10 k1 , (46)
 ∗ 
k3 ≡ |a2 |2 − |a1 |2 c11 − c10 k1 + h.c. . (47)
The conditions to have a stable local minimum at X = 0 with a supersymmetry-breaking scale
equal to MS are
|a1 | = MS , k1 = 0, |k3 | > 2|k2 |. (48)
For cpq = O(1), these conditions require that |a1,2 | = O(MS ) and |a3 |  MS . If the 3 complex
parameters a1,2,3 scan uniformly, then we obtain
N (MS ) ∝ MS6 . (49)
Note though that we can get different powers with different assumptions about the landscape
sector. For instance, if it has a CP symmetry that makes all the parameters real, then we would
have N (MS ) ∝ MS3 , while if it also has the discrete R symmetry of the previous example we
would have N(MS ) ∝ MS2 .
Of course the non-supersymmetric minima we are considering are unstable to decaying to
a supersymmetric vacuum, but the lifetime can be exponentially long. Indeed, we expect the
bounce action to scale like S ∼ ( X)4 / V where X and V are the shifts in field expectation
value and potential energy between the two minima. In our case, V = |a1 |2 and the location of
G.F. Giudice, R. Rattazzi / Nuclear Physics B 757 (2006) 19–46 37

the closest supersymmetric minimum is determined by the quartic term a4 in W and therefore
X ∼ |a1 |1/3 . This gives a lifetime Γ ∼ exp[−(M∗ /MS )2/3 ], which is extremely small as soon
as supersymmetry is broken below M∗ .
We have phrased the discussion as though the X sector is separate from the landscape sec-
tor, but in fact our conclusions apply to supersymmetry breaking on a generic supersymmetric
landscape. It is clear that in order to find supersymmetry-breaking extrema, some fields in the
theory must become light; indeed, there must be a massless goldstino. However with a com-
pletely generic superpotential, we expect that all the fields are heavy with O(1) masses. In some
places in field space, though, it may happen that one field X is light while the remaining fields
φi are heavy. We can then expand W as

1 1
W = W0 (φ) + W1 (φ)X + W2 (φ)X 2 + W3 (φ)X 3 + · · · (50)
2 3
with the remaining fields φ having an exponentially large number of supersymmetric minima. In
these minima, the parameters in the X theory will scan, and again, when these parameters are
small, the scanning can be taken to be flat.
Note that the scanning for the vacuum energy is completely independent of the supersym-
metry breaking scale [28]. In all vacua (supersymmetric and non-supersymmetric), we will in
general have W  ≡ a0 = 0. When gravity is turned on, this gives a negative contribution to the
cosmological constant ΛSUSY = −3|W |2 /MP2 , which is parametrically unrelated to the scale of
supersymmetry breaking. For a uniformly-distributed complex parameter a0 , the scanning mea-
sure is d 2 a0 = |W |d|W | = dΛSUSY , i.e., there is uniform measure on cosmological-constant
space. Therefore, the request of a small cosmological constant does not impose additional con-
straints on the statistical distribution of MS .
As we have already mentioned there will also be an exponentially large number of supersym-
metric stationary points, where all the landscape fields have generically O(1) masses. What role
can these vacua play in our argument? If these landscape vacua remain exactly supersymmetric
even after including the possible infrared dynamics of some hidden gauge group, then they do
not play any role in any selection criteria. Our universe does not appear to be supersymmetric.
It is still possible, however, that at these minima some low-energy group with a supersymmetry-
breaking infrared dynamics survives. It is natural to expect the distribution of MS from these
vacua to be roughly logarithmic dN ∝ d ln MS [29], very much like the case of ΛQCD con-
sidered in Section 3. If the total number of vacua from this branch were large enough, then it
would swamp the distribution of MS coming from the branch of local supersymmetry-breaking
minima we focused on so far. In that case, the total distribution of MS would be essentially
dN ∝ d ln MS , and the weak principle would predict MS  Qc . However the relative weight
in vacuum statistics of this branch with dynamical supersymmetry breaking very strongly de-
pends on microphysics inputs we do not control. On one side, it is not clear how generic it is
that these hidden gauge groups lead to dynamical supersymmetry breaking. Also, there is no
universal rationale to count the supersymmetric versus non-supersymmetric local minima, even
with a simple landscape model with N chiral fields φi (i = 1, . . . , N ). Assume, for instance, the
superpotential is a generic polynomial of degree M + 1 in φi . Then the number of classically
supersymmetric vacua determined by the equation

∂i W = 0 (51)
38 G.F. Giudice, R. Rattazzi / Nuclear Physics B 757 (2006) 19–46

scales like M N . The non-supersymmetric stationary points are determined by the equation

∂i W
∂j = 0. (52)
∂i ∂ k K
It is easy to see that, by suitably choosing the Kähler potential, there can be more solutions to
Eq. (52) than to Eq. (51). For instance, take the case of just one superfield X with superpotential
and Kähler potential given by
 
W = exp(−λX), K = 2X † X − Si X † X , (53)
where λ is a coupling constant and Si(x) is the sine-integral function, such that the Kähler metric
is non-singular and positive definite. This gives a scalar potential

λ2 e−λ(X+X )

V= , (54)
2 − cos(X † X)
which can reach its supersymmetric vacuum only at X → ∞, but has an infinite number of non-
supersymmetric local minima. This result can be generalized to the case of N fields. Therefore,
depending on the properties of the Kähler potential, there may or there may not be more non-
supersymmetric than supersymmetric vacua.
The results of this paper depend on the assumption that the tree-level supersymmetry-breaking
vacua dominate in number over those with dynamical supersymmetry breaking. It is however
remarkable that once this assumption is made the distribution of supersymmetry-breaking vacua
depends on a few universal and basic ingredients. Our conclusion is that, for a “generic” theory
with a large N number of fields, there can be a huge number of non-supersymmetric vacua. In
the neighborhood of any one of these vacua, the breaking of supersymmetry can be characterized
by a single field X getting an F -component, and MS has a distribution

N(MS ) ∝ MSn , (55)


where n can run from 2 to 6 depending on assumptions on the structure of the landscape sector,
with n = 6 the most “generic”.
Note that for all n > 2, we have a huge preference for high-scale supersymmetry breaking;
in fact, the tuning it takes to get low-energy supersymmetry with mZ ∼ MS is much bigger than
the standard hierarchy problem ∼ m2Z . For n = 2, it is about as tuned as the usual hierarchy
problem, although if we manage to argue that MS is a loop factor bigger than mZ , we win in
tuning by a factor (MS /mZ )2 . However, as we argued in this paper, if Qc has a maximum, then
the statistical preference for high scale MS is eliminated by the anthropic prior that electroweak
symmetry be broken (weak principle). The little hierarchy remains as the only detectable signal
of an extremely atypical choice of vacuum, which is dictated by the anthropic prior.
The main result of our paper relies on the assumption of softly-broken supersymmetry. For
instance if supersymmetry were broken at the cutoff scale, our minimal scenario, where only
the overall value of MS scans, would hardly be tenable. In that case, all higher supercovariant
derivative terms in the action would affect the Higgs potential and there could be plenty of other
vacua where H  is not controlled by radiative electroweak breaking. However, if we have such
a preference for breaking supersymmetry at a high scale, what stops us from going all the way up
to the fundamental scale M∗ ? The answer is metastability of the non-supersymmetric vacuum.
As we said, one source of instability is given by tunneling of X to the closest supersymmetric
minimum with the parameters of its potential fixed. However, the local minimum for X requires
G.F. Giudice, R. Rattazzi / Nuclear Physics B 757 (2006) 19–46 39

a special choice for the landscape fields φ, in order to tune the parameters a1,2,3 to small sizes.
So another, potentially more important, source of vacuum decay is given by tunneling in φ space.
It is reasonable to expect that the euclidean action for these processes will also be proportional
to an inverse power of MS . Thus, the total decay rate for the non-supersymmetric minimum is
expected to be

Γ ∼ M∗4 e−(M∗ /MS )


q
(56)
for some q. For large MS , the decay rate is unsuppressed and therefore metastability (corre-
sponding to Γ < H04 , where H0 is the present value of the Hubble constant) puts a cutoff on the
highest MS
M∗
MSmax ∼ . (57)
[ln(M∗ /H0 )]1/q
This tells us that the only non-supersymmetric vacua that are cosmologically stable will indeed
be approximately supersymmetric. In particular, this means that it is at least consistent to imagine
that the only thing that scans is the overall scale of supersymmetry breaking. We do not have to
worry about higher-derivative operators that would effectively make all the ratios of soft terms
to scan, even with only a single source of supersymmetry breaking in X.

7. Phenomenological consequences

The proximity of the critical scale Qc to the supersymmetric mass MS can be empirically
tested at collider experiments. When the new-particle spectrum is known, one will be able to
reconstruct the running of the Higgs mass parameters and observe if the critical condition for
electroweak breaking is immediately achieved. However, even without a complete knowledge
of the supersymmetric spectrum, we can obtain, under certain assumptions on the ratios of soft
terms, some predictions on the Higgs and stop masses.
Let us work in the limit of large tan β (which is the most favourable case with respect to
the Higgs mass bound), where the critical scale is determined by the condition m22 (Qc ) = 0.
Since we expect a little hierarchy between the supersymmetric and the weak scale, in order
to accurately compute the relevant physical quantities, we match to the one Higgs SM at the
supersymmetric scale and take into account the leading RG evolution effects down to the weak
scale. As it is convenient and customary, we choose the geometric average of the physical stop
masses (Mt˜2 = Mt˜1 Mt˜2 ) as the matching scale from which to compute the infrared logarithms.
Notice that we do not need to specify a relation between MS and Mt˜ since, as discussed in
Section 3, MS appears in our equations only through the scheme-independent ratio Qc /MS .
After integrating out all supersymmetric particles and the additional neutral and charged Higgs
bosons, the Higgs sector is described by the familiar SM scalar potential
λ
V = m2 |H |2 + |H |4 . (58)
2
At the scale Mt˜, the Higgs parameters m2 and λ are determined by matching the supersymmetric
theory with the SM:


 2  g12  2   2 
2 3LS
m (Mt˜) = m2 = − λt mt˜ + mt˜ + |At | −
2 2 2 2 2
M1 + μ − g2 M2 + μ
2 2
,
L R 5 8π 2
(59)
40 G.F. Giudice, R. Rattazzi / Nuclear Physics B 757 (2006) 19–46


g 2 + g 2 3λ4t 
λ(Mt˜) = + X 
t , (60)
4 16π 2
Mt˜

where Xt is defined in the appendix and LS = ln Qc /MS . In Eq. (60) we have also included a
term which is formally a one-loop correction, but which can be numerically very important when
the trilinear coupling At is large. Consistently with our hypothesis, we can drop terms suppressed
by inverse powers of tan β or proportional to μ. Indeed, for natural values of the soft parameters
of order MS , the higgsino mass is expected to be of order μ = O(MS / tan β).
Next, we renormalize the parameters in Eq. (58) to the scale of the top mass mt , and also
express the result in terms of the MS top Yukawa ht computed at the top scale

m2 (mt ) = m2 (Mt˜)K1 , (61)


g2 + g 2
λ(mt ) = K2 , (62)
4
3h2 tS
K1 ≡ 1 − t 2 , (63)

3h2t tS
K2 ≡ K12 (1 + δ)  1 − + δ, (64)
4π 2
Mt˜
tS ≡ ln . (65)
mt
Here K1 is just the Higgs wavefunction renormalization due to top loops, while δ is the full RG
improved top-stop additive correction to the Higgs quartic coupling, given in Appendix A. We
have used the SM RG evolution, including only effects from top-Yukawa and strong interactions.
We kept linear terms in tS and Xt , and quadratic terms enhanced by m2t /m2Z .
The minimization of the potential in Eq. (58) allows us to express the Higgs and the Z masses
in terms of m2 and λ: m2h = −2m2 and m2Z = −m2 (g 2 + g 2 )/(2λ), where parameters are evalu-
ated at the scale mt . Using these equations, we can compute the Higgs mass and Mt˜ in terms of
mZ , LS and the ratios among soft terms:

m2h = K2 m2Z , (66)


 2
 M  K2 2
Mt˜2 =  2t˜  mZ . (67)
m2 2K1
Here Mt˜2 /m22 , with m22 given by Eq. (59), obviously depends only on ratios of soft terms.
In Fig. 5 we show the values of the Higgs mass mh and of the lightest stop mass Mt˜1 , ob-
tained from Eqs. (66)–(67) by fixing LS = 1/6 (which, as shown in Section 6, is the most
“generic” landscape prediction) and by varying the ratios of low-energy soft parameters in
the range 1/2 < m2 /m2 < 2, 0.8 < At /mQ̃ < 1, 1/2 < M32 /m2 < 2, 1/10 < M1,2 2 /m2 < 1.
Ũ Q̃ Q̃ Q̃
These ratios are varied independently, and therefore we are making no assumption of scalar
universality or gaugino mass unification. The three regions shown in Fig. 5 correspond to the
top mass equal to its present central value ±2σ , and satisfy the requirement M3 > 200 GeV,
M2 > 100 GeV, M1 > 50 GeV. The restricted range of At /mQ̃ is a natural consequence of
the RG running of soft parameters up to a large scale. Indeed, the gluino mass gives a large
renormalization correction to both parameters, focusing the low-energy value of this ratio, very
much independently of the initial values of the various soft parameters at the high scale. For in-
stance, taking the top Yukawa corresponding to large tan β and running up to the GUT scale, we
G.F. Giudice, R. Rattazzi / Nuclear Physics B 757 (2006) 19–46 41

Fig. 5. The solid lines are the boundaries of the regions of Higgs mass (mh ) and lightest stop mass (Mt˜ ) ob-
1
tained by requiring ln Qc /MS = 1/6 and by scanning the ratios of soft parameters in the range 1/2 < m2 /m2 < 2,
Ũ Q̃
0.8 < At /mQ̃ < 1, 1/2 < M32 /m2 < 2, 1/10 < M1,22 /m2 < 1, under the constraint M > 200 GeV, M > 100 GeV,
3 2
Q̃ Q̃
M1 > 50 GeV, for the three values of mt indicated in the figure. The purple dot-dashed line is the boundary of the
analogous region obtained for 1 < At /mQ̃ < 3 and mt = 172.7 GeV. The dashed line is the present lower bound on a
SM Higgs-boson mass mh > 114.4 GeV. (For interpretation of the references to colour in this figure legend, the reader
is referred to the web version of this article.)

find that 0.79(0.61) < At /mQ̃ < 0.97(1.14) for any initial condition of universal scalar masses
m, of unified gaugino masses M and trilinear couplings A, such that 0 < m2 /M 2 < 1(3) and
|A/M| < 1(3).
The results in Fig. 5 show how, for a small value of ln Qc /MS , the Higgs mass is predicted
to be very close to its experimental lower bound. On one hand, this can justify why searches
for Higgs and supersymmetric particles have failed so far; on the other hand, it shows that
Higgs and supersymmetric particles lie rather close to their experimental limits. With large
values of At , heavier Higgs bosons and lighter stops can be obtained, as illustrated by the re-
gion in Fig. 5 corresponding to a parameter scan in the range 1 < At /mQ̃ < 3 (shown only in
the case mt = 172.7 GeV). However, as shown above, such large values of this ratio are un-
natural from the point of view of the high-energy theory. Very small values of At /mQ̃ would
further lower the prediction for mh , but this also requires a tuning of soft-term boundary con-
ditions at the high scale. The prediction on mh is rather sensitive on the precise value of mt ,
as it is well known and as illustrated in Fig. 5. A fixed value of ln Qc /MS also selects a lim-
ited range of stop masses. Of course, the precise values of the allowed Mt˜1 depend on the
choice of the interval in which the ratios of soft parameters are varied. The prediction shown
in Fig. 5 corresponds to the natural hypothesis that these ratios are not very different from
unity.
42 G.F. Giudice, R. Rattazzi / Nuclear Physics B 757 (2006) 19–46

8. Conclusions

Low-energy supersymmetry still remains the best known candidate to solve the hierarchy
problem, although its natural prediction for new particles with masses around mZ has not been
confirmed by LEP. The resulting necessity to push MS , the scale of supersymmetric particle
masses, almost an order of magnitude above mZ leads to an apparent fine tuning of few per-
cent or worse. Although this is a much smaller problem than the original hierarchy, it is still
worrisome for at least two reasons. First, the absence of fine tuning was, after all, the starting
motivation for low-energy supersymmetry. Second, the necessary post-LEP mild tuning puts in
question the chances of discovery at the LHC. Indeed the naturalness criterion, in spite of its
intrinsic arbitrariness, is necessary to guarantee that supersymmetric particles are accessible to
LHC energies, while the more quantitative requirement of a thermal relic density appropriate for
dark matter is, by itself, not sufficient. Actually, taking into account LEP bounds and WMAP
data, supersymmetric thermal dark matter requires rather uncharacteristic choices of parameters,
raising the issue of a further source of tuning [30].
Different approaches have been proposed in the literature to reduce the amount of tuning or to
explain a little hierarchy between MS and mZ [11–14]. Large trilinear A terms and a low scale
for the original supersymmetry breaking alleviate the problem, but a complete solution may
require a real modification of the minimal supersymmetric SM dynamics at the TeV. Here we
have followed a drastically different approach, appealing to anthropic considerations to predict
the most probable value of MS , the scale of supersymmetric particle masses.
Symmetry principles have been so successful in particle physics that a general consensus has
grown on the idea that nature is described by a final unique theory, completely determined by
symmetry properties, possibly allowing no logically consistent modifications. More recently, this
view has been challenged, as a result of both experimental observations and theoretical specu-
lations. On one side, the evidence for dark energy reopened the question of the cosmological
constant, which has a satisfactory anthropic justification [18], but no successful explanations
based on symmetry or on dynamics. Also, the negative LEP searches for new physics have cre-
ated some conflict in essentially all known models that can naturally explain the weak scale. On
the theoretical side, the formulation of the string landscape [17] together with an inflationary
picture has given a more solid justification for a multiverse description, where some of the prop-
erties of our universe are determined by environmental selection. If true, this description would
represent the ultimate Copernican revolution, since neither the earth nor our observed universe
have a central and unique role in nature.
From a scientific point of view, the great limitation of anthropic considerations, as opposed
to speculations based on symmetry or dynamics, is the dearth of testable predictions. This is es-
pecially true when we cannot directly probe the properties of the statistical ensemble on which
the anthropic principle is applied, as it is the case of the multiverse picture. Still, it is false that
no physical consequences can be obtained. Predictions can be obtained, although they are differ-
ent in nature from those derived by dynamics and can usually be expressed only in probabilistic
terms. A celebrated example is the expectation that the cosmological constant is of the order of
the critical density of our universe [18]. Another use of the anthropic principle is a change of per-
spective (as, e.g., in split supersymmetry [23]) where arguments based on symmetry properties
(e.g., the hierarchy problem) are abandoned in favor of mere observational facts. In this paper
we have offered a new example of an application of the anthropic principle to particle physics
that can lead to testable predictions and we have derived, under certain assumptions, the most
probable values of supersymmetric particle masses.
G.F. Giudice, R. Rattazzi / Nuclear Physics B 757 (2006) 19–46 43

First of all, we have recast the hierarchy problem in terms of a criticality condition. Then,
assuming a distribution of vacua where MS changes and imposing the anthropic request that
electroweak symmetry must be broken by the Higgs field, we have obtained that MS is pushed
close to Qc , justifying with a statistical argument the quasi-criticality of low-energy supersym-
metry. In this way we have derived a little hierarchy between mZ and MS , a posteriori explaining
why LEP has not discovered supersymmetry, while maintaining the prediction of discovery at
the LHC. We have also discussed how our conclusions change as we modify the anthropic priors
or the number of scanning parameters.
An interesting conclusion is found when the higgsino mass μ is allowed to scan independently
of MS . The anthropic argument shows that values of μ of the order of MS are preferred, giving
a statistical (rather than dynamical) explanation for the approximate coincidence between the
higgsino and gaugino masses. Actually, for moderate values of B, we predict that higgsinos are
somewhat lighter and that tan β is moderately large. Once again, we recall that all predictions
based on the anthropic principle refer to probability distributions. Indeed, we have found that,
for the considered observables, the variance is of the order of the average (X 2  ∼ X2 for
an observable X) and therefore large statistical fluctuations are possible. In other words, our
predictions suffer from a “cosmic variance” problem since, unfortunately, we can measure only
the properties of a single universe, which is actually part of a large statistical ensemble.
Supersymmetry plays a crucial role in the mechanism we have presented, because it provides
a dynamical explanation for the separation of scales between Qc and MP . Here, like in ordinary
low-energy supersymmetry, we take advantage of this natural hierarchy, but we are not trying
to derive the absolute value of MS . However, for a fixed value of the weak scale, we obtain a
statistical distribution of the relative location of supersymmetry breaking, i.e., of MS /mZ , favor-
ing a little hierarchy. Notice that in this respect, our mechanism could be applied to any theory
with radiative electroweak breaking that predicts a separation between Qc and the fundamental
high-energy scale. Our approach is less radical than split supersymmetry, where even the large
hierarchy is attributed to anthropic considerations. However, while in split supersymmetry there
is no justification for the proximity of the dark-matter particle mass to the weak scale, here we
retain a dynamical explanation of this coincidence.
Our result essentially follows from the observation that electroweak breaking implies a max-
imum value of the supersymmetry-breaking scale, MS < Qc . On the other hand, the vacuum
statistics prefer to break supersymmetry at the highest possible scale. Therefore, the combination
of the two effects stabilizes MS very near the critical value. In other words, electroweak breaking
is a rare phenomenon within the landscape and therefore, once we impose the prior H  = 0, the
most likely situation is that SU(2) × U (1) is only barely broken, and supersymmetry has to live
dangerously close to the critical line of unbroken symmetry.

Acknowledgements

We thank Nima Arkani-Hamed for collaboration throughout this project and for numerous
suggestions. We also thank Savas Dimopoulos, Michael Douglas, and Andrea Romanino for
discussions.

Appendix A

Here we derive some of the equations for the Higgs parameters used in this paper. First con-
sider the Higgs potential improved by the addition of the leading top-stop correction δ to the
44 G.F. Giudice, R. Rattazzi / Nuclear Physics B 757 (2006) 19–46

quartic coupling

g2  † 2
V = m21 |H1 |2 + m22 |H2 |2 − m23 (H1 H2 + h.c.) + H1 σ H1 + H2† σ H2
8
g 2  2 (g 2 + g 2 )
+ |H1 |2 − |H2 |2 + δ|H2 |4 , (A.1)
8 8


3h4t Xt tS  2 
δ= tS + + 3ht − 16gs (Xt + 2tS ) ,
2
(A.2)
(g 2 + g 2 )π 2 sin4 β 4 32π 2


2(At − μ/ tan β)2 (At − μ/ tan β)2
Xt ≡ 1− , Mt˜2 ≡ Mt˜1 Mt˜2 , (A.3)
Mt˜2 12Mt˜2
Mt˜
tS ≡ ln . (A.4)
mt
Here ht is the MS top Yukawa at Q = mt in the SM effective theory.
In the presence of a hierarchy between mZ and the pseudoscalar Higgs mass mA , the eigen-
values of the mass matrix M2 defined by Eq. (A.1) are
  
m2lar = m21 + m22 1 + O(Z ) , (A.5)
m21 m22 − m43  
m2sma = 1 + O(Z ) , (A.6)
m21 + m22

where Z ≡ m2sma /m2lar ∼ m2Z /m2A will be our expansion parameter. It is convenient to diagonal-
ize M2 by redefining
  
H1 cos β − sin β H
= . (A.7)
H2 sin β cos β H
By integrating out H in Eq. (A.1) we find the effective potential for H

m21 m22 − m43 g2 + g 2  2 


Veff = |H |2 + cos 2β + δ sin4 β |H |4 + O(Z ), (A.8)
m21 + m22 8
where, given that H /H  = O(Z ), the leading result simply amounts to setting H = 0. No-
tice also that tan β ≡ H2 /H1  is equal to tan β at leading order in Z . By minimizing Eq. (A.8)
and expanding M2 around its zero at leading order in LS one obtains Eq. (7).
We will now instead study the Higgs spectrum for arbitrary mA , but, again, including the
leading top-stop correction. This corresponds to finding the mass eigenvalues and mixing angles
from the Higgs potential in Eq. (A.1). Defining
Δ
Δ = m2Z sin2 βδ, m̂22 = m22 + , (A.9)
2
the result is
2(m21 − m̂22 tan2 β) 2m23
m2Z = , sin 2β = , (A.10)
tan2 β − 1 m2A
m2A = m21 + m̂22 , mH + = m2A + m2W , (A.11)
G.F. Giudice, R. Rattazzi / Nuclear Physics B 757 (2006) 19–46 45


1 2    2  2 
m2h,H = m + m2Z + Δ ± cos 2β m2A − m2Z + Δ + sin2 2β m2A + m2Z .
2 A
(A.12)
In the limit → ∞ the lightest Higgs mass becomes
m2A
 
m2h = m2Z cos2 2β + δ sin4 β , (A.13)
in agreement with Eq. (A.8). Notice that only the CP-even Higgs masses are formally affected
by the presence of Δ. The tree level relation m2h < min(m2A , m2Z ) now becomes
  δ
m2h < min m2A , m2Z + Δ for m2A > m2Z , (A.14)
4+δ
δ
m2A < m2h < m2Z + Δ for m2A < m2Z . (A.15)
4+δ
Finally, the square of the ZZh coupling λ2ZZh is suppressed with respect to the SM value by a
factor


1 m2 − m2Z cos 4β + Δ cos 2β
sin2 (β − α) = 1 + A , (A.16)
2 m2H − m2h
while the ZZH and ZAh couplings are proportional to cos(β − α) like in the minimal super-
symmetric SM. As can be seen from the above equations, in order to have a significant reduction
in sin2 (β − α) one needs m2A < m2Z + Δ as well as tan β 1. However in this region the ZAh
coupling is sizeable and moreover, for small Δ, the threshold for Ah production would be sig-
nificantly below the maximal LEP2 energy. The experimental bound from Ah production then
requires a non-negligible Δ for this region to be viable. This is the reason why the allowed area
in Fig. 3 does not extend far away from the critical line: a sizeable top-stop contribution to Δ is
needed.

References

[1] E. Witten, Nucl. Phys. B 188 (1981) 513.


[2] L.E. Ibanez, G.G. Ross, Phys. Lett. B 110 (1982) 215;
L. Alvarez-Gaume, J. Polchinski, M.B. Wise, Nucl. Phys. B 221 (1983) 495.
[3] R. Barbieri, G.F. Giudice, Nucl. Phys. B 306 (1988) 63;
G.W. Anderson, D.J. Castano, Phys. Lett. B 347 (1995) 300, hep-ph/9409419.
[4] Y. Okada, M. Yamaguchi, T. Yanagida, Prog. Theor. Phys. 85 (1991) 1;
J.R. Ellis, G. Ridolfi, F. Zwirner, Phys. Lett. B 257 (1991) 83.
[5] J.L. Feng, K.T. Matchev, T. Moroi, Phys. Rev. D 61 (2000) 075005, hep-ph/9909334.
[6] G.F. Giudice, A. Masiero, Phys. Lett. B 206 (1988) 480.
[7] LEP SUSY Working Group, note LEPSUSYWG/01-03 (2001).
[8] LEP Higgs Working Group, note LHWG/05-01 (2005).
[9] B. de Carlos, J.A. Casas, Phys. Lett. B 309 (1993) 320, hep-ph/9303291;
S. Dimopoulos, G.F. Giudice, Phys. Lett. B 357 (1995) 573, hep-ph/9507282;
L. Giusti, A. Romanino, A. Strumia, Nucl. Phys. B 550 (1999) 3, hep-ph/9811386;
J.A. Casas, J.R. Espinosa, I. Hidalgo, JHEP 0411 (2004) 057, hep-ph/0410298.
[10] LEP Higgs Working Group, R. Barate, et al., Phys. Lett. B 565 (2003) 61, hep-ex/0306033.
[11] Z. Berezhiani, P.H. Chankowski, A. Falkowski, S. Pokorski, Phys. Rev. Lett. 96 (2006) 031801, hep-ph/0509311;
T. Roy, M. Schmaltz, JHEP 0601 (2006) 149, hep-ph/0509357;
C. Csaki, G. Marandella, Y. Shirman, A. Strumia, Phys. Rev. D 73 (2006) 035006, hep-ph/0510294.
[12] P. Batra, A. Delgado, D.E. Kaplan, T.M.P. Tait, JHEP 0402 (2004) 043, hep-ph/0309149;
J.A. Casas, J.R. Espinosa, I. Hidalgo, JHEP 0401 (2004) 008, hep-ph/0310137.
46 G.F. Giudice, R. Rattazzi / Nuclear Physics B 757 (2006) 19–46

[13] K. Choi, A. Falkowski, H.P. Nilles, M. Olechowski, Nucl. Phys. B 718 (2005) 113, hep-th/0503216;
K. Choi, K.S. Jeong, T. Kobayashi, K.I. Okumura, Phys. Lett. B 633 (2006) 355, hep-ph/0508029;
R. Kitano, Y. Nomura, Phys. Lett. B 631 (2005) 58, hep-ph/0509039.
[14] R. Barbieri, A. Strumia, Phys. Lett. B 490 (2000) 247, hep-ph/0005203.
[15] P.C. Schuster, N. Toro, hep-ph/0512189.
[16] A. Pierce, J. Thaler, hep-ph/0604192.
[17] R. Bousso, J. Polchinski, JHEP 0006 (2000) 006, hep-th/0004134;
S.B. Giddings, S. Kachru, J. Polchinski, Phys. Rev. D 66 (2002) 106006, hep-th/0105097;
A. Maloney, E. Silverstein, A. Strominger, hep-th/0205316;
S. Kachru, R. Kallosh, A. Linde, S.P. Trivedi, Phys. Rev. D 68 (2003) 046005, hep-th/0301240;
L. Susskind, hep-th/0302219.
[18] S. Weinberg, Phys. Rev. Lett. 59 (1987) 2607;
A. Vilenkin, Phys. Rev. Lett. 74 (1995) 846, gr-qc/9406010.
[19] V. Agrawal, S.M. Barr, J.F. Donoghue, D. Seckel, Phys. Rev. D 57 (1998) 5480, hep-ph/9707380;
T.E. Jeltema, M. Sher, Phys. Rev. D 61 (2000) 017301, hep-ph/9905494;
H. Oberhummer, A. Csoto, H. Schlattl, Nucl. Phys. A 689 (2001) 269, nucl-th/0009046;
C.J. Hogan, astro-ph/0602104.
[20] N. Arkani-Hamed, S. Dimopoulos, S. Kachru, hep-th/0501082.
[21] S. Ashok, M.R. Douglas, JHEP 0401 (2004) 060, hep-th/0307049;
F. Denef, M.R. Douglas, JHEP 0405 (2004) 072, hep-th/0404116.
[22] R. Harnik, G.D. Kribs, G. Perez, hep-ph/0604027.
[23] N. Arkani-Hamed, S. Dimopoulos, JHEP 0506 (2005) 073, hep-th/0405159;
G.F. Giudice, A. Romanino, Nucl. Phys. B 699 (2004) 65, hep-ph/0406088;
G.F. Giudice, A. Romanino, Nucl. Phys. B 706 (2005) 65, Erratum;
N. Arkani-Hamed, S. Dimopoulos, G.F. Giudice, A. Romanino, Nucl. Phys. B 709 (2005) 3, hep-ph/0409232.
[24] J.L. Feng, A. Rajaraman, B.T. Smith, hep-ph/0512172;
R. Dermisek, H.D. Kim, hep-ph/0601036.
[25] E. Cremmer, S. Ferrara, C. Kounnas, D.V. Nanopoulos, Phys. Lett. B 133 (1983) 61;
J.R. Ellis, A.B. Lahanas, D.V. Nanopoulos, K. Tamvakis, Phys. Lett. B 134 (1984) 429.
[26] C. Kounnas, F. Zwirner, I. Pavel, Phys. Lett. B 335 (1994) 403.
[27] M. Dine, D. O’Neil, Z. Sun, JHEP 0507 (2005) 014, hep-th/0501214.
[28] L. Susskind, hep-th/0405189;
M.R. Douglas, hep-th/0405279.
[29] M. Dine, E. Gorbatov, S.D. Thomas, hep-th/0407043.
[30] N. Arkani-Hamed, A. Delgado, G.F. Giudice, Nucl. Phys. B 741 (2006) 108, hep-ph/0601041.
Nuclear Physics B 757 (2006) 47–78

MSGUT: From bloom to doom


Charanjit Singh Aulakh ∗ , Sumit Kumar Garg
Department of Physics, Panjab University, Chandigarh 160014, India
Received 4 July 2006; accepted 19 July 2006
Available online 12 September 2006

Abstract
By a systematic survey of the parameter space we confirm our surmise [C.S. Aulakh, MSGUTs from
germ to bloom: Towards falsifiability and beyond, hep-ph/0506291] that the Minimal Supersymmetric GUT
(MSGUT) based on the 210 ⊕ 126 ⊕ 126 ⊕ 10 Higgs system is incompatible with the generic Type I and
Type II seesaw mechanisms. The incompatibility of the Type II seesaw mechanism with this MSGUT is
due to its generic extreme sub-dominance with respect to the Type I contribution. The Type I mechanism
although dominant over Type II is itself unable to provide neutrino masses larger than ∼ 10−3 eV anywhere
in the parameter space. Our Renormalization Group based analysis shows the origin of these difficulties to
lie in a conflict between baryon stability and neutrino oscillation. The MSGUT completed with a 120-plet
Higgs is the natural next to minimal candidate. We propose a scenario where the 120-plet collaborates with
the 10-plet to fit the charged fermion masses. The freed 120-plet couplings can then give sub-dominant
contributions to charged fermion masses and enhance the Type I seesaw masses sufficiently to provide a
viable seesaw mechanism. We give formulae required to verify this scenario.
© 2006 Elsevier B.V. All rights reserved.

1. Introduction

SO(10) GUTs accommodate complete fermion families together with the superheavy right
handed neutrinos (required by the seesaw mechanism [2,3]) in a 16-plet spinorial irreps. They
also provide very natural Higgs multiplets (126) (or 16H × 16H ) to generate the (U (1)B−L break-
ing) right handed neutrino Majorana masses and the (SU(2)L × U (1) breaking) triplet vevs re-
quired to implement the Type I [2] and Type II [3] mechanisms respectively. Thus models which
incorporate high scale breaking of the SU(2)R × U (1)B−L components of the SO(10) gauge

* Corresponding author.
E-mail address: aulakh@pu.ac.in (C.S. Aulakh).

0550-3213/$ – see front matter © 2006 Elsevier B.V. All rights reserved.
doi:10.1016/j.nuclphysb.2006.07.030
48 C.S. Aulakh, S.K. Garg / Nuclear Physics B 757 (2006) 47–78

symmetry provide an elegant and natural context to understand the clear indication of a seesaw
connection between neutrino mass and GUT mass scales provided by the discovery of neutrino
masses in the 50 milli-eV range (Δm2ν ∼ 2.2 × 10−3 eV2 ∼ ((102 GeV)2 /(1015 GeV))2 ). In su-
persymmetric models with left right symmetric (SU(3) × SU(2)L × SU(2)R × U (1)B−L ⊂ G)
gauge groups, R-parity (Rp = (−)3(B−L)+2S ) [4] is a part of the Gauge symmetry and is nat-
urally preserved [5–10] till low energies in models using only B–L even vevs for the seesaw.
They thus predict a stable LSP which is welcome as cosmological dark matter. With such a com-
plement of virtues Supersymmetric SO(10) GUTs of some variety or the other are considered
leading contenders by a majority of workers. The Supersymmetric SO(10) GUT based on the
120 ⊕ 126 ⊕ 210 Higgs multiplets proposed long ago [11,12] shares all these manifest virtues
of LR supersymmetric models. It has the additional virtue of maximal simplicity from a pa-
rameter counting and representation economy point of view [10–13]. It thus lays claim to the
role of contributing the correct GUT symmetry breaking sector (AM Higgs) for a fully realistic
Minimal Supersymmetric GUT (MSGUT). The other crucial component required to fully de-
fine a MSGUT is however the complement of fermion mass (FM) Higgs used to fit the charged
fermion masses and mixings and the neutrino oscillation mass and mixing parameters which con-
stitute the actual ‘data’ that a GUT must confront. SO(10) permits an FM Higgs system based
on 10 ⊕ 120 ⊕ 126 irreps. The first and third of these are have Yukawa couplings to the mat-
ter 16-plets which are symmetric under interchange of family indices and have been extensively
considered. The 120-plet with its antisymmetric Yukawa couplings has so far been conceded
much less importance, acting either as a perturbation to the main structure determined by the 10
and 126 plets [17,24] or else somewhat cursorily [14,16,18,19]. In this paper we shall develop a
line of argument that culminates in a suggestion that a central role for the 120-plet is natural.
In 1992, with LEP data in hand, Babu and Mohapatra [13] proposed that if only the 10 ⊕ 126
irreps were used to fit the charged fermion masses and mixing then the matter fermion Yukawa
couplings of the GUT would be completely determined. Hence the model would become pre-
dictive in the neutrino sector. The compatibility of this scenario with the (now better known)
neutrino mass-mixing data has been the subject of an accelerating succession of works since
their proposal [15,20–22,24,25]. These works have demonstrated the compatibility of the data
with the generic (Georgi–Jarlskog [26] plus Type I and Type II seesaw) structural form of the
fermion mass formulae dictated by SO(10) clebsches and restricted FM Higgs content. Sym-
metric Yukawa couplings alone are quite successful but the possibility of improving the fits,
particularly as regards the CKM CP phase, by using a subdominant contribution of the 120-plet
has also been demonstrated [17,24]. However the generic fitting procedure does not fix the over
all mass scale of the neutrino masses or the relative strength of the Type I and Type II masses.
On the other hand these quantities are fixed in a fully specified GUT model and it is thus cru-
cial to investigate whether any given GUT model of this general type is compatible with both
the generic fits and the actual neutrino mass data. In the 1990’s the construction of natural and
fully consistent minimal left right supersymmetric models (MSLRMs) was accomplished [5–7].
Moreover these analyses showed clearly [9,28] what was already noted at the beginning [11,27]
of the study of multi-scale Susy GUTs: that in LR Susy GUTs there are light multiplets which
violate the conventional wisdom of the “survival principle”. Thus it was clear [9,28] that Susy
GUTs required the use of calculated rather than merely (“survival principle”) estimated masses
for RG analyses. The RG analysis carried out in this work already indicated that the use of cal-
culated spectra forces together the various possible intermediate scales into a narrow range close
to the GUT scale resulting in an effective “SU(5) conspiracy” i.e. the necessity for single step
breaking of SO(10).
C.S. Aulakh, S.K. Garg / Nuclear Physics B 757 (2006) 47–78 49

A calculation of the full GUT spectrum and couplings in various SO(10) GUT models
required complete knowledge of the “Clebsches” relating SO(10) and GSM × U (1)R group
labels on SO(10) fields—at least for tensorial irreps—before progress in the Renormalization
Group (RG) analysis could be made. Furthermore the compatibility of fermion fits with GUTs
could also be investigated only if Clebsches were also available for the spinorial (16-plet) ir-
rep. The requisite computations (based on decomposition of SO(10) labels into those of the
SU(4) × SU(2)L × SU (2)R (“Pati–Salam”) sub-group) were presented by us in [30]. Using
these Clebsches first the mass matrix of the MSSM type doublets ([1, 2, ±1]) and classic GUT
ΔB = 0 process mediating triplets [3, 1, ± 23 ] [30] and then a complete calculation [31], of all
the couplings and mass matrices of the MSGUT was given by us. With similar motivations two
calculations, one in parallel and cross checking with ours [32], and another [33] quite separate,
which both used the same somewhat abstract [34] method (but different phase conventions) to
calculate the “Clebsches” of tensor (but not spinor) multiplets, appeared. In [30,31] we provided
the complete (chiral and gauge) spectra, neutrino mass matrices, gauge and chiral couplings and
the effective d = 5 operators for Baryon violation in terms of GUT parameters and MSSM fields.
Thus the stage was laid both for a completely explicit RG based analysis of the MSGUT and a
completely specified investigation of the compatibility of Type I and Type II seesaw mechanisms.
The calculation of threshold effects based on these spectra was described in [1,31] and shown to
controvert the conventional wisdom [35] which doubted the stability of the successful unification
(and mt , sin2 θW prediction [36]) of the one-loop MSSM coupling evolution [36–38].
The question of the fitting of the charged fermion masses, the relative size of the neutrino
mass splittings and the quark and lepton mixing matrices had been extensively analyzed [13,14,
20–22,24,25] using the generic formulae valid in SO(10) GUTs with only [13] 10, 126 FM Higgs
representations. However this procedure does not constrain the relative size of the Type I and
Type II contributions or the overall scale of the neutrino masses. Within a fully specified theory
like the MSGUT, however, both these parameters are specified. In [23] the question of assuring
Type II (over Type I) dominance on the basis of the Yukawa values found in the successful generic
Type II fits [21,22] was considered. The authors concluded on the basis of order of magnitude
estimates that the Type I mass would emerge as too large unless the B–L breaking scale was
raised. Furthermore they found that the mass of the SU(2)L triplet field whose tadpole vev gives
rise to the Type II seesaw must be lowered to enhance Type II seesaw masses. To ensure these
they considered extending the model by modifying the GUT scale breaking to be via SU(5) (this
required introducing a 54-plet) or else by perturbing it by introducing a 120-plet.
Our point of view, program and results on these questions are related but still quite distinct
from the estimates and scenarios of [23]. In [1] we took up the question of the compatibility
of the MSGUT with the successful generic 10 + 126 FM Higgs fits. We surveyed the variation
of the relative strength of Type I and Type II seesaw mechanisms, as well as their absolute
magnitudes while keeping in view the viability of the relevant regions of the GUT parameter
space i.e. with the stability of the basic unification scenario. Our graphical survey yielded the
striking observations that:

• For generic as well as special values of MSGUT couplings and FM Yukawas taken from
the generic Type II seesaw fit, the Type II seesaw itself implies that it is highly subdominant to
Type I seesaw in the MSGUT unless the relative strength of Type I versus Type II is adjusted to
a very small value. The primary reason for this relative dominance of Type I is not the value of
the symmetry breaking scale but rather the value of the effective coupling in the Type I seesaw
50 C.S. Aulakh, S.K. Garg / Nuclear Physics B 757 (2006) 47–78

formula implied by the Yukawas found in the Type II generic fit (which simply assumes Type II
is dominant);
• The maximal values of the Type I seesaw masses attainable in the MSGUT are at least one
order of magnitude shorter of those required by atmospheric neutrino oscillations.

Note that on both accounts our conclusions are at variance with those of [23] although they
also find difficulties in ensuring Type II dominance over Type I.
In the present paper we present a detailed survey of the parameter space to definitely con-
firm the above conclusions and thus pose a stringent challenge to the viability of the orig-
inal proposal of Babu and Mohapatra [13] in the context of the MSGUT. Already in [31]
we noted that the variation of the unification stability monitoring parameters (USMPs) (i.e
−1
Δ(Log MX ), Δ(Sin2 θW ), Δ(αG )) with the fast control parameter ξ of the MSGUT exhibited
very striking sharp peaks and dips (with the characteristic appearance of poles of a function of
ξ ). In many cases these spikes were at the values of points of increased symmetry uncovered by
the analytic solution of the AM spontaneous symmetry breaking problem in the MSGUT [10,
31,32]. In addition there were other spikes (not obviously related to points of extended symme-
try) whose investigation was called for by the “tomogram” of the MSGUT parameter space that
we had provided [31]. In [10,31,32] an analytic parametrization of the SSB in the MSGUT in
terms of the solutions of a cubic equation (linear in a fast control parameter ξ ) was achieved.
In particular in [10,32,43] the parameter ξ was eliminated—using the cubic equation—for the
solution values x and thus a very direct, elegant and unified parametrization of the pole and zero
structures characterizing the MSGUT AM-SSB was achieved in terms of the variation of a sin-
gle complex variable x. This allowed the identification of additional special points. In particular,
motivated by our observations [1], of a set of points associated with possible growth of the Type I
and Type II seesaw contributions [10,32,43]. Such growth had already been previously observed
at some points in [1] but found to give unviable USMPs. Thus to complete the program of [1]
we have now investigated the behaviour of the Type I and Type II coefficient functions defined
by the precise Clebsches and fermion mass matrices of the MSGUT [1,31] over the complex x
(equivalently to ξ ) plane: including all the special points described above. The loopholes offered
[1,43] by the special points are illusory rather than real: typically some or all of the USMPs ex-
plode due to the very growth which is being invoked to strengthen one or the other seesaw. Our
conclusions are thus unchanged from those of [1]: the MSGUT has suffered a failure of its most
critical features and is thus now defunct.
A way out of this impasse appears by allowing in the 120-plet which was earlier discriminated
against on grounds of convenience and simplicity alone. ‘Lamppost logic’ may have failed here
as so often before. Nature does not heed the wishful thinking behind such logic: the morning
often reveals that the keys sought for all night lay in the gutter just beyond the lamppost! Another
alternative is that SSB in the MSGUT may actually be controlled [39,40] by the explosive growth
of the gauge coupling above MX due to gaugino condensation in the coset SO(10)/G123 which
drives [40] chiral condensates of the AM Higgs fields whose value may be calculable due to the
power of supersymmetric holomorphy. In such a scenario the crucial composition of the zero
mode Higgs doublets must be evaluated without using the “cubic” solution mentioned above.
In Section 2, referring the reader to the original papers [1,30,31] for all derivations, we present
the basic formulae [1,31] governing the Type I and II seesaw mechanisms in the MSGUT and
identify the seesaw monitoring parameters (SMPs) (R, FI , FII ) that we will use for our graphi-
cal presentation of the seesaw relevant topography of the parameter space. We list the obvious
points of enhanced gauge symmetry. This is accompanied by an Appendix A where these for-
C.S. Aulakh, S.K. Garg / Nuclear Physics B 757 (2006) 47–78 51

mulae are also evaluated in the (rational) function (of x) form advocated by [43]. This allows the
identification of an additional list of exceptional points requiring special treatment. In Section 3
we describe the analytic and graphical investigation of the behaviour of the USMPs and SMPs
as we scan over the complex x plane for generic and for exceptionally favourable values of the
‘slow’ AM Higgs parameters [31] i.e. x (equivalently ξ ). By covering the behaviour at the en-
hanced symmetry points and exceptional seesaw points (collectively called ESPs) and scanning
the whole x-plane we eliminate all escape routes and pin down the MSGUT to its doom. In Sec-
tion 4 we propose that the inclusion of the 120-plet in a very specific role can solve the problem
with the small neutrino masses found for the MSGUT and give the fermion mass formulae for
the NMSGUT along with the doublet Higgs mass matrix H. This specifies the composition of
the massless doublets once det H = 0 is imposed. Thus the stage is set for the accomplishment
of a program analogous to that accomplished for the MSGUT [1,10–12,19,30–33,42,44], which
can be implemented without ambiguity. We conclude with a discussion of the outlook for future
work.

2. Seesaw formulae and SMPs for the MSGUT

2.1. MSGUT couplings, vevs and masses

The MSGUT is the renormalizable globally supersymmetric SO(10) GUT whose Higgs
chiral supermultiplets consist of Adjoint Multiplet (AM) type totally antisymmetric tensors:
210(Φij kl ), 126(μ̄Σ̄ij klm ), 126(μ̄Σij klm ) (i, j = 1, . . . , 10) which break the GUT symmetry to
the MSSM, together with Fermion mass (FM) Higgs 10-plet (Hi ). The 126 plays a dual or AM-
FM role since it also enables the generation of realistic charged fermion and neutrino masses and
mixings (via the Type I and/or Type II seesaw mechanisms); three 16-plets μ̄ΨA (A = 1, 2, 3)
contain the matter including the three conjugate neutrinos (ν̄LA ).
The superpotential (see [10,30–32] for comprehensive details) contains the mass parameters

m: 2102 , M: 126 · 126, MH : 102 (1)


and trilinear couplings

λ: 2103 , η: 210 · 126 · 126, γ ⊕ γ̄ : 10 · 210 · (126 ⊕ 126). (2)


The GUT scale vevs that break the gauge symmetry down to the SM symmetry (in the notation
of [30]) are [11,12]
  a
(15, 1, 1) 210 : φabcd  = abcdef ef , (3)
  2
(15, 1, 3) 210 : φabα̃ β̃  = ωab α̃ β̃ , (4)
 
(1, 1, 1) 210 : φα̃ β̃ γ̃ δ̃  = pα̃ β̃ γ̃ δ̃ , (5)
 
(10, 1, 3) 126 : Σ̄1̂3̂5̂8̂0̂  = σ̄ , (6)
 
(10, 1, 3) 126 : Σ2̂4̂6̂7̂9̂  = σ. (7)
The vanishing of the D-terms of the SO(10) gauge sector potential imposes only the condition
|σ | = |σ̄ |. Except for the simpler cases corresponding to enhanced unbroken gauge symmetry
(SU(5) × U (1), SU(5), G3,2,2,B−L , G3,2,R,B−L etc.) [10,32], this system of equations is essen-
tially cubic and can be reduced to the single equation [10] for a variable x = −λω/m, in terms
52 C.S. Aulakh, S.K. Garg / Nuclear Physics B 757 (2006) 47–78

of which the vevs a, ω, p, σ, σ̄ are specified:

8x 3 − 15x 2 + 14x − 3 = −ξ(1 − x)2 , (8)


where ξ = λM
ηm . Then the dimensionless vevs in units of (m/λ) are ω̃ = −x [10] and

(x 2 + 2x − 1) x(5x 2 − 1) 2 λx(1 − 3x)(1 + x 2 )


ã = , p̃ = , σ̃ σ̄˜ = . (9)
(1 − x) (1 − x)2 η (1 − x)2
This exhibits the crucial importance of the parameters ξ, x. Note that one can trade [10,32,43] the
parameter ξ for x with advantage (using Eq. (8)) since ξ is uniquely fixed given x. By a survey
of the behaviour of the theory as a function of the complex parameter x we are simultaneously
covering the behaviour of the three different solutions possible for each complex value of ξ .
We thus change our graphical presentation in terms of three plots versus ξ [1,31] to a single
plot versus x for each USMP or SMP. Moreover, as emphasized by [43] analysis of the spikes
observed in [1,31] is also facilitated by the use of the parameter x. Using the above vevs and
the methods of [30] we calculated the complete gauge and chiral multiplet GUT scale spectra
and couplings for the 52 different MSSM multiplet sets falling into 26 different MSSM multiplet
types (prompting a natural alphabetization of their naming convention [30]!) of which 18 are
unmixed while the other 8 types occur in multiple copies. The (full details of these) spectra may
be found in [30,31] and equivalent results (with slightly differing conventions) are presented in
[32]. A related calculation with very different conventions has been reported in [33,41]. The
initially controversial relation between the overlapping parts of these papers was discussed and
resolved in [44].
Among the mass matrices is the all important 4 × 4 Higgs doublet mass matrix [30,31] H:
⎛ √ √ ⎞
−MH
√ +γ̄ 3(ω − a) −γ 3(ω + a) −γ̄ σ̄
⎜ −γ̄ 3(ω + a) 0 −(2M + 4η(a + ω)) 0√ ⎟
H=⎝ √ ⎠.
γ 3(ω − a) −(2M + 4η(a
√ − ω)) 0 −2ησ̄ 3
−σ γ −2ησ 3 0 −2m + 6λ(ω − a)
(10)
H can be diagonalized by a bi-unitary transformation [10,31,32]: from the 4 pairs of Higgs
doublets h(i) , h̄(i) arising from the SO(10) fields to a new set H (i) , H̄ (i) of fields in terms of
which the doublet mass terms are diagonal.


Ū T HU = Diag m(1) (2)
H , mH , . . . ,
h(i) = Uij H (j ) , h̄(i) = Ūij H̄ (j ) . (11)
To keep one pair of these doublets light one tunes MH so that Det H = 0. This matrix can
then be diagonalized by a bi-unitary transformation yielding thereby the coefficients describing
the proportion of the doublet fields in the 10, 126, 126, 210 GUT multiplets present in the light
doublets: which proportions are important for many phenomena. In the effective theory at low
energies the GUT Higgs doublets h(i) , h̄(i) are present in the massless doublets H (1) , H̄ (1) in a
proportion determined by the first columns of the matrices U, Ū :
E
MX : h(i) → αi H (1) , αi = Ui1 ,
h̄(i) → ᾱi H̄ (1) , ᾱi = Ūi1 . (12)
The all important normalized 4-tuples α, ᾱ can be easily determined by solving the zero mode
conditions: Hα = 0; ᾱ T H = 0.
C.S. Aulakh, S.K. Garg / Nuclear Physics B 757 (2006) 47–78 53

2.2. RG analysis and USMPs

In [1,31] we discussed at length the use of plots of the USMPs versus the fast parameter ξ to
investigate the question:
Are the one loop values of αG (MX ), Sin2 θW and MX generically stable against superheavy
threshold corrections?
We followed the approach of Hall [45] in which the mass (MX = mg λ 4|ã + w̃|2 + 2|p̃ + ω̃|2 )
of the baryon number violating superheavy ([3, 2, ± 3 ] or X-type) gauge bosons is chosen as the
5

transition scale between the effective MSSM and the full SO(10) GUT with all superheavy fields
retained. Thus MX is used as common physical superheavy matching point (Mi = MV = MX )
in the equations relating the MSSM couplings to the SO(10) coupling:
1 1 MX bij
= + 8πbi ln + 4π ln Xj − 4πλi (MX ). (13)
αi (MS ) αG (MX ) MS bj
j

See [31,45] for details. In this approach it is recognized that above the scale MX the effective
theory changes from the MSSM to Susy SO(10) model structured by the complex superheavy
spectra which we have computed so that the theory appears as unbroken SO(10) only to leading
−1
order. Thus we compute the corrections to the three USMPs Log10 MX , sin2 θW (MS ), αG (MX )
as a function of the MSGUT parameters and the answer to the question of stability of the per-
turbative unification is determined by the ranges of GUT parameters where these corrections are
consistent with the known or surmised data on Log10 MX , sin2 θW (MS ), αG . The consistency
requirement that the SO(10) theory remains perturbative after threshold and two loop correction
and αG not decrease so much as to invalidate the neglect of one-loop effects in the couplings λ,
η, γ , γ̄ is also appropriate. We implement these by the requirements that the USMPs obey

−1 
|ΔG | = Δ αG (MX )   10,
ΔX = Δ(Log10 MX )  −1,


|ΔW | = Δ sin2 θW (MS )  < 0.02. (14)
See [1,31] for a detailed discussion. We find the corrections
M
Δ(th) (Log10 MX ) = 0.0217 + 0.0167(5b̄1 + 3b̄2 − 8b̄3 ) Log10 ,
MX

M
Δ(th) sin2 θW (MS ) = 0.00004 − 0.00024(4b̄1 − 9.6b̄2 + 5.6b̄3 ) Log10 ,
MX

−1 M
Δ(th) αG (MX ) = 0.1565 + 0.01832(5b̄1 + 3b̄2 + 12b̄3 ) Log10 . (15)
MX
Where b̄i = 16π 2 bi are 1-loop β function coefficients (βi = bi gi3 ) for multiplets with mass M
(a sum over representations is implicit). Note a minor, but crucial correction, in the above equa-
M
tions relative to [1,31]: There and here, we actually used M X
as the argument of the logarithms
in these formulae. This is the value given by the algebra, which is then conventionally approx-

imated by M0 since the difference is of order g [45]. However in our case we can retain the
MX
“exact” form and still calculate the threshold corrections since those depend on ratios of masses.
This improves the analysis and makes it more convenient. Moreover (see below) it means that
the overall scale parameter that we use, namely the coefficient m of 2102 in the superpotential,
54 C.S. Aulakh, S.K. Garg / Nuclear Physics B 757 (2006) 47–78

is not fixed at the one loop unification scale 1016.25 GeV but slides with the corrections to MX .
These corrections are to be added to the one loop values corresponding to the successful gauge
unification of the MSSM: Using the values
0
αG (MX )−1 = 25.6, MX0 = 1016.25 GeV, MS = 1 TeV,
α1−1 (MS ) = 57.45, α2−1 (MS ) = 30.8, α3−1 (MS ) = 11.04 (16)
the two loop corrections are
 
2-loop MX

Δ log10 = −0.08, Δ2-loop sin2 θW (MS ) = 0.0026,
MS
Δ2-loop α −1 (M ) = −0.546.
G X (17)
We see that in comparison with the large threshold effects that might be expected [35] in view
of the large number (506) of heavy superfields the 2 loop corrections are quite small [31]. The
striking new result of [31] was thus the explicit demonstration that the conjecture [35] concerning
the “Futility of high precision SO(10) calculations” was in fact false and that a proper calculation
based on calculated spectra yields well defined and consistent results for significant regions of
the MSGUT parameter space (see below).
The parameter ξ = λM/ηm is the only numerical parameter that enters into the cubic Eq. (8)
that determines the parameter x in terms of which all the superheavy vevs are given. It is thus the
most crucial determinant of the mass spectrum. The dependence of the threshold corrections on
the parameters λ, η, γ , γ̄ is comparatively mild except when coherent e.g when many masses are
lowered together leading to αG explosion, Log MX collapse or large changes in sin2 θW (MS ).
The typical behaviours of the USMPs can be seen in Figs. 1–6.
The parameter ratio m/λ can be extracted as the overall scale of the vevs. Since the threshold
corrections we calculate are dependent only on (logarithms of) ratios of masses, the parameter
m is fixed in terms of the mass MV = MX of the lightest superheavy vector particles mediating
proton decay by:
 
MX
ΔX = Δ Log10 ,
1 GeV
|λ|
|m| = 1016.25+ΔX GeV. (18)
g 4|ã + w̃|2 + 2|p̃ + ω̃|2
The presence of the factor 10ΔX in the formulae for the neutrino masses will play a crucial role
as one of the hands of the “scissor” operated by Baryon Decay and constraints and neutrino
oscillation data. The lowering of the “slow diagonal” parameters λ, η tends to make fields light
and thus give large negative corrections to ΔG (as can be seen in the expansion of the darker areas
in Fig. 4 relative to Fig. 3). Such growth of the gauge coupling can invalidate the self consistency
of the perturbative approximation used through out. Conversely when αG decreases too much
the neglect of Yukawa loops becomes moot. Thus λ, η are effectively restricted to magnitudes
of order 1. Note, however, that ΔW , ΔX are insensitive to the precise values—whether real or
complex—of λ, η as long as their magnitudes are ∼ 1: as can be seen in the pairs Fig. 1 and Fig. 6
and Fig. 2 and Fig. 5. In addition there are the “slow off diagional” parameters γ , γ̄ whose effect
is quite mild, and can be taken to be any values ∼ 0.1 to 1.0, without any appreciable change in
the numerical plots.
C.S. Aulakh, S.K. Garg / Nuclear Physics B 757 (2006) 47–78 55

Fig. 1. Contour plot of the threshold corrections to Sin2 θW on the complex x plane. Contours at ΔW = −0.02, −0.01,
0.01, 0.02. The shading progresses from black (< −0.02) to white > 0.02.

2.3. Fermion mass formulae and SMPs

The Dirac masses of matter fermions in the MSGUT at scale MX are


mu = v(ĥ + fˆ),
mν = v(ĥ − 3fˆ),
md = v(r1 ĥ + r2 fˆ),
ml = v(r1 ĥ − 3r2 fˆ). (19)
Here we have extracted v = 174 GeV. Note the characteristic (Georgi–Jarlskog) [26] factors of
(−3) in the lepton masses relative to the quark masses. Here [1,30,31]

√ 2
ĥ = 2 2hα1 sin β, ˆ
f = −4 if α2 sin β,
3
α¯1 α¯2
r1 = cot β, r2 = cot β (20)
α1 α2
56 C.S. Aulakh, S.K. Garg / Nuclear Physics B 757 (2006) 47–78

Fig. 2. Contour plot of the threshold corrections ΔX to Log10 (MX /1 GeV) on the complex x plane. Contours at
ΔX = −1, −0.5, 0.5, 1. The shading progresses from black (< −1) to white > 1.

and (h, f ) are the couplings in the MSGUT superpotential. They are 3 × 3 symmetric matrices
with rows and columns indexed by the SO(10) family indices A, B = 1, 2, 3 [1,10,30,31]. Notice
that since |αi |  1 by unitarity, and |h, f | < 0.5 by perturbativity, the inclusion of α1 , α2 in the
definitions can yield subtleties only at zeros of α1 , α2 . The Majorana mass parameters Mν , Mν̄
of the left and right handed neutrinos defined as the coefficients of νν, ν̄ ν̄ in the superpotential
are [1,31]

Mν̄ = fˆσ̄ˆ ,
Mν = r3 fˆ, (21)
where [1,31]

i σ̄ 3
σ̄ˆ = ,
α2 sin β


MO = 2 M + η(3a − p) ,
C.S. Aulakh, S.K. Garg / Nuclear Physics B 757 (2006) 47–78 57

−1
Fig. 3. Contour plot of the threshold corrections ΔG to αG (MX ) on the complex x plane. Contours at ΔG = −10, −5,
5, 10. The shading progresses from black (< −10) to white > 10.
  
√ √ α4 v
r3 = −2i 3(α1 γ + 2 3ηα2 ) sin β. (22)
α2 MO
From these one obtains [31] the Type I [2] and Type II seesaw [3] Majorana masses (in the
conventional normalization W = Mν νν/2 + · · ·) of the low energy neutrinos:
MνI = vr4 n̂, MνII = 2vr3 fˆ,
iα2 sin β v
r4 = √ ,
2 3 σ̄
n̂ = (ĥ − 3fˆ)fˆ−1 (ĥ − 3fˆ). (23)
For investigation of the Type I seesaw we will find it useful to define functions FI , FII by ((1.70 ×
10−3 eV) = v 2 /MX0 and we have omitted irrelevant phases)


MνI = 1.70 × 10−3 eV sin βFI n̂,


MνII = 1.70 × 10−3 eV sin βFII fˆ, (24)
58 C.S. Aulakh, S.K. Garg / Nuclear Physics B 757 (2006) 47–78

−1
Fig. 4. Contour plot of the threshold corrections to αG (MX ) on the complex x plane illustrating the drastic effect of
decreasing the value of the diagonal parameter λ. Contours at ΔG = −10, −5, 5, 10 the shading progresses from black
(< −10) to white > 10.

where

x(1 − 3x)(1 + x 2 )
σ̂ = ,
(1 − x)2
  
−ΔX iα2 η 4|ã + ω̃|2 + 2|p̃ + ω̃|2
FI = 10 √ g ,
2 6 λ σ̂

√ (α1 γ + 2 3α2 η) 4|ã + ω̃|2 + 2|p̃ + ω̃|2
FII = 10−ΔX (−2i 3gα4 ) . (25)
α2 η (ξ + 3ã − p̃)
In these formulae the scale parameter m has been eliminated in favour of ΔX using Eq. (18).
Explicit forms for FI , FII in terms of the parameter x are given in Appendix A. We also define
the ratio R = |FI /FII | to capture the relative strength of the two seesaw contributions. A crucial
point is that if one wishes to maintain perturbativity and work with GUT Yukawa couplings  1
then the large tan β scenario is necessary. The Type I and Type II fits of the fermion mass-mixing
C.S. Aulakh, S.K. Garg / Nuclear Physics B 757 (2006) 47–78 59

Fig. 5. Contour plot of the threshold corrections ΔX to Log10 (MX /1 GeV) on the complex x plane showing the minor
effect of reducing the value of the diagonal parameter λ. Contours at ΔX = −1, −0.5, 0.5, 1 the shading progresses from
black (< −1) to white > 1.

data are carried out [14,20,22,24,25] assuming only the characteristic form of Eqs. (19), (21) to
derive “sum rules” [13,14,20,21] among the different mass matrices at the GUT scale. In these
generic fits, the freedom to choose the coefficients r1 , r2 , r3 , r4 is assumed to be compatible with
the theory in which the mechanisms are embedded. The fits then yield the neutrino masses only
up to an overall scale and the required relative strength of Type I and Type II is simply assumed.
In [1] we showed that these assumptions are unjustified for the MSGUT which seems to impose
extreme dominance of Type I over Type II fits and moreover is incapable of achieving neutrino
masses as large as 0.05 eV as required by experiment. Our argument was based on the following
observations:

• When a pure Type II fit is assumed one finds that [22,24,25] the maximal value of fˆ eigen-
values is ∼ 10−2 while the corresponding values for ĥ are about 102 times larger. This is a
structural feature due to the approximate equality of mb (MX ) and mτ (MX ) which ensures that
no Georgi–Jarlskog type contribution can be important for the third generation. As a result
60 C.S. Aulakh, S.K. Garg / Nuclear Physics B 757 (2006) 47–78

Fig. 6. Contour plot of the threshold corrections to Sin2 θw on the complex x plane showing the minor effect of reducing
the value of the diagonal parameter λ. Contours at ΔW = −0.02, −0.01, 0.01, 0.02 the shading progresses from black
(< −0.02) to white > 0.02.

n̂ ∼ 102 fˆ. This implies that the ratio R defined above must obey R  10−3 for the pure Type II
not to be overwhelmed by the Type I values it implies. Our initial survey of the MSGUT para-
meter space showed that, generically, this did not happen and pure Type II could not work. Even
near exceptional points no escape was allowed by the need to maintain viable USMPs. In the next
section we describe a much more exhaustive survey of the MSGUT to support this conclusion.
Note that this is the crux of the reason for the failure of the Type II seesaw and it applies equally
to “mixed Type I–Type II” fits [25].
• The Type I fits [15,25] yield, typically, n̂ ∼ 5fˆ ∼ 0.3 and thus we see that the magnitude of
the function FI will need to be greater than about 102 in order that neutrino masses as large as the
0.05 eV required for compatibility with neutrino mass squared splittings as large as Δ(mν )223 ∼
0.002 eV2 be achievable. This was seen to be generically un-achievable in [1]. In the next section
we prove that this requirement combined with that of viable USMPs makes it impossible for the
Type I fit to reach viability. This will be interpreted as indicating a missing piece to the puzzle,
the omission of the 120-plet.
C.S. Aulakh, S.K. Garg / Nuclear Physics B 757 (2006) 47–78 61

The estimate of the magnitude of n̂ chosen is obviously critical to the disqualification of the
MSGUT Type I masses. A value n̂ ∼ 10 would drastically modify our conclusions. The 3 gen-
eration fits performed so far are all simply numerical—although guided by analytic insight into
where in the parameter space a fit might occur [15,25]. Therefore every fit will have its own value
of the eigenvalues of n̂n̂† which might in principle have widely different magnitudes. However
no such exceptional Type I fit has been found so far. Although an analytical proof of this is
still elusive—because of the very complexity that obstructs the determination of an analytic fit,
there are heuristic arguments why a large n̂ fit is unlikely to exist. The charged fermion fit suc-
ceeds because the second generation couplings of the 126 dominate those of the 10 so that the
Georgi–Jarlskog mechanism can operate. Although the opposite (h33 f33 ) is true for the third
generation so that no such mechanism is effective (as it should not since at MX the b, τ masses
are roughly equal) yet the values of fˆ in the 2–3 sector are not so small that the eigenvalues of
n̂ become larger than 1. The Type I neutrino masses are found to obey a normal hierarchy where
the second and third generation neutrino masses are much larger than the first generation so that
the 2–3 block of n̂ is dominant over the rest of n̂. The large neutrino mixing in the 2–3 sector im-
plies that the matrix elements n̂ in that block are roughly of the same order of magnitude. These
elements are generically found to be of magnitude less than 1: The inverse hierarchical structure
of fˆ ensures that the normal hierarchy in ĥ − 3fˆ does not get amplified by the fˆ−1 factor in the
Type I seesaw formula to the extent of increasing the maximal eigenvalue of n̂ beyond 0.5 or so.
An analytic proof of this is still lacking. Thus in principle there is a loophole which is still un-
closed: Exceptional Type I fits with maximal eigenvalue greater than 10, if found, will need to be
reanalyzed separately. We emphasize that the question is not one of finding improved or optimal
fits but only of the order of magnitude of the maximal eigenvalue of n̂. We shall assume that such
exceptional fits do not exist. In other words we assume that the Type I fits of [15,25] are generic
and representative as regards the order of magnitude of n̂. It is important to build a data base or
other more compact description of all passable achieved fits that may be used to survey all the
h2
possibilities. Thus a naive estimate for the typical maximal magnitude of n̂: Max n̂ ∼ f3333
1
is actually far off the mark. As explained above the characteristic off diagonal structure of the
successful Type I fits in the generic Babu–Mohapatra proposal, seems to be intimately bound
up with this difference between the naive and actual magnitudes of (the eigenvalues of) n̂. All
these remarks are illustrated by the values taken from the published example in [25] (which is,
however, not a completely satisfactory fit since the electron mass is fit only to an accuracy of
about 10%).
⎛ ⎞
−0.00016 − 0.00013i −0.000456 − 0.00035i −0.00642 + 0.002i
ĥ = ⎝ −0.00046 − 0.00035i −0.00584 − 0.00101 0.03142 ⎠,
−0.00642 + 0.002i 0.03142 −0.55633 + 0.095531i
(26)
 
0.00023 + 0.00018i 0.00018 + 0.00017i 0.00285 − 0.00053i
fˆ = 0.00018 + 0.00017i 0.00697 + 0.00164i −0.01347 − 0.00158i , (27)
0.00285 − 0.00053i −0.01347 − 0.00158i 0.00053 − 0.037858i
 
2997.76 + 347.393i 115.886 − 418.543i 125.607 − 173.883i
ˆ−1
f = 115.886 − 418.543i 51.4218 + 16.1651i −40.8494 + 15.3566i ,
125.607 − 173.883i −40.8494 + 15.3566i −18.5738 + 4.48526i
(28)
ˆ 2 ˆ
n̂33 = (ĥ33 − 3f33 ) /f33 = 6.37485 + 6.98329i,
est
(29)
62 C.S. Aulakh, S.K. Garg / Nuclear Physics B 757 (2006) 47–78

 
0.00278 + 0.00284i 0.00712 + 0.00287i 0.02697 + 0.00393i
n̂ = 0.00712 + 0.00287i 0.09225 + 0.02520i −0.12975 − 0.08165i , (30)
0.02697 + 0.00393i −0.12975 − 0.08165i 0.00256 + 0.15193i

1
Eigenvalues [n̂.n̂† ] 2 = {0.2771, 0.06298, 0.01088}. (31)

The magnitude of the maximum eigenvalue of n̂ is seen to be almost 35 times smaller than the
“generic argument” one! Thus generic arguments for this magnitude need to be taken with a large
lump of salt!

3. USMP and SMP plots: Gloomy pictures?

We wish to survey the USMP and SMP variation over the complex x plane. We will use
generic but fixed values with magnitudes ∼ 1 for the ‘diagonal’ slow parameters [1,31] {λ, η}.
We have checked that for any such (real or complex) values there is no significant modification
of our conclusions. If one lowers λ, η below about 0.2 or so the USMPs exhibit pathologies due
to mass scale lowering. Specifically, an increase of ΔG beyond a value of 10 or a decrease below
(say) −10 makes the perturbative formulae used throughout unreliable. Similarly varying the
‘non-diagonal’ slow parameters γ , γ̄ to small values, though not forbidden by USMPs [1,31] is
also infructuous for purposes of viable SMPs.
Besides the survey across generic regions of the complex x (equivalently the ξ ) plane we must
also examine the possibility [1,10,31,32,43] that the behaviour at certain exceptional points may
invalidate the generic trend of seesaw exclusion. For this purpose we provide in Table 1 a list
of the exceptional points. These consist on the one hand of a list of points of enhanced gauge
symmetry where the USMPs typically show spikes [31] and on the other of special points where
the SMPs may exhibit spikes [1,43]. The complete analytic [10,32,43] identification of the latter
set of points using the convenient parametrization of vevs in terms of rational functions of the
variable x alone (i.e. with ξ eliminated in favour of x) is related in Appendix A.

Table 1
Special points with extended gauge symmetry and/or candidate seesaw enhancement
x ξ Symmetry/remarks
1/2 −5 SU(5), zero of p2
−1 10 SU(5) flipped, zero of p2
0 3 SU(3) × SU(2)L × SU(2)R × U (1)B−L
1/3 −2/3 Flipped SU(5) × U (1)
±i 3 ∓ 6i SU(3) × SU(2)L × U (1)R × U (1)B−L
0.123765 1.93008 zero of p3
0.646451 − 0.505389i 4.86831 + 3.29252i zero of p3
0.646451 + 0.505389i 4.86831 − 3.29252i zero of p3
√ 3 (1 ∓ i √7 )
(3 ± i 7 )/8 2 zero of q2
−3.46301 28.2958 zero of p5
0.262212 ± 0.388123i 2.13469 ∓ 3.38025i zero of p5
0.358184 ± 0.133971i −0.449266 ∓ 2.30098i zero of p5
0.1984 1.167 zero of q3
−0.099 ± 2.24i 1.596 ∓ 15.572 zero of q3
C.S. Aulakh, S.K. Garg / Nuclear Physics B 757 (2006) 47–78 63

3.1. The fatal weakness of Type II

Let us begin with a discussion of the viability of the Type II mechanism. The SMP R we
proposed [1] to monitor the strength of Type I relative to the Type II seesaw is particularly
simple to analyze. It is the magnitude of a rational analytic function (the ratio of a polynomial of
degree 10 to a polynomial of degree 9) of x and is dependent on no other parameters.
     
 FI   (x − 1)(3x − 1)q2 q3 2   P (x) 

R= =    =   (32)
FII 4(4x − 1)(x 2 + 1)q32   Q(x) 
(the polynomials qi , qi are defined in Table 2 in Appendix A).
Therefore R is bounded from below by 0 and its minimae are isolated: they are the zeros
of P (x). As |x| → ∞ it grows as |x| (with coefficient 3/64). For |x|  3 ⇒ R  10−1 or so.
This is clearly visible in the 3D plot shown in Fig. 7. Thus only for |x|  3 do we need to even
consider the possibility of dominant Type √ II seesaw. This is in fact the region which contains the
zeros of R namely x = {1/3, 1, {(3 ± i 7)/8}, {0.198437, −0.0992186 ± 2.24266i}} of which
the latter two sets are the zeros of q2 and q3 respectively. A picture of the variation of R in
this region of the x plane is given by the 3D plot of R −1 in Fig. 8 which clearly shows that, as
expected, the zeros of the numerator dominate the behaviour.
The real zeros of R show up as “cliffs” (instead of narrow peaks ) due to periodicity in Arg x
which slices them vertically, while “ridges” observed in these graphs are associated with the
complex zeros of q2 , q3 (one needs to magnify the graph to see the ridge form of the zeros of

Fig. 7. 3D plot of the relative strength R of Type I and Type II seesaw mechanisms on the complex x plane showing that
the region |x| > 3 is irrelevant for promoting Type II dominance.
64 C.S. Aulakh, S.K. Garg / Nuclear Physics B 757 (2006) 47–78

Fig. 8. 3D plot of the relative strength R −1 of Type II versus Type I seesaw mechanisms on the complex x plane showing
the zeros of R in the region |x| > 3.

q2 ). There is a strong preference for a narrow range of Arg x for minimizing R (as also seen
directly in the plots of R in the neighbourhood of these zeros discussed below). Thus the growth
of R −1 is seen to be well localized to the regions around the zeros of R. From this survey as
well as the theory of analytic functions it is clear that the behaviour of R and the USMPs in the
neighbourhood of its zeros will be sufficient to decide whether any Type II solutions are viable.
We now discuss this in greater detail.
Since Type II dominance desires as small a value as possible of R we need to approach as
closely as viable to the zeros of P (x). Among the zeros of R, however, are also the zeros of
α2 : {x = {1/3, 1}, {0.1984, −0.099 ± 2.24i} (zeros of q3 )}. The decrease of |α2 | is limited by
the requirement of maintaining perturbativity in the superpotential Yukawa coupling ∼ 3fˆ/α2 <
1. Type II dominant fits yield fˆ about 10−2 [1,22,24,25]. Even if we are generous with the
coefficients (∼ 3) that occur in the relation, α2 cannot decrease below 10−2 : so we must maintain
at least such distance from the zeros of α2 so that: |α2 | > 10−2 . The remaining zeros of R are
just those of q2 . We shall see below that approach to these zeros is also limited by the USMPs.
Thus R cannot viably decrease below R̄min = Min{R̄(xi ), i = 1, . . . , 7} the values of R at closest
permissible approach to its zeros xi .
No further survey of the parameter space is thus required than to evaluate R̄min . To esti-
mate R̄min (xi ) is straightforward given our established USMP-SMP numerical techniques [1,31].
We allow an approach to the zeros of R as closely as will permit α2  10−2 . This limits the
radius of the circle of closest approach to the zeros x = {1/3, 1, {0.198437, −0.0992186 ±
2.24266i}} of α2 , for the 3 real zeros and the pair of complex zeros respectively, to be i =
C.S. Aulakh, S.K. Garg / Nuclear Physics B 757 (2006) 47–78 65

Fig. 9. Plot of R on a circle of radius 0.02 around x0 = −0.099219 + 2.2426i.

10−2 {0.7, 3.9, 4.0, 0.53}−1 = {0.014, 0.0025, 0.0025, 0.02}. We observe values of
 
xi , i , R̄min (xi ), {U SMP s}
 
= 1, 0.014, 0.007, {ΔG ∼ −14, ΔW ∼ −0.11}
 
= 0.3333, 0.0025, 0.015, {U SMP sOK}
 
= 0.198437, 0.0025, 6 × 10−5 , {ΔX < −2.9}
 
= 0.198437, 0.025, 0.016, {ΔX < −2.4}
 
= −0.099219 + 2.2426i, 0.02, 2 × 10−5 , {U SMP sOK, FII ∼ 2.4} . (33)
The closest one comes to the required values of R ∼ 10−4 or smaller is in the case of the complex
zeros of q3 . However even in this case the value of FII is simply too small and yields a maximal
neutrino mass of about 4 × 10−5 eV (since fˆ ∼ 10−2 ). To illustrate the behaviour in this most
favourable but still unviable case we exhibit the variation of R, ΔX , FII on a circle of radius 0.02
around a complex zero of q3 as Figs. 9–11.
Finally we turn to the zeros of q2 . They are not zeros of α2 so there is no constraint on
approach to the zero from perturbativity but the conclusions are equally negative:
 
x, , R̄min (xi ), {U SMP s}
 √ 
= (3 + i 7)/8, 0.06, 0.015, {ΔX < −0.9, FII ∼ 80}
 √ 
(3 + i 7)/8, 0.005, 0.0014, {ΔX < −1.3, FII ∼ 2100} . (34)
The rapid fall of ΔX (which also promotes FII growth) as one approaches the zeros excludes
them long before the required small values of R are reached. The strong decrease of ΔX for
values of x near x = 0 is obvious from Figs. 2 and 5.
66 C.S. Aulakh, S.K. Garg / Nuclear Physics B 757 (2006) 47–78

Fig. 10. Plot of ΔX on a circle of radius 0.02 around x0 = −0.099219 + 2.2426i.

Fig. 11. Plot of FII on a circle of radius 0.02 around x0 = −0.099219 + 2.2426i.
C.S. Aulakh, S.K. Garg / Nuclear Physics B 757 (2006) 47–78 67

To sum up: the zeros of R are ringed by zones of exclusion that ensure the parameters of the
MSGUT can never be chosen to ensure Type II domination while maintaining viable USMPs.

3.2. Type I fails: Collapse on the final lap

We next turn to a consideration of the overall magnitude of the Type I seesaw masses which
we have shown to dominate the Type II contribution almost everywhere. The Type I mass formula
reads—up to an irrelevant overall phase


MIν = 1.7 × 10−3 eV FI n̂ ∼ sin β,
 √
γg (1 − 3x) |p2 p3 | z2 −ΔX
FI = √ √ 10 q3 (35)
2 2ηλ x(x 2 + 1) z16
= F̂I 10−ΔX .
The complicated effect of the RG thresholds and evolution are all contained inside ΔX . How-
ever this quantity is itself limited to lie above −1 by Baryon lifetime > 1033 yrs! Before actually
inspecting the effect of ΔX in detail we should therefore inspect whether the almost completely
known function F̂I could ever achieve the magnitude of 10 or so it would require to yield neu-
trino masses of around 0.05 eV (recall n̂ ∼ 0.3 and ΔX > −1) in the most favourable case. This
seems a rather mild requirement but we shall show that it cannot be satisfied viably anywhere in
the parameter space of the MSGUT!
The limiting behaviour of |F̂I | is
|x| → ∞ ⇒ |F̂I | →∼ 0.02,
|x| → {0, ±i} ⇒ |F̂I | → ∞. (36)
The approach to the asymptotic value is rather rapid while the singularities as x ∼ {0, ±i} are
1
quite mild: ∼ |Δx|− 2 . Figs. 12, 13 are 3D plots of |F̂I | over the complex x plane for |x| < 10
(more than sufficient to reach the limiting form) and |x| < 1.5 (to show the singularities at x =
0, ±i more clearly).
The dependence of F̂I on the ‘slow AM’ parameters {γ , γ̄ , η, λ} is rather marginal—as long
as we do not consider pathological cases where η or λ are very small (forbidden by the USMP
variation [1]) or pathological cases where perturbativity in any of these couplings is lost. We
have verified this by examining 3-dimensional plots of FI over the complex x plane for random
complex values of these couplings without finding any visible variation. Apart from the con-
straint that their magnitudes are greater than 0.2 their values can be chosen randomly without
any appreciable effect.
In Figs. 12 and 13 the rapid approach to the asymptotic value of |F̂I | beyond |x| = 1.2, the
ridge associated with x = 0 (due to its indeterminate phase), and the pointed ridges associated
with x = ±i, as well as the boundedness by about 0.1 except right close to the singularities,
are the relevant and crucial features. Plots which have very different coupling values are almost
indistinguishable from Figs. 12, 13. This is easy to understand in terms of the expression for Z16
(Table 2) since that is where the dependence on these parameters enters. It is evident that the
variation of the slow parameters has no appreciable effect at all. So they can be set to have any
values with magnitudes ∼ 1 without any loss of generality. The behaviour of F̂ implies that only
very close to the singular points x = {0, ±i} is there any hope that |F̂I | will achieve the required
values  10. Elsewhere it is manifestly bounded by 0.1 or so and is thus a factor of 102 short
68 C.S. Aulakh, S.K. Garg / Nuclear Physics B 757 (2006) 47–78

Fig. 12. 3D plot of |F̂I | on the x-plane.

of the required magnitude (in the most favourable case of ΔX = −1: Which value is itself only
borderline compatible with Super-Kamiokande data on baryon lifetime).
In Figs. 14, 15 we exhibit contour plots of F̂I and FI respectively on the complex x-plane.
Evidently only around x = 0, ±i is there any chance of large enough F̂I . However neither region
gives a viable result. Firstly, it is evident from Fig. 2 that the region around x = 0 is forbidden
since ΔX < −1. A ‘scissors’ operates between the exploding effect of ΔX as one approaches
x = 0 and the rapidly decreasing |F̂I | as one recedes from it.
Around x = ±i, ΔX is > 1 and hence will suppress rather than enhance the favourable value
of F̂I . This is evident in Fig. 15 where x = ±i are out of the running altogether. One is again
caught in a scissors. At this point (of enhanced G3211 symmetry) ΔX is large and positive at the
singularity. So it completely overwhelms the growth of F̂I due to the singularity.
Thus we have shown that even via Type I seesaw there is no possibility of achieving neutrino
masses larger than about 5 × 10−3 eV in the MSGUT!
Although the argument given above is complete we also studied the behaviour at other special
points such as x = {−1, 1/3, 0.5} and the zeros of {q2 (x), q3 , p3 (x), p5 (x)}, some of which have
been claimed (and some not!) to represent candidate points for Type I enhancement [43] but
are not in fact so. We did not find any appreciable growth of FI with tolerable values of ΔX at
any such point: exactly as expected from our arguments given above. Some growth is observed
for the x values (digits truncated ) 0.1237 (real zero of p3 ), 1/3 (enhanced symmetry point
with symmetry SU(5)flipped × U (1)) and 0.198 (real zero of q3 ). In all three cases it is due to
C.S. Aulakh, S.K. Garg / Nuclear Physics B 757 (2006) 47–78 69

Fig. 13. 3D plot of |F̂I | on the x-plane for |x| < 1.5.

catastrophic fall in ΔX . The contour plot of FI neatly summarizes all this behaviour since only
the forbidden region around x = 0 ever gives large enough FI .
To trace the reason for the divergence in our candidates from those of [43] is easy: if we
translate the expressions of [43] which are written in terms of ml , md − ml into our expressions
written in terms of ĥ, fˆ we immediately obtain that the functions fI , fII of [43] need to be
multiplied by (in our notation) {r12 /(4r2 ), r2 } ∼ {(p3 p4 )/(p5 p2 ), (p5 p2 )/(p3 p4 )} respectively
and this then removes the aforementioned points from the list of candidates, leaving, after all
repeated factors are cancelled, only {0, ±i} as candidates for strong Type I seesaw and the zeros
of R namely x = {1/3, 1} and the zeros of q2 and q3 .

4. The next MSGUT

We have seen that the Type II seesaw mechanism does not work when only the 10 and 126
FM Higgs irreps are employed and even the Type I mechanism fails even though not as badly as
Type II. It is natural then to look for a role for the remaining FM Higgs multiplet type allowed
by SO(10): The long neglected 120-plet irrep. This irrep has a Yukawa coupling matrix that
obeys gAB = −gBA i.e. is family index antisymmetric. As such it introduces novel properties and
problems into the fitting matrix problem. It has so far been cast in a minor supporting role [17,24]
in the story of the seesaw. Only a few authors [14,16,18,19] have considered the 120 plet in a large
role and even they did not analyse the fitting problem comprehensively. The 10 and 120 irreps
may together successfully explain the charged fermion spectrum, particularly since the 120-plet
also contains, besides a (1, 2, 2) sub-representation, a (15, 2, 2) subrep (w.r.t. the Pati–Salam
70 C.S. Aulakh, S.K. Garg / Nuclear Physics B 757 (2006) 47–78

Fig. 14. Contour plot of |F̂I | on the x-plane with contours at |F̂I | = 0.01, 0.1, 0.5. Lower values are shaded darker. The
small white islands are around x = 0, ±i.

subgroup). Thus it can also be used to implement—albeit in a novel way due to its antisymmetric
Yukawa couplings—a Georgi–Jarlskog [26] mechanism to explain the approximate equality of
ms and mμ at MS in the MSSM. This relieves the 126 of the multiple loads it has shouldered
so far and which, finally, the present papers show it tends to cast down on the last lap of tests.
Once the 126-plet coupling fAB is relieved of the necessity of being comparable to the couplings
of the 10-plet the Type I seesaw mechanism can receive the required further 10–100 fold boost
in magnitude since the Type I seesaw masses are proportional to fˆ−1 . It should be emphasized
that the smallness of fˆ in this scenario must be considered as a structurally defining condition
rather than as the smallness of a perturbation because it leads to light right handed neutrinos.
A similar scenario was considered in [16] but analyzed by making various somewhat arbitrary
assumptions. Thus no generic analysis like that for the 10 − 126 [21,22,24,29] system is yet [46]
available to compare with.
In this way we propose that our results provide a hint of the direction to proceed: every mem-
ber of the SO(10) FM Higgs family will receive its characteristic function, role and gainful
C.S. Aulakh, S.K. Garg / Nuclear Physics B 757 (2006) 47–78 71

Fig. 15. Contour plot of |FI | on the x-plane with contour at |FI | = 10 lower values are shaded darker. Note that the
required large values can only lie right close to the forbidden region around x = 0 where ΔX < −1.

employment. It seems that the Good God may be just at least to the different breeds of Higgs
particles!
To implement our (long felt [48]) intuition of a proper share and role for each Higgs type in
the intricately wrought elegance of SO(10) unification we have calculated [46] the complete set
of couplings and masses of the Next to Minimal Supersymmetric GUT or NMSGUT defined by
the inclusion of a 120-plet Higgs into the MSGUT Higgs complement. Here we restrict ourselves
to a brief survey of the essential features.
In the notation of [30] the real vector indices of the upper left block embedding of SO(6)
in SO(10) are denoted by a, b = 1, 2, . . . , 6 and of the lower right block embedding of SO(4)
in SO(10) by α̃, β̃ = 7, 8, 9, 10. The doublet indices of SU(2)L (SU(2)R ) are denoted by
α, β = 1, 2 (α̇ β̇ = 1̇, 2̇). The index of the fundamental 4-plet of SU(4) is denoted by a (lower)
μ, ν = 1, 2, 3, 4 and its upper-left block SU(3) subgroup indices are μ̄, ν̄ = 1, 2, 3.
The decomposition of the 120-plet w.r.t. SU(4) × SU(2)L × SU(2)R is as follows:

Oij k (120) = Oabc (10 + 10, 1, 1) + Oabα̃ (15, 2, 2)




+ Oa α̃ β̃ (6, 1, 3) + (6, 3, 1) + Oα̃ β̃ γ̃ (1, 2, 2)
μν
= Oμν
(s)
(10, 1, 1) + Ō(s) (10, 1, 1) + Oνα α̇ μ (15, 2, 2)
(a)
+ Oμν α̇ β̇ (6, 1, 3) + Oμν
(a)
αβ (6, 3, 1) + Oα α̇ (1, 2, 2) (37)
72 C.S. Aulakh, S.K. Garg / Nuclear Physics B 757 (2006) 47–78

(where we have used the superscripts (s), (a) to discriminate the symmetric 10-plet from the anti-
symmetric 6-plet).
The decomposition of its coupling to 16-plets of SO(10) is [30]
1
ψC2(5) γi γj γk χOij k
(3!)

μν √

= −2 Ō(s) ψμα χνα + Oμν ψ̂ χ̂α̇ − 2 2Oνμα α̇ ψ̂α̇ν χμα − ψμα χ̂α̇ν
(s) μα̇ ν


(a) α̇ β̇ μ ν μναβ √
μ μ
− 2 Oμν ψ̂α̇ χ̂β̇ + Õ(a) ψμα χνβ + 2O α α̇ +ψ̂α̇ χμα − ψμα χ̂α̇ . (38)

The SU(4) × SU(2)L × SU(2)R labels in the above [30] equations are trivially decompos-
able to SM labels and thus such expressions—available for every coupling in the MSGUT and
NMSGUT—constitute a complete solution for the Clebsches in these MSGUTs.
(15)α (15)α
The 120-plet contributes two pairs hα5 = O1̇α , h̄α5 = O2̇α , hα6 = O1̇ , h̄α6 = O2̇ of MSSM
([1, 2, 1] ⊕ [1, 2, −1]) doublets. Their Yukawa couplings to MSSM fermions follow immediately
from the above unitary decompositions:
ᾱ1 ᾱ2
mu = v(ĥ + fˆ + ĝ), r1 = cot β, r2 = cot β,
α1 α2


4i 3α5
ˆ
mν = v (ĥ − 3f ) + (r5 − 3)ĝ , r5 = √ ,
α6 + i 3α5

ᾱ6 + i 3ᾱ5
md = v(r1 ĥ + r2 fˆ + r6 ĝ), r6 = √ cot β,
α6 + i 3α5

4i 3ᾱ5
m = v(r1 ĥ − 3r2 fˆ + r7 ĝ),
l
r̄5 = √ cot β,
α6 + i 3α5

2 √
ĝ = 2ig (α6 + i 3α5 ) sin β, r7 = (r̄5 − 3r6 ) (39)
3
which clearly exhibits the analogy with the Georgi–Jarlskog type couplings in the MSGUT. We
propose to fit the charged fermion masses using light doublet components of predominantly 10-
and 120-plet origin. The additional terms in the superpotential do not modify the AM SSB and
moreover still contain only the 26 MSSM multiplet types that we described [31] for the MSGUT.
The corresponding mass matrices—some of whose dimensions, like the doublet mass matrix
given below, are raised relative to the MSGUT—will be given, in our notation, and along with
the corresponding RG and USMP-SMP analysis in [46]. A corresponding calculation of mass
matrices only—with very different conventions—is already available [44].
The extra terms contributed by the 120-plet to the superpotential are:
M0 k ρ
W120 = Oij k Oij k + Hi Oj kl Φij kl + Oij m Oklm Φij kl
2(3!) 3! 4!
ζ ζ̄
+ Oij k Σij lmn Φklmn + Oij k Σ̄ij lmn Φklmn . (40)
2(3!) 2(3!)
The coupling of the P([3, 3, ± 23 ]) and K([3, 1, ± 83 ])-type color triplets contained in the 120-
plet leads to quite novel, family index antisymmetric and even SU(2)L triplet mediated, con-
tributions to the effective LLLL and RRRR type ΔB = 0 violating d = 5 operators in the low
C.S. Aulakh, S.K. Garg / Nuclear Physics B 757 (2006) 47–78 73

energy effective theory [46]. These may lend additional texture to the emerging story of the deep
connections between neutrino mass and Baryon violation [31,49].
The composition of the MSSM [1, 2, ±1] doublet sector can be deduced from the left and
right null eigenvectors of the 6 × 6 mass matrix of the NMSGUT analogously to those defined
for the MSGUT in Section 2.
√ √ √
⎛ −MH γ̄ 3(ω − a) −γ 3(ω + a) −γ̄ σ̄ kp − 3ikω ⎞
√ √
−γ̄ 3(ω + a) 0 −(2M + 4η(a + ω)) 0 − 3ζ̄ ω i(p + 2ω)ζ̄
⎜ √ √ √ ⎟
⎜ γ 3(ω − a) −(2M + 4η(a − ω))

0 −2ησ̄ 3 3ζ ω −i(p − 2ω)ζ
√ ⎟
⎜ −σ γ −2ησ 3 −2m + 6λ(ω − a) ⎟.
⎜ √ √
0 ζσ 3iζ σ ⎟
⎝ pk 3ζ̄ ω − 3ωζ ζ̄ σ̄ −Mo √ρ iω
3

√ √
3ikω i(p − 2ω)ζ̄ −i(p + 2ω)ζ − 3i ζ̄ σ̄ − √ρ iω −M0 − 2ρ
3 a
3

The left and right null eigenvectors are calculated after imposing the light doublet condition
Det H = 0.
The relative strength of the contributions, of the 10- and 120-plets vis à vis the 126-plet, to the
charged fermion mass matrices can be studied by examining the variation of α2 /α1,5,6 , ᾱ2 /ᾱ1,5,6
in exactly the same manner as already done by us in this paper. The pure form of this scenario
can thus be examined directly in terms of fitting the charged fermion masses and mixings in
terms of 10- and 120-plet contributions only. If this succeeds to a within about 10%, the effect
of including the 126-plet contributions can be examined [46]. Naturally, the investigation is best
begun with the 2 × 2 case as for the 10 + 126 system [19,21,29]. This is reported in [47].

5. Conclusions and outlook

The Supersymmetric GUT based on the 210 ⊕ 126 ⊕ 126 AM Higgs system is the simplest
Supersymmetric GUT that elegantly (almost, v.supra!) realizes the classic program of Grand Uni-
fication. Its symmetry breaking structure is so simple as to permit an explicit analysis of its mass
spectrum at the GUT scale and an evaluation therefrom of the threshold corrections and mix-
ing matrices relevant to various important quantities. It implements in a fascinating and elegant
way an intimate (“ouroborotic”) connection between the physics of Lepton number and Baryon
number violation. Since it has the least number of parameters of any theory that accomplishes as
much this theory merited the name of the minimal supersymmetric GUT or MSGUT. The same
simplicity and analyzability of GUT scale structure also applies to the theory with an additional
120-plet, since it contains no standard model singlets, and thus justifies calling it the next to min-
imal (or new minimal) Susy GUT (NMSGUT). The small number of Yukawa couplings of the
MSGUT makes the fit to the now well characterized fermion mass spectra very tight. In [1] we
initially observed and in this paper we have shown that the combined constraints of the seesaw
fit and the preservation of MSSM one loop gauge unification are enough to rule out the MSGUT.
We now argue that the very natural inclusion of the third possible SO(10) FM Higgs type,
i.e. the 120-plet, which was somewhat arbitrarily [9]—but very fruitfully [13–15,17,20–22,24,
25,29]—excluded from a leading role in the ‘SO(10) party on the GUT scene’, offers a natural
resolution of the difficulty of the MSGUT in achieving large enough Type I seesaw masses. In
this scenario the 10-plet and 120-plet are primarily responsible for the charged fermion mass fit
due to a relatively suppressed contribution of the 126-plet to the charged fermion masses due
to a small value of the Yukawa coupling 16 · 16 · 126. This very suppression implies that the
Type I seesaw will become enhanced essentially due to lowering of the right handed neutrino
masses. Thus far from being a perturbation the role of the 126 is rather to define a very specific
class of Grand Desert in which the right handed neutrinos have masses considerably less than the
74 C.S. Aulakh, S.K. Garg / Nuclear Physics B 757 (2006) 47–78

generically expected GUT scale masses. Thus this proposal, which was thrust upon us by detailed
analysis of compatibility between GUT and Neutrino mass scale hierarchies, is distinct from
previous scenarios and, moreover, has very clear and distinct phenomenological consequences.
Furthermore the 126-plet Yukawa coupling is itself released from the Type I charged fermion
constraints and is subject only to the relatively mild constraints coming from the right handed
Majorana neutrino mass limits from cosmology. If this does not conflict with the requirement to
achieve the small–large/quark–lepton mixing duality then the one degree of magnitude failure of
the seesaw in the MSGUT should be quite surmountable. In a slogan:

The MSGUT is dead! Long live the (Next)MSGUT!

Note added

Our surmise was announced in May/June 2005 (PLANCK05, Trieste, May 2005 and hep-
ph/0506291) and our proof appeared as hep-ph/0512224v1. In May 2006 a calculation appeared
[50] confirming that optimal FM fits found by the “downhill simplex” method are not viable in
the MSGUT, for exactly the reasons given by us.

Acknowledgements

C.S.A. is grateful to S. Bertolini and M. Malinsky for providing a complete set of data of
one of their Fermion Mass fits and to G. Senjanovic, Alejandra Melfo for correspondence and
encouragement. S.K.G. acknowledges financial support from the University Grants Commission
of the Government of India. C.S.A. is very grateful to Sukhdeep Randhawa and Amarjit S. Mundi
for hospitality and help with graphics issues and to his family, Satbir Kaur, Simran K. Aulakh
and Noorvir S. Aulakh, for their patience and good cheer during the very trying completion of
this paper while travelling on vacation.

Appendix A

In this appendix we express our formulae [1,31] for neutrino masses etc in terms of the vari-
able x which parameterizes [10] the fast variation of AM scale mass matrices in an elegant
and simplifying manner, as has been particularly emphasized by [32,43]. This allows the easy
analytic localization of putative points of seesaw growth [43]. It enhances confidence in the
completeness of the exclusion by survey [1] of any possible seesaw fit. The use of the vari-
able x also significantly simplifies the presentation of data since considering variation over the
x-plane unifies consideration of the three solutions for x corresponding to a given value of ξ
while retaining complete clarity due to the rational (3 to one) map from the x-plane to the
ξ -plane. The process of translation merely makes explicit in the mass matrices what was al-
ready done [10,31,32] when parameterizing the AM vevs in terms of x. If we substitute for the
AM vevs in terms of x, the equations for the null eigen-vectors of H—the doublet mass ma-
trix of the MSGUT—subject to the condition Det H = 0 become: (note the labelling of doublets
 (15)α 2̇α
(15)α φ44 (15) (15) φ 441̇
h h̄ = (H2̇α , Σ̄2̇ , Σ2̇ , √ ) ⊕ (Hα 1̇ , Σ̄α 1̇ , Σα 1̇ , √α ))
2 2

A = {α1 , α2 , α3 , α4 } = N{α̂1 , α̂2 , α̂3 , α̂4 } = N Â,


ˆ
Ā = {ᾱ , ᾱ , ᾱ , ᾱ } = N̄{ᾱˆ , ᾱˆ , ᾱˆ , ᾱˆ } = N̄ Ā,
1 2 3 4 1 2 3 4
C.S. Aulakh, S.K. Garg / Nuclear Physics B 757 (2006) 47–78 75

Table 2
Functions of x entering the expressions for the USMPs and SMPs
q2 4x 2 − 3x + 1
q3 4x 3 − 9x 2 + 9x − 2
q3 p4 /(3x − 1) = x 3 + 5x − 1
p2 (2x − 1)(x + 1)
p3 12x 3 − 17x 2 + 10x − 1
p4 (3x − 1)(x 3 + 5x − 1)
p5 9x 5 + 20x 4 − 32x 3 + 21x 2 − 7x + 1
z2 2|x|2 + |1 − x|2

 γ (x−1)p p 2 3  γ̄ (x−1)p p 2 1  x(1−3x)(1+x 2 ) 2
z16 |p3 p5 |2 + 34  η
3 4 + 
4 η
2 5  + γ (x − 1)p q
2 3 3 λη

 γ̄ (x−1)p p 2 3  γ (x−1)p p 2 1  x(1−3x)(1+x 2 ) 2
z̄16 2
|p3 p5 | + 4 3  3 4  +4  2 5  
+ 2 γ̄ (x − 1)p3 q3
η η λη

N |p3 p5 |/ z16

N̄ |p3 p5 |/ z̄16

 √ √ 
( 3 γp4 (1 − x)) ( 3 γ̄ 2(1 − x)) γ q3 σ̃ (x − 1)
 = 1, , , √ ,
2ηp5 2ηp3 2 ηλp5
 √ √ 
ˆ ( 3 γp4 (1 − x)) ( 3 γ̄ p2 (1 − x)) γ̄ q3 σ̄˜ (x − 1)
Ā = 1, , , √ ,
2ηp5 2ηp3 2 ηλp5
|p3 p5 |
N= √ ,
z16
|p3 p5 |
N̄ = √ . (A.1)
z̄16
The crucial functions FI , FII , R used as SMPs [1] in our survey now become (up to irrelevant
overall parameter phases)
 
10−ΔX γ g z2 (1 − 3x) q3
FI = √ √ |p2 p3 p5 | ,
2 2 ηλ z16 x(1 + x 2 ) p5
√   2
2 2γ g |p p p | z2 (x + 1) (4x − 1)q32
FII = 10−ΔX √
2 3 5
,
ηλ (x − 1) z16 x(1 − 3x) q3 q2 p5
   
 FI   (x − 1)(3x − 1)q2 q3 2 
R =   =  . (A.2)
FII 8(4x − 1)(x 2 + 1)q32 
The various polynomials defined in [32,43] together with some additional definitions are given
in Table 2.
A number of remarks are in order
• The asymptotic behaviour of {F̂I , F̂II , R} as |x| → ∞ is {|x|0 , |x|−1 , |x|} and as x → 0 is
1 1
{|x|− 2 , |x|− 2 , |x|0 } so it is obvious that only a bounded range of (say) |x| < 10 need be consid-
ered to check if R is small or F̂I is large.
• For Type II not to be dominated by the Type I implies the ratio R should be very small
(∼ 10−4 ). So it √ would seem that the enhanced symmetry points x = 0, 1/3 and the additional
points x = { 8 }, {0.198437, −0.0992186 ± 2.24266i}, which are the zeros of q2 , q3 may
3±i 7
76 C.S. Aulakh, S.K. Garg / Nuclear Physics B 757 (2006) 47–78

help to strengthen the Type II seesaw and should be added to our list of exceptional points. If
one extracts a factor α2−2 to define R = Rα2−2 one sees that only the zeros of q2 are of any
relevance to strengthening Type II seesaw. Our different parametrization of the problem of dom-
inance together with the extraction of the hidden dependence on the threshold corrections to MX
which is coded in ΔX makes our analysis [1] different from [43]. Differences from [43] in the
points relevant for Type II have emerged: we exclude x = 0 and the zeros of p5 , p2 since they
cancel with the normalizations (which we have not left implicit as in [32]). Moreover we have
used expressions in terms of the actual Yukawa couplings rather than mass eigenvalues which
can conceal compensating growth of Yukawas and coefficients αi necessitating separate consid-
eration of whether one has ventured into strong Yukawa coupling regions. Nevertheless we have
included all such putative special points in our survey and checked that they do not provide the
required exceptional behaviour.
• The functions z2 , z16 have no zeros even when γ , γ̄ are both zero. When γ , γ̄ are non-zero
the normalizing factors never diverge.
• The points x = 0, ±i are the only remaining candidates to strengthen Type I and are already
included in the list of enhanced symmetry points.

References

[1] C.S. Aulakh, MSGUTs from germ to bloom: Towards falsifiability and beyond, hep-ph/0506291.
[2] P. Minkowski, Phys. Lett. B 67 (1977) 110;
M. Gell-Mann, P. Ramond, R. Slansky, in: P. van Niewenhuizen, D.Z. Freedman (Eds.), Supergravity, North-
Holland, Amsterdam, 1979;
T. Yanagida, in: O. Sawada, A. Sugamoto (Eds.), Proceedings of Workshop on Unified Theory and Baryon Number
in the Universe, KEK, 1979;
R.N. Mohapatra, G. Senjanović, Phys. Rev. Lett. 44 (1980) 912.
[3] R.N. Mohapatra, G. Senjanović, Phys. Rev. D 23 (1981) 165;
G. Lazarides, Q. Shafi, C. Wetterich, Nucl. Phys. B 181 (1981) 287.
[4] R.N. Mohapatra, Phys. Rev. D 34 (1986) 3457.
[5] C.S. Aulakh, K. Benakli, G. Senjanović, Reconciling supersymmetry and left–right symmetry, Phys. Rev. Lett. 79
(1997) 2188, hep-ph/9703434.
[6] C.S. Aulakh, A. Melfo, G. Senjanović, Minimal supersymmetric left–right model, Phys. Rev. D 57 (1998) 4174–
4178, hep-ph/9707256.
[7] C.S. Aulakh, A. Melfo, A. Rasin, G. Senjanović, Supersymmetry and large scale left–right symmetry, Phys. Rev.
D 58 (1998) 115007, hep-ph/9712551.
[8] C.S. Aulakh, GUT genealogies for susy seesaw Higgs, in: Dubna 2001, Supersymmetry and Unification of Funda-
mental Interactions, hep-ph/0204098.
[9] C.S. Aulakh, B. Bajc, A. Melfo, A. Rašin, G. Senjanović, SO(10) theory of R-parity and neutrino mass, Nucl. Phys.
B 597 (2001) 89, hep-ph/0004031.
[10] C.S. Aulakh, B. Bajc, A. Melfo, G. Senjanović, F. Vissani, The minimal supersymmetric grand unified theory, Phys.
Lett. B 588 (2004) 196, hep-ph/0306242.
[11] C.S. Aulakh, R.N. Mohapatra, CCNY-HEP-82-4 April 1982, CCNY-HEP-82-4-REV, June 1982, Phys. Rev. D 28
(1983) 217.
[12] T.E. Clark, T.K. Kuo, N. Nakagawa, Phys. Lett. B 115 (1982) 26.
[13] K.S. Babu, R.N. Mohapatra, Phys. Rev. Lett. 70 (1993) 2845.
[14] K. Matsuda, Y. Koide, T. Fukuyama, Can the SO(10) model with two Higgs doublets reproduce the observed
fermion masses?, Phys. Rev. D 64 (2001) 053015, hep-ph/0010026.
[15] K. Matsuda, Y. Koide, T. Fukuyama, H. Nishiura, How far can the SO(10) two Higgs model describe the observed
neutrino masses and mixings?, Phys. Rev. D 65 (2002) 033008;
K. Matsuda, Y. Koide, T. Fukuyama, H. Nishiura, Phys. Rev. D 65 (2002) 079904, hep-ph/0108202, Erratum;
T. Fukuyama, N. Okada, Neutrino oscillation data versus minimal supersymmetric SO(10) model, JHEP 0211
(2002) 011, hep-ph/0205066.
C.S. Aulakh, S.K. Garg / Nuclear Physics B 757 (2006) 47–78 77

[16] N. Oshimo, Antisymmetric Higgs representation in SO(10) for neutrinos, Phys. Rev. D 66 (2002) 095010, hep-
ph/0206239;
N. Oshimo, Model for neutrino mixing based on SO(10), Nucl. Phys. B 668 (2003) 258, hep-ph/0305166.
[17] B. Dutta, Y. Mimura, R.N. Mohapatra, Suppressing proton decay in the minimal SO(10) model, Phys. Rev. Lett. 94
(2005) 091804, hep-ph/0412105;
B. Dutta, Y. Mimura, R.N. Mohapatra, Phys. Lett. B 603 (2004) 35, hep-ph/0406262.
[18] B. Bajc, G. Senjanović, Radiative seesaw: A case for split supersymmetry, Phys. Lett. B 610 (2005) 80, hep-
ph/0411193.
[19] B. Bajc, A. Melfo, G. Senjanović, F. Vissani, Yukawa sector in non-supersymmetric renormalizable SO(10), hep-
ph/0510139.
[20] K.Y. Oda, E. Takasugi, M. Tanaka, M. Yoshimura, Unified explanation of quark and lepton masses and mixings in
the supersymmetric SO(10) model, Phys. Rev. D 59 (1999) 055001, hep-ph/9808241.
[21] B. Bajc, G. Senjanović, F. Vissani, b–tau unification and large atmospheric mixing: A case for non-canonical see-
saw, Phys. Rev. Lett. 90 (2003) 051802, hep-ph/0210207.
[22] H.S. Goh, R.N. Mohapatra, S.P. Ng, Phys. Lett. B 570 (2003) 215;
H.S. Goh, R.N. Mohapatra, S.P. Ng, Phys. Rev. D 68 (2003) 115008.
[23] H.S. Goh, R.N. Mohapatra, S. Nasri, SO(10) symmetry breaking and type II seesaw, Phys. Rev. D 70 (2004) 075022,
hep-ph/0408139.
[24] S. Bertolini, M. Frigerio, M. Malinsky, Fermion masses in SUSY SO(10) with type II seesaw: A non-minimal
predictive scenario, Phys. Rev. D 70 (2004) 095002, hep-ph/0406117;
S. Bertolini, M. Malinsky, On CP violation in minimal renormalizable SUSY SO(10) and beyond, hep-ph/0504241.
[25] K.S. Babu, C. Macesanu, Neutrino masses and mixings in a minimal SO(10) model, hep-ph/0505200.
[26] H. Georgi, C. Jarlskog, Phys. Lett. B 86 (1979) 297.
[27] C.S. Aulakh, PhD Thesis CCNY-CUNY (1983), Intermediate mass scales in supersymmetric gauge theories, UMI
84-01477.
[28] C.S. Aulakh, B. Bajc, A. Melfo, A. Rasin, G. Senjanović, Intermediate scales in supersymmetric GUTs: The survival
of the fittest, Phys. Lett. B 460 (1999) 325, hep-ph/9904352.
[29] B. Bajc, G. Senjanović, F. Vissani, Probing the nature of the seesaw in renormalizable SO(10), Phys. Rev. D 70
(2004) 093002, hep-ph/0402140;
B. Bajc, G. Senjanović, F. Vissani, How neutrino and charged fermion masses are connected within minimal super-
symmetric SO(10), hep-ph/0110310.
[30] C.S. Aulakh, A. Girdhar, SO(10) a la Pati–Salam, hep-ph/0204097, v2 August 2003, v4, 9 February, 2004;
C.S. Aulakh, A. Girdhar, Int. J. Mod. Phys. A 20 (2005) 865.
[31] C.S. Aulakh, A. Girdhar, SO(10) MSGUT: Spectra, couplings and threshold effects, Nucl. Phys. B 711 (2005) 275,
MSGUT a la Pati–Salam: From futility to precision, hep-ph/0405074.
[32] B. Bajc, A. Melfo, G. Senjanović, F. Vissani, The minimal supersymmetric grand unified theory. I: Symmetry
breaking and the particle spectrum, Phys. Rev. D 70 (2004) 035007, hep-ph/0402122.
[33] T. Fukuyama, A. Ilakovac, T. Kikuchi, S. Meljanac, N. Okada, General formulation for proton decay rate in minimal
supersymmetric SO(10) GUT, Eur. Phys. J. C 42 (2005) 191, hep-ph/0401213.
[34] X.G. He, S. Meljanac, Phys. Rev. D 41 (1990) 1620;
D.G. Lee, Phys. Rev. D 49 (1994) 1417.
[35] V.V. Dixit, M. Sher, Futility of high precision SO(10) calculations, Phys. Rev. D 40 (1989) 3765.
[36] W. Marciano, G. Senjanović, Phys. Rev. D 25 (1982) 3092.
[37] M.B. Einhorn, D.R.T. Jones, Nucl. Phys. B 196 (1982) 475.
[38] U. Amaldi, W. de Boer, H. Furstenau, Phys. Lett. B 260 (1991) 447.
[39] C.S. Aulakh, Truly minimal unification: Asymptotically strong panacea?, hep-ph/0207150.
[40] C.S. Aulakh, Taming asymptotic strength, hep-ph/0210337.
[41] T. Fukuyama, A. Ilakovac, T. Kikuchi, S. Meljanac, N. Okada, SO(10) group theory for the unified model building,
J. Math. Phys. 46 (2005) 033505, hep-ph/0405300.
[42] C.S. Aulakh, MSGUT: From futility to precision, hep-ph/0410308.
[43] B. Bajc, A. Melfo, G. Senjanović, F. Vissani, Fermion mass relations in a supersymmetric SO(10) theory, hep-
ph/0511352.
[44] C.S. Aulakh, Consistency of the minimal supersymmetric GUT spectra, Phys. Rev. D 72 (2005) 051702, hep-
ph/0501025.
[45] L.J. Hall, Nucl. Phys. B 178 (1981) 75.
[46] C.S. Aulakh, S.K. Garg, in preparation.
78 C.S. Aulakh, S.K. Garg / Nuclear Physics B 757 (2006) 47–78

[47] C.S. Aulakh, Fermion mass hierarchy in the Nu MSGUT. I: The real core, hep-ph/0602132.
[48] M.D. Tiwana, Higgs mass matrices in the NMSGUT, MSc. Thesis, Panjab University, April 2004.
[49] K.S. Babu, J.C. Pati, F. Wilczek, Suggested new modes in supersymmetric proton decay, Phys. Lett. B 423 (1998)
337, hep-ph/9712307.
[50] S. Bertolini, T. Schwetz, M. Malinsky, Fermion masses and mixings in SO(10) models and the neutrino challenge
to SUSY GUTs, hep-ph/0605006.
Nuclear Physics B 757 (2006) 79–116

Ten-dimensional supersymmetric Janus solutions


Eric D’Hoker, John Estes, Michael Gutperle ∗
Department of Physics and Astronomy, University of California, Los Angeles, CA 90095, USA
Received 3 April 2006; received in revised form 3 August 2006; accepted 18 August 2006
Available online 11 September 2006

Abstract
The reduced field equations and BPS conditions are derived in Type IIB supergravity for configurations
of the Janus type, characterized by an AdS4 -slicing of AdS5 , and various degrees of internal symmetry and
supersymmetry. A generalization of the Janus solution, which includes a varying axion along with a varying
dilaton, and has SO(6) internal symmetry, but completely broken supersymmetry, is obtained analytically in
terms of elliptic functions. A two-parameter family of solutions with 4 real supersymmetries, SU(3) internal
symmetry, a varying axion along with a varying dilaton, and non-trivial B(2) field, is derived analytically
in terms of genus 3 hyper-elliptic integrals. This supersymmetric solution is the 10-dimensional Type IIB
dual to the N = 1 interface super-Yang–Mills theory with SU(3) internal symmetry previously found in the
literature.
© 2006 Published by Elsevier B.V.

1. Introduction

The AdS/CFT correspondence relates string theories on anti-de Sitter space–times (AdS) to
conformal field theories (CFT) on the boundary of the AdS space–time [1–3]. The case which is
understood perhaps best relates Type IIB string theory on AdS5 × S5 to Yang–Mills theory with
N = 4 supersymmetry and gauge group SU(N ). (For reviews, see [4,5].)
The AdS/CFT correspondence is expected to hold as well in situations with less or no super-
symmetry, and with space–times which are only asymptotically AdS. Thus, a suitable deforma-
tion on the gauge theory side of the correspondence should be dual to an associated deformation
on the string theory side and vice versa. Many interesting such cases are known and have been
studied extensively. These include the holographic representation of renormalization group-flows

* Corresponding author.
E-mail address: gutperle@physics.ucla.edu (M. Gutperle).

0550-3213/$ – see front matter © 2006 Published by Elsevier B.V.


doi:10.1016/j.nuclphysb.2006.08.017
80 E. D’Hoker et al. / Nuclear Physics B 757 (2006) 79–116

[6], and the solutions of Klebanov–Strassler [7], Polchinski–Strassler [8] and Maldacena–Nunez
[9].
In [10], a dilatonic deformation of the Type IIB background AdS5 × S5 was found.1 The Janus
solution breaks all supersymmetries, but is nevertheless stable against small and a large class
of large perturbations [11–13]. It can be viewed as a curved dilatonic domain wall [14]. The
holographic dual to the Janus solution is N = 4 super-Yang–Mills theory in 3 + 1 space–time
dimensions, with a planar (2 + 1)-dimensional interface, across which the gauge coupling varies
discontinuously. The interface carries no additional degrees of freedom. Conformal invariance
in 2+1 dimensions is preserved by the interface at the classical level, and holds in conformal
perturbation theory to first non-trivial order as well [15].
The Janus solution is remarkably simple.2 In fact, in this paper, we shall show that, even when
generalized to include a varying axion in addition to a varying dilaton, the Janus solution admits
an analytic form in terms of elliptic functions. This raises the hope that correlation functions in
the Janus background may be studied using analytic methods, a topic which we plan to address
in a later publication.
Furthermore, the remarkable simplicity of the non-supersymmetric Janus solution suggests
that Janus may possess supersymmetric generalizations available in analytic form as well.
Clearly, such analytic solutions would be valuable as starting points for the analytic study of
correlators in the corresponding backgrounds. (Note that, on the one hand, the backgrounds of
Klebanov–Strassler [7] or Maldacena–Nunez [9] are not asymptotically AdS while, on the other
hand, the Polchinski–Strassler [8] solution is only known approximately, in an expansion in the
strength of the fluxes.)
The fact that interesting supersymmetric generalizations of Janus should exist is further sug-
gested by considering its CFT dual, namely (3 + 1)-dimensional super-Yang–Mills theory with a
(2 + 1)-dimensional planar interface. In [15], it was found that 2 Poincaré supersymmetries can
be preserved by adding “interface operators” whose support is confined to the interface. In the
conformal limit, 2 conformal supersymmeries emerge as well. The interface operators break the
R-symmetry from SO(6) down to SU(3).
In [19], a complete classification of supersymmetry restoring interface operators on the gauge
theory side is given. In particular, it is established there that, for interface theories with SU(3)
internal symmetry, 4 is the maximum number of conformal supersymmetries. The corresponding
theory is constructed explicitly, and coincides with the one presented in [15] in terms of N = 1
off-shell fields. It is further established in [19] that interface theories with extended supersym-
metry exist as well. One interface theory has 8 conformal supersymmetries and SO(2) × SU(2)
internal symmetry, while another has 16 conformal supersymmetries and SU(2) × SU(2) internal
symmetry.
Further evidence for the existence of 10-dimensional supersymmetric generalizations of Janus
is provided in [20], where a supersymmetric Janus solution of five-dimensional gauged super-
gravity was found (building on previous work on curved domain walls in AdS5 [21–25]). The
starting point of [20] was an SU(3) invariant gauging of the universal hypermultiplet of five-
dimensional N = 2 supergravity [26]. While it is believed that the resulting theory is a consistent

1 It was named the Janus solution, after the two-faced Roman god of gates, doors, beginnings, endings and, now, also,
string dualities.
2 Other dilatonic deformations found in the literature are singular and their physical interpretation remains more ob-
scure [16–18].
E. D’Hoker et al. / Nuclear Physics B 757 (2006) 79–116 81

truncation of the ten-dimensional supergravity, the details of this truncation, and hence of any
possible lift of the solution to ten dimensions, are unknown.
In the present paper, a family of supersymmetric Janus solutions is derived directly in ten-
dimensional Type IIB supergravity. We focus here on Janus solutions which provide holographic
duals to the supersymmetric interface CFTs of [15], with N = 1 interface supersymmetry and 3
chiral multiplets related by SU(3) internal symmetry. Our starting point is the construction of the
most general Ansatz for Type IIB supergravity fields which preserves SO(2, 3) × SU(3) symme-
try, and which transforms covariantly under the SL(2, R) symmetry of Type IIB supergravity, as
well as under the unique U (1)β ⊂ SU(4) which commutes with SU(3). The reduced field equa-
tions are then solved subject to the BPS conditions, namely the conditions for the vanishing of
the supersymmetry variations of the dilatino and gravitino fields. The resulting family of solu-
tions contains a subset that is of the Janus type, and the solutions in this subset may be expressed
analytically via hyper-elliptic integrals of genus 3.
The Ansatz that will be obtained in this paper for the construction of supersymmetric Janus
solutions is based on an AdS4 slicing of AdS5 , just as the original non-supersymmetric Janus
Ansatz was. If a slicing of AdS5 by 4-dimensional Minkowski space were used instead, one
would recover an Ansatz used by Romans [42] to construct SU(3)-symmetric compactifications
of Type IIB supergravity, but without supersymmetry. (In eleven-dimensional supergravity, the
corresponding solutions were constructed in [43].) In AdS/CFT such Minkowski-sliced solutions
have a natural interpretation in terms of RG flows [6,31,45,47]. Some techniques used for the
study of the Minkowski slicings can be applied to the AdS4 slicings, and the resulting reduced
field equations are related.3
The remainder of the paper is organized as follows.
In Section 2, the field equations of Type IIB supergravity, as well as the supersymmetry varia-
tions (both for vanishing fermion fields) are summarized. In Section 3, the original Janus solution
is reviewed, extended to include a varying axion along with a varying dilaton, and expressed ana-
lytically in terms of elliptic functions. In Section 4, the main results on supersymmetric interface
CFT are collected.
In Section 5, the most general Ansatz for Type IIB supergravity fields, subject to SO(2, 3) ×
SU(3) × U (1)β × SL(2, R) symmetry is constructed. The reduced Bianchi identities and field
equations for this Ansatz are derived in Section 6, where it is shown that these equations may
also be obtained from a reduced action, which is computed, and a vanishing Hamiltonian con-
straint, which amounts to the reduced Wheeler–De Wit equation. In Section 7, the supersymmetry
variations for the Ansatz are derived. In Section 8, it is demonstrated that every supersymmet-
ric solution with varying axion and varying dilaton is actually the SL(2, R) image of a “real”
solution with vanishing axion.
In Section 9, it is shown that, for supersymmetric solutions, the vanishing condition of the
Hamiltonian constraint factorizes into a product of two factors. The vanishing of the first factor
leads to a family of degenerate solutions (for which the second factor vanishes as well), and it
is these solutions which are obtained analytically in terms of genus 3 hyper-elliptic integrals. In
Section 10, it is shown that these degenerate solutions are of the Janus type, and thus asymp-
totically AdS. The equations for the non-degenerate solutions are more involved and have not
yet been solved analytically. Numerical evidence suggests that these solutions may not be of the
Janus type. In Section 11, the holographic dual CFT is interpreted in terms of deformations of

3 We thank the referee for pointing out this relationship.


82 E. D’Hoker et al. / Nuclear Physics B 757 (2006) 79–116

the AdS5 × S 5 background, while in Section 12, some concluding remarks are offered. Finally, a
convenient basis of Dirac matrices is presented in Appendix A.

2. Type IIB supergravity

In this section, we present the field equations, Bianchi identities, and supersymmetry varia-
tions for Type IIB supergravity, which were originally derived in [27,28]. The metric signature
used here is (−+ · · · +) in contrast with [27], where the signature is (+− · · · −). We restrict to
vanishing fermion fields, which will suffice for our analysis.
The bosonic fields of Type IIB supergravity are: the metric gMN ; the axion-dilaton complex
scalar B which takes values in the coset SU(1, 1)/U (1); and the antisymmetric tensors B(2)
(which is complex) and C(4) (which is real). It is standard to introduce composite fields in terms
of which the field equations are expressed simply. They are as follows,4
1
P = f 2 dB, f= ,
1 − |B|2
Q = f 2 Im(B d B̄) (2.1)
and the field strengths F(3) = dB(2) , and
 
G = f F(3) − B F̄(3) ,
i
F(5) = dC(4) + (B(2) ∧ F̄(3) − B̄(2) ∧ F(3) ). (2.2)
16
The scalar field B is related to the axion χ and dilaton φ fields by
1 + iτ
B= , τ = τ1 + iτ2 = χ + ie−φ . (2.3)
1 − iτ
In terms of the composite fields P , Q, and G, there are “Bianchi identities” given as follows,

dP − 2iQ ∧ P = 0, (2.4)
dG − iQ ∧ G + P ∧ Ḡ = 0, (2.5)
dQ + iP ∧ P̄ = 0, (2.6)
1
dF(5) − i G ∧ Ḡ = 0. (2.7)
8
The field strength F(5) is required to be self-dual,

F(5) = ∗F(5) . (2.8)


The field equations are given by5
1
0 = ∇ M PM − 2iQM PM + GMNP GMNP , (2.9)
24

4 Throughout, we shall pass freely between tensor and form notations, with a differential form ω of rank n associated

with tensor components ωM1 ···Mn by the relation ω = n! M Mn .
M1 ···Mn dx 1 ∧ · · · ∧ dx
5 The sign of the term GG in (2.9) has been corrected compared to the original equation in [27]; the need for this
correction was noted independently in [29–31].
E. D’Hoker et al. / Nuclear Physics B 757 (2006) 79–116 83

2
0 = ∇ P GMNP − iQP GMNP − P P ḠMNP + iF(5)MN P QR GP QR , (2.10)
3
1 2 
0 = RMN − PM P̄N − P̄M PN − F(5) MN
6
1  1
− GM P Q ḠN P Q + ḠM P Q GN P Q + gMN GP QR ḠP QR . (2.11)
8 48
The fermionic fields are the dilatino λ and the gravitino ψM , both of which are complex Weyl
spinors with opposite 10-dimensional chiralities, given by Γ11 λ = λ, and Γ11 ψM = −ψM . The
supersymmetry variations of the fermions (still in a purely bosonic background) are
i MNP
δλ = iPM Γ M B −1 ε ∗ − Γ GMNP ε,
24
i
δψM = Dμ ε + F(5)N P QRS Γ N P QRS ΓM ε
480
1  NP Q 
+ Γ GN P Q − 9Γ N P GMNP B −1 ε ∗ , (2.12)
96 M
where B is the charge conjugation matrix of the ten-dimensional Clifford algebra.6
Type IIB supergravity is invariant under SU(1, 1) ∼ SL(2, R) symmetry, which leaves gμν
and C(4) invariant, acts by Möbius transformation on the field τ , and linearly on B(2) ,
    
aτ + b Im B(2) a b Im B(2)
τ→ , → (2.13)
cτ + d Re B(2) c d Re B(2)
with a, b, c, d ∈ R and ad − bc = 1. In this non-linear realization of SL(2, R), the field B takes
values in the coset SU(1, 1)/U (1) ∼ SL(2, R)/U (1), and the fermions λ and ψμ transform lin-
early under the isotropy gauge group U (1) with composite gauge field Q.
The field equations derive from an action, (we omit the overall prefactor 1/2κ10 2 ),

  
√ M
1 ∂M τ ∂ τ̄ 1
S = dx g R − 2 − G MNP Ḡ MNP
− 4|F (5) |2
− i C(4) ∧ F(3) ∧ F̄(3)
(Im τ )2 12
(2.14)
in the following sense. The field equations are derived by first requiring that S be extremal under
arbitrary variations of the fields gMN , τ , B(2) and C(4) ; and second by imposing the self-duality
condition (2.8) on F(5) as a supplementary equation.

3. The generalized non-supersymmetric Janus solution

In this section, the original Janus solution of [10] is reviewed, extended to include a varying
axion along with a varying dilaton, and expressed analytically in terms of elliptic functions. The
Ansatz is required to have SO(2, 3) × SO(6) symmetry. Since AdS4 × R × S 5 admits no SO(6)-
invariant 2-forms, the symmetry requires B(2) = 0. The metric gMN and 5-form F(5) are given
by an AdS4 -slicing of AdS5 , consistent with SO(2, 3) × SO(6) symmetry,
ds 2 = h dμ2 + h dsAdS
2
4
+ h1 dsS25 ,
5/2
F5 = 2h5/2 dμ ∧ ωAdS4 + 2h1 ωS 5 , (3.1)

6 It is defined by BB ∗ = I and BΓ M B −1 = (Γ M )∗ ; see Appendix A for our Γ -matrix conventions. Throughout,


complex conjugation of functions will be denoted by bar, while that of spinors will be denoted by star.
84 E. D’Hoker et al. / Nuclear Physics B 757 (2006) 79–116

where ωAdS4 and ωS 5 are the canonical volume forms on the corresponding manifolds. The dila-
ton φ, axion χ and the functions h, h1 depend on μ only. Remarkably, the original Janus solution
was found by setting h1 = 1, thereby leaving the S 5 metric unchanged. Numerical evidence sug-
gests that solutions with varying h1 always have singularities. In Section 3.3, we shall present
arguments, based on the AdS/CFT correspondence, that such breathing modes which vary h1 ,
must be absent lest the corresponding solution become singular.
The reduced field equations (for h1 = 1) may be expressed in terms of τ and h,
τ

i 3 h

+ τ
+ = 0, (3.2)
τ τ2 2h

|2
4(h
)2 − 4hh

+ 8h3 = h2 , (3.3)
τ22
12h2 + (h
)2 + 2hh

− 16h3 = 0. (3.4)
Since h is real, the imaginary part of (3.2) is independent of h and may be integrated to give
|τ − p|2 = r 2 , p, r ∈ R. (3.5)
This means that as μ varies, τ evolves along a segment of a geodesic in the upper half plane
equipped with the SL(2, R)-invariant Poincaré metric7 |dτ |2 /τ22 . Integrating also the real part of
(3.2), we find that |τ
|2 /τ22 = c02 / h3 . for some constant c0 ∈ R. This relation gives the velocity
of τ along the geodesic as a function of h. Finally, using this result in (3.3) leads to an equation
that is consistent with (3.4) and has a first integral given by,
c02
(h
)2 = 4h3 − 4h2 + (3.6)
6h
which is the same equation as the first integral for the original Janus solution [10].
The interpretation of the above results is as follows. In the original Janus solution, the axion
field vanished, and the dilaton evolution spanned a rather special geodesic in the upper half
plane: a vertical line segment with τ1 = 0. Since the action of SL(2, R) on the upper half plane
is transitive on points as well as on connected geodesic segments of equal length, solutions
with varying axion may be obtained as SL(2, R) images of solutions with vanishing axion. The
remarkable result obtained above is that all solutions with varying axion may be obtained as
SL(2, R) images of solutions with vanishing axion. (See Fig. 1.)

3.1. Analytical solution in terms of elliptic functions

In [10], the quadrature of (3.6) was obtained numerically. Actually, the general solution may
be expressed analytically in terms of elliptic functions. To do so, we consider a coordinate inde-
pendent object, namely the 1-form,
dμ dh
ν=√ =
. (3.7)
h 4h4 − 4h3 + c02 /6
The 1-form ν is proportional to the holomorphic Abelian differential on a torus. We represent the
constant c0 by c02 = 24γ03 (1 − γ0 ), and parametrize the Abelian differential ν and the function h

7 Its geodesics are the half circles with arbitrary center p on the real axis and arbitrary radius r.
E. D’Hoker et al. / Nuclear Physics B 757 (2006) 79–116 85

Fig. 1. Mapping geodesic segments under SL(2, R) in the dilaton/axion upper half plane.

in terms of γ0 and the Weierstrass ℘-function, expressed in terms of the canonical coordinate z
of the torus. We find the following expressions,
dz γ0 (3 − 4γ0 )
ν=√ , h(μ) = γ0 + . (3.8)
γ0 ℘ (z) + 2γ0 − 1
Here, the Weierstrass ℘-function has been normalized to standard form, (∂z ℘)2 = 4℘ 3 −
16γ0 (1 − γ0 )℘ − 4(1 − γ0 ). The discriminant of the corresponding curve Δ = 64c04 (32c02 −
81)/27 vanishes at c02 = 0 and at the critical value c02 = 81/32, identified in [10] as the value
where the range of the dilaton begins to diverge. The dilaton axion equation reduces to
ν
|dτ |/τ2 = c0 (3.9)
h
which may be integrated by standard elliptic function methods.

3.2. Structure of the AdS/CFT dual

The Janus solution can be viewed as a dilatonic domain wall in which the dilaton varies
with the coordinate μ, which parameterizes the AdS4 slicing of AdS5 . (See Fig. 2.) It follows
from the dilatino supersymmetry variation (2.12), that no supersymmetries are preserved for the
Janus solution with a varying dilaton and vanishing B(2) . Nevertheless, in [11–13] convincing
arguments were presented that the Janus solution is stable against all small and a certain class of
large perturbations.
The angular coordinate μ covers a range μ ∈ [−μ0 , μ0 ] where μ0 > π/2. The structure of
the boundary of this space can be analyzed using global coordinates for the AdS4 slices. Near
μ = ±μ0 the non-compact part of the metric has the following asymptotic behavior,
1  2 
ds 2 ∼ cos λ dμ2 − dt 2 + dλ2 + sin2 λ dΩS22 (3.10)
(μ ∓ μ0 )2 cos2 λ
with 0  λ < π2 . The constant time section of the boundary is a non-singular geometry, consisting
of two halves of S 3 at μ = ±μ0 , joined at the pole of S3 where λ = π/2. In Poincaré coordinates
for the AdS4 slices, the spatial section of the boundary consists of two three-dimensional half
planes joined by a two-dimensional interface. The dilaton varies continuously with μ and takes
86 E. D’Hoker et al. / Nuclear Physics B 757 (2006) 79–116

Fig. 2. Sketch of the boundary geometry of the Janus solution in global and Poincaré coordinates for the AdS4 slices.

two different constant values at the boundary,


(0) (1)
lim φ(μ) = φ± + φ± (μ ∓ μ0 )4 + O (μ ∓ μ0 )8 . (3.11)
μ→±μ0

The holographic dual gauge theory CFT of the Janus solution was proposed in [10] and analyzed
in detail in [15]. The CFT dual is a planar interface theory. The action on both sides of the inter-
face is the standard N = 4 SYM action but the coupling constant varies discontinuously across
the interface. The symmetry SO(2, 3) of the Janus solution maps to the conformal symmetry of
a planar interface on the CFT side. This symmetry is manifest at the classical level, but was also
shown to persist at the first non-trivial quantum level [15]. The SO(6) symmetry of the Janus
solution maps to an (accidental) internal symmetry on the CFT side. Note that, in contrast to the
defect conformal field theories examined earlier in the context of AdS/CFT [32–36], the CFT
dual to the Janus solution is characterized by an interface that carries no degrees of freedom in
addition to the ones inherited from the bulk N = 4, whence the name “interface”, as opposed to
“defect”.

3.3. The absence of breathing modes

The AdS/CFT dictionary relates the breathing mode of S 5 , described by the function h1 in the
(20)
Janus metric of (3.1), to a dimension 8 operator Ok=0 = tr(F+2 F−2 ) (in the notation of [4]). The
argument below will show that the breathing mode cannot be excited in the Janus solution, and
thus the function h1 must be a constant. On the one hand, a non-vanishing source for the operator
Ok=0 would correspond to a behavior (μ − μ0 )−4 near the AdS boundary, as μ → μ0 . Such a
(20)

behavior would lead to a singular 10-dimensional metric. Since the Janus solution is regular, this
source must be absent. On the other hand, a non-vanishing expectation value would correspond to
a behavior (μ − μ0 )8 near the boundary. A power series expansion near the boundary of AdS for
the Janus solution, and allowing for the presence of a breathing mode, reveals that such a term is
forced to vanish. By standard AdS/CFT arguments the breathing mode is therefore exactly zero.

4. Supersymmetric interface CFT

In [15], it was shown that preserving supersymmetry in a planar interface Yang–Mills theory
necessarily leads to the breaking of the internal SO(6). In turn, arguments are presented in [15]
that 2 Poincaré supercharges may be preserved upon reducing SO(6) to SU(3), and including
E. D’Hoker et al. / Nuclear Physics B 757 (2006) 79–116 87

certain “interface counterterms” which are built out of the fields of the bulk N = 4 super-Yang–
Mills theory.
Specifically, in [15], the Yang–Mills coupling g(x π ) is assumed to be a function of the coor-
dinate x π (the CFT side coordinate corresponding to the coordinate μ of the AdS side) and to
vary across the interface located at x π = 0. The Lagrangian is formulated with N = 1 auxiliary
fields for a single chiral multiplet and a single gauge multipet,
 
i
i

 
Lchiral = −∂μ φ̄∂ φ − ψ̄γ ∂μ ψ + F̄ F + W F − W ψ̄ 1 + γ ψ + c.c. ,
μ μ 5
2 4
1 a aμν i 1
Lgauge = − 2 Fμν F − 2 λ̄a γ μ Dμ λa + 2 D a D a . (4.1)
4g 2g 2g
Here, ψ and λa are Majorana spinors, φ is a complex scalar, F and D a are auxiliary fields, and
W is the superpotential of the chiral multiplet. In Lchiral , all dependence on the coupling g is
contained in W . Upon adding to Lchiral and Lgauge the following “interface counterterms”,
 
∂W ∂ W̄
δLchiral = i∂π g − ,
∂g ∂g
 
  i π μν a 1
δLgauge = −∂π g −2 ε̄γ γ λ Fμν + ε̄γ π γ 5 λa D a (4.2)
4 2
the combined Lagrangians are invariant under the supersymmetry generated by spinors ε satis-
fying the interface projection relation, 1/2(1 + iγ 5 γ π )ε = ε. For the N = 4 theory, we have
W ∼  ij k Φ i Φ j Φ k , where Φ k are 3 complex fields, Φ k = φ 2k−1 + iφ 2k . This theory has SU(3)
internal symmetry, which rotates the three chiral multiplets into one another.
Surprisingly, the derivation of the existence of the N = 1 interface supersymmetry presented
in [15] seems to depend on the precise normalizations of the chiral and gauge multiplet La-
grangians: canonical for the chiral multiplet, but with the gauge coupling factored out for the
gauge multiplet.
In a companion paper [19], the existence of supersymmetry in the N = 4 super-Yang–Mills
theory with an interface and with “interface counterterms” is solved in generality. We confirm that
the model of [15] indeed possesses N = 1 interface supersymmetry and SU(3) global symmetry,
independently of any normalization issues. Further models with supersymmetry, including one
with N = 4 interface supersymmetry and SO(4) internal symmetry, are also discovered in [19]
along with a complete classification of all possible supersymmetries.
It may be helpful to clarify the counting of the number of preserved supersymmetries. Since
the interface field theory has only (2 + 1)-dimensional Poincaré invariance, the counting of su-
persymmetries is conveniently carried out from a three-dimensional point of view. An N = 1
interface supersymmetry corresponds to 2 real Poincaré supercharges. When, in addition, the
interface field theory is conformal invariant, the number of supercharges is double the number
of Poincaré supercharges. Thus, the N = 1 interface CFT has altogether 4 real supercharges. In
the dual supergravity this means that we look for, and find, a two-parameter family of solutions
which preserve 4 real supersymmetries.

5. The ten-dimensional Janus Ansatz

In this section, we shall construct the most general Ansatz for Type IIB supergravity fields,
consistent with the symmetries of the expected CFT dual theory with N = 1 interface super-
88 E. D’Hoker et al. / Nuclear Physics B 757 (2006) 79–116

symmetry and global SU(3) symmetry relating the 3 chiral multiplets inherited from the parent
N = 4 theory.

5.1. Symmetries of the Ansatz

For given gauge and interface couplings, the dual CFT has SO(2, 3) conformal and SU(3)
internal symmetry, along with 2 Poincaré and 2 conformal supersymmetries. Therefore, on the
AdS side, we shall seek an Ansatz for the supersymmetric generalization of the Janus solution
which is invariant under the following bosonic symmetries,

SO(2, 3) × SU(3). (5.1)


The supersymmetries will be achieved later by enforcing the BPS conditions. In analogy with
Janus, the solution is expected to depend continuously on at least one parameter (for Janus, this
is the constant c0 ) and include the undeformed AdS5 × S 5 as a limiting case (for Janus, c0 → 0).
Therefore, the topology of the internal space is expected to remain the same and equal to the
topology of S 5 .
The interface CFT naturally consists of a family of theories. This is familiar from the parent
N = 4 theories, which are labeled by the gauge coupling g and the instanton angle θ , and are
mapped into one another by the standard action of SL(2, Z) on g and θ . Montonen–Olive duality
states that two theories related by SL(2, Z) are physically the same. On the AdS side, SL(2, Z)
degenerates to SL(2, R), which constitutes a symmetry of Type IIB supergravity. Therefore, on
the AdS side, we shall seek an Ansatz which forms a family on which SL(2, R) acts consistently
as well. For example, our previous generalization of Janus, which includes the axion and the
dilaton, is such a family of Ansätze, while the original Janus Ansatz clearly is not, since SL(2, R)
does not act consistently on it.
The interface CFT is also naturally a family of theories in terms of its interface couplings.
Indeed, the interface couplings of the CFT dual theory of [15] and [19] are mapped into one
another by SU(4). Theories with different interface couplings related in this way by SU(4) are
physically the same. On the AdS side, not all of these SU(4) transformations can be implemented
in a useful way. Once an Ansatz has been forced to be invariant under SU(3), the embedding of
SU(3) in SU(4) is fixed, and the only useful transformations of SU(4) that can be implemented
on the SU(3)-invariant Ansatz are those that commute with SU(3). This is a single generator,
spanning a group U (1)β , the notation of which will be motivated later on. To summarize, we shall
seek an Ansatz for the supersymmetric generalization of the Janus solution which is invariant
under the group

SO(2, 3) × SU(3) × U (1)β × SL(2, R) (5.2)


in the sense described above.
The presence of the SO(2, 3) factor requires the Ansatz to be an AdS4 slicing of AdS5 , as was
already the case for the original Janus. The group SU(3) × U (1)β must be realized as an isometry
of the 5-dimensional internal space. The only 4-dimensional manifold with SU(3) isometry is
CP2 = SU(3)/SU(2) × U (1). The product space CP2 × S 1 thus has the correct isometry SU(3) ×
U (1)β , but not the correct topology of S 5 . The topology of S 5 is easily recovered by recalling that
S 5 is the total space of a S 1 bundle over CP2 [42–45]. A natural candidate internal space for our
Ansatz is CP2 ×q S 1 , where ×q produces the S 1 bundle over CP2 with integer first Chern class
q ∈ Z. The requirement that this space have the topology of S 5 reduces this choice to q = ±1,
E. D’Hoker et al. / Nuclear Physics B 757 (2006) 79–116 89

so we may simply set q = 1. (Note that another candidate internal space SU(3)/SO(3) has the
topology of S 5 , but its isometry does not include the U (1)β factor.)
Based on these symmetry considerations, we conclude that the Ansatz must be constructed on
the space

R × AdS4 × CP2 ×1 S 1 , (5.3)


where CP2 ×1 S 1 is the sphere S 5 , deformed while preserving SU(3) × U (1)β isometry.

5.2. Invariant metrics and frames on CP2

In this section, we summarize important properties of CP2 and its S 1 fiber bundle S 5 , and
derive all the invariants needed for the construction of the invariant Ansatz. Recall that both
spaces may be viewed as symmetric spaces via the following cosets S 5 = SO(6)/SO(5) and
CP2 = SU(3)/S(U (2) × U (1)). But, S 5 may also be viewed as a non-symmetric homogeneous
space via the coset S 5 = SU(3)/SU(2), from which it is clear that S 5 is the total space of a
S 1 = U (1) bundle over CP2 .
The space CP2 may be parametrized locally by two complex coordinates ζ1 , ζ2 ∈ C or, equiv-
alently, by four real angles α, θ, φ, ψ, related to one another by
i
ζ1 = tan(α) cos(θ/2)e 2 (ψ+φ) ,
i
ζ2 = tan(α) sin(θ/2)e 2 (ψ−φ) . (5.4)
The Fubini–Study metric is given by
¯  
2
dsCP2
= gi j¯ dζ i ⊗ d ζ̄ j , gi j¯ = ∂i ∂j¯ ln 1 + |ζ1 |2 + |ζ2 |2 Q . (5.5)
It is useful to work with an orthonormal frame êa , a = 6, 7, 8, 9 on CP2 which may be expressed
in terms of the angular coordinates,8

ê6 = dα,
1
ê7 = sin(2α)σ3 , σ3 = dψ + cos(θ ) dφ,
4
1
ê8 = sin(α)σ1 , σ1 = − sin(ψ) dθ + cos(ψ) sin(θ ) dφ,
2
1
ê = sin(α)σ2 ,
9
σ2 = cos(ψ) dθ + sin(ψ) sin(θ ) dφ. (5.6)
2
The 1-forms σ 1 , σ 2 , σ 3 span the frame of S 3 . We shall also make use of a basis in which the
complex structure of CP2 is manifest, and introduce a complex splitting of the frames,

êz1 = ê6 + i ê7 , êz2 = ê8 + i ê9 ,


êz̄1 = ê6 − i ê7 , êz̄2 = ê8 − i ê9 . (5.7)
The metric may be expressed in terms of the above coordinates and frames by,

8 We choose the labels 6, 7, 8, 9 for internal labels, as it is in this manner that the indices will be embedded in
10-dimensional space–time.
90 E. D’Hoker et al. / Nuclear Physics B 757 (2006) 79–116

1 2  2 
2
dsCP2
= dα 2 + sin (α) σ1 + σ22 + cos2 (α)σ32
4

9
= êa ⊗ êa = êz1 ⊗ êz̄1 + êz2 ⊗ êz̄2 . (5.8)
a=6

We shall also need the torsion-free connection associated with this frame, satisfying

d êa + ω̂a b ∧ êb = 0. (5.9)


These connection components are given by,
1 1
ω̂6 7 = − cos(2α)σ3 , ω̂6 8 = +ω̂7 9 = − cos(α)σ1 ,
2  2
1 1
ω̂ 9 = 1 − cos2 (α) σ3
8
ω̂ 9 = −ω̂ 8 = − cos(α)σ2 .
6 7
(5.10)
2 2

5.3. Invariant 2-forms on CP2 and S 5

In this subsection, we shall obtain the most general 2-forms on S 5 , invariant under SU(3), and
use them to build an SU(3) × U (1)β -invariant Ansatz for the antisymmetric tensor field B(2) in
the subsequent section. It is well know that the cohomology of CP2 is generated by the Kähler
form K, which derives from a U (1)-connection A1 by K = dA1 . In terms of the coordinates of
CP2 introduced earlier, these forms are given by,
¯1 1
K = igi j¯ dζ i ∧ d ζ̄ j = sin(2α)dα ∧ σ3 + sin2 (α)σ1 ∧ σ2 ,
2 2
i ζ̄ i dζ i − ζ i d ζ̄ i 1
A1 = − = sin2 (α)σ3 . (5.11)
2 1 + |ζ1 | + |ζ2 |
2 2 2
In terms of the frames êa , we have

K = 2ê6 ∧ ê7 + 2ê8 ∧ ê9 = i êz1 ∧ êz̄1 + i êz2 ∧ êz̄2 . (5.12)


The Kähler form is simply related to the volume form as follows, K 2 = 8ê6 ∧ ê7 ∧ ê8 ∧ ê9 .
The five sphere S 5 is constructed as a S 1 fibration over CP2 . Introducing the S 1 coordinate
β, the isometry group U (1)β acts on S 1 by shifts of β. The round metric on S 5 is then given in
terms of this fibration by,

dsS25 = (dβ + A1 )2 + dsCP


2
2
, (5.13)
where A1 is the Kähler one form defined in (5.11). We introduce a fifth frame component

ê5 = dβ + A1 . (5.14)
Under the action of SU(3), the connection A1 shifts by a non-trivial exact form, which may be
compensated for by the opposite shift in β, so that the combination dβ + A1 is invariant under
SU(3). It is of course also invariant under constant shifts of β, forming the group U (1)β . The
frame ê5 , ê6 , ê7 , ê8 , ê9 is an orthonormal frame for S 5 .
Clearly, the Kähler form K on CP2 is invariant under SU(3) and will be a candidate for an
invariant 2-form in the Ansatz for the antisymmetric tensor field B(2) . To ensure a consistent
Ansatz, however, we shall need the most general 2-form invariant under SU(3) on the deformed
E. D’Hoker et al. / Nuclear Physics B 757 (2006) 79–116 91

sphere S 5 , and this requires an exhaustive study of all SU(3)-invariant 2-forms on S 5 . In this
task, we are helped by two theorems (see for example the Corollaries 1.18 and 1.14 of [37], as
well as [38]), valid for compact, connected G and H ,

Theorem 1. The ring of G-invariant n-forms on a homogeneous space G/H is obtained from a
basis of left-invariant 1-forms θ a , a = 1, . . . , dim G on G, by constant tensors ωa1 a2 ···an , which
vanish whenever ai ∈ H, and are invariant under H, by the expression,

ω(n) = ωa1 a2 ···an θ a1 ∧ · · · ∧ θ an . (5.15)

Theorem 2. On a symmetric space G/H , every G-invariant form is closed.

Given that S 5 = SO(6)/SO(5) is a symmetric space coset, and that H 2 (S 5 , R) = 0, there are
no SO(6)-invariant 2-forms on S 5 , a fact that was used to set B(2) = 0 in the original Janus
solutions. Given, that the coset CP2 = SU(3)/S(U (2) × U (1)) is a symmetric space, and that
H 2 (CP2 , R) is generated by a single element, namely the Kähler form K, we conclude that K is
the unique SU(3)-invariant 2-form on CP2 . These facts are standard.
Finally, we wish to obtain all SU(3)-invariant 2-forms on S 5 . To this end, we express S 5 as
the coset S 5 = SU(3)/SU(2), where SU(2) is embedded in SU(3) in the 2-dimensional rep-
resentation. This coset is not a symmetric space, so that SU(3)-invariant forms need not be
closed. It is now straightforward to obtain all such invariant forms, using Theorem 1. We need
all SU(2)-invariant tensors on the left-invariant 1-forms θ a on SU(3), which vanish on SU(2).
These 1-forms are precisely êz1 , êz2 , êz̄1 , êz̄2 , ê5 constructed earlier. Clearly, we recover the
Kähler form of (5.12) following the construction of Theorem 1 in this manner. There is also the
following 2-form (and its complex conjugate),

1
êz1 ∧ êz2 = εij êzi ∧ êzj (5.16)
2
i,j =1,2

which is not invariant under the action of SU(3) isometries on CP2 , because the form is not
invariant under the U (1) factor of the isotropy group of CP2 . It can be made into a well-defined
SU(3)-invariant form Â2 on S 5 by compensating for the phase factor,

idζ 1 ∧ dζ 2 e3iβ
Â2 ≡ êz1 ∧ êz2 e3iβ = . (5.17)
(1 + |ζ1 |2 + |ζ2 |2 )3/2

Some further useful properties involving Â2 are collected below.

d Â2 = 3i(dβ + A1 ) ∧ Â2 = 3i ê5 ∧ Â2 ,


¯ 1
Â2 ∧ Â2 = K 2 = 4ê6 ∧ ê7 ∧ ê8 ∧ ê9 . (5.18)
2
In particular, the formula for the differential shows that Â2 is indeed the solution to an SU(3)-
invariant equation (d − 3i ê5 )Â2 = 0, which is consistent with the fact that Â2 itself is invariant.
By contrast, the form êz1 ∧ êz2 satisfies a differential equation (d − 3iA1 )(êz1 ∧ êz2 ) = 0 which
is not invariant. Under U (1)β , the form Â2 transforms with a constant phase factor.
92 E. D’Hoker et al. / Nuclear Physics B 757 (2006) 79–116

5.4. The Ansatz for the metric

The Ansatz for the metric follows from the symmetry considerations above,
 
ds 2 = f42 dμ2 + dsAdS
2
4
+ f12 (dβ + A1 )2 + f22 dsCP
2
2
. (5.19)
Invariance of the metric under SO(2, 3) × SU(3) × U (1)β requires that the functions f1 , f2 and
f4 depend only on μ. According to the transformation rules of Type IIB supergravity, the metric,
in the Einstein frame, must be invariant under SL(2, R), which requires that the functions f1 , f2 ,
and f4 are invariant under SL(2, R). The associated orthonormal frame is given by the following
set of 1-forms,

ei = f4 êi i = 0, 1, 2, 3,
e = f4 dμ,
4

e5 = f1 ê5 = f1 (dβ + A1 ),
ea = f2 êa , a = 6, 7, 8, 9. (5.20)
For i = 0, 1, 2, 3, the êi span the orthonormal frame for AdS4 and may be chosen as follows,

ê0 = r −1 dr, êi = r −1 dx i , i = 1, 2, 3. (5.21)


For a = 6, 7, 8, 9, the êa span the orthonormal frame on CP2 of (5.6). The volume form9 is

e0123456789 = f1 f24 f45 dμ ∧ ê0123 ∧ ê56789 . (5.22)

5.5. The Ansatz for the antisymmetric tensor fields

Invariance under SO(2, 3) × SU(3) × U (1)β requires the self-dual 5-form to be of the form,
 
F(5) = f5 −e01234 + e56789 , (5.23)
where f5 is a scalar function that depends only on μ, by the same arguments as used for the
metric. Recall that F(5) and thus f5 must also be invariant under SL(2, R).
To construct a 2-form B(2) which is invariant under SO(2, 3) × SU(3) × U (1)β × SL(2, R),
we make use of the SU(3) invariant 2-forms K,  , and ¯ . Since B is complex, we include
2 2 (2)
Â2 and ¯2 with independent complex coefficient functions f3 and ḡ3 ,

B(2) = if3 Â2 − i ḡ3 ¯2 + f6 K. (5.24)


It will turn out that the Type IIB field equation (2.10) for Gμνρ force f6
= 0 and render this term
pure gauge; therefore we shall set f6 = 0 in the sequel. The field strength F(3) is then,
f3
ḡ3
f3 5 ḡ3 5
F(3) = i 2
e 4 ∧ A2 − i e4 ∧ Ā2 − 3 e ∧ A2 − 3 e ∧ Ā2 . (5.25)
f 4 f2 f4 f22 2
f 1 f2 f1 f22
The associated composite G is given by

G = ae5 ∧ A2 − ibe4 ∧ A2 + ce5 ∧ Ā2 − ide4 ∧ Ā2 , (5.26)

9 We shall introduce the following notation, ei1 i2 ···ip ≡ ei1 ∧ ei2 ∧ · · · ∧ eip and use it throughout.
E. D’Hoker et al. / Nuclear Physics B 757 (2006) 79–116 93

where the coefficient functions are given by


3 3
a=− f (f3 − Bg3 ), c=− f (ḡ3 − B f¯3 ),
f1 f22 f1 f22
1 1
b=− f (f3
− Bg3
), d =+ f (ḡ3
− B f¯3
). (5.27)
f4 f22 f4 f22
For later convenience, we have here expressed these forms in terms of the frame ea , for which
   
A2 = f22 Â2 = e6 + ie7 ∧ e8 + ie9 e3iβ . (5.28)
With f3 and g3 functions only of μ, the Ansatz in (5.24) is invariant under SO(2, 3) × SU(3).
Under U (1)β , the forms A2 and Ā2 transform with constant opposite phases. Thus, U (1)β is not
a symmetry of any one Ansatz, but rather relates one Ansatz to another.

5.6. Transformation properties under SL(2, R)

Under SL(2, R), the metric (in the Einstein frame) and the 5-form F(5) are invariant. The
dilaton/axion field B, and the associated function f , transform as
uB + v  s
Bs = , f s = |v̄B + ū|f, f 2B
= e2iθ f 2 B
, (5.29)
v̄B + ū
where the superscripts s indicate the transformed objects, u, v ∈ C and ūu − v̄v = 1. The func-
tions f3 and g3 transform linearly, according to (2.13), and we have,
 
f3s = uf3 + vg3 , f s f3s − B s g3s = eiθ f (f3 − Bg3 ),
 
g3s = v̄f3 + ūg3 , f s ḡ3s − B s f¯3s = eiθ f (ḡ3 − B f¯3 ), (5.30)
where the phase θ is a field-dependent transformation parameter, given by
 
v B̄ + u 1/2
eiθ = . (5.31)
v̄B + ū
Since the phases of all terms in G are the same, we get a covariant formula, Gs = eiθ G.

6. The reduced Bianchi identities and field equations

In this section, we reduce the Type IIB Bianchi identities and field equations, given in Sec-
tion 2, to the Ansatz constructed in Section 5. The Bianchi identities (2.4), (2.5), and (2.6) are
automatically satisfied. The Bianchi identity (2.7) for F(5) reduces to
f2
f
1
f5
= −4 f5 − 1 f5 + f4 (a b̄ + āb + cd̄ + c̄d) (6.1)
f2 f1 2
and is solved by

3 |f3 |2 − |g3 |2 + C1
f5 = (6.2)
2 f1 f24
which fixes f5 in terms of f1 , f2 , f3 , g3 , and the (real) integration constant C1 .
94 E. D’Hoker et al. / Nuclear Physics B 757 (2006) 79–116

6.1. The reduced field equations for B and G

The field equation for the complex scalar B reduces to






f4 f1
f2
f2
B +B 3 + +4 + 2f 2 B̄B
B
+ 2 42 (ac − bd) = 0, (6.3)
f4 f1 f2 f
where the variables a, b, c, d were introduced in (5.27).
To reduce the field equations of the antisymmetric tensor field B(2) , it is convenient to first
recast (2.10) in terms of differential forms,10
∗d(∗G) + i(iQ G) + (iP Ḡ) − 4i(iG F(5) ) = 0. (6.4)
Here iV G stands for the contraction of G with V . To calculate this equation, it is helpful to
have the following contractions, ie5 ∧A2 e56789 = A2 , as well as the complex conjugate relation.
Identifying terms in A2 , we find, after some simplification,


f2 f1 f

f3

− Bg3

− 9 42 (f3 − Bg3 ) + + 3 4 (f3


− Bg3
)
f1 f1 f4
f42 f5
− 2f 2 B
(g3
− B̄f3
) − 12 (f3 − Bg3 ) = 0, (6.5)
f1
while in Ā2 , the equation is obtained from (6.5) by letting f3 → ḡ3 , and f5 → −f5 , leaving
all other functions unchanged. It is actually more convenient to express these field equations in
terms of the coefficient functions a, b, c, d of G, and we find (with Q = Qμ dμ),



f1 f2
f4
a − iQμ a = − +2 a + 3 b − f 2 B
c̄,
f1 f2 f1




f f f f4
b
− iQμ b = − 4 4 + 2 2 + 1 b + 3 a − f 2 B
d̄ + 4f4 f5 a,
f4 f2 f1 f1



f f f
c
− iQμ c = − 1 + 2 2 c − 3 d − f 2 B
ā,
4
f1 f2 f1


f f
f
f4
d
− iQμ d = − 4 4 + 2 2 + 1 d − 3 c + 4f4 f5 c − f 2 B
b̄. (6.6)
f4 f2 f1 f1

6.2. Reducing Einstein’s equations

To reduce the Einstein equations in (2.11), we first obtain the Ricci curvature tensor. It is con-
venient to carry out all calculations using the orthonormal frame of (5.20), which we shall denote
collectively by eA where A = (i, 4, 5, a) with i = 0, 1, 2, 3, and a = 6, 7, 8, 9. The torsion-free
connection ωA B , the associated curvature Ω A B , the Riemann tensor R A BCD , and the Ricci ten-
sor (all expressed in frame indices) are then defined by the relations,
0 = deA + ωA B ∧ eB ,

10 Our conventions for the Poincaré dual are given via the following pairing relation between two arbitrary rank p

differential forms S(p) and T(p) , by S(p) ∧ ∗T(p) = p! 1 S a1 ···ap T 0123456789 . In particular, we have ∗∗S
(p) (p)a1 ···ap e (p) =
(−1) p+1 S(p) , and the duals ∗e 01234 = −e 56789 , and ∗e 6789 =e 012345 , which will be useful later on.
E. D’Hoker et al. / Nuclear Physics B 757 (2006) 79–116 95

Ω A B = dωA B + ωA C ∧ ωC B ,
1
Ω A B = R A BCD eC ∧ eD ,
2
RBD = R A BAD , (6.7)
where A, B, C, D = 0, 1, 2, . . . , 9. The corresponding objects for the unwarped AdS4 and CP2
geometries will be denoted with hats; they are the frames êi , êa , the connections ω̂i j , ω̂a b , the
curvatures Ω̂ i j , Ω̂ a b , the Riemann tensors R̂ i j kl , R̂ a bcd , and the Ricci tensors R̂j l , R̂bd . They
all obey Eqs. (6.7) with hatted objects and for the ranges of indices i, j, k, l = 0, 1, 2, 3 and
a, b, c, d = 6, 7, 8, 9. (The unwarped geometry also has structure in the direction 5, where it can
be viewed as the S 1 fiber of S 5 over CP2 ; we shall treat this direction as separate.) The AdS4
curvature Ω̂ is calculated using the unwarped frame in (5.21), and is given by11
1
Ω̂ij = − (ηik ηj l − ηil ηj k )êk ∧ êl ,
2
R̂ij kl = −(ηik ηj l − ηil ηj k ),
R̂j l = −3ηj l . (6.8)
The CP2 curvature is calculated using the unwarped frame in (5.6) and is given by
Ω̂ab = (δac δbd + kac kbd + kab kcd )êc ∧ êd ,
R̂abcd = δac δbd − δad δbc + kac kbd − kad kbc + 2kab kcd ,
R̂bd = 6δbd , (6.9)
where the tensor k is anti-symmetric and defined by kcd êc ∧ êd = 2K = 2ê6 ∧ ê7 + 2ê8 ∧ ê9 .
The connection components of the full geometry are needed to compute the curvature as well
as the supersymmetry variation equations. They are given by ωAB = −ωBA and
ωij = ω̂ij , ωi5 = 0,
f

ωi 4 = 42 ei , ωib = 0,
f4
f1
5 f1
ω5 4 = e , ω5a = kab eb ,
f1 f4 f22
f2
a f1
ωa 4 = e , ωab = ω̂ab − kab e5 . (6.10)
f2 f4 f22
The components of the curvature form Ω A B and of the Riemann tensor R A BCD are needed only
to evaluate the Ricci tensor, and will not be exhibited here. The non-vanishing components of the
Ricci tensor are given by,
 
3 (f4
)2 f4

f1
f4
f2
f4

Rij = −ηij + 2 + + + 4 ,
f42 f44 f43 f1 f43 f2 f43
f4

(f4
)2 f1
f4
f1

f2
f4
f2

R44 = −4 +4 + − +4 −4 ,
f43 f44 f1 f43 f1 f42 f2 f43 f2 f42

11 The explicit forms of the AdS and CP unwarped connections will not be needed here.
4 2
96 E. D’Hoker et al. / Nuclear Physics B 757 (2006) 79–116

f1

f1
f4
f
f
f12
R55 = − −3 −4 1 22 +4 ,
f1 f42 f1 f4 3 f 1 f 2 f4 f24
 
6 f2

f
f
f1
f2
f12 (f2
)2
Rab = δab − − 3 2 43 − −2 −3 . (6.11)
f22 f 2 f42 f 2 f4 f1 f2 f42 f24 f22 f42
Finally, for later use, we record also the Ricci scalar,

4 (f2
)2 (f4
)2 f1
f2
f1
f4
f2
f4
f12 f42 f42
R= 2 3 2 +3 2 −3+2 +2 +8 − +6 2 (6.12)
f4 f2 f4 f1 f2 f1 f4 f2 f4 f24 f2
which will be used to evaluate the reduced action.
It is straightforward to evaluate the “matter” contributions to Einstein’s equations, which leads
to the following reduced Einstein equations.
For i, j = 0, 1, 2, 3, we have,

(f4
)2
f4

f1
f4
f2
f4

3+2 + + + 4 − 4f52 f42


f42 f4 f1 f4 f2 f4
1  
+ f42 −|a|2 − |b|2 − |c|2 − |d|2 = 0. (6.13)
2
For 44, we have,
f4

(f
)2 f
f
f

f
f
f

4 − 4 42 − 1 4 + 1 − 4 2 4 + 4 2 − 4f52 f42 + 2f 4 |B
|2
f4 f4 f1 f4 f1 f2 f4 f2
1  
+ f42 −|a|2 + 3|b|2 − |c|2 + 3|d|2 = 0. (6.14)
2
For 55, we have
f1

f
f
f
f
f 2f 2
+ 3 1 4 + 4 1 2 − 4 1 44 + 4f52 f42
f1 f1 f4 f1 f2 f2
1  
+ f42 3|a|2 − |b|2 + 3|c|2 − |d|2 = 0 (6.15)
2
and finally for a, b = 6, 7, 8, 9, we have,

6f42 f2

f2
f4
f1
f2
f12 f42 (f2
)2
− + + 3 + + 2 + 3 + 4f52 f42
f22 f2 f2 f4 f1 f2 f24 f22
1  
+ f42 |a|2 + |b|2 + |c|2 + |d|2 = 0. (6.16)
2
Recall that here, as before, the variables a, b, c, d are given in terms of the independent functions
f3 , g3 by the definitions (5.27).

6.3. Integral of motion

The 4 Einstein equations possess a single integral of motion, which may be viewed as the
(vanishing) Hamiltonian, or Wheeler–De Witt equation. It is obtained by adding the 4 equations
with the following multiplicative factors, +4, −1, +1, +4, and is given by
E. D’Hoker et al. / Nuclear Physics B 757 (2006) 79–116 97

(f2
)2 (f
)2 f
f
f
f
f
f
f 2f 2
12 + 12 2
+ 12 42 + 8 1 2 + 8 1 4 + 32 2 4 + 4 1 44
f2 f4 f1 f2 f1 f4 f2 f4 f2
f2  
− 24 42 + 8f42 f52 − 2f 4 |B
|2 + 2f42 |a|2 − |b|2 + |c|2 − |d|2 = 0. (6.17)
f2
In general, the above system of reduced equations (including the reduced equations for B and
for a, b, c, d) does not appear to possess any further first integrals.

6.4. Reduced action principle

In this subsection, we shall show that the reduced field equations, listed above, may be derived
from the Type IIB action (2.14), reduced to our Ansatz. To show this, we shall need one more
ingredient in the construction of the Ansatz that was not needed for the field equations, but is
needed for the action, namely the anti-symmetric tensor C(4) . The starting point is the relation
between C(4) and F(5) in (2.2). Using the Ansatz for F(5) , B(2) and F(3) , as well as the Bianchi
identities (6.2), we obtain the following on-shell expression for dC(4) ,
3 i
dC(4) = −f45 f5 ê01234 + C1 ê56789 − (f3 f¯3
− f¯3 f3
− g3 ḡ3
+ ḡ3 g3
)ê46789 . (6.18)
2 4
Therefore, the most general Ansatz for C(4) is as follows,
3
C(4) = g5 ê0123 + C1 β ê6789 + h5 ê6789 , (6.19)
2
where g5 and h5 are functions of μ only. The term proportional to C1 accounts for the term
proportional to C1 in (6.18), via the fact that ê56789 = dβ ∧ ê6789 .
To obtain an off-shell formulation, for use in the action, we postulate (6.19) as the Ansatz for
C(4) . This is clearly the most general SO(2, 3) × SU(3) × U (1)β invariant 4-form we can write
down.12 Evaluating F(5) now from the invariant Ansatz for C(4) in (6.19), we find an expression
which is not necessarily self-dual,
3
F(5) = g5
ê01234 + C1 ê56789 + h
5 ê46789
2
i
+ (f3 f¯3 − f¯3 f3
− g3 ḡ3
+ ḡ3 g3
)ê46789

4
3 
+ |f3 |2 − |g3 |2 ê56789 . (6.20)
2
In turn, self-duality of F(5) requires the following relations,

3 |f3 |2 − |g3 |2 + C1
0 = f4−5 g5
+ ,
2 f1 f24
i
0 = h
5 + (f3 f¯3
− f¯3 f3
− g3 ḡ3
+ ḡ3 g3
), (6.21)
4
which reproduce the on-shell relations (6.18).

12 Notice that under the action of U (1) , the form C


β (4) is not strictly invariant, but changes by a gauge transformation
since ê6789 is a closed form.
98 E. D’Hoker et al. / Nuclear Physics B 757 (2006) 79–116

We are now ready to reduce the Type IIB supergravity action of (2.14). It will be expressed
here in terms of our complex fields B, and G, and is given by, 13
 
√ 1
S = dx g R − 2f ∂M B∂ B̄ − GMNP Ḡ
4 M MNP
− 4|F(5) |2
12

− i C(4) ∧ G ∧ Ḡ. (6.22)

Since, after variation of the fields, we are to further enforce the self-duality relation, this action is
not quite unique. Indeed, we are free to add to the Lagrangian density a term proportional to the
square of the self-duality relation |F(5) − ∗F(5) |2 , as its variation will vanish on self-dual fields.
In reducing the action over our Ansatz, it will be convenient to add the term 4|F(5) − ∗F(5) |2 to
the Lagrangian density, as this will eliminate terms in the reduction that are quartic in f3 and g3 .
Evaluating this modified action on our Ansatz, omitting the overall volume factors of ê0123 ∧
ê 56789 of AdS4 × S 5 , and integrating all second derivatives by part to convert all terms to involve
only first derivatives, we obtain a
 
Sreduced = dμ 12(f2
)2 f1 f22 f43 + 12(f4
)2 f1 f24 f4 + 8f1
f2
f23 f43 + 8f1
f4
f24 f42

+ 32f2
f4
f1 f23 f42 − 12f1 f24 f43 − 4f13 f45 + 24f1 f22 f45 − 2f 4 f1 f24 f43 |B
|2
 
− 2f1 f43 f 2 |f3
− Bg3
|2 + f 2 |ḡ3
− B f¯3
|2
f45  2 
− 18 f |f3 − Bg3 |2 + f 2 |ḡ3 − B f¯3 |2
f1

 
+ 8f1 f24 f4−5 (g5
)2 + 24g5
|f3 |2 − |g3 |2 . (6.23)

We have shown that the reduced field equations follow from the action Sred .
The only further relations implied by the self-duality constraint (2.8) are the value of the
constant C1 and the expression for the function h5 which yields (6.21).

7. Supersymmetry variations

In this section, we reduce the BPS equations δλ = δψμ = 0 expressing the vanishing super-
symmetry variation of the dilatino and gravitino fields. Fields satisfying these reduced equations
with non-vanishing supersymmetry parameter ε will exhibit some degree of residual supersym-
metry. It will be convenient to express the gravitino equation in differential form notation. The
dilatino and gravitino equations are given respectively by,
1
(Γ · P )B −1  ∗ − (Γ · G) = 0, (7.1)
24
dε + ωε + ϕ (1) ε + ϕ (2) B −1 ε ∗ = 0, (7.2)

13 Compared to the conventions of [39], we omit an overall factor of 1/2κ 2 , divide C


10 (4) by a factor of 4, and divide
F(5) by a factor of 4. The relative normalization of the Chern–Simons term against the |F(5) |2 term is checked using the
self-duality of F(5) and the Bianchi identity for F(5) . The absolute normalization of the |F(5) |2 term may be checked
from Einstein’s equations, while that of the Chern–Simons term may be checked independently against the field equation
for G.
E. D’Hoker et al. / Nuclear Physics B 757 (2006) 79–116 99

where the connection components are as follows,14


1
ω = Γ MN ωMN ,
4
i i
ϕ =− Q+
(1)
(Γ · F(5) )eA Γ A ,
2 480
1  
ϕ (2) = − eA Γ A (Γ · G) + 2(Γ · G)Γ A . (7.3)
96
Here, eA is the frame of (5.20), ω its torsion-free connection, B is the complex conjugation
matrix, and we have used the following relation
Γ MNP Q GN P Q − 9ηMN Γ P Q GN P Q = −Γ M (Γ · G) − 2(Γ · G)Γ M (7.4)
to relate the gravitino equation of (2.12) to that of (7.2) and (7.3).

7.1. Γ -matrices

It will be useful to choose a well-adapted basis for Γ -matrices,15


Γ μ = σ1 ⊗ γ μ ⊗ I4 , μ = 0, 1, 2, 3, 4,
Γ m
= σ 2 ⊗ I4 ⊗ γ ,m
m = 5, 6, 7, 8, 9 (7.5)
for which we have
Γ 11 = Γ 0123456789 = σ3 ⊗ I4 ⊗ I2 ⊗ I2 ,
B = Γ 2568 = σ3 ⊗ γ 2 ⊗ σ1 ⊗ σ2 . (7.6)

7.2. Reducing the dilatino equation

Evaluating the dilatino equation (7.1) on the Ansatz, we find,


f 2 B
4 −1 ∗  5   
4 Γ B  = aΓ − ibΓ 4 Γ z1 z2 e+3iβ ε + cΓ 5 − idΓ 4 Γ z̄1 z̄2 e−3iβ ε. (7.7)
f4
To solve this equation, we multiply both sides to the left by Γ 4 and use the fact that Γ 45 com-
mutes with Γ z1 z2 . The rhs then automatically projects onto the subspace of spinors corresponding
to eigenvalue +1 of the Dirac matrix I2 ⊗ I4 ⊗ γ 5 . Since B commutes with this matrix, both
ε and B −1 ε ∗ must have eigenvalue +1. Introducing a basis of 2-component spinors u± , with
σ1 u± = +u∓ , and σ± u± = 0, we parametrize of the eigenvalue +1 subspace for chiral spinors
ε satisfying Γ 11 ε = −ε, in terms of two 4-dimensional complex spinors ζ± , and their complex
conjugates ζ±∗ ,
ε = u− ⊗ (ζ+ ⊗ u+ ⊗ u+ + ζ− ⊗ u− ⊗ u− ),
 
B −1 ε ∗ = u− ⊗ iγ 2 ζ−∗ ⊗ u+ ⊗ u+ − iγ 2 ζ+∗ ⊗ u− ⊗ u− . (7.8)

14 Throughout, we shall use the notation Γ · T ≡ Γ M1 ···Mn T


M1 ···Mn for the contraction of rank n totally anti-symmetric
Γ -matrices and tensors.
15 A useful set of conventions for the 4 × 4 matrices γ A , as well as some further formulas for combinations of

Γ -matrices, such as Γ zi , and Γ AB , are presented in Appendix A.


100 E. D’Hoker et al. / Nuclear Physics B 757 (2006) 79–116

In this basis, the dilatino equation reduces to,


 4 
aγ + b ζ− = e−3iβ f 2 B
(f4 )−1 γ 2 ζ−∗ ,
 4 
cγ + d ζ+ = e+3iβ f 2 B
(f4 )−1 γ 2 ζ+∗ . (7.9)
Notice that the equations for ζ+ and ζ− can be satisfied independently of one another. Non-trivial
solutions require the following relations,

ζ+ = 0, (c + d)(c̄ − d̄) = f 4 |B
|2 (f4 )−2 ,
ζ− = 0, (a + b)(ā − b̄) = f 4 |B
|2 (f4 )−2 . (7.10)
The β-dependence of ζ± is readily solved for, since a, b, c, d, f4 and f 2 B
are all β-independent.
The general solution is given by
3  4 
ζ± = e± 2 iβ ζ̂± , aγ + b ζ̂− = f 2 B
(f4 )−1 γ 2 ζ̂−∗ ,
 4 
cγ + d ζ̂+ = f 2 B
(f4 )−1 γ 2 ζ̂+∗ , (7.11)

where ζ̂± is independent of β.

7.3. Reducing the gravitino equation

We begin by decomposing the gravitino equations (7.2) onto the eigenvalue +1 and −1 of
the matrix Γ 6789 = −I8 ⊗ σ3 ⊗ σ3 . Since ε and B −1 ε ∗ belong to the subspace with definite
eigenvalue −1, this decomposition may be achieved by decomposing the covariant derivative ac-
cording to components that commute or anti-commute with Γ 6789 . This decomposition is unique
since Γ 6789 is invertible. We have,

dε + ω+ ε + ϕ+ ε + ϕ+ B −1 ε ∗ = 0,
(1) (2)

(1) (2) −1 ∗
ω− ε + ϕ− ε + ϕ− B ε = 0. (7.12)
Here, the components that commute with Γ 6789 are given by,
1 1 1 1
ω+ = ωij Γ ij + ωi4 Γ i4 + ω45 Γ 45 + ωab Γ ab ,
4 2 2 4
(1) i i  
ϕ+ = − Q + (Γ · F(5) ) ei Γ i + e4 Γ 4 + e5 Γ 5 ,
2 480
(2) 1  
ϕ+ = − eA Γ A (Γ · G) + 2(Γ · G)Γ A , (7.13)
96
A=0,1,2,3,4,5

while the components that anti-commute with Γ 6789 are given by


1 1
ω− = ωa4 Γ a4 + ωa5 Γ a5 ,
2 2
(1) i
ϕ− = (Γ · F(5) )ea Γ a ,
480
(2) 1  
ϕ− = − ea Γ a (Γ · G) + 2(Γ · G)Γ a . (7.14)
96
Here, the indices i, j continue to run over the range 0, 1, 2, 3, while a, b run over 6, 7, 8, 9.
E. D’Hoker et al. / Nuclear Physics B 757 (2006) 79–116 101

7.3.1. Reducing the algebraic equation


The second equation in (7.12) is purely algebraic in ε. Using the expressions for the connec-
tion components of (6.10), and the anti-symmetric tensor fields F(5) of (5.23) and G of (5.26), we
find one equation involving both ζ+ and γ 2 ζ+∗ , and another equation involving both ζ− and γ 2 ζ−∗ .
We proceed by assuming that B
= 0 since we shall be interested in obtaining Janus solutions
with varying dilaton field. We may then eliminate γ 2 ζ±∗ using the dilatino equation (7.9). The
final result is expressed in terms of ζ̂± , using (7.11), by the following two independent equations,


f f1 f 4 f2   
2 2 γ 4 ± 2 2 − 2f5 f4 − 24
γ 4 a + bγ 4 c + dγ 4 ζ̂± = 0, (7.15)
f2 f2 f B

where ± is correlated with the subscript of ζ̂± .


The above equations imply that, still under the assumption that B
= 0, either ζ̂+ or ζ̂− (or
both) must vanish. To show this, we begin by assuming that ζ̂+ = 0, and decompose the equation
for ζ̂+ onto the two γ 4 chiralities of ζ̂+ . Note that the equations must hold for both chiralities,
since the dilatino equation (7.9) chirality,
f2
f1 f 4 f2
+2 + 2 2 − 2f5 − 24
(a + b)(c + d) = 0,
f2 f2 f B
f2
f1 f 4 f2
−2 + 2 2 − 2f5 + 24
(a − b)(c − d) = 0, (7.16)
f2 f2 f B

where the ± sign refers to the γ 4 -chirality of ζ̂+ . Using these two relations, required by ζ̂+ = 0
in the equation for ζ̂− gives f1 f4 f2−2 ζ̂− = 0. For non-degenerate geometries, f1 f4 f2−2 = 0 and
thus we must have ζ̂− = 0, as announced.
The equations for ζ̂+ and ζ̂− are mapped into one another by parity, which maps μ → −μ,
and are equivalent to one another. In view of the result above, we may choose ζ̂− = 0.

7.3.2. Reducing the differential equation


To reduce the differential equation in (7.12), we begin by calculating the connection ωab Γ ab ,
using Γ ab = I8 ⊗ γ ab . Furthermore, we use the relations
γ 67 u± ⊗ u± = +γ 89 u± ⊗ u± = ±iu± ⊗ u± ,
γ 68 u± ⊗ u± = −γ 79 u± ⊗ u± = ±u∓ ⊗ u∓ ,
γ 69 u± ⊗ u± = +γ 78 u± ⊗ u± = iu∓ ⊗ u∓ (7.17)
and the fact that ω̂68 = ω̂79 , and ω̂69 = −ω̂78 , as found in (5.10), to derive
1 3i f1
ωab γ ab = A1 σ3 − i 2 e5 σ3 . (7.18)
4 2 f2
Here, σ3 acts on ζ± by σ3 ζ± = ±ζ± . All connection components now commute with σ3 and
we may decompose the reduced gravitino equation onto its decoupled ζ+ and ζ− components.
Finally, the simple β-dependence of ζ± found in (7.11) is used to recast the equations in terms of
ζ̂± . The differential dβ picked up in the process combines with the U (1) connection A1 in (7.18)
and forms the gauge invariant combination dβ + A1 = ê5 , so that we have effectively
1 3i 5 f1
ωab γ ab → e σ3 − i 2 e 5 σ3 . (7.19)
4 2f1 f2
102 E. D’Hoker et al. / Nuclear Physics B 757 (2006) 79–116

Finally, retaining only the ζ̂+ -component of the differential equation, we obtain
 
1 1 i 3i 5 f1 5
0 = d ζ̂+ + ω̂ij γ + ωi4 γ − ω45 γ +
ij i4 4
e − i 2 e ζ̂+
4 2 2 2f1 f2
 
i 1  i 
+ − Q + f5 ei γ + e4 γ 4 − ie5 ζ̂+
2 2
1 i     
+ ei γ a + bγ 4 + e4 γ 4 a − 3bγ 4 + e5 −3ia + ibγ 4 γ 2 ζ̂+∗ . (7.20)
4

7.3.3. Analyzing the reduced differential equation


Since ζ̂+ is β-independent in (7.20), no ∂β ζ̂+ dβ contribution will appear in d ζ̂+ . As a result,
we must have ie5 d ζ̂+ = 0, which yields a further algebraic equation,


f1 4 f4 f1 f 4 1  
γ + 3 − 2 2 − f4 f5 ζ̂+ = f4 3a − bγ 4 γ 2 ζ̂+∗ . (7.21)
f1 f1 f2 2

The decomposition of (7.20) onto the directions e4 and êi , yields two differential equations,
 
1 f4
i 1 1
∇ˆ i + γ γ4 + f5 f4 γ i ζ̂+ + f4 (a − bγ4 )γ i γ2 ζ̂+∗ = 0, (7.22)
2 f4 2 4
 
i 1 1
∂μ − Q + f5 f4 γ4 ζ̂+ + f4 (aγ4 − 3b)γ2 ζ̂+∗ = 0, (7.23)
2 2 4
where ∇ˆ i is the covariant derivative with connection ω̂ij γ ij /4. Finally, we proceed to eliminating
the spinor γ 2 ζ̂+∗ using the dilatino equation (7.9), and to further simplifying the equations. From
the outset, we use the fact that ζ̂− = 0, as derived earlier. The results for all supersymmetry
variation equations are summarized below.

7.4. Summary of the reduced dilatino/gravitino equations

• The implications of the dilatino equations,

(c + d)(c̄ − d̄) = f 4 |B
|2 (f4 )−2 ; (7.24)
• The algebraic integrability equations,
f2
f42
= (ac + bd), (7.25)
f2 2f 2 B

f1 f 4 f42
− f 4 f 5 = (ad + bc); (7.26)
f22 2f 2 B

• The e5 component equation,


f1
f42
= (3ac − bd), (7.27)
f1 2f 2 B

f4 f1 f 4 f42
3 − 2 2 − f4 f 5 = (3ad − bc); (7.28)
f1 f2 2f 2 B

E. D’Hoker et al. / Nuclear Physics B 757 (2006) 79–116 103

• The AdS4 components’ equation,




f12 f42 f2 f4
2
− + = 1. (7.29)
f24 f2 f4

This summary of results is obtained as follows. Eq. (7.24) is just the original dilatino equation
of (7.10) for the case where ζ̂− = 0. Eqs. (7.25) and (7.26) are obtained by taking the sum and
the difference of the two algebraic equations (7.16). Eqs. (7.27) and (7.28) are obtained (7.21)
by eliminating γ 2 ζ̂+∗ using the dilatino equation (7.9), and taking the sum and difference of the
resulting equations for ±1 eigenvalues of the γ 4 matrix. To obtain Eq. (7.29) requires more
work. We begin by eliminating the spinor γ 2 ζ̂+∗ using the dilatino equation (7.9). The resulting
equation takes the form,
 
∇ˆ i + κγi + λγi γ 4 ζ̂+ = 0, (7.30)
where
1 f42 1 f1 f4
κ = f 4 f5 + (ad + bc) = ,
2 4f 2 B
2 f22
1 f4
f42 1 f2
1 f4

λ= + (ac + bd) = + . (7.31)


2 f4 4f 2 B
2 f2 2 f4
In deriving the second equalities on the rhs of the equations above, we have eliminated the
combination ad + bc using (7.26), and eliminated the combination ac + bd using (7.25). The
coefficients κ and λ depend upon μ but not upon the variables of AdS4 , and may thus be viewed
as constants in this differential equations.
The integrability of (7.30) requires κ 2 − λ2 = 1/4, which yields (7.29). Finally, equation
(7.23) is not reproduced in the summary because it only governs the μ-evolution of ζ̂+ , is always
integrable, and imposes no new condition of the dynamical variables of the system, namely
f1 , f2 , f3 , g3 , f4 and B.
The solution of equations (7.24) to (7.29) gives a background which preserves four real su-
percharges. The counting proceeds as follows: In Type IIB supergravity the supersymmetry
transformation parameter  has 32 real components. The compatibility with the CP2 fibration
leads to (7.8) reducing  to 16 real components parameterized by two complex four-dimensional
spinors ζ+ and ζ− . The analysis of Section 7.3 implies that one of the two ζ± is zero, leaving
four complex components. Finally the dilatino supersymmetry condition leads to a reality condi-
tion leaving four real unbroken supersymmetries, which is the degree of supersymmetry of our
solution. In the sequel, we shall not need to explicitly solve for the spinor ζ̂+ ; instead it will suf-
fice to solve the integrability conditions which guarantee the existence of the four real unbroken
supersymmetry.

8. Reality properties of supersymmetric solutions

We shall now search for supersymmetric solutions. This requires satisfying the field equations
(6.3), (6.6), and (6.13)–(6.16), (one special combination of which is the constraint (6.17)), as
well as the supersymmetry conditions (7.24)–(7.29) of the summary. Assuming that all these
equations are simultaneously satisfied guarantees that we will have a true solution to the field
equations, which also is supersymmetric. The simultaneous consideration of the susy equations
and the field equations leads to many simplifications.
104 E. D’Hoker et al. / Nuclear Physics B 757 (2006) 79–116

8.1. Solving the equations for B and τ

The difference between (7.25) and (7.27) yields


f1
f2
f2
− = 24
(ac − bd) (8.1)
f1 f2 f B
which upon eliminating ac − bd in (6.3), and dividing the resulting equation by B
gives
B

f1
f2
f4

0= + 3 + 2 + 3 + 2f 2 B̄B
. (8.2)
B
f1 f2 f4
Taking the real part of (8.2), and using the fact that f 2 (B̄B
+ B B̄
) = (ln f 2 )
, we get
C2
f 2 |B
| = , (8.3)
f1 f22 f43
3

where C2 is a constant. Converting the imaginary part of (8.2), into an equation for τ = τ1 + iτ2 ,
using (2.3), we get precisely the imaginary part of the τ -equation for the non-supersymmetric
Janus solution of (3.2),
τ

τ̄

τ1

0= − + 2i . (8.4)
τ
τ̄
τ2
Just as in the case of the non-supersymmetric Janus solution, its solution requires the ax-
ion/dilaton field τ to flow along a geodesic in the τ -upper-half-plane,
|τ − p|2 = r 2 , (8.5)
where p, r are arbitrary real constants.

8.2. Mapping to real solutions

Since B flows along a geodesic, and SL(2, R) acts transitively on the space of all geodesic
segments of the same hyperbolic length,16 we may use an SL(2, R) transformation to map any
one geodesic segment into a geodesic segment of B real, or equivalently τ purely imaginary.
Since SL(2, R) is a symmetry of the Type IIB supergravity equations, the most general solution
is obtained by taking B real, and then applying the most general SL(2, R) transformation to the
solution. Henceforth, we restrict to B real. The reality of B implies that the following quantities
are real,
ac, bd, ad, bc ∈ R. (8.6)
As a result, a/b and c/d are real, and thus have pairwise identical phases. The reality of the
products then further imposes the following phase arrangements,
a = ar eiθ , c = cr e−iθ ,
b = br eiθ , d = dr e−iθ , (8.7)
where ar , br , cr , dr , θ are all real. It follows from (7.24) that, unless the dilaton is constant, c
cannot vanish identically. Now substitute the above phase relations in the field equation for c and

16 By a geodesic segment we understand a connected segment of a single geodesic.


E. D’Hoker et al. / Nuclear Physics B 757 (2006) 79–116 105

derive that θ
= 0. Hence θ is a constant phase. Changing this phase is equivalent to making a
U (1)β rotation. Thus, we may now take also a, b, c, d ∈ R. The general solution will be obtained
from the real solution by making the inverse U (1)β × SL(2, R) rotation.

8.3. Reduced equations for real B, a, b, c, d

The field equations simplify for real field, and we have,




f f
f4
a
= − 1 + 2 2 a + 3 b − f 2 B
c,
f1 f2 f1




f f f f4
b
= − 4 4 + 2 2 + 1 b + 3 a − f 2 B
d + 4f4 f5 a,
f4 f2 f1 f1



f1 f2
f4
c =− +2 c − 3 d − f 2 B
a,
f1 f2 f1



f4 f2
f1
f4
d =− 4 +2 + d − 3 c + 4f4 f5 c − f 2 B
b. (8.8)
f4 f2 f1 f1
Einstein’s equations are arranged as follows: (6.14) is a linear combination of for (6.13), (6.15)–
(6.17), and will therefore be omitted,
f4

(f
)2 f
f
f
f
1  
0= + 3 + 2 42 + 1 4 + 4 2 4 − 4f52 f42 − f42 a 2 + c2 + b2 + d 2 ,
f4 f4 f1 f4 f2 f4 2
f1

f
f
f
f
f 2f 2 1  
0= + 3 1 4 + 4 1 2 − 4 1 44 + 4f52 f42 + f42 3a 2 − b2 + 3c2 − d 2 ,
f1 f1 f4 f1 f2 f2 2
f2

6f42 f
f
f
f
f 2f 2 (f
)2
0= − 2 + 3 2 4 + 1 2 + 2 1 44 + 3 22 + 4f52 f42
f2 f2 f2 f4 f1 f2 f2 f2
1  
+ f42 a 2 + c2 + b2 + d 2 . (8.9)
2
The constraint is given by
(f2
)2 (f4
)2 f1
f2
f
f
f
f

0 = 12 + 12 +8 + 8 1 4 + 32 2 4 − 2f 4 (B
)2
f22 f42 f1 f2 f1 f4 f2 f4
f12 f42 f42  
+ 12 + 4 − 24 + 8f42 f52 + 2f42 a 2 − b2 + c2 − d 2 . (8.10)
f24 f22
The Bianchi identity for F(5) is solved in terms of a single constant C1 ,
1   3 C1
f5 = f1 a 2 − c 2 + . (8.11)
6 2 f1 f24
The dilatino equation simplifies as follows,

c2 − d 2 = f 4 (B
)2 (f4 )−2 . (8.12)
The remaining supersymmetry variation equations continue to be given by (7.25)–(7.29), but
now for a, b, c, d and B real.
106 E. D’Hoker et al. / Nuclear Physics B 757 (2006) 79–116

9. Solving the field and supersymmetry equations

We shall now apply the following procedure to the solution of the susy variation equations and
the field equations: (1) Take the solution of the B-equation from the results above; (2) Solve the
susy variation equations and the Hamiltonian constraint (which will restrict the possible initial
data); (3) Use those to solve the field equations.

9.1. Solving for real B, a, b, c, d

From the linear combinations of (7.25)–(7.28) that expose the combinations c ± d we obtain
the following equations for a ± b,
 
f 2 B
f2
f1 f4
(a + b)(c + d) = 2 2 + 2 − f4 f 5 ,
f4 f2 f2
2


f B f2 f1 f4
(a − b)(c − d) = 2 2 − 2 + f4 f 5 . (9.1)
f4 f2 f2
The remaining equations may be expressed as follows,

f4 f1 f4 2f 2
3 − 2 − 2f4 f5 = 2 4
ad, (9.2)
f1 f2 f B
f1
f2
2f 2
+ = 2 4
ac. (9.3)
f1 f2 f B
They will be important in the next subsection.

9.2. Solving the constraint

From the constraint (8.10), the combination involving a, b, c, d and f 2 B


, is eliminated by
using the dilatino equation (8.12) for c, d and the f 2 B
term, and the product of the two equations
in (9.1) for a, b, yielding
 2   
f
f
f
f
f1
f2
f12 f42 f42 f1 f42 f5
3 2+ 4 +2 2 + 4 + +3− −6 +4 = 0. (9.4)
f2 f4 f2 f4 f1 f2 f24 f22 f22
Next, we use (7.29) to replace the first term,
  
f2
f4
f1
f2
f12 f42 f42 f1 f42 f5
+ + + −3 +2 = 0. (9.5)
f2 f4 f1 f2 f24 f22 f22
Using the lhs of (9.2), and the fact that c = 0 in the equation above, we find,
 

f f
f1 f4
a c 2 + 4 − d 2 = 0. (9.6)
f2 f4 f2
Thus, the constraint factorizes when reduced to the subspace of supersymmetric solutions. We
study the vanishing of each factor in turn.
E. D’Hoker et al. / Nuclear Physics B 757 (2006) 79–116 107

9.3. The case a = 0

The vanishing of a in (9.3) and (9.2) together with the c = 0 implies that,
3 1 f1
f1 f2 = ρ, f5 = − (9.7)
2f1 2 f22
with ρ a constant, which will be fixed later. The remaining differential equation for f2 is,
 
f2
f 1 f4 f4
2c = 3d − . (9.8)
f2 f22 f1
The reduced field equation for a in (8.8) for a = 0 simplifies to,
f1 2

b= f B c. (9.9)
3f4
Using this expression to eliminate b from (7.25) and (7.26) yields c and d,
 
1 1 1/2
c=3 − , (9.10)
f22 f12
f2

cd = 6 . (9.11)
f1 f2 f4
Combining the result for c with the f5 equation in (8.11), we find that f12 f22 = ρ 2 = 3C1 /2.
The field equations for a, b, c, d of (8.8) are now satisfied as follows. The equation for a is
automatic using (9.9); the equation for b is automatic using (9.9); the equation for c is automatic
using (9.11). Finally, the equation for d is handled as follows. Start with the dilatino equation
(8.12) as a definition of d, take the derivative and use the field equations in (8.8) for c, d. One
arrives at
 f
f

4 c2 − d 2 2 − 8c2 4 + 8f4 f5 cd − 2f 2 B
bd = 0. (9.12)
f2 f4
Using (9.9) to eliminate b implies that the first and last terms in (9.12) cancel one another, leaving
the following linear relation between c and d, after eliminating f5 using (9.7),
 
f
3 f4 1 f1 f4
c 4= − d. (9.13)
f4 2 f1 2 f22
Eliminating now f4 /f1 using (9.8), we get


f f
f1 f4
c 2 + 4 − d 2 = 0. (9.14)
f2 f4 f2
This is the second factor in the relation (9.6).

9.4. Exact solution of the case a = 0 via hyper-elliptic integrals

Combining the relations (9.14) and (8.12), we may solve completely for c and d,
 
C22 C22 f24
c = 4 8 6,
2
d = 4 8 6 1− 2 2 .
2
(9.15)
f 1 f2 f4 f1 f2 f4 f 1 f4
108 E. D’Hoker et al. / Nuclear Physics B 757 (2006) 79–116

Eliminating c between the expression above and the one already obtained in (9.10), we obtain a
polynomial relation between f1 , f2 , and f4 ,
  1
f14 f26 − f12 f28 f46 = C22 . (9.16)
9
The function f1 may be eliminated using (9.7), and the resulting equations (7.29) and (9.16) may
be expressed in terms of f4 and the composite ψ defined by f2 f4 = ρ/ψ , so that
ρ2 C22 2
f44 = + ψ , (9.17)
ψ 4 9ρ 6

2
ψ
= ρ −4 f48 ψ 6 − 1. (9.18)
ψ
Eliminating f4 gives a single genus 5 decoupled hyper-elliptic equation for ψ, or equivalently
gives a genus 3 decoupled equation for Ψ ≡ ψ 2 ,
 2
C2

)2 = 1 + 28 ψ 6 − ψ 2 , (9.19)

 
1
2 C22 3 2
(Ψ ) = Ψ 1 + 8 Ψ − Ψ 2. (9.20)
4 9ρ

9.5. The general case

The general case is specified by the vanishing of the second factor in (9.6) only. This readily
allows for the solution of c and d in terms of the functions f1 , f2 , f4 , and we find,
 
C2 f 4 1/2
c = 2 4 3, d = vc, v = 1 − 22 2 . (9.21)
f1 f 2 f4 f1 f 4
The remaining susy variation equations are equivalent to
f1 f42 f1
f2
f1 f42 f4 f 1 f 4
2a =+ + , 2av = +3 − 2 − 2f4 f5 ,
f22 f1 f2 f22 f1 f2
f1 f42 f1
f
f1 f 2 f4 f1 f 4
2bv =− +3 2, 2b 24 = −3 + 5 2 − 2f4 f5 . (9.22)
f22 f1 f2 f2 f1 f2
Eliminating a, b gives
 
f1
f2
f4 f1 f4
v + + = +3 − 2 − 2f4 f5 ,
f1 f2 f1 f2



1 f f f4 f1 f4
− 1 + 3 2 = −3 + 5 2 − 2f4 f5 . (9.23)
v f1 f2 f1 f2
Finally, eliminating f5 and a using (8.11) gives the following equations,
     
1 f1
3 f2
f 4 f1 f4
v+ + v− =6 − 2 ,
v f1 v f2 f1 f2


f1 f2
f1 f 3 2 f 2f 6 f 2f 4 C22 C1 f44
+ + 6 44 = 36 1 84 − 12 1 64 + 4 2 12 − 36 . (9.24)
f1 f2 f2 f2 f2 f1 f2 f42 f28
E. D’Hoker et al. / Nuclear Physics B 757 (2006) 79–116 109

Fig. 3. The potential V (ψ) for C22 /ρ 8 = 0.36.

In attempting to separate these three equations, it appears natural to define the following new
combinations of the functions f1 , f2 , f3 ,

f 1 f4
ψ1 ≡ f 2 f4 , ψ2 ≡ f 1 f2 , ψ3 ≡ 2 , v = 1 − ψ3−2 . (9.25)
f2
Together with equation (7.29), the equations then become,
ψ1

= ψ3 v,
ψ1
ψ
1 ψ3
ψ1
v 2+ = −5ψ3 + 6 ,
ψ2 v ψ3 ψ2


ψ2 ψ1 ψ32 2 ψ 2ψ 4 ψ1 ψ33 C2ψ 2 ψ1 ψ33
+6 v = 36 1 2 3 − 12 + 64 24 34 − 36C1 . (9.26)
ψ2 ψ2 ψ2 ψ2 ψ1 ψ2 ψ23
So far, we have not succeeded in decoupling these equations.

10. Numerical results

For the special case a = 0 it was shown in Section 9.3 that the existence of an unbroken
supersymmetry is equivalent to the existence of solutions to Eq. (9.19) which can be viewed as
describing the motion if a particle with coordinate ψ in a potential V (ψ). (See Fig. 3.)

)2 + V (ψ) = 0, (10.1)
where the potential is given by
 2
C2
V (ψ) = − 1 + 28 ψ 6 + ψ 2 . (10.2)

The complete form of the solution is determined once ψ is found. Near the boundary of AdS
the metric function f4 behaves like f4 ∼ 1/(μ ∓ μ0 ) as μ → ±μ0 . It follows from (9.17) that
ψ vanishes as ψ ∼ (μ ∓ μ0 ) in this limit. On the other hand the limit ψ → ∞ corresponds
to the metric becoming singular. A regular Janus like solution can only exist if the potential is
somewhere positive, since in this case there are two allowed regions 0 < ψ < ψ1 and ψ2 < ψ <
∞. For the nonsingular Janus solution ψ takes values in the first region. It is easy to see that this
110 E. D’Hoker et al. / Nuclear Physics B 757 (2006) 79–116

Fig. 4. (a) Value of μ0 as a function of C22 /ρ 8 , (b) dilaton for three values of parameters.

Fig. 5. (a) f4 for three values of parameters, (b) f1 and f2 for three values of parameters.

implies a condition on the parameters of the potential


C22 55
8
< 6 4. (10.3)
ρ 2 3
The metric for the undeformed AdS5 corresponds to f4 (μ) = 1/ cos(μ) with range μ ∈
[−π/2, π/2]. For the Janus solution the range μ is increasing μ ∈ [−μ0 , μ0 ], where μ0 > π/2.
The range depends on C22 /ρ 8 and diverges when this parameter approaches its critical value
(10.3) (see Fig. 4(a)).
In the following we present the results for the solution for three choices of parameters ρ and
C2 . We have fixed ρ = 1.4 and picked C2 = 0.445, C2 = 0.245 and C2 = 0.045. The dilaton can
be determined by numerically integrating (8.3), the plot of the dilaton for the three choices of
parameters is given in Fig. 4(b),
The metric function f4 is related to ψ by Eq. (9.17), it diverges at μ = ±μ0 which corresponds
to the two AdS boundary components. The metric function f2 and f1 are determined by the
relations f2 f4 = ρ/ψ and f1 f2 = ρ. (See Fig. 5.)
Note that since f1 is the scale factor of the S 1 fiber and f2 is the scale factor of the CP2 base
in the squashed sphere, the fact that f1 → f2 as one approaches the AdS boundary components
E. D’Hoker et al. / Nuclear Physics B 757 (2006) 79–116 111

Fig. 6. (a) AST function c for three values of parameters, (b) AST function d for three values of parameters.

μ → ±μ0 , means that the sphere becomes ‘un-squashed’ and the SO(6) symmetry is restored at
the AdS boundary. The third rank anti-symmetric tensor is determined by ψ via Eq. (9.15) and
the plot for the two function c and d is given by Fig. 6.

11. Holographic dual

The AdS/CFT correspondence relates 10-dimensional Type IIB supergravity fields to gauge
invariant operators on the N = 4 super-Yang–Mills side. In the following we briefly review some
aspects of this map. The Poincaré metric of Euclidean AdS5 is given by
 
1
ds = 2 dz +
2 2 2
dxi . (11.1)
z
i
Near the boundary of AdS5 , where z → 0, a scalar field Φm of mass m behaves as

Φm (z, x) ∼ φnon-norm (x)z4−Δ + φnorm (x)zΔ , (11.2)


where m2 = Δ(Δ − 4). The non-normalizable mode corresponds to insertion in the Lagrangian
of an operator OΔ with scaling dimension Δ. The boundary source can be determined from
(11.2) by

φnon-norm (x) = lim zΔ−4 Φ(z, x). (11.3)


z→0

If φnon-norm vanishes a non-zero φnorm corresponds to a non-vanishing expectation value OΔ  =


φnorm of the operators OΔ on the Yang–Mills side.
As reviewed in Section 3.2 for the Janus solution, the boundary geometry, and hence the
holographic dual, are more complicated. The asymptotic behavior of the solution obtained in
Section 9.3 is readily obtained using power series expansion near the boundary components.
Employing Poincaré coordinates for the AdS4 slices, the asymptotic form of the non-compact
part of the ten dimensional metric can be obtained by expanding f4 (μ) near the two boundary
components μ = ±μ0
 
1 3

ds =
2
dz +
2
dxi + z dμ + O (μ ∓ μ0 )0 .
2 2 2
(11.4)
(μ ∓ μ0 ) z
2 2
i=1
112 E. D’Hoker et al. / Nuclear Physics B 757 (2006) 79–116

The boundary is reached when (μ ∓ μ0 )z → 0, and consists of one component with μ = μ0 and
one component with μ = −μ0 . The complete boundary corresponds to two 4-dimensional half
spaces joined by a R3 interface located at z = 0.
The asymptotic behavior of the dilaton near the boundaries is given by
C2
φ(μ) = φ0± + (μ ∓ μ0 )4 + O (μ ∓ μ0 )6 . (11.5)
2ρ 4
The dilaton corresponds to a dimension Δ = 4 operator. Hence it follows from (11.3) that there
is a source φ0± of the operator dual to the dilaton in the two boundary components. This is
interpreted on the Yang–Mills side as a theory where the gauge coupling takes two different
values on the half spaces separated by a planar interface.
The asymptotic behavior of the anti-symmetric tensor field components c and d is,
C2
c(μ) = 9/2
(μ0 ∓ μ)3 + O (μ ∓ μ0 )5 ,
ρ
C2
d(μ) = 9/2 (μ ∓ μ0 )3 + O (μ ∓ μ0 )5 . (11.6)
ρ
The (μ0 ∓ μ)3 behavior of the functions c and d is in agreement with the fact that the low-
est Kaluza–Klein modes on the S 5 of the anti-symmetric rank 2 tensor field is associated with
dimension Δ = 3 operator [40,41].
In the light of (11.3), the c ∼ (μ ∓ μ0 )Δ behavior of (11.6) seems to suggest that there is no
source for the dual operator of the anti-symmetric tensor. However this conclusion is not correct.
For the Janus metric the appropriate rescaling of the field to extract the non-normalizable mode
is given by

cnon-norm = lim  Δ−4 c(μ)


→0
1 C2
= lim (μ ∓ μ0 )3
→0  ρ 9/2

(μ ∓ μ0 )2 C2
= lim , (11.7)
(μ∓μ0 )z→0 z ρ 9/2
where  = (μ ∓ μ0 )z was used. For a point on the boundary which is away from the three-
dimensional interface one has z = 0 and it follows from (11.7) that the source for the dual
operator vanishes away from the interface. However for the interface one has z = 0 and cnon-norm
in (11.7) diverges. This behavior indicates the presence of a delta function source for the
dual
 Δ = 3 operator on the interface, since the integral over a small disk around the interface
dμ dz zcnon-norm is finite.
The metric functions f1 and f2 behave as
 
√ C2
f1 (μ) = ρ 1 + 2 8 (μ ∓ μ0 )6 + O (μ ∓ μ0 )8 ,
36ρ
 
√ C22
f2 (μ) = ρ 1 − (μ ∓ μ0 ) + O (μ ∓ μ0 ) .
6 8
(11.8)
36ρ 8
Repeating the analysis of the anti-symmetric tensor for the metric fields reveals that there no
additional operator sources turned on by f1 and f2 .
E. D’Hoker et al. / Nuclear Physics B 757 (2006) 79–116 113

12. Conclusions

In this paper, a supersymmetric generalization of the Janus solution of ten-dimensional type


IIB superstring theory was found. The solution was constructed by imposing the condition
of the existence of a preserved supersymmetry on the most general ansatz compatible with
SO(2, 3) × SU(3) × U (1)β × SL(2, R) symmetry. The supergravity solution is in agreement with
the field theoretical analysis of [15,19]. The restoration of supersymmetry requires a non-trivial
antisymmetric tensor B(2) . The dual holographic interpretation of the presence of the B(2) ten-
sor field is given by turning on a dimension Δ = 3 operator on the interface. There are several
questions relating to our work which are worth pursuing:
The generalized non-supersymmetric Janus solution we have found can be represented exactly
using elliptic functions. Therefore, it may be possible to calculate correlation functions in our
background exactly, and compare supergravity and field theory predictions beyond conformal
perturbation theory.
In [19] a detailed analysis of the interface field theory found that there are additional pos-
sibilities to obtain a larger unbroken supersymmetry. In particular a case with N = 4 interface
supersymmetry was found. The internal symmetry of this theory is SU(2) × SU(2). A possible
Ansatz in this case replaces the squashed five sphere used in this paper with the T 1,1 manifold
[42,46] which realizes the SU(2) × SU(2) symmetry.
The AdS5 × S5 background is famously obtained as the near horizon limit of N D3-branes.
Various defect conformal field theories can holographically be obtained via near horizon limits of
intersecting D-brane systems. The brane interpretation of the non-supersymmetric Janus, as well
as our supersymmetric solution, is unknown at this point. The fact that both the dilaton as well as
the AST field are sourced by fivebranes suggests that a brane realization of the supersymmetric
Janus solution could be given by the near horizon limit of intersecting D3/D5 branes. However,
the only known solutions of this kind treat the five branes in the probe approximation [32,33].
Whether a fully back reacted solution is related to Janus is a very interesting question, which we
plan to investigate in the future.

Acknowledgements

It is a pleasure to acknowledge helpful conversations with Iosif Bena, Per Kraus, and Norisuke
Sakai. This work was supported in part by National Science Foundation (NSF) grant PHY-04-
56200. E.D. is grateful to the Kavli Institute for Theoretical Physics (KITP) for their hospitality
and support under NSF grant PHY-99-07949. M.G. is grateful to the Harvard Particle Theory
group for hospitality while this work was being completed.

Appendix A. Realization of the Γ -matrices

It will be useful to choose a well-adapted basis for Γ -matrices,


Γ μ = σ1 ⊗ γ μ ⊗ I4 , μ = 0, 1, 2, 3, 4,
Γ m
= σ 2 ⊗ I4 ⊗ γ ,
m
m = 5, 6, 7, 8, 9 (A.1)
for which we have
Γ 11 = Γ 0123456789 = σ3 ⊗ I4 ⊗ I2 ⊗ I2 ,
B = Γ 2568 = σ3 ⊗ γ 2 ⊗ σ1 ⊗ σ2 . (A.2)
114 E. D’Hoker et al. / Nuclear Physics B 757 (2006) 79–116

The following convention for the 4 × 4 matrices γ A will be adopted,

iγ 0 = σ2 ⊗ I2 , γ 5 = σ3 ⊗ σ3 ,
γ 1 = σ1 ⊗ I2 , γ 6 = σ1 ⊗ I2 ,
γ 2 = σ3 ⊗ σ2 , γ 7 = σ2 ⊗ I2 ,
γ 3 = σ3 ⊗ σ1 , γ 8 = σ3 ⊗ σ1 ,
γ 4 = σ3 ⊗ σ3 , γ 9 = σ3 ⊗ σ2 . (A.3)
The Γ -matrices in the complex frame associated with CP2 and (5.7), are as follows,

Γ z1 = Γ 6 + iΓ 7 = 2σ2 ⊗ I4 ⊗ σ+ ⊗ I2 ,
Γ z̄1 = Γ 6 − iΓ 7 = 2σ2 ⊗ I4 ⊗ σ− ⊗ I2 ,
Γ z2 = Γ 8 + iΓ 9 = 2σ2 ⊗ I4 ⊗ σ3 ⊗ σ+ ,
Γ z̄2 = Γ 8 − iΓ 9 = 2σ2 ⊗ I4 ⊗ σ3 ⊗ σ− . (A.4)
The following combinations will also be useful in evaluating the connection form,

Γ ij = I2 ⊗ γ ij ⊗ I4 , i, j = 0, 1, 2, 3,
Γ i4
= I 2 ⊗ γ γ ⊗ I4 ,
i 4

Γ 45
= iσ3 ⊗ γ 4 ⊗ γ 5 ,
Γ 4a = iσ3 ⊗ γ 4 ⊗ γ a ,
Γ 5a = I2 ⊗ I4 ⊗ γ 5 γ a ,
Γ ab = I2 ⊗ I4 ⊗ γ ab , a, b = 6, 7, 8, 9. (A.5)

References

[1] J.M. Maldacena, The large N limit of superconformal field theories and supergravity, Adv. Theor. Math. Phys. 2
(1998) 231, Int. J. Theor. Phys. 38 (1999) 1113, hep-th/9711200.
[2] S.S. Gubser, I.R. Klebanov, A.M. Polyakov, Gauge theory correlators from non-critical string theory, Phys. Lett.
B 428 (1998) 105, hep-th/9802109.
[3] E. Witten, Anti-de Sitter space and holography, Adv. Theor. Math. Phys. 2 (1998) 253, hep-th/9802150.
[4] E. D’Hoker, D.Z. Freedman, Supersymmetric gauge theories and the AdS/CFT correspondence, in: S.S. Gubser,
J.D. Lykken (Eds.), Strings, Branes, and Extra Dimensions, World Scientific, Singapore, 2004, hep-th/0201253.
[5] O. Aharony, S.S. Gubser, J.M. Maldacena, H. Ooguri, Y. Oz, Large N field theories, string theory and gravity, Phys.
Rep. 323 (2000) 183, hep-th/9905111.
[6] D.Z. Freedman, S.S. Gubser, K. Pilch, N.P. Warner, Renormalization group flows from holography supersymmetry
and a c-theorem, Adv. Theor. Math. Phys. 3 (1999) 363, hep-th/9904017.
[7] I.R. Klebanov, M.J. Strassler, Supergravity and a confining gauge theory: Duality cascades and chiSB-resolution of
naked singularities, JHEP 0008 (2000) 052, hep-th/0007191.
[8] J. Polchinski, M.J. Strassler, The string dual of a confining four-dimensional gauge theory, hep-th/0003136.
[9] J.M. Maldacena, C. Nunez, Towards the large N limit of pure N = 1 super-Yang–Mills, Phys. Rev. Lett. 86 (2001)
588, hep-th/0008001.
[10] D. Bak, M. Gutperle, S. Hirano, A dilatonic deformation of AdS(5) and its field theory dual, JHEP 0305 (2003)
072, hep-th/0304129.
[11] D.Z. Freedman, C. Nunez, M. Schnabl, K. Skenderis, Fake supergravity and domain wall stability, Phys. Rev. D 69
(2004) 104027, hep-th/0312055.
[12] I. Papadimitriou, K. Skenderis, Correlation functions in holographic RG flows, JHEP 0410 (2004) 075, hep-
th/0407071.
E. D’Hoker et al. / Nuclear Physics B 757 (2006) 79–116 115

[13] A. Celi, A. Ceresole, G. Dall’Agata, A. Van Proeyen, M. Zagermann, On the fakeness of fake supergravity, Phys.
Rev. D 71 (2005) 045009, hep-th/0410126.
[14] J. Sonner, P.K. Townsend, Dilaton domain walls and dynamical systems, Class. Quantum Grav. 23 (2006) 441,
hep-th/0510115.
[15] A.B. Clark, D.Z. Freedman, A. Karch, M. Schnabl, The dual of Janus—an interface CFT, Phys. Rev. D 71 (2005)
066003, hep-th/0407073.
[16] S.S. Gubser, Dilaton-driven confinement, hep-th/9902155.
[17] D. Bak, M. Gutperle, S. Hirano, N. Ohta, Dilatonic repulsons and confinement via the AdS/CFT correspondence,
Phys. Rev. D 70 (2004) 086004, hep-th/0403249.
[18] A. Kehagias, K. Sfetsos, On running couplings in gauge theories from type-IIB supergravity, Phys. Lett. B 454
(1999) 270, hep-th/9902125.
[19] E. D’Hoker, J. Estes, M. Gutperle, Interface Yang–Mills, supersymmetry, and Janus, UCLA/06/TEP/03 preprint
(2006), hep-th/0603013.
[20] A. Clark, A. Karch, Super Janus, JHEP 0510 (2005) 094, hep-th/0506265.
[21] K. Behrndt, M. Cvetic, Bent BPS domain walls of D = 5, N = 2 gauged supergravity coupled to hypermultiplets,
Phys. Rev. D 65 (2002) 126007, hep-th/0201272.
[22] A. Ceresole, G. Dall’Agata, R. Kallosh, A. Van Proeyen, Hypermultiplets, domain walls and supersymmetric at-
tractors, Phys. Rev. D 64 (2001) 104006, hep-th/0104056.
[23] G.L. Cardoso, G. Dall’Agata, D. Lust, Curved BPS domain walls and RG flow in five dimensions, JHEP 0203
(2002) 044, hep-th/0201270.
[24] G. Lopes Cardoso, G. Dall’Agata, D. Lust, Curved BPS domain wall solutions in five-dimensional gauged super-
gravity, JHEP 0107 (2001) 026, hep-th/0104156.
[25] G.L. Cardoso, D. Lust, The holographic RG flow in a field theory on a curved background, JHEP 0209 (2002) 028,
hep-th/0207024.
[26] A. Ceresole, G. Dall’Agata, General matter coupled N = 2, D = 5 gauged supergravity, Nucl. Phys. B 585 (2000)
143, hep-th/0004111.
[27] J.H. Schwarz, Covariant field equations of chiral N = 2, D = 10 supergravity, Nucl. Phys. B 226 (1983) 269.
[28] P.S. Howe, P.C. West, The complete N = 2, D = 10 supergravity, Nucl. Phys. B 238 (1984) 181.
[29] J.P. Gauntlett, D. Martelli, J. Sparks, D. Waldram, Supersymmetric AdS(5) solutions of type IIB supergravity, hep-
th/0510125.
[30] U. Gran, J. Gutowski, G. Papadopoulos, D. Roest, Systematics of IIB spinorial geometry, hep-th/0507087.
[31] K. Pilch, N.P. Warner, N = 2 supersymmetric RG flows and the IIB dilaton, Nucl. Phys. B 594 (2001) 209, hep-
th/0004063.
[32] A. Karch, L. Randall, Open and closed string interpretation of SUSY CFT’s on branes with boundaries, JHEP 0106
(2001) 063, hep-th/0105132.
[33] O. DeWolfe, D.Z. Freedman, H. Ooguri, Holography and defect conformal field theories, Phys. Rev. D 66 (2002)
025009, hep-th/0111135.
[34] O. Aharony, O. DeWolfe, D.Z. Freedman, A. Karch, Defect conformal field theory and locally localized gravity,
JHEP 0307 (2003) 030, hep-th/0303249.
[35] J. Erdmenger, Z. Guralnik, I. Kirsch, Four-dimensional superconformal theories with interacting boundaries or
defects, Phys. Rev. D 66 (2002) 025020, hep-th/0203020.
[36] S. Yamaguchi, AdS branes corresponding to superconformal defects, JHEP 0306 (2003) 002, hep-th/0305007.
[37] B.A. Dubrovin, A.T. Fomenko, S.P. Novikov, Modern Geometry—Methods and Applications, Part III, Introduction
to Homology Theory, Graduate Text in Mathematics, vol. 124, Springer-Verlag, Verlag, 1990.
[38] E. D’Hoker, Invariant effective actions, cohomology of homogeneous spaces and anomalies, Nucl. Phys. B 451
(1995) 725, hep-th/9502162;
E. D’Hoker, S. Weinberg, General effective actions, Phys. Rev. D 50 (1994) 6050, hep-ph/9409402.
[39] J. Polchinski, String Theory, vol II, Cambridge Univ. Press, Cambridge, 1998, pp. 90, 91, see corresponding errata
on http://www.itp.ucsb.edu/joep/bigbook.html.
[40] H.J. Kim, L.J. Romans, P. van Nieuwenhuizen, The mass spectrum of chiral N = 2, D = 10 supergravity on S 5 ,
Phys. Rev. D 32 (1985) 389.
[41] M. Gunaydin, N. Marcus, The spectrum of the S 5 compactification of the chiral N = 2, D = 10 supergravity and
the unitary supermultiplets of U(2, 2/4), Class. Quantum Grav. 2 (1985) L11.
[42] L.J. Romans, New compactifications of chiral N = 2, D = 10 supergravity, Phys. Lett. B 153 (1985) 392.
[43] C.N. Pope, N.P. Warner, Two new classes of compactifications of D = 11 supergravity, Class. Quantum Grav. 2
(1985) L1.
116 E. D’Hoker et al. / Nuclear Physics B 757 (2006) 79–116

[44] J.T. Liu, H. Sati, Breathing mode compactifications and supersymmetry of the brane-world, Nucl. Phys. B 605
(2001) 116, hep-th/0009184.
[45] K. Pilch, N.P. Warner, N = 1 supersymmetric renormalization group flows from IIB supergravity, Adv. Theor.
Math. Phys. 4 (2002) 627, hep-th/0006066.
[46] I.R. Klebanov, E. Witten, Superconformal field theory on threebranes at a Calabi–Yau singularity, Nucl. Phys. B 536
(1998) 199, hep-th/9807080.
[47] O. DeWolfe, D.Z. Freedman, S.S. Gubser, A. Karch, Modeling the fifth dimension with scalars and gravity, Phys.
Rev. D 62 (2000) 046008, hep-th/9909134.
Nuclear Physics B 757 (2006) 117–145

Quantum mechanical sectors in thermal N = 4


super-Yang–Mills on R × S 3
Troels Harmark ∗ , Marta Orselli
The Niels Bohr Institute and Nordita, Blegdamsvej 17, 2100 Copenhagen Ø, Denmark
Received 15 June 2006; accepted 25 August 2006
Available online 15 September 2006

Abstract
We study the thermodynamics of U (N ) N = 4 super-Yang–Mills (SYM) on R × S 3 with non-zero
chemical potentials for the SU(4) R-symmetry. We find that when we are near a point with zero temperature
and critical chemical potential, N = 4 SYM on R×S 3 reduces to a quantum mechanical theory. We identify
three such critical regions giving rise to three different quantum mechanical theories. Two of them have a
Hilbert space given by the SU(2) and SU(2|3) sectors of N = 4 SYM of recent interest in the study of
integrability, while the third one is the half-BPS sector dual to bubbling AdS geometries. In the planar limit
the three quantum mechanical theories can be seen as spin chains. In particular, we identify a near-critical
region in which N = 4 SYM on R × S 3 essentially reduces to the ferromagnetic XXX1/2 Heisenberg spin
chain. We find furthermore a limit in which this relation becomes exact.
© 2006 Elsevier B.V. All rights reserved.

1. Introduction

The thermodynamics of large N U (N) N = 4 super-Yang–Mills (SYM) on R×S 3 has proven


to be interesting for several reasons. It has a confinement/deconfinement phase transition like in
QCD that can be studied even at weak coupling [1]. This phase transition is conjectured to corre-
spond to the Hagedorn phase transition for the dual type IIB string theory on AdS5 × S 5 , which is
in accordance with the fact that the large N N = 4 SYM theory has a Hagedorn spectrum [2–4].
This is very interesting since it means that we can study what happens beyond the Hagedorn
transition on the weakly coupled gauge theory side. For large coupling the same phase transition
corresponds to the Hawking–Page phase transition for black holes in anti-de Sitter space, which

* Corresponding author.
E-mail addresses: harmark@nbi.dk (T. Harmark), orselli@nbi.dk (M. Orselli).

0550-3213/$ – see front matter © 2006 Elsevier B.V. All rights reserved.
doi:10.1016/j.nuclphysb.2006.08.022
118 T. Harmark, M. Orselli / Nuclear Physics B 757 (2006) 117–145

is a phase transition in semi-classical gravity [1,5]. Thus, by studying the thermodynamics of


N = 4 SYM we can hope to learn about such important subjects as what is beyond the Hagedorn
transition, confinement in QCD and phase transitions in gravity.
In this paper we find that thermal U (N ) N = 4 SYM has quantum mechanical sectors, by
which we mean that near certain critical points most of the degrees of freedom of N = 4 SYM
can be integrated out and only a small subset, that we can regard as quantum mechanical, re-
mains. These critical points arise in the study of the thermodynamics of U (N ) N = 4 SYM on
R × S 3 with non-zero chemical potentials corresponding to the three R-charges for the SU(4) R-
symmetry of N = 4 SYM. Our main result is that when we are near a point with zero temperature
and critical chemical potentials, N = 4 SYM reduces to one out of three simple quantum me-
chanical theories. Furthermore, for large N these three quantum mechanical theories are mapped
in a precise way to spin chain theories.
Denoting the three chemical potentials of N = 4 SYM as Ω1 , Ω2 , Ω3 and setting Ω1 =
Ω2 = Ω, Ω3 = 0, we can write one of the near-critical regions that we study as
T  1, 1 − Ω  1, λ  1, (1.1)
where T is the temperature and λ is the ’t Hooft coupling of N = 4 SYM. In this region we
are close to the critical point (T , Ω1 , Ω2 , Ω3 ) = (0, 1, 1, 0). We show in this paper that in the
region (1.1) N = 4 SYM on R × S 3 reduces to a quantum mechanical theory with the Hilbert
space consisting of all multi-trace operators made out of the letters Z and X, where Z and X are
complex scalars of N = 4 SYM with R-symmetry weights (1, 0, 0) and (0, 1, 0). This is precisely
the so-called SU(2) sector that has been discussed in recent developments on the integrability of
N = 4 SYM [6–10].
We find that it is natural to reformulate N = 4 SYM in the region (1.1) in terms of the rescaled
temperature T̃ ≡ T /(1 − Ω). Writing the dilatation operator of N = 4 SYM as D = D0 + λD2 +
O(λ2 ) where D0 is the zeroth order dilatation operator and λ is the ’t Hooft coupling, we can
write the leading terms of the Hamiltonian of our quantum mechanical theory as
H = D0 + λ̃D2 , (1.2)
where λ̃ ≡ λ/(1 − Ω) is a rescaled coupling. This resembles the leading terms of the dilatation
operator of the SU(2) sector except for the rescaled coupling λ̃. The first correction to (1.2) is of
order λ̃λ. Our result is thus that in the near-critical region (1.1) N = 4 SYM on R × S 3 reduces
to a quantum mechanical theory with temperature T̃ , Hamiltonian (1.2) (for the leading terms)
and with the Hilbert space corresponding to the SU(2) sector of N = 4 SYM.
For large N we can focus on single-trace operators of a certain length L. Such operators can
be thought of as periodic spin chains of length L. The Hamiltonian (1.2) is then L + λ̃D2 and
D2 is known to correspond to the ferromagnetic XXX1/2 Heisenberg spin chain Hamiltonian.
Thus, for N = ∞ our result is that thermal N = 4 SYM on R × S 3 reduces to the ferromagnetic
XXX1/2 Heisenberg spin chain, in the sense that we have a precise relation between the partition
functions of the two theories.
A further result of this paper is that if we take the limit
T → 0, Ω → 1, λ → 0,
T λ
T̃ = fixed, λ̃ = fixed (1.3)
1−Ω 1−Ω
the Hamiltonian (1.2) becomes exact with the Hilbert space being the SU(2) sector. Hence, for
N = ∞ and in the limit (1.3) we have that the relation between the partition function of N = 4
T. Harmark, M. Orselli / Nuclear Physics B 757 (2006) 117–145 119

SYM on R × S 3 and that of the ferromagnetic XXX1/2 Heisenberg spin chain is exact, i.e. we
find that
∞ 
 ∞
1 −nL/T̃ (XXX)
log ZN =4 (T̃ ) = e ZL (T̃ /n), (1.4)
n
n=1 L=1
(XXX)
where ZN =4 is the partition function for N = 4 SYM on R×S 3 and ZL is the partition func-
tion for the ferromagnetic XXX1/2 Heisenberg spin chain of length L with Hamiltonian λ̃D2 .
We consider furthermore two other near-critical regions. Near (T , Ω1 , Ω2 , Ω3 ) = (0, 1, 1, 1)
we find that N = 4 SYM on R × S 3 reduces to a quantum mechanical theory in the so-called
SU(2|3) sector of N = 4 SYM which also recently has been considered in the study of integra-
bility [8,11]. This sector consists of three complex scalars and two complex fermions. We find
similar results in this sector as for the SU(2) sector.
Near (T , Ω1 , Ω2 , Ω3 ) = (0, 1, 0, 0) we find instead that N = 4 SYM on R × S 3 reduces to
the half-BPS sector consisting of multi-trace operators made of a single complex scalar Z. This
sector is precisely the half-BPS sector dual to the bubbling AdS geometries of [12]. As part of
this, it also contains the states dual to the vacuum of the maximally supersymmetric pp-wave
background [13,14], to AdS5 × S 5 [15], and to giant gravitons in AdS5 × S 5 [16]. The reduction
of N = 4 SYM to the half-BPS sector was previously considered in [17].
Finally, we consider the one-loop partition function for planar N = 4 SYM on R × S 3 with
non-zero chemical potentials and we find the corrected Hagedorn temperature, generalizing [18].
We find furthermore the explicit form of the corrected partition functions and Hagedorn tempera-
ture for the SU(2) and SU(2|3) sectors. As a consistency check, we verify that one gets the same
result by taking the limit of the full partition function as what one gets from the reduced partition
functions.
This paper is structured as follows. In Section 2 we consider free N = 4 SYM on R × S 3 . We
compute the partition function with non-zero chemical potentials in Section 2.1 and we find the
Hagedorn temperature in Section 2.2. In Section 2.3 we identify the three near-critical regions
and we show the reductions to the half-BPS sector, the SU(2) sector and the SU(2|3) sector.
We consider furthermore these reductions in the oscillator basis of N = 4 SYM in Appendix A.
Finally in Section 2.4 we consider the thermodynamics above the Hagedorn temperature.
In Section 3 we consider the three near-critical regions for interacting N = 4 SYM on R × S 3
and find that we still have the reductions to the half-BPS sector, the SU(2) sector and the SU(2|3)
sector, but now with a non-trivial Hamiltonian. For N = ∞ we relate this Hamiltonian to spin
chain Hamiltonians, in particular we find that the SU(2) sector has a Hamiltonian with the leading
part given by the ferromagnetic XXX1/2 Heisenberg spin chain. We briefly review the XXX1/2
Heisenberg spin chain in Appendix B.
In Section 4 we consider the low temperature limit for the near-critical region in which N = 4
SYM reduces to the SU(2) sector. In this case we find for large N that the ferromagnetic XXX1/2
Heisenberg spin chain governs the dynamic and from this we can find which states we are driven
towards as we take the temperature to zero.
In Section 5 we write down the decoupling limit mentioned above, from which it follows
for the SU(2) sector that we have an exact relation between N = 4 SYM and the XXX1/2
Heisenberg spin chain for N = ∞.
In Section 6 we consider the one-loop correction to the thermal partition function of large N
U (N ) N = 4 SYM on R × S 3 . We show how to compute the partition function with non-zero
chemical potentials, following [18]. We have put part of this computation in Appendix C. We
120 T. Harmark, M. Orselli / Nuclear Physics B 757 (2006) 117–145

find the one-loop corrected Hagedorn temperature both for small chemical potential and near the
critical points. Near the critical points we also find the partition function explicitly, and we find
that the one-loop partition function of N = 4 SYM on R × S 3 indeed correctly reduces to the
one of the reduced theories.
In Section 7 we present our conclusions and discuss future directions.
Note on related work: We note that during the work on this paper the article [19] appeared
with results that overlap with Sections 2.1 and 2.2.

2. Free thermal N = 4 SYM on R × S 3

We consider in this section the thermal partition function of N = 4 SYM on R × S 3 with


chemical potentials at zero coupling.

2.1. Calculation of the partition function

In this section we consider the generalization of the computation of the partition function
for U (N ) N = 4 SYM on R × S 3 at zero coupling gYM 2 = 0 in [2–4] to include the three chemical

potentials associated with the SU(4) R-symmetry of N = 4 SYM.


The partition function of U (N ) N = 4 SYM on R × S 3 is given by the trace of e−βH over all
of the physical states, where β = 1/T is the inverse temperature and H is the Hamiltonian. From
the state/operator correspondence we have that any state of U (N ) N = 4 SYM on R × S 3 can
be mapped to a gauge invariant operator of U (N ) N = 4 SYM on R4 . The Hamiltonian is then
mapped to the dilatation operator D (here and in the following we set the radius of S 3 to one).
The Gauss constraint for a U (N ) gauge theory on R × S 3 means that we can only have states
which are singlets of U (N ). For operators, this means that the set of operators we have are made
by combining single-trace operators, where each single-trace operator is made from combining
individual letters, which are the individual operator one can make using a single field of N = 4
SYM and the covariant derivative [2–4].
To include the chemical potential associated with the SU(4) R-symmetry of N = 4 SYM
we need to introduce the R-charges. Let R1 , R2 and R3 denote the Cartan generators of SU(4)
(corresponding to the standard Cartan generators of SO(6)). Then R1 , R2 and R3 are the three
R-charges of N = 4 SYM and corresponding to these we have three chemical potentials Ω1 ,
Ω2 and Ω3 . When computing a partition function in the grand canonical ensemble one should
compute the trace of e−βH +βΩ1 R1 +βΩ2 R2 +βΩ2 R2 over all the physical states.
For the free N = 4 SYM theory we should use the zeroth order dilatation operator D0 as the
Hamiltonian. We can then schematically write the full partition function in the grand canonical
ensemble as
 R 
Z(x, y1 , y2 , y3 ) = TrM x D0 y1R1 y2R2 y3 3 . (2.1)
Here we write M for the set of multi-trace operators (or rather the corresponding states) and we
introduce the useful book keeping devices
x ≡ e−β , yi ≡ eβΩi , i = 1, 2, 3. (2.2)
We note the important point that for finite N not all multi-trace operators are linearly indepen-
dent. Certain single-trace operators can for example by written in terms of multi-trace operators.
We therefore assume M to be defined such that all of the multi-trace operators in M are linearly
independent, since otherwise we would count too many states [2,4].
T. Harmark, M. Orselli / Nuclear Physics B 757 (2006) 117–145 121

To compute the partition function one should first find the partition function for a single letter.
To do this, we need to understand the possible letters one can have and what their conformal
dimensions and R-charges are. The field content of N = 4 SYM consists of 6 real scalars φa , a =
1, . . . , 6, a gauge boson Aμ and the complex fermionic fields ψAα , ψ̄α̇A , α, α̇ = 1, 2, A = 1, 2, 3, 4,
corresponding to 16 real fermionic components. The scalars all have conformal dimension 1, the
gauge boson also have dimension one while the fermions have dimension 3/2. With respect to the
SU(4) R-symmetry we have that the 6 scalars correspond to a [0, 1, 0] representation, the gauge
boson is a singlet under SU(4) R-symmetry, while the fermions correspond to a [1, 0, 0] and
a [0, 0, 1] representation of SU(4). With respect to (R1 , R2 , R3 ) we then have that for instance
the [0, 1, 0] representation corresponding to the 6 scalars have weights (±1, 0, 0), (0, ±1, 0) and
(0, 0, ±1). For use in following sections of this paper we define here the three complex scalars
Z = φ1 + iφ2 , X = φ3 + iφ4 and W = φ5 + iφ6 , corresponding to the weights (1, 0, 0), (0, 1, 0)
and (0, 0, 1), respectively.
The set of letters of N = 4 SYM, here denoted by A, is the set of all the different operators
on R4 that one can form by applying the covariant derivative an arbitrary number of times on
either one of the scalars φa , on the gauge field strength Fμν or on one of the fermions ψAα , ψ̄α̇A .
These operators should be independent of each other in the sense that two operators which are
related by the EOMs count as the same operator. It is well known [2–4] that a scalar on R × S 3
has letter partition function (x + x 2 )/(1 − x)3 , a fermion 2x 3/2 /(1 − x)3 and a gauge boson
(6x 2 − 2x 3 )/(1 − x)3 . Using this, we get the following letter partition function for N = 4 SYM
on R × S 3

z(x, y1 , y2 , y3 )
 R 
= TrA x D0 y1R1 y2R2 y3 3

x + x 2  2x 3/2  12
3 3
6x 2 − 2x 3 −1  −1 
= + y i + y + yi + yi 2 . (2.3)
(1 − x)3 (1 − x) 3 i
(1 − x)3
i=1 i=1

If we consider the large N case, we can for small enough energies E  N 2 ignore the
non-trivial relations between multi-trace operators, e.g., the set of single-trace operators is well
defined in this case. This enables us to make a purely combinatorical computation of the partition
function. One begins by computing the single-trace partition function. The single trace operators
are Tr(A1 A2 · · · AL ) with Ai ∈ A. Note that here and in the following we take the U (N ) trace to
be in the adjoint representation of U (N ). One can then use standard combinatorical techniques
to find the single-trace partition function as [2–4]

 ϕ(k)   
ZST (x, y1 , y2 , y3 ) = − log 1 − z ωk+1 x k , y1k , y2k , y3k , (2.4)
k
k=1

where we introduced the useful quantity ω = e2πi which is −1 if uplifted to a half-integer power,
following [18]. In this way we ensure that the fermionic part of the partition function has the
correct sign corresponding to fermionic statistics. In (2.4) ϕ(k) is the Euler totient function which
appears here due to the combinatorical complication that the single-trace operators have a cyclic
symmetry.
The complete partition function Z(x, y1 , y2 , y3 ) for U (N ) N = 4 SYM on R × S 3 with
N = ∞, which traces over all the multi-trace operators build from the single-trace operators,
122 T. Harmark, M. Orselli / Nuclear Physics B 757 (2006) 117–145

can then be found as



 1  
log Z(x, y1 , y2 , y3 ) = ZST ωn+1 x n , y1n , y2n , y3n
n
n=1
∞
  
=− log 1 − z ωk+1 x k , y1k , y2k , y3k . (2.5)
k=1
By a more careful analysis one can find the partition function for finite N , in which case
there are non-trivial relations between the multi-trace operators. The partition function for U (N )
N = 4 SYM on R × S 3 with chemical potentials is [4,20]
 ∞

1    k   † k 
Z(x, y1 , y2 , y3 ) = [dU ] exp k+1 k k k k
z ω x , y1 , y2 , y3 Tr U Tr U . (2.6)
k
k=1

Here [dU ] is the integral over the group U (N ) normalized such that [dU ] = 1. As mentioned
above, we take the trace over U (N ) to be in the adjoint representation.

2.2. Hagedorn temperature for non-zero chemical potentials

If we consider the N = ∞ partition function Eq. (2.5) for U (N ) N = 4 SYM on R × S 3 it is


clear that there is a singularity when

z(x, y1 , y2 , y3 ) = 1. (2.7)
This is the Hagedorn singularity of the partition function (2.5) [2–4] here generalized to in-
clude non-zero chemical potentials. It is easy to see that (2.7) with (2.3) for given chemical
potentials Ωi defines a critical temperature TH (Ω1 , Ω2 , Ω3 ). One can check from the partition
function (2.5) that there are no singularities for T < TH (Ω1 , Ω2 , Ω3 ).
For temperatures just below the Hagedorn temperature, write
TH − T  
z(x, y1 , y2 , y3 ) = 1 − + O (TH − T )2 (2.8)
TH C
for 0  TH (Ω1 , Ω2 , Ω3 ) − T  TH (Ω1 , Ω2 , Ω3 ) with C = C(Ω1 , Ω2 , Ω3 ). Then the partition
function for temperatures just below the Hagedorn temperature has the behavior
TH C
Z(T , Ω1 , Ω2 , Ω3 )  . (2.9)
TH − T
From this one can find that the density of states for single-trace operators is E −1 eE/TH [4].
Therefore, when N = ∞ we have a Hagedorn density of states for large energies.
For small chemical potentials it is straightforward to compute that the Hagedorn temperature
is

1  3    3
TH (Ω1 , Ω2 , Ω3 ) = + p1 Ωi2 + p2 Ωi2 Ωj2 + p3 Ωi4 + O Ωi6 ,
β0
i=1 i<j i=1

β0 = − log(7 − 4 3 ),
√ √
1 (18 − 5 3 ) (18 − 11 3 )
p1 = − √ , p2 = β0 , p 3 = β0 . (2.10)
6 3 1296 2592
T. Harmark, M. Orselli / Nuclear Physics B 757 (2006) 117–145 123

Fig. 1. The Hagedorn temperature TH as function of Ω in the case (Ω1 , Ω2 , Ω3 ) = (Ω, 0, 0).

Fig. 2. The Hagedorn temperature TH as function of Ω in the two cases (Ω1 , Ω2 , Ω3 ) = (Ω, Ω, 0), displayed on the
left, and (Ω1 , Ω2 , Ω3 ) = (Ω, Ω, Ω), displayed on the right.

In Figs. 1 and 2 we have displayed TH as a function of Ω for the three particular cases given by
(Ω1 , Ω2 , Ω3 ) = (Ω, 0, 0), (Ω1 , Ω2 , Ω3 ) = (Ω, Ω, 0) and (Ω1 , Ω2 , Ω3 ) = (Ω, Ω, Ω). As we
shall see in the following those three special cases are highly relevant for this paper. Note that
if we define R as being the charge related to the chemical potential Ω we have that R = R1 ,
R = R1 + R2 and R = R1 + R2 + R3 corresponds to the three cases, respectively.
For the case (Ω1 , Ω2 , Ω3 ) = (Ω, 0, 0) depicted in Fig. 1 we see that the behavior near the
critical point (T , Ω) = (0, 1) is

1 log(− log(1 − Ω))
TH (Ω) = − 1− + ··· (2.11)
log(1 − Ω) log(1 − Ω)
for 1 − Ω  1. Thus, the slope of the Hagedorn curve in the (T , Ω) diagram is zero in the critical
point (T , Ω) = (0, 1), as is also clear from Fig. 1.
For the case (Ω1 , Ω2 , Ω3 ) = (Ω, Ω, 0) depicted in the left part of Fig. 2 we have instead that
the behavior near the critical point (T , Ω) = (0, 1) is

1−Ω 2 − 1 log 2/(1−Ω)  
TH (Ω) = 1− e 2 + O e− log 2/(1−Ω) (2.12)
log 2 log 2
124 T. Harmark, M. Orselli / Nuclear Physics B 757 (2006) 117–145

for 1 − Ω  1. We see from this that the slope of the Hagedorn curve at the critical point
(T , Ω) = (0, 1) is − log 2.
Finally for the case (Ω1 , Ω2 , Ω3 ) = (Ω, Ω, Ω) depicted in the right part of Fig. 2 we have
that the behavior near the critical point (T , Ω) = (0, 1) is

1−Ω 6 − log 4/(1−Ω)  −2 log 4/(1−Ω) 
TH (Ω) = 1− e +O e (2.13)
log 4 log 4
for 1 − Ω  1. We see from this that the slope of the Hagedorn curve at the critical point
(T , Ω) = (0, 1) is − log 4.

2.3. Decoupling for near-critical chemical potentials

We now turn to examine what happens when the chemical potentials are near-critical, i.e.
when one or more of the chemical potentials Ωi are close to 1. From Figs. 1 and 2 we see
that to zoom in to a region where the chemical potentials are near-critical we also need to send
the temperature to zero. For x → 0 it is clear that z(x, yi ) → 0 unless we send one or more
of the yi to infinity (we restrict ourselves here to positive chemical potentials without loss of
generality). Write now yi = y αi where αi , i = 1, 2, 3, are numbers. Assume without loss of
generality 0  α3  α2  α1 = 1. From Eq. (2.3) we see then that we should take the limit

x → 0, xy = fixed. (2.14)
One can now see that we get three different limits depending on if one, two or three of the αi ,
i = 1, 2, 3, are equal to one. It is easy to see that this corresponds to sending either one, two or
three of the Ωi , i = 1, 2, 3, to 1 as T → 0. We can therefore restrict ourselves in the following to
the three cases (α1 , α2 , α3 ) ∈ {(1, 0, 0), (1, 1, 0), (1, 1, 1)}.
Writing y = exp(βΩ) we see that the limit (2.14) means that T → 0 and Ω → 1 such that
T /(1 − Ω) is fixed. In fact, it is useful to define
T
T̃ ≡ , x̃ ≡ xy, x̃ = exp(−1/T̃ ). (2.15)
1−Ω
As we shall see, T̃ can be thought of as a temperature in the decoupled sectorafter taking the
limit (2.14). The R-charge that corresponds to the chemical potential Ω is R = 3i=1 αi Ri . With
R
this, we have y1R1 y2R2 y3 3 = y R .

Case I: R = R1 . The half-BPS sector


We take (α1 , α2 , α3 ) = (1, 0, 0) and hence R = R1 . From the letter partition function (2.3) we
see that in the limit (2.14) we have

z(x, yi ) = xy = x̃ (2.16)
up to corrections of order x. Therefore, we see that the set of possible letters reduces to just the
single letter Z, which is the complex scalar in N = 4 SYM with weight (1, 0, 0). The multi-trace
operators in this sector are of the form
     
Tr Z L1 Tr Z L2 · · · Tr Z Lk . (2.17)
Thus, the limit we are considering corresponds to being in the well-known half-BPS sector of
N = 4 SYM spanned by operators of the form (2.17). All the operators of the form (2.17) are
T. Harmark, M. Orselli / Nuclear Physics B 757 (2006) 117–145 125

chiral primaries of N = 4 SYM and preserve at least half of the supersymmetries. By considering
the partition function (2.6) for any N we see that the partition function of U (N ) N = 4 SYM
on R × S 3 reduces to the one of the half-BPS sector given by (2.17). The limit thus reduces the
N = 4 SYM to the quantum mechanical theory with (2.17) as the states in the Hilbert space.
This was previously discussed in [17].
If we consider the thermodynamics of the half-BPS sector (2.17) for large N it is easy to see
from (2.16) that we never reach the Hagedorn singularity: T̃ can be arbitrarily large.
We note here that the half-BPS sector (2.17) is interesting for various reasons; it contains the
states dual to the vacuum of the maximally supersymmetric pp-wave [13,14], to AdS5 × S 5 [15],
and to giant gravitons in AdS5 × S 5 [16], and a correspondence between states in this sector and
half-BPS backgrounds of type IIB string theory has been found in [12].

Case II: R = R1 + R2 . The SU(2) sector


For this case we take (α1 , α2 , α3 ) = (1, 1, 0) so that R = R1 + R2 and (Ω1 , Ω2 , Ω3 ) =
(Ω, Ω, 0). Taking the limit (2.14) the letter partition function (2.3) now becomes

z(x, yi ) = 2xy = 2x̃ (2.18)


up to corrections of order x. In this case, the set of possible letters reduces to the two complex
scalars Z and X with weights (1, 0, 0) and (0, 1, 0), respectively. This is due to the fact that these
two letters are the only letters for which the conformal dimension is equal to the eigenvalue of
R = R1 + R2 . For all other letters the conformal dimension is greater than the eigenvalue of R.
Thus, the set of multi-operators consist of all operators of the form
 (1)   (2) (2) (2)   (k) (k) (k) 
Tr A(1) (1) (i)
1 A2 · · · AL1 Tr A1 A2 · · · AL2 · · · Tr A1 A2 · · · ALk , Aj = Z, X. (2.19)

From (2.6) we see that the partition function for free U (N ) N = 4 SYM on R × S 3 in the
limit (2.14) is
 ∞

 2x̃ k    k 
Z(x, yi ) = [dU ] exp k
Tr U Tr U †
. (2.20)
k
k=1

As for the half-BPS sector we see that N = 4 SYM in the limit (2.14) is reduced to a quantum
mechanical theory, with the multi-trace operators (2.19) as the Hilbert-space. It is not hard to see
that precisely the fact that x → 0 means that the more covariant derivatives an operator has the
more decoupled it becomes. Thus, we remove all the modes coming from having a field theory
on a space, i.e., in this case the Kaluza–Klein modes on S 3 . In this sense we lose the locality of
the field theory and the system becomes instead quantum mechanical.
In Appendix A we take the limit (2.14) in the oscillator representation of N = 4 SYM. This
is an alternative way of showing that we get the SU(2) sector in the limit (2.14).
For N = ∞ it is easy to see from (2.18) that we have a Hagedorn singularity for x̃ = 12 , which
corresponds to
TH (Ω) 1
T̃H = = . (2.21)
1−Ω log 2
We note that this precisely corresponds to the leading part of (2.12). Indeed, viewing the
limit (2.14) as zooming into the region T  1 and 1 − Ω  1 we see that corresponds to the
linear slope of the Hagedorn curve near the critical point (T , Ω) = (0, 1) in the left part of Fig. 2.
126 T. Harmark, M. Orselli / Nuclear Physics B 757 (2006) 117–145

In conclusion, we see that the SU(2) sector captures the leading features of N = 4 SYM on
R × S 3 near the critical point (T , Ω) = (0, 1). We also see that despite the fact that N = 4 SYM
reduces from a field theory to a quantum mechanical theory we keep the interesting physics such
as the Hagedorn transition for large N . Finally we note that using the partition function (2.20)
when T  1 and 1 − Ω  1 instead of the full partition function (2.6) for free N = 4 SYM
on R × S 3 is a very good approximation. Indeed, if Ω = 0.99, the correction on the Hagedorn
temperature is of order 10−15 .

Case III: R = R1 + R2 + R3 . The SU(2|3) sector


This case has (α1 , α2 , α3 ) = (1, 1, 1) and hence R = R1 + R2 + R3 . Taking the limit (2.14)
the letter partition function (2.3) reduces to
3 3
z(x, yi ) = 3xy + 2(xy) 2 = 3x̃ + 2x̃ 2 (2.22)
up to corrections of order x. Thus, the set of possible letters reduces to the three complex
scalars Z, X and W with weights (1, 0, 0), (0, 1, 0) and (0, 0, 1), respectively, and two com-
plex fermions χ1 and χ2 both of weight ( 12 , 12 , 12 ).1 This is precisely the SU(2|3) sector of N = 4
SYM as defined in [8,11]. In Appendix A we have shown this using the oscillator representation
of N = 4 SYM. In this way we show directly that we obtain the SU(2|3) sector as it is defined
in [8] in terms of the oscillator representation of N = 4 SYM.
The Hilbert space of the SU(2|3) sector consists of the multi-trace operators
 (1)   (2) (2) (2)   (k) (k) (k) 
Tr A(1) (1)
1 A2 · · · AL1 Tr A1 A2 · · · AL2 · · · Tr A1 A2 · · · ALk ,
(i)
Aj = Z, X, W, χ1 , χ2 . (2.23)

From (2.6) we see that the partition function for free U (N ) N = 4 SYM on R × S 3 in the
limit (2.14) is
 ∞

 3x̃ k + 2(−1)k+1 x̃ 32 k    k 
Z(x, yi ) = [dU ] exp k
Tr U Tr U †
. (2.24)
k
k=1

For N = ∞ we see from (2.22) that the Hagedorn singularity occurs at


TH (Ω) 1
T̃H = = . (2.25)
1−Ω log 4

2.4. Above the Hagedorn temperature

In this section we consider the behavior of free N = 4 SYM on R × S 3 above the Hagedorn
temperature, following [4].2 Since the N = ∞ partition function is singular at the Hagedorn tem-
perature we should instead use the exact partition function (2.6) which takes non-trivial relations
between multi-trace operators into account. Now, the eigenvalues of the U (N ) group element U

1 Note here that we started with 16 real fermionic components. Picking out a particular weight then leaves us with

two real fermionic components, corresponding to the 2x̃ 3/2 term in the partition function. This can also be seen as two
complex fermions χ1 and χ2 in the sense that their complex conjugates are not present in this sector, just as the complex
conjugates of the three complex scalars Z, X and W are not present in this sector.
2 See also [21].
T. Harmark, M. Orselli / Nuclear Physics B 757 (2006) 117–145 127

are elements eiθ on the unit circle. For large N these eigenvalues become a continuous distri-
bution
π
and we write ρ(θ) for the density of eigenvalues at the angle θ normalized such that
−π dθ ρ(θ ) = 1. Using this, we find from (2.6) the effective action for the eigenvalues [4]

 1 − z(ωn+1 x n , yin )
I = N2 |ρn |2 an , an (x, yi ) = (2.26)
n
n=1
π
with ρn = −π dθ cos(nθ )ρ(θ ). To find the correct eigenvalue distribution we should minimize I .
For temperatures below the Hagedorn temperature we have that an > 0 and hence the minimum
distribution of eigenvalues is the uniform distribution. This is easily seen to give the N = ∞
partition function (2.5) [4].
When we reach the Hagedorn temperature we have that a1 = 0, and this means that the min-
imum of I appears for a non-uniform distribution of the eigenvalues when we are above the
Hagedorn temperature. Using the same procedure as in [4] we determine the behavior of the free
energy near the transition as a perturbative expansion in T ≡ T − TH (Ωi ) when we are slightly
above the Hagedorn temperature. Following [4], the expression for the partition function can be
written as
 ∞

log Z 2 3 4 3 1  n(n2 − 1)z(x n , yin )  5
− 2 =− − − − + O  , (2.27)
N 4 3 8 4 1 − z(x n , yi )
n
n=2

where  = cos (θ0 /2), the angle θ0 is defined by sin2 (θ0 /2) = 1− 1 − 1/z(x, yi ) and z(x, yi ) is
2

given in Eq. (2.3). The Gibbs free energy F = F (T , Ωi ) slightly above the Hagedorn temperature
is then given by
   3/2
F 1 ∂z(x, yi )  1 ∂z(x, yi )   
2
= −  T H T −  TH T 3/2 + O T 2 (2.28)
N 4 ∂T T =TH 3 ∂T T =TH
with T ≡ T − TH (Ωi )  0. Using (2.28) with (2.10) we get the explicit expansion
 √ 
(2 3 + β0 )  2
3
F 3  4
= −β0 1 − β0 Ωi + O Ωi T
N2 8 36
i=1
  √ 
(4 + 3 β0 )  2
3
2 3
 4  
− β0 1 − β0 √ Ωi + O Ωi T 3/2 + O T 2 (2.29)
8 24 3 i=1
for 0  T  1. When the chemical potentials are set to zero we recover the result of [4].
Note from the above that while F /N 2 in the large N limit is finite for temperatures above
the Hagedorn temperature, it is zero for temperatures below the Hagedorn temperature. Thus, we
can regard F /N 2 as an order parameter for the Hagedorn phase transition. Since the derivative
of the free energy is discontinuous at the Hagedorn temperature we see that free U (N ) N = 4
SYM on R × S 3 has a first order phase transition at the Hagedorn temperature [4].
We now turn to the behavior of the free energy slightly above the Hagedorn tempera-
ture in the case of near-critical chemical potential. We examine the two cases corresponding
to (Ω1 , Ω2 , Ω3 ) = (Ω, Ω, 0) and (Ω1 , Ω2 , Ω3 ) = (Ω, Ω, Ω) sending Ω → 1 by taking the
limit (2.14) described in Section 2.3. In this limit we get a rescaled temperature T̃ = T /(1−Ω) as
defined in (2.15). From this we see that we naturally get a rescaled free energy F̃ = −T̃ log Z =
F /(1 − Ω) where F is the Gibbs free energy. In the limit Ω → 1 with T̃ fixed, we get that
F̃ = F̃ (T̃ ), i.e. the rescaled free energy depends only on T̃ .
128 T. Harmark, M. Orselli / Nuclear Physics B 757 (2006) 117–145

Considering the case (Ω1 , Ω2 , Ω3 ) = (Ω, Ω, 0) we have from Section 2.3 that free N = 4
SYM decouples to the SU(2) sector (2.19) in the limit Ω → 1 with T̃ fixed. From (2.21) we
have that the Hagedorn temperature is T̃H = 1/ log 2. Using (2.28), it is straightforward to show
that the free energy slightly above the Hagedorn temperature is
F̃ log 2 (log 2)2  
2
= − (T̃ − T̃ H ) − (T̃ − T̃H )3/2 + O (T̃ − T̃H )2 (2.30)
N 4 3
for 0  T̃ − T̃H  1. One can either derive this using the full letter partition function (2.3) and
then take the limit Ω → 1 with T̃ fixed, or alternatively derive it directly using the letter partition
function (2.18) for the SU(2) sector.
Similarly we can proceed in the case (Ω1 , Ω2 , Ω3 ) = (Ω, Ω, Ω) where we have from Sec-
tion 2.3 that free N = 4 SYM decouples to the SU(2|3) sector (2.23) in the limit Ω → 1 with T̃
fixed. We know from Eq. (2.25) that the Hagedorn temperature is T̃H = 1/ log 4 and using (2.28)
we have that the free energy slightly above the Hagedorn temperature is
F̃ 9 log 2 9(log 2)2  
2
=− (T̃ − T̃H ) − √ (T̃ − T̃H )3/2 + O (T̃ − T̃H )2 (2.31)
N 16 4 2
for 0  T̃ − T̃H  1. Again, as in the SU(2) sector, this result can be found in two different ways
corresponding to either starting from the letter partition function (2.3) and then take the limit on
the final result, or starting with the SU(2|3) letter partition function (2.22).

High temperatures
If we consider instead the high temperature regime the eigenvalue distribution
 becomes almost
like a delta-function [4]. Therefore, ρn = 1 and we get that I = N 2 ∞ n=1 an . If we consider a
high-temperature limit with the chemical potentials being fixed, we get the Gibbs free energy
π 2  3 2 4  
F =− V S N T +O T3 . (2.32)
6
This is precisely the free energy of free N = 4 SYM, i.e. it is the result that one would get
from N 2 times the free energy of U (1) N = 4 SYM. Thus while free N = 4 SYM on R × S 3
behaves as a confined theory for low temperature, it behaves as a deconfined theory at high
temperatures [2,4].
If we instead consider the case in which Ωi /T does not go to zero for T → ∞ for at least one
of the chemical potentials, we get the free energy
 3 

1 2 2  
3
2 π
2 1  3
F = −VS 3 N T + T
4
Ωi − Ω 4
i − 2 Ω i
2 2
Ω j + O T .
6 4 32π 2
i=1 i=1 i<j
(2.33)
This is the same result as in [22,23] where the free energy is computed as N 2 times the free
energy of free U (1) N = 4 SYM. Note that the regularization procedure for obtaining (2.33) is
the same as in [22,23].

3. Quantum mechanical sectors for near-critical chemical potential

In Section 2.3 we saw for free N = 4 SYM on R × S 3 that regions with small temperature
and near-critical chemical potential are very interesting since the free N = 4 SYM effectively
T. Harmark, M. Orselli / Nuclear Physics B 757 (2006) 117–145 129

reduces to free quantum mechanical systems in such regions. In this section we continue to
examine these quantum mechanical sectors of thermal N = 4 SYM on R × S 3 but now in the full
interacting theory. We show in the following that the interacting N = 4 SYM on R × S 3 reduces
to well-defined interacting quantum mechanical systems in such regions of small temperature
and near-critical chemical potential.
Consider the partition function
 
Z(β, Ω) = TrM e−βD+βΩR . (3.1)
Here D is the dilatation operator of N = 4 on R × S 3 which for weak coupling λ  1 can be
expanded as [7,24]


D = D0 + λn/2 Dn , (3.2)
n=2

where we define for convenience the ’t Hooft coupling as


2 N
gYM
λ= . (3.3)
4π 2
Furthermore, R is a linear combination of the three R-charges R1 , R2 and R3 , with Ω as the
corresponding chemical potential. We restrict in the following to the three cases R = R1 , R =
R1 + R2 and R = R1 + R2 + R3 . Clearly we have that D0  R for the three choices of R.
We can rewrite the partition function (3.1) as follows
 ∞



Z(β, Ω) = TrM exp −β(D0 − R) − β(1 − Ω)R − βλD2 − β n/2
λ Dn . (3.4)
n=3

Consider the region

T  1, 1 − Ω  1, λ  1. (3.5)
We now argue that one can neglect all states with D0 − R > 0 in the partition function (3.4). First
we observe that since β
1 and D0 − R is a non-negative integer the states with D0 − R > 0
would have an exceedingly small weight factor. However, one should also ensure then that the
D0 = R states does not have an equally small weight factor. This is precisely ensured by having
1 − Ω and λ  1. We can therefore write the partition function (3.1) in the region (3.5) as
 ∞



Z(β, Ω) = TrH exp −β(1 − Ω)D0 − βλD2 − β n/2
λ Dn (3.6)
n=3

with
  
H = α ∈ M  (D0 − R)α = 0 , (3.7)
i.e., we have restricted the trace to be only over states with D0 = R. Comparing this to Sec-
tion 2.3, we see that restricting to states in H corresponds to the reduction of N = 4 SYM on
R × S 3 found in the free theory. Defining
λ
β̃ = β(1 − Ω), λ̃ = (3.8)
1−Ω
130 T. Harmark, M. Orselli / Nuclear Physics B 757 (2006) 117–145

we can write (3.6) as


 
Z(β̃) = TrH e−β̃H (3.9)
with H being the Hamiltonian
√ ∞
H = D0 + λ̃D2 + λ̃ λ λn/2 Dn+3 . (3.10)
n=0
Considering the three cases R = R1 , R = R1 + R2 and R = R1 + R2 + R3 we have from Sec-
tion 2.3 that H in those three cases corresponds to the half-BPS-sector given by (2.17), the SU(2)
sector given by (2.19) and the SU(2|3) sector given by (2.23). We have thus shown that inter-
acting N = 4 SYM on R × S 3 reduces to those sectors in the region (3.5) with the Hamiltonian
given by (3.10).
Note that we have not assumed anything about N , thus the above considerations work equally
well for finite N and in the large N limit. If we assume N = ∞, we can ignore the non-trivial
relations between multi-trace operators and work instead with single-trace operators. We can
then think of the Hamiltonian (3.10) as the Hamiltonian of a periodic one-dimensional spin-
chain. Below we consider the three possible cases and identify the spin-chain models.

Case I: R = R1 . The half-BPS sector


For R = R1 the interacting thermal N = 4 SYM on R × S 3 is reduced to the Hamiltonian
(3.10) acting on the multi-trace operators of the form (2.17). Since these operators are chiral
primaries of N = 4 SYM all the interaction terms are zero on these states, and hence the Hamil-
tonian (3.10) is H = D0 for this sector.

Case II: R = R1 + R2 . The SU(2) sector


With R = R1 + R2 the interacting thermal N = 4 SYM on R × S 3 is reduced to a quantum
mechanical theory with Hamiltonian (3.10) acting on the SU(2) sector of N = 4 SYM on R × S 3
which is spanned by operators of the form (2.19). Note that in the SU(2) sector the half-integer
powers of λ in (3.10) are not present and we have instead a Hamiltonian of the form [7]


H = D0 + λ̃D2 + λ̃λ λn D2n+4 . (3.11)
n=0
For N = ∞ we can restrict ourselves to consider the single-trace operators, since they are
a well-defined subset of the operators. In the SU(2) sector the single-trace operators are of the
form
Tr(A1 A2 · · · AL ), Ai ∈ {X, Z}. (3.12)
Such single-trace operators can be regarded as spin-chains. In particular a single-trace of length L
corresponds to a periodic spin chain of length L. For a chain of length L the leading interaction
term D2 in the Hamiltonian (3.11) is given by [6,7]

1
L
D2 = (Ii,i+1 − Pi,i+1 ). (3.13)
2
i=1
Here Pi,i+1 is the permutation operator and Ii,i+1 is the identity oprator acting on the letters at
positions i and i + 1. This term of the Hamiltonian (3.11) corresponds precisely to the Hamil-
tonian of the ferromagnetic XXX1/2 Heisenberg spin chain reviewed in Appendix B, where we
T. Harmark, M. Orselli / Nuclear Physics B 757 (2006) 117–145 131

think of the letters Z and X as spin up and spin down.3 Some of the higher terms in (3.11) are
known as well [7,10], but as will be clear in the following they will not play a role for our con-
siderations since they are much weaker coupled than the D2 term. Finally we note that there is
considerable evidence that the Hamiltonian (3.11) is integrable [6,7,9,10].
For N = ∞ we can thus conclude that the thermodynamics of N = 4 SYM on R × S 3 in
the region (3.5) with R = R1 + R2 can be understood from the thermodynamics of the XXX1/2
Heisenberg spin chain.

Case III: R = R1 + R2 + R3 . The SU(2|3) sector


For the case R = R1 + R2 + R3 the interacting thermal N = 4 SYM on R × S 3 in the
region (3.5) reduces to a quantum mechanical theory with Hamiltonian (3.10) acting on the
SU(2|3) sector of N = 4 SYM spanned by operators of the form (2.23).
When N = ∞ we can again restrict to the single-trace operators which in this sectors are of
the form

Tr(A1 A2 · · · AL ), Ai ∈ {X, Z, W, χ1 , χ2 }. (3.14)


Then a single-trace operator of length L can be regarded as a periodic spin-chain of length L.
The leading interaction term D2 in the Hamiltonian (3.10) can then be written as [8,11]

1
L
D2 = (Ii,i+1 − Πi,i+1 ), (3.15)
2
i=1
where Πi,i+1 is the graded permutation operator which permutes the fields at sites i and i + 1
picking up a minus sign if the exchange involves two fermions.
In conclusion we have found that for N = ∞ the thermodynamics of N = 4 SYM on R × S 3
in the region (3.5) with R = R1 + R2 + R3 can be understood from the thermodynamics of the
SU(2|3) spin chain with Hamiltonian (3.15).4

4. Low temperature limit and the Heisenberg spin chain

In this section we consider what happens as we approach the critical point (T , Ω) = (0, 1) in
the specific case of the SU(2) model, i.e. the case with R = R1 + R2 .
We saw in Section 3 that the thermal partition function of N = 4 SYM on R × S 3 in the
region (3.5) with R = R1 + R2 reduces to the partition function (3.9) with the Hamiltonian (3.11).
For N = ∞ we have that a single-trace of fixed length L corresponds to periodic spin-chain of
length L and the Hamiltonian (3.11) is a spin-chain Hamiltonian, with the leading interaction
term D2 corresponding to an XXX1/2 Heisenberg spin chain Hamiltonian.
Consider now being in the region (3.5). Take then the zero temperature limit T → 0 keeping
λ and T̃ = T /(1 − Ω) fixed. In the (T , Ω) diagram depicted in the left part of Fig. 2 this cor-
responds to moving towards the critical point (T , Ω) = (0, 1) in a straight line with slope 1/T̃ .
In terms of the partition function (3.9) and Hamiltonian (3.11) we see that this corresponds to
fixing the temperature while increasing the λ̃ coupling. Since we have that λ  1 and since λ̃

3 Note that J = −λ̃ in comparing with the Hamiltonian (B.1).


4 Note that for this sector the spin-chain is dynamic since it can change the length through the D term [11]. However,
3
we can ignore this higher-loop effect here since we are mostly concerned with the one-loop interaction which corresponds
to the D2 term (3.15).
132 T. Harmark, M. Orselli / Nuclear Physics B 757 (2006) 117–145

is growing towards infinity, we can ignore the higher terms in (3.11) and instead work with the
Hamiltonian

H = D0 + λ̃D2 (4.1)

with D2 given by (3.13). For a fixed length L of the chain (or for the single-trace operators)
this is precisely the ferromagnetic XXX1/2 Heisenberg spin chain Hamiltonian (plus a constant
term). Therefore, we see that the approach to the critical point (T , Ω) = (0, 1) is governed com-
pletely by the XXX1/2 Heisenberg spin chain. Note that letting λ̃ go to infinity does not spoil
our approximations of Section 3 since we always have that β
β̃.
Since we are keeping β̃ fixed we see from the weight factor e−β̃L−β̃ λ̃D2 that it is reasonable
to consider the limit for a chain of fixed length since the coupling in front of L is constant. The
remaining part of the weight factor is e−β̃ λ̃D2 and thus we see that our limit corresponds to taking
the zero temperature limit of the XXX1/2 Heisenberg spin chain.
As reviewed in Appendix B we have that the states with the lowest energy of the ferromag-
netic XXX1/2 Heisenberg spin chain are the zero eigenvalue states of D2 , which when written
as single-trace operators are of the form
  
Tr sym Z L−M X M , (4.2)

where ‘sym’ means total symmetrization. It is clear that any state which is totally symmetrized
has eigenvalue one under the permutation operator, hence the eigenvalue of D2 is zero on such
states. Since 0  M  L we have L + 1 different vacuum states for a chain of length L. Now,
since our limit corresponds to taking the zero temperature limit of the XXX1/2 Heisenberg spin
chain, and since the zero temperature limit means that the states with lowest energy dominates,
we can conclude that we are driven towards the vacuum states (4.2) as we approach the critical
point (T , Ω) = (0, 1).
That we are driven towards the states (4.2) makes sense also from another point of view,
namely that (4.2) corresponds to chiral primaries of N = 4 SYM, and thus the zero temperature
limit that we are taking is driving us towards a 1/2 BPS sector of N = 4 SYM.
There is also another zero temperature limit which is natural to consider. Start again in the
region (3.5). Let then T → 0 with Ω and λ being fixed. In this limit we have that the rescaled
temperature T̃ decreases, while the couplings λ̃ and λ both are fixed. This means that this limit
corresponds to keeping the Hamiltonian (3.11) fixed while changing the temperature T̃ of the
decoupled theory. Thus, in this limit we are moving towards the ground states of the quantum
mechanical theory given by the Hamiltonian (3.11). For N = ∞ we can consider the single-
trace operators of a fixed length. Then the D0 term in (3.11) can be ignored and to leading order
(neglecting the D4 term and higher terms) we have a zero temperature limit of the ferromag-
netic XXX1/2 Heisenberg spin chain. As for the previous limit considered above, this means we
are driven towards the ferromagnetic vacuum states (4.2), which are chiral primaries of N = 4
SYM.
We considered in the above two zero temperature limits of the SU(2) sector. It is not hard
to see that we get similar results for the corresponding zero temperature limits in the SU(2|3)
sector. In particular, we are driven towards the vacuum states of the SU(2|3) spin chain given
by (3.15) which are the states that have zero eigenvalue for D2 . Moreover, these states are chiral
primaries of N = 4 SYM.
T. Harmark, M. Orselli / Nuclear Physics B 757 (2006) 117–145 133

5. Decoupling limit to exact quantum mechanical Hamiltonian

We show in the following that we can take decoupling limits of the thermal interacting N = 4
SYM on R × S 3 to a quantum mechanical system which is described exactly by the one-loop
corrected Hamiltonian in that sector. For N = ∞ the Hamiltonians are the ones of the well-
known spin-chain models.
Consider the partition function (3.1) with the full dilatation operator (3.2). We consider here
again the cases R = R1 , R = R1 + R2 and R = R1 + R2 + R3 . Consider then the following
decoupling limit
T λ
T → 0, T̃ ≡ fixed, λ̃ ≡ fixed. (5.1)
1−Ω 1−Ω
Clearly Ω → 1 and λ → 0 in this limit. From the partition function (3.1) it is clear that we can
ignore states with D0 > R, and hence we only have states with D0 = R. Applying the arguments
of Section 3 we get that the limit (5.1) of the full partition function (3.1) reduces to the limit (5.1)
of the reduced partition function (3.6). Since λ → 0 we see that all the higher-loop terms drop
out, and only the D0 and D2 terms remain. The limit (5.1) of the partition function (3.1) therefore
gives the result
 
Z(β̃) = TrH e−β̃H , (5.2)
where H is the Hamiltonian
H = D0 + λ̃D2 . (5.3)
Thus, thermal interacting N = 4 SYM on R × S3in the limit (5.1) is described exactly by the
Hamiltonian (5.3). Note here that this is true for any N . Furthermore, it is interesting to note that
λ̃ can take any value. One can thus end up with a strongly coupled D2 term in the Hamiltonian
as a good description of N = 4 SYM, as we in fact already saw in Section 4.
For N = ∞, we get as above that we can think of the single-trace operators as spin-chains.
We thus have that the thermodynamics of interacting N = 4 SYM on R × S 3 in the decoupling
limit (5.1) can be described exactly by a spin-chain model with Hamiltonian (5.3).
If we consider the case R = R1 + R2 for N = ∞ we see that the thermodynamics of
N = 4 SYM on R × S 3 in the decoupling limit (5.1) can be described exactly by the fer-
romagnetic XXX1/2 Heisenberg spin chain (see Appendix B). This is easily seen from the
Hamiltonian (5.3) with D2 given in (3.13). Written explicitly, we have that the full partition
function for N = 4 SYM in the limit (5.1) is
∞ 
 ∞
1 −nL/T̃ (XXX)
log Z(T̃ ) = e ZL (T̃ /n), (5.4)
n
n=1 L=1
(XXX)
where ZL is the partition function for the ferromagnetic XXX1/2 Heisenberg spin chain of
length L with Hamiltonian λ̃D2 .
Similarly, for the case R = R1 + R2 + R3 we have that the thermodynamics of N = 4 SYM
on R × S 3 in the decoupling limit (5.1) can be described exactly by the spin chain model given
by the D2 term (3.15).
In conclusion we have found limits in which planar thermal N = 4 SYM is described ex-
actly by well-defined spin-chain models. The spin-chain models involved are short-range and the
coupling λ̃ in front of the spin chain term D2 in the Hamiltonian can take any value.
134 T. Harmark, M. Orselli / Nuclear Physics B 757 (2006) 117–145

6. One-loop partition function

In this section we consider the one-loop correction to the partition function for U (N ) N = 4
SYM on R × S 3 with non-zero chemical potentials in planar limit N = ∞, generalizing the
procedure in [18]. We use this to find the one-loop correction to the Hagedorn temperature. We
consider subsequently the one-loop correction to the partition function and Hagedorn tempera-
ture for the near-critical regions where N = 4 SYM reduces to the SU(2) and SU(2|3) sectors.

6.1. One-loop correction to partition function and Hagedorn temperature


 R
Consider the complete single-trace partition function ZST = Tr(x D 3i=1 yi i ) for U (N )
N = 4 SYM on R × S 3 in the planar limit. Up to the first order in the ’t Hooft coupling λ
(0) (1) (0)
the single-trace partition function can be written as ZST = ZST + λZST + O(λ2 ) where ZST is
the zeroth order single-trace partition function given in (2.4) and with the first-order contribution
given by
3

(1)
 R
ZST (x, yi ) = log x Tr i D0
yi x D2 . (6.1)
i=1
This follows from the expansion (3.2) of the dilatation operator and from the fact that the R-
charges commute with the dilatation operator. Applying the arguments of [18] where it is used
that one can refrase (6.1) as a spin-chain partition function, we arrive at the following expression
for the one-loop single trace partition function

(1)
∞   D2 (ωL+1 x L , y L )
L−1
ZST (x, yi ) = log x i

L=1 k=0
1 − z(ω L+1 x L , y L )
i
(k,L)=1

  (L−k) 
+ δL =1 P D2 ωL−k+1 x L−k , yi , ωk+1 x k , yik (6.2)

with
   
3
Ri (A1 )+Ri (A2 )
D2 (x, yi ) = x d(A1 )+d(A2 ) yi A1 A2 |D2 |A1 A2 , (6.3)
A1 ,A2 ∈A i=1

   
3
Ri (A1 ) Ri (A2 )
P D2 (w, yi , w̄, ȳi ) = w d(A1 ) yi w̄ d(A2 ) ȳi A1 A2 |D2 |A2 A1 . (6.4)
A1 ,A2 ∈A i=1

Here L can be seen as the length of the spin chain and (k, L) = 1 means that k and L are relatively
prime. We have also included the fermion contribution. We note that Eq. (6.2) is a direct general-
ization of the result of [18]. From (6.2) it is in principle straightforward to compute the one-loop
correction (6.1), once the two expectation values (6.3) and (6.4) are known. From this one gets
the corrected multi-trace partition function using the general prescription in (2.5). In Appendix C
we computed D2 and we sketched how to compute P D2 . We have not computed the corrected
partition function here explicitly since we do not need it for the purposes of this paper. However,
below we compute it explicitly in the near-critical regions giving the SU(2) and SU(2|3) sectors.
We use now the result (6.2) to compute the one-loop correction to the Hagedorn tempera-
ture. From Eq. (2.9) we have that the zeroth order contribution to the partition function goes
T. Harmark, M. Orselli / Nuclear Physics B 757 (2006) 117–145 135

like (TH − T )−1 near the Hagedorn temperature, for fixed chemical potentials. This behavior
resists also for the corrected partition function where now the value of the Hagedorn temper-
ature is shifted by the higher loop corrections. One can then compute the one-loop corrected
(1)
Hagedorn temperature by considering the pole of ZST in (6.2) at the zeroth order Hagedorn tem-
perature TH(0) . As in the case of zero chemical potentials [18] the term proportional to P D2 does
not give rise to divergences. Hence, we get the following formula for the one-loop correction to
the Hagedorn temperature

D2 
δTH = λ ∂z  (6.5)
T ∂T T =TH(0)
for given chemical potentials Ωi .
Using Eq. (6.5), we compute now the one-loop corrected Hagedorn temperature for small
values of the chemical potentials Ωi . To this end, we use the results on D2 of Appendix C to
find the following expression for D2 evaluated at the Hagedorn temperature for small chemical
potentials

β2  2 √ √  2 2
3
3 β3 
D2 = 1− 0 Ωi − 0 72 − 56 3 + 3β0 (41 − 26 3 ) Ωi Ωj
4 18 864
i=1 i<j

β03  √ √  3
 
− 72 − 56 3 + 3β0 (45 − 26 3 ) Ωi4 + O Ωi6 . (6.6)
1296
i=1

To compute this we used the zeroth order Hagedorn temperature for small chemical potentials
given in Eq. (2.10). Inserting Eq. (6.6) in Eq. (6.5), we find that the one-loop corrected Hagedorn
temperature for small chemical potentials is


3  
3
 
TH (Ωi ) = p0 + p1 Ωi2 + p2 Ωi2 Ωj2 + p3 Ωi4 + O Ωi6 ,
i=1 i<j i=1
 
1 λ  
p0 = 1+ + O λ2 ,
β0 2
 
1 λ √  
p1 = − √ 1 − (11 − β0 3 ) + O λ2 ,
6 3 2
β0  √   √ √   
p2 = 18 − 5 3 + λ 60 − β0 72 − 35 3 − β0 (69 3 − 113) + O λ2 ,
1296
β0  √   √ √   
p3 = 18 − 11 3 + λ 60 − β0 72 − 47 3 − β0 (69 3 − 122) + O λ2 . (6.7)
2592
Note that for zero chemical potentials in Eqs. (6.6) and (6.7) we recover the result of [18].

6.2. The SU(2) sector

We consider now the near-critical region (3.5) with (Ω1 , Ω2 , Ω3 ) = (Ω, Ω, 0). From Sec-
tion 3 we know that the single-trace sector of the planar limit of U (N ) N = 4 SYM on R × S 3
reduces to the SU(2) sector with single-traces of the form (3.12). From Section 3 we have fur-
thermore that we can consider T̃ = T /(1 − Ω) as the effective temperature and that the one-loop
136 T. Harmark, M. Orselli / Nuclear Physics B 757 (2006) 117–145

corrected Hamiltonian becomes H = D0 + λ̃D2 with λ̃ = λ/(1 − Ω). In the following we em-
ploy these results to find the corrected partition function and Hagedorn temperature for this
near-critical region. Note that we assume in the following that λ̃  1.
From Section 2 we have that the zeroth order contribution to the partition function for the
SU(2) sector is


(0) ϕ(k)  
ZST (x̃) = − log 1 − 2x̃ k . (6.8)
k
k=1

We now consider the first correction in λ̃ to this partition function when λ̃  1.5 To this end, we
use the formula [18]

(1)

   D2 (ωL+1 x̃ L )
L−1
  L−k+1 L−k k+1 k 

ZST (x̃) = log x̃ + δ L =1 P D 2 ω x̃ , ω x̃ .
1 − z(ωL+1 x̃ L )
L=1 k=0
(k,L)=1
(6.9)
In the SU(2) sector the expectation values of D2 and P D2 are given by [18]
   
D2 (x̃) = x̃ 2 , P D2 (x̃1 , x̃2 ) = −x̃1 x̃2 . (6.10)
Substituting now those expressions into the formula (6.9), we recover the known result for the
one-loop partition function in the SU(2) sector [18]



 n
(1) n 1 − 3x̃
ZST (x̃) = log x̃ x̃ − ϕ(n)x̃ . (6.11)
1 − 2x̃ n
n=1
Similarly to Eq. (6.5), we have that the correction to the Hagedorn temperature is

D2 
δ T̃H = λ̃ ∂z(x̃)  , (6.12)
(0)
T̃ T̃ =T̃H
∂ T̃
(0)
where T̃H = 1/ log 2. We used here that P D2 is not divergent, as one can see from (6.10).
From Eqs. (6.10) and (6.12) we get then that the corrected Hagedorn temperature for λ̃  1 is
 
1 1  
T̃H = 1 + λ̃ + O λ̃2 . (6.13)
log 2 4
It is important to notice that starting instead from the general expressions for D2 (x, yi ) and
P D2 (x, yi ) for N = 4 SYM given in Appendix C and taking the limit (5.1) precisely gives the
result (6.10).6 From this fact one can in turn see that both the one-loop corrected partition func-
tion and Hagedorn temperature reduces to (6.11) and (6.13) found above. This is in accordance
with our derivation of the interacting Hamiltonian in Section 3.
Finally we note that the two loop corrected Hagedorn temperature in the SU(2) sector has
been considered in [25]. However, their result is not directly applicable in our case, since the two
Hamiltonians for the corrections are different.

5 The one-loop partition function for the SU(2) sector is computed previously in [18], but we review it here for com-

pleteness, and since we use the same technique below to compute the first correction for λ̃  1 for the SU(2|3) sector.
6 For D we have from (C.8) and (C.4)–(C.6) in Appendix C that V does not contribute and V
2 0 j 2 → 0 in the
(0,0)
limit (5.1), while V1 = x 2 y 2 since only the x 2 F[1,0,1] term contributes. For P D2 one can take the limit on Eq. (C.9)
and see that it reduces to the correct answer.
T. Harmark, M. Orselli / Nuclear Physics B 757 (2006) 117–145 137

6.3. The SU(2|3) sector

In the SU(2|3) sector the story is very similar to the one for the SU(2) sector. We are consider-
ing the near-critical region (3.5) with (Ω1 , Ω2 , Ω3 ) = (Ω, Ω, Ω). From Section 3 we know that
the single-trace sector of the planar limit of U (N ) N = 4 SYM on R × S 3 reduces to the SU(2|3)
sector with single-traces of the form (3.14). The zeroth order single-trace partition function is


(0) ϕ(k)  
ZST (β̃) = − log 1 − 3x̃ k − 2(−1)k+1 x̃ 3k/2 . (6.14)
k
k=1

We compute the first correction in λ̃ to this partition function when λ̃  1 using again Eq. (6.9).
Using that the dilatation operator is given by (3.15) we find
 
D2 (x̃) = 3x̃ 2 + 6x̃ 5/2 + 3x̃ 3 ,
  3/2 3/2  3/2 3/2 
P D2 (x̃1 , x̃2 ) = −3x̃1 x̃2 + 3x̃1 x̃2 − 3 x̃1 x̃2 + x̃1 x̃2 . (6.15)
As for the SU(2) sector, these results can be recovered using the expressions for D2 (x, yi )
and P D2 (x, yi ) for N = 4 SYM given in Appendix C and taking the limit (5.1). Inserting the
previous expressions in Eq. (6.9) we get that the one-loop partition function in the SU(2|3) sector
is given by

(1)
  
L−1

ZST (x̃) = log x̃ 3x̃ + 3x̃ 3/2 − 3 (−1)L−k+1 x̃ (3L−k)/2 + (−1)k+1 x̃ L+k/2
L=2 k=0
(k,L)=1


n 1 − (−1) x̃ − 4x̃ n + 7(−1)n x̃ 3n/2 − 3x̃ 2n
n n/2
−3 ϕ(n)x̃ . (6.16)
1 − 3x̃ n − 2(−1)n+1 x̃ 3n/2
n=1
Using now Eqs. (6.12) and (6.15) we get for the one-loop corrected Hagedorn temperature the
following result
 
1 3  2
T̃H = 1 + λ̃ + O λ̃ . (6.17)
log 4 8
One can check that only the D2 (x̃) part of the one-loop partition function contributes to this.

7. Discussion and conclusions

In this paper we have found that thermal N = 4 SYM on R × S 3 greatly reduces near the crit-
ical points (T , Ω1 , Ω2 , Ω3 ) = (0, 1, 0, 0), (T , Ω1 , Ω2 , Ω3 ) = (0, 1, 1, 0) and (T , Ω1 , Ω2 , Ω3 ) =
(0, 1, 1, 1). We identified the three quantum mechanical theories that N = 4 SYM reduces to,
and in particular we showed that the Hilbert spaces correspond to a half-BPS sector and the
SU(2) and SU(2|3) sectors of N = 4 SYM. We found the Hamiltonian for these three theories
and we saw that the one-loop correction to the dilatation operator has a special significance in
this. The existence of these quantum mechanical sectors of N = 4 SYM could prove highly use-
ful. Through the AdS/CFT correspondence the thermodynamics of N = 4 SYM is linked to the
Hagedorn transition in string theory, and since for instance the SU(2) sector is greatly reduced in
complexity compared to the full N = 4 SYM, we can get a much better handle on the behavior
of N = 4 SYM in this particular near-critical region than on N = 4 SYM with zero chemical
potentials.
138 T. Harmark, M. Orselli / Nuclear Physics B 757 (2006) 117–145

For N = ∞ we found that the near-critical regions giving the SU(2) and SU(2|3) sectors
can be described in terms of spin chain theories. In particular the SU(2) sector corresponds
to a ferromagnetic XXX1/2 Heisenberg spin chain to leading order (or exactly, if we take the
limit of Section 5). This provides a very different realization of spin chains for the planar limit
of N = 4 SYM on R × S 3 than in the study of integrability [6,7,9]. In terms of integrability,
the SU(2) and SU(2|3) sectors are closed subsectors of the conjectured complete N = 4 spin
chain, i.e. they decouple to all orders in perturbation theory [11]. However, it is not clear that
this decoupling holds at strong coupling [26,27]. Instead, in the limit of this paper we have an
effective reduction of N = 4 SYM to the SU(2) and SU(2|3) sectors which does not rely on
the SU(2) and SU(2|3) sectors being closed in the sense of having interactions with the other
operators of N = 4 SYM. Any such interaction would in any case be suppressed in the near-
critical regions that we consider. It would therefore be interesting to consider if our decoupling
of the SU(2) and SU(2|3) sectors corresponds to a similar decoupling for thermal string theory
on AdS5 × S 5 with near-critical chemical potentials.
It is intriguing to compare our limit to the pp-wave limits of AdS5 × S 5 [14]. It is not hard to
see that the near-critical region giving us the reduction to the SU(2) sector has some similarities
with the pp-wave limit of [28] since we keep only states with D0 = R1 + R2 . It is clear that
to connect to the limit of N = 4 SYM found in [28] we need to consider only a subsector of
the pp-wave string theory of [28]. This seems possible to achieve by turning on the appropriate
chemical potential. This would be interesting to study since we have a Hagedorn transition both
in the gauge theory side and on the pp-wave side [29].
Another interesting direction to pursue would be to compare our results on the Hagedorn
temperature as a function of the chemical potential to the Hawking–Page transition [1,5] with
chemical potentials [30]. With the chemical potentials set to zero we have a consistent picture
that the Hagedorn transition is a first order transition both for weak coupling λ  1 [4,31] and
for strong coupling λ
1 [32] where it is mapped to the Hawking–Page transition. It would be
interesting to see whether the picture is equally consistent once the chemical potentials are turned
on.
Finally, we note that we expect similar decoupled quantum mechanical sectors in other super-
symmetric gauge theories with R-symmetry, in regions with near-critical chemical potentials.

Acknowledgements

We thank P. Di Vecchia, G. Grignani, C. Kristjansen and N. Obers for useful discussions and
H. Osborn for useful correspondence. We thank KITP for hospitality while part of this work was
completed. This research was supported in part by the National Science Foundation under Grant
No. PHY99-0794. The work of M.O. is supported in part by the European Community’s Human
Potential Programme under contract MRTN-CT-2004-005104 ‘Constituents, fundamental forces
and symmetries of the universe’.

Appendix A. Oscillator representation of N = 4 SYM

In the oscillator representation of N = 4 SYM [8,33] we can write all the gauge-invariant
operators using two bosonic oscillators aα , bα̇ , α, α̇ = 1, 2, and one fermionic oscillator ca ,
a = 1, 2, 3, 4, with the commutation relations
 α †  α̇ †   a †
a , aβ = δβα , b , bβ̇ = δβ̇α̇ , c , cb = δba . (A.1)
T. Harmark, M. Orselli / Nuclear Physics B 757 (2006) 117–145 139

In terms of these oscillators, the set of letters A of N = 4 SYM is given by


 2
φ: c† |0 repr. [0, 1, 0](0,0) ,
 3
ψ: a† c† |0 repr. [0, 0, 1]( 1 ,0) , ψ̄: b† c† |0 repr. [1, 0, 0](0, 1 ) ,
2 2
 † 2  † 2
F : a |0 repr. [0, 0, 0](1,0) , F̄ : b |0 repr. [0, 0, 0](0,1) ,
D: a† b † repr. [0, 0, 0]( 1 , 1 ) , (A.2)
2 2

where F is the field strength, ψ the fermions and φ the scalars. Moreover D is the covariant
derivative. One can then generate A by acting with D k . Note that we also specified the represen-
tation under SU(4) × SO(4) that the fields are in, for example [0, 0, 1](1/2,0) corresponds to the
[0, 0, 1] of SU(4) and the ( 12 , 0) of SO(4).
Write now the number operators as a α = a†α aα , bα̇ = b†α̇ bα̇ and ca = c†a ca , where it should be
understood that there are no sums over the indices. We define then the operators
1 1  1  1 
C =1− a + a 2 + b1 + b2 − c1 + c2 + c3 + c4 ,
2 2 2
1 
D0 = 1 + a 1 + a 2 + b 1 + b 2 . (A.3)
2
Here C is the central charge which should be annihilated on physical states, while D0 is the
dilatation operator in free N = 4 SYM. The three R-charges are
1 1  1 
R1 = c − c2 − c3 + c4 , R2 = −c1 + c2 − c3 + c4 ,
2 2
1 
R3 = −c1 − c2 + c3 + c4 . (A.4)
2
We can now write the letter partition function as

z(x, y1 , y2 , y3 )
 R 
= TrA x D0 y1R1 y2R2 y3 3

 
1
R
= δ(C)x D0 y1R1 y2R2 y3 3
a1 ,a2 ,b1 ,b2 =0 c1 ,c2 ,c3 ,c4 =0

 
 
1 
4
1 R
= (a + 1)(b + 1) δ 2−a+b− c a
x 1+ 2 (a+b) y1R1 y2R2 y3 3 . (A.5)
a,b=0 c1 ,c2 ,c3 ,c4 =0 a=1

It is straightforward to see that this gives the letter partition function (2.3) computed in Sec-
tion 2.1. Note that we defined a = a 1 + a 2 and b = b1 + b2 in (A.5).
We consider now the decoupling limits of Section 2.3. Consider first the case in which Ω1 =
Ω2 = Ω, Ω3 = 0 and hence R = R1 + R2 . Taking the limit (2.14), i.e., with x̃ ≡ xy fixed and
x → 0, y = exp(βΩ), it is easy to see that only the sector with D0 = R survives. Using the above
formulas we see that since R = R1 + R2 = −c3 + c4 the limit (2.14) corresponds to inserting the
Kronecker delta δ(2 + a + b + 2c3 − 2c4 ) into the sum in (A.5). This Kronecker delta-function
can clearly only be 1 provided a = b = c3 = 0 and c4 = 1, since all the number operators are
positive and the fermionic number operators only take the values 0 and 1. We are thus in the
140 T. Harmark, M. Orselli / Nuclear Physics B 757 (2006) 117–145

sector given by
a 1 = a 2 = b1 = b2 = c3 = 0, c4 = 1 (A.6)
and it is easy to see that the only states in this sector are c†1 c†4 |0
and c†2 c†4 |0 ,
corresponding to
the two complex scalars Z and X. This is clearly the SU(2) sector, as defined in [8], and the
partition function is indeed easily found from (A.5) to reduce to (2.18).
Consider instead the case in which Ω1 = Ω2 = Ω3 = Ω and hence R = R1 + R2 + R3 . Taking
the limit (2.14) we see again that only the sector with D0 = R remains. Using that R = R1 +
R2 + R3 = 12 (−c1 − c2 − c3 + 3c4 ) we see that this limit corresponds to inserting δ(2 + a + b +
c1 + c2 + c3 − 3c4 ) into the sum in (A.5). It is clear that this Kronecker delta only can be non-zero
provided c4 = 1. If we consider the case b = 1 we see that then we need a = c1 = c2 = c3 = 0,
but that is not a physical state. This means that b = 0 and that a + c1 + c2 + c3 = 1, which is
equivalent to stating that b = 0 and C = 0. We are thus in the sector given by
b1 = b2 = 0, c4 = 1. (A.7)
The physical states in this sector are c†1 c†4 |0 , c†2 c†4 |0 and c†3 c†4 |0 , corresponding to the three
complex scalars Z, X and W , and a†1 c†4 |0 and a†2 c†4 |0 corresponding to the two complex fermi-
ons χ1 and χ2 . This is clearly the SU(2|3) sector defined in [8]. Furthermore, it is straightforward
to find that the partition function (A.5) reduces to (2.22).

Appendix B. The XXX1/2 Heisenberg spin chain

For convenience we briefly review here some essential facts of the XXX1/2 Heisenberg spin
chain. We are considering a periodic spin chain of length L, so that the Hilbert space of the spin
chain is spanned by states with M down-spins and L − M up-spins, 0  M  L. Thus, the Hilbert
space has dimension 2L . The Hamiltonian of a one-dimensional XXX1/2 Heisenberg spin chain
is traditionally defined as
L  
1
H =J Si · Si+1 − , (B.1)
4
i=1

where Si acts on the ith spin as σ /2, i.e. with σ being the Pauli matrices. To find the eigen-
values and eigenstates of the Hamiltonian one uses the Bethe ansatz [34] (see for example [35]
for specific examples of spectra for L = 4, 6). In [36] the full spectrum has been found in the
thermodynamic limit L → ∞. Defining the total spin

L
S = Si (B.2)
i=1

we have that [H, S]  = 0. This means that any eigenstate of H is part of a spin multiplet with

respect to S.
If we have J < 0 the Hamiltonian (B.1) is describing a ferromagnet. The ferromagnetic vacua
are the states with eigenvalue zero of H . These are totally symmetrized states with M down-
spins and L − M up-spins, 0  M  L. Clearly there are L + 1 such states and they in fact make
up a L + 1 dimensional representation with respect to S.
If we have instead that J > 0 the Hamiltonian (B.1) is describing an antiferromagnet. The an-
tiferromagnetic vacuum state is a unique state with L/2 up-spins and L/2 down-spins (assuming

L even). It is a singlet with respect to S.
T. Harmark, M. Orselli / Nuclear Physics B 757 (2006) 117–145 141

Appendix C. Computations for one-loop partition function

In this appendix we derive the expression for D2 (x, yi ) used in Section 6 to compute
the one-loop correction to the Hagedorn temperature. We also briefly discuss how to compute
P D2 (x1 , yi(1) , x2 , yi(2) ) .
From the definition (6.3) of D2 (x, yi ) we have that it corresponds to the expectation value
of D2 acting on the product of two copies of the singleton representation A × A. To com-
pute D2 (x, yi ) we can then employ the fact that it commutes with the two-letter Casimir of
PSU(2, 2|4) on A × A [8]. To this end, we use the following modules of PSU(2, 2|4) [37,38]
1 1 1 1 1 1
, , ,
A = B[0,1,0]
2 2
(0,0)
, V0 = B[0,2,0]
2 2
(0,0)
, V1 = B[1,0,1]
4 4
(0,0)
,
1,1
Vj = C[0,0,0] j j
for j  2. (C.1)
( 2 −1, 2 −1)

Here we wrote the modules in the notation of [37]. For each module it is written what super-
conformal primary operator the representation is generated from, e.g., for V1 it is [1, 0, 1](0,0)
which is the primary operator in the [1, 0, 1] representation
 of SU(4) and in the singlet (0, 0) of
SU(2) × SU (2). We have then that A × A = ∞ j =0 Vj and that the eigenvalue of D2 in Vj is
j 1
given by the harmonic number h(j ) = n=1 n [8]. We can therefore compute D2 (x, yi ) by

computing TrVj (x D0 3i=1 yiRi ). This can be done using the tables for the modules (C.1) pre-
sented in [18,37,38]. We define
(j1 ,j2 )  
F[k,p,q] = (2j1 + 1)(2j2 + 1)W[k,p,q] , W[k,p,q] ≡ Tr[k,p,q] yiRi . (C.2)
We see that W[k,p,q] is the weighted sum of the weights of [k, p, q]. For the specific representa-
tions we have

3
  
3
 −1/2 
yi + yi−1 ,
1/2
W[0,0,0] = 1, W[0,1,0] = W([1,0,0]+[0,0,1]) = yi + yi ,
i=1 i=1


3
 −1/2 

3

1/2 −1 
W([1,1,0]+[0,1,1]) = yi + yi yj + y j −1 ,
i=1 j =1
   
W[0,2,0] = yi + yi−1 yj + yj−1 − 4,
1ij 3
   
W[1,0,1] = yi + yi−1 yj + yj−1 + 3,
1i<j 3


3
  3
 
W([2,0,0]+[0,0,2]) = 2 yi + yi−1 + yi + yi−1 . (C.3)
i=1 i=1
3 Ri
From the above we can now compute Vj (x, yi ) ≡ (1 − x)4 TrVj (x D0 i=1 yi ). We get

(0,0) 5  (0, 1 ) ( 12 ,0) 


V0 = x 2 F[0,2,0] + x 2 F[1,1,0]
2
+ F[0,1,1]
 (0,0) (0,0) (1,0) (0,1) ( 12 , 12 ) 
+ x 3 F[2,0,0] + F[0,0,2] + F[0,1,0] + F[0,1,0] + F[1,0,1]
7  ( 1 ,0) (0, 12 ) (1, 12 ) ( 12 ,1)   (0,0) (1,1) (0,0) 
+ x 2 F[0,0,1]
2
+ F[1,0,0] + F[1,0,0] + F[0,0,1] + x 4 2F[0,0,0] + F[0,0,0] − F[1,0,1]
142 T. Harmark, M. Orselli / Nuclear Physics B 757 (2006) 117–145

9  ( 1 ,0) (0, 12 )  ( 12 , 12 )
− x 2 F[1,0,0]
2
+ F[0,0,1] − x 5 F[0,0,0] , (C.4)

(0,0) 5  ( 1 ,0) (0, 12 ) (0, 12 ) ( 12 ,0) 


V1 = x 2 F[1,0,1] + x 2 F[1,0,0] 2
+ F[0,0,1] + F[1,1,0] + F[0,1,1]
 (0,0) (1,0) (0,1) (1,0) (0,1) ( 12 , 12 )
+ x 3 2F[0,1,0] + F[0,1,0] + F[0,1,0] + F[0,0,2] + F[2,0,0] + F[0,0,0]
( 12 , 12 ) ( 12 , 12 ) 
+ F[1,0,1] + F[0,2,0]
7  ( 1 ,0) (0, 12 ) ( 32 ,0) (0, 32 ) (1, 12 ) ( 12 ,1)
+ x 2 F[0,0,1] 2
+ F[1,0,0] + F[0,0,1] + F[1,0,0] + F[1,0,0] + F[0,0,1]
(1, 12 ) ( 12 ,1) 
+ F[0,1,1] + F[1,1,0]
 (1,0) (0,1) ( 12 , 32 ) ( 32 , 12 ) (1,1) (1,1)
+ x 4 F[0,0,0] + F[0,0,0] + F[0,1,0] + F[0,1,0] + F[0,0,0] + F[1,0,1]
(0,0) (0,0) (0,0) 
− F[0,0,0] − F[1,0,1] − F[0,2,0]
9  ( 3 ,1) (1, 32 ) (0, 12 ) ( 12 ,0) (0, 12 ) ( 12 ,0) 
+ x 2 F[1,0,0] 2
+ F[0,0,1] − F[0,0,1] − F[1,0,0] − F[1,1,0] − F[0,1,1]
 ( 32 , 32 ) (0,1) (1,0) ( 12 , 12 ) ( 12 , 12 ) 
+ x 5 F[0,0,0] − F[0,1,0] − F[0,1,0] − F[0,0,0] − F[1,0,1]
11  ( 1 ,1) (1, 12 )  (1,1)
− x 2 F[0,0,1] 2
+ F[1,0,0] − x 6 F[0,0,0] , (C.5)

( j2 −1, j2 −1) 1  ( j , j −1 ) ( j −1 j 
2 ,2)
Vj = x j F[0,0,0] + x j + 2 F[1,0,0] 2 2
+ F[0,0,1]
 ( j2 , j2 −1) ( j2 −1, j2 ) ( j −1 j −1
2 , 2 ) ( j −1 j −1
2 , 2 ) ( j −3 j −3 
2 , 2 )
+ x j +1 F[0,1,0] + F[0,1,0] + F[1,0,1] + F[0,0,0] − F[0,0,0]
3  ( j +1 , j −1) ( j2 −1, j +1 ( j −1 j
( j2 , j −1 ( j2 , j −1
2 ) 2 ,2) 2 ) 2 )
+ x j + 2 F[0,0,1] 2 2
+ F[1,0,0] + F[0,0,1] + F[1,0,0] + F[0,1,1]
( j −1 j
2 ,2) ( j2 −1, j −3
2 ) ( j −3 j
2 , 2 −1)

+ F[1,1,0] − F[1,0,0] − F[0,0,1]
 ( j2 +1, j2 −1) ( j2 −1, j2 +1) ( j +1 j −1
2 , 2 ) ( j −1 j +1
2 , 2 ) ( j2 , j2 ) ( j2 , j2 )
+ x j +2 F[0,0,0] + F[0,0,0] + F[0,1,0] + F[0,1,0] + F[0,0,0] + F[1,0,1]
(j ,j ) ( j +1 , j −1
2 ) ( j −1 , j +1
2 ) ( j −1, j −1) ( j −1, j −1)
+ F[0,2,0]
2 2
+ F[0,0,2]
2
+ F[2,0,0]
2
− F[0,0,0]
2 2
− F[1,0,1]
2 2

j −1 
( j −1 , j −3 ) ( j −3
2 , 2 )
− F[0,1,0]2 2
− F[0,1,0]
5  ( j −1 , j +1) ( j2 +1, j −1 ( j +1 j
( j2 , j +1 ( j +1 j
( j2 , j +1
2 ) 2 ,2) 2 ) 2 ,2) 2 )
+ x j + 2 F[1,0,0] 2 2
+ F[0,0,1] + F[0,1,1] + F[1,1,0] + F[1,0,0] + F[0,0,1]
( j2 , j −3
2 ) ( j −3 j
2 ,2) ( j2 −1, j −1
2 ) ( j −1 j
2 , 2 −1) ( j2 −1, j −1
2 ) ( j −1 j
2 , 2 −1)

− F[0,0,1] − F[1,0,0] − F[1,1,0] − F[0,1,1] − F[0,0,1] − F[1,0,0]
 ( j2 +1, j2 ) ( j2 , j2 +1) ( j +1 j +1
2 , 2 ) ( j +1 j +1
2 , 2 ) ( j +1 j −3
2 , 2 )
+ x j +3 F[0,1,0] + F[0,1,0] + F[1,0,1] + F[0,0,0] − F[0,0,0]
( j −3 , j +1
2 ) ( j , j −1) ( j −1, j ) ( j , j −1) ( j −1, j ) ( j −1 , j −1
2 )
− F[0,0,0]
2
− F[0,1,0]
2 2
− F[0,1,0]
2 2
− F[0,0,2]
2 2
− F[2,0,0]
2 2
− F[0,2,0]
2

j −1 
( j −1 , j −1 ) ( j −1
2 , 2 )
− F[0,0,0]
2 2
− F[1,0,1]
7  ( j +1 , j +1) ( j2 +1, j +1 ( j +1 j
( j2 −1, j +1 ( j2 , j −1
2 ) 2 , 2 −1) 2 ) 2 )
− x j + 2 F[0,0,1]
2 2
+ F[1,0,0] − F[0,0,1] − F[1,0,0] − F[1,0,0]
j 
( j −1 j
2 ,2) ( j2 , j −1
2 ) ( j −1
2 ,2)
− F[0,0,1] − F[0,1,1] − F[1,1,0]
T. Harmark, M. Orselli / Nuclear Physics B 757 (2006) 117–145 143

 ( j2 +1, j2 +1) ( j +1 j −1
2 , 2 ) ( j −1 j +1
2 , 2 ) ( j2 , j2 ) ( j2 , j2 ) 
− x j +4 F[0,0,0] − F[0,1,0] − F[0,1,0] − F[1,0,1] − F[0,0,0]
9  ( j +1 , j ) ( j2 , j +1  ( j +1 j +1
2 ) 2 , 2 )
− x j + 2 F[1,0,0]
2 2
+ F[0,0,1] − x j +5 F[0,0,0] . (C.6)
A nice check of the formulas derived for Vj (x, yi ) is given by the following equality

 Vj (x, yi )
= z(x, yi )2 , (C.7)
(1 − x)4
j =0

where z(x, yi ) is the letter partition function (2.3). From the above, we have then


  3
Ri
 Vj (x, yi )
D2 (x, yi ) = TrA×A x D0
yi D2 = h(j ) . (C.8)
(1 − x)4
i=1 j =0

Using Eqs. (C.4), (C.5) and (C.6) for Vj one can then obtain the expression for D2 (x, yi ) .
We do not write the result here, since it is a highly complicated expression. Instead we use in
Section 6 Eq. (C.8) to find D2 (x, yi ) for small chemical potentials and for near-critical chemical
potentials.

C.1. Oscillator representation of P D2

We explain here briefly how to compute P D2 (x1 , yi(1) , x2 , yi(2) ) defined in (6.4). We do not
compute the resulting expression here due to the fact that it does not contribute to the correction
to the Hagedorn temperature.
It is not possible to employ the same technique used above for D2 to compute P D2 , since
P D2 , unlike D2 , does not commute with the two-letter PSU(2, 2|4) Casimir [18]. Instead, we
use the oscillator representation of N = 4 reviewed in Appendix A to write down an expression
for P D2 . Following [18], we find
 
P D2 (x1 , yi(1) , x2 , yi(2) )

 
1 
2
D R R R
= δ(C(i) )x(i)0(i) y1(i)
1(i) 2(i)
y2(i) 3(i)
y3(i)
α ,bα̇ =0 ca =0 i=1
a(i) (i) (i)


  1  2  1  2 
a(i) a(i) b(i) b(i)
×
k,k  ,p,p  =0
k k p p


1 4  a 
 c
× (i) C(n, n12 , n21 ), (C.9)
la
l1 ,l2 ,l3 ,l4 =0 a=1

where the coefficient C(n, n12 , n21 ) are given by [8]


( 12 (n12 + n21 ))(1 + 12 (n − n12 − n21 ))
C(n, n12 , n21 ) = (−1)(1+n12 n21 ) , (C.10)
(1 + n2 )
with C(n, 0, 0) = h(n/2). Moreover,

2
 
n= 1
a(i) + a(i)
2
+ b(i)
1
+ b(i)
2
+ c(i)
1
+ c(i)
2
+ c(i)
3
+ c(i)
4
,
i=1
144 T. Harmark, M. Orselli / Nuclear Physics B 757 (2006) 117–145


2 
2 
4 
4
n12 = α
a(1) + α̇
b(1) + a
c(1) − k − k  − p − p − la ,
α=1 α̇=1 a=1 a=1

2 
2 
4 
4
n21 = α
a(2) + α̇
b(2) + a
c(2) − k − k  − p − p − la . (C.11)
α=1 α̇=1 a=1 a=1

References

[1] E. Witten, Anti-de Sitter space, thermal phase transition, and confinement in gauge theories, Adv. Theor. Math.
Phys. 2 (1998) 505, hep-th/9803131.
[2] B. Sundborg, The Hagedorn transition, deconfinement and N = 4 SYM theory, Nucl. Phys. B 573 (2000) 349–363,
hep-th/9908001.
[3] A.M. Polyakov, Gauge fields and space–time, Int. J. Mod. Phys. A 17 (S1) (2002) 119–136, hep-th/0110196.
[4] O. Aharony, J. Marsano, S. Minwalla, K. Papadodimas, M. Van Raamsdonk, The Hagedorn/deconfinement phase
transition in weakly coupled large N gauge theories, hep-th/0310285.
[5] S.W. Hawking, D.N. Page, Thermodynamics of black holes in anti-de Sitter space, Commun. Math. Phys. 87 (1983)
577.
[6] J.A. Minahan, K. Zarembo, The Bethe-ansatz for N = 4 super-Yang–Mills, JHEP 0303 (2003) 013, hep-th/
0212208.
[7] N. Beisert, C. Kristjansen, M. Staudacher, The dilatation operator of N = 4 super-Yang–Mills theory, Nucl. Phys.
B 664 (2003) 131–184, hep-th/0303060.
[8] N. Beisert, The complete one-loop dilatation operator of N = 4 super-Yang–Mills theory, Nucl. Phys. B 676 (2004)
3–42, hep-th/0307015.
[9] N. Beisert, M. Staudacher, The N = 4 SYM integrable super spin chain, Nucl. Phys. B 670 (2003) 439–463, hep-
th/0307042.
[10] N. Beisert, V. Dippel, M. Staudacher, A novel long range spin chain and planar N = 4 super-Yang–Mills,
JHEP 0407 (2004) 075, hep-th/0405001.
[11] N. Beisert, The su(2|3) dynamic spin chain, Nucl. Phys. B 682 (2004) 487–520, hep-th/0310252.
[12] H. Lin, O. Lunin, J.M. Maldacena, Bubbling AdS space and 1/2 BPS geometries, JHEP 0410 (2004) 025, hep-
th/0409174.
[13] M. Blau, J. Figueroa-O’Farrill, C. Hull, G. Papadopoulos, A new maximally supersymmetric background of IIB
superstring theory, JHEP 0201 (2002) 047, hep-th/0110242.
[14] D. Berenstein, J.M. Maldacena, H. Nastase, Strings in flat space and pp waves from N = 4 super-Yang–Mills,
JHEP 0204 (2002) 013, hep-th/0202021.
[15] O. Aharony, S.S. Gubser, J. Maldacena, H. Ooguri, Y. Oz, Large N field theories, string theory and gravity, Phys.
Rep. 323 (2000) 183, hep-th/9905111.
[16] J. McGreevy, L. Susskind, N. Toumbas, Invasion of the giant gravitons from anti-de Sitter space, JHEP 0006 (2000)
008, hep-th/0003075.
[17] D. Yamada, Quantum mechanics of lowest Landau level derived from N = 4 SYM with chemical potential, hep-
th/0509215.
[18] M. Spradlin, A. Volovich, A pendant for Polya: The one-loop partition function of N = 4 SYM on R × S 3 , Nucl.
Phys. B 711 (2005) 199–230, hep-th/0408178.
[19] D. Yamada, L.G. Yaffe, Phase diagram of N = 4 super-Yang–Mills theory with R-symmetry chemical potentials,
hep-th/0602074.
[20] P. Basu, S.R. Wadia, R-charged AdS5 black holes and large N unitary matrix models, Phys. Rev. D 73 (2006)
045022, hep-th/0506203.
[21] H. Liu, Fine structure of Hagedorn transitions, hep-th/0408001.
[22] M. Cvetic, S.S. Gubser, Thermodynamic stability and phases of general spinning branes, JHEP 9907 (1999) 010,
hep-th/9903132.
[23] T. Harmark, N.A. Obers, Thermodynamics of spinning branes and their dual field theories, JHEP 0001 (2000) 008,
hep-th/9910036.
[24] N. Beisert, The dilatation operator of N = 4 super-Yang–Mills theory and integrability, Phys. Rep. 405 (2005)
1–202, hep-th/0407277.
[25] M. Gomez-Reino, S.G. Naculich, H.J. Schnitzer, More pendants for Polya: Two loops in the SU(2) sector,
JHEP 0507 (2005) 055, hep-th/0504222.
T. Harmark, M. Orselli / Nuclear Physics B 757 (2006) 117–145 145

[26] J. Callan, G. Curtis, et al., Quantizing string theory in AdS5 × S 5 : Beyond the pp-wave, Nucl. Phys. B 673 (2003)
3–40, hep-th/0307032.
[27] J.A. Minahan, The SU(2) sector in AdS/CFT, Fortschr. Phys. 53 (2005) 828–838, hep-th/0503143.
[28] M. Bertolini, J. de Boer, T. Harmark, E. Imeroni, N.A. Obers, Gauge theory description of compactified pp-waves,
JHEP 0301 (2003) 016, hep-th/0209201.
[29] L.A. Pando Zayas, D. Vaman, Strings in RR plane wave background at finite temperature, Phys. Rev. D 67 (2003)
106006, hep-th/0208066;
B.R. Greene, K. Schalm, G. Shiu, On the Hagedorn behaviour of pp-wave strings and N = 4 SYM theory at finite
R-charge density, Nucl. Phys. B 652 (2003) 105–126, hep-th/0208163;
Y. Sugawara, Thermal amplitudes in DLCQ superstrings on pp-waves, Nucl. Phys. B 650 (2003) 75–113, hep-
th/0209145;
R.C. Brower, D.A. Lowe, C.-I. Tan, Hagedorn transition for strings on pp-waves and tori with chemical potentials,
Nucl. Phys. B 652 (2003) 127–141, hep-th/0211201;
Y. Sugawara, Thermal partition function of superstring on compactified pp-wave, Nucl. Phys. B 661 (2003) 191–
208, hep-th/0301035;
G. Grignani, M. Orselli, G.W. Semenoff, D. Trancanelli, The superstring Hagedorn temperature in a pp-wave back-
ground, JHEP 0306 (2003) 006, hep-th/0301186;
F. Bigazzi, A.L. Cotrone, On zero-point energy, stability and Hagedorn behavior of type IIB strings on pp-waves,
JHEP 0308 (2003) 052, hep-th/0306102.
[30] A. Buchel, L.A. Pando Zayas, Hagedorn vs. Hawking–Page transition in string theory, Phys. Rev. D 68 (2003)
066012, hep-th/0305179.
[31] O. Aharony, J. Marsano, S. Minwalla, K. Papadodimas, M. Van Raamsdonk, A first order deconfinement transition
in large N Yang–Mills theory on a small S 3 , Phys. Rev. D 71 (2005) 125018, hep-th/0502149.
[32] J.L.F. Barbon, E. Rabinovici, Extensivity versus holography in anti-de Sitter spaces, Nucl. Phys. B 545 (1999)
371–384;
J.L.F. Barbon, I. Kogan, E. Rabinovici, On stringy thresholds in SYM/AdS thermodynamics, Nucl. Phys. B 544
(1999) 104–144, hep-th/9809033.
[33] M. Gunaydin, N. Marcus, The spectrum of the S 5 compactification of the chiral N = 2, D = 10 supergravity and
the unitary supermultiplets of U (2, 2|4), Class. Quantum Grav. 2 (1985) L11.
[34] H. Bethe, On the theory of metals. 1. Eigenvalues and eigenfunctions for the linear atomic chain, Z. Phys. 71 (1931)
205–226.
[35] F. Berruto, G. Grignani, G.W. Semenoff, P. Sodano, On the correspondence between the strongly coupled 2-flavor
lattice Schwinger model and the Heisenberg antiferromagnetic chain, Ann. Phys. 275 (1999) 254–296, hep-
th/9901142.
[36] L.D. Faddeev, L.A. Takhtajan, What is the spin of a spin wave? Phys. Lett. A 85 (1981) 375–377;
L.D. Faddeev, L.A. Takhtajan, Spectrum and scattering of excitations in the one-dimensional isotropic Heisenberg
model, J. Sov. Math. 24 (1984) 241–267.
[37] F.A. Dolan, H. Osborn, On short and semi-short representations for four dimensional superconformal symmetry,
Ann. Phys. 307 (2003) 41–89, hep-th/0209056.
[38] M. Bianchi, J.F. Morales, H. Samtleben, On stringy AdS5 × S 5 and higher spin holography, JHEP 0307 (2003) 062,
hep-th/0305052.
Nuclear Physics B 757 (2006) 146–171

Heterotic non-linear sigma models with anti-de Sitter


target spaces
Georgios Michalogiorgakis ∗ , Steven S. Gubser
Joseph Henry Laboratories, Princeton University, Princeton, NJ 08544, USA
Received 11 August 2006; accepted 30 August 2006
Available online 29 September 2006

Abstract
We calculate the beta function of non-linear sigma models with S D+1 and AdSD+1 target spaces in a 1/D
expansion up to order 1/D 2 and to all orders in α  . This beta function encodes partial information about
the spacetime effective action for the heterotic string to all orders in α  . We argue that a zero of the beta
function, corresponding to a worldsheet CFT with AdSD+1 target space, arises from competition between
the one-loop and higher-loop terms, similarly to the bosonic and supersymmetric cases studied previously
in [J.J. Friess, S.S. Gubser, Non-linear sigma models with anti-de Sitter target spaces, Nucl. Phys. B 750
(2006) 111–141]. Various critical exponents of the non-linear sigma model are calculated, and checks of
the calculation are presented.
© 2006 Elsevier B.V. All rights reserved.

1. Introduction

Particular interest attaches to backgrounds of string theory involving AdSD+1 because of their
relation to conformal field theories in D dimensions [2–4] (for a review see [5]). But because
these geometries (with some exceptions) arise from the near-horizon geometry of D-branes, for-
mulating a closed string description is complicated by the presence of Ramond–Ramond fields.
It was recently proposed [1] that AdSD+1 vacua might exist without any matter fields at all.
Instead of relying upon the stress–energy of matter fields to curve space, the proposal is that
higher powers of the curvature compete with the Einstein–Hilbert term to produce string-scale
AdSD+1 backgrounds. The main support for this proposal comes from large D computations of

* Corresponding author.
E-mail address: gmichalo@princeton.edu (G. Michalogiorgakis).

0550-3213/$ – see front matter © 2006 Elsevier B.V. All rights reserved.
doi:10.1016/j.nuclphysb.2006.08.026
G. Michalogiorgakis, S.S. Gubser / Nuclear Physics B 757 (2006) 146–171 147

the beta function for the quantum field theory on the string worldsheet. Before discussing these
computations, let us review the lowest-order corrections to the beta function in an α  expansion:
α 2  
bosonic: βij = α  Rij + Riklm Rj klm + O α  3 ,
2
α 2  
heterotic: βij = α  Rij + Riklm Rj klm + O α  3 ,
4
ζ (3)α  4    
type II: βij = α  Rij + Rmhki Rj rt m R k qs r R tqsh + R k qs t R hrsq + O α  5 .
2
(1)
These expressions are obtained using dimensional regularization with minimal subtraction, and
all derivatives of curvature are assumed to vanish as well as all matter fields. Derivatives of
curvature indeed vanish for symmetric spaces: for example,
1
Rij kl = − (gik gj l − gil gj k ) (2)
L2
in the case of AdSD+1 . One indeed finds non-trivial zeroes for AdSD+1 from all three beta func-
tions in (1). An examination of higher order corrections in the bosonic and type II cases shows
that the zero persists in the most accurate expressions for the beta function that are available at
present; however its location changes significantly, converging to α  D/L2 = 1 as D becomes
large. One aim of the present paper is to pursue similar large D computations in the heterotic
case.
It should be clear from the outset that the question of the existence of AdSD+1 vacua with
α  D/L2 close to unity is a difficult one to settle perturbatively. Fixed order computations are not
reliable guides because the scale of curvature is close to the string scale. Large D computations
with finite α  D/L2 seem to be a better guide, but they too could be misleading, mainly because
higher order effects in 1/D than we are able to compute could change the behavior of the beta
function significantly. These difficulties were discussed at some length in [1]. Also, the vanishing
of a beta function such as the ones in (1) is only a necessary condition for constructing a string
theory: one must also cancel the Weyl anomaly and formulate a GSO projection that ensures
modular invariance and the stability of the vacuum.
There is a more general reason to be interested in high-order computations of the beta function
on symmetric spaces: from them we can extract information about the structure of high powers
of the curvature that is quite different from what is available from expansions of the Virasoro
amplitude. While the latter tells us about terms involving many derivatives but only four powers
of the curvature (because only four gravitons are involved in the collision), the former tells us
about many powers of the curvature with no extra derivatives.
The organization of the rest of this paper is as follows. In Section 2, some general properties
of the heterotic NLσ M are discussed. In Section 3, the formalism and the results at 1/D order
are presented. In Section 4, the critical exponents at 1/D 2 , the beta function, and the central
charge of the CFT are computed. The appendices include a brief explanation of the method of
the calculation for the diagrams needed and the values of these diagrams.

2. The heterotic non-linear sigma model

As in [1], much will be made of a connection through analytic continuation of the NLσ M on
AdSD+1 and the NLσ M on S D+1 . If L is the radius of S D+1 and g = α  /L2 , then continuing to
148 G. Michalogiorgakis, S.S. Gubser / Nuclear Physics B 757 (2006) 146–171

Fig. 1. The Feynman rules for the heterotic sigma model. The shaded circles indicate a dressed propagator. The circles
indicate that a loop involving only the components of the Φ superfield receives a factor of N . We have suppressed the
tensor structure of the rules since only δμν appears.

negative g leads to the AdSD+1 NLσ M. The argument in [1] is slightly formal because it relies
on an order-by-order perturbative evaluation of the partition function.
The action for the S D+1 heterotic NLσ M is
 
1    1
S= d 2 x d θ̄ D+ Φ∂− Φ + Λ Φ 2 − 1 + d 2 x λA ∂+ λA , (3)
4πg 4πg
where
∂ ∂ ∂
Φ = S + θ̄ Ψ, Λ = u + θ̄ σ, D+ = + θ̄ , ∂± = . (4)
∂ θ̄ ∂x+ ∂x∓
Λ is a spinorial superfield, and u and Ψ have opposite chirality. This leads to the action

1    
S= d 2 x (∂S)2 + Ψ̄ i∂Ψ + σ S 2 − 1 + 2ūΨ S . (5)
4πg
We have omitted the fermions λA from (5) because they decouple from the gravitational action
when the gauge field is set to zero [6,7] as in our case. The Feynman rules for the theory (5) can
be seen in Fig. 1. There is also a tadpole for σ , but we omit it because it does not contribute to
the Dyson equations for the scaling parts of the dressed propagators, as in [8].
After a change of variables that renders the kinetic terms canonical, we can continue to neg-
ative values of g as in [1] to obtain an AdSD+1 heterotic NLσ M. Quantities that are computed
locally and perturbatively, such as n-point functions, cannot distinguish between a space of pos-
itive or negative curvature. As the beta function is derived from such quantities, it too can be
continued to negative g, at least order by order in perturbation theory.
G. Michalogiorgakis, S.S. Gubser / Nuclear Physics B 757 (2006) 146–171 149

The heterotic NLσ M on S D+1 is a generalization of the O(D + 2) model, and much of the
relevant literature concentrates on an expansion in 1/(D + 2) rather than 1/D. We will therefore
set

N =D+2 (6)
and work with N or D, according to convenience, in the rest of this paper.

2.1. Some properties of the heterotic NLσ M for large D

It is known that in the bosonic sigma model a mass appears [9] in the 1/N expansion. The
same phenomenon appears in the supersymmetric extension of the sigma model [10,11] where
also the fermions acquire the same mass, signaling chiral symmetry breaking. In the heterotic
case the bosons S also acquire the same mass, showing that the interaction term does not destroy
this effect. To understand this, let us start from our action (5), and in the partition function inte-
grate first the fermionic fields and then the bosons, since the action is quadratic in these. We have
omitted normalization factors of the partition function in the following.
   
i   2  
Z = DS DΨ Dσ Du exp − d x (∂S) + i Ψ̄ ∂/ Ψ + σ S − 1 + 2ū(S · Ψ )
2 2
4πg
   

 N/4 i   1
= DS Dσ Du det(i/ ∂) exp d 2 x S ∂ 2 − σ S − S i ū uS i
4πg i/

  
−N/2   
 N/4 1 i
= Dσ Du det(i/ ∂) det −∂ 2 − σ ū ū exp d 2x σ
i/
∂ 4πg

⇒ Z = Dσ Du eiSeff , (7)

where the effective action for the Lagrange multiplier fields is given by
  

1 N N 1
Seff = d 2 x σ − Tr log(i/ ∂ ) + Tr log −∂ 2 − σ − ū u . (8)
4πg 4 2 i/

Since we are taking the limit N → ∞ with g0 N finite, we see that all terms in the action are
of order N . We can evaluate this integral by the method of steepest descent, i.e. by finding the
classical value of σ , u that minimizes the exponent, as is done for instance in [12,13]. This gives
the variational equations

1 1
x| |x = ,
−∂ 2 − σ (x) − ū i1∂/ u 2πNg
1
∂/ u
x| |x = 0. (9)
−∂ 2 − σ (x) − ū i1∂/ u

Because the right-hand sides are constant, the left-hand sides must also be constant. A solution
to these equations is given by

u(x) = 0, σ (x) = −m2 . (10)


150 G. Michalogiorgakis, S.S. Gubser / Nuclear Physics B 757 (2006) 146–171

It is easy to see that ∂1 u(x) = const has as its only solution u = 0. This is in contrast to the
supersymmetric case [12], where there are three solutions. Now m2 must satisfy

1 1
d 2k 2 = . (11)
k + m2 2πg0
Using a simple-momentum cutoff, 1
2πg0 N = 1
2π log Λ
m . By renormalizing at a scale M we get

1 1 1 Λ
= + log . (12)
2πg0 N 2πgN 2π M
Solving for the mass m we get m = M exp[−1/gN ]. Since this is a physical mass we expect that
it does not depend on the renormalization scale. Using the Callan–Symanzik equation for m,
 
∂ ∂
M + β(g) m(g, M) = 0, (13)
∂M ∂g
gives the beta function β(g) = −g 2 N . The mass is the same in the bosonic, supersymmetric,
[9,11] and heterotic model. This could have been predicted since the first order β function is
the same in all models, βij = α  Rij . Another way to get the same result is to calculate the ef-
fective potential for the σ field and see that the minimum of the potential is not at zero but at
σ = Me−1/gN . One can go further and examine the effective action (8). It is easy to evaluate
the counterterms needed for one-loop renormalization, as we have already computed the wave
function renormalization of the σ field, and doing so one finds
   
1 Λ2 N N 1
L0,eff = 1 + gN log 2 σ − Tr log ∂/ + Tr log −∂ 2 − σ − ū u . (14)
4πg M 4 2 i/

The bare and the dressed quantities are related by
gN Λ2
σ0 = Zσ, u0 = Z 1/2 u, g0 = Z −1 g, Z=1+ log 2 . (15)
2 M
Calculating the quadratic terms in the fields u, σ will give us the propagators for these fields. We
can easily find
N 1 1 N 1 1
Seff = Tr ū u − Tr σ σ. (16)
2 −∂ − m i/
2 2 ∂ −m 4 −∂ − m −∂ − m2
2 2 2

The next step is to evaluate the propagators. One finds [11,12]


2i   2i  2   
Su (k) = − (/k − 2m)V k 2 , Dσ (k) = 4m − k 2 V k 2 , (17)
N N
where in d dimensions
 2    −1
  (4π)d/2 4m − k 2 1−d/2 k2
V k2 = F
2 1 2 − d/2, 1/2, 3/2; . (18)
4(2 − d/2) 4 k 2 − 4m2
In two dimensions this simplifies to

 −1
 2 k2 k2
V k =π arctanh . (19)
k 2 − 4m2 k 2 − 4m2
It is easy to see that the u propagator in d = 2 dimensions has a pole at k 2 = 4m2 . This means
that there is a particle with this mass. Since the classical equations give u = −iS/ ∂ Ψ , there is
G. Michalogiorgakis, S.S. Gubser / Nuclear Physics B 757 (2006) 146–171 151

a boson–fermion bound state, created by the operator S/ ∂ Ψ . This is in agreement with [11], but
there is no supersymmetric corresponding fermion–fermion bound state, since the Gross–Neveu
interaction that is responsible for it is absent.
One can also extend the calculation of [12,14] to show that there is no multi-particle produc-
tion in the heterotic sigma model. For example, the process 2 → 4 particles can be shown to
vanish. The reasoning is that the formalism of [14], valid for the bosonic case, can be extended
to include superfields.

3. Critical exponents in the 1/D expansion

3.1. General discussion

The method used to determine the critical exponents is the one developed in [8,15] for the
bosonic model and extended to the supersymmetric case in [16,17]. To this end one writes ex-
pressions for the propagators of the fields near the critical point. In keeping with the notation of
[8,15,18] we assign dimensions to the fields
1
dim[S] = dim[Ψ ] − = ΔS = (d − 2 + η)/2,
2
1
dim[σ ] = dim[u] + = Δσ = 2 − η − χ. (20)
2
For small but non-zero x, the two-point functions may be expanded as follows:
ΓSS   2λ
 1 + γP ΓΨ Ψ x/  
GSS (x) = 1 + ΓSS x , GΨ Ψ (z) = 1 + ΓΨ Ψ x 2λ ,
x 2ΔS 2 x 2ΔS +2
Γσ σ   1 − γP Γuux/  
Gσ σ (x) = 2Δ 1 + Γσ σ x 2λ , Guu (x) = 2Δ
 2λ
1 + Γuu x , (21)
x σ 2 x σ

where
 
1 0
γP = ρ ρ =
0 1
(22)
0 −1
is the chirality matrix in 2 dimensions. We have omitted the O(N) indices because both the
propagators and the vertex are proportional to δμν . So all Green’s functions can be expressed as
a scalar function times products of δμν , and one does not have to worry about tensor structures
of the form (x − y)μ (x − y)ν . Then one can write the Dyson equations in a 1/D expansion for
the propagators. Graphical expressions of these equations are shown in Figs. 2 through 5. The
Dyson equations impose consistency conditions on the critical exponents that determine them
completely. The graphs that appear in the Dyson equations are the 1PI graphs with the excep-
tion of graphs which contain subgraphs that already appear in the Dyson equations at a lower
effective loop order: in other words, we exclude diagrams that are already taken into account by
expressions for the corrected propagators. The effective loop order is the number of loops minus
the number of loops involving only the components of the Φ superfield.
The left-hand side of each Dyson equation is a 1PI propagator, which is the inverse of the
connected two-point function. These inverse propagators are computed by first passing to Fourier
space using

e−ik·x π μ α(Δ)22(μ−Δ)
d d k 2Δ = . (23)
k x 2(μ−Δ)
152 G. Michalogiorgakis, S.S. Gubser / Nuclear Physics B 757 (2006) 146–171

The inverse propagators are found to be1


p(ΔS )  
G−1
SS (x) = 2(2μ−Δ )
1 − q(ΔS , λ)ΓSS  2λ
x ,
ΓSS x S

1 − γ p(Δ S )/
x  
G−1 1 − s(ΔS , λ)ΓΨ Ψ x 2λ ,
P
Ψ Ψ (x) = 2(2μ−Δ )
2 ΓΨ Ψ x S

p(Δσ )  
G−1
σ σ (x) = 1 − q(Δσ , λ)Γσ σ x 2λ ,
Γσ σ x 2(2μ−Δσ )
1 + γP r(Δσ − 1)/ x  
G−1
uu (x) = 2(2μ−Δ +1)
1 − s(Δσ − 1, λ)Γuu  2λ
x , (24)
2 Γuu x σ

where

μ = d/2, ΔS = μ − 1 + η/2, Δσ = 2 − η − χ, (25)


and for arbitrary y,
(μ − y) α(y − μ) yp(y)
α(y) = , p(y) = 2μ , r(y) = ,
(y) π α(y) μ−y
α(y − λ)α(y + λ − μ) y(y − μ)q(y, λ)
q(y, λ) = , s(y, λ) = . (26)
α(y)α(y − μ) (y − λ)(y + λ − μ)
To calculate the beta function, one first evaluates the critical exponents of the model at the fixed
point in d = 2 +  dimensions and then uses the relation
1
λ = − β  (gc ), (27)
2
valid at the critical point, to extract β(g). This is possible because, in dimensional regularization
with minimal subtraction, the only  dependence in β(g) is an overall additive term: See [1] for
details. Expanding in 1/D with κ = gD held fixed, one finds

bi (κ) ∞
λi () β(g)
λ() = , =−κ + , (28)
Di g Di
i=0 i=1
where

κc λ1 (ξ )
λ0 (κc ) = , b1 (κ) = −2κ dξ ,
2 ξ2
0

λ2 (ξ ) − b1 (ξ )λ1 (ξ )
b2 (κ) = −2κ dξ . (29)
ξ2
0
Note that the critical exponent λ is a measurable quantity and as such it should not depend on
the renormalization scheme used. Passing from the λi (κ) to the bi (κ) does introduce significant
scheme dependence.

1 Note that G −1 1+γp  1 0


Ψ Ψ · GΨ Ψ does not strictly give the unit matrix but instead 2 = 0 0 in two dimensions, which is
what we want. Also note that in the Dyson equation for Ψ in the right-hand side one encounters u propagators that give
the right chiral structure, and vice versa for the Dyson equation of the u field.
G. Michalogiorgakis, S.S. Gubser / Nuclear Physics B 757 (2006) 146–171 153

Fig. 2. The Dyson equations for the S propagator.

Fig. 3. The Dyson equations for the Ψ propagator.

3.2. Critical exponents at order 1/D

The Dyson equations can be expressed in terms of parameters

2 Γ
ΓSS σσ ΓSS ΓΨ Ψ Γuu
w= , v= , (30)
(2πg)2 (2πg)2
which can be regarded as dressed vertex factors for the two vertices shown in Fig. 1. The lead-
ing non-trivial Dyson equations come from the graphs labeled Σ0,A , Σ0,B , Φ0 , Π0 , and F0 in
Figs. 2 through 5: The tree-level graphs make no contribution to the leading scaling behavior.
The quantities ΓSS , Γ  , Γ  , and Γ  describe how far one is removed from the fixed point;
ΨΨ σσ uu
so in particular one must be able to set them all to zero and get a self-consistent set of equations.
Then the dependence of each graph on the position-space separation x is just an overall power
of x. Matching these overall powers leads simply to the constraint χ = 0. Matching other factors
154 G. Michalogiorgakis, S.S. Gubser / Nuclear Physics B 757 (2006) 146–171

Fig. 4. The Dyson equations for the σ propator.

Fig. 5. The Dyson equations for the u propagator.

leads to the equations


p(ΔS ) + w + v = 0, r(ΔS ) + v = 0,
1 2
r(Δσ − 1) + v = 0, p(Δσ ) + w = 0, (31)
N N
which determine the quantities ΔS , Δσ , w, and v as functions of μ and N . The system is in
fact over-determined if we recall the relations (25) and the constraint χ = 0. But we will see in
Section 4.1 that χ = 0 is only a leading order result; thus to solve (31) we expand
ηi χi wi vi
η= , χ = , w = , v = . (32)
Di Di Di Di
i0 i0 i0 i0

Then one straightforwardly extracts from (31) the coefficients


η0 = 0, χ0 = 0, w0 = 0, v0 = 0,
(2μ − 1)
η1 = −2 ,
(μ − 1)2 (1 − μ) 2 (μ − 1)(μ + 1)
(2 − μ)(μ − 1)(μ + 1)
w1 = η1 ,
2π 2μ
(1 − μ)(μ − 1)(μ + 1)
v1 = − η1 . (33)
2π 2μ
G. Michalogiorgakis, S.S. Gubser / Nuclear Physics B 757 (2006) 146–171 155

Higher order coefficients receive contributions from higher order graphs. Note that χ0 = 0 could
be obtained either from matching overall powers of x or from Eqs. (31).
Now consider non-zero coefficients ΓSS  , Γ  , Γ  , and Γ  : This corresponds to moving
ΨΨ σσ uu
away from the fixed point. Linearizing the Dyson equations leads to the constraints
 
−p(ΔS )q(ΔS , λ)ΓSS + w(ΓSS + Γσ σ ) + v(ΓΨ Ψ + Γuu

) = 0,
−r(ΔS )s(ΔS , λ)ΓΨ Ψ + v(ΓSS
 
+ Γuu ) = 0,
 
−r(Δσ − 1)s(Δσ − 1, λ)Γuu + N v(ΓSS + ΓΨ Ψ ) = 0,
−p(Δσ )q(Δσ , λ)Γσ σ + N ΓSS

w = 0. (34)
Graphically, these equations arise from using the leading power-law expressions (e.g. ΓSS /x 2ΔS
rather than GSS (x)) for all propagators except one, chosen arbitrarily; and for that one, use the
correction term (e.g. ΓSS ΓSS /x 2ΔS −2λ ). The linear Eqs. (34) must admit a non-zero solution for
   
ΓSS , ΓΨ Ψ , Γσ σ , Γuu in order for the correction terms to describe a genuine deformation of the
critical point. So the corresponding determinant must vanish, which leads to
 

2 v  
(w + v)q(ΔS , λ) + w 1 − − 1 − s(Δσ , λ)s(Δσ − 1, λ)
q(Δσ , λ) s(Δσ − 1, λ)
(1 − s(Δσ − 1, λ))2
=v . (35)
s(Δσ − 1, λ)
Note that setting v = 0 gives the equation valid for the bosonic model as expected. To simplify
2−μ
(35), one can use (33) and note that w1 = v1 μ−1 . So far we have not used any expansion in 1/D.
Using the expansions (32) and
λi
λ= (36)
Di
i0

we can determine
1
λ0 = μ − 1, λ1 = (2μ − 1)(μ − 1)η1 . (37)
2
Another way to compute the original determinant, is the following: One notes that r(ΔS )s(ΔS ,
λ) ∼ 1/D 0 , while all other terms scale at least as 1/D 1 . This means that, in expanding the 4 × 4
determinant of (34), the first order contribution comes only from the determinant of the 3 × 3
matrix
⎛ ⎞
−p(ΔS )q(ΔS , λ) + w w v
AI ≡ ⎝ w −p(Δσ )q(Δσ , λ)/N 0 ⎠,
v 0 −r(Δσ − 1)s(Δσ − 1, λ)/N
(38)
where in each element of the matrix we only keep the first term in the 1/D expansion. This
determinant provides the first 1/D term of (35) and thus reproduces (37).
1 = 2 λ1 , even
1 bos
It is interesting to compare with the bosonic case. It is easily seen that λhet
het bos
though η1 , w1 , and v1 are not so simply related to η1 and the corresponding vertex factor for
sup
1 = 2 λ1 is expected: We know that λ1 = 0 in the type II case,
1 bos
the bosonic case. The relation λhet
and having half the fermions in the heterotic case will cancel only half the bosonic contribution.
156 G. Michalogiorgakis, S.S. Gubser / Nuclear Physics B 757 (2006) 146–171

3.3. A check of the calculation

Following [18], we see that η is the anomalous dimension of the S propagator. So far we have
computed it in the 1/D expansion using techniques in position space. One can also straightfor-
wardly compute η1 in momentum space using the expressions for the propagators that we have
previously found (17). Firstly we note that for small k,
 
−2+η −2+ηo η1  
G̃SS (k) ∼ k ∼k 1+ log k + O 1/N 2
, (39)
N
with G̃SS the Fourier transform of the S propagator. But G̃SS (k) can also be determined from
the one-particle irreducible diagrams Σ(k 2 ):
 
 2 η1
G̃−1
SS (k) = k 2
+ Σ k − Σ(0) ∼ k 2
1 − log k , (40)
N
since η0 = 0. Having calculated the propagators of the Lagrange multiplier fields it is straight-
forward to compute the Σ(k 2 ) from the one-loop diagrams. We find
 
 2 dd p i  2 dd p i
Σ k = D u p − tr Su (/k)
(2π)d (p − k)2 (2π)d (/k − p /)

2 / )V (p 2 )
d d p tr(/kp
=
N (2π)d (p + k)2
√ M π
2d−2 /N π p 2 k cos θ sind−2 θ
=  d−1  dp dθ 2 ,
2 F1 (2 − d/2, 1/2, 3/2, 1)(2 − d/2) 2 p + k 2 + 2pk cos θ
0 0
(41)
where we have put m2 = 0 as usual [17], and M is the cutoff. One notes that, for small k the
k 2 log k behavior comes from the small p region [18]. The integral is trivial to do, and for the
k 2 log k part it gives
  2d−1 1
Σ k2 = −     k 2 log k. (42)
Nd 2 F1 (2 − d/2, 1/2, 3/2, 1) 2 − d2  d2
It is easy to see using d = 2μ and properties of the gamma function that this coincides with
the expression for η1 in (33). An easier way to do the checking is the following. One can write
similar expressions for Σ(k 2 ) for the bosonic and supersymmetric models [8,17]. Then it is easy
to observe that
     
Σbos k 2 − Σbos (0) = −Σsup k 2 + 2Σhet k 2 . (43)
This easily gives2
2ηhet = ηbos + ηsup . (44)
With [8,17]
(2 − μ) 4 (2μ − 2)
ηbos = ηsup , ηsup = (45)
μ N  (μ − 1)(2 − μ)(μ)
2

we find ηhet = μ1 ηsup which agrees with (33).

2 We have used that Σ (0) = 0 and Σ (0) = 0.


sup het
G. Michalogiorgakis, S.S. Gubser / Nuclear Physics B 757 (2006) 146–171 157

4. Results at order 1/D 2

Each graph in Figs. 2–5 carries an overall factor 1/D M where M is the number loops minus
the number of loops containing only S and Ψ . To see this, first note that each propagator GXX
carries a factor ΓXX (where X = S, Ψ , σ , or u). Next note that the amplitude for each graph
must contain an overall factor which is a product of the factors w = ΓSS 2 Γ /(2πg)2 and v =
σσ
2
ΓSS ΓΨ Ψ Γuu /(2πg) , one for each vertex in the graph. The overall factor 1/D M arises because
w and v scale as 1/D and because each loop containing only S and Ψ carries a factor of N . The
graphs in Figs. 2 and 3 are those with M  2, and the ones in Figs. 4 and 5 are those with M  1.
Together, these are all the graphs that can contribute to η, w, v, and λ through order 1/D 2 , and
they also determine χ through order 1/D. Because we quote final results in terms of 1/D, we
must keep in mind the relation between expansions in 1/N and 1/D:
w̃i wi
w= = , w1 = w̃1 , w2 = w̃2 − 2w̃1 , (46)
Ni Di
i0 i0
with similar relations for other quantities.

4.1. Calculation of η2

A technical complication arises in the 1/D 2 corrections to the Dyson equations that was
explained and resolved in [8,15]. The problem is that the higher-loop graphs diverge when χ = 0.
In fact, χ = 0 only up to 1/D corrections. But it convenient to regularize the “divergence” and
extract finite expressions for the two-loop Dyson equations through the following steps:

(1) Shift χ → χ + Δ.
(2) Expand the amplitudes for individual graphs in powers of Δ.
(3) Cancel 1/Δ terms against certain counter-terms in the Lagrangian.
(4) Fix χ1 by setting to zero certain terms in the Dyson equation which depend logarithmically
on the position-space separation x and which, if non-zero, would spoil self-consistency.

We will now go through these steps in detail for the Ψ propagator. The reader who wishes to
bypass the technical details can skip to (56) and (57), which are the two-loop Dyson equations
with all divergences removed. But the results (54) and (58) for χ1 provide important consistency
checks.
When χ = 0, x dependence cannot be canceled out of the Dyson equations in a simple way:
 =Γ  
Setting ΓSS Ψ Ψ = Γσ σ = Γuu = 0, one obtains for the Ψ propagator’s Dyson equation
 χ  2χ  3χ
r(ΔS ) + v x 2 + v 2 x 2 Φ1 + N wv 2 x 2 Φ2 = 0. (47)
Here Φ1 and Φ2 are functions of ΔS , Δσ , and μ which diverge when 2ΔS + Δσ − 2μ = −χ = 0.
Although these are in some sense an artifact of a limit (χ → 0) which one cannot take indepen-
dently of the large N limit, it is convenient nevertheless to regulate them, as explained above, by
shifting
χ → χ + Δ. (48)
The amplitudes Φ1,2 may then be expanded as
Xi
Φi = + Φi + O(Δ), (49)
Δ
158 G. Michalogiorgakis, S.S. Gubser / Nuclear Physics B 757 (2006) 146–171

where both Xi and Φi are functions of ΔS , Δσ , and μ, subject to 2ΔS + Δσ − 2μ = 0. In


Appendix C we exhibit Φ1,2 in the form (49), as well as a number of related quantities that enter
into other Dyson equations. To cancel the divergent 1/Δ terms in Φ1,2 , one may rescale the
Lagrange multiplier fields in the original action (3). This rescaling amounts to adding counter-
terms to the action, and it can be expressed, to the relevant order, as
   
m1 m1
v→ 1+ v, w→ 1+ w. (50)
N N
(The factor on v and w is the same because of supersymmetry.) Subjecting (47) to the shift (48)
and the rescaling (50), it becomes, keeping terms up to 1/N 2 ,
 χ  
r(ΔS ) + x 2 v + v 2 Φ1 + N v 2 wΦ2
 
 χ m1  χ X1  2χ X2
+ x2 v1 + v12 x 2 + N v12 w1 x 2 = 0. (51)
N Δ Δ
The last line contains all the divergent pieces. Setting χ = 0 [8] and taking the limit Δ → 0
determines m1 as
m1 X1 X2
−v1 = v12 + N v12 w1 . (52)
N Δ Δ
Plugging (52) back into (51), and now considering a finite χ we get
 2 2χ 
 χ   (x ) − (x 2 )χ
r(ΔS ) + x 2 v + v 2 Φ1 + N v 2 wΦ2 + v12 X1
χ
 2 3χ 
(x ) − (x ) 2 χ
+ Nv12 w1 X2 = 0. (53)
χ
When one expands χ = χ1 /N + O(N −2 ), there are terms that behave as log x 2 . One gets rid of
these if χ1 obeys

χ1 = −v1 X1 − 2v1 w1 X2 . (54)


We could have derived (52), (54) purely within the N → ∞ limit. In this setup there is no need
for Δ and χ is taken to be finite. Note that it behaves as χ ∼ 1/N since χ0 = 0 (33). The Dyson
equation after the rescaling (50) is (51) with Δ replaced with χ . In taking the N → ∞ limit there
are terms that diverge linearly with N and terms that behave as log x 2 . Respectively these are
 
 2 χ m1 2 X1 X2
x v1 + v1 + N v 1 w1
2
,
N χ χ
   
m1
log x 2 χv1 1 + + 2v12 X1 + 3N v12 w1 X2 . (55)
N
Setting these to zero fixes m1 , χ1 as in (52), (54), with Δ replaced by χ . We choose to keep
the Δ shift, as is common in the literature [8,15,19,20]. If one wishes to translate our results, in
the N → ∞ formalism, only a simple substitution of Δ → χ1 /N is needed in the values of the
diagrams given in Appendix C. What is left is the finite correction to the leading Dyson equation
for the Ψ propagator:

r(ΔS ) + v + v 2 Φ1 + N wv 2 Φ2 = 0. (56)


G. Michalogiorgakis, S.S. Gubser / Nuclear Physics B 757 (2006) 146–171 159

Following the same procedure for the S, σ , and u Dyson equations, one gets the finite equations

p(ΔS ) + w + v + w 2 Σ1 − v 2 Σ2 + N w 3 Σ3 − 2N v 2 wΣ4 = 0,


N N N2 3  N2 2 
p(Δσ ) + w + w 2 Π1 + w Π2 − wv Π3 = 0,
2 2 2 2
r(Δσ − 1) + N v + N v 2 F1 + N 2 wv 2 F2 = 0, (57)
where Σi , Πi , and Fi
are the finite parts of Σi , Πi , and Fi , listed in Appendix C. The minus signs
in (57) come from fermion loops. From each Dyson equation one also gets a new determination
of m1 and χ1 :

1 w12 S1 − v12 S2 + w13 S3 − 2w1 v12 S4 −w12 S1 + v12 S2 − 2w13 S3 + 4w1 v12 S4
m1 = , χ1 = ,
Δ w1 + v 1 w1 + v 1
1 
m1 = − w1 P1 + w12 P2 − v12 P3 , χ1 = −w1 P1 − 2w12 P2 + 2v12 P3 ,
Δ
1
m1 = − (v1 Y1 + w1 v1 Y2 ), χ1 = −v1 Y1 − 2w1 v1 Y2 , (58)
Δ
where Pi , Si , Yi are the residues of Πi , Σi , and Fi , respectively.
Fortunately, the four seemingly independent determinations of m1 and χ1 all agree, as one
can check by explicitly evaluating (52), (54), and (58) using expressions from Appendix C with
ΔS = μ − 1 and Δσ = 2. This provides a check that the renormalization procedure we have
chosen to cancel the divergences of higher-loop graphs is consistent. Other schemes change the
values for individual amplitudes, but the critical exponents remain the same [16].
Interestingly, there is yet another consistency check on χ1 . One can show from (54) or (58)
that

χ1 = μ(2μ − 3)η1 . (59)


This is seen to comply with a scaling law formulated for the bosonic model in [18]:

2λ = 2μ − Δσ . (60)
That this relation is also valid in our case can be seen by applying the Callan–Symanzik equation
near the critical point for σ (p)σ (−p) or ū(p)u(−p) propagator.
Now we can solve (56)–(57) by eliminating w and v:

1 v2 w1 v12 
r(ΔS ) = r(Δσ − 1) + 12 (F1 − Φ1 ) + (F2 − Φ2 ). (61)
N N N2
Expanding ΔS , Δσ in (61), we can determine η˜2
η̃2 1   μ 1
= + (μ − 1)(2μ − 1) −1 + π cot μπ + H (2μ − 2) + + −1
η12 2μ μ − 2 2(μ − 1)
 
1 1
− μ(μ − 2) B(2μ − 3) − B(μ − 1) − + −2 , (62)
μ − 1 2μ − 3
where H (x) = ψ(x + 1) − ψ(1) and the B(x) function is defined in the appendix. The first line
just comes from the Hartree–Fock diagrams, i.e. by iterating the 1/N Dyson equation to the next
order, the second is the contribution of Φ1 , F1 and the third comes from Φ2 , F2 . We also can
160 G. Michalogiorgakis, S.S. Gubser / Nuclear Physics B 757 (2006) 146–171

determine the values of w2 , v2 as


w̃2 (2μ − 1)(μ − 1)  
= 3 − μ + (μ − 2)π cot μπ + (μ − 2)H (2μ − 3)
η1 w1 μ−2
 
− μ (7μ − 9)B(μ − 1) + (13 − 10μ)B(2) + (3μ − 4)B(2μ − 3)
μ(μ − 1)
− μ + 2μ(μ − 1) − , (63)
2μ − 3
ṽ2 η˜2 1 μ  
= 2− − + μ(μ − 2) + 2μ(μ − 2) B(2) − B(μ − 1) . (64)
η1 v1 η1 2μ 2
In (63), the first three terms come from iteration of the first order equations, while in (64), the
first two terms come from such iteration.

4.2. Calculation of λ2

As in Section 3.2, the calculation of λ2 through order 1/D 2 requires evaluating each graph
with one propagator altered from its leading power behavior (e.g. ΓSS /x 2ΔS for an S propagator)
to its sub-leading power behavior (e.g. ΓSS ΓSS  /x 2ΔS −2λ ). The four Dyson equations lead to four
 
linear equations in the quantities ΓSS , ΓΨ Ψ , Γσ σ , and Γuu
 :

  
−p(ΔS )q(ΔS , λ) + w + ΣS ΓSS + (w + Σσ )Γσ σ

+ (v + Σu )Γuu + (v + ΣΨ )ΓΨ Ψ = 0,
 
−r(ΔS )s(ΔS , λ) + ΦΨ ΓΨ Ψ + (v + ΦS )ΓSS 
+ (v + Φu )Γuu = 0,
 
p(Δσ )q(Δσ , λ)
− + Πσ Γσ σ + (w + ΠS )ΓSS
+ ΠΨ ΓΨ Ψ + Πu Γuu
= 0,
N
 
1  
− r(Δσ − 1)s(Δσ − 1, λ) + Fu Γuu + (v + FS )ΓSS + (v + FΨ )ΓΨ Ψ = 0, (65)
N
where for example we denote by ΣΨ all the diagrams that appear in the S propagator where the
Ψ propagator is corrected. As in Section 4.1, the amplitudes diverge when 2ΔS + Δσ − 2μ → 0,
and the same procedure described there to regulate and subtract the divergences and to remove
terms proportional to log x 2 carries over to the present case. The finite parts of all the quantities
in (65) are given in Appendix D, as well as some further remarks on their evaluation.
The system (65) must have a non-zero solution for the Γ ’s, so the determinant must be zero.
This determines λ2 . A way to calculate the determinant to sufficient accuracy is to note that
r(ΔS )s(ΔS , λ) ∼ 1/N 0 , and then expand the determinant into three 3 × 3 determinants, i.e.
expanding in the line of Ψ field Dyson equation. All terms have to be expanded up to 1/N 2
accuracy. One also notes that λ2 only appears in the expansion of p(ΔS )q(ΔS , λ) at this order.
So λ2 is going to be a linear combination of the various sums of diagrams given in the appendix,
factors of w2 and v2 , and terms that come from iterating the 1/N equations. The final result is
quite involved and we prefer to give it implicitly as
λ̃2 1 25 5(μ + 1) 5 μ−2
=− − + 25 − − − + 50(μ − 1)
η12 2(μ − 1)2 2(μ − 1) 2(μ − 2) 4(2μ − 3) 2
 
19 45 3 μ2 (2μ − 3)2
+ (μ − 1) − (μ − 2) + (μ − 2) − (2μ − 3) −
2
8 2 2 8(μ − 1)
G. Michalogiorgakis, S.S. Gubser / Nuclear Physics B 757 (2006) 146–171 161

 
21 9 25 10
+ (μ − 1) 2
− (μ − 2) − 75(2μ − 3) − −
4 4 (μ − 2) (2μ − 3)
 
μ F u + FS
− 2μ(μ − 1)ṽ2 − + 2μ(2μ − 3) w̃2 − 2μ(μ − 1)(μ − 2)
2 η1 v 1
 
(2μ − 3)2 Πσ μ(μ − 1) ΣS μ−1 Σσ
− μ(μ − 1) −2 − μ − 2μ
μ − 2 η1 v 1 (μ − 2) η1 v1
2 μ−2 η1 v 1
Fσ 3 
− 3μ(2μ − 3)(μ − 1) − π cot μπ + H (2μ − 4) . (66)
η1 v 1 2
The right-hand side is a function of μ which can be obtained explicitly by substituting the ex-
pressions (33), (63), (64), (D.31), (D.33), (D.34), (D.37), (D.38), and (D.39) into (66).

4.3. Calculation of the beta function

As explained in Section 3.1, we can calculate the beta function once we know λ. Noting that
(66) gives the 1/N 2 expansion term and subtracting 2λ1 we find
2 3 4  
λ0 = /2, λ1 () = + − + O 5 , (67)
  4 8 16
5 9  5
λ2 = + ζ (3)  + O  .
4
(68)
16 4
Using (29), we compute the beta function for the heterotic string in a constant curvature back-
ground:
   
1 3 g4D D g5D2 3 D
β(g) = −Dg − Dg −
2
1+ − −
2 4 2 4 2 3
 
3 1
− ζ (3)g 5 D 2 + O . (69)
2 D3
It is obvious that there is agreement with the first two loops of the expression (1), where we use
∂g g
β(g) = M =− g ij βij (70)
∂M N −1
and (2). We do not know of any calculation of the beta function of the heterotic string in the min-
imal subtraction scheme that goes beyond two loops. In [21,22] the beta function was computed
using the background field method, and found to be in three loops
 
(3) α 3 3 1 1
βij = Rikj l R kmnp R l mnp − Rlm Ri lnp Rj m np − Rj l R lmnp Rimnp . (71)
8 2 2 2
The appearance of the Ricci tensor means that it is not minimal subtraction. Divergences involv-
ing the Ricci tensor can only appear through closed loops where at least one propagator starts
and ends at the same vertex. Within the minimal subtraction scheme, at more than one loop these
terms combined with their counterterms never produce a simple pole [23,24]. A small check
4
of our result comes from the famous ζ (3) term, ζ (3)α m k r tqsh + R k t R hrsq ).
2 Rmhki Rj rt (R qs R qs
This term is identical in the bosonic [25–27], supersymmetric [23,24], and heterotic [28] cases.
In an expansion of the Virasoro amplitude, it is associated with the constant term in an expan-
sion in the Mandelstam variables s, t , and u. At loop order n + 1 in NLσ M calculations, it seems
162 G. Michalogiorgakis, S.S. Gubser / Nuclear Physics B 757 (2006) 146–171

likely that the coefficient of ζ (n) is the same for the bosonic, supersymmetric, and heterotic cases
(see [16] for a comparison of the bosonic and supersymmetric cases).
In [28–30], the absence of a three-loop term of the form α  3 R 3 was noted. The three-point
scattering amplitudes suggest that there are also no RF 2 or F 3 in the effective action. One knows
that identifying the gauge connection with the spin connection in the heterotic string effective
action will give the superstring effective action, where there is no α  3 term. So if there were any
R 3 terms in the heterotic case it would not be possible to cancel them. All this seems in conflict
with the (69), where the term proportional to g 4 would seem to correspond to an R 3 term in the
effective action. But it should be noted that the relation between the effective action and the beta
function is [31]
2  δSeff
2κD α = Kijkl βkl , (72)
δgij
where Kijkl can be computed perturbatively. In the bosonic case, this was done in the minimal
subtraction scheme in [31,32]; in the heterotic case, this was done in a different scheme in [21,
22]; but we do not know of a minimal subtraction calculation of Kijkl in the heterotic case. In the
bosonic case, Kijkl receives contributions starting at two loops, and it can be shown that this is
compatible with an independent calculation of the effective action using scattering amplitudes.
The same thing may happen in the heterotic case: in particular, R 3 terms could indeed be absent
from Seff , and the g 4 term in (69) could come entirely from Kijkl . A similar conclusion is reached
in [33,34] where it is shown that the beta function of the heterotic string in the presence of back-
ground gauge fields has a term at three loops that behaves as F 3 , even though no corresponding
term is present in the effective action.
Finally, it is possible to make a statement about the three-loop structure of the beta function
in the α  expansion. Excluding the Ricci tensor and the Ricci scalar, since the beta function is
computed within the minimal subtraction scheme, the terms that are third order in the Riemann
tensor and are compatible with the g 4 terms in (69) are given by
 
1 1
α  3 Rklmn Ri mlr Rj k lr − Riklj R kmnr R l mnr . (73)
8 16

4.4. Singularities of the critical exponents; central charge of the CFT

Because λ involves products of Γ functions it is natural to investigate the location of its


1 = 2 λ1 , the location of the pole of
1 bos
singularities closest to the origin, as in [1]. Because λhet
λ1 coincides with the pole in the bosonic case, with half the residue. One also has to note that
η1 (μ) behaves as η1 ∼ − π42 2μ−1 1
i.e. it has a simple pole at  = −1. But λ1 ’s first singularity is
at  = −3, since the pole of η1 is canceled by a similar pole of χ1 . Examining term by term the
structure of λ2 it is easy to see that the singularities of λ2 come from the η12 factor that multiplies
the whole expression (66) and from the three-loop diagrams that have the Lagrange multiplier
field propagator corrected, i.e. Π2σ , Π3u , F2u , and F2σ . Since
−1 1
R3 (μ) ∼ , R2 (μ) ∼ (74)
2μ − 1 (2μ − 1)2
and λ2 has terms that behave as ∼ R32 η12 and R2 η12 times a μ polynomial with no zero at μ = 1/2,
we see that it has a fourth order pole. In all, one finds
3/(4π 2 ) log(3 + κ)
λ1 = − + O(1), b1 = − + O(1), (75)
+3 2π 2
G. Michalogiorgakis, S.S. Gubser / Nuclear Physics B 757 (2006) 146–171 163

8/π 4   16/3π 4  
λ2 = + O ( + 1)−3 , b2 = − + O (κ + 1)−2 . (76)
( + 1) 4 (κ + 1)3

The 1/(κ + 1)2 term in b2 comes only from the factors R3 η12 and from the Hartree–Fock dia-
grams. The singularities in the heterotic case are at the same locations and of the same order as
in the bosonic case.
Because of the sign of b2 , there is clearly a zero of β(g) (computed through order 1/D 2 ) for
negative g, close to κ = −1. The same caveats discussed in [1] apply: Higher order terms in
1/D could conceivably cause this zero to disappear or move significantly. In Section 5 we will
comment further on higher-order corrections. For the remainder of this section we will assume
that the computation of β(g) that we have carried out is precise enough to describe the zero
correctly.
The zero of β(g) arises through competition between the one-loop term (corresponding to
Einstein gravity) and b2 (corresponding to a combination of all α  corrections to Einstein grav-
ity). Because the geometry has string scale curvatures (more precisely, L2 ∼ Dα  ) there is no
reason to think that the worldsheet central charges are particularly close to the flat-space results.
Fortunately, one can calculate the central charges using Zamolodchikov’s c-theorem:
∂c 3(D + 1)
= β(g). (77)
∂g 2g 2
The result (77) holds for both the holomorphic and the anti-holomorphic sides: c and c̃ differ
by a constant. To derive the prefactor on the right-hand side of (77), one can consider two-point
functions of the graviton perturbation Oij = 2πα
1 ¯
 ∂Xi ∂Xj + 4π Ψ ∂Ψ around flat space, as is done
1

in [1]. This prefactor receives higher loop corrections, and knowing Kijkl in higher loops, one
3

can in principal compute them. As in the bosonic and supersymmetric cases, the results suggest
that with increasing D the critical point moves closer to κ = −1: Integrating (77) leads to
κc    
3(D + 1) 1 b1 (κ) b2 (κ) 3
c = (D + 1) + dκ −κ + + ≈ (D + 1) 1 − κc ,
2 κ D D2 2
0
κc  
3 3(D + 1) 1 b1 (κ) b2 (κ) 3
c̃ = (D + 1) + dκ −κ + + 2
≈ (D + 1)(1 − κc ), (79)
2 2 κ D D 2
0
where we have noted that the central charge of the holomorphic side in flat space is c = D + 1,
while for the anti-holomorphic side it is c̃ = 32 (D + 1). The approximate equalities arise from
dropping the b1 (κ) and b2 (κ) terms from the integrand: Their only role at this level of approxi-
mation is to set κc . As κc gets closer to −1 (i.e. as D becomes large), the central charges converge
to
5
c = (D + 1), c̃ = 3(D + 1). (80)
2

3 Another way to derive (77) is to use the relation of the central charge to the spacetime effective action. At least up to

two-loop order, the effective action at the fixed point is equal to −c/2κD2 α  [35–37]. Also up to two loops, the K kl of
ij
(72) is simply given by a product of Kronecker δ’s, as in the bosonic case [32]. Using the fact that in symmetric spaces
∂ ∂
β(g)gij = −gβij , g ij =g , (78)
∂g ij ∂g
one indeed ends up with (77).
164 G. Michalogiorgakis, S.S. Gubser / Nuclear Physics B 757 (2006) 146–171

The result (80) for c is the same as in the bosonic case, while for c̃ it is the same as the type II case
[1]. As in [1], (80) appears to set only an approximate upper bound on the central charges. The
dominant error in the calculation (79) is from the uncertainty in the prefactor in (77). Analogous
to the speculations in [1], it is conceivable that the expressions (80) might in fact be exact. But
this would require a significant conspiracy between the prefactor in (77) and the beta function.
The fact that the location of the critical point at finite D is so close to the singularity of λ
means that the critical exponent λ evaluated at the critical point is large and positive. This leads
to an operator with a large and negative dimension, which appears to violate unitarity. However,
one could hope that a consistent GSO projection would project this operator out of the spectrum.

5. Discussion

The existence of the AdSD+1 critical point depends on competition between one-loop and
1/D 2 effects. It would therefore be instructive to compute the beta function through order 1/D 3
and see whether the fixed point persists. Given that the number of diagrams needed for the com-
putation at the next order grows significantly, the shortest path seems to be calculating χ3 and
using (60) to deduce λ3 . However, note that for the calculation of χ3 one needs to derive the
residues of diagrams at order 1/D 4 , which include some six-loop diagrams.
There is some reason to think that the singularities of λ at order 1/D 3 are no worse than at
order 1/D 2 : Examining the diagrams needed for the Dyson equations of the Lagrange multiplier
fields, we see that at order 1/D 3 these come from either inserting a σ or u propagator in the
1/D 2 diagrams or inserting a loop of S or Ψ in the middle of the diagram. The computation for
the diagrams that come from inserting a σ or u propagator can easily be seen to be reduced to
the sum of diagrams similar to Π2 or Π3 with one different exponent. A naive calculation does
not produce any worse singularities than the ones already contained in Π2 and Π3 . However, one
also has to compute the more difficult diagrams with the additional S or Ψ loop.
It is evident that the methods of [8,15], has many advantages over calculating Feynman
diagrams in momentum space. In the latter approach one encounters difficulties already in cal-
culating second order diagrams, since the propagators of the Lagrange multiplier fields are in
general hypergeometric functions. It is noteworthy that even though we start from d = 2 +  di-
mensions, one can calculate critical exponents of the O(N) model in any dimension 2 < d < 4,
and there is agreement with the results in three dimensions [38] in the bosonic case.
Perhaps the methods of [8,15] could be applied to a related quantum field theory:
  
1 1 λN 2
S=    
d x ∇ Φ · ∇ Φ + λσ Φ · Φ −
d
σ , (81)
2 2 4
which for d = 3 is the proposed dual of an AdS4 vacuum of a theory with arbitrarily high spin
gauge fields [39]. What makes (81) susceptible to a position-space treatment analogous to those
in [8,15] is that only cubic vertices are involved. It would be interesting to compute, for example,
the four point function of σ to order 1/N 2 and compare it to the corresponding AdS4 calculation,
as is done for example in [40] at order 1/N .

Acknowledgements

G.M. would like to thank J. Friess for valuable conversations.


G. Michalogiorgakis, S.S. Gubser / Nuclear Physics B 757 (2006) 146–171 165

Appendix A. Anomalies

Since we have coupled only the right moving fermions to gravity it is natural to investigate
whether there are anomalies. These are related to a breakdown of general coordinate invariance
or local Lorentz invariance. We will investigate only the latter, as is usually done [41]. Indeed
when there is a coordinate anomaly one can add a counterterm to the action and convert it to a
Lorentz anomaly [42]. It is convenient to use the tetrad formalism. Local Lorentz transformations
in this formalism are given by
 p
eμ (x) = eμ q (x)Θ p q (x). (A.1)
p
The Riemann tensor can be written Rμν q , with mixed spacetime and tangent space indices,
and can be regarded as a two form R2 . We only have to worry about the massless fields of the
supergravity sector [41]. The anomaly polynomials for the spinor and the gravitino contain only
terms that are proportional to polynomials in Tr R2∧2m . In the AdS space that we are interested in
we can calculate
1  
Rμν a b Rκλ c a = 2 δ c k Rμνλb + δ c λ Rμνbκ , (A.2)
L
which when anti-symmetrizing to get the wedge product returns zero. So R2 ∧ R2 = 0 in our
case, and we do not have to worry about the gravitational anomalies. Another way to view this is
to say that the field strength H3 = dB2 obeys the modified Bianchi identity
1
dH3 = (Tr R2 ∧ R2 − Tr F2 ∧ F2 ). (A.3)

A three form H3 obeying (A.3) is required for cancelation of perturbative heterotic string world-
sheet anomalies, as is briefly reviewed in [43]. Because Tr R2 ∧ R2 = 0, this is trivially satisfied.

Appendix B. Position space methods for calculating graphs

There are 11 diagrams needed for the calculation of η at 1/D 2 . We designate them by Σ , Π ,
Φ, F . The way to compute them was developed in [8,15] for the bosonic graphs and extended
to include fermionic graphs in [16,17]. The main advantage of the method is that there is no
need to explicitly evaluate any Feynman diagram. Here we will only give a few key observa-
tions that facilitate the evaluation of the diagrams. The first observation is that the chain of two
propagators is equalto a propagator times a prefactor. Graphically this is shown in Fig. 6, where
= π μ 3i=1 α(xi ). The third exponent x3
ν(x1 , x2 , x3 )  is determined by the “uniqueness” [15]
requirement i xi = 2μ, for the bosonic graphs and i xi = 2μ − 1 if there are one or more
fermion lines in the graph. An identity exists for a three point vertex (see Fig. 7), which
 is simi-
larly related to a “unique” triangle, where now the uniqueness  requirement is that i xi = μ. If
there are one or more fermion lines the uniqueness changes to i xi = μ + 1, and the results of
[16] are unchanged in our case. There is no similar identity for a four point function, and for the
(1, 1) supersymmetric model that means that one has to retain the auxiliary field F . In computing
the values at order 1/D 2 , χ is set to zero [15]. For a non-zero Δ the diagrams, for example the
self energy of Ψ designated A, lose their uniqueness. However one can subtract from A a graph
B that has the same divergent substructure as A, but can be calculated for an arbitrary Δ. So one
has to compute (A − B) + B. Since B can be calculated for arbitrary Δ and contains the diver-
gence, one can evaluate (A − B) at zero Δ when both diagrams become unique. A valuable first
166 G. Michalogiorgakis, S.S. Gubser / Nuclear Physics B 757 (2006) 146–171

Fig. 6. Products of two propagators are related to a single propagator. So, in a two-loop diagram (Σ1 for example)
inserting a point in one of the three propagators that connect to one internal vertex can make this vertex unique. The
dotted lines denote fermions. Since we are dealing with chiral fermions, taking the trace in the third graph only produces

one half the full result. ν is equal to ν(x1 , x2 , x3 ) = π 3i=1 α(xi ).

Fig. 7. An identity that allows the integration of a unique vertex. Only bosonic propagators are shown. Similar identities
with fermions can which are used in our calculation can be found in [20].

step in the calculation is the evaluation of all the self-energy graphs, which we will not include
here since it was done in detail in [15,17]. We just note that the most basic tool is the insertion
of a point facilitated by the fact that we can write a propagator as a product of two. Then one
can choose one of the exponents in such a way that the vertex that the propagator is attached to,
becomes unique.

Appendix C. Calculation of the graphs needed for η2

We give the results for the various graphs occurring in the 1/D 2 calculation. We only give
the simple pole term and the constant term in an expansion in Δ. The purely bosonic graphs
were calculated in [15]. Compared to the calculation of the fermionic graphs in [20], there are
differences that have to do with taking the trace of fermion loops, i.e. some factors of two in
bosonic diagrams with fermion loops. Otherwise the calculation is almost identical. We find
small discrepancies with [20] in some of the diagrams, mostly factors of 2 and some minus
signs. The most notable difference is Φ1 , where the residue has a different denominator. We
believe that our value is correct, since it leads to the same χ1 as the evaluations from the other
Eqs. (58).

B(x) = ψ(x) + ψ(μ − x), (C.1)


 
2μ 2
2π α (ΔS )α(Δσ ) Δ 
Σ1 = 1+ B(Δσ ) − B(ΔS ) , (C.2)
Δ(μ) 2


2π 2μ α 2 (ΔS )α(Δσ − 1) Δ 1 1
Σ2 = 1+ B(Δσ − 1) − B(ΔS ) + − , (C.3)
Δα(Δσ − 1)(μ) 2 Δσ − 1 ΔS
G. Michalogiorgakis, S.S. Gubser / Nuclear Physics B 757 (2006) 146–171 167

 
2π 4μ α 3 (ΔS )α 3 (Δσ )α(μ + ΔS − Δσ ) 1  
Σ3 = + Δ B(Δσ ) − B(ΔS ) , (C.4)
Δ(μ) 2
π 4μ α 3 (ΔS )α 2 (Δσ − 1)α(Δσ )α(ΔS + μ − Δσ )
Σ4 =
ΔΔS (ΔS + μ − Δσ )(Δσ − 1)2 (μ)


Δ 3 2
× 1+ B(Δσ ) + 3B(Δσ − 1) − 4B(ΔS ) + − , (C.5)
2 Δσ − 1 ΔS
2π 2μ α 2 (ΔS )α(Δσ )   
Π1 = 1 + Δ B(Δσ ) − B(ΔS ) , (C.6)
Δ(μ)
π 4μ α 3 (ΔS )α 3 (Δσ )α(ΔS + μ − Δσ )
Π2 =
Δ(μ)
  
× 1 + Δ 4B(Δσ ) − 3B(ΔS ) − B(μ + ΔS − Δσ ) , (C.7)
− 1)α(ΔS + μ − Δσ )
π 4μ α 3 (ΔS )α 2 (Δσ
Π3 =
2Δα(μ − Δσ )ΔS (ΔS + μ − Δσ )(Δσ − 1)2 (μ)

× 1 + Δ 2B(Δσ ) − 3B(ΔS ) + 2B(Δσ − 1) − B(ΔS + μ − Δσ )


1 2 1
− + − , (C.8)
ΔS Δσ − 1 ΔS − Δσ + μ


π 2μ α 2 (ΔS − 1)α(Δσ ) 1
Φ1 = − 1 + Δ B(Δσ ) − B(ΔS − 1) − , (C.9)
ΔΔS (ΔS − 1)(μ) ΔS − 1
π 4μ α 3 (ΔS )α 2 (Δσ − 1)α(Δσ )α(ΔS + μ − Δσ )
Φ2 = −
ΔΔS (ΔS + μ − Δσ )(Δσ − 1)2 (μ)


1
× 1 + Δ B(Δσ ) − 2B(ΔS ) + B(Δσ − 1) + , (C.10)
Δσ − 1
2π 2μ α 2 (ΔS )α(Δσ − 1)
F1 = −
ΔΔS (Δσ − 1)(μ)


1 1
× 1 + Δ B(Δσ − 1) − B(ΔS ) − + , (C.11)
2ΔS Δσ − 1
π 4μ α 3 (ΔS )α(Δσ )α 2 (Δσ − 1)α(ΔS + μ − Δσ )
F2 = −
ΔΔS (ΔS + μ − Δσ )(Δσ − 1)2 (μ)

× 1 + Δ B(Δσ ) + 3B(Δσ − 1) − B(ΔS + μ − Δσ ) − 3B(ΔS )


1 3 1
− + − . (C.12)
ΔS Δσ − 1 ΔS + μ − Δσ
Note that there is a similar three-loop diagram with Π3 with the role of the Ψ , u propagators
interchanged. But it scales as 1/N 3 . A very useful identity for the evaluation of various quantities
given in the text is B(x) = B(μ − x).

Appendix D. Calculation of the graphs need for λ2

In this section we give the formal expressions for the sums of the corrected diagrams (D.2)–
(D.11), the values of the 43 individual diagrams that contribute (D.12)–(D.30), and finally the
168 G. Michalogiorgakis, S.S. Gubser / Nuclear Physics B 757 (2006) 146–171

explicit form of the sums (D.31)–(D.39). Beforehand one has to define the functions appearing
as

R1 = ψ  (μ − 1) − ψ  (μ),
R2 = ψ  (2μ − 3) − ψ  (2 − μ) − ψ  (μ − 1) + ψ  (1),
R3 = ψ(2μ − 3) + ψ(2 − μ) − ψ(μ − 1) − ψ(1), (D.1)
where ψ(z) = d log (z)/dz.

ΣS = 2w 2 Σ1Sa + w 2 Σ1Sb − v 2 Σ2S + N w 3 (2Σ3Sa + 2Σ3Sb + Σ3Sc )


− 2Nv 2 w(Σ4Sa + Σ4Sb + Σ4Sc ), (D.2)
ΣΨ = −2v Σ2Ψ − 2N v w(Σ4Ψ a + Σ4Ψ b ),
2 2
(D.3)
Σσ = 2w 2 Σ1σ + N w 3 (2Σ3σ a + Σ3σ b ) − 2N v 2 wΣ4σ , (D.4)
Σu = −2v Σ2u − 2N v w(Σ4ua + Σ4ub ),
2 2
(D.5)
ΦΨ = v Φ1Ψ + N v wΦ2Ψ ,
2 2
Φσ = N v wΦ2σ ,
2
(D.6)
ΦS = 2v Φ1S + 2N v w(Φ1Sa + Φ2Sb ),
2 2
Φu = 2v Φ1u + N v wΦ2u ,
2 2
(D.7)
ΠS = 4w Π1S + N w (4Π2Sa + 2Π2Sb ) − 4N wv Π3S ,
2 3 2
(D.8)
Πσ = w Π1σ + 2N w Π2σ ,
2 3
ΠΨ = −2N wv Π3Ψ , 2
Πu = −2N wv Π3u ,
2
(D.9)
FΨ = 2v F1Ψ + 2N v w2F2Ψ ,
2 2
FS = 2v F1S + 2N v w(F2Sa + F2Sb ),
2 2
(D.10)
Fu = v F1u + N v wF2u ,
2 2
Fσ = N v wF2σ .
2
(D.11)
There are 19 diagrams associated with the S-propagator and eight for the propagator of the other
fields. The value of each graph is given below for completeness. The bosonic ones were calcu-
lated in [15], while similar to the fermionic ones were done in [19].
π 2μ π 2μ
Σ1Sa = , Σ 1Sb = , (D.12)
(μ − 2) 2 (μ) (μ − 2)2  2 (μ)
2(μ + 1)π 2μ (μ2 − 3μ + 1)(1 − μ)π 4μ
Σ2S = − , Σ3Sa = , (D.13)
μ(μ − 1) (μ)
2 (μ − 2)3 (μ)(2μ − 3)
 
π 4μ (2 − μ) 2μ − 3
Σ3Sb = 3R1 + , (D.14)
(2 − μ)(μ − 1)(2μ − 2) (μ − 2)2
π 4μ (4 − μ) 2π 4μ (1 − μ)
Σ3Sc = , Σ 4Sa = , (D.15)
(μ − 2)3 (μ − 1)(2μ − 4) (μ)(2μ − 2)
 
π 4μ (2 − μ) 2μ − 3
Σ4Sb = − 3R1 + , (D.16)
(μ)(2μ − 2) (μ − 1)(μ − 2)
2(μ − 3)(2μ − 3)(1 − μ)π 4μ π 2μ
Σ4Sc = , Σ2Ψ = = −Φ2S , (D.17)
(2 − μ)(2μ − 2)(μ) (μ − 1) 2 (μ)
π 4μ (1 − μ)(2μ2 − 5μ + 1)
Σ4Ψ a = = −Φ2Sa ,
2(μ − 2)(μ)(2μ − 2)
π 4μ 3(μ − 3)(2 − μ)R1
Σ4Ψ b = = −Φ2Sb , (D.18)
2(2 − μ)(μ)(2μ − 2)
G. Michalogiorgakis, S.S. Gubser / Nuclear Physics B 757 (2006) 146–171 169

π 2μ (μ2 − 3μ + 1)
Σ1σ = = Π1S ,
(μ − 2)2  2 (μ)
π 4μ (2μ2 − 7μ + 4)(1 − μ)
Σ3σ a = = Π2Sa , (D.19)
(μ − 2)3 (μ)(2μ − 3)
3π 4μ (3 − μ)R1
Σ3σ b = = Π2Sb ,
(2 − μ)3 (μ − 1)(2μ − 2)
π 4μ (2μ − 5)(1 − μ)
Σ4σ = = Π3S , (D.20)
(μ − 2)(μ)(2μ − 2)
π 2μ π 4μ (4μ − 9)(1 − μ)
Σ2u = , Σ4ua = = −F2sa , (D.21)
 2 (μ) 2(μ − 2)(μ)(2μ − 2)
3π 4μ (2 − μ)R1
Σ4ub = = −F2Sb , (D.22)
2(μ − 2)(μ)(2μ − 2)
π 2μ π 2μ μ
Φ1S = − , Φ1u = , (D.23)
(μ − 1) 2 (μ) (1 − μ) 2 (μ)
π 4μ (1 − μ)(4μ2 − 11m + 5) 3π 2μ R1
Φ2u = , Π1σ = , (D.24)
2(1 − μ)(μ − 2)(μ)(2μ − 2) (2 − μ)(2μ − 3) 2 (μ − 1)
π 4μ (2 − μ)
Π2σ =
2(2 − μ)3 (μ − 1)(2μ − 1)
  
2(μ − 2) 1
× 6R1 − R2 − R32 + R3 − , (D.25)
(μ − 1)(3μ − 2) μ−2
(2 − μ)π 4μ
Π3u = −
4(μ − 2)2 (μ)(2μ − 2)
 
2R3 (μ − 2) − 2
× 6R1 − R2 − R3 +
2 2
= −F2σ , (D.26)
(μ − 1)(2μ − 3)
3(2 − μ)π 4μ R1 μπ 2μ
Π3Ψ = = Φ2σ , F1Ψ = , (D.27)
(2μ − 3)(μ)(2μ − 1) (1 − μ) 2 (μ)
π 4μ (4μ2 − 11μ + 5)(1 − μ)
F2Ψ = , (D.28)
2(1 − μ)(μ − 2)(μ)(2μ − 2)
π 2μ 3(μ − 1)π 2μ
F1S = − , F1u = , (D.29)
 2 (μ) 2(μ − 2) 2 (μ)
π 4μ (2 − μ)
F2u =
4(μ − 1)2 (μ − 2)2 (μ)(2μ − 2)
 
  2(μ − 2)R3 2 4
× (μ − 1) 6R1 − R2 − R32 + + − . (D.30)
2μ − 3 2μ − 3 μ − 1

D.1. Summing up graphs

In this subsection we give the explicit values for the various sums appearing in the λ2 calcu-
lation. We have omitted the 1/N 2 factor that multiplies all of the diagrams so that the expression
170 G. Michalogiorgakis, S.S. Gubser / Nuclear Physics B 757 (2006) 146–171

(66) for λ̃2 , does not contain any factors of N .


ΣS μ+2 2  
= −1 + + + 2μ(2μ − 5) + 2μ(2μ − 3)(μ − 3) 2 − (μ − 2)2
η1 v 1 μ − 1 (μ − 1) 2

− 4μ(μ − 2) − 6μ(μ − 2)R1 , (D.31)


 2 
μ (μ − 2)(2μ − 3)
ΣΨ = −η1 v1 + 3(μ − 1)(μ − 3)R1 = ΦS , (D.32)
μ−1
 
μ(3 + μ(3μ − 7)) ΠS
Σσ = −η1 v1 + 3μ(μ − 2)R 1 = , (D.33)
(μ − 1)2 2
 
Σu = η1 v1 μ(8 − 4μ) + 3(μ − 1)R1 = FS , (D.34)
3μ(μ − 1)(μ − 2)R1
Φσ = η1 v 1 = (ΠΨ )/2, (D.35)
2μ − 3
Φu 1 FΨ
= 1 + 6μ − 4μ2 − =− , (D.36)
η1 v 1 μ−1 η v
 1 1

Πσ 3μ(μ − 2) μ 2((μ − 2)R3 − 1)
= R1 − 6R1 − R2 − R32 + , (D.37)
η1 v 1 2(2μ − 3) μ−1 (2μ − 3)(μ − 1)
 
Πu μ(μ − 1) (μ − 2)R3 − 1 2Fσ
= 6R1 − R2 − R32 + 2 =− , (D.38)
η1 v 1 μ−2 (2μ − 3)(μ − 1) η1 v 1

Fu 3 μ(μ − 1) μ 4 2
= + 6R1 − R2 + R32 − +
η1 v 1 4 μ − 2 4 (μ − 1)2 (2μ − 3)(μ − 1)

2(μ − 2)
+ R3 . (D.39)
2μ − 3

References

[1] J.J. Friess, S.S. Gubser, Non-linear sigma models with anti-de Sitter target spaces, Nucl. Phys. B 750 (2006) 111–
141, hep-th/0512355.
[2] J.M. Maldacena, The large N limit of superconformal field theories and supergravity, Adv. Theor. Math. Phys. 2
(1998) 231–252, hep-th/9711200.
[3] S.S. Gubser, I.R. Klebanov, A.M. Polyakov, Gauge theory correlators from non-critical string theory, Phys. Lett.
B 428 (1998) 105–114, hep-th/9802109.
[4] E. Witten, Anti-de Sitter space and holography, Adv. Theor. Math. Phys. 2 (1998) 253–291, hep-th/9802150.
[5] O. Aharony, S.S. Gubser, J.M. Maldacena, H. Ooguri, Y. Oz, Large N field theories, string theory and gravity, Phys.
Rep. 323 (2000) 183–386, hep-th/9905111.
[6] A. Sen, The heterotic string in arbitrary background field, Phys. Rev. D 32 (1985) 2102.
[7] C.M. Hull, E. Witten, Supersymmetric sigma models and the heterotic string, Phys. Lett. B 160 (1985) 398–402.
[8] A.N. Vasiliev, Y.M. Pismak, Y.R. Khonkonen, 1/N expansion: Calculation of the exponents η and ν in the order
1/N 2 for arbitrary number of dimensions, Theor. Math. Phys. 47 (1981) 465–475.
[9] A.M. Polyakov, Interaction of Goldstone particles in two dimensions. Applications to ferromagnets and massive
Yang–Mills fields, Phys. Lett. B 59 (1975) 79–81.
[10] E. Witten, A supersymmetric form of the nonlinear sigma model in two dimensions, Phys. Rev. D 16 (1977) 2991.
[11] O. Alvarez, Dynamical symmetry breakdown in the supersymmetric nonlinear sigma model, Phys. Rev. D 17 (1978)
1123.
[12] I. Aref’eva, V. Krivoshchekov, P. Medvedev, 1/N perturbation theory and quantum conservation laws for the su-
persymmetric chiral field, Theor. Math. Phys. 40 (1980) 565.
[13] D.J. Gross, A. Neveu, Dynamical symmetry breaking in asymptotically free field theories, Phys. Rev. D 10 (1974)
3235.
[14] A. Zamolodchikov, A. Zamolodchikov, Relativistic S-matrix in two dimensions having O(N ) isotopic symmetry,
Nucl. Phys. B 133 (1978) 525, also JETP 26 (1977) 457.
G. Michalogiorgakis, S.S. Gubser / Nuclear Physics B 757 (2006) 146–171 171

[15] A.N. Vasiliev, Y.M. Pismak, Y.R. Khonkonen, Simple method of calculating the critical indices in the 1/N expan-
sion, Theor. Math. Phys. 46 (1981) 104–113.
[16] J.A. Gracey, On the beta function for sigma models with N = 1 supersymmetry, Phys. Lett. B 246 (1990) 114–118.
[17] J.A. Gracey, Critical exponents for the supersymmetric sigma model, J. Phys. A 23 (1990) 2183–2194.
[18] S.-k. Ma, Critical exponents above Tc to O(1/n), Phys. Rev. A 7 (1973) 2172.
[19] J.A. Gracey, Probing the supersymmetric O(N ) sigma model to O(1/N 2 ): Critical exponent ν, Nucl. Phys. B 352
(1991) 183–214.
[20] J.A. Gracey, Probing the supersymmetric O(N ) sigma model to O(1/N 2 ): Critical exponent η, Nucl. Phys. B 348
(1991) 737–756.
[21] A.P. Foakes, N. Mohammedi, D.A. Ross, Three loop beta functions for the superstring and heterotic string, Nucl.
Phys. B 310 (1988) 335.
[22] A.P. Foakes, N. Mohammedi, D.A. Ross, Effective action and beta functions for the heterotic string, Phys. Lett.
B 206 (1988) 57.
[23] M.T. Grisaru, A.E.M. van de Ven, D. Zanon, Four loop divergences for the N = 1 supersymmetric nonlinear sigma
models in two-dimensions, Nucl. Phys. B 277 (1986) 409.
[24] M.T. Grisaru, A.E.M. van de Ven, D. Zanon, Four loop β function for the N = 1 and N = 2 supersymmetric
nonlinear sigma model in two dimensions, Phys. Lett. B 173 (1986) 423.
[25] F. Wegner, Four loop order beta function of nonlinear sigma models in symmetric spaces, Nucl. Phys. B 316 (1989)
663–678.
[26] I. Jack, D.R.T. Jones, N. Mohammedi, The four loop metric beta function for the bosonic sigma model, Phys. Lett.
B 220 (1989) 171.
[27] I. Jack, D.R.T. Jones, N. Mohammedi, A four loop calculation of the metric beta function for the bosonic sigma
model and the string effective action, Nucl. Phys. B 322 (1989) 431.
[28] D.J. Gross, J.H. Sloan, The quartic effective action for the heterotic string, Nucl. Phys. B 291 (1987) 41.
[29] R.R. Metsaev, A.A. Tseytlin, Curvature cubed terms in string theory effective actions, Phys. Lett. B 185 (1987) 52.
[30] Y. Cai, C.A. Nunez, Heterotic string covariant amplitudes and low-energy effective action, Nucl. Phys. B 287 (1987)
279.
[31] A.A. Tseytlin, Vector field effective action in the open superstring theory, Nucl. Phys. B 276 (1986) 391.
[32] I. Jack, D.R.T. Jones, D.A. Ross, On the relationship between string low-energy effective actions and O(α  3 ) sigma
model beta functions, Nucl. Phys. B 307 (1988) 130.
[33] U. Ellwanger, J. Fuchs, M.G. Schmidt, The heterotic sigma model with background gauge fields, Nucl. Phys. B 314
(1989) 175.
[34] U. Ellwanger, J. Fuchs, M.G. Schmidt, The heterotic sigma model in the presence of background gauge fields: Three
loop results, Phys. Lett. B 203 (1988) 244.
[35] H. Osborn, String theory effective actions from bosonic sigma models, Nucl. Phys. B 308 (1988) 629.
[36] H. Osborn, General bosonic sigma models and string effective actions, Ann. Phys. 200 (1990) 1.
[37] S.P. de Alwis, Strings in background fields beta functions and vertex operators, Phys. Rev. D 34 (1986) 3760.
[38] Y. Okabe, M. Oku, 1/N expansion up to order 1/N 2 . 3. Critical exponents gamma and nu for D = 3, Prog. Theor.
Phys. 60 (1978) 1287–1297.
[39] I.R. Klebanov, A.M. Polyakov, AdS dual of the critical O(N ) vector model, Phys. Lett. B 550 (2002) 213–219,
hep-th/0210114.
[40] T. Leonhardt, A. Meziane, W. Ruhl, On the proposed AdS dual of the critical O(N ) sigma model for any dimension
2 < d < 4, Phys. Lett. B 555 (2003) 271–278, hep-th/0211092.
[41] M.B. Green, J.H. Schwarz, Anomaly cancelation in supersymmetric D = 10 gauge theory and superstring theory,
Phys. Lett. B 149 (1984) 117–122.
[42] W.A. Bardeen, B. Zumino, Consistent and covariant anomalies in gauge and gravitational theories, Nucl. Phys.
B 244 (1984) 421.
[43] E. Witten, World-sheet corrections via D-instantons, JHEP 0002 (2000) 030, hep-th/9907041.
Nuclear Physics B 757 (2006) 172–196

Electroweak phase transition and baryogenesis


in the nMSSM
Stephan J. Huber a , Thomas Konstandin b,∗ , Tomislav Prokopec c ,
Michael G. Schmidt d
a Theory Division, CERN, CH-1211 Geneva 23, Switzerland
b Department of Physics, Royal Institute of Technology (KTH), AlbaNova University Center,
Roslagstullsbacken 11, 106 91 Stockholm, Sweden
c Institute for Theoretical Physics (ITF) & Spinoza Institute, Utrecht University,
Leuvenlaan 4, Postbus 80.195, 3508 TD Utrecht, The Netherlands
d Institut für Theoretische Physik, Heidelberg University, Philosophenweg 16, D-69120 Heidelberg, Germany

Received 1 August 2006; accepted 1 September 2006

Available online 22 September 2006

Abstract
We analyze the nMSSM with CP violation in the singlet sector. We study the static and dynamical proper-
ties of the electroweak phase transition. We conclude that electroweak baryogenesis in this model is generic
in the sense that if the present limits on the mass spectrum are applied, no severe additional tuning is re-
quired to obtain a strong first-order phase transition and to generate a sufficient baryon asymmetry. For this
we determine the shape of the nucleating bubbles, including the profiles of CP-violating phases. The baryon
asymmetry is calculated using the advanced transport theory to first and second order in gradient expansion
presented recently. Still, first and second generation sfermions must be heavy to avoid large electric dipole
moments.
© 2006 Elsevier B.V. All rights reserved.

PACS: 98.80.Cq; 11.30.Er; 11.30.Fs

* Corresponding author.
E-mail addresses: stephan.huber@cern.ch (S.J. Huber), konstand@theophys.kth.se, konstand@kth.de
(T. Konstandin), t.prokopec@phys.uu.nl (T. Prokopec), m.g.schmidt@thphys.uni-heidelberg.de (M.G. Schmidt).

0550-3213/$ – see front matter © 2006 Elsevier B.V. All rights reserved.
doi:10.1016/j.nuclphysb.2006.09.003
S.J. Huber et al. / Nuclear Physics B 757 (2006) 172–196 173

1. Introduction

Lately, electroweak baryogenesis (EWBG) [1] has again attracted more attention, not least be-
cause new collider data will hopefully provide information about the relevant (supersymmetric?)
physical parameters in the next years.
Electroweak baryogenesis relies on a strong first-order electroweak phase transition as the
source of out-of-equilibrium effects. During a first-order phase transition, bubbles of the low-
temperature (broken) phase nucleate and expand to fill all space. An important aspect in the
determination of the baryon asymmetry is the impact of transport phenomena [2]. Without trans-
port, CP-violating currents would only be generated near the bubble wall profile of the Higgs
vevs. Close to the wall, sphaleron transitions are already strongly suppressed by the mass of
the W-bosons, so that B-violating processes are inefficient in producing the observed baryon
asymmetry (BAU). Early works that incorporated transport effects into the EWBG calculus were
based on the WKB approach [3]. In this framework, a CP-violating shift in the dispersion relation
induces a force of second order in the gradient expansion in the Boltzmann equation and leads
to CP-violating fermion densities in the symmetric phase. Later on, this formalism was applied
to the MSSM, where CP violation results from mixing effects in the chargino sector [4]. In this
context, the formalism had to be extended to the case of mixing fermions. In the MSSM, the
second-order source is too weak to yield a successful baryogenesis [5,6].
One disadvantage of the WKB method is the neglect of dynamical flavor mixing effects. While
the shift in the dispersion relation is due to mixing of left- and right-handed components of the
fermions and already present in the one-flavor case, flavor mixing contributions have been com-
pletely neglected after a flavor basis transformation to the mass eigenbasis. A series of papers
[7,8] aimed at improving on this point by including flavor mixing by using a perturbative ex-
pansion of the Kadanoff–Baym equations. Here, the deviations of the Green function have been
interpreted as sources in the diffusion equation. This approach—like the WKB-approach—has
the weakness that the transport equations only describe the dynamics of two classical quasi-
particles in the chargino sector. CP violation is communicated from the charginos to the SM
particles by their interactions. Therefore, the authors of Refs. [7,8] used the winos and higgsinos
as quasi-particles in the interaction basis, where the interactions take a particularly simple form.
In the WKB-approach the natural choice is the mass eigenbasis. This dependence on a flavor ba-
sis is unsatisfactory, especially since the flavor mixing CP-violating source vanishes in the mass
eigenbasis completely. A numerical analysis, making use of the first-order source in the interac-
tion basis in Refs. [7,8], leads to successful electroweak baryogenesis for a certain range in the
parameter space, even though at least some fine-tuning is required to fulfill the electron electric
dipole moment (EDM) constraints.
Recently, some of the authors have derived semiclassical transport equations for the chargino
sector from first principles [5,9]. The derivation is based on the Kadanoff–Baym equations and
does not depend on classical reasoning to fill the gap between CP violation and transport effects.
Technically, the two main improvements on the resulting transport equations are independence of
the flavor basis and the absence of the source strength ambiguities. First, since the semiclassical
transport equations describe the dynamics of a 2 × 2 matrix in flavor space, the transformation
properties of the transport equations under flavor basis changes are explicit, and no restriction to
quasi-particles has to be used. Secondly, because the CP-violating sources appear naturally and
uniquely as higher order terms in the gradient expansion of the mass background fields, there
are no ambiguities in the source strength. The mixing effects lead to an additional force of first
order of the gradient expansion in the transport equations. In contrast, in the work reported on
174 S.J. Huber et al. / Nuclear Physics B 757 (2006) 172–196

in [8], for dimensional reasons, the sources had to be multiplied by a typical thermalization time
τ , while in [7] the removal of the ambiguity in the source strength was based on the (classical)
Fick’s law. In the case of second-order effects, the first principle derivation confirms the WKB
approach if the latter is handled carefully [10,11].
Applying this advanced transport theory to the MSSM [6], the quantitative analysis shows
two distinct features, which are less definitive in the results of the former approach [7,8]:
First, mixing effects are strongly suppressed away from mass degeneracy in the chargino sec-
tor, |m2χ̃ ± − m2χ̃ ± | > (20 GeV)2 . So mixing in that sector is only effective if the a priori unrelated
1 2
wino mass parameter M2 and the higgsino mass parameter μ are tied together. Secondly, the pro-
duced BAU suffers from an exponential Boltzmann suppression in the case of heavy charginos.
In this formalism, successful electroweak baryogenesis requires rather large CP violation in the
chargino sector, sin(δCP ) > 0.25, even for the most favorable choice of the other model parame-
ters. In comparison, the approach followed in Refs. [7,8] leads to viable baryogenesis for less
constrained chargino masses and CP violation of order sin(δCP ) > 0.1. Hence, if the advanced
transport theory is used, not only the parameter space of viable baryogenesis is much more re-
stricted, but also the maximally achievable BAU is smaller. Because of the necessity of large
CP-violating phases, additional arguments (cancellations, or a large value of the CP-odd Higgs
mass parameter) are required to suppress the electron EDM by a factor 5-6. On top of this, a light
right-handed stop and a light Higgs are needed to allow for a strong first-order phase transition
[12] (light stop scenario). Thus, electroweak baryogenesis in the MSSM is severely constrained.
In this paper we study electroweak baryogenesis in a singlet extension of the MSSM,
where the divergences of the singlet tadpole are tamed by a discrete R-symmetry [13–16]. The
R-symmetry is violated by the supersymmetry breaking terms. A singlet tadpole is then induced
at some high loop order, which is too small to destabilize the weak scale, but large enough to
evade the cosmological domain wall problem [17].
Our analysis supports the result of Ref. [18] that a strong first-order phase transition is quite
generic, once experimental constraints on the Higgs and sparticle spectrum are taken into ac-
count. The phase transition is induced by tree-level terms in the Higgs potential without the need
of a light stop. Going beyond Ref. [18], we actually compute the baryon asymmetry and the
bubble wall properties. We find that the observed baryon asymmetry can be produced with mild
tuning of the model parameters. The first and second generation squarks and sleptons have to be
heavy (a few TeV) to suppress the one-loop contributions to the EDMs. In contrast to the MSSM,
there are no strong constraints from the two-loop EDMs, since tan β is usually small [19,20].
Several variants of MSSM singlet extensions have been studied in the literature with respect to
their impact on electroweak baryogenesis [21–27]. The most detailed of these studies is Ref. [26],
where a general singlet model without discrete symmetries was considered. This general model
supports electroweak baryogenesis in a large part of its parameter space. In the current work,
the R-symmetry forbids a self coupling of the singlet, leading to a quite constrained Higgs and
neutralino phenomenology. Still, it is encouraging to see that even this restricted framework
allows for successful electroweak baryogenesis.
The paper is organized as follows. In Section 2 we will present the model and clarify notation.
In Sections 4 and 5, we will discuss the mechanism that drives baryogenesis in the nMSSM and
the dynamics of the phase transition. In Section 6, numerical results will be presented before we
conclude in Section 7.
S.J. Huber et al. / Nuclear Physics B 757 (2006) 172–196 175

2. The nearly minimal supersymmetric Standard Model

2.1. The model

The notation in this section follows Refs. [15,18], including, however, a generalization to an
additional CP-violating phase in the singlet sector. The superpotential, including the multi-loop
generated tadpole term is
m212
WnMSSM = λŜ Ĥ1 · Ĥ2 − Ŝ + yt Q̂ · Ĥ2 Û c + · · · , (1)
λ
where the dots denote the remaining terms in the MSSM superpotential, Ĥ1 = (Ĥ10 , Ĥ1− ), Ĥ2 =
(Ĥ2+ , Ĥ20 ), Ŝ is the singlet superfield and, A · B =  ab Aa Bb = A1 B2 − A2 B1 .
The tree-level potential consists of
V0 = VF + VD + Vsoft , (2)
where, restricting to third generation quarks,
 
 m212 2  2

VF = λH1 · H2 − + λSH10 + yt t˜L t˜R∗  + |λSH1− + yt b̃L t˜R∗ |2 + |λS|2 H2† H2
λ 
 ∗ 2 †
 
+ yt t˜ H H2 + |yt Q̃ · H2 |2 ,
R 2
ḡ 2  † 2 g 2  2
VD = H2 H2 − H1† H1 + H1† H2  ,
8 2
Vsoft = m21 H1† H1 + m22 H2† H2 + m2s |S|2 + (ts S + h.c.) + (aλ SH1 · H2 + h.c.)
 
+ m2Q Q̃† Q̃ + m2U |t˜R |2 + at Q̃ · H2 t˜R∗ + h.c. . (3)
The Higgs sector of this potential has only one physical CP-violating phase, which after some
redefinition of the fields can be attributed to the parameter ts . We assume that this phase is the
only source of CP violation in the model, i.e. the gaugino masses and squark and slepton soft
terms are taken to be real as well as the parameter λ.
In the case when the squarks have vanishing vevs, the tree-level Higgs potential is
V0 = m21 H1† H1 + m22 H2† H2 + m2s |S|2 + λ2 |H1 · H2 |2
  ḡ 2  † 2 g 2  2
+ λ2 |S|2 H1† H1 + H2† H2 + H2 H2 − H1† H1 + H1† H2 
8 2
+ ts (S + h.c.) + (aλ SH1 · H2 + h.c.) − m12 (H1 · H2 + h.c.).
2
(4)
We define the vevs as H10 
= φ1 = φ2
eiq1 , H20 
S = φs
eiq2 , eiqs .
We use the gauge freedom
to set H1−  = 0 and choose the phase convention q1 = q2 = q/2. The absence of a charged
condensate H2+  = 0 will be ensured by the positivity of the squared charged Higgs mass [18]
that is determined in the numerical analysis.
Using the definition of the β angle
φ1 = φ cos(β), φ2 = φ sin(β), (5)
and ts = |ts |eiqt , the tree-level potential reads finally
V0 = M 2 φ 2 + m2s φs2 + 2|ts |φs cos(qt + qs ) + 2ãφ 2 φs + λ2 φ 2 φs2 + λ̃2 φ 4 , (6)
176 S.J. Huber et al. / Nuclear Physics B 757 (2006) 172–196

where we have used


M 2 = m21 cos2 β + m22 sin2 β − m212 sin 2β cos q,

ã = sin 2β cos(q + qs ),
2
λ2 2 ḡ 2
λ̃2 = sin 2β + cos2 2β, (7)
4 8
to shorten the notation.

2.2. Effective potential at zero temperature

At zero temperature we take into account, in addition to the tree-level potential, the Coleman–
Weinberg one-loop contributions
 
1   2   2
V = gb h mb − gf h mf , (8)
16π 2
b f

where the two sums run over the bosons and the fermions with the degrees of freedom gb and gf
respectively, and
  2 
  m4 m 3
h m2 = ln − . (9)
4 Q2 2
We choose the renormalization point to be Q = 150 GeV in the DR-scheme and suitable counter-
terms, such that the one-loop contributions to the potential preserve the location of the tree-level
minimum. This leads to the following shifts in the bare mass parameters
1 ∂ V 1 ∂ V 1 ∂ V
m21 = − , m22 = − , m2s = − . (10)
2φ1 ∂φ1 2φ2 ∂φ2 2φs ∂φs
As degrees of freedom of the relevant one-loop contributions, we take
gW = 6, gZ = 3, gt = 12, gt˜L = gt˜R = 6. (11)
Contributions from the charginos and neutralinos are not taken into account. The masses used in
the one-loop potential are listed in Appendix A. The neutral Higgs masses are computed from
the second derivatives of the one-loop potential.

2.3. Effective potential at finite temperature

Taking into account temperature effects, we correct the effective potential by the thermal one-
loop contributions, which read
 
T4   2 2   2 2
V T = g J
b + mb /T − g J
f − m f /T , (12)
2π 2
b f

with the definitions




 2
J± y = dx x log 1 ∓ exp − x 2 + y 2 .
2
(13)
0
S.J. Huber et al. / Nuclear Physics B 757 (2006) 172–196 177

Table 1
Mass constraints on the spectrum
Species Mass bound
Charginos χ̃ ±  104 GeV
Neutralinos χ̃ 0  25 GeV
Charged Higgses H ±  90 GeV
Neutral Higgses H 0  114 GeV

Discussing a strong first-order phase transition in the MSSM, it is important to modify this
expression by undergoing a two-step procedure, first deriving a 3D effective action and then
treating this further with two-loop perturbation theory [28] or, more safely, with lattice numerical
methods [29]. A simplified prescription, the above one-loop expression modified by “daisy”
resummation, follows the same direction. It allowed the formulation of the postulate of a “light”
stop, in order to obtain a strong phase transition, in a transparent way [12]. In the nMSSM model
the strong first-order phase transition should be triggered by the tree-level Lagrangian and we do
not need this refined analysis. Just adding the “daisy” correction does not necessarily improve
the analysis.

3. Constraints at zero temperature

Before we analyze the phase transition and compute the produced baryon asymmetry, we con-
front the model with constraints coming from collider physics. The present limits are summarized
in Table 1.
In models with extended Higgs sectors, the Higgs couplings deviate from the SM values. Of
particular importance is the ZZHi vertex, where Hi denote the neutral Higgs mass eigenstates,
as computed from the one-loop potential. The size of this vertex is reduced by a factor

ξi = cos(β)Oi1 + sin(β)Oi2 , (14)


with respect to the SM value. The matrix Oik relates the mass eigenstates with the two CP-even
flavor eigenstates S1 , S2 . In the CP-violating case, O is a 5 × 5 matrix without block-diagonal
structure, and the special form of Eq. (14) is due to our convention for CP-even Higgs states
(see Appendix A). If the neutral Higgs mass is below the value given in Table 1, the LEP bound
translates into an upper bound on ξ , as given in Ref. [30].
We do not implement any constraints on the squark spectrum, but choose the following stop
mass parameters as used in Ref. [18]:

m2Q = m2U = 500 GeV, at = 100 GeV. (15)


The nMSSM suffers from a light singlino state. Because of the missing singlet self coupling,
this state acquires its mass only by mixing with the higgsinos. This is an important differ-
ence from more general singlet models, such as the one discussed in Ref. [26]. If this lightest
neutralino has a mass mχ̃ 0 < mZ /2, it contributes to the invisible Z width, leading to the con-
straint [30]
 4m2χ̃ 0 32
  g 2 (|U13 |2 − |U14 |2 )2 mZ
BR Z → χ̃ χ̃ =
0 0
1− < 0.8 × 10−3 . (16)
4π 24 cos2 (θW ) ΓZ m2Z
178 S.J. Huber et al. / Nuclear Physics B 757 (2006) 172–196

Here, U denotes the unitary matrix that diagonalizes the neutralinos as defined in Appendix A
and ΓZ = 2.5 GeV denotes the Z width.
A light neutralino can be avoided if the Higgs singlet coupling λ is taken to be large. For
large values of λ a Landau pole is encountered below the GUT scale. Avoiding this Landau pole
requires tan β > 1.3 and λ < 0.8 [18], but we will also consider larger values of λ. This can be
motivated by the so-called “Fat Higgs” models, where the Higgs becomes composite at some
intermediate scale [31,32].
In Ref. [18] the model was further constrained by the relic neutralino density. Here we will
not impose this constraint directly. However, from Ref. [18] we take the bound mχ̃ 0 > 25 GeV,
which ensures that the dark matter density remains below the observed value. For much larger
masses the relic neutralino density is quite small, so that neutralinos will only provide a fraction
of the total dark matter in the Universe.

4. Electroweak baryogenesis

4.1. Sources of CP violation in the nMSSM

As discussed earlier, in the nMSSM there is the possibility of additional CP violation in the
singlet sector. As in the MSSM, the relevant source in the transport equations usually comes from
the charginos, even though the neutralinos can in certain cases contribute sizable effects as well
[33]. In the interaction basis, where the higgsinos and winos are the quasi-particles, their mass
matrix takes the following form:
  q(z)
0 Yχ̃T± M2 gφ2 (z)e−i 2
Mχ̃ ± = , Yχ̃ ± = q(z) , (17)
Yχ̃ ± 0 gφ1 (z)e−i 2 μ(z)
where z denotes the direction along which a nearly planar bubble wall is moving. Unlike what
happens in the MSSM, the effective μ term acquires a z-dependence

μ(z) = −λφs (z) eiqs (z) . (18)


The leading contribution (to the left-handed current) to second order in the gradient expansion
is proportional to (see Eq. (92) in Ref. [9], and Ref. [5])
 D
S (2) ∼ m† m − m† m , ∂kz ĝ eq , (19)
where ĝ eq denotes the zero component of the vector part of the chargino Green function in
thermal equilibrium and in the interaction basis. The superscript D indicates that the diagonal
entries in the mass eigenbasis are projected out following the conventions of Ref. [9]. The first-
order sources that are used to calculate the BAU in the MSSM and nMSSM in Refs. [9] and [7,8]
become prominent when chargino mass eigenstates are nearly degenerate. Here these sources are
generally expected to be suppressed with respect to the second-order source (19), because we are
mainly interested in the generic non-degenerate case. Note that gradient expansion applies when
the typical momentum of the particles is large with respect to the inverse wall thickness, k
−1
w .
Since k ∼ T , this condition is reasonably well satisfied even for rather thin walls considered in
this paper. Therefore we expect that the sources that are not captured by the gradient expansion
Eq. (19)—an important example of which is quantum mechanical CP-violating reflection [2]—
are subdominant. This then suggests that the source (19) provides quite generically the main
contribution to the chargino-mediated BAU in the nMSSM.
S.J. Huber et al. / Nuclear Physics B 757 (2006) 172–196 179

In the MSSM, where q(z) = 0 and ∂z μ = 0, the evaluation of the second-order source (19)
leads to
  D
(2) 0 M2 φ2 − μ∗ φ1
SMSSM ∼ g 2 , ∂ ĝ eq
; (20)
−M2∗ φ2 + μφ1 0 kz

using the conventions of Ref. [9], this can be written as


Im(M2 μ)  
(2)
Tr SMSSM ∼ g2 φ1 φ2 + φ1 φ2 ∂kz ĝ eq,L , (21)
Λ
where Λ denotes the difference of the eigenvalues of the matrix m† m.
In the nMSSM, there are various additional contributions from the derivatives acting on μ in
the source (19), especially a novel diagonal term of the following form:
 D
(2) 0 0
SnMSSM ∼ , ∂kz ĝ eq
. (22)
0 μ ∗ μ − μ∗ μ
This contribution dominates if μ M2 , which is usually the case in the nMSSM since μ is
related to the singlet vev, while the Wino mass parameter M2 is not related to the parameters of
the singlet sector and expected to be of the SUSY scale.
Hence, we will consider two scenarios. First, we consider the case of large M2 ≈ 1 TeV.
In this region the contribution (22) will almost coincide with the full expression (19). Second,
we will choose a rather small wino-mass parameter M2 ≈ 200 GeV. In this case the additional
contributions in Eq. (19) can lead to an enhancement or a cancellation in the BAU, and one
should keep in mind that the neglected mixing effects could contribute as well. In both cases,
using the full second-order source (19), the baryon to entropy ratio η is determined as was done
in Ref. [6] for the MSSM.
We use a system of diffusion equations that was first derived in Ref. [34] and later adapted
in Refs. [4,6–8,11,35]. This system describes how the CP violation is communicated from the
chargino sector to the left-handed quarks and finally biases the sphaleron processes. These
diffusion equations rely on certain assumptions, e.g. that the supergauge interactions are in equi-
librium, and hence lead to sizable uncertainties. Furthermore, we do not take into account recent
developments in the determination of interaction rates presented in Refs. [36,37], but employ the
parameters of Ref. [6]. Nevertheless, the accuracy of the determined BAU should be sufficient
for the analysis in this work.
In the following we will briefly focus on the term (22), which is prominent in the limit of
large M2 . Beside the critical temperature Tc , the generated BAU only depends on the profile of
the higgsino mass parameter μ(z) during the phase transition, namely the change of the phase
qs , the wall thickness lw and the profile of |μ(z)|.
With good accuracy, the baryon to entropy ratio, η ≡ nB /s, scales as
qs
η∝ . (23)
lw Tc
The dependence on the profile of |μ(z)| is shown in Fig. 1. In this example, the profile is para-
metrized by

 
μ(z) = μ0 − μ 1 + 1 tanh(z/ lw ) , (24)
2 2

1 1
qs (z) = qs + tanh(z/ lw ) , (25)
2 2
180 S.J. Huber et al. / Nuclear Physics B 757 (2006) 172–196

Fig. 1. The dependence of η10 ≡ 1010 η on the chargino mass parameter μ = μ0 . The parameters used are lw = 10/Tc
and qs = π/10.

and the values μ0 = μ, lw = 10/Tc , qs = π/10 and several values of Tc have been chosen.
To give some feeling about the produced BAU, we note that a good estimate of the predicted
η10 is in this case given by the formula
 3
qs 1 μ0 2 μ
η10 ≈ c(Tc ) exp(−μ0 /τ Tc ), (26)
π lw Tc τ Tc τ Tc
with c(Tc ) ≈ 1.6Tc /GeV and τ ≈ 0.78. This formula characterizes the BAU in the case of large
M2 ≈ 1 TeV and μ0 ≈ μ.

4.2. EDM constraints in the nMSSM

The most severe experimental constraints on CP violation come from measurements of the
EDM of the electron, de < 1.6 × 10−27 e cm [38], and neutron, dn < 3.0 × 10−26 e cm [39]. Al-
ready at the one-loop level, contributions of the superpartners give sizable effects in the case of
CP-violating phases of O(1). These one-loop diagrams are the same in the nMSSM and MSSM,
setting μ = −λS. Minimizing the Higgs potential, we can compute the phase of the effective
μ parameter. Of course, this phase could be neutralized by introducing a compensating phase in
the parameter M2 . Such a tuning would allow us to eliminate the one-loop EDMs completely,
without much affecting the generated baryon asymmetry, since the dominating source is propor-
tional to the change in the phase qs and not sensitive to the value of qs = arg(μ) in the broken
phase. This is not possible in the MSSM, because there the produced BAU is, like the electron
EDM contribution from the charginos, proportional to the combination Im(μM2 ).
However, here we take a different approach, using only the phase qt in the Higgs potential as
sole source of CP violation. The one-loop EDMs then induce mass bounds for the first and second
generation squarks and sleptons; depending on the model parameters these are in the range from a
few TeV up to 50 TeV [40]. Since the constraints from the neutron EDM are usually less stringent
than the ones coming from the electron EDMs, we will focus in the following on the latter. In
S.J. Huber et al. / Nuclear Physics B 757 (2006) 172–196 181

single cases we also calculated the Barr–Zee type contributions to the neutron EDM [41], but
they barely reach the most recent experimental bounds [39].
Besides heavy sfermion masses, there is another possibility to suppress the one-loop electron
EDM in our model. Notice that the absolute value of the CP-violating phase (qs + q) can be
smaller in the broken phase than in the symmetric phase. This phase (qs + q) is the only CP-
odd combination that enters the electron EDMs on the one-loop level in our simplified nMSSM
model, with CP violation only in the ts parameter. Hence, there is the possibility that today’s
observed sin(qs + q) 1, even though the phase qs greatly changed during the phase transition
thus producing the BAU. We will analyze this possibility in detail in Section 6. The explicit form
of the one-loop electron EDM contributions is given in Appendix B. This possibility entails a
certain amount of tuning.
Additional EDMs can be generated from two-loop chargino or Higgs graphs (see, for in-
stance, [40] and references therein). Notice that the MSSM two-loop chargino contribution to
the electron EDM [19,20] is proportional to tan β and hence subleading in our model that usually
predicts tan β ∼ O(1). Potentially harmful diagrams, including the additional CP-odd scalar in
the nMSSM, are small as well because of the modest tan β and because only the Higgs com-
ponent of the CP-odd scalars couples to the charginos, while the singlet component delivers the
additional CP violation (for a calculation of these contributions see [42]).

5. Electroweak phase transition

One of the parameters entering our baryogenesis analysis is the thickness of the Higgs wall
profile during the electroweak phase transition lw . Since our CP-violating source is a second-
order effect in gradients, the integrated BAU scales as η ∼ 1/ lw , as already mentioned in
Eq. (23).
To determine the wall thickness lw one has to examine the dynamics of the phase transition
[43]. This has been done for the MSSM in Ref. [12] and for the NMSSM in Ref. [26]. Typical
values for the MSSM seem to be close to lw = 10/Tc . In the nMSSM, we expect rather thin wall
profiles, since the linear singlet term and the trilinear singlet Higgs term in the effective potential
will make the phase transition much stronger than the loop-suppressed stop corrections that are
responsible for the first-order phase transition in the MSSM. In the light of Eq. (23) this will
further enhance the produced asymmetry with respect to the MSSM case.
To determine the dynamical parameters of the wall, we solved for the classical bounce solu-
tion of the Higgs and singlet fields (φ1 , φ2 , φs , q, qs ) at the critical temperature (where the two
minima of the potential are degenerate). Our numerical approach is based on the variation of the
classical action and is discussed in detail in Ref. [44].
Another parameter relevant to the dominating source in Eq. (22) is the profile of the CP-
violating phase of the singlet field. A typical solution is displayed in Fig. 2 corresponding to the
parameters qs = 0.119 and lw = 4.81Tc−1 .
To illuminate a little bit the nature of the phase transition, we will recall some approximate
analytical criterion for the occurrence of a first-order phase transition, first given in a slightly
different way in Ref. [18]. Consider the tree-level potential in Eq. (6) without CP violation,
qt = 0. Additionally assume that the temperature effects give rise to the effective potential

V T = αφ 2 T 2 , (27)
182 S.J. Huber et al. / Nuclear Physics B 757 (2006) 172–196

Fig. 2. A typical wall profile for the parameter qs , corresponding to the parameters qs = 0.119 and lw = 4.81Tc−1 .

where α is some unspecified positive constant. In Ref. [18] it was shown that a necessary condi-
tion for a first-order phase transition is approximately given by
 
1  λ2 ts 
2
ms <  − ms ã . (28)
λ̃ m s
This can be seen in the following way. Given a certain value for φ, φs can easily be evaluated
to be

ts + ãφ 2
φs = − . (29)
m2s + λ2 φ 2
Using this in our potential, and expanding around φ = 0, we obtain
ts2   (ts + ãφ 2 )2
V + V T = − + M 2 + αT 2 φ 2 + λ̃2 φ 4 − 2
ms 2 ms + λ2 φ 2
= c 0 + c1 φ 2 + c 2 φ 4 + c 3 φ 6 + · · · , (30)
with the coefficients
ts2 2ãts λ2 ts2
c0 = − , c1 = M 2 + αT 2 − + ,
m2s m2s m4s
 
1 λ 2 ts 2 λ2 λ 2 ts 2
c2 = λ̃ − 2 ã − 2 ,
2
c3 = 4 ã − 2 . (31)
ms ms ms ms
If the symmetric minimum is absent at zero temperature, c1 (T = 0) < 0, a temperature T2 can be
found, such that c1 (T2 ) = 0. For this temperature there exists a lower-lying potential minimum in
the case c2 < 0, which is equivalent to the condition in Eq. (28). Since for temperatures T  T2 a
potential well develops between the symmetric and the lower broken vacuum, a first-order phase
transition is possible, given the vacuum decay rate is large enough such that the transition occurs
S.J. Huber et al. / Nuclear Physics B 757 (2006) 172–196 183

Table 2
Dimensionless parameter ranges used for the numerical analysis
Lower bound Parameter Upper bound
0.1 < m1 < 1
0.1 < m2 < 1
0.1 < m12 < 2
0 < ms < 2
−2 < λ < 2
0 < aλ < 2
1/3
0 < ts < 2
0 < qt < 2π

before the temperature T2 is reached. Hence, it is possible in the nMSSM to obtain a first-order
phase transition due to tree-level dynamics, in contrast to the MSSM. Analyzing the numerical
results, we will see that the constraint in Eq. (28) is usually fulfilled in viable models, even if the
one-loop contributions to the potential and the CP phase are included and hence that the phase
transition is dominated by the tree-level dynamics.
The argument just presented is a concrete realization of the general effective field theory
approach recently discussed in Refs. [45,46]. There it was shown that a strong first-order phase
transition can be induced at tree-level by the interplay of a negative φ 4 term and a positive φ 6
term which stabilizes the potential. The suppression scale of the φ 6 term should be somewhat
below a TeV for the mechanism to work. Here the relevant φ 4 and φ 6 operators are generated by
integrating out the singlet field. This generalizes the usual situation, where the phase transition
is induced by a negative φ 3 term and a positive φ 4 term.

6. Numerical analysis

To inspect the parameter space of the nMSSM we proceed as follows. First, we choose ran-
dom parameters for the Higgs potential in the ranges displayed in Table 2. To ensure maximal
numerical stability, all chosen parameters are of O(1), and can be thought of as dimensionless
parameters. Those parameters then lead not to the physical Higgs vev, but to some dimension-
less Higgs vev φ0 . Finally, all dimensionful quantities, such as the critical temperature or the
mass spectrum have to be scaled with (173.458 GeV)/φ0 to yield the physical values. During the
minimization of the potential, stable and metastable broken minima were analyzed. Depending
on the parameters, metastable minima occur but, in the CP-conserving case qt = 0, no transi-
tional (spontaneous) CP violation was observed in contrast to the NMSSM [26] that contains an
additional cubic singlet term.
Next, we correct the bare parameters with the one-loop contributions of Eq. (8) and confront
them with the constraints on the mass spectrum from Table 1 and on the Z-width from Eq. (16).
If the parameter set passes these constraints, we add the temperature dependent contributions
to the effective potential as explained in Section 2.3 and examine the phase transition. We require
that the models have a first-order phase transition of sufficient strength [18,47], φ0 /Tc > 0.9.
Before we discuss baryogenesis in our model, we would like to examine restrictions on the
parameters imposed by the constraints on the mass spectrum and comment on the criterion for a
first-order phase transition given in the last section in Eq. (28).
First, the eight parameters given in Table 2 have to lead to the correct Higgs vev, which is
achieved by a rescaling of the dimensionful parameters. Hence our parameter space is effectively
184 S.J. Huber et al. / Nuclear Physics B 757 (2006) 172–196

Fig. 3. The critical temperature Tc , the Higgs vev φ in the broken phase at Tc and one Higgs mass as functions of aλ .

only seven dimensional. One restriction on the parameters is that aλ cannot be chosen arbitrarily
large, since this destabilizes the potential in the negative φs direction. Analyzing the parameter
sets that fulfill the mass constraints, one observes that the parameters λ and qt are not distrib-
uted homogeneously. Small values of λ make it seemingly difficult to fulfill the mass bound of
the chargino, since one of the diagonal entries of the chargino mass matrix is −λφs . To have
a potential with extremely large vev φs requires at least some fine-tuning, since the one-loop
contribution tends to yield an effective potential that is unbounded from below, if the tree-level
parameters are chosen to provide a large vev φs . Large values of λ hence seem to be the more
natural choice, even though they can lead to a Landau pole [18]. Usually, the mass constraints on
the neutralinos are automatically fulfilled, if the charginos surpass their more restrictive bounds,
but additional constraints on the parameters enter through the spectrum of the Higgs particles. In
many cases, a range of values for the parameter aλ can be found, where off-diagonal elements
in the Higgs mass matrix cancel, which tends to enlarge the lightest Higgs mass. In addition the
parameter aλ has a strong influence on the phase transition according to Eq. (28).
This situation is demonstrated in Fig. 3 for a parameter set with a rather small parameter λ. The
parameters m1 , m2 , m12 , ms are chosen such that tan(β) = 2.0, φs = −250 GeV, φ = 173 GeV
and Ma = 500 GeV at the tree-level, where the CP-odd Higgs mass parameter is defined by
1  2 
Ma2 = m − aλ φs . (32)
sin β cos β 12
1/3
The remaining parameters are λ = 0.55, ts = 70 GeV, qs = 0.3, while aλ is varied. For
172 GeV < aλ < 178 GeV, this model develops a strong first-order phase transition and gen-
erates more than the observed BAU. For lower values of aλ , the model has no stable broken
phase, while for larger values of aλ , the phase transition is too weak. The plotted Higgs mass is
that of the third lightest Higgs, but the two lighter states would have escaped detection at LEP
because of the suppressed coupling to the Z-boson. This example demonstrates that even though
there exist viable models without Landau pole in λ, this possibility entails a certain amount of
tuning.
S.J. Huber et al. / Nuclear Physics B 757 (2006) 172–196 185

Fig. 4. The parameters λ and qt for a set of random models that fulfill the mass constraints.

Fig. 5. The plots show the combinations of the parameters that enter the tree-level condition for a first-order phase
transition, Eq. (28). The left plot contains parameter sets that fulfill only the mass constraints, while the right plot contains
parameter sets that have in addition a strong first-order phase transition.

The second parameter that is restricted by the mass constraints is the CP-violating phase qt .
The reason for this effect is that values with cos(qt ) ≈ −1 lead to smaller Higgs masses. Fig. 4
displays the parameters λ and qt for a set of random models that fulfill the mass constraints.
Demanding a strong first-order phase transition further restricts the parameter space. In Fig. 5
we plot the left-hand side versus the right-hand side of the criterion in Eq. (28) (both sides are
scaled by 1/(m1 m2 ) to make them dimensionless). In the left plot we use random models that
fulfill the mass constraints on the spectrum, but are unconstrained otherwise; in the right plot we
impose the mass constraints on the model and require a strong first-order phase transition. In the
latter case, most of the parameter sets are in accordance with Eq. (28), while the parameter sets
in the former case are evenly distributed. Hence, the tree-level criterion for the phase transition
seems to be applicable even if the one-loop contributions to the effective potential and the CP
phase are taken into account.
186 S.J. Huber et al. / Nuclear Physics B 757 (2006) 172–196

Fig. 6. The binned result of the parameter tan(β) for the parameter sets with a strong first-order phase transition.

Fig. 7. The binned result of the parameter lw for the parameter sets with a strong first-order phase transition.

In the nMSSM, for several reasons, we expect a much larger BAU than in the MSSM. First,
the parameter tan(β), which needs to be large in the MSSM, is naturally O(1) in the nMSSM as
depicted in Fig. 6. This does not only help to suppress the two-loop contributions to the EDMs, as
discussed in the previous section, but also enhances the contributions from the source in Eq. (21).
Secondly, the wall thickness, which in the MSSM usually is of order 20/Tc –30/Tc [12], can be
much smaller, since the phase transition is strengthened by the linear and trilinear terms in the
effective potential. This claim is supported by Fig. 7. The third reason for the enhancement of
the BAU in the nMSSM with respect to the MSSM is the additional source in Eq. (22), which is
in many cases dominating the generation of the BAU.
Finally, we calculate the generated baryon asymmetry and compare the result with the ex-
perimental observation, η = (0.87 ± 0.03) × 10−10 [48]. The result is shown in Figs. 8 and 9.
Approximately 50% of the parameter sets predict a higher value than the observed baryon asym-
S.J. Huber et al. / Nuclear Physics B 757 (2006) 172–196 187

Fig. 8. The binned result for the BAU analysis for large M2 = 1 TeV. Approximately 50% of the parameter sets predict
a value of the baryon asymmetry higher than the observed one. The bottom line corresponds to parameter sets that fulfill
current bounds at the electron EDM with 1 TeV sfermions (4.8%).

Fig. 9. The binned result for the BAU analysis for small M2 = 200 GeV. Approximately 63% of the parameter sets
predict a value of the baryon asymmetry higher than the observed one. The bottom line corresponds to models that fulfill
current bounds at the electron EDM with 1 TeV sfermions (6.2%).

metry in the model with large M2 = 1 TeV, while this number increases to 63% for small
M2 = 200 GeV.
In addition, we plotted the BAU generated by parameter sets that fulfill the experimental
bounds on the electron EDM with sfermion masses of 1 TeV in the first and second generation.
Some of them predict a BAU in accordance with observation, and hence give the possibility to
construct nMSSM models that contain less constrained sfermions (lighter than 1 TeV) being at
the same time consistent with EDM constraints and baryogenesis. In some cases the electron
EDM is small because of a random cancellation between the neutralino and chargino contribu-
188 S.J. Huber et al. / Nuclear Physics B 757 (2006) 172–196

tions, but occasionally the suppression of the electron EDM is due to the fact that the combination
sin(qs + q) is relatively small in the broken phase.

7. Conclusion

We have analyzed the phase transition and baryogenesis in the nMSSM (1)–(4) with CP vi-
olation in the singlet sector. We have shown that the singlet field enhances the strength of the
phase transition in such a way that one typically obtains a strong phase transition, as required
for successful baryogenesis. This is to be contrasted with the MSSM, in which the mass of the
lightest Higgs field must not be greater than about 120 GeV, and the right-handed stop must be
light, mt˜R ∈ (120, 160) GeV.
Next we performed the calculation of baryogenesis mediated by charginos in the nMSSM.
After calculating the CP-violating sources in the gradient expansion, we argued that in most of
the parameter space the dominant source comes from the second-order semiclassical force in the
Boltzmann transport equation for charginos. The source related to flavor mixing, of first order in
the gradient expansion, tends to be smaller, because the bubble wall is rather thin. For a generic
choice of parameters, one is far from the chargino mass degeneracy, where the first-order source
may be important. To come to this conclusion, we used an approach to the calculation of the
first-order sources [6,9] that differs from earlier work [7,8,18] in the sense that our treatment of
sources is basis independent, and the magnitude of the source in the transport equation is unam-
biguous. Using this advanced transport theory, the first-order sources are of a somewhat lower
amplitude and exhibit a much narrower resonance near the chargino mass degeneracy, with the
effect that in most of the parameter space the second-order source dominates. In the MSSM, this
is not the case for the chargino-mediated baryogenesis because the wall tends to be thicker, thus
weakening the second-order (semiclassical force) source, while leaving the first-order source
more or less unchanged. Furthermore, the dominant second-order source of baryogenesis in the
nMSSM, Eq. (22), is not present in the MSSM. Owing to these differences, successful baryo-
genesis in the MSSM is only possible near the resonance (chargino mass degeneracy), and with
nearly maximum CP violation, which is in conflict with the current EDM bounds, unless a tuning
of parameters is invoked [6].
On the other hand our analysis of the baryon production in the nMSSM looks promising.
When we restrict the CP-violating phase in the singlet sector to be about qt ∼ 0.3, and take
tan(β) ∼ 1, we still get baryon production consistent with the observed value. This choice
certainly does not violate any of the current EDM bounds. When qt is chosen randomly,
approximately 50% (63%) models predict more than the observed BAU when M2 = 1 TeV
(M2 = 200 GeV), which indicates that baryogenesis in the nMSSM is generic.
It is finally interesting to compare baryogenesis in the nMSSM and in the general NMSSM
formerly analyzed in Ref. [26]. Because of the presence of a singlet self coupling and an explicit
μ term, the NMSSM allows for a much richer Higgs phenomenology. (However, some additional
assumptions about the structure of the higher dimensional operators have to be made, in order to
prevent the destabilization of the electroweak scale by corrections to the singlet tadpole.) There
is no danger of a light singlino state, so that the model can account for the observed baryon
asymmetry also for values of λ  0.1 and large tan β. It also allows for transitional CP violation,
which means an electroweak phase transition that connects a high-temperature CP-broken phase
with a low-temperature CP-symmetric phase. This way there are varying complex phases in the
bubble wall, without leaving a trace at low temperatures. In the present model this possibility is
prevented by the strongly constrained Higgs potential.
S.J. Huber et al. / Nuclear Physics B 757 (2006) 172–196 189

In Ref. [26] the source terms were computed by solving the Dirac equation for the charginos in
the WKB approximation. As has been recently shown in Ref. [11], this formalism can reproduce
the second-order source of Ref. [5], used in the present work, if the Lorentz transformation to a
general Lorentz frame is done carefully (see also Ref. [49]). These results suggest that in Ref. [26]
the baryon asymmetry was underestimated by a factor of 2 to 5. It seems to be interesting to
update this analysis to meet the present experimental constraints. It also seems promising to
apply the presented techniques to more general supersymmetric models, such as models with
extra U (1) symmetries [50].

Acknowledgements

We would like to thank M. Laine for discussing some subtleties of the electroweak phase
transition in the MSSM with us. T.K. is supported by the Swedish Research Council (Vetenskap-
srådet), Contract No. 621-2001-1611.

Appendix A. Mass spectrum of the nMSSM

In the following we collect the mass matrices that have been used in the one-loop potentials
in Section 2.
We used the physical constants
sin(θW ) = 0.2312, αEW = 1/127.907, (A.1)
mZ = 91.1876 GeV, mt = 165.0 GeV, (A.2)
which give rise to the values
g = 0.357, g = 0.652, φ0 = 173.458 GeV, (A.3)
where φ0 denotes the T = 0 value of the vev φ.
If not stated differently, we have used the SUSY-breaking parameters
mU = mQ = 500 GeV, at = 100 GeV,
M2 = 2M1 = 200 GeV, mE = 500 GeV, (A.4)
which enter in the mass matrices that we will define subsequently.

A.1. Higgs bosons

For the neutral Higgs bosons we use the notation


Hi0 = eiqi (φi + Si + iPi ), S = eiqs (φs + Ss + iPs ), (A.5)
and q1 = q2 = q/2. The corresponding mass matrix has the following form
 2 2
MSS MSP
MH2
= , (A.6)
MP2 S MP2 P
with the matrices MSS , MSP = MP† S , and MP P given by the following entries. The CP-even
entries at the tree-level read
  ḡ 2  2 
2
MSS11 = m21 + λ2 φ22 + φs2 + 3φ1 − φ22 ,
4
190 S.J. Huber et al. / Nuclear Physics B 757 (2006) 172–196

1  
2
MSS12 = −m212 cos(q) + aλ φs cos(q + qs ) + φ1 φ2 4λ2 − ḡ 2 ,
2
2
MSS13 = aλ φ2 cos(q + qs ) + 2λ2 φ1 φs ,
  ḡ 2  2 
2
MSS22 = m22 + λ2 φ12 + φs2 + 3φ2 − φ12 ,
4
2
MSS23 = aλ φ1 cos(q + qs ) + 2λ2 φ2 φs ,
2
MSS33 = m2s + λ2 φ 2 . (A.7)
The CP-odd entries are
  ḡ 2  2 
MP2 P 11 = m21 + λ2 φ22 + φs2 + φ1 − φ22 ,
4
MP2 P 12 = m212 cos(q) − aλ φs cos(q + qs ),
MP2 P 13 = −aλ φ2 cos(q + qs ),
  ḡ 2  2 
MP2 P 22 = m22 + λ2 φ12 + φs2 + φ2 − φ12 ,
4
MP2 P 23 = −aλ φ1 cos(q + qs ),
MP2 P 33 = m2s + λ2 φ 2 . (A.8)
Finally the CP-mixed entries yield

12 = m12 sin(q) − aλ φs sin(q + qs ),


2 2
MSP
13 = −aλ φ2 sin(q + qs ),
2
MSP
23 = −aλ φ1 sin(q + qs ).
2
MSP (A.9)
In the CP-conserving case the submatrix (A.9) vanishes so that CP-even and CP-odd states do
not mix.
If the one-loop effective potential is included, we determine the masses of the neutral higgses
by the second derivatives of the effective potential.
The mass matrix of the charged Higgs bosons in the basis (H1− , H̄2+ ) contains the complex
entries
ḡ 2  2  g2
± = m1 + λ φ − φ2 − φ12 + φ22 ,
2 2 2
MH
11 4 2
1  2 
2
MH ± = − φ1 φ2 2λ − g + m12 e − aλ φs ei(q+qs ) ,
2 2 iq
12 2
ḡ 2  2  g2
2
MH ± = m22 + λ2 φ − φ1 − φ22 + φ12 . (A.10)
22 4 2

A.2. Charginos and neutralinos

The chargino mass matrix reads


  q
0 Yχ̃T± M2 gφ2 e−i 2
Mχ̃ ± = , Yχ̃ ± = q , (A.11)
Yχ̃ ± 0 gφ1 e−i 2 −λφs eiqs
and is diagonalized by the biunitary transformation Yχ̃ ± = U ∗ Yχ̃ ± V † .
diag
S.J. Huber et al. / Nuclear Physics B 757 (2006) 172–196 191

The symmetric mass matrix of the neutralinos in the basis χ̃ 0 = (i B̃, i W̃ 3 , H̃10 , H̃20 , S̃) yields
⎛ M ... ... ... ...⎞
1
⎜ 0 M2 ... ... ...⎟
⎜ g q q ⎟
⎜ − √ φ1 e−i 2 √g φ1 e−i 2 0 ... ...⎟,
Mχ̃ 0 =⎜ 2 2 ⎟ (A.12)
⎜ g q q ⎟
⎝ − √ φ2 e−i 2 − √g φ2 e−i 2 λφs eiqs 0 ...⎠
2 2
i q2 q
0 0 λφ2 e λφ1 ei 2 0
diag
and can be diagonalized using a unitary matrix, Mχ̃ 0 = X T Mχ̃ 0 X.

A.3. Gauge bosons

The mass of the W -boson is given by


1
m2W = g 2 φ 2 , (A.13)
2
while the photon and the Z-boson share the following Hermitian mass matrix
 1 2 2
2g φ − 12 gg φ 2
2
MZγ = (A.14)
− 12 gg φ 2 1 2 2
2g φ

that leads to the Z-boson mass mZ = (ḡ/ 2 )φ.

A.4. Tops and stops

The top quark has the mass

m2t = yt2 φ22 , (A.15)


and the masses of the stops are given by the following Hermitian matrix
 2 iq
mQ + m2t + 14 (g 2 − 13 g 2 )(φ12 − φ22 ) at φ2 e 2 + yt λφs φ1 e−i(qs +q/2)
Mt˜ =
2
. (A.16)
at φ2 e−iq/2 + yt λφs φ1 ei(qs +q/2) m2U + m2t + 13 g 2 (φ12 − φ22 )

A.5. Sneutrinos and selectrons

For the selectrons we have



mE −λφs eiqs me tan(β)
Mẽ =
2
, (A.17)
−λφs e−iqs me tan(β) mE
2
that is diagonalized via the transformation Mẽ,diag = D † Mẽ2 D.
For the electron EDM contributions from the charginos we use a sneutrino of mass mν̃ =
1 TeV.

Appendix B. One-loop contributions to the electron EDM

In this appendix we briefly discuss the one-loop contributions to the electron EDM coming
from chargino and neutralino exchange. For the sfermions of the first two generations we assume
192 S.J. Huber et al. / Nuclear Physics B 757 (2006) 172–196

masses of 1 TeV. The contribution from the charginos is given by [51]:

2 m +  m2 +
de-chargino αEM χ̃i χ̃i
= 2 2
Im(Γei )A , (B.1)
e 4π sin (θW ) i=1 mν̃e m2ν̃e
where
∗ me eiq/2
Γei = κe Ui2 Vi1 , κe = √ . (B.2)
2mW cos(β)
Here, the matrices U and V diagonalize the chargino mass matrix as defined in the previous
section and

1 2
A(r) = 3−r + ln(r) (B.3)
2(1 − r)2 1−r
denotes the loop function.
Analogously, the contribution from the neutralinos is [51,52]

de-neutralino αEM 2  4 m 0
χ̃i  
= 2 2
Im(ηeik )B m2χ̃ 0 /m2ẽ , (B.4)
e 4π sin (θW ) k=1 i=1 mẽ i

with
 
  ∗ √ ∗
 ∗ κe ∗
ηeik = tan(θW )X1i + X1i D1k + 2κe X3i D2k tan(θW )X1i D2k + √ X3i D1k . (B.5)
2
In this equation, X and D diagonalize the neutralinos and selectrons, respectively. The loop
function B is defined by

1 2r
B(r) = 1+r + ln(r) . (B.6)
2(1 − r)2 1−r
Following the discussion in Ref. [51], it can be shown that if qt is the only CP-violating phase
in the Higgs sector, both contributions depend only on the CP-odd combination (qs + q). The
current experimental bound on the electron EDM is de < 1.6 × 10−27 e cm [38].

Appendix C. Example sets

In this appendix we give some examples of parameters that develop a strong first-order phase
transition. The examples are not chosen arbitrarily, but they represent specific cases. The first
set, shown in Tables 3–10, accomplishes the generation of a large BAU thanks to a relatively thin

Table 3
Parameter examples used for the numerical analysis
1/3
Set m1 in GeV m2 in GeV m12 in GeV ms in GeV λ aλ in GeV ts in GeV qt
1 157.2 93.0 170.6 55.8 1.2642 268.2 98.4 2.113
2 124.2 149.2 215.0 127.2 1.4320 254.1 152.2 5.050
3 248.9 230.8 243.3 160.5 −0.8937 214.2 190.5 0.046
4 397.1 251.9 375.5 342.1 −1.2991 292.0 310.5 0.282
5 66.4 98.5 111.8 65.5 0.7656 95.3 102.0 6.193
6 425.1 165.7 240.1 26.8 0.5500 177.3 70.9 0.300
S.J. Huber et al. / Nuclear Physics B 757 (2006) 172–196 193

Table 4
The mass spectrum of the charginos, charged Higgses and stops (in GeV)
Set χ̃ ± H± t˜1/2
1 221.77 107.40 219.56 521.43 529.30
2 221.64 131.22 236.41 537.16 514.91
3 270.45 109.31 410.84 485.74 563.03
4 146.21 315.37 588.45 479.29 567.75
5 221.06 105.94 148.97 530.64 521.67
6 225.24 144.62 494.87 520.50 528.89

Table 5
The mass spectrum of the neutral Higgses (in GeV)
Set S, P
1 142.72 210.53 217.41 273.78 357.83
2 177.85 260.59 292.50 296.78 396.55
3 119.12 202.13 251.16 432.44 455.77
4 121.65 395.09 448.97 602.98 641.00
5 115.07 118.67 169.87 211.08 237.02
6 76.74 89.86 114.76 504.68 506.91

Table 6
The mass spectrum of the neutralinos (in GeV)
Set χ̃ 0
1 267.76 105.80 113.09 221.89 184.03
2 313.48 105.14 138.79 222.79 198.12
3 135.40 68.73 89.48 275.43 269.04
4 389.60 323.06 81.50 161.87 123.56
5 222.16 181.83 116.61 107.33 94.72
6 227.33 42.25 105.37 165.39 175.04

Table 7
The vevs in the broken phase at temperature T = 0
Set φ in GeV β q φs in GeV qs
1 173.46 0.926 0.157 −71.0 −0.392
2 173.46 0.739 −0.214 −81.8 0.687
3 173.46 0.808 0.015 −202.9 −0.036
4 173.46 0.915 0.074 −202.6 −0.256
5 173.46 0.714 −0.024 −112.6 0.051
6 173.46 1.108 0.029 −251.9 −0.067

wall. On the other hand the electron EDM is extremely small, partly because the combination
q + qs is rather small in the broken phase and partly due to a coincidental cancellation between
the neutralino and the chargino contributions to the electron EDM. Set number 6 describes a
model with small λ, taken from Fig. 3 with aλ = 177.3 GeV. This model generates more than
the observed BAU, but the calculated electron EDM is slightly too big, so that the sfermions of
the first two generations have to be heavier than 1 TeV. Starting from qt ∼ 0.1 instead of 0.3
should yield η10 ∼ 1 and an electron EDM within the experimental bound. The remaining sets
194 S.J. Huber et al. / Nuclear Physics B 757 (2006) 172–196

Table 8
The vevs in the broken phase at temperature T = Tc
Set φ in GeV β q φs in GeV qs
1 165.4 0.900 0.175 −71.4 −0.437
2 170.0 0.726 −0.223 −83.2 0.708
3 150.3 0.803 0.016 −214.9 −0.038
4 164.5 0.911 0.076 −207.4 −0.259
5 151.6 0.687 −0.029 −120.2 0.058
6 141.9 1.102 0.041 −261.5 −0.092

Table 9
The vevs in the symmetric phase at temperature T = Tc
Set φ in GeV β q φs in GeV qs
1 0.0 – – −281.4 −2.113
2 0.0 – – −241.6 1.233
3 0.0 – – −283.3 −0.046
4 0.0 – – −259.9 −0.282
5 0.0 – – −266.0 0.090
6 0.0 – – −471.5 −0.300

Table 10
The parameters of the phase transition and the generated BAU
Set Tc in GeV de in 10−27 e cm lw in Tc−1 η10
1 113.5 0.002 2.43 29.443
2 99.1 0.796 6.38 −3.214
3 109.1 0.499 7.82 0.014
4 78.6 5.894 34.01 0.005
5 115.8 −0.893 3.05 −0.398
6 105.6 −2.054 2.39 2.717

are randomly chosen. Set number 6 demonstrates that even a sizable value of qt does not lead to
a large baryon asymmetry if qs is small along the bubble wall. The rather thick wall induces an
additional suppression of η10 in this case.

References

[1] V.A. Kuzmin, V.A. Rubakov, M.E. Shaposhnikov, On the anomalous electroweak baryon number nonconservation
in the early universe, Phys. Lett. B 155 (1985) 36.
[2] A.G. Cohen, D.B. Kaplan, A.E. Nelson, Diffusion enhances spontaneous electroweak baryogenesis, Phys. Lett.
B 336 (1994) 41, hep-ph/9406345.
[3] M. Joyce, T. Prokopec, N. Turok, Electroweak baryogenesis from a classical force, Phys. Rev. Lett. 75 (1995) 1695,
hep-ph/9408339;
M. Joyce, T. Prokopec, N. Turok, Phys. Rev. Lett. 75 (1995) 3375, Erratum;
M. Joyce, T. Prokopec, N. Turok, Nonlocal electroweak baryogenesis. Part 2: The classical regime, Phys. Rev. D 53
(1996) 2958, hep-ph/9410282.
[4] J.M. Cline, M. Joyce, K. Kainulainen, Supersymmetric electroweak baryogenesis in the WKB approximation, Phys.
Lett. B 417 (1998) 79, hep-ph/9708393;
J.M. Cline, M. Joyce, K. Kainulainen, Phys. Lett. B 448 (1999) 321, Erratum;
J.M. Cline, M. Joyce, K. Kainulainen, Supersymmetric electroweak baryogenesis, JHEP 0007 (2000) 018, hep-
ph/0006119;
S.J. Huber et al. / Nuclear Physics B 757 (2006) 172–196 195

J.M. Cline, M. Joyce, K. Kainulainen, Supersymmetric electroweak baryogenesis, hep-ph/0110031, Erratum.


[5] K. Kainulainen, T. Prokopec, M.G. Schmidt, S. Weinstock, First principle derivation of semiclassical force for
electroweak baryogenesis, JHEP 0106 (2001) 031, hep-ph/0105295;
T. Prokopec, M.G. Schmidt, S. Weinstock, Transport equations for chiral fermions to order h-bar and electroweak
baryogenesis, Ann. Phys. 314 (2004) 208, hep-ph/0312110;
T. Prokopec, M.G. Schmidt, S. Weinstock, Transport equations for chiral fermions to order h-bar and electroweak
baryogenesis. II, Ann. Phys. 314 (2004) 267, hep-ph/0406140.
[6] T. Konstandin, T. Prokopec, M.G. Schmidt, M. Seco, MSSM electroweak baryogenesis and flavour mixing in trans-
port equations, Nucl. Phys. B 738 (2006) 1, hep-ph/0505103.
[7] M. Carena, M. Quiros, M. Seco, C.E.M. Wagner, Improved results in supersymmetric electroweak baryogenesis,
Nucl. Phys. B 650 (2003) 24, hep-ph/0208043.
[8] M. Carena, J.M. Moreno, M. Quiros, M. Seco, C.E.M. Wagner, Supersymmetric CP-violating currents and elec-
troweak baryogenesis, Nucl. Phys. B 599 (2001) 158, hep-ph/0011055.
[9] T. Konstandin, T. Prokopec, M.G. Schmidt, Kinetic description of fermion flavor mixing and CP-violating sources
for baryogenesis, Nucl. Phys. B 716 (2005) 373, hep-ph/0410135.
[10] M. Joyce, K. Kainulainen, T. Prokopec, The semiclassical propagator in field theory, Phys. Lett. B 468 (1999) 128,
hep-ph/9906411.
[11] L. Fromme, S.J. Huber, Top transport in electroweak baryogenesis, hep-ph/0604159.
[12] J.M. Moreno, M. Quiros, M. Seco, Bubbles in the supersymmetric standard model, Nucl. Phys. B 526 (1998) 489,
hep-ph/9801272.
[13] C. Panagiotakopoulos, K. Tamvakis, Stabilized NMSSM without domain walls, Phys. Lett. B 446 (1999) 224, hep-
ph/9809475.
[14] C. Panagiotakopoulos, K. Tamvakis, New minimal extension of MSSM, Phys. Lett. B 469 (1999) 145, hep-
ph/9908351.
[15] C. Panagiotakopoulos, A. Pilaftsis, Higgs scalars in the minimal non-minimal supersymmetric standard model,
Phys. Rev. D 63 (2001) 055003, hep-ph/0008268.
[16] A. Dedes, C. Hugonie, S. Moretti, K. Tamvakis, Phenomenology of a new minimal supersymmetric extension of
the standard model, Phys. Rev. D 63 (2001) 055009, hep-ph/0009125.
[17] S.A. Abel, S. Sarkar, P.L. White, On the cosmological domain wall problem for the minimally extended supersym-
metric standard model, Nucl. Phys. B 454 (1995) 663, hep-ph/9506359.
[18] A. Menon, D.E. Morrissey, C.E.M. Wagner, Electroweak baryogenesis and dark matter in the nMSSM, Phys. Rev.
D 70 (2004) 035005, hep-ph/0404184.
[19] D. Chang, W.F. Chang, W.Y. Keung, New constraint from electric dipole moments on chargino baryogenesis in
MSSM, Phys. Rev. D 66 (2002) 116008, hep-ph/0205084.
[20] A. Pilaftsis, Higgs-mediated electric dipole moments in the MSSM: An application to baryogenesis and Higgs
searches, Nucl. Phys. B 644 (2002) 263, hep-ph/0207277.
[21] M. Pietroni, The electroweak phase transition in a nonminimal supersymmetric model, Nucl. Phys. B 402 (1993)
27, hep-ph/9207227.
[22] A.T. Davies, C.D. Froggatt, R.G. Moorhouse, Electroweak baryogenesis in the next-to-minimal supersymmetric
model, Phys. Lett. B 372 (1996) 88, hep-ph/9603388.
[23] S.J. Huber, M.G. Schmidt, SUSY variants of the electroweak phase transition, Eur. Phys. J. C 10 (1999) 473, hep-
ph/9809506.
[24] M. Bastero-Gil, C. Hugonie, S.F. King, D.P. Roy, S. Vempati, Does LEP prefer the NMSSM?, Phys. Lett. B 489
(2000) 359, hep-ph/0006198.
[25] J. Kang, P. Langacker, T. Li, Electroweak baryogenesis in a supersymmetric U (1) model, Phys. Rev. Lett. 94 (2005)
061801, hep-ph/0402086.
[26] S.J. Huber, M.G. Schmidt, Electroweak baryogenesis: concrete in a SUSY model with a gauge singlet, Nucl. Phys.
B 606 (2001) 183, hep-ph/0003122;
S.J. Huber, P. John, M. Laine, M.G. Schmidt, CP violating bubble wall profiles, Phys. Lett. B 475 (2000) 104,
hep-ph/9912278.
[27] K. Funakubo, S. Tao, F. Toyoda, Phase transitions in the NMSSM, Prog. Theor. Phys. 114 (2005) 369, hep-
ph/0501052.
[28] D. Bodeker, P. John, M. Laine, M.G. Schmidt, The 2-loop MSSM finite temperature effective potential with stop
condensation, Nucl. Phys. B 497 (1997) 387, hep-ph/9612364.
[29] M. Laine, K. Rummukainen, Two Higgs doublet dynamics at the electroweak phase transition: A non-perturbative
study, Nucl. Phys. B 597 (2001) 23, hep-lat/0009025.
196 S.J. Huber et al. / Nuclear Physics B 757 (2006) 172–196

[30] LEP Collaborations, ALEPH Collaboration, DELPHI Collaboration, L3 Collaboration, OPAL Collaboration, Line
Shape Sub-Group of the LEP Electroweak Working Group, Combination procedure for the precise determination
of Z boson parameters from results of the LEP experiments, hep-ex/0101027.
[31] R. Harnik, G.D. Kribs, D.T. Larson, H. Murayama, The minimal supersymmetric fat Higgs model, Phys. Rev. D 70
(2004) 015002, hep-ph/0311349.
[32] A. Delgado, T.M.P. Tait, A fat Higgs with a fat top, JHEP 0507 (2005) 023, hep-ph/0504224.
[33] V. Cirigliano, S. Profumo, M.J. Ramsey-Musolf, Baryogenesis, electric dipole moments and dark matter in the
MSSM, hep-ph/0603246.
[34] P. Huet, A.E. Nelson, Electroweak baryogenesis in supersymmetric models, Phys. Rev. D 53 (1996) 4578, hep-
ph/9506477.
[35] M. Carena, M. Quiros, A. Riotto, I. Vilja, C.E.M. Wagner, Electroweak baryogenesis and low energy supersymme-
try, Nucl. Phys. B 503 (1997) 387, hep-ph/9702409.
[36] C. Lee, V. Cirigliano, M.J. Ramsey-Musolf, Resonant relaxation in electroweak baryogenesis, Phys. Rev. D 71
(2005) 075010, hep-ph/0412354.
[37] V. Cirigliano, M.J. Ramsey-Musolf, S. Tulin, C. Lee, Yukawa and tri-scalar processes in electroweak baryogenesis,
hep-ph/0603058.
[38] B.C. Regan, E.D. Commins, C.J. Schmidt, D. DeMille, New limit on the electron electric dipole moment, Phys.
Rev. Lett. 88 (2002) 071805.
[39] C.A. Baker, et al., An improved experimental limit on the electric dipole moment of the neutron, hep-ex/0602020.
[40] M. Pospelov, A. Ritz, Electric dipole moments as probes of new physics, Ann. Phys. 318 (2005) 119, hep-
ph/0504231.
[41] S.M. Barr, A. Zee, Electric dipole moment of the electron and of the neutron, Phys. Rev. Lett. 65 (1990) 21;
S.M. Barr, A. Zee, Phys. Rev. Lett. 65 (1990) 2920, Erratum.
[42] D. Chang, W.Y. Keung, A. Pilaftsis, New two-loop contribution to electric dipole moment in supersymmetric theo-
ries, Phys. Rev. Lett. 82 (1999) 900, hep-ph/9811202;
D. Chang, W.Y. Keung, A. Pilaftsis, Phys. Rev. Lett. 83 (1999) 3972, Erratum.
[43] G.D. Moore, T. Prokopec, How fast can the wall move? A study of the electroweak phase transition dynamics, Phys.
Rev. D 52 (1995) 7182, hep-ph/9506475;
G.D. Moore, T. Prokopec, Bubble wall velocity in a first order electroweak phase transition, Phys. Rev. Lett. 75
(1995) 777, hep-ph/9503296.
[44] T. Konstandin, S.J. Huber, Numerical approach to multi dimensional phase transitions, JCAP 0606 (2006) 021,
hep-ph/0603081.
[45] C. Grojean, G. Servant, J.D. Wells, First-order electroweak phase transition in the standard model with a low cutoff,
Phys. Rev. D 71 (2005) 036001, hep-ph/0407019.
[46] D. Bodeker, L. Fromme, S.J. Huber, M. Seniuch, The baryon asymmetry in the standard model with a low cut-off,
JHEP 0502 (2005) 026, hep-ph/0412366.
[47] M.E. Shaposhnikov, Structure of the high temperature gauge ground state and electroweak production of the baryon
asymmetry, Nucl. Phys. B 299 (1988) 797.
[48] D.N. Spergel, et al., astro-ph/0603449.
[49] K. Kainulainen, T. Prokopec, M.G. Schmidt, S. Weinstock, Semiclassical force for electroweak baryogenesis:
Three-dimensional derivation, Phys. Rev. D 66 (2002) 043502, hep-ph/0202177.
[50] J. Kang, P. Langacker, T.j. Li, T. Liu, Electroweak baryogenesis in a supersymmetric U (1) model, Phys. Rev.
Lett. 94 (2005) 061801, hep-ph/0402086.
[51] T. Ibrahim, P. Nath, The neutron and the lepton EDMs in MSSM, large CP violating phases, and the cancellation
mechanism, Phys. Rev. D 58 (1998) 111301, hep-ph/9807501;
T. Ibrahim, P. Nath, Phys. Rev. D 60 (1999) 099902, Erratum.
[52] F. del Aguila, M.B. Gavela, J.A. Grifols, A. Mendez, Specifically supersymmetric contribution to electric dipole
moments, Phys. Lett. B 126 (1983) 71;
F. del Aguila, M.B. Gavela, J.A. Grifols, A. Mendez, Phys. Lett. B 129 (1983) 473, Erratum.
Nuclear Physics B 757 (2006) 197–210

Complete Higgs mass dependence of top quark pair


threshold production to order ααs
D. Eiras, M. Steinhauser ∗
Institut für Theoretische Teilchenphysik, Universität Karlsruhe, 76128 Karlsruhe, Germany
Received 22 May 2006; accepted 4 September 2006
Available online 2 October 2006

Abstract
In this paper we consider the production of top quark pairs close to threshold and compute the dependence
on the Higgs boson mass to order ααs . This requires the evaluation of the matching coefficient of the
vector current to two-loop level and the inclusion of Higgs mass dependent operators in the non-relativistic
effective theory. For Higgs boson masses below 200 GeV moderate contributions to the top quark mass and
the peak cross section are observed. We also provide additional information to the on-shell wave function
renormalization which is relevant for the matching coefficient.
© 2006 Elsevier B.V. All rights reserved.

PACS: 14.65.Ha; 14.80.Bn

1. Introduction

One of the most important tasks of a future international linear collider (ILC) [1] is the precise
measurement of the cross section for the production of top quark pairs close to threshold. The
comparison with theoretical calculations will lead to a determination of the top quark mass with
an unrivaled precision leading to an uncertainty below 100 MeV [2]. Furthermore, also the width
of the top quark, the strong coupling and the Yukawa coupling between the top quark and Higgs
boson can be extracted from such a measurement. The findings of this paper are important in
this context. However, the final success of such an enterprise depends crucially on whether the
theoretical precision can match the expected experimental one. The complete next-to-next-to-
leading analysis of Ref. [3] has shown that it is important to obtain the third-order result within

* Corresponding author.
E-mail address: matthias.steinhauser@uka.de (M. Steinhauser).

0550-3213/$ – see front matter © 2006 Elsevier B.V. All rights reserved.
doi:10.1016/j.nuclphysb.2006.09.010
198 D. Eiras, M. Steinhauser / Nuclear Physics B 757 (2006) 197–210

QCD. This constitutes a long-term calculation which has already been started (see, e.g., Refs. [4,
5] for the most recent results).
The naive scaling rule α ∼ αs2 , where α is the fine structure and αs is the strong coupling con-
stant, shows that the next-to-next-to-next-to-leading terms are parametrically of the same order
as the mixed corrections proportional to ααs . In this paper we take the first step to a complete
order ααs result and compute the full dependence on the Higgs boson mass. On one hand this
requires the evaluation of two-loop vertex corrections within the Standard Model (SM) involving
a Higgs boson and a gluon. Furthermore, a new operator depending on the Higgs mass has to be
considered on the effective theory side. As we will explicitly see, the two individual pieces are
divergent, however, the physical quantities formed by the proper combination are finite.
Effects from the Higgs boson mass have already been studied in Ref. [6] (see also Ref. [7])
where a Yukawa-type potential has been used and the Schrödinger equation has been solved nu-
merically in order to obtain the imaginary part of the Green function and finally the total cross
section. In this paper we will consider the Higgs boson mass to be much larger than the soft
scales involved in the process and include it in a systematic way in the non-relativistic effective
theory. The importance of a systematic and careful treatment of electroweak corrections to the
top quark pair production has been stressed in Refs. [8,9], where certain next-to-next-to-leading
logarithmic electroweak effects associated to the instability of the top quark have been consid-
ered.
The production of top quark pairs in electron positron collisions is mediated both by the
exchange of a photon and a Z boson. Note, however, that in the threshold region the axial-vector
contribution of the Z boson exchange is suppressed by a factor v 2 ∼ αs2 as compared to the vector
contribution, where v is the velocity of the produced quarks. The latter can be obtained from the
photon-exchange diagrams in a straightforward way. In the case of electroweak corrections there
are contributions to the Z boson mediated diagrams which involve a coupling of the Z to the
Higgs boson. It has been shown in Ref. [10] that these contributions are numerically quite small
which is the reason why we do not include such diagrams in our analysis. Thus in this paper only
the photon-exchange diagrams are considered.
The paper is organized as follows: In the next section we consider the new Higgs mass de-
pendent operator included into non-relativistic QCD (NRQCD) and evaluate the corrections to
the energy level and the wave function at the origin. In Section 3 the Higgs mass dependent cor-
rections to the matching coefficient are discussed and Section 4 contains a brief account of the
phenomenological applications of our results. We conclude in Section 5. In Appendix A we dis-
cuss the results for the matching coefficients induced by vertex corrections involving a Z boson
and Appendix B contains additional information on the wave function renormalization constant
which is relevant in connection to the matching coefficient.

2. Effective theory

The standard theoretical framework for the threshold production of top quark pairs is given
by NRQCD [11,12] which ensures the automatic resummation of terms like (αs /v)n (n =
1, 2, 3, . . .) where v is the velocity of the produced quarks. NRQCD is constructed from QCD
by integrating out the hard scales associated to the mass of the heavy quark, m. Thus it contains
only degrees of freedom of the order mv and mv 2 .
In this paper we consider next to the top quark, which takes over the role of the heavy quark,
also the Higgs boson with mass Mh . Since Mh  mt v both mass scales are integrated out simul-
taneously. In Ref. [13] all operators which appear within QCD up to third order in perturbation
D. Eiras, M. Steinhauser / Nuclear Physics B 757 (2006) 197–210 199

theory have been classified. The only operator which has to be added to the Hamiltonian in
Eq. (6) of Ref. [13] is given by
απm2t
δHH = − 2 M2 M2
, (1)
sW W H
where sW is the sine of the Weinberg angle and MW is the W boson mass. The corresponding
expression in coordinate space is proportional to the delta function. If we employ the counting
rule α ∼ αs2 it is easy to see that δHH gives contributions which are parametrically of the same
order as the ones from the third-order QCD Hamiltonian.
Let us in a first step evaluate the corrections to the energy levels induced by the operator δHH .
Using first order perturbation theory we obtain
    ααs CF m4t
δEnH = ψnC δHH ψnC = EnC 2 2 M2
, (2)
2sW nMW h

with EnC = −CF2 αs2 mt /(4n2 ) and ψnC is the Coulomb wave function. n is the principal quantum
number.
In order to compute the correction to the wave function at the origin the operator of Eq. (1)
has to be inserted into the standard formulae of non-relativistic perturbation theory

δψn (0) = − d3 r GC (0, r, E)δHH ψnC (r ), (3)

where GC is the Coulomb Green function. Following Ref. [14] we split GC into a contribution
with zero, one and infinitely many gluon exchanges. Since only the one-gluon-exchange part is
divergent it is convenient to perform the corresponding calculation in momentum space whereas
the other contributions are evaluated in coordinate space. As a final result we obtain
  
ααs m4t 1 αs CF mt 3
δψn (0) = ψn (0) 2 2 2 CF − ln
H C
+ , (4)
sW MW Mh 4 μ 8
with |ψnC (0)|2 = CF3 αs3 m3t /(8πn3 ). The divergence in Eq. (4) has been subtracted minimally
which will also be done for the coefficient function considered in the next section.
We want to mention that the formulae of this section are adapted for the top quark. However,
they also apply to other quark masses, in particular to the bottom system, by simply exchanging
the top quark mass.

3. Higgs mass dependence of the matching coefficient

The matching coefficient of the vector current j μ = t¯γ μ t is defined by


 
1
jvk = cv φ † σ i χ + O , (5)
m2t
where φ and χ are two-component Pauli spinors for quark and anti-quark, respectively, and
k = 1, 2, 3 denote the spacial components. Note that there is no contribution to our order from
the time-component.
For the practical computation of cv it is convenient to consider the t t¯γ vertex in the limit
where for the photon energy, s, we have s ≈ 4m2t . In this case we can establish the equation
Z2 Γv = cv Z̃2 Z̃v−1 Γ˜v + · · · , (6)
200 D. Eiras, M. Steinhauser / Nuclear Physics B 757 (2006) 197–210

where we have on the left- and right-hand side quantities of the full and effective theory, re-
spectively. The latter are marked by a tilde and the ellipses denote terms suppressed by inverse
powers of the top quark mass. Γv denotes the t t¯γ vertex corrections where it is understood that
the couplings and masses are renormalized. The two-loop mixed correction to the on-shell wave
function renormalization, Z2 , has been computed recently in Ref. [15]. In particular, for the
Higgs boson contributions both the exact result and the expansions in three kinematic regions
have been provided.
In a next step the so-called threshold expansion [16,17] is applied to Eq. (6). In our case this
leads to four regions which have to be considered [16,17]: the hard, the soft, the potential and
the ultra-soft region. They are characterized by the scaling behaviour of the energy and three-
momentum component of the individual four-vectors of the particles. All contributions except
the hard one cancels between the left- and right-hand side of Eq. (6). As a consequence Γv has
to be evaluated for s = 4m2t . Furthermore, on the right-hand side only tree contributions have to
be considered. In particular we have Z̃2 = 1 since only scaleless integrals—which are set to zero
within dimensional regularization—contribute at loop level.
The one-loop corrections to cv are finite [10]. However, starting from two-loop order, the
matching coefficient exhibits infra-red divergences which are compensated by ultra-violet diver-
gences of the effective theory rendering physical quantities finite. In Eq. (6) the renormalization
constant Z̃v which generates the anomalous dimension of j˜v takes over this part. Note that the
vector current in the full theory does not get renormalized.
It is convenient to introduce the perturbative decomposition of the matching coefficient
 2
αs αs α ααs
cv = 1 + CF cv
QCD,1
+ CF cvQCD,2 + 2 cvew + 2
CF cvmix , (7)
π π πsW π 2 sW
where analog formulae also hold for Z2 and Γv . Since we consider in this paper only the contri-
bution from the Higgs boson the corresponding quantities obtain an additional superscript H .
The two-loop QCD corrections, cvQCD,2 , have been computed in Refs. [18,19] (see also
Ref. [20]). The complete SM corrections to one-loop order can be found in Ref. [10] where next
to the vertex corrections also the box diagrams contributing to e+ e− → t t¯ have been considered.
The latter, however, get no contributions from the Higgs boson.
From the perturbative expansion of Eq. (6) it is easy to see that the one-loop results cvQCD,1
and cvew are simply given by the sum of the one-loop expressions for Z2 and Γv . Whereas the
individual pieces are still divergent the proper sum is finite. For convenience we repeat the QCD
corrections and the Higgs mass dependent term of cvew which are given by

cvQCD,1 = −2,
 2
m2t 3yH − 1 2 − 9yH
2 + 12y 4 2 − 6y 4 )
(−2 + 5yH
cvH,ew
= 2 2
− 4
H
ln yH −
2
2
H
Ψ (yH ) , (8)
MW 12yH 48yH 24yH
where
⎧√ √

⎨ 4x
2 −1
arctan 4x 2 − 1 for x  12 ,
x2
Ψ (x) = √ √ (9)

⎩ 1−4x
2
ln 1−√1−4x
2
for x < 12 ,
2 2x 1+ 1−4x 2

and yH = mt /Mh .
D. Eiras, M. Steinhauser / Nuclear Physics B 757 (2006) 197–210 201

At two-loop order it is quite difficult to obtain a closed analytic result valid for all Higgs and
top quark masses. However, it is possible to get compact formulae valid in various kinematical
regions which—when combined—cover the whole Higgs mass range. This strategy has been
successfully applied in Ref. [15] to the on-shell wave function renormalization (see also Appen-
dix B). Thus, let us consider cvH,ew in the limits mt MH , mt ≈ MH and mt  MH where it is
given by
    2   2 
H,ew m2 31 5 3 ln yH 307 5 ln yH
cv,0 = 2t yH 2
− ln yH2
+ yH 4
− − + yH
6
− −
MW 144 24 16 4 480 8
 2 
737 11 ln yH
+ yH8
− − + ··· ,
360 6
    
H,ew m2 1 π 1 π 1 π
cv,1a = 2t + √ + yH,1a + √ + yH,1a 2
− + √
MW 6 8 3 24 8 3 12 6 3
   
1 13π 1 5π
+ yH,1a − +
3
√ + yH,1a 4
+ √ + ··· ,
36 108 3 288 54 3
    
H,ew m2 1 π 1 π 1 π
cv,1b = 2t + √ − yH,1b + √ + yH,1b 2
− + √
MW 6 8 3 24 8 3 8 24 3
   
13 19π 59 23π
− yH,1b
3
− √ + yH,1b − 4
+ √ + ··· ,
72 216 3 288 216 3
  
m2t π ln yH 23π 1 7 3 1 55π 1
cv,∞ = 2
H,ew
yH − − + + ln yH 2
+ 3
MW 4 2 96 yH 48 8 yH 512 yH
 
71 ln yH 1
+ − − 4
+ ··· , (10)
720 12 yH
H,ew H,ew
with yH,1a = (1−1/yH 2 ) and y
H,1b = (1−yH ). cv,1a and cv,1b are two different representations
2

of the same information which turn out to be useful for different values of yH .
In Fig. 1 the exact result for cvew /(m2t /MW2 ) (full black line, cf. Eq. (8)) is shown together
H,ew H,ew H,ew
with the expansions from Eq. (10), cv,0 (dash-dotted), cv,1a (short dashed), cv,1b (dotted) and
H,ew
cv,∞ (long dashed). One can see that the approximations nicely cover the whole region of yH .
Note that the expansion around mt = MH shows better convergence properties in case yH,1b is
used as a parameter.
Let us draw the attention of the reader to the expansion terms for large top quark masses which
starts with an enhancement factor m3t /(MW 2 M ). Two out of the three powers in m come from
h t
the Yukawa coupling between the Higgs boson and the top quark. Furthermore, the factor mt /Mh
is the indication of the Coulomb singularity which would be present for a massless Higgs boson.
The next-to-leading term is quadratic in mt accompanied by a logarithm in mt /Mh . Both terms
indicate that even for the leading terms it is not possible to nullify the Higgs boson mass and non-
trivial integration regions have to be considered which makes the evaluation of the corresponding
expansion terms quite tedious. In particular, this is true for the two-loop order where due to the
Coulomb divergence one will have a further factor mt /Mh as compared to the one-loop term.
Thus the expansion starts with √ a quartic top quark mass dependence. Furthermore, there are
momentum regions which have mt /MH as expansion parameter.
On the other hand, as can be seen in Fig. 1, the result from this region is phenomenologically
less important since for Higgs boson masses above 115 GeV there is perfect agreement between
202 D. Eiras, M. Steinhauser / Nuclear Physics B 757 (2006) 197–210

Fig. 1. cvH,ew /(m2t /MW


2 ) as a function of 1/y . The solid (black) line represents the exact result. The approximations in
H
the three regions are shown as dotted, dashed and dash-dotted lines where for the expansion around yH ≈ 1 two different
approximations have been chosen.

the exact result and the approximations one obtains for mt ≈ Mh and mt Mh . Higgs boson
masses below approximately 115 GeV are excluded by the direct search at the CERN Large
Electron Positron Collider (LEP). For this reason we do not consider the limit mt  Mh at two-
loop order.
Let us now turn to two loops. The Feynman diagrams are shown in Fig. 2. They are gener-
ated with QGRAF [21]. The application of q2e and exp [22,23] identifies the topology of the
individual diagrams and adopts the notation in order to match one of the following functions

J ± (n1 , n2 , n3 , n4 , n5 )

e2 γE dd k dd l
= ,
(iπ d/2 )2 (k 2 + 2kq)n1 (l 2 ± 2lq)n2 (k 2 )n3 ((k − l)2 )n4 (l 2 − M 2 )n5
HN± (n1 , n2 , n3 , n4 , n5 )

e2 γE dd k dd l
= ,
(iπ )d/2 2 (k + 2kq) 1 (l ± 2lq) 2 (k 2 )n3 ((k − l)2 − M 2 )n4 (l 2 )n5
2 n 2 n

YN± (n1 , n2 , n3 , n4 , n5 )

e2 γE d d k dd l
= ,
(iπ )d/2 2 (k + 2kq) 1 (l ± 2lq) 2 ((k − l)2 ∓ 2q(k − l))n3 (k 2 )n4 (l 2 − M 2 )n5
2 n 2 n
+
ZN (n1 , n2 , n3 , n4 , n5 )

e2 γE d d k dd l
= ,
(iπ d/2 )2 (k 2 + 2kq)n1 (l 2 + 2lq)n2 ((k − l)2 − 2q(k − l))n3 (k 2 − M 2 )n4 (l 2 )n5
(11)
where d = 4 − 2 is the space–time dimension and the ni are integer indices. The choice of the
five-line integrals of Eq. (11) is possible due to the special kinematics we have at hand. Note that
J + , H + and Y − correspond to self energies whereas J − , H − , Y + and Z + to vertex diagrams.
D. Eiras, M. Steinhauser / Nuclear Physics B 757 (2006) 197–210 203

Fig. 2. Two-loop Feynman diagrams contributing to cvH,mix .

In a next step we use the program AIR [24] in order to reduce the integrals to about 30
master integrals. They range from very simple two-point expressions up to complicated two-
scale integrals with five lines. At this point an asymptotic expansion in the various kinematical
regions is performed. From the one-loop result (cf. Fig. 1) and from the considerations in the
context of the on-shell wave function renormalization (see Ref. [15] and Appendix B) one can
see that for the phenomenological interesting Higgs boson masses it is sufficient to have the
expansion around mt ≈ MH and the one for mt MH at hand. Thus the master integrals are
expanded in these two limits. In the case mt ≈ MH this reduces to a simple Taylor expansion
whereas for mt MH the well-established hard-mass procedure [17] is applied. The latter is
actually automated in the program exp [22,23]. Hence as an independent check we applied with
the help of exp the hard-mass procedure to each diagram which immediately leads to simpler
expressions and avoids the reduction to the master integrals. In this way the result for mt Mh
could be checked. We want to mention that the calculation has been performed for general QCD
gauge parameter, ξ . The cancellation of ξ in the final result serves as a welcome check for our
calculation.
As already mentioned in the Introduction, even after the proper combination of Z2 and Γv
is formed, as prescribed by Eq. (6), there remains an infra-red divergence which is plugged
into Z̃v . We subtract this divergence in the MS scheme and introduce the anomalous dimension
γv = ddlnlnZ̃μv which is given by

ααs π 2 m4t
γv = − 2
C F 2 M2
. (12)
π 2 sW 4 MW h
204 D. Eiras, M. Steinhauser / Nuclear Physics B 757 (2006) 197–210

In the regions mt MH and mt ≈ MH three, respectively, six expansion terms have been
evaluated. Our results read
 2
H,mix m2t π m2 29 277π 2 π 2 ln 2 21ζ3 139
cv,0 = 2 ln 2t − − − − + 2
ln yH
MW 8 μ 216 2304 8 16 216
  
103 2 2 583 875π 2 151 17 2 2
− ln yH yH +2
− + ln yH −
2
ln yH yH 4
288 576 6912 192 16
 
1533691 27103π 2 66647 2251 2 2
+ − − ln yH −
2
ln yH yH + · · · ,
6
432000 138240 43200 720

H,mix m2 π 2 1 m2
cv,1a = 2t ln 2t − 5.760 − 5.533yH,1a − 5.704yH,1a2
− 5.888yH,1a
3
MW 8 1 − yH,1a μ

− 6.053yH,1a − 6.200yH,1a + · · · ,
4 5


H,mix m2t π 2 m2
cv,1b = 2
(1 − yH,1b ) ln 2t − 5.760 + 5.533yH,1b − 0.171yH,1b
2
+ 0.0124yH,1b
3
MW 8 μ

+ 0.0304yH,1b + 0.0296yH,1b + · · · ,
4 5
(13)

where mt is the on-shell mass and ζ3 ≈ 1.20206 is Riemann’s zeta-function. The ln(m2t /μ2 ) term
reminds on the divergence which has been subtracted minimally. Since in the limit mt MH
some coefficients of the -expansion of the master integrals could only be computed numerically
H,mix H,mix
the results for the finite parts of cv,1a and cv,1b are presented in numerical form. We want to
H,mix H,mix
stress once again that cv,1a and cv,1b contain the same information. However, expressed in
terms of yH,1b the convergence properties are much better.
In Fig. 3 the results are shown where the same notation as for the one-loop result has been
adopted and μ = mt has been chosen. Next to the highest expansion terms we also include
the lower-order terms as thin lines which nicely demonstrate the convergence properties in the
individual regions. Comparing the two parameterizations of the expansion around Mt = Mh one
observes that the one in terms of yH,1b demonstrates a much better convergence behaviour: the
thin dots in Fig. 3 are practically indistinguishable from the fat ones. This can also be seen in
Eq. (13) where the coefficients become quite small starting from the third term which is not the
H,mix H,mix H,mix
case for cv,1a . Thus, the combination of cv,1b and cv,0 provides a very good approximation
to cv for Higgs boson masses above Mh = 115 GeV.
H

4. Phenomenological application

Let us in a first step discuss the effect on the top quark mass. The connection between the
position of the peak in the threshold cross section, Eres , and the top quark mass is given by
p.t.
Eres = 2mt + E1 + δ Γt Eres , (14)
p.t.
where E1 is the perturbative part of the ground state energy and δ Γt Eres = 100 ± 10 MeV [25]
takes into account the effect of the finite width, the higher order resonances and the continuum.
p.t.
Non-perturbative effects are negligible for the top quark system. E1 up to third order in QCD
D. Eiras, M. Steinhauser / Nuclear Physics B 757 (2006) 197–210 205

Fig. 3. cvmix /(m2t /MW


2 ) as a function of 1/y . The approximations are shown as dotted, dashed and dash-dotted lines
H
where for the expansion around yH ≈ 1 two different parameterizations have been chosen. Lower-order terms are plotted
using thin lines. For the renormalization scale μ = mt has been adopted.

has been computed in Ref. [25]. The contribution from the Higgs boson is given in Eq. (2) where
αs has to be evaluated at the typical soft scale given by μs = CF αs (μs )mt ≈ 30 GeV.
p.t.
According to Eq. (14) a shift in E1 can directly be translated into a shift in the top quark
mass, δmt . Choosing for the input values mt = 175 GeV, MW = 80.41 GeV, sW 2 = 0.23,

α = 1/127 [26], αs (MZ ) = 0.118, MH = 120 GeV and μs = 30 GeV we observe for δmt ≈
−E1H /2 ≈ 26 MeV. This reduces to 9(1) MeV for MH = 200(500) GeV. If MH = 120 GeV is
adopted and μs = 15(60) GeV is chosen one obtain δmt ≈ 38(18) MeV. Thus, for light Higgs
masses relatively large effects are observed whereas for larger MH the numerical effect on mt is
small. Note that our findings are in agreement with the more qualitative analysis of Ref. [6] (see
also Ref. [7]).
As a second application it is interesting to consider the effect of the new corrections to the
normalized cross section R = σ (e+ e− → t t¯ )/σ (e+ e− → μ+ μ− ) at the resonance energy which
is dominated by the contribution from the would-be toponium ground-state. It is of the form
|ψ1 (0)|2
R1 = R1LO cv2 (mt ) + ··· (15)
|ψ1C (0)|2
with R1LO = 6πNc Q2t |ψ1C (0)|2 /(m2t Γt ). The ellipses in Eq. (15) denote contributions from
higher order operators. Note that the divergences in cv and ψ1 (0) cancel in the combination
of Eq. (15). The most advanced evaluation of R1 is provided in Ref. [4] where all logarithmi-
cally enhanced third-order corrections and the ones proportional to β03 are included. Let us for
completeness repeat the final numerical result which is given by
 (1) (2) 
R1 ≈ R1LO 1 − 0.244NLO + 0.449NNLO − 0.277N3 LO + δH + δH + · · · , (16)
where μs = 30 GeV has been chosen. The prime reminds that the N3 LO corrections are not
(1) (2)
complete and δH and δH parameterize the one- and two-loop corrections due to the Higgs
boson considered in the present paper. Using MH = 120/200/500 GeV and μs = 30 GeV we
(1) (2)
obtain δH = 0.067/0.034/0.009 and δH = 0.036/0.011/0.0002. Thus, moderate effects are
206 D. Eiras, M. Steinhauser / Nuclear Physics B 757 (2006) 197–210

observed for Higgs boson masses below about 200 GeV. However, in contrast to the pure QCD
(2)
effects of Eq. (16) the convergence seems to be much better as δH it is substantially smaller than
(1)
δH .
In principle it is possible to apply the formulae of this section also to the bottom system.
However, due to the suppression factor m2b /MH 2 the numerical effect is very small for Higgs

boson masses above 100 GeV.

5. Conclusions

In this paper we take an important step towards the evaluation of the mixed QCD/electroweak
corrections to the threshold production of top quark pairs. Within the framework of NRQCD
we consider all contributions involving the Higgs boson to the production process mediated by
a photon exchange. This includes a new operator to be introduced in the effective theory and
the two-loop corrections to the matching coefficients. For the latter, expansion terms have been
computed around the limits Mt ≈ MH and Mt MH . It has been argued that they provide a very
good approximation in the whole range of Higgs boson masses above 115 GeV. Thus, for any
practical purpose the complete Higgs boson mass dependence is available. Moderate numerical
effects on the top quark mass and the peak cross section are observed for Higgs boson masses
below 200 GeV. Considering the anticipated precision of an ILC it is certainly necessary to take
into account these corrections.

Acknowledgements

We would like to thanks J.H. Kühn, A.A. Penin, V.A. Smirnov and O.V. Tarasov for useful
discussions and A.A. Penin for carefully reading the manuscript. This work was supported by
the “Impuls- und Vernetzungsfonds” of the Helmholtz Association, contract number VH-NG-
008 and the SFB/TR 9.

Appendix A. Matching coefficients for the Z boson exchange diagrams

In the following we present the results for the matching coefficients induced by vertex cor-
rections involving the Z boson and the corresponding Goldstone boson, respectively. The calcu-
lation is similar to the diagrams involving a Higgs boson which is presented in the main part of
the paper. Note that there are also box diagrams involving the Z boson which are not considered
here. The one-loop result is given by [10]
 2
2 2 5yZ − 2 3yZ4 − 4yZ2 + 1 (yZ2 − 1)2
Z,ew
cv = at sW − ln yZ +
2
Ψ (yZ )
12yZ2 12yZ4 6yZ2

3y 2 + 2 1
+ vt sW − Z 2 −
2 2
ln yZ2
12yZ 12yZ4
 
7 1 yZ2 3
+ + + − Ψ (yZ ) , (A.1)
48 6yZ2 4 16(4yZ2 − 1)

with yZ = mt /Mz , at = 1/(2sW cW ), vt = (1/2 − 4sW 2 /3)/(s c ) and c
W W W = 2 =
1 − sW
MW /MZ . The function Ψ (x) can be found in Eq. (9). It is straightforward to obtain the expansion
D. Eiras, M. Steinhauser / Nuclear Physics B 757 (2006) 197–210 207

Fig. 4. (a) cvZ,ew and (b) cvZ,mix as a function of 1/yZ . The same notation as in Figs. 1 and 3 is adopted. The vertical line
indicates the physical value of 1/yZ = MZ /mt .

in the various kinematical regions. Thus we refrain from listing them explicitly. In Fig. 4(a) the
exact result is plotted together with the expansions where in each region the same depth is chosen
as in Eq. (10). The vertical dashed line marks the phenomenological result for yZ which is nicely
approximated both from the mt  MZ and the mt ≈ MZ region. However, it should be noted
that both parameterizations for the mt = Mh expansion become instable below 1/yZ ≈ 0.6.
At two-loop order we consider the diagrams of Fig. 2 where the Higgs boson is replaced by
the Z boson. The reduction of the occurring integrals is in close analogy and in fact the same
set of master integrals is necessary. Although the region mt MZ is phenomenologically not
relevant we nevertheless present the results since it constitutes an important cross checks for the
other kinematical limits. Furthermore, it is possible to obtain from this limit the result for the case
of the bottom quark. We adopt the notation from Eq. (7) and denote the contribution from the Z
boson as cvmix by cvZ,mix . The additional subscript “0”, “1a” or “1b” reminds on the kinematical
region considered. Our results read

Z,mix   2 π 2 m2t
cv,0 = vt2 + at2 sW ln 2
8 μ
 
587 131π 2 π 2 ln 2 21 28 11 2 2 2
+ vt2 sW
2
− + − − ζ3 + ln yZ2 − ln yZ yZ
432 576 8 16 27 72
 2 
403 205π 2299 5 2 2 4
+ − + + ln yZ2 − ln yZ yZ
64 576 864 18
 
21091951 3951π 2 28301 469 2 2 6
+ − + + ln yZ2 + ln yZ yZ + · · ·
864000 2560 7200 160
 2 2 
281 85π π ln 2 21 29 11 2 2 2
+ at2 sW
2
− + − − ζ3 + ln yZ2 − ln yZ yZ
432 192 8 16 54 72
 2 
709 11π 521 79 2 2 4
+ + + ln yZ2 − ln yZ yZ
576 432 288 72
 
260687 16013π 2 32453 5479 2 2 6
+ − + ln yZ −
2
ln yZ yZ + · · · ,
32000 69120 21600 1440
208 D. Eiras, M. Steinhauser / Nuclear Physics B 757 (2006) 197–210

Fig. 5. (a) One-loop, (b) 1/ pole and (c) constant part of the two-loop corrections to Z2V ,OS as a function of
1/yH = MH /mt . The solid (black) line represents the exact result. The approximations in the three regions are shown
as dotted, dashed and dash-dotted lines where for the expansion around yH ≈ 1 two different parameterizations (short
dashes and dots) have been chosen.

Z,mix   2 π 2 m2t 1 
cv,1a = vt2 + at2 sW ln 2 + vt2 sW2
−4.564 − 4.699yZ,1a − 5.009yZ,1a2
8 μ 1 − yZ,1a
 
− 5.277yZ,1a
3
− 5.502yZ,1a
4
− 5.693yZ,1a
5
+ · · · + at2 sW
2
−1.324 − 1.504yZ,1a

− 1.740yZ,1a − 1.930yZ,1a − 2.083yZ,1a − 2.212yZ,1a + · · · ,
2 3 4 5

Z,mix   2 π 2 m2t 
cv,1b = vt2 + at2 sW ln 2 (1 − yZ,1b ) + vt2 sW
2
−4.564 + 4.699yZ,1b − 0.311yZ,1b
2
8 μ
 
− 0.043yZ,1b
3
+ 0.001yZ,1b
4
+ 0.011yZ,1b
5
+ · · · + at2 sW
2
−1.324 + 1.504yZ,1b

− 0.236yZ,1b − 0.047yZ,1b − 0.011yZ,1b − 0.001yZ,1b + · · · ,
2 3 4 5
(A.2)
where the divergence has been subtracted in a minimal way and yZ = mt /MZ , yZ,1a = (1 −
1/yZ2 ) and yZ,1b = (1 − yZ2 ). In Fig. 4(b) the results of Eq. (A.2) are shown as a function in
Z,mix
1/yZ . Similarly to the two-loop case one observes a rapid convergence for cv,1b whereas the
Z,mix
magnitude of the coefficients of the higher order terms in cv,1a even increase. This gives us quite
D. Eiras, M. Steinhauser / Nuclear Physics B 757 (2006) 197–210 209

Z,mix
some confidence that cv,1b evaluated at the physical scale provides a very good approximation
to the unknown exact result. Using 1/yZ = 91.19/175 ≈ 0.52 we get

cvZ,mix μ2 =m2 ≈ −3.1 ± 0.3, (A.3)
t

where we assign a generous uncertainty of 10% which results from the comparison of lower-order
approximations. The accuracy of the result for cvZ,mix as given in Eq. (A.3) is more than sufficient
as far as the foreseen precision of the measurement of the threshold top quark production is
concerned.

Appendix B. Addendum to Ref. [15]

In this appendix we want to provide an update of the results of Ref. [15]. In Fig. 4 of that
reference the Higgs boson mass dependence of the on-shell wave function renormalization con-
stant to one- and two-loop order has been plotted as a function of 1/yH = MH /mt . Next to the
exact results also the ones for mt MH , mt ≈ MH and mt  MH have been shown and good
agreement over almost the whole range in yH has been found. However, there was a small gap
around 1/yH ≈ 2 which was not covered very well. This deficit is removed in Fig. 5 where the
result obtained in the limit mt ≈ MH is plotted both in terms of yH,1a = (1 − 1/yH 2 ) (dashed)

and yH,1b = (1 − yH 2 ) (dotted). One can see that the former is valid down to quite small values

of 1/yH whereas the validity of the latter expansion reaches up to larger values of 1/yH leading
to a significant overlap with the expansions for large and small values of yH . Thus, in the whole
yH range the simple expansions approximate the exact result to a very high precision.
Let us also mention that there is a misprint in the definition of vt in Ref. [15] after Eq. (17):
it is too small by a factor two. Furthermore, the analytic expression for Y in Eq. (22) has to be
multiplied by (−1) and in Eq. (27) a factor 1/sW 2 is missing in the terms containing α /π .
s

References

[1] See, e.g., http://linearcollider.org/.


[2] M. Martinez, R. Miquel, Eur. Phys. J. C 27 (2003) 49.
[3] A.H. Hoang, et al., Eur. Phys. J. C 2 (2000) 1, hep-ph/0001286.
[4] A.A. Penin, V.A. Smirnov, M. Steinhauser, Nucl. Phys. B 716 (2005) 303, hep-ph/0501042.
[5] M. Beneke, Y. Kiyo, K. Schuller, Nucl. Phys. B 714 (2005) 67, hep-ph/0501289.
[6] M.J. Strassler, M.E. Peskin, Phys. Rev. D 43 (1991) 1500.
[7] R. Harlander, M. Jezabek, J.H. Kühn, Acta Phys. Pol. 27 (1996) 1781, hep-ph/9506292.
[8] A.H. Hoang, C.J. Reisser, Phys. Rev. D 71 (2005) 074022, hep-ph/0412258.
[9] A.H. Hoang, C.J. Reisser, hep-ph/0604104.
[10] R.J. Guth, J.H. Kühn, Nucl. Phys. B 368 (1992) 38.
[11] W.E. Caswell, G.P. Lepage, Phys. Lett. B 167 (1986) 437.
[12] G.T. Bodwin, E. Braaten, G.P. Lepage, Phys. Rev. D 51 (1995) 1125, hep-ph/9407339;
G.T. Bodwin, E. Braaten, G.P. Lepage, Phys. Rev. D 55 (1997) 5853, Erratum.
[13] B.A. Kniehl, A.A. Penin, V.A. Smirnov, M. Steinhauser, Nucl. Phys. B 635 (2002) 357, hep-ph/0203166.
[14] B.A. Kniehl, A.A. Penin, M. Steinhauser, V.A. Smirnov, Phys. Rev. Lett. 90 (2003) 212001, hep-ph/0210161.
[15] D. Eiras, M. Steinhauser, JHEP 0602 (2006) 010, hep-ph/0512099.
[16] M. Beneke, V.A. Smirnov, Nucl. Phys. B 522 (1998) 321, hep-ph/9711391.
[17] V.A. Smirnov, Applied Asymptotic Expansions in Momenta and Masses, Springer-Verlag, Berlin, 2001.
[18] A. Czarnecki, K. Melnikov, Phys. Rev. Lett. 80 (1998) 2531, hep-ph/9712222.
[19] M. Beneke, A. Signer, V.A. Smirnov, Phys. Rev. Lett. 80 (1998) 2535, hep-ph/9712302.
[20] B.A. Kniehl, A. Onishchenko, J.H. Piclum, M. Steinhauser, Phys. Lett. B 638 (2006) 209, hep-ph/0604072.
[21] P. Nogueira, J. Comput. Phys. 105 (1993) 279.
210 D. Eiras, M. Steinhauser / Nuclear Physics B 757 (2006) 197–210

[22] R. Harlander, T. Seidensticker, M. Steinhauser, Phys. Lett. B 426 (1998) 125, hep-ph/9712228.
[23] T. Seidensticker, hep-ph/9905298.
[24] C. Anastasiou, A. Lazopoulos, JHEP 0407 (2004) 046, hep-ph/0404258.
[25] A.A. Penin, M. Steinhauser, Phys. Lett. B 538 (2002) 335, hep-ph/0204290.
[26] F. Jegerlehner, hep-ph/0105283.
Nuclear Physics B 757 (2006) 211–232

Verification of bootstrap conditions for amplitudes


with quark exchanges in QMRK ✩
A.V. Bogdan, A.V. Grabovsky
Institute for Nuclear Physics, 630090 Novosibirsk, Russia
Novosibirsk State University, 630090 Novosibirsk, Russia
Received 11 June 2006; received in revised form 8 September 2006; accepted 8 September 2006
Available online 26 September 2006

Abstract
The compatibility of the multi-Regge form of QCD amplitudes in the quasi-multi-Regge kinematics
(QMRK) and the s-channel unitarity imposes some constraints on the effective jet-production vertices. We
demonstrate that these constraints known as bootstrap conditions are satisfied for the amplitudes with the
Reggeized quark exchanges.
© 2006 Elsevier B.V. All rights reserved.

1. Introduction

This article continues the development of the quark Reggeization theory [1] in QCD. A no-
ticeable progress has been recently achieved here, in particular, the quark Regge trajectory in the
next-to-leading approximation (NLA) in D dimensions was found [2,3] and the next-to-leading
order (NLO) corrections to the effective vertices appearing in the leading logarithmic approxima-
tion (LLA) were calculated [4,5]. All these results were obtained assuming the Reggeized form
for amplitudes in the multi-Regge kinematics (MRK) in the NLA. It is clear that this assumption,
called the quark Reggeization hypothesis, must be proved. However, in the NLO it is tested only
in αs2 order [2,3] so far. Moreover, its complete proof in the LLA for any quark–gluon inelastic
process in all orders of αs was given only recently [6]. This proof is based on the relations re-
quired by compatibility of the multi-Regge form of QCD amplitudes with the s-channel unitarity
(bootstrap relations). The fulfillment of the bootstrap relations is secured by several conditions


Work supported in part by the Russian Fund of Basic Researches, project 04-02-16685-a, and in part by the Dynasty
Foundation and Russian Science Support Foundation.
E-mail addresses: a.v.bogdan@inp.nsk.su (A.V. Bogdan), grab@nsunet.ru (A.V. Grabovsky).

0550-3213/$ – see front matter © 2006 Elsevier B.V. All rights reserved.
doi:10.1016/j.nuclphysb.2006.09.004
212 A.V. Bogdan, A.V. Grabovsky / Nuclear Physics B 757 (2006) 211–232

(bootstrap conditions) on Reggeon vertices and trajectories. An analogous proof can (and has to)
be constructed in the NLA as well.
The only kinematics essential in the LLA is MRK, which means that all particles produced in a
high-energy process have limited transverse momenta and are well separated in rapidity space. In
the NLA, production amplitudes in this kinematics can be obtained by taking one of the effective
vertices or Regge trajectory in the NLO. But in the NLA another, quasi-multi-Regge kinematics
(QMRK) becomes also important. In this case one of the produced particles is replaced by a
jet containing two particles with similar rapidities. At this moment all multi-particle Reggeon
vertices required in the NLA are obtained [7]. Therefore, a proof of the quark Reggeization
hypothesis concerning the QMRK may be given.
In this paper we prove the quark Reggeization in the QMRK. Our method is the direct con-
tinuation of the one used to prove this hypothesis in the LLA. The bootstrap conditions for
amplitudes in the QMRK are the same as in the LLA with the substitution of jet production ver-
tices for particle production ones. Hereafter we demonstrate that all these conditions are fulfilled.
For convenience we work in the operator formalism, which was introduced in [8], developed in
another form in [9] and extended to inelastic amplitudes and quark exchanges in [6].
The paper is organized as follows. The next section contains all necessary denotations and the
definition of the QMRK. Section 3 presents the bootstrap conditions in operator formalism. In
Section 4 we prove these conditions for impact factors and Reggeon–Reggeon–jet (RRJ) effective
vertices. Section 5 concludes the paper.

2. Quasi-multi-Regge form of QCD amplitudes

Considering the QMRK we talk about a multiparticle production amplitude as about the
amplitude of jet production where one of the jets consists of two particles. Such a jet can be
produced either in the fragmentation regions of initial particles, or in the central region, i.e.
with the rapidity far away from the rapidities of colliding particles. Let us consider the process
A + B → A + P1 + · · · + Pn + B  in the QMRK. Using the same denotations as in [6] we
introduce light-cone momenta n1 and n2 : n21 = n22 = 0, (n1 n2 ) = 1 and denote (pn2 ) ≡ p + ,
(pn1 ) ≡ p − , so that pq = p + q − + p − q + + p⊥ q⊥ . Here the sign ⊥ means transverse to the
(n1 , n2 ) plane components. We assume that initial momenta pA and pB have predominant
components along n1 and n2 , respectively. For generality we do not demand that transverse
components pA⊥ and pB⊥ should be zero, but assume |pA⊥ 2 | ∼ |p 2 | ∼ p 2 ∼ p 2  p + p −
B⊥ A B A B
+ −
and remain limited (not grow) at pA pB → ∞. For the final jet momenta pi , i = 0, . . . , n + 1,
we assume the MRK conditions:
p0−  p1−  · · ·  pn−  pn+1

,
+
pn+1  pn+  · · ·  p1+  p0+ , (1)
where pi⊥ are limited. It ensures that the squared invariant masses sij = (pi + pj )2 are large
compared with the squared transverse momenta and invariant masses of the jets. At i < j they
have the form
pi+   pj−  
sij ≈ 2pi+ pj− = pj2 − pj2⊥ = pi2 − pi⊥
2
, (2)
pj+ pi−
and at i < l < j submit to relations
 
sil slj ≈ sij pl2 − pl⊥
2
. (3)
A.V. Bogdan, A.V. Grabovsky / Nuclear Physics B 757 (2006) 211–232 213

For the momentum transfers qi , i = 1, . . . , n + 1,

q1 = p0 − pA , qj +1 = qj + pj (j = 1, . . . , n), (4)
we have

qi2 ≈ q⊥
2
. (5)
In the LO the amplitude A2→n+2 of the process A + B → A + P1 + · · · + Pn + B has the
multi-Regge form
ω
R1 s1ω1 P1 s2ω2 n+1
sn+1 R
AR
2→n+2 = Γ A A γR 1 R 2 Pn
· · · γR n Rn+1 d
ΓB  Bn+1 , (6)
d1 d2 n+1

where Γ R R
A A and ΓB  B are the particle–particle–Reggeon (PPR) effective vertices, describing
particle–particle P → P  transitions due to interaction with Reggeons R. For gluon quantum
numbers in qi channel, ωi = ωG (qi ) is the gluon Regge trajectory and di ≡ di (qi ) = qi⊥ 2 ; for
Pi
quark numbers, ωi = ωQ (qi ) is the quark Regge trajectory and di ≡ di (qi ) = m − q̂i⊥ . γR i Ri+1
are the Reggeon–Reggeon–particle (RRP) effective vertices describing production of particle
Pi at Reggeon transitions Ri+1 → Ri . In order to be definite we do not consider here anti-
quark quantum numbers in any of qi channels. It determines the order of the multipliers in (6).
Nonetheless, our consideration does not lose generality because amplitudes with quark and anti-
quark exchanges are related by charge conjugation.
Since we come to the QMRK replacing one of the particles Pi in the MRK with a pair with
fixed invariant mass, QMRK amplitudes have the same form (6) as LO MRK ones, where one
{P1 P2 }
P
of the vertices γRR or ΓPR P is substituted with the jet production vertex γRR R
or Γ{P  
1 P2 }P
respectively. Note, that because the QMRK leads to the loss of a large logarithm in the unitarity
relations, energy scales in (6) are unimportant in the NLA. Moreover, we need trajectories and
vertices only in the LO there. Assuming similar ordering of longitudinal components one can
obtain the more general multi-jet amplitudes AR R
2+n1 →2+n2 from A2→n+2 by usual crossing rules.
Note, that as in (6) we can neglect imaginary parts of these amplitudes since in the QMRK they
are next-to-next-to-leading. Therefore, as well as for the amplitudes in the LO, crossing rules
connecting the QMRK amplitudes do not affect the Regge factors siωi .
Hereafter we work in the physical light-cone gauge
(ep)⊥
(ep) = (en1 ) = 0, e = e⊥ − n1 , (7)
p−
where e is the polarization vector of a gluon with momentum p.
We use the PPR vertices in the LO in this gauge from [6]:
− G  ∗ 
ΓGG G = −2gpG TG G eG eG⊥ , ΓQG Q = g ūQ t G γ − uQ ,

+ G  ∗ 
ΓGG G = −2gpG TG G eG eG⊥ , ΓG G +
Q Q = g ūQ t γ uQ ,


ΓQ̄G Q̄ = −g ῡQ̄ t G γ − υQ̄ , ΓG


Q̄ Q̄
= −g ῡQ̄ t G γ + υQ̄ , (8)
 ∗
ΓGQ Q = −gt G êG  ⊥ uQ , ΓQ̄Q G = −gt G êG⊥ υQ̄ ,

ΓQ
G Q̄

= −g ῡQ̄ t G êG ⊥, ΓQ Q G = −g ūQ t êG⊥ .

G
(9)
214 A.V. Bogdan, A.V. Grabovsky / Nuclear Physics B 757 (2006) 211–232

Here we denote particles and Reggeons by symbols which accumulate all their quantum numbers.
We use the letter P for particles (jets) and the letter R for Reggeons independently of their nature,
letters G and Q for ordinary gluons and quarks and G and Q for the Reggeized ones. In the gauge
(7) the RRP vertices for gluon, quark and antiquark production with momentum p = q2 − q1 at
the transition of Reggeon R2 (with momentum q2 ) to Reggeon R1 (with momentum q1 ) are as
follows [6]:
 2 
∗ q2⊥
γG1 G2 = 2gTG1 G2 e⊥ q2⊥ − pG⊥ 2
G G
, (10)
pG⊥
 
G ∗ pG⊥
G
γQ 1 Q2
= −gt e ⊥ γ ⊥ − 2(q̂ 2⊥ − m) 2
, (11)
pG⊥
q̂1⊥ q̂2⊥
γG1 Q2 = g ūQ + t G1 , γQ1 G2 = −gt G2 − υQ̄ .
Q Q̄
(12)
pQ pQ̄

The particle–jet–Reggeon (PJR) and Reggeon–Reggeon–jet (RRJ) effective vertices taken from
[10] and [7] can be presented in different forms and our goal is to find the presentation in which
subsequent calculations become trivial. Taking this in mind we introduce a set of functions Fi ,
Ki , Vi (see below Eqs. (20)–(33)) which determine all the PJR and RRJ effective vertices. The
selection of these functions is a nontrivial task and a result of the analysis of cancellations during
the verification of all QMRK bootstrap conditions.
Firstly, we introduce xi = ki− /k − , with ki being the momentum of the final particle and k =
k1 + k2 the momentum of the jet, so x1 + x2 1, we also use v = x2 k1⊥ − x1 k2⊥ . The commonly
arising denominators may be expressed via

D(p, q) = x1 p⊥
2
+ x2 q⊥
2
and d(p, q) = (x1 p⊥ − x2 q⊥ )2 . (13)
We rewrite all RRP and PJR effective vertices in terms of the functions Fi , Ki , Vi by means of
the following identities
1 x2 x1
= + , (14)
2 (k 2
k2⊥ 1⊥ − m2 ) k2⊥ (D(k2 , k1 ) − x2 m2 )
2 2
(k1⊥ − m )(D(k2 , k1 ) − x2 m2 )
2
x 1 x2
(d(k2 , k1 ) − x2 m2 )(D(k2 , k1 ) − x2 m2 )
2
 
1 1 1
= − , (15)
(k1 + k2 )2⊥ − m2 d(k2 , k1 ) − x22 m2 D(k2 , k1 ) − x2 m2
x12 k2⊥
2
− x22 k1⊥
2
= (x1 − x2 )d(k2 , k1 ) − 2x1 x2 (k1 + k2 , v)⊥
 2 
= x1 x2 k2⊥ − k1⊥
2
+ (x1 − x2 )D(k2 , k1 ), (16)
(e1 , v)⊥ (e2 k2 )⊥ + (e2 , v)⊥ (e1 k1 )⊥
= x2 (e2 , k1 + k2 )⊥ (e1 k1 )⊥ − x1 (e1 , k1 + k2 )⊥ (e2 k2 )⊥
= x1 (e2 , v)⊥ (e1 , k1 + k2 )⊥ + x2 (e1 , v)⊥ (e2 , k1 + k2 )⊥ , (17)
 
k̂1⊥ + m μ μ k̂2⊥ + m
υk2 = 2k − ūk1 γ⊥ υk2 ,
μ
ūk1 n̂1 γ⊥ + γ⊥ (18)
x1 x2
 
k̂1⊥ + m μ μ k̂2⊥ − m
uk2 = 2k − ūk1 γ⊥ uk2 ,
μ
ūk1 n̂1 γ⊥ + γ⊥ (19)
x1 x2
A.V. Bogdan, A.V. Grabovsky / Nuclear Physics B 757 (2006) 211–232 215

where e1 and e2 are the polarization vectors of emitted gluons. Our functions are:
μ  μ 
F1 (k2 , k1 ) = ê1⊥ 2x1 k2⊥ − x2 (k̂1⊥ + m)γ μ ,
μ μ
F̃1 (k2 , k1 ) = F1 (k2 , k1 )|1↔2 , (20)
μ  
F2 (k2 , k1 ) = γ μ 2x1 (e2 , v)⊥ + x2 ê2⊥ (v̂ + x2 m) , (21)
μ  μ 
F3 (k2 , k1 ) = ê1⊥ 2v μ − x2 γ⊥ (v̂ − x2 m) , (22)
μ  μ 
F̃3 (k2 , k1 ) = ê2⊥ −2v μ + x1 γ⊥ (v̂ + x1 m) , (23)
μ  
F4 (k2 , k1 ) = ê2 γ μ (v̂ + m) − 2x1 v μ , (24)
μ  
F̃4 (k2 , k1 ) = ê1 −γ μ (v̂ − m) + 2x2 v μ , (25)
F5 (k2 , k1 ) = (k̂1⊥ + m)(k̂2⊥ + m), (26)
μ
μ 2x1 ê1⊥ (k̂1⊥ + m)k2⊥
K1 (k2 , k1 ) = , (27)
(D(k2 , k1 ) − x2 m2 )(k1⊥
2 − m2 )

μ μ
K̃1 (k2 , k1 ) = K1 (k2 , k1 )|1↔2 , (28)
μ
μ 2x2 ê1⊥ (k̂1⊥ + m)k2⊥ μ μ
K2 (k2 , k1 ) = , K̃2 (k2 , k1 ) = K2 (k2 , k1 )|1↔2 , (29)
(D(k2 , k1 ) − x2 m2 )k2⊥
2

μ μ
μ 4x1 k1⊥ (e2 , k2 )⊥ μ 4x2 k1⊥ (e2 , k2 )⊥
K3 (k2 , k1 ) = 2
, K̃3 (k2 , k1 ) = 2
, (30)
D(k2 , k1 )k1⊥ D(k2 , k1 )k2⊥
μ μ μ
V1 (k2 , k1 ) = 2x1 x2 (e1 e2 )⊥ v μ − 2x1 (e2 , v)⊥ e1⊥ − 2x2 (e1 , v)⊥ e2⊥ , (31)
μ μ  2 
e1⊥ V2 (k2 , k1 ) = x1 x2 (e1 e2 )⊥ k2⊥ − k1⊥
2
+ 2(e1 , v)⊥ (e2 k2 )⊥
+ 2(e2 , v)⊥ (e1 , k1 )⊥ , (32)
μ μ μ
V3 (k2 , k1 ) = 2x1 x2 ê2⊥ v μ − 2x1 (e2 , v)⊥ γ⊥ − 2x2 v̂⊥ e2⊥ , (33)

where 1 ↔ 2 means k1 ↔ k2 , x1 ↔ x2 , e1 ↔ e2 .
Notice that
μ μ μ μ
e1⊥ F̃3,4 (k2 , k1 ) = e2⊥ F3,4 (k2 , k1 )|1↔2 (34)

and
μ μ μ μ
e1 K̃3 (k2 , k1 )|1↔2 = e1 K3 (k2 , k1 ). (35)
μ μ
The functions V1 (k2 , k1 ), e1⊥ V2 (k2 , k1 ) are antisymmetric under (1 ↔ 2) replacement. One can
μ μ
see that V3 (k2 , k1 ) equals V1 (k2 , k1 ), where e1⊥ is changed to γ⊥ . We also need some of these
functions with the substitution e1 → n1 . We mark them with a “−” sign, e.g.:
−μ  
F̃4 (k2 , k1 ) = n̂1 −γ μ (v̂ − m) + 2x2 v μ . (36)

Now we are ready to present the PJR effective vertices describing the transition of a particle with
momentum k + q to a jet with momentum k = k1 + k2 and a Reggeon carrying momentum q
216 A.V. Bogdan, A.V. Grabovsky / Nuclear Physics B 757 (2006) 211–232

(everywhere below (1 ↔ 2) means in addition G1 ↔ G2 ):


 μ μ
Q G2 G1 γ⊥ V1 (k2 , k1 )
Γ{G1 G2 }Q
=g 2
t t
d(k2 , k1 )
μ μ 
e2⊥ F3 (k2 , k1 + q)
− t G1 t G2 + (1 ↔ 2) uQ , (37)
d(k2 , k1 + q) − x22 m2
 μ
G 2 − μ V1 (k2 , k1 )
Γ{G 1 G2 }G
= g 2k e T a
T a
GG G1 G2
2d(k2 , k1 )
μ 
  V1 (k2 + q, k1 )
+ T GT G G + (1 ↔ 2) , (38)
1 G2 d(k + q, k )
2 1

where eμ is the polarization vector of the incoming gluon.


 −μ −μ
G F3 (k2 , k1 )   F3 (k2 + q, k1 )
t Gt G + tGtG
μ
Γ{Q 1 G}Q
= −g 2 e2⊥ ūQ1
d(k2 , k1 ) − x22 m2 d(k2 + q, k1 ) − x22 m2
−μ 
F3 (k2 , k1 + q)
− tGtG uQ , (39)
d(k2 , k1 + q) − x22 m2
μ
where k1 and k2 are momenta of the emitted quark and gluon and e2 is the polarization vector of
the outgoing gluon.
 μ μ
F4 (k2 + q, k1 ) F2 (k2 , k1 )
Γ{Q
Q̄G2 }G
= −g e t t 2 μ G2 G
+ t G t G2
d(k2 + q, k1 ) − m2 d(k2 , k1 ) − x22 m2

  V μ (k2 , k1 + q)
+ t G t G2 3 υ , (40)
x1 d(k2 , k1 + q) Q̄
where eμ is the polarization vector of the incoming gluon and k1 and k2 are momenta of the
emitted antiquark and gluon.
−μ
Q g2 μ t a F̃4 (k2 , k1 )
Γ{Q = − − t a γ⊥ uQ ⊗ ūQ1 υ
1 Q̄2 }Q 2k d(k2 , k1 ) − m2 Q̄2
−μ
g2 a μ t a F3 (k2 + q, k1 )
+ − t γ⊥ υQ̄2 ⊗ ūQ1 uQ , (41)
2k2 d(k2 + q, k1 ) − x22 m2
where k1 and k2 are momenta of the emitted quark and antiquark.
 −μ −μ
G F̃4 (k2 + q, k1 )
G G F̃ (k2 , k1 + q)
Γ{Q = g e ūQ −t t
2 μ
+ tGtG 4
1 Q̄2 }G d(k2 + q, k1 ) − m2 d(k2 , k1 + q) − m2

  F̃4−μ (k2 , k1 )
+ t Gt G υ , (42)
d(k2 , k1 ) − m2 Q̄
where eμ is the polarization vector of the incoming gluon and k1 and k2 are momenta of the emit-
ted quark and antiquark. Quite analogously, the RRJ effective vertices describing the production
of a jet {P1 P2 } with momentum k = k1 + k2 at the collision of two Reggeons with momenta −q1
A.V. Bogdan, A.V. Grabovsky / Nuclear Physics B 757 (2006) 211–232 217

and q2 look as follows


μ   2 μ 
{G1 G2 } V1 (k2 , k1 ) q2⊥ (k1 + k2 )⊥ μ
γG1 G2 =g 2
TGa 2 G1 TGa2 G1 − q2⊥
d(k2 , k1 ) (k1 + k2 )2⊥
μ μ
e1⊥ V2 (k2 , k1 )q2⊥2  G G
 μ
2V2 (q2 − k1 , k1 )
μ
+ +g T T
2 1 2
G2 G1 e1⊥ − D(q − k , k )
D(k2 , k1 )(k1 + k2 )2⊥ 2 1 1

 μ μ 
+ q2⊥
2
K3 (k2 , k1 ) − K3 (q2 − k1 , k1 ) + (1 ↔ 2),
μ μ μ μ 
{G1 G2 }   e1⊥ V2 (k2 , k1 ) (k1 + k2 )⊥ V1 (k2 , k1 ) q̂2⊥ − m
γQ1 Q = g 2 t G1 t G2 +
2 D(k2 , k1 ) d(k2 , k1 ) (k1 + k2 )2⊥
μ μ  μ μ
γ V (k2 , k1 ) e2⊥ F1 (k2 , q2 − k2 )
− ⊥ 1 + g 2 t G1 t G2
2d(k2 , k1 ) D(k2 , q2 − k2 ) − x2 m2

 μ μ μ μ 
+ e2⊥ K2 (k2 , q2 − k2 ) − e1⊥ K̃3 (k2 , k1 ) (q̂2⊥ − m) + (1 ↔ 2), (43)

where e1 , e2 are the polarization vectors of the emitted gluons.


 μ 
{Q̄G}   μ
2 G G γ⊥ 2V2 (k2 , q2 − k2 ) μ
γQG = g t t − + q2⊥ K̃3 (k2 , q2 − k2 )
2
2k1− D(k2 , q2 − k2 )
 μ μ  
g 2 G G q2⊥ F2 (k2 , k1 ) G G F5 (k1 , q2 − k1 )
− − t t − t t ê2⊥ +1 , (44)
k d(k2 , k1 ) − x22 m2 D(q2 − k1 , k1 ) − m2
where k1 and k2 are momenta of the emitted antiquark and gluon and e2 is the emitted gluon
polarization;
  −μ
{QG}   F1 (q2 − k1 , k1 ) −μ
γGQ = g 2 e2⊥ ūQ t G t G
μ
− K (q 2 − k ,
1 1k )(q̂ 2⊥ − m)
D(q2 − k1 , k1 ) − x2 m2 1
−μ −μ
F1 (k2 , q2 − k2 ) F3 (k2 , k1 )
− tGtG − t Gt G
D(k2 , q2 − k2 ) − x2 m 2 d(k2 , k1 ) − x22 m2
 −μ −μ −μ 
+ t G t G K1 (k2 , k1 ) + K2 (k2 , k1 ) − K2 (k2 , q2 − k2 ) (q̂2⊥ − m)
 −μ −μ 
F3 (k2 , k1 ) F1 (k2 , k1 ) 1
+ t Gt G + (q̂2⊥ − m)
d(k2 , k1 ) − x22 m2 D(k2 , k1 ) − x2 m2 k̂1⊥ + k̂2⊥ − m

G G −μ
− t t K1 (k2 , k1 )(q̂2⊥ − m) , (45)

where k1 and k2 are momenta of the emitted quark and gluon and e2 is the emitted gluon polar-
ization;
 
{QQ̄} g2 n̂1 F5 (k2 , q2 − k2 ) G2 G1 n̂1 F5 (q2 − k1 , k1 )
γG1 G2 = − ūQ t G1 t G2 − t t υ
k D(k2 , q2 − k2 ) − m2 D(q2 − k1 , k1 ) − m2 Q̄

g 2  G2 G1  2
q2⊥ n̂1 F5 (k2 , k1 )
+ − ūQ t t n̂1 +
k (k1 + k2 )⊥ 2 D(k2 , k1 ) − m2
μ −μ   μ −μ 
(k1 + k2 )⊥ F̃4 (k2 , k1 ) g 2  G2 G1  q2⊥ F̃4 (k2 , k1 )
− υQ̄ + − ūQ t t −n̂1 + υQ̄ ;
d(k2 , k1 ) − m2 k d(k2 , k1 ) − m2
218 A.V. Bogdan, A.V. Grabovsky / Nuclear Physics B 757 (2006) 211–232

 −μ 
{QQ̄} g2 a μ F1 (q2 − k1 , k1 ) μ−
γQ1 Q2 = − − t γ⊥ υQ̄ ⊗ ūQ t a
− K1 (q2 − k1 , k1 )(q̂2⊥ − m)
2k2 D(q2 − k1 , k1 ) − x2 m2

q̂2⊥ − m n̂1 F5 (k2 , k1 )
− g2t a ⊗ ūQ t a n̂1 +

k (k1 + k2 )⊥ 2 D(k2 , k1 ) − m2
μ −μ 
(k1 + k2 )⊥ F̃4 (k2 , k1 )
− υQ̄
d(k2 , k1 ) − m2
−μ
μ F̃4 (k2 , k1 )
− g 2 t a γ⊥ ⊗ ūQ t a υ ,
2k − (d(k2 , q2 − k2 ) − m2 ) Q̄
where k1 and k2 are momenta of the emitted quark and antiquark.

3. Bootstrap conditions in QMRK

In this article we prove a part of the quark Reggeization hypothesis in the NLA for the QMRK
or in other words the multi-Regge form (6) for the QMRK. Here it seems sensible to make two re-
marks concerning “Reggeization” and “signaturization”. These notions were carefully discussed
in [6] for the case of the LLA. In the QMRK all conclusions are valid as well with the corre-
sponding substitution of “jets” for “particles”, so the reader is referred to the end of Section 2 of
[6].
The proof of the form (6) in the QMRK may be done by obvious extension of the corre-
sponding argumentation presented in [6] for the LLA. It is based on the relations required by the
compatibility of the quasi-multi-Regge form of the QCD amplitudes with the s-channel unitarity
(bootstrap relations). Fulfillment of the bootstrap relations impose some constraints on the Regge
vertices and the trajectory (bootstrap conditions). The bootstrap conditions are formulated in this
Section and checked in the next one.
It is worth mentioning that besides all, the argumentation in [6] uses the amplitude in Born
approximation to calculate loop-by-loop all radiative corrections to Born amplitudes and examine
the formula (6). Therefore, the Born form for the QMRK has to be proved first. It is verified for
{P1 P2 } R
the corresponding tree amplitudes from which the vertices γRR and Γ{P   were extracted.
1 P2 }P
For the amplitude with an arbitrary number n of emitted particles the proof may be performed
via the t-channel unitarity as it was done in [1,11] for the MRK.
In order to present the bootstrap conditions in a compact way similar to one in [6] we use
operator formalism as in [6] slightly adopting it for jet production. So, Gi | and |Gi  are “bra”-
and “ket”-vectors for t-channel states of the Reggeized gluon with transverse momentum ri⊥ and
color index Gi . Their scalar product is

Gi |Gj  = ri⊥


2
δ(ri⊥ − rj ⊥ )δGi Gj . (46)
Similarly, we introduce Qi | and |Qi  denoting the t -channel states of the Reggeized quark with
transverse momentum ri⊥ , color index Qi and spinor index ρi and their scalar product

Qi |Qj  = (m − r̂i⊥ )ρi ρj δ(ri⊥ − rj ⊥ )δQi Qj . (47)


Two-Reggeon states are built from the above ones. It is useful to distinguish the states |Ri Rj 
(the corresponding “bra”-vector Ri Rj |) and |Rj Ri  (“bra”-vector Rj Ri |). We associate the
first of them with the case when Reggeon Ri is located in the lower part of Fig. 1, i.e. when it
belongs to ARAB→n+2 in the unitarity relation, and the second with the case when it is in the upper
A.V. Bogdan, A.V. Grabovsky / Nuclear Physics B 757 (2006) 211–232 219

Fig. 1. Schematic representation of the s-channel discontinuity discS


s AAB→A B  .

part of Fig. 1, i.e. in the amplitude AR


n+2→A B  . We define three types of states
|Gi Gj  = |Gi |Gj , |Gi Qj  = |Gi |Qj , |Qi Gj  = |Qi |Gj . (48)
States of different types are orthogonal one another. All of them create a complete set, i.e.

d D−2 r1⊥ d D−2 r2⊥


Ψ |Φ = Ψ |G1 G2  2 r2
G1 G2 |Φ
r1⊥ 2⊥

d D−2 r1⊥ d D−2 r2⊥


+ Ψ |Q1 G2  Q1 G2 |Φ
(m − r̂1⊥ )r2⊥
2

d D−2 r1⊥ d D−2 r2⊥


+ Ψ |G1 Q2  G1 Q2 |Φ, (49)
(m − r̂2⊥ )r1⊥
2

where summation over color and spin indices is assumed.


Bootstrap conditions relate jet production-operators, impact-factors and jet production effec-
tive vertices. We define impact-factors describing jet production in the fragmentation regions
of initial particles as the projections of the t-channel states |{B̄1 B̄2 }B or {A1 A2 }Ā| on two-
Reggeon states (cf. Eqs. (32), (33) in [6])
   1  R R1 R2 R1 

R1 R2 {B̄1 B̄2 }B = δ(r1⊥ + r2⊥ − qB⊥ ) − Γ{B1 B2 }P ΓP B ± Γ B̄P Γ P {B̄  B̄  }
2

P
2pB 1 2

1  R2 R1 R2 R1 
+ − ΓB1 P Γ{P B2 }B ± Γ B̄{P B̄  } Γ P B̄ 
2pB  2 1
1

1  R2 R1 R2 R1 
+ − ΓB  P Γ{P B  }B ± Γ B̄{P B̄  } Γ P B̄  , (50)
2pB  2 1 1 2
2
R
where Γ{B  B  }B are the particle–jet–Reggeon (PJR) effective vertices describing particle–jet
1 2
P → {P1 P2 } transition due to the interaction with Reggeon R; the + (−) sign stands for the
fermion (boson) state in the t -channel, qB = pB − pB1 − pB2 , the sum is taken over quantum
numbers of particles P (they can be different in different terms) and the factor 1/pB−i comes from
the phase space element in the unitarity relation (see Eqs. (27), (38) in [6]). The factor 1/2 and
the last term in each brackets in (50) stand on account of the “signaturization”. The bar over
particle symbol means, as usual, antiparticle while Γ R 2
B̄P
, ΓR 2
B̄{P B̄  }
and Γ R 1
P B̄ 
,ΓR 1
P {B̄  B̄  }
are ob-
i 1 2
R2 R2
tained from ΓB̄P , ΓB̄{P B̄  }
and ΓPRB̄1 , ΓPR{B̄1  B̄  } correspondingly taking instead of wave functions
i 1 2
220 A.V. Bogdan, A.V. Grabovsky / Nuclear Physics B 757 (2006) 211–232

(polarization vectors and Dirac spinors) of B̄ and B̄i the wave functions of B and Bi from the
first term in brackets of (50).
Quite analogously,
   1  R R1 R2 R1 
{A1 A2 }Ā R1 R2 = δ(r1⊥ + r2⊥ − qA⊥ ) + Γ {A1 A2 }P Γ P A ± Γ ĀP Γ P {Ā1 Ā2 }
2

P
2pA
1  R2 R1 R2 R1 
+ + Γ A1 P Γ {P A2 }A ± Γ Ā{P Ā2 } Γ P Ā1
2pA
1

1  R2 R1 R2 R1 
+ + Γ A P Γ {P A }A ± Γ Ā{P Γ
Ā1 } P Ā2
, (51)
2pA 2 1
2

where qA = pA + pA − pA .


1 2
The strong bootstrap conditions resulting from the quasi-elastic (one final particle is a two-
particle jet) amplitudes impose the following constraints on the impact factors
   
{B̄ B̄ }B = g Rω (qB⊥ ) Γ R  , {A1 A2 }Ā = gΓ R
1 2 {B B }B 1 2 {A A }A Rω (qA⊥ ) ,
1 2
(52)

where |Rω (q⊥ ) are universal (process independent) eigenstates of the kernel K̂ (see Eq. (43) in
[6]) with the eigenvalues ωR (q). From calculations in leading order we know that

G1 G2 Gω (q⊥ ) = δ(r1⊥ + r2⊥ − q⊥ )TGG1 G2 ,

G1 Q2 Qω (q⊥ ) = δ(r1⊥ + r2⊥ − q⊥ )t G1 ,

Q1 G2 Qω (q⊥ ) = −δ(r1⊥ + r2⊥ − q⊥ )t G2 . (53)
Similarly to the impact factors for scattering jets we define the impact factors for Reggeon–jet
transitions (compare with (51)) as

R1 R2 {P̄1 P̄2 }Rj +1
 1  R R R 
= δ(r1⊥ + r2⊥ − qj ⊥ ) Γ 2 γP ±ΓR 1
γ 2 j +1
2k − {P1 P2 }P R1 Rj +1 P {P̄1 P̄2 } P
P
1  R2 Rj +1 
+ − ΓPR1 P2 γR1 R2j +1 ± Γ R
{P P } 1
γ
P P̄1 {P P̄2 }
2k1

1  R2 {P P1 } R1 R2 Rj +1 
+ − ΓP2 P γR1 Rj +1 ± Γ P P̄ γ{P P̄ } , (54)
2k2 2 1

{P P }
where γRi1Ri+1 2
are the RRJ effective vertices, describing production of jets {P1 P2 } at the
Reggeon transition Ri+1 → Ri ; k = k1 + k2 is the jet momentum, qj ⊥ = q(j +1)⊥ − k⊥ , the
+ (−) sign stands for the case when the Reggeon quantum numbers (i.e. quark or gluon) in the
j and j + 1 channels are equal (different). Analogously

{P1 P2 }Ri R1 R2
 1  R R 
= δ(r1⊥ + r2⊥ − q(i+1)⊥ ) Γ 2 γP ± Γ P {1P̄ P̄ } γPRi R2
2k + {P1 P2 }P Ri R1 1 2
P

1  R2 {P P2 } R Ri R2  1  R2 {P P1 } R1 Ri R2 
+ + Γ P1 P γRi R1 ± Γ P P̄1 γ{P + Γ γ ± Γ γ , (55)
2k1 1 P̄2 } 2k2+ P2 P Ri R1 P P̄2 {P P̄1 }
A.V. Bogdan, A.V. Grabovsky / Nuclear Physics B 757 (2006) 211–232 221

where q(i+1)⊥ = qi⊥ + k⊥ .


RR
In expressions (54), (55) we introduced the effective vertices γ{P̄ i P̄ j} . One can obtain them
1 2
{P P }
from γRi1Rj2 replacing the wave-functions of the emitted particles with the wave-functions of
the corresponding incoming antiparticles:

ūQ → ῡQ̄ , υQ̄ → uQ , eG → eG (56)
and inverting the momenta kPi → −kP̄i . In fermion case we also have to change the overall
sign due to the operator ordering. There is no uniformity in literature on the matter of includ-
RR
ing this factor (−1) into the definition of the corresponding effective vertices γ{P̄ i P̄ j} or into the
1 2
impact-factor definition. We define all pair-production vertices without it. As for the RRJ effec-
tive vertices, we follow the denotations from [6], where γQ̄GQ is defined with (−1). Thus, when
this factor arises and it is not included in the RRJ or PPR effective vertex we explicitly write it
in the impact-factor.
Finally, we introduce the operator {P 1 P2 } for the production of a jet {P1 P2 } with the overall
momentum k = k1 + k2 as having the following matrix elements:

R1 R2 |{P  
1 P2 }|R1 R2 
= δ(q(l+1)⊥ − k⊥ − ql⊥ )
 {P P }  {P P } 
× γR 1R2 δR2 R δ(r2⊥ − r2⊥ )dR2 + γR 1R2 δR1 R δ(r1⊥ − r1⊥ )dR1
1 1 2 2 2 1
P1 P2  P2 P1 

+ γR R γR R δ(k1 + r1⊥ − r1⊥ ) + γR γ  δ(k2 + r1⊥ − r1⊥ ) , (57)
1 1 2 2 1 R1 R2 R2

where ql⊥ = r1⊥ + r2⊥ , q(l+1)⊥ = r1⊥  + r .


2⊥
An additional bootstrap condition, which may be obtained from the bootstrap relation for
the amplitudes of a process A + B → A + {P1 P2 } + B  , is analogous to the LLA one [6]. For
“ket”-vectors it reads as follows
{P1 P2 }

{P
1 P2 } Rω (q(i+1)⊥ ) gdi+1 (q(i+1)⊥ ) + {P̄1 P̄2 }Ri+1 = Rω (qi⊥ ) gγRi Ri+1 , (58)
where qi⊥ = q(i+1)⊥ − k⊥ , while for “bra”-vectors this condition has the form
{P1 P2 }
gdi (qi⊥ ) Rω (qi⊥ ) {P

1 P2 } + {P1 P2 }Ri = gγRi Ri+1 Rω (q(i+1)⊥ ) , (59)
where q(i+1)⊥ = qi⊥ + k⊥ .
Note that the conditions for “ket”- and “bra”-vectors in (58), (59) and (52) are not independent
since these vectors are interrelated. Indeed, the replacement of “+” and “−” momenta compo-
nents turns any of them into the other, so we consider only “ket”-vectors herein.
Jet production in the central region in the Reggeized gluon collision and in the fragmentation
region with the Reggeized gluon was considered in [12]. We investigate here the bootstrap
conditions for Reggeized quarks.

4. Verification of bootstrap conditions on Reggeon vertices

In this section we explicitly show that bootstrap conditions (52) and (58), (59) are satisfied by
the known expressions for the effective vertices presented in Section 2. For this purpose we have
chosen such a spacial parametrization of these Regge vertices that their following insertion into
the bootstrap conditions leads to trivial cancellations.
222 A.V. Bogdan, A.V. Grabovsky / Nuclear Physics B 757 (2006) 211–232

Each bootstrap condition on concrete Regge vertex we check only for G1 Q2 | (or for Q1 G2 |).
Looking at the diagrams one can see that in case of Q1 G2 | (G1 Q2 |) the calculation is quite
analogous being different only in an overall “−” sign and the replacement r2 ↔ r1 .

4.1. Two-gluon jet production in fragmentation region

We begin with two-gluon jet production in the fragmentation region. In this and the next
subsection let us denote the momentum of the incoming quark QB as pB and the momenta of
the outgoing Reggeized gluon G1 and quark Q2 as r1 and r2 respectively, r1 + r2 = q. Here the
momenta of the emitted gluons are k1 and k2 . We also often meet shifted momenta
   
k1⊥ = k1⊥ + r1⊥ , k2⊥ = k2⊥ + r2⊥ , k2⊥ = k2⊥ + r1⊥ , k1⊥ = k1⊥ + r2⊥ .
(60)
Q
The bootstrap condition for Γ{G has the form:
1 G2 }Q
Q
G1 Q2 {Ḡ1 Ḡ2 }QB = G1 Q2 Qω (q⊥ ) gΓ{G1 G2 }QB
, (61)
where

G1 Q2 {Ḡ1 Ḡ2 }QB
= δ(q⊥ + k⊥ − pB⊥ )
   G 
1 Q2 G1 Q1
× Γ Γ
{G1 G2 }Q QQB + Γ 1
Γ
4k − G{Ḡ1 Ḡ2 } Q̄B G
Q G
  G  
1
+ − ΓGQ12Q Γ{G
G1
2 Q}QB
+ Γ 1
Γ Q2
GḠ1 Q̄B {GḠ2 }
+ (1 ↔ 2) . (62)
2k1 Q G

As in the definition of the effective vertices we use the denotation (1 ↔ 2) assuming the replace-
ment {e1⊥ , k1⊥ , x1 , G1 } ↔ {e2⊥ , k2⊥ , x2 , G2 }. The parts of (62) yield:
1  Q2 G1
Γ{G1 G2 }Q ΓQQ
2k − B
Q
 μ μ  μ μ 
G2 G1 G1 e1⊥ F̃3 (k2 , k1 ) G2 G1 G1 γ⊥ V1 (k2 , k1 )
= g −t t t
3
+t t t + (1 ↔ 2) uQB ,
d(k2 , k1 ) − x12 m2 d(k2 , k1 )
1  G1 (63)

Γ G{Ḡ Ḡ } Γ Q 1
Q̄B G
2k 1 2
G
 
 G G G  V1μ (k2 , k1 )  G  G G  V1μ (k2 , k1 ) μ
= −g 3 2
t t ,t 1 1 + t t t1 2 1 + (1 ↔ 2) γ⊥ uQB ,
d(k2 , k1 ) d(k2 , k1 )
1  Q2 G1 (64)
− ΓG1 Q Γ{G2 Q}QB
2k1 Q

e F (k2 , k1 ) F3 (k2 , k1 )
μ μ μ μ
 G G  e2⊥
=g t3 G1
t G2 t G1 2⊥ 3 + t 1t 2
d(k2 , k1 ) − x22 m2 d(k2 , k1 ) − x22 m2
μ μ 
e F (k2 , pB − k2 )
− t G1 t G2 2⊥ 3 uQB , (65)
d(k2 , pB − k2 ) − x22 m2
A.V. Bogdan, A.V. Grabovsky / Nuclear Physics B 757 (2006) 211–232 223

1  G1
Γ Γ Q2
2k1− G GḠ1 Q̄B {GḠ2 }

F̃3 (k2 , k1 )
μ μ μ μ
  e1⊥  G G  G e2⊥ F3 (k2 , pB − k2 )
= g 3 −t G2 t G1 t G1 − t 1t 1 t 2
d(k2 , k1 ) − x12 m2 d(k2 , pB − k2 ) − x22 m2

   γ⊥μ V1μ (k2 , k1 )
+ t G2 t G1 t G1 uQB . (66)
d(k2 , k1 )
−μ μ
Here we use the relation ê1⊥ n̂2 F3 = 2F3 − n̂1 (· · ·) to obtain (65). Now one can clearly see
that (61) is an identity.
The bootstrap condition for antiquark–gluon production reads as

G1 Q2 {QḠ}GB = G1 Q2 Qω (q⊥ ) gΓ{Q
Q̄G}G
, (67)
B

but we need not check it because Γ{Q


Q̄G}GB
is connected with the two-gluon production effective
Q
vertex Γ{G1 G2 }QB
by crossing rules.

4.2. Quark–antiquark jet production in fragmentation region

We denote the momenta of the emitted quark Q1 and antiquark Q̄2 as k1 and k2 . The bootstrap
Q
condition for Γ{Q has the form:
1 Q̄2 }Q
Q
G1 Q2 {Q̄1 Q2 }QB = G1 Q2 Qω (q⊥ ) gΓ{Q Q̄ }Q
, (68)
1 2

where

G1 Q2 {Q̄1 Q2 }QB = δ(q⊥ + k⊥ − pB⊥ )
   G 
1 Q2 G1 Q2
× Γ Γ + Γ 1
Γ
2k − {Q1 Q̄2 }Q QQB G{Q̄1 Q2 } Q̄B G
Q G
1  G1
+ − Γ Q̄Q̄ Γ Q 2
Q̄B {Q̄Q2 }
2k1 1

  G 
1 Q2 G1 Q2
+ − ΓQ̄ G Γ{GQ1 }QB + 1
Γ QQ2 Γ Q̄ {Q̄ Q} . (69)
2k2 G
2
Q
B 1

The parts of (69) yield:


 −μ
1  Q2 G1 g3 a G1 μ t a F̃4 (k2 , k1 )
Γ Γ = −t t γ u
⊥ QB ⊗ ūQ1 υ
2k − {Q1 Q̄2 }Q QQB 2k − d(k2 , k1 ) − m2 Q̄2
Q
−μ 
t a γ⊥
μ
t a t G1 F3 (k2 , k1 )
+ υQ̄2 ⊗ ūQ1 uQB , (70)
x2 d(k2 , k1 ) − x22 m2
 −μ 
1  G1 Q2 g3 a μ a G1 F̃4 (k2 , k1 )
Γ Γ = − t γ ⊥ uQ ⊗ ū Q −t t
2k − G{Q̄1 Q2 } Q̄B G 2k − B
d(k2 , k1 ) − m2
G
−μ 
F̃4 (k2 , k1 )
G1 a  a G  F̃4−μ (k2 , k1 )
+t t + t t 1 υ , (71)
d(k2 , k1 ) − m2 d(k2 , k1 ) − m2 Q̄2
224 A.V. Bogdan, A.V. Grabovsky / Nuclear Physics B 757 (2006) 211–232

 −μ
1  G1 Q2 g3 a μ t G1 t a F̃4 (k2 , k1 )
Γ Q̄Q̄ Γ Q̄ {Q̄Q } = − − −t γ⊥ uQB ⊗ ūQ1 υ
2k1− 1 B 2 2k d(k2 , k1 ) − m2 Q̄2

−μ 
t a γ⊥
μ
t G1 t a F3 (k2 , k1 )
+ υQ̄2 ⊗ ūQ1 u QB , (72)
x2 d(k2 , k1 ) − x22 m2
 −μ
1  Q2 G1 g3 a μ a G1 F3 (k2 , k1 )
Γ Γ = − t γ υ
⊥ Q̄2 ⊗ ū t t
2k2− G Q̄2 G {GQ1 }QB
Q1
2x2 k − d(k2 , k1 ) − x22 m2
−μ
  F3 (pB − k1 , k1 )
+ t G1 t a
d(pB − k1 , k1 ) − x22 m2
−μ 
F3 (k2 , k1 )
− t G1 t a uQ , (73)
d(k2 , k1 ) − x22 m2
 −μ
1  G1 Q2 g3 t a t G1 F̃4 (k2 , k1 ) μ
Γ Γ = − ū Q υ ⊗ t a γ⊥ uQB
2k2− Q QQ2 Q̄B {Q̄1 Q} 2k −
1
d(k2 , k1 ) − m2 Q̄2
−μ 
t a t G1 γ⊥
μ
t a F3 (pB − k1 , k1 )
+ υQ̄2 ⊗ ūQ1 uQB . (74)
x2 d(pB − k1 , k1 ) − x22 m2
Now one can clearly see that (68) is an identity.

4.3. Two-gluon jet production in central region

Here and in the following subsections we denote the momentum of the incoming Reggeon
as q2 and the momenta of the outgoing Reggeons R1 and R2 (the corresponding “bra”-vector
R1 R2 |) as r1 and r2 correspondingly, r1 + r2 = q1 . The momenta of the emitted particles (here
{G1 G2 }
they are gluons G1 and G2 ) are k1 and k2 (k1 + k2 = k). The bootstrap condition for γQ1 Q 2
reads as follows


Q1 G2 |{G
1 G2 } Qω (q2⊥ ) g(m − q̂2⊥ ) + Q1 G2 {Ḡ1 Ḡ2 }Q2
{G G }
= −δ(q1⊥ + k⊥ − q2⊥ )t G2 gγQ1 Q
1 2
2
, (75)

where {G1 G2 } is the operator of two gluon production with the matrix element


Q1 G2 |{G  
1 G2 }|Q1 G2 
= δ(q1⊥ + k⊥ − q2⊥ )
 {G1 G2 }  {G1 G2 } 
× γQ Q  δG2 G  δ(r2⊥ − r2⊥ )r2⊥ + γG G 
2
2
δQ1 Q1 δ(r1⊥ − r1⊥ )(m − r̂1⊥ )
1 1 2 2
 

+ γGG2G  γQG1
Q
δ(r2⊥ + k2⊥ − r2⊥ ) + γGG1G  γQG2
Q
δ(r2⊥ + k1⊥ − r2⊥ ) , (76)
2 2 1 1 2 2 1 1

and

Q1 G2 {Ḡ1 Ḡ2 }Q2
   Q 
1 G2 G2 Q2
= δ(q1⊥ + k⊥ − q2⊥ ) Γ{G1 G2 }G γQ1 Q2 +
G 1
Γ Q̄{Ḡ Ḡ } γQ̄
4k − 1 2
G Q̄
  Q  
1
ΓGG12G γQ1 Q22 + (−1) γ{GQ̄G
2 Q2
{GG }
+ − 1
Γ Q̄Ḡ + (1 ↔ 2) . (77)
2k1 1 2}
G Q̄
A.V. Bogdan, A.V. Grabovsky / Nuclear Physics B 757 (2006) 211–232 225

The term γGG2G  γQ


G1  ) in bootstrap condition (76) is transformed via (14) as
δ(r2⊥ + k2⊥ − r2⊥
2 2 Q
1 1


γGG2G  γQ
G1
 δ(r2⊥ + k2⊥ − r2⊥ )
2 2 1 Q1
 
= δ(q1⊥ + k⊥ − q2⊥ )g 3 t G1 t G2 t G2
 μ  μ 
× e2⊥ K2 (k2 , k1 ) + K1 (k2 , k1 ) − K2 (k2 , k1 ) − K1 (k2 , k1 )
μ μ μ

μ  μ 
+ e1⊥ K3 (k2 , k1 ) + K̃3 (k2 , k1 ) − K3 (k2 , k1 ) − K̃3 (k2 , k1 ) (q̂2⊥ − m).
μ μ μ
(78)

The similar procedure helps us to rewrite the contribution of γGG1G  γQ G2


 . Performing cancella-
2 2 1 Q1
tions (which are trivial due to our effective vertex presentation) inside formulas (76) and (77) we
get


Q1 G2 |{G
1 G2 } Qω (q2⊥ ) g(m − q̂2⊥ )

   (q2 − r1 )μ μ
⊥ V1 (k2 , k1 )
= g 3 δ(q1⊥ + k⊥ − q2 ) t G2 t G1 t G2
d(k2 , k1 )(q2 − r1 )2⊥
e1⊥ V2 (k2 , k1 )
μ μ
  
+ 2 t G2 t G1 t G2
D(k2 , k1 )(q2 − r1 )2⊥
 
t G2 [t G1 t G2 ] (k1 + k2 )⊥ V1 (k2 , k1 ) e1⊥ V2 (k2 , k1 )
μ μ μ μ
− +
(k1 + k2 )2⊥ d(k2 , k1 ) D(k2 , k1 )
  γ μ V1μ (k2 , k1 ) 1
+ t G1 t G2 t G2
2d(k2 , k1 ) q̂2⊥ − r̂2⊥ − m
+ t G2 t G2 t G1 e1⊥ K3 (k2 , k1 ) − t G1 t G2 t G2 e2⊥ K2 (k2 , k1 )
μ μ μ μ
    μ μ
+ t G2 t G2 t G1 e1⊥ K3 (k2 , k1 ) − t G2 t G2 t G1 e1⊥ K3 (k2 , k1 )
μ μ
  μ  μ 
+ t G1 t G2 t G2 e2⊥ K2 (k2 , k1 ) + K1 (k2 , k1 ) − K1 (k2 , k1 )
μ μ

μ μ 
G1 G2 G2 e2⊥ F1 (k2 , k1 ) 1
−t t t (q̂2⊥ − m) + (1 ↔ 2) (79)
D(k2 , k1 ) − x2 m2 q̂2⊥ − r̂2⊥ − m
and

Q1 G2 |{Ḡ1 Ḡ2 }Q2 



  γ μ V1μ (k2 , k1 )
= g 3 δ(q1⊥ + k⊥ − q2⊥ ) t G2 t G1 t G2
2d(k2 , k1 )
 μ 
G2 G1 G2 μ F1 (k2 , q2 − k2 ) μ
− t t t e2⊥ + K2 (k2 , q2 − k2 )(q̂2⊥ − m)
D(k2 , q2 − k2 ) − x2 m2

 G G  G  (q2 − r1 )μ μ
⊥ V1 (k2 , k1 )
+ t 1t 2 t 2
d(k2 , k1 )(q2 − r1 )2⊥
  γ μ V1μ (k2 , k1 ) 1
− t G1 t G2 t G2
2d(k2 , k1 ) q̂2⊥ − r̂2⊥ − m
V2 (k2 , k1 )
μ μ
 G  G G  e1⊥
− 2 t 1 t 2t 2
D(k2 , k1 )(q2 − r1 )2⊥
226 A.V. Bogdan, A.V. Grabovsky / Nuclear Physics B 757 (2006) 211–232

e2⊥ F1 (k2 , k1 )


μ μ
1
+ t G1 t G2 t G2
D(k2 , k1 ) − x2 m2 q̂2⊥ − r̂2⊥ − m
  μ μ
+ t G1 t G2 t G2 e2⊥ K2 (k2 , k1 ) + t G1 t G2 t G2 e1⊥ K̃3 (k2 , k1 )
μ μ

  μ  μ 
− t G1 t G2 t G2 e2⊥ K2 (k2 , k1 ) + K1 (k2 , k1 ) − K1 (k2 , k1 )
μ μ

 
+ t G2 t G2 t G1 e1⊥ K3 (k2 , k1 ) (q̂2⊥ − m) + (1 ↔ 2).
μ μ
(80)

One can easily check that the sum of (79) and (80) gives the r.h.s. of (75) fulfilling bootstrap
condition (58) for this case.

4.4. Quark–antiquark jet production in central region

We denote the momenta of the emitted quark and antiquark as k1 and k2 respectively. The
{Q1 Q̄2 }
bootstrap condition for γQ1 Q 2
has the form:


Q1 G2 |{Q
1 Q̄2 } Qω (q2⊥ ) g(m − q̂2⊥ ) + Q1 G2 {Q̄1 Q2 }Q2
{Q1 Q̄2 }
= Q1 G2 Qω (q1⊥ ) gγQ1 Q 2
, (81)


where {Q1 Q̄2 } is the operator of quark–antiquark production with the matrix elements


Q1 G2 |{Q  
1 Q̄2 }|Q1 G2 
 {Q1 Q̄2 } 
= δ(q1⊥ + k⊥ − q2⊥ ) γQ Q  δG2 G  δ(r2⊥ − r2⊥ )r2⊥
2
2
1 1
{Q1 Q̄2 } 

+ γG G  δQ1 Q1 δ(r1⊥ − r1⊥ )(m − r̂1⊥ ) , (82)
2 2


Q1 G2 |{Q   Q̄2 Q1 
1 Q̄2 }|G1 Q2  = δ(q1⊥ + k⊥ − q2 )γQ G  γG Q δ(r2⊥ + k1⊥ − r2⊥ ) (83)
1 1 2 2

and

Q1 G2 |{Q̄1 Q2 }Q2 


   Q 
1 G2 G2 Q2
= δ(q1⊥ + k⊥ − q2⊥ ) Γ γ G
{Q1 Q̄2 }G Q1 Q2
+ Γ 1
γ
2k − Q̄{Q̄1 Q2 } Q̄
G Q̄
 
1  G2 {QQ̄2 } 1 G2 {Q1 Q̄}
 Q G2 Q2
+ − Γ γ + − ΓQ̄ Q̄ γQ1 Q2 + 1
Γ GQ2 γ{Q̄ G} . (84)
2k1 Q Q1 Q Q1 Q2 2k2 2
G
1

Q̄2 Q1 
With the help of (14) the contribution of γQ γ  into Q1 G2 |{Q1 Q̄2 }|Qω (q2⊥ )g(m − q̂2⊥ )
1 G1 G2 Q2
yields
μ
 γ⊥ G2 G1  −μ −μ
δ(q1⊥ + k⊥ − q2⊥ )g 3 t G1 υ ⊗ ūQ1 t t K1 (k2 , k1 ) + K2 (k2 , k1 )
2k2− Q̄2
−μ −μ 
− K1 (k2 , k1 ) − K2 (k2 , k1 ) (q̂2⊥ − m). (85)
A.V. Bogdan, A.V. Grabovsky / Nuclear Physics B 757 (2006) 211–232 227

We can present the result for Q1 G2 |{Q̄1 Q2 }Q2  in the following form:

Q1 G2 {Q̄1 Q2 }Q2
 μ   −μ
g 3 a γ⊥ G2 a F1 (q2 − k1 , k1 )
= δ(q1⊥ + k⊥ − q2⊥ ) − t υQ̄2 ⊗ ūQ1 t t
k 2x2 D(q2 − k1 , k1 ) − x2 m2

−μ
− K1 (q2 − k1 , k1 )(q̂2⊥ − m)
 −μ −μ −μ −μ 
− t G2 t a K1 (k2 , k1 ) + K2 (k2 , k1 ) − K2 (k2 , k1 ) − K1 (k2 , k1 ) (q̂2⊥ − m)
 −μ 
F1 (k2 , k1 ) 1 −μ 
− t a t G2 − K (k ,
2 1k )(q̂ 2⊥ − m)
D(k2 , k1 ) − x2 m2 q̂2⊥ − r̂2⊥ − m 1
 
ūQ1 n̂1  a G  t a t G2 F5 (k2 , k1 ) t G2 t a F5 (k2 , k1 )
+ (q̂2⊥ − m)t a ⊗ t t 2 + − υ
(q2 − r1 )2⊥ D(k2 , k1 ) − m2 D(k2 , k1 ) − m2 Q̄2
 μ 
γ t a t G2   (q̂2⊥ − m)(q2 − r1 )μ
+ ⊥ t G2 t a − (q̂2⊥ − m) − t G2 t a ⊥
2 q̂2⊥ − r̂2⊥ − m (q2 − r1 )2⊥
−μ 
t a F̃4 (k2 , k1 )
⊗ ūQ1 υ . (86)
d(k2 , k1 ) − m2 Q̄2

The contribution of Q1 G2 |{Q 


1 Q̄2 }|Qω (q2⊥ )g(m − q̂2⊥ ) into the bootstrap relation reads as
follows:

Q1 G2 |{Q
1 Q̄2 } Qω (q2⊥ ) g(m − q̂2⊥ )
 μ
g3 γ
= δ(q1⊥ + k⊥ − q2⊥ ) − t a ⊥ υQ̄2
k 2x2
  −μ  
F1 (k2 , k1 ) 1 −μ 
⊗ ūQ1 t a t G2 − K (k ,
2 1k )(q̂ 2⊥ − m)
D(k2 , k1 ) − x2 m2 q̂2⊥ − r̂2⊥ − m 1

G2 a  −μ  −μ  −μ   −μ   
+ t t K1 (k2 , k1 ) + K2 (k2 , k1 ) − K1 (k2 , k1 ) − K2 (k2 , k1 )

× (q̂2⊥ − m) − (q̂2⊥ − m)t a


 
ūQ1 n̂1  a G  t a t G2 F5 (k2 , k1 ) t G2 t a F5 (k2 , k1 )
⊗ t t 2 + − υ
(q2 − r1 )2⊥ D(k2 , k1 ) − m2 D(k2 , k1 ) − m2 Q̄2
 μ
γ⊥ t a t G2  G a  (q2 − r1 )μ ⊥
+ + t t 2 (q̂2⊥ − m)
2 q̂2⊥ − r̂2⊥ − m (q2 − r1 )2⊥
−μ
t a F̃4 (k2 , k1 )
⊗ ūQ1 υ + (q̂2⊥ − m)t G2 t a
d(k2 , k1 ) − m2 Q̄2
 μ −μ 
ūQ1 t a n̂1 F5 (k2 , k1 ) (k1 + k2 )⊥ F̃4 (k2 , k1 )
⊗ n̂1 + − υQ̄2 . (87)
(k1 + k2 )2⊥ D(k2 , k1 ) − m2 d(k2 , k1 ) − m2
The sum of the two last expressions gives the r.h.s. of (81) fulfilling bootstrap condition for
{Q1 Q̄2 }
γQ1 Q 2
.
228 A.V. Bogdan, A.V. Grabovsky / Nuclear Physics B 757 (2006) 211–232

4.5. Quark–gluon jet production in central region

We denote the momenta of the emitted quark and gluon as k1 and k2 . The bootstrap condition
{Q G }
for γG1 Q1 2 2 has the form:


G1 G2 |{Q
1 G2 } Qω (q2⊥ ) g(m − q̂2⊥ ) + G1 G2 {Q̄1 Ḡ2 }Q2
{Q G }
= G1 G2 Gω (q1⊥ ) gγG1 Q1 2 2 , (88)


where {Q 1 G2 } is quark–gluon production operator with the matrix elements


G1 G2 |{Q  
1 G2 }|Q1 G2 
= δ(q1⊥ + k⊥ − q2⊥ )
 {Q G }  Q 

× γG Q1  2 δG2 G2 δ(r2⊥ − r2⊥ 2
)r2⊥ + γGG2G  γG 1Q δ(r2⊥ + k2⊥ − r2⊥ ), (89)
1 1 2 2 1 1

G1 G2 |{Q  
1 G2 }|G1 Q2 
= δ(q1⊥ + k⊥ − q2⊥ )
 {Q G }  Q 

× γG Q1  2 δG1 G1 δ(r1⊥ − r1⊥ 2
)r1⊥ + γGG2G  γG 1Q δ(r2⊥ + k1⊥ − r2⊥ ) (90)
2 2 1 1 2 2

and

G1 G2 |{Q̄1 Ḡ2 }Q2 


   G 
1 G2 Q G2 Q2
= δ(q1⊥ + k⊥ − q2⊥ ) Γ{Q1 G2 }Q γG1 Q2 − 1
Γ Q̄{Q̄ Ḡ } γQ̄
2k − 1 2
Q Q̄
  G 
1 G2 {QG2 } G2 Q2
+ − ΓQ1 Q γG1 Q2 − 1
Γ Q̄Q̄ γ{Q̄Ḡ }
2k1 Q
1 2


1  G2 {Q1 G} G2 Q2 
+ − ΓG2 G γG1 Q2 − Γ G 1
γ
GḠ2 {Q̄G}
. (91)
2k2 G
Q
The contribution of γGG2G  γG 1Q into G1 G2 |{Q  1 G2 }|Qω (q2⊥ )g(m − q̂2⊥ ) yields
1 1 2 2
 
δ(q1⊥ + k⊥ − q2⊥ )g 3 t G2 t G1 t G2
 μ
× e2⊥ ūQ1 K2 (k2 , k1 ) + K1 (k2 , k1 ) − K2 (k2 , k1 ) − K1 (k2 , k1 )
μ μ μ μ

− K2 (k2 , k1 ) − K1 (k2 , k1 ) + K2 (k2 , k1 ) + K1 (k2 , k1 ) (q̂2⊥ − m)
μ μ μ μ
(92)
Q
and the contribution of γGG2G  γG 1Q can be obtained from (92) by the substitution r1 ↔ r2 ,
2 2 1 1
G1 ↔ G2 .
We can present the result for G1 G2 |{Q̄1 Ḡ2 }Q2  in the following form:

G1 G2 {Q̄1 Ḡ2 }Q2

  F3μ (k2 , k1 )
= δ(q1⊥ + k⊥ − q2⊥ )g 3 e2⊥ ūQ1 −t G2 t G2 t G1
μ
d(k2 , k1 ) − x22 m2
 μ 
 G G  G  F1 (q2 − k1 , k1 ) μ
+ t t t 2 1 2 − K1 (q2 − k1 , k1 )(q̂2⊥ − m)
D(q2 − k1 , k1 ) − x2 m2
A.V. Bogdan, A.V. Grabovsky / Nuclear Physics B 757 (2006) 211–232 229

 μ 
  F1 (k2 , q2 − k2 )
− t G2 t G1 t G2
μ
+ K (k
2 2 2 , q − k2 )(q̂ 2⊥ − m)
D(k2 , q2 − k2 ) − x2 m2

t G2 t G2 t G1 t G2 t G1 t G2
μ μ
F3 (k2 , k1 ) F3 (k2 , k1 )
+ −
d(k2 , k1 ) − x22 m2 q̂2⊥ − r̂1⊥ − m d(k2 , k1 ) − x22 m2 q̂2⊥ − r̂2⊥ − m
  μ
+ t G2 t G1 t G2 K1 (k2 , k1 ) − K1 (k2 , k1 )
μ

+ K1 (k2 , k1 ) + K2 (k2 , k1 ) − K2 (k2 , k1 )
μ μ μ

F1 (k2 , k1 ) t G2 t G2 t G1 F1 (k2 , k1 ) t G1 t G2 t G2


μ μ
+ −
D(k2 , k1 ) − x2 m2 q̂2⊥ − r̂1⊥ − m D(k2 , k1 ) − x2 m2 q̂2⊥ − r̂2⊥ − m
 
F1 (k2 , k1 )
μ
  1
+ t G2 t G2 t G1
μ 
− K1 (k2 , k1 )
D(k2 , k1 ) − x2 m2 q̂2⊥ − r̂1⊥ − m
 
F1 (k2 , k1 )
μ
  1
− t G2 t G1 t G2
μ 
− K (k , k 1 )
D(k2 , k1 ) − x2 m2 q̂2⊥ − r̂2⊥ − m 1 2

+ t G2 t G1 t G2 K2 (k2 , k1 ) − t G1 t G2 t G2 K2 (k2 , k1 )


μ μ
  μ
− t G1 t G2 t G2 K1 (k2 , k1 ) − K1 (k2 , k1 ) + K1 (k2 , k1 )
μ μ


+ K2 (k2 , k1 ) − K2 (k2 , k1 ) (q̂2⊥ − m) .
μ μ
(93)

The contribution of G1 G2 |{Q 1 G2 }|Qω (q2⊥ )g(m − q̂2⊥ ) into the bootstrap relation reads as
follows:


G1 G2 |{Q
1 G2 } Qω (q2⊥ ) g(m − q̂2⊥ )
μ
= δ(q1⊥ + k⊥ − q2⊥ )g 3 e2⊥ ūQ1

   μ 
× t G2 t G1 t G2 K1 (k2 , k1 ) + K2 (k2 , k1 )
μ

  μ 
− t G2 t G1 t G2 K1 (k2 , k1 ) − K1 (k2 , k1 ) + K1 (k2 , k1 ) + K2 (k2 , k1 ) − K2 (k2 , k1 )
μ μ μ μ
  μ 
+ t G1 t G2 t G2 K1 (k2 , k1 ) − K1 (k2 , k1 ) + K1 (k2 , k1 ) + K2 (k2 , k1 ) − K2 (k2 , k1 )
μ μ μ μ
 μ μ 
  F3 (k2 , k1 ) F1 (k2 , k1 ) 1
+ t G2 t G2 t G1
μ
+ − K (k
1 2 1 , k )
d(k2 , k1 ) − x22 m2 D(k2 , k1 ) − x2 m2 k̂⊥ − m
− t G2 t G1 t G2 K2 (k2 , k1 ) + t G1 t G2 t G2 K2 (k2 , k1 )
μ μ

t G2 t G2 t G1 t G2 t G1 t G2
μ μ
F3 (k2 , k1 ) F3 (k2 , k1 )
− +
d(k2 , k1 ) − x22 m2 q̂2⊥ − r̂1⊥ − m d(k2 , k1 ) − x22 m2 q̂2⊥ − r̂2⊥ − m
F1 (k2 , k1 ) t G2 t G2 t G1 F1 (k2 , k1 ) t G1 t G2 t G2
μ μ
−  + 
D(k2 , k1 ) − x2 m2 q̂2⊥ − r̂1⊥ − m D(k2 , k1 ) − x2 m2 q̂2⊥ − r̂2⊥ − m
 
F1 (k2 , k1 )
μ
  1
− t G2 t G2 t G1
μ 
− K (k , k 1 )
D(k2 , k1 ) − x2 m2 q̂2⊥ − r̂1⊥ − m 1 2
 
F1 (k2 , k1 )
μ
G G G 1 μ 
+ t t t
2 1 2 − K1 (k2 , k1 ) (q̂2⊥ − m).
D(k2 , k1 ) − x2 m2 q̂2⊥ − r̂2⊥ − m
The sum of the two last expressions precisely gives the r.h.s. of (88).
230 A.V. Bogdan, A.V. Grabovsky / Nuclear Physics B 757 (2006) 211–232

4.6. Antiquark–gluon jet production in central region

We denote the momenta of the emitted antiquark and gluon as k1 and k2 respectively. The
bootstrap condition has the form:

 2 {Q̄1 G2 }
G1 Q2 |{Q̄
1 G2 } Gω (q2⊥ ) gq2⊥ + G1 Q2 {Q1 Ḡ2 }G2 = G1 Q2 Qω (q1⊥ ) gγQ2 G2 , (94)


where {Q̄1 G2 } is the operator of antiquark–gluon production with the matrix element


G1 Q2 |{Q̄  
1 G2 }|G1 G2 

= δ(q1⊥ + k⊥ − q2⊥ )
 {Q̄ G }  Q̄ 

× γQ G1  2 δG1 G1 δ(r1⊥ − r1⊥ 2
)r1⊥ + γGG2G  γQ 1G  δ(r2⊥ + k1⊥ − r2⊥ ) (95)
2 2 1 1 2 2

and

G1 Q2 {Q1 Ḡ2 }G2
   QG G 
1 Q2
= δ(q1⊥ + k⊥ − q2⊥ ) Γ γ G
{Q̄1 G2 }G G1 G2
− γ 2 2
Γ 1
2k − Q Q{Q1 Ḡ2 }
G Q
  QG G 
1
ΓQ̄Q2G γG1 G22 −
{GG }
+ − γ{Q2Ḡ 2} Γ QQ1
1
2k1 G
1
Q
2

  G 
1
ΓGQ22Q γG1 G12 − Q2 G2
{Q̄ Q}
+ − Γ G1Ḡ γ{Q 1 G}
. (96)
2k2 Q G
2

Q̄ 
The contribution of γGG2G  γQ 1G  into G1 Q2 |{Q̄1 G2 }|Gω (q2⊥ )gq2⊥ yields
2
1 1 2 2

   q2
−δ(q1⊥ + k⊥ − q2⊥ )g 3 t G1 t G2 t G2 2⊥
2k1−
 μ 
× K̃3 (k2 , k1 ) + K3 (k2 , k1 ) − K̃3 (k2 , k1 ) − K3 (k2 , k1 ) γ μ υQ̄1 .
μ μ μ
(97)

We can present the result for G1 Q2 |{Q1 Ḡ2 }G2  in the following form:

G1 Q2 |{Q1 Ḡ2 }G2 



g3 [t G2 t G1 ]t G2 q2⊥
2 (q − r )μ F μ (k , k )
2 1 ⊥ 2 2 1
= δ(q1⊥ + k⊥ − q2⊥ ) −
k− (q2 − r1 )2⊥ d(k2 , k1 ) − x22 m2
μ μ
q F (k2 , k1 ) 2 
t G2 [t G1 t G2 ]ê2 q2⊥ F5 (k1 , k2 )

− t G1 t G2 t G2 2⊥ 2 + 1 +
d(k2 , k1 ) − x22 m2 (q2 − r1 )2⊥ D(k2 , k1 ) − m2
 μ 
  γμ 2V (k2 , q2 − k2 )
+ t G1 t G2 t G2
μ
− 2 + K̃3 (k2 , q2 − k2 )q2⊥ 2
2x1 D(k2 , q2 − k2 )
 
G1 G2 G2 F5 (k1 , q2 − k1 )  G  G G  γ μ q2⊥ 2
K̃3 (k2 , k1 )
μ
− t t t ê2 1 + + t 1 t 2t 2
D(q2 − k1 , k1 ) − m2 2x1
A.V. Bogdan, A.V. Grabovsky / Nuclear Physics B 757 (2006) 211–232 231

   γ μ  μ    2
− t G2 t G1 t G2 K̃3 (k2 , k1 ) + K3 (k2 , k1 ) − K3 (k2 , k1 ) q2⊥
μ μ
2x1

 G  G G  γ μ V2μ (k2 , k1 ) 2
q2⊥
+ t t t
2 1 2 υ . (98)
x1 D(k2 , k1 ) (q2 − r1 )2⊥ Q̄1


The contribution of G1 Q2 |{Q̄ 1 G2 }|Gω (q2⊥ )gq2⊥ into the bootstrap relation reads as follows:
2

 2
G1 Q2 |{Q̄
1 G2 } Gω (q2⊥ ) gq2⊥

g 3 [t G2 t G1 ]t G2 q2⊥
2 (q − r )μ F μ (k , k )
2 1 ⊥ 2 2 1
= δ(q1⊥ + k⊥ − q2⊥ ) −
k (q2 − r1 )⊥ 2 d(k2 , k1 ) − x2 m2
2

 
  F5 (k1 , k2 ) 2
q2⊥
− t G2 t G1 t G2 ê2 1 +
D(k2 , k1 ) − m2 (q2 − r1 )2⊥
[t G2 [t G1 t G2 ]]q2⊥
2 γ μ V (k , k  ) [t G1 [t G2 t G2 ]]γ μ q2⊥
μ 2
K̃3 (k2 , k1 )
2 1 μ
− 2
 −
x1 (q2 − r1 )⊥ 2 D(k2 , k1 ) 2x1

[[t G1 t G2 ]t G2 ]q2⊥
2 γμ
μ   μ   μ 

− K̃3 (k2 , k1 ) + K3 (k2 , k1 ) − K3 (k2 , k1 ) υQ̄1 . (99)
2x1
The sum of the two last expressions gives the r.h.s. of (94). It concludes the proof of the bootstrap
conditions.

5. Summary

The further development of the quark Reggeization theory in QCD demands a proof of the
quark Reggeization hypothesis in the NLO. Besides the MRK the NLA also includes another,
quasi-multi-Regge kinematics, in which one of the produced particles is replaced by a jet con-
taining two particles with similar rapidities. In this paper we have proved the quark Reggeization
in the QMRK by means of the method analogous to the proof performed in the LO. It is based
on explicit verification of the so-called bootstrap conditions—the constraints on the effective
Reggeon vertices. These conditions are imposed by the bootstrap relations which are required by
the compatibility of the s-channel unitarity with the QMRK form of amplitude (6). We formulate
these conditions in the operator formalism in the transverse momentum, color and spin space.
This formalism was firstly introduced in [8], developed in another form in [9], then extended to
inelastic amplitudes in [6] and adopted here for the QMRK. The direct insertion of the effective
vertices into the bootstrap conditions leads to extremely cumbersome and tedious calculations,
which are almost completely cancellations. We make these cancellations transparent for verifica-
tion introducing nontrivial parametrization of the Reggeon vertices through the set of functions
Fi , Ki , Vi Eqs. (20)–(33).

Acknowledgement

We would like to thank V.S. Fadin for drawing our attention to this work and for helpful
comments and discussions.
232 A.V. Bogdan, A.V. Grabovsky / Nuclear Physics B 757 (2006) 211–232

References

[1] V.S. Fadin, V.E. Sherman, Pis’ma Zh. Eksp. Teor. Fiz. 23 (1976) 599, Sov. Phys. JETP Lett. 23 (1976) 548;
V.S. Fadin, V.E. Sherman, Zh. Eksp. Teor. Fiz. 72 (1977) 1640, Sov. Phys. JETP 45 (1977) 861.
[2] A.V. Bogdan, V. Del Duca, V.S. Fadin, E.W.N. Glover, JHEP 0203 (2002) 32, hep-ph/0201240.
[3] A.V. Bogdan, V.S. Fadin, Phys. At. Nucl. 68 (2005) 1599, Yad. Fiz. 68 (2005) 1659, hep-ph/0408127.
[4] V.S. Fadin, R. Fiore, Phys. Rev. D 64 (2001) 114012, hep-ph/0107010.
[5] M.I. Kotsky, L.N. Lipatov, A. Principe, M.I. Vyazovsky, Nucl. Phys. B 648 (2003) 277, hep-ph/0207169.
[6] A.V. Bogdan, V.S. Fadin, Nucl. Phys. B 740 (2006) 36, hep-ph/0601117.
[7] L.N. Lipatov, M.I. Vyazovsky, Nucl. Phys. B 597 (2001) 399, hep-ph/0009340.
[8] M. Braun, G.P. Vacca, Phys. Lett. B 454 (1999) 319, hep-ph/9810454;
M. Braun, G.P. Vacca, Phys. Lett. B 477 (2000) 156, hep-ph/9910432.
[9] V.S. Fadin, R. Fiore, M.I. Kotsky, A. Papa, Phys. Lett. B 495 (2000) 329, hep-ph/0008057.
[10] V.S. Fadin, Phys. At. Nucl. 66 (2003) 2017.
[11] V.S. Fadin, E.A. Kuraev, L.N. Lipatov, Phys. Lett. B 60 (1975) 50;
E.A. Kuraev, L.N. Lipatov, V.S. Fadin, Zh. Eksp. Teor. Fiz. 71 (1976) 840, Sov. Phys. JETP 44 (1976) 443;
E.A. Kuraev, L.N. Lipatov, V.S. Fadin, Zh. Eksp. Teor. Fiz. 72 (1977) 377, Sov. Phys. JETP 45 (1977) 199.
[12] V.S. Fadin, M.G. Kozlov, A.V. Reznichenko, Phys. At. Nucl. 67 (2004) 359, Yad. Fiz. 67 (2004) 377, hep-
ph/0302224.
Nuclear Physics B 757 [FS] (2006) 233–258

Semiclassical limit of the FZZT Liouville theory


Leszek Hadasz a,b,∗,1 , Zbigniew Jaskólski c
a Physikalisches Institut, Rheinische Friedrich-Wilhelms-Universität, Nußallee 12, 53115 Bonn, Germany
b M. Smoluchowski Institute of Physics, Jagiellonian University, W. Reymonta 4, 30-059 Kraków, Poland
c Institute of Theoretical Physics, University of Wrocław, pl. M. Borna 9, 50-204 Wrocław, Poland

Received 3 April 2006; received in revised form 2 July 2006; accepted 28 August 2006
Available online 3 October 2006

Abstract
The semiclassical limit of the FZZT Liouville theory on the upper half plane with bulk operators of
arbitrary type and with elliptic boundary operators is analyzed. We prove the Polyakov conjecture for an
appropriate classical Liouville action. This action is calculated in a number of cases: One bulk operator of
arbitrary type, one bulk and one boundary, and two boundary elliptic operators. The results are in agreement
with the classical limits of the corresponding quantum correlators.
© 2006 Elsevier B.V. All rights reserved.

1. Introduction

In the last few years the quantum Liouville theory attracted again a considerable attention,
mainly for its application in describing D-brane dynamics in non-compact curved or time de-
pendent backgrounds (for review and references see [1]). This renewed interest was supported
by a recent progress in the quantum Liouville theory on bordered surfaces, where exact ana-
lytic forms of all basic correlators were derived [2–7] by powerful conformal bootstrap methods.
These methods were previously successfully applied in rederiving [8] the DOZZ bulk three point
function [9,10] and proving its consistency [11–13].
In the case of the DOZZ three point function on the sphere the exact result agrees with the
perturbative calculations [14] based on the Hamiltonian approach [15–17]. A perturbative check
of the exact formula for the one point function on the pseudosphere was given in [4,18]. This

* Corresponding author.
E-mail addresses: hadasz@th.if.uj.edu.pl (L. Hadasz), jask@ift.uni.wroc.pl (Z. Jaskólski).
1 Alexander von Humboldt Fellow.

0550-3213/$ – see front matter © 2006 Elsevier B.V. All rights reserved.
doi:10.1016/j.nuclphysb.2006.08.027
234 L. Hadasz, Z. Jaskólski / Nuclear Physics B 757 [FS] (2006) 233–258

suggests that the bootstrap solution should have a well defined path integral representation. In all
the cases in which a classical solution exists one then could expect that the quantum expressions
arise as a result of integrating fluctuations of classical background geometries.
The path integral formulation of the Liouville theory goes back to Polyakov’s celebrated pa-
per on conformal anomaly in string theory [19]. The main problem of the Polyakov original
formulation was the field-dependent functional measure arising from the gauge fixing of the path
integral over 2-dim metrics. A solution to this problem was proposed by David [20] and Dis-
tler and Kawai [21]. It was shown that the original Weyl-invariant measure can be replaced by
the translation-invariant one provided that renormalized parameters in the Liouville action are
properly chosen. This gave rise to the so-called standard functional formulation of the Liouville
theory [22–24] where the translation-invariant functional measure is defined with respect to the
flat reference metric and the correlation functions are constructed in terms of vertex functionals.
In this approach the Liouville action explicitly depends on the coupling constant b related to
the central charge by c = 1 + 6(b + b1 )2 . The standard formulation breaks down in the so-called
“strong coupling regime” 1 < c < 25 where the coupling constant b and the Liouville action
become complex.
On the other hand, in the case of “heavy” elliptic weights on compact surfaces there exists
so-called geometric approach originally proposed by Polyakov [25] and further developed by
Takhtajan [26,27] (see [28] for recent generalizations). In contrast to the standard formulation the
correlators of primary fields with elliptic and parabolic weights are expressed in terms of path in-
tegral over conformal class of Riemannian metrics with prescribed singularities at the punctures.
In this approach one tacitely assumes that the functional measure is translation-invariant and de-
fined with respect to the classical hyperbolic geometry. Since the classical action depends on b
only via Q = b + b1 the c = 1 barrier is not visible in this approach. The geometric approach has
been recently compared with the exact bootstrap solution on the pseudosphere [29]. It was shown
in particular that the choice of regularization is crucial for the agreement with the bootstrap ap-
proach. A functional representation of Liouville correlators with heavy elliptic charges on the
sphere [30], the pseudosphere [31], and the disc [32] has been developed. This representation
agrees with the bootstrap solution at least up to one loop calculations.
In the present paper we address the problem of calculating the classical limit of the FZZT
Liouville theory for heavy charges. Our motivation is twofold. In spite of the recent progress
in constructing the path integral representation of Liouville correlators for elliptic weights there
are still open questions of a consistent functional formulation of DOZZ theory. This for instance
concerns an interesting problem of factorization. Due to the global character of the Liouville
action related to its specific dependence on the background metric it is not simply described by
the standard cutting-open procedure. It seems that in order to handle this problem one should
first construct a functional representations for Liouville correlators with hyperbolic weights. The
FZZT theory provides simple cases where exact solutions are known and both problems can be
relatively easily analyzed. The classical limit is just the first step of such calculations.
The second, probably more important motivation is the semiclassical limit itself. It turned out
that analyzing the quantum Liouville theory in this limit one gets essentially new insight into
the classical hyperbolic geometry. One of the first results of this type was the so-called Polyakov
conjecture, originally obtained as a classical limit of the Ward identity and proved latter as an
exact theorem [33–38]. It says that the classical Liouville action is a generating function for the
accessory parameters of the Fuchsian uniformization of the punctured sphere.  
More intriguing results are related with the classical conformal block fδ δδ34 δδ21 (x), defined
by the classical limit of the BPZ quantum conformal block [39] with heavy weights Δ = Q2 δ,
L. Hadasz, Z. Jaskólski / Nuclear Physics B 757 [FS] (2006) 233–258 235

Δi = Q2 δi , [10,40,41]:
     
Δ3 Δ2 δ3 δ2
F1+6Q2 ,Δ (x) ∼ exp Q fδ
2
(x) . (1.1)
Δ4 Δ1 δ4 δ1
It was argued in [10] that the 4-point Liouville classical action can be expressed in terms of 3-
point actions and the classical conformal block in a given channel calculated for a saddle point
value of the intermediate weight. Since there are three different decompositions of the 4-point
action one gets consistency conditions called the classical bootstrap equations [42]. It was also
conjectured [42] that the saddle point weight is closely related to the length of the closed geodesic
in the corresponding channel. These statements are far from being proved in a rigorous way, but
there are many nontrivial numerical checks instead [42]. Let us finally stress that once the 4-point
Liouville action is known one can in principle calculate the uniformization of the 4-punctured
sphere [43], which is a long standing open mathematical problem. Analyzing various classical
limits one may hope to gain some new information on the classical conformal block, which up
to now is only available through term by term symbolic calculations from the limit (1.1).
The organization of the paper is as follows. In Section 2, following the ideas of [44], we
formulate the SL(2, R)-monodromy problem with boundary. In Section 3 we introduce an ap-
propriate classical Liouville action and derive the formulae for its partial derivatives with respect
to bulk and boundary conformal weights, and bulk and boundary cosmological constants. A
novel technical result is the formula for the partial derivative of the action with respect to the
mj
ratio ωj = √m in terms of special values of some conformal map. Details of its derivation are
explained in Appendix A.
In Section 4 we present a proof of the Polyakov conjecture for elliptic boundary weights and
arbitrary bulk weights. This extends the previous results [33–38] to the FZZT Liouville theory
and is one of the main results of this paper.
The last section contains four examples of explicit calculations of the classical actions both
from classical solutions and from the classical limit of exact quantum expressions. The simplest
one is the case of one bulk elliptic singularity considered in Section 5.1. The case of one bulk hy-
perbolic singularity is considered in Section 5.2. This calculations provide an additional support
for the construction of the classical action for hyperbolic singularities proposed and analyzed in
[38,45]. The cases of two boundary, and one bulk one boundary elliptic weights are calculated in
Section 5.3 and Section 5.4, respectively. Both ways of calculation the classical Liouville action
are rather complicated in these cases. The full agreement of the results provides an additional
strong evidence that the classical limit of the DOZZ bootstrap solution exists and is properly
described by a classical action satisfying the equations derived in Sections 3 and 4.

2. The monodromy problem with boundary

Let us consider the Fuchs equation

∂ 2 ψ(z) + T (z)ψ(z) = 0, (2.1)


with the energy–momentum tensor of the form
m 
  n  δjB cjB

δi δi ci c̄i
T (z) = + + + + + ,
(z − zi )2 (z − z̄i )2 z − zi z − z̄i (z − xj )2 z − xj
i=1 j =1
(2.2)
236 L. Hadasz, Z. Jaskólski / Nuclear Physics B 757 [FS] (2006) 233–258

where δi are bulk conformal weights located at the points z1 , . . . , zm of the upper half-plane, δjB
are boundary conformal weights located at the points x1 < · · · < xn of the real axis, and ci , cjB
are accessory parameters. We assume that δi , δjB , cjB are real so that

T (z) = T (z̄). (2.3)


The requirement that T (z) is regular at the infinity implies the relations

m 
n
2 ci + cjB = 0,
i=1 j =1

m 
n 
m 
n
2 (zi ci ) + xj cjB = −2 δi − δjB ,
i=1 j =1 i=1 j =1

m


n 
m 
n
2  zi2 ci + xj2 cjB = −4 zi δi − 2 xj δjB .
i=1 j =1 i=1 j =1
ψ − 
A fundamental systems of solutions Ψ = ψ+
to the Fuchs equation (2.1) is normalized if

ψ − ∂ψ + − ψ + ∂ψ − = 1.
Let Lj denote the part of the real axis between xj and xj +1 (with the exception of Ln denoting
the set of points on the real axis to the right of xn and to the left of x1 ). It follows from (2.3) that
for each boundary segment Lj there exist normalized solutions Ψj to the Fuchs equation (2.1)
regular and real along Lj . For any other normalized solution Ψ we define the matrices

Mj = Σ −1 BjT Σ̄ B̄j , j = 1, . . . , n,
 0 1
where Σ = −1 0
and the matrix Bj is determined by the relation Ψ = Bj Ψj .
We are interested in the following version of the Riemann–Hilbert problem [44]. For given sets
of positive weights {δi }m
i=1 , {δj }j =1 and real numbers {ωj }j =1 one has to adjust the accessory
B n n

parameters in such a way that the Fuchs equation (2.1) admits a normalized fundamental system
Ψ of solutions, such that:

(1) monodromies around all singularities zi of the upper half plane belong to SL(2, R);
(2) the function −iΨ T · Σ · Ψ̄ is strictly positive or strictly negative on the upper half plane
except the points zi , i = 1, . . . , m and xj , j = 1, . . . , n;
(3) for each boundary segment Lj the boundary condition

+ωj if −iΨ T · Σ · Ψ̄ > 0,
Tr Mj =
−ωj if −iΨ T · Σ · Ψ̄ < 0,
is satisfied.

If Ψ is a solution to the monodromy problem above then the relation


√ 2
m T
e−ϕ = Ψ (z) · Σ · Ψ (z) (2.4)
2i
L. Hadasz, Z. Jaskólski / Nuclear Physics B 757 [FS] (2006) 233–258 237

determines a conformal factor ϕ on the upper half plane satisfying the Liouville equation

¯ = m ϕ
∂ ∂ϕ e (2.5)
2
and the boundary conditions

ϕ m
∂y e− 2 Lj = − ωj , j = 1, . . . , n. (2.6)
2
They are usually written in the form
 √
∂y ϕ L = mj e 2 , mj = ωj m,
ϕ
(2.7)
j

where mj are so-called boundary cosmological constants. Note that T (z) of (2.2) is the classical
energy momentum tensor of the solution ϕ:
1 1 ϕ ϕ
T (z) = Tcl (z) = − (∂ϕ)2 + ∂ 2 ϕ = −e 2 ∂ 2 e− 2 . (2.8)
4 2
The conformal factor ϕ is a regular single valued function on the upper half plane z  0
except the singular points zi , xj . It defines the hyperbolic metric eϕ dz d z̄ with the constant
negative scalar curvature
¯ = −2m,
R = −e−ϕ Δϕ = −4e−ϕ ∂ ∂ϕ
and with the constant geodesic curvature of each boundary sector Lj :
 √
1 −ϕ a  mj m
κj = e n ∂a ϕ  = −
2  =− ωj .
2 Lj 2 2

3. The classical Liouville action

In order to simplify our considerations we assume that all weights are elliptic:

1 − ξi2 1 − νj2
δi = , j = 1, . . . , m, δjB = , j = 1, . . . , n, 0 < ξi , νj < 1.
4 4
It follows from (2.2), (2.3), and (2.8) that the most singular terms in the expansions of the Liou-
ville field around the locations of elliptic weights read

ϕ(z, z̄) ∼ −2(1 − ξi ) log |z − zi |, ϕ(z, z̄) ∼ −2(1 − νj ) log |z − xj |. (3.1)


Let X be the upper half plane with the discs of radii around the points zi , i = 1, . . . , m and
the semi-discs of radii around the points xj , j = 1, . . . , n removed. Denote by Si the boundary
of the disc around zi and by sj the semicircle forming a boundary of the semi-discs around xi .
Finally, let li denote the part of the real axis between xi + and xi+1 − (with the exception of
ln denoting the set [−R, x1 − ] ∪ [xn + , R]). The regularized action functional is defined by

S[φ] = lim S [φ],


→0
S [φ] = lim S R [φ],
R→∞
238 L. Hadasz, Z. Jaskólski / Nuclear Physics B 757 [FS] (2006) 233–258


1  
S R [φ] = ¯ + meφ
d 2 z ∂φ ∂φ

X
m 
  
1 − ξi (1 − ξi )2
+ κz |dz|φ − log
4π 2
i=1 Si

n  n 
  
mj φ 1 − νj (1 − νj )2
+ dx e +
2 κz |dz|φ − log
4π 4π 4
j =1 lj j =1 sj

1
+ |dz|φ + log R, (3.2)
2πR
sR

where sR is the semicircle |z| = R on the upper half plane. The action is constructed in such a
way that for fields satisfying

φ ∼ −2(1 − ξi ) log |z − zi | for z → zi ,


φ ∼ −2(1 − νj ) log |z − xj | for z → xj ,
φ ∼ −4 log |z| for z → ∞,
the limit in (3.2) exists and the equation δS[φ] = 0 gives (2.5) and (2.7).
Let ϕ(z, z̄) denote a solution of (2.5) and (2.7) with some specified values of m, ξi , zi and
mj , νj , xj . We define the classical Liouville action Scl as the value of the action functional (3.2)
calculated on this solution. Using the equations of motion and the boundary condition satisfied
by ϕ(z, z̄) one immediately gets
  
∂Scl 1
= lim − κz |dz|ϕ + (1 − ξi ) log , (3.3)
∂ξi →0 4π
Si
  
∂Scl 1 1 − νj
= lim − κz |dz|ϕ + log . (3.4)
∂νj →0 4π 2
sj

Shifting the classical solution ϕ = ϕ̃ − log m one obtains



mj
Scl (m, ξi , zi , mj , νj , xj ) = Scl 1, ξi , zi , √ , νj , xj
m
 m 
 1 − ξi  n
1 − νj 1
+ +ε − log m. (3.5)
2 4 2
i=1 j =1

It is convenient to regard the classical action as a function of the variables m and ωj (instead
of m and mj ). One than has

∂Scl 1 ϕ̃
= dx e 2 , j = 1, . . . , n. (3.6)
∂ωj 4π
Lj

Eqs. (3.3) and (3.4) express the derivatives of the classical action in terms of the classical
metric in the vicinity of the singularities. It is useful to work out a similar “local” expression for
the integral in the r.h.s. of (3.6). To this end note that the solution Ψ to the monodromy problem
L. Hadasz, Z. Jaskólski / Nuclear Physics B 757 [FS] (2006) 233–258 239

described in the previous section determines a multivalued analytic function from


M = {z ∈ C: z  0} ∪ {∞} \ {z1 , . . . , zn , x1 , . . . , xj }
to the upper half plane H = {z ∈ C: z > 0},
ψ − (z)
M z → ρ(z) = ∈ H. (3.7)
ψ + (z)
The pull-back of the standard hyperbolic metric on H by any branch of ρ yields a regular hyper-
bolic metric on M:
1
eϕ(z,z̄) dz d z̄ = dρ d ρ̄. (3.8)
m(ρ)2
One can always chose a branch of ρ such that the image ρ(Lj ) of the boundary segment Lj is a
connected open curve in H joining the points
ρj = lim ρ(x), ρj +1 = lim ρ(x).
x→xj + x→xj +1 −
ω
This curve has the constant geodesic curvature − 2j with respect to the Poincaré metric on H (the
sign being determined with respect to ρ(M)). It admits a simple description on the Lobachevsky
plane as the arc of the Euclidean circle containing the points ρj , ρj +1 and with its radius R and
its center O determined by the condition that the Euclidean distance of O from the real axis is
ω
equal to | 2j |R. For |ωj |  2 there are two and for |ωj | > 2 four different arcs with this property.
It is shown in Appendix A that in both cases the hyperbolic length depends only on the location of
the endpoints ρj , ρj +1 and |ωj |. It follows from (3.8) that the length of the boundary component
Lj with respect to the metric eϕ(z,z̄) is equal to the length of its image ρ(Lj ) in the Lobachevsky
plane. Using the explicit expressions (A.1) and (A.2) one gets

 2 
∂Scl 1 ωj
=  arcsinh 1 − β(ρj , ρj +1 ) (3.9)
∂ωj π 4 − ω2 2
j

for |ωj | < 2 and



 2 
∂Scl 1 ωj
=  arcsin − 1 β(ρj , ρj +1 ) (3.10)
∂ωj π ωj2 − 4 2

for |ωj | > 2, where


def |z − w|
β(z, w) = √ .
2 zw

4. Polyakov conjecture

The Polyakov conjecture states that


∂Scl
= −ci , i = 1, . . . , m; (4.1)
∂zi
∂Scl
= −cjB , j = 1, . . . , n. (4.2)
∂xj
240 L. Hadasz, Z. Jaskólski / Nuclear Physics B 757 [FS] (2006) 233–258

Eqs. (4.1) can be derived by essentially the same methods one uses in proving the Polyakov
conjecture for the Riemann sphere and we shall skip here the derivation. Let us only mention that
Eqs. (4.1) are valid in the case of parabolic and hyperbolic bulk singularities as well, although
the classical Liouville action is different in those cases.
We shall prove Eqs. (4.2). Using the Liouville equation (2.5) and the boundary conditions
(2.7) one gets
∂Scl
= lim D [ϕ],
∂xj →0
 
i ∂ϕ 1 − νj ∂ϕ
D [ϕ] = ¯
(∂ϕ d z̄ − ∂ϕ dz) + κz |dz|
8π ∂xj 4π ∂xj
sj sj
 
i
1 − νj
+ ¯ + meϕ +
(d z̄ − dz) ∂ϕ ∂ϕ κz |dz|∂x ϕ
8π 4π
sj sj
 
mk−1 ϕ  mk ϕ 
+ e2 − e2 . (4.3)
4π z=xj − 4π z=xj +
Applying the identity
∂ϕ
= −∂x ϕ + hj ,
∂xj
where
 ∂ϕ  m
∂ϕ ∂ϕ
hj = − − + ,
∂xk ∂zi ∂ z̄i
k =j i=1

one has
  
1 − νj ∂ϕ 1 − νj 1 − νj
κz |dz| + κz |dz|∂x ϕ = κz |dz|hj ,
4π ∂xj 4π 4π
sj sj sj

and

i ¯ d z̄ − ∂ϕ dz) ∂ϕ
(∂ϕ
8π ∂xj
sj
 
i ¯ + i
¯ d z̄ − ∂ϕ dz)(∂ϕ + ∂ϕ) ¯ d z̄ − ∂ϕ dz)hj .
=− (∂ϕ (∂ϕ
8π 8π
sj sj

Note that the function hj is regular for z → xj . Taking into account the asymptotic behavior of
ϕ for z → xj one thus gets in the limit → 0:
 
1 − νj i
κz |dz|hj + ¯ d z̄ − ∂ϕ dz)hj = o(1).
(∂ϕ
4π 8π
sj sj

It follows that up to terms vanishing in the limit → 0 the first two lines on the r.h.s. of (4.3)
yield:
L. Hadasz, Z. Jaskólski / Nuclear Physics B 757 [FS] (2006) 233–258 241

 
i

− (∂ϕ ¯ + i
¯ d z̄ − ∂ϕ dz)(∂ϕ + ∂ϕ) ¯ + meϕ
(d z̄ − dz) ∂ϕ ∂ϕ
8π 8π
sj sj
 
i
i

= dz (∂ϕ)2 − meϕ − ¯ 2 − meϕ


d z̄ (∂ϕ)
8π 8π
sj sj
 
i

= ¯ − 2Tcl (z) − i
dz ∂ 2 ϕ − ∂ ∂ϕ d z̄ ∂¯ 2 ϕ − ∂ ∂ϕ
¯ − 2T̄cl (z̄) ,
4π 4π
sj sj

where we have used the expression (2.8) for the classical energy–momentum tensor and its com-
plex conjugate T̄cl (z̄) along with the Liouville equation (2.5). Since T̄cl (z̄) = Tcl (z̄) we have
  
i i i
− dz Tcl (z) + d z̄ T̄cl (z̄) = dz Tcl (z) = −cjB .
2π 2π 2π
sj sj |z−xj |=

On the other hand


 
i

¯ − i
dz ∂ 2 ϕ − ∂ ∂ϕ d z̄ ∂¯ 2 ϕ − ∂ ∂ϕ
¯
4π 4π
sj sj
 
=
i ¯
(dz∂ + d z̄∂)(∂ϕ ¯ = 1
− ∂ϕ) dϑ

∂y ϕ
4π 4π ∂ϑ
sj sj
 
1  
mk ϕ  mk−1 ϕ 
= ∂y ϕ x=x + − ∂y ϕ x=x − = e 2  − e2  ,
4π j j 4π z=xj + 4π z=xj −

so that one finally gets

D [ϕ] = −cjB + o(1),

and
∂Scl
= −cjB .
∂xj
Let us note that the proof presented above applies to the bulk singularities as well.

5. Classical solutions

5.1. One bulk elliptic singularity

In the case of an elliptic singularity located at z1 on the upper half-plane z1 > 0 the energy-
momentum takes the form

1 − ξ2 1 1 2
T (z) = + − , 0 < ξ < 1. (5.1)
4 (z − z1 )2 (z − z̄1 )2 (z − z1 )(z − z̄1 )
242 L. Hadasz, Z. Jaskólski / Nuclear Physics B 757 [FS] (2006) 233–258

Since the singularity is elliptic and there are no singularities on the real axis it is more convenient
to work with the normalized solutions to the Fuchs equation (2.1) with a diagonal SU(1, 1)
monodromy:
⎡ 1+ξ ⎤
  √ 1 (z − z1 )
1−ξ
2 (z − z̄1 ) 2
ψ1 ξ(z1 −z̄1 )
Ψ= =⎣ 1+ξ 1−ξ
⎦. (5.2)
ψ2 √ 1 (z − z1 ) 2 (z − z̄1 ) 2
ξ(z −z̄ ) 1 1
In this setting a real, single valued conformal factor of the hyperbolic geometry with constant
scalar curvature −2m is given by

m t

e−ϕ(z,z̄)/2 = e ψ1 (z)ψ1 (z) − e−t ψ2 (z)ψ2 (z)


2
√      ξ 
m  (z − z1 )(z − z̄1 )  t  z − z1 −ξ 
−t  z − z1 

=   e  −e  , (5.3)
2ξ z1 − z̄1 z − z̄1 z − z̄1 
where t is real and t  0 so that the r.h.s. is non-negative on the upper half plane and the real
axis. By direct calculations one gets

ϕ m t
√ mB
∂y e− 2 R = e + e−t = m cosh t = − .
2 2
Thus a regular metric can be constructed if and only if the bulk and boundary cosmological
constants satisfy the conditions
mB < 0, m2B  4m.
This in particular means that the geodesic curvature κ of the boundary is bounded from below,
κ  m.
For the solution (5.3) the r.h.s. of Eqs. (3.3), (3.6), (4.1), and (3.5) can be easily calculated,

∂Scl m
= log ξ − log − t − ξ log |z1 − z̄1 |,
∂ξ 2
∂Scl ξ
= ,
∂ωB ω2 − 4 B

∂Scl 1 − ξ 2 1
= ,
∂z1 2 z1 − z̄1
where ωB = −2 cosh t . The solution reads2
ξ 1 − ξ2
Scl = ξ log ξ − ξ − log m + ξ log 2 − ξ t − log |z1 − z̄1 | + const. (5.4)
2 2
Let us note in passing that using the form of the elliptic basis (5.2) one gets for the map (3.7)
t t
e 2 ψ1 (z) − e− 2 ψ2 (z) et (z − z̄1 )ξ − (z − z1 )ξ
ρ(z) = i =i ,
t
e ψ1 (z) − e
2 − 2t
ψ2 (z) et (z − z̄1 )ξ − (z − z1 )ξ
so that
sinh t + i sin 2πξ sinh t
ρ1 = lim ρ(z) = i , ρ2 = lim ρ(z) = i .
z→−∞ cosh t + cos 2πξ z→∞ cosh t + 1

2 In the case under consideration the classical action could be also calculated by explicit integration in (3.2) (see [32]).
L. Hadasz, Z. Jaskólski / Nuclear Physics B 757 [FS] (2006) 233–258 243

This immediately gives


|ρj +1 − ρj | sin πξ sin πξ
 = = 

2 ρj +1 ρj | sinh t| ωB 2


−1 2

and from (3.10) we get as above


∂Scl ξ
= ,
∂ωB ωB2 − 4

which confirms formulae (3.9) and (3.10).


We now shall compare (5.4) with the semi-classical limit of the FZZ 1-point function:
  U (α, μB )
Vα (z1 ) = ,
|z1 − z̄1 |2Δα
2

Q−2α

U (α, μB ) = πμγ b2 2b  2bα − b2  2α b −


1
b2
− 1 cosh(2α − Q)πs, (5.5)
b
where s is defined by the relation

μ2B
cosh2 πbs = sin πb2 . (5.6)
μ
In our notation
m mB Q 1
μ= , μB = , α= (1 − ξ ), Q=b+ , (5.7)
4πb2 4πb2 2 b
and
2

1 1+ 1
ξ

U (ξ, mB ) = πμγ b2 2 b2  b2 (1 − ξ ) + 1 − ξ
b



×  − 1 + b12 ξ cosh − b + b1 ξ πs .
In the limit b → 0 the relation (5.6) yields πbs = t , hence
2

1 (1+ 1 )ξ

πμγ b2 2 b2  b2 (1 − ξ ) + 1 − ξ cosh − b + b1 ξ πs
b
1
− [−ξ +ξ log ξ −ξ log b2 −tξ ]
≈e b2 .
Due to the poles of the gamma function along the negative real axis the b → 0 asymptotic of the
term (−(1 + b12 )ξ ) is more subtle. Within the path integral approach the poles in (5.5) arise due
to the integration over the zero mode of the Liouville field. If the classical solution exists, one
should not expect any pole structure in the quasi-classical limit. In order to show that the poles
can be regarded as a sub-leading correction one may use the formula
π
(−x) = − ,
x(x) sin πx
along with the Stirling asymptotic expansion
1

x(x) ≈ e−x+ x+ 2 log x


.
244 L. Hadasz, Z. Jaskólski / Nuclear Physics B 757 [FS] (2006) 233–258

In the limit b → 0 one obtains:



1 1
ξ − 12 ξ + 12 log 12 ξ −log − π1 sin π2 ξ
 − 1 + 2 ξ ≈ e b2 b b b
b
1  π

− −ξ +ξ log ξ −ξ log b2 + 12 b2 log 1
ξ +b2 log − π1 sin ξ
=e b2 b2 b2 .
Keeping only the leading terms one thus have
1 ξ

− −2 log m+ξ log 2−ξ +ξ log ξ −tξ


U (ξ, mB ) ≈ e b2 ,
in perfect agreement with the classical action (5.4).

5.2. One bulk hyperbolic singularity

Hyperbolic weight corresponds to the energy–momentum of the form



1 + λ2 1 1 2
T (z) = + − . (5.8)
4 (z − z1 )2 (z − z̄1 )2 (z − z1 )(z − z̄1 )
Repeating (with obvious modifications) the calculations from the previous subsection one gets
the metric
√      
m  (z − z1 )(z − z̄1 )   z − z1 
 −t ,
e−ϕ(z,z̄)/2 = sin λ log (5.9)
λ  z − z̄ 1

1
 z − z̄ 
1
where
ωB mB
cos t = − ≡− √ . (5.10)
2 2 m
The metric can be constructed only if |ωb |  2. In the coordinates
z − z1
w = τ + iσ = log
z − z̄1
it takes the form

−ϕ(w,w̄)/2 m
e = sin(λτ − t). (5.11)
λ
The conformal factor is singular along the lines
λτ = t + πk, k ∈ Z,
and the metric eϕ(w,w̄) |dw|2 has closed geodesics located at
π
λτ = t + (2k + 1), k ∈ Z.
2
For positive ωb there exists a solution of (5.10) satisfying −π < t < − π2 . The metric between
the real axis τ = 0 and the geodesic corresponding to k = 0,
π
λτg = + t < 0,
2
is then regular. As a final step of the construction of the C 1 metric on the upper half-plane we
shall “fill in” the hole τ < τg with a flat metric determined by

−ϕ0 (w,w̄)/2 m
e =
λ
L. Hadasz, Z. Jaskólski / Nuclear Physics B 757 [FS] (2006) 233–258 245

or, in the z coordinates,


√  
m  (z − z1 )(z − z̄1 ) 
e−ϕ0 (z,z̄)/2 = .
λ  z1 − z̄1
The classical Liouville action in the presence of hyperbolic singularities [38,45] is constructed
as a sum of the standard Liouville action functional calculated on the conformal factor of the
hyperbolic metric in the region “between the holes” and the actions for “holes” around each
hyperbolic singularity. In our case the first contribution is

S1 (m, mB , λ, z1 )
  R  
1  
= lim d z |∂ϕ| + me + mB dy eϕ/2 +
2 2 ϕ
κz |dz|ϕ , (5.12)
4π R→∞
MR −R ∂ MR

where MR is a part of the upper half plane outside the hole, log | z−zz−z̄1 | > λ ( 2 + t), bounded by
1 1 π

the semi-circle of radius R. The second one is a regularized Liouville action functional, calcu-
lated for the flat metric ϕ0 on the hole around z = z1 with a small disc of radius removed,
   
 z − z1  1 π
H = z ∈ H: log < + t ∧ |z − z | < .
z − z̄1  λ 2
1

It reads

S2 (m, mB , λ, z1 ) = lim S2, (m, mB , λ, z1 ),


→0
 
1   1

S2, (m, mB , λ, z1 ) = d 2 z |∂ϕ0 |2 + meϕ0 + κz |dz|ϕ0 + λ2 − 1 log .


4π 4π
H ∂H
(5.13)
Shifting

ϕ = ϕ̃ − log m, ϕ0 = ϕ̃0 − log m,

one checks that the classical action depends on m only trough ωB ,


mB

Scl = S √
m
, λ, z1 ≡ S(ωB , λ, z1 ).

From the Polyakov conjecture and Eq. (5.8) one gets

∂Scl 1 + λ2 1
= . (5.14)
∂z1 2 z1 − z̄1
One also has (using the form of functionals (5.12) and (5.13))

∂Scl 1 λ
= dy eϕ̃/2 =  , (5.15)
∂ωB 4π 4 − ω 2
R B

and finally
246 L. Hadasz, Z. Jaskólski / Nuclear Physics B 757 [FS] (2006) 233–258

  
∂Scl 1 2 ∂ ϕ̃0
= lim d z e + λ log
∂λ →0 4π ∂λ
H
 τg 
1 ∂ 2 π
= lim dτ λ + λ log = t + + λ log |z1 − z̄1 |, (5.16)
→0 2 ∂λ 2
τ0

where the change of the integration variables from z to w and the fact that |z − ξ | = corresponds
to τ = log − log |z1 − z̄1 | have been used.
Integrating Eqs. (5.14), (5.15) and (5.16) we get

1 + λ2 π
S(m, mB , λ, z1 ) = log |z1 − z̄1 | + λ t +
2
+ const, (5.17)
4 2
where the constant is independent of m, mB , λ, z1 .
2
For hyperbolic weights Δ = α(Q − α) > Q4 , i.e. for α of the form
Q
α= (1 + iλ), λ ∈ R,
2
the Liouville one-point coupling constant U (α) given by (5.5) is complex. Let us write

U (α) = Φ(α)Us (α),


where Φ is the phase and Us the modulus3 of U . The phase Φ coincides with a square root of
the Liouville reflection amplitude,

Φ 2 (α) = SL (α),

2α−Q (−(2α − Q)/b)(1 − b(2α − Q))


SL (α) = πμγ b2 b ,
((2α − Q)/b)(1 + b(2α − Q))
and



Q−2α

2 cosh[(2α − Q)πs]
Us (α) =  1 + (2α − Q)b  1 − (2α − Q)b  2α−Qb  b .
b
Using (5.6), (5.7) and the fact that t < 0, we get for b → 0
 
Us (α) 1 1 + λ2 π
∼ exp − 2 log |z1 − z̄1 | + λ t +
2
,
|z1 − z̄1 |2Δα b 4 2
in perfect agreement with the classical action (5.17) again.

5.3. Two boundary elliptic singularities

In the case of two singularities the conformal weights must be the same, ν1 = ν2 = ν. Using
the SL(2, R) symmetry one can place them at x1 = 0 and x2 = ∞. This corresponds to the
following energy–momentum tensor
1 − ν2 1
T (z) = .
4 z2

3 The subscript s is meant to remind that U (α) is symmetric under reflection α → Q − α


s
L. Hadasz, Z. Jaskólski / Nuclear Physics B 757 [FS] (2006) 233–258 247

Normalized solutions, regular and real along the positive and the negative semi-axes, are given
by
⎡ 1−ν ⎤ ⎡ 1+ν ⎤
 −  √1 z 2  −  √1 (−z) 2
ψ1 (z) ψ (z)
=⎣
ν
Ψ1 = ⎦, Ψ2 = 2
=⎣
ν
⎦, (5.18)
ψ1+ (z) 1
√ z 2
1+ν
ψ +
2 (z) 1
√ (−z) 2
1−ν
ν ν

respectively. They are related on the upper half plane by


 iπ(1−ν) 
0 e 2
Ψ1 = iπ(1+ν) Ψ2 . (5.19)
e 2 0
In terms of Ψ1 the solution to the Liouville equation reads

−ϕ/2 m T
e = Ψ (z)ΣMΨ1 (z), (5.20)
2i 1
where the matrix
 
a iβ
M= , |a|2 + βγ = 1, γ , β ∈ R, a ∈ C,
iγ ā
can be chosen such that the r.h.s. of (5.20) is positive on the upper half plane. This implies in
particular that γ and −β are positive. The boundary conditions
 
∂y ϕ y=0 = m1 eϕ/2 for x > 0, ∂y ϕ y=0 = m2 eϕ/2 for x < 0, (5.21)

imply
m1 m2
a + ā = √ = ω1 , αeiπν + ᾱe−iπν = − √ = −ω2 .
m m
Solving for a one gets

−ω2 − e−iπν ω1
a= .
2i sin πν
From the Gauss–Bonnet theorem one may expect that the geodesic curvature of boundary
components should be positive. It can be easily checked that this is really so in the symmetric
case ω1 = ω2 = ω, when the hyperbolic metric exists if the condition
1
−ω > 2 sin πν
2
is satisfied. We assume in the following that the boundary geodesic curvatures are positive and
such that the function

−ϕ/2 m  
e = |z| γ |z|−ν − β|z|ν − 2aeiνθ (5.22)

is positive for z = |z|eiθ in the upper half plane. Introducing parametrization

ω1 = −2 cosh t1 , ω2 = −2 cosh t2 , t1 , t2  0,
we have for the solution (5.22)
248 L. Hadasz, Z. Jaskólski / Nuclear Physics B 757 [FS] (2006) 233–258

   
∂Scl 1 1−ν 1 1+ν
= lim − κ|dz|ϕ + log + κ|dz|ϕ − log
∂ν →0 4π 2 4π 2
|z|= ,z>0 |z|= 1 ,z>0

m 1

= log ν − log − log |a|2 − 1


2 2

= log ν − log m + log sin πν

1 π(1 − ν) + i(t1 + t2 ) 1 π(1 − ν) − i(t1 + t2 )
− log sin − log sin
2 2 2 2

1 π(1 − ν) + i(t1 − t2 ) 1 π(1 − ν) − i(t1 − t2 )
− log sin − log sin ,
2 2 2 2
√ ∞  π(1−ν)+i(t1 +t2 )
π(1−ν)+i(t1 −t2 )

∂Scl m ϕ i sin sin ∂t1
= dx e 2 = log 2
π(1−ν)−i(t +t )
2
π(1−ν)−i(t −t )
,
∂ω1 4π 2π sin 1 2
sin 1 2 ∂ω1
0 2 2

√ 0  π(1−ν)+i(t1 +t2 )
π(1−ν)−i(t1 −t2 )

∂Scl m ϕ i sin sin ∂t2
= dx e 2 = log 2
π(1−ν)−i(t +t
2
π(1−ν)+i(t
.
∂ω2 4π 2π 1 2 ) 1 −t2 ) ∂ω
sin 2 sin 2 2
−∞

Integrating these equations one gets (up to a constant)


  π(1 − ν) + i(τ1 t1 + τ2 t2 )
1
Scl = ν log ν − log m − 1 + s(πν) + s ,
2 τ1 =± τ2 =±
2
(5.23)
where
x
1
def
s(x) = dy log sin y.
π
π/2

The quantum boundary two-point coupling constant has the form [3]:

[πμγ (b2 )b2−2b ](Q−2β)/2b b (2β − Q)b−1 (Q − 2β)


2

d(β|s1 , s2 ) =



, (5.24)
Sb β + i s1 +s
2
2
Sb β + i s1 −s
2
2
Sb β − i s1 −s
2
2
Sb β − i s1 +s
2
2

with appropriate counterparts of relations (5.6), (5.7) assumed. Taking into account

1
ti = πbsi , β∼ (1 + ν),
2b
and the asymptotic behavior

1 log 2 1
log Sb (x) ∼ s(πbx) + xb − , (5.25)
b2 b2 2
one can check that the semiclassical asymptotic of (5.24) is given by the classical action (5.23)
indeed.
L. Hadasz, Z. Jaskólski / Nuclear Physics B 757 [FS] (2006) 233–258 249

5.4. One bulk, one boundary elliptic singularities

For T (z) having a single pole at the real axis and a single pole in the interior of the upper half
plane one can always choose the “boundary” pole to be located at z = ∞. The second pole we
shall take at z = z1 , z1 > 0. The energy–momentum tensor takes the form
δ δ δb − 2δ
T (z) = + + (5.26)
(z − z1 )2 (z − z̄1 )2 (z − z1 )(z − z̄1 )
with elliptic weights
1 − ξ2 1 − ν2
δ= , δb = .
2 2
A normalized basis in the space of solutions of (2.1), with diagonal SU(1, 1) monodromy around
z = z1 , can be chosen in the form
1 1−ξ 1+ξ z1 −z

ψ1 (z) = √ (z − z1 ) 2 (z − z̄1 ) 2 2 F1 1−ν


2 , 2 , 1 − ξ, z1 −z̄1 ,
1+ν
ξ(z1 − z̄1 )
1 1+ξ 1−ξ z1 −z

ψ2 (z) = √ (z − z1 ) 2 (z − z̄1 ) 2 2 F1 1−ν


2 , 2 , 1 + ξ, z1 −z̄1 .
1+ν
(5.27)
ξ(z1 − z̄1 )
The functions ψ1,2 (z) are analytic in the vicinity of the real axis (the cuts between the branching
points z = z1 , z = ∞ and z = z̄1 can be chosen such that they do not intersect the real axis).
A real, single-valued around z = z1 solution to the Liouville equation (2.5) can be expressed
through ψ1 (z), ψ2 (z) as

m  2  2

e−ϕ/2 = a ψ1 (z) − a −1 ψ2 (z) , (5.28)


2
with a (real) constant a to be determined from (2.6).
To this end it is convenient to express ψi (z) in terms of ψi (z̄). Using the formulae for analytic
continuation of the hypergeometric functions one gets:
   
ψ1 (z) ψ1 (z̄)
=C , (5.29)
ψ2 (z) ψ2 (z̄)
where
⎡ (1−ξ )(ξ ) (1−ξ )(−ξ ) ⎤
( 1−ν 1+ν
2 )( 2 ) ( 1−ν 1+ν
2 −ξ )( 2 −ξ )
C =i⎣ (1+ξ )(ξ ) (1+ξ )(−ξ )
⎦. (5.30)
( 1−ν 1+ν
2 +ξ )( 2 +ξ ) ( 1−ν 1+ν
2 )( 2 )
Hence

−ϕ/2 m
e = Ψ (z)T · A · C · Ψ (z̄), (5.31)
2
where A = diag(a, −a −1 ). The boundary condition

 mB m
∂y e−ϕ/2 y=0 = − ≡− ωB
2 2
yields
(1 − ξ )(−ξ ) (1 + ξ )(ξ )
ωB = −a 1−ν

− a −1 1−ν
1+ν
. (5.32)
 2 − ξ  1+ν 2 − ξ  2 +ξ  2 +ξ
250 L. Hadasz, Z. Jaskólski / Nuclear Physics B 757 [FS] (2006) 233–258

Solving with respect to a one gets



1+ν

 1−ν2 − ξ  2 − ξ (1 + ξ )
a± =
π(1 − ξ )
 2
ωB πν ωB
× sin πξ ± cos 2 + − 1 sin πξ .
2
(5.33)
2 2 4
Let us note that the change of sign in a is equivalent to the change of sign of ωB and the
change of sing of the r.h.s. of (5.28). It does not lead therefore to any new solutions of the
Liouville equation. With no loss of generality we can then work with a = a+ .
It should be stressed that not for all parameters ωB , ξ, ν, for which a is real, the formula (5.28)
yields a regular solution for the Liouville equation. Indeed, the r.h.s. of (5.28) may change sign
on the upper half plane and the zero lines appearing in this situation give rise to singular lines
of the corresponding hyperbolic geometry. Even in the simple situation at hand, the problem of
determining the ranges of parameters for which a regular solution exists and its classical action
is well defined is rather involved and we are not aware of any compete solution to it. In the
following we simply assume that ωB , ξ, ν are such that a regular solution does exist.
Taking into account the asymptotic behavior of the solution Ψ (5.27) for z → z1 and (5.33)
one gets from (3.3):
∂Scl 1
= − log m − ξ log |z1 − z̄1 | + log 2
∂ξ 2


2 −ξ
(ξ ) 1+ν ν
− log
+ log cos π +ξ
(1 − ξ ) 1+ν
2 +ξ
2
  2 
ωB πν ωB
− log sin πξ + cos 2 + − 1 sin πξ .
2
(5.34)
2 2 4
Using the asymptotic behavior of the hypergeometric function for large arguments one finds
for z → ∞:
ν
(z1 − z̄1 )− 2 (ν)(1 − ξ ) ν−ξ 1+ξ
ψ1 (z) ∼ √ 1+ν
1+ν
(z − z1 ) 2 (z − z̄1 ) 2 ,
ξ  2  2 −ξ
ν
(z1 − z̄1 )− 2 (ν)(1 + ξ ) ν+ξ 1−ξ
ψ2 (z) ∼ √ 1+ν
1+ν
(z − z1 ) 2 (z − z̄1 ) 2 , (5.35)
ξ  2  2 +ξ
and
 1+ν 
∂Scl 1 ν ( 2 ) 1
= − log m + log |z1 − z̄1 | + log − log π
∂ν 4 2 (ν) 2
1  1+ν
1+ν
 1
+ log  2 − ξ  2 + ξ + log 2
2  2 
1 B+ B−
− log

, (5.36)
2 sin πξ cos π ν2 + ξ sin πξ cos π ν2 − ξ
where
 2
ωB πν ωB
B± = ± sin πξ + cos2 + − 1 sin2 πξ .
2 2 4
L. Hadasz, Z. Jaskólski / Nuclear Physics B 757 [FS] (2006) 233–258 251

In order to calculate the ω-derivative one needs the map ρ (3.7) defined in terms of the
SL(2, R)-monodromy solution to the Fuchs equation, which can be easily constructed from the
SU(1, 1)-monodromy one:
   √ 
ψ̃1 (z) 1 i a − √ia  ψ1 (z) 
Ψ̃ (z) = =√ √ .
ψ̃2 (z) 2i a √1 ψ2 (z)
a
Using (5.35) one finds the z → ∞ asymptotic
ψ̃1 (z) er (z − z1 )−ξ (z − z̄1 )ξ − 1
ρ(z) ≡ ∼i ,
ψ̃2 (z) er (z − z1 )−ξ (z − z̄1 )ξ + 1
where
1+ν

(1 − ξ ) +ξ
e ≡ r
2
a
(1 + ξ ) 1+ν
2 −ξ
  2 
1 ωB πν ωB
=
sin πξ + cos2 + − 1 sin πξ .
2
2 + πξ
cos πν 2 2 4
Calculating the limits along the real axis
sinh r + i sin 2πξ sinh r
ρ1 = lim ρ(x) = i , ρ2 = lim ρ(x) = i ,
x→−∞ cosh r + cos 2πξ x→∞ cosh r + 1
one obtains
|ρj +1 − ρj | sin πξ
 = .
2 ρj +1 ρj sinh r
Hence for ωB > 2 formula (3.10) implies

 2 
∂Scl 1 sin πξ ωB
=  arcsin −1 ,
∂ωB π ω2 − 4 sinh r 2
B
or
 
∂Scl 1 sinh t sin πξ
= arcsin , (5.37)
∂t π sinh r
where ωB = 2 cosh t .
Checking integrability conditions or direct integration of Eqs. (5.34), (5.36) and (5.37) is
rather involved and not especially illuminating. We shall check instead that (5.34), (5.36) and
(5.37) coincide with the corresponding derivatives of the classical action obtained from the clas-
sical limit of the exact quantum expression.
The bulk-boundary correlation function for the location of the bulk operator at z = z1 and the
boundary operator at z = ∞ is given by [6]:
  P (α, β|s)I (α, β|s)
Vα (z1 )Bβss (∞) = ,
|z1 − z̄1 |2Δα −Δβ

2  1 (Q−2α−β)
P (α, β|s) = −2πi πμγ b2 b2−2b 2b
b3 (Q − β)b (2α − β)b (2Q − 2α − β)
× ,
b (Q)b (Q − 2β)b (β)b (Q − 2α)b (2α)
252 L. Hadasz, Z. Jaskólski / Nuclear Physics B 757 [FS] (2006) 233–258


Sb (u + β/2 + α − Q/2)Sb (u + β/2 − α + Q/2)
I (α, β|s) = du e−2πus ,
Sb (u − β/2 − α + 3Q/2)Sb (u − β/2 + α + Q/2)
iR

where the relations (5.6) and (5.7) are still assumed.


Using the asymptotic of the Barnes gamma function
 
x 1 1 1 1 1 2
log b ∼ − 2 g(x) + x− log 2π + x− log b ,
b b 2 2 2 2
where
x
g(x) = dy log (y),
1
2

one obtains

1 1 1−ν 1−ν
log P (α, β|s) ∼ − − ξ log m + − ξ log 2 + ν log 2π
b2 2 2 2

1+ν 1−ν
− 3g +g + g(ν) + g(ξ ) + g(1 − ξ )
2 2

1+ν 1+ν
−g −ξ −g +ξ .
2 2
Rescaling the integration variable u → y = bu and using (5.25) one has
  
1 1
I (α, β|s) ∼ dy exp 2 f(y, t, ξ, ν) ,
b b
iR

where

f(y, t, ξ, ν) = −(1 + ν) log 2 − 2ty (5.38)



π π
+ s πy + (1 − ν − 2ξ ) + s πy + (1 − ν + 2ξ )
4 4

π π π π
− s πy + (1 + ν − 2ξ ) + − s πy + (1 + ν + 2ξ ) + . (5.39)
4 2 4 2
This integral can be approximated by its saddle point value. The saddle point equation


sin πy + π4 (1 − ν − 2ξ ) sin πy + π4 (1 − ν + 2ξ ) cos πξ + sin 2πy − πν


e =
2t

= 2

cos πy + π4 (1 + ν − 2ξ ) cos πy + π4 (1 + ν + 2ξ ) cos πξ − sin 2πy + πν


2
yields two solutions

1 πν
cos 2πys± = sinh2 t cos πξ sin
cosh t2
− sin2 πν
2
2
 2 
πν πν ωB
± cosh t cos cos2 + − 1 sin2 πξ . (5.40)
2 2 4
L. Hadasz, Z. Jaskólski / Nuclear Physics B 757 [FS] (2006) 233–258 253

The appropriate solution could be in principle selected by a careful analysis of the position of the
contour with respect to poles located on the real axis. In the present calculations we have chosen
ys = ys+ on the basis of numerical checks of the final result instead.
The asymptotic takes the form
1

log I (α, β|s) ∼ 2


f ys (t, ξ, ν), t, ξ, ν ,
b
where ys (t, ξ, ν) = ys± (t, ξ, ν). The classical action calculated from the classical limit of the
quantum expression
  b→0 − 1 S
Vα (z1 )Bβss (∞) ∼ e b2 cl
reads

1 1−ν 1 ν2 ξ 2 1−ν
Scl = − ξ log m + + − log |z1 − z̄1 | − − ξ log 2
2 2 4 4 2 2
+ log 2 + g(ξ ) + g(1 − ξ )

1+ν 1−ν
− ν log π + 3g −g − g(ν)
2 2

1+ν 1+ν
+g −ξ +g + ξ + 2tys
2 2

π π
− s πys + (1 − ν − 2ξ ) − s πys + (1 − ν + 2ξ )
4 4

π π π π
+ s πys + (1 + ν − 2ξ ) + + s πys + (1 + ν + 2ξ ) + ,
4 2 4 2
and therefore
∂Scl 1
= − log m + −ξ log |z1 − z̄1 | + log 2
∂ξ 2

 

2 −ξ
(ξ ) 1+ν 1 cos 2πys − sin πξ + πν

,
− log
+ log 2
(1 − ξ ) 1+ν
2 +ξ
2 cos 2πys + sin πξ − πν2
(1)  1+ν 
∂Scl 1 ν ( 2 ) 1
= − log m + log |z1 − z̄1 | + log − log π
∂ν 4 2 (ν) 2
 
1 1+ν 1+ν 1 1+ν
+ log  −ξ  +ξ − log cos π
2 2 2 2 2
 
1 πν πν
+ log cos 2πys − sin πξ + cos 2πys + sin πξ − ,
4 2 2
(1) 
∂Scl 1
= 2ys = arcsin 1 − cos2 (2πys ).
∂t π
These equations have to be compared with Eqs. (5.34), (5.36) and (5.37). We have checked the
corresponding equalities by Mathematica 5.2 obtaining a perfect agreement in all three cases.
Similar calculations can be also performed in the case of the Liouville boundary three point
function derived in [11]. This function is particularly interesting due to its relation with the
braiding matrix for the BPZ conformal block [11,12,39]. The resulting equations for the classical
254 L. Hadasz, Z. Jaskólski / Nuclear Physics B 757 [FS] (2006) 233–258

action are however complicated and we have not succeeded in finding a parametrization that
would allow to integrate them in a closed form.

Acknowledgements

This work was partially supported by the Polish State Research Committee (KBN) grant No. 1
P03B 025 28. The research of L.H. is supported by the Alexander von Humboldt Foundation
scholarship. Part of the ideas contained in this work emerged during the 2005 Simons Work-
shop in Physics and Mathematics. L.H. thanks the organizers of the Workshop for their warm
hospitality, which created at Stony Brook a highly stimulating environment and for the financial
support.

Appendix A

We shall calculate the length of the j th boundary component


 
|dρ|
j = dx e ϕ̃(z,z̄)/2
= ,

Lj ρ(Lj )

in terms of the endpoints ρj = limx→xj + ρ(x), ρj +1 = limx→xj +1 − ρ(x). For arbitrary two
points ρj , ρj +1 ∈ H one can always find an SL(2, R) transformation w(ρ) such that

w(ρj +1 ) = −w(ρj ) > 0, w(ρj +1 ) = w(ρj ).


It is then sufficient to calculate the hyperbolic length of the curve w ◦ ρ(Lj ) connecting the
points qj = w(ρj ), qj +1 = w(ρj +1 ). Let us note that the sign of the boundary geodesic curva-
ture depends on the location of ρ(M) with respect to ρ(Lj ), so one can assume ωj > 0 in the
following.
ω
For 0 < ωj  2 there are always two curves with the geodesic curvature 2j connecting points
qj , qj +1 (see Fig. 1). These are arcs of two circles with their centers on the imaginary axis and
their radii determined by the condition that the Euclidean distance of the center from the real axis
ω
equals 2j R:

2qj  

R± = ±ωj + 4 − ωj2 β 2 + 4 ,
4 − ωj2

where β = β(qj , qj +1 ) and

|z − w|
β(z, w)= √
2 zw
is an SL(2, R)-invariant.
The hyperbolic lengths of the corresponding arcs can be easily calculated

 ±
π−ϑ  
2dϑ 8 2 ± ωj π θ±
± = = arctanh tan − ,
2 sin ϑ ∓ ωj 4 − ωj2 2 ∓ ωj 4 2
ϑ±
L. Hadasz, Z. Jaskólski / Nuclear Physics B 757 [FS] (2006) 233–258 255

Fig. 1. The geometry involved in the determination of the boundary length, ωj < 2.

ωj
where θ± = Arg(qj ± i 2 R± ). It follows from Fig. 1 that
|qj +1 −qj |
π θ± (2 ∓ ωj )β
tan − =  2
=  .
4 2 |qj +1 −qj |2 2 + (4 − ωj2 )β 2 + 4
R± + 2 −
R± 4

Using the identity


x
2 arctanh √ = arcsinh x
1 + 1 + x2
one gets

 2 
4 ωj
± =  arcsinh 1 − β(qj , qj +1 ) . (A.1)
4 − ωj2 2

Since the lengths of both arcs are the same one obtains j = ± which yields formula (3.9).
In the case ωj > 2 the points qj , qj +1 can be connected by curves with the geodesic curvature
ωj
2 if and only if (ωj − 4)β  4 (which is satisfied for β = β(qj , qj +1 ) by construction). One
2 2

has two circles with the radii


2qj  

R± = 2 ωj ± 4 − ωj2 β 2 + 4 ,
ω −4
and four different arcs (see Fig. 2).
256 L. Hadasz, Z. Jaskólski / Nuclear Physics B 757 [FS] (2006) 233–258

Fig. 2. The geometry involved in the determination of the boundary length, ωj > 2.

Let + −
± , ± be the lengths of the upper, and lower arcs of the circle with radius R± , respec-
tively. Following essentially the same calculations as above and using
x
2 arctan √ = arcsin x,
1 + 1 − x2
x
2 arctan √ = π − arcsin x (0 < x < 1),
1 − 1 − x2
one gets

 2 
+ − 4 ωj
− = + =  arcsin − 1 β(qj , qj +1 ) ,
ω2 − 4 2
j

 2 
4 4 ωj
−
− = +
+ = π− arcsin − 1 β(qj , qj +1 ) . (A.2)
ωj2 − 4 ωj2 − 4 2

As the image w ◦ ρ(Lj ) of the j th boundary component is one of the four arcs, its hyperbolic
length is either + −
− or − . On the other hand, the length of the Lj depends analytically on ωj , so
the formula for |ωj | > 2 should be an analytic continuation of that for |ωj | < 2. Taking this into
account one finally gets j = + −
− = + , what proves the formula (3.10).

References

[1] Y. Nakayama, Liouville field theory: A decade after the revolution, Int. J. Mod. Phys. A 19 (2004) 2771, hep-
th/0402009.
[2] J. Teschner, Remarks on Liouville theory with boundary, hep-th/0009138.
[3] V. Fateev, A.B. Zamolodchikov, A.B. Zamolodchikov, Boundary Liouville field theory. I: Boundary state and bound-
ary two-point function, hep-th/0001012.
L. Hadasz, Z. Jaskólski / Nuclear Physics B 757 [FS] (2006) 233–258 257

[4] A.B. Zamolodchikov, A.B. Zamolodchikov, Liouville field theory on a pseudosphere, hep-th/0101152.
[5] B. Ponsot, J. Teschner, Boundary Liouville field theory: Boundary three point function, Nucl. Phys. B 622 (2002)
309, hep-th/0110244.
[6] K. Hosomichi, Bulk-boundary propagator in Liouville theory on a disc, JHEP 0111 (2001) 044, hep-th/0108093.
[7] B. Ponsot, Liouville theory on the pseudosphere: Bulk-boundary structure constant, Phys. Lett. B 588 (2004) 105,
hep-th/0309211.
[8] J. Teschner, On the Liouville three point function, Phys. Lett. B 363 (1995) 65, hep-th/9507109.
[9] H. Dorn, H.J. Otto, Two and three point functions in Liouville theory, Nucl. Phys. B 429 (1994) 375, hep-
th/9403141.
[10] A.B. Zamolodchikov, A.B. Zamolodchikov, Structure constants and conformal bootstrap in Liouville field theory,
Nucl. Phys. B 477 (1996) 577.
[11] B. Ponsot, J. Teschner, Liouville bootstrap via harmonic analysis on a noncompact quantum group, hep-th/9911110.
[12] B. Ponsot, J. Teschner, Clebsch–Gordan and Racah–Wigner coefficients for a continuous series of representations
of Uq (sl(2, R)), Commun. Math. Phys. 224 (2001) 613, math.QA/0007097.
[13] J. Teschner, Liouville theory revisited, Class. Quantum Grav. 18 (2001) R153, hep-th/0104158.
[14] C.B. Thorn, Liouville perturbation theory, Phys. Rev. D 66 (2002) 027702, hep-th/0204142.
[15] E. Braaten, T. Curtright, C.B. Thorn, An exact operator solution of the quantum Liouville field theory, Ann.
Phys. 147 (1983) 365.
[16] E. Braaten, T. Curtright, G. Ghandour, C.B. Thorn, Nonperturbative weak coupling analysis of the Liouville quan-
tum field theory, Phys. Rev. Lett. 51 (1983) 19.
[17] E. Braaten, T. Curtright, G. Ghandour, C.B. Thorn, Nonperturbative weak coupling analysis of the quantum Liou-
ville field theory, Ann. Phys. 153 (1984) 147.
[18] P. Menotti, E. Tonni, The tetrahedron graph in Liouville theory on the pseudosphere, Phys. Lett. B 586 (2004) 425,
hep-th/0311234.
[19] A.M. Polyakov, Quantum geometry of bosonic strings, Phys. Lett. B 103 (1981) 207.
[20] F. David, Conformal field theories coupled to 2-D gravity in the conformal gauge, Mod. Phys. Lett. A 3 (1988)
1651.
[21] J. Distler, H. Kawai, Conformal field theory and 2-D quantum gravity or who’s afraid of Joseph Liouville?, Nucl.
Phys. B 321 (1989) 509.
[22] N. Seiberg, Notes on quantum Liouville theory and quantum gravity, in: Common Trends in Mathematics and
Quantum Field Theory, Proceedings of the 1990 Yukawa International Seminar, Prog. Theor. Phys. Suppl. 102
(1990) 319.
[23] J. Polchinski, A two-dimensional model for quantum gravity, Nucl. Phys. B 324 (1989) 123.
[24] P.H. Ginsparg, G.W. Moore, Lectures on 2-D gravity and 2-D string theory, hep-th/9304011.
[25] A.M. Polyakov, as reported in Refs. [26,27,33,34].
[26] L.A. Takhtajan, Semiclassical Liouville theory, complex geometry of moduli spaces, and uniformization of Riemann
surfaces, in: Cargese 1991, Proceedings, New Symmetry Principles in Quantum Field Theory, pp. 383–406.
[27] L.A. Takhtajan, Topics in quantum geometry of Riemann surfaces: Two-dimensional quantum gravity, in: Quantum
Groups and Their Applications in Physics, IOS Press, 1994, pp. 541–580, hep-th/9409088.
[28] L.A. Takhtajan, L.P. Teo, Quantum Liouville theory in the background field formalism. I: Compact Riemann sur-
faces, hep-th/0508188.
[29] P. Menotti, E. Tonni, Standard and geometric approaches to quantum Liouville theory on the pseudosphere, Nucl.
Phys. B 707 (2005) 321, hep-th/0406014.
[30] P. Menotti, G. Vajente, Semiclassical and quantum Liouville theory on the sphere, Nucl. Phys. B 709 (2005) 465,
hep-th/0411003.
[31] P. Menotti, E. Tonni, Liouville field theory with heavy charges. I: The pseudosphere, hep-th/0602206.
[32] P. Menotti, E. Tonni, Liouville field theory with heavy charges. II: The conformal boundary case, hep-th/0602221.
[33] P.G. Zograf, L.A. Takhtajan, On Liouville equation, accessory parameters and the geometry of Teichmüller space
for Riemann surface of genus 0, Math. USSR Sbornik 60 (1988) 143.
[34] P.G. Zograf, L.A. Takhtajan, On uiformization of Riemann surfaces and the Weil–Petersson metric on Teichmüller
and Schottky spaces, Math. USSR Sbornik 60 (1988) 297.
[35] P.G. Zograf, The Liouville action on moduli spaces, and uniformization of degenerating Riemann surfaces,
Leningrad Math. J. 1 (1990) 941.
[36] L. Cantini, P. Menotti, D. Seminara, Proof of Polyakov conjecture for general elliptic singularities, Phys. Lett. B 517
(2001) 203, hep-th/0105081.
[37] L. Takhtajan, P. Zograf, Hyperbolic 2-spheres with conical singularities, accessory parameters and Kähler metrics
on M0,n , Trans. Amer. Math. Soc. 355 (2003) 1857–1867, math.CV/0112170.
258 L. Hadasz, Z. Jaskólski / Nuclear Physics B 757 [FS] (2006) 233–258

[38] L. Hadasz, Z. Jaskolski, Polyakov conjecture for hyperbolic singularities, Phys. Lett. B 574 (2003) 129, hep-
th/0308131.
[39] A.A. Belavin, A.M. Polyakov, A.B. Zamolodchikov, Infinite conformal symmetry in two-dimensional quantum
field theory, Nucl. Phys. B 241 (1984) 333.
[40] A.B. Zamolodchikov, Two-dimensional conformal symmetry and critical four-spin correlation functions in the
Ashkin–Teller model, Sov. Phys. JEPT 63 (5) (1986) 1061.
[41] A.B. Zamolodchikov, Conformal symmetry in two-dimensional space: Recursion representation of conformal block,
Theor. Math. Phys. 73 (1987) 1088.
[42] L. Hadasz, Z. Jaskolski, M. Piatek, Classical geometry from the quantum Liouville theory, Nucl. Phys. B 724 (2005)
529, hep-th/0504204.
[43] L. Hadasz, Z. Jaskolski, Liouville theory and uniformization of four-punctured sphere, hep-th/0604187.
[44] A. Bilal, J.L. Gervais, Exact quantum three point function of Liouville highest weight states, Nucl. Phys. B 305
(1988) 33.
[45] L. Hadasz, Z. Jaskolski, Classical Liouville action on the sphere with three hyperbolic singularities, Nucl. Phys.
B 694 (2004) 493, hep-th/0309267.
Nuclear Physics B 757 [FS] (2006) 259–279

QCD with bosonic quarks at nonzero chemical potential


K. Splittorff a,∗ , J.J.M. Verbaarschot b
a The Niels Bohr Institute, Blegdamsvej 17, DK-2100, Copenhagen Ø, Denmark
b Department of Physics and Astronomy, SUNY, Stony Brook, NY 11794, USA

Received 31 May 2006; accepted 4 September 2006


Available online 4 October 2006

Abstract
We formulate the low energy limit of QCD like partition functions with bosonic quarks at nonzero chem-
ical potential. The partition functions are evaluated in the parameter domain that is dominated by the zero
momentum modes of the Goldstone fields. We find that partition functions with bosonic quarks differ struc-
turally from partition functions with fermionic quarks. Contrary to the theory with one fermionic flavor,
where the partition function in this domain does not depend on the chemical potential, a phase transition
takes place in the theory with one bosonic flavor when the chemical potential is equal to mπ /2. For a pair
of conjugate bosonic flavors the partition function shows no phase transition, whereas the fermionic coun-
terpart has a phase transition at μ = mπ /2. The difference between the bosonic theories and the fermionic
ones originates from the convergence requirements of bosonic integrals resulting in a noncompact Gold-
stone manifold and a covariant derivative with the commutator replaced by an anti-commutator. For one
bosonic flavor the partition function is evaluated using a random matrix representation.
© 2006 Published by Elsevier B.V.

1. Introduction

Bosonic quarks appear in QCD whenever the weight of the partition function includes an
inverse determinant of the Dirac operator. Ratios of determinants occur frequently in analytical
approaches to QCD. Inverse determinants are for example used to quench a flavor from the
theory [1,2]. This form of quenching is useful in describing quenched and partially quenched
[3,4] lattice results and is an essential ingredient when computing the spectral properties of the
QCD Dirac operator. In particular, if we consider the spectral correlation functions of the Dirac

* Corresponding author.
E-mail address: split@alf.nbi.dk (K. Splittorff).

0550-3213/$ – see front matter © 2006 Published by Elsevier B.V.


doi:10.1016/j.nuclphysb.2006.09.011
260 K. Splittorff, J.J.M. Verbaarschot / Nuclear Physics B 757 [FS] (2006) 259–279

operator in the microscopic limit [5,6], the theories with bosonic flavors are not only a part of the
calculation, they are also part of the result [7–9].
The microscopic limit is an extreme version of the chiral limit where the sea and valence
quark masses times the volume are kept fixed as the volume is taken to infinity. In this limit the
zero modes of the Goldstone bosons associated with chiral symmetry breaking dominate the low
energy partition function which reduces to a group integral uniquely determined by the pattern
of chiral symmetry breaking [10]. Sometimes this limit is referred to as the -limit but since the
term microscopic limit appeared earlier in the literature (see for example [11,12]) we prefer to
use this term. For a review of QCD in the microscopic limit see [13].
In this paper we will consider the low energy limit of QCD at nonzero chemical potential. For
fermionic quark flavors the baryon chemical potential does not affect the low energy effective
partition function at zero temperature. The reason is that the lightest degrees of freedom (the
pions) have zero baryon charge. Because the fermion determinant is complex, QCD at nonzero
baryon chemical potential cannot be simulated directly on the lattice by probabilistic methods.
That is why sometimes a theory is considered where the fermion determinant is replaced by
its absolute value. For an even number of flavors this corresponds to the product of a fermion
determinant and its complex conjugate in the weight of the partition function. The low energy
partition function for QCD with flavors and conjugate flavors depends on the chemical potential
[14]. The conjugate fermionic flavors correspond to ordinary fermionic flavors with the opposite
sign of the chemical potential [15]. A theory with one fermionic flavor and a conjugate fermionic
flavor is therefore identical to a theory with two fermionic flavors at nonzero isospin chemical
potential. Since the pions have nonzero isospin charge this low energy effective partition function
depends on the chemical potential. In particular, a phase transition to a Bose–Einstein condensed
phase takes place at zero temperature when the isospin chemical potential reaches mπ /2, see
[14,16,17,19,20].
In this paper we examine the properties of the low energy QCD partition function with bosonic
quarks at nonzero chemical potential. We find that the bosonic theories behave completely dif-
ferent from their fermionic counterparts. The bosonic partition functions depend on the chemical
potential even in the absence of conjugate bosonic flavors. By explicit computation of the par-
tition function with one bosonic flavor in the microscopic limit we find that this theory has a
phase transition when the chemical potential, μ, reaches mπ /2. On the contrary, in the theory
with one conjugate pair of bosonic quarks we find that the free energy as well as its derivative are
continuous at μ = mπ /2 and zero temperature. The difference in the phase structure between the
bosonic and the fermionic theories has its origin in the convergence requirements of bosonic in-
tegrals which forces us to rewrite the determinants in the partition function as the determinant of
a Hermitian matrix (also known as hermitization [21–23]). This procedure leads to a noncompact
Goldstone manifold and, as will be shown here, a covariant derivative where the commutator is
replaced by an anti-commutator. A comparison of bosonic and fermionic partition functions is
given in Table 1.
In order to compute the partition function with one bosonic flavor we make use of the random
matrix representation. All other partition functions discussed here have been derived directly
from the low energy effective theory.
The outline of this paper is as follows. We first discuss in general terms the low energy physics
of QCD with bosonic quarks at nonzero chemical potential. As a warm up exercise we review
results for fermionic quarks at nonzero chemical potential. We then turn to the theories with
conjugate pairs of quarks. In Section 6 we present a random matrix model and perform the
calculation of the partition function with one bosonic flavor in the microscopic limit. Taking the
K. Splittorff, J.J.M. Verbaarschot / Nuclear Physics B 757 [FS] (2006) 259–279 261

thermodynamic limit of this result then allows us to examine the phase diagram of QCD with
one bosonic quark.

2. General discussion

In this section we give a general discussion of QCD-like partition functions at nonzero chemi-
cal potential. The Euclidean QCD partition function at nonzero chemical potential for Nf flavors
with mass m is given by
 
Z Nf (m; μ) = detNf (D + μγ0 + m) . (1)
Here and below · · · denotes the average with respect to the Yang–Mills action. If Nf > 0
the quarks are fermions while for negative Nf they are interpreted as bosons. With fermionic
quarks at zero temperature this partition function is independent of the chemical potential for
μ < mN /3, which immediately follows from the definition of the grand canonical partition func-
tion as an average over Boltzmann factors and fugacities. In terms of the representation (1), the
μ-independence of the free energy in the thermodynamic limit can be understood from the gauge
transformation
D + μγ0 + m = e−μt [D + m]eμt , (2)
so that the factor exp(±μt) can be absorbed in the boundary conditions of the fermionic fields.
In the phase that is not sensitive to the boundaries, the partition function does not depend on μ.
A nonzero baryon density is obtained by baryons winding around the torus in the time direc-
tion. In this phase the boundary conditions are important and the partition function becomes
μ-dependent.
The second partition function we consider is the phase quenched partition function (the su-
perscript n counts the pairs of conjugate fermionic, n > 0, or bosonic, n < 0, determinants)
   
Z n=1 m, m∗ ; μ = det(D + μγ0 + m) det(D + μγ0 + m)† (3)
  ∗

= det(D + μγ0 + m) det −D + μγ0 + m ,
so that μ can be interpreted as the isospin chemical potential [15]. Also in this case the chemical
potential can be gauged into the boundary conditions, this time by a gauge transformation in
isospin space. For μ < mπ /2 the partition function is μ-independent at zero temperature. Again,
this also follows from the zero temperature limit of the Boltzmann factors.
Next, let us consider the bosonic partition function
 
  1
Z n=−1 m, m∗ ; μ = (4)
det(D + μγ0 + m) det(D + μγ0 + m)†
 
1
= .
det(D + μγ0 + m) det(−D + μγ0 + m∗ )
Because of the nonhermiticity, the inverse determinants cannot be written directly as a convergent
bosonic integral. However, this can be achieved by introducing the infinitesimal regulator 

−1
 ∗
  D + μγ0 + m
Z n=−1
m, m ; μ = det . (5)
−D + μγ0 + m∗ 
The parameter  may be regarded as source for pions composed of bosonic anti-quarks and
conjugate bosonic quarks. Expressing the partition function as an integral over the eigenvalues
262 K. Splittorff, J.J.M. Verbaarschot / Nuclear Physics B 757 [FS] (2006) 259–279

of the Hermitian matrix in (5) one can easily convince oneself that the partition function diverges
logarithmically in . In [8] this was shown explicitly by performing the integrals in the low-
energy limit of this partition function. Because of the -term it is not possible to gauge away the
μ-dependence of the partition function, and we do not expect to find a phase where the partition
function is μ-independent. Indeed, we show below that the partition function (5) is μ-dependent
for all values of μ. Therefore the pion condensate, which has  as source term, is nonzero for all
values of μ. In particular this implies that there must be a massless mode in the spectrum. We
identify this mode explicitly below.
Finally, we discuss the partition function
 
1
Z Nf =−1 (m; μ) = . (6)
det(D + μγ0 + m)
It is not possible to write this partition function as a convergent bosonic quark integral. To prop-
erly define the partition function we have to rewrite it as
 
Nf =−1 det(−D + μγ0 + m∗ )
Z (m; μ) = (7)
det(D + μγ0 + m) det(−D + μγ0 + m∗ )
so that the inverse determinant can be represented as a bosonic integral after it has been regular-
ized as in (5). The particle content of this partition function is one conjugate fermionic quark,
one bosonic quark, and one conjugate bosonic quark. Because of the extra determinant in the nu-
merator, the limit  → 0 is finite in this case, i.e. the leading term is of order  0 . In the partition
function the gauge symmetry therefore allows us to transform the μ dependence to the bound-
aries allowing for a phase transition to occur. As we will show below a phase exists where this
partition function does not depend on μ. For μ > mπ /2 this phase gives way to a μ dependent
phase.
A summary of the different partition functions discussed in this section is given in Table 1.
Notice that the regularization enters because the chemical potential breaks the hermiticity
properties of the Dirac operator. If we where to consider an imaginary chemical potential, the
Dirac spectrum would remain on the imaginary axis and no regularization of the bosonic theories
would be required [24].

3. Fermionic partition functions

In the microscopic limit the zero momentum modes of the pions dominate the low energy
effective partition function of QCD in the sense that the nonzero momentum modes factorize
from the partition function leaving us with a group integral over the zero momentum modes.
This integral is uniquely determined by the pattern of chiral symmetry breaking, and in the case
of bosonic quarks, by the convergence of the integrals. Before turning to the theories with bosonic

Table 1
Summary of properties of low energy QCD at nonzero chemical potential and zero temperature
Theory Number of charged Goldstone modes for μ < μc Critical chemical potential
det(D + μγ0 + m) 0 μc = 1m
3 N
| det(D + μγ0 + m)|2  2 μc = 1m
  2 π
1 4 μc = 1m
det(D+μγ0 +m) 2 π
 1 
nonapplicable μc = 0
| det(D+μγ0 +m)|2
K. Splittorff, J.J.M. Verbaarschot / Nuclear Physics B 757 [FS] (2006) 259–279 263

quarks, as a warm-up exercise, we recall results obtained for fermionic quarks. The emergence
of the phase structure is discussed as well.
With fermionic quarks the partition function is automatically convergent and we need only
worry about the symmetries. The Lagrangian is determined by local gauge invariance in isospin
space [16]. For Nf + n ordinary fermionic quarks and n conjugate fermionic quarks it is given
by [18]
1 Σ  
L = Fπ2 Tr ∇ν U ∇ν U −1 − Tr M U + U −1 (8)
4 2
with U ∈ SU(Nf + 2n), and the quark mass matrix is given by M = diag(m1 , . . . , mNf , {z}n ,
{z∗ }n ). The charge matrix B is the diagonal matrix B = diag(1, . . . , 1Nf +n , −1, . . . , −1n ). The
chiral Lagrangian is parameterized by two low energy constants, the chiral condensate, Σ , and
the pion decay constant Fπ . The covariant derivatives are defined by

∇ν U = ∂ν U + μδν 0 [U, B], ∇ν U −1 = ∂ν U −1 + μδν 0 U −1 , B . (9)
In the microscopic limit the nonzero momentum modes factorize from the partition function
[10] so that the zero momentum part of the partition function in the sector of zero topological
charge is given by1

  V 2 2 −1 1 −1
Z Nf ,n {mf }, z, z∗ ; μ = dU e− 4 Fπ μ Tr[U,B][U ,B] + 2 ΣV Tr M(U +U ) . (10)
U ∈U (Nf +2n)

Note that the chemical potential and the quark masses only appear through the dimensionless
combinations

μ̂ ≡ μFπ V and M̂ ≡ MΣV , (11)
where V is the volume of space-time. In the absence of conjugate quarks the charge matrix B is
the unit matrix and the fermionic partition function is independent of the chemical potential. For
example the partition function with one fermionic flavor is [5]

Z Nf =1 (m̂) = I0 (m̂). (12)


An explicit expression for fermionic partition functions was derived in [8,29]. Here we take
a closer look at the simplest nontrivial example which is given by the theory with one fermionic
quark and one conjugate fermionic quark both with real mass m,
1
dt te−2μ̂
2 2t 2
Z n=1 (m̂, m̂; μ̂) = 2e2μ̂ I0 (m̂t)2 . (13)
0
The chiral condensate follows by differentiation of the free energy,
Σ n=1 (m̂; μ̂) 1
= ∂m̂ log Z n=1 (m̂, m̂; μ̂). (14)
Σ 2
In Fig. 1 we have plotted the chiral condensate function of m (in units of mc = 2μ2 F 2 /Σ )
√ as a √
and as a function of μ (in units of μc = mΣ/( 2Fπ )). The full curves display a smooth

1 Without loss of generality we consider the partition function in the trivial topological sector.
264 K. Splittorff, J.J.M. Verbaarschot / Nuclear Physics B 757 [FS] (2006) 259–279

dependence as was to be expected for a finite volume. In the thermodynamic limit, the would-be
phase transition at m = mc in the left figure or at μ = μc in the right figure becomes sharp. In
terms of the dimensionless variables this limit is reached for m̂ → ∞ or μ̂ → ∞ and a kink
develops at the expected value of m̂/(2μ̂2 ) = 1.
In order to see how the kink is recovered we take the thermodynamic limit of (13) which is
given by a leading order saddle point approximation. Using the asymptotic form of the Bessel
function we obtain

1
2μ̂2 1 −2μ̂2 t 2 2t m̂
Z n=1
(m̂, m̂; μ̂) ∼ e dt e e . (15)

0

A phase transition occurs when the saddle point,


t= , (16)
2μ̂2
hits the integration boundary at t = 1, that is when m̂ = 2μ̂2 . In the phase where the saddle point
is inside the boundary we obtain

Σ n=1 (m̂; μ̂) m̂


∼ . (17)
Σ 2μ̂2
If the saddle point is outside the integration domain the chiral condensate is simply given by

Σ n=1 (m̂; μ̂)


= 1. (18)
Σ
These results are shown by the dashed curves in Fig. 1 and are in agreement with the results
obtained from chiral perturbation theory [16–19].


Fig. 1. The chiral condensate for n = 1 (one fermion and one conjugate fermion). Left: The mass dependence for μ̂ = 5
is shown by the full curve. In the thermodynamic limit (μ̂ → ∞, dashed curve) the condensate grows linearly with m
until it reaches a plateau at m = mc = 2μ2 Fπ2 /Σ . Right: The full curve displays the chemical potential dependence for
m̂ = 10 and the dashed curve shows the result in the thermodynamic limit (m̂ → ∞).
K. Splittorff, J.J.M. Verbaarschot / Nuclear Physics B 757 [FS] (2006) 259–279 265

4. The theory with a pair of conjugate bosonic quarks

We now turn to QCD with bosonic quarks at nonzero chemical potential and start off with the
theory with a pair of conjugate bosonic flavors (4). This theory is simpler than the theory with a
single bosonic quark since the latter involves both fermionic and bosonic flavors (see (7)).

4.1. Covariant derivatives

As discussed in Section 2, the n = −1 partition function diverges. To derive the effective


partition function we therefore need to take into account the convergence of the integrals in
addition to the symmetries. If we write the Dirac operator as

0 d
D= (19)
−d † 0
the determinant in (5) can be rearranged as
   
  0 z d + μ    z d + μ 0 

 0  −d † + μ z   z∗  0 d − μ 
 = † . (20)
 z∗ −d + μ  
0  d +μ 0  −z∗ 

 d† + μ z∗ 0    0 d † − μ −z  
For the purpose of studying the transformation properties, we rewrite the determinant of the
right-hand side in a 2 × 2 block notation as

ζ1 d + B1
. (21)
d † + B2 ζ2
This operator becomes locally gauge invariant under time dependent but spatially constant flavor
gauge transformations

−1

ζ1 d + B1 v 0 ζ1 d + B1 u 0
→ (22)
d † + B2 ζ2 0 u−1 d † + B2 ζ2 0 v
if B1 and B2 are transformed as
B1 → vB1 v −1 − [∂0 v]v −1 ,
B2 → uB2 u−1 + [∂0 u]u−1 (23)
and the mass matrices are transformed as
ζ1 → vζ1 u−1 ,
ζ2 → uζ2 v −1 . (24)
In [8] we showed that the Goldstone manifold is this case is given by Gl(2)/U (2). The transfor-
mation (22) induces the following transformation on the Goldstone fields
Q → uQv −1 . (25)
Therefore,
∂0 Q → ∂0 uQv −1 + u∂0 Qv −1 + uQ∂0 v −1 ,
∂0 Q−1 → ∂0 vQ−1 u−1 + v∂0 Q−1 u−1 + vQ−1 ∂0 u−1 . (26)
266 K. Splittorff, J.J.M. Verbaarschot / Nuclear Physics B 757 [FS] (2006) 259–279

One immediately sees that the covariant combinations are

∂0 Q − QB1 − B2 Q,
∂0 Q−1 + B1 Q−1 + Q−1 B2 . (27)

Of course B1,ν = B2,ν = μδν 0 σ3 ≡ μδν 0 B. We thus obtain the covariant derivatives
 
∇ν Q = ∂ν Q − μδν 0 {Q, B}, ∇ν Q−1 = ∂ν Q−1 + μδν 0 Q−1 , B . (28)

The chiral Lagrangian is the low-energy limit of QCD and should have the same covariance
properties as were derived above. Taking into account terms to order p2 in momentum counting
we find the Lagrangian

F2 i  
L=− Tr ∇ν Q∇ν Q−1 − Σ Tr M T Q − I Q−1 I (29)
4 2
with

 z 0 1
M= and I = . (30)
z∗  −1 0
As before, the pion decay constant is denoted by Fπ and the absolute value of the chiral con-
densate is given by Σ. We emphasize that the sign before the kinetic term is opposite to the sign
of the kinetic term in the fermionic Lagrangian. This sign enters to compensate the sign due to
the noncompactness of the modes in Gl(2)/U (2) so that the kinetic terms pion fields have the
correct sign [25,26].
In the case of the fermionic partition function with one flavor and one conjugate flavor, we also
could have rearranged the fermion determinant as in (20), and we would have obtained covariant
derivatives as in (28). Indeed, if we make the transformation

U → U σ1 (31)

in the Lagrangian (8), the covariant derivatives in (9) change to


 
∂ν U + μδν 0 [U, B] → ∂ν U − μδν 0 {U, B} σ1 ,
  
∂ν U −1 + μδν 0 U −1 , B → σ1 ∂ν U −1 + μδν 0 U −1 , B . (32)

Notice that both U and U σ1 are unitary. The mass term in (8) becomes

z 0  −1
  −1
 z 0  
Tr U +U = Tr M U + σ3 U σ3 → Tr U − σ1 σ3 U −1 σ3 σ1 σ1
0 z∗ 0 z∗

0 z∗  
= Tr U + I U −1 I . (33)
z 0
The invariance properties of this term are consistent with those of the representation (20). Be-
cause both U and −U belong to U (2) this term has the discrete symmetry M → −I MI . In the
bosonic case this symmetry dictates that the relative sign of the two terms in the mass term has
to be negative.
K. Splittorff, J.J.M. Verbaarschot / Nuclear Physics B 757 [FS] (2006) 259–279 267

4.2. Microscopic limit of the partition function

In the microscopic limit, the partition function is dominated by the zero momentum modes
and in the sector of zero topological charge it is given by

 ∗  dQ V 2 2 −1 i −1
θ (Q)e− 4 Fπ μ Tr{Q,B}{Q ,B} + 2 V Σ Tr M (Q−I Q I ) ,
T
Z n=−1
z, z ; μ = lim 2
→0 det Q
(34)
where dQ θ (Q)/det2 Q is the integration measure on positive definite 2 × 2 Hermitian matrices.2
In the limit  → 0 the integrals can be performed analytically resulting in [8]
2
2

e−2μ̂ ẑ + ẑ∗ 2
2
  |ẑ|
Z n=−1 ẑ, ẑ∗ ; μ̂ = log  2
exp − 2
K0 . (35)
μ̂ 8μ̂ 4μ̂2
Using the Toda lattice equation, the product of this partition function and its fermionic coun-
terpart gives the quenched microscopic spectral density at nonzero chemical potential. This has
been confirmed beautifully by recent quenched lattice QCD simulations at nonzero chemical
potential with a staggered Dirac operator [27] and an overlap Dirac operator [28].
The volume dependence has dropped from the partition function (35) and no kink will develop
in the thermodynamic limit. The chiral condensate is given by
Σ n=−1 (m̂; μ̂) 1
= − ∂m̂ log Z n=−1 (m̂, m̂; μ̂)
Σ 2
m̂ m̂
∼ for m̂ → ∞, μ̂ → ∞ and ∼ 1. (36)
2μ̂2 2μ̂2
It follows that the low energy theory with one bosonic flavor and its conjugate does not have a
phase transition as function of μ̂. This is to be contrasted with the fermionic counterpart which,
as discussed in Section 3, has a phase transition at m̂ = 2μ̂2 . Plots of the chiral condensate for
n = −1 are shown in Fig. 2.
The reason for the absence of a phase transition at m̂ = 2μ̂2 becomes clear from the calcu-
lation of the mass spectrum of the Lagrangian (29) along the lines of [17] (see Appendix A).
For√all values of the chemical potential we find a charged massless mode (in fact with a mass
∼ ) which condenses  for any μ > 0. The masses of the remaining three modes are given by
2μ, m2π /2μ and 2μ 1 + 3(mπ /2μ)4 , exactly the same as in other QCD-like theories at nonzero
chemical potential [17,19].

5. The effective partition function and chiral random matrix theory

We now turn to the computation of the partition function with one bosonic quark, that is
Nf = −1. As follows from Eq. (7) the Goldstone bosons result from the breaking of chiral
symmetry in the theory with one fermionic and two bosonic quarks. The partition function is
therefore given by an integral over the super group Ĝl(1|2). Such integrals are technically rather

2 In [36] we incorrectly used commutators in (34) instead of anti-commutators. This resulted in an extra overall factor

of exp(2V Fπ2 μ2 ) which was corrected for by hand in the calculation of the spectral density. In [8], the correct form
of Z n=−1 (z, z∗ ; μ) was obtained because the overall μ-dependent constant was obtained by matching to the spectral
density via the Toda lattice equation.
268 K. Splittorff, J.J.M. Verbaarschot / Nuclear Physics B 757 [FS] (2006) 259–279


Fig. 2. The chiral condensate for n = −1 (one boson and one conjugate boson). Left: The mass dependence for μ̂ = 5
(full curve). Right: The chemical potential dependence for m̂ = 10 (full curve). In this case there is no phase transition at
μ = mπ /2 as can be seen from the result in the thermodynamic limit (dashed curves, obtained for μ̂ → ∞ or m̂ → ∞,
respectively).

complicated, and at this point we have not succeeded to evaluate the Nf = −1 partition function
along these lines.
Fortunately, the effective low energy partition function has an alternative representation as the
large N limit of a random matrix theory with the same chiral symmetries. This has been shown
explicitly in the fermionic case at zero [6] and nonzero [8,29,30] chemical potential as well as
for bosonic [31] and supersymmetric [32–35] partition functions at zero chemical potential. For
bosonic quarks at nonzero chemical potential the equivalence between the effective theories and
random matrix theory in the microscopic limit has been established for n = −1 in [36].
As before, we only consider the theory in the sector of zero topological charge.

6. The random matrix model

The random matrix partition function with Nf quark flavors of mass m and n pairs of regular
and conjugate quarks with masses y and z∗ , respectively, is defined by [30]

Nf ,n  ∗
  
ZN {mf }, y, z ; μ ≡ dΦ dΨ wG (Φ)wG (Ψ )detNf D(μ) + mf
   
× detn D(μ) + y detn D† (μ) + z∗ , (37)
where the non-Hermitian Dirac operator is given by

0 iΦ + μΨ
D(μ) = . (38)
iΦ † + μΨ † 0
Here, Φ and Ψ are complex N × N matrices with the same Gaussian weight function
 
wG (X) = exp −N Tr X † X . (39)
Bosonic quarks appear as inverse determinants and notationally we simply allow Nf and n to
take negative values.
K. Splittorff, J.J.M. Verbaarschot / Nuclear Physics B 757 [FS] (2006) 259–279 269

Of course, in general the QCD partition function and the random matrix partition function are
different. However, when we consider the microscopic limit where the variables

m̂ = 2mN and μ̂2 = 2μ2 N (40)


are fixed as N → ∞ the random matrix partition function and the QCD partition function coin-
cide provided that we identify (see the discussion in [36])

m̂ = 2mN → mV Σ,
μ̂ = 2μ2 N → μ2 Fπ2 V . (41)
In this section we will work within the random matrix framework and use the identifications (41)
in the final results.
In [30] it was shown that the random matrix partition function (37) can be rewritten in the
eigenvalue representation,
 
N
N ,n      
ZN f m, y, z∗ ; μ ∼ d 2 zk P Nf ,n {zi }, zi∗ ; μ , (42)
C k=1

where the integration extends over the full complex plane and the joint probability distribution
of the eigenvalues is given by
    1   2
P Nf ,n {zi }, zi∗ ; μ = 2N ΔN zl2 
μ

N
  N  n  n
× w zk , zk∗ ; μ m2 − zk2 f y 2 − zk2 z∗ 2 − zk∗ 2 . (43)
k=1

The Vandermonde determinant is defined as

 2  
N
 2 
ΔN zl ≡ zi − zj2 , (44)
i>j =1

and the weight function reads



 ∗
 N (1 + μ2 ) N (1 − μ2 )  2 ∗2

w zk , zk ; μ = |zk | K0
2
|zk | exp −
2
zk + zk . (45)
2μ2 4μ2
The eigenvalue representation makes it possible to employ the method of orthogonal polyno-
mials in the complex plane [37–41] to compute the spectral density and eigenvalue correlation
functions [30].

6.1. Orthogonal polynomials and their Cauchy transform

In order to evaluate the partition function with Nf = −1 we will make use of orthogonal
polynomials and their Cauchy transform. The complex Laguerre polynomials given by [30]

1 − μ2 k N z2
pk (z; μ) = k!Lk − (46)
N 1 − μ2
270 K. Splittorff, J.J.M. Verbaarschot / Nuclear Physics B 757 [FS] (2006) 259–279

are the orthogonal polynomials corresponding to the weight w(z, z∗ ; μ) given in (45). To be
specific, the polynomials satisfy the orthogonality condition [42]

 
d 2 z w z, z∗ ; μ pk (z; μ)pl (z; μ)∗ = rk δkl , (47)
C
with the norm
πμ2 (1 + μ2 )2k k!k!
rk = . (48)
N 2k+2
The Cauchy transform of the orthogonal polynomials is defined as

1  
hk (m; μ) = d 2 z 2 w z, z∗ ; μ pk∗ (z; μ), (49)
z −m 2
C
where C indicates that the integration extends over the complex plane. Using that the weight
function and polynomials are even functions of z the Cauchy transform can be written as

1  
hk (m; μ) = d 2 z w z, z∗ ; μ pk∗ (z; μ). (50)
z(z − m)
C

7. QCD with one bosonic flavor

It was shown in [36] that


N =−1 1
ZN f (m; μ) = − hN−1 (m; μ). (51)
rN−1
Therefore, studying the properties of bosonic partition functions is equivalent to analyzing the
properties of the Cauchy transform. While the relation between the partition function and the
Cauchy transform was established in [36], no explicit evaluation of the Cauchy transform was
given; this evaluation follows below. We will find that the partition function with one bosonic
quark depends on the chemical potential for μ > mπ /2.
We are interested in the microscopic limit where N → ∞ for fixed m̂ = 2N m and fixed
μ̂2 = 2μ2 N . (This scaling of μ is the analogue of the weak nonhermiticity limit introduced in
[43].) In this limit the polynomial pN−1 /rN−1 is given by

pN−1 (ẑ; μ̂) 1 (N − 1)! μ̂2 N−1 ẑ2


= 1 − L N−1 −
rN−1 rN−1 N N−1 2N 4N
2N 5/2 eN − 3 μ̂2
∼√ e 2 I0 (ẑ), (52)
2ππ μ̂2
where we have used that
π μ̂2 (1 + μ̂2 /2N )2(N−1) (N − 1)!(N − 1)!
rN−1 = . (53)
2N 2N +1
The microscopic limit of the partition function is thus given by
 2
z2 +z∗ 2
2N 1/2 eN − 3 μ̂2 z∗ |z| −  
Z Nf =−1 (m̂; μ̂) = − e 2 d 2
z K 0 e 8μ̂2 I0 z∗ . (54)
π μ̂ 2 z − m̂ 4μ̂ 2
C
K. Splittorff, J.J.M. Verbaarschot / Nuclear Physics B 757 [FS] (2006) 259–279 271

To calculate this integral we write (with z = x + iy)


  1   ∗  
I0 z∗ = Sgn(y) K0 z − K0 −z∗ (55)
πi
and first calculate the integral over y by a complex contour integration. The first term in (55) is
exponentially damped in the upper half of the complex y-plane and the second term in the lower
half-plane. This allows us to close the integration contour of the y variable, and by Cauchy’s
theorem we obtain for the real part of the partition function
∞  
 N =−1  2x 2 − 2x m̂ + m̂2
Re Z f (m̂; μ̂) = 2πcN (μ̂) dx (2x − m̂) exp −
4μ̂2
−∞


(2x − m̂)m̂
× θ (m̂ − x)θ (2x − m̂)K0 I0 (2x − m̂)
4μ̂2


(2x − m̂)m̂
− θ (m̂ − x)θ (m̂ − 2x)I0 K0 (m̂ − 2x) . (56)
4μ̂2
The imaginary part of the partition function is given by
  
 N =−1  2x 2 − 2x m̂ + m̂2
Im Z f (m̂; μ̂) = −2cN (μ̂) dx (2x − m̂) exp −
4μ̂2


(2x − m̂)m̂
× θ (x − m̂)K0 K0 (2x − m̂)
4μ̂2

(2x − m̂)m̂
+ θ (m̂ − x)θ (2x − m̂)K0 K0 (2x − m̂)
4μ̂2


(m̂ − 2x)m̂
+ θ (m̂ − x)θ (m̂ − 2x)K0 K0 (m̂ − 2x) . (57)
4μ̂2
The overall constant cN (μ̂) is defined by
2N 1/2 eN − 3 μ̂2
cN (μ̂) = e 2 . (58)
π μ̂2
The integrand of the imaginary part is odd about x = m̂/2 so that the integral vanishes. Using
s = x − m̂/2 as new integration variable we obtain the final expression for the partition function
with Nf = −1 at nonzero chemical potential
m̂/2
m̂2
 
Nf =−1 − s2
Z (m̂; μ̂) = 4πcN (μ̂)e 8μ̂2 ds s exp − 2
2μ̂
−∞



s m̂ s m̂
× θ (s)K0 I 0 (2s) − θ (−s)I 0 K 0 (−2s)
2μ̂2 2μ̂2
 m̂/2  

2
− m̂ 2 s2 s m̂
= 4πcN (μ̂)e 8μ̂ ds s exp − 2 K0 I0 (2s)
2μ̂ 2μ̂2
0
∞  

s2 s m̂
+ ds s exp − 2 I0 K0 (2s) . (59)
2μ̂ 2μ̂2
0
272 K. Splittorff, J.J.M. Verbaarschot / Nuclear Physics B 757 [FS] (2006) 259–279


Fig. 3. The chiral condensate in the theory with one bosonic flavor. Left: the mass dependence for μ̂ = 5 (full curve).
In the thermodynamic limit (μ̂ → ∞, dashed curve) Σ Nf =−1 changes from a linear dependence on m to become m
independent for m > mc = 2μ2 Fπ2 /Σ . Right: The dependence on the chemical potential for m̂ = 10 (full curve) and in
the thermodynamic limit (m̂ → ∞, dashed curve).

7.1. The chiral condensate for one bosonic flavor

The chiral condensate given by


Σ Nf =−1 (m̂; μ̂)
= −∂m̂ log Z Nf =−1 (m̂; μ̂) (60)
Σ
is plotted in Fig. 3 (full curve). Notice that the chiral condensate depends on the chemical
potential, and that for m̂ 2μ̂2 , it approaches the value Σ . Below we will show that in the
thermodynamic limit Σ Nf =−1 develops a kink at m̂ = 2μ̂2 .

7.2. Phase transition with one bosonic flavor

Above we have determined the partition function in the microscopic limit, where the source
m times the volume N , is kept fixed, and as a consequence, the phase transition at m̂ = 2μ̂2 is
smeared. A cusp in the derivative of the free energy only appears in the thermodynamic limit
where N → ∞ for fixed m. This limit can also be approached by taking m̂ = 2mN → ∞ and
μ̂2 = 2μ2 N → ∞ in the microscopic results as will be done below.
In the thermodynamic limit the integral (59) can be calculated by a saddle point approxima-
tion. The saddle point of the first integral is at

s=− + 2μ̂2 , (61)
2
and the saddle point of the second integral is at

s = −2μ̂2 + . (62)
2
In the thermodynamic limit the partition function for m̂ > 2μ̂2 is therefore given by

Z Nf =−1 (m̂; μ̂) = √ cN (μ̂)μ2 e2μ̂ −m̂ .
2
(63)
2m̂
K. Splittorff, J.J.M. Verbaarschot / Nuclear Physics B 757 [FS] (2006) 259–279 273

The exponent depends linearly on m̂ resulting in a chiral condensate that is equal to 1. In the nor-
malization for which the partition function is μ-independent for small μ (see [30]), the exponent
in cN (μ̂) (see Eq. (58)) becomes exp(−2μ̂2 ) instead of exp(− 32 μ̂2 ). A phase transition occurs
when the dominant saddle point hits the boundary of the integration domain, i.e. at m̂ = 2μ̂2 , cf.
(61). For m̂ < 2μ̂2 the saddle point is outside of the domain of the first integral and the main
contribution to the first integral comes from the region close to m̂/2 so that

Z Nf =−1 (m̂; μ̂) ∼ em̂−m̂
2 /2μ̂2 −2μ̂2
for .μ̂2 > (64)
2
The condensate in the thermodynamic limit is therefore given by

Σ Nf =−1 (m̂; μ̂) −1 + μ̂m̂2 for m̂ < 2μ̂2 ,
= (65)
Σ 1 for m̂ > 2μ̂2 .
In Fig. 3, this result is shown by the dashed curves.
As promised, we have shown that in the thermodynamic limit the chiral condensate has a kink
at m̂ = 2μ̂2 . Using (41) and the GOR-relation (2mΣ = Fπ2 m2π ) we see that the phase transition
takes place at μ = mπ /2 with

Σ Nf =−1 (mπ ; μ) m2
−1 + 2μπ2 for mπ < 2μ,
= (66)
Σ 1 for mπ > 2μ.
This behavior of the chiral condensate can be explained from the structure in (7). We first note
that the average mass derivatives of fermionic and bosonic determinants give factors that are
equal in magnitude but have opposite sign. At μ = 0 the contribution from the fermionic quark
thus cancels against the contribution from one of the bosonic quarks resulting in a total chiral
condensate of magnitude Σ. For μ < mπ /2 the partition function is μ independent and the
chiral condensate therefore remains at the value Σ . As μ exceeds mπ /2 it becomes energetically
favorable to create pions made out a bosonic anti-quark and a conjugate bosonic quark. As a
consequence the contribution to the chiral condensate from the bosonic quarks starts rotating to
zero for μ > mπ /2 so that the total chiral condensate for large values of μ/mπ is given by the
contribution of the fermionic flavor alone. For μ mπ /2 the chiral condensate therefore is equal
to the chiral condensate at μ = 0 but with the opposite sign.
The second order phase transition can also be seen in the baryon density given by

nNf =−1 (m̂; μ̂) Nf =−1 0 for μ̂2 < m̂2 ,
= −∂μ̂ log Z (m̂; μ̂) = 2 (67)
Fπ 4μ̂(1 − 4m̂μ̂4 ) for μ̂2 > m̂2 .

In contrast, the free energy for Nf = 1 does not depend on μ̂.


When m̂ ∼ 1 the finite volume effects for the chiral condensate are strong and the micro-
scopic prediction differs significantly from the result in the thermodynamic limit. For m̂ 1 the
partition function is given by
m̂2  
Z Nf =−1 (m̂; μ̂) = A + log m̂ + O m̂2 (68)
8
resulting in a chiral condensate
Σ Nf =−1 (m̂; μ̂) 1
= m̂ log m̂ for m̂ 1, (69)
Σ 4A
274 K. Splittorff, J.J.M. Verbaarschot / Nuclear Physics B 757 [FS] (2006) 259–279

where A is a constant. This explains the approach to zero for m̂ → 0 of the solid curve in the left
panel of Fig. 3.

8. The phase structure and the eigenvalue spectrum

The chiral phases of QCD reflect themselves in the eigenvalue spectrum of the Dirac operator.
For zero chemical potential the eigenvalue density near the origin of the imaginary axis is an
order parameter for spontaneous breaking of chiral symmetry [44]. When the chemical potential
is nonzero the eigenvalues move into the complex plane and the chiral condensate in full QCD
is then linked to the oscillations of the eigenvalue density [45]. For nonzero isospin chemical
potential (n = 1) the transition into the Bose–Einstein condensed phase takes place when the
quark mass is at the boundary of the support of the eigenvalue density (see for example [18]).
Also in the case with one pair of conjugate bosons, n = −1, the phase structure is closely
related to the Dirac spectrum. Setting n = −1 in (42) we obtain the following eigenvalue repre-
sentation of the partition function,
N =0,n=−1  
ZN f z, z∗ ; μ
 
1
N
  2 N
  1
∼ 2N d 2 zk ΔN zl2  w zk , zk∗ ; μ . (70)
μ
k=1 k=1
(z − zk )(z∗ 2 − zk∗ 2 )
2 2
C
We observe that the probability to find an eigenvalue, zj , has a modulus squared pole at the mass
z. This gives rise to the logarithmic divergence discussed in Section 2,

1 1
d 2u 2 ∼ 2 log . (71)
|z − u |2 2 |z|
C (u)

Here, C (u) is an annulus centered at u with outer radius 1 and inner radius  1. Because of
the Vandermonde determinant, the probability that two or more eigenvalues are close to the mass
z does not diverge. Therefore the logarithmically divergent contribution to the partition function
is from eigenvalue configurations with exactly one eigenvalue close to the mass. Setting zN = z
in (70) and using that

  2 N 2  
  2 N−1 2 N−1 
Δ z   = Δ z  z2 − z2 2 , (72)
l l=1 zN =z l l=1 l
l=1
we find that the product on the right-hand side exactly cancels the two bosonic determinants. The
remaining integral over N − 1 eigenvalues is equal to the quenched partition function which does
not depend on the mass. The only mass dependence is from the weight function, w, evaluated at
the mass z and from the divergence (71). We thus find
N =0,n=−1   w(z, z∗ ; μ)
ZN f z, z∗ ; μ ∼ , (73)
|z|2
which is exactly the result (35) given in Section 4.
The absolute value squared pole responsible for attracting the eigenvalue zN to the mass z
appears for all nonzero values of μ. Hence the n = −1 theory must be regulated for all nonzero
values of μ. The transition to μ → 0 is discontinuous because if μ = 0 the probability to find an
eigenvalue near z is zero unless z is on the imaginary axis.
K. Splittorff, J.J.M. Verbaarschot / Nuclear Physics B 757 [FS] (2006) 259–279 275

9. Summary and discussion

We have computed the microscopic limit (also known as the -limit) of the QCD partition
function at nonzero chemical potential with one bosonic quark. For the computation we made use
of the random matrix representation of the partition function. Contrary to the partition function
with one fermionic flavor this partition function depends on the chemical potential. In particular
it has been shown that a phase transition takes place at μ = mπ /2. The theory with a pair of
conjugate bosonic flavors also behaves differently from its fermionic counterpart. While a phase
transition signaled by Bose–Einstein condensation of pions takes place in the fermionic theory, its
bosonic variant always remains in a μ dependent phase. The main differences between fermionic
and bosonic partition functions have been summarized in Table 1.
Bosonic partition functions as studied in this paper also appear in the closely related Hatano–
Nelson model [46]. This is a model for a disordered system in an imaginary vector potential.
The random matrix limit of this model is a non-Hermitian random matrix model just like that in
(37) except for the chiral block structure in (38). The partition function with Nf = −1 can again
be expressed as the Cauchy transform of the orthogonal polynomials. However, because this
model is technically simpler than the random matrix model for the QCD partition function, this
partition function can also be evaluated directly by means of standard random matrix techniques.
The agreement of both approaches [47] confirms our understanding of the Cauchy transform
representation of the partition function. Moreover, as in the chiral case, we find that the partition
function of the Hatano–Nelson model with Nf = −1 depends on the chemical potential and a
phase transition appears in the thermodynamic limit.
These results generalize to theories with more fermionic and bosonic flavors, and we expect
similar surprises for 2 color QCD and for QCD with quarks in the adjoint representation.
Our results might be relevant for a better understanding of the sign problem in QCD at nonzero
baryon chemical potential. The expectation value of the phase of the fermion determinant is given
by
 
  det(D + μγ0 + m)
exp(2iθ) =
det(D + μγ0 + m)∗
 
det2 (D + μγ0 + m)
= . (74)
det(D + μγ0 + m)∗ det(D + μγ0 + m)

It thus corresponds to a partition function with a bosonic quark and its conjugate and two fermi-
onic quarks. If the same arguments apply as in the case of the partition function with one
bosonic flavor the vacuum expectation value of exp(2iθ) will be nonanalytic at μ = mπ /2, in
the quenched as well as in the unquenched theory. We speculate that this nonanalyticity persist
at nonzero temperatures and that beyond this point the severity of the sign problem increases
drastically.

Note added in proof

The expectation that exp(2iθ) is nonanalytic at μ = mπ /2, in the quenched as well as in the
unquenched theory has now been verified [48].
276 K. Splittorff, J.J.M. Verbaarschot / Nuclear Physics B 757 [FS] (2006) 259–279

Acknowledgements

We wish to thank Gernot Akemann, James Osborn, Dennis Dietrich, Dominique Toublan,
Poul Henrik Damgaard and Frank Wilczek for valuable discussions. Gernot Akemann is thanked
for pointing out a mistake in an earlier version of this paper. This work was supported in part by
US DOE Grant No. DE-FG-88ER40388. K.S. was supported by the Carslberg Foundation.

Appendix A. Calculation of the particle spectrum of the chiral Lagrangian for a pair of
bosonic quarks

In this appendix we calculate the particle spectrum of the chiral Lagrangian (29) for QCD
with a pair of conjugate bosonic quarks. We only work out in detail the case that z = x is real
and  = 0.
The quark fields in the Goldstone manifold Gl(2)/U (2) can be parameterized as

cosh ser+t sinh seiθ+t −1 cosh se−r−t − sinh seiθ−t


Q= , Q = .
sinh se−iθ+t cosh se−r+t − sinh se−iθ−t cosh ser−t
(A.1)
For the covariant derivatives we find

∇0 Q ≡ ∂0 Q − μ{Q, B}
 r+t 
e [∂0 cosh s + cosh s∂0 (r + t) − 2μ cosh s] eiθ +t [∂0 sinh s + sinh s∂0 (iθ + t)]
= e −iθ +t [∂0 sinh s + sinh s∂0 (−iθ + t)] e−r+t [∂0 cosh s + cosh s∂0 (−r + t) + 2μ cosh s]
(A.2)
and
 
∇0 Q−1 ≡ ∂0 Q−1 + μ Q−1 , B
 −r−t 
e [∂0 cosh s − cosh s∂0 (r + t) + 2μ cosh s] −eiθ−t [∂0 sinh s + sinh s∂0 (iθ − t)]
= −e−iθ−t [∂ sinh s − sinh s∂ (iθ + t)] er−t [∂0 cosh s+ cosh s∂0 (r − t) − 2μ cosh s]
.
0 0
(A.3)
For the time derivatives in the kinetic term we obtain

Tr ∇0 Q∇0 Q−1 = −2(∂0 s)2 − 2(∂0 t)2 − 2 cosh2 s(∂0 r)2 − 2 sinh2 s(∂0 θ )2
+ 8μ cosh2 s∂0 r − 8μ2 cosh2 s. (A.4)
The mass term is given by
1  
−i Σ Tr M T Q − I Q−1 I
2
1
= −iΣ (4x sinh s cosh t cos θ − 4y sinh s sinh t sin θ + 4 cosh r cosh s cosh t). (A.5)
2
The total Lagrangian is

F2 1  
− Tr ∇μ Q∇Q†μ − i Σ Tr M T Q − I Q−1 I (A.6)
4 2
with M and I defined in (30). Notice that because of the noncompact degrees of freedom, the
sign of the kinetic term is opposite to the usual sign.
K. Splittorff, J.J.M. Verbaarschot / Nuclear Physics B 757 [FS] (2006) 259–279 277

The static part of the Lagrangian for y = 0 and  = 0 is given by

Lstatic = 2F 2 μ2 cosh2 s − i2Σx sinh s cosh t cos θ − 4 cosh r cosh s cosh t. (A.7)
The minimum is at (note that we can always divide out cosh s, so there is only one minimum)

xΣ m2π
sinh s = i = i (A.8)
2F 2 μ2 4μ2
and

t = 0, r = 0, θ = 0. (A.9)
We expand the chiral Lagrangian (A.6) to second order in the small fluctuations about the saddle
point. The second order terms are given by

1 1 1 1
F 2 − (∂0 δs)2 − (∂0 δt)2 − cosh2 s(∂0 δr)2 − sinh2 s(∂0 δθ)2
2 2 2 2

 
+ 4μ sinh s cosh sδs∂0 δr − 2μ cosh s + sinh s (δs)
2 2 2 2


+ iΣx sinh s (δs)2 + (δt)2 − (δθ )2 . (A.10)
We obtain the uncoupled masses

2xΣ m4π
m2t = −i sinh s = , (A.11)
F2 4μ2
m2θ = 4μ2 . (A.12)
The remaining two masses of follow from solving
2

E − 4μ2 cosh2 s +4μE sinh s cosh s


det =0
−4μE sinh s cosh s E 2 cosh2 s

since cosh s = 0 we can divide by cosh2 s and obtain


 
E 2 E 2 − 4μ2 cosh2 s + 16μ2 E 2 sinh2 s = 0 (A.13)
with solutions

3m2π
E 2 = 0 or E 2 = 4μ2 cosh2 s − 16μ2 sinh2 s = 4μ2 1 + . (A.14)
16μ4

√ μ > 0. For  = 0 one can show along the same lines that
Note that one mode is massless for all
the massless mode obtains a mass ∼ .

References

[1] K.B. Efetov, Phys. Rev. Lett. 79 (1997) 491;


K.B. Efetov, Adv. Phys. 32 (1983) 53;
K.B. Efetov, Supersymmetry in Disorder and Chaos, Cambridge Univ. Press, Cambridge, 1997.
278 K. Splittorff, J.J.M. Verbaarschot / Nuclear Physics B 757 [FS] (2006) 259–279

[2] S.F. Edwards, P.W. Anderson, J. Phys. F 5 (1975) 965;


S.F. Edwards, R.C. Jones, J. Phys. A 9 (1976) 1595;
J.J.M. Verbaarschot, M.R. Zirnbauer, Ann. Phys. 158 (1984) 78;
J.J.M. Verbaarschot, M.R. Zirnbauer, J. Phys. A 18 (1985) 1093;
A. Kamenev, M. Mézard, J. Phys. A 32 (1999) 4373;
A. Kamenev, M. Mézard, Phys. Rev. B 60 (1999) 3944;
I.V. Yurkevich, I.V. Lerner, Phys. Rev. B 60 (1999) 3955;
M.R. Zirnbauer, cond-mat/9903338.
[3] C.W. Bernard, M.F.L. Golterman, Phys. Rev. D 46 (1992) 853;
C.W. Bernard, M.F.L. Golterman, Phys. Rev. D 49 (1994) 486.
[4] P.H. Damgaard, K. Splittorff, Nucl. Phys. B 572 (2000) 478;
P.H. Damgaard, K. Splittorff, Phys. Rev. D 62 (2000) 054509;
P.H. Damgaard, Phys. Lett. B 476 (2000) 465.
[5] H. Leutwyler, A. Smilga, Phys. Rev. D 46 (1992) 5607.
[6] E.V. Shuryak, J.J.M. Verbaarschot, Nucl. Phys. A 560 (1993) 306;
J.J.M. Verbaarschot, Phys. Rev. Lett. 72 (1994) 2531.
[7] K. Splittorff, J.J.M. Verbaarschot, Phys. Rev. Lett. 90 (2003) 041601.
[8] K. Splittorff, J.J.M. Verbaarschot, Nucl. Phys. B 683 (2004) 467.
[9] K. Splittorff, J.J.M. Verbaarschot, Nucl. Phys. B 695 (2004) 84.
[10] J. Gasser, H. Leutwyler, Phys. Lett. B 188 (1987) 477.
[11] K.B. Efetov, Adv. Phys. 32 (1983) 53.
[12] J.J.M. Verbaarschot, M.R. Zirnbauer, Ann. Phys. 158 (1984) 78.
[13] J.J.M. Verbaarschot, T. Wettig, Annu. Rev. Nucl. Part. Sci. 50 (2000) 343.
[14] M. Stephanov, Phys. Rev. Lett. 76 (1996) 4472.
[15] M. Alford, A. Kapustin, F. Wilczek, Phys. Rev. D 59 (1999) 054502.
[16] J.B. Kogut, M.A. Stephanov, D. Toublan, Phys. Lett. B 464 (1999) 183.
[17] J.B. Kogut, M.A. Stephanov, D. Toublan, J.J.M. Verbaarschot, A. Zhitnitsky, Nucl. Phys. B 582 (2000) 477.
[18] D. Toublan, J.J.M. Verbaarschot, Int. J. Mod. Phys. B 15 (2001) 1404.
[19] D.T. Son, M.A. Stephanov, Phys. Rev. Lett. 86 (2001) 592;
K. Splittorff, D.T. Son, M.A. Stephanov, Phys. Rev. D 64 (2001) 016003;
J.B. Kogut, D. Toublan, Phys. Rev. D 64 (2001) 034007;
K. Splittorff, D. Toublan, J.J.M. Verbaarschot, Nucl. Phys. B 620 (2002) 290;
K. Splittorff, D. Toublan, J.J.M. Verbaarschot, Nucl. Phys. B 639 (2002) 524;
J. Wirstam, J.T. Lenaghan, K. Splittorff, Phys. Rev. D 67 (2003) 034021;
J.T. Lenaghan, F. Sannino, K. Splittorff, Phys. Rev. D 65 (2002) 054002.
[20] J.B. Kogut, D.K. Sinclair, Phys. Rev. D 66 (2002) 034505.
[21] J. Feinberg, A. Zee, Nucl. Phys. B 504 (1997) 579.
[22] R.A. Janik, M.A. Nowak, G. Papp, J. Wambach, I. Zahed, Phys. Rev. E 55 (1997) 4100;
R.A. Janik, M.A. Nowak, G. Papp, I. Zahed, Nucl. Phys. B 501 (1997) 603.
[23] K.B. Efetov, Phys. Rev. Lett. 79 (1997) 491;
K.B. Efetov, Phys. Rev. B 56 (1997) 9630.
[24] P.H. Damgaard, U.M. Heller, K. Splittorff, B. Svetitsky, Phys. Rev. D 72 (2005) 091501;
P.H. Damgaard, U.M. Heller, K. Splittorff, B. Svetitsky, D. Toublan, Phys. Rev. D 73 (2006) 074023;
P.H. Damgaard, U.M. Heller, K. Splittorff, B. Svetitsky, D. Toublan, Phys. Rev. D 73 (2006) 105016.
[25] H. Leutwyler, Ann. Phys. 235 (1994) 165.
[26] M.R. Zirnbauer, J. Math. Phys. 37 (1996) 4986.
[27] G. Akemann, T. Wettig, Phys. Rev. Lett. 92 (2004) 102002;
G. Akemann, T. Wettig, Phys. Rev. Lett. 96 (2006) 029902, Erratum;
J.C. Osborn, T. Wettig, PoS LAT2005 (2005) 200, hep-lat/0510115.
[28] J. Bloch, T. Wettig, hep-lat/0604020.
[29] G. Akemann, Y.V. Fyodorov, G. Vernizzi, Nucl. Phys. B 694 (2004) 59.
[30] J.C. Osborn, Phys. Rev. Lett. 93 (2004) 222001.
[31] D. Dalmazi, J.J.M. Verbaarschot, Nucl. Phys. B 592 (2001) 419.
[32] J.C. Osborn, D. Toublan, J.J.M. Verbaarschot, Nucl. Phys. B 540 (1999) 317.
[33] P.H. Damgaard, J.C. Osborn, D. Toublan, J.J.M. Verbaarschot, Nucl. Phys. B 547 (1999) 305.
[34] G. Akemann, Y.V. Fyodorov, Nucl. Phys. B 664 (2003) 457.
K. Splittorff, J.J.M. Verbaarschot / Nuclear Physics B 757 [FS] (2006) 259–279 279

[35] Y.V. Fyodorov, G. Akemann, JETP Lett. 77 (2003) 438, Pis’ma Zh. Eksp. Teor. Fiz. 77 (2003) 513.
[36] G. Akemann, J.C. Osborn, K. Splittorff, J.J.M. Verbaarschot, Nucl. Phys. B 712 (2005) 287.
[37] G. Akemann, G. Vernizzi, Nucl. Phys. B 660 (2003) 532.
[38] G. Akemann, Acta Phys. Pol. B 34 (2003) 4653.
[39] M.C. Bergère, hep-th/0311227.
[40] M.C. Bergère, hep-th/0404126.
[41] G. Akemann, A. Pottier, J. Phys. A 37 (2004) L453.
[42] G. Akemann, Nucl. Phys. B 730 (2005) 253.
[43] Y.V. Fyodorov, B.A. Khoruzhenko, H.-J. Sommers, Phys. Lett. A 226 (1997) 46;
Y.V. Fyodorov, B.A. Khoruzhenko, H.-J. Sommers, Phys. Rev. Lett. 79 (1997) 557.
[44] T. Banks, A. Casher, Nucl. Phys. B 169 (1980) 103.
[45] J.C. Osborn, K. Splittorff, J.J.M. Verbaarschot, Phys. Rev. Lett. 94 (2005) 202001.
[46] J. Miller, J. Wang, Phys. Rev. Lett. 76 (1996) 1461;
N. Hatano, D.R. Nelson, Phys. Rev. Lett. 77 (1996) 570.
[47] K. Splittorff, J.J.M. Verbaarschot, in preparation.
[48] K. Splittorff, J.J.M. Verbaarschot, hep-lat/0609076.
Nuclear Physics B 757 [FS] (2006) 280–302

Exact partition function of SU(m|n) supersymmetric


Haldane–Shastry spin chain
B. Basu-Mallick ∗ , Nilanjan Bondyopadhaya
Theory Group, Saha Institute of Nuclear Physics, 1/AF Bidhan Nagar, Kolkata 700 064, India
Received 1 August 2006; accepted 13 September 2006
Available online 25 September 2006

Abstract
By taking the freezing limit of a spin Calogero–Sutherland model containing ‘anyon like’ representation
of the permutation algebra, we derive the exact partition function of SU(m|n) supersymmetric Haldane–
Shastry (HS) spin chain. This partition function allows us to study global properties of the spectrum like
level density distribution and nearest neighbour spacing distribution. It is found that, for supersymmetric
HS spin chains with large number of lattice sites, continuous part of the energy level density obeys Gaussian
distribution with a high degree of accuracy. The mean value and standard deviation of such Gaussian distri-
bution can be calculated exactly. We also conjecture that the partition function of supersymmetric HS spin
chain satisfies a duality relation under the exchange of bosonic and fermionic spin degrees of freedom.
© 2006 Elsevier B.V. All rights reserved.

1. Introduction

Haldane–Shastry (HS) spin chain is a well-known quantum integrable model, where equally
spaced spins on a circle interact with each other through pairwise exchange interactions inversely
proportional to the square of their chord distances. Study of such HS spin- 12 chain with long-
range interaction was originally motivated from the fact that the exact ground state wavefunction
of this model coincides with the U → ∞ limit of Gutzwiller’s variational wave function for
the Hubbard model, and also with the one-dimensional version of the resonating valence bond
state proposed by Anderson [1,2]. Remarkably, HS spin chain can be explicitly solved in much

* Corresponding author.
E-mail addresses: bireswar.basumallick@saha.ac.in, biru@theory.saha.ernet.in (B. Basu-Mallick),
nilanjan.bondyopadhaya@saha.ac.in (N. Bondyopadhaya).

0550-3213/$ – see front matter © 2006 Elsevier B.V. All rights reserved.
doi:10.1016/j.nuclphysb.2006.09.009
B. Basu-Mallick, N. Bondyopadhaya / Nuclear Physics B 757 [FS] (2006) 280–302 281

greater detail than integrable spin chains with short-range interactions, has a Yangian quantum
group symmetry and interestingly shares many of the characteristics of an ideal gas, but with
fractional statistics [3–5]. The Hamiltonian of SU(m) HS model with N number of lattice sites
is given by
1  (1 + Pj k )
H= , (1.1)
2
1j <kN
sin2 (ξj − ξk )
where ξj = j π/N and Pj k is the exchange operator interchanging the ‘spins’ (taking m possible
values) on j th and kth lattice sites.
By using the motif representations associated with Y (glm ) Yangian symmetry of HS spin
chain (1.1), one can find out its complete spectrum including the degeneracy factor for each
energy level [6–8]. However, in practice, the computation of such degeneracy factors becomes
very cumbersome for m > 2 and large values of N . Therefore, it is difficult to express the par-
tition function of HS spin chain in a simple form (for arbitrary values of N and m) with the
help of motif representations. Due to this reason, it is worthwhile to explore other approaches
for calculating the partition function of HS spin chain. In fact, a rather simple expression for
the exact partition function of SU(m) HS spin chain has been obtained recently [9] by applying
the so-called freezing trick [10–12]. This freezing trick utilizes the connection between SU(m)
HS spin Hamiltonian and SU(m) spin Calogero–Sutherland (CS) model which has dynamical
as well as spin degrees of freedom. More precisely, one takes the strong coupling limit of spin
CS Hamiltonian, so that the particles freeze at their classical equilibrium positions of the scalar
part of the potential and spins get decoupled from the dynamical degree of freedom. As a result,
one can derive the partition function of HS spin chain by ‘modding out’ the partition function
of spinless CS model from that of the spin CS model. By using this partition function of SU(m)
HS spin chain, it is possible to study the energy level density distribution and the nearest neigh-
bour spacing distribution for fairly large values of N [9]. Interestingly, it has been found that,
the continuous part of such energy level density follows Gaussian distribution to a high degree
of approximation.
In this context it may be noted that, there exists a SU(m|n) supersymmetric extension of HS
spin chain [3], where each site is occupied by either one of the m type of bosonic states or one
of the n type of fermionic states. Such supersymmetric spin chains play an important role in
describing some correlated systems of condensed matter physics, where holes moving in the dy-
namical background of spin behave as bosons and spin- 12 electrons behave as fermions [13]. It is
worth noting that the supersymmetric SU(m|n) HS spin chain exhibits Y (gl(m|n) ) super-Yangian
symmetry [3], which is also the quantum group symmetry of supersymmetric SU(m|n) Poly-
chronakos spin chain [14,15]. Consequently, by using the motif representations and skew-Young
diagrams associated with supersymmetric Polychronakos spin chain [15], one can in principle
calculate the degeneracy factors for all energy eigenvalues of supersymmetric HS spin chain.
However, similar to the nonsupersymmetric case, this method for finding the full spectrum and
related partition function becomes very complicated for large values of N .
The aim of the present article is to find out the exact partition function for supersymmetric
SU(m|n) HS model by applying the freezing trick and also to study global properties like level
density distribution of the corresponding spectrum. For this purpose, it is convenient to map the
supersymmetric HS model to a usual spin chain containing an ‘anyon like’ representation of
the permutation algebra as spin dependent interactions [16,17]. In Section 2 we describe this
mapping and also show how the freezing trick can be applied for the case of SU(m|n) HS spin
chain by embedding it in a spin CS model containing the same anyon like representation of the
282 B. Basu-Mallick, N. Bondyopadhaya / Nuclear Physics B 757 [FS] (2006) 280–302

permutation algebra. In Section 3, we find out the complete spectrum of such spin CS model
including the degeneracy factors for all energy levels. In Section 4, we calculate the partition
function of this spin CS model at the strong coupling limit and divide it by that of the spinless
CS model to finally obtain the partition function of SU(m|n) HS spin chain. In this section, we
also discuss about the motif representation for SU(m|n) HS spin chain and find that, due to the
lifting of a selection rule, some extra energy levels appear in the spectrum in comparison with the
case of SU(m) spin chain. Subsequently, we conjecture that the partition function of SU(m|n)
HS model satisfies a duality relation under the exchange of bosonic and fermionic spin degrees of
freedom. In Section 5, we study the level density distribution and the nearest neighbour spacing
distribution for the spectrum of SU(m|n) HS spin chain by using its exact partition function. It is
found that, for sufficiently large values of N , continuous part of the energy level density satisfies
the Gaussian distribution with a high degree of accuracy. We also derive exact expressions for
the mean value and standard deviation which characterize such Gaussian distribution. Section 6
is the concluding section.

2. Application of the freezing trick

For the purpose of defining the SU(m|n) supersymmetric HS spin chain, let us consider a set
of operators like Cj†α (Cj α ) which creates (annihilates) a particle of species α on the j th lattice
site. These creation (annihilation) operators are assumed to be bosonic when α ∈ [1, 2, . . . , m]
and fermionic when α ∈ [m + 1, m + 2, . . . , m + n]. Thus, the parity of Cj†α (Cj α ) is defined as
 
p(Cj α ) = p Cj†α = 0 for α ∈ [1, 2, . . . , m],
 
p(Cj α ) = p Cj†α = 1 for α ∈ [m + 1, m + 2, . . . , m + n].
These operators satisfy commutation (anticommutation) relations like
 † †   † 
[Cj α , Ckβ ]± = 0, Cj α , Ckβ ±
= 0, Cj α , Ckβ ±
= δj k δαβ , (2.1)
where [A, B]± ≡ AB − (−1)p(A)p(B) BA. Next, we focus our attention to a subspace of the
related Fock space, for which the total number of particles per site is always one:

m+n
Cj†α Cj α = 1, (2.2)
α=1
for all j . On the above mentioned subspace, one can define supersymmetric exchange operators
as
(m|n)
 † †
m+n
P̂j k ≡ Cj α Ckβ Cjβ Ckα , (2.3)
α,β=1
(m|n)
where 1  j < k  N . These P̂j k ’s yield a realization of the permutation algebra given by

Pj2k = 1, Pj k Pkl = Pj l Pj k = Pkl Pj l , [Pj k , Plm ] = 0, (2.4)


(m|n)
where j, k, l, m are all distinct indices. Replacing Pj k by P̂j k in Eq. (1.1), one obtains the
Hamiltonian of SU(m|n) supersymmetric HS model as [3]
(m|n)
1  (1 + P̂j k )
H (m|n)
= . (2.5)
2
1j <kN
sin2 (ξj − ξk )
B. Basu-Mallick, N. Bondyopadhaya / Nuclear Physics B 757 [FS] (2006) 280–302 283

Now we want to describe how this SU(m|n) supersymmetric HS model (2.5), containing
creation–annihilation operators, can be transformed to a spin chain. To this end, we consider a
particular type of anyon like representation of permutation algebra (2.4), which acts on a spin
state like |α1 α2 . . . αN  (with αj ∈ [1, 2, . . . , m + n]) as [16,17]
(m|n)
P̃j k |α1 · · · αj · · · αk · · · αN  = eiΦ(αj ,αj +1 ,···,αk ) |α1 · · · αk · · · αj . . . αN , (2.6)

where eiΦ(αj ,αj +1 ,...,αk ) = 1 if αj , αk ∈ [1, 2, . . . , m], eiΦ(αj ,αj +1 ,...,αk ) = −1 if αj , αk ∈ [m + 1,


k−1 m+n
m + 2, . . . , m + n], and eiΦ(αj ,αj +1 ,...,αk ) = (−1)π p=j +1 τ =m+1 δαp ,τ if αj ∈ [1, 2, . . . , m] and
αk ∈ [m + 1, m + 2, . . . , m + n] or vice versa. For the purpose of interpreting the phase
factor eiΦ(αj ,αj +1 ,...,αk ) in a physical way, it is convenient to call αi a ‘bosonic’ spin when
αi ∈ [1, 2, . . . , m] and a ‘fermionic’ spin when αi ∈ [m + 1, m + 2, . . . , m + n]. From Eq. (2.6)
it follows that, the exchange of two bosonic (fermionic) spins produces a phase factor of 1 (−1)
irrespective of the nature of spins situated in between the j th and kth lattice sites. However, if we
exchange one bosonic spin with one fermionic spin, then the phase factor becomes (−1)ρ where
ρ is the total number of fermionic spins situated in between the j th and kth lattice sites.
Next we observe that, due to the constraint (2.2), the Hilbert space associated with SU(m|n)
HS Hamiltonian (2.5) can be spanned through the following orthonormal basis vectors:

C1α C † · · · CN
1 2α2

αN |0, where |0 is the vacuum state and αj ∈ [1, 2, . . . , m + n]. Consequently,
it is possible to define a one-to-one mapping between these basis vectors and those of the above
mentioned spin chain as

|α1 α2 · · · αN  ↔ C1α C † · · · CN
1 2α2

αN |0. (2.7)
With the help of commutation (anticommutation) relations (2.1), one can easily verify that
(m|n) †
P̂j k C1α1
· · · Cj†αj · · · Ckα

k

· · · CN αN |0

= eiΦ(αj ,...,αk ) C1α 1
· · · Cj†αk · · · Ckα

j

· · · CN αN |0, (2.8)

where eiΦ(αj ,...,αk ) is the same phase factor which appeared in Eq. (2.6). Comparison of Eq. (2.8)
(m|n)
with Eq. (2.6) through the mapping (2.7) reveals that, the anyon like representation P̃j k is
(m|n)
equivalent to the supersymmetric exchange operator P̂j k . Hence, if we define a spin chain
(m|n)
Hamiltonian through P̃j k as
(m|n)
1  (1 + P̃j k )
H (m|n)
= , (2.9)
2
1j <kN
sin (ξj − ξk )
2

that would be completely equivalent to the supersymmetric SU(m|n) HS model (2.5) [16].
(m|n)
Clearly, for the special case n = 0, P̃j k reproduces the original spin exchange operator Pj k
and H (m|n) (2.9) reduces to the Hamiltonian of SU(m) HS spin chain (1.1). Since it is convenient
to apply the freezing trick to the spin chain Hamiltonian (2.9), for the rest of this article we shall
deal with this form of supersymmetric SU(m|n) HS model instead of its original form (2.5).
(m|n)
By using the anyon like representation P̃j k , one can construct a spin CS model like
(m|n)
N
∂2  (a + P̃j k )

H =− + 2a , (2.10)
j =1
∂xj2 1j <kN
sin2 (xj − xk )
284 B. Basu-Mallick, N. Bondyopadhaya / Nuclear Physics B 757 [FS] (2006) 280–302

which contains spin as well as dynamical degrees of freedom and the positive parameter a as
coupling constant. With the help of mapping (2.7) it can be shown that, this spin CS model is
equivalent to a supersymmetric spin CS model [18] with Y (gl(m|n) ) super-Yangian symmetry.
The spin CS Hamiltonian H ∗ (2.10) might be formally written as
H ∗ = H0 + 4a H(m|n) , (2.11)
where H0 is the Hamiltonian of spinless CS model given by [19]
N
∂2  1
H0 = − + 2a(a − 1) , (2.12)
j =1
∂xj2
1j <kN
sin (xj − xk )
2

and H(m|n) is obtained from H (m|n) (2.9) by the replacement ξj → xj . Now the decoupling of the
dynamical degrees of freedom of H ∗ (2.10) from its spin degrees of freedom can be achieved by
using the freezing trick [10–12]. This trick is based on the fact that in the limit a → ∞, particles
freeze at the equilibrium positions of H0 , which are simply the lattice points (ξj ) of the spin
chain in Eq. (2.9). Consequently, by using Eq. (2.11) at freezing limit, we find that the energy
levels of H ∗ are approximately given by
Ej∗k E0,j + 4aEk , (2.13)
where E0,j and Ek are any two levels of H0 and H (m|n) respectively. Hence, we obtain a relation
like
Z ∗ (4aT )
Z (m|n) (T ) = lim , (2.14)
a→∞ Z0 (4aT )

where Z (m|n) , Z ∗ and Z0 denote the partition functions corresponding to the Hamiltonians
H (m|n) , H ∗ and H0 respectively. Thus the freezing trick allows us to compute the partition func-
tion of SU(m|n) supersymmetric HS spin chain, by modding out the contribution of spinless CS
model from the partition function of spin CS model (2.10). Due to the Gallelian invariance of H ∗
and H0 it follows
that, if ψ is an eigenstate of any one of these Hamiltonians with momentum p,
N
then ψ
= e2iτ j =1 xj ψ will also be an eigenstate of the same Hamiltonian with momentum
(p + 2τ N). As a result, we can always adjust the parameter τ such that ψ
will be an eigenfunc-
tion of H ∗ or H0 with zero momentum. In this article, we shall always consider eigenstates of
these Hamiltonians with zero momenta and evaluate the partition functions Z ∗ as well as Z0 at
the center of mass frame. Since both Z ∗ and Z0 get modified by the same multiplicative factor
due to a Gallelian transformation, Z (m|n) does not depend on the choice of the reference frame.

3. Spectrum of spin CS model

In this section our aim is to find out the complete spectrum of spin CS model (2.10) containing
anyon like representation of the permutation algebra. Even though the spectrum of such spin CS
model has been studied earlier [17], multiplicities of degenerate eigenfunctions corresponding
to all energy levels have not been found. Since these numbers are required for calculating the
partition function of this model, here we want to derive a general expression for the degeneracy
factors of all energy levels. It is well known that the eigenfunctions of spin CS Hamiltonian
(2.10) can be written in a factorised form like
ψ(x1 , . . . , xN ; α1 , . . . , αN ) = Γ a φ(x1 , . . . , xN ; α1 , . . . , α2 ), (3.1)
B. Basu-Mallick, N. Bondyopadhaya / Nuclear Physics B 757 [FS] (2006) 280–302 285


where Γ = i<j sin(xi − xj ). By operating H ∗ (2.10) on the above form of ψ , we find that

H ∗ ψ = Γ a H̃ ∗ φ, (3.2)
where

 zk + zj

∗ ∂ 2 ∂
H̃ = 4 zj +a zk − zj
∂zj z − zj ∂zk ∂zj
j k<j k

(m|n)  zj zk a2  2 
− 2a 1 + P̃j k + N N −1 , (3.3)
(zj − zk )2 12
k<j

with zj = e2ixj . Eq. (3.2) implies that, if φ is an eigenvector of H̃ ∗ with eigenvalue E, then Γ a φ
would be an eigenvector of H ∗ with the same eigenvalue. Thus the diagonalisation problem of
H ∗ is reduced to the diagonalisation problem of H̃ ∗ .
For solving H̃ ∗ , it is convenient to introduce another operator H which acts only on the
coordinate degree of freedom and may be given by [7]

 zk + zj

∂ 2 ∂
H=4 zj +a zk − zj
∂zj z − zj ∂zk ∂zj
j k<j k

zj zk a2  2 
− 2a (1 − Kj k ) + N N −1 , (3.4)
(zj − zk )2 12
k<j

where the Kj k is the coordinate exchange operator which exchanges the coordinates of j th and
kth particle:

Kj k f (x1 , . . . , xj , . . . , xk , . . . , xN ) = f (x1 , . . . , xk , . . . , xj , . . . , xN ).

It may be observed that, H̃ ∗ (3.3) can be reproduced from the expression of H (3.4) through the
substitution Kj k → −P̃j k . This connection between H̃ ∗ and H will play a crucial role in our
(m|n)

calculation for finding the spectrum of H̃ ∗ . Let us first consider state vectors given by monomials
like
p p p
ξp = z1 1 z2 2 · · · zNN , (3.5)

where p ≡ {p1 , p2 , . . . , pN } ∈ RN satisfies the constraints: (i) (pi − pj ) are integers for all i, j

and (ii) N i=1 pi = 0. The last condition implies that these monomials represent state vectors
with zero total momentum. In particular, one can consider ξp̂ corresponding to a nonincreasing
vector p̂ ≡ (p̂1 , p̂2 , . . . , p̂N ), whose elements satisfy
 the conditions: (i) li ≡ p̂i − p̂i+1 is a non-
negative integer for i ∈ [1, . . . , N − 1] and (ii) N i=1 p̂i = 0. It is evident that, N − 1 number of
nonnegative integers (li ’s) are sufficient to specify a nonincreasing vector p̂. Given two distinct
nonincreasing vectors p̂ and p̂
, we shall write p̂ ≺ p̂
if p̂1 − p̂1
= · · · = p̂i−1 − p̂i−1
= 0 and

p̂i < p̂i . A partial ordering can be defined on monomials like ξp (3.5) in the following way. By
permuting the elements of any p, one can always construct a unique nonincreasing vector p̂. The
basis element ξp would precede ξp
if p̂ ≺ p̂
, where p̂ and p̂
are nonincreasing vectors obtained
from p and p
respectively by permuting their components. The above defined ordering is effec-
tively a partial ordering, since it does not induce an ordering between ξp and ξp
when p̂ = p̂
. It
286 B. Basu-Mallick, N. Bondyopadhaya / Nuclear Physics B 757 [FS] (2006) 280–302

can be shown that the action of H (3.4) on the state vector ξp yields [7,9]

Hξp = Ep̂ ξp + cpp
ξp
, (3.6)
p

(p̂
<p̂)

where

N
2
Ep̂ = 2p̂i + a(N + 1 − 2i) . (3.7)
i=1
Thus it is clear that, if one constructs a Hilbert space through basis vectors of the form (3.5) and
partially order them in the above mentioned way, then H will act as a triangular matrix on this
space.
Next we want to construct another partially ordered Hilbert space, on which H̃ ∗ (3.3) can
be represented as a triangular matrix. To this end, we define a set of permutation operators as
(m|n) (m|n) (m|n)
Πj k = P̃j k Kj k . Since both P̃j k and Kj k satisfy an algebra of the form (2.4), while act-
(m|n)
ing on the spin and coordinate spaces respectively, the newly defined operator Πj k also yields
a representation of the same permutation algebra on the direct product of coordinate and spin
spaces. Hence, by using this representation of permutation algebra, we can construct a ‘general-
ized’ antisymmetric projection operator Λ(m|n) satisfying the relation
(m|n) (m|n)
Πj k Λ(m|n) = Λ(m|n) Πj k = −Λ(m|n) , (3.8)
(m|n)
or, equivalently, P̃j k Λ(m|n) = −Kj k Λ(m|n) [16,17]. Even though Λ(m|n) can be expressed as
(m|n)
a function of Πj k , explicit form of this projection operator is not necessary for our present
(m|n)
purpose. However it may be noted that, since both Kj k and P̃j k commute with H (3.4), Λ(m|n)
also satisfies the relation
 
H, Λ(m|n) = 0. (3.9)
With the help of projection operator Λ(m|n) ,
we define a state vector on the direct product of
coordinate and spin spaces as

φpα ≡ φpα11 ...p
...αN
N
= Λ(m|n) ξp |α1 · · · αN  . (3.10)
Using Eqs. (3.8) and (2.6) it can be shown that
α ...α ...α ...α
φp11 ...pjj ...pkk ...pNN
(m|n) p1 p p p 
= −Λ(m|n) Kj k P̃j k z1 · · · zj j · · · zk k · · · zNN |α1 · · · αj · · · αk · · · αN 
p p p p 
= −eiΦ(αj ,...,αk ) Λ(m|n) z1 1 · · · zj k · · · zk j · · · zNN |α1 · · · αk · · · αj · · · αN 
α ...α ...α ...α
= −eiΦ(αj ,...,αk ) φp11 ...pkk ...pjj ...pNN . (3.11)
By repeatedly using the above equation we find that

φpα = (α, p)φp̂α , (3.12)


where (α, p) = ±1, p̂ is the nonincreasing vector corresponding to p and α
is a spin vector
which is obtained by permuting the components of α. Hence, all state vectors of the form (3.10)
can be obtained by choosing p from the set of nonincreasing vectors only.
B. Basu-Mallick, N. Bondyopadhaya / Nuclear Physics B 757 [FS] (2006) 280–302 287

Corresponding to any given nonincreasing vector p̂, one can define a vector space as
  
Vp̂ ≡ φ α α1 , . . . , αN ∈ [1, 2, . . . , m + n] .
p̂ (3.13)
It is important to note that, different values of α may lead to the same φp̂α which is a basis element
of Vp̂ . To see this thing in a simple way, let us take a nonincreasing sequence p̂ satisfying the
condition p̂i = p̂j = p̂ (say). For this special case, Eq. (3.11) reduces to
α ...α ...α ...αN α ...α ...α ...αN
i
φp̂1 ...p̂... j
p̂...p̂
= −eiΦ(αi ,...,αj ) φp̂1 ...p̂...
j i
p̂...p̂
. (3.14)
1 N 1 N
α ...α ...α ...α α ...α ...α ...α
i
Clearly φp̂1 ...p̂... j
p̂...p̂N
N j
and φp̂1 ...p̂... i
p̂...p̂N
N
represent the same state vector (up to a phase factor),
1 1
although they correspond to different values of α. For a given φp̂α , we say that two spin compo-
nents of α belong to the same ‘sector’ if the corresponding two components of p̂ are equal to
α ...αi ...αj ...αN
each other. For example, the spin components αi and αj appearing in the state φp̂1 ...p̂... p̂...p̂ 1 N
belong to the same sector according to this convention. Since eiΦ = 1, for αi , αj ∈ [1, 2, . . . , m],
it is clear from Eq. (3.14) that bosonic spins within the same sector obey ‘fermionic statistics’
after antisymmetrisation. In particular, two bosonic spins of same flavour cannot coexist within
a single sector. Similarly, since eiΦ = −1 for αi , αj ∈ [m + 1, m + 2, . . . , m + n], one can find
from Eq. (3.14) that fermionic spins within the same sector obey ‘bosonic statistics’ after an-
tisymmetrisation. Therefore, any number of fermionic spins having the same flavour can be
accommodated within a single sector.
Now we want to find out the dimensionality of the space Vp̂ . For this purpose, it is useful to
write p̂ in the form
k1 ki kr
        
p̂ ≡ ρ1 , . . . , ρ1 , . . . , ρi , . . . , ρi , . . . , ρr , . . . , ρr , (3.15)
r
where ρ1 > · · · > ρi > · · · > ρr , i=1 ki = N , and r is an integer which can take any value from
1 to N . It is obvious that k ≡ {k1 , . . . , kr }, which belongs to the set PN of ordered partitions of N ,
may be treated as a function of p̂. For a given φp̂α , clearly the components of α are separated
into r different sectors where the ith sector contains ki number of spins. It is evident that the
dimensionality of the space Vp̂ may be obtained by counting the number of independent ways
one can distribute total N number of spins within r sectors. To this end, let us first try to find out
the number of independent ways of filling up the ith sector through j1 number of bosonic spins
and j2 number of fermionic spins, where j1 + j2 = ki . Using Eq. (3.14) we have already seen
that, bosonic and fermionic spins within the same sector obey fermionic and bosonic statistics
respectively. Therefore, we can pick up j1 number of bosonic spins from m different flavours
in m Cj1 different ways and j2 number of fermionic spins from n different flavours in j2 +n−1 Cj2
p!
ways, where p Cl = l!(p−l)! for l  p and p Cl = 0 for l > p. Thus the number of independent
ways of filling up the ith sector through j1 number of bosonic spins and j2 number of fermionic
spins is given by
m
Cj1 n+j2 −1 Cj2 .
Summing up these numbers for all possible values of j1 and j2 , we obtain the total number of
independent ways of filling up the ith sector through ki number of spins as

ki
m   i)
min(m,k
m!(n + ki − j1 − 1)!
n+ki −j1 −1
d (m|n)
(ki ) = Cj1 Cki −j1 = .
j1 !(m − j1 )!(n − 1)!(ki − j1 )!
j1 =0 j1 =0
(3.16)
288 B. Basu-Mallick, N. Bondyopadhaya / Nuclear Physics B 757 [FS] (2006) 280–302

Since two spins belonging to different sectors do not follow any exchange relation like (3.14),
the number of independent ways we can distribute total N number of spins within r different
sectors is given by the product of all d (m|n) (ki ). Therefore, by using (3.16), we finally obtain the
dimension of Vp̂ as
 min(m,k ) 

r 
r i
n+ki −j −1
d (m|n)
(k) = d (m|n)
(ki ) = m
Cj Cki −j . (3.17)
i=1 i=1 j =0

Even though this expression is derived by assuming that bosonic and fermionic spin degrees
of freedom (i.e., m and n respectively) take nonzero values, it is also possible to obtain the
dimension of Vp̂ for the SU(n) fermionic case by putting m = 0 in Eq. (3.17):


r
n+ki −1
d (0|n) (k) = Cki . (3.18)
i=1

Furthermore, by putting n = 0 in Eq. (3.17), subsequently using the relation p Cl = 0 for l >
p  0, and also assuming that −1 C0 = 1, one can reproduce the dimension of Vp̂ for the SU(m)
bosonic case [9] as


r
d (m|0) (k) = m
Cki . (3.19)
i=1

It is interesting to observe that, while d (m|0) (k) (3.19) can take a nonzero value only if ki  m
for all i, both d (m|n) (k) (3.17) and d (0|n) (k) (3.18) take nonzero values for any k ∈ PN . Conse-
quently, Vp̂ will represent a nontrivial vector space for the SU(m) bosonic case only if at most
m components of p̂ take the same value. On the other hand, Vp̂ will represent a nontrivial vector
space for all possible values of p̂ when at least one fermionic spin degrees of freedom is present.
The Hilbert space associated with H̃ ∗ (3.3) may now be defined by taking the direct sum of
Vp̂ (3.13) for all allowed values of p̂:

V= Vp̂ . (3.20)

We define a partial ordering in this Hilbert space by saying that the basis element φp̂α precedes

φp̂α
if p̂ ≺ p̂
. By consecutively applying the relations (3.10), (3.8), (3.9), (3.6) and (3.12), it is
easy to check that
 
H̃ ∗ φp̂α = Λ(m|n) Hξp̂ |α1 · · · αN 


=Λ (m|n)
Ep̂ ξp̂ + cp̂p ξp |α1 · · · αN 

(p̂
<p̂)
 
= Ep̂ φp̂α + (α, p
)cp̂p
φp̂α
. (3.21)
p

(p̂
<p̂)
B. Basu-Mallick, N. Bondyopadhaya / Nuclear Physics B 757 [FS] (2006) 280–302 289

Hence H̃ ∗ is represented as a triangular matrix on V. Diagonal elements of this triangular matrix


yield the eigenvalues of H̃ ∗ as

N
2
E ∗ (p̂, α) ≡ Ep̂ = 2p̂i + a(N + 1 − 2i) . (3.22)
i=1

Consequently, the eigenvalues of spin CS Hamiltonian H ∗ (2.10) are also given by E ∗ (p̂, α) in
the above equation.
Since E ∗ (p̂, α) in Eq. (3.22) does not really depend on the spin vector α, the number of
degenerate energy eigenstates associated with the quantum number p̂ would coincide with the
dimension of the space Vp̂ . Thus the degeneracy factor of the energy eigenvalue E ∗ (p̂, α) cor-
responding to the quantum number p̂ is given by d (m|n) (k) appearing in Eq. (3.17). We have
already seen that, in contrast to the pure bosonic case, d (m|n) (k) takes nonzero values for all
possible k ∈ PN when at least one fermionic spin degrees of freedom is present. Consequently,
the presence of fermionic spin degrees of freedom in H ∗ (2.10) would lead to a spectrum with
many additional energy levels in comparison with the spectrum of bosonic spin CS model.
Finally let us briefly comment about the known spectrum of spinless CS Hamiltonian H0
(2.12) [19]. Using the fact that the eigenfunctions of H0 can be written in a factorised form like
ψ0 = Γ a φ0 (x1 , . . . , xN ), it is possible to transform H0 into H̃0 as
H0 ψ0 = Γ a H̃0 φ0 ,
where H̃0 can be obtained from H (3.4) through the substitution Kij → 1. For constructing the
Hilbert space associated with H̃0 , one may consider elements like φp ≡ Λ0 (ξp ), where Λ0 is the
symmetriser in the coordinate space: Kj k Λ0 = Λ0 . Since φp = φp̂ , where p̂ is the nonincreasing
vector corresponding to p, the Hilbert space of H̃0 is defined through independent basis vectors
φp̂ for all values of p̂. An ordering can be defined among these state vectors by saying that
φp̂ precedes φp̂
if p̂ ≺ p̂
. Using Eq. (3.6) it can be shown that, H̃0 acts as a triangular matrix
on these completely ordered basis vectors and the eigenvalues of H0 are also given by Ep̂ in
Eq. (3.22). However, due to the absence of spin degrees of freedom, only one energy eigenstate
is obtained corresponding to each quantum number p̂ in this case.

4. Partition function of SU(m|n) HS spin chain

By using the freezing trick we have seen that, the partition function of supersymmetric
SU(m|n) HS spin chain can be obtained by dividing the partition function of spin CS model
(2.10) at the strong coupling limit through that of the spinless CS model (2.12). To execute this
program, let us first briefly recapitulate the calculation for the partition function of spinless CS
model (2.12) at a → ∞ limit [9]. It should be noted that, the eigenvalues in Eq. (3.22) can be
expanded in powers of a as

N

E (p̂, α) ≡ E(p̂) = a E0 + 4a
2
(N + 1 − 2i)p̂i + O(1), (4.1)
i=1

where E0 = 13 N (N 2 − 1). Since E0 does not depend on p̂ or α, the effect of this E0 will be
manifested as the same overall multiplicative factor in the partition functions of spin CS model
and its spinless counterpart. Hence, by dropping the first term in Eq. (4.1), and neglecting the
O(1) term in the limit a → ∞, one can write down the partition function of spinless CS model
290 B. Basu-Mallick, N. Bondyopadhaya / Nuclear Physics B 757 [FS] (2006) 280–302

(2.12) as
 
Z0 (4aT ) q i p̂i (N+1−2i)
, (4.2)

where q = e−1/(kB T ) . Using N − 1 number of nonnegative integers (li ’s) which uniquely deter-
mine p̂, one can further simplify this partition function as [9]

 
N−1 
N−1
1
Z0 (4aT ) q j (N−j )lj = . (4.3)
1 − q j (N−j )
l1 ,...,lN−1 0 j =1 j =1

Next, we want to calculate the partition function of spin CS Hamiltonian (2.10) at a → ∞


limit. Dropping again the first term as well as O(1) term from the right-hand side of expansion
(4.1), and expressing the nonincreasing vector p̂ through Eq. (3.15), E ∗ (p̂, α) can be written as


r 
Ki
E ∗ (p̂, α) 4a ρi (N + 1 − 2j ), (4.4)
i=1 j =Ki−1 +1

where Ki = ij =1 kj denote the partial sums corresponding to the partition k ∈ PN and K0 = 0.
Using a set of variables like νj ≡ ρj − ρj +1 for j ∈ [1, . . . , r − 1] (since ρj > ρj +1 , all νj ’s are
positive integers), one can express the energy eigenvalue in Eq. (4.4) as


r−1
E ∗ (p̂, α) 4a νj Nj , (4.5)
j =1

where Nj = Kj (N − Kj ). It may be noted that, due to the condition N i=1 p̂i = 0, r − 1 number
of νj’s uniquely determine the nonincreasing  vector
 p̂ in Eq. (3.15). Consequently, the single
sum p̂ can be replaced by the double sum k∈PN ν1 ,...,νr−1 >0 in the expression of the parti-
tion function. By using the eigenvalue relation (4.5) and the corresponding degeneracy factor
d (m|n) (k) (3.17), we obtain the partition function of spin CS Hamiltonian (2.10) at a → ∞
limit as
 
r−1
Z ∗ (4aT ) d (m|n) (k) q Nj ν j
p̂ j =1

  
r−1
= d (m|n) (k) q Nj ν j
k∈PN ν1 ,...,νr−1 >0 j =1

 
r−1
q Nj
= d (m|n)
(k) . (4.6)
k∈PN j =1
1 − q Nj

Using Eqs. (2.14), (4.3) and (4.6), we finally obtain the partition function of SU(m|n) HS spin
chain as

N−1
   
r−1
q Nj
Z (m|n) (T ) = 1 − q l(N−l) d (m|n) (k) . (4.7)
l=1 k∈PN j =1
(1 − q Nj )
B. Basu-Mallick, N. Bondyopadhaya / Nuclear Physics B 757 [FS] (2006) 280–302 291

Since the partial sums K1 , K2 , . . . , Kr−1 associated with k are natural numbers obeying 1 
K1 < · · · < Kr−1  N − 1, one can define their complements (Ki
’s) as elements of the set:
{1, . . . , N − 1} − {K1 , . . . , Kr−1 }, which satisfy the ordering K1
< · · · < KN
−r . Hence one can

rearrange the product N−1 l=1 (1 − q
l(N−l) ) into two terms as [9]


N−1
 
 r−1 
 N−r

1 − q l(N−l) = 1 − q Nj 1 − q Ni , (4.8)
l=1 j =1 i=1

where Ni
= Ki
(N − Ki
). By substituting this relation to Eq. (4.7), we get a simplified expression
for the partition function of SU(m|n) HS model as

r−1
 
Nj N−r

Z (m|n)
(T ) = d (m|n)
(k)q j =1
1 − q Ni . (4.9)
k∈PN i=1

We have already seen that, both d (m|n) (k) (3.17) and d (0|n) (k) (3.18) take nonzero values for any
k ∈ PN . Consequently, in contrast to the restricted choice of k for the case of bosonic spin chain
[9], all possible k ∈ PN will contribute to the partition function (4.9) in the case of supersym-
metric as well as fermionic HS spin chain.
It is well known that the spectrum of bosonic SU(m) HS spin chain (1.1) containing N number
of lattice sites can be obtained from motifs like δ ≡ (0, δ1 , . . . , δN−1 , 0), where each δj is either
0 or 1 [6–8]. The form of these motifs and corresponding eigenvalues can be reproduced by
using the partition function of bosonic SU(m) HS spin chain [9]. Now we want to explore how
the motifs associated with SU(m|n) supersymmetric HS spin chain emerge naturally from the
expression of partition function (4.9). To this end, we define a motif corresponding to the partition
k by using the following rule: δj = 0 if j coincides with one of the partial sums Ki and δj = 1
otherwise. Furthermore, it is assumed that the lowest power of q in Eq. (4.9) for the partition k
gives the energy eigenvalue E(δ) of the above motif δ. In this way we obtain the energy levels
of SU(m|n) supersymmetric HS spin chain as

r−1
N (N 2 − 1) 
N−1
E(δ) = Ni = + δj j (j − N ), (4.10)
6
i=1 j =1

which apparently coincides with that of the bosonic HS spin chain. However it should be ob-
served that, for the case of SU(m) spin chain, only those k would contribute in the partition
function for which Kj − Kj −1 = kj  m [9]. This leads to a selection rule which prohibits the
occurrence of m or more consecutive 1’s within the corresponding motifs. On the other hand,
since all k ∈ PN contribute to the partition function (4.9) of supersymmetric HS spin chain, it is
possible to place any number of consecutive 1’s or 0’s within a motif δ. Consequently, the se-
lection rule occurring in the bosonic case is lifted for the case of supersymmetric HS spin chain
and many extra energy levels appear in the corresponding spectrum. This absence of selection
rule in the spectrum of supersymmetric HS spin chain was previously observed by Haldane on
the basis of numerical calculations [3]. By using the expression of E(δ) in Eq. (4.10), we can
easily evaluate the maximum and minimum energy eigenvalues of this system. From the expres-
sion of E(δ) it is evident that, the motif δ ≡ (0, 0, . . . , 0, 0) would correspond to the maximum
energy Emax = N (N6 −1) . Similarly for the motif δ ≡ (0, 1, . . . , 1, 0), we obtain the minimum en-
2


ergy of the system as Emin = N (N6 −1) + N−1
2
j =1 j (j − N ) = 0. It is interesting to note that these
292 B. Basu-Mallick, N. Bondyopadhaya / Nuclear Physics B 757 [FS] (2006) 280–302

maximum and minimum energy eigenvalues of SU(m|n) supersymmetric HS spin chain do not
depend on the values of m and n. Moreover, the lifting of the selection rule is responsible for the
zero minimum energy of supersymmetric HS spin chain.
Using Mathematica we find that, for a wide range of values of m, n and N , the partition
function (4.9) of SU(m|n) HS model satisfies a duality relation of the form
N(N 2 −1)  
Z (m|n) (q) = q 6 Z (n|m) q −1 . (4.11)
This result motivates us to conjecture that the above duality relation, involving the interchange
of bosonic and fermionic spin degrees of freedom, is valid for all possible values of m, n and
N . It may be noted that, for the particular case n = 0, Eq. (4.11) relates the partition function of
SU(m) bosonic HS spin chain to that of SU(m) fermionic spin chain. By applying the relation

= N (N6 −1) [9,20], we find
(m|0) (0|m) 2
P̃j k = −P̃j k and the summation formula 1j <kN 2 1
sin (ξj −ξk )
that the Hamiltonians of bosonic and fermionic spin chains are connected as
N (N 2 − 1)
H (m|0) = − H (0|m) .
6
Using the above relation along with the definition of partition function given by Z (m|n) (q) =
(m|n)
tr[q H ], one can easily prove Eq. (4.11) for the particular case n = 0. It would be interesting to
explore whether Eq. (4.11) can be also proved for the general case by establishing some relation
(m|n) (n|m)
between P̃j k and P̃j k . Comparing the coefficients of same power of q from both sides of
Eq. (4.11), we find that the energy levels of SU(n|m) spin chain can be obtained from those of
SU(m|n) spin chain through the transformation Ei → N (N6 −1) − Ei and also get the relation
2



(n|m) N (N − 1)
2
D (m|n)
(Ei ) = D − Ei , (4.12)
6
where D(m|n) (Ei ) denotes the degeneracy factor corresponding to energy Ei of SU(m|n) HS
spin chain. Thus it is evident that, the spectrum of SU(n|m) spin chain can be obtained from that
of SU(m|n) spin chain through an inversion and overall shift of all energy levels. Such relation
between the spectra of supersymmetric HS spin chains was empirically found by Haldane with
the help of numerical analysis [3].

5. Spectral properties of SU(m|n) HS spin chain

In this section we shall explore some spectral properties of supersymmetric HS model by


using its exact partition function Z (m|n) (T ) (4.9). It has been already mentioned that, calculation
for the degeneracy factors associated with the energy eigenvalues of this spin chain becomes very
cumbersome by using the motif representations for large values of N . However, with the help
of a symbolic software package like Mathematica, it is possible to express the partition function
(4.9) as a polynomial of q and explicitly find out the degeneracy factors of all energy levels for
relatively large values of N . In this way, we can study properties like level density distribution
and nearest-neighbour spacing (NNS) distribution for the spectrum of supersymmetric HS spin
chain.
For the case of SU(m) bosonic spin chain, it has been found earlier that the continuous part of
the energy level density obeys Gaussian distribution to a very high degree of accuracy for N 1
[9]. At present, our aim is to study the level density distribution in the spectrum of SU(m|n)
B. Basu-Mallick, N. Bondyopadhaya / Nuclear Physics B 757 [FS] (2006) 280–302 293

Fig. 1. Energy levels Ei and degeneracies D(1|1) (Ei ) of the SU(1|1) HS spin chain for N = 15.

supersymmetric HS spin chain and investigate whether it exhibits a similar behaviour. To begin
with, let us consider the simplest case of SU(1|1) supersymmetric HS spin chain. In this case, the
degeneracy factor in Eq. (3.17) reduces to a simple form given by d (1|1) (k) = 2r . By substituting
this degeneracy factor to Eq. (4.9), taking some specific value for the number of lattice sites
like N = 15 and using Mathematica, we express the partition function of SU(1|1) spin chain
as a polynomial of q. The coefficient of q Ei in such polynomial evidently gives the degeneracy
factor D(1|1) (Ei ) corresponding to the energy eigenvalue Ei , which we plot in Fig. 1. This figure
clearly indicates that the energy level distribution obey Gaussian approximation but with some
local fluctuations. Similar behaviour of energy level distribution has been found by studying
SU(m|n) HS spin chain with other values of m, n and sufficiently large values of N .
From the above discussion it is apparent that, if we decompose the energy level density associ-
ated with SU(m|n) HS spin chain as a sum of continuous part and fluctuating part, the continuous
part will obey Gaussian distribution for large values of N . This behaviour of the continuous part
can be measured in a quantitative way by studying the cumulative level density [9], which elimi-
nates the fluctuating part of the level density distribution. For the case of SU(m|n) HS spin chain,
cumulative level density of the spectrum is defined as

1 
F (E) = D(m|n) (Ei ). (5.1)
(m + n)N
Ei E

Obviously, this F (E) can also be obtained by expressing the exact partition function (4.9) as
a polynomial of q. We want to check whether this F (E) agrees well with the error function
294 B. Basu-Mallick, N. Bondyopadhaya / Nuclear Physics B 757 [FS] (2006) 280–302

given by


1 E−μ
G(E) = 1 + erf √ , (5.2)
2 2σ
where μ and σ are respectively the mean value and the standard deviation associated with the en-
ergy level density distribution. These parameters are related to the Hamiltonian H (m|n) (2.9) as
tr[H (m|n) ]
μ= , (5.3a)
(m + n)N
tr[(H (m|n) )2 ]
σ2 = − μ2 . (5.3b)
(m + n)N
For the purpose of comparing F (E) with G(E), it is necessary to express the parameters μ
and σ as some functions of m, n and N . To this end, we need the following trace formulas:
 (m|n) 2 
tr P̃ij = tr[1] = s N , (5.4a)
 (m|n) 
tr P̃ij = s N−2 t, (5.4b)
 (m|n) (m|n)   (m|n) (m|n)   (m|n) (m|n) 
tr P̃ij P̃il = tr P̃ij P̃j l = tr P̃ij P̃kj =s N−2
, (5.4c)
 (m|n) (m|n) 
tr P̃ij P̃kl =s N−4 2
t , (5.4d)
where s = m + n, t = m − n and i, j, k, l are all different indices. Derivation of these trace
formulas is given in Appendix A of this article. For obtaining the functional form of μ and σ , it
is also required to evaluate summations like
 1
R0 ≡ , (5.5a)
i<j
sin (ξi − ξj )
2

 1
R1 ≡ , (5.5b)
i<j
sin (ξi − ξj )
4

  1
R2 ≡ , (5.5c)
i<j k<l
sin (ξi − ξj ) sin2 (ξk − ξl )
2
(k,l=i,j )
 1  1
R3 ≡ 2 +
i<j j <l
sin (ξi − ξj ) sin (ξj − ξl )
2 2
i<j i<l
sin (ξi − ξj ) sin2 (ξi − ξl )
2

 1
+ . (5.5d)
i<j k<j
sin (ξi − ξj ) sin2 (ξk − ξj )
2

It is easy to see that the above defined R0 , R1 , R2 and R3 satisfy the relation
R02 = R1 + R2 + R3 . (5.6)
Using some summation formulas given in Ref. [20], it can be shown that
N(N 2 − 1)
R0 = , (5.7a)
6
N(N 2 − 1)(N 2 + 11)
R1 = , (5.7b)
90
B. Basu-Mallick, N. Bondyopadhaya / Nuclear Physics B 757 [FS] (2006) 280–302 295

N(N 2 − 1)2 (N − 4) N (N 2 − 1)(N 2 + 11)


R2 = + , (5.7c)
36 90
4N (N 2 − 1)(N 2 − 4)
R3 = . (5.7d)
45
Derivation of these relations is discussed in Appendix B of this article.
Now, by using Eqs. (5.3a), (5.4a), (5.4b) and (5.7a), we can express μ as a function of m, n
and N given by
s2 + t s2 + t  2 
μ= 2
R0 = 2
N N −1 . (5.8)
2s 12s
Next, by using the trace formulas (5.4a)–(5.4d), we obtain
 2  s N−2 (s 2 + 2t) 2 s N s N−4 t 2 s N−2
tr H (m|n) = R0 + R1 + R2 + R3 . (5.9)
4 4 4 4
Substituting the expressions for μ in Eq. (5.8) and tr[(H (m|n) )2 ] in Eq. (5.9) to Eq. (5.3b), and
subsequently using (5.6), it can be shown that
s4 − t 2 s2 − t 2
σ2 = R 1 + R3 . (5.10)
4s 4 4s 4
Finally, by substituting the values of R1 (5.7b) and R3 (5.7d) to Eq. (5.10), we can express σ as
a function of m, n and N given by
 4 
s − t2  2  2  s2 − t 2  2  2  1/2
σ= N N − 1 N + 11 + N N − 1 N − 4 . (5.11)
360s 4 45s 4
Thus we are able to find out the functional forms of the parameters μ and σ for the case of
SU(m|n) HS spin chain. It may be observed that, in the special case n = 0 (for which one gets
s = t = m), Eqs. (5.8) and (5.11) reproduce the forms of μ and σ corresponding to the SU(m)
bosonic HS spin chain [9].
Now we can compare the cumulative level density F (E) (5.1) with the error function G(E)
(5.2), where the values of μ and σ are obtained from Eqs. (5.8) and (5.11) respectively for any
given m, n and N . In Fig. 2, we plot such F (E) and G(E) for the particular case of SU(1|1)
spin chain with N = 15 lattice sites. From this figure it is evident that F (E) follows G(E)
to a high degree of approximation. One can also quantify the agreement between F (E) and
G(E) by calculating the corresponding mean square error (MSE), which for the above mentioned
case is given by 8.46 × 10−6 . It may be noted that, the agreement between F (E) and G(E)
improves rapidly with increasing values of N . For example, in the case of SU(1|1) model, MSE
between F (E) and G(E) decreases from 5.17 × 10−5 to 8.46 × 10−6 when the value of N is
increased from 10 to 15. Next, we consider the particular cases of SU(1|2) as well as SU(2|1)
supersymmetric spin chain with N = 15 lattice sites and also the SU(3) bosonic spin chain with
same number of lattice sites for the sake of comparison. In Fig. 3, we plot F (E) and G(E) for
SU(1|2), SU(2|1) and SU(3) HS spin chains and find that the corresponding MSEs are given by
2.38 × 10−5 , 1.2 × 10−5 and 8.61 × 10−6 respectively. Again F (E) shows very good agreement
with G(E) for all of these cases. Such agreement also improves rapidly with increasing values
of N . For example, in the case of SU(1|2) spin chain, MSE between F (E) and G(E) decreases
from 8.23 × 10−5 to 2.38 × 10−5 when the value of N is increased from 10 to 15. Analysing
many other particular cases with different values of m, n and sufficiently large values of N , we
find that F (E) follows G(E) with a high degree of approximation for all of these cases.
296 B. Basu-Mallick, N. Bondyopadhaya / Nuclear Physics B 757 [FS] (2006) 280–302

Fig. 2. Continuous curve represents the error function G(E) and crosses represent the cumulative distribution function
F (E) (at its discontinuity points) for SU(1|1) spin chain with N = 15.

Fig. 3. Left continuous curve represents G(E) for SU(1|2) spin chain with N = 15 and the corresponding F (E) is given
by crosses. Middle continuous curve represents G(E) for SU(2|1) spin chain with N = 15 and the corresponding F (E)
is given by dots. Right continuous curve represents G(E) for SU(3) spin chain with N = 15 and the corresponding F (E)
is given by triangles.
B. Basu-Mallick, N. Bondyopadhaya / Nuclear Physics B 757 [FS] (2006) 280–302 297

From the above discussion it is evident that the local fluctuations in energy level distribution,
as shown in Fig. 1 for the particular case of SU(1|1) spin chain, get cancelled very rapidly
whenever we take the cumulative sum of such distribution. Furthermore, for sufficiently large
values of N , continuous part of the level density distribution in the spectrum of supersymmetric
HS spin chain satisfies the Gaussian approximation at the same high level of accuracy as in the
pure bosonic case. It may be noted that, the level density of embedded Gaussian orthogonal
ensemble (GOE) also follows Gaussian distribution at the limit N → ∞, provided the number
of one-particle states tends to infinity faster than N [21]. However, in our case of SU(m|n) HS
spin chain, the number of one-particle states (i.e., m + n) remains fixed for all values of N .
Next we want to study the NNS distribution in the spectrum of SU(m|n) HS spin chain. To
eliminate the effect of level density variation in the calculation of NNS distribution for the full
energy range, it is necessary to apply an unfolding mapping to the ‘raw’ spectrum [22]. This
unfolding mapping may be defined by using the continuous part of the cumulative level density
distribution. We have already seen that, for the case of SU(m|n) HS spin chain, the continuous
part of cumulative level density is given by G(E) (5.2) with a high degree of approximation. So
we transform each energy Ei , i = 1, . . . , l, into an unfolded energy ξi ≡ G(Ei ). The function
p(u) is defined as the density of the normalized spacings ui = (ξi+1 − ξi )/Δ, where Δ = (ξl −
ξ1 )/(l − 1) is the mean spacing of the unfolded energy. To get rid of local fluctuations
u occurring
in p(u), again we study the cumulative NNS distribution given by P (u) = 0 p(x) dx, instead
of p(u). In this context it may be noted that, NNS distributions corresponding to the cases of
classical GOE as well as embedded GOE obey the Wigner’s law [23]:
π  
p(u) = u exp −πu2 /4 .
2
On the other hand, from the conjecture of Berry and Tabor one may expect that the NNS distrib-
ution for an integrable model will obey Poisson’s law given by p(u) = exp(−u) [24]. However,
it has been found that the NNS distribution for SU(m) bosonic HS spin chain does not follow
either Wigner’s law or Poisson’s law within a wide range of N [9]. Instead, the cumulative NNS
distribution for this bosonic spin chain can be well approximated by a function like
 
P̃ (u) = v α 1 − γ (1 − v)β , (5.12)
where v = u/umax with umax being the largest normalized spacing, α and β are two free para-
meters taking values within the range 0 < α, β < 1, and the value of γ is fixed by requiring that
the average normalized spacing be equal to 1. In our study we also find that, NNS distribution
in the spectrum of SU(m|n) HS spin chain does not follow either Wigner’s law or Poisson’s law
within a wide range of N . In particular, it is observed that the slope of cumulative NNS dis-
tribution diverges for both u → 0 and u → umax , which cannot be explained from Wigner’s or
Poisson’s distribution. Furthermore, we find that the cumulative NNS distribution for SU(m|n)
HS spin chain can be fitted well by P̃ (u) in Eq. (5.12) within a range of N . For example, in the
particular case of SU(1|1) spin chain, it is checked that P (u) agrees well with P̃ (u) within the
range N  20. In Fig. 4, we plot such P (u) and P̃ (u) for N = 17 lattice sites (umax = 2.626
in this case) and found a good agreement with MSE = 0.0234 when the values of free parame-
ters are taken as α = 0.39 and β = 0.29. However, it is possible that the appearance of such
non-Poissonian NNS distribution in the spectrum of SU(m|n) HS spin chain is an artifact of
finite-size effect, which requires further investigation.
Finally we want to make a comment about the behaviour of parameters μ in Eq. (5.8) and σ
in Eq. (5.11) under the exchange of bosonic and fermionic spin degrees of freedom. Since s → s
298 B. Basu-Mallick, N. Bondyopadhaya / Nuclear Physics B 757 [FS] (2006) 280–302

Fig. 4. Dotted curve represents the cumulative NNS distribution P (u) for SU(1|1) spin chain with N = 17 and the
continuous curve represents the approximate distribution function P̃ (u).

and t → −t under this exchange, we find that σ remains invariant and μ changes to μ̄ given by
μ̄ = N (N6 −1) − μ. It is interesting to note that this relation between μ and μ̄ can also be obtained
2

by applying Eq. (4.12):




1  (m|n) 1  (n|m) N (N 2 − 1) N (N 2 − 1)
μ= N D (Ei )Ei = N D − E i Ei = − μ̄.
s s 6 6
Ei Ei

This agreement clearly gives a support to our conjecture (4.11). By using this conjecture we
have found in Section 4 that, the spectrum of SU(n|m) spin chain can be obtained from that of
SU(m|n) spin chain through an inversion and overall shift of all energy levels. Since none of
these operations change the standard deviation of level density distribution, σ should take the
same value for SU(m|n) and SU(n|m) HS spin chain. Hence, the observation that σ in Eq. (5.11)
remains invariant under the exchange of bosonic and fermionic spin degrees of freedom, is also
consistent with our conjecture (4.11).

6. Conclusion

Here we derive an exact expression for the partition function of SU(m|n) supersymmetric HS
spin chain by using the freezing trick and also study some properties of the related spectrum. For
applying the freezing trick, we consider a spin CS model containing an anyon like representation
of the permutation algebra as spin dependent interaction. We find out the complete spectrum of
such spin CS model including the degeneracy factors of all energy eigenvalues. At the strong
coupling limit, this spin CS model reduces to the sum of spinless CS model with only dynamical
degrees of freedom and SU(m|n) supersymmetric HS spin chain. Consequently, by factoring out
the contribution due to dynamical degrees of freedom from partition function of this spin CS
B. Basu-Mallick, N. Bondyopadhaya / Nuclear Physics B 757 [FS] (2006) 280–302 299

model, we obtain the partition function of SU(m|n) supersymmetric HS spin chain. By using
this partition function, we study the motif representation for SU(m|n) HS spin chain and find
that, due to the lifting of a selection rule, some additional energy levels appear in the spectrum
in comparison with the case of SU(m) bosonic spin chain.
By using Mathematica we observe that, the partition function of SU(m|n) HS model satisfies
the duality relation (4.11) for many values of m, n and N . This observation motivates us to
conjecture that this duality relation, involving the interchange of bosonic and fermionic spin
degrees of freedom, is valid for all possible values of m, n and N . It would be interesting if
this duality relation can be proved analytically by using the motif representations and skew-
Young diagrams associated with the Y (gl(m|n) ) quantum group. Furthermore, it is known that, the
partition functions of SU(m) and SU(m|n) Polychronakos spin chains are intimately connected
with Rogers–Szegö (RS) polynomial [8,15], which appears in the theory of partitions [25]. Since,
HS spin chain share the same quantum group symmetry with Polychronakos spin chain, it might
be promising to investigate mathematical structures connected with the partition functions of
SU(m) as well as SU(m|n) HS spin chain and explore whether some new RS type polynomials
can be generated in this way.
By using the partition function of SU(m|n) HS spin chain, we study global properties of
its spectrum like level density distribution and NNS distribution. It is found that, similar to the
case of SU(m) bosonic HS spin chain, continuous part of the energy level density satisfies the
Gaussian distribution with a high degree of accuracy for sufficiently large values of N . We also
derive exact expressions for the mean value and the standard deviation which characterize such
Gaussian distribution. It would be interesting to provide an explanation for this behaviour of
energy level density distribution in the framework of random matrix theory and explore whether
the underlying quantum group symmetry of HS spin chain plays some role in this matter.

Acknowledgements

We would like to thank Palash B. Pal for some helpful discussions.

Appendix A. Evaluation of trace formulas

Here we shall derive the trace formulas (5.4a)–(5.4d) by assuming that |α1 · · · αN  (with αj ∈
[1, . . . , s]) are orthonormal set of vectors. Since the trace of identity operator is given by the
dimension of the Hilbert space, Eq. (5.4a) is really a trivial relation. Using Eq. (2.6), it can be
shown that
(m|n)
α1 · · · αi · · · αj · · · αN |P̃ij |α1 · · · αi · · · αj · · · αN  = (−1)(αi ) δαi αj ,
where (αi ) = 0 (1) when αi is a bosonic (fermionic) spin. With the help of above equation, we
derive the trace relation (5.4b) as
 (m|n) 
tr P̃ij

s
(m|n)
= α1 · · · αi · · · αj · · · αN |P̃ij |α1 · · · αi · · · αj · · · αN 
α1 ...αN =1
s
(αi ,αj ) 
s
(m|n)
= α1 · · · αi · · · αj · · · αN |P̃ij |α1 · · · αi · · · αj · · · αN 
α1 ...αN =1 αi ,αj =1
300 B. Basu-Mallick, N. Bondyopadhaya / Nuclear Physics B 757 [FS] (2006) 280–302

s
(αi ,αj ) 
s
= (−1)(αi ) = s N−2 t,
α1 ...αN =1 αi =1

where the notation


s
(αi ,αj )

α1 ...αN =1

represents summation over all spin components α1 , . . . , αN except αi and αj .


Next, by using Eq. (2.6), it is found that
(m|n) (m|n)
α1 · · · αi · · · αj · · · αl · · · αN |P̃ij P̃il |α1 · · · αi · · · αj · · · αl · · · αN  = δαi αj δαi αl .
Applying the above equation, we obtain a trace relation in Eq. (5.4c) as
 (m|n) (m|n) 
tr P̃ij P̃il

s
(m|n) (m|n)
= α1 · · · αi · · · αj · · · αl · · · αN |P̃ij P̃il |α1 · · · αi · · · αj · · · αl · · · αN 
α1 ...αN =1
s
(αi ,αj ,αl ) 
s
= 1 = s N−2 .
α1 ...αN =1 αi =1

Other trace relations in (5.4c) can be proved in a similar way.


By using Eq. (2.6), it is also found that
(m|n) (m|n)
α1 · · · αi · · · αj · · · αk · · · αl · · · αN |P̃ij P̃kl |α1 · · · αi · · · αj · · · αk · · · αl · · · αN 
= (−1)(αi )+(αk ) δαi αj δαk αl .
With the help of this equation, we obtain the trace relation (5.4d) as
 (m|n) (m|n) 
tr P̃ij P̃kl

s
(m|n) (m|n)
= α1 · · · αi · · · αj · · · αk · · · αl · · · αN |P̃ij P̃kl
α1 ...αN =1
× |α1 · · · αi · · · αj · · · αk · · · αl · · · αN 
s
(αi ,αj ,αk ,αl ) 
s
= (−1)(αi )+(αk ) = s N−4 t 2 .
α1 ...αN =1 αi ,αk =1

Appendix B. Evaluation of summation formulas

Here we briefly describe the way of calculating known summation formulas (5.7a) and (5.7b)
[9], and subsequently present our derivation for new ones like (5.7c) and (5.7d). From the work
of Calogero et al. [20], it is known that

N−1
1 N2 − 1
= , (B.1)
j =1 sin2 ( jNπ ) 3
B. Basu-Mallick, N. Bondyopadhaya / Nuclear Physics B 757 [FS] (2006) 280–302 301

and

N−1
1 (N 2 − 1)(N 2 + 11)
= . (B.2)
j =1 sin4 ( jNπ ) 45

Using the translational invariance on a circular lattice and summation relation (B.1), one can
obtain Eq. (5.7a) as

1 N 
N−1
1 1 N (N 2 − 1)
R0 = = = .
2
i=j
sin2 (ξi − ξj ) 2 2 jπ
j =1 sin ( N )
6

Similarly, by using (B.2), one obtains Eq. (5.7b) as

1 N 
N−1
1 1 N (N 2 − 1)(N 2 + 11)
R1 = = = .
2
i=j
sin4 (ξi − ξj ) 2 4 jπ
j =1 sin ( N )
90

For the purpose of calculating R2 in Eq. (5.5c), we note that R0 in Eq. (5.5a) can be expressed
as
 1 
N
1 
N
1 1
R0 = + + − .
k<l
sin (ξk − ξl )
2
r=1
sin (ξi − ξr )
2
r=1
sin (ξj − ξr )
2
sin (ξi − ξj )
2
(k,l=i,j ) (r=i) (r=j )

Substituting the value of R0 given in (5.7a) to the above relation and also using (B.1), we find
that
 1 (N 2 − 1)(N − 4) 1
= + 2 .
k<l
sin (ξk − ξl )
2 6 sin (ξi − ξj )
(k,l=i,j )

By substituting this expression to R2 in Eq. (5.5c), and subsequently using Eqs. (5.7a) as well as
(5.7b), we derive the value of R2 given in Eq. (5.7c) as
 2
1 (N − 1)(N − 4) 1
R2 = + 2
i<j
sin2 (ξi − ξj ) 6 sin (ξi − ξj )
N(N 2 − 1)2 (N − 4) N (N 2 − 1)(N 2 + 11)
= + .
36 90
Finally, by substituting the values of R0 (5.7a), R1 (5.7b) and R2 (5.7c) to the relation (5.6), we
easily obtain the value of R3 appearing in Eq. (5.7d).

References

[1] F.D.M. Haldane, Phys. Rev. Lett. 60 (1988) 635.


[2] B.S. Shastry, Phys. Rev. Lett. 60 (1988) 639.
[3] F.D.M. Haldane, in: A. Okiji, N. Kawakami (Eds.), Proceedings of the 16th Taniguchi Symposium, Kashikojima,
Japan, 1993, Springer-Verlag, Berlin, 1994.
[4] Z.N.C. Ha, Quantum Many-Body Systems in One Dimension, Series on Advances in Statistical Mechanics, vol. 12,
World Scientific, Singapore, 1996.
[5] A.P. Polychronakos, Generalized statistics in one dimension, in: Les Houches Lectures, 1998, hep-th/9902157.
[6] F.D.M. Haldane, Z.N.C. Ha, J.C. Talstra, D. Benard, V. Pasquier, Phys. Rev. Lett. 69 (1992) 2021.
302 B. Basu-Mallick, N. Bondyopadhaya / Nuclear Physics B 757 [FS] (2006) 280–302

[7] D. Benard, M. Gaudin, F.D.M. Haldane, V. Pasquier, J. Phys. A 26 (1993) 5219.


[8] K. Hikami, Nucl. Phys. B 441 (1995) 530.
[9] F. Finkel, A. González-López, Phys. Rev. B 72 (2005) 174411.
[10] A.P. Polychronakos, Phys. Rev. Lett. 70 (1993) 2329;
A.P. Polychronakos, Nucl. Phys. B 419 (1994) 553.
[11] B. Sutherland, B.S. Shastry, Phys. Rev. Lett. 71 (1993) 5.
[12] A.P. Polychronakos, Physics and mathematics of Calogero particles, hep-th/0607033.
[13] P. Schlottmann, Int. J. Mod. Phys. B 11 (1997) 355.
[14] B. Basu-Mallick, H. Ujino, M. Wadati, J. Phys. Soc. Jpn. 68 (1999) 3219.
[15] K. Hikami, B. Basu-Mallick, Nucl. Phys. B 566 (2000) 511.
[16] B. Basu-Mallick, Nucl. Phys. B 540 (1999) 679.
[17] B. Basu-Mallick, Nucl. Phys. B 482 (1996) 713.
[18] C. Ahn, W.M. Koo, Phys. Lett. B 365 (1996) 105.
[19] B. Sutherland, Phys. Rev. A 5 (1972) 1372.
[20] F. Calogero, A.M. Perelomov, Commun. Math. Phys. 59 (1978) 109.
[21] K.K. Mon, J.B. French, Ann. Phys. 95 (1975) 90.
[22] F. Haake, Quantum Signatures of Chaos, Springer-Verlag, Berlin, 2001.
[23] V.K.B. Kota, Phys. Rep. 347 (2001) 223.
[24] M.V. Berry, M. Tabor, Proc. R. Soc. London A 356 (1977) 375.
[25] G.E. Andrews, The Theory of Partitions, Addison–Wesley, Reading, MA, 1976.
Nuclear Physics B 757 [FS] (2006) 303–343

Logarithmic extensions of minimal models:


Characters and modular transformations
B.L. Feigin a , A.M. Gainutdinov b , A.M. Semikhatov b,∗ , I.Yu. Tipunin b
a Landau Institute for Theoretical Physics, Russian Federation
b Lebedev Physics Institute, 117924 Moscow, Russian Federation

Received 6 July 2006; accepted 25 September 2006


Available online 20 October 2006

Abstract
We study logarithmic conformal field models that extend the (p, q) Virasoro minimal models. For co-
prime positive integers p and q, the model is defined as the kernel of the two minimal-model screening
operators. We identify the field content, construct the W -algebra Wp,q that is the model symmetry (the
maximal local algebra in the kernel), describe its irreducible modules, and find their characters. We then
derive the SL(2, Z)-representation on the space of torus amplitudes and study its properties. From the ac-
tion of the screenings, we also identify the quantum group that is Kazhdan–Lusztig-dual to the logarithmic
model.
© 2006 Elsevier B.V. All rights reserved.

1. Introduction

1.1. Logarithmic conformal field theories can be interesting for two reasons at least. The first
is their possible applications in condensed-matter systems: the quantum Hall effect [1–3], self-
organized critical phenomena [4,5], the two-dimensional percolation problem [6–8], and others
(see, e.g., [9] and the references therein). The second is the general category-theory aspects
of conformal field theory involving vertex-operator algebras with nonsemisimple representation
categories [10,11] (also see [12] and the references therein). Unfortunately, there are few well-
investigated examples in logarithmic conformal field theory. Presently, only the (2, 1) model

* Corresponding author.
E-mail addresses: feigin@mccme.ru (B.L. Feigin), azot@mccme.ru (A.M. Gainutdinov), ams@sci.lebedev.ru
(A.M. Semikhatov), tipunin@td.lpi.ru (I.Yu. Tipunin).

0550-3213/$ – see front matter © 2006 Elsevier B.V. All rights reserved.
doi:10.1016/j.nuclphysb.2006.09.019
304 B.L. Feigin et al. / Nuclear Physics B 757 [FS] (2006) 303–343

[13–18] is formulated with sufficient completeness. A number of results exist on the (p, 1) mod-
els [19–22], but no model has been investigated so completely as the Ising model, for example.
An essential new feature of nonsemisimple (logarithmic) conformal field theories, in compar-
ison with semisimple (rational) theories, already occurs in constructing the space of states. In
the semisimple case, it suffices to take the sum of all irreducible representations in each chiral
sector. But in the nonsemisimple case, there are various indecomposable representations, con-
structed beginning with first extensions of irreducible representations, then taking their further
extensions, and so on, ending up with projective modules (the largest possible indecomposable
extensions, and hence the modules with the largest Jordan cells, for the scaling dimension opera-
tor L0 , that can be constructed for a given set of irreducible representations). The space of states
is therefore given by the sum over all nonisomorphic indecomposable projective modules,

P= Pι . (1.1)
ι
This affects not only theory but also applications: the physically important operators (thermal,
magnetic, and so on) in specific models may often be identified with the field corresponding
to the “highest-weight” vector in a projective module Pι , not with the primary field (of the
same dimension) corresponding to the highest-weight vector of the irreducible submodule of
Pι . Also, whenever indecomposable representations are involved, there are more possibilities for
constructing modular invariants by combining the chiral and antichiral spaces of states.
But an even more essential point about nonsemisimple theories is that before speaking of
the representations, their characters, fusion, etc., one must find the symmetry algebra whose
representations, characters, fusion, etc., are to be considered; this algebra is typically not the
“naive”, manifest symmetry algebra. This point was expounded in [21]. The symmetry algebras
of logarithmic conformal field theory models are typically nonlinear extensions of the naive
symmetry algebra (e.g., Virasoro), i.e., are some W -algebras. The first examples of W -algebras
arising in this context were studied in [13–15], also see [23] and the references therein.
In a given logarithmic model, the chiral W -algebra must be big enough to ensure that only
a finite number of its irreducible representations are realized in the model. Only then can one
expect to have finite-dimensional fusion rules and modular group representation (we recall that a
finite-dimensional modular group representation is a good test of the consistency of the model).
Also, once finitely many irreducible representations are involved, there are finitely many projec-
tive modules, and the sum in (1.1) makes sense.
In this paper, we systematically investigate the logarithmic extensions of the (p, q) models of
conformal field theory. This task was already set in [22], in the line of an appropriate extension
of the results obtained for the (p, 1) models. The method in [21,22] is to define and construct
logarithmic conformal field models in terms of free fields and screenings. The chiral algebra
W that is the symmetry algebra of the model is then derived from the kernel of the screenings
in the vacuum representation of a lattice vertex-operator algebra L. Irreducible representations
of W are identified with the images and cohomology of (certain powers of) the screenings in
irreducible representations of L. The projective W -modules in (1.1) are then to be constructed as
the projective covers of the irreducible representations (this construction may be rather involved;
a notable exception is provided by the (2, 1) model, see [16–18]).
In what follows, we set p = p+ and q = p− , a fixed pair of coprime positive integers.
In the rational (p+ , p− ) model, the chiral symmetry algebra is the vertex-operator algebra
Mp+ ,p− defined as the cohomology of screenings that act in the vacuum representation Lp+ ,p−
of the appropriate lattice vertex-operator algebra Lp+ ,p− ; irreducible representations of Mp+ ,p−
B.L. Feigin et al. / Nuclear Physics B 757 [FS] (2006) 303–343 305

can then be identified with the cohomology of (powers of) the screenings in irreducible repre-
sentations of Lp+ ,p− .
In the logarithmic (p+ , p− ) model (the vacuum representation of) the chiral algebra Wp+ ,p−
extends Mp+ ,p− such that Mp+ ,p− = Wp+ ,p− /R, where R is the maximal vertex-operator
ideal in Wp+ ,p− . The W -algebra Wp+ ,p− can be defined as the intersection of the screen-
ing kernels in Lp+ ,p− . (As we see in what follows, Wp+ ,p− is generated by the energy–
momentum tensor T (z) and two Virasoro primaries W + (z) and W − (z) of conformal dimension
(2p+ −1)(2p− − 1).) The irreducible representations of Wp+ ,p− are of two different kinds. The
first are the 12 (p+ − 1)(p− − 1) irreducible modules of the Virasoro minimal model or, in other
words, the modules annihilated by R. The second are the 2p+ p− modules that admit a non-
trivial action of R. They can be identified with the images of (powers of) the screenings in the
respective irreducible representations of the lattice vertex-operator algebra Lp+ ,p− and decom-
pose into infinitely many irreducible Virasoro modules. (The Virasoro embedding structure in
the (3, 2) and (5, 2) logarithmic models was recently arduously explored in [24].)
The characters of the 12 (p+ − 1)(p− − 1) + 2p+ p− irreducible representations of Wp+ ,p−
(or, in slightly different terms, the Grothendieck ring1 G) do not exhaust the space C of torus
amplitudes of the logarithmic (p+ , p− ) model: we only have that C ⊃ G. This is a characteristic
feature of nonsemisimple (logarithmic) conformal field theories, cf. [21,22,26–28]. Because C
carries a representation of SL(2, Z), a much better approximation to this space is C̄, the SL(2, Z)-
representation generated from G:
C ⊃ C̄ ⊃ G
(where most probably C = C̄).

1.1.1. Theorem. In the logarithmic (p+ , p− ) model,

(1) dim G = 2p+ p− + 12 (p+ − 1)(p− − 1);


(2) dim C̄ = 12 (3p+ − 1)(3p− − 1);
(3) the SL(2, Z)-representation π on C̄ has the structure
C̄ = Rmin ⊕ Rproj ⊕ C2 ⊗ (R ⊕ R ) ⊕ C3 ⊗ Rmin , (1.2)
where C2
is the standard two-dimensional representation, ∼ 2 2 C3
= S (C ) is its symmetrized
square, Rmin is the 12 (p+ − 1)(p− − 1)-dimensional SL(2, Z)-representation on the char-
acters of the (p+ , p− ) Virasoro minimal model, and Rproj , R , and R are SL(2, Z)-
representations of the respective dimensions 12 (p+ + 1)(p− + 1), 12 (p+ − 1)(p− + 1), and
2 (p+ + 1)(p− − 1).
1

We note that the space Rmin ⊕ Rproj is spanned by the characters of irreducible modules of
the lattice vertex-operator algebra Lp+ ,p− and coincides with the space of theta functions.2 The
space Rproj can be identified with the characters of projective Wp+ ,p− -modules.

1 The free Abelian group generated by symbols [M], where M ranges over all representations subject to relations
[M] = [M  ] + [M  ] for all exact sequences 0 → M  → M → M  → 0.
2 Technically, the extension from R
min ⊕ Rproj to C̄ involves derivatives of the theta functions, which gives rise to the
explicit occurrences of the modular parameter τ equalizing the modular weights. The highest theta-function derivative
and simultaneously the top power of τ thus occurring, n, is equal to 2 for (p+ , p− ) models and to 1 for (p, 1) models
[22], which can be considered the origin of the corresponding Cn+1 in (1.2) and in a similar formula in [22].
306 B.L. Feigin et al. / Nuclear Physics B 757 [FS] (2006) 303–343

1.1.2. Remark. The 2p+ p− + 12 (p+ − 1)(p− − 1) irreducible Wp+ ,p− -representations are rather
naturally arranged into a Kac table as follows. First, the 12 (p+ − 1)(p− − 1) rational-model rep-
resentations occupy the standard positions, with the standard identifications, in the boxes of the
standard (p+ − 1) × (p− − 1) Kac table. Next, each box of the extended p+ × p− Kac table
contains two, a “plus” and a “minus”, of the Wp+ ,p− -representations labeled (Xr,s ±)
1rp+ in
1sp−
what follows. We note that a logarithmic conformal field theory containing only finitely many
irreducible or indecomposable Virasoro representations does not seem to exist; in particular, each
(p+ , p− ) model considered in this paper contains infinitely many indecomposable Virasoro rep-
resentations, usually with multiplicity greater than 1, and we use the term “Kac table” exclusively
to refer to a finite set of irreducible representations of the Wp+ ,p− algebra.

Theorem 1.1.1 also implies that there exist SL(2, Z)-representations π̄ and π ∗ on C̄ such that

π(γ ) = π ∗ (γ )π̄ (γ ), π̄ (γ )π ∗ (γ  ) = π ∗ (γ  )π̄ (γ ), γ , γ  ∈ SL(2, Z).


The representation π̄ can be restricted to G, which then decomposes in terms of SL(2, Z) repre-
sentations as

G = Rmin ⊕ Rproj ⊕ R ⊕ R ⊕ Rmin .


This decomposition can be taken as the starting point for constructing a logarithmic Verlinde
formula as in [21].
In view of the fundamental importance of the SL(2, Z) action, this theorem gives a very strong
indication regarding the field content of a consistent conformal field theory model: the 12 (3p+ −
1)(3p− − 1)-dimensional space of generalized characters C̄ is a strong candidate for the space
of torus amplitudes of the logarithmic (p+ , p− ) model. The following conjecture appears to be
highly probable.

1.1.3. Conjecture. The SL(2, Z)-representation generated from G coincides with the space of
torus amplitudes:

C̄ = C
as SL(2, Z) representations.

1.2. To move further in constructing the full space of states of the logarithmic theory as in
(1.1), one must construct projective covers of all the 2p+ p− + 12 (p+ − 1)(p− − 1) irreducible
Wp+ ,p− -representations. This is a separate, quite interesting task.3
Some useful information on the structure of the chiral-algebra projective modules can be
obtained from the Kazhdan–Lusztig correspondence. In general, it is a correspondence between
a chiral algebra W and its representation category W realized in conformal field theory, on the
one hand, and some “dual” quantum group and its representation category on the other hand.
In some “well-behaved” cases, the occurrence of the Kazhdan–Lusztig-dual quantum group can

3 In particular, the question about logarithmic partners of the energy–momentum tensor T (z) and other fields takes the
form of a well-posed mathematical problem about the structure of projective Wp+ ,p− -modules. Logarithmic partners
have been discussed, e.g., in [9,29–32] in the case c = 0, where the differential polynomial in T (z) whose logarithmic
partner is sought coincides with T (z) itself.
B.L. Feigin et al. / Nuclear Physics B 757 [FS] (2006) 303–343 307


be seen by taking ι dι Pι , a direct sum of projective W -modules with some multiplicities dι
chosen such that they are additive with respect to the direct sum4 and multiplicative with respect
to the quasitensor product in W. If such a choice of the dι is possible, then
 
g = EndW dι Pι (1.3)
ι
can be endowed with the structure of a (Hopf) algebra and dι are the dimensions of its irre-
ducible representations (constructing the comultiplication requires certain conditions, which we
do not discuss here). In this case, the category W is equivalent (as a quasitensor category) to
the category of finite-dimensional representations of g. Such an extremely well-behaved case
is realized in (p, 1) logarithmic models [18]: classification of indecomposable representation of
the Kazhdan–Lusztig-dual quantum group g, which is not difficult to obtain using quite stan-
dard means, gives the classification of indecomposable Wp,1 -representations. In particular, the
structure of projective Wp,1 -modules is thus known.
In the (p+ , p− ) logarithmic models, the Kazhdan–Lusztig-dual quantum group gp+ ,p− (ob-
tained as a subalgebra in the quotient of the Drinfeld double of the algebra of screenings for
the Wp+ ,p− algebra) is not Morita-equivalent to Wp+ ,p− , but nevertheless provides impor-
tant information on the structure of indecomposable Wp+ ,p− modules. On the one hand, the
quantum group gp+ ,p− and its representation category give only an “approximation” to the
structure of the Wp+ ,p− -representation category (the representation categories are certainly not
equivalent, in contrast to the (p, 1) case [18]; in particular, there are 2p+ p− indecompos-
able projective gp+ ,p− -modules but 2p+ p− + 12 (p+ − 1)(p− − 1) indecomposable projective
Wp+ ,p− -modules). On the other hand, this “approximation” becomes the precise correspondence
as regards the modular group representations: naturally associated with gp+ ,p− is the SL(2, Z)-
representation on its center [34], which turns out to be equivalent to the SL(2, Z)-representation
on the Wp+ ,p− -algebra characters and generalized characters in (1.2).
The gp+ ,p− quantum group acts in the space F obtained as a certain extension (“dressing”)
of the irreducible Lp+ ,p− -modules. Moreover, F is in fact a (Wp+ ,p− , gp+ ,p− )-bimodule. This
bimodule structure plays an essential role in the description of the full conformal field theory on
Riemann surfaces of different genera, defect lines, boundary conditions, etc.

1.3. From the standpoint of applications in condensed-mater physics, the definition of both
the W -algebra Wp+ ,p− and the quantum group gp+ ,p− refers to the Coulomb-gas picture [35,36],
where the starting point is a two-dimensional scalar field ϕ with the action written in complex
coordinates as

1 ¯ dz d z̄.
S0 = − ∂ϕ ∂ϕ

(The normalization is chosen such that the propagator has the form
ϕ(z, z̄)ϕ(0, 0) = log |z| and
vertex operators exp(αϕ(z, z̄)) do not involve an i in the exponent.) Furthermore, the field is

4 For a given chiral algebra W , such sums are assumed to be finite; from a somewhat more general standpoint, this
must follow from a set of fundamental requirements on W in a given nonsemisimple model. First, the algebra itself
must be generated by a finite number of fields Wi (z). The category W of W -modules with locally nilpotent action of
the positive modes of Wi (z) is then well defined. Second, the W -modules must be finitely generated, such that for any
collection of positive integers (Ni ), the coinvariants with respect to the subalgebra W (N1 , . . . , Nm ) ⊂ W generated by
the modes Wi [ni ] with ni  Ni be finite-dimensional in any module from W. In particular, this means that the category
W contains a finite number of irreducible representations, and hence a finite number of projective modules (cf. [27,33]).
308 B.L. Feigin et al. / Nuclear Physics B 757 [FS] (2006) 303–343


taken to be compactified
 to a circle (of the radius 2p+ p− ), which just means that the fields
ϕ and ϕ + 2iπ 2p+ p− are considered equivalent. Then the observables are given by vertex
operators exp( √ n ϕ(z, z̄)) with n ∈ Z.
2p+ p−
This model with central charge c = 1 has a large symmetry algebra, the lattice vertex-operator
algebra Lp+ ,p− mentioned above. The minimal models can be regarded as conformal points with
c < 1 [37] to which the system renormalizes after the perturbation
 
S = S0 + λ+ e α+ ϕ(z,z̄)
dz d z̄ + λ− eα− ϕ(z,z̄) dz d z̄

with the appropriate α+ and α− (and some constant λ± ). The symmetry of the model thus ob-
tained is the vertex-operator algebra Mp+ ,p− , which is “much smaller” than Lp+ ,p− .
Logarithmic conformal points occur in this approach for quenched random systems [38]
whose action is given by
 
S = S0 + λ+ (z, z̄)eα+ ϕ(z,z̄) dz d z̄ + λ− (z, z̄)eα− ϕ(z,z̄) dz d z̄,

where λ± (z, z̄) are quenched random variables with appropriately chosen correlators λ± (z, z̄),
λ± (z1 , z̄1 )λ± (z2 , z̄2 ), and so on. The parameters involved in these fixed correlators renormalize
such that the system occurs in a new infrared fixed point with the same c < 1 as in the minimal
model, but with the symmetry algebra given by a W -algebra (a subalgebra of Lp+ ,p− ). The
entire system can thus be regarded as the tensor product of the original Coulomb-gas model
and an additional model describing a quenched disorder through the chosen correlators of the
λ± (z, z̄). We do not know how these correlators must be chosen in order to produce just the
(p+ , p− ) logarithmic conformal field theory model; instead, we take a more algebraic stand and
study the W -algebra Wp+ ,p− of the expected fixed point. It is then possible to make contact with
quenched disorder by studying the projective modules of this W -algebra.
This paper is organized as follows. In Section 2, we introduce the basic notation and de-
scribe some facts in the free-field description of minimal models. We introduce vertex operators
in Section 2.1, define the free-field and Virasoro modules in Section 2.2, and introduce the
lattice vertex-operator algebra Lp+ ,p− and its modules that are important ingredients in con-
structing the Wp+ ,p− -representations in Section 2.3. In Section 3, we describe the action of
screenings on the vertices in Section 3.1 and introduce the Kazhdan–Lusztig-dual quantum
group gp+ ,p− : we reformulate the action of the screenings in terms of a Hopf algebra H in
Section 3.2, and then construct its Drinfeld double gp+ ,p− in Section 3.3. In Section 4, we finally
construct the Wp+ ,p− algebra (we introduce it in Section 4.1, formulate the structural result in
Section 4.2 and describe the irreducible and Verma Wp+ ,p− -modules in Sections 4.3 and 4.4;
a (Wp+ ,p− , gp+ ,p− )-bimodule structure of the space of states in given Section 4.5). In Section 5,
we calculate the SL(2, Z)-representation generated from the Wp+ ,p− -characters; on the result-
ing space of generalized characters (as noted above, most probably the torus amplitudes), we
then decompose the SL(2, Z)-action as in (1.2). In Section 5.1, we calculate characters of the
irreducible Wp+ ,p− -modules; in Section 5.2, we introduce generalized characters, calculate the
SL(2, Z) action on C̄, and give a direct proof of Theorem 1.1.1. Several series of modular invari-
ants are considered in Section 5.3. Some implications of the Kazhdan–Lusztig correspondence
and several open problems are mentioned in the conclusions.
B.L. Feigin et al. / Nuclear Physics B 757 [FS] (2006) 303–343 309

Notation

We fix two coprime positive integers p+ and p− and set


 
2p+ 2p−
α− = − , α+ = , α0 = α+ + α− . (1.4)
p− p+
With the pair (p+ , p− ), we associate the sets of indices


I0 = (r, s)
0  r  p+ , 0  s  p− , p− r + p+ s  p+ p− , (r, s) = (0, p− ) , (1.5)


I1 = (r, s)
1  r  p+ − 1, 1  s  p− − 1, p− r + p+ s  p+ p− , (1.6)


I = (r, s)
0  r  p+ − 1, 1  s  p− − 1, p− r + p+ s  p+ p− , (1.7)


I = (r, s)
1  r  p+ − 1, 0  s  p− − 1, p− r + p+ s  p+ p− . (1.8)
The numbers of elements in these sets are |I0 | = 12 (p+ + 1)(p− + 1), |I1 | = 12 (p+ − 1)(p− − 1),
|I | = 12 (p+ + 1)(p− − 1), and |I | = 12 (p+ − 1)(p− + 1).
(1)
The building blocks for the characters are the so-called theta-constants θs,p (q), θs,p (q), and
(2)
θs,p (q), where
 

∂ m

θs,p (q) = θs,p (q, 1), θs,p (q) = z


(m)
θs,p (q, z)

, m ∈ N, (1.9)
∂z z=1
and the theta function is defined as
2
θs,p (q, z) = q pj zpj , |q| < 1, z ∈ C, p ∈ N, s ∈ Z.
s
j ∈Z+ 2p

To further simplify the notation, we resort to the standard abuse by writing f (τ ) for f (e2iπτ ),
with τ in the upper complex half-plane; it is tacitly assumed that q = e2iπτ .
There are the easily verified properties
(m) (m)
θs+2pa,p (τ ) = θs,p
(m)
(τ ), θ−s,p (τ ) = (−1)m θs,p
(m)
(τ ), p ∈ N, m ∈ N0 , a ∈ Z
 (τ ) = θ  (τ ) = 0. We often write
and θ0,p p,p

θs,p+ p− (τ ) ≡ θs , (1)
θs,p + p−
(τ ) ≡ θs , (2)
θs,p + p−
(τ ) ≡ θs .
Similar abbreviations are used for the characters: we write
± ±
χr,s (τ ) ≡ χr,s , χr,s (τ ) ≡ χr,s .
Finally, we use the η-function


1  
η(q) = q 24 1 − qn .
n=1

2. Free-field preliminaries

We introduce a free bosonic field and describe its energy–momentum tensor, vertex operators,
screenings, and representations of the Virasoro algebra.
310 B.L. Feigin et al. / Nuclear Physics B 757 [FS] (2006) 303–343

2.1. Free field and vertices

Let ϕ denote a free scalar field with the OPE


1
∂ϕ(z)∂ϕ(w) = (2.1)
(z − w)2
and the mode expansion

∂ϕ(z) = ϕn z−n−1 . (2.2)
n∈Z

The energy–momentum tensor is given by

1 α0
T (z) = ∂ϕ(z)∂ϕ(z) + ∂ 2 ϕ(z) (2.3)
2 2
(see (1.4)). The modes of ∂ϕ(z) span the Heisenberg algebra and the modes of T (z) span the
Virasoro algebra Virp+ /p− with the central charge

(p+ − p− )2
c=1−6 . (2.4)
p+ p−

The vertex operators are given by ej (r,s)ϕ(z) with j (r, s) = 1−s


2 α− + 1−r
2 α+ , r, s ∈ Z. Equiv-
alently, these vertex operators can be parameterized as
p− (1−r)−p+ (1−s)+p+ p− n
√ ϕ(z)
Vr,s;n (z) = e 2p+ p−
, 1  r  p+ , 1  s  p− , n ∈ Z. (2.5)
The conformal dimension of Vr,s;n (z) assigned by the energy–momentum tensor is

(p+ s − p− r + p+ p− n)2 − (p+ − p− )2


Δr,s;n = . (2.6)
4p+ p−
We note that

Δ−r,−s;−n = Δr,s;n , Δr+kp+ ,s+kp− ;n = Δr,s;n , Δr,s+kp− ;n = Δr,s;n+k . (2.7)


The vertex operators satisfy the braiding relations

Vr,s;n (z1 )Vr  ,s  ;n (z2 )


  
= q(p− (1−r)−p+ (1−s)+p+ p− n)(p− (1−r )−p+ (1−s )+p+ p− n ) Vr  ,s  ;n (z2 )Vr,s;n (z1 ), (2.8)
where

q = e 2p+ p− . (2.9)
(The convention is to take the OPE of the operators in the left-hand side with |z1 | > |z2 | and
make an analytic continuation to |z1 | < |z2 | by moving z1 along a contour passing below z2 , i.e.,
as z1 − z2 → eiπ (z1 − z2 ).)
In what follows, we also write Vr,s;0 (z) ≡ Vr,s (z) and Δr,s;0 ≡ Δr,s .
B.L. Feigin et al. / Nuclear Physics B 757 [FS] (2006) 303–343 311

2.2. Definition of modules

For 1  r  p+ , 1  s  p− , and n ∈ Z, let Fr,s;n denote the Fock module of the Heisenberg
algebra generated
 from (the state corresponding to) the vertex operator Vr,s;n (z). The zero mode
ϕ0 = 2iπ
1
dz ∂ϕ(z) acts in Fr,s;n by multiplication with the number

p− (1 − r) − p+ (1 − s) + p+ p− n
ϕ0 v = √ v, v ∈ Fr,s;n .
2p+ p−

We write Fr,s ≡ Fr,s;0 . For convenience of notation, we identify F0,s;n ≡ Fp+ ,s;n+1 and Fr,0;n ≡
Fr,p− ;n−1 .
Let Yr,s;n with 1  r  p+ , 1  s  p− , and n ∈ Z denote the Virasoro module that coincides
with Fr,s;n as a linear space, with the Virasoro algebra action given by (2.3) (see [39]). As with
the Fr,s;n , we also write Yr,s ≡ Yr,s;0 .

2.2.1. Subquotient structure of the modules Yr,s;n


We recall the subquotient structure of the Virasoro modules Yr,s;n [40]. We let Jr,s;n denote
the irreducible Virasoro module with the highest weight Δr,s;n (as before, 1  r  p+ , 1  s 
p− , and n ∈ Z). Evidently, Jr,s;n  Jp+ −r,p− −s;−n . The 12 (p+ − 1)(p− − 1) nonisomorphic
modules among the Jr,s;0 with 1  r  p+ − 1 and 1  s  p− − 1 are the irreducible modules
from the Virasoro (p+ , p− ) minimal model. We also write Jr,s ≡ Jr,s;0 . For convenience of
notation, we identify J0,s;n ≡ Jp+ ,s;n+1 and Jr,0;n ≡ Jr,p− ;n−1 .
The well-known structure of Yr,s for 1  r  p+ − 1 and 1  s  p− − 1 is recalled in Fig. 1.

Fig. 1. Embedding structure of the Feigin–Fuchs module Yr,s . The notation is as follows. The cross × corresponds to the
subquotient Jr,s , the filled dots • to Jr,p− −s;2n+1 with n ∈ N0 , the triangles  to Jp+ −r,s;2n+1 with n ∈ N0 , the open
dots ◦ to Jr,s;2n with n ∈ N, and the squares  to Jp+ −r,p− −s;2n with n ∈ N. The arrows correspond to embeddings of
the Virasoro modules and are directed toward submodules. The notation [a, b; n] in square brackets is for subquotients
isomorphic to Ja,b;n . The filled dots constitute the socle of Yr,s .
312 B.L. Feigin et al. / Nuclear Physics B 757 [FS] (2006) 303–343

2.2.2. The Fock spaces introduced above constitute a free-field module



p+ 
p−
F= Fr,s;n . (2.10)
n∈Z r=1 s=1

It can be regarded as (the chiral sector


√ of) the space of states of the Gaussian Coulomb gas model
compactified on the circle of radius 2p+ p− . This model has c = 1 and its symmetry is a lattice
vertex operator algebra, which is described in the next subsection.

2.3. The lattice vertex-operator algebra

Let Lp+ ,p− be the lattice vertex-operator algebra (see [41–43]) generated by the vertex oper-
ators

V1,1;2n (z) = en 2p+ p− ϕ(z)
, n ∈ Z.
The underlying vector space (the vacuum representation) of Lp+ ,p− is

Lp+ ,p− = F1,1;2n . (2.11)
n∈Z

The vertex-operator algebra Lp+ ,p− has 2p+ p− irreducible modules, denoted in what follows
± with 1  r  p and 1  s  p . Their Fock-module decompositions are given by
as Vr,s + −
 
+ −
Vr,s = Fp+ −r,p− −s;2n , Vr,s = Fp+ −r,p− −s;2n+1 ,
n∈Z n∈Z
1  r  p+ , 1  s  p− . (2.12)
+
In this numerology, the vacuum representation Lp+ ,p− coincides not with the usual V1,1
but with
+
Vp+ −1,p− −1 ; such a “twist of notation” turns out to be convenient in what follows.
The model with the space of states (2.10) is rational with respect to Lp+ ,p− , i.e., the space F
is a finite sum of irreducible Lp+ ,p− -modules:

 
p+ p−

+ −
F= Vr,s ⊕ Vr,s .
r=1 s=1

Evidently, each Vr,s ± is a Virasoro module by virtue of the free-field construction (2.3). We

need to recall a number of standard facts known in the representation theory of the Virasoro
algebra. With the same notation as in Fig. 1, we describe the Virasoro structure of the Vr,s ± in
+ − + −
Fig. 2 (Vp+ −r,p− −s ), Fig. 3 (Vr,p− −s ), Fig. 4 (Vr,p− ), and Fig. 5 (Vp+ −r,p− ).
At the moment, the reader is asked to ignore the dotted lines in the figures (they are to be de-
scribed in Section 4). The action of Lp+ ,p− in Vr,s ± is as follows. In Figs. 2 and 3, the Heisenberg

subalgebra acts in each two-strand “braid” (labeled by the last integer in the square brackets at
the top of each braid, even in Fig. 2 and odd in Fig. 3) as in the corresponding Fock module.
Vertex operators V1,1;2n (z) map between the braids over the distance n (to the left for n > 0). In
Figs. 4 and 5, the Heisenberg subalgebra acts in each vertical strand and V1,1;2n (z) act between
strands similarly.
We next consider the socle (the maximal semisimple submodule) of Vr,s ± (the irreducible

Lp+ ,p− -module defined in (2.12)), viewed as a Virasoro module.


B.L. Feigin et al. / Nuclear Physics B 757 [FS] (2006) 303–343 313

Fig. 2. Structure of the modules Vp++ −r,p− −s . The notation is the same as in Fig. 1. Filled dots denote Virasoro repre-
+
sentations that are combined into Xr,s . Levels of conformal dimensions are chosen for (r, s) ∈ I1 , and p− > p+ . The
+
dotted lines show the action of the W ± (z) generators of Wp+ ,p− in Kr,s , see Eq. (3.21) below.

± modules regarded as Virasoro modules, we set


2.3.1. Definition. With the Vr,s
± = soc V ±
Xr,s 1  r  p+ − 1, 1  s  p− − 1,
p+ −r,p− −s ,
± = soc V ∓
Xr,p 1  r  p+ − 1,
− p+ −r,p− ,
Xp+ ,s = soc Vp∓+ ,p− −s ,
± 1  s  p− − 1,
Xp±+ ,p− = soc Vp±+ ,p− .

We note that Xp±+ ,p− = Vp±+ ,p− . Therefore, in particular,

 
p+ p−

+ −
soc F = Xr,s ⊕ Xr,s . (2.13)
r=1 s=1

The space Xr,s+ is represented by the collection of filled dots in Figs. 2 and 5, X −
p+ −r,s by

the collection of boxes in Fig. 3, and Xp+ −r,p− by the collection of open dots in Fig. 4. They
decompose into direct sums of irreducible Virasoro modules as follows.
314 B.L. Feigin et al. / Nuclear Physics B 757 [FS] (2006) 303–343


Fig. 3. Structure of the modules Vr,p − −s
. The notation is the same as in Fig. 1. Boxes  denote Virasoro representations
that are combined into Xp−+ −r,s . Levels of conformal dimensions are chosen for (r, s) ∈ I1 , and p− > p+ . The dotted
lines show the action of the W ± (z) generators of Wp+ ,p− in Xp−+ −r,s .

± for 1  r  p and 1  s  p decomposes


2.3.2. Lemma. As a Virasoro module, the space Xr,s + −
as
 
+ −
Xr,s  (2a + 1)Jr,p− −s;2a+1 , Xr,s  2aJr,p− −s;2a , (2.14)
a0 a1

with the identification J0,s;n ≡ Jp+ ,s;n+1 and Jr,0;n ≡ Jr,p− ;n−1 introduced above.

As we see in what follows, the Xr,s± become W -algebra representations. Describing the W -

algebra requires studying the action of screenings, which we consider in the next section.

3. Screening operators and the quantum group

In this section, we introduce screening operators and study the quantum group associated with
them. Because the screenings do not act in the free-field module F , we have to extend F to a
larger space F in Section 3.1. In Section 3.2, we then reformulate the action of the screenings
as a representation of a Hopf algebra, with the result formulated in Section 3.3. A subalgebra in
a quotient of the Drinfeld double of this Hopf algebra is the quantum group that is Kazhdan–
± introduced above
Lusztig dual to the W -algebra. In Section 3.4, we next describe the spaces Xr,s
B.L. Feigin et al. / Nuclear Physics B 757 [FS] (2006) 303–343 315

+
Fig. 4. Structure of the modules Vr,p − . The notation is the same as in Fig. 1. Open dots ◦ denote Virasoro representations
that are combined into Xp+ −r,p− . The dotted lines show the action of the W ± (z) generators of Wp+ ,p− in Xp−+ −r,p− .

in terms of the screenings. We also show in Section 3.5 that the relevant spaces carry a represen-
tation of the s(2, C) algebra commuting with the Virasoro action. All these ingredients are to
be used in the next section in the study of the Wp+ ,p− algebra.

3.1. Screening operators and dressing

Free-field construction of both the minimal model and its logarithmic extension involves
screening operators e+ and f− that commute with the energy–momentum tensor, [e+ , T (z)] =
[f− , T (z)] = 0. They have the standard form
 
e+ = dz e α+ ϕ(z)
, f− = dz eα− ϕ(z) . (3.1)

The operators e+ and f− do not act in the space F (see (2.10)), and to complete their defi-
nition, we must extend this space appropriately. For this, we introduce the space F spanned by
dressed fields (cf. [25,44])

dz1 · · · dzj dw1 · · · dwm eα+ ϕ(z1 ) · · · eα+ ϕ(zj ) eα− ϕ(w1 ) · · · eα− ϕ(wm ) P(∂ϕ)Vr,s;n (z),
Γˆ
0  j  p+ − 1, 0  m  p− − 1, (3.2)
where Γˆ is a local system defined as in [44] and P(∂ϕ) is a differential polynomial in ∂ϕ.
316 B.L. Feigin et al. / Nuclear Physics B 757 [FS] (2006) 303–343

Fig. 5. Structure of the modules Vp−+ −r,p− . The notation is the same as in Fig. 1. Filled dots • denote Virasoro rep-
+ ±
resentations that are combined into Xr,p − . The dotted lines show the action of the W (z) generators of Wp+ ,p− in
+
Xr,p− .

The action of e+ and f− on each vector (3.2) can then be evaluated by manipulations with
contour integrals, as described in [25,44]. In particular,
r
e+ : Fr,s;n → Fp+ −r,s;n+1 ,
f−s : Fr,s;n → Fr,p− −s;n−1 .
Evidently, these spaces and the maps between them constitute the Felder complexes [25] whose
cohomology gives the standard minimal model.

3.2. The Hopf algebra of screenings

The action of screening operators in F gives rise to a Hopf algebra representation. The fol-
lowing lemma basically restates some known facts about the screenings [25,44].

3.2.1. Lemma. The space F defined in Section 3.1 admits the action of the operators
√ iπ
ϕ0
e+ , f− , k=e 2p+ p−
, (3.3)
which satisfy the relations
p p
e++ = f− − = 0, k 4p+ p− = 1, e + f− = f − e+ ,
−1 −1 −1
ke+ k = q + e+ , kf− k = q − f− , (3.4)
B.L. Feigin et al. / Nuclear Physics B 757 [FS] (2006) 303–343 317

where (see (2.9))


iπ iπ
q+ = q2p− = e p+ , q− = q2p+ = e p− .

Let H denote the associative algebra generated by e+ , f− , and k with relations (3.4). Clearly,
the PBW basis in H is given by
ej mn = k j e+
m n
f− , 0  j  4p+ p− − 1, 0  m  p+ − 1, 0  n  p− − 1. (3.5)

3.2.2. Lemma. The algebra H is a Hopf algebra with the counit


(e+ ) = (f− ) = 0, (k) = 1, (3.6)
comultiplication
Δ(k) = k ⊗ k, Δ(e+ ) = e+ ⊗ 1 + k 2p− ⊗ e+ ,
Δ(f− ) = f− ⊗ 1 + k −2p+ ⊗ f− , (3.7)
and antipode
S(k) = k −1 , S(e+ ) = −k −2p− e+ , S(f− ) = −k 2p+ f− . (3.8)

We note that the counit is given by the trivial representation on the vertex V1,1 = 1. The co-
multiplication is taken from operator product expansion; for example, the standard manipulations
lead to
 
e+ ej1 ϕ(w) ej2 ϕ(u)

 
= eα+ ϕ(z) ej1 ϕ(w) ej2 ϕ(u) dz
   
= e α+ ϕ(z) j1 ϕ(w)
e dz e j2 ϕ(u)
+e iπα+ j1 j1 ϕ(w)
e e α+ ϕ(z) j2 ϕ(u)
e dz

= e+ ej1 ϕ(w) ej2 ϕ(u) + k 2p− ej1 ϕ(w) e+ ej2 ϕ(u) ,


which gives the comultiplication for e+ , and similarly for f− and k. With the counit and comul-
tiplication thus fixed, the antipode S is uniquely calculated from the comultiplication, with the
result in (3.8).
The full quantum group realized in a given conformal field theory model involves not only
the screening but also the contour-removal operators. A convenient procedure for introducing the
latter is to take Drinfeld’s double (see [45,46] and Appendix B.1) of the algebra H generated by
the screenings.

3.3. Structure of the double

3.3.1. Theorem. The double D(H) of H is a Hopf algebra generated by e± , f± , k, and κ with
the relations
p p
e±± = f± ± = 0, k 4p+ p− = κ 4p+ p− = 1, (3.9)
ke± k −1 = q± e± , kf± k −1 = q−1
± f± , κe± κ −1 = q−1
± e± ,
κf± κ −1 = q± f± , (3.10)
318 B.L. Feigin et al. / Nuclear Physics B 757 [FS] (2006) 303–343

kκ = κk, e+ e − = e − e + , f+ f− = f − f+ ,
e+ f − = f − e + , e− f + = f + e− , (3.11)

k ±2p∓ − κ ±2p∓
[e± , f± ] = ±p∓ ∓p∓ , (3.12)
q± − q±
Δ(k) = k ⊗ k, Δ(e+ ) = e+ ⊗ 1 + k 2p− ⊗ e+ ,
Δ(f− ) = f− ⊗ 1 + k −2p+ ⊗ f− , (3.13)

Δ(κ) = κ ⊗ κ, Δ(f+ ) = f+ ⊗ κ 2p− + 1 ⊗ f+ ,


Δ(e− ) = e− ⊗ κ −2p+ + 1 ⊗ e− , (3.14)

S(k) = k −1 , S(e+ ) = −k −2p− e+ , S(f− ) = −k 2p+ f− , (3.15)


S(κ) = κ −1 , S(f+ ) = −f+ κ −2p− , S(e− ) = −e− κ 2p+ , (3.16)
(e± ) = (f± ) = 0, (k) = (κ) = 1. (3.17)

We prove this in Appendix B.2 by a routine application of the standard construction.


The double thus introduces contour-removal operators e− and f+ , dual to e+ and f− in the
sense that is fully clarified in the proof in Appendix B. But the doubling procedure also yields
the dual κ to the Cartan element k in H, which is to be eliminated by passing to the quotient

D̄(H) = D(H)/(kκ − 1) (3.18)


over the Hopf ideal generated by the central element kκ − 1. We next take a subalgebra in D̄(H)
(which, unlike D̄(H), is a factorizable ribbon quantum group, see [34]).

3.3.2. Definition. Let gp+ ,p− be the subalgebra in D̄(H) generated by e+ , f+ , e− , f− , and

K = k2. (3.19)

Explicit relations among the generators are a mere rewriting of the corresponding formulas in
the theorem,
p p
e±± = f± ± = 0, K 2p+ p− = 1,
Ke± K −1 = q2± e± , Kf± K −1 = q−2
± f± ,
e+ e − = e − e + , f+ f− = f − f+ , e+ f − = f − e+ , e− f + = f + e− ,
K ±p∓ − K ∓p∓
[e± , f± ] = ±p∓ ∓p∓ ,
q± − q±
with the comultiplication, antipode, and counit defined in (3.13)–(3.17). The structure of gp+ ,p−
and its relation to the logarithmic (p+ , p− ) model are investigated in [34].
This quantum group has 2p+ p− irreducible representations, which we label as X± r,s with 1 
r  p+ and 1  s  p− . The gp+ ,p− -module X± r,s is generated by e + and f − from the eigenvector
r−1 −s+1 ±
of K with the eigenvalue ±q+ q− , and we have dim Xr,s = rs. In what follows, we use the
gp+ ,p− -modules to describe the bimodule structure of F.
B.L. Feigin et al. / Nuclear Physics B 757 [FS] (2006) 303–343 319

3.4. Kernels and images of the screenings

± , so far
Our aim in what follows is to introduce some other module structures in the spaces Xr,s
defined as Virasoro modules in Definition 2.3.1. For this, we first describe them an intersection
of the images of screenings.
The socle of the space F (see (2.13)), still viewed as a Virasoro module, can also be written
as
p −1 p −1
soc F = im e++ ∩ im f− − ⊂ F.
± that constitute soc F
Equivalently (and somewhat more convenient technically), the spaces Xr,s
are described as intersections of the images of the screenings as
p −r p −s
± = im e +
Xr,s + ∩ im f− − in Vp±+ −r,p− −s for 1  r  p+ − 1,
1  s  p− − 1,
p −r
± = im e +
Xr,p − + in Vp∓+ −r,p− for 1  r  p+ − 1,
p −s
Xp±+ ,s = im f− − in Vp∓+ ,p− −s for 1  s  p− − 1. (3.20)
p −r
In Fig. 2, the image of e++ is given by the collection of filled and open dots and the image of
p −s
f− − by the collection of filled dots and boxes.
In addition to the images, we also consider the kernels, namely, ker e+ ∩ ker f− ⊂ F, which
decomposes similarly to (2.13) as

p+ 
p−
 
+ −
ker e+ ∩ ker f− = Kr,s ⊕ Kr,s ,
r=1 s=1

where the ±
Kr,s can equivalently be identified in Vp±+ −r,p− −s as
± = ker er ∩ ker f s
Kr,s + − in Vp±+ −r,p− −s for 1  r  p+ − 1,
1  s  p− − 1,
±
Kr,p −
= ker e+
r in Vp∓+ −r,p− for 1  r  p+ − 1,
Kp±+ ,s = ker f−s in Vp∓+ ,p− −s for 1  s  p− − 1 (3.21)
(and Kp±+ ,p− = Vp±+ ,p− for uniformity of notation). Clearly, the Kr,s
± are Virasoro modules.
r
In Fig. 2, the kernel of e+ coincides with the collection of filled and open dots and the cross,
and the kernel of f−s coincides with the collection of filled dots, boxes, and the cross. It is easy
to see that
+ + + +
Kr,s ⊃ Xr,s with Kr,s /Xr,s = Jr,s for 1  r  p+ − 1, 1  s  p− − 1, (3.22)
− −
Kr,s = Xr,s for 1  r  p+ , 1  s  p− , (3.23)
+ +
Kr,s = Xr,s whenever r = p+ or s = p− . (3.24)
± are in fact irreducible W
In what follows, we see that the spaces Xr,s p+ ,p− -representations; on
±
the other hand, the Kr,s , which are also Wp+ ,p− -representations, are not all irreducible. However,
+
they play a crucial role in the Wp+ ,p− representation theory; K1,1 is the vacuum representation
±
of the Wp+ ,p− algebra, and the Kr,s can be identified with the preferred basis of the Wp+ ,p−
fusion algebra G.
320 B.L. Feigin et al. / Nuclear Physics B 757 [FS] (2006) 303–343

3.5. A Lusztig extension of the quantum group

± and K± admit an action of the s(2, C) algebra.


Virasoro modules Xr,s r,s

3.5.1. Lemma. The spaces Xr,s± and K± admit an s(2, C)-action that is a derivation of the op-
r,s
erator product expansion. The Virasoro algebra Virp+ /p− , see (2.3), commutes with this s(2, C).

The construction of the s(2, C) action is based on Lusztig’s divided powers and can be briefly
p
described as follows [47]. Morally, the e and f generators of s(2, C) are given by e++ /[p+ ]+ !
p−
and f− /[p− ]− !, where we use the standard notation
qn − q−n
[n]q = , [n]q ! = [1]q [2]q · · · [n]q
q − q−1
for q-integers and q-factorials and set [m]+ = [m]qp− and [m]− = [m]qp+ . But if taken liter-
+ −
p p
ally, both e++ /[p+ ]+ ! and f− − /[p− ]− ! are given by 0/0 ambiguities for e+ and f− acting as
described above and for q in Eq. (2.9) (because [p± ]± = 0). To resolve the ambiguities, we
consider a deformation of q in Eq. (2.9) as

+
q = e 2p+ p−
or, equivalently, a deformation of the parameter α0 . We thus obtain e+ (), f− (), and q-factorials
[n]± ! depending on , such that [p± ]± ! = 0. The limits
e+ ()p+ f− ()p−
e = lim and f = lim
→0 [p+ ]+ ! →0 [p− ]− !
± and are independent of the deformation details, and are therefore well-defined opera-
act in Kr,s
± . Because e and f commute with the Virasoro algebra action in K± , the spaces X ±
tors in Kr,s r,s r,s
are also representations of the s(2, C) algebra generated by e and f .
+ /X + (where 1  r  p − 1 and 1  s  p − 1) are s(2, C)-
We also note that Jr,s = Kr,s r,s + −
singlets.

± and X ± have the structure of (Vir


3.5.2. Lemma. The spaces Kr,s r,s p+ /p− , s(2, C))-bimodules
given by

+
Kr,s = Cn ⊗ 2n−1 , 1  r  p+ − 1, 1  s  p− − 1 (3.25)
n∈N
and
 
+ −
Xr,s  Jr,p− −s;2n−1 ⊗ 2n−1 , Xr,s  Jr,p− −s;2n ⊗ 2n , (3.26)
n∈N n∈N
1  r  p+ , 1  s  p− ,
where n is the n-dimensional irreducible s(2, C) representation and

Jr,p− −s;2n−1 , n  2,
Cn = Jr,s Jr,p− −s;1
× −→ • , n = 1,
where we use the same notation as in Fig. 1 for the extension of the Virasoro module Jr,s by the
Virasoro module Jr,p− −s;1 .
B.L. Feigin et al. / Nuclear Physics B 757 [FS] (2006) 303–343 321

Each direct summand in the decomposition of Xr,s + in (3.26) corresponds to a horizontal row

of filled dots in Figs. 2 and 5, and each direct summand in the decomposition of Xp−+ −r,s corre-
sponds to a horizontal row of boxes and open dots in Figs. 3 and 4, respectively, with the s(2, C)
algebra acting in each row. We thus have 1, 3, 5, . . . -dimensional s(2, C)-representations in
Figs. 2 and 5, and 2, 4, 6, . . . -dimensional representations in Figs. 3 and 4.
We now have all the ingredients needed for constructing the W -algebra Wp+ ,p− of the
(p+ , p− ) logarithmic conformal field theory model.

4. Vertex-operator algebra for the (p+ , p− )-conformal field theory and its representations

In this section, we define the chiral symmetry algebra Wp+ ,p− of the (p+ , p− ) logarithmic
model and describe irreducible and Verma Wp+ ,p− -modules.

4.1. Definition of Wp+ ,p−

4.1.1. Definition. The algebra Wp+ ,p− is the subalgebra of Lp+ ,p− with the underlying vector
+
space Wp+ ,p− = K1,1 ⊂ Lp+ ,p− (see (3.21)).

The (Virp+ /p− , s(2, C))-bimodule structure of the vacuum Wp+ ,p− -representation Wp+ ,p−
is described in Lemma 3.5.2 and can be understood better using Fig. 2, as the part of the figure
consisting of the cross and filled dots.5

4.2. The W algebra of (p+ , p− ) logarithmic models

4.2.1. Theorem.

(1) The algebra Wp+ ,p− is generated by T (z) in (2.3) and the two currents W + (z) and W − (z)
given by

W + (z) = (f− )p− −1 V1,p− −1;3 (z), W − (z) = (e+ )p+ −1 Vp+ −1,1;−3 (z), (4.1)
which are Virasoro primaries of conformal dimension (2p+ −1)(2p− − 1).
(2) The maximal (and the only nontrivial) vertex-operator ideal R of Wp+ ,p− is generated by
W + (z) and W − (z).
(3) The quotient Wp+ ,p− /R is the vertex-operator algebra Mp+ ,p− of the (p+ , p− ) Virasoro
minimal model.

4.2.2. Remark. In accordance with the definition of the screening action in (4.1), the W ± (z)
currents can be written as (with normal ordering understood)
+
W + (z) = P3p+ p− −3p+ −p− +1
ep+ α+ ϕ (z),

W − (z) = P3p+ p− −p+ −3p− +1
ep− α− ϕ (z),

5 In the Coulomb-gas picture, W


p+ ,p− can be viewed as the symmetry algebra of a fixed-point system with quenched
disorder.
322 B.L. Feigin et al. / Nuclear Physics B 757 [FS] (2006) 303–343

where Pj± are polynomials of degree j in ∂ n ϕ, n  1 (with deg ∂ n ϕ = n). The OPE of W + (z)
and W − (z) is
Sp+ ,p− (T )
W + (z)W − (w) = + less singular terms, (4.2)
(z − w)7p+ p− −3p+ −3p− +1
where Sp+ ,p− (T ) is the vacuum singular vector—the polynomial of degree 12 (p+ − 1)(p− − 1) in
T and ∂ n T , n  1, such that Sp+ ,p− (T ) = 0 is the polynomial relation for the energy–momentum
tensor in the (p+ , p− ) Virasoro minimal model. We note that this OPE is already quite difficult
for direct analysis in the simplest case of (3, 2) model, see Appendix A.

4.2.3. Remark. As follows from Definition 4.1.1 and Lemma 3.5.2, Wp+ ,p− admits an s(2, C)
action and the W ± (z) generators are highest- and lowest-weight vectors of an s(2, C)-triplet.
We note that the W algebra of (p, 1) logarithmic models bears the name triplet [14,16] because
it admits such an s(2, C) action and the corresponding W ± (z) generators are also the highest-
and lowest-weight vectors in a triplet.

Proof of Theorem 4.2.1. With decomposition (3.25) taken into account, it suffices to construct
(Virp+ /p− , s(2, C)) highest-weight vectors in each direct summand in order to show part (1).
This can be done as follows. The highest-weight vector of each component Cn ⊗ 2n−1 (cor-
responding to the rightmost dot in each row of filled dots in Fig. 2) is identified with the field
W −,n (z), n  1, defined as the first nonzero term in the OPE of W − (z) with W −,n−1 (z); the
recursion base is W −,1 ≡ 1. The highest-weight vector of C1 ⊗ 1 is identified with the identity
operator 1. This shows part (1).
+ p −1 p −1
Next, it follows from Lemma 3.5.2 that X1,1 = im e++ ∩ im f− − is invariant under
+ +
Wp+ ,p− . The space X1,1 is the maximal Wp+ ,p− -submodule in Wp+ ,p− = K1,1 , and therefore
+ +
there exists a map Wp+ ,p− → Mp+ ,p− with the kernel X1,1 . We set R = X1,1 , which is the max-
imal vertex-operator ideal in Wp+ ,p− . Evidently, W ± (z) generate some subspace R in R, but
the decomposition into the direct sum of s(2, C)-representations allows calculating the charac-
p −1 p −1
ter of R , which coincides with the character of im e++ ∩ im f− − calculated from the Felder
resolution, thus showing parts (2) and (3). 2

4.3. Irreducible Wp+ ,p− -modules

We now describe irreducible Wp+ ,p− -modules. First, the irreducible Virasoro modules Jr,s
with (r, s) ∈ I1 are at the same time Wp+ ,p− -modules, with the ideal R acting by zero. We write
Xr,s for the Jr,s regarded as Wp+ ,p− -modules. Second, there are 2p+ p− irreducible Wp+ ,p− -
modules where R acts nontrivially. These are the spaces Xr,s ± introduced in Definition 2.3.1,

which are evidently Wp+ ,p− -modules because of their description in (3.20).
± is an irreducible W
4.3.1. Proposition. Xr,s p+ ,p− -module.

Proof. Taking decomposition (3.26) into account, we show the irreducibility of Xr,s± by literally

repeating the construction for the (Virp+ /p− , s(2, C)) highest-weight vectors in the proof of
Theorem 4.2.1.
+ is V
The lowest-dimension Virasoro primary in Xr,s r,p− −s;1 (of dimension Δr,p− −s;1 ) and two
lowest-dimension Virasoro primaries in Xr,s are (e+ )p+ −r Vp+ −r,s;−2 and (f− )p− −s Vr,p− −s;2

B.L. Feigin et al. / Nuclear Physics B 757 [FS] (2006) 303–343 323

Table 1
The p+ × p− W -algebra Kac table for p+ = 3 and p− = 2. Each (r, s)
+
box contains the dimension of the highest-weight vectors of Xr,s and

Xr,s , in this order. The (p+ − 1) × (p− − 1) = 2 × 1 subtable also con-
tains the dimensions of the Xr,s (which are X1,1 in this case), shown in
parentheses. The (infinite) Virasoro content follows from the decompo-
sitions in Lemma 2.3.2
5 , 33 1 21 1 , 35
− 24
8 8 8, 8 24
2, 7 (0) 1, 5 (0) 1 , 10
3 3

(of dimension Δp+ −r,s;−2 ). The irreducible representations can be arranged into a p+ × p− Kac
+ and X − in each box (in addition, the standard (p − 1) × (p − 1) Kac table
table, with Xr,s r,s + −
containing 12 (p+ − 1)(p− − 1) distinct representations is inherited from the minimal model); the
± is shown by dotted lines in Figs. 2–
(3, 2) example is given in Table 1. The Wp+ ,p− -action in Xr,s
5: for Xr,s with 1  r  p+ − 1 and 1  s  p− − 1 in Fig. 2, for Xp−+ −r,s with 1  r  p+ − 1
+

and 1  s  p− − 1 in Fig. 3, for Xp−+ −r,p− with 1  r  p+ − 1 in Fig. 4, and for Xr,p + with

1  r  p+ − 1 in Fig. 5.

4.4. “Verma” modules

The Lp+ ,p− -modules Vr,s± (see (2.12)) are W


p+ ,p− -modules (simply because Wp+ ,p− is a
± as W
subalgebra in Lp+ ,p− ). In referring to the modules Vr,s p+ ,p− -modules, it is convenient
to call them the Verma modules of the Wp+ ,p− algebra. (Their counterparts for the Kazhdan–
Lusztig-dual quantum group [34] are indeed Verma modules, but investigation of the Verma
± is a separate problem, which we do not consider here and only use the conve-
properties of Vr,s
nient and suggestive name for these modules.) We now describe their subquotients.
± is as follows.
4.4.1. Proposition. The subquotient structure of the Vr,s

(1) Vp±+ ,p−  Xp±+ ,p− .


(2) For 1  r  p+ − 1, Xp∓+ −r,p− ⊂ Vr,p
±

± /X ∓
and Vr,p−
±
p+ −r,p−  Xr,p− .
∓ ∓
(3) For 1  s  p− − 1, Xp+ ,p− −s ⊂ Vp±+ ,s and Vp±+ ,s /Xp+ ,p− −s  Xp±+ ,s .
(4) For 1  r  p+ − 1 and 1  s  p− − 1, Vr,s + admits a filtration H ⊂ H ⊂ V + , where
0 1 r,s
+ − − + /H  X + ; and V − ad-
H0  Xp+ −r,p− −s , H1 /H0  Xp+ −r,s ⊕ Xr,s ⊕ Xr,p − −s
, and Vr,s 1 r,s r,s
mits a filtration

H0 ⊂ H1 ⊂ Vr,s ,
where H0  Xp−+ −r,p− −s , H1 /H0  Xp++ −r,s ⊕ Xr,p
+
− −s
+ /H  X − .
, and Vr,s 1 r,s

The subquotient structure is clear from Figs. 2, 3, 4, and 5. For example, we consider Fig. 2.
+ , the cross × is X , the ◦ and  are combined
The filled dots • constitute the subquotient Xr,s r,s
into Xr,p− −s and Xp+ −r,s respectively, and the  are combined into Xp++ −r,p− −s .
− −

4.5. Bimodule structure of F

The quantum group gp+ ,p− acts in the space F, which is in fact a (Wp+ ,p− , gp+ ,p− )-
bimodule.
324 B.L. Feigin et al. / Nuclear Physics B 757 [FS] (2006) 303–343

Fig. 6. Subquotient structure of Qr,0 .  denotes the external tensor product. Solid lines denote the Wp+ ,p− -action and
dashed lines denote the gp+ ,p− -action.

The subquotient structure of the bimodule F shows the origin of the multiplicities in decom-
position (4.4): each indecomposable Wp+ ,p− -module enters in several copies produced by the
action of H.

4.5.1. Proposition. As a (Wp+ ,p− , gp+ ,p− )-bimodule, the space F decomposes as

F= Qr,s (4.3)
(r,s)∈I0

where Qr,s are indecomposable (Wp+ ,p− , gp+ ,p− )-bimodules with the following structure.

(1) If (r, s) = (0, 0), then Q0,0 = Xp++ ,p−  X+p+ ,p− ;
(2) if (r, s) = (p+ , 0), then Qp+ ,0 = Xp−+ ,p−  X−p+ ,p− ;
(3) if 1  r  p+ − 1 and s = 0, then Qr,0 has the subquotient structure described in Fig. 6;
(4) if r = 0 and 1  s  p− − 1, then Q0,s has the subquotient structure obtained from Fig. 6 by
changing Xr,p + → X+ , X+ + + + + +
− p+ ,s p+ −r,p− → Xp+ ,p− −s , Xr,p− → Xp+ ,s , Xp+ −r,p− → Xp+ ,p− −s ,
and e+ → f− ;
(5) if (r, s) ∈ I1 , then Qr,s has the subquotient structure described in Fig. 7.

4.5.1. In particular, forgetting the quantum-group structure, we obtain a decomposition of


F into a finite sum of Wp+ ,p− Verma modules with multiplicities (dimensions of irreducible
gp+ ,p− -modules) as
p+ −1 p− −1 p+ −1
   +    + 
− −
F= (p+ − r)(p− − s) Vr,s ⊕ Vr,s ⊕ (p+ − r)p− Vr,p−
⊕ Vr,p−
r=1 s=1 r=1
p− −1
    
⊕ (p− − s)p+ Vp++ ,s ⊕ Vp−+ ,s ⊕ p+ p− Vp++ ,p− ⊕ Vp−+ ,p− . (4.4)
s=1
The structure of F as a (Wp+ ,p− , gp+ ,p− )-bimodule can be taken as a starting point for con-
structing the projective Wp+ ,p− -modules by the method in [17].

5. The space of torus amplitudes

In this section, we find the SL(2, Z)-representation generated from the space of characters G
in the logarithmic (p+ , p− ) model. With Conjecture 1.1.3 assumed, this gives the space of torus
amplitudes. In Section 5.1, we calculate the characters of irreducible Wp+ ,p− -modules in terms
of theta-constants, see (1.9). It is then straightforward to calculate modular transformations of the
B.L. Feigin et al. / Nuclear Physics B 757 [FS] (2006) 303–343 325

+ −
Fig. 7. Subquotient structure of Qr,s . We use the notation • = Xr,s , ◦ = Xr,p − −s
,  = Xp−+ −r,s ,  = Xp++ −r,p− −s , and
× = Xr,s for Wp+ ,p− -modules, and ♣ = X+ − − +
r,s , ♦ = Xr,p− −s , ♥ = Xp+ −r,s , and ♠ = Xp+ −r,p− −s for quantum-group
modules. Solid lines show the action of Wp+ ,p− -generators and dashed lines show the action of gp+ ,p− -generators,
and  denotes external tensor product.

characters, which we do in Section 5.2. This extends the space of characters by C[τ ]/τ 3 , giving
rise to the space of generalized characters C̄ (which, as noted above, most probably coincides
with the space C of torus amplitudes). (Theta-function derivatives enter the characters through
the second order; this can be considered a “technical” reason for the explicit occurrences of τ
and τ 2 in the modular transformation properties.) The analysis of the SL(2, Z)-representation on
C̄ then yields the results in Theorem 1.1.1. In Section 5.3, we use the established decomposition
of the SL(2, Z) action to briefly consider modular invariants.

5.1. Wp+ ,p− -characters

Let
c
± L −
χr,s (q) = TrXr,s
± q 0 24 , 1  r  p+ , 1  s  p − ,
c
χr,s (q) = TrXr,s q L0 − 24 , (r, s) ∈ I1 ,
± and X
with c given by (2.4), be the characters of the irreducible Wp+ ,p− -representations Xr,s r,s
(see Section 4.3).

5.1.1. Proposition. The irreducible Wp+ ,p− -representation characters are given by

1
χr,s = (θp+ s−p− r − θp+ s+p− r ), (r, s) ∈ I1 , (5.1)
η
326 B.L. Feigin et al. / Nuclear Physics B 757 [FS] (2006) 303–343


+ 1
χr,s = θ  − θp+ s−p− r − (p+ s + p− r)θp + s+p− r
(p+ p− )2 η p+ s+p− r

(p+ s + p− r)2 (p+ s − p− r)2
+ (p+ s − p− r)θp + s−p− r + θp+ s+p− r − θp+ s−p− r ,
4 4
1  r  p+ , 1  s  p − , (5.2)


− 1
χr,s = θ  − θp+ p− +p+ s−p− r + (p+ s + p− r)θp + p− −p+ s−p− r
(p+ p− )2 η p+ p− −p+ s−p− r
(p+ s + p− r)2 − (p+ p− )2
+ (p+ s − p− r)θp + p− +p+ s−p− r + θp+ p− −p+ s−p− r
4

(p+ s − p− r)2 − (p+ p− )2
− θp+ p− +p+ s−p− r ,
4
1  r  p+ , 1  s  p − . (5.3)

Proof. We first recall the well-known irreducible Virasoro characters [48]

char Jr,s;n (q)


1−c 
q 24 Δr,s;n+2m Δr,s;−n−2m Δr,p −s;n+2m+1
= q + q − q −
η(q)
m0 m1 m0

− q Δr,p− −s;−n−2m−1 . (5.4)
m0

In particular, for the characters of Xr,s , we immediately have

q 24  Δr,s;2m
1−c

char Xr,s (q) = q − q Δr,p− −s;2m+1 ,
η(q)
m∈Z

which gives (5.1) when rewritten in terms of the theta-constants.


Next, from Lemma 2.3.2, we have that the character of each Xr,s + for 1  r  p and 1  s 
+
p− is given by

+ L −c
χr,s (q) = TrXr,s
+ q 0 24 = (2a + 1) char Jr,p− −s;2a+1 (q).
a0

Substituting (5.4) here, we have


+
χr,s (q) = Σ1 + Σ2 ,
where
 
1−c Δr,p− −s;2a+2m+1 Δr,p− −s;−2a−2m−1
Σ1 = (2a + 1)q 24 q + q ,
a0 m0 m1
 
1−c
Σ2 = − (2a + 1)q 24 q Δr,s;2a+2m+2 + q Δr,s;−2a−2m−2 .
a0 m0 m0
B.L. Feigin et al. / Nuclear Physics B 757 [FS] (2006) 303–343 327

Setting Δ̄r,s;n = Δr,s;n + 1−c


24 for brevity, we obtain


m −m−2

Σ1 = (2a + 1)q Δ̄r,p− −s;2m+1 + (2a + 1)q Δ̄r,p− −s;2m+1
m0 a=0 m−2 a=0

2 Δ̄r,p− −s;2m+1
= (m + 1) q
m∈Z
p+ p− (m−
p+ s+p− r
= m2 q 2p+ p− )

m∈Z
 
p+ s + p− r 2 p+ p− (m− p2p
+ s+p− r )2
= m− q + p−
2p+ p−
m∈Z
 
p+ s + p− r p+ s + p− r p+ p− (m− p2p
+ s+p− r )2
+ m− q + p−
p+ p− 2p+ p−
m∈Z
(p+ s + p− r)2 p+ p− (m− p2p
+ s+p− r )2
+ q + p−
(2p+ p− )2
m∈Z
1  p+ s + p− r  (p+ s + p− r)2
= θ −p + s−p− r + θ −p+ s−p− r + θ−p+ s−p− r ,
(p+ p− )2 (p+ p− )2 (2p+ p− )2
and a similar calculation gives

1  p+ s − p− r  (p+ s − p− r)2
Σ2 = − θ p s−p r + θ p s−p r − θp+ s−p− r ,
(p+ p− )2 + − (p+ p− )2 + −
(2p+ p− )2
whence (5.2) follows. Similar calculations give the character in (5.3). 2

± in Eqs. (2.12), we calculate their


5.1.2. Remark. From the definition of the Verma modules Vr,s
characters as

q 24 Δr,s,2n
1−c
+ 1
char Vr,s (q) = q = θp s−p r, p p (q),
η(q) η(q) + − + −
n∈Z

q
1−c
24 1

char Vr,s (q) = q Δr,s,2n+1 = θp p +p s−p r, p p (q).
η(q) η(q) + − + − + −
n∈Z

On the other hand, from the subquotient structure of Verma modules described in Proposi-
tion 4.4.1, we obtain the identities

+
char Vr,s +
(q) = χr,s (q) + χr,s (q) + χr,p − −s
(q) + χp−+ −r,s (q) + χp++ −r,p− −s (q),
+

char Vr,s −
(q) = χr,s (q) + χr,p − −s
(q) + χp++ −r,s (q) + χp−+ −r,p− −s (q)
for (r, s) ∈ I1 , which shows that, as could be expected, the space of Verma-module characters
coincides with the space of theta-functions (not including their derivatives). We also note that
similar expressions for characters involving second derivatives of theta-functions were proposed
in [49].
328 B.L. Feigin et al. / Nuclear Physics B 757 [FS] (2006) 303–343

5.2. The space C̄

We now construct the SL(2, Z)-representation C̄ generated from the space G of the irreducible-
representation characters and simultaneously obtain the decomposition in Theorem 1.1.1.

5.2.1. First, G contains the subspace Rmin spanned by χr,s with (r, s) ∈ I1 —the Virasoro
minimal-model characters, which evidently carry an SL(2, Z)-representation. Second, as we see
shortly, the 12 (p+ + 1)(p− + 1)-dimensional space Rproj ⊂ G linearly spanned by the functions
+ + 2χ −
r,s = χr,s + 2χr,s − +
r,p− −s + 2χp+ −r,s + 2χp+ −r,p− −s , (r, s) ∈ I1 ,
0,s = 2χp++ ,p− −s + 2χp−+ ,s , 1  s  p− − 1,
r,0 = 2χp++ −r,p− + 2χr,p
− ,

1  r  p+ − 1,
0,0 = 2χp++ ,p− ,
p+ ,0 = 2χp−+ ,p− (5.5)
(which can be identified with the characters of projective Wp+ ,p− -modules) also carries an
SL(2, Z)-representation.
Next, the (p+ p− − 1)-dimensional complement of Rmin ⊕ Rproj in G contains no more sub-
spaces closed under the SL(2, Z)-action. It is convenient to choose the functions
p+ s − p− r  +   

ρr,s = χr,s − p+ (p− − s) χr,s + χp−+ −r,s + p+ s χp++ −r,p− −s + χr,p−
− −s
,
2
(r, s) ∈ I1 ,
  
ρ0,s = p+ sχp++ ,p− −s − (p− − s)χp−+ ,s , 1  s  p− − 1,
p− r − p+ s  + −   

ρr,s = χr,s + p− (r − p+ ) χr,s + χr,p − −s
+ p− r χp++ −r,p− −s + χp−+ −r,s ,
2
(r, s) ∈ I1 ,
  
ρr,0 = p− rχp++ −r,p− − (p+ − r)χr,p−

, 1  r  p+ − 1,

(p+ s − p− r)2
ρr,s = p+ p− (p+ − r)(p− − s)χr,s +
+ rsχp++ −r,p− −s − χr,s
4p+ p−


− (p+ − r)sχr,p − −s
− r(p− − s)χp−+ −r,s , (r, s) ∈ I1

as a basis in this complement.


We then define
 (τ ) = τρ  (τ ),
ϕr,s (r, s) ∈ I ,
r,s
 (τ ) = τρ  (τ ),
ϕr,s (r, s) ∈ I ,
r,s
ψr,s (τ ) = 2τρr,s (τ ) + iπp+ p− χr,s (τ ), (r, s) ∈ I1 ,
ϕr,s (τ ) = τ 2 ρr,s (τ ) + iπp+ p− τ χr,s (τ ), (r, s) ∈ I1 .
(5.6)

5.2.1.1. Proposition. The SL(2, Z)-representation C̄ generated from G is the linear span of the
2 (3p+ − 1)(3p− − 1) functions
1

χr,s (τ ), ρr,s (τ ), ψr,s (τ ), and ϕr,s (τ ) with (r, s) ∈ I1 ,


B.L. Feigin et al. / Nuclear Physics B 757 [FS] (2006) 303–343 329

r,s (τ ) with (r, s) ∈ I0 ,


 (τ )
ρr,s and ϕr,s  (τ ) with (r, s) ∈ I , and

 (τ )
ρr,s and ϕr,s  (τ ) with (r, s) ∈ I .


This is summarized in Table 2, where we also indicate the SL(2, Z)-representation and its
dimension spanned by each group of functions.
The proof amounts to the following two lemmas, the first of which shows that C̄ is closed
under the S-transformation and the second that it is closed under the T-transformation.

5.2.1.2. Lemma.
  √
1 2 2   πp− rr  πp+ ss 
χr,s − = − (−1)rs +r s sin sin χr  ,s  (τ ),
τ p+ p− p+ p−
(r  ,s  )∈I1
(r, s) ∈ I1 , (5.7)
  √ 
1 2   πp− rr  πp+ ss 
r,s − = 2(−1)rs +r s cos cos r  ,s  (τ )
τ p+ p− p+ p−
(r  ,s  )∈I1
p+ −1 p− −1
 πp− rr   πp+ ss 
+ (−1)r s cos r  ,0 (τ ) + (−1)rs cos 0,s  (τ )
p+ p−
r  =1 s  =1

1 1
+ 0,0 (τ ) + (−1)p− r+p+ s p+ ,0 (τ ) , (r, s) ∈ I0 , (5.8)
2 2
  √ 
1 2   πp+ ss  πp− rr  

ρr,s − = −i  2(−1)rs +r s sin cos ϕr  ,s  (τ )
τ p+ p− (r  ,s  )∈I p− p+
1
p− −1 
πp+ ss  
rs 
+ (−1) sin ϕ0,s  (τ ) , (r, s) ∈ I , (5.9)
p−
s  =1
  √ 
1 2   πp+ ss  πp− rr  

ϕr,s − =i 2(−1)rs +r s sin cos ρr  ,s  (τ )
τ p+ p− (r  ,s  )∈I p− p+
1
p− −1 
πp+ ss  
rs 
+ (−1) sin ρ0,s  (τ ) , (r, s) ∈ I , (5.10)

p−
s =1

Table 2
The basis in C̄
Subrepresentation Dimension Basis
Rmin 1 (p − 1)(p − 1) χr,s , (r, s) ∈ I1
2 + −
Rproj 1 (p + 1)(p + 1) r,s , (r, s) ∈ I0
2 + −
 
C2 ⊗ R 2 · 12 (p+ − 1)(p− + 1) ρr,s , ϕr,s , (r, s) ∈ I
 
C2 ⊗ R 2 · 12 (p+ + 1)(p− − 1) ρr,s , ϕr,s , (r, s) ∈ I
C3 ⊗ Rmin 3 · 12 (p+ − 1)(p− − 1) ρr,s , ψr,s , ϕr,s , (r, s) ∈ I1
330 B.L. Feigin et al. / Nuclear Physics B 757 [FS] (2006) 303–343

  √ 
1 2   πp− rr  πp+ ss  

ρr,s − = −i  2(−1)rs +r s sin cos ϕr  ,s  (τ )
τ p+ p− p+ p−
(r  ,s  )∈I1
p+ −1 
πp− rr  
rs
+ (−1) sin ϕr  ,0 (τ ) , (r, s) ∈ I , (5.11)
p+
r  =1

  √ 
1 2   πp− rr  πp+ ss  

ϕr,s − =i 2(−1)rs +r s sin cos ρr  ,s  (τ )
τ p+ p− p+ p−
(r  ,s  )∈I1
p+ −1 
πp− rr  
rs
+ (−1) sin ρr  ,0 (τ ) , (r, s) ∈ I , (5.12)
p+
r  =1

  √
1 2 2   πp− rr  πp+ ss 
ρr,s − = − (−1)rs +r s sin sin ϕr  ,s  (τ ),
τ p+ p− p+ p−
(r  ,s  )∈I1
(r, s) ∈ I1 , (5.13)
  √
1 2 2   πp− rr  πp+ ss 
ψr,s − = (−1)rs +r s sin sin ψr  ,s  (τ ),
τ p+ p− p+ p−
(r  ,s  )∈I1
(r, s) ∈ I1 , (5.14)
  √
1 2 2   πp− rr  πp+ ss 
ϕr,s − = − (−1)rs +r s sin sin ρr  ,s  (τ ),
τ p+ p− p+ p−
(r  ,s  )∈I1
(r, s) ∈ I1 . (5.15)

Proof. The standard modular transformation formulas for the theta-constants are
  
−iτ −iπ rsp
2p−1
1
θs,p − = e θr,p (τ ),
τ 2p
r=0
  
−iτ −iπ rsp 
2p−1
 1
θs,p − =τ e θr,p (τ ),
τ 2p
r=0
  
−iτ −iπ rsp  2 
2p−1
 1 
θs,p − = e τ θr,p (τ ) + iπpτ θr,p (τ )
τ 2p
r=0

and the eta function transforms as


 
iπ 1 √
η(τ + 1) = e 12 η(τ ), η − = −iτ η(τ ).
τ
We next use the resummation formula (C.2). Substituting
−iπ p rs 1 θr (τ )
g(r) = e + p− , h+ (r) = √ ,
2p+ p− η(τ )
u+
+ (r, s) = r,s , u−
+ (r, s) = χr,s
B.L. Feigin et al. / Nuclear Physics B 757 [FS] (2006) 303–343 331

there, we find the transformation of the theta-constants as


2p+ p− −1
θs (− τ1 ) 1 1 −iπ p rsp
1
= e + − θ (τ )
r
η(− τ ) 2p+ p− η(τ ) r=0
 + −1
p
1 1 (−1)s πr  s
=√ 0,0 (τ ) + p+ ,0 (τ ) + cos r  ,0 (τ )
2p+ p− 2 2 
p+
r =1
p− −1
πs  s
+ cos
0,s  (τ )
p−
s  =1
 
πr  s πs  s πr  s πs  s
+2 cos cos r  ,s  (τ ) − sin sin χr  ,s  (τ ) .
 
p+ p− p+ p−
(r ,s )∈I1

Taking the sum and the difference of these formulas with s → p− r − p+ s and s → p− r + p+ s,
we obtain (5.7) and (5.8).
Substituting
−iπ p rs τ θr (τ )
g(r) = e + p− , h+ (r) =  ,
2p+ p− η(τ )
u+ 
− (r, s) = ρr,s , u− 
− (r, s) = ρr,s
in (C.2), we find the transformation of the theta-constant derivatives as

θs (− τ1 ) −2iτ  πr  s πs  s  πs  s πr  s 

= √ sin cos ρ   (τ ) + sin cos ρ   (τ )
η(− τ1 ) 2p+ p− p+ p− r ,s p− p+ r ,s
(r  ,s  )∈I1
p+ −1 p− −1 
1 πr  s  1 πs  s 
+ sin ρ  (τ ) + sin ρ  (τ ) .
2  p+ r ,0 2  p− 0,s
r =1 s =1
Taking the sum and the difference of these formulas with s → p− r − p+ s and s → p− r + p+ s,
we obtain (5.9), (5.10) and (5.11), (5.12).
For the second derivatives of the theta-constants, we have
    
1  1  1
θ p+ s+p− r − − θ p+ s−p− r −
η(− τ1 ) τ τ

2 1   πr  s πs  s
= (−1)rs +sr +1 sin sin
p+ p− η(τ ) (r  ,s  )∈I p+ p−
1
 2     
× τ θp+ s  +p− r  (τ ) − θp+ s  −p− r  (τ ) + iπp+ p− τ θp+ s  +p− r  (τ ) − θp+ s  −p− r  (τ )
whence (5.13), (5.14), and (5.15) immediately follow. 2

5.2.1.3. Lemma.
χr,s (τ + 1) = λr,s χr,s (τ ), r,s (τ + 1) = λr,s r,s (τ ), (5.16)
  
  

ρr,s (τ + 1) = λr,s ρr,s (τ ), ϕr,s (τ + 1) = λr,s ϕr,s (τ ) + ρr,s (τ ) , (5.17)
  
  

ρr,s (τ + 1) = λr,s ρr,s (τ ), ϕr,s (τ + 1) = λr,s ϕr,s (τ ) + ρr,s (τ ) , (5.18)
332 B.L. Feigin et al. / Nuclear Physics B 757 [FS] (2006) 303–343

 
ρr,s (τ + 1) = λr,s ρr,s (τ ), ψr,s (τ + 1) = λr,s ψr,s (τ ) + 2ρr,s (τ ) , (5.19)
 
ϕr,s (τ + 1) = λr,s ϕr,s (τ ) + ψr,s (τ ) + ρr,s (τ ) , (5.20)
where
iπ( 2p− r 2 + 2p+ s 2 − 12
c p p 1
λr,s = e2iπ(Δr,s − 24 ) = (−1)rs e
)
+ − . (5.21)

Proof. Elementary calculation. 2

5.2.2. It may be useful to give the formulas inverse to those in Section 5.2.1. Let I denote
the set of indices


I = (r, s)
1  r  p+ − 1, 1  s  p− − 1

(actually labeling the Kac-table boxes) and let Ī1 = I\I1 . For (r, s) ∈ Ī1 , it is convenient to set

ρr,s = ρp+ −r,p− −s , 


ρr,s = −ρp+ −r,p− −s , 
ρr,s = −ρp+ −r,p− −s ,
r,s = p+ −r,p− −s , χr,s = χp+ −r,p− −s .
Then the formulas that invert those in Section 5.2.1 are
 2 s 2 + p2 r 2 
+ 1   p+ p− rs p+ −
χr,s = ρr,s − p− rρr,s − p+ sρr,s + r,s − χr,s ,
(p+ p− )2 2 4
(r, s) ∈ I,
 
1  p+ s
χp++ ,s = ρ0,p − −s
+  0,p − −s , 1  s  p− − 1,
p+ p− 2
 
+ 1  p− r
χr,p = ρ +  −r,0 , 1  r  p+ − 1,
p+ p− p+ −r,0
− p +
2
1
χp++ ,p− = 0,0 ,
2 
1

χr,s = −ρp+ −r,s − p− rρp+ −r,s + p+ sρp+ −r,s
(p+ p− )2
2 s 2 + p 2 r 2 − (p p )2 
p+ p− rs p+ − + −
+ p+ −r,s + χp+ −r,s , (r, s) ∈ I,
2 4
 
1  p+ s
χp−+ ,s = −ρ0,s + 0,s , 1  s  p− − 1,
p+ p− 2
 
− 1  p− r
χr,p −
= −ρ r,0 +  r,0 , 1  r  p+ − 1,
p+ p− 2
1
χp−+ ,p− = p+ ,0 .
2

5.3. Modular invariants

The decomposition of the SL(2, Z) action established in Theorem 1.1.1 considerably simpli-
fies finding sesquilinear modular invariants. We illustrate this by giving several easily constructed
series.
B.L. Feigin et al. / Nuclear Physics B 757 [FS] (2006) 303–343 333

First, in the ρ  , ρ  , and ρ sectors (see Table 2), modular invariants necessarily involve τ
explicitly: they are given by
p+ −1


2

2

ρ (τ, τ̄ ) = im τ
ρr,0 (τ )
+ 2 im τ
ρr,s (τ )
,
r=1 (r,s)∈I1

a “symmetric” expression with the ,


and
ρr,s
 
ρ(τ, τ̄ ) = ρ̄r,s (τ̄ ) 8(im τ )2 ρr,s (τ ) + 4p+ p− im τ πχr,s (τ )
(r,s)∈I1
 
+ χ̄r,s (τ̄ ) 4p+ p− im τ πρr,s (τ ) + (πp+ p− )2 χr,s (τ ) .
± and χ
All these are expressed through the characters χr,s r,s in accordance with the formulas in
Section 5.2.1.
Next, in the  sector, we have the A-series invariants
+ −1
p − −1
p


2

2

2

 [A] (τ, τ̄ ) =
0,0 (τ )
+
p + ,0 (τ )

+2

r,0 (τ ) + 2
0,s (τ )
2
r=1 s=1

+4
r,s (τ )
2
(r,s)∈I1

and additional D-series invariants in the case where p− ≡ 0 mod 4:


+ −1


2 p

 [D] (τ, τ̄ ) =
0,0 (τ ) + p+ ,0 (τ )
+
r,0 (τ ) + p −r,0 (τ )
2
+
r=1


2

2
+
0,s (τ ) + 0,p (τ )
+ 2
r,s (τ ) + r,p− −s (τ )
.
− −s
2sp− −1 (r,s)∈I1
s even s even
Of course, “symmetric” D-invariants exist whenever p+ ≡ 0 mod 4. Again, all these invariants
are expressed through the characters via the formulas in Section 5.2.1.
We also note an E6 -like invariant for (p+ , p− ) = (5, 12):


2

2

2
 [E6 ] (τ, τ̄ ) =
0,1 (τ ) − 0,7 (τ )
+
0,2 (τ ) − 0,10 (τ )
+
0,5 (τ ) − 0,11 (τ )



2

2

2
+ 2
1,1 (τ ) − 1,7 (τ )
+ 2
2,1 (τ ) − 2,7 (τ )
+ 2
2,5 (τ ) − 3,1 (τ )



2

2

2
+ 2
2,2 (τ ) − 3,2 (τ )
+ 2
1,5 (τ ) − 4,1 (τ )
+ 2
1,2 (τ ) − 4,2 (τ )
.
It would be quite interesting to systematically obtain all modular invariants as quantum-group
invariants.

6. Conclusions

In the logarithmically extended (p+ , p− ) minimal models, we have constructed the chiral
vertex-operator algebra Wp+ ,p− , its irreducible representations and Verma modules, and calcu-
lated the SL(2, Z)-representation generated by the characters. The space of generalized characters
carrying that representation most probably coincides with the space of torus amplitudes. A great
deal of work remains to be done, however.
334 B.L. Feigin et al. / Nuclear Physics B 757 [FS] (2006) 303–343

Fig. 8. Sixteen subquotients, which do not suffice to build a projective Wp+ ,p− -module.

6.1. First and foremost, constructing the chiral sector of the space of states requires building
projective modules of Wp+ ,p− . Unfortunately, little is known about their structure (some indirect
but quite useful information has recently become available in [24]). One of the clues is the known
structure [34] of projective modules of the Kazhdan–Lusztig-dual quantum group gp+ ,p− , which
must be a piece of the structure of projective Wp+ ,p− -modules. That is, taking an irreducible
Wp+ ,p− -module Xr,s + (with 1  r  p − 1 and 1  s  p − 1), replacing it with the g
+ − p+ ,p− -
module Xr,s , taking the universal projective cover P+
+
r,s of the latter, and translating the result back
into the Wp+ ,p− -language, we obtain the structure of sixteen subquotients in Fig. 8, where solid
and dotted lines denote the elements of Ext1 and the symbolic notation for the modules is as in
+,  = X−
Fig. 7, i.e., • = Xr,s − +
p+ −r,s , ◦ = Xr,p− −s , and  = Xp+ −r,p− −s . However, in addition to
these subquotients and embeddings, the projective Wp+ ,p− -module Pr,s + (the universal projective
+
cover of Xr,s ) must also contain subquotients isomorphic to the minimal-model representations
Xr,s . Furthermore, the Wp+ ,p− -representation category contains the projective covers Pr,s of the
irreducible representations Xr,s , which have no projective-module counterparts in the gp+ ,p− -
representation category (the Pr,s may be interesting because of their relation to the boundary-
condition-changing operator whose 4-point correlation function gives the Cardy formula [6] for
the crossing probability).

6.2. The foregoing is directly relevant to establishing the relation between our construction,
specialized to (p+ = 3, p− = 2), and the analysis in [9,29,38,50], where a family of c = 0 models
was considered, labeled by a parameter b. The approach in [9], in principle, allows adding various
primary fields to the theory and, in this sense, is not aimed at fixing a particular chiral algebra.
The logarithmic minimal models in this paper, on the other hand, are minimal extensions of the
rational (p+ , p− ) models, minimal in the sense that consistent “intermediate” models—with a
subalgebra of Wp+ ,p− but with a finite number of fields—do not seem to exist.6 There are various
ways to construct “larger” logarithmic extensions of minimal models, for example, by taking the
kernel not of two but of one screening (cf. [17]). In our “minimal” setting, in particular, the W3,2 -
primary fields are just those whose dimensions are in Table 1, and similarly for all the (p+ , p− )
models, as stated in Theorem 1.1.1. (We note once again that the Virasoro field content, which

6 To avoid misunderstanding, we note that logarithmic models may of course be constructed based on different
W -algebras, overlapping with Wp+ ,p− only over the Virasoro algebra.
B.L. Feigin et al. / Nuclear Physics B 757 [FS] (2006) 303–343 335

appears to have been the subject of some discussion in other approaches, then follows uniquely
from the decompositions in Lemma 2.3.2—in fact, from the structure of the relevant complexes.
In particular, the number of Virasoro primary fields is not limited to an integer multiple of either
the standard (p+ − 1) × (p− − 1) or the extended p+ × p− Kac table.) In the setting in this paper,
fixing the value of b or another similar parameter requires constructing projective W -modules,
which will be addressed elsewhere (among other things, it will completely settle the “logarithmic
partner” issues).

6.3. The SL(2, Z)-representation on the generalized characters (presumably, torus ampli-
tudes) coincides with the SL(2, Z)-representation on the center of the Kazhdan–Lusztig-dual
quantum group gp+ ,p− [34]. This remarkable correspondence deserves further study, as do other
aspects of the Kazhdan–Lusztig correspondence.
The Kazhdan–Lusztig correspondence suggests that in some generality, the space of torus
amplitudes C and the conformal field theory center Zcft are related by conformal-field-theory
analogues of the Radford and Drinfeld maps known in the theory of quantum groups. Then, under
the identification of Zcft with the space of boundary conditions preserved by the corresponding
W -algebra, the images of irreducible characters under the “W -Radford map” are the Ishibashi
states and the images under the “W -Drinfeld map” are the Cardy states.
We also give a fusion algebra suggested by the Kazhdan–Lusztig correspondence. This fusion
is the gp+ ,p− Grothendieck ring G, with the preferred basis of 2p+ p− irreducible representa-
tions. In terms of the Wp+ ,p− algebra, we identify the preferred basis elements with the 2p+ p−
± defined in (3.21). For all 1  r, r   p , 1  s, s   p , and α, β = ±, we
representations Kr,s + −
thus have the algebra
 
β
−1
r+r −1
s+s
Kr,s
α
Kr  ,s  = K̃u,v
αβ
, (6.1)
u=|r−r  |+1 v=|s−s  |+1
step=2 step=2

where
⎧ α

⎪ Kr,s , 1  r  p+ , 1  s  p− ,

⎪ −α

⎨ K2p+ −r,s + 2Kr−p+ ,s , p+ + 1  r  2p+ − 1, 1  s  p− ,
α

K̃r,s
α
= Kr,2p −α
α
− −s
+ 2Kr,s−p , 1  r  p+ , p− + 1  s  2p− − 1,




⎪ Kα −α −α
+ 2K2p+ −r,s−p− + 2Kr−p + 4Kr−p
α

⎪ ,
⎩ 2p+ −r,2p− −s + ,2p− −s + ,s−p−
p+ + 1  r  2p+ − 1, p− + 1  s  2p− − 1.
This algebra has several noteworthy properties:
+ +
(1) it is generated by two elements K1,2 and K2,1 ;
(2) its radical is generated by the algebra action on Kp++ ,p− ; the quotient over the radical coin-
cides with the fusion of the (p+ , p− ) Virasoro minimal models;
+
(3) K1,1 is the identity;
− − α = K−α .
(4) K1,1 acts as a simple current, K1,1 Kr,s r,s

A functor between the representation categories of Wp+ ,p− and gp+ ,p− is not yet known, and
± is based on in-
the identification of the basis of irreducible gp+ ,p− -representation with the Kr,s
direct arguments. First, it is clear that the minimal-model representations Xr,s act by zero on the
336 B.L. Feigin et al. / Nuclear Physics B 757 [FS] (2006) 303–343

other basis elements in the fusion algebra, simply because of the vanishing of three-point func-
tions involving two minimal-model representations and one representation on which the ideal R
acts nontrivially. We next recall that the 2p+ p− representations Kr,s ± are defined as kernels of the

screenings. Some of them are reducible, see (3.22), “inasmuch as” the Felder complexes have
a nonzero cohomology (the minimal-model representations Jr,s ≡ Xr,s ). Their role as counter-
parts of the irreducible gp+ ,p− -representations under the Kazhdan–Lusztig correspondence is
supported by the numerical evidence in [24]: decomposing Kr,s ± into the Virasoro modules (see

Lemma 3.5.2) shows that the (3, 2) and (5, 2) results in [24] agree with (6.1). Needless to say, it
would be quite interesting to properly define the Wp+ ,p− -fusion and establish (6.1) by some “rel-
atively direct” (or just numerical, as the first step) calculation of the Wp+ ,p− (not just Virasoro)
coinvariants.7
We also note that the quantum group gp+ ,p− can be quite useful in constructing the full
(holomorphic + antiholomorphic) space of states, by taking gp+ ,p− -invariants in the product of
(Wp+ ,p− , gp+ ,p− )-bimodules.

Acknowledgements

We are grateful to A. Belavin, J. Fuchs, A. Isaev, S. Parkhomenko, and P. Pyatov for the useful
discussions and comments. This paper was supported in part by the RFBR Grant 04-01-00303
and the RFBR-JSPS Grant 05-01-02934YaF_a. The work of AMG, AMS and IYuT is supported
in part by the LSS-4401.2006.2 grant. The work of IYuT was supported in part by the RFBR
Grant 05-02-17217 and the “Dynasty” foundation.

Appendix A. (3, 2)-model examples

In the simplest case of the (3, 2) model, the minimal-model character is trivial, χ1,1 (q) = 1,
and 12 nontrivial characters are expressed through theta-constants in accordance with Eqs. (5.2)–
±
(5.3). Explicitly, for example, the characters of X1,1 are
+
q −2 χ1,1 (q) = 1 + q + 2q 2 + 2q 3 + 4q 4 + 4q 5 + 7q 6 + 8q 7 + 12q 8 + 14q 9
+ 21q 10 + 24q 11 + 34q 12 + 44q 13 + 58q 14 + 72q 15 + · · ·
+ +
(with the character of K1,1 given by 1 + χ1,1 (q)), and

q −7 χ1,1 (q) = 2 + 2q + 4q 2 + 6q 3 + 10q 4 + 12q 5 + 20q 6 + 26q 7 + 36q 8 + 48q 9
+ 66q 10 + 84q 11 + 114q 12 + 144q 13 + 188q 14 + 240q 15 + · · ·
(the dimensions of the respective highest-weight vectors are Δ1,1;1 = 2 and Δ2,1;−2 = 7). The
modular transformation properties of the characters follow by expressing them through the basis
in Table 2 (where now |I1 | = 1, |I0 | = 6, |I | = 2, and |I | = 3) in accordance with the for-
mulas in Section 5.2.2 and using Lemma 5.2.1.2. Under the SL(2, Z) action, the 13 characters
(including χ1,1 ) give rise to the dimension-20 space of generalized characters. To give examples
of modular transformations, we use Eqs. (5.6) and express the S-transformed characters through

7 The fusion algebra in (6.1) also follows from the above SL(2, Z) action on the characters via a procedure generalizing
the Verlinde formula, similar to that in [21]; the (somewhat bulky) details will be given elsewhere.
B.L. Feigin et al. / Nuclear Physics B 757 [FS] (2006) 303–343 337

the characters with τ -dependent coefficients:


   √   
+ 1 13 i( 3 + 18π)τ τ2 iτ τ2 −
χ1,1 − = − + − χ1,1 (τ ) + √ − χ1,1 (τ )
τ 144 108 144 6 3 3
   
1 iτ − iτ τ2 −
+ √ + χ1,2 (τ ) − √ + χ2,1 (τ )
12 3 6 6 3 6
 
iτ 1 − iτ − 1 −
+ − √ χ2,2 (τ ) − √ χ3,1 (τ ) − √ χ3,2 (τ )
12 12 3 6 3 12 3
 2     
τ iτ + 1 iτ + iτ τ2 +
+ − √ χ1,1 (τ ) − √ + χ1,2 (τ ) + √ + χ2,1 (τ )
3 6 3 12 3 6 6 3 6
 
1 iτ + iτ + 1 +
+ √ − χ2,2 (τ ) + √ χ3,1 (τ ) + √ χ3,2 (τ )
12 3 12 6 3 12 3
and
   √   2 
− 1 23 i( 3 + 18π)τ τ2 τ iτ −
χ1,1 − = − − + χ1,1 (τ ) + − √ χ1,1 (τ )
τ 144 108 144 3 6 3
   
1 iτ − iτ τ2 −
+ √ + χ1,2 (τ ) + √ + χ2,1 (τ )
12 3 6 6 3 6
 
iτ 1 iτ − 1
+ − √ χ − (τ ) + √ χ3,1 (τ ) − √ χ3,2 −
(τ )
12 12 3 2,2 6 3 12 3
     
iτ τ2 + 1 iτ + iτ τ2 +
+ √ − χ1,1 (τ ) − √ + χ1,2 (τ ) − √ + χ2,1 (τ )
6 3 3 12 3 6 6 3 6
 
1 iτ + iτ + 1 +
+ √ − χ2,2 (τ ) − √ χ3,1 (τ ) + √ χ3,2 (τ ).
12 3 12 6 3 12 3
We next consider fusion relations (6.1). To write them explicitly for (p+ , p− ) = (3, 2), we
− α K− = K−α , and therefore the entire 12 × 12 multiplication table
recall that K1,1 acts as Kr,s 1,1 r,s
+
essentially reduces to its 6 × 6 block, where (recalling that K1,1 acts as identity) the 12 · 5 · 6
independent relations are
+ + − + + + + + + − +
K1,2 K1,2 = 2K1,1 + 2K1,1 , K1,2 K2,1 = K2,2 , K1,2 K2,2 = 2K2,1 + 2K2,1 ,
+ + + + + − +
K1,2 K3,1 = K3,2 , K1,2 K3,2 = 2K3,1 + 2K3,1 ,
+ + + + + + + + + + − +
K2,1 K2,1 = K1,1 + K3,1 , K2,1 K2,2 = K1,2 + K3,2 , K2,1 K3,1 = 2K1,1 + 2K2,1 ,
+ + − +
K2,1 K3,2 = 2K1,2 + 2K2,2 ,
+ + − − + + + + − +
K2,2 K2,2 = 2K1,1 + 2K3,1 + 2K1,1 + 2K3,1 , K2,2 K3,1 = 2K1,2 + 2K2,2 ,
+ + − − + +
K2,2 K3,2 = 4K1,1 + 4K2,1 + 4K1,1 + 4K2,1 ,
+ + − + + + + − + +
K3,1 K3,1 = 2K2,1 + 2K1,1 + K3,1 , K3,1 K3,2 = 2K2,2 + 2K1,2 + K3,2 ,
+ + − − − + + +
K3,2 K3,2 = 4K1,1 + 4K2,1 + 2K3,1 + 4K1,1 + 4K2,1 + 2K3,1 .
Finally, more as a curiosity than for any practical purposes, we give explicit free-field expres-
sions for the W3,2 -algebra generators in (4.1) (arbitrarily normalized):
338 B.L. Feigin et al. / Nuclear Physics B 757 [FS] (2006) 303–343


+ 35  4 2 56 5 3 28 8 280  2
W = ∂ ϕ + ∂ ϕ∂ ϕ + ∂ 6 ϕ∂ 2 ϕ + ∂ 7 ϕ∂ϕ − √ ∂ 3 ϕ ∂ 2 ϕ
27 27 27 27 9 3
70 4  2 2 280 4 3 56 5 2 28
− √ ∂ ϕ ∂ ϕ − √ ∂ ϕ∂ ϕ∂ϕ − √ ∂ ϕ∂ ϕ∂ϕ − √ ∂ 6 ϕ(∂ϕ)2
3 3 9 3 3 3 9 3
35  2 4 280 3  2 2 280  3 2 140
+ ∂ ϕ + ∂ ϕ ∂ ϕ ∂ϕ + ∂ ϕ (∂ϕ)2 + ∂ 4 ϕ∂ 2 ϕ(∂ϕ)2
3 3 9 3
56 140  3 560 70
+ ∂ 5 ϕ(∂ϕ)3 − √ ∂ 2 ϕ (∂ϕ)2 − √ ∂ 3 ϕ∂ 2 ϕ(∂ϕ)2 − √ ∂ 4 ϕ(∂ϕ)4
9 3 3 3 3 3
 √
 2 2 56 28 1
+ 70 ∂ ϕ (∂ϕ) + ∂ ϕ(∂ϕ) − √ ∂ ϕ(∂ϕ) + (∂ϕ) − √ ∂ ϕ e2 3ϕ ,
4 3 5 2 6 8 8
3 3 27 3

217  5 2 2653 6 4 23 7 3 11 8 2 1 9
W− = ∂ ϕ − ∂ ϕ∂ ϕ − ∂ ϕ∂ ϕ − ∂ ϕ∂ ϕ − ∂ ϕ∂ϕ
192 3456 384 1152 768
1225  2 13475  4 2 2 2695
− √ ∂ 4ϕ ∂ 3ϕ − √ ∂ ϕ ∂ ϕ + √ ∂ 5 ϕ∂ 3 ϕ∂ 2 ϕ
64 3 576 3 64 3
2555 5 4 2891 6  2 2 1351 6 3
+ √ ∂ ϕ∂ ϕ∂ϕ − √ ∂ ϕ ∂ ϕ − √ ∂ ϕ∂ ϕ∂ϕ
192 3 576 3 192 3
103 7 2 13 3535  3 2  2 2
− √ ∂ ϕ∂ ϕ∂ϕ − √ ∂ 8 ϕ(∂ϕ)2 + ∂ ϕ ∂ ϕ
192 3 384 3 32
735  3 3 3395 4  2 3 245 4 3 2
− ∂ ϕ ∂ϕ − ∂ ϕ ∂ ϕ + ∂ ϕ∂ ϕ∂ ϕ∂ϕ
16 54 24
12635  4 2 245 5  2 2 105 5 3
+ ∂ ϕ (∂ϕ)2 + ∂ ϕ ∂ ϕ ∂ϕ + ∂ ϕ∂ ϕ(∂ϕ)2
576 12 32
2443 6 2 19 13405  2 5 8225  3
− ∂ ϕ∂ ϕ(∂ϕ)2 − ∂ 7 ϕ(∂ϕ)3 − √ ∂ ϕ + √ ∂ 3 ϕ ∂ 2 ϕ ∂ϕ
288 96 144 3 24 3

105 3  3 2 2 665  2 245
− ∂ ϕ ∂ ϕ(∂ϕ)2 + √ ∂ 4 ϕ ∂ 2 ϕ (∂ϕ)2 + √ ∂ 4 ϕ∂ 3 ϕ(∂ϕ)3
4 24 3 2 3
245 5 2 91 16205  4
− √ ∂ ϕ∂ ϕ(∂ϕ)3 − √ ∂ 6 ϕ(∂ϕ)4 + ∂ 2 ϕ (∂ϕ)2
8 3 24 3 144
385 3  2 2 525  3 2 35
+ ∂ ϕ ∂ ϕ (∂ϕ)3 + ∂ ϕ (∂ϕ)4 + ∂ 4 ϕ∂ 2 ϕ(∂ϕ)4 − 7∂ 5 ϕ(∂ϕ)5
4 8 3

665  3 105 3 3 2 35
+ √ ∂ 2 ϕ (∂ϕ)4 + ∂ ϕ∂ ϕ(∂ϕ)5 − √ ∂ 4 ϕ(∂ϕ)6
3 3 2 3 3
455  2 2 25
+ ∂ ϕ (∂ϕ)6 + 5∂ 3 ϕ(∂ϕ)7 + √ ∂ 2 ϕ(∂ϕ)8 + (∂ϕ)10
6 3
 √
1
− √ ∂ 10 ϕ e−2 3ϕ ,
13824 3
where, despite the brackets introduced for the compactness

of notation, the nested normal or-
4 2 2
dering is from right to left, e.g., ∂ ϕ(∂ ϕ(∂ ϕe 2 3ϕ )). These dimension-15 operators have the
OPE
T (w)
W + (z)W − (w) = 27 · 3 · 53 · 72 · 11 · 17 + ···,
(z − w)28
B.L. Feigin et al. / Nuclear Physics B 757 [FS] (2006) 303–343 339

with the energy–momentum tensor given by (2.3), that is,


1 1
T (z) = ∂ϕ(z)∂ϕ(z) − √ ∂ 2 ϕ(z).
2 2 3
In the (3, 2) model, the minimal-model vertex-operator algebra Mp+ ,p− is trivial, or in other
words, T (z) is in the ideal R.

Appendix B. Quantum-group technicalities

B.1. Drinfeld double of a quantum group

We recall [45,46] that the space H ∗ of linear functions on a Hopf algebra H is a Hopf algebra
with the multiplication, comultiplication, unit, counit, and antipode given by
 

βγ , x =
β, x 
γ , x  , Δ(β), x ⊗ y =
β, yx ,
(x)
   

1, x = (x), (β) =
β, 1 , S(β), x = β, S −1 (x) (B.1)
for any β, γ ∈ H ∗ and x, y ∈ H . The Drinfeld double [45,46] D(H ) is a Hopf algebra with the
underlying vector space H ∗ ⊗ H and with the multiplication, comultiplication, unit, counit, and
antipode given (in addition to the formulas for H and H ∗ ) by
 
xβ = β S −1 (x  )?x  x  , x ∈ H, β ∈ H ∗ . (B.2)
(x)

B.2. Proof of Theorem 3.3.1

By induction, it is easy to see that the comultiplication in the PBW basis in H is given by

Δ(ej mn )
m n    
m n p r(r−m) p+ s(s−n)
= q − q− ej +2p− (m−r)−2p+ (n−s),r,s ⊗ ej,m−r,n−s .
r + s − +
r=0 s=0
(B.3)
With (3.5), we define κ, f+ , e− ∈ H∗ by the relations


κ, ej mn = δm,0 δn,0 qj ,
j −j
q+ q−

f+ , ej mn = −δm,1 δn,0 p −p− ,
e− , ej mn = −δm,0 δn,1 p −p+ (B.4)
q+− − q+ q−+ − q−
and then follow the standard step-by-step construction of a Drinfeld double, based on Eqs. (B.1)
and (B.2), which now become
     
kβ = β k −1 ?k k, e+ β = β(?e+ )1 + β ?k 2p− e+ − β e+ k −2p− ?k 2p− k 2p− ,
   
f− β = β(?f− )1 + β ?k −2p+ f− − β f− k 2p+ ?k −2p+ k −2p+ . (B.5)
340 B.L. Feigin et al. / Nuclear Physics B 757 [FS] (2006) 303–343

We here use (B.3). The following formulas are then obtained by direct calculation:
   
κ k −1 ?k = κ, κ(?e+ ) = 0, κ ?k 2p− = q+ κ,
   
κ e+ k −2p− ?k 2p− = 0, κ(?f− ) = 0, κ ?k −2p+ = q−1
− κ,
 −2p

2p
κ f− k ?k + + = 0,
 −1  −κ 2p−  
f+ k ?k = q−1 + f + , f + (?e + ) = p− −p− , f+ ?k 2p− = f+ ,
q + − q+
  −1  
f+ e+ k −2p− ?k 2p− = p− −p− , f+ (?f− ) = 0, f+ ?k −2p+ = f+ ,
q + − q+
 −2p

2p
f+ f− k ?k + + = 0,
   
e− k −1 ?k = q− e− , e− (?e+ ) = 0, e− ?k 2p− = e− ,
  −κ −2p+  
e− e+ k −2p− ?k 2p− = 0, e− (?f− ) = p+ −p+ , e− ?k −2p+ = e− ,
q − − q−
  −1
e− f− k 2p+ ?k −2p+ = p+ −p .
q − − q− +
Applying the first relation in (B.1) to (B.3), we subsequently obtain
 a 
κ , ej mn = qaj δm0 δn0 ,
 a  [a]+ ! aj +p− a(a−1)
f+ , ej mn = (−1)a δma δn0 p− −p− a q+
2
,
(q+ − q+ )
 a  [a]− ! −aj +p+ a(a−1)
e− , ej mn = (−1)a δm0 δna p+ −p+ a q−
2
,
(q− − q− )
  (−1)a+b [a]+ ![b]− ! aj +p− a(a−1) −bj +p+ b(b−1)
f+a e− κ , ej mn = δma δnb
b c
q
b +
2
q− 2
qj c .
p −p− a p −p+
(q+− − q+ ) (q−+ − q− )
2 p 2 elements {f a eb κ c } with 0  a  p − 1,
It is now straightforward to prove that the 4p+ − + − +
0  b  p− − 1, and 0  c  4p+ p− − 1 are linearly independent (cf. [22]) and that the relations
claimed in the theorem are indeed satisfied.

Appendix C. Summation over 2p+ p− consecutive values

Here, we isolate elementary but bulky formulas needed in the derivation of modular transfor-
mations. For any f satisfying f (r + 2p+ p− ) = f (r), we have the obvious identity
2p+ p− −1 p+ −1 p− −1 p+ −1 p−

f (r) = f (p− r  + p+ s  ) + f (p− r  − p+ s  )
r=0 r  =0 s  =0 r  =0 s  =1
+ −1 p
p − −1
 
= f (0) + f (−p+ p− ) + f (p− r  + p+ s  ) + f (p− r  − p+ s  )
r  =1 s  =1
+ −1
p p− −1
   
+ f (p− r  ) + f (−p− r  ) + f (p+ s  ) + f (−p+ s  ) .
r  =1 s  =1
B.L. Feigin et al. / Nuclear Physics B 757 [FS] (2006) 303–343 341

Next, for f (r) = g(r)h± (r), where g(r + 2p+ p− ) = g(r), h± (r + 2p+ p− ) = h± (r), and in
addition h± (−r) = ±h± (r), we have
p− −1
2p+
g(r)h± (r)
r=0
= g(0)h± (0) + g(−p+ p− )h± (−p+ p− )
 
+ g(p− r  + p+ s  ) ± g(−p− r  − p+ s  ) h± (p− r  + p+ s  )
(r  ,s  )∈I1
  
+ g(p− r  − p+ s  ) ± g(−p− r  + p+ s  ) h± (p− r  − p+ s  )
+ −1
p − −1
p
   
+ g(p− r  ) ± g(−p− r  ) h± (p− r  ) + g(p+ s  ) ± g(−p+ s  ) h± (p− s  ).
r  =1 s  =1
(C.1)
In terms of the combinations

u+
± (r, s) = h± (p− r + p+ s) + h± (p− r − p+ s),
u−
± (r, s) = h± (p− r + p+ s) − h± (p− r − p+ s), (r, s) ∈ I1 ,
Eq. (C.1) is written as
p− −1
2p+
g(r)h± (r)
r=0
= g(0)h± (0) + g(−p+ p− )h± (−p+ p− )
1 
+ g(p− r  + p+ s  ) ± g(−p− r  − p+ s  )
2  
(r ,s )∈I1

+ g(p− r  − p+ s  ) ± g(−p− r  + p+ s  ) u+  
± (r , s )

+ g(p− r  + p+ s  ) ± g(−p− r  − p+ s  )
 
− g(p− r  − p+ s  ) ∓ g(−p− r  + p+ s  ) u−  
± (r , s )
+ −1
p − −1
p
   
+ g(p− r  ) ± g(−p− r  ) h± (p− r  ) + g(p+ s  ) ± g(−p+ s  ) h± (p− s  ).
r  =1 s  =1
(C.2)
This rearrangement of the sum of 2p+ p− -periodic functions over 2p+ p− consecutive values is
an efficient way to find modular transformations of the (p+ , p− )-model characters.

References

[1] V. Gurarie, M. Flohr, C. Nayak, The Haldane–Rezayi quantum Hall state and conformal field theory, Nucl. Phys.
B 498 (1997) 513–538, cond-mat/9701212.
[2] F.H.L. Essler, H. Frahm, H. Saleur, Continuum limit of the integrable sl(2/1) 3–3̄ superspin chain, Nucl. Phys.
B 712 (2005) 513–572, cond-mat/0501197.
[3] H. Saleur, Lectures on non-perturbative field theory and quantum impurity problems, cond-mat/9812110.
[4] G. Piroux, P. Ruelle, Pre-logarithmic and logarithmic fields in a sandpile model, J. Stat. Mech. 0401 (2004) P005,
hep-th/0407143.
342 B.L. Feigin et al. / Nuclear Physics B 757 [FS] (2006) 303–343

[5] M. Jeng, Conformal field theory correlations in the Abelian sandpile model, Phys. Rev. E 71 (2005) 016140, cond-
mat/0407115.
[6] J. Cardy, Critical percolation in finite geometries, J. Phys. A 25 (1992) L201–L206, hep-th/9111026.
[7] G.A.M. Watts, A crossing probability for critical percolation in two dimensions, J. Phys. A 29 (1996) L363, cond-
mat/9603167.
[8] H. Saleur, B. Duplantier, Exact determination of the percolation hull exponent in two dimensions, Phys. Rev. Lett. 58
(1987) 2325–2328.
[9] V. Gurarie, A.W.W. Ludwig, Conformal field theory at central charge c = 0 and two-dimensional critical systems
with quenched disorder, hep-th/00409105.
[10] Y.-Z. Huang, J. Lepowsky, L. Zhang, A logarithmic generalization of tensor product theory for modules for a vertex
operator algebra, math.QA/0311235.
[11] M. Miyamoto, Modular invariance of vertex operator algebras satisfying C2 -cofiniteness, math.QA/0209101.
[12] J. Fuchs, On non-semisimple fusion rules and tensor categories, hep-th/0602051.
[13] H.G. Kausch, Extended conformal algebras generated by a multiplet of primary fields, Phys. Lett. B 259 (1991)
448.
[14] M.R. Gaberdiel, H.G. Kausch, A rational logarithmic conformal field theory, Phys. Lett. B 386 (1996) 131–137,
hep-th/9606050.
[15] M.R. Gaberdiel, H.G. Kausch, A local logarithmic conformal field theory, Nucl. Phys. B 538 (1999) 631–658,
hep-th/9807091.
[16] H.G. Kausch, Symplectic fermions, Nucl. Phys. B 583 (2000) 513–541, hep-th/0003029.
[17] J. Fjelstad, J. Fuchs, S. Hwang, A.M. Semikhatov, I.Yu. Tipunin, Logarithmic conformal field theories via logarith-
mic deformations, Nucl. Phys. B 633 (2002) 379–413, hep-th/0201091.
[18] B.L. Feigin, A.M. Gainutdinov, A.M. Semikhatov, I.Yu. Tipunin, Kazhdan–Lusztig correspondence for the repre-
sentation category of the triplet W-algebra in logarithmic CFT, math.QA/0512621.
[19] M. Flohr, Bits and pieces in logarithmic conformal field theory, Int. J. Mod. Phys. A 18 (2003) 4497–4592, hep-
th/0111228.
[20] M.R. Gaberdiel, H.G. Kausch, Indecomposable fusion products, Nucl. Phys. B 477 (1996) 293–318, hep-
th/9604026.
[21] J. Fuchs, S. Hwang, A.M. Semikhatov, I.Yu. Tipunin, Nonsemisimple fusion algebras and the Verlinde formula,
Commun. Math. Phys. 247 (2004) 713–742, hep-th/0306274.
[22] B.L. Feigin, A.M. Gainutdinov, A.M. Semikhatov, I.Yu. Tipunin, Modular group representations and fusion in
logarithmic conformal field theories and in the quantum group center, Commun. Math. Phys. 265 (2006) 47–93,
hep-th/0504093.
[23] N. Carqueville, M. Flohr, Nonmeromorphic operator product expansion and C2 -cofiniteness for a family of
W-algebras, J. Phys. A 39 (2006) 951–966, math-ph/0508015.
[24] H. Eberle, M. Flohr, Virasoro representations and fusion for general augmented minimal models, hep-th/0604097.
[25] G. Felder, BRST approach to minimal models, Nucl. Phys. B 317 (1989) 215–236.
[26] T. Kerler, V.V. Lyubashenko, Non-semisimple Topological Quantum Field Theories for 3-Manifolds with Corners,
Springer Lecture Notes in Mathematics, vol. 1765, Springer-Verlag, 2001.
[27] M. Miyamoto, A theory of tensor products for vertex operator algebra satisfying C2 -cofiniteness,
math.QA/0309350.
[28] M. Flohr, M.R. Gaberdiel, Logarithmic torus amplitudes, hep-th/0509075.
[29] V. Gurarie, A.W.W. Ludwig, Conformal algebras of 2d disordered systems, J. Phys. A: Math. Gen. 35 (2002) L377–
L384, cond-mat/9911392.
[30] I.I. Kogan, A. Nichols, Stress energy tensor in LCFT and the logarithmic Sugawara construction, Int. J. Mod. Phys.
A 18 (2003) 4771–4788, hep-th/0112008.
[31] I.I. Kogan, A. Nichols, Stress energy tensor in c = 0 logarithmic conformal field theory, hep-th/0203207.
[32] M. Flohr, A. Müller-Lohmann, Notes on non-trivial and logarithmic CFTs with c = 0, J. Stat. Mech. 0604 (2006)
P002, hep-th/0510096.
[33] C. Dong, H. Li, G. Mason, Twisted representations of vertex operator algebras, Math. Ann. 310 (1998) 571–600.
[34] B.L. Feigin, A.M. Gainutdinov, A.M. Semikhatov, I.Yu. Tipunin, Kazhdan–Lusztig-dual quantum group for loga-
rithmic extensions of Virasoro minimal models, math.QA/0606506.
[35] B. Nienhuis, Critical behavior of two-dimensional spin models and charge asymmetry in the Coulomb gas, J. Stat.
Phys. 34 (1984) 731–761.
[36] Vl.S. Dotsenko, V.A. Fateev, Conformal algebra and multipoint correlation functions in 2D statistical models, Nucl.
Phys. B 240 (1984) 312–348.
B.L. Feigin et al. / Nuclear Physics B 757 [FS] (2006) 303–343 343

[37] A.B. Zamolodchikov, ‘Irreversibility’ of the flux of the renormalization group in a 2D field theory, JETP Lett. 43
(1986) 730–732.
[38] J. Cardy, The stress tensor in quenched random systems, cond-mat/0111031.
[39] B.L. Feigin, D.B. Fuchs, Representations of Infinite-Dimensional Lie Groups and Lie Algebras, Gordon and Breach,
New York, 1989.
[40] B.L. Feigin, D.B. Fuks, Verma modules over the Virasoro algebra, Funct. Anal. Appl. 17 (1983) 241.
[41] V. Kac, Vertex Operator Algebras for Beginners, University Lecture Series, vol. 10, American Mathematical Society,
Boston, 1998.
[42] I.B. Frenkel, Y.-Z. Huang, J. Lepowsky, On axiomatic approach to vertex operator algebras and modules, Memoirs
Amer. Math. Soc. 104 (1989).
[43] J. Lepowsky, H. Li, Introduction to Vertex Operator Algebras and Their Representations, Progress in Mathematics,
Birkhäuser, 2004.
[44] P. Bouwknegt, J. McCarthy, K. Pilch, Fock space resolutions of the Virasoro highest weight modules with c  1,
Lett. Math. Phys. 23 (1991) 193–204, hep-th/9108023.
[45] C. Kassel, Quantum Groups, Springer-Verlag, New York, 1995.
[46] V. Chari, A. Pressley, A Guide to Quantum Groups, Cambridge Univ. Press, 1994.
[47] B.L. Feigin, I.Yu. Tipunin, unpublished.
[48] A. Rocha-Caridi, Vacuum vector representations of the Virasoro algebra, in: Vertex Operators in Mathematics and
Physics, in: MSRI Publications, vol. 3, Springer, Heidelberg, 1984, pp. 451–473.
[49] A. Nichols, Extended chiral algebras and the emergence of SU(2) quantum numbers in the Coulomb gas, hep-
th/0302075.
[50] J. Cardy, Logarithmic correlations in quenched random magnets and polymers, cond-mat/9911024.
Nuclear Physics B 757 (2006) 344–346

CUMULATIVE AUTHOR INDEX B751–B757

Abdalla, E. B752 (2006) 40 Calcagni, G. B752 (2006) 404


Allés, B. B752 (2006) 124 Callin, P. B752 (2006) 60
Álvarez, E. B756 (2006) 148 Chen, C.-M. B751 (2006) 260
Álvarez-Gaumé, L. B753 (2006) 92 Cieza Montalvo, J.E. B756 (2006) 1
Ananth, S. B753 (2006) 195 Cirelli, M. B753 (2006) 178
Andrews, R.P. B751 (2006) 304 Cirigliano, V. B752 (2006) 18
Aoyama, T. B754 (2006) 48 Cirigliano, V. B753 (2006) 139
Aulakh, C.S. B757 (2006) 47 Clark, S.S. B756 (2006) 38
Cornou, J.L. B756 (2006) 16
Bailin, D. B755 (2006) 79 Cortez Jr., N.V. B756 (2006) 1
Baseilhac, P. B754 (2006) 309 Cuadros-Melgar, B. B752 (2006) 40
Basu-Mallick, B. B757 (2006) 280 Czakon, M. B751 (2006) 1
Baulieu, L. B753 (2006) 252 Czakon, M. B755 (2006) 221
Baulieu, L. B753 (2006) 273
Becker, K. B751 (2006) 108
Danilov, G.S. B754 (2006) 187
Becker, M. B751 (2006) 108
Dasgupta, K. B755 (2006) 21
Beneke, M. B751 (2006) 160
de Carlos, B. B752 (2006) 404
Berenstein, D. B753 (2006) 69
De Felice, A. B752 (2006) 404
Bhattacharjee, P. B752 (2006) 280
De Freitas, A. B755 (2006) 272
Bietenholz, W. B754 (2006) 17
Delduc, F. B753 (2006) 211
Blas, D. B756 (2006) 148
D’Elia, M. B752 (2006) 124
Blumenhagen, R. B751 (2006) 186
Deppisch, F. B752 (2006) 80
Blümlein, J. B755 (2006) 112
D’Hoker, E. B753 (2006) 16
Blümlein, J. B755 (2006) 272
Bogdan, A.V. B757 (2006) 211 D’Hoker, E. B757 (2006) 79
Bolognesi, S. B752 (2006) 93 Diaconescu, D.-E. B752 (2006) 329
Bolognesi, S. B754 (2006) 293 Dijkgraaf, R. B752 (2006) 329
Bondyopadhaya, N. B757 (2006) 280 Dobashi, S. B756 (2006) 171
Bornyakov, V.G. B756 (2006) 71 Dolgov, A.D. B752 (2006) 297
Bossard, G. B753 (2006) 252 Donagi, R. B752 (2006) 329
Bossard, G. B753 (2006) 273 Dorey, N. B751 (2006) 304
Boughezal, R. B755 (2006) 221
Boyko, P.Yu. B756 (2006) 71 Ecker, G. B753 (2006) 139
Brandenburg, A. B754 (2006) 107 Eguchi, H. B752 (2006) 1
Brandt, F.T. B754 (2006) 146 Eidemüller, M. B753 (2006) 139
Brink, L. B753 (2006) 195 Eiras, D. B757 (2006) 197
Brümmer, F. B755 (2006) 186 Enciso, A. B751 (2006) 376
Burgess, C.P. B752 (2006) 60 Engquist, J. B752 (2006) 206
Burrington, B.A. B757 (2006) 1 Estes, J. B753 (2006) 16

0550-3213/2006 Published by Elsevier B.V.


doi:10.1016/S0550-3213(06)00829-7
Nuclear Physics B 757 (2006) 344–346 345

Estes, J. B757 (2006) 79 Ji, X. B753 (2006) 42


Eto, M. B752 (2006) 140 Jing, J. B755 (2006) 313

Fehér, L. B751 (2006) 436 Kaiser, R. B753 (2006) 139


Feigin, B.L. B757 (2006) 303 Katz, S. B755 (2006) 21
Feng, B. B754 (2006) 351 Kitazawa, N. B755 (2006) 254
Forgács, P. B751 (2006) 390 Klein, S. B755 (2006) 272
Fornengo, N. B753 (2006) 178 Knauf, A. B755 (2006) 21
Fré, P. B751 (2006) 343 Koike, Y. B752 (2006) 1
Frenkel, J. B754 (2006) 146 Konstandin, T. B757 (2006) 172
Fu, J.-X. B751 (2006) 108 Kosmas, T.S. B752 (2006) 80
Kovalenko, A.V. B756 (2006) 71
Gaillard, M.K. B751 (2006) 75 Kozlovskii, M.P. B753 (2006) 242
Gainutdinov, A.M. B757 (2006) 303 Kühböck, H. B754 (2006) 1
Garavuso, R.S. B755 (2006) 329 Kühn, J.H. B752 (2006) 327
Garg, S.K. B757 (2006) 47
Garriga, J. B756 (2006) 148 Lavignac, S. B755 (2006) 137
Geyer, B. B755 (2006) 112 Lavoura, L. B754 (2006) 1
Giudice, G.F. B757 (2006) 19 Li, J.-r. B754 (2006) 351
Giusto, S. B754 (2006) 233 Li, M. B755 (2006) 286
Glück, M. B754 (2006) 178 Li, T. B751 (2006) 260
Gluza, J. B751 (2006) 1 Lipatov, L.N. B754 (2006) 187
Grabovsky, A.V. B757 (2006) 211 Liu, J.T. B757 (2006) 1
Gran, U. B753 (2006) 118 Loll, R. B751 (2006) 419
Grassi, P.A. B751 (2006) 53 Lombardo, M.P. B752 (2006) 124
Grimus, W. B754 (2006) 1 Love, A. B755 (2006) 79
Grinstein, B. B752 (2006) 18
Grinstein, B. B755 (2006) 199 Macêdo, A.M.S. B752 (2006) 439
Grisaru, M. B755 (2006) 21 Macedo-Junior, A.F. B752 (2006) 439
Gubser, S.S. B757 (2006) 146 Mahato, M. B757 (2006) 1
Gudnason, S.B. B754 (2006) 293 Mangano, G. B756 (2006) 100
Gutowski, J. B753 (2006) 118 Manvelyan, R. B751 (2006) 285
Gutperle, M. B753 (2006) 16 Martemyanov, B.V. B756 (2006) 71
Gutperle, M. B757 (2006) 79 Maru, N. B754 (2006) 127
Gwyn, R. B755 (2006) 21 Masina, I. B755 (2006) 1
Mathews, P. B753 (2006) 1
Hadasz, L. B757 (2006) 233 Mathur, S.D. B754 (2006) 233
Harmark, T. B757 (2006) 117 McInnes, B. B754 (2006) 91
He, X. B755 (2006) 313 Meyer, F. B753 (2006) 92
Hebecker, A. B755 (2006) 186 Michalogiorgakis, G. B757 (2006) 146
Heise, R. B753 (2006) 195 Miele, G. B756 (2006) 100
Hofman, C. B752 (2006) 329 Minakami, S. B752 (2006) 391
Hosteins, P. B755 (2006) 137 Mitov, A. B751 (2006) 18
Hou, D.-f. B754 (2006) 351 Moch, S. B751 (2006) 18
Huang, Q.-G. B755 (2006) 286 Molina, C. B752 (2006) 40
Huber, S.J. B757 (2006) 172 Morales Morera, J.F. B751 (2006) 53
Moster, S. B751 (2006) 186
Idilbi, A. B753 (2006) 42 Müller-Preussker, M. B756 (2006) 71
Ilgenfritz, E.-M. B756 (2006) 71 Muramoto, C. B754 (2006) 146
Inami, T. B752 (2006) 391
Isozumi, Y. B752 (2006) 140 Naculich, S.G. B755 (2006) 164
Ivanov, E. B753 (2006) 211 Nakayama, Y. B755 (2006) 295
Nanopoulos, D.V. B751 (2006) 260
Jäger, S. B751 (2006) 160 Nelson, B.D. B751 (2006) 75
Jantzen, B. B752 (2006) 327 Nitta, M. B752 (2006) 140
Jaskólski, Z. B757 (2006) 233 Nitta, M. B752 (2006) 391
346 Nuclear Physics B 757 (2006) 344–346

Ohashi, K. B752 (2006) 140 Semikhatov, A.M. B757 (2006) 303


Olesen, P. B752 (2006) 197 Serpico, P.D. B756 (2006) 100
Ooguri, H. B755 (2006) 239 Shapiro, I. B756 (2006) 207
Ookouchi, Y. B755 (2006) 239 Shcheredin, S. B754 (2006) 17
Orselli, M. B757 (2006) 117 Shibusa, Y. B754 (2006) 48
Skvortsov, E.D. B756 (2006) 117
Paccetti Correia, F. B751 (2006) 222 Smirnov, V.A. B752 (2006) 327
Pajer, E. B756 (2006) 16 Sokatchev, E. B754 (2006) 329
Pallis, C. B751 (2006) 129 Sorella, S.P. B753 (2006) 252
Pando Zayas, L.A. B757 (2006) 1 Sorella, S.P. B753 (2006) 273
Pantev, T. B752 (2006) 329 Splittorff, K. B757 (2006) 259
Papadopoulos, G. B753 (2006) 118 Srivastava, Y.K. B754 (2006) 233
Pastor, S. B756 (2006) 100 Stanev, Ya.S. B754 (2006) 329
Pavan, A.B. B752 (2006) 40 Steinhauser, M. B757 (2006) 197
Penin, A.A. B752 (2006) 327 Strumia, A. B753 (2006) 178
Pich, A. B753 (2006) 139 Sturani, R. B756 (2006) 16
Pinansky, S. B753 (2006) 69 Sugiyama, K. B753 (2006) 295
Pinto, T. B756 (2006) 100 Sundell, P. B752 (2006) 206
Pisanti, O. B756 (2006) 100 Svendsen, H.G. B753 (2006) 195
Polikarpov, M.I. B756 (2006) 71
Polychronakos, A.P. B751 (2006) 376 Tanaka, K. B752 (2006) 1
Portolés, J. B753 (2006) 139 Tatar, R. B755 (2006) 21
Prokopec, T. B757 (2006) 172 Tavartkiladze, Z. B751 (2006) 222
Prytula, O.O. B753 (2006) 242 Tipunin, I.Yu. B757 (2006) 303
Pusztai, B.G. B751 (2006) 436 Tokunaga, T. B753 (2006) 295
Pylyuk, I.V. B753 (2006) 242 Tonasse, M.D. B756 (2006) 1
Trapletti, M. B755 (2006) 186
Rajabpour, M.A. B754 (2006) 283
Trigiante, M. B751 (2006) 343
Rattazzi, R. B757 (2006) 19
Tseng, L.-S. B751 (2006) 108
Ravindran, V. B752 (2006) 173
Ravindran, V. B753 (2006) 1
Urban, F.R. B752 (2006) 297
Ren, H.-c. B754 (2006) 351
Utermann, A. B754 (2006) 107
Reuillon, S. B751 (2006) 390
Reya, E. B754 (2006) 178
Riemann, T. B751 (2006) 1 Valle, J.W.F. B752 (2006) 80
Ringwald, A. B754 (2006) 107 van Neerven, W.L. B755 (2006) 272
Robaschik, D. B755 (2006) 112 Vasiliev, M.A. B756 (2006) 117
Rodríguez-Gómez, D. B752 (2006) 316 Vázquez-Mozo, M.A. B753 (2006) 92
Roest, D. B753 (2006) 118 Verbaarschot, J.J.M. B757 (2006) 259
Rossi, G.C. B754 (2006) 329 Verdaguer, E. B756 (2006) 148
Rouhani, S. B754 (2006) 283 Veselov, A.I. B756 (2006) 71
Rühl, W. B751 (2006) 285 Volkov, M.S. B751 (2006) 390

Sá Borges, J. B756 (2006) 1 Wang, T. B756 (2006) 86


Sahu, N. B752 (2006) 280 Weigand, T. B751 (2006) 186
Savoy, C.A. B755 (2006) 1 Westra, W. B751 (2006) 419
Savoy, C.A. B755 (2006) 137
Schmidt, M.G. B751 (2006) 222 Yajnik, U.A. B752 (2006) 280
Schmidt, M.G. B757 (2006) 172 Yamashita, T. B754 (2006) 127
Schnitzer, H.J. B755 (2006) 164 Yau, S.-T. B751 (2006) 108
Schuck, C. B754 (2006) 178 Yuan, F. B753 (2006) 42
Schwarz, A. B756 (2006) 207
Seki, S. B753 (2006) 295 Zohren, S. B751 (2006) 419

Potrebbero piacerti anche