You are on page 1of 804

Nuclear Physics B 584 (2000) 345

www.elsevier.nl/locate/npe

Natural effective supersymmetry


Junji Hisano a , Kiichi Kurosawa b,c , Yasunori Nomura c
a Theory Group, KEK, Tsukuba, Ibaraki 305-0801, Japan
b YITP, Kyoto University, Kyoto 606-8502, Japan
c Department of Physics, University of Tokyo, Tokyo 113-0033, Japan

Received 28 February 2000; accepted 23 May 2000

Abstract
Much heavier sfermions of the first-two generations than the other superparticles provide a natural
explanation for the flavor and CP problems in the supersymmetric Standard Model (SUSY SM).
However, the heavy sfermions may drive the mass squareds for the light third generation sfermions
to be negative through two-loop renormalization group (RG) equations, breaking color and charge.
Introducing extra matters to the SUSY SM, it is possible to construct models where the sfermion
masses are RG invariant at the two-loop level in the limit of vanishing gaugino-mass and Yukawacoupling contributions. We calculate the finite corrections to the light sfermion masses at the twoloop level in the models. We find that the finite corrections to the light-squark mass squareds are
negative and can be less than (0.31)% of the heavy-squark mass squareds, depending on the number
and the parameters of the extra matters. We also discuss whether such models realized by the U(1)X
gauge interaction at the GUT scale can satisfy the constraints from 1mK and K naturally. When
both the left- and right-handed down-type squarks of the first-two generations have common U(1)X
charges, the supersymmetric contributions to 1mK and K are sufficiently suppressed without
spoiling naturalness, even if the flavor-violating supergravity contributions to the sfermion mass
matrices are included. When only the right-handed squarks of the first-two generations have a
common U(1)X charge, we can still satisfy the constraint from 1mK naturally, but evading the
bound from K requires the CP phase smaller than 102 . 2000 Elsevier Science B.V. All rights
reserved.
PACS: 14.80.Ly; 12.60.Jv; 12.15.Ff

1. Introduction
The Standard Model (SM) has enjoyed remarkable successes in describing physics
down to a scale of 1018 m, the weak scale. However, the quadratically divergent
radiative correction to the Higgs boson mass leads to the well-known naturalness problem.
Supersymmetry (SUSY) provides a solution to this problem by extending the chiral
symmetry to the scalar partners. At present, the minimal supersymmetric standard model
0550-3213/00/$ see front matter 2000 Elsevier Science B.V. All rights reserved.
PII: S 0 5 5 0 - 3 2 1 3 ( 0 0 ) 0 0 3 4 3 - 6

J. Hisano et al. / Nuclear Physics B 584 (2000) 345

(MSSM) is considered to be the most promising candidate beyond the SM (for reviews see,
[1,2]).
The MSSM is severely constrained by flavor physics. Introduction of SUSY-breaking
terms provides new sources of the flavor-changing neutral currents (FCNCs) and CP
violation. For instance, the experimental value of the mass difference between KL and
KS , 1mK , leads to the upper bound on the squark contribution written as
!2
1m2  30 TeV 2
d
2
. 1,
(1)
sin d
m
d
m
2
d

in a basis where the quark mass matrices are diagonal (see, for example, [3,4]). Here, m
2
d

is the averaged squark mass, and 1m2 and sin2 d are the mass-squared difference and
d
the mixing angle between the down-type squarks of the first-two generations. In order to
satisfy this constraint, (i) 1m2 ' 0 (degeneracy [5,6]), (ii) sin d ' 0 (alignment [79]),
d
(iii) m
d & 30 TeV (decoupling [1013]), or the hybrid of them is required.
In this article, we discuss the third possibility, the decoupling scenario. In this scenario,
the masses for the third generation squarks, Higgsinos, and gauginos are of the order of the
weak scale while the other squarks and sleptons are heavy enough to suppress the FCNC
and CP-violating processes. This is still natural at the one-loop level, since the squarks
and sleptons of the first-two generations are not strongly coupled with the Higgs boson.
This mass spectrum is sometimes called as effective supersymmetry or the more minimal
supersymmetric standard model. It is realized in the anomalous U(1) SUSY-breaking
models [1424], U(1)0 models [2527], composite models of the first-two generation
particles [28,29], or radiatively-driven models with specific boundary conditions [3032].
However, this scenario generically suffers from a problem that the third generation
sfermions obtain the vacuum expectation value (VEV), breaking color and charge [33,
34]. At the two-loop level the heavy first-two generation sfermions contribute to the the
renormalization group (RG) equations for the third generation sfermion masses through the
gauge interactions. The RG equations for the SUSY-breaking sfermion masses m
e2r through
the SM gauge interactions are given by [35]
 
dm
e2r X A 2 A X A 2
=
8
Cr
Ts m
es

d
4
s
A
  X
X  Y  A  X
Y
2
+2
Ys m
es +
8
Ys CsA m
e2s ,
(2)
Yr
Yr
4
4
4
s
s
A

at the two-loop level in the DR scheme [36]. Here, we have taken a limit where the
gaugino masses vanish. An index A (= 13) represents the SM gauge groups, and we
have adopted the SU(5)GUT normalization for the U(1)Y gauge coupling (1 (5/3) Y ).
TrA , CrA and Yr are the Dynkin index, the quadratic Casimir coefficient and the hypercharge
for the sfermion r, respectively. We find that all sfermion mass squareds contribute to the
RG equations at the two-loop level. Thus, the contribution from the heavy sfermions may
destroy the mass spectrum by driving the light-sfermion mass squareds to be negative at

J. Hisano et al. / Nuclear Physics B 584 (2000) 345

low energies. This fact makes it difficult to construct models where such a hierarchical
sfermion mass spectrum is generated at much higher energy scale than the weak scale.
However, the hierarchical mass spectrum can be stabilized against the RG evolution if
the following relations among the SUSY-breaking mass squareds for the heavy sfermions
are imposed [37,38]:
X
TrA m
e2r = 0,
(3)
r

X
X

Yr m
e2r = 0,

(4)

Yr CrA m
e2r = 0,

(5)

for A = 13. These relations cannot be satisfied by the heavy MSSM sfermions alone,
since all of them have positive SUSY-breaking mass squareds. However, the relations
can be satisfied if we introduce extra fields with negative SUSY-breaking mass squareds
which transform nontrivially under the SM gauge groups, since the sum is taken over
the extra fields as well as the heavy MSSM sfermions. The extra fields should have the
supersymmetric masses to avoid the breaking of the SM gauge groups caused by their
condensations. Thus, the extra fields have to be in vector-like representations of the SM
gauge groups, and we call them extra matters. The supersymmetric masses for the extra
matters should not be much larger than the SUSY-breaking scalar masses. Otherwise, the
large radiative correction is generated below the energy scale where the extra matters
are decoupled. We refer this extension of the MSSM to the Natural Effective SUSY SM
(NESSM), hereafter.
The NESSM can be realized by assuming that the SUSY-breaking masses for the
heavy sfermions and extra matters are generated by a D-term VEV of some U(1)X gauge
interaction such that m
e2r = QX
r hDX i. The U(1)X should be broken in order to give the
supersymmetric masses for the extra matters. In terms of the U(1)X charges, Eqs. (3)(5)
are written as
X
TrA QX
(6)
r = 0,
r

X
X

Yr QX
r = 0,

(7)

Yr CrA QX
r = 0.

(8)

The first equation corresponds to vanishing mixed anomalies between U(1)X and the SM
gauge groups. The second and third ones mean that U(1)X has no kinetic mixing with
U(1)Y at the one- and two-loop levels. If U(1)X is anomaly-free and all fields charged under
U(1)X are embedded in the SU(5)GUT multiplets, Eqs. (6), (7) are satisfied automatically.
In this case, we have only to choose the U(1)X charges for the extra matters to satisfy
Eq. (8). Even if U(1)X is anomalous, however, we can still construct a model in which
some of the fields are decoupled at the U(1)X breaking scale and Eqs. (6)(8) are satisfied
at low energies.

J. Hisano et al. / Nuclear Physics B 584 (2000) 345

In this article, we calculate the finite corrections to the light-sfermion mass squareds at
the two-loop level in the NESSM, assuming that all fields are embedded in the SU(5)GUT
multiplets at the GUT scale. In the NESSM, while the dangerous contributions which lead
to color and charge breaking do not exist in the RG equations for the SUSY-breaking
scalar masses, the finite corrections from the heavy sfermions and the extra matters at the
two-loop level may still drive the light-sfermion mass squareds to be negative. We find
that the finite corrections to the light-squark mass squareds depend on the number and the
parameters of the extra matters and they are less than (0.31)% of the heavy-squark mass
squareds.
We also discuss whether the models in which the NESSM is realized by the U(1)X gauge
interaction at the GUT scale (MGUT ' 2 1016 GeV) are viable or not in the light of the
experimental constraints. The anomalous U(1) SUSY-breaking model given in Refs. [37,
38] is an explicit example for such models. On this setup, the supergravity contributions
to the SUSY-breaking terms, which are generically non-universal in flavor space, are
suppressed compared with the U(1)X D-term contribution. Moreover, the breaking of the
U(1)X symmetry can naturally explain the hierarchical structure of the quark and lepton
masses by the FroggattNielsen mechanism [39]. In this paper, we consider the constraints
from 1mK and K , which lead to the severest bound on the heavy-sfermion mass scale. We
find that, when both the left- and right-handed down-type squarks (only the right-handed
squarks) of the first-two generations have common U(1)X charges, the contributions to
1mK and K (1mK ) are sufficiently suppressed without spoiling naturalness even if the
flavor-violating supergravity contributions to the sfermion mass matrices are included.
This paper is organized as follows. In the next section, we show our setup and the U(1)X
charge assignments. We explain that the squark and slepton masses are inversely related
to the quark and lepton masses through their U(1)X charges. In Sections 3, we derive the
formula for the finite corrections to the light sfermion masses at the two-loop level in the
NESSM, and evaluate them numerically. In Section 4, using the constraints from 1mK
and K the lower bounds on the third generation sfermion masses at the GUT scale are
derived. We neglect the effects of the Yukawa interactions in Sections 3 and 4, since they
are model-dependent. The effects are evaluated in Section 5. We also discuss the three-loop
RG contributions to the light sfermion masses there. Section 6 is devoted to the conclusions
and discussion. In Appendix A, we review the anomalous U(1) SUSY-breaking model
given in Refs. [37,38] as one of the explicit realizations of the NESSM. In Appendix B,
we show the complete formulae for the contributions to the light sfermion masses from the
heavy ones at the two-loop level. The formulae for 1mK and K are given in Appendix C.

2. Setup
In this section, we show the setup of the NESSM on which we perform our analyses in
later sections. In the NESSM, the SUSY-breaking scalar masses for the heavy sfermions
and extra scalars are given by the U(1)X D-term. We here assume that the nonzero U(1)X
D-term is generated associated with the breaking of the U(1)X gauge symmetry caused

J. Hisano et al. / Nuclear Physics B 584 (2000) 345

by the VEV of a field which has a U(1)X charge of 1 (QX


= 1). In this case, the
U(1)X D-term is related to the F -term of the field at a (local) minimum of the potential
as follows:
DX =

2
gX

MX2

|F |2 ,

(9)

where MX (= gX |hi|) is the U(1)X gauge boson mass. Here, we have assumed that the
VEVs and F -terms of the other fields are negligibly small for simplicity. Then, the relation

between the supergravity contribution m0 and the U(1)X D-term (mD |DX | ) is given
by
m0 '

|F | |hi|
=
mD ,
Mpl
Mpl

(10)

where Mpl is the reduced Planck scale (Mpl ' 2 1018 GeV). This shows that the
supergravity contributions to the sfermion masses are, in general, suppressed compared
with the D-term contribution, mD , as long as |hi| is smaller than Mpl . Then, the U(1)X
charged sfermions have large masses of the order of the U(1)X D-term, mD , while
the neutral sfermions only receive smaller masses of order m0 from the supergravity
contributions. In order to obtain desired mass spectrum, we set m0 to be of the order of
the weak scale and |hi|/Mpl ' (101 102 ).
Next, we consider the U(1)X charge assignment. We assume that all the fields of the
NESSM are embedded in SU(5)GUT multiplets at the GUT scale. It not only maintains
the successful gauge coupling unification but also guarantees vanishing U(1)Y D-term
contribution at the one-loop level (Eq. (7)). We also assume that one pair of the extra
matters are introduced in (5 + 5? ) representation of the SU(5)GUT for simplicity. The
extension to the other cases is straightforward. Then, the U(1)X charges for the 5 and
X
5? extra matters, QX
5ex and Q5? ex , are determined from those for the quarks and leptons in
the SM to satisfy Eqs. (6), (8) as
QX
5ex

= 2

3
X

QX
10i ,

(11)


X
QX
10i + Q5? i ,

(12)

i=1

QX
5? ex =

3
X
i=1

X
where QX
10i and Q5? i (i = 13) are the U(1)X charges for the quarks and leptons embedded
?
in the 10 and 5 representations of the SU(5)GUT , respectively. An index i represents the
generation.
The U(1)X charges for the SM matter multiplets are determined by the following
considerations. The naturalness argument tells us that the superparticles which are strongly
coupled with the Higgs boson are necessary to have masses around the weak scale and thus
should have zero U(1)X charges. This requires that QX
103 = 0, since the top squarks and
left-handed bottom squark are coupled with the Higgs boson through large top Yukawa
coupling. On the other hand, the superpartners of the light quarks and leptons have to

J. Hisano et al. / Nuclear Physics B 584 (2000) 345

be heavy enough to sufficiently suppress the FCNC processes, so that they should have
positive U(1)X charges. This explains the hierarchy of the quark and lepton masses
naturally by the FroggattNielsen mechanism, since the Yukawa matrices are generated
through the VEV of the field suppressed by suitable powers of hi/Mpl [39] as
W=

3
X


(fu )ij

i,j =1

3
X
i,j =1

hi
Mpl


(fd )ij

QX

X
10i +Q10j

hi
Mpl

10i 10j Hu

QX

X
10i +Q5? j

10i 5? j Hd ,

(13)

where Hu and Hd are the Higgs doublets for which we have assumed the vanishing U(1)X
charges. Thus, it is possible to reproduce the observed quark and lepton mass matrices
X
with (fu )ij , (fd )ij = O(1) if we appropriately choose the U(1)X charges QX
10i and Q5? i
[4043].
According to the above arguments, we consider the following U(1)X charge assignments:
Model

II

III

IV

QX
101

QX
102

QX
103

QX
5? 1

QX
5? 2
QX
5? 3

QX
5ex

QX
5? ex

Here, the U(1)X charges for the Higgs multiplets are zero in all the models and the
charges for the extra-matter multiplets were determined by Eqs. (11), (12). Model I is
the simplest possibility realizing the decoupling scenario. In this model, the FCNCs are
strongly suppressed due to the degeneracy of the SUSY-breaking masses between the firsttwo generation sfermions, while the hierarchy among the quark and lepton masses cannot
be explained completely. Models IIIV are motivated to explain the fermion mass hierarchy
by the FroggattNielsen mechanism. In particular, in Models II, IV the second and third
X
generation doublet leptons have the same U(1)X charges, QX
5? 2 = Q5? 3 , so that the observed
1
large mixing between and [44] is naturally explained. In these models, the FCNC
1 Models II and IV prefer large and small angle MSW solutions to the solar neutrino problem, respectively
[45,46].

J. Hisano et al. / Nuclear Physics B 584 (2000) 345

processes are less suppressed than in Model I. Models II, III have non-degenerate SUSYbreaking masses for the left-handed down-type squarks in the first-two generations. Thus,
K 0 K 0 oscillation has dominant contribution from the off-diagonal elements in the lefthanded squark mass matrix. Model IV has non-degenerate SUSY-breaking masses for both
the left- and right-handed down-type squarks in the first-two generations, so that K 0 K 0
oscillation receives contributions from off-diagonal elements of both the left- and righthanded squark mass matrices. As a result, Model IV is more severely constrained from
K 0 K 0 mixing than Models II, III, as will be discussed later.
We now discuss the other SUSY-breaking parameters. Since the gaugino masses are
constrained by naturalness argument, we consider them to be of the order of the weak scale.
Indeed, in the anomalous U(1) SUSY-breaking model given in Appendix A, the gaugino
masses arise from the F -term of the dilaton field [48] and their sizes can be of the order
of the weak scale [49]. We treat the gaugino mass (at the GUT scale) as a free parameter
in the phenomenological analyses below. Also, the superpotential Eq. (13) generates the
SUSY-breaking trilinear scalar couplings,
X
(Au )ij = QX
10i + Q10j

(Ad )ij = QX
10i

 F

,
hi
 F
.
+ QX
5? j
hi

(14)
(15)

The (Au )ij and (Ad )ij are defined by Lsoft = (yu )ij (Au )ij 10i 10j hu + (yd )ij (Ad )ij
10i 5? j hd . The (yu )ij and (yd )ij are the Yukawa couplings W = (yu )ij 10i 10j Hu +
(yd )ij 10i 5? j Hd , and and h represent the scalar components of the corresponding
supermultiplets and H , respectively. From Eq. (10), we find that Au ' Ad ' mD except
for the ones which involve only the light sfermions.
The supersymmetric masses and the holomorphic SUSY-breaking masses for the extra
matters come from the U(1)X symmetry breaking. The supersymmetric masses should be
of the same order with mD as we mentioned in the introduction. While these features
are naturally explained in the explicit model given in Refs. [37,38] (see Appendix A), we
mainly adopt a phenomenological approach in this article and take the masses as input
parameters. We parameterize the supersymmetric and holomorphic SUSY-breaking mass
parameters as
Wex = m 3ex 3? ex + m0 2ex 2? ex ,
Lex, soft = F 3ex 3? ex + F0 2ex 2? ex


+ h.c. .

(16)
(17)

Here, the triplet extra matters 3ex and 3? ex and the doublet extra matters 2ex and 2? ex
are embedded in the 5ex and 5? ex. In this article, we impose the SU(5)GUT relations on the
supersymmetric and the holomorphic SUSY-breaking masses, m = m0 and F = F0 , at
the GUT scale unless otherwise stated. Then, the masses for the heavy sfermions and extra
matters are completely determined by three free parameters mD , m and F once a U(1)X
charge assignment is specified.

10

J. Hisano et al. / Nuclear Physics B 584 (2000) 345

3. Corrections to the light sfermions


In the NESSM, the two-loop RG contributions to the light sfermions are canceled
between the heavy sfermions and the extra-matter multiplets. However, there are still
negative finite corrections. In this section, we numerically estimate the effects of the finite
corrections in various cases, using their explicit form which is completely calculated at
the two-loop level in Appendix B. We conclude this section by showing how hierarchical
the sfermion mass spectrum can be in the NESSM without breaking color and charge,
comparing the NESSM with the original effective SUSY in which there is no extra-matter
multiplet.
Now, let us numerically estimate the finite corrections in the models discussed in
the previous section. We first consider Models I, II. In these models, the U(1)X charge
assignments are consistent with the SU(5)GUT and the extra-matter multiplets have an
invariance under the parity 5ex 5? ex , so that the finite corrections to the light
, take relatively simple forms. The explicit expressions for m2
are
sfermions, m2
f , finite
f , finite
given in Eqs. (B.47)(B.49) in Appendix B. These expressions can be further simplified in
the case of one-pair extra-matter multiplets as follows:



X A 2 
m2
1 2
2

CfA TrF TFA


m
+
log
m2f, finite = 4

F
4
3
m2
A
F

1
+ m2 G(y1 , y2 )
2



h
i
X Y
A
(18)
Yf TrF YF CFA m2F log m2F ,
+4
4
4
A

where






1
1
y2
+ y1 Li2 1
+ (y1 y2 ), (19)
G(y1 , y2 ) = y1 log y1 2y1 Li2 1
y1
2
y1
y1

m21
m2

y2

m22
m2

(20)

where F denotes the heavy sfermions; m1 and m2 are the mass eigenvalues for the
extra scalars (see Appendix B). Here, we have neglected the difference between the
supersymmetric masses (and the holomorphic SUSY-breaking masses) of the triplet and
doublet extra matters for demonstrational purpose, though it will be included in later
numerical calculations.
The first term of the first line in Eq. (18) gives dominant contributions to the lightsfermion mass squareds. They are negative and could cause color and charge breaking
since m > mF . The second term has the same form as the scalar mass squared generated
by integrating out the messenger fields in the gauge-mediated SUSY breaking mechanism
[50]. Therefore, it gives the positive contributions almost proportional to the SUSYbreaking bilinear coupling F of the extra scalars. Even if F = 0, however, G(y1 , y2 )

J. Hisano et al. / Nuclear Physics B 584 (2000) 345

11

remains positive, since m21 and m22 are smaller than m2 by the U(1)X D-term. The last
term which is generated through U(1)Y D-term does not give significant contributions due
to the smallness of the hypercharge gauge coupling. The renormalization-point dependence
of the term is canceled out due to Eq. (8).
In order to see the dependence of the finite corrections on m , mD and F , we fix
the heavy-sfermion mass scale mD = 10 TeV as an overall scale and take m /mD and
F /(m mD ) as remaining two dimensionless free parameters. The ratio m /mD must
be of order unity in the NESSM, since otherwise the heavy sfermions generate the large
radiative corrections below the scale m where the extra-matter multiplets decouple. Then,
F /(m mD ) should not be much larger than one to avoid the negative mass squared for
the extra scalar. Indeed, F /(m mD ) is of order unity in the explicit example constructed.
Fig. 1 illustrates the dependence of the finite corrections on these two parameters. As
an example, we consider the corrections to the light squarks in Models I, II and plot the
ratio of the finite corrections to the heavy sfermion mass-scale squared m2D . Here, we
only take into account the dominant two-loop contribution through the strong interaction,
(3)
that is, we drop the last term and set A = 3 and C = 4/3 in Eq. (18). The minimum
f
of the ratio is 0.36% (m /mD = 2.2) and 0.54% (m /mD = 2.7) in Models I and II,
respectively, when F = 0. The difference between Models I and II in Fig. 1 comes mainly




m2 /m2D = TrF T A
QX , which is
from the group-theoretical factor in Eq. (18), TrF T A
F
F
F
F
4 in Model I and 6 in Model II. Thus, the corrections in Model II are larger than in Model I
approximately by a factor of 1.5.
The ratio given in Fig. 1 determines how hierarchical the sfermion mass spectrum can
be without breaking color and charge. Since the bare mass mq,0
for the light squarks has
to be larger than these negative corrections, it gives a lower bound on mq,0
. In Model I, for
example, we obtain
2
2
m2q,0
> mq,
finite ' 0.4% mD ,

(21)

Fig. 1. Finite corrections to the light squarks through the strong interaction. The percentages on the
figures represent the ratio of the finite corrections to the heavy-sfermion mass scale, m2q,
/m2D .
finite
The left figure is in Model I and the right is in Model II. The left side of the solid line is excluded
since the mass squared for the extra scalar is negative and color is broken. Here, we have taken
mD = 10 TeV and evaluated all the parameters at = 10 TeV.

12

J. Hisano et al. / Nuclear Physics B 584 (2000) 345

when the supersymmetric mass for the extra matter, m , is not much larger than mD as
discussed in the previous section. 2 Since mq,0
set the scale for the light sfermion masses,
we find that the splitting between the heavy and light sfermions cannot far exceed an
order of magnitude. In the case of the original effective SUSY models (models without
extra-matter multiplets), the RG contributions to the light sfermions, m2 , are written
f , log
by solving the RG equation Eq. (2) as follows:

h
i ()
A
GUT
2

,
TrF TFA
m

F
bA
4
4
A

m2f, log() = 4

X Cf
A

(22)

where bA is the coefficient for the one-loop beta function of the SM gauge coupling and
GUT is the gauge coupling constant at the GUT scale. Here, the contributions from the
U(1)Y D-term are neglected. In Model I, for example, radiative corrections to the light
squarks through the strong interaction are estimated as
2
m2q,
log ( = 10 TeV) ' 2.3% mD ,

(23)

2
so that the bare light-squark mass squared m2q,0
should be larger than about mq,
log ( =

10 TeV) ' 2.3% m2D . The rough estimations in Eqs. (21), (23) show that the splitting
between the heavy and light sfermion masses in the NESSM can be more than twice larger
than in the original effective SUSY. This conclusion remains true even if full radiative
corrections are taken into account, as will be shown below.
We here comment on the effect of introducing more than one pairs of extra-matter
multiplets. If we introduce more extra-matter pairs, the extra matters could have smaller
U(1)X charges satisfying Eqs. (6), (8) and thus smaller supersymmetric masses. Thus, we
can reduce the finite corrections to the light-sfermion mass squareds by introducing more
extra matters, since it is the supersymmetric mass for the extra matters that determines
the size of the negative finite corrections. For example, the minimum size of the finite
corrections with F = 0 in Model I are reduced to 0.36%, 0.30%, 0.26%, and
0.24% by introducing one, two, three and four pairs of 5ex and 5? ex , respectively.
However, introducing extra matters changes the beta functions, so that the behavior of
the gauge couplings becomes worse at high energy in this case. In this article, we limit
ourselves to the case of one pair of 5ex and 5? ex , hereafter.
We now include the effects from the gauginos. Since the gauginos give positive RG
contributions at the one-loop level, they relax the lower bounds on the light sfermion
masses. The lower bounds on the light sfermion masses can be determined as follows. We
first set the boundary conditions at the GUT scale 2 1016 GeV: we, for simplicity, take the
universal gaugino mass m1/2 and give the bare mass m0 to the light sfermions at the GUT
scale. Then, we run all the soft SUSY-breaking masses using two-loop RG equations. We
2 In the region where m /m & O(10) and F /(m m ) = O(10), the finite corrections can be much smaller

D
due to the accidental cancellation between the negative contribution from the heavy-sfermion loops and the
gauge-mediated contribution from the extra-matter multiplets. However, this does not necessarily mean that more
hierarchical superparticle mass spectrum can be realized, since the gluino mass receives the correction from the
extra-matter loops (mg ' (3 /4 )(F /m ) 0.1mD ).

J. Hisano et al. / Nuclear Physics B 584 (2000) 345

13

neglect the effects of the Yukawa couplings, since they are rather model-dependent without
the assumption of universal scalar masses. The effects will be discussed in Section 5. The
heavy sfermions are decoupled at the scale mD = 10 TeV, where the full finite corrections
are added to the light sfermions. The positive contributions from the gauginos are included
from the GUT scale to the scale where the gluino decouples. After these steps, if any light
sfermions have negative mass squareds at the weak scale, the parameter region is excluded
and we need larger value of m0 at the GUT scale.
In Figs. 2 and 3, we have plotted the lower bounds on the light sfermion masses m0 as
a function of the gaugino mass m1/2 in Models I and II. The m0 and m1/2 are the GUTscale values and normalized by the heavy-sfermion mass scale mD . In each figure, the

Fig. 2. Lower bounds on the mass ratios of the light sfermions to the heavy sfermions in Model I. m0
and m1/2 represent the light sfermion and gaugino masses at the GUT scale, respectively. The regions
below the curves are excluded, since the light sfermions indicated by various lines have negative
mass squareds at the weak scale, breaking color and charge. Here, we have set mD = 10 TeV and
m = 2.0 mD at the GUT scale, and neglected the Yukawa couplings.

Fig. 3. Lower bounds on the mass ratios of the light sfermions to the heavy sfermions in Model II.
We have set mD = 10 TeV and m = 2.5 mD at the GUT scale.

14

J. Hisano et al. / Nuclear Physics B 584 (2000) 345

Fig. 4. Lower bounds on the mass ratios of the light sfermions to the heavy sfermions in Model III.
We have set mD = 10 TeV and m = 2.5 mD at the GUT scale.

left graph (ESSM) is in the case of the original Effective SUSY SM (without extra-matter
multiplets) and the right one (NESSM) is of the Natural Effective SUSY SM. We have
taken the supersymmetric mass m for the extra-matter multiplets as m = 2.0 mD in
Model I and m = 2.5 mD in Model II at the GUT scale. The various curves represent
the lower bounds on the bare masses for the indicated light sfermions, below which they
have negative mass squareds at the weak scale. From Figs. 2 and 3 we find that in the
NESSM the light sfermion masses and/or the gaugino masses can take values less than
half in the ESSM without breaking color and charge. This in turn means that we can obtain
more hierarchical mass spectrum in the NESSM than in the ESSM.
Note that we have assumed that the triplet and doublet extra matters have the same
supersymmetric masses at the GUT scale. Then, at low energy the supersymmetric mass
for the triplet extra matter becomes about twice as large as that for the doublet one due to
the RG evolution. As a result, we cannot take the supersymmetric mass which minimizes
the finite corrections in Fig. 1, since then the doublet extra scalar has negative mass
squared. Thus, if we allow different supersymmetric masses for the triplet and doublet
extra matters at the GUT scale, we can further reduce the size of finite corrections. 3 In
this case, however, the running masses for the triplet extra scalars take negative values at
the high-energy scale, which means that the scalar potential has another minimum at the
nonzero values of the triplet extra scalars.
In the rest of this section, we consider Models III, IV. The finite corrections in these
models are given by Eqs. (B.44)(B.46) in Appendix B. They take less simple forms due
to the absence of the parity between the extra-matter multiplets, 5ex 5? ex . In Figs. 4
and 5, we have plotted the lower bounds on m0 in Models III and IV, respectively. We have
taken the supersymmetric mass for the extra-matter multiplet as m = 2.5 mD in Model
III and m = 2.7 mD in Model IV at the GUT scale. All the other parameters have been
taken the same as in Models I, II.
3 The different supersymmetric masses for the triplet and doublet extra matters do not necessarily contradict
with the GUT, since we can make them split using GUT-breaking VEV such as h6(24)i.

J. Hisano et al. / Nuclear Physics B 584 (2000) 345

15

Fig. 5. Lower bounds on the mass ratios of the light sfermions to the heavy sfermions in Model IV.
We have set mD = 10 TeV and m = 2.7 mD at the GUT scale.

The main difference of Models III, IV from Models I, II is the existence of the one-loop
U(1)Y D-term,
 
h
i
Y
(24)
Yf TrR YR cos 2 m21 log m21 m22 log m22 ,
m2f, FI-1loop =
4
where R denotes the extra matters. This term arises because the RG evolution from the
GUT to low-energy scale splits the supersymmetric masses for the triplet and doublet extra
matters. However, the corrections Eq. (24) do not dominate the two-loop contributions,
since the hypercharge gauge coupling is much smaller
than

 the strong coupling and there
2 in Eq. (18). As a consequence,
m
is no enhancement by factors such as 4 or TrF T A
F
F
Figs. 4 and 5 are roughly the same as Figs. 2 and 3. Nonetheless, we can still find the small
effects of the one-loop U(1)Y D-term. In Model III, the lower bound on the mass for u 3
is tighter than that for q3 , since a positive one-loop U(1)Y D-term is generated. In Model
IV, the lower bound on the mass for e3 is much larger than that in Model II because of a
negative one-loop U(1)Y D-term.
4. Constraints from K 0 K 0 mixing
K 0 K 0 mixing gives the severest constraints on the masses for the first-two generation
sfermions in flavor-changing processes. In this section, we calculate the lower bounds on
the heavy-sfermion mass scale mD from the KL KS mass difference 1mK and the CPviolating parameter K . Combining these bounds on mD with the bounds on the mass ratio
m0 /mD given in the previous section, we obtain the lower bounds on the light sfermion
masses m0 at the GUT scale. It turns out that no severe fine tuning is needed in the NESSM,
compared with the original Effective SUSY SM (ESSM).
We first discuss the structure of the mass matrix for the down-type squarks, since it
induces a dominant flavor violation in K 0 K 0 mixing through the gluinosquark box
diagram. We restrict our attention to the contribution from the heavy first-two generation

16

J. Hisano et al. / Nuclear Physics B 584 (2000) 345

squarks below. The mass matrix for the down-type squarks in the U(1)X gauge basis is
given by
!
2
0
MLL
2
,
(25)
M =
2
0 MRR
where we have set the leftright mixing mass terms zero due to the smallness of the
2
corresponding Yukawa couplings. The left- and right-handed squark mass matrices MLL
2 consist of two parts as follows:
and MRR
2
2
2
= MLL,D
+ MLL,0
,
MLL

2
2
2
MRR
= MRR,D
+ MRR,0
.

(26)

2
2
and MRR,D
come from the U(1)X D-term and are proportional to their U(1)X
MLL,D
charges,
!
!
QX
QX
0
0
101
5? 1
2
2
2
MRR,D =
(27)
mD ,
m2D .
MLL,D =
X
0 QX
0
Q
102
5? 2
2
2
and MRR,0
are contributions from supergravity, whose components are of the order
MLL,0
of the light-sfermion mass squared m20 and induce flavor violation, 4
!
!
2
2
2
2
O(m
O(m
)
O(m
)
)
O(m
)
0
0
0
0
2
2
MLL,0
=
=
,
MRR,0
.
(28)
O(m20 ) O(m20 )
O(m20 ) O(m20 )

The SUSY contribution to K 0 K 0 mixing is controlled by two parameters qualitatively,


as shown in Appendix C: the averaged squark masses m
2LL, RR and the off-diagonal
2LL and m
2RR are given by
elements LL, RR . 5 The averaged squark masses m
q
q
2
2
X m2 ,
X
Q
m

=
QX
(29)
m
2LL = QX
RR
101 102 D
5? 1 Q5? 2 mD ,
where we have neglected O(m20 ) contributions. On the other hand, LL and RR are the
off-diagonal elements of the squark mass matrices normalized with the averaged squark
masses, in a basis where the quark mass matrix is diagonal (see also Appendix C). In order
to change the U(1)X gauge basis to this basis, we introduce unitary matrices V L and V R
which diagonalize the down-type quark Yukawa matrix given in the U(1)X gauge basis.
We parameterize them as


cos L eiL sin L
L
,
(30)
V =
eiL sin L
cos L
and V R with the replacement L R in V L . Then, the squark mass matrices, M2LL and
M2RR , in a basis where the quark mass matrix is diagonal are given by
4 If the U(1) charges for the first and second generations are different, the off-diagonal elements of M 2
X
LL,0
X
|QX
|QX
QX
|
? Q5? |
2
10
10
5
1
2 and (hi/Mpl )
1
2 , respectively.
and MRR,0 are suppressed by (hi/Mpl )
5 In the following figures, we use the exact formula for the K 0 K 0 mixing given in Appendix C.

J. Hisano et al. / Nuclear Physics B 584 (2000) 345


2
M2LL = V L MLL
V L,

2
M2RR = V R MRR
V R.

Consequently, the off-diagonal element LL is represented as follows:


 2 
 m2D
m0
X

Q
+
O
LL = eiL sin L cos L QX
,
101
102
2
m
LL
m
2LL

17

(31)

(32)

X
X
X
and RR is obtained with the replacement L R and QX
10i Q5? i . If Q101 6= Q102 , LL
is dominated by the U(1)X D-term contribution and is given definitely up to the mixing
X
angle sin L . On the other hand, if QX
101 = Q102 , the U(1)X D-term contribution vanishes
and the second term dominates LL . In this case, the off-diagonal element has a large model
dependence since we cannot calculate the supergravity contributions.
The contribution to K 0 K 0 mixing from the gluinosquark box diagram is calculated
in Appendix C. We have included the leading-order QCD corrections [51,34] used the bag
parameters obtained by lattice calculations [52,53]. The details are given in Appendix C.
The constraints from 1mK are summarized as follows:

m
LL, RR
,
(2535) TeV
(m
LL m
RR )1/2
.
(LL RR )1/2 <
(150250) TeV
LL, RR <

(33)
(34)

This shows that if LL and RR are of the same order, Eq. (34) gives about ten times as
severe bounds on the heavy squark masses as Eq. (33). Furthermore, if LL and/or RR
have CP-violating phases of order unity, the constraints from K give twelve times severer
bounds,
m
LL, RR
,
(35)
(310430) TeV
(m
LL m
RR )1/2
,
(36)
(LL RR )1/2 <
(19003100) TeV
as discussed in Appendix C.
Now, let us discuss Models IIV in turn. In Model I, the averaged squark masses are
LL, RR <

2RR = m2D .
m
2LL = m

(37)

Since the U(1)X charges for the first-two generations are the same, the D-term contribution
does not induce any flavor violation. Thus, Model I is the hybrid scenario of the decoupling
and degeneracy in a sense. Then, the flavor violation comes from the supergravity
contributions. Although we cannot calculate the off-diagonal elements in this case, we
LL, RR . For simplicity, we here assume
expect LL,RR to be of order (0.11)m20 /m
LL =

m20
m2D

RR =

m20
m2D

Then, we obtain the bound on the light sfermion mass m0 from Eq. (34) as


m0 3
TeV,
m0 > (150250)
mD

(38)

(39)

18

J. Hisano et al. / Nuclear Physics B 584 (2000) 345

if LL and RR do not have CP-violating phases. The ratio m0 /mD is bounded from below
in Fig. 2 in the previous section, so that the lower bounds on m0 can be estimated as
follows:
ESSM
m0 /mD > 0.18
m0 > (0.91.5) TeV

NESSM
m0 /mD > 0.07
m0 > (60100) GeV

(40)

Here, the bounds on m0 /mD are those for zero gaugino mass. Note that the lower bound
on m0 depends on the third power of the ratio m0 /mD in Eq. (39). As a result, although the
bound on m0 /mD in the NESSM is only 1/2.5 of that in the ESSM, the bound on the light
sfermion mass m0 in the NESSM becomes much smaller than in the ESSM by a factor of
(1/2.5)3 1/15.
If LL and RR have CP phases of order unity, the constraint from K gives twelve times
larger bounds on m0 than that from 1mK as,
ESSM
m0 /mD > 0.18
m0 > (1118) TeV

NESSM
m0 /mD > 0.07
m0 > (0.81.3) TeV

(41)

Thus, it seems to require some tuning of the phases or electroweak symmetry breaking
even in the NESSM. In order to reduce the bound by a factor of 3, for example, the phase
of LL RR must be tuned to be 1/32 0.1. However, including the gaugino contributions
reduces the lower bound on m0 , so that we can, in fact, realize the hierarchical spectrum
without tuning as shown in Fig. 6.

Fig. 6. Lower bounds on the bare masses for the light sfermions in Model I. The horizontal axis
represents the gluino running mass evaluated at the gluino decoupling scale, and the vertical axis the
indicated light sfermion mass at the GUT scale. The regions below the curves are excluded due to the
negative mass squareds for the corresponding sfermions. Bold lines represent the constraints from
1mK . Fine lines represent the constraints from K with O(1) CP phases.

J. Hisano et al. / Nuclear Physics B 584 (2000) 345

19

In Fig. 6, we have plotted the lower bounds on m0 , the masses for the various light
squarks and sleptons at the GUT scale, including the gaugino contributions. The horizontal
axis represents the running gluino mass evaluated at the gluino decoupling scale, which is
twice as large as the GUT-scale gaugino mass m1/2 in the previous section. The boundary
conditions at the GUT scale, such as m = 2.0 mD , are the same as in the previous
section. The only difference from the previous section is that mD is not fixed to 10 TeV
but varies with the constraints from K 0 K 0 mixing.
Fig. 6 shows that there is no constraints from 1mK in the NESSM. Moreover, the
NESSM solves the CP problem in K 0 K 0 mixing without tuning of the phases or
electroweak symmetry breaking. Note that the shape of the excluded region in Fig. 6 is
very different from that in Fig. 2. This is because mD can take a smaller value as m0
decreases,
mD > const (m0 )2/3 ,

(42)

as can be seen from Eq. (39). This is peculiar to the case where the flavor violation comes
from the supergravity contributions. In contrast, in the case where the U(1)X D-term
contributes to the flavor violation such as in Models IIIV, the lower bound on mD does
not change with m0 , so that the shape of the excluded region of m0 is similar to that of
m0 /mD in the previous section.
Now, we turn to the other models. In Model II, the averaged squark masses are given by

m
2RR = m2D .
(43)
m
2LL = 2 m2D ,
The left-handed down-type squarks of the first and second generations have different
U(1)X charges, so that the U(1)X D-term contribution induces flavor violation. In this
case, the mixing angle L is necessary to determine LL . Since the product of V L and the
diagonalizing matrix for the left-handed up-type quark gives the Cabbibo angle sin C =
0.22, it is natural to take sin L sin C . Thus, we here take sin L = 0.22 and set the
off-diagonal element LL as
sin L cos L
0.15.
(44)

2
On the other hand, RR is determined from the supergravity contributions and is more
model-dependent than LL . Therefore, we here restrict our attention to the flavor violation
from the U(1)X D-term contribution and give the lower bounds on the light sfermion
masses using only the constraint on LL . The supergravity contributions are discussed later.
Then, we obtain the following bound from Eq. (33):


m0
TeV.
(45)
m0 > (3.24.4)
mD
Since the bounds on the ratio m0 /mD are given in Fig. 3 in the previous section, we can
estimate the lower bounds on m0 from 1mK as
LL =

ESSM
m0 /mD > 0.22
m0 > (700970) GeV

NESSM
m0 /mD > 0.09
m0 > (290400) GeV

(46)

20

J. Hisano et al. / Nuclear Physics B 584 (2000) 345

Fig. 7. Lower bounds on the bare masses for the light sfermions in Model II. The constraints from
1mK (Eq. (33)) are plotted.

In Fig. 7, we have plotted the lower bounds on m0 in Model II including the effect of the
gaugino masses. It shows that we can take more natural mass scale for the light sfermions
in the NESSM than in the ESSM. If LL has a phase of order unity, however, the lower
bounds on the light sfermion masses become higher by a factor of 12. Therefore, the phase
is necessary to be less than 102 to avoid extreme fine tuning of the electroweak symmetry
breaking.
Before turning to Model III, we comment on the supergravity contributions. If we take
RR = m20 /m2D as we did in Model I, we obtain the following lower bounds on m0 from
Eq. (34): m0 > (2.64.3) TeV in the ESSM and m0 > (430720) GeV in the NESSM.
Thus, it seems to require a slight tuning of the electroweak symmetry breaking even in the
NESSM. However, the size of RR is strongly model-dependent as we emphasized, and if
we take RR = (1/4) m20 /m2D , for example, the bounds on m0 become half. Moreover,
the bounds are greatly lowered by including the effect of the gaugino masses as in Fig. 6.
Therefore, the supergravity contributions are less important to restrict the parameters in
Model II.
In Fig. 8, we have plotted the lower bounds on m0 in Model III.Here, we have taken
sin L = 0.22 as in Model II. Since the U(1)X charge assignment for the first-two generations in Model III is the same as in Models II, K 0 K 0 mixing gives the same constraints
on both models. The lower bounds on the mass ratios m0 /mD are also almost the same
between Models II and III as shown in the previous section. Therefore, the bounds on m0
in Fig. 8 are similar to those in Fig. 7.
In Model IV, both LL and RR come from the large U(1)X D-term contribution, so that
the constraint is the severest among Models IIV. The averaged squark masses are given by

m
2RR = 2 m2D .
(47)
m
2LL = 2 m2D ,
If we set both angles L and R equal to the Cabbibo angle for simplicity, then the LL and
RR are given by

J. Hisano et al. / Nuclear Physics B 584 (2000) 345

21

Fig. 8. Lower bounds on the bare masses for the light sfermions in Model III. The constraints from
1mK (Eq. (33)) are plotted.

Fig. 9. Lower bounds on the bare masses for the light sfermions in Model IV. The constraints from
1mK (Eq. (34)) are plotted.

LL =

sin L cos L
0.15,

RR =

sin R cos R
0.15.

Substituting these values into Eq. (34), we obtain the following bound:


m0
TeV.
m0 > (1932)
mD

(48)

(49)

The bounds on the ratio m0 /mD are given in Fig. 5, so that the lower bounds on m0 can be
estimated as
ESSM
m0 /mD > 0.23
m0 > (4.47.4) TeV

NESSM
m0 /mD > 0.10
m0 > (1.93.2) TeV

(50)

22

J. Hisano et al. / Nuclear Physics B 584 (2000) 345

In Fig 9, we have plotted the lower bounds on m0 in Model IV including the effect of the
gaugino mass. It shows that an extreme fine tuning is required in Model IV. We can reduce
the degree of fine tuning by taking smaller values for the mixing angles sin L and sin R .
However, in order to reduce the bounds on the masses for both the light sfermions and
the gluino to 500 GeV, we must take sin R less than 0.03 and it causes another tuning
problem. In consequence, it seems difficult to realize Model IV without fine tuning in a
framework of the NESSM, even if there is no CP phase.

5. Other renormalization-group effects


In this section, we discuss other RG effects on the light sfermion masses which we
have not considered in the previous sections. There are three types of contributions
which potentially affect the previous analyses: the three-loop contribution through the
gauge couplings, the contribution from the bottom and tau Yukawa couplings, and the
contribution from the top Yukawa coupling.
First, we consider the correction to the light sfermions from the RG equations at the
three-loop level. We here adopt the SDR scheme defined by the analytic continuation into
superspace [54]. The RG equations for the SUSY-breaking scalar mass squareds in this
0
scheme coincide with those in the DR scheme at the two-loop level, so that SDR scheme
0
is considered to be all-order definition of the DR scheme.
When the SUSY-breaking scalar mass squareds are regarded as a D-term of an external
U(1)X gauge multiplet, the RG contributions from the gauge interactions can be divided
into two classes in the limit of vanishing gaugino mass in this scheme. One is the
contribution from mixed anomalies between the external U(1)X and the internal gauge
symmetries. The other is the contribution from the kinetic-term mixing between U(1)
symmetries, which exists only if there is an internal U(1) gauge symmetry. Since we have
imposed Eqs. (3)(5) in the NESSM, the RG contributions to the light sfermion masses
from the heavy sfermions and extra matters only come from the U(1)Y and U(1)X mixing
contribution at the three-loop level [37,38,55]. Then, the contributions are suppressed
by small Y and negligible compared with the two-loop finite correction. Furthermore,
in models where the extra matters are the fundamental representation of SU(3)C and
SU(2)L , the contributions from the heavy sfermions and extra matters in the three-loop
RG equations are proportional to 3 2 Y or Y3 , since the terms proportional to 32 Y and
22 Y vanish due to Eqs. (4), (5) in this case. Thus, the three-loop RG contributions are
more strongly suppressed in such models.
Next, we consider the effect of the bottom and tau Yukawa couplings. In Models II,
IV the right-handed bottom squark and the left-handed slepton of the third generation
obtain masses from the U(1)X D-term due to QX
5? 3 6= 0. Then, the bottom and tau Yukawa
couplings may drive the mass squareds for the doublet squark of the third generation and
the right-handed stau to be negative at the one-loop level. The RG contributions to the
doublet squark of the third generation and the right-handed stau, m2q3 , yukawa and me23 , yukawa ,
are given as

J. Hisano et al. / Nuclear Physics B 584 (2000) 345



m2q3 , yukawa ' 0.003% 1 + tan2 m2D + |Ab |2 ,


me23 , yukawa ' 0.005% 1 + tan2 m2D + |A |2 .

23

(51)
(52)

Here, we have assumed that tan is not so large, and Ab and A are the soft SUSY-breaking
trilinear scalar couplings associated with the bottom and tau Yukawa couplings. Then, the
RG contribution to the doublet quark of the third generation is negligible if tan < 10,
compared with the finite correction. On the other hand, the RG contribution to the righthanded stau can be larger than the finite correction. However, the bound on the bare mass
given by these contributions is still looser than the bounds on the light squarks from the
finite correction, as long as tan is sufficiently small. Since Models II, IV predict the small
bottom and tau Yukawa coupling constants due to QX
5? 3 6= 0 and thus require small tan ,
these contributions are sufficiently suppressed.
Finally, we discuss the effect of the top Yukawa coupling. In the previous sections, we
have neglected it, since the effect of the top Yukawa coupling depends on the supergravity
contributions, which are model-dependent. There are two negative contributions to the topsquark mass squared through the top Yukawa coupling. One is the RG contribution which
mainly comes from the up-type Higgs mass mhu . The other is the contribution from the
leftright mixing term, which gives the negative contribution in diagonalizing the topsquark mass matrix. The leftright mixing term for top squark is given by

(53)
Mt2 LR = mt (At + cot ),
where mt is the top quark mass, At the soft SUSY-breaking trilinear coupling associated
with the top Yukawa coupling and the supersymmetric mass for the Higgs doublets.
Thus, it turns out that and At play an important role in calculating the negative
contributions. However, At at the weak scale is almost saturated by the radiative correction
from the gluino, so that we take At = 0 at the GUT scale for simplicity in the following
calculations. The size of is determined by the electroweak symmetry breaking.
In Figs. 10 and 11, we have plotted the lower bounds on the bare masses for top squarks.
Here, we have assumed the universal scalar mass at the GUT scale except for the up-type
Higgs mass, and taken At = 0 at the GUT scale for simplicity. In the figures, we have
shown two extreme cases for the up-type Higgs mass in each choice for the sign of the
parameter: one is the universal case mhu = m0 and the other is mhu = 0. If is positive,
the At and are added up constructively in the leftright mixing term, since At receive
the positive contribution from the gaugino mass. Thus, the lower bound on the top squark
mass is somewhat severe in this case. On the other hand, if is negative, the cancellation
between two contributions from At and occurs, so that the bound is weaker than the case
with > 0.
In the region where m0 is much larger than the gaugino masses, the up-type Higgs mass
mhu gives the dominant effect. Thus, the effect of the top Yukawa coupling is much modeldependent in this region, since we cannot predict the up-type Higgs mass at the GUT scale.
For instance, if the up-type Higgs mass is much smaller than the top squark mass at the
GUT scale, the effect is negligible as shown in Figs. 10 and 11. On the other hand, in the
region where m0 is small enough, the dominant effect of the top Yukawa coupling comes

24

J. Hisano et al. / Nuclear Physics B 584 (2000) 345

Fig. 10. Lower bounds on the bare masses for the light sfermions in Model I are plotted by combining
the constraint from K (Eq. (36)) with the condition that mt1 , mt2 > 0, where t1 and t2 represent the
mass eigenstates for top squarks. The regions below the curves are excluded due to the negative
mass-squared eigenvalues for top squarks. Here, we have set tan = 3.0.

Fig. 11. Lower bounds on the bare masses for the light sfermions in Model II are plotted by
combining the constraint from 1mK (Eq. (33)) with the condition that m2 , m2 > 0. The regions
t1
t2
below the curves are excluded due to the negative mass-squared eigenvalues for top squarks. Here,
we have set tan = 3.0.

from the leftright mixing term through diagonalization of the top-squark mass matrix.
Since At term is generated radiatively by the gluino as At mg , it gives the negative
contribution to the top-squark mass eigenvalue of order mt At mt mg . Thus, in order
to obtain the positive top-squark mass squared at the weak scale, we have to take a larger
mass for the gluino than the case without top-Yukawa contributions. The required increase
mg of the gluino mass is determined by the following inequality: 2mg mg & mt mg , since

J. Hisano et al. / Nuclear Physics B 584 (2000) 345

25

the positive contribution from the gluino is estimated as of order m2g . That is, we have to
take a larger mass for the gluino by the order of 100 GeV compared with the previous case
as shown in Figs. 10 and 11.
To summarize, the lower bounds on the light sfermion masses become somewhat severer
by including the effects of the top Yukawa coupling. However, we have found that we
can still take the light sfermion and gluino masses as small as 400 GeV in the natural
effective SUSY, while we have to take them larger than 1 TeV in the original effective
SUSY.

6. Conclusions and discussion


In this article, we have calculated the finite corrections to the light sfermion masses
from the heavy sfermions and extra matters at the two-loop level in the NESSM, assuming
that all fields are embedded in the SU(5)GUT multiplets at the GUT scale. While the
sfermion mass squareds in the NESSM are RG invariant at the two-loop level in the limit
of vanishing gaugino-mass and Yukawa-coupling contributions, the finite corrections may
still drive the light-sfermion mass squareds to be negative. The finite corrections increase
(decrease) if the supersymmetric masses (the holomorphic SUSY-breaking masses) of the
extra matters become large. We have found that the corrections can be less than (0.31)%
of the heavy-squark mass squareds.
We have also discussed whether the models in which the NESSM is realized by the
U(1)X gauge interaction at the GUT scale are viable or not in the light of the experimental
constraints from 1mK and K . On this setup, the supergravity contributions to the
SUSY-breaking terms, which are generically non-universal in flavor space, are suppressed
compared with the U(1)X D-term contribution. We have found that, when both the left- and
right-handed down-type squarks of the first-two generations have common U(1)X charges,
the supersymmetric contributions to 1mK and K are sufficiently suppressed without
breaking naturalness, even if the flavor-violating supergravity contributions to the sfermion
mass matrices are included. On the other hand, when only the right-handed squarks of
the first and second generations have a common U(1)X charge, we can still satisfy the
constraint from 1mK naturally, but evading the bound from K requires somewhat small
CP phase of order 102 .
The formulae we have given in Appendix B are applicable to any models which have
hierarchical mass spectrum for the sfermions such as effective SUSY. In particular, the
formulae for the finite corrections can also be applied to the models where the U(1)X Dterm is generated at much lower energy than the GUT scale [2527], since they do not have
explicit renormalization-point dependence. Then, it gives the lower bounds on the lightsfermion bare masses similar to the ones we have derived in this article. Thus, we believe
that our result gives the most conservative bound for the naturalness in the effective SUSY
models.

26

J. Hisano et al. / Nuclear Physics B 584 (2000) 345

Acknowledgements
We would like to thank Hitoshi Murayama and Youichi Yamada for useful discussions.
K.K. and Y.N. thank the Japan Society for the Promotion of Science for financial support.
This work was supported in part by the Grant-in-Aid for Scientific Research from the
Ministry of Education, Science, Sports and Culture of Japan, on Priority Area 707
Supersymmetry and Unified Theory of Elementary Particles and by the Grant-in-Aid
No.10740133 (J.H.).

Appendix A. Anomalous U(1) model


In this appendix, we briefly review the model given in Refs. [37,38] and show that it
naturally realizes the setup of the NESSM. The model is based on the anomalous U(1)
SUSY-breaking model [18,19].
The anomalous U(1) gauge symmetry frequently appears in low-energy effective
theories of string theory. The matter content is anomalous under this U(1) symmetry, but its
anomalies are canceled by a nonlinear transformation of the dilaton chiral multiplet [56].
This leads to the generation of a nonzero FayetIliopoulos (FI) D-term which is smaller
2 . We
than the reduced Planck scale Mpl by a one-loop factor, 2 ' (g 2 /192 2 ) Tr QMpl
parameterize it as Mpl with  = O(0.1), and take the sign convention of U(1) charges
such that 2 > 0.
We identify this anomalous U(1) gauge symmetry with U(1)X given in the text. Thus,
X
we assign the U(1)X charges QX
10i and Q5? i to the quark and lepton chiral multiplets 10i
X
X
and 5? i , respectively (Q10i , Q5? i > 0). We also introduce extra-matter chiral multiplets
X
X
X
5ex and 5? ex with U(1)X charges QX
5ex and Q5? ex (Q5ex , Q5? ex < 0), in order to satisfy
Eqs. (6), (8). Then, the SUSY-breaking model is constructed as follows.
We consider the SU(Nc ) SUSY gauge theory with Nf flavors Qa and Q a (Nc /2 <
Nf < Nc ), and introduce two singlet chiral superfields 1 and 2 . Here, a, a = 1, . . . , Nf
X
represent flavor indices. We assign the U(1)X charges as Qa ((QX
5ex + Q5? ex )/2),
X
X
X
X
Q a ((Q5ex + Q5? ex )/2), 1 (1) and 2 (1 + Q5ex + Q5? ex ). The tree-level superpotential
of the model is given by

Qa Q a
(f )aa 1 2 + (f )aa 5ex 5? ex .
(A.1)
Wtree =
Mpl
Then, the dynamical superpotential is generated by nonperturbative effects of the SU(Nc )
gauge interaction, and the effective superpotential of the model is exactly given by

1
1 2 Tr (f M) + 5ex 5? ex Tr (f M)
Weff =
Mpl
 3Nc Nf  1
Nc Nf

,
(A.2)
+ (Nc Nf )
det M
in terms of gauge-invariant composite fields M a = Qa Q a [57]. Here, is the dynamical
a

scale of the SU(Nc ) gauge theory. The D-term potential for U(1)X is given by

J. Hisano et al. / Nuclear Physics B 584 (2000) 345

VD =

2
 p
gX
X

2 QX
5ex + Q5? ex Tr M M
2


X
2
X
2
X
2 2
.
|1 |2 + 1 + QX
5ex + Q5? ex |2 | + Q5ex |5ex | + Q5? ex |5? ex |

27

(A.3)

In this appendix, we use the same letter for the chiral superfield and its scalar component.
The dynamics of the model can be understood as follows. First, nonperturbatively
generated superpotential forces M to have VEVs, which gives supersymmetric mass terms
for the singlet fields and the extra matter fields. The large FI D-term, , is absorbed by
the shift of the singlet fields 1 and 2 . The point is that if f is larger than f , only
the singlet fields shift to absorb the FI D-term and the extra-matter fields do not develop
VEVs, avoiding phenomenologically disastrous large breaking of the SM gauge groups.
We assume that this condition, f > f , is satisfied. Then, since the singlet fields have the
supersymmetric mass induced by the VEVs of M, they cannot absorb completely and
nonzero U(1)X D-term remains. This nonvanishing D-term gives the MSSM sfermions
and extra scalars the SUSY-breaking mass squareds proportional to their U(1)X charges.
Now, let us minimize the potential explicitly. The minimum of the potential can be
obtained by making an expansion in the small parameter (/Mpl )(3Nc Nf )/Nc  1. To
the leading order, the VEVs of the fields 1 , 2 and M at the minimum are given by
solving the following equations:

X
2
2
(A.4)
|1 |2 1 + QX
5ex + Q5? ex |2 | = ,

2
2
2
|1 | (Nf Nc )|1 | + Nf |2 |


X
2
2
2
(A.5)
= 1 + QX
5ex + Q5? ex |2 | (Nf Nc )|2 | + Nf |1 | ,
 3Nc Nf  1
Nc Nf

(f )aa
=
1 2 .
(A.6)
M 1 aa
det M
Mpl
From Eqs. (A.4), (A.5), we find that both the 1 and 2 fields have nonvanishing VEVs
of order , and the VEV of the M field is given by Eq. (A.6) as
h1 i ' h2 i ' ,

hMi '

3Nc Nf
Nc

Mpl
2

 Nc Nf
Nc

(A.7)

Then, the Yukawa matrices for the quarks and leptons are generated through the VEV of
the 1 field suppressed by suitable powers of h1 i/Mpl ' /Mpl =  (see Eq. (13)).
Calculating higher order in (/Mpl )(3Nc Nf )/Nc , we find that the VEVs of the auxiliary
fields can be written in terms of the VEVs of the scalar fields as
FM

a
a

F
1

F
2

2

Nc Nf 1 |1 |2 + |2 |2
=
Tr(f M) f1 aa ,

Nc Nf Mpl
1 2


2
=
Tr(f M),
Mpl


1
=
Tr(f M),
Mpl

(A.8)
(A.9)
(A.10)

28

J. Hisano et al. / Nuclear Physics B 584 (2000) 345

DX =



2Nf Nc 1
|1 |2 + |2 |2
Tr(f M) 2 .
X
X
2
Nc
Mpl |1 |2 (1 + Q5ex + Q5? ex )|2 |2

(A.11)

Substituting Eq. (A.7) into Eqs. (A.8)(A.11), we obtain


FM '

2(3Nc Nf )

1/Nc

,
2N Nc
4(Nc Nf ) Mpl f

1/Nc
3Nc Nf

,
F ' F '
N
1
2
2(Nc Nf ) Mpl f

2/Nc
3Nc Nf
.
DX '
N
2(Nc Nf ) Mpl f

(A.12)

(A.13)

(A.14)

The dynamical scale is determined so that |DX | ' (110) TeV to give the heavy
sfermions multi-TeV masses. From Eqs. (A.12)(A.14), we find the following useful
relation:






F1 2 F2 2 FM 2
'
'

(A.15)
|DX | '
h i
hMi .
h1 i
2
This relation ensures that flavor-breaking supergravity contributions are an order of
magnitude smaller than the contribution from the U(1)X D-term, since their sizes are
estimated as 6




p

F1



' F1 '  |DX | ' 102 103 GeV.
(A.16)
M M h i
pl
pl
1

From Eqs. (A.7), (A.14), we also find that |DX | ' hMi/Mpl . Thus, the supersymmetric
mass for the extra matters is actually the same order with the SUSY-breaking masses for
the heavy sfermions.
The gaugino masses arise from the F -term of the dilaton field [48]. Their sizes can
be of the order of the weak scale [49], and then it does not much affect the preceding

analysis of the dynamics. The SUSY-breaking trilinear scalar couplings of order |DX |
are also generated by the superpotential which generates Yukawa matrices for the quarks
and leptons, except for the ones which involve only the light sfermions (see Eqs. (14),
(15)).
Finally, we comment on the anomaly. We have identified the U(1)X in the text with the
anomalous U(1) gauge symmetry. Then, Eq. (6) might seem contradicted by the fact that
the anomalous U(1) gauge symmetry has mixed anomalies for all the other gauge groups
including the SM ones. However, it is not a contradiction. Since U(1)X is broken down at
very high energy scale of order , it does not necessarily mean that the matter content is
6 Since the gravitino mass is of the order of the weak scale, we can induce -term (supersymmetric mass
term for the Higgs field, W = Hu Hd ) of the desired size (the weak scale) by introducing a holomorphic term
K = Hu Hd in the Khler potential [58].

J. Hisano et al. / Nuclear Physics B 584 (2000) 345

29

anomalous below scale. That is, if we introduce fields 5anom and 5? anom of masses of
order with the superpotential
W h1 i5anom 5? anom ,

(A.17)

we can match the anomalies as required by the anomalous U(1) symmetry, keeping

Eqs. (6), (8) satisfied between two scales |DX | and . Then, the large two-loop RG
contributions are absent below the scale as we have explained above. 7

Appendix B. Finite corrections


In this appendix, we calculate the contributions to the gaugino and the light sfermion
masses arising from loops of the extra-matter multiplets and the heavy sfermions. We use
0
the DR scheme [36] to regularize the theory, since we have adopted the SDR scheme
0
[54] which is all-order definition of the DR scheme in the text (see Section 5). In this
scheme, the -scalar mass mA is zero at the tree level and it does not appear in the relation
between physical quantities. However, if the supertrace of the matter fields is nonzero, mA
receives divergent radiative correction at one loop through loops of the sfermions, so that
the counterterm is needed to cancel this divergence. The insertion of this counterterm gives
divergent contribution to the sfermion masses at one loop. Thus, we have to carefully treat
the -scalar in order to obtain the two-loop contribution to the light sfermion masses when
the supertrace is nonvanishing.
First, we explain our notations. The superpotential of the vector-like extra matters, ex
and ex , is given by
Wex = m ex ex .

(B.1)

We denote the scalar and fermion components of ex ( ex ) as ex and ex ( ex and ex ),


respectively. In addition, the extra scalars have the SUSY-breaking mass terms


2 ex ex .
e2 ex
(B.2)
ex m
Lex, soft = F ex ex + h.c. m
Then, the mass terms for the extra scalars are written as
!
 2 ex

e
Lex = ex , ex Mex ;
ex
!
e2
F
|m |2 + m
2
eex =
.
M
2
F
|m |2 + m

(B.3)
(B.4)

2 can be diagonalized by the unitary matrix V as


eex
The mass matrix M
7 The light-sfermion mass squareds may receive negative RG and finite contributions above and at the
decoupling scale of 5anom and 5? anom . However, we can make these negative contributions smaller than
the supergravity contributions by choosing their group-theoretical factors to be small, since there is no large
log-factor in the contributions. We include these contributions in the supergravity contributions when we make
phenomenological analyses in the text.

30

J. Hisano et al. / Nuclear Physics B 584 (2000) 345


2
eex
diag m21 , m22 = V M
V .
We parameterize V as


cos ei sin
,
V=
ei sin cos

(B.5)

(B.6)

and define
y1

m21
m2

y2

m22
m2

(B.7)

We calculate the gaugino and the light sfermion masses arising from loops of the extramatter multiplets in terms of these parameters. Then, the contribution from the heavy
sfermions are obtained by taking the limit m 0 and 0.

B.1. The gaugino masses


The gauginos acquire their masses through the one-loop diagram of the extra-matter
multiplet shown in Fig. 12. If there is a pair of vector-like extra matters, it is given by
 2
2
X
gA
m
m2
A
T m
V2 V1
log
,
(B.8)
mgA =
2
2
2
4
m m
m2
=1,2
where A = 13 represents the standard-model gauge groups, and we have adopted the

SU(5) GUT normalization for the U(1)Y gauge coupling (g1 5/3 gY ). Here, T A is
the Dynkin index of the extra-matter representation, in a normalization T A = 1/2 for a
fundamental of SU(N ) and T 1 = (3/5)Y 2 for U(1)Y . Note that T A is the Dynkin index of
the extra matter and not the sum of the index of the extra matter and anti-matter. That is,
we use T A = 1/2 for a vector-like pair of fundamental extra matters.
Then, the gaugino masses are given by

 

A
y1 log y1 y2 log y2 y1 y2 log(y1 /y2 )
A
i
TrR TR m e
,
sin 2
mgA = 2
4
(y1 1)(y2 1)
(B.9)
where the trace is taken over pairs of the extra matters R. (Note that m , y1 , y2 , and
depend on R.) This is in agreement with the result given in Ref. [59].

Fig. 12. Diagram contributing to the gaugino masses.

J. Hisano et al. / Nuclear Physics B 584 (2000) 345

31

Obviously, the heavy sfermions do not contribute to the gaugino masses. It can also be
seen by taking the limit m 0 and 0 in Eq. (B.9).
B.2. The sfermion masses
We now calculate the light sfermion masses at the two-loop level. There are two types
of contributions which generate the light sfermion masses. One is the contribution directly
arising from the loop diagrams: the one-loop graph involving the -scalar and the two-loop
2 . The other contribution is
graphs involving heavy particles, both of which are of order A
that through the generation of U(1)Y FI D-term at one loop and two loops, which are of
order Y and Y A , respectively.
We first consider the direct contribution from the extra-matter multiplets. One diagram
contributing to the light sfermion masses is the one-loop -scalar graph shown in Fig. 13.
It contains the counterterm m2 for the -scalar mass, L = (1/2)m2 A a A a , and gives
A

the light-sfermion mass squared,


m2f, A =

2 0() 2  2 X 2 A

gA Cf m2 ,

(4)2 IR
A

(B.10)

in D = 4 2 dimensions. Here, C A is the quadratic Casimir coefficient for the light


f

sfermion f, is the renormalization scale, and IR is the infra-red cutoff. The counterterm
m2 is determined to cancel the divergence of the -scalar mass arising from one loop of
A

the extra-matter multiplets,

(B.11)
as
m2

 
2
h
i 1
2gA
A
2
2
2
=
TrR TR m1 + m2 2m
.
(4)2


Substituting Eq. (B.12) into Eq. (B.10), we obtain



i 1
h
X A 2
2IR
2
A
A 2
log 2 .
Cf TrR TR m (y1 + y2 2)
mf, A = 4
4


(B.12)

(B.13)

Fig. 13. One-loop diagram contributing to the light sfermion masses which involves the -scalar A.

32

J. Hisano et al. / Nuclear Physics B 584 (2000) 345

Here, the combination + log(4) has been absorbed by an appropriate redefinition

of the renormalization scale, e /4 (where is the Euler number). The result


2 , so that it contributes to the two-loop RG equations for the light
has 1/ pole of order A
sfermion masses. We will omit , hereafter.
The remaining diagram consists of two-loop graphs involving the extra-matter multiplets
given in Fig. 14. These graphs are identical to those considered in Ref. [50] in the
case of vanishing supertrace. Their contribution is ultra-violet finite even in the case of
nonvanishing supertrace. Together with Eq. (B.13), we obtain the light sfermion masses
induced directly by loops of the extra-matter multiplets as
i 1 
h
X A 2

2
A
A 2

Cf TrR TR m (y1 + y2 2)
mf, direct R = 4
4

A




X A 2

CfA TrR TRA m2 (y1 + y2 2) log m2 + 2
+4
4
A





1
1
+ (y1 log y1 + y2 log y2 ) 2 y1 Li2 1
+ y2 Li2 1
y1
y2





y2
y1
1
+ y2 Li2 1
,
(B.14)
+ sin2 2 y1 Li2 1
2
y1
y2
R1
which is in agreement with the result given in Ref. [59]. Here, Li2 (x) = 0 dt log(1
xt)/t is the dilogarithm function. Note that 2IR is canceled between the diagrams of
Figs. 13 and 14. Taking the limit m 0 and 0 and regarding m21 as the heavy
sfermion masses m2 , we obtain the light sfermion masses induced by loops of the heavy
F
sfermions as
h
i 1 
X A 2

2
CfA TrF TFA
m
m2f, direct F = 4

F
4

A





X A 2
1 2
A
A 2
2
Cf TrF TF mF log mF + 2
+4
,
(B.15)
4
3
A

to the light
where F denotes the heavy sfermions. The direct contribution m2
f , direct
sfermion masses is given by summing up that from the extra-matter multiplets and that
from the heavy sfermions,


(B.16)
m2f, direct = m2f, direct R + m2f, direct F .
We next consider the light sfermion masses induced via the generation of U(1)Y FI
D-term. At the one-loop level, the light sfermion masses generated by loops of the extramatter multiplets are given by (see Fig. 15)

 

h

i 1
Y
2
2
2

Y TrR YR cos 2 m1 m2
+1
mf, FI-1loop R =
4 f

 
h
i
Y
(B.17)
Yf TrR YR cos 2 m21 log m21 m22 log m22 .
+
4

J. Hisano et al. / Nuclear Physics B 584 (2000) 345

33

Fig. 14. Two-loop diagram contributing to the light sfermion masses which involves the extra-matter
multiplets ex and ex .

Fig. 15. One-loop diagram contributing to the light sfermion masses through the generation of U(1)Y
FI D-term.

34

J. Hisano et al. / Nuclear Physics B 584 (2000) 345

Fig. 16. Two-loop diagram contributing to the light sfermion masses through the generation of U(1)Y
FI D-term.

This contribution vanishes if the extra-matter multiplets have an invariance under the
parity ex ex , as can be readily seen by taking = /4 in the expression. The heavy
sfermions also generate one-loop U(1)Y FI D-term, which gives

 
h
i 1

Y
Yf TrF YF m2F
+1
m2f, FI-1loop F =
4

 
h
i
Y
(B.18)
Yf TrF YF m2F log m2F .
+
4
The total contribution through the generation of U(1)Y FI D-term at one loop is


m2f, FI-1loop = m2f, FI-1loop R +m2f, FI-1loop F .

(B.19)

U(1)Y FI D-term is also generated at two loops through the diagram shown in Fig. 16.
The resulting light sfermion masses from extra-matter loops are written as

X
X




2
2 2
A

gY gA Yf TrR YR CR
V2 V2
V1 V1
I1 m2
mf, FI-2loop R =
A

+4

V1 V1



V2 V2
I2 m2


V2 V2
V2 V2
V1 V1
V1 V1

J. Hisano et al. / Nuclear Physics B 584 (2000) 345

35


V 1 V1
V 2 V2
I3 m2 , m2 , m2

X



2
2
V1 V1 V2 V2 (m2 ) Ic m , m
,
+

(B.20)

in the Feynman gauge. Here, functions I s are defined as


Z 4 Z 4

d p
d k
1
(2p k)2
1
2
, (B.21)
I1 m =
(2)4 (2)4 (p2 m2 )2
k2
(p k)2 m2
Z 4 Z 4

d p
1
k2 k p
1
d k
, (B.22)
I2 m2 =
(2)4 (2)4 (p2 m2 )2
k2
(p k)2 m2
Z 4
Z 4

d p
d q
1
1
1
,
(B.23)
I3 m2 , m2 , m2 =
4
2
2
2
4
2
2
(2) p m p m (2) q m2
Z 4

1
1
d p
2
2
.
(B.24)
Ic m , m = i
(2)4 p2 m2 p2 m2
P 2 A
CR (m2 ) ( ex )
The counterterm (m2 ) for the extra-scalar masses, L = A gA

(ex ) , is determined to cancel the divergence,

(B.25)
as
(m2 ) =

 
1
1
2
(4) 

sin2 2 (m21 m22 ) 4m2 sin 2 cos 2 (m21 m22 ) ei

sin 2 cos 2 (m21 m22 ) ei sin2 2 (m21 m22 ) 4m2

!
. (B.26)

Substituting Eq. (B.26) into Eq. (B.20), we obtain



m2f, FI-2loop R =


h
X Y  A 
i 1
A
2
2
Y TrR YR CR cos 2 m1 m2
+3
2
4
4 f

A


X Y  A 
1
1
Yf TrR YR CRA cos 2 m2 log2 m21 + m2 log2 m22
+4
4
4
2
2
A


+ m21 + m2 log m21 m22 + m2 log m22




m2
m2


+ m21 m2 Li2 1 2 m22 m2 Li2 1 2
m1
m2

36

J. Hisano et al. / Nuclear Physics B 584 (2000) 345


(m2 log m21 m22 log m22 )2
1
+ YR CRA cos 2 sin2 2 2 1
4
m21 m22


4 m21 log m21 m22 log m22 m21 log2 m21 m22 log2 m22



+ m21 m22 2 + log m21 + log m22 log m21 log m22
.
(B.27)
This contribution vanishes when = /4 as it should be. The contribution from the
heavy sfermions is read off by taking the limit m 0 and 0 as

m2f, FI-2loop F =

h
i 1
X Y  A 
A 2
Y Tr Y C m
+3
2
4
4 f F F F F 
A



X Y  A 
1 2
A 2
2
Y Tr Y C m log mF +
,
(B.28)
+4
4
4 f F F F F
6
A

reproducing the earlier result derived in Ref. [34]. 8 In addition, there is another diagram
which contributes to the light sfermion masses at the two-loop level. The diagram is shown
in Fig. 17 and its contribution is
m2f, FI-2(1loop) = gY4 Yf J1 J2 + igY4 Yf m2 J2 + igY4 Yf J1 g 4 ,
Y

where

J1 = TrR YR

V1 V1


V2 V2

 Z 4

1
d p
,
+ TrF YF
(2)4 p2 m2

d 4p
1
4
2
(2) p m2

(B.29)

(B.30)

Fig. 17. Two-loop diagram contributing to the light sfermion masses.


8 Their result is different from ours by a finite part proportional to + log(4 ), since they subtracted
divergences not in the MS scheme.

J. Hisano et al. / Nuclear Physics B 584 (2000) 345

37

 X

Z 4
1
1
d q


J2 = TrR YR2
V 2 V2
V2 V2
V 1 V1
V1 V1
(2)4 q 2 m2 q 2 m2

 Z 4

 Z 4

1
1
d q
d q
2
+
Tr
Y
.
(B.31)
+ TrF YF2

f
f
(2)4 (q 2 m2 )2
(2)4 (q 2 m2 )2
F

The counterterms m2 and g 4 are defined by


Y
Y
X



 
2

V1 V1 V2 V2 ex YR ex + F YF F + f Yf f
L = gY m2
Y

gY4 g 4 f Yf f
Y

X

V1 V1


V2 V2





ex YR ex + F YF F . (B.32)

They are determined by the conditions,

(B.33)
and

(B.34)
respectively, which lead to
h
h
io 1 
i
1 n
2
2
2
cos
2
m

m
m
Tr
Y
Y
+
Tr
,
m2 =
R R
1
2
F
F F
Y
(4)2

h i
h i
h io 1 
1 n
2
2
2
.
2
Tr
Y
+
Tr
Y
+
Tr
g 4 =
R R
F
f Yf
F
Y

(4)2

(B.35)
(B.36)

From Eqs. (B.29), (B.35), (B.36), we obtain


m2f, FI-2(1loop) =


Y
4

Y
4

2

Yf

2

 2 n
h
h
io
i
1
TrR YR cos 2 m21 m22 + TrF YF m2F

h i
h i
h io
n
2 TrR YR2 + TrF YF2 + Trf Yf2

n
h
h
i
i
Yf TrR YR cos 2 m21 m22 + TrF YF m2F
h
h
io
i
TrR YR cos 2 m21 log m21 m22 log m22 TrF YF m2F log m2F




TrR YR2 cos2 2 log m21 + log m22

38

J. Hisano et al. / Nuclear Physics B 584 (2000) 345



m2 log m21 m22 log m22
2YR2 sin2 2 1 1
m21 m22
h
i
h
i
+ TrF YF2 log m2F + Trf Yf2 log m2f .

(B.37)

Thus, the light sfermion masses induced at two loops through the generation of U(1)Y FI
D-term are given by


(B.38)
m2f, FI-2loop = m2f, FI-2loop R + m2f, FI-2loop F + m2f, FI-2(1loop).
Altogether, the light sfermion masses at the two-loop level are given by
m2f, total = m2f, direct + m2f, FI-1loop + m2f, FI-2loop,

(B.39)

combining Eqs. (B.16), (B.19), (B.38).

B.3. The finite case


In the previous subsection, we have calculated the gaugino and the light sfermion
0
masses at the two-loop level in the DR scheme. The light sfermion masses generically
have divergent contribution, so that they receive large negative contribution from the RG
evolution. However, the divergences are canceled among loops of various heavy particles
under the conditions Eqs. (3)(5) discussed in the text. These conditions are written as
i
h
h
i
2
(B.40)
TrR TRA m2 (y1 + y2 2) + TrF TFA
mF = 0,
i
h
h
i

(B.41)
TrR YR cos 2 m21 m22 + TrF YF m2F = 0,
h
h
i
i

(B.42)
TrR YR CRA cos 2 m21 m22 + TrF YF CFA m2F = 0,
in the notation of this appendix. Using these relations in Eqs. (B.16), (B.19), (B.38), we
obtain the finite contribution to the light sfermion masses,
m2f, total = m2f, direct + m2f, FI-1loop + m2f, FI-2loop;

(B.43)

m2f, direct =


X A 2
CfA TrR TRA m2 (y1 + y2 2) log m2 + (y1 log y1 + y2 log y2 )
4
4
A





1
1
+ y2 Li2 1
2 y1 Li2 1
y1
y2





y2
y1
1
+y2 Li2 1
+ sin2 2 y1 Li2 1
2
y1
y2



X A 2
1 2
2
2
CfA TrF TFA
,
(B.44)
+4
mF log mF
4
3
A

J. Hisano et al. / Nuclear Physics B 584 (2000) 345

m2f, FI-1loop =
 
h
i
Y
Yf TrR YR cos 2 m21 log m21 m22 log m22
4
 
h
i
Y
Yf TrF YF m2F log m2F ,
+
4

39

(B.45)

m2f, FI-2loop =


X Y  A 
1
1
Yf TrR YR CRA cos 2 m2 log2 m21 + m2 log2 m22
4
4
4
2
2
A


+ m21 + m2 log m21 m22 + m2 log m22




m2
m2


+ m21 m2 Li2 1 2 m22 m2 Li2 1 2
m1
m2

2
2
2
2
2
(m log m1 m2 log m2 )
1
+ YR CRA cos 2 sin2 2 2 1
4
m21 m22


4 m21 log m21 m22 log m22 m21 log2 m21 m22 log2 m22



+ m21 m22 2 + log m21 + log m22 log m21 log m22



X Y  A 
1
Yf TrF YF CFA m2F log m2F + 2
+4
4
4
6
A
 2 n
h
i
Y
Yf TrR YR cos 2 m21 log m21 m22 log m22
+
4
h
io
+ TrF YF m2F log m2F



TrR YR2 cos2 2 log m21 + log m22


m21 log m21 m22 log m22
1
m21 m22

h
i
+ TrF YF2 log m2F .
2YR2 sin2 2

(B.46)

It should be understood that the coupling constants A and Y are the renormalized
ones at the scale , A () and Y (), and the logarithms log m2 of various masses are
2
2
normalized
 2 at the
 scale , log(m / ). Here, we have dropped the term proportional
2
to Trf Y log m which is the contribution from the RG evolution below . Note that
f

from one-loop U(1)Y FI D-term is generically nonvanishing


the contribution m2
f , FI-1loop
even if we assign U(1)X charges consistent with the SU(5)GUT, since the extra scalars
embedded in a common SU(5)GUT multiplet could have different masses due to the
difference between their supersymmetric masses caused by RG evolution below the GUT
scale.

40

J. Hisano et al. / Nuclear Physics B 584 (2000) 345

Furthermore, the above expression is considerably simplified when the following two
conditions are satisfied:
1. The U(1)X charge assignment is consistent with SU(5)GUT: TrF [YF f (m2 )] = 0.
F
2. The extra-matter multiplets have an invariance under the parity ex ex : =
4

These conditions are satisfied in the case of Model I and Model II considered in the text.
Then, Eqs. (B.44)(B.46) are reduced to
X A 2
CfA
m2f, direct = 4
4
A


TrR TRA m2 (y1 + y2 2) log m2 + (y1 log y1 + y2 log y2 )





1
1
+ y2 Li2 1
2 y1 Li2 1
y1
y2





y2
y1
1
y1 Li2 1
+ y2 Li2 1
+
2
y1
y2



X A 2
1 2
A
A 2
2
Cf TrF TF mF log mF
,
(B.47)
+4
4
3
A

m2f, FI-1loop = 0,

h
i
X Y  A 
Yf TrF YF CFA m2F log m2F .
m2f, FI-2loop = 4
4
4

(B.48)
(B.49)

Appendix C. Constraints from 1mK and K


In this appendix, we calculate the constraints on the first-two generation sfermion masses
from 1mK and K in effective SUSY, where the gluino is much lighter than the firsttwo generation sfermions. In the following calculation, we only consider the gluino box
diagram since it gives a dominant contribution. According to Refs. [5153], we not only
take into account the leading QCD corrections but also make use of B parameters instead
of the vacuum insertion approximation.
The mass matrix for the down-type squark is relevant to the gluino box diagram. In a
basis where the down-type Yukawa matrix is diagonal, the mass matrix is
!

2
2
(M
)
(M
)

ij
ij
dLj
LL
LR

,
(C.1)
Lmass = dLi dRi
2
dRj
(M2
LR )ij (MRR )ij
where the subscripts i, j are the indices of the generation. We here restrict the subscript i
to i = 1, 2, since we calculate the constraints on the first-two generation sfermion masses.
From now we take the leftright mixing mass term M2LR = 0 because of the small Yukawa
couplings. In consequence, M2LL and M2RR are diagonalized separately by the sfermion
mixing matrices (U L )XL ,i and (U R )XR ,i as follows:

J. Hisano et al. / Nuclear Physics B 584 (2000) 345


U L M2LL U L = diag m2XL ,

U R M2RR U R = diag m2XR ,

41

(C.2)

where XL = 1L , 2L and XR = 1R , 2R .
The 1S = 2 effective Lagrangian at the scale mD , where the heavy sfermions decouple,
is written as


(C.3)
Leff = 32 (mD ) C1 O1 + C 1 O 1 + C4 O4 + C5 O5 .
The operators O1,4,5 and their coefficients C1,4,5 are defined as follows:


O1 = d PL s d PL s ,


O4 = d PL s d PR s ,


O5 = d PL s d PR s ,


11
1
L
L
L
L
C1 = U1,XL UXL ,2 U1,YL UYL ,2 IXL ,YL + IXL ,YL ,
9
36


1
7
L
R
L
R
C4 = U1,XL UXL ,2 U1,YR UYR ,2 IXL ,YR IXL ,YR ,
3
3


5
1
L
R
L
R
C5 = U1,XL UXL ,2 U1,YR UYR ,2 IXL ,YR + IXL ,YR .
9
9

(C.4)
(C.5)
(C.6)
(C.7)
(C.8)
(C.9)

O 1 and C 1 are obtained with the replacement L R in O1 and C1 , respectively. The loop
integrals IX,Y and IX,Y in the above equations are
IX,Y = 0,

IX,Y =

log m2X log m2Y


m2X m2Y

(C.10)

where we have neglected the contributions dependent on the gluino mass mg , since they
are suppressed by m2g /m2X ( 1).
The above effective Lagrangian is valid for an arbitrary down-type squark mass matrix.
We here give a convenient expression by using the so-called mass insertion method [60,
61]. We introduce two parameters essential for the mass insertion method. One is the
2RR ) for the left- (right-) handed down-type squarks in the first-two
averaged mass m
2LL (m
generations, which is defined by a geometric mean as
q
q
m
2RR m21R m22R .
(C.11)
m
2LL m21L m22L ,
The other is the off-diagonal element LL (RR ) of M2LL (M2RR ), which is normalized
with the averaged mass as
LL

(M2LL)1L ,2L
m
2LL

RR

(M2RR )1R ,2R


m
2RR

(C.12)

By taking a leading order of the mass differences, (m21L m22L ) and (m21R m22R ), the
coefficients in Eqs. (C.7)(C.9) become

42

J. Hisano et al. / Nuclear Physics B 584 (2000) 345


2
11 LL
,
108 m
2LL
1 LL RR
,
C4 =
9m
LL m
RR
5 LL RR
.
C5 =
27 m
LL m
RR

C1 =

(C.13)
(C.14)
(C.15)

C 1 is obtained with the replacement L R in C1 . In Section 4, we often use the


expression in the mass insertion method in estimating the bounds on the light sfermion
masses. However, Figs. 611 are obtained by making use of the exact coefficients given in
Eqs. (C.7)(C.9).
Since the above effective Lagrangian is obtained at the heavy-sfermion mass scale mD ,
we must evolve it using RG equations to the hadronic scale had , where hadronic matrix
elements are evaluated. The coefficients at had are calculated at the one-loop level as
follows [51]:
C1 (had ) = 1 C1 (mD ),
C 1 (had ) = 1 C 1 (mD ),

(C.16)
(C.17)

1
C4 (had ) = 4 C4 (mD ) + (4 5 )C5 (mD ),
3
C5 (had ) = 5 C5 (mD ),

(C.18)
(C.19)

where

1 =

3 (mb )
3 (had )

6/25

3 (mt )
3 (mb )

6/23 

3 (mg )
3 (mt )

6/21 

3 (mD )
3 (mg )

4 = 14 ,
1/2

5 = 1 .

2/b3
,

(C.20)
(C.21)
(C.22)

Here, we have taken the hadronic scale had = 2 GeV according to Ref. [52] and assumed
that all the light sfermions and the gauginos decouple at the gluino mass scale for
simplicity. b3 is the coefficient of the one-loop beta function for the strong coupling
between the gluino mass mg and the heavy-sfermion mass scale mD .
Instead of using the vacuum insertion approximation, we represent the hadronic matrix
elements of the renormalized operators O() at the renormalization scale in terms of the
corresponding B parameters as follows:


1

K 0 O1 () K 0 = K 0 O 1 () K 0 = mK fK2 B1 (),
3

2
0 1

0
mK
mK fK2 B4 (),
K O4 () K =
4 ms () + md ()

2
0

0
1
mK

mK fK2 B5 (),
K O5 () K =
12 ms () + md ()

(C.23)
(C.24)
(C.25)

J. Hisano et al. / Nuclear Physics B 584 (2000) 345

43

where mK = 497.7 MeV, fK = 160 MeV, ms (2 GeV) = 125 MeV and md (2 GeV) =
7 MeV. In our calculation, we use the following values for the B parameters obtained by
lattice calculations at = 2 GeV [62,63]:
B1 ( = 2 GeV) = 0.60(6),

(C.26)

B4 ( = 2 GeV) = 1.03(6),

(C.27)

B5 ( = 2 GeV) = 0.73(10).

(C.28)

We have now explained all the elements necessary for calculating the SUSY contribution
to the KL KS mass difference,


(C.29)
1mK,SUSY = 2 Re K 0 Leff K 0 .
Using them, we obtain constraints from 1mK by imposing a condition that the SUSY
contribution does not saturate the experimental value,



(C.30)
2 K 0 Leff K 0 < 1mK = 3.49 1015 GeV,
in the case where the LL and RR have no CP-violating phases. The constraints can be
expressed in a simple form using the mass insertion as follows:
m
LL, RR
,
(2535) TeV
(m
LL m
RR )1/2
.
(LL RR )1/2 <
(150250) TeV
LL, RR <

(C.31)
(C.32)

If LL and/or RR have CP-violating phases, there is another constraint from K ,



0

1
Im K Leff K 0 < K = 2.3 103 .

2 1mK

(C.33)

The constraint is the severest when hK 0 |Leff |K 0 i is pure imaginary, and then the above
constraint Eq. (C.33) is rewritten as




(C.34)
2 K 0 Leff K 0 < 2 2 K 1mK = 6.5 103 1mK .
Thus, the constraints from K can be severer than those from 1mK by a factor of

(2 2 K )1/2 12.4.
References
[1]
[2]
[3]
[4]
[5]
[6]
[7]
[8]

H.P. Nilles, Phys. Rep. 110 (1984) 1.


H.E. Haber, G.L. Kane, Phys. Rep. 117 (1985) 75.
J.S. Hagelin, S. Kelley, T. Tanaka, Nucl. Phys. B 415 (1994) 293.
F. Gabbiani, E. Gabrielli, A. Masiero, L. Silvestrini, Nucl. Phys. B 477 (1996) 321.
J. Ellis, D.V. Nanopoulos, Phys. Lett. B 110 (1982) 44.
R. Barbieri, R. Gatto, Phys. Lett. B 110 (1982) 211.
L.J. Hall, L. Randall, Phys. Rev. Lett. 65 (1990) 2939.
M. Dine, R. Leigh, A. Kagan, Phys. Rev. D 48 (1993) 4269.

44

[9]
[10]
[11]
[12]
[13]
[14]
[15]
[16]
[17]
[18]
[19]
[20]
[21]
[22]
[23]
[24]
[25]
[26]
[27]
[28]
[29]
[30]
[31]
[32]
[33]
[34]
[35]
[36]
[37]
[38]
[39]
[40]
[41]
[42]
[43]
[44]
[45]
[46]
[47]
[48]
[49]
[50]
[51]
[52]
[53]
[54]
[55]
[56]
[57]
[58]
[59]

J. Hisano et al. / Nuclear Physics B 584 (2000) 345

Y. Nir, N. Seiberg, Phys. Lett. B 309 (1993) 337.


M. Dine, A. Kagan, S. Samuel, Phys. Lett. B 243 (1990) 250.
S. Dimopoulos, G.F. Giudice, Phys. Lett. B 357 (1995) 573.
A. Pomarol, D. Tommasini, Nucl. Phys. B 466 (1996) 3.
A.G. Cohen, D.B. Kaplan, A.E. Nelson, Phys. Lett. B 388 (1996) 588.
H. Nakano, hep-th/9404033.
E. Dudas, S. Pokorski, C.A. Savoy, Phys. Lett. B 369 (1996) 255.
E. Dudas, C. Grojean, S. Pokorski, C.A. Savoy, Nucl. Phys. B 481 (1996) 85.
Y. Kawamura, T. Kobayashi, Phys. Lett. B 375 (1996) 141.
P. Binetruy, E. Dudas, Phys. Lett. B 389 (1996) 503.
G. Dvali, A. Pomarol, Phys. Rev. Lett. 77 (1996) 3728.
R.N. Mohapatra, A. Riotto, Phys. Rev. D 55 (1997) 1138; Phys. Rev. D 55 (1997) 4262.
R.-J. Zhang, Phys. Lett. B 402 (1997) 101.
A.E. Nelson, D. Wright, Phys. Rev. D 56 (1997) 1598.
Q. Shafi, Z. Tavartkiladze, Phys. Lett. B 473 (2000) 272.
S. Komine, Y. Yamada, M. Yamaguchi, Phys. Lett. B 481 (2000) 67.
D.E. Kaplan, F. Lepeintre, A. Masiero, A.E. Nelson, A. Riotto, Phys. Rev. D 60 (1999) 055003.
D.E. Kaplan, G.D. Kribs, Phys. Rev. D 61 (2000) 075011.
L. Everett, P. Langacker, M. Plmacher, J. Wang, Phys. Lett. B 477 (2000) 233.
N. Arkani-Hamed, M.A. Luty, J. Terning, Phys. Rev. D 58 (1998) 015004.
M.A. Luty, J. Terning, hep-ph/9812290.
J.L. Feng, C. Kolda, N. Polonsky, Nucl. Phys. B 546 (1999) 3.
J. Bagger, J.L. Feng, N. Polonsky, Nucl. Phys. B 563 (1999 ) 3.
J.A. Bagger, J.L. Feng, N. Polonsky, R.-J. Zhang, Phys. Lett. B 473 (2000) 264.
N. Arkani-Hamed, H. Murayama, Phys. Rev. D 56 (1997) 6733.
K. Agashe, M. Graesser, Phys. Rev. D 59 (1999) 015007.
S.P. Martin, M.T. Vaughn, Phys. Rev. D 50 (1994) 2282.
I. Jack, D.R.T. Jones, S.P. Martin, M.T. Vaughn, Y. Yamada, Phys. Rev. D 50 (1994) 5481.
J. Hisano, K. Kurosawa, Y. Nomura, Phys. Lett. B 445 (1999) 316.
Y. Nomura, hep-ph/9909281.
C.D. Froggatt, H.B. Nielsen, Nucl. Phys. B 147 (1979) 277.
L. Ibanez, G.G. Ross, Phys. Lett. B 332 (1994) 100.
P. Binetruy, P. Ramond, Phys. Lett. B 350 (1995) 49.
E. Dudas, S. Pokorski, C.A. Savoy, Phys. Lett. B 356 (1995) 45.
P. Binetruy, S. Lavignac, P. Ramond, Nucl. Phys. B 477 (1996) 353.
Super-Kamiokande Collaboration, Y. Fukuda et al., Phys. Rev. Lett. 81 (1998) 1562.
L. Hall, H. Murayama, N. Weiner, Phys. Rev. Lett. 84 (2000) 2572.
T. Yanagida, J. Sato, Nucl. Phys. Proc. Suppl. 77 (1999) 293.
P. Ramond, Nucl. Phys. Proc. Suppl. 77 (1999) 3.
N. Arkani-Hamed, M. Dine, S.P. Martin, Phys. Lett. B 431 (1998) 329.
T. Barreiro, B. de Carlos, J.A. Casas, J.M. Moreno, Phys. Lett. B 445 (1998) 82.
S.P. Martin, Phys. Rev. D 55 (1997) 3177.
J.A. Bagger, K.T. Matchev, R.-J. Zhang, Phys. Lett. B 412 (1997) 77.
M. Ciuchini et al., JHEP 9810 (1998) 008.
R. Contino, I. Scimemi, Eur. Phys. J. C 10 (1999) 347.
N. Arkani-Hamed, G.F. Giudice, M.A. Luty, R. Rattazzi, Phys. Rev. D 58 (1998) 115005.
N. Arkani-Hamed, R. Rattazzi, Phys. Lett. B 454 (1999) 290.
M.B. Green, J.H. Schwarz, Phys. Lett. B 149 (1984) 117.
I. Affleck, M. Dine, N. Seiberg, Nucl. Phys. B 256 (1985) 557.
G.F. Giudice, A. Masiero, Phys. Lett. B 206 (1988) 480.
E. Poppitz, S.P. Trivedi, Phys. Lett. B 401 (1997) 38.

J. Hisano et al. / Nuclear Physics B 584 (2000) 345

[60]
[61]
[62]
[63]

M. Dugan, B. Grinstein, L. Hall, Nucl. Phys. B 255 (1985) 413.


L.J. Hall, V.A. Kostelecky, S. Raby, Nucl. Phys. B 267 (1986) 415.
S.R. Sharpe, Nucl. Phys. Proc. Suppl. 53 (1997) 181.
C.R. Allton et al., Phys. Lett. B 453 (1999) 30.

45

Nuclear Physics B 584 (2000) 4668


www.elsevier.nl/locate/npe

Large quark rotations, neutrino mixing


and proton decay
Yoav Achiman a,b, , Carsten Merten a,1
a Department of Physics, University of Wuppertal, Gaustr. 20, D-42097 Wuppertal, Germany
b School of Physics and Astronomy, Tel Aviv University, 69978 Tel Aviv, Israel

Received 7 April 2000; revised 31 May 2000; accepted 5 June 2000

Abstract
Right-handed (RH) rotations do not play a role in the Standard Model, and only the differences
of the LH mixing angles are involved in VCKM . This leads to the huge freedom in the fermionic
mass matrices. However, that is no more true in extensions of the Standard Model. For example in
GUTs large RH rotations of the quarks can be related to the observed large neutrino mixing or in
particular, all mixing angles are relevant for the proton decay. We present a simple realistic nonSUSY SO(10) GUT with large RH and LH mixing and study the corresponding nucleon decay rates.
2000 Elsevier Science B.V. All rights reserved.
Keywords: Neutrinos; Proton decay; GUTs; Mass matrices; Right-handed rotations

1. Introduction
What is the origin of the fermionic masses? This is one of the open questions in the
Standard Model (SM). The mass matrix entries are arbitrary parameters of the model and
only the neutrinos are restricted to be massless.
One can add conjectures for the structure of the fermionic mass matrices to the SM.
Many different conjectures are known to give the right masses of the charged fermions
and VCKM (within the experimental errors) and this is clearly an indication that the mass
problem is far from being solved.
The main reason for this large freedom is that right-handed (RH) rotations 2 are nonobservable in the SM. Also the observed left-handed (LH) mixing matrix VCKM involves
Corresponding author. E-mail: achiman@theorie.physik.uni-wuppertal.de
1 merten@theorie.physik.uni-wuppertal.de
2 A general non-hermitian matrix is diagonalized by a bi-unitary transformation. This means that two unitary
matrices, one from the left and one from the right, are needed. Those matrices are equal only for hermitian (or
symmetric) matrices [1].

0550-3213/00/$ see front matter 2000 Elsevier Science B.V. All rights reserved.
PII: S 0 5 5 0 - 3 2 1 3 ( 0 0 ) 0 0 3 5 3 - 9

Y. Achiman, C. Merten / Nuclear Physics B 584 (2000) 4668

47

only the differences between the mixing angles of the u-like and d-like quarks and the
individual mixing can be large. Actually, there is already a strong indication from the
neutrino sector that large rotations are required (for the atmospheric neutrino anomaly [2
5], for the solar neutrino puzzle [611]).
The considerable freedom in the mass matrices of the SM does not exist in its extensions.
We know that the SM must be extended for many reasons if not alone to explain the
fermionic masses and mixing angles. In particular, in its most popular extension, the Grand
Unified Theories (GUTs) [1215], all mixing angles are relevant, as will be explained later.
In such a framework one cannot change the weak basis without changing the physics.
The aim of this paper is to present a simple realistic SO(10) GUT [1618] model with
large LH as well as RH mixing angles and to use them to study the consequences for
nucleon decay. Mixing effects were generally neglected in the conventional proton decay
models [1923] by assuming that the mixing angles are small.
Large rotations in the quark sector could be a natural reason for the large mixing
observed in the leptonic sector in terms of neutrino oscillations. In particular, there is a
kind of duality between the RH mixing of the quarks and the leptonic LH rotations [24
31].
A predictive model for the fermionic masses must involve a family symmetry which
dictates the texture of the mass matrices and protects it from getting large radiative
corrections. We used a global U (1)F [32,33] in the framework of an SO(10) GUT that
will add relations between the matrix elements. Our aim was to look for the simplest
possible realization of a realistic model with large mixing angles. We therefore used the
famous SO(10) paper of Harvey et al. [34] and generalized it to asymmetric matrices. Such
matrices give usually large LH and/or RH mixing angles.
We did not use non-renormalizable contributions to the mass matrices la Froggatt
and Nielsen [35,36] because this method assumes ad hoc physics beyond the GUT and
many new particles. In addition, the resulting matrix elements are given there in orders
of magnitude only. This can explain the hierarchy of the masses but not the light seesaw neutrino properties. Those are obtained from a product of three matrices and hence
predicted up to a factor of [O(1)]3 which may be quite large [37].
Most recent models use SUSY GUTs [38]. 3 However, the available parameter space
of low-energy SUSY shrinked recently so much that MSSM is on the verge of loosing its
naturality [4042]. At the same time solutions without low energy SUSY have emerged
quite naturally [43] in superstring and M theory, and also the hierarchy problem can
be solved adding extra dimensions [4446]. We think therefore that it is worthwhile to
look for a non-SUSY GUT which is consistent with all observed experimental facts. The
hope is that the fine tuning (hierarchy) problem will be solved in the more fundamental
theory. Note also that conventional GUTs are relatively simple and give much more reliable
predictions for nucleon decay than SUSY theories.
Unification of the gauge coupling constants is obtained using an intermediate breaking
scale [4755] MI ' 1011 GeV. This is very useful because MI is also the right mass scale
3 We shall use it also in a forthcoming paper [39].

48

Y. Achiman, C. Merten / Nuclear Physics B 584 (2000) 4668

for the RH neutrinos needed for the see-saw mechanism as well as for leptogenesis as the
origin of the baryon asymmetry [5659] and the (invisible) axion window [6063].
The observed neutrino oscillations teach us about the neutrino masses and mixing
angles. This is the first evidence for physics beyond the SM. Our claim is that observation
of other phenomena like RH currents, leptoquarks, baryon asymmetry induced by
leptogenesis and especially nucleon decay can reveal the unknown mixing angles and
reduce considerably the freedom in the fermionic mass matrices.
The plan of the paper is as follows. In Section 2 we discuss the symmetry breaking that
is dictated by the requirement of gauge unification and the Higgs representations needed to
give the correct fermion masses. The mass matrices and our global U (1)F details are given
in Section 3. In Section 4 the numerical solutions for the mass matrices are obtained by the
use of the renormalization group equations (RGEs), and a fit to the observed properties of
the charged fermions is elaborated. Three solutions are found which are consistent with the
observed neutrino anomalies [211,64,65] (except for LSND [66,67]). All mixing angles
for those solutions are obtained explicitly and this allows for the calculation of the nucleon
decay rates in Section 5. Section 6 is devoted to the conclusions.

2. Symmetry breaking of the SO(10) GUT


We use an SO(10) GUT [1618] which is broken down to the SM via an intermediate
scale MI as follows
MU

MI

MZ

SO(10) GI GSM SU(3)C U (1)em ,

(1)

where the intermediate symmetry group is the PatiSalam one [68], GI = GPS
SU(4)C SU(2)L SU(2)R .
The breaking at MU is done using the Higgs representations 210 (or 54 in models
with D parity) while for the breaking at MI we use the SM singlet of a 126 , i.e., the GPS
representation (10, 1, 3)126 . For the masses of the Higgs scalars the extended survival
hypothesis [69] is assumed.
In view of the representation content of the mass terms
16 16 = (10 126)symm 120antisymm

(2)

we give the light fermions masses via the VEVs of (1, 2, 2)10/120 and (15, 2, 2)120/126.
The RH neutrino masses of order MR MI will be induced via the VEV of the above
mentioned (10, 1, 3)126 .
The Higgs doublet of the SM is therefore a linear combination of the SU(2)L doublets
in the representations (1, 2, 2) and (15, 2, 2). The exact number of representations needed
for the fermion mass matrices will be given later when we will discuss those matrices. This
number is needed however to fix the RGEs used to calculate MI , MU and U (MU ) as well
as the values of the mass matrices at MI . 4 As will be explained in the next chapter, we
4 Note that due to the quark-lepton symmetry of G the SO(10) mass relations are valid also at the scale M .
I
PS

Y. Achiman, C. Merten / Nuclear Physics B 584 (2000) 4668

49

used in the RGEs N1 = number of (1, 2, 2) = 4 and N15 = number of (15, 2, 2) = 2 as


well as L = number of (10, 3, 1)126 = 0 and R = number of (10, 1, 3)126 = 1 for the
GPS symmetry breaking.
For the matching conditions at MI we took
1
(MI ) = 31 (MI ) +
4C

1
,
12

(3)

1
2L
(MI ) = 21 (MI ),
5
2
1
1
,
(MI ) = 11 (MI ) 31 (MI ) +
2R
3
3
3
while at MU the conditions are

(4)
(5)

1
,
3
1
1
,
(MU ) +
= 2L
2
1
1
.
(MU ) +
= 2R
2
In general the matching conditions between gauge couplings belonging to theories
symmetry groups Gj and Gk can be written as
1
1
(MU ) = 4C
(MU ) +
U

(6)
(7)
(8)
with

1
1
S2 (Gj ) = k1 (MI /U )
S2 (Gk ),
(9)
12
12
where S2 (Gj ) is the Dynkin index of the adjoint representation of the group Gj . Details
can be found in [70], the results are given in Table 1. We checked also the RGEs for the
intermediate symmetry GI = GPS D, where D is the discrete D-parity which requires
2L = 2R between MU and MI . In this case however MU (GPS D) = 1.16 1015 GeV,
a value which would lead to a too fast proton decay.
j1 (MI /U )

Table 1
Symmetry breaking scales and gauge couplings for N1 = 4 and N15 = 2
Quantity
Value

MI
6.14 1010 GeV

1 (MI )
(46.00)1

2 (MI )
(39.86)1

Quantity
Value

3 (MI )
(31.39)1

2R (MI )
(55.85)1

2L (MI )
(39.86)1

Quantity
Value

4C (MI )
(31.42)1

MU
1.31 1016 GeV

U (MI )
(20.08)1

3. The mass matrices


The aim of this paper is to present models with large RH and LH mixing angles of
the quarks that lead to large mixing of the leptons and to study the predictions of those

50

Y. Achiman, C. Merten / Nuclear Physics B 584 (2000) 4668

models for the nucleons decay ratios. To generate the mass matrices we use the method
of Harvey et al. [34] as it was realized in [71] for symmetric mass matrices. However, in
order to have large mixing angles one should rather use asymmetric mass matrices. For
asymmetric matrices we need to make certain changes in the above scenario and add also
the antisymmetric Higgs representation 120 . Practically speaking, we shall use a global
family symmetry U (1)F or Zn that will dictate the neutrino properties in terms of the
observed masses and mixings of the charged fermions. This symmetry will be chosen in
such a way that the predictive Fritzsch texture [72] will be realized. However, as it is
well known the symmetric version of this texture cannot account for the large top quark
mass [73,74]. As we need anyhow asymmetric mass matrices we shall use the asymmetric
Fritzsch texture which is also known under the name nearest neighbour interaction model
(NNI) [75]. Namely,

0 A 0
M = B 0 C .
(10)
0 D E
In view of the fact that we are actually mainly interested in the predictions for the nucleon
decay rates which are not sensitive to the details of CP violation we will use for simplicity
real mass matrices. 5
The three fermion families and the different Higgs representations in 16i k 16j
transform under the global U (1)F as follows
16j exp(ij )16j ,

(11)

k exp(ik )k .

(12)

The invariance under U (1)F requires therefore that the k must obey i + j = k . Hence,
the fermionic part of the mass matrices has the following quantum numbers:

1 + 1 1 + 2 1 + 3
(13)
Mf 1 + 2 2 + 2 2 + 3 .
1 + 3 2 + 3 3 + 3
To realize the NNI texture (10) one sees that only Higgs representations with the charges
= 1 + 2 , 2 + 3 and 3 + 3 can couple to the fermions. Also, we still have the
possibility to couple one Higgs representation to two different combinations, i.e., 1 +2 =
23 .
Taking all this into account the Yukawa coupling matrices (at energies & MI ) can
have the structure

0 x1 0
0 y1 0
0
z1 0
(1)
(1)
(1)
Y10 = x1 0 0 , Y126 = y1 0 0 , Y120 = z1 0 0 ,
0 0 x1
0 0 y1
0
0 0

0 0 0
0 0 0
0
0
0
(2)
(2)
(2)
(14)
0
z2 .
Y10 = 0 0 x2 , Y126 = 0 0 y2 , Y120 = 0
0 x2 0
0 y2 0
0 z2 0
5 This can help to solve the strong CP problem while the observed CP violation in the K-decay can come from
a different origin than the CKM matrix.

Y. Achiman, C. Merten / Nuclear Physics B 584 (2000) 4668

51

On top of that, we need at least one 126 with a large VEV in (10, 1, 3) to break the
SO(10) gauge symmetry at MI . This VEV will generate also the RH neutrinos masses
MR MI . Actually, as is discussed in Appendix A we will use only one 126 in addition
to two 10 and two 120 to generate the asymmetrical mass matrices. The detailed fits
required four VEVs in the direction (1, 2, 2) and two in that of (15, 2, 2). Those give also
the right unification as discussed before.
In terms of the notation and discussion of Appendix A we obtained the following
expressions for the mass matrix elements in terms of the VEVs and Yukawa couplings:
(Md )12 = x1 d(1) + y1 d(1) + z1 d(1),

(15)

(Md )21 = x1 d(1)


(Md )23 = x2 d(2)
(2)
(Md )32 = x2 d

(16)

+ y1 d(1) z1 d(1),

+ z2 d(2) + d(2) ,
(2)
(2) 
z2 d + d ,


(17)
(18)

 

x1
y1
x1 d(1) +
y1 d(1) ,
x1
y1

(19)

(Me )12 = x1 d(1) 3y1d(1) + z1 d(1) = (Md )12 4y1 d(1),

(20)

(Me )21 = x1 d(1)


(Me )23 = x2 d(2)
(Me )32 = x2 d(2)

(21)

(Md )33 = x1 d(1) + y1 d(1) =

3y1d(1) z1 d(1) = (Md )21 4 y1d(1) ,



+ z2 d(2) 3 d(2) = (Md )23 4z2 d(2) ,

z2 d(2) 3 d(2) = (Md )32 + 4 z2 d(2) ,



y1
(1)
(1)
y1 d(1) ,
(Me )33 = x1 d 3 y 1d = (Md )33 4
y1

(22)
(23)
(24)

(Mu )12 = x1 u(1) + y1 u(1) + z1 u(1),

(25)

(Mu )21 = x1 u(1)


(Mu )23 = x2 u(2)
(Mu )32 = x2 u(2)

(26)

+ y1 u(1) z1 u(1),

+ z2 u(2) + u(2) ,

z2 u(2) + u(2) ,


 

x1
y1
x1 u(1) +
y1 u(1) ,
x1
y1

(Mu )33 = x1 u(1) + y1 u(1) =



= x1 u(1) 3 y1u(1) + z1 u(1) = (Mu )12 4 y1u(1) ,
M(Dir)

12

= x1 u(1) 3y1u(1) z1 u(1) = (Mu )21 4y1 u(1) ,
M(Dir)

21


= x2 u(2) + z2 u(2) 3 u(2) = (Mu )23 4z2 u(2) ,
M(Dir)

23


= x2 u(2) z2 u(2) 3 u(2) = (Mu )32 + 4z2 u(2) ,
M(Dir)

32
 

y1
(Dir)
(1)
(1)
= x1 u 3y1u = (Mu )33 4
M
y1 u(1) .
33
y1

(27)
(28)
(29)
(30)
(31)
(32)
(33)
(34)

All those elements are obtained from 14 independent Higgs parameters which are products
of VEVs and Yukawa couplings. They also fix the matrix elements of the RH neutrino
Majorana mass matrix (in terms of the VEV of (10, 1, 3) that breaks GPS GSM ), with
MR MI being a quasi-free parameter:

52

Y. Achiman, C. Merten / Nuclear Physics B 584 (2000) 4668

(Maj)
MR

= MR y1
0

y1
0
0

0
0 1

0 = y1 MR 1 0
0 0
y1

0
0 .
y1 /y1

(35)

By diagonalizing the charged fermion mass matrices


UL Mu UR = M(D)
u ,

DL Md DR = M(D)
d ,

EL Me ER = M(D)
e ,

UL DL = VCKM ,

(36)

and the light see-saw neutrino mass matrix


T
light
(Maj) 1
MR
M(Dir)
M ' M(Dir)

(37)

we obtained all masses of the charged fermions and those of the three light neutrinos as
well as the mixing matrices UL,R , DL,R , EL,R and N . 6
Neglecting phases (as explained before) we have 12 masses and 21 mixing parameters.
All those fix the proton decay rates and other GUT scale effects (like baryon asymmetry
induced by leptogenesis [5659]). However, in the framework of the SM (with massive
neutrinos) the only observable mixing matrices would be VCKM = UL DL and U = EL N .
This gives 18 mixing observables. 7

4. Numerical solutions for the mass matrices


The first step will be to calculate numerically the mass and mixing matrices in (36).
This is done using the SO(10) relations (1534) for the mass matrices at = MI =
6.14 1010 GeV. On the RHS of (36) we use the calculated masses of the charged fermions
(see Table 2) and the (real) CKM matrix at = MI . Those values are obtained using
the RGEs as is described in detail in [70]. We then run the mass matrices from MI to
MZ , diagonalize them there and compare the obtained masses and mixings with their
experimentally observed values.
Now, without using the neutrino sector, only 13 of our 14 parameters are independent
(2)
(2)
because only the combination z2 ( u + u ) appears in the charged fermion equations.
Table 2
Fermion masses at MI
Mass
Value

mu (MI )
1.16 MeV

md (MI )
2.38 MeV

ms (MI )
47.4 MeV

Mass
Value

mc (MI )
337.6 MeV

mb (MI )
1360 MeV

mt (MI )
101.2 GeV

Mass
Value

me (MI )
513 keV

m (MI )
108.14 MeV

m (MI )
1838.3 MeV

6 There is only one neutrino mixing matrix as Mlight is always symmetric.

7 Note however, given the mass matrices we have definite predictions for the proton decay, the baryon

asymmetry and RH currents which can be measured in principle by future experiments.

Y. Achiman, C. Merten / Nuclear Physics B 584 (2000) 4668

53

(2)

Fig. 1. Solution 1 (large mixing MSW) in the y1 /y1 z2 u parameter space.

We have in (36) 30 nonlinear equations. On the other hand each of the (real) 6 mixing
matrices is parametrized by 3 angles, so altogether there are 18 mixing angles. With the
13 Higgs parameters we have 31 unknowns. One of them must therefore be given to
be able to search for solutions numerically. For the given quantity we use the ratio y1 /y1
which determines the RH neutrino mass matrix (35) except for a global factor and we vary
its value between 1 6 |y1 /y1 | 6 1000. By studying the equations in detail one can see that,
once the value of y1 /y1 is given, the neutrino sector of our model is uniquely fixed up to
(2)
the two parameters MR MI and z2 u . We therefore look if there are solutions of the
model for reasonable values of y1 /y1 , MR and z2 u(2) which predict neutrino properties
lying in the range allowed by oscillation experiments [211,64,65]:
|(U)13 | 6 0.05,

(38)

0.49 6 |(U)23 | 6 0.71,

(39)

0.03 6 |(U)12 | 6 0.05 (small angle MSW),


or 0.35 6 |(U)12 | 6 0.49 (large angle MSW ),
 2

m2 m21 6 1000.
50 6 1m2atm /1m2sun m23 m22

(40)
(41)

The value of y1 MR will be fixed at the end to give the exact absolute scale of the neutrino
masses.
Given these three parameters the model predicts the three neutrino masses and three
(2)
lepton mixing angles in U. We found two regions in the y1 /y1 z2 u parameter space
which obey the atmospheric neutrino requirements together with small angle MSW [76,77]
explanation for the solar neutrino puzzle (models 2a,b) and one with the large angle MSW
(model 1) as can be seen in the Figs. 1, 2 and 3. For each region we fixed a representative
solution (see Table 3) and used it to calculate the corresponding neutrino properties. The
explicit results for the neutrinos are given in Table 4.

54

Y. Achiman, C. Merten / Nuclear Physics B 584 (2000) 4668

(2)

Fig. 2. Solution 2a (small mixing MSW) in the y1 /y1 z2 u parameter space.

(2)

Fig. 3. Solution 2b (small mixing MSW) in the y1 /y1 z2 u parameter space.


Table 3
(2)
Values of y1 /y1 and z2 u for the three representative solutions
Solution

MSW effect

Model 1
Model 2a
Model 2b

Large mixing
Small mixing
Small mixing

y1 /y1
19
24
18

(2)

z2 u (MeV)
3450
9175
9925

In this way the explicit LH and RH mixing matrices are also fixed. The corresponding
mixing angles are given in Table 5. They are used to calculate the branching ratios of the
nucleon decays.

Y. Achiman, C. Merten / Nuclear Physics B 584 (2000) 4668

55

Table 4
Masses and mixing angles of the light neutrinos at MZ
Parameter

Value in model 1

Value in model 2a

Value in model 2b

(y1 MR /MI ) m1
(y1 MR /MI ) m2
(y1 MR /MI ) m3

0.0245 eV
0.0876 eV
2.402 eV

8.73104 eV
0.355 eV
3.031 eV

1.16103 eV
0.467 eV
4.365 eV

()

0.487

0.050

0.051

()
23
()
31

0.205

0.506

0.496

0.004

0.003

0.003

12

m2 /m1
m3 /m2

 2

m3 m22
m22 m21

3.57
27.43

406.3
8.54

815.4

71.9

401.2
9.35
86.4

5. Calculation of the nucleon decay rates


The partial decay rate for a given process nucleon meson + antilepton is expressed
as follows:
!
X
X
1 2
2
2
2
2
2
2
m j |S| |A| |AL |
|Al Ml | + |AR |
|Ar Mr | ,
(42)
j =
16 nucl
r
l

where Ml and Mr are the hadronic transition matrix elements for the relevant decay
process. l and r denote the chirality of the corresponding antilepton. Al and Ar are the
relevant coefficients of the effective Lagrangian (67) given in Appendix B. A, AL and AR
are factors which result from the renormalization of the four fermion operators as follows:


1 (MZ ) 23/82
,
(43)
AL =
1 (MI )


1 (MZ ) 11/82
,
(44)
AR =
1 (MI )








4C (MI ) 5/8 2L (MI ) 27/100 2R (MI ) 3/20 2 (MZ ) 27/38
A=
4C (MU )
2L (MU )
2R (MU )
2 (MI )
2/7 
6/23
6/25 


3 (MZ )
3 (mb )
3 (mc )
3 (1 GeV) 2/9

.
(45)
3 (MI )
3 (MZ )
3 (mb )
3 (mc )
Using then [78]
3 (1 GeV) = 0.544,

3 (mc ) = 0.412,

3 (mb ) = 0.226

(46)

one obtains
|AL |2 = 1.155,

|AR |2 = 1.071,

|A|2 = 23.59.

(47)

56

Y. Achiman, C. Merten / Nuclear Physics B 584 (2000) 4668

Table 5
Mixing angles in the three different models
Parameter

Value in model 1

Value in model 2a

Value in model 2b

L12

(u)

0.627

0.540

0.275

L23

(u)

0.055

0.056

0.075

L31

(u)

0.000

0.000

0.000

(u)
R12
(u)
R23
(u)
R31
(d)
L12
(d)
L23
(d)
L31
(d)
R12
(d)
R23
(d)
R31
(e)
L12
(e)
L23
(e)
L31
(e)
R12
(e)
R23
(e)
R31
()
12
()
23
()
31

0.005

0.006

0.012

0.049

0.051

0.043

0.000

0.000

0.000

0.404

0.317

0.498

0.025

0.024

0.041

0.018

0.016

0.012

0.117

0.152

0.092

0.979

1.011

0.679

0.000

0.000

0.000

0.018

0.015

0.015

0.733

0.060

0.089

0.000

0.001

0.001

0.262

0.314

0.301

0.063

0.747

0.561

0.014

0.001

0.002

0.487

0.050

0.051

0.205

0.506

0.496

0.004

0.003

0.003

s (E
s (E
|S|2 = hNucl
r1 , rE2 , rE3 )|(Er1 rE2 )|Nucl
r1 , rE2 , rE3 )i is the probability to find two valence
quarks of the nucleon at one point in space. We used here the value |S|2 = 0.012 GeV3
given in [79]. j (1 j2 )(1 j4 ) with j = mMeson /mNucl is an SU(6) spin-flavour
symmetry breaking phase space factor.
The resulting branching ratios are given in the Tables 6 and 8. Our model also gives the
total decay rates. The predicted rates in the different models are given in the Tables 7 and 9.
Note however that these predictions have a relatively large uncertainty. This was estimated
by Langacker [80] to be
+0.5

pe+ = 100.71.03.0 yrs.

(48)

Y. Achiman, C. Merten / Nuclear Physics B 584 (2000) 4668

57

Table 6
Partial decay rates i / of the proton for the case of vanishing fermionic mixing in comparison with
the rates obtained in our solutions
Decay channel
of the proton

Rates in %
(no mixing)

Rates in %
in model 1

Rates in %
in model 2a

Rates in %
in model 2b

p e+ 0
p e+ K 0
p e+
p + 0
p + K 0
p +
p e+ 0
p e+
p e+ K 0
p + 0
p +
p eC +
p eC K +
C+
p
C K+
p
p eC +
p eC K +
C +
p
C K +
p
p C +
p C K +
p C +
p C K +

33.6

1.2

5.8

5.1
16.9

32.3

0.1
4.9

0.1

21.4
3.1
0.8
8.5
2.6
0.3
3.3
10.8
0.0
1.3
4.3
25.6
2.0
8.9
1.3
3.9
0.4
1.4
0.0
0.1
0.1
0.0
0.0

25.1
2.6
0.9
5.7
0.9
0.2
3.8
12.7
0.0
0.9
2.9
35.6
2.0
0.5
0.2
5.4
0.3
0.1
0.0
0.1
0.1
0.0
0.0

27.8
4.5
1.0
5.6
1.8
0.2
4.2
14.0
0.0
0.8
2.8
27.7
4.1
0.3
0.3
4.2
0.6
0.0
0.0
0.0
0.0
0.0
0.0

p e+ X 0
p + X0
p C X+

56.8
5.8
37.4

39.4
17.0
43.7

45.1
10.6
44.3

51.5
11.2
37.2

Table 7
Total decay rate and lifetime of the proton for the representative solutions
Quantity
p
p

Value in model 1

Value in model 2a

Value in model 2b

2.54 1035 yr1


3.94 1034 yr

2.57 1035 yr1


3.89 1034 yr

2.83 1035 yr1


3.53 1034 yr

Taking this into account our predicted representative rates have therefore a chance to be
observed by SuperKamiokande [81,82] and ICARUS [83].

58

Y. Achiman, C. Merten / Nuclear Physics B 584 (2000) 4668

Table 8
Partial decay rates i / of the bound neutron for the case of vanishing fermionic mixing in
comparison with the rates obtained in our solutions
Decay channel
of the neutron

Rates in %
(no mixing)

Rates in %
in model 1

Rates in %
in model 2a

Rates in %
in model 2b

n e+
n +
n e+
n +
n eC 0
n eC K 0
n eC
C0
n
C K0
n
C
n
n eC 0
n eC
n eC K 0
C 0
n
C
n
C K 0
n
n C 0
n C K 0
n C
n C 0
n C
n C K 0

62.9

9.7

15.1

0.6

1.7

2.3
7.7

0.0

40.1
15.8
6.2
2.4
12.0
6.8
0.4
4.2
0.1
0.2
1.8
6.1
0.2
0.6
2.1
0.1
0.0
0.7
0.0
0.0
0.0
0.0

46.2
10.4
7.1
1.6
16.4
4.9
0.6
0.2
1.0
0.0
2.5
8.4
0.2
0.0
0.1
0.0
0.0
0.4
0.0
0.0
0.0
0.0

49.9
10.0
7.7
1.5
12.5
8.2
0.5
0.1
0.9
0.0
1.9
6.4
0.4
0.0
0.1
0.0
0.0
0.0
0.0
0.0
0.0
0.0

n e+ X
n + X
n C X0

72.6

27.4

46.3
18.2
35.3

53.3
12.0
34.7

57.6
11.5
31.0

Table 9
Total decay rate and lifetime of the bound neutron for the representative solutions
Quantity
n
n

Value in model 1

Value in model 2a

Value in model 2b

2.72 1035 yr1


3.68 1034 yr

2.80 1035 yr1


3.57 1034 yr

3.14 1035 yr1


3.18 1034 yr

Yet the essential predictions of the model are actually the branching ratios. Comparing
our branching ratios with those of the conventional SO(10) (i.e., without large mixings)
one sees clearly the suppression of the p, n e+ X channels relative to the channels
p, n + X, c X. In particular p, n + , c K are prominent. Note that in SUSY

Y. Achiman, C. Merten / Nuclear Physics B 584 (2000) 4668

59

Table 10
Ratios of some partial nucleon decay rates
Ratio
(p e+ K 0 )
(p e+ 0 )
(p + 0 )
(p + K 0 )
(p C K + )
(p C + )
(p e+ 0 )
(p C + )
(n + )
(n e+ )
(n + )
(n e+ )
(n C K 0 )
(n C 0 )
(n e+ )
(n C 0 )

No mixing

Model 1

Model 2a

Model 2b

0.145

0.104

0.162

3.27

6.33

3.11

0.003

0.098

0.064

0.157

1.040

0.618

0.693

0.993

0.394

0.225

0.200

0.387

0.225

0.195

0.113

0.469

0.347

0.722

4.16

2.48

2.78

3.96

GUTs [38] the dominant decays are into final states involving K mesons [8486]. 8 The
special properties of our model are clearly reflected in the comparison of ratios as is done
in Table 10.

6. Conclusions
We presented an SO(10) GUT with a global U (1)F family symmetry that is consistent
with all experimental observations. The special thing about this model is that it involves
large mixing angles for the quarks, in contrast with the conventional expectations. The
large mixing in the lepton (neutrino) sector results therefore naturally.
Large rotations have considerable consequences for observables outside the Standard
Model. In particular calculations of the nucleon decay rates require the knowledge of all
mixing angles. We found rates which are obviously different from the conventional GUTs.
Our model is however only one simple example for the effects of large mixing angles. We
are studying now a SUSY GUT with all possible phases and its consequences also for other
GUT observables or RH currents. The hope is that by restricting the values of the mixing
angles, we will be able to reduce the large freedom in the present models for fermionic
masses.
8 This is also the case for non-SUSY GUTs with maximal RH mixings [8789].

60

Y. Achiman, C. Merten / Nuclear Physics B 584 (2000) 4668

Table 11
Higgs couplings and vacuum expectation values in SO(10) GUTs
(i)

(j )

(k)

(l)

Higgs representation

(1, 2, 2)10

(1, 2, 2)120

(15, 2, 2)120

(15, 2, 2)126

Yukawa coupling matrix

(i)
Y10
(i) (i)
u , d

(j )
Y120
(j ) (j )
u , d

(k)
Y120
(k) (k)
u , d

(l)
Y126
(l) (l)
u , d

Vacuum expectation values

Acknowledgements
We would like to thank M.K. Parida for discussions and especially for helping us with
threshold effects. One of us (Y.A.) is also grateful to the School of Physics and Astronomy
of the Tel Aviv University for the kind hospitality during his present visit.

Appendix A. Fermionic masses in SO(10) GUTs


Dirac masses for the charged fermions are of the form m(LC )T CL and therefore
transform under SO(10) as follows:
16 16 = (10 126)symm 120antisymm.

(49)

The SO(10) Higgs representations which can give masses to the fermions have the
representation content
10 (1, 2, 2) (6, 1, 1),

(50)

120 (1, 2, 2) (15, 2, 2) (6, 3, 1) (6, 1, 3) (10, 1, 1) (10, 1, 1),

(51)

126 (15, 2, 2) (10, 3, 1) (10, 1, 3) (6, 1, 1)

(52)

under GPS = SU(4)C SU(2)L SU(2)R , while for the fermions in the 16
16 (4, 2, 1) (4, 1, 2).

(53)

Hence the Dirac mass terms transform as follows


1, 2) = (15, 2, 2) (1, 2, 2)
(4, 2, 1) (4,

(54)

The different Yukawa couplings and the corresponding VEVs of 10 , 120 and 126 are
given in Table 11.
Taking now the ClebschGordan coefficients into account the mass matrices get the
following contributions [90]
Md = d Y10 + d Y126 + ( d + d )Y120 ,

(55)

Me = d Y10 3d Y126 + ( d 3 d )Y120 ,

(56)

= u Y10 + u Y126 + ( u + u )Y120,

(57)

= u Y10 3u Y126 + ( u 3 u )Y120.

(58)

Mu
(Dir)
M

The RH neutrinos can acquire also Majorana masses in term of the (10, 1, 3) component
of 126 :

Y. Achiman, C. Merten / Nuclear Physics B 584 (2000) 4668

61

1, 2) (4,
1, 2) (10, 1, 3) = (1, 1, 1) ,
(4,

(59)

(Maj)
MR

(60)

= MR Y126 MI Y126 .

The corresponding VEV MI is responsible for the symmetry breaking step GPS GSM .
The neutrinos will have therefore the 6 6 mass martix


0
M(Dir)


(61)
M=
T
(Maj) .
MR
M(Dir)

Using the fact that the non-vanishing entries of the Majorana mass matrix are much larger
than those of the Dirac matrix one can approximately block diagonalize the 6 6 matrix
and obtain the see-saw matrix [9193]
T
light
(Maj) 1
(62)
MR
M(Dir)
M M(Dir)

as well as
heavy

M
light

(Maj)

MR .

(63)

is symmetric and therefore can be diagonalized using one unitary matrix N


light

NT M

light(D)

N = M

(64)

Neutrino oscillations are induced via the leptonic analogue to the CKM matrix
U = EL N .

(65)

Appendix B. Effective SO(10) Lagrangian for nucleon decays


The baryon number violating part of the SO(10) Lagrangian (without fermionic mixing)
is known to be [1215]
g
C

+
+
+ dR eR
LB6=0 = U X u L uL + dL eL
2
g
C

+
C
u L eL
+ U Y u L dL dR eR
2

gU 0
C

C
C
u R eR
+ X dL dL u L eL
2
gU 0
C

+
C
u R eR
+ Y dL uL dL eL
2

gU

+ dR eR
+ u L eL + u R eR
+ X3 dL eL
2
+ h.c.
(66)
Taking all possible fermion mixings into account one obtains [89] the effective four
fermion Lagrangian


C
 +
C
 +
dL + A2 u L uL eR
dR
Leff = A1 u L uL eL


C

C


L uL +
+ A3 u L uL +
L dL + A4 u
R d R

62

Y. Achiman, C. Merten / Nuclear Physics B 584 (2000) 4668



C
 +
C
 +
+ A5 u L uL eL
sL + A6 u L uL eR
sR


C

C


L uL +
+ A7 u L uL +
L sL + A8 u
R sR


C
 C
C
 C
dR + A10 u L dL R
dR
+ A9 u L dL eR


C
 C
C
 C
sR + A12 u L dL R
sR
+ A11 u L dL eR


C
C
 C
 C
dR + A14 u L sL R
dR
+ A13 u L sL eR


C
C


+ A15 u L dL CR dR + A16 u L dL CR sR

C

+ A17 u L sL CR dR
+ (terms with two s quarks)
+ (terms with c, b and t quarks)
+
C
and e,,
+ (terms with L,R
L)

+ h.c.,

(67)

where the coefficients Ai are given as follows [70]:


e (UR )11 (UL )11 + (UR )21 (UL )21 + (UR )31 (UL )31
A1 = G


(ER )11 (DL )11 + (ER )21 (DL )21 + (ER )31 (DL )31

e (UR )11 (DL )11 + (UR )21 (DL )21 + (UR )31 (DL )31
+G

(ER )11 (UL )11 + (ER )21 (UL )21 + (ER )31 (UL )31 ,

e (UR )11 (UL )11 + (UR )21 (UL )21 + (UR )31 (UL )31
A2 = G

(EL )11 (DR )11 + (EL )21 (DR )21 + (EL )31 (DR )31

e0 (DR )11 (UL )11 + (DR )21 (UL )21 + (DR )31 (UL )31
+G

(EL )11 (UR )11 + (EL )21 (UR )21 + (EL )31 (UR )31 ,

e (UR )11 (UL )11 + (UR )21 (UL )21 + (UR )31 (UL )31
A3 = G

(ER )12 (DL )11 + (ER )22 (DL )21 + (ER )32 (DL )31

e (UR )11 (DL )11 + (UR )21 (DL )21 + (UR )31 (DL )31
+G

(ER )12 (UL )11 + (ER )22 (UL )21 + (ER )32 (UL )31 ,

e (UR )11 (UL )11 + (UR )21 (UL )21 + (UR )31 (UL )31
A4 = G

(EL )12 (DR )11 + (EL )22 (DR )21 + (EL )32 (DR )31

e0 (DR )11 (UL )11 + (DR )21 (UL )21 + (DR )31 (UL )31
+G

(EL )12 (UR )11 + (EL )22 (UR )21 + (EL )32 (UR )31 ,

e (UR )11 (UL )11 + (UR )21 (UL )21 + (UR )31 (UL )31
A5 = G

(ER )11 (DL )12 + (ER )21 (DL )22 + (ER )31 (DL )32

e (UR )11 (DL )12 + (UR )21 (DL )22 + (UR )31 (DL )32
+G

(ER )11 (UL )11 + (ER )21 (UL )21 + (ER )31 (UL )31 ,

e (UR )11 (UL )11 + (UR )21 (UL )21 + (UR )31 (UL )31
A6 = G

Y. Achiman, C. Merten / Nuclear Physics B 584 (2000) 4668


(EL )11 (DR )12 + (EL )21 (DR )22 + (EL )31 (DR )32

e0 (DR )12 (UL )11 + (DR )22 (UL )21 + (DR )32 (UL )31
+G

(EL )11 (UR )11 + (EL )21 (UR )21 + (EL )31 (UR )31 ,

e (UR )11 (UL )11 + (UR )21 (UL )21 + (UR )31 (UL )31
A7 = G

(ER )12 (DL )12 + (ER )22 (DL )22 + (ER )32 (DL )32

e (UR )11 (DL )12 + (UR )21 (DL )22 + (UR )31 (DL )32
+G

(ER )12 (UL )11 + (ER )22 (UL )21 + (ER )32 (UL )31 ,

e (UR )11 (UL )11 + (UR )21 (UL )21 + (UR )31 (UL )31
A8 = G

(EL )12 (DR )12 + (EL )22 (DR )22 + (EL )32 (DR )32

e0 (DR )12 (UL )11 + (DR )22 (UL )21 + (DR )32 (UL )31
+G

(EL )12 (UR )11 + (EL )22 (UR )21 + (EL )32 (UR )31 ,

e (UR )11 (DL )11 + (UR )21 (DL )21 + (UR )31 (DL )31
A 9 = G

(NL )11 (DR )11 + (NL )21 (DR )21 + (NL )31 (DR )31

e0 (DR )11 (DL )11 + (DR )21 (DL )21 + (DR )31 (DL )31
G

(NL )11 (UR )11 + (NL )21 (UR )21 + (NL )31 (UR )31 ,

e (UR )11 (DL )11 + (UR )21 (DL )21 + (UR )31 (DL )31
A10 = G

(NL )12 (DR )11 + (NL )22 (DR )21 + (NL )32 (DR )31

e0 (DR )11 (DL )11 + (DR )21 (DL )21 + (DR )31 (DL )31
G

(NL )12 (UR )11 + (NL )22 (UR )21 + (NL )32 (UR )31 ,

e (UR )11 (DL )11 + (UR )21 (DL )21 + (UR )31 (DL )31
A11 = G

(NL )11 (DR )12 + (NL )21 (DR )22 + (NL )31 (DR )32

e0 (DR )12 (DL )11 + (DR )22 (DL )21 + (DR )32 (DL )31
G

(NL )11 (UR )11 + (NL )21 (UR )21 + (NL )31 (UR )31 ,

e (UR )11 (DL )11 + (UR )21 (DL )21 + (UR )31 (DL )31
A12 = G

(NL )12 (DR )12 + (NL )22 (DR )22 + (NL )32 (DR )32

e0 (DR )12 (DL )11 + (DR )22 (DL )21 + (DR )32 (DL )31
G

(NL )12 (UR )11 + (NL )22 (UR )21 + (NL )32 (UR )31 ,

e (UR )11 (DL )12 + (UR )21 (DL )22 + (UR )31 (DL )32
A13 = G

(NL )11 (DR )11 + (NL )21 (DR )21 + (NL )31 (DR )31

e0 (DR )11 (DL )12 + (DR )21 (DL )22 + (DR )31 (DL )32
G

(NL )11 (UR )11 + (NL )21 (UR )21 + (NL )31 (UR )31 ,

e (UR )11 (DL )12 + (UR )21 (DL )22 + (UR )31 (DL )32
A14 = G

(NL )12 (DR )11 + (NL )22 (DR )21 + (NL )32 (DR )31

e0 (DR )11 (DL )12 + (DR )21 (DL )22 + (DR )31 (DL )32
G

(NL )12 (UR )11 + (NL )22 (UR )21 + (NL )32 (UR )31 ,

63

64

Y. Achiman, C. Merten / Nuclear Physics B 584 (2000) 4668


e (UR )11 (DL )11 + (UR )21 (DL )21 + (UR )31 (DL )31
A15 = G

(NL )13 (DR )11 + (NL )23 (DR )21 + (NL )33 (DR )31

e0 (DR )11 (DL )11 + (DR )21 (DL )21 + (DR )31 (DL )31
G

(NL )13 (UR )11 + (NL )23 (UR )21 + (NL )33 (UR )31 ,

e (UR )11 (DL )11 + (UR )21 (DL )21 + (UR )31 (DL )31
A16 = G

(NL )13 (DR )12 + (NL )23 (DR )22 + (NL )33 (DR )32

e0 (DR )12 (DL )11 + (DR )22 (DL )21 + (DR )32 (DL )31
G

(NL )13 (UR )11 + (NL )23 (UR )21 + (NL )33 (UR )31 ,

e (UR )11 (DL )12 + (UR )21 (DL )22 + (UR )31 (DL )32
A17 = G

(NL )13 (DR )11 + (NL )23 (DR )21 + (NL )33 (DR )31

e0 (DR )11 (DL )12 + (DR )21 (DL )22 + (DR )31 (DL )32
G

(NL )13 (UR )11 + (NL )23 (UR )21 + (NL )33 (UR )31 .
e0 = g 2 /2M 2 0 0 , where M 2 =
e = g 2 /2M 2 and G
We used here the definitions G
U
X,Y
U
X,Y
X ,Y
MX2 0 ,Y 0 MU2 is assumed.
The coefficients Ai are connected to the hadronic transition amplitudes of the elementary
processes responsible for the nucleon decays. The independent amplitudes are given in
Tables 12 and 13.
Table 12
Decay amplitudes for the elementary processes of proton decays
Decay process
+ C
p eR
u u

Lagrangian term


u C
L uL



+
eR
dR

Amplitude ( 30 )

Coefficient

A2

+ C
p eR
u u
+ C
p eR
d d
+ C
p eR d d

+
u C uL eR
dR


 L
+
u
e

u C

L
R
 R

 L
+
u
e

u C

L
R
L
R

A2

A2

+2

A2

+ C
p eR
u u

+
u C uL eR
dR


 L
+
C

u L uL eR dR



+

u C
L uL eR sR



+
u
e

u C

L
R
L
R



+
C

u L uL eR sR



C

u C
L dL eR dR



C d
d

u C

L
R
L
eR

10

A2

A2

A6

+2

A6

A6

+4

A9

10

A9

+ C
p eR
d d
+ C
p eR
s d
+ C
p eR
s d
+ C
p eR
s d
C d C u
p eR
C d C u
p eR





Y. Achiman, C. Merten / Nuclear Physics B 584 (2000) 4668

65

Table 12 continued
Decay process
C d C u
p eR
C s C u
p eR
C s C u
p eR
C s C u
p eR
C s C u
p eR
C s C u
p eR
C s C u
p eR

Lagrangian term

Amplitude ( 30 )

Coefficient

A9

+4

A11

+2

A11

A13

12

A13

+4

A11

12

A13

Amplitude ( 30 )

Coefficient

A2

+10

A2

+8

A2

+8

A9

A9

+4

A9

+8

A9

+2

A9

+10

A9

+4

A11

A11

A13

+12

A13

A11

+12

A13



C

u C
L dL eR dR



C

u C
L dL eR sR



C s
d

u C

L
R
 eR

 L
C d
s

u C

L
L
eR R



C

u C
L sL eR dR



C

u C
L dL eR sR



C

u C
L sL eR dR

Table 13
Decay amplitudes for the elementary processes of neutron decays
Decay process
+ C
n eR
u d
+ C
n eR
u d
+ C
n eR
u d
C uC u
n eR
C uC u
n eR
C d C d
n eR
C d C d
n eR
C uC u
n eR
C d C d
n eR
C sC d
n eR
C sC d
n eR
C sC d
n eR
C sC d
n eR
C sC d
n eR
C sC d
n eR

Lagrangian term




u C
L uL

u C
L uL





+
eR
dR
+
eR
dR





u C
L uL eR dR



C

u C
L dL eR dR



C d
d

u C

L
R
 eR

 L
C d
d

u C

R
L
L
eR



C

u C
L dL eR dR



C d
d

u C

L
R
 eR

 L
C d
d

u C

L
R
L
eR



C

u C
L dL eR sR



C s
d

u C

L
R
 eR

 L
C d
s

u C

L
L
eR R



C d
s

u C

L
R
L
eR



C
C

u L dL eR sR



C

u C
L sL eR dR

66

Y. Achiman, C. Merten / Nuclear Physics B 584 (2000) 4668

References
[1] Y. Achiman, PRHEP-Corfu 98/026, RH Mixing Observable?, Talk at the Corfu Summ. Inst. on
Elem. Part. Phys., 1998, hep-ph/981238.
[2] M. Nahakata, SuperKamiokande Collaboration, Talk presented at TAUP 99, Paris 69
September 1999.
[3] Y. Fukuda et al., SuperKamiokande Collaboration, Phys. Lett. B 436 (1998) 33; Phys. Rev.
Lett. 81 (1998) 1562.
[4] M. Ambrosio et al., MACRO Collaboration, Phys. Lett. B 434 (1998) 451.
[5] M. Spurio, MACRO Collaboration, hep-ex/9908066.
[6] B.T. Cleveland et al., Homestake Collaboration, Astrophys. J. 496 (1998) 505.
[7] K.S. Hirata et al., Kamiokande Collaboration, Phys. Rev. Lett. 77 (1996) 1683.
[8] W. Hampel et al., GALLEX Collaboration, Phys. Lett. B 447 (1999) 127.
[9] J.N. Abdurashitov et al., SAGE Collaboration, Phys. Rev. Lett. 77 (1996) 4708; Phys. Rev. C 60
(1999) 055801.
[10] Y. Fukuda et al., SuperKamiokande Collaboration, Phys. Rev. Lett. 81 (1998) 1158; Phys. Rev.
Lett. 82 (1999) 2430.
[11] M.B. Smy, SuperKamiokande Collaboration, hep-ex/9903034.
[12] P. Langacker, Phys. Rep. C 72 (1981) 185.
[13] A. Masiero, in: Grand Unification with and without Supersymmetry and Cosmological
Implications, World Scientific, Singapur, 1984.
[14] G.G. Ross, Grand Unified Theories, Benjamin/Cummings, Menlo Park, 1985.
[15] R.N. Mohapatra, Unification and Supersymmetry, 2nd edn., Springer-Verlag, New York, 1992.
[16] H. Georgi, in: C.E. Carlson (Ed.), Particles and Fields, AIP Press, New York, 1975.
[17] H. Fritzsch, P. Minkowski, Ann. Phys. 93 (1975) 193.
[18] M.S. Chanowitz, J. Ellis, M.K. Gaillard, Nucl. Phys. B 128 (1977) 506.
[19] P. Langacker, Phys. Rep. C 72 (1981) 187; in: E. Kolb et al. (Eds.), Inner Space and Outer
Space, University of Chicago Press, Chicago, 1986.
[20] D.G. Lee, R.N. Mohapatra, M.K. Parida, M. Rani, Phys. Rev. D 51 (1995) 229.
[21] M.S. Chanowitz, J. Ellis, M.K. Gaillard, Nucl. Phys. B 128 (1977) 506.
[22] G. Kane, G. Karl, Phys. Rev. D 22 (1980) 2808.
[23] M.B. Gavela, A. Le Yaouanc, L. Oliver, O. Pne, J.C. Raynal, Phys. Rev. D 23 (1981) 1580.
[24] G. Altarelli, F. Feruglio, Phys. Lett. B 451 (1998) 388.
[25] Z. Berezhiani, A. Rossi, hep-ph/9811447.
[26] S. Lola, G.G. Ross, hep-ph/9902283.
[27] C.H. Albright, K.S. Babu, S.M. Barr, Phys. Rev. Lett. 81 (1998) 1167.
[28] C.H. Albright, S.M. Barr, Phys. Rev. D 58 (1998) 013002.
[29] Y. Nomura, T. Yanagida, hep-ph/9807325.
[30] M. Bando, T. Kugo, hep-ph/9902204.
[31] Y. Achiman, C. Merten, Talk presented at TAUP 99, Paris 69 September 1999 (to be published
in Nucl. Phys. Proc. Suppl.), hep-ph/9911314.
[32] A. Davidson, V.P. Nair, K.C. Wali, Phys. Rev. D 29 (1984) 1504; Phys. Rev. D 29 (1984) 1513.
[33] Y. Achiman, T. Greiner, Nucl. Phys. B 443 (1995) 3.
[34] J.A. Harvey, P. Ramond, D.B. Reiss, Nucl. Phys. B 199 (1982) 223.
[35] C.D. Froggatt, H.B. Nielsen, Nucl. Phys. B 147 (1979) 277.
[36] P. Ramond, hep-ph/0001006.
[37] Y. Achiman, Nucl. Phys. Proc. Suppl. 66 (1998) 400.
[38] R.N. Mohapatra, hep-ph/9911272.
[39] Y. Achiman, M. Richter, work in progress.
[40] G.F. Giudice, hep-ph/9912279.
[41] L. Giusti, A. Romanino, A. Strumia, Nucl. Phys. B 550 (1999) 3.

Y. Achiman, C. Merten / Nuclear Physics B 584 (2000) 4668

[42]
[43]
[44]
[45]
[46]
[47]
[48]
[49]
[50]
[51]
[52]
[53]
[54]
[55]
[56]
[57]
[58]
[59]
[60]
[61]
[62]
[63]
[64]
[65]
[66]
[67]
[68]
[69]
[70]
[71]
[72]
[73]
[74]
[75]
[76]
[77]
[78]
[79]
[80]
[81]
[82]
[83]
[84]
[85]

67

A. Romanino, A. Strumia, hep-ph/9912301.


L.E. Ibaez, FTUAM-99-37, hep-ph/9911499.
M. Chaichian, A.B. Kobakidze, hep-th/9912193.
Z. Checko, A.E. Nelson, hep-ph79912186.
I. Antoniadis, hep-ph/9909212.
S. Rajpoot, Phys. Rev. D 22 (1980) 2244.
R.W. Robinett, J.L. Rosner, Phys. Rev. D 26 (1982) 2396.
Y. Tosa, G.C. Branco, R.E. Marshak, Phys. Rev. D 28 (1983) 1731.
D. Chang, R.N. Mohapatra, M.K. Parida, Phys. Rev. D 30 (1984) 1052.
J.M. Gipson, R.E. Marshak, Phys. Rev. D 31 (1985) 1705.
D. Chang, R.N. Mohapatra, J.M. Gipson, R.E. Marshak, M.K. Parida, Phys. Rev. D 31 (1985)
1718.
N.G. Deshpande, E. Keith, P.B. Pal, Phys. Rev. D 46 (1992) 2261; Phys. Rev. D 47 (1993) 2892.
V.A. Kuzmin, M.E. Shaposhnikov, Phys. Lett. B 92 (1980) 115.
F. Acampora, G. Amelino-Camelia, F. Buccella, O. Pisanti, L. Rosa, T. Tuzi, Nuovo Cimento
A 108 (1995) 375.
M. Fukugita, T. Yanagida, Phys. Lett. B 174 (1986) 45.
W. Buchmller, M. Plmacher, Phys. Lett. B 389 (1996) 73.
M. Flanz, E.A. Paschos, U. Sarkar, J. Weiss, Phys. Lett. B 389 (1996) 693.
A. Pilaftsis, hep-ph/9812256.
R.D. Peccei, H.R. Quinn, Phys. Rev. D 16 (1977) 1791.
J.E. Kim, Phys. Rev. Lett. 43 (1979) 103.
M.A. Shifman, V.I. Vainstein, V.I. Zakharov, Nucl. Phys. B 166 (1980) 4933.
M. Dine, W. Fisher, M. Strednicki, Phys. Lett. B 104 (1981) 199.
S.M. Bilenky, C. Giunti, W. Grimus, Prog. Part. Nucl. Phys. 43 (1999) 1.
C. Bemporad et al., CHOOZ Collaboration, Nucl. Phys. Proc. Suppl. 77 (1999) 159.
C. Athanassopoulos et al., LSND Collaboration, Phys. Rev. Lett. 77 (1996) 3082; Phys. Rev.
Lett. 81 (1998) 1774.
K. Eitel et al., KARMEN Collaboration, Nucl. Phys. Proc. Suppl. 77 (1999) 212.
J. Pati, A. Salam, Phys. Rev. D 10 (1974) 275.
F. del Aguila, L.E. Ibaez, Nucl. Phys. B 177 (1981) 60.
C. Merten, PhD Thesis WUB-DIS 99-14, University of Wuppertal, 1999, (in german), hepph/0002301.
Y. Achiman, T. Greiner, Nucl. Phys. B 443 (1995) 3.
H. Fritzsch, Phys. Lett. B 70 (1977) 436; Phys. Lett. B 73 (1978) 317; Nucl. Phys. B 155 (1979)
189.
K.S. Babu, Q. Shafi, Phys. Lett. B 357 (1995) 365.
Y. Achiman, T. Greiner, Phys. Lett. B 324 (1994) 33.
G.C. Branco, L. Lavoura, F. Mota, Phys. Rev. D 39 (1989) 3443.
L. Wolfenstein, Phys. Rev. D 17 (1978) 2369; Phys. Rev. D 20 (1979) 2634.
S.P. Mikheyev, A.Yu. Smirnov, Sov. J. Nucl. Phys. 42 (1986) 913.
A.J. Buras, Les Houches Lectures, in: F. David, R. Gupta (Eds.), Probing the Standard Model
of Particle Interactions, Elsevier Science B.V., 1998, hep-ph/9806471.
J. Bolz, P. Kroll, Z. Phys. A 356 (1996) 327.
P. Langacker, in: E.W. Kolb (Ed.), Inner Space, Outer Space, Univ. Chicago Press, Chicago,
1986.
M. Shiozawa et al., SuperKamiokande Collaboration, Phys. Rev. Lett. 81 (1998) 3319.
Y. Hayato et al., SuperKamiokande Collaboration, Phys. Rev. Lett. 83 (1999).
F. Arneodo et al., ICARUS Collaboration, Nucl. Phys. Proc. Suppl. 70 (1999) 453.
V. Lucas, S. Raby, Phys. Rev. D 55 (1997) 6986.
P. Nath, R. Arnowitt, hep-ph/9708469.

68

Y. Achiman, C. Merten / Nuclear Physics B 584 (2000) 4668

[86] J. Pati, K.S. Babu, F. Wilczek, hep-ph/9912301.


[87] Y. Achiman, J. Keymer, University of Wuppertal Preprint WUB 83-16, contr. paper #275 to Int.
Symp. on Lepton and Photon Interactions, Cornell 1983.
[88] J. Keymer, Diploma thesis WUD 84-23, University of Wuppertal, 1983, (in German).
[89] Y. Achiman, S. Bielefeld, Phys. Lett. B 412 (1997) 320.
[90] H. Georgi, D.V. Nanopoulos, Nucl. Phys. B 159 (1979) 16.
[91] M. Gell-Mann, P. Ramond, R. Slansky, in: P. van Nieuwenhuizen, D.Z. Freedman (Eds.),
Supergravity, North-Holland, Amsterdam, 1979.
[92] T. Yanagida, in: O. Sawada, A. Sugamoto (Eds.), Proceedings of the Workshop on the Unified
Theory and the Baryon Number of the Universe, KEK report 79-18, Tsukuba, 1979.
[93] R.N. Mohapatra, G. Senjanowic, PRL4480912.

Nuclear Physics B 584 (2000) 69108


www.elsevier.nl/locate/npe

CFTs from CalabiYau four-folds


Sergei Gukov a, , Cumrun Vafa b , Edward Witten c
a Department of Physics, Princeton University, Princeton, NJ 08544, USA
b Jefferson Laboratory of Physics, Harvard University, Cambridge, MA 02138, USA
c School of Natural Sciences, Institute for Advanced Study, Olden Lane, Princeton, NJ 08540, USA

Received 25 May 2000; accepted 13 June 2000

Abstract
We consider F/M/Type IIA theory compactified to four, three, or two dimensions on a Calabi
Yau four-fold, and study the behavior near an isolated singularity in the presence of appropriate
fluxes and branes. We analyze the vacuum and soliton structure of these models, and show that near
an isolated singularity, one often generates massless chiral superfields and a superpotential, and in
many instances in two or three dimensions one obtains nontrivial superconformal field theories. In
the case of two dimensions, we identify some of these theories with certain KazamaSuzuki coset
models, such as the N = 2 minimal models. 2000 Elsevier Science B.V. All rights reserved.

1. Introduction
We have learned in recent years that it is fruitful to study singular limits of string
compactifications. In this paper, we consider theories with four supercharges in four,
three, and two dimensions, constructed by considering F-theory, M-theory, and type IIA
string theory on a CalabiYau four-fold with an isolated complex singularity. We can
connect these theories to each other by circle compactifications from four to three to
two dimensions. In addition to the choice of singularity, the description of these theories
depends on certain additional data involving the four-form flux and membrane charge in
M-theory (and related objects in the other theories).
We will analyze the vacuum structure of these theories and the domain walls connecting
the possible vacua. We argue that in many cases, the nonperturbative physics near a
singularity generates massless chiral superfields with a superpotential, leading in many
instances, especially in two dimensions or in three dimensions with large membrane
charge, to an infrared flow to a nontrivial conformal field theory. In some cases, we can
identify the theories in question with known superconformal models; for example, type IIA
Corresponding author. E-mail: gukov@feynman.princeton.edu

0550-3213/00/$ see front matter 2000 Elsevier Science B.V. All rights reserved.
PII: S 0 5 5 0 - 3 2 1 3 ( 0 0 ) 0 0 3 7 3 - 4

70

S. Gukov et al. / Nuclear Physics B 584 (2000) 69108

at a four-fold An singularity gives an N = 2 KazamaSuzuki model in two dimensions,


as we argue using the vacuum and soliton structure. More generally, from the ADE fourfold hypersurface singularities with appropriate fluxes, we obtain all the N = 2 Kazama
Suzuki models [1] at level one. This is a large list of exactly solvable conformal theories,
which includes the N = 2 unitary minimal models. It is quite satisfying that strings in the
presence of singularities captures such a large class of known conformal theories in two
dimensions, and suggests that maybe even in higher dimensions, strings propagating in
singular geometries yield an equally large subspace of conformal theories. Moreover, since
the KazamaSuzuki models are exactly solvable conformal theories in two dimensions, it
would be interesting to see to what extent its known spectrum and correlation functions can
be extracted from string theory. This would be a natural testing grounds in view of potential
application to higher dimensions where the conformal theories are less well understood.
Our result gives a relation between singularity theory, as in the Landau-Ginzburg approach to conformal theories in two dimensions [24], and the singularity of internal compactification geometry. This relation may well extend to non-supersymmetric examples; it
would certainly be interesting to explore this.
In Section 2, we analyze the fluxes, branes, and vacuum states near a four-fold
singularity. In Section 3, we show how to compute the spectrum of domain walls (which
can also be viewed as strings and kinks in the three and two-dimensional cases) for a special
class of singularities. In Section 4, we identify the models derived from ADE singularities
with KazamaSuzuki models at level 1. In Section 5, we discuss some additional classes
of singularities on four-folds, and in Section 6, we discuss the reinterpretation of some of
our results in terms of branes.

2. Classification of vacua
2.1. The G-field and domain walls
For our starting point, we take M-theory on R3 Y , where Y is a compact eightmanifold. Soon, we will specialize to the case that Y is a CalabiYau four-fold, so as to
achieve supersymmetry. We also will note in Section 2.5 the generalization of our remarks
to type IIA or F-theory compactification on Y .
To fully specify a vacuum on Y , one must specify not just Y but also the topological
class of the three-form potential C of M-theory, whose field strength is G = dC. Roughly
speaking, C-fields are classified topologically by a characteristic class H 4 (Y ; Z). At
the level of de Rham cohomology, is measured by the differential form G/2 , and we
sometimes write it informally as = [G/2]. 1
1 The assertion that takes values in H 4 (Y ; Z) is a bit imprecise, since in general [5] the G-field is shifted
from standard Dirac quantization and is not an element of H 4 (Y ; Z). But the difference between two C-fields is
always measured by a difference 0 H 4 (Y ; Z). itself takes values in a principal homogeneous space
for the group H 4 (Y ; Z); the relation between H 4 (Y ; Z) and is just analogous to the relation between H 2 (Y ; Z)
and the set of Spinc structures on Y . The shift in the quantization law of G arises precisely when the intersection

S. Gukov et al. / Nuclear Physics B 584 (2000) 69108

71

Without breaking the three-dimensional Poincar symmetries, this model can be


generalized by picking N points Pi Y and including N membranes with worldvolumes
of the form R3 Pi . More generally, we include both membranes and antimembranes and
let N be the difference between the number of membranes and antimembranes; thus it can
be a positive or negative integer. With Y being compact, the net source of the C-field must
vanish; this gives a relation [5,6]
Z
1 GG

.
(2.1)
N=
24 2
(2)2
Y

If G obeys the shifted quantization condition mentioned in the last footnote, then the right
hand side of (2.1) is always integral [5].
In this construction, models defined using the same Y but different are actually
different states of the same model. 2 To show this, it suffices to describe a domain wall
interpolating between models with the same Y and with C-fields of arbitrary characteristic
classes 1 and 2 . By Poincar duality, H4 (Y ; Z) = H 4 (Y ; Z). Hence, there is a four-cycle
S Y , representing an element of H4 (Y ; Z), such that if [S] H 4 (Y, Z) is the class that
is Poincar dual to S, then 2 1 = [S]. Now, consider a fivebrane in R3 Y whose
worldvolume is W = R2 S, with R2 a linear subspace of R3 . Being of codimension one
in spacetime, this fivebrane looks macroscopically like a domain wall. Moreover, because
the fivebrane is a magnetic source of G, the characteristic class = [G/2] jumps by [S]
in crossing this domain wall. Hence if it equals 1 on one side, then it equals 2 = 1 + [S]
on the other side.
Eq. (2.1) implies that if jumps in crossing a domain wall, then N must also jump.
Let us see how this comes about. The key is that there is a self-dual three-form T on the
fivebrane with a relation
dT = G|W 2(M),

(2.2)

where G|W is the restriction of the G-field to the worldvolume W , M is the union of all
boundaries of membrane worldvolumes that terminate on W , and (M) is a four-form
with delta function support on M. Because the G-field actually jumps in crossing the
fivebrane, it is not completely obvious how to interpret the term G|W . We will assume
that this should be understood as the average of the G-field on the two sides: G|W =
(G1 + G2 )/2. Since the left hand side of (2.2) is zero in cohomology, we get a relation in
cohomology
[M] =

G1 + G2
,
2(2)

(2.3)

where [M] is the cohomology class dual to M. We are interested in the case that the
membrane worldvolumes are of the form R3 Pi , so that their boundaries on W are of the
form on H 4 (Y ; Z) is not even. In our examples, this will occur only in Section 5, and we will ignore this issue
until that point.
2 This may also be true of models with different Y as suggested by results of [7] in the threefold case but
this issue is much harder to explore.

72

S. Gukov et al. / Nuclear Physics B 584 (2000) 69108

form R2 Pi . In evaluating (2.3), we can suppress the common R2 factor, and integrate
over S to get a cohomology relation. The integration converts [M] into N1 N2 , the
change in N in crossing the domain wall or in other words the net number of membranes
whose boundary is on the fivebrane. So
Z
Z
1 G21 G22
1 G1 + G2
=
.
(2.4)
N1 N2 =
2
2
2
(2)2
S

Here we have used the fact that [S] = [(G2 G1 )/2] to convert an integral over S to one
over Y . Clearly, (2.4) is compatible with the requirement that (2.1) should hold on both
sides of the domain wall.
Incorporation of supersymmetry
Now we wish to specialize to the case that Y is a CalabiYau four-fold and to look for
vacua with unbroken supersymmetry.
For this, several restrictions must be imposed. The requirements for G, assuming that
one wants unbroken supersymmetry with zero cosmological constant, have been obtained
in an elegant computation [8]. The result is that G must be of type (2, 2) and must be
primitive, that is, it must obey
K G = 0,

(2.5)

where K is the Khler form. We will analyze this condition in Appendix A, but for now
we note that it implies that G is self-dual and hence in particular that
Z
GG
> 0,
(2.6)
(2)2
Y

with equality only if G = 0.


The second basic consequence of supersymmetry is that N must be positive. Only
membranes and not anti-membranes on R3 Pi preserve the same supersymmetry that
is preserved by the compactification on the complex four-fold Y .
Given that N must be positive, the relation (2.1) implies that
Z

GG
(2.7)
6 .
2
12
(2)
Y

This inequality together with self-duality implies that, for compact Y , there are only finitely
many choices of G that are compatible with unbroken supersymmetry. For negative,
there are none at all.
Energetic considerations
R
There is another way to understand the result that = N + 81 2 G G should not
change in crossing the domain wall. This is based on a finite energy condition. The
condition that the domain wall be flat and have finite tension requires that the energy
density on the two sides of the domain wall should be equal far away from the domain
wall.
R
The energy density in the bulk gets contribution from the G flux given by 81 2 G G

S. Gukov et al. / Nuclear Physics B 584 (2000) 69108

73

and from the membranes by N . For the supersymmetric situation we are considering, G
is self-dual (as
R explained in Appendix A), i.e., G = G so the energy density is given
by N + 81 2 G G, and so its constancy across a domain wall is a consequence of the
finite energy of the domain wall. This is also important in our applications later, as we
will use a BPS formula for the mass of domain walls. Due to boundary terms at infinity,
such formulas are generally not valid for objects of very low codimension in space (the
codimension is one in our case). The cancellation of the boundary terms in question in our
case is again precisely the condition of constancy of across the domain wall.
2.2. Behavior near a singularity
So far we have considered the case of a compact smooth manifold Y . Our main interest
in the present paper, however, is to study the behavior as Y develops a singularity. For
practical purposes, it is convenient then to omit the part of Y that is far from the singularity
and to consider a complete but not compact CalabiYau four-fold that is developing a
singularity. In fact, some of the singularities we will study like the An singularities
of a complex surface for very large n probably cannot be embedded in a compact
CalabiYau manifold. Our discussion will apply directly to an isolated singularity of a
non-compact variety.
Hypersurface singularities are an important example and will be our focus in this
paper except in Section 5. For example, one of our important applications will be to
quasihomogeneous hypersurface singularities. In this case, we begin with five complex
variables za , a = 1, . . . , 5 of degree ra > 0 and a polynomial F (z1 , . . . , z5 ) that is
homogeneous of degree 1. We assume that F is such that the hypersurface F = 0 is
smooth except for an isolated singularity at z1 = = z5 = 0. Then we let X be a smooth
deformation of this singular hypersurface such as
F (z1 , . . . , z5 ) = 

(2.8)

with  a constant, or more generally


X
ti Ai (z1 , . . . , z5 ),
F (z1 , . . . , z5 ) =

(2.9)

with complex parameters ti and polynomials Ai that describe relevant perturbations of the
singularity F = 0. The singular hypersurface F = 0 admits the U (1) symmetry group
za eira za .
Under this transformation, the holomorphic four-form
dz2 dz3 dz4 dz5
=
F /z1
has charge
X
ra 1.
r =

(2.10)

(2.11)

(2.12)

The U (1) symmetry in (2.10) is an R-symmetry group if r 6= 0. If the model is to


flow in the infrared to a superconformal field theory, an R-symmetry must appear in the

74

S. Gukov et al. / Nuclear Physics B 584 (2000) 69108

superconformal algebra; we propose that it is the symmetry just identified. If the Ai have
degrees ri , then the dimensions of the corresponding operators are proportional to ri /r
(in other words, the R-charges normalized so that has R-charge 1). Since the ri are
positive, requiring that the dimensions be positive gives a condition r > 0:
X
ra > 1.
(2.13)
a

We will see the importance of this condition from several points of view.
In what sense is such an X a CalabiYau manifold? The holomorphic four-form
defined in (2.11) has no zeroes or poles on X, though it has in a certain sense (using the
compactification described in the next paragraph) a pole at infinity. A theorem of Tian and
Yau [9] asserts, assuming (2.13), that there is a CalabiYau metric on X with volume form

| |,

(2.14)

and moreover (see the precise statement in Eq. (2.3) of [9]) this metric is asymptotically
conical, that is it looks near infinity like
2
.
ds 2 = dr 2 + r 2 ds

(2.15)

Here ds is an angular metric, and r is a radial coordinate near infinity which scales
under za ra za as r (a ra 1)/4 r. This exponent ensures that the volume form
derived from (2.15) scales like ||2 .
To apply the TianYau theorem to the hypersurface X and deduce the statements
in the last paragraph, one writes ra = ba /d with ba relatively prime integers and d a
positive integer. Then one introduces another complex variable w of degree 1/d, and one
compactifies X to the compact variety Y 0 defined by the equation F (zi ) wd = 0 in a
weighted projective space. The discussion of [9] applies to this situation, with D the divisor
w = 0, and identifies r with a fractional power of |w|.
It is very plausible that if a compact CalabiYau manifold Y develops an isolated
hypersurface singularity that is at finite distance on the moduli space, then the Calabi
Yau metric on Y looks locally like the conelike metric that we have just described on the
hypersurface X. 3 For our purposes, we do not strictly need to know that this is true, but
the physical applications are certainly rather natural if it is.
Flux at infinity
Noncompactness of X leads to several important novelties in the specification of the
model. First of all, flux can escape to infinity, and hence (2.1) no longer holds. Rather, an
extra term appears in (2.1), namely the flux measured at infinity. This flux is a constant of
the motion, invariant under the dynamics which occurs in the interior of X. If we absorb
the constant /24 in the definition of , 4 then we can write the conserved quantity as
3 There is no claim here that X can be globally embedded in Y , only that the behavior of Y near its singularity
can be modeled by X. Note that the variety Y 0 used in the last paragraph in relation to the TianYau theorem is
not a CalabiYau manifold.
4 When X is not compact, must in any event be defined by a curvature integral and need not coincide with
the topological Euler characteristic.

S. Gukov et al. / Nuclear Physics B 584 (2000) 69108

=N +

1
2

GG
.
(2)2

75

(2.16)

We can think of as a constant that must be specified (in addition to giving X) in order
to determine the model. A model with given has various vacuum states, determined by
the values of N and G. For unbroken supersymmetry, both terms on the right-hand side of
(2.16) are positive (for the same reasons as in the case of compact X), so there are only
finitely many possible choices of N and G for fixed .
In addition to , there is another quantity that characterizes the definition of the model
and commutes with the dynamics. For finiteness of the energy, it is reasonable to require
that the flux G vanishes if restricted to X, the region near infinity in X. This does not
imply that the cohomology class vanishes if restricted to X, but only that its restriction
is a torsion class. Thus, the C-field at infinity is flat, but perhaps topologically nontrivial.
Local dynamics cannot change the behavior at infinity, so the restriction of to X is
another invariant of the local dynamics, which must be specified in defining a model.
There is another way to see more explicitly how this invariant comes about. For this,
we have to look at precisely what Poincar duality says in the case of a noncompact
manifold X. Domain walls of the type introduced in Section 2.1 are classified by H4 (X; Z),
which classifies the cycles S on which a fivebrane can wrap to make a domain wall.
4 (X; Z) (where H 4 denotes cohomology
Poincar duality says that this is the same as Hcpct
cpct
5
with compact support), the rough idea being that if S is a four-cycle determining an
element of H4 (X; Z) (so in particular S is by definition compact), then the Poincar dual
class [S] is represented by a delta function (S) that has compact support. C-fields on
4 (X; Z)
X are classified topologically by H 4 (X; Z). The groups H 4 (X; Z) and Hcpct
that classify, respectively, topological classes of C-fields and of changes in C-fields in
crossing domain walls are in general different for non-compact X. However, there is always
a natural map
4
(X; Z) H 4 (X; Z)
i : Hcpct

(2.17)

(by forgetting that a class has compact support). Moreover, for hypersurface singulari4 (X; Z) and H 4 (X; Z) are lattices, which we will call and , respectively.
ties, Hcpct
Poincar duality in the noncompact case says that and are dual lattices. When the
4 (X; Z) has no null vectors, the map i is an embedding, and
intersection form on Hcpct
can be regarded as a sublattice of its dual lattice . This makes things very simple.
The lattice can actually be described rather simply. In fact, topologically, the
hypersurface X is homotopic to a bouquet of four-spheres. 6 H4 (X; Z) is a lattice
with one generator for every four-sphere in the bouquet.
5 The cohomology of X with compact support is generated by closed forms on X with compact support,
subject to the equivalence relation that
= + d if  has compact support.
6 Such a bouquet is, by definition, associated with a tree diagram in which the vertices represent four-spheres
and two vertices are connected by a line if and only if the corresponding four-spheres intersect. Such a diagram
has the form of a simply-laced Dynkin diagram (with vertices for four-spheres and lines for intersections of them),
except that the Cartan matrix may not be positive definite and thus one is not restricted to the ADE case.

76

S. Gukov et al. / Nuclear Physics B 584 (2000) 69108

In crossing a domain wall, cannot change by an arbitrary amount, but only by


something of the form i([S]) where [S] is a class with compact support. The possible
values of modulo changes due to the dynamics, that is due to crossing domain walls, are
thus classified by

4
(X; Z) = /.
(2.18)
H 4 (X; Z)/i Hcpct
This can be reinterpreted as follows. The exact sequence of the pair (X, X) reads in
part
i

H 4 (X, X; Z) H 4 (X; Z) H 4 (X; Z) H 5 (X, X; Z) . (2.19)


i (X; Z). By Poincar duality, H 5 (X, X; Z) =
Here H i (X, X; Z) is the same as Hcpct
H3 (X; Z), and this is zero on dimensional grounds for a bouquet of four-spheres. So the
exact sequence implies that


4
(X; Z) = /.
(2.20)
H 4 (X; Z) = H 4 (X; Z) i Hcpct

Thus the value of modulo changes in crossing domain walls (the right-hand side of
(2.20)) can be identified with the restriction of to X (the left hand side).
If the intersection pairing on H4 (X; Z) is degenerate, then we should define to be the
quotient of H 4 (X, X; Z) by the group of null vectors (which can be shown to be precisely
the image in H 4 (X, X; Z) of H 3 (X; Z)). The dual is the subgroup of H 4 (X; Z)
consisting of elements whose restriction to X is torsion. With these definitions of and
, everything that we have described above carries over (i embeds as a finite index
sublattice of ; the G-fields, with appropriate boundary conditions, take values in ,
and jump in crossing a domain wall by elements of ).
Examples
We will now illustrate these perhaps slightly abstract ideas with examples that will be
important later.
Consider first the simple case that X is a deformation of a quadric singularity:
5
X

za2 = .

(2.21)

a=1
2
2
If we assume that
p  is real and write za = xa + iya , we get xE yE =  and xE yE = 0.
2
Setting uE = xE /  + yE , we see that uE is a unit vector. The subset of X with yE = 0 is a
four-sphere S; since yE uE = 0, X is the cotangent bundle of S. In particular, X is homotopic
to the four-sphere S. This is the case in which the bouquet of spheres is made from just a
4 (X; Z)
single sphere. The self-intersection number of S is S S = 2. 7 The lattice = Hcpct
1

4
is generated by [S], but the dual lattice = H (X; Z) is generated by 2 [S] (whose scalar
4 (X; Z) = 1 / = Z .
product with S is 1). So H 4 (X; Z) = H 4 (X; Z)/Hcpct
2
2
7 To compute this, deform S to the four-sphere S 0 defined by y = u , y = u , y = u , y = u , y =
1
2 2
1 3
4 4
3 5
0. Then S 0 intersects S at the two points u1 = = u4 = 0, u5 = 1, and each point contributes +1 to the
intersection number. Hence S S = S S 0 = 2.

S. Gukov et al. / Nuclear Physics B 584 (2000) 69108

77

A somewhat more sophisticated example is the An1 singularity in complex dimension


four:
Pn (z1 ) +

5
X

za2 = 0.

(2.22)

a=2

Here Pn (z1 ) is a polynomial of degree n. For simplicity we take


Pn (z1 ) =

n
Y

(z1 bi )

(2.23)

i=1

with real bi , b1 < b2 < < bn . For i = 1, . . . , n 1, we define a four-sphere Si by


requiring that z1 is real with bi < z1 < bi+1 , and that the zj for j > 1 are all real or
all imaginary depending on the value of i modulo two. The Si generate the lattice =
4 (X; Z). The intersection numbers of the S are S 2 = 2, S S
Hcpct
i
i
i+1 = 1, with others
i
vanishing. (Si intersects Si+1 at the single point z1 = bi+1 , zj = 0 for j > 1; Si does not
intersect Sj if |j i| > 1.) Endowed with this intersection form, is the root lattice of the
Lie group An1 = SU(n), while the dual lattice = H 4 (X; Z) is the weight lattice. The
quotient is H 4 (X; Z) = / = Zn . It can be shown that X is homotopic to the union
of the Si , which form a bouquet. In this case, the bouquet is associated with the Dynkin
diagram of An .
More generally, if F (z1 , z2 , z3 ) is a polynomial in three complex variables that describes
a deformation of an ADE surface singularity, we can consider the corresponding surface
singularity in complex dimension four:
H (z1 , z2 , z3 ) + z42 + z52 = 0.

(2.24)

The case just considered, with H (z1 , z2 , z3 ) = Pn (z1 ) + z22 + z32 , corresponds to An1 . (The
appropriate H s for the other cases are written at the end of Section 2.5.) For any of the
ADE examples, is the root lattice of the appropriate simply-connected ADE group G,
is the weight lattice of G, and the quotient H 4 (X; Z) = / is isomorphic to the
center of G. One approach to proving these assertions is to show that they are true for
the middle-dimensional cohomology of the surface H (z1 , z2 , z3 ) = 0, and are unaffected
by stabilizing the singularity by adding two more variables with the quadratic terms
z42 + z52 .
2.3. Distance to singularity and hodge structure of cohomology
In the present subsection, we return to the case of a compact CalabiYau four-fold Y .
We suppose that, when some complex parameters ti are varied, Y develops a singularity
that looks like a quasihomogeneous hypersurface singularity F (z1 , . . . , z5 ) = 0, where the
za have degrees ra > 0 and F is of degree 1. Upon varying the complex structure of Y , the
hypersurface is deformed to a smooth one which looks locally like
X
ti Ai (z1 , . . . , z5 ) = 0.
(2.25)
F (z1 , . . . , z5 ) +
i

Here the ti are complex parameters, and the Ai are perturbations of the equation.

78

S. Gukov et al. / Nuclear Physics B 584 (2000) 69108

The first question to examine is whether the singularity at ti = 0 can arise at finite
distance in CalabiYau moduli space. The Khler form on the parameter space is
2
K,
(2.26)
t i tj
where K is the Khler potential K. On the parameter space of a compact CalabiYau
manifold, the Khler potential of the WeilPeterson metric is 8
Z

K = ln .
(2.27)
= dt i d tj

We want to analyze a possible singularity of this integral near za = 0 in the limit that the
ti go to zero. If and only if there is such a singularity, the distance to ti = 0 will be infinite
in the metric (2.26). For analyzing this question, the large za behavior, which depends on
how the singularity is embedded in a compact variety Y , is immaterial (as long as there
is some cutoff to avoid a divergence at large za ); we can, for instance, replace Y by the
hypersurface in (2.25) and restrict the integral to the region |za | < 1.
To determine the small za behavior of the integral, we use a simple scaling. Under za
ra za , scales like a ra 1 and so the integral in (2.27) scales like ||2a ra 2 . Small za
corresponds to small . Hence the condition that the integral converges at small za is
X
ra 1 > 0.
(2.28)
a

This is a satisfying result, in that this is the same condition that was needed to get an
R-symmetry with positive charges and to apply the TianYau theorem on existence of
asymptotically cone-like CalabiYau metrics.
We will now apply this kind of reasoning to address the following question, whose
importance will become clear: as Y becomes singular, what is the Hodge type of the
vanishing cohomology, that is, of the part of the cohomology that disappears at
the singularity? We only have to look at middle-dimensional cohomology, because the
deformation of a hypersurface singularity has cohomology only in the middle dimension.
First let us ask if there is vanishing cohomology of type (4, 0). For this, we normalize
the holomorphic (4, 0)-form of Y in such a way that far from za = 0 it has a limit as
ti 0. Then we ask if the integral
Z

(2.29)
Y

converges as ti 0. If the answer is no, then to make the integral converge as ti 0, we


would have to rescale so that in the limit it vanishes pointwise away from the singularity.
Then in the limit ti 0, would be a closed four-form that is non-zero but vanishes
away from the singularity. There would thus be vanishing cohomology of type (4, 0). If
the answer is yes, there is no vanishing cohomology of type (4, 0).
8 The derivation of this formula is just as in the three-fold case; see [10] for an exposition.

S. Gukov et al. / Nuclear Physics B 584 (2000) 69108

79

We have already seen that convergence of the integral in (2.29) is the condition that the
singularity is at finite distance in moduli space. Hence, singularities that can arise in the
dynamics of a compact CalabiYau four-fold have no vanishing cohomology of type (4, 0).
Now let us look for vanishing cohomology of type (3, 1). The (3, 1) cohomology is
generated by i = D/Dti , where D/Dti is the covariant derivative computed using the
GaussManin connection. To determine if i is a vanishing cycle, we need to examine the
integral
Z
i i ,
(2.30)
Y

and ask if it is finite as all tj 0. If not, then to make the integral converge, we would
have to rescale i by a function of the tj , and in the limit tj 0, i would represent
a nonzero (3, 1) cohomology class that vanishes away from the singularity, or in other
words a piece of the vanishing cohomology of type (3, 1). The integral (2.30) is more
conveniently written as
Z
2

.
(2.31)
ti ti
Y

Whether this integral converges can, again, be determined by scaling. If the function Ai in
(2.25) scales under za ra za as si , then ti scales like 1si and (2.31) scales like ||wi
with
!
X
ra + si 2 .
(2.32)
wi = 2
a

Vanishing (3, 1) cohomology arises when wi 6 0, so that the integral in (2.31) diverges
near z = 0. The most dangerous case is for Ai = 1, si = 0. The condition that wi > 0 for
all i, so that there is no vanishing (3, 1) cohomology, is thus
X
ra > 2.
(2.33)
a

We can classify the models that obey this condition. Consider a LandauGinzburg model
with chiral superfields a , a = 1, . . . , 5 and superpotential F (1 , . . . , 5 ). If a have
P
c = 5a=1 (1
degree ra and F has degree one, then the central charge of this model is b
P
2ra ) = 5 2 a ra . The condition (2.33) thus amounts to 9
b
c < 1.

(2.34)

The singularities that obey this condition are the ADE singularities. They are given, in a
suitable set of coordinates, by
F (z1 , . . . , z5 ) = H (z1 , z2 , z3 ) + z42 + z52 ,

(2.35)

where H (z1 , z2 , z3 ) = 0 is the equation of an ADE surface singularity.


9 Note that in terms of b
c the condition that the local singularity of the fourfold be at finite distance in moduli
space (2.28) is that b
c < 3, which generalizes for an n-fold singularity to b
c < n 1.

80

S. Gukov et al. / Nuclear Physics B 584 (2000) 69108

Application to hypersurface
We have developed this discussion for the case of a compact CalabiYau manifold Y
that develops a hypersurface singularity, but it is more in the spirit of the present paper
to decompactify Y and focus on the hypersurface itself, that is to consider M-theory on
R3 X, where X is a hypersurface that develops the given singularity. This is the natural
framework for studying M-theory near a singularity, with extraneous degrees of freedom
decoupled. Let us therefore now explain the significance of the above results for this case.
P
If we work on the noncompact hypersurface, the condition that a ra + si > 2, which
R
ensures that there is not a divergence of |i |2 near za = 0, also ensures that there is
such a divergence near za = . The large za divergence means that, in M-theory on
R3 X, the modes that deform the singularity of X have divergent kinetic energy and
are not dynamical. They correspond, instead, to coupling constants of the theory near the
singularity; they can be specified externally as part of the definition of the problem.
In the ADE examples, the complex structure modes are all non-dynamical in this sense.
For other examples, positivity of (2.32) does not hold for all i, and therefore some of the
complex structure deformations of X are dynamical; they vary quantum mechanically in
the theory at the singularity. Only those modes for which wi > 0 can be specified externally
and represent coupling constants.
Now let us consider the Hodge type of the G-field in the hypersurface case. For unbroken
supersymmetry in flat spacetime, G must be a harmonic L2 form of type (2, 2) [8]. It
must, as well, be integral and primitive.
For hypersurface singularities with asymptotically conical metrics of the type predicted
by the TianYau theorem, the condition that G be a harmonic L2 form is a mild one in the
following sense. For an asymptotically conical metric on a manifold X, one expects the
i (X; R) in
space of L2 harmonic forms of degree i to be isomorphic to the image of Hcpct
H i (X; R). For hypersurface singularities of complex dimension four, there is only fourdimensional cohomology, so we expect L2 harmonic forms of degree four only. Assuming
4 (X), the image of H i (X; R) in H i (X; R) is all of
there are no null vectors in Hcpct
cpct
H i (X; R), so one expects that all of the four-dimensional cohomology is realized by L2
harmonic forms.
What about the requirement that G be primitive? Primitiveness means that K G = 0,
where K is the Khler form. If G is an L2 harmonic four-form on a manifold whose L2
harmonic forms are all four-forms, then K G is automatically zero (if not zero, it would
be an L2 harmonic six-form). Thus, for singularities of this type, the condition that G
should be primitive is automatically obeyed. 10 In Section 5, we will examine a singularity
of a different sort for which primitiveness of G is an important constraint.
The remaining constraint that we have not examined yet is a severe constraint in the
case of hypersurface singularities. This is the condition that G should be of type (2, 2). For
10 A different explanation of this is as follows. In Section 2.2, we compactified X to a complete but non-Calabi
Yau variety Y 0 by adding a divisor D at infinity. D is an ample divisor, and the primitive cohomology in this
situation is the cohomology that vanishes when restricted to D. This is certainly so for the vanishing cohomology,
whose support is far from D.

S. Gukov et al. / Nuclear Physics B 584 (2000) 69108

81

ADE singularities, as we have seen above, the vanishing cohomology is all of type (2, 2),
P
so the L2 harmonic forms have this property. For other singularities, with a ra < 2, there
is vanishing cohomology of types (3, 1), (2, 2), and (1, 3). Under such conditions, it is
generically very hard to find a non-zero four-form that is of type (2, 2) and integral. Once
an integral four-form G is picked, requiring that it be of type (2, 2) will generally put
restrictions on the complex structure of X. Since some of the complex structure modes
are dynamical whenever there is vanishing (3, 1) cohomology, the restriction on complex
structure that is entailed in making G be of type (2, 2) is likely to play an important role
in the dynamics of these models. In this paper, to avoid having to deal with the dynamical
complex structure modes and the Hodge structure of the singularity, we will study in detail
only the ADE singularities. (For fourfold examples where moduli are dynamically frozen
see [11].)
Here is another way to see the distinguished nature of the ADE singularities. As we
explain in Appendix A, the intersection form on H 4 (X, Z) is positive definite on the
primitive cohomology of type (2, 2), and negative definite on the primitive cohomology of
types (3, 1) and (1, 3). Hence, in particular, having the primitive cohomology be entirely of
type (2, 2) is equivalent to positive definiteness of the intersection form on H 4 (X; Z). For
an intersection form specified by a bouquet of spheres to be positive definite is a condition
that singles out the ADE Dynkin diagrams, so again we see that the ADE singularities are
the ones with vanishing cohomology that is entirely of type (2, 2).
2.4. Interpretation of constraints on G
Since the constraints on G found in [8] have played an important role in this discussion,
we will pause here to attempt to gain a better understanding of these constraints.
We consider compactification of M-theory on a compact four-fold Y . We first suppose
that G is zero. Variations of the CalabiYau metric of Y arise either from variations of
the complex structure or variations of the Khler structure. The variations of the complex
structure are parametrized classically by complex parameters ti , which we promote to
chiral superfields Ti . If hp,q is the dimension of the Hodge group H p,q (Y ), then the
number of Ti is h3,1 . The Khler structure is parametrized classically by h1,1 real
parameters ki . Compactification of the C-field on Y gives rise to h1,1 U (1) gauge fields ai
on R3 whose duals are scalars i that combine with the ki to make h1,1 chiral superfields
that we may call Ki .
If G = 0, the expectation values of the Ti and Ki are arbitrary, in the supergravity
approximation to M-theory. (Instantons can lift this degeneracy [12].) For non-zero
G, this
R
is not so. After picking an integral four-form G (which must be such that X G G > 0
or the equations we are about to write will have no solutions), we must adjust the complex
structure of X so that
G0,4 = G1,3 = 0,

(2.36)

and the Khler structure of X so that


G K = 0.

(2.37)

82

S. Gukov et al. / Nuclear Physics B 584 (2000) 69108

In (2.36), Gp,q denotes the (p, q) part of G.


We want to describe an effective action for the Ti and Kj that accounts for (2.36)
R
and (2.37). Since supersymmetric actions of the general kind d 4 (. . .) do not lift vacuum
degeneracies, we look for superpotential interactions. Thus, we want a superpotential
e (Ki ) that will account for (2.37).
W (Ti ) that will account for (2.36), and a superpotential W
To obtain (2.36), we propose to let be a holomorphic four-form on Y , and take
Z
1
G.
(2.38)
W (Ti ) =
2
Y

This object is not, strictly speaking, a function of the Ti but a section of a line bundle over
the moduli space M of complex structures on Y (on which the Ti are coordinates), since
it is proportional to the choice of . Let L be the line bundle over M whose fiber is the
space of holomorphic four-forms on Y . The Khler form of M can be written
Z

,
(2.39)
= ln
Y

in other words it is ln ||2 for any section of L, and this means [13] that W should
be a section of L. Thus, the linear dependence on in (2.38) is the right behavior
of a superpotential. In supergravity with four supercharges, the condition for unbroken
supersymmetry in flat space is W = dW = 0. With W as in (2.38), the condition W = 0 is
that G0,4 = 0. Also, since the objects d/dti generate H 3,1(Y ), the condition dW = 0 is
that G1,3 = 0. So we have found the supersymmetric interaction that accounts for (2.36).
Another way to justify (2.36) is to consider supersymmetric domain walls. The tension
of a domain wall obtained by wrapping a brane on a four-cycle S is the absolute value of
R
S . If G changes from G1 to G2 in crossing the domain wall, then G2 G1 = 2[S],
so this integral is
Z
1
(G2 G1 ).
(2.40)
2
X

In a theory with four supercharges, the tension of a supersymmetric domain wall is the
absolute value of the change in the superpotential W . So (2.42) should be the change in W
in crossing the domain wall, a statement that is clearly compatible with (2.38).
In a similar spirit, one can readily guess the interaction responsible for (2.37):
Z
e (Ki ) = K K G.
(2.41)
W
X

Here K is a complexified Khler class whose real part is the ordinary Khler class K. The
e = 0 is K G = 0, whose real part is (2.37). W
e = 0 is a consequence of this,
condition d W
and imposes no further condition.
At this point, we may conjecture a generalization of the discussion of [8]. In that
paper, supersymmetric compactifications on R3 Y were considered. More generally,
a supersymmetric compactification might have a (negative) cosmological constant in the

S. Gukov et al. / Nuclear Physics B 584 (2000) 69108

83

noncompact dimensions, leading to a supersymmetric compactification on AdS3 Y . The


condition for such an AdS compactification with unbroken supersymmetry is that W 6= 0
but DW/Dti = 0 (here, since W 6= 0, we must be careful and use the appropriate covariant
derivative D/Dti ). In view of the form of (2.39), one may guess that this should be done by
keeping the condition G1,3 = 0 but dropping the condition G0,4 = 0. This is demonstrated
in Appendix B.
In M-theory on a compact CalabiYau four-fold Y , near a hypersurface singularity,
the relation of the change in the superpotential in crossing a domain wall to (2.42)
shows that W cannot vanish in all vacua. Hence, most vacua will actually lead to AdS3
compactifications in the case of compact Y . Going to a non-compact manifold has the
effect of decoupling gravity and lets us avoid inducing a cosmological constant.
Going back to supersymmetric compactifications to R3 , it is interesting to compactify
one of the directions in R3 on a circle and consider type IIA on R2 Y . The above analysis
carries over immediately for supersymmetric vacua with a nonzero value of the Ramond
Ramond four-form G. However, in type IIA string theory, in view of mirror symmetry and
other T-dualities, one naturally thinks that one should construct a more general effective
superpotential to incorporate the possibility of turning on a full set of RamondRamond
fields, and not just the four-form. Indeed, the mirror of G0,4 would be the RR zero-form
(responsible [14] for the massive deformation of type IIA supergravity), and the mirror of
G1,3 would be the RR two-form. This is under investigation [15].
Physical interpretation
We will now discuss the physical interpretation of the superpotentials that we have
computed.
We have computed the superpotential as a function of the superfields Ti and Kj with all
other degrees of freedom integrated out. For Y a large, smooth CalabiYau four-fold, this is
a very natural thing to do, since the superfields Ti and Kj are massless if G = 0, while other
superfields are massive. However, we have argued that as one approaches a singularity,
there are different vacuum states in the theory at the singularity that are specified by
different choices of the G-field. We will interpret the theory near the singularity as a
theory of dynamical chiral fields such that the critical points of the superpotential as
a function of are given by the possible choices of G-field. Thus, a more complete
b ( ; Ti , Kj ), such
description of the theory would involve a superpotential function W
e
b with respect
that the function W (Ti , Kj ) = W (Ti ) + W (Kj ) is obtained by extremizing W
to the . For fixed choices of Ti and Kj , the extremization with respect to has
different solutions, corresponding to the different choices of G. It is very difficult to see
the superfields explicitly, but for suitable examples we will identify the superpotential
b ( ; Ti , Kj ) in section 3 by studying the soliton structure.
function W
2.5. Analogs for type IIA and F-theory
We have formulated the discussion so far in terms of M-theory on R3 Y , with Y a
CalabiYau four-fold, but there are close analogs for type IIA on R2 Y and (if Y is

84

S. Gukov et al. / Nuclear Physics B 584 (2000) 69108

elliptically fibered) for F-theory on R4 Y .


The analysis of [8] carries over to type IIA, with G now understood as the Ramond
Ramond four-form field. Our analysis of the vacuum structure also carries over readily to
this case. One obvious change is that domain walls are now constructed from four-branes
(with worldvolume R S R2 Y ). Another obvious change is that, in type IIA, the
space-filling membranes that contribute to the formula (2.16) for the flux at infinity are
replaced by space-filling fundamental strings. Also, in the type IIA case, alongside the
RamondRamond four-form, one would want to incorporate the RamondRamond zeroform and two-form, as we have discussed briefly in Section 2.4.
In going to F-theory, the space-filling membranes that contribute to the flux at infinity
are replaced by space-filling threebranes. Also we need to discuss the F-theory analog of
the G-field. Let Y be a four-fold that is elliptically fibered over a base B. Let i , i = 1, 2,
be a basis of integral harmonic one-forms on the fibers, and let be an integral two-form
generating the two-dimensional cohomology of the fibers. Then a four-form G on Y has at
the level of cohomology an expansion
X
Hi i ,
(2.42)
G=g+p +
i

where g, p, and Hi are respectively forms of degree 4, 2, and 3 on B. (Hi is a three-form


on B with values in the one-dimensional cohomology of the fibers, while g and p are
ordinary four- and two-forms on B.) If G is primitive, then it is in particular self-dual (see
Appendix A). For G to be integral, g and p must be integral. Self-duality of G gives a
relation between g and p which, in the limit that the area of the fibers of Y B is very
small, is impossible to obey if g and p are non-zero and integral. Hence, the surviving part
of G in the F-theory limit is contained in the Hi , which are interpreted physically as the
NeveuSchwarz and RamondRamond three-form field strengths of type IIB superstrings.
With g = p = 0, G is an element of the primitive cohomology of Y that is odd under the
involution that acts as 1 on the elliptic fibers and trivially on the base.
In terms of a type IIB description, we have the following structure. Let H NS , H R denote
the NS and Ramond three-form field strengths. Let B denote the base of F-theory visible
to type IIB. Consider
H + = H R H NS ,
H = H R H NS .
We view as varying over B with monodromies around the loci of seven-branes by
SL(2, Z) transformations

a + b
.
c + d

Under such transformations


H + (c + d)1 H + ,
H (c + d)1 H .

S. Gukov et al. / Nuclear Physics B 584 (2000) 69108

85

A supersymmetric configuration in this context is obtained by choosing an integral (1, 2)


form on the base, H , well defined modulo transformation by (c + d)1 around the
7-branes. Alternatively, H is a section of 1,2 L where L is a line bundle over B whose
first chern class is c1 (L) = 12c1 (B). Then we identify
H + = H,

H = H .

Moreover we require that H k = 0 where k denotes the Khler class of B. In this case a
given model is specified by fixing
Z
1
1
H H ,
=N +
2
4
2
B

where N denotes the number of D3-branes.


To describe the domain walls, recall that one can interpret F-theory on R4 Y in terms
of type IIB on R4 B with (p, q)-sevenbranes on a certain locus L B. Domain walls
across which the Hi jump are described by a five-brane wrapped on R3 V R4 B
with V a three-cycle in B. The (p, q) type of the five-brane varies as V wraps around the
discriminant locus in B.
This is a rather complicated structure in general, but to study the local behavior near a
singularity, it simplifies considerably. One reason for this is that near an isolated singularity,
one can replace B by C3 . If we pick coordinates z1 , z2 , z3 on C3 , then the elliptic fibration
over C3 can be described very explicitly by a Weierstrass equation for additional complex
variables x, y:
y 2 = x 3 + f (z1 , z2 , z3 )x + g(z1 , z2 , z3 ).

(2.43)

The fibers degenerate over a singular locus L which is the discriminant of the cubic, given
by = 0, where
= 4f 3 (z1 , z2 , z3 ) + 27g 2 (z1 , z2 , z3 ).

(2.44)

A singular behavior of the elliptic fibration Y just corresponds in this language to a


singularity of the hypersurface L C3 . We are interested in a singular point of L at which
4f 3 + 27g 2 = 0 but f and g are not both zero. 11 Near such a singular point, the detailed
construction of in terms of f and g is irrelevant, and and one can regard L as a fairly
generic deformation of a hypersurface singularity = 0.
Actually, the full structure of (p, q) sevenbranes also simplifies in this situation. The
deformation of an isolated surface singularity is topologically a bouquet of two-spheres,
and in particular simply-connected. Hence, there is no monodromy around which the type
of brane can change; the (p, q) type of the sevenbrane is fixed, and one can think of it (for
example) as a D7-brane. Thus, the F-theory analog of a CalabiYau fourfold singularity is
a more elementary-sounding problem: the study of a D7-brane in R10 = R4 C3 whose
worldvolume is R4 L, where L C3 is developing a singularity.
11 Singularities with f and g both zero are composite 7-branes of various types (the order of vanishing of
discriminant would be bigger than 1). For such cases the simplifications described in the text do not arise and the
full structure of (p, q) sevenbranes is relevant.

86

S. Gukov et al. / Nuclear Physics B 584 (2000) 69108

Now, let us describe the vacuum states and domain walls in this context. The D7brane supports a U (1) gauge field, whose first Chern class is an element of H 2 (L; Z).
This group is a lattice , whose rank is the number of two-spheres in the bouquet.
A D5-brane can end on a D7-brane; its boundary couples magnetically to the gauge field
on the D7-brane. Hence the domain walls across which the first Chern class jumps are
built from fivebranes of topology R3 V , where V is a three-manifold in C3 whose
boundary lies in L. In crossing such a domain wall, the first Chern class jumps by the
2 (L; Z). Poincar
cohomology class [V ], which is an element of = H2 (L; Z) = Hcpct
duality for noncompact manifolds asserts that and are dual, and 12 the natural map
2 (L; Z) H 2 (L; Z) gives an embedding of in . Thus we have a familiar
i : Hcpct
situation: the vacuum is determined by a point in a lattice , and in crossing a domain
wall it can jump by an element of a sublattice . is endowed with an integral quadratic
form (the intersection pairing), and as the notation suggests, is the dual lattice of
with respect to this pairing.
The ADE singularities will furnish important examples in the present paper, for reasons
that we have already explained. Thus, let us explain how they arise in the F-theory context.
An example of an elliptic four-fold fibration Y with an isolated singularity is given by the
following Weierstrass equation:

y 2 = x 3 3a 2x + H (z1 , z2 , z3 ) + 2a 3 .

(2.45)

Here a is an arbitrary non-zero constant, and H is a quasihomogeneous polynomial


describing a singularity in three variables at z1 = z2 = z3 = 0. If we shift x to x + a,
the equation becomes
y 2 = x 3 + 3ax 2 + H (z1 , z2 , z3 ),

(2.46)

and this makes it obvious that the singularity of the elliptic fibration is obtained by
stabilizing the surface singularity H = 0 by adding the quadratic terms 3a 2x 2 y 2 (the
x 3 term is irrelevant near the singularity, which is at x = y = 0). The equation for the
discriminant locus L C3 reduces to H = 0 (plus higher order terms that are irrelevant
near the singularity). So the singularity of the elliptic four-fold is just the stabilization
of the singularity L. To obtain the ADE singularities, for both the surface L and the fourfold Y , we need only select the appropriate H :
H = z1n + z22 + z32 ,
H
H
H
H

An1 ,
n
2
2
= z1 + z1 z2 + z3 , Dn+1 ,
= z13 + z24 + z32 , E6 ,
= z13 + z1 z23 + z32 , E7 ,
= z13 + z25 + z32 , E8 .

12 If there are null vectors in , there is a slightly more elaborate story as mentioned in Section 2.2.

(2.47)

S. Gukov et al. / Nuclear Physics B 584 (2000) 69108

87

2.6. Conformal field theory: first results


Given type IIA, M-theory, or F-theory on a singular geometry, one natural question is
whether a nontrivial conformal field theory arises in the infrared.
In one situation, an affirmative answer to this question is strongly suggested by recent
literature. This is the case of M-theory at a quasihomogeneous four-fold singularity (for
the present discussion this need not be a hypersurface singularity) with a large value of the
conserved quantity that was introduced in Section 2.2:
Z
GG
1
.
(2.48)
=N +
2
(2)2
X

We suppose that the four-fold X is a cone over a seven-manifold Q. Consider M-theory on


R3 X, with a specified (flat) C-field at infinity that we call C , and with a very large
number of membranes near the singularity, such that the total membrane charge (including
the contribution of the C-field) is . This system is believed [1618] to be described in
the infrared by a conformal field theory that is dual to M-theory on AdS4 Q, with a
constant curvature (but topologically trivial) C-field on AdS4 that depends on , and a flat
but topologically nontrivial C-field on Q that is equal to C . For a special case in which
the role of C has been analyzed (for Q = RP7 ) see [19].
The AdS4 dual of this CFT depends only on what one can measure on Q, that is C
and , and not the detailed way of decomposing in terms of N and G as in (2.48).
That decomposition arises if one makes a deformation of the theory, deforming X to a
smooth hypersurface. M-theory on R3 X with X such a smooth hypersurface has vacua
corresponding to all choices of N and G obeying (2.48). When X develops a singularity,
the G-field apparently disappears at the singularity, and the decomposition of into
membrane and G-field terms is lost.
Note that the vacua with N 6= 0 do not have a mass gap even after deforming to
smooth X. There are at least massless modes associated with the motion of the membranes
on X. To get a theory that after deformation of the parameters flows in the infrared to
massive vacua only, one must set to the smallest possible value for a given value of C ,
so that after deforming to a smooth X, N will be zero for all vacua. We recall that C
determines a coset in / . To get a massive theory, must equal the minimum of
Z
1 GG
,
(2.49)
2
(2)2
X

with G running over the elements of the coset of / determined by C ; the massive
vacua are in correspondence with the choices of G that achieve the minimum.
Our goal in the next two sections will be to analyze, for the ADE singularities, the
massive models just described. The analysis will be made by analyzing the domain wall
structure, or, as it is usually called in two dimensions, the soliton structure. To justify
the analysis, we need to know that there are no quantum corrections to the classical
geometry (which we will use to find the solitons). Such corrections would come from
appropriate instantons. For example, for type IIA on a CalabiYau threefold near the

88

S. Gukov et al. / Nuclear Physics B 584 (2000) 69108

conifold singularity, the Euclidean D2-brane instantons wrapped around the S3 in the
conifold smooth out the singular classical geometry [2022]. Likewise, in M-theory
compactifications on suitable CalabiYau four-folds, a superpotential is generated by
wrapped Euclidean fivebranes [12]. Such effects, however, are absent in the examples
we are considering. For example, in the F-theory, we are really studying, as we have
explained above, a sevenbrane on L C3 . Since the C3 has no nontrivial cycles, the
relevant instantons will have to end on L, in order to have finite action. For the instanton
to affect the quantum moduli space it has to be BPS, which in particular requires that the
boundary of the instanton be a nontrivial compact cycle in L. In type IIB string theory
the only possible candidate instantons which could end on a sevenbrane are fivebranes
and onebranes (of appropriate (p, q) type). Viewing them as instantons, their boundaries
would be five- and one-dimensional, respectively. So if L has no nontrivial compact fiveor one-dimensional cycles, then there are no instantons, and quantum corrections do not
modify the singular classical geometry. In our case, L, whose compact geometry consists
of a bouquet of two-spheres, has only two-cycles, so there are no instantons. This is to be
contrasted with the seemingly similar problem of F-theory on a CalabiYau threefold. In
that case, L is a complex curve, with nontrivial one-cycles; instanton one-branes can and
do modify the classical geometry. This is in fact the F-theory version of the description of
the corrections to conifold geometry in type IIA compactification (and reduces to it upon
compactification on T2 ). For F-theory on a four-fold, if the singularity of the surface L
is not isolated, then it would generically also have nontrivial one-cycles and would thus
receive corrections.
For M-theory or type IIA near a four-fold hypersurface singularity, a similar statement
holds. In this case, the local geometry of the deformed singularity has nontrivial four-cycles
only. Thus there is no room for instantons, i.e., wrapped Euclidean M2- or M5-branes,
which would require nontrivial three or six-cycles on X. Thus the classical singularity
survives quantum corrections.
Thus, to analyze the small theories, we will look for supersymmetric domain walls
using the classical geometry near the singularity. The domain walls are constructed from
branes whose volumes vanish as the hypersurface X becomes singular, so their tensions
go to zero. Thus one can reasonably hope to get a description in terms of an effective
theory that contains only light degrees of freedom and generates these domain walls. In
fact, for the massive models derived from ADE singularities, we will propose a description
in terms of an effective superpotential for a certain set of chiral superfields that generate
the same soliton structure. This description will make clear that one should expect flow to
a nontrivial IR conformal field theory in the two-dimensional cases, and in a few cases in
three dimensions.
The basic strategy for identifying a supersymmetric theory based on its BPS soliton
structure is the classification approach of [23] to N = 2 supersymmetric theories in two
dimensions. Consider a theory with N = 2 supersymmetry in two dimensions with k
vacua, and consider the integral k k matrix S given by
S = 1 A,

(2.50)

S. Gukov et al. / Nuclear Physics B 584 (2000) 69108

89

where 1 represents the identity matrix and A is a strictly upper triangular matrix whose Aij
entry for i < j is the number of nearly massless BPS solitons interpolating between the ith
sector and the j th sector weighted with the index (1)F F [24], i.e.,
Aij = Trij solitons(1)F F.
It was argued that this massive deformation comes from a CFT in the UV limit with central
charge b
c and k chiral fields with R-charges qi which satisfy



c)
Eigenvalues S t S = exp 2i(qi 12b
(even the integral part of qi can be determined from Aij [23]). This is a strong restriction,
and in case of deformations of minimal models, the solitons completely characterize the
conformal theory. In other words, any theory which upon mass deformation has the same
solitonic structure as that for a massive deformation of a minimal model is equivalent to it!
For non-minimal models, the relation above between the spectrum of the solitons and the
charges of chiral fields is still a very powerful connection and in particular fixes the central
charge of the corresponding conformal theory.
Above two dimensions, the domain wall or soliton analysis still identifies an effective
superpotential, but it is less common for a theory with a given superpotential to flow to a
nontrivial IR conformal field theory. For instance, a theory with a single chiral superfield
and superpotential W = n is believed to flow to a nontrivial CFT in two dimensions
for all n > 2, while in three dimensions this is expected only for n = 3 [25], and in four
dimensions, it is believed to flow to a trivial IR theory for all n. In any event, our analysis
will identify the nonperturbative massless fields and superpotential near the singularity
also in three and four dimensions. Also, even in four dimensions, a n superpotential can
become relevant as a perturbation to certain fixed points [26], so with some modification
of our construction, the superpotential we find may eventually be important in analyzing
four dimensional CFTs that arise from string theory.
The soliton analysis will give detailed information about the behavior for small
membrane charge, which is the opposite limit from the AdS description discussed above
that governs the large charge behavior at least for the M-theory compactifications. For
the type IIA and F-theory compactifications, the description of the large charge behavior
appears to be less simple.

3. Geometry of domain walls


As explained in Section 2.6, our task now is to analyze the soliton structure for certain
hypersurface singularities. In fact, we will consider the Ak singularities which were
introduced in Section 2.2.
Instead of specializing to four-folds, it proves insightful to consider the more general
problem of identifying BPS states of wrapped n-branes in a CalabiYau n-fold near an
isolated singularity. To study the behavior near an Ak singularity, we consider a local model
for a CalabiYau n-fold given by
2
= 0,
Pm (z1 ) + z22 + + zn+1

90

S. Gukov et al. / Nuclear Physics B 584 (2000) 69108

where Pm (z1 ) is a polynomial of degree m = k + 1 in z1 . When Pm has two equal roots,


we get a singular geometry. The most singular geometry arises when Pm (z1 ) = z1m . For
a generic polynomial Pm (z1 ), the geometry is not singular and the compact homology
of this manifold has a basis made of m 1 spheres of real dimension n intersecting
each other according to the Am1 Dynkin diagram. For a particular choice of Pm , we
explained how to construct these spheres in Section 2.2. The intersection form on the
compact homology is symmetric if n is even and antisymmetric if n is odd. We would
like to consider minimal wrapped n-branes, i.e., minimal supersymmetric cycles, in this
geometry. A supersymmetric cycle is a Lagrangian submanifold (that is, the Khler form
vanishes on it). Moreover, on a minimal supersymmetric n-cycle the holomorphic n-form
of CalabiYau is real (with a suitably chosen overall phase) and gives the volume of the
n-brane. For a minimal supersymmetric cycle the quantity
Z
V = ||,
C

which is the volume of cycle C, is minimized and is given by


Z
V =
C

for some choice of phase . Or stated equivalently, the condition is that


Z

Z




|| = ,


C

which is the condition for minimizing the volume of C among the Lagrangian submanifolds in a given homology class.
The holomorphic n-form , up to an overall complex scale factor, is given by
=

idz1 dzn
dz1 dzn
.
=q
zn+1
z22 + + zn2 Pm (z1 )

We would like to minimize the volume form given by ||. To count the minimal
supersymmetric cycles, we follow the strategy in [27] and decompose the geometry to the
fiber and the base as follows. Suppose C is a supersymmetric minimal cycle. Consider
the image of C on z1 . This is a one-dimensional subspace, because for a fixed z1 , the
P
manifold (being defined by j >1 zj2 = Pm (z1 )) has for its only nontrivial cycle a sphere
1/2 , from which one can deduce by
Sn1
z1 . Note that the radius of this sphere is |Pm (z1 )|
scaling that

Z

dz2 dzn1

(3.1)
= |Pm (z1 )|(n2)/2



zn
Sn1
z1

up to an irrelevant multiplicative constant. The inverse image of C over a point in the z1


plane must, if not empty, be a minimal cycle, and so must be (n 1)-dimensional; hence

S. Gukov et al. / Nuclear Physics B 584 (2000) 69108

91

the image of C in the z1 plane must be one-dimensional. The minimization of || will be


done in two steps: We first consider Lagrangian submanifolds Cf (z1 ) for a fixed z1 which
R
minimize the Cf (z1 ) || and next consider the minimization of the volume interval over
an interval I in z1 . In this way we get using (3.1)
Z
Z Z
Z

n2


|| =
= |Pm (z1 )| 2 dz1 .
(3.2)


I

Cf (z1 )I

S n1 (z1 )

We now minimize the volume of the supersymmetric n-cycle with respect to the choice of
the one-dimensional line segment I representing the image of the supersymmetric cycle
on the z1 plane. One can allow the line segment to end at some special points on z1
where Pm (z1 ) = 0, and these are the only allowed boundaries. In fact, precisely if the
line segment terminates at zeroes of Pm , the D-brane worldvolume is closed and smooth.
Indeed the topology of the cycle is an Sn which can be viewed as an Sn1 sphere fibered
over the interval, where at the boundaries of the interval the radius of Sn1 vanishes. The
expression (3.2) is minimized if along the segment on z1 plane the condition |Pm (z)| =
Pm (z) is satisfied for some z-independent phase . Let us define a function W with
n2

dW = Pm 2 dz1 .

(3.3)

In terms of W , the condition for minimal volume is that the image of the curve in the W
plane be a straight line along the direction specified by 1 . Moreover, the end-points of
the segment in the z1 plane correspond to critical points in W , i.e., dW = 0 (for n = 2
the endpoints are defined by the condition that Pm (z1 ) = 0). These conditions are identical
[28] for finding solitons in an N = 2 LandauGinzburg theory in two dimension (or more
generally, BPS domain walls in theories with four supercharges in dimensions two, three,
or four) with superpotential given by W ! If n is even, (3.3) corresponds to a well defined
function of z1 . If n is odd, it gives rise to a well defined (meromorphic) one-form on a
hyperelliptic cover of the z1 plane, branched over the zeroes of Pm (z1 ).
Strictly speaking we have constructed the supersymmetric cycle by assuming that the
condition that the cycle Cf be Lagrangian is the same as being Lagrangian relative to the
Khler form induced on the fiber. This is not necessarily true. For example if the Khler
form has a piece of the form
k = + fi dz1 dzi +
the condition of Lagrangian gets modified. In special cases, like when the polynomial Pm
has real coefficient one can use a Z2 antiholomorphic involution to argue that the cycles we
constructed are both Lagrangian and supersymmetric. In the more general case we proceed
as follows: consider a generic
Y
Pm (z1 ) = (z1 ai ).
i

Consider a one parameter family of CalabiYau metrics where


ai (t) = tai .

92

S. Gukov et al. / Nuclear Physics B 584 (2000) 69108

Note that the BPS states will be the same for all t, because the condition of the BPS charges
getting aligned does not change as we change t (the BPS charges only receive an overall
rescaling). However to construct the Khler metric as a function of t we note that it can be
mapped to the previous metric by defining
e
z1 = tz1 ,
e
zi = t n/2 zi ,

for i 6= 1.

Thus we use the z variables but rescale the Khler form accordingly. In this way as t
the mixed terms in the Khler form are dominated by the terms purely in the fiber direction
(for n > 2 which is the case of main interest). Therefore in this limit the condition of
Lagrangian submanifold in the fiber that we have used becomes accurate.
Let us consider some special cases. The cases for a K3 surface and for a CalabiYau
threefold have already been considered in [27] (see also [29,30]), and will be reviewed
below.
Solitons for K3
In the case n = 2, the above geometry is the complex deformation of the Am1
singularity. For any choice of the polynomial Pm (z1 ), we expect m(m 1)/2 solitons
(up to the choice of orientation) to complete the adjoint representation of U (1)m1 to the
SU(m) gauge multiplet. From (3.3), we see that in this case W = z1 . There are m roots for
Pm (z), and the solitons correspond to straight lines between the roots. Note that this gives
m(m 1)/2 solitons up to the choice of the orientation, as was anticipated.
Solitons for CY3
For the case of CalabiYau threefolds, n = 3. In this case W is defined by
1

dW = Pm (z1 ) 2 dz1 .
Here dW can be viewed as a meromorphic one-form on a hyperelliptic Riemann surface
over z1 branched over the roots of Pm (z1 ). The geometry of these solitons for this class
of conformal theories would correspond to straight lines on the Jacobian of this surface
defined by the integrals dW and is presently under study [31].
Solitons for CY4
For the case of four-folds, which are of course our main focus in the present paper, the
definition of W in (3.3) shows that W is a polynomial of degree m + 1 in the z1 plane.
We have already shown that the conditions for finding the solitons in this geometry are
the same as those in an LG theory with the superpotential W . In this case, however, if
we use our four-fold in type IIA superstring theory, the analogy becomes more precise:
compactification on the four-fold leads to a theory in two dimensions with N = 2, and it
is natural to identify the corresponding W with the superpotential of an N = 2 Landau
Ginzburg theory. We will indeed argue that for a certain choice of the membrane charge,
the type IIA on a deformed A1 singularity leads to an N = 2 theory with the same W for its

S. Gukov et al. / Nuclear Physics B 584 (2000) 69108

93

superpotential. For more general choices of the membrane charge, we find closely related
KazamaSuzuki coset models at level 1.
Before we discuss these subtleties, note that even though we have m critical points, it is
no longer true in general that we have m(m 1)/2 solitons. In general the pre-image of a
straight line connecting the images of critical points in the W plane will not connect the
critical points in the z1 space. In fact as we change the polynomial Pm (z1 ), it is known that
the number of BPS states jumps [24]. For some choices of Pm (z1 ) we do have exactly the
same number of solitons as in the K3 case. For example, it has been shown [28] that for
Pm (z1 ) = z1m n
for any constant , there is one soliton for each pair of m critical points z1 = with
m = 1, though, unlike the K3 case the image in the z1 plane is not a straight line.
It would be interesting to compare the formula for W that we have deduced from
the soliton structure to the analysis of Section 2.4. Although this is guaranteed to work,
because both capture the mass of the BPS soliton, we have not attempted to check this
correspondence explicitly.

4. Identifications with KazamaSuzuki models


Let us consider in more detail type IIA strings propagating on a smooth hypersurface X
obtained by deforming the Am1 singularity:
Pm (z1 ) + z22 + z32 + z42 + z52 = 0.
As explained in Section 2, in order to specify the problem fully, we must fix the value C
of the C-field at infinity and also the flux
= N + 12 2 ,

(4.1)

where = [G/2] is the characteristic class of the C-field.


As we explained in Section 2.2, is restricted to a fixed coset in / , where and

are the root and weight lattices of the Lie group SU(m). The coset in which takes
values is determined by C . For the theory to have a mass gap, as discussed in Section 2.6,
we set equal to the minimum value of 12 2 (for all in the given coset), so that N = 0
for all vacua.
This can be made very explicit in the case of the Am1 singularity. C takes values
in / , which is isomorphic to the center of SU(m), or Zm . For k = 0, . . . , m 1, if
C = k, then to minimize 2 , must be a weight of the k-fold antisymmetric tensor
product of the fundamental representation of SU(m). We denote that representation as Rk .
The number of choices of is the dimension of Rk or m!/k!(m k)!. This is the number
of vacuum states of the k t h model, if Pm is such that the hypersurface X is smooth.
For k = 0, there is only one vacuum ( = 0), and the theory is trivial and massive. Let
us consider the next simplest case, where k = 1 and is a weight of the fundamental
representation of SU(m). In this case, we have m vacua. To find the degeneracy of the
solitons between these vacua, we use the analysis of Section 3. We found that the solitons

94

S. Gukov et al. / Nuclear Physics B 584 (2000) 69108

are exactly the same as those for an N = 2 LG theory with a chiral field and a
superpotential W obeying dW/d = Pm (). In fact we can identify the m vacua with
the m critical points of Pm , and as we found in Section 3, the condition for the existence
of a soliton in the LG theory is exactly the same as the condition for a BPS wrapped fourbrane in type IIA near the Am1 singularity. In this case the soliton data are enough to
determine the theory [23] as discussed at the end of Section 2; the two-dimensional theory
with superpotential W is the Am minimal model [3,4]. So the type IIA theory near an Am1
four-fold singularity is governed by the Am N = 2 minimal model. This model can also be
viewed as an N = 2 KazamaSuzuki coset model at level one, of the form
SU(m)
.
SU(m 1) U (1)
For the M-theory or F-theory near an Am1 four-fold singularity, we still get a description
in terms of a chiral field with the same superpotential, but in most instances (the exception
being 3 in three dimensions), a theory in three or four dimensions with a m+1
superpotential is believed to flow to a free theory in the infrared.
We now consider the other choices of C , so that is a weight of the k-fold antisymmetric product of the fundamental representation of SU(m) with some k > 1. We argue
that it has exactly the same solitonic spectrum as a deformation of the following LG theory,
which we will call the k-fold symmetric combination of the k = 1 model. Consider the
function of k variables
W (z1 , . . . , zk ) = z1m+1 + z2m+1 + + zkm+1 .
It is invariant under permutations of the zi , and so can be expressed as a polynomial in the
elementary symmetric functions
X
zi1 zil .
xl =
i1 <<il

The superpotential we consider is thus


W (x1 , . . . , xk ) = W (z1 ) + + W (zk ).

(4.2)

The LG model with superpotential (4.2) has been conjectured in [32] to be equivalent to
the following KazamaSuzuki coset model at level 1:
SU(m)
.
SU(m k) SU(k) U (1)
For the deformed singularity, with W = Pm , we claim that the deformed LG superpotential is given by
W (x1 , . . . , xk ) = W (z1 ) + + W (zk ),
where again what we mean by this expression is that the superpotential is W (z1 ) + +
W (zk ) regarded as a polynomial in the elementary symmetric functions x1 , . . . , xk . Let us
see why this LG superpotential has exactly the same solitonic spectrum that we get for
type IIA at an Am1 four-fold singularity with C = k mod m. It is not too difficult to
show [23,33] that the set of vacua of a LG theory made of a k-fold symmetric combination

S. Gukov et al. / Nuclear Physics B 584 (2000) 69108

95

of a given LG theory (in the sense introduced above) can be identified with the k-fold
antisymmetric tensor product of the space of vacua of the original LG theory. 13 As we
already discussed, the k = 1 model has a one-variable superpotential W (z), and its vacua
correspond to the fundamental weights of the SU(m) lattice. Thus we can identify the
vacua of the k-fold symmetric combination of the k = 1 model with the weights of the
k-fold anti-symmetric tensor representation Rk . As far as the allowed solitons, on the LG
side, they can be constructed in the decoupled theory with superpotential W (z1 , . . . , zk ) =
W (z1 ) + + W (zk ) before re-expressing this in terms of the symmetric functions xi . In
this description, it is clear that soliton states are just the products of soliton states in the
individual one-variable theories, and that irreducible solitons (which cannot break up into
several widely separated mutually BPS solitons) are solitons in just one of the variables zi .
So if we label the vacua by |i1 , . . . , ik i, with is denoting a vacuum in the sth one-particle
theory, then the allowed solitons only change one vacuum index at a time. So the solitons
of a LG theory that is constructed as a k-fold symmetric combination of a one-variable
theory are in natural correspondence with the solitons of the one-variable theory. This is
the same result that we get from type IIA near the Am1 singularity with C = k mod m.
Indeed, for this model, the solitons are constructed by finding supersymmetric four-cycles.
The analysis of those cycles in Section 3 depends only on the geometry of the hypersurface
and makes no reference to C . Hence the solitons for any k are in a natural sense the same
as the solitons of the k = 1 model.
In other words the solitons are in 11 correspondence with those roots of SU(m) that
appear as solitons for the one-variable LG superpotential given by W . Whichever roots
appear act in the natural way on the weights of the representation Rk . In the case W =
zm+1 az, all the roots appear with multiplicity 1. This structure for the solitons of the
deformed KazamaSuzuki model was suggested in [34] where it was argued to correspond
(with a specific choice of Khler potential) to an integrable model.
4.1. Other ADE singularities
So far we have mainly considered the local singularity to be
H (z1 , z2 , z3 ) + z42 + z52 = 0
with H being an Am1 singularity. Here we would like to generalize this to the case where
H determines a D or E type singularity.
The general structure is quite like what we have seen for Am1 . C takes values in

/ , where is the root lattice of the appropriate simply-connected ADE group G, and
is the weight lattice of G. The quotient / is isomorphic to the center of G.
Just as in the SU(m) case, to make possible a deformation to a massive theory, we need
to pick so that ranges over the weights of the smallest representation with a given
13 The main point that must be shown is that the vacua in the different factors must be distinct. To illustrate
why, it suffices to consider the case that k = 2 and that each individual model has only one vacuum. So we start
with m = 2: W (z) = z2 . Then we write W (z1 , z2 ) = z12 + z22 in terms of the symmetric functions x1 = z1 + z2 ,
x2 = z1 z2 , getting W (x1 , x2 ) = x12 2x2 . This function has no critical points, so the combined model has no
supersymmetric vacua, as expected.

96

S. Gukov et al. / Nuclear Physics B 584 (2000) 69108

nontrivial action of the center of G. (If we pick the trivial action of the center, we will
get the trivial representation and a massive free theory.) The appropriate representations
are the representations with Dynkin label 1. In the Dn case, the relevant choices are the
vector and spinor representations. For D2n , there are two different spinor representations,
but they differ by an outer automorphism of D2n and give equivalent theories. So there are
essentially two choices of C leading to massive theories based on the Dn singularity.
For the E6 theory, there is only the fundamental 27-dimensional representation (and
its conjugate, which gives an equivalent theory); for E7 there is the fundamental 56dimensional representation. So E6 and E7 lead to one massive theory each. E8 is simplyconnected with trivial center, so we cannot use it to get a conformal theory with a massive
deformation.
The distinguished representations that we have described are in one-to-one correspondence with nodes of index 1 on the ADE Dynkin diagram and thence with Hermitian
symmetric spaces G/H (where H is obtained by omitting the given node from the Dynkin
diagram). Apart from the Grassmannians SU(m)/SU(k) SU(m k) U (1) that we have
already encountered, these Hermitian symmetric spaces are as follows. For the Dn case,
there are two inequivalent choices, given by
SO(2n)
,
SO(2n 2) SO(2)
SO(2n)
,
U (n)

for fundamental rep.,

for spinor rep.,

and for the E6 and E7 cases one has


E6
,
SO(10) U (1)
E7
,
E6 U (1)

for fundamental rep.,

for fundamental rep.

Such a Hermitian symmetric space determines a series of N = 2 KazamaSuzuki models


(at level 1, 2, 3, . . .). It is natural to conjecture that, as we have found for Am1 , the massive
models obtained from type IIA at an ADE singularity are the massive deformations of the
corresponding level 1 KazamaSuzuki (or KS) models. As a first check, it is known that
for a level one G/H KS model, the dimension of the chiral ring is equal to the dimension
of the corresponding representation of G. This in turn is equal to the dimension of the
cohomology of G/H and it was conjectured in [32] that the chiral ring is isomorphic to the
cohomology ring, which in turn was shown to arise from the ring of an LG theory. Thus
the G/H theories at level 1 were identified with specific LG models. 14 Moreover, the
structure of the solitons for a special (integrable) deformation of the KS model at level 1
was studied in [34] and it was conjectured that the solitons exist precisely for each allowed
single root acting on the corresponding weight diagram. Though we have not analyzed the
BPS spectrum of the D4-branes in this case to find the multiplicity of the solitons for each
14 The higher level KS models do not generally admit an LG description.

S. Gukov et al. / Nuclear Physics B 584 (2000) 69108

97

root, it is natural to expect that at least for specific deformations, just as in the Am1 case,
the solitons are given by the root lattice of the corresponding group with multiplicity 1. In
this case we would reproduce the solitonic structure anticipated in [34]. It is quite satisfying
that we apparently get all the Hermitian symmetric space KS models at level 1 in such a
uniform way from considering singularities of CY four-folds.

5. Other types of singularity


The only four-fold singularities that we have so far considered in any detail are
hypersurface singularities. A CalabiYau four-fold can, however, develop singularities of
many different types. We cannot offer any sort of overview of the possibilities, but will
briefly analyze two cases in the present section.
5.1. Hyper-Khler singularities
First we will consider what one might call hyper-Khler singularities singularities
near which Y admits a hyper-Khler structure, though Y may not be globally a hyperKhler manifold. An important fact here is that M-theory compactification on R3 Y
with Y hyper-Khler has N = 3 supersymmetry in three dimensions, because the space of
covariantly constant spinors on a hyper-Khler eight-manifold is three-dimensional.
To isolate the behavior near the singularity, we replace Y by an asymptotically conical
hyper-Khler manifold X that is developing a singularity. We will focus on a very concrete
example, with X = T CP2 , the cotangent bundle of complex projective two-space. This
hyper-Khler manifold is conveniently obtained by considering a U (1) gauge theory with
eight supercharges, and three hypermultiplets Ai , i = 1, 2, 3, of charge 1. 15 There is an
SU(3) global symmetry group, with the Ai transforming as the 3. There is also an SU(2)
R-symmetry group, and it is possible to add an SU(2) triplet of FayetIliopoulos terms
dE to the D-flatness equations. A manifestly SU(2)-invariant way to exhibit the D-flatness
equations is as follows. The bosonic parts of the Ai can be regarded as a complex field Ai ,
= 1, 2, transforming as (3, 2) under SU(3) SU(2)R . The D-flatness condition is
X

Ai A i = dE E ,
(5.1)
i

with E the traceless 2 2 Pauli matrices. The moduli space X of zero energy states of the
classical gauge theory is the space of solutions of (5.1) divided by the action of the gauge
theory. In this description, it is manifest that if dE = 0, then X has an SU(3) SU(2)R
symmetry, broken if dE 6= 0 to SU(3) U (1)R . The SU(3) preserves the hyper-Khler
structure, and SU(2)R rotates the three complex structures on X. If dE = 0, X is a cone
over a seven-manifold Q described by
X

Ai A i = .
(5.2)
i
15 We will present this gauge theory as a formal device, but it may have a physical interpretation in terms of a
membrane probe of the singularity.

98

S. Gukov et al. / Nuclear Physics B 584 (2000) 69108

It is fairly easy to see that this manifold is a copy of SU(3)/U (1), where the U (1) acts by
right multiplication by

(5.3)
diag ei , ei , e2i .
SU(3) acts on SU(3)/U (1) by left multiplication, and SU(2)R acts by right multiplication
by SU(3) elements that commute with (5.3). Even if dE 6= 0, X is asymptotic to a cone
over Q at big distances. The R-symmetry group that acts faithfully on X is actually
SO(3)R = SU(2)R /Z2 . That is because the center of SU(2)R is equivalent to a U (1) gauge
transformation. In M-theory on R3 X, the three spacetime supersymmetries transform
as a vector of SO(3)R .
Now let us explain why for dE 6= 0, X is equivalent to T CP2 . In terms of a description
that makes manifest only half the supersymmetry of the gauge theory, one can break
up the bosonic part of the Ai into pairs of complex fields B i , Ci , transforming as 3
and 3 of an SU(3) symmetry group, and with charges 1 and 1 under the U (1) gauge
group. (Compared to the previous description, B i = Ai1 and Ci = A i 2 .) This description
breaks SO(3)R to SO(2)R = U (1)R , with dE splitting as a real component dR and complex
component dC . The D-flatness equations of the U (1) gauge theory are in this description
X
X
|B i |2
|Cj |2 = dR ,
i
X

B i Ci = dC .

(5.4)

One must also divide by the actionpof U (1). By an SO(3)R rotation, one can set dC = 0 and
P
P
ei = B i / dR + j |Cj |2 obey i |B
ei |2 = 1 and, after dividing
dR > 0. The quantities B
2
by the gauge group, define a point in CP . With dC = 0, the second equation in (5.4) can be
interpreted to mean that Ci lies in the cotangent space to CP2 , at the point determined by
ei . Thus X is isomorphic to T CP2 . For any manifold W , regarded as the zero section
the B
of T W , the self-intersection number W W is equal to the Euler characteristic of W . The
Euler characteristic of CP2 is 3, so in our example
W W =3

(5.5)

with W = [CP2 ].
Though turning on dE breaks the SO(3)R symmetry of X to SO(2), it preserves the hyperKhler structure and all of the supersymmetry of M-theory on R3 X.
The appearance of an SO(3)R symmetry at dE = 0 is a hint that M-theory on R3 X
flows to a superconformal field theory in the infrared as dE 0. Indeed, in three space
time dimensions with N supercharges, the superconformal algebra contains an SO(N )R
R-symmetry group.
To get more insight, let us now analyze the possible G-fields on the smooth manifold
X with dE 6= 0. Since X is contractible to CP2 , one has H 4 (X; R) = H 4 (CP2 ; R). The
non-zero Betti numbers are h0 = h2 = h4 = 1. The cohomology with compact support
is, by Poincar duality, the dual of this, so the non-zero Betti numbers with compact
support are h4cpct = h6cpct = h8cpct . Hence, just on dimensional grounds, the natural map
k (X; R) H k (X; R) is zero except for k = 4. For k = 4, H 4 (X; R) is generated
i : Hcpct
cpct

S. Gukov et al. / Nuclear Physics B 584 (2000) 69108

99

by the class [W ] = [CP2 ], and the nonzero intersection number (5.5) implies that i 6= 0.
For an asymptotically conical manifold, one expects the space of L2 harmonic forms
to coincide with the image of i, so in the present example we expect precisely one L2
harmonic form , in dimension four. is necessarily primitive with respect to all of the
complex structures, since if K is any of the Khler forms, then K, if not zero, would
be an L2 harmonic six-form.
Hence, turning on a nonzero G-field, proportional to , preserves all of the supersymmetries. In fact, we must turn on such a G-field, for the following reason. According to [5],
on a spacetime X, the general flux quantization law for G is not that G/2 has integral
periods but that the periods of G/2 coincide with the periods of c2 (X)/2 mod integers.
(There is a slightly more general formulation if X is not a complex manifold.) In our situation, the integral of c2 (X)/2 over CP2 is a half-integer, 16 so we need
Z
1
G
Z+ ,
(5.6)
2
2
CP2

and in particular G cannot be zero.


If we normalize the four-form to represent the class [CP2 ], then generates
4
Hcpct (X; Z) (or rather its image in real cohomology). Also, = 3, so the dual lattice
H 4 (X; Z) is generated by /3. Hence, we require
 


G
= k + 12 , with k Z.
(5.7)
2
3
4 (X; Z) = Z . The different possibilities for the
One also has H 4 (Q; Z) = H 4 (X; Z)/Hcpct
3
restriction of the C-field to X = Q are determined by the value of k modulo three.
In the presence of N membranes and a G-field, the membrane flux at infinity is
Z
(k + 12 )2
GG
1
.
(5.8)
=
N
+
=N +
2
6
(2)2
X

In evaluating the integral, we used (5.7) and the fact that = 3. A check on (5.8) is
that if k is shifted by an integer multiple of 3 (the 3 is needed so as to leave fixed the
restriction of G to Q), changes by an integer. According to the discussion in Section 2,
a model is specified by fixing the value of and also by fixing the value of k modulo 3.
A supersymmetric vacuum is then found by finding a nonnegative N and an integer k in
the given mod 3 coset such that (5.8) is obeyed. There is precisely one case of a model
having more than one vacuum, with all vacua having N = 0. This arises for = 3/8, with
k = 1 and k = 2. We do not know a LandauGinzburg or other semiclassical description
for this N = 3 model with two vacua (but see below).
For sufficiently large , this model (at dE = 0) is expected to flow to a nontrivial
superconformal field theory in the infrared. Indeed, the standard conjectures would suggest
16 Let the total Chern class of the tangent bundle of CP2 be 1 + c + c . The total Chern class of the cotangent
1
2
2

2
bundle of CP2 is then 1 c1 + c2 . The total Chern class of T CP2 , restricted
R to CP T CP , is hence
(1 c1 + c2 )(1 + c1 + c2 ) = 1 c12 + 2c2 , so c2 (T CP2 ) = c12 + 2c2 . Since CP2 c12 = 9, which is odd, the
claim follows.

100

S. Gukov et al. / Nuclear Physics B 584 (2000) 69108

that the SCFT in question is dual to M-theory on AdS3 Q, with the C-field on Q being
determined by the value of k modulo three. We have no good way at present to determine
if the model flows to a nontrivial SCFT also for small .
We expect that instead of T CP2 , one could in a similar way analyze T F , where F is
a two-dimensional Fano surface. One can also consider a collection of intersecting CP2 s
(with a suitable normal bundle over it) and carry out a similar analysis.
Physical interpretation of gauge theory?
So far the U (1) gauge theory with three charged hypermultiplets has been considered
just as a mathematical device. It is natural to wonder whether, in fact, this gauge theory
can be interpreted physically as the long wavelength theory of a membrane probe of the
R3 T CP2 solution of M-theory. More generally, we would like to find an effective action
for N membranes probing the R3 T CP2 singularity (in the limit that CP2 is blown
down) that will give a gauge theory dual of M-theory on AdS4 Q. In the spirit of [17,
18], such a description might be roughly as follows. Consider an N = 4 supersymmetric
gauge theory in three dimensions with gauge group S(U (N) U (N)) and hypermultiplets
Break N = 4 to N = 3 with some ChernSimons
consisting of three copies of (N, N).
interaction, determined by the C-field. (Gauge theories with ChernSimons interactions are
essentially the only known classical field theories in three spacetime dimensions without
gravity with N = 3 supersymmetry. For a study of their dynamics in the abelian case,
see [35].) Such a model might have roughly the right properties.
5.2. Blowup of orbifold singularity
The other kind of singularity that we will briefly examine is a simple orbifold singularity.
We begin with C4 , with complex coordinates z1 , . . . , z4 , and consider the Z4 symmetry
za iza . The quotient C4 /Z4 is a CalabiYau orbifold.
If one analyzes this type of orbifold in type IIA string theory, one finds that there is one
blow-up mode and no complex structure deformation. The blow-up corresponds to a very
simple resolution of the singularity, in which it is replaced by the total space W of a line
bundle L = O(4) over CP3 . Thus, CP3 is embedded in W as an exceptional divisor, the
zero section of L. W admits a CalabiYau metric, asymptotic in closed form to the flat
metric on C4 /Z4 ; because of the SU(4) symmetry of W , it is actually possible to describe
this metric by quadrature, though we will not do so here.
One might at first think that one could approach the C4 /Z4 orbifold singularity in
M-theory by a motion in Khler moduli space, leading to a blow-down of the exceptional
divisor CP3 W . However, since the Hodge numbers hi,0 (CP3 ) are zero for i > 0,
fivebrane wrapping on CP3 will produce a superpotential [12], proportional roughly to
eV with V the volume of CP3 . Moreover, though the multiple cover formula for multiple
fivebrane wrapping in M-theory is not known, analogy with other multiple cover formulas
(such as the formula for multiple covers by fundamental strings [36,37]) suggests that the
sum over multiple covers of CP3 will produce a pole at V = 0. If this is so, there will
not be interesting long distance physics associated with the behavior of M-theory near a

S. Gukov et al. / Nuclear Physics B 584 (2000) 69108

101

C4 /Z4 singularity. At any rate, one certainly cannot expect to study M-theory on C4 /Z4
while ignoring the superpotential.
Is it possible to include a G-field on W while preserving supersymmetry? If so, then
since the G-field must vanish in cohomology on a fivebrane worldvolume (because of the
existence of a field T on the fivebrane with dT = G), in the presence of the G-field the
superpotential would be absent, and the question of the behavior near the C4 /Z4 singularity
would be restored.
The answer to the question of whether a supersymmetric G-field is possible turns out,
however, to be no, in the following interesting way. First of all, W is contractible to CP3 ,
so its nonzero Betti numbers are h0 = h2 = h4 = h6 = 1. For cohomology with compact
support, one has the dual Betti numbers h2cpct = h4cpct = h6cpct = h8cpct = 1. This suggests that
k (W ; R) H k (W ; R) may be nonzero for k = 2, 4, and 6. A topological
the map i : Hcpct
analysis, using the fact that c1 (L)3 |W 6= 0, shows that this is so. Consequently, given the
asymptotically conical nature of the CalabiYau metric on W , we expect the space of L2
harmonic forms on W to be three-dimensional, with one class each in dimension 2, 4,
and 6. Given this, there is only one option for how the SU(2) group R that acts on the
cohomology of a Khler manifold (see Appendix A) can act on the L2 harmonic forms
on W : they transform with spin 1. Hence, though there is an L2 harmonic four-form on W ,
it is not primitive, and one cannot turn on a G-field without breaking supersymmetry.

6. Brane perspective
We will conclude this paper by pointing out a reinterpretation of the problem in terms
of singularities of branes. We already explained in Section 2.6 that F-theory on a four-fold
singularity can be reinterpreted as type IIB with a D7-brane that has a worldvolume R4 L,
where L C3 is a singular complex surface. By successive circle compactifications, it
follows that M-theory or type IIA at a four-fold singularity can be described by type IIA
with a singular sixbrane R3 L, or type IIB with a singular fivebrane R2 L.
Analogous phenomena have been noted in the past in the context of N = 2 conformal
theories with NS or M5-branes worldvolumes with singular geometry R4 where
is a Riemann surface which develops a singularity, say of the form y 2 = x n locally (for
n > 2). These give models for studying type II strings at a CalabiYau threefold singularity
capturing ArgyresDouglas points of N = 2 conformal theories [38] and are presently
under study [31].
As we discussed in Section 2.6, it is important that in the cases that we have looked at,
there are no corrections to the classical Rn L geometry. Let us raise the general question
of this sort. Suppose we have a p-brane of some kind, with worldvolume
Rn Xp+1n
embedded in R10 or R11 depending on whether we are dealing with string theory or
M-theory. We assume that this geometry preserves some number of supersymmetries in
Rn . Let us assume X develops a singularity. Is this singular geometry smoothed out in the

102

S. Gukov et al. / Nuclear Physics B 584 (2000) 69108

quantum theory? A necessary condition for that is the existence of instantons which end
on X. So if q-branes can end on this particular p-brane, the condition is the absence of
compact q-cycles in X. So for Dp-branes in type IIA or IIB string theory, since D(p 2)branes and fundamental one-branes can end on them, the condition is the absence of
topologically nontrivial compact one-cycles and p 2-cycles in the geometry of X. For
M5-branes, the condition is the absence of compact two-cycles in X.

Appendix A. Primitive forms


The de Rham cohomology of a compact Khler manifold X, or the space of L2
harmonic forms on any Khler manifold, admits an SU(2) action which is as follows.
(See [39], pp. 122126, for a mathematical introduction.) A diagonal generator J3 of SU(2)
multiplies an (n p)-form by (n p)/2, where n is the complex dimension of X. The
lowering operator J acts by wedge product with the Khler form K:
G K G.

(A.1)

And the raising operator J+ is the adjoint operation of contraction with K:


Gi1 i2 ...in K i1 i2 Gi1 i2 ...in .

(A.2)

Conceptually, this SU(2) action arises as follows. Begin with the supersymmetric
nonlinear sigma model in four dimensions with target space X, and dimensionally reduce
it to 0 + 1 dimensions. This gives a supersymmetric system in which the Hilbert space is
the space of differential forms on X, the four supercharges are are the and operators
and their adjoints, and there is an SU(2) symmetry that comes from rotations of the three
extra dimensions. From this point of view, the SU(2) arises as an R-symmetry group, so
we denote it as R.
Since an (n p)-form has J3 eigenvalue (n p)/2, it clearly transforms under R with
spin at least |n p|/2. For n p > 0, we declare the primitive part of H np (X; R) to
consist of the harmonic forms that transform with spin precisely (n p)/2. For a middledimensional form, with p = n, this definition means that the primitive part of H n (X; R)
consists precisely of the R-invariants.
For a noncompact Khler manifold X, if all of the L2 harmonic forms are in the
middle dimension, then they are all automatically primitive. For a middle-dimensional L2
harmonic form that is not R-invariant can be raised and lowered to make L2 harmonic
forms of other dimensions.
For a middle-dimensional L2 harmonic form G, such as the G-field on a CalabiYau
four-fold, primitiveness is equivalent to either 0 = J G, which is the condition on G given
in [8], or 0 = J+ G = K G.
If there is a projective embedding such that K is the class of a hyperplane section H ,
then K is cohomologous to a form supported on H , and K G = 0 says that G has support
disjoint from H .
An important illustrative case is that of a complex surface W . The middle-dimensional
cohomology of W is two-dimensional and can be decomposed as follows. The space of

S. Gukov et al. / Nuclear Physics B 584 (2000) 69108

103

self-dual forms at a given point is three-dimensional; the self-dual forms are the (2, 0)
and (0, 2) forms and the multiples of the Khler class K. The (2, 0) and (0, 2) forms are
clearly primitive (the lowering operator would map them to (3, 1) and (1, 3)-forms) but
the Khler class K is not (as K K 6= 0). The anti-self dual two-forms are of type (1, 1)
and are of the form = ai j dzi d z j where ai j is traceless. Tracelessness of a means
that is annihilated by the raising and lowering operators and so is primitive. So for a
complex surface, the middle-dimensional primitive cohomology is of type (2, 0) or (0, 2)
and self-dual, or of type (1, 1) and anti-self-dual.
Closer to our needs in this paper is the case of a complex four-fold X. At any point
P X, the holonomy group U (4) acts on the differential forms at P . Since the generator
of the center of U (4) simply multiplies a (p, q) form by p q, we focus on the SU(4)
action. We look first at the (p, p) forms for p = 0, 1, 2, . . . , since they are closed under the
action of R. The (0, 0) forms transform in the trivial representation 1 of SU(4). The (1, 1)
forms ai j dzi dzj transform as 4 4 = 1 15, with 15 the adjoint representation. Since

h d z i d z j transforms as the 6, the (2, 2)-forms


a (2, 0) or (0, 2)-form hij dzi dzj or e
ij

transform as 6 6 = 1 15 20. From this, it follows that (2, 2)-forms that transform as
20 of SU(4) have R = 0, those that transform as 15 have R = 1, and those that transform
as 1 have R = 2.
In particular, the primitive (2, 2)-forms transform in an irreducible representation of
SU(4). From this, it follows that they all transform with the same eigenvalue under the
Hodge operator. To determine the sign, it suffices to consider the case that X = W1 W2 ,
with the Wi complex surfaces, and to consider on X the primitive (2, 2)-form G = 1 2 ,
where for i = 1, 2, i is a primitive (1, 1)-form on Wi . Since the i are anti-self-dual, G is
self-dual.
We can similarly analyze the primitive (3, 1) cohomology. The (2, 0)-forms transform as
6 under SU(4), while the (3, 1)-forms transform as 6 10. (3, 1)-forms that transform as
10 of SU(4) have R = 0 and so are primitive, while those that transform as 6 have R = 1.
Since the primitive (3, 1)-forms transform irreducibly under SU(4), they again all have
the same eigenvalue of . Indeed, by considering the case G = , with a primitive
(1, 1)-form on a surface W1 and a primitive (2, 0)-form on another surface W2 , we learn
that the primitive (3, 1) cohomology of a four-fold is anti-self-dual.
Finally, the (4, 0) cohomology of a four-fold transforms trivially under SU(4) and is
primitive. By setting G = 1 2 with i a (2, 0) form on Wi for i = 1, 2, we learn that
the (4, 0) cohomology on a four-fold is self-dual.
In sum, the Hodge operator acts on the primitive (p, 4 p) cohomology of a four-fold
as (1)p . This type of argument can clearly be generalized to other dimensions.

Appendix B. Four-fold compactifications to AdS3


The purpose of this appendix is to extend the analysis in [8] of supersymmetric
compactification to R3 Y with a G-field to the case of AdS3 Y . In Section 2.4, we

104

S. Gukov et al. / Nuclear Physics B 584 (2000) 69108

argued that the (4, 0) part of G is the superpotential, and accordingly here we will find that
it determines the AdS3 cosmological constant.
We follow the notations of [8] where capital letters M, N , . . . run from 0 to 10 and
denote eleven-dimensional indices; m, n, . . . are real indices tangent to X; and Greek letters
, , . . . , stand for the three-dimensional Lorentzian indices 0, 1, 2. Finally, lower case
letters a, b, . . . , from the beginning of the alphabet denote holomorphic indices tangent
to X.
The bosonic part of the eleven-dimensional effective action, corrected by the -model
anomaly on the five-brane worldvolume, has the following form:

Z
Z 

1
1
1
1
d 11 x gR
G G + C G G (2)4 C I8 . (B.1)
S11 =
2
2
2
6
The eight-form anomaly polynomial can be expressed in terms of the Riemann tensor [40]:



1
1
1
2 2
4
trR
trR
+

.
(B.2)
I8 =
(2)4
768
192
In these units the five-brane tension T6 = 1/(2)3 . The field equation for G that follows
from the action (B.1) looks like:
1
d G = G G + (2)4 I8 .
2
Let us take the following ansatz for the metric:

2
= 1 ds32 (x ) + ds82 (x m )
ds11

(B.3)

(B.4)

and for the three-dimensional components of G:


Gm =  m f (x m ).

(B.5)

This compactification leads to the maximally symmetric three-dimensional spacetime.


Here we introduced the warp factor (x m ) and a scalar function f (x m ), both of which
depend only on the coordinates on X. Below we show that these two functions are related
by the supersymmetry conditions that we are going to formulate in a moment. We also
allow for arbitrary internal components Gmnpq , the form of which will be fixed by the field
equation and supersymmetry conditions. We assume that the other components of G, as
well as the gravitino M , vanish.
Now we examine conditions for the configuration described above to be supersymmetric.
Since M vanishes in the background, we only have to check that the variations of the
gravitino vanish for some Majorana spinor :
1
M M M N N (log )
4

1 3/2
P QRS

(B.6)
M P QRS 8M
GP QRS = 0,

288
where the first two terms come from the covariant derivative in the eleven-dimensional
metric (B.4).

S. Gukov et al. / Nuclear Physics B 584 (2000) 69108

105

Following [8], we make the 11 = 3 + 8 split:


= 9 ,

m = 1 m ,

where the eleven-dimensional gamma-matrices M are hermitian for M = 1, . . . , 10 and


anti-hermitian for M = 0. They satisfy:
{M , N } = 2gMN .

(B.7)

We use the standard notation


M1 ...Mn = [M1 . . . Mn ]

(B.8)

for the antisymmetrized product of gamma-matrices. We decompose the supersymmetry


parameter as:
=  +  ,

(B.9)

where  is an anti-commuting Killing spinor in three dimensions:

(B.10)
 = ,
2
and is a commuting eight-dimensional complex spinor of definite chirality. Without loss
of generality we can take:
9 = .

(B.11)

Here 9 is the eight-dimensional chirality operator that anti-commutes with all the m s and
satisfies 92 = 1. The sign in (B.11) determines whether it is space-filling membranes or
space-filling antimembranes that can be included without breaking supersymmetry. If the
sign is changed, the corresponding supersymmetric vacuum (with the same cosmological
constant ) can be obtained from that with 9 = + by changing the sign of the function
f and the chirality of .
The -component of the supersymmetry condition (B.6) takes the form:

1
n (log ) 9 n
4


1
1 3/2

9 mnpq Gmnpq + 3/2 (m f ) m = 0. (B.12)

288
6
From the decomposition (B.9) and the equations (B.10), (B.11), we obtain the solution
to (B.12):
f = 3/2 ,

1443/2 = G.
/

(B.13)

/ for the total contraction Gmnpq mnpq . In particular, threeHere we have written G
dimensional components of the four-form field strength:
Gm =  m 3/2

(B.14)

have the form similar to the membrane solution with the effective membrane charge
density 12 G2 (2)4 I8 , as follows from the field equation (B.3) for the internal
components. Substituting (B.14) into (B.3), we obtain the additional equation
 1

m 3/2  mnpqrst u Grst u,
(B.15)
m 3/2 Gmnpq =
4!

106

S. Gukov et al. / Nuclear Physics B 584 (2000) 69108

which uniquely determines G given its cohomology class, and is identically obeyed if G
is self-dual (as we will find) and closed.
Now we return to the supersymmetry condition (B.6) and consider its m-component:

1
m + 3/2 npq Gmnpq
2
24
3
1
(B.16)
+ m (log ) n (log )m n + c.c. = 0,
4
8
where we used the explicit form of the solution (B.13) and standard properties of gammamatrices. By means of the rescaling transformations:
m m

gmn ,
gmn gmn = 3/2e

= 1/4e

the equation (B.16) can be written in the compact form:


3/4
1
e
me
npqe
+ 3/4Gmnpq e
+ c.c. = 0.
(B.17)
2
24
Then, following [8], we choose e
to be a covariantly constant spinor of unit norm, and
eme

use it to define an almost complex structure:


,
e
mne
Jemn = ie

(B.18)

which is actually integrable. 17 Hence X is a complex manifold. Furthermore, because Je


is covariantly constant:
ep Jemn = 0

the four-fold X is a Khler manifold with


ga b
Jea b = ie
being a Khler form. Because the metric is of type (1, 1), it is convenient to think of a
and a as creation and annihilation operators that satisfy the algebra:

 a b

, = 2g a b .
{ a , b } = a , b = 0,
a act on the Fock vacuum as annihilation operators:
Namely, e
a and e
ea = 0,

a = 0.
e

(B.19)

To obtain the algebraic constraints on the field G, we multiply the differential


equation (B.17) by a which kills the first term in that equation. To avoid cluttering we
omit tilde from the notation and, finally, obtain:
123/2 a m Gmnpq a npq + c.c. = 0.

(B.20)

Components of this equation with different gamma-matrix structure must vanish separately. For example, if we choose m to be an anti-holomorphic index and use (B.19), we
find that the (1, 3) piece of the field G must be zero:
Ga bcd
= 0.
17 Since the proof is very standard (see, e.g., [8]), we omit it in the present discussion.

S. Gukov et al. / Nuclear Physics B 584 (2000) 69108

107

And G(2,2) , the (2,2) piece of G, must be primitive:


cd

= 0.
Gab
cd
J

(B.21)

Finally, taking the trace over the holomorphic index a in the main supersymmetry
condition (B.20), we get the relation:
963/2 = Gabcd abcd .

(B.22)

This extends the corresponding formula (2.45) in [8] to compactifications with non-zero
cosmological constant, and shows that is proportional to G4,0 , which is natural given
the discussion in Section 2.4.

Acknowledgements
We have benefited from discussions with M. Bershadsky, N. Hitchin, S. Katz,
R. MacPherson, D. Morrison, T. Pantev, N. Seiberg, A. Shapere, and S.-T. Yau.
The work of S.G. was supported in part by grant RFBR No 98-02-16575 and Russian
Presidents grant No 96-15-96939. The work of C.V. was supported in part by NSF grant
PHY-98-02709 and that of E.W. was supported in part by NSF grant PHY-9513835.

References
[1] Y. Kazama, H. Suzuki, New N = 2 superconformal field theories and superstring compactification, Mod. Phys. Lett. A 4 (1989) 325.
[2] A.B. Zamolodchikov, JETP Lett. 46 (1987) 160.
[3] C. Vafa, N.P. Warner, Catastrophes and the classification of conformal theories, Phys. Lett.
B 218 (1989) 51.
[4] E. Martinec, Algebraic geometry and effective Lagrangians, Phys. Lett. B 217 (1989) 431.
[5] E. Witten, On flux quantization in M-theory and the effective action, J. Geom. Phys. 22
(1997) 1.
[6] S. Sethi, C. Vafa, E. Witten, Constraints on low-dimensional string compactifications, Nucl.
Phys. B 480 (1996) 213.
[7] B. Greene, D. Morrison, A. Strominger, Black hole condensation and the unification of string
vacua, Nucl. Phys. B 451 (1995) 109.
[8] K. Becker, M. Becker, M-theory on eight manifolds, Nucl. Phys. B 477 (1996) 155.
[9] G. Tian, S.-T. Yau, Complete Khler manifolds with zero Ricci curvature II, Inv. Math. 106
(1991) 27.
[10] P. Candela, X.C. de la Ossa, Moduli space of CalabiYau manifolds, Nucl. Phys. B 355 (1991)
455.
[11] W. Lerche, FayetIliopoulos potentials from four-folds, JHEP 9711 (1997) 004.
[12] E. Witten, Non-perturbative superpotentials in string theory, Nucl. Phys. B 474 (1996) 343.
[13] J. Bagger, E. Witten, Quantization of Newtons constant in certain supergravity theories, Phys.
Lett. B 115 (1982) 202.
[14] J. Polchinski, A. Strominger, New vacua for type II string theory, Phys. Lett. B 388 (1996) 736.
[15] S. Gukov, to appear.
[16] A. Kehagias, New type IIB vacua and their F-theory interpretation, hep-th/9805131.

108

S. Gukov et al. / Nuclear Physics B 584 (2000) 69108

[17] I. Klebanov, E. Witten, Superconformal field theory on three-branes at a CalabiYau singularity,


Nucl. Phys. B 536 (1998) 199.
[18] D. Morrison, R. Plesser, Nonspherical horizons, hep-th/9810201.
[19] S. Sethi, A relation between N = 8 gauge theories in three dimensions, JHEP 9811 (1998) 003,
hep-th/9809162.
[20] K. Becker, M. Becker, A. Strominger, Five-brane, membranes and nonperturbative string
theory, Nucl. Phys. B 456 (1995) 130.
[21] H. Ooguri, C. Vafa, Summing up D-instantons, Phys. Rev. Lett. 77 (1996) 3296.
[22] N. Seiberg, S. Shenker, Hypermultiplet moduli space and string compactification to threedimensions, Phys. Lett. B 388 (1996) 521.
[23] S. Cecotti, C. Vafa, On classification of N = 2 supersymmetric theories, Commun. Math.
Phys. 158 (1993) 569.
[24] S. Cecotti, P. Fendley, K. Intriligator, C. Vafa, A new supersymmetric index, Nucl. Phys. 386
(1992) 405.
[25] O. Aharony, A. Hanany, K. Intriligator, N. Seiberg, M.J. Strassler, Aspects of N = 2
supersymmetric gauge theories in three-dimensions, Nucl. Phys. B 499 (1997) 67.
[26] D. Kutasov, A comment on duality in N = 1 supersymmetric non-abelian gauge theories, Phys.
Lett. B 351 (1995) 230.
[27] A. Klemm, W. Lerche, P. Mayr, C. Vafa, N. Warner, Self-dual strings and N = 2 supersymmetric
field theory, Nucl. Phys. B 477 (1996) 746.
[28] P. Fendley, S.D. Mathur, C. Vafa, N.P. Warner, Integrable deformations and scattering matrices
for the N = 2 supersymmetric discrete series, Phys. Lett. 243 (1990) 257.
[29] J.M. Rabin, Phys. Lett. B 411 (1997) 274.
[30] J. Schulze, N.P. Warner, BPS geodesics in N = 2 supersymmetric YangMills theory, Nucl.
Phys. B 498 (1997) 101.
[31] A. Shapere, C.Vafa, work in progress.
[32] W. Lerche, C. Vafa, N.P. Warner, Chiral rings in N = 2 superconformal theories, Nucl.
Phys. B 324 (1989) 427.
[33] K. Intriligator, Fusion residues, Mod. Phys. Lett. A 6 (1991) 3543.
[34] W. Lerche, N.P. Warner, Polytopes and solitons in integrable N = 2 supersymmetric Landau
Ginzburge theories, Nucl. Phys. B 358 (1991) 571.
[35] A. Kapustin, M. Strassler, On mirror symmetry in three-dimensional Abelian gauge theories,
JHEP 9904 (1999) 021, hep-th/9902033.
[36] P. Candelas, X.C. De la Ossa, P.S. Green, L. Parkes, A pair of CalabiYau manifolds as an
exactly soluble superconformal theory, Nucl. Phys. B 359 (1991) 21.
[37] P. Aspinwall, D. Morrison, Topological field theory and rational curves, Commun. Math.
Phys. 151 (1993) 245.
[38] P.C. Argyres, M.R. Douglas, New phenomena in SU(3) supersymmetric gauge theory, Nucl.
Phys. B 448 (1995) 93.
[39] P. Griffiths, J. Harris, Principles of Algebraic Geometry, Wiley Interscience, 1978.
[40] L. Alvarez-Gaum, E. Witten, Nucl. Phys. B 234 (1983) 269.

Nuclear Physics B 584 (2000) 109148


www.elsevier.nl/locate/npe

Two two-dimensional supergravity theories from


CalabiYau four-folds
S. James Gates Jr. a,1 , Sergei Gukov b,c, , Edward Witten b,d
a Department of Physics, University of Maryland at College Park, College Park, MD 20742-4111, USA
b California Institute of Technology, CIT-USC Center For Theoretical Physics, Pasadena, CA 91125, USA
c Joseph Henry Laboratories, Princeton University, Princeton, NJ 08544, USA
d School of Natural Sciences, Institute for Advanced Study, Olden Lane, Princeton, NJ 08540, USA

Received 25 May 2000; accepted 13 June 2000

Abstract
We consider two-dimensional supergravity theories with four supercharges constructed from
compactification of Type II string theory on a generic CalabiYau four-fold. In type IIA and type IIB
cases, respectively, new superspace formulations of N = (2, 2) and N = (0, 4) dilaton supergravities
are found and their coupling to matter multiplets is discussed. 2000 Elsevier Science B.V. All rights
reserved.

I found a way to make it work.


Stanislaw Ulam

1. Introduction
For a long time, compactification of heterotic string theory on CalabiYau manifolds
was the primary candidate for constructing realistic models in four dimensions with N = 1
supersymmetry. This was also a strong motivation to study type II superstrings on Calabi
Yau three-folds which share many common properties with the corresponding heterotic
compactifications. At the same time substantial progress has been made in understanding
the mathematical aspects of CalabiYau three-folds, such as quantum cohomology and
mirror symmetry [1,2]. It was not until the discovery of F-theory [3] that it was realized that
N = 1 four-dimensional heterotic string vacua can be equivalently described as F-theory
compactifications on elliptically fibered CalabiYau four-folds. Since then, the study of
CalabiYau four-fold compactifications has become of particular importance for physical
Corresponding author. E-mail: gukov@feynman.princeton.edu
1 gatess@wam.umd.edu

0550-3213/00/$ see front matter 2000 Elsevier Science B.V. All rights reserved.
PII: S 0 5 5 0 - 3 2 1 3 ( 0 0 ) 0 0 3 7 4 - 6

110

S.J. Gates Jr. et al. / Nuclear Physics B 584 (2000) 109148

applications. Compactification of F-theory on an elliptically fibered CalabiYau fourfold is closely related to the corresponding compactifications of type IIA and M-theory.
Namely, when the area of the elliptic fiber shrinks to zero, M-theory compactification
on a CalabiYau four-fold is well described by F-theory compactification on the same
CalabiYau manifold. On the other hand, type IIA string theory is related to M-theory via
compactification on an extra circle (of small radius).
One of the most striking outcomes in the study of compactifications on CalabiYau
three-folds is the great success in understanding non-perturbative phenomena in N = 2
field theories in four dimensions, see, e.g., [4] for introduction and references. On the
other hand, understanding of CalabiYau four-fold compactifications still is quite far from
that stage, so we shall not discuss non-perturbative phenomena in this paper. Instead,
we consider classical supergravity theories interacting with two-dimensional non-linear
sigma-models that can be constructed from CalabiYau four-folds. Surprisingly, it turns
out that manifestly supersymmetric formulations of such theories has not been given
previously.
On general grounds, a compactification of type IIA (IIB) string theory on a Calabi
Yau four-fold leads to a N = (2, 2) (respectively, N = (0, 4)) effective field theory in
two dimensions. In the low-energy limit the theory is described by supergravity coupled
to matter. For example, from the KaluzaKlein reduction of type IIA string theory on a
CalabiYau four-fold in Section 3 we find that a suitable low-energy theory is N = (2, 2)
dilaton supergravity interacting with some number of chiral and twisted chiral multiplets. It
is invariant under the mirror transformation which, acting on the matter fields, exchanges
chiral multiplets and twisted chiral multiplets. We thus generalize the proposal of [5] where
it was suggested that the kinematic structure of the mirror transformation has its origin in
a mapping between chiral and twisted chiral multiplets when these superfields are regarded
as the fundamental degrees of N = (2, 2) superstring theories. The generalization posits
that this mapping also applies to the effective action. Another characteristic feature of
this supergravity theory is that the supergravity multiplet contains a real dilaton field. In
a special case, when all matter multiplets are chiral and massless, a component action of
this N = 2 dilaton supergravity was constructed in [6]. However, to describe type IIA
compactifications on CalabiYau four-folds we need a generalization of this theory that
includes interaction with twisted chiral multiplets and the possibility to turn on the
superpotential and as well the twisted superpotential. Thus, in Sections 4 and 5 we present
superspace construction of general N = (2, 2) dilaton supergravity coupled to matter.
The construction in Section 4 is based on the Goldstone mechanism in the superspace
formulation of non-minimal gauged N = (2, 2) supergravity. Coupling of the new N =
(2, 2) dilaton supergravity to matter multiplets is the subject of Section 5, where we discuss
local integration in superspace. In Section 6 we perform the KaluzaKlein reduction of
type IIB string theory on a CalabiYau four-fold and describe the component action of
the resulting N = (0, 4) dilaton supergravity. Most of this section, as well as Section 3,
is not new and presented for the sake of completeness. The superspace formulation of
the new N = (0, 4) dilaton supergravity is presented in Section 7. In the Appendix A
we present a straightforward but technical worldsheet calculation of string amplitudes

S.J. Gates Jr. et al. / Nuclear Physics B 584 (2000) 109148

111

corresponding to the target space metric of the effective two-dimensional theory, and in
the Appendix B we list extra derivative constraints arising from de-gauging N = (2, 2)
non-minimal supergravity. Appendix C contains components of the covariant derivative in
N = (2, 2) dilaton supergravity needed in Section 5. Finally, in Appendix D we repeat the
derivation [7] of the chiral density projection formula in N = (2, 2) dilaton supergravity.
We begin in the next section with a summary of notations and definitions used throughout
the paper.

2. CalabiYau four-folds: some conventions and definitions


We study compactification of type II string theory on M(2) X where M(2) is a
maximally symmetric homogeneous two-dimensional spacetime and X is a Calabi
Yau four-fold. We use the following notations for the spacetime indices. Capital letters
M, N, . . . , run from 0 to 9 and denote ten-dimensional Lorentz indices. Latin letters
m, n, . . . , and a, b, . . . , represent, respectively, real and holomorphic indices tangent to
X. Greek letters , , . . . , and , , . . . , are used for the two-dimensional spinor indices
+ and , and light-cone indices and , correspondingly. Sometimes we also use
capital latin letters A, B, . . . , to denote both spinor and vector indices.
A CalabiYau space X is a compact Khler manifold with complex dimension four
and SU(4) holonomy group. It follows that X is a Ricci-flat manifold and, therefore, it
can be used as a background for type II string compactification. As a topological space,
X is classified by the Hodge numbers hp,q which count the number of harmonic (p, q)(p,q)
H p,q (X), i = 1, . . . , hp,q . The non-vanishing cohomology groups have the
forms i
following dimensions [8]:
h1,1 = h3,3 ,

h3,1 = h1,3 ,

h2,1 = h1,2 = h3,2 = h2,3 ,


h0,0 = h4,4 = h4,0 = h0,4 = 1,
h2,2 = 2(22 + 2h1,1 + 2h3,1 h2,1 ).

(2.1)

For the Euler number of X we have:

= 8 + h1,1 + h3,1 h2,1 .


(2.2)
6
We denote by a covariantly constant (4, 0)-form. The (1, 1)- and (3, 1)-forms are
related to the deformation parameters of the Khler form and the complex structure of X,
respectively. Namely, an arbitrary variation of the metric of the CalabiYau four-fold X
that respects SU(4) holonomy looks like:

ga b dza d z b + gab dza dzb + c.c.,


where
3,1

gab =

h
X
j =1

1,1

j wab ,

iga b =

h
X
i=1

s i ai b .

(2.3)

112

S.J. Gates Jr. et al. / Nuclear Physics B 584 (2000) 109148

from the forms w we can construct elements in


By the appropriate contraction with ,
ab
H 1,3(X):
j

(1,3)

= a b cd g dd wdf d z a d z b d z c dzf .
j

In what follows we will use some integrals over the CalabiYau space X [9,10]:
Z
Z
1

K K K K,
V = d 8z g =
4!
X
X
Z
1
8
d z g wi ab w jab
Gi =
,
j
4V
X
Z
1

d 8 z g ak b l a b ,
Gk l =
2V
X
Z
(1,1)
(1,2)
(2,1)
m
n ,
Yimn = i

(2.4)

(2.5)

(2.6)

dij kl =

(1,1)

(1,1)

(1,1)

(1,1)

(2.7)

The notations (2.4) and (2.5) will become clear in the next section where we identify
expectation values of the fields si and j with the Khler and complex structure moduli,
respectively. In particular, we write:
1,1

K = iga b dz d z =
a

h
X

(1,1)

hsi i i

(2.8)

i=1

for the Khler form on X. The moduli space of a CalabiYau space is locally a product of
the moduli space of complex deformations, Mc (X), and the (complexified) moduli space
of Khler structure, MK (X). Notice, the metric Gi defined above is the WeilPetersson
j
metric on the moduli space of complex structure of X, with the Khler potential, cf. [9]:
!
Z
.
(2.9)
K(i , i ) = ln
X

3. Compactification of type IIA string theory on CalabiYau four-folds


In this section we describe the effective two-dimensional theory constructed from
compactification of type IIA string theory on a CalabiYau four-fold X. When the volume
of X is large compared to the string scale, type IIA supergravity is a good low-energy
approximation to type IIA string theory. Therefore, in the large volume limit we may
describe the low-energy effective theory studying compactification of type IIA supergravity
on the CalabiYau space X. With this motivation, let us start this section recalling some
facts about type IIA supergravity itself.

S.J. Gates Jr. et al. / Nuclear Physics B 584 (2000) 109148

113

The bosonic field content of type IIA supergravity contains the metric gMN , the
dilaton , a vector field AM , and tensor fields BMN and CMNP . The bosonic part of the
Lagrangian (in string frame) looks like:





1
1
1
1
L(10) = g e2 R (10) + 4()2 H 2 F 2 G0 2 + , (3.1)
2
12
4
48
where we introduced the gauge-invariant field strengths:
F = dA,

H = dB,

G = dC,

G0 = G + A H.

With this choice of normalization the fields A, B, and C transform in a natural, dilatonindependent way under gauge transformations.
The dots in the Lagrangian (3.1) stand for higher order terms among which we find the
ChernSimons term B G G and the anomaly term B I8 , where the eight-form I8 is
proportional to the Euler density of X. After integration over a compact eight-manifold X
these topological terms produce a global anomaly [11,12]:
Z
1

G G.
(3.2)
N=
24 2(2)2
X

To cancel the tadpole for the B-field one has to introduce N fundamental strings filling
two-dimensional non-compact space.
The action of type IIA supergravity is invariant under 16 left and 16 right supersymmetry transformations, such that the left supersymmetries are chiral while the right supersymmetries are anti-chiral with respect to the ten-dimensional chirality operator 11 .
Since X admits a nowhere vanishing complex spinor of definite chirality, compactification
of type IIA string theory on X is described by N = (2, 2) supergravity theory coupled
to matter. With the appropriate choice of orientation, the fundamental strings filling twodimensional spacetime do not break supersymmetry further.
To find the spectrum of the effective low-energy theory we perform KaluzaKlein reduction of type IIA supergravity to two dimensions. Below we describe the decomposition of type IIA bosonic fields in harmonics of X. By supersymmetry, incorporation of
fermionic zero-modes completes the resulting spectrum into appropriate N = (2, 2) supermultiplets. Via dimensional reduction type IIA dilaton becomes a real scalar field in
the two-dimensional theory. The ten-dimensional metric gMN decomposes into the twodimensional metric g , h3,1 complex scalars i and h1,1 real scalars sj defined in (2.3).
The antisymmetric tensor fields BMN and CMNP can be expanded into harmonic modes
as follows:
1,1

B=

h
X

r i i(1,1),

(3.3)

i=1
1,1

C=

h
X
j =1

2,1

(1,1)

Aj j

h
X
k=1

(2,1)

zk k

+ c.c.

(3.4)

114

S.J. Gates Jr. et al. / Nuclear Physics B 584 (2000) 109148

It is convenient to combine real fields s i and r i into complex scalars i . Taking into account
the vector field A from the RamondRamond sector of type IIA theory, we end up with
the following list of N = (2, 2) supermultiplets:
a gravitational multiplet :
h3,1 chiral multiplets :
1,1
h twisted chiral multiplets :
2,1
h (twisted) chiral multiplets :

g ,
i ,
j,
zk ,

A ,
,
i

j ,

z k .

,
Aj ,

Vector fields A and Ai do not have propagating degrees of freedom in two dimensions
and play the role of auxiliary fields in the supergravity multiplet and twisted chiral
multiplets, respectively.
The complex scalar fields zk which come from (2, 1)-modes take value in a torus. When
(2,1)
(2,1)
= 0 and H k
= 0 there is no superpotential for
background fluxes satisfy G k
the corresponding harmonics zk , so these fields are massless. A T-duality transformation on
the torus then converts them from ordinary chiral superfields to twisted chiral superfields.
It is natural to choose zk to be scalar components of chiral superfields. Indeed, if we
start in eleven dimensions, the reduction of the C-field (3.4) yields h2,1 complex scalar
modes in three dimensions. These modes are scalar components of chiral superfields since
there is no notion of twisted chiral superfields in three dimensions. After a further
compactification on a circle they naturally remain as chiral superfields in two dimensions,
but now a T -duality becomes possible and zk can be alternatively described if one prefers
as twisted chiral superfields.
To find the effective action for the light fields we have to substitute (2.3), (3.3) and (3.4)
in the Lagrangian (3.1) and integrate over the internal space X. Using the formulas (2.4),
(2.5) and (2.6) we obtain the following effective action for the bosonic modes, cf. [6,10]:



L(2) = e2 V R (2) + 4()2 Gi i j
j


 1


1
(3.5)
Gi j i j Yimn i D zm D z n + ,
2
4
where the covariant derivative D acting on zm contains a connection corresponding to the
holomorphic dependence of the basis of (2, 1)-forms on the complex structure [10].
To summarize, we find that in the large volume limit compactification of type IIA string
theory on a CalabiYau four-fold X leads to N = (2, 2) dilaton supergravity coupled to
a non-linear sigma-model. The target space of this sigma-model is parametrized by some
number of chiral and twisted chiral multiplets. This agrees with the result of [13], where
it was found that the most general N = (2, 2) non-linear sigma-model is based on a target
space with two non-commuting complex structures J , so that the space ker(J+ J ) is
parametrized by chiral superfields, while ker(J+ + J ) is parametrized by twisted chiral
superfields.
This kind of dilaton supergravity coupled to N = (2, 2) chiral matter was studied
some time ago [6]. However, for our purposes we need to generalize the component
construction of [6] to include twisted chiral multiplets. Furthermore, background fluxes
of RamondRamond field strengths induce effective superpotential [14,15] and/or twisted

S.J. Gates Jr. et al. / Nuclear Physics B 584 (2000) 109148

115

chiral superpotential [1416] in the two-dimensional theory. Hence, we have to incorporate


these terms in the construction as well. The most elegant and convenient way to do
this is in N = (2, 2) superspace where the supersymmetry becomes manifest [17,18]. In
addition to the usual spacetime coordinates x , N = (2, 2) superspace is parametrized by
.
anti-commuting coordinates = ( + , ) and their complex conjugates = ( u , ).
Then, we expect that the action of the matter fields (3.5) can be written in a compact form,
similar to the action of matter coupled N = 1 supergravity in four dimensions:
Z
Z
2
(3.6)
S = d x d 2 d 2 E 1 exp(K).
We postpone the discussion of the superspace measure E till the next sections where
superspace formulation will be discussed in detail. Now we simply assume that the suitable
measure exists. The main advantage of the superspace formulation is that due to the
extended supersymmetry, all the term in the action (3.5) with up to two derivatives or four
fermions are determined by a single real function K(i , i , j , j , zk , z k , ), the Khler
potential [19]. It is invariant under the generalized Khler transformation:
K K + 1 (i , j , zk ) + 1 ( i , j , z k ) + 2 (i , j , z k ) + 2 ( i , j , zk ).
(3.7)
The target space metric is given by the second derivative of the Khler potential.
For the sake of simplicity, let us assume for a moment that h2,1 = 0. Then the metric is
block diagonal:
Gi j =

2K
= 0.
i j

(3.8)

From the condition (3.8) it follows that locally we can write the Khler potential that gives
the effective action (3.5) as:
K = Kc (i , i ) + KK (j , j ),

(3.9)

where Kc is the Khler potential (2.9) on the moduli space of the complex structure:
!
Z
.
Kc (i , i ) = ln
X

Similar to the case of CalabiYau three-folds [9], one can verify that the metric Gi j can
be obtained from the Khler potential:
!
Z
KKKK .
(3.10)
KK (j , j ) = ln
X

(2)

is a harmonic 2-form on a CalabiYau four-fold X, its Hodge dual is given


Indeed, if
by the following neat formula:
R
1 (2)
2 ( X (2) K K K)
(2)
R
K K K.
(3.11)
= KK+
2
3 ( X K K K K)
Therefore, we can write (2.5) in the following form:

116

S.J. Gates Jr. et al. / Nuclear Physics B 584 (2000) 109148

1
(1,1)
(1,1)
k l
2V
X
Z
1
=
k(1,1) l(1,1) K K
4V

Gk l =

72V 2

k(1,1) K K K

! Z

!
l(1,1) K K K

Using the explicit expression (2.8) for the Khler form K, it is easy to see that the above
metric indeed follows from the Khler potential (3.10):
Gk l =

1 2 KK (j , j )
.
2
k l

Hence, to the leading order the metric on the target space is Khler, torsionless, and equal
to the metric on the moduli space of the CalabiYau space X, Mc (X) MK (X).
The classical action (3.5) is invariant under two U (1) R-symmetries. We will denote
their linear combinations as U (1)A and U (1)V . The action of these symmetries on the
supercharges can be represented as:
Q
Q+

Q u ,
Q . ,

(3.12)

where the upper (lower) row is assigned a U (1)A charge +1 (1) while the right (left)
column is assigned a U (1)V charge +1 (1). These R-symmetries are not symmetries of
the string theory there is no way to assign R-transformations to massive string modes
to preserve them. Even though we will not explicitly include massive string modes in
the present paper, we will include a superpotential and twisted chiral superpotential that
violate the R-symmetries. Even in the absence of the superpotentials, higher derivative
interactions among the massless fields obtained by integrating out massive string states
would be expected to violate the R-symmetry.
The explicit expression for the chiral superpotentials generated by the most general
P
RamondRamond flux F = (RR fields) in terms of the CalabiYau moduli was derived
in [15]:
Z
1
G,
(3.13)
W (i ) =
2
X

and for the twisted chiral superpotential:


Z
e (j ) = 1
eK F .
W
2

(3.14)

e (j ) are holomorphic functions


The superpotential W (i ) and the twisted superpotential W
of the fields i and j , respectively. Taking into account the superpotential terms, the action
of the matter fields reads as:

S.J. Gates Jr. et al. / Nuclear Physics B 584 (2000) 109148

S(2) =

117

d 2 x d 2 d 2 E 1 eK + d 2 x d 2 E 1 W (i )
Z
Z
.
e (j ) + c.c.
+ d 2 x d + d Ee1 W

(3.15)

Generic values of RamondRamond fluxes completely break the N = (2, 2) supersymmetry. 2 In the two-dimensional theory this effect corresponds to generation of a superpotential that lifts (part of) supersymmetric vacua. However, if the vacuum values of the
fields i and j satisfy:
DW
=0
Di

and

e
DW
= 0,
Di

(3.16)

then type IIA compactification on the corresponding CalabiYau manifold is supersymmetric [14,15]. From the formulas (2.2), (3.2) and the quantization condition of the G-flux
[12] it follows that there is a finite number of choices for [G] H 4 (X) corresponding to
supersymmetric vacua. In particular, if h2,1 > 8 + h1,1 + h3,1 , then there are no such vacua
at all.
From the superspace construction in Section 5 it follows that in the equations (3.16) we
should use the appropriate covariant derivatives:
Kc (i , i )
W
DW
=
+
W,
Di
i
i

e W
e KK (j , j )
DW
e,
=
+
W
Di
i
i

(3.17)

where Kc and KK are given by the tree-level formulas (2.9) and (3.10), respectively.
A simple way to see that one has to use the covariant derivatives instead of ordinary ones
is to consider first compactification of F-theory on the same CalabiYau space 3 X. In the
component action of the effective N = 1 four-dimensional theory there is a scalar potential:



(3.18)
eK G i j (Di W ) D W 3|W |2 ,
j

where the covariant derivative Di W = DW/Di is defined in (3.17). After a further


compactification on a torus T 2 , this theory is dual to compactification of type IIA string
theory on X. It is clear that after the dimensional reduction of the four-dimensional
component action the covariant derivatives (3.17) also appear in the component action
of the two-dimensional theory in question.
These models can have a variety of T-duality symmetries. Of particular interest are
mirror symmetries [21,22]. A mirror symmetry is, of course, a symmetry that maps
2 Investigating the supersymmetry conditions as in [14,15,20], one can also show that any H -field flux breaks
all the supersymmetry. A simple way to see this is to assume, on the contrary, that there exists a supersymmetric
vacuum corresponding to a non-zero H -flux and consider a BPS soliton connecting such a vacuum to the vacuum
with zero H -flux. In type IIA string theory this soliton would correspond to an N S5-brane wrapped over a
Poincar dual supersymmetric 5-cycle. However, there is a contradiction since CalabiYau 4-folds do not have
supersymmetric 5-cycles, see, e.g., [15]. Therefore, a non-zero H -flux
R lifts all the supersymmetric vacua. It is
natural to interpret this in terms of the effective superpotential W C H for the scalar fields zk .
3 Of course, here we assume that X is elliptically fibered. The result, however, is independent of this
assumption.

118

S.J. Gates Jr. et al. / Nuclear Physics B 584 (2000) 109148

type IIA string theory on a four-fold X to type IIA string theory on the mirror variety
e such that:
X,
e
hp,q (X) = h4p,q (X)

(3.19)

e are equivalent. This operation


and the conformal field theories associated with X and X
corresponds [5] to a transformation which exchanges chiral multiplets and twisted chiral
multiplets. It can be interpreted in terms of the supergeometrical coordinate transformation
.
that also exchanges chiral multiplets and twisted chiral multiplets and changes
the superspace measure in a way consistent with other definitions of mirror symmetry. In
particular, the latter implies that under the mirror symmetry we have i j which is
consistent with our interpretation of (vevs of) these fields as the Khler and the complex
structure moduli of the CalabiYau space X. Therefore, the mirror symmetry relates
different quantum N = (2, 2) theories also interchanging:
U (1)A U (1)V ,
i j ,
e (j ).
W (i ) W

(3.20)

The mirror map has no effect on the N = (2, 2) dilaton supergravity itself, so that
in the absence of matter fields it must be mirror-symmetric. It may seem that twisted
chiral fields zk violate the invariance under (3.20). Recall that via a spacetime T-duality
transformation, those fields can be described by either chiral or twisted chiral superfields.
Hence mirror symmetry just exchanges these two descriptions.
By definition, the low-energy effective action (3.15) describes dynamics of the light
modes in type IIA string theory on X in the large volume limit. In other words, tree-level
amplitudes in type IIA string theory must agree with the corresponding amplitudes in the
effective two-dimensional theory. In the Appendix A we illustrate this by a worldsheet
calculation which independently proves that the target space metric is block-diagonal, cf.
(3.8).
A superspace formulation of N = (2, 2) dilaton supergravity that includes chiral and
twisted chiral multiplets on equal footing does not seem to exist in the literature. Although
a superspace model of N = (2, 2) supergravity where the dilaton is a complex field was
constructed in [24], we are interested in a theory where the supergravity multiplet contains
a real dilaton field. A superspace formulation of such a supergravity theory is presented in
the next section.
4. Superspace formulation of N = (2, 2) dilaton supergravity
In this section we present a superspace construction of N = (2, 2) dilaton supergravity
without gauged symmetry. This last property is a distinguishing feature of the new
formulation since all the known N = (2, 2) gravity theories have at least one gauged U (1)
R-symmetry (see [23] for a general presentation). Theories where the entire U (1)A
U (1)V symmetry group is gauged are called non-minimal (or reducible), as opposed to

S.J. Gates Jr. et al. / Nuclear Physics B 584 (2000) 109148

119

minimal theories where only U (1)A or U (1)V factor is gauged. It is very well known
how to obtain one supergravity theory with a smaller holonomy group from a supergravity
theory with a larger holonomy group. This process has been for a long time called degauging (see [17], Section 5.3.b.7).
The basic idea of de-gauging is to break the gauge symmetry introducing extra matter
field in a Goldstone-like mechanism. For example, consider an abelian vector multiplet in
a four-dimensional N = 1 gauge theory. It contains a U (1) gauge vector field, gaugino
and an auxiliary field. All the other fields can be set to zero by a supersymmetric choice
of gauge, the so-called WessZumino gauge. On the other hand, a massive gauge multiplet
contains some extra component fields which could be eliminated in the massless multiplet.
The reason is that the massive system no longer possess the U (1) gauge invariance. This
toy model teaches us that when a symmetry is broken in superspace, extra component
fields not present in the gauge symmetric phase begin to appear. In other words, Goldstone
supermultiplets must appear. And their component fields come from that part of the vector
multiplet that was ignored in the symmetric phase. Following these steps, we construct
N = (2, 2) dilaton supergravity via de-gauging U (1)A U (1)V non-minimal gauged
supergravity. An advantage of this approach is that both the original and the resulting
theories are manifestly invariant under the mirror symmetry (3.20). We also find that the
new N = (2, 2) supergravity multiplet contains a real dilaton field , in accordance with
the results of Section 3 where we studied compactification of type IIA string theory on
CalabiYau four-folds. We hope that apart from this obvious application there may also be
many other aspects of the new supergravity to explore. For example, it would be interesting
to study black hole solutions in this dilaton supergravity, cf. [25].
The relation between different supergravity theories can be schematically represented in
the form of the following diagram:
U (1)A U (1)V
.
&

U (1)V ,
(4.1)
U (1)A
&
.
1
where the theory with the trivial holonomy group is the new N = (2, 2) dilaton supergravity we are going to construct. Notice, however, that various arrows in this diagram have
different meaning. For example, minimal theories with either of the U (1) R-symmetries
gauged can be obtained by truncation of the non-minimal U (1)A U (1)V gauged supergravity [23]. On the other hand, the vertical arrow corresponds to de-gauging U (1)A
U (1)V symmetry, so that the total number of degrees of freedom increases. More precisely, it has to be a combination of consistent truncation of the non-minimal N = (2, 2)
supergravity to a minimal one plus a de-gauging of the latter. To see this, let us count the
number of real Goldstone scalars. Since the broken U (1)A U (1)V phase of the theory
has exactly the same field content as the gauge symmetric phase plus the field content
of the Goldstone multiplets minus the parts that go into the longitudinal components of
the U (1)A U (1)V gauge fields, we find that in total Goldstone multiplets should have
three real scalars. However, there are no mirror-symmetric N = (2, 2) multiplets with such

120

S.J. Gates Jr. et al. / Nuclear Physics B 584 (2000) 109148

a field content. Therefore, we conclude that the vertical arrow should be a more economical
de-gauging. Indeed, if we were following another route via a minimal gauged supergravity,
at the first step we would have to make a consistent truncation that would eliminate one of
the gauge fields. In order to de-gauge the resulting minimal gauged supergravity we would
have to introduce extra chiral (or twisted chiral) Goldstone superfield. In any case, one real
scalar from this multiplet would become the longitudinal component of the gauge vector
field, and the other would become a dilaton, in agreement with what we expect. Assuming
that the latter route (which is not, unfortunately, mirror-symmetric) is equivalent to the vertical arrow on the above diagram, we expect that there is an economical de-gauging of the
non-minimal N = (2, 2) supergravity that leads to only one massless scalar (the dilaton).
To start the construction, let us arrange component fields found in the previous section
into superfields. First we give the definitions in flat superspace and then extend them
to curved superspace. Left-right symmetric N = 2 superspace is parametrized by the
bosonic coordinates x = (x , x ), two anti-commuting complex spinor coordinates =
.
( + , ) and their complex conjugates = ( u , ). The spinor derivatives D+ , D ,
Du and D. satisfy {D+ , Du } = and {D , D. } = , with all other (anti-)commutators
vanishing.
Irreducible matter superfields are defined by imposing some constraints on general
complex superfields. The simplest constraints are linear in derivatives and look like:
Du = D. = D+ = D = 0

(4.2)

for a chiral superfield , and:


Du = D = D+ = D. = 0

(4.3)

for a twisted chiral superfield .


In what follows we promote the complex scalar fields i and j defined in the
previous section to the chiral and twisted chiral superfields i and j , respectively.
Similarly, we will regard the compact fields zk as the scalar components of (twisted) chiral
superfields Zk .
Components of the chiral superfield i can be obtained using the projection method:
i | = i ,
i | = i ,
i
,
D+ i | = +

i
Du i | = u
,
i.
.

D i | = ,

i
,
D i | =
i
i
[D+ , D ]i | = Ai ,
[Du , D. ] i | = A i ,
(4.4)
2
2
where, for example, i | denotes the leading component of the superfield i , with all the coordinates put to zero. Similarly, we find the components of a twisted chiral multiplet j :

j | = j ,

j | = j ,

j
D+ j | = + ,
j
D. j | = . ,

j
Du j | = u ,

i
[D+ , D. ]j | = Bj ,
2

D j | = ,
i
[Du , D ] j | = B j .
2
j

(4.5)

S.J. Gates Jr. et al. / Nuclear Physics B 584 (2000) 109148

121

To define chiral and twisted chiral superfields in curved N = (2, 2) superspace, the spinor
derivatives D must be appropriately replaced by the covariant derivatives .
In our construction of N = (2, 2) dilaton supergravity we start with non-minimal gauged
supergravity and then de-gauge U (1)A U (1)V symmetry. If we introduce superfields A
and A0 representing U (1)V and U (1)A gauge connections, and denote by the Lorentz
spin-connection, then the covariant derivative in this theory has the form:
= E B DB + X + A Y + A0 Y 0 ,

(4.6)

where E B is the supervielbein.


The Lorentz generators, X , U (1)V symmetry generators, Y, and U (1)A symmetry
generators, Y 0 , act on the covariant derivative in the following way [23]:
1
1
[X , . ] = . ,
[X , ] = ,
2
2
i
i .
.
[Y, ] = + ,
[Y, ] = ,
2
2
i
i
0
0
[Y , . ] = . .
(4.7)
[Y , ] = ,
2
2
The constraints which define non-minimal N = 2 U (1)A U (1)V supergravity are
given by:
{+ , + } = 0,

{ , } = 0,

{ , . } = i ,
{+ , u } = i ,
1
1

(4.8)
{+ , . } = F (X i Y),
{+ , } = R(X i Y 0 ),
2
2
where the chiral superfield R and the twisted chiral superfield F are related to the twodimensional curvature R (2) and the abelian field strengths of the graviphoton gauge fields.
The easiest way to see this is to compute the commutator [23,26]:


1
1
1
2
.
( R) R R + (+ F ) F F X
[ , ] =
2
2
2
i
i
(4.9)
+ (+ . F )Y + ( 2 R)Y 0 + + c.c.,
2
2
where the dots stand for the covariant derivative terms like ( F )u , etc. If we set F = 0
in the constraints (4.8), we obtain U (1)A minimal gauged supergravity theory. On the
other hand, if we set R = 0, we end up with U (1)V minimal theory. We denote the leading
components of the superfields F and R as follows:
F | = G,

R| = H.

(4.10)

In order to obtain N = (2, 2) supergravity theory without gauged symmetry, we replace


b which includes only derivatives and Lorentz generator,
the covariant derivative by
so that:
b + A Y + A0 Y 0
=

(4.11)

122

S.J. Gates Jr. et al. / Nuclear Physics B 584 (2000) 109148

b contains neither
and the remaining derivatives are real. The new covariant derivative
the U (1)V gauge connection nor the U (1)A gauge connection, so it can describe twodimensional N = (2, 2) supergravity without gauged symmetry.
Substituting (4.11) into (4.8), and using (4.7), we derive the form of the commutator
b operators:
algebra for the
b+ } = i(+ + e
b+ ,
b ,
b } = i( e
b ,
b+ ,
+ )
{
)
{
b } = 1 RX
b+ + i (+ e
b ,
+ i ( + e
b+ ,
)
+ )
{
2
2
2
i
1
i
b. } = F X + (. + e
b+ (+ e
b. ,
b+ ,
. )
+ )
{
2
2
2
i
i
bu } = i
b + (u + e
b+ (+ + e
bu ,
b+ ,
u )
+ )
{
2
2
i
i
b. } = i
b + (. e
b ( e
b. ,
b ,
. )
)
(4.12)
{
2
2
have appeared. They are components of the non-minimal
where the new fields and e
gauged supergravity multiplet that could be eliminated by a gauge transformation in the
U (1)A U (1)V gauge-symmetric phase. In the theory we are constructing the U (1)A
are dynamical. They have
U (1)V gauge symmetry is broken, so that the fields and e
to be identified with the spinorial derivatives of a new matter Goldstone multiplet. To the
can also be identified with the dilatino field of the
leading order in , the fields and e
new supergravity multiplet.
If we introduce four linear independent spinors (along with their conjugates):
+ ,
+ + + e
+ ,
e
+ + e

e
,
e
+ e

then the conditions (4.12) which define N = (2, 2) dilaton supergravity can be written in
the following simple form:
b+ } = i+
b ,
b } = i
b+ ,
b ,
b+ ,
{
{
1
i
i
b } = RX
+ e
b+ + e
b ,
b+ ,

+
{
2
2
2
b. } = 1 F X + i e
b+ i e
b. ,
b+ ,
.
+
{
2
2
2
bu } = i
b + i u
b+ i +
bu ,
b+ ,
{
2
2
b. } = i
b + i .
b i
b. .
b ,
{
2
2
Note that these equations, as well as the conditions (4.12), are manifestly invariant under
the mirror symmetry transformation (3.20).
To summarize, there exists a unique mirror-symmetric two-dimensional N = (2, 2)
b .
dilaton supergravity defined by the set of constraints (4.12) on the covariant derivative
The new supergravity theory does not have gauged symmetry and contains a real dilaton
field , as we will show in a moment.

S.J. Gates Jr. et al. / Nuclear Physics B 584 (2000) 109148

123

Since we define new N = (2, 2) dilaton supergravity theory imposing constraints (4.12)
b , the Bianchi identities in this theory may lead to further
on the covariant derivatives
b , the
constraints on some fields. 4 For theories described by covariant derivatives
Bianchi identities are simply Jacobi identities:


b ,
b ) } = 0,
b[ , [
(4.13)

where [ , } is the graded commutator, and [ , ) stands for the graded antisymmetrization
directly
symbol. Instead of deriving derivative constraints on the spinor fields and e
from the Jacobi identities (4.13) we use an equivalent approach which is much easier.
While substituting (4.11) into (4.8) one also finds terms proportional to gauge symmetry
generators Y and Y 0 . Vanishing of these terms leads to a set of constraints which is
equivalent to the set of constraints obtained from the Jacobi identities (4.13). We outline
the result in Appendix B.
There are two simple solutions to the Jacobi identities corresponding to either or
b antie
put to zero. An advantage of the first solution is that the covariant derivative
2
b
commutes with itself, = 0, like in the usual gauged supergravity theories [23]. On the
other hand, in the second case we find especially simple form of the anti-commutators
b } and {
b+ ,
b. }. In both cases the remaining spinor superfields can be expressed
b+ ,
{
in terms of an unconstraint real superfield V :
V = V

(4.14)

so that the Jacobi identities (4.13) are satisfied. This means that (4.13) impose no further
constraints on V , and only define the other superfields (like F and R) in terms of the
derivatives of V . It is natural to identify the dilaton field with the leading scalar component
of V :
= V |.

(4.15)

Below we present more evidence for this identification. One might notice that a real
superfield V contains one massless vector field, in agreement with the result of the previous
section. 5 It is also worthwhile to stress here that massless superfield V is not a Goldstone
multiplet itself, but rather what remains after the Goldstone mechanism takes place. A nice
property of this solution is that V is manifestly mirror-symmetric.
Since local integration measures of the new N = (2, 2) dilaton supergravity can be
nicely derived from the corresponding expressions of the U (1)A U (1)V theory only
for the solution corresponding to e
= 0, in what follows we discuss in detail only this
case. Namely, we take the following ansatz for the spinors :
b+ V ),
+ = i(
+ = e
bu V ),
u = i(
u = e

b V ),
= e
= i(
b. V ),
. = e
. = i(

(4.16)

4 In ordinary field theories, the fields satisfy Bianchi identities because they are expressed in terms of the
potentials; they are identities and impose no extra constraints.
5 However, massless vector fields in two dimensions do not have propagating degrees of freedom. For the
same reason two-dimensional superfield V does not have an irreducible transverse component, unlike a similar
four-dimensional superfield.

124

S.J. Gates Jr. et al. / Nuclear Physics B 584 (2000) 109148

b V and e
which implies = 2i
= 0.
Substituting (4.16) into (4.12), we find the following supergravity algebra:
b+ } = 2(
b+ V )
b+ ,
b ,
b } = 2(
b V )
b ,
b+ ,
{
{
1
1
,
b+ ,
b } = RX
b. } = F X ,
b+ ,
{
{
2
2
bu } = i
b + (
bu V )
b+ + (
b+ V )
bu ,
b+ ,
{
b. } = i
b + (
b. V )
b + (
b V )
b. ,
b ,
{

(4.17)

where the superfields R and F can be obtained from the Bianchi identities. Solving the set
of constraints in Appendix B we get:
b+ V ,
b
R = 4

b.
b+ V .
F = 4

(4.18)

We also find the following expressions for the gauge connection:


b V 2
b+
b+ V )(
bu V ),
bu V + 4(
= i
=e
b V 2
b
b V )(
b. V ).
b. V + 4(
= i
= e
Further commutators of the covariant derivatives with vector indices follow from the
consistency of the Bianchi identities (4.13):

b ] = 2i+ u + (
bu + ) + (
b+ u )
b+ ,
b+ ,
[

b ] = 2i+ u + (
bu + ) + (
b+ u )
bu ,
bu ,
[

b ] = 2i . + (
b. ) + (
b . )
b ,
b ,
[

b ] = 2i . + (
b. ) + (
b . )
b. ,
b. ,
(4.19)
[


b. R)
b F ) + 1 . R 1 F X i R
b ] = i (
b. i F
b ,
+ i (
b+ ,
[
2
2
2
2
2
2


i
i
i
i
b R) + (
b. F ) 1 R + 1 . F X + R
b + F
b. ,
b ]=
bu ,
(
[
2
2
2
2
2
2


bu R)
b+ F ) + 1 u R 1 + F X + i R
b+ ,
b ] = i (
bu + i F
+ i (
b ,
[
2
2
2
2
2
2


b+ R) + i (
bu F ) 1 + R + 1 u F X i R
b+ i F
b ] = i (
bu ,
b. ,
[
2
2
2
2
2
2
where we used (4.17). It is worthwhile to stress here that one would obtain a different
result de-gauging the corresponding commutators in the U (1)A U (1)V non-minimal
supergravity [27].
5. Lagrangians for matter multiplets coupled to N = 2 dilaton supergravity
In order to couple matter fields to new N = (2, 2) dilaton supergravity we have to repeat
the analysis of [23]. Up to terms with two derivatives or four fermions, the most general
action of N = (2, 2) supergravity coupled to chiral superfields i and Zk , and twisted
chiral superfields j looks like (3.15):

S.J. Gates Jr. et al. / Nuclear Physics B 584 (2000) 109148

S=

125

d 2 x d 2 d 2 E 1 L(i , i , j , j , Zk , Zk )
Z
Z
Z
Z
.
e (j ) + c.c.
+ d 2 x d 2 E 1 W (i ) + d 2 x d + d Ee1 W

(5.1)

In order to obtain the component action corresponding to (5.1), one needs the appropriate
projection formulas. For gauged N = (2, 2) supergravity theories such formulas were
derived by Grisaru and Wehlau [7]. In the case of the minimal U (1)A theory the local
density projection formula has the following form:
Z
Z

.
2
4
1
d x d E L = d 2 x e1 2 + i + i u

.
.
+ 12 H u + u 2 L|.
(5.2)
Here is the gravitino field and L is an arbitrary scalar function of superfields. In fact,
the same projection formula is also valid in the non-minimal U (1)A U (1)V supergravity
theory [23]. Although (5.2) is a D-type superinvariant, sometimes it is called a chiral
density projector because in the non-minimal N = (2, 2) supergravity 2 L is a chiral
superfield (for a general L). Therefore, replacing 2 L by an arbitrary covariantly chiral
Lagrangian Lc , we can obtain the component projection formula for any chiral superspace
integral:
Z
Z

.
d 2 x d 2 E 1 Lc = d 2 x e1 2 + i + i u

.
.
+ 12 H u + u Lc |.
(5.3)
In particular, the superspace measures E 1 and E 1 are related as follows:
Z
Z
d 2 x d 4 E 1 L = d 2 x d 2 E 1 2 L|.

(5.4)

By mirror symmetry the twisted chiral density projection formula in the U (1)V gauged
supergravity theory has the following form:
Z
Z

d 2 x d 4 E 1 L = d 2 x e1 . + i + + i u

+ u + u u L|.
+ 12 G
(5.5)
As explained in [23], the derivation of this formula goes through as in [7] for the
case of U (1)A theory. In the case of the U (1)V U (1)A gauged supergravity the
symmetry between chiral and twisted chiral fields is restored by the contribution of the
anticommutator term {+ , . } F |. More explicitly, the twisted chiral density projection
formula (5.5) can be derived using the methods of [7] or [28].
The projection formulas in the two-dimensional N = (2, 2) dilaton supergravity can
b . Thus,
be obtained from (5.2) and (5.5) replacing by a new covariant derivative
substituting (4.11) and (4.16) into (5.2) and using the commutation relations (4.7)(4.8)
we obtain the following density projection formula:

126

S.J. Gates Jr. et al. / Nuclear Physics B 584 (2000) 109148

Z
2

d xd E




b+ V )
b (
b V )
b+ (
d 2 x e1

.
b+ V )
b+ (
+ i

b 2
.
.
b V ) + 1 H u + u
b (
L|
i u
2
Z



.
b + i
b+
b+ i u +
b
= d 2 x e1


.
.
b. L. (5.6)
bu
+ 14 H u + u +

L=

In order to convince even hard boiled sceptics that (5.6) is the right projector, in
Appendix D we repeat the calculation of Grisaru and Wehlau [7] in the new N = (2, 2)
dilaton supergravity. As expected, the result is equivalent to (5.6).
Similarly, the twisted chiral density projection formula (5.5) yields:
Z
Z



2
4
1
b. V )
b+ V )
b+ (
b. (
d x d E L = d 2 x e1


b+ V ) + i u
b. V )
b+ (
b. (
i

+ u + u
b L|.
bu
(5.7)
+ 12 G
We note that for a given superspace Lagrangian L, both projection formulas (5.6) and (5.7)
lead to the same result:
Z
Z



b+ V )
b V )
b (
b+ (
d 2 x d 4 E 1 L = d 2 x e1

.
b+ V )
b+ (
+ i

b 2
.
.
b V ) + 1 H u + u
b (
L|
i u
2
Z




b. V )
b+ (
b+ V ) i
b+ V )
b. (
b+ (
= d 2 x e1


u
u b b

b V ) + 1 G
b. (
u L|.
+ i u
2 + +
This follows from the corresponding property of the local density projectors in gauged
N = (2, 2) supergravity theory [7], and also can be verified explicitly using the
commutation relations (4.17) in the new N = (2, 2) dilaton supergravity.
Now we are ready to derive component actions for various superspace Lagrangians L.
Let us start with a simple example corresponding to pure dilaton supergravity. Obviously,
in order to reproduce the right exponential dependence on the dilaton field in (3.5), we
have to take the function L in the form:
Lgrav = exp(2V ).
Substituting this in the projection formula (5.6) (or (5.7)) and seting all the fermions to
zero, we obtain the following action for bosonic fields:
Z


(5.8)
Sgrav = d 2 x e1 exp(2) R (2) + 4( )( ) .

S.J. Gates Jr. et al. / Nuclear Physics B 584 (2000) 109148

127

Here we used (4.16) and the formulas for the other components of the real superfield V
derived in Appendix C. Clearly, the action (5.8) for dilaton and graviton fields agrees 6
with the first two terms in the effective action (3.5) of type IIA theory on a CalabiYau
four-fold. Moreover, we note that bosonic action (5.8) has exactly the same form as the
action of N = 0 dilaton gravity studied long time ago, see, e.g., [25].
Now we consider coupling of N = (2, 2) dilaton supergravity to matter fields. In
particular, we are interested in superspace form of the effective action (3.5) describing
compactification of type IIA string theory on a CalabiYau four-fold X. Once again, to
reproduce the exponential dependence on = V | we take the superspace action in the
form:
Z
(5.9)
d 2 x d 4 E 1 exp(2V )L,
where L is a function of all matter superfields but V . It is convenient to absorb exp(2V )
in the definition of the supervielbein determinant E01 = E 1 exp(2V ), so that L = 1
corresponds to pure supergravity action (5.8), like in N = 1 four-dimensional theory.
Commuting exp(2V ) to the left in (5.6), we find the modified projection formula:
Z
d 2 x d 4 E01 L
Z



b+ 3(
b+ V )
b 3(
b V )
= d 2 x e1 exp(2)


.
b+ V ) i u
b V )
b+ 3(
b 3(
+ i

.
.
+ 12 H u + u


bu V )
b. 2(
b. V ) L|.
bu 2(
(5.10)

Applying this projection formula to an arbitrary function L(i , i , j , j ) of chiral
superfields i and twisted chiral superfields j we get the action of the bosonic fields
(the vielbein determinant e1 is suppressed):
h


L = e2 L R (2) + 4 12 log L 12 log L


+ 12 (log L)i j ( i ) j 12 (log L)i j ( i ) j
i
+ 12  (log L)i j ( i )( j ) + (log L)j i ( i )( j ) . (5.11)
The subscripts on L denote derivatives with respect to the scalar fields, e.g., (log L)i j =
( 2 /i j ) log L. Deriving (5.11) one may find helpful some formulas from Appendix D
where we discuss in detail the component action of a free chiral superfield. A careful
reader may notice that (5.11) has the structure reminiscent of N = 1 supergravity in four
dimensions. In particular, it is convenient to introduce the Khler potential K:
L = exp(K)
6 We will account for the extra volume factor V in a moment.

(5.12)

128

S.J. Gates Jr. et al. / Nuclear Physics B 584 (2000) 109148

so that the superspace action (5.9) takes the form:


Z
Z
d 2 x d 4 E01 eK = d 2 x d 4 E 1 e2V eK .

(5.13)

Performing the superspace integration, one finds Lagrangian for the bosonic fields:
h


)( e
) 12 Ki j ( i ) j
R (2) + 4( e
L = e2e

+ 12 Ki j ( i ) j 12  Ki j ( i )( j )
i
(5.14)
+ Kj i ( i )( j ) ,
where we introduced a new dilaton field e
= + 12 K invariant under generalized Khler
transformations (3.7):
K K + 1 (i , j ) + 1 ( i , j ) + 2 (i , j ) + 2 ( i , j ).

(5.15)

Under this Khler transformation the original dilaton field is shifted in the opposite
way, so that the linear combination e
= + 12 K remains invariant. Since the Khler
metric is invariant under (5.15) as well, both the superspace action (5.13) and the
corresponding component action (5.14) are manifestly invariant under the generalized
Khler transformations (5.15).
Now we are in position to identify the function K that would reproduce the effective
action (3.5) of type IIA theory on a CalabiYau four-fold. Namely, the superspace action
(5.13) gives the effective action (3.5) if K is the total Khler potential (3.9). This form of
the superspace action might be expected for a number of reasons. First of all, it is similar
to the superspace action of N = 1 supergravity in four dimensions. Moreover, type IIA
supergravity on a CalabiYau four-fold has a breathing mode corresponding to rescaling
of the volume V c2 V and simultaneous shift of the dilaton + log c, cf. (3.1).
Therefore, the superspace action is expected to be a function of 2V + K, where K is the
total Khler potential given by (3.9)(3.10).
To summarize, we constructed superspace Lagrangians describing N = (2, 2) dilaton
supergravity coupled to matter superfields and found projection formulas that allow one
to rewrite integrals over the entire superspace in terms of component fields. Therefore,
we provide a superspace formulation of the effective field theories constructed from
compactification of type IIA string theory on CalabiYau four-folds, as well as more
general N = (2, 2) sigma-models with torsion coupled to dilaton supergravity, cf. [6].
Incorporation of superpotential terms is more subtle. These terms are superinvariants
obtained by integration only over a half of the superspace, cf. (5.1). Unfortunately, unlike
(5.2) and (5.5), the chiral and twisted chiral density projectors in N = (2, 2) dilaton
supergravity do not simply follow from the full superspace projector (5.10). However, by
dimensional arguments and from an examination of the index structure of the possible
terms, the chiral density projection formula must look like:
Z
Z
Z
 2

.
.
b + 1 H u + u W | (5.16)
d 2 x d 2 E 1 W = d 2 x e1
2

S.J. Gates Jr. et al. / Nuclear Physics B 584 (2000) 109148

129

where the dots stand for term containing or terms linear in covariant derivatives. 7 By
the similar reasoning, the twisted chiral density projection formula must look like (5.5):
Z
Z
e
d 2 x d 2 Ee1 W
Z


u
u e
b+ + 1 G

b.
(5.17)
= d 2 x e1
2 + + W |.
Although normalization and coefficients in (5.16) and (5.17) may not be correct, the terms
quadratic in the gravitino are rather general and, in particular, are independent on degauging. So, we infer that one effect of the superpotential is to produce a mass term for the
gravitino fields [23]:
m . W,

e.
m W

(5.18)

It is this property of N = (2, 2) supergravity that was needed in [14,15] in order to find
the superpotentials (3.13) and (3.14) induced by RamondRamond fluxes. Moreover, let
us demonstrate that (5.16) and (5.17) lead to the expected structure of the scalar potential
(3.18). Extending the computation of (5.11), from (5.1) we get the action of the auxiliary
fields:
2
2

Laux L 12 H i A i (log L)i + 12 G i B j (log L)j + (log L)i j Ai A j


(log L)i j Bi B j i(Wi (log L)i )Ai W 12 H + iAi (log L)i


+ iBi (log L)i + c.c.
e 1G
ei (log L)i Bi W
(5.19)
i W
2

Integrating out the auxiliary fields and using (5.12), we find the expected scalar potential
(3.18):



 |W |2 |W
e
e D j W
e |2
Di W
Laux eK K1 (Di W ) Dj W K1
i j
i j

with the covariant derivatives (3.17).

6. Compactification of type IIB string theory on CalabiYau four-folds


Compactification of type IIB string theory on a CalabiYau four-fold X leads to a chiral
N = (0, 4) supersymmetric effective field theory in two non-compact dimensions. In the
low-energy limit this theory is described by N = (0, 4) supergravity coupled to scalar
superfields. In this section we perform a KaluzaKlein reduction on a CalabiYau fourfold X and, in particular, find that the supergravity multiplet includes a real scalar dilaton
field instead of an SU(2) gauge field [29,30]. A manifestly supersymmetric formulation of
such N = (0, 4) supergravity theory will be presented in the next section. In this section
we show that KaluzaKlein harmonics combine into N = (0, 4) scalar superfields [30,
31]. Furthermore, since we deal with N = (0, 4) supersymmetry, there is a significant
difference between right-movers and left-movers. Namely, all left-moving modes are
7 These terms will not affect the action of bosonic fields.

130

S.J. Gates Jr. et al. / Nuclear Physics B 584 (2000) 109148

supersymmetry singlets. We will also see that the difference between the zero-point energy
of the left-movers and the right-movers is proportional to the Euler number of X [32].
In the large volume limit the bosonic spectrum of light modes in type IIB string theory
includes the metric gMN , the dilaton , the axion l, the 4-form tensor DMNP Q and two
RR and B NS . Together with the fermionic superpartners all these fields fit
tensor fields BMN
MN
into type IIB supergravity multiplet. Non-perturbative type IIB string theory is invariant
under SL(2, Z) duality group. In order to see the action of this group on the supergravity
fields, it is convenient to define the following quantities:
= l + ie ,


1
||2
Re
,
M=
1
Im Re

NS
[M BNP
]
,
HMNP =
RR
[M BNP ]

(6.1)
(6.2)

(6.3)

and
3 NS
RR
P BQR]
.
(6.4)
FMNP QR = [M DNP QR] + B[MN
4
Then, the field strength FMNP QR is a singlet under the SL(2, Z) duality group, while H
transforms as a vector. Finally, SL(2, Z) acts on a complex scalar in the usual way:

a + b
,
c + d

(6.5)

where the integer numbers a, b, c and d satisfy ad bc = 1.


The five-form field strength F is self-dual:
F = F.

(6.6)

Although this equation can not be derived from any action, for a moment we ignore this
subtlety and write bosonic type IIB supergravity Lagrangian simply as:


1
1
Tr(M M)
L(10) = g R (10) +
4
16

5 2
3 T
(6.7)
+ H M H + F + ,
16
24
where the dots stand for higher derivative terms.
As in type IIA theory, in order to find the zero-mode spectrum we have to expand type
IIB supergravity fields in harmonic (p, q)-forms on the space X. The metric modes are
exactly the same as in (2.3). Namely, from the reduction of gMN we find h1,1 real scalars si ,
h3,1 complex scalars j and the two-dimensional metric g . It turns out that one of the
scalars si comes into the two-dimensional supergravity multiplet. Namely, it is the Kaluza
Klein mode corresponding to the volume of the CalabiYau four-fold:
V = dij kl s i s j s k s l ,

S.J. Gates Jr. et al. / Nuclear Physics B 584 (2000) 109148

131

where dij kl are the intersection numbers of X given by (2.7). With this mode excluded,
the Khler deformations of the metric yield h1,1 1 scalars si = V 4 si satisfying the
condition:
1

dij kl si s j sk s l = 1.
Expanding the doublet of tensor fields as:
1,1

NS

h
X

1,1

(1,1)
ri i ,

RR

i=1

h
X

(1,1)

ti i

(6.8)

i=1

we get pairs of real scalars ri and ti , h1,1 in number. All these modes are both right-moving
and left-moving.
Expansion of the self-dual field D is a bit subtle. Namely, instead of D one has
P
to expand the field strength F i ui i(4) and impose the self-duality condition
(4)
(6.6). Depending on whether the form i H 4 (X, R) is self-dual or anti-self-dual
the scalar field ui is left-moving or right-moving, respectively. Therefore, we have to
distinguish carefully self-dual and anti-self-dual harmonics of F . To this end we recall
some topological properties of CalabiYau four-folds. There is a decomposition of the
space of the middle-dimensional forms on X:
H 4 (X, R) = B+ (X) B (X),
where we denote by B+ (X) (respectively B (X)) the space of (anti-)self-dual 4-forms
on X. Let us call the corresponding dimensions b = dim B (X). Then b+ and b are
related by the Hirzebruch signature 8 (see, e.g., [8]):
Z

1
7p2 p12 = + 32
(Q) = b+ b =
45
3
X

R
of the quadratic form Q(1 , 2 ) = X 1 2 . On the other hand, we also have b4 =
b+ + b = 2 + 2h3,1 + h2,2 . Hence, using (2.1) and (2.2) we find:
b+ = 47 + 3h1,1 + 4h3,1 2h2,1

(6.9)

and
b = 1 + h1,1 + 2h3,1 .

(6.10)

Actually, we can be a little bit more precise. All the forms of Hodge type (3, 1) or (1, 3)
are anti-self-dual, while the (4, 0)- and (0, 4)-forms on a CalabiYau four-fold are self-dual
[14]. Therefore, from (6.9) and (6.10) we find that:
(2,2)

= 45 + 3h1,1 + 4h3,1 2h2,1

(6.11)

(2,2)

= h1,1 1.

(6.12)

b+
and
b

8 The explicit form of the Pontryagin classes is given by: p = 1 tr R 2 , p = 1 tr R 4 + 1 (tr R 2 )2 .


1
2
2
4
8

132

S.J. Gates Jr. et al. / Nuclear Physics B 584 (2000) 109148

Now we expand the self-dual field strength F as:


2,2

2,2

3,1

i=1

j =1

k=1

b+
b
h
X
X
X
(+)
( ui )i +
( vj )j() +
( pk )k(3,1) + ( q) + c.c., (6.13)
F=

where the scalar fields ui and vj are real, while pk and q are complex. As we explained
above, ui and q must be left-moving, vj and pk must be right-moving. In particular, the
former are singlets with respect to four left supercharges satisfying:
 i
j
Q+ , Q+ = ij P+ .
Note that the KaluzaKlein modes of the self-dual field F associated with (2, 1)-harmonic
forms give 2-form field strengths of massless vector fields in two dimensions and,
therefore, do not lead to new propagating degrees of freedom.
Unlike N = (2, 2) theory constructed from compactification of type IIA string theory,
in type IIB compactification on on a CalabiYau four-fold X not all the fermionic
modes can be determined by N = (0, 4) supersymmetry. In the right sector we still can
use supersymmetry arguments to conclude that two-dimensional supergravity multiplet
/ fermions and 2 ind(D)
/ RaritaSchwinger fields that come from the
contains 2 ind(D)
corresponding spin- 21 and spin- 23 fields in type IIB supergravity. On a CalabiYau fourfold the Dirac index is given by:

Z 
1
7 2
p p2 = 2
/ =
(6.14)
ind(D)
1440
4 1
X

in accordance with N = (0, 4) supersymmetry. Furthermore, all the right-moving scalars


found above (i , i , pi , pi , sj , rj , tj , vj , and l) are accompanied by right-moving
fermions. Simple counting gives:
n+ = 4h3,1 + 4h1,1

(6.15)

for the total number of the right-moving fermions. 9 There are also left-moving fermions
which are supersymmetry singlets. The number of left-moving fermions, however, is not
determined by supersymmetry. So, it has to be computed separately. Since the fermions
in question come from the type IIB gravitinos, their number (minus the number of rightmoving fermions) is given by the RaritaSchwinger index:

/ 3/2 ).
n n+ = 2 ind(D
Using (6.15) and the explicit expression for the RaritaSchwinger index on a CalabiYau
four-fold X:

Z 
1
37 2
p1 31p2 = 4h1,1 4h3,1 + 4h2,1
/ 3/2) =
(6.16)
ind(D
180
4
X

we obtain:
n = 4h2,1 .
9 Note that n is divisible by 4.
+

(6.17)

S.J. Gates Jr. et al. / Nuclear Physics B 584 (2000) 109148

133

Now we are ready to assemble the supermultiplets. Combining the left-moving bosonic
modes with the fermion fields we get the following supermultiplets:
a gravitational multiplet :
h3,1 scalar multiplets i :
h1,1 scalar multiplets j :

g ,
i ,
sj ,

V,
i ,
rj ,

pi ,
tj ,

pi ,
vj ,

l.

Performing a reduction of the ten-dimensional supersymmetry conditions one can easily


check that these fields indeed represent bosonic components of the supermultiplets
as stated. Note, all the matter multiplets include four real scalar fields in accordance
with the general classification of scalar superfields [2931,33]. However, the content
of the gravitational multiplet is different from what was usually studied in N = (0, 4)
supergravity theories [29,30]. In compactification of type IIB string theory on a Calabi
Yau four-fold we find that the gravitational multiplet includes a real scalar V instead of
SU(2) gauge field. In the next section we describe the superspace formulation of this N =
(0, 4) supergravity using the Goldstone approach.
In order to find the low-energy effective action one has to substitute (2.3), (6.8) and
(6.13) into (6.7). Integrating over the internal space by means of the formulas (2.4), (2.5)
we get the following two-dimensional action for bosonic fields:




L(2) = g V 14 R (2) + 12 (+ )( ) + 12 Gi + i j
j


+ 12 Gi 0 pi + p j
j









1
+ 2 Gi j + s i s j + + r i r j + + t i t j + 0 v i + v j


 1
+ 12 Qij 0 ui uj + (0 q)( q) + c.c. .
2

(6.18)

This Lagrangian describes non-linear sigma-model interacting with N = (0, 4) supergravity. The target space of the left-moving fields is the cotangent bundle to the moduli space
of the CalabiYau manifold X, T Mc (X) T MK (X), cf. [30]. Since the moduli space
itself is a Khler space, this result agrees with the general analysis of N = (0, 4) supersymmetric sigma-models. According to [33], N = (0, 4) supersymmetric sigma-model is
based on a target space which has three covariantly constant (with respect to + ) complex
structures which obey the quaternionic algebra:
Jr Js = rs + frs t Jt .
Another interesting feature that we expect to see in this N = (0, 4) theory is SL(2, R)
symmetry of classical type IIB supergravity. Apart from , l, ri and ti , all the scalar
fields listed above are singlets with respect to this symmetry. The complex scalar =
l + ie transforms as (6.5) under SL(2, R) duality transformation, with a, b, c and d
real numbers obeying ad bc = 1. The doublet of real scalar fields (ri , ti ) transforms as a
vector under the general SL(2, R) transformation. In other words, only the scalar multiplets
i transform nontrivially under this symmetry, while all the other fields, including the
supergravity itself, are SL(2, R)-singlets.

134

S.J. Gates Jr. et al. / Nuclear Physics B 584 (2000) 109148

Finally, we remark that from T-duality with type IIA string theory on a CalabiYau fourfold we expect an anomaly similar to (3.2) in type IIB compactification on a CalabiYau
four-fold. Recall that due to the global anomaly (3.2), in type IIA vacuum we had to include
N = /24 fundamental strings filling two-dimensional spacetime to cancel the tadpole.
Under a T-duality in one of the spacetime directions the winding modes of these strings
transform into /24 momentum modes:

(6.19)
P+ P =
24
in the type IIB vacuum corresponding to compactification on the same CalabiYau fourfold X. Here, for the sake of simplicity, we assumed that there are no background fluxes.
One can interpret (6.19) as the difference in the zero-point energy of the left-moving and
the right-moving KaluzaKlein modes [32]. In order to see this, we note that a free boson
on a circle has vacuum energy 1/24 and a periodic fermion has vacuum energy +1/24.
Therefore, due to N = (0, 4) supersymmetry, in the right sector bosonic and fermionic
contributions cancel each other, i.e., P+ = 0. In the left sector we have 48 + 6h1,1 + 6h3,1
and l along
2h2,1 bosonic modes corresponding to the fields i , i , sj , rj , tj , ui , q, q,
with n = 4h1,1 fermionic modes, cf. (6.17). Hence, the total vacuum momentum in the
left sector is non-zero and is given by the following formula:

1
48 + 6h1,1 + 6h3,1 6h2,1 .
(6.20)
P =
24
Using P+ = 0 and the explicit expression (2.2) for the Euler number, one can easily obtain
the formula (6.19).
7. Superspace formulation of N = (0, 4) dilaton supergravity
In this section we construct N = (0, 4) dilaton supergravity that arise, for example,
in type IIB superstring compactification on CalabiYau four-folds. As we demonstrated
in the previous section such a theory has a number of distinct features which are absent
in the existing superspace formulations of two-dimensional N = (0, 4) supergravities.
Namely, unlike the standard formulations with gauged SU(2) R-symmetry [29,30], new
supergravity does not have a gauged symmetry and the supergravity multiplet contains a
real dilaton field V. Below we present a superspace construction of this theory obtained via
de-gauging SU(2) symmetry.
First let us remind that gauged N = (0, 4) supergravity is defined in superspace by the
following set of constraints [34]:
[+i , +j } = 0,

[+i , u j } = i2i ,


[+i , } = i + i X + j Yi j ,
[+i , } = 0,


[ , } = 12 +i +i + + i u i + RX + iFi j Yj i

(7.1)

on the covariant derivatives A (+i , u i , , ):


A = EA B DB + A X + iAAi j Yj i .

(7.2)

S.J. Gates Jr. et al. / Nuclear Physics B 584 (2000) 109148

135

Here EA B is the supervielbein, and X and Yji are the Lorentz and SU(2) symmetry
generators, respectively. The superfield AAi j is SU(2) gauge connection, while A stands
for the Lorentz spin-connection. We write [ , } for the graded (anti-)commutator. Finally,
DA denotes the flat space fermi and bose derivatives DA (D +i , D+ i , , ).
The Lorentz generators act on A as follows:
1
[X , +i } = +i ,
2
[X , } = ,

1
X , u i = u i ,
2
[X , } = .

Similarly, for the action of SU(2) gauge symmetry generators we have:



 k
Yj , +i = ik +j 12 jk +i ,

 k
Yj , u i = ji + k + 12 jk u i ,

 k

 k
Yj , = 0.
Yj , = 0,
The constraints (7.1) lead to a set of Bianchi identities that are solved if:
u i +j = 0,

+i +j = 12 i R + iFi j ,

+i R = i2 + i ,

+i Fj k = 2ik + j + jk + i .

The first step in obtaining two-dimensional N = (0, 4) supergravity theory that does not
contain a gauged SU(2) is to note that the covariant derivative in (7.2) can be split as:
bA + iAAk l Yl k .
A =

(7.3)

bA does not contain the SU(2) connection nor


Since the superspace covariant derivative
the generator, it can not describe two-dimensional N = (0, 4) supergravity with gauged
SU(2) symmetry. We next use (7.3) to derive the form of the commutator algebra for the
bA operators.

A straightforward set of calculations leads to:




b+j } = i A+i j k + A+j i k
b+k ,
b+i ,
[


bu j = i2 j
b + i A + j i k
b+k + iA+i k j
bu k ,
b+i ,

i


b = iA i k
b+i ,
b } = iA i k
b+k ,
b+k i + i X ,
b+i ,
[
(7.4)



i
1
+i
+
b }=
b+i + i
bu + RX ,
b ,
[
2
where the connection superfields now explicitly appear on the right-hand side of the
equations. The leading component of A+i j k is a component of the gauged supergravity
multiplet that could be eliminated by a gauge transformation in the SU(2) gauge-symmetric
phase.
At this stage, we have completed half of the de-gauging process. The second half
consists of specifying the spinorial SU(2) connections in terms of some components
of another (matter) multiplet that is consistent with the two-dimensional N = (0, 4)
supergravity theory. For this purpose, we introduce the second of the four distinct N =
(0, 4) scalar multiplets (SM-II) that were discussed in [30,31]. In the case of rigid
supersymmetry this multiplet is described by:

136

S.J. Gates Jr. et al. / Nuclear Physics B 584 (2000) 109148

D+i V = i i ,

V = V,

D+i j k = 2ik j jk i ,
Du i j = ji V + i j i ,

i i = 0,
j i = i j ,
D+i j = 0.

In the locally supersymmetric theory one has to replace D by .


The triplet of spin-0 fields j i can be eaten by the triplet of spin-1 fields in the minimal
N = (0, 4) supergravity multiplet and thus become their longitudinal components via the
usual Goldstone mechanism. This eliminates the local SU(2) symmetry. The scalar and
spinor field of the matter multiplet become the dilaton and dilatino.
We are thus led to conjecture that the new form of two-dimensional N = (0, 4)
supergravity with a component spectrum given by:
e
+i
Ai j
i
V

graviton,
SU(2)-doublet, gravitino,
SU(2)-triplet, vector auxiliary fields,
SU(2)-doublet, dilatino field and
real dilaton field

may be constructed with (7.4) as its starting point. We note that the chirality of the dilatino
is opposite to that of the gravitino.
In order to gain a control over the component field content of the theory, we must impose
the following constraints:


A+ij k = 2 j ik i jk ,


A+ i j k = 2 j ik i jk .
At lowest order in , the field A+ij k is a component field that is absent in the SU(2)
gauge-symmetric phase of the theory; it can be set to zero in the WessZumino gauge. On
the other hand, when the SU(2) symmetry is broken, part of this field becomes dynamical.
In general, A+ij k contains SU(2) representations of spin- 21 and spin- 23 . However, the above
constraints eliminate the pure spin- 23 representation of SU(2).
With this result substituted into (7.4),
b+j } = i[ i
b+j + j
b+i ],
b+i ,
[



 k
j
j
bu j } = i2
b + i 2 k j ik
b+i ,
b+k + i 2 k j i j
bu ,
[
i
i
i
k
b } = iA i k
b+i ,
b } = iA i k
b+k ,
b+k i + i X ,
b+i ,
[
[


b } = 1 +i
b ,
b+i + + i
bu i + RX
[
2
we can calculate the Bianchi identities:




b+j },
b+k + [
b+i },
b+j + [
b+k },
b+i = 0,
b+k ,
b+j ,
b+i ,
[




b+j },
bu k + [
bu k },
b+j + [
bu k },
b+i = 0.
b+i ,
b+j ,
b+i ,
[
These will be satisfied if:
b+i j = i j ,

b+i j = i i j iA

b V,
+ ji

etc.

(7.5)

S.J. Gates Jr. et al. / Nuclear Physics B 584 (2000) 109148

137

Let us now briefly comment on density projectors in the new N = (0, 4) dilaton
supergravity theory. For a general superspace Lagrangian L, the component action of N =
(0, 4) gauged supergravity can be obtained by means of the following projection formula:
Z
Z


1
bu i
bu j L|
d 2 x d 2 E 1 12 Cij
d 2 x d 2 d 2 E 1 L =
2
Z


1
b+i
b+j L|,
(7.6)
d 2 x d 2 E1 12 C ij
+
2
where the corresponding chiral and anti-chiral density projector formulas look like:
Z
Z


b+i + i4e + i
b+j L|,
d 2 x d 2 E 1 L| = i d 2 x 12 e1 C ij
Z
Z

 j
bu i + i4e +i
bu L|.
d 2 x d 2 E1 L| = i d 2 x 12 e1 Cij
Similar formulas also hold in the new two-dimensional N = (0, 4) dilaton supergravity.
The explicit expressions for the density projectors can be obtained by a straightforward but
tedious computation substituting (7.3) into (7.6).

Acknowledgements
We are grateful to Marc Grisaru, Martin Rocek and John H. Schwarz for useful
discussions. The research of S.J.G. is supported by the NSF grant No PHY-98-02551;
S.G. is supported in part by the Caltech Discovery Fund, grant RFBR No 98-02-16575 and
Russian Presidents grant No 96-15-96939. The work of E.W. is supported in part by NSF
Grant PHY-9513835 and the Caltech Discovery Fund.

Appendix A. Worldsheet calculation of type IIA string amplitudes


Consider compactification of type IIA string theory on a CalabiYau four-fold X. Let
us further assume that there are no background RamondRamond fluxes, so that twodimensional spacetime is flat. From the worldsheet viewpoint, this compactification
corresponds to adjoining c = (12, 12), N = (2, 2) superconformal field theory (SCFT)
to free conformal theory with central charge c = (3, 3) that is responsible for the twodimensional spacetime. It is the first part that will be interesting to us. Namely, we are
going to show that two-point correlation function of vertex operators corresponding to
chiral and twisted chiral superfields is zero, i.e., that the Zamolodchikov metric on the
moduli space of CalabiYau four-folds is block diagonal (3.8).
N = (2, 2) superconformal algebra consists of two N = 2 supervirasoro algebras one
left-moving and one right-moving each generated by an energymomentum tensor T ,
a current J and two weight 3/2 supercurrents G with J -charge Q = 1. Recall that in
a KaluzaKlein reduction two-dimensional chiral superfields come from harmonic (3, 1)forms on X, while twisted chiral superfields correspond to harmonic (1, 1)-forms. Similar
to the three-fold case, we identify these fields with marginal operators in (c, c) and (a, c)

138

S.J. Gates Jr. et al. / Nuclear Physics B 584 (2000) 109148

multiplets, respectively. Let us call this operators (1,1) and (1,1) . They are neutral,
= (0, 0), and have conformal weight 1/2. The lowest components of (anti-)chiral
(Q, Q)
so (1,1) and (1,1) are not the lowest
multiplets must satisfy 2h = Q and 2h = Q,
components in the corresponding multiplets. They can be obtained in the operator product
expansion of the supercurrents with operators :
(1,1) (z, z ) =
2G (w, w)

1
(1,1) (z, z ) + reg,
wz

(A.1)

where reg stands for the regular part. Another operator product expansion that will be
useful to us is the following:
(1,1) (z, z ) = reg.
2G (w, w)

(A.2)

Now we are ready to demonstrate (3.8). Consider a matrix element of the target space
metric Gi j that mixes (1, 1) and (3, 1) moduli:
Gi j
|z z0 |4

i

j
= (1,1)
(z, z ) (1,1)
(z0 , z 0 )
I

dw

j
i
2G (w, w)
(1,1)
(z, z ) (1,1)
(z0 , z 0 ) = 0.
=
2i

(A.3)

and
In the last equality we used the fact (A.2) that the operator product of G (w, w)
j
(z0 , z 0 ) has no singularity as w z0 .
(1,1)
One might think that Gi j would be non-zero since the corresponding OPE has a
singular part:


(1,1) (z, z )
(1,1) (z, z ) =
+ reg.
2G (w, w)
z
wz
However, repeating the above arguments in the right sector one can easily see that Gi j is

j
i
(1,1)
i
also zero. In fact, OPE is singular in both left and right sectors only for h(1,1)
j
i

i which correspond to the metric G for chiral multiplets and the
and h

(1,1)

(1,1)

i j

metric Gi j for the twisted chiral multiplets, respectively. Therefore, we conclude that the
target space is locally a product of the manifold Mc (X) parametrized by the chiral fields
i and the manifold MK (X) spanned by the twisted chiral fields j .

Appendix B. Extra derivative constraints arising from de-gauging N = (2, 2)


non-minimal supergravity
In this appendix we collect some more technical formulas that arise in the construction of
N = (2, 2) dilaton supergravity via de-gauging U (1)A U (1)V non-minimal supergravity.
Consistency of the de-gauging procedure requires that all terms in (4.8) with gauge
symmetry generators Y and Y 0 vanish, so that the commutator algebra of the new covariant
b has the form (4.12), e.g.:
derivative

S.J. Gates Jr. et al. / Nuclear Physics B 584 (2000) 109148

139

b+ ,
b+ } = {+ , + } 2{+ , + Y + e
{
+ Y 0 }
+ )Y 0 + i+ + + ie
+ +
= 2(+ + )Y 2(+e
b+ + + Y + e
+ )Y 0
= i(+ + e
+ )(
+ Y 0 ) 2(+ + )Y 2(+e




b+ + 2(+ + ) + ie
= i(+ + e
+ )
+ + Y + 2(+e
+ ) + i+e
+ Y 0


b+ + 2(
b+ + ) + ie
= i(+ + e
+ )
+ + + ie
+ + Y


b+e
+ ) + i+e
+ + i+e
+ Y 0
+ 2(




b+ + 2(
b+ + ) + 2ie
b+e
+ ) + 2i+e
+ Y 0 .
= i(+ + e
+ )
+ + Y + 2(
Therefore, we obtain the following commutation relation:
b+ } = i(+ + e
b+
b+ ,
+ )
{
plus two constraints (Jacobi identities):
b+ + ) ie
+ + = 0,
(

b+e
+ ) i+e
+ = 0.
(

Similar calculations lead to the following set of constraints:


b+ + ie
+ + = 0,

b
e
+ i = 0,

b+e
+ i+e
+ = 0,

b
e
e
i = 0,

b + + i(e
b+ +
+ e
+ ) = 0,

b. e
b+e
. +
+ + i(+e
. . e
+ ) = 0,

i
be
b+e
+e
+
+ + 2ie
i(+e
+ e
+ ) R = 0,

2
i
b. + + 2i+ . i(e
b+ . +
+ . + e
. + ) F = 0,

2
b+e
bue
e
+e
+ u ),
u +
+ ) + 2e
u + (+e
u + e
= i(
b
b
e
e
= i(+ u + u + ) + 2+ u + (+ u + + u ),

be
b. e
e
. +
) 2e
. + (e
. + e
e
. ),
= i(
b . +
b. ) + 2 . (e
. + e
. ).
= i(

By virtue of the above relations, the Jacobi identities (4.13) are automatically satisfied.
Furthermore, only half of the spinor fields are independent. In Section 4 we discuss two
are put to zero. In both cases the
natural solutions to these constraints: when either or e
remaining spinor superfields can be expressed in terms of an unconstrained real superfield.

Appendix C. Components of covariant derivatives in N = (2, 2) dilaton


supergravity
The expressions for the covariant derivatives in N = (2, 2) dilaton supergravity
evaluated at = 0 are, cf. [17]:

140

S.J. Gates Jr. et al. / Nuclear Physics B 584 (2000) 109148

b | = ,

b | +
b |
b | = D +

= D + + ,

(C.1)

where D is the fully covariant gravitational derivative with the Lorentz connection =
| that includes, in addition to the ordinary connection, extra terms that are bilinear in
the gravitini , . Specifically, D is defined to be: 10
D = e + X .
We also need expressions for the higher components of the covariant derivatives. The
b is defined by [17]:
component of
b }|
b ,
b | = 1 {
b

(C.2)

b is:
while the component of
b ,
b ]| +
b
b | = [
b |
b

b ]| + D
b | +
b
b | +
b
b
b ,
= [
|.
From (C.2) and (4.17) we obtain the following results:
b+ | = i+ + ,
b+

bu | = iu u ,
bu

b
b | = i ,

b. | = i. . ,
b.

b+
,
b | = 1 H X ,
b. | = 1 GX
b+

4
4
bu
b. | = 1 H X ,
b | = 1 GX ,
bu

4
4


.
i
i
+
bu | = D + + u + + i + i u + u + i . ,
b+

2
2
2
2
2


.
i
i
i
i
i
+

b. | = D + + + + . + u + .
b

2
2
2
2
2
and from (C.3) we derive the series of identities that appears below:
. 
X
b |=D
b+ | + i u D | 1 H + G
b+

2
4


i
+ i + + + u + + u
2

i
b
b
2i+ u (u + ) (+ u ) + + u
2

i u u
i u . .
+ + u + ,
2

 2

.
i
i b
1
1 +
i b

b
b
b
u + H
| = D | + D | (u R) + (+ F ) +
2
2
2
2
2
10 In the notations of [23] this would correspond to = = .

(C.3)

S.J. Gates Jr. et al. / Nuclear Physics B 584 (2000) 109148


 
1
+ u G X
2


.
.
i
G + + + + 2 + ( + . )
2

. u
.
i .

H + u + .
2


.
i
1
u
bu | + D | G + H X + i + + + u +
|=D
2
4
2

i +
bu + ) + (
b+ u )
+ + 2i+ u + (
2


.
i
i + u
u
+ + i u u + + . ,
2
2

1
i
i
i
1
.

b+ R) + (
bu F ) + H + u G
b
| = D | + D | + (
2
2
2
2
2

1 + 1 u
G H X
4
4


i
i
+ + H + + + .
2
2


.
.
i u
u + i 2 . . ,
+ G
2
2
.
i
b. R)
b F ) + 1 . H 1 G
b+ | + D | i (
+ i (

|=D
2
2
2
2
2

1 + 1 u
+ H+ G X
4
4

.
.
i
i . +
+ + . + i u u
+ + + 2 G
2
2
2


.
.
i
+ H . ,
2

.
.
b | + i D | 1 + H + u G X + i + +
|=D
2
4
2



i .

b. ) (
b . )
+ i + + . 2i . (
2

.
i . u
i .
+ u + . ,
2
2
hi
i +
i
1
1
b R) + (
b. F ) H + . G
bu | + D | + (
|=D
2
2
2
2
2
1 . i
1
G H X
4
4



i
i
i + +
+ + u + + H + + + + u +
2
2
2

.
i
2 u u u + G + + . ,
2
1
2
i
+
2
i
+
2

bu
b

b
b.

b
b+

b
b

b
bu

141

142

S.J. Gates Jr. et al. / Nuclear Physics B 584 (2000) 109148


+ u H X + i + +
b.
b |=D
b. | + i D | 1 + G

2
4
2

i
+ + .
2
h
i
b. ) + (
b . )
+ u u + 2i . + (
2
i

.
.
i
+ i . , .
2
where we also used the commutation relations (4.19).
b is given by:
b component of
Furthermore, the
b |
b | +
b |=
b
b | +
b
b |,
b

so that:
b ]| = [
b |,
b |] +
b
b | +
b
b |
b
b |
b
b |
b ,
[

b
b | +
b
b |
= [D , D ] + D[ ] + D[ ( ] ) +
b
b |
b
b |.

b ,
b ]|, we get:
b | into [
b
Substituting the above expressions for the components of
b ]|
b ,
[



= [D , D ] + D[ ] + D[ ]


i +
i b
i b.
1
+
b
+ D u | + D | + (
R) + (
F ) H
2
2
2
2

1
1 .
1 .
+ G G H X
2
4
4


i
i
+ + + + u + + H + +
2
2



.
i + u
i
u
+
.
+ + 2 u u + G +
2
2


.
.
i
1
b | + D | + H + u G X + i + +
+ D
2
4
2



i .

.
.
.
.
b
b
+ i + + 2i ( ) ( )
2


i . u
i . .
.
+ u +
2
2

bu | + i + D |
+ u D
2


i b.
1
1 .
1
1 .
i b
+ ( R) + ( F ) H + G G H X
2
2
2
2
4
4



i
i
i
+ + + + u + + H + + + + u +
2
2
2

S.J. Gates Jr. et al. / Nuclear Physics B 584 (2000) 109148

143



.
i
2 u u u + G + + .
2


.
b. | + i D | 1 + G
+ u H X + i + +
+ D
2
4
2

i
i
+ + . + u u
2
2
 

.
.
i
b. ) + (
b . ) + i . .
+ 2i . + (
2

. 
i
1
b+ | + u D | H + G
X
+ D
2
4



i
bu + ) (
b+ u ) +
+ i + + + u + + u 2i+ u (
2


.
i u
i
i
+ + u u + u + u .
2
2
2

.
i
b | + D |
D
2



 

i b
1
1
1
1
i b
u + + H
+ u G X
(u R) + (+ F ) +
2
2
2
2
2
2



.
.
.
i
i
i
+ G + + + + 2 + + . + H + u u
2
2
2

i . .
+ .
2


.
bu | + i u D | 1 G + H X
u D
2
4

i +
i + +
+ + u + +
2
2
bu + ) + (
b+ u )
+ 2i+ u + (



.
i
i
+ + u + i u u u + + .
2
2

.
b. | + i D |
D
2


1
i
1
i
1 +G
1 u H X
b+ R) + (
bu F ) + H + u G
+ (
2
2
2
2
4
4



i
i
i +
u
+ H + + + . + u G
2
2
2


.
i .

+ 2 . . .
2

(C.4)

Now one can compare this huge formula with the leading component of the commutator
b ] computed directly from the Bianchi identities (4.13):
b ,
[

144

S.J. Gates Jr. et al. / Nuclear Physics B 584 (2000) 109148


b ,
b ] = u S + S + i(
b+ S)
bu S) + i(
X
[


1 b.
1 b b
i . i
R F (
(

R)

F
)
u
+

2
2
2
2


i
i
1 b
1 b.
b+
R . F + (
(

+
F
)
R)
+

2
2
2
2


i
1 b
i
1 b
b.
u R + + F (
(

R)
+
+
F
)

+
u
2
2
2
2


i
1 b
i
1 b
b ,
u F + + R (
(

+
R)
F
)
+

+
u
2
2
2
2

(C.5)

where we denoted:
i b
1 . 1
i b.
R F .
S = (
F ) +
R) + (
2
2
2
2
b+ | = + in (C.4) and (C.5), for example, we find:
By comparing the coefficients of
i
1 b
i
1 b.
R . F + (
R) + (
F )
2
2
2
2

i
= D[ +] + + + + + u
2
 i .
.
i
i
+ + + u + + + u + +
2 
2
2


i u +
+
+
bu + ) (
b+ u )
i + + + u 2i+ u (
2

 i
 i .
.
i
G + + u + + + u + H .
2
2
2
In the same fashion we obtain the relations defining the other components of the superfields
R and F :
1 b
i
1 b
i
u F + + R (
u F ) + (+ R)
2
2
2
2


i +

= D[ ] + H + + i
2



i .
i
b
b
.
.
.
.
+ + 2i ( ) ( ) + u H + +
2
2

i .
+ + .
2
 i
.
i
i
+ u 2 + + . u +
2
2
2

i .
+ . ,
2
1 b.
1 b
i . i
R F (
F )
R) (
2
2
2
2

 i
.
i
= D[ u] + + + u + 2 u u + u
2
2

S.J. Gates Jr. et al. / Nuclear Physics B 584 (2000) 109148

145



i
+ u + u + 2 u +
2

i
i .
+ u + u u +
2
2

. u
i

bu + )
H + u 2i+ u + (
2



.
,
b+ u ) + i + u + i u u i u G
+(
2
2
i
1
1
i
+ (
bu R)
b+ F )
u R + + F (
2
2
2
2
 i
.
.
.
.
.
i
i
= D[ ] + + G + + + + u G + +
2
2
 2


.
.
.
i

b
b
.
.
.
.
+ 2i + ( ) + ( ) + i
2
 i u + .
i . .
i + u

2
2
2

.
i . .
2 . ,
2
b+ S)
bu S) + i(

u S + S + i(

i b.
1
1 .
i b
G
= [D , D ]X + + (
R) + (
F ) H +
2
2
2
2


1 .
1
1
G H + H + u G
4
4
4

i
1
i
b R) + (
b. F ) H
+ u (
2
2
2

1
1 .
1 .
+ G G H
2
4
4
 1
. 
1 . +

G + u H + + H + G
4 
4



 
1
1 +
1 u
i b
1
i b
(

F
)
+
+

G
R)
+
+ (
+
+
u
u
2
2
2
2
2
2

.
1
+ u G + H
4 

. i
b+ R) + i (
bu F ) 1 + H + 1 u G
1 +G
1 u H .
(
2
2
2
2
4
4

Appendix D. Derivation of the projection formula in N = (2, 2) dilaton supergravity


Here we repeat the derivation [7] of the local density projection formula in N = (2, 2)
dilaton supergravity. Namely, we start by writing the most general expression for the chiral
projector with the right dimension and index structure:

146

S.J. Gates Jr. et al. / Nuclear Physics B 584 (2000) 109148

Z
2

d xd E

Z
L=

b
b+
b + X+
b+ + X
b + Y ]
2 L|,
d 2 x e1 [

(D.1)

where the coefficients X and Y are to be determined. Following [7], we evaluate (D.1) for
of a free chiral multiplet:
the kinetic action L =
Z
Z
h



2
4
1
b
b
b2
b+
b |
2 | +
2
d x d E = d 2 x e1


 2 
b
b
2
2
b+ | +
b |
b




b
b
b+
b+ |
2 | + X+
2
+ X+



 
b
b
b | + Y
b2 | .
b
2 | + X
2
+ X
(D.2)
Clearly, the left-hand side of this formula is invariant under complex conjugation. So,
the right-hand side must be invariant as well. As we will see in a moment, this condition
completely determines the unknown coefficients X and Y . It suffices to consider only
bosonic terms. Using the formulas in Appendix C along with the definition (4.4), one can
easily compute the relevant components:

bu
b. |

b+
D + u + A,


.
+

bu
b. |
b
D + A,


b
bu
b. |
D D + i u + + + D
b+


.
.
i
+ i D H A + i u A
4


.

i u + A.
Substituting these into (D.2), we find that the resulting component action is invariant under
complex conjugation only if we put:

.
X+ = i ,

X = i u + ,


.
.
1
Y = H u + u + .
4
With these expressions for X and Y the result does not depend on whether we use the
chiral projector (D.1) or its complex conjugate, of course, as it should be. Specifically, the
chiral projection formula in N = (2, 2) dilaton supergravity takes the following form:

Z
Z


.
2
4
1
2
1 b b
b+ i u +
b
+ + i
d xd E L= d xe


 u
 b2
. u
.
1

+ H + +
L|. (D.3)
4
One can easily check that this expression is equivalent to the chiral density projector
(5.6) obtained by de-gauging the corresponding projector in the non-minimal N = (2, 2)
supergravity.

S.J. Gates Jr. et al. / Nuclear Physics B 584 (2000) 109148

147

References
[1]
[2]
[3]
[4]
[5]
[6]
[7]
[8]
[9]
[10]
[11]
[12]
[13]
[14]
[15]
[16]
[17]
[18]
[19]
[20]
[21]
[22]
[23]
[24]
[25]
[26]
[27]
[28]
[29]
[30]

S.-T. Yau (Ed.), Essays on Mirror Manifolds, International Press, 1992.


B. Greene, S.-T. Yau (Eds.), Mirror Symmetry II, International Press, 1997.
C. Vafa, Evidence for F-Theory, Nucl. Phys. B 496 (1996) 403.
A. Klemm, On the geometry behind N = 2 supersymmetric effective actions in four dimensions,
hep-th/9705131.
S.J. Gates Jr., Vector multiplets and the phases of N = 2 theories in 2D through the looking
glass, Phys. Lett. B 352 (1995) 43.
B. de Wit, M.T. Grisaru, E. Rabinovici, H. Nicolai, Two-loop finiteness of D = 2 supergravity,
Phys. Lett. B 286 (1992) 78.
M.T. Grisaru, M.E. Wehlau, Superspace measures, invariant actions, and component projection
formulae for (2, 2) supergravity, Nucl. Phys. B 457 (1995) 219.
A. Klemm, B. Lian, S.-S. Roan, S.-T. Yau, CalabiYau fourfolds for M- and F-theory
compactifications, Nucl. Phys. B 518 (1998) 515.
P. Candelas, X.C. de la Ossa, Moduli space of CalabiYau manifolds, Nucl. Phys. B 355 (1991)
455.
M. Haack, J. Louis, Duality in heterotic vacua with four supercharges, hep-th/9912181.
S. Sethi, C. Vafa, E. Witten, Constraints on low-dimensional string compactifications, Nucl.
Phys. B 480 (1996) 213.
E. Witten, On flux quantization in M-theory and the effective action, J. Geom. Phys. 22
(1997) 1.
A. Sevrin, J. Troost, Off-shell formulation of N = 2 non-linear sigma-models, Nucl. Phys. B 492
(1997) 623.
S. Gukov, C. Vafa, E. Witten, CFTs from CalabiYau four-folds, Nucl. Phys. B 584 (2000) 69,
preceding paper in this issue.
S. Gukov, Solitons, superpotentials and calibrations, hep-th/9911011.
W. Lerche, FayetIliopoulos potentials from four-folds, JHEP 9711 (1997) 004.
S.J. Gates, M.T. Grisaru, M. Rocek, W. Siegel, Superspace, Benjamin-Cummings, 1983.
P.S. Howe, G. Papadopoulos, N = 2, D = 2 supergeometry, Class. Quant. Grav. 4 (1987) 11.
S.J. Gates, C.M. Hull, M. Rocek, Twisted multiplets and new supersymmetric nonlinear sigma
models, Nucl. Phys. B 248 (1984) 157.
K. Becker, M. Becker, M-theory on eight-manifolds, Nucl. Phys. B 477 (1996) 155.
W. Lerche, C. Vafa, N.P. Warner, Chiral rings in N = 2 superconformal theories, Nucl. Phys.
B 324 (1989) 427.
L. Dixon, Lectures at the 1987 ICTP summer Workshop in High Energy Physics and
Cosmology.
S.J. Gates Jr., M.T. Grisaru, M.E. Wehlau, A study of general 2D, N = 2 matter coupled to
supergravity in superspace, Nucl. Phys. B 460 (1996) 579.
S.J. Gates, T. Kadoyoshi, S. Nojiri, S.D. Odintsov, Quantum cosmology in the models of 2d and
4d dilatonic supergravity with WZ matter, Phys. Rev. D 58 (1998) 084026.
C. Callan, S. Giddings, J. Harvey, A. Strominger, Evanescent black holes, Phys. Rev. D 45
(1992) 1005.
A. Alnowaiser, Supergravity with N = 2 in two dimensions, Class. Quant. Grav. 7 (1990) 1033.
M.T. Grisaru, M.E. Wehlau, Prepotentials for (2, 2) supergravity, Int. J. Mod. Phys. A 10 (1995)
753.
S.J. Gates, Ectoplasm has no topology, Nucl. Phys. B 541 (1999) 615.
E. Bergshoeff, E. Sezgin, The (4, 0) heterotic string with WessZumino term, Mod. Phys. Lett.
A 1 (1986) 191.
H. Nishino, Alternative N = (4, 0) superstring and -models, Phys. Lett. B 355 (1995) 117.

148

S.J. Gates Jr. et al. / Nuclear Physics B 584 (2000) 109148

[31] S.J. Gates Jr., L. Rana, Manifest (4, 0) supersymmetry, sigma models and the ADHM instaton
construction, Phys. Lett. B 345 (1995) 233.
[32] K. Dasgupta, S. Mukhi, A note on low-dimensional string compactifications, Phys. Lett. B 398
(1997) 285.
[33] C.M. Hull, G. Papadopoulos, P.K. Townsend, Potentials for (p, 0) and (1, 1) supersymmetric
sigma models with torsion, Phys. Lett. B 316 (1993) 291.
[34] R. Dhanawittayapol, S.J. Gates Jr., L. Rana, A canticle on (4, 0) supergravity-scalar multiplet
systems for a cognoscente, Phys. Lett. B 389 (1996) 264.

Nuclear Physics B 584 (2000) 149170


www.elsevier.nl/locate/npe

Geometry of the embedding of supergravity scalar


manifolds in D = 11 and D = 10
M. Cvetic a,1 , H. L a,1 , C.N. Pope b,2
a Department of Physics and Astronomy, University of Pennsylvania, Philadelphia, PA 19104, USA
b Center for Theoretical Physics, Texas A&M University, College Station, TX 77843, USA

Received 15 March 2000; accepted 4 April 2000

Abstract
Several recent papers have made considerable progress in proving the existence of remarkable
consistent KaluzaKlein sphere reductions of D = 10 and D = 11 supergravities, to give gauged
supergravities in lower dimensions. A proof of the consistency of the full gauged SO(8) reduction
on S 7 from D = 11 was given many years ago, but from a practical viewpoint a reduction to a
smaller subset of the fields can be more manageable and explicit, for the purposes of lifting lowerdimensional solutions back to the higher dimension. The major complexity of the spherical reduction
Anstze comes from the spin-0 fields, and of these, it is the pseudoscalars that are the most difficult
to handle. In this paper we address this problem in two cases. One arises in a truncation of SO(8)
gauged supergravity in four dimensions to U (1)4 , where there are three pairs of dilatons and axions in
the scalar sector. The other example involves the truncation of SO(6) gauged supergravity in D = 5
to a subsector containing a scalar and a pseudoscalar field, with a potential that admits a second
supersymmetric vacuum aside from the maximally-supersymmetric one. We briefly discuss the use
of these embedding anstze for the lifting of solutions back to the higher dimension. 2000 Elsevier
Science B.V. All rights reserved.

1. Introduction
In this paper, we discuss two examples where non-trivial subsets of the scalar sectors
of gauged supergravities are obtained by spherical reduction from a higher dimension.
The first example is the embedding of the scalars in the U (1)4 maximal abelian
truncation of SO(8) gauged N = 8 maximal supergravity in D = 4, arising from D =
11 via compactification on S 7 . The consistency of the full SO(8) reduction on S 7 was
proven in [1], although at a somewhat implicit level. The N = 2 truncation includes a
total of six scalar fields, comprising three dilaton/axion pairs. In terms of the original
1 Research supported in part by DOE grant DE-FG02-95ER40893.
2 Research supported in part by DOE grant DE-FG03-95ER40917.

0550-3213/00/$ see front matter 2000 Elsevier Science B.V. All rights reserved.
PII: S 0 5 5 0 - 3 2 1 3 ( 0 0 ) 0 0 2 1 5 - 7

150

M. Cvetic et al. / Nuclear Physics B 584 (2000) 149170

SO(8) representations of the full theory, where there are 35 scalars in the 35v , and 35
pseudoscalars in the 35c of SO(8), the three dilatons come from the 35v , and the three
axions come from the 35c . In [2], a further simplifying truncation was performed, in
which the three axions were set to zero. The reduction ansatz becomes considerably
more complicated when axions are included, as was already seen in the case of the
single dilaton/axion pair of the N = 4 gauged SO(4) truncation, discussed in [3]. In the
present example, the inclusion of the three axions as well as the three dilatons leads to a
considerably more complicated structure in the reduction ansatz.
The second example is a truncation of the SO(6) gauged N = 8 maximal supergravity
in D = 5, arising from type IIB via compactification on S 5 . In this case there are 42 spin-0
fields in total, comprising 20 scalars in the 200 of SO(6), 20 pseudoscalars in the 10 and 10,
and two singlets corresponding to the original dilaton and axion of the type IIB theory. The
truncation we shall consider retains two spin-0 fields, comprising one scalar from the 200 ,
and one pseudoscalar from the 10 and 10. This particular truncation is of interest because
it is large enough to include the fields that participate in two distinct supersymmetric vacua
of the D = 5 gauged theory [4], one with maximal N = 8, SO(6) symmetry, and the other
with N = 2, SU(1) U (1) symmetry. Although an explicit interpolating solution is not
known it is in principle describable within the truncation we are making.
In both of our examples, we shall concentrate on elucidating the geometrical structure
of the embedding in D = 11 or type IIB supergravity. Specifically, we shall concentrate
on the ansatz for the KaluzaKlein reduction of the metric tensor. Strictly speaking, one
can only be sure that the reduction is fully consistent with all the equations of motion of
the higher-dimensional theory if one has the complete ansatz for all the higher-dimensional
fields, including the antisymmetric tensor field strengths. (Or, alternatively, if an existence
proof for the consistency of the reduction ansatz has independently been constructed.)
Obtaining the ansatz for the antisymmetric tensor fields is notoriously difficult, and we
shall not complete this part of the analysis in this paper. In the case of our D = 4 example,
we can appeal to the results of [1], in which a complete proof of the consistency of the S 7
reduction is exhibited. In principle it allows one to read off the ansatz for the 4-form field
strength, although only an implicit procedure for its construction is presented. On the other
hand, the general ansatz for the metric tensor is rather explicit, and it is by making use of
this expression that we are able to obtain the D = 4 results in this paper. These results can
be used in order to study the eleven-dimensional geometrical structure of general domainwall solutions in D = 4 supported both by the three dilatonic scalars and also the three
accompanying axions. Such solutions can be constructed from the purely dilatonic ones by
means of SL(2, R) transformations.
In D = 5 the situation is less clear, since no proof for the consistency of the full S 5
reduction to SO(6) gauged N = 8 maximal supergravity currently exists. A conjecture
for the metric reduction ansatz appears in [4], which is closely analogous to the known
construction in D = 4 given in [1], and it is this that we use in order to obtain an explicit
expression for the metric embedding for our 2-scalar truncation. Again, the complexities
of the antisymmetric tensor embedding have prevented us from obtaining a full nonlinear result in that sector. Thus the status of our D = 5 embedding is that, subject to

M. Cvetic et al. / Nuclear Physics B 584 (2000) 149170

151

the assumption of an ultimate consistency of the S 5 reduction scheme, 3 and subject


to the assumption that the conjecture for the metric ansatz in [4] is correct, then our
explicit results for the 2-scalar metric ansatz is valid. In principle, our result can then
be used to study the geometry of the RG flow describing the transition between the two
supersymmetric extrema of the associated scalar potential.
2. N = 2 U (1)4 gauged supergravity in D = 4 from D = 11
2.1. The three dilaton/axion pairs in D = 4
The 35v + 35c of spin-0 fields in SO(8) gauged supergravity in D = 4 are described in
terms of a 56-vielbein V, with the block-diagonal form

 IJ
vij KL
uij
,
(1)
V=
v k`I J uk` KL
which transforms under local SU (8) and rigid E7 [10,11]. In terms of the quantities uij I J
and vij KL , it was shown in [1] (having been previously proposed in [12]) that the ansatz
for the inverse of the internal S 7 compactifying metric is
g mn (x, y) 1 g mn (x, y)
=
K mI J

1
2




K mI J K nKL + K nI J K mKL uij I J + vij I J uij KL + v ij KL ,

(2)

round S 7 ,

are the 28 Killing vectors on the


and
det(g
(x,
y))
mn
,
(3)
2 =
det(gmn (y))
where gmn (x, y) is the inverse of g mn (x, y), and gmn (y) is gmn (x, y) with the scalar fields
all set to zero, so that it becomes the round S 7 metric. The eleven-dimensional metric
ansatz will be given by [1,12]

2
= 1 ds42 + gmn (x, y) dy m dy n = 1 ds42 + g mn (x, y) dy m dy n ,
(4)
d s11

where

mn(x,y) is the inverse of g mn (x, y). 4


where gmn (x, y) = g
We use the parameterisation of the uij I J and vij KL matrices described in [13]. In
particular, we introduce three scalars i , and three associated pseudoscalars i , whose
kinetic Lagrangian is

1X
(i )2 + sinh2 i (i )2 .
(5)
L=
2
i

3 Further evidence for the consistency of the S 5 reduction was obtained in [5], where certain scalar plus gravity
truncations in KaluzaKlein sphere reductions were proved to be consistent. Additionally, the complete consistent
reductions of D = 11 supergravity on S 4 [6,7] and massive type IIA supergravity on S 4 [8] have been constructed.
Recently, more evidence for the consistency of the S 5 reduction was presented in [9].
4 For now, we shall leave out the KaluzaKlein gauge fields from the construction of the metric. As discussed
in [2,13], the truncation to three dilaton/axion pairs is naturally accompanied by the four U (1) gauge fields of
the maximal abelian U (1)4 subgroup of SO(8). These gauge fields are easily incorporated in the KaluzaKlein
ansatz, and we shall add them in at the end of the derivation. We shall also set the gauge coupling constant g
equal to 1 for now, and restore it later.

152

M. Cvetic et al. / Nuclear Physics B 584 (2000) 149170

To shorten the subsequent formulae, we make the following definitions:


ci cosh i ,

si sinh i .

(6)

Also, for future convenience, we introduce the standard dilaton/axion pairs (i , i ),


related to (i , i ) by
cosh i = cosh i + 12 i2 ei ,
cos i sinh i = sinh i 12 i2 ei ,

(7)

sin i sinh i = i e .
i

In terms of these fields, the scalar kinetic terms are



1X
(i )2 + e2i (i )2 .
L=
2

(8)

After some algebra, we find that uij I J and vij KL are given by
KL
1
Pij QKL
4 uij

= c1 (Pa1 a2 Qa1 a2 + Pa3 a4 Qa3 a4 ) + c2 (Pa1 a3 Qa1 a3 + Pa2 a4 Qa2 a4 )


+ c3 (Pa1 a4 Qa1 a4 + Pa2 a3 Qa2 a3 )
+ P12 c1 c2 c3 Q12 + c1 s2 s3 ei(2+3 ) Q34
+ c2 s1 s3 ei(1+3 ) Q56 + c3 s1 s2 ei(1 +2 ) Q78

+ P34 c1 c2 c3 Q34 + c1 s2 s3 ei(2+3 ) Q12 + c2 s1 s3 ei(1 3 ) Q78



+ c3 s1 s2 ei(12 ) Q56
+ P56 c1 c2 c3 Q56 + c1 s2 s3 ei(23 ) Q78
+ c2 s1 s3 ei(1 +3 ) Q12 + c3 s1 s2 ei(1+2 ) Q34

+ P78 c1 c2 c3 Q78 + c1 s2 s3 ei(2+3 ) Q56


+ c2 s1 s3 ei(1 +3 ) Q34 + c3 s1 s2 ei(1+2 ) Q12 ,

1
4 vij KL Pij QKL

= s1 ei1 a1 b1 a2 b2 Pa1 a2 Qb1 b2 + ei1 a3 b3 a4 b4 Pa3 a4 Qb3 b4

s2 ei2 a1 b1 a3 b3 Pa1 a3 Qb1 b3 + ei2 a2 b2 a4 b4 Pa2 a4 Qb2 b4

s3 ei3 a1 b1 a4 b4 Pa1 a4 Qb1 b4 + ei3 a2 b2 a3 b3 Pa2 a3 Qb2 b3
+ P12 s1 s2 s3 ei(1+2 +3 ) Q12 + s1 c2 c3 ei1 Q34

+ s2 c1 c3 ei2 Q56 + s3 c1 c2 ei3 Q78
+ P34 s1 s2 s3 ei(12 3 ) Q34 + s1 c2 c3 ei1 Q12

+ s2 c1 c3 ei2 Q78 + s3 c1 c2 ei3 Q56
+ P56 s1 s2 s3 ei(1+2 3 ) Q56 + s1 c2 c3 ei1 Q78

+ s2 c1 c3 ei2 Q12 + s3 c1 c2 ei3 Q34
+ P78 s1 s2 s3 ei(12 +3 ) Q78 + s1 c2 c3 ei1 Q56

(9)

M. Cvetic et al. / Nuclear Physics B 584 (2000) 149170


+ s2 c1 c3 ei2 Q34 + s3 c1 c2 ei3 Q12 .

153

(10)

Here, we have introduced P and Q simply as arbitrary antisymmetric tensors, in order to


provide a compact way of summarising all the components of the uij I J and vij KL matrices.
The index notation is as follows. Indices with a 1 subscript, such as a1 , range over the
values (1, 2); similarly a2 ranges over (3, 4), a3 ranges over (5, 6) and a4 ranges over
(7, 8).
Next, we substitute these results into the ansatz (2) for the inverse S 7 metric. It is
advantageous to introduce a new parameterisation for the dilaton/axion pairs, as follows:

ei 1 + i2 Yi4 1/2 Y 1 ,
Y
bi i Yi2 ,
(11)
Yi e(1/2)i ,
i
and so
cosh i =
cos i sinh i =

1
2
1
2


ei2 ,
Yi2 + Y

ei2 ,
Yi2 Y

(12)

sin i sinh i = bi .
It is also advantageous to redefine the SO(8) basis relative to the one we have used so
far. The action of transformation, which amounts to a triality rotation under which Kij
1
k`
2 (ij ) Kk` , is given explicitly in Appendix A. After doing this, we find that the inverse
internal metric (2) takes the form 5
s2 g mn m n



2
2
2
2
2
2
2
2
e12 K57
= Y12 K13
+ K14
+ K23
+ K24
+ K58
+ K67
+ K68
+Y


2
2
2
2
2
2
2
2
e22 K37
+ K16
+ K25
+ K26
+ K38
+ K47
+ K48
+Y
+ Y22 K15


2
2
2
2
2
2
2
2
e32 K35
+ K18
+ K27
+ K28
+ K36
+ K45
+ K46
+Y
+ Y32 K17
2
2
2
2
e12 Y22 Y
e12 Y
e22 Y
e32 K34
e32 K56
e22 Y32 K78
+ Y12 Y
+Y
+Y
+ Y12 Y22 Y32 K12

e12 K56 K78
2b2b3 Y12 K12 K34 Y

e22 K34 K78
2b1b3 Y22 K12 K56 Y

e32 K34 K56 .
2b1b2 Y32 K12 K78 Y

(14)

In order to proceed further, it is useful to look at the geometry of the 7-sphere in some
detail. Some useful results on this topic are collected in Appendix B.
2.2. The metric ansatz for the three dilaton/axion pairs
From the results in Appendix B, it follows that the inverse metric (14) for the system
with 3 dilatons and 3 axions is a direct sum of a 4 4 part involving the i basis vectors,
5 The notation for writing the inverse metric is 2 g mn . The derivatives do not act on any other objects
m n
s
here; it is just a convenient way of writing all the components of g mn in one formula, exactly analogous to
writing the downstairs metric as ds 2 = gmn dy m dy n . For example, the inverse of the 2-sphere metric ds 2 =
d 2 + sin2 d 2 is written as

s2 = 2 +

1
sin2

2 .

(13)

154

M. Cvetic et al. / Nuclear Physics B 584 (2000) 149170

and a 3 3 part involving the i basis vectors (which are constrained by the fact that
i i = 1):
s2 = s24 + s23 .

(15)

For the 4 4 inverse metric, we find


X

2
2
e2
2
s24 =
i Qi i 2b2 b3 Y1 1 2 Y1 3 4
i



e22 2 4 2b1 b2 Y32 1 4 Y
e32 2 3 ,
2b1b3 Y22 1 3 Y

(16)

where
Q1 = Y12 Y22 Y32 21 + Y12 22 + Y22 23 + Y32 24 ,
e32 23 + Y
e22 24 ,
e22 Y
e32 22 + Y12 21 + Y
Q2 = Y12 Y
e32 22 + Y
e12 24 ,
e12 Y
e32 23 + Y22 21 + Y
Q3 = Y22 Y
e22 22 + Y
e12 23 .
e12 Y
e22 24 + Y32 21 + Y
Q4 = Y32 Y

(17)

For the 3 3 part, we find


s23 = Y12 (1 2 2 1 )2 + Y22 (1 3 3 1 )2
e12 (3 4 4 3 )2
+ Y32 (1 4 4 1 )2 + Y
e22 (2 4 4 2 )2 + Y
e32 (2 3 3 2 )2 ,
+Y

(18)

where i i = 1.
Because of the block-diagonal structure, we can invert the two parts separately. For the
4 4 part, we straightforwardly invert the inverse metric to obtain


1 X 2
2
i Zi di2 + 2b2 b3 21 22 d1 d2 23 24 d3 d4
d s4 =

i

+ 2b1 b3 21 23 d1 d3 22 24 d2 d4


2 2
2 2
(19)
+ 2b1 b2 1 4 d1 d4 2 3 d2 d3 ,
where
e22 Y
e12 Y
e12 Y
e32 22 + Y
e32 23 + Y
e22 24 ,
Z1 = 21 + Y
e2 Y 2 2 + Y
e2 Y 2 2 ,
Z 2 = 2 + Y 2 Y 2 2 + Y
2

2 3

1 2

1 3

e22 22 + Y32 Y
e22 24 ,
Z3 = 23 + Y12 Y32 21 + Y12 Y
e32 22 + Y22 Y
e32 23 .
Z4 = 24 + Y12 Y22 21 + Y12 Y

(20)

The function is given by


e22 Y
e32 42 + Y
e32 43 + Y
e22 Y32 44
e12 Y22 Y
e12 Y
= Y12 Y22 Y32 41 + Y12 Y


e22 + Y32 Y
e32 Y12 21 22 + Y
e12 23 24
+ Y22 Y


e22 22 24
e12 + Y32 Y
e32 Y22 21 23 + Y
+ Y12 Y


e32 22 23 .
e12 + Y22 Y
e22 Y32 21 24 + Y
+ Y12 Y

(21)

M. Cvetic et al. / Nuclear Physics B 584 (2000) 149170

155

There remains the problem of inverting the 3 3 part s23 of the inverse metric. Since we
know the inverse metric in the form (18), expressed in terms of the four i basis vectors
formed from the the constrained i , it is helpful first to solve the constraint i i = 1
explicitly, by introducing three angular coordinates as follows:
1 = c cos 12 ,
3 = s cos 1 ,
2

2 = c sin 12 ,
4 = s sin 1 ,
2

(22)

where c = cos , s = sin . It then follows that


= 12 (1 2 2 1 ),
= 12 (3 4 4 3 ),
= sc1 (1 1 + 2 2 ) + cs 1 (3 3 + 4 4 ).

(23)

Substituting into (18), the inverse metric is then expressed in terms of the three
unconstrained basis vectors ( , , ), and hence it can be straightforwardly inverted.
Having done so, the downstairs metric can then be re-expressed elegantly in terms of the
redundant set of four di differentials, in the form


1 X
2
Zi d2i + 12 b12 (1 d1 + 2 d2 )2 + (3 d3 + 4 d4 )2
d s3 =

i

+ 12 b22 (1 d1 + 3 d3 )2 + (2 d2 + 4 d4 )2


2
2
1 2
(24)
+ 2 b3 (1 d1 + 4 d4 ) + (2 d2 + 3 d3 ) .
Finally, adding this to the 4 4 metric d s42 given in (19), we obtain the result for the
downstairs 7-metric, d s72 = d s42 + d s32 :



1 X
Zi d2i + 2i di2 + 2b2 b3 21 22 d1 d2 23 24 d3 d4
d s72 =

i

+ 2b1 b3 21 23 d1 d3 22 24 d2 d4

+ 2b1 b2 21 24 d1 d4 22 23 d2 d3

+ 12 b12 (1 d1 + 2 d2 )2 + (3 d3 + 4 d4 )2

+ 12 b22 (1 d1 + 3 d3 )2 + (2 d2 + 4 d4 )2


1 2
2
2
(25)
+ 2 b3 (1 d1 + 4 d4 ) + (2 d2 + 3 d3 ) .
We can now work out the eleven-dimensional metric ansatz, given by (4). To do this, we
first note that the determinant of (25), where it is understood that the i coordinates are
expressed in terms of (, , ) using (22), is
 2 2 2 2  2 2 
1 2 3 4
21 22 23 24 s 2 c2
s c
,
(26)
=
det(gmn ) =
2
16
16 3

156

M. Cvetic et al. / Nuclear Physics B 584 (2000) 149170

where in the first expression, the first factor is the determinant of 4 4 block involving
the i coordinates, and the second factor is from the 3 3 block involving the (, , )
coordinates. From (3), it follows that
= 1/3 ,

(27)

and hence from (4) that the ansatz for the eleven-dimensional metric takes the following
rather explicit form:
2
= 1/3 ds42 + 1/3 d s72
d s11

= 1/3 ds42 + g 2 2/3


X



Zi d2i + 2i di2 + 2b2 b3 21 22 d1 d2 23 24 d3 d4


i

+ 2b1b3 21 23 d1 d3 22 24 d2 d4
+ 2b1b2 21 24 d1 d4 22 23 d2 d3




+ 12 b12 (1 d1 + 2 d2 )2 + (3 d3 + 4 d4 )2
+ 12 b22 (1 d1 + 3 d3 )2 + (2 d2 + 4 d4 )2
+ 12 b32 (1 d1 + 4 d4 )2 + (2 d2 + 3 d3 )2






.

(28)

Note that we have reinstated the gauge-coupling constant g in this expression.


Having obtained the KaluzaKlein metric ansatz for the three dilaton/axion pairs, it
is a simple matter to incorporate also the associated U (1)4 gauge fields that naturally
accompany this truncation of the maximal supergravity. Denoting their potentials by Ai(1) ,
for i = 1, 2, 3, 4, we simply replace each occurrence of di in (28) by
di di gAi(1) .

(29)

Finally in this section, we may note that our result (28) is consistent with previouslyobtained special cases. In particular, if we set the three axions i to zero, then the function
reduces to
= 2 ,

(30)

where
Y1 2
Y2 2
Y3 2
+
+
.
Y2 Y3 2 Y1 Y3 3 Y1 Y2 4
In the absence of axions, it is natural to define
Y1
Y2
Y3
, X3 =
, X4 =
,
X1 = Y1 Y2 Y3 , X2 =
Y2 Y3
Y1 Y3
Y1 Y2
implying that we shall have
X
Xi 2i ,
Zi = Xi1 .
=
= Y1 Y2 Y3 21 +

(31)

(32)

(33)

It can be seen that the metric ansatz (28) therefore indeed reduces to the one given in [2] if
the axions are set to zero.

M. Cvetic et al. / Nuclear Physics B 584 (2000) 149170

157

2.3. The ansatz for the 4-form field strength


b(4) of
In principle, we should like to obtain also the ansatz for the 4-form field strength F
eleven-dimensional supergravity. In spherical KaluzaKlein reductions it is always much
more difficult to obtain the ansatz for antisymmetric tensors than for the metric, and the
present case is no exception. Unfortunately, one can only obtain limited guidance from
those results that are presented in [1]. In other truncations, simpler than the case in hand, it
has been possible to determine the field-strength ansatz by brute-force methods, and up to
a point, this technique is still useful here. (This method was used successfully in [3], where
the complete and explicit ansatz for the S 7 reduction to the bosonic sector of N = 4, SO(4)
gauged supergravity in D = 4 were obtained.) The contributions to the 4-form ansatz can
be organised into different sectors, and in all except one of these we have obtained complete
results. Since these are instructive and useful in their own right, it seems to be worthwhile
to present those results that we have obtained here.
We begin with a summary of the four-dimensional theory comprising gravity, the three
dilaton/axion pairs, and the associated U (1)4 gauge fields.
2.3.1. D = 4 Lagrangian
The complete Lagrangian for four-dimensional N = 8 SO(8)-gauged supergravity was
obtained in [10,11]. In [2,13], the truncation to the N = 2 U (1)4 -gauged subsector was
discussed. Adapting these results to the notation of this paper, we find that the fourdimensional bosonic Lagrangian for this N = 2 truncation is given by

1X
di di + e2i di di V 1 + LKin + LCS ,
2
3

L4 = R1

(34)

i=1

where V is the potential for the scalar fields, and LKin and LCS are the kinetic terms and
i
= dAi(1) . The scalar potential
the ChernSimons terms for the four U (1) gauge fields F(2)
is given by
V = 4g 2

3
X


ei2 .
Yi2 + Y

(35)

i=1

The kinetic terms for the gauge fields are


h
e22 Y
e32 F 1 F 1 + Y
e12 Y22 Y32 F 2 F 2
e12 Y
LKin = 12 |W |2 P0 Y
(2)
(2)
(2)
(2)
4
4
e22 Y32 F 3 F 3 + Y12 Y22 Y
e32 F(2)
+ Y12 Y
F(2)
(2)
(2)

e12 F 1 F 2 Y12 F 3 F 4
+ 2P1 b2 b3 Y
(2)
(2)
(2)
(2)

2
1
3
2
2
4
e
+ 2P2 b1 b3 Y2 F(2) F(2) Y2 F(2) F(2)
i
e32 F 1 F 4 Y32 F 2 F 3 ,
+ 2P3 b1 b2 Y
(2)
(2)
(2)
(2)

where
P0 1 + b12 + b22 + b32,

W P0 2ib1 b2 b3 ,

(36)

158

M. Cvetic et al. / Nuclear Physics B 584 (2000) 149170

P1 1 b12 + b22 + b32,

P2 1 + b12 b22 + b32 ,

P3 1 + b12 + b22 b32.


Finally, the ChernSimons terms for the gauge fields are
h
1
2
e2 e2 1
e2 2 2 2
e2 Y
LCS = |W |2 b1 b2 b3 Y
1 2 Y3 F(2) F(2) + Y1 Y2 Y3 F(2) F(2)
4
4
e22 Y32 F 3 F 3 + Y12 Y22 Y
e32 F(2)
+ Y12 Y
F(2)
(2)
(2)


e12 F 1 F 2 Y12 F 3 F 4
+ b1 P0 + 2b22 b32 Y
(2)
(2)
(2)
(2)


1
3
2
4
e22 F(2)
F(2)
Y22 F(2)
F(2)
+ b2 P0 + 2b12 b32 Y
 2 1
i
4
2
3
e3 F(2) F(2)
Y32 F(2)
F(2)
.
+ b3 P0 + 2b12 b22 Y

From (34), we find that the equations of motion for the gauge fields are

d |W |2 Ri = 0,
for i = 1, 2, 3, 4, where


e12 Y
e22 Y
e32 P0 F 1 + 2b1 b2 b3 F 1
R1 = Y
(2)
(2)

 2 
2
e12 P1 b2 b3 F(2)
+ b1 P0 + 2b22 b32 F(2)
+Y



e22 P2 b1 b3 F 3 + b2 P0 + 2b12 b32 F 3
+Y
(2)
(2)

 4 
4
e32 P3 b1 b2 F(2)
+ b3 P0 + 2b12 b22 F(2)
,
+Y


e12 Y22 Y32 P0 F 2 + 2b1 b2 b3 F 2
R2 = Y
(2)
(2)

 1 
2
1
e
+ Y1 P1 b2 b3 F(2) + b1 P0 + 2b22 b32 F(2)

 4 
4
+ b2 P0 + 2b12 b32 F(2)
Y22 P2 b1 b3 F(2)

 3 
3
+ b3 P0 + 2b12 b22 F(2)
Y32 P3 b1 b2 F(2)
,


2 e2 2
3
3
R3 = Y1 Y2 Y3 P0 F(2) + 2b1 b2 b3 F(2)

 4 
4
+ b1 P0 + 2b22 b32 F(2)
Y12 P1 b2 b3 F(2)

 1
1
e22 P2 b1 b3 F(2)
+ b2 P0 + 2b12 b32 F(2)
]
+Y


2
2
2 2
2
Y3 P3 b1 b2 F(2) + b3 (P0 + 2b1 b2 )F(2) ,


4
4
e32 P0 F(2)
+ 2b1 b2 b3 F(2)
R4 = Y12 Y22 Y

 3 
3
+ b1 P0 + 2b22 b32 F(2)
Y12 P1 b2 b3 F(2)

 2
2
+ b2 P0 + 2b12 b32 F(2)
]
Y22 P2 b1 b3 F(2)



e32 P3 b1 b2 F 1 + b3 P0 + 2b12 b22 F 1 .
+Y
(2)
(2)

(37)

(38)

(39)

(40)

b(4)
2.3.2. The ansatz for F
b(4) was obtained for the
In previous papers the ansatz for the 4-form field strength F
4
U (1) truncation in absence of the three axions [2], and for the N = 4 gauged SO(4)
truncation, in which there is one scalar and one axion [3]. Based on those results, it can be

M. Cvetic et al. / Nuclear Physics B 584 (2000) 149170

159

b(4) as the sum of three terms, each with its own


seen to be natural to write the ansatz for F
characteristic contribution to the whole.
Thus we are led to the following construction for the 4-form field strength:
b0 + F
b00
b(4) = 2gU (4) + F
F
(4)
(4)


1
2Y11 dY1 1 Y14 d1 d 21 + 22
2g


1
2Y21 dY2 2 Y24 d2 d 21 + 23
+
2g


1
2Y31 dY3 3 Y34 d3 d 21 + 24 ,
+
2g
+

(41)

where



e12 23 + 24 + Y22 21 + 23
U = Y12 21 + 22 + Y



e32 22 + 23 ,
e22 22 + 24 + Y32 21 + 24 + Y
+Y
and (4) denotes the volume form on the four-dimensional spacetime. The term
be given by
X

2
i
b00 = 1 |W |2
d

gA
Ri .
F
i
i
(4)
(1)
2g 2

(42)
b00
F
(4)

will

(43)

b0 . This will be written


(We shall justify these expressions below.) The remaining term is F
(4)
b0 = d A 0 . It will be the determination of A 0 that
in terms of a potential A 0 , as F
(3)

(4)

(3)

(3)

presents the greatest difficulty.


b(4) = 0. This feature was already
b(4) does not identically satisfy d F
It will be noted that F
seen in the truncations in [2] and [3]. It is not possible, at least within the usual secondb(4) in the S 7
order formulation of eleven-dimensional supergravity, to write an ansatz for F
b
reduction that identically satisfies d F(4) = 0. An implication from this is that one cannot
write the ansatz directly on the potential A (3) , which in turn means that one cannot write an
ansatz that can be substituted directly into the eleven-dimensional action. One must work
at the level of the equations of motion.
b(4) = 0 provides
b(4) must satisfy the Bianchi identity d F
In fact the requirement that F
us with very important clues as to the correct form of the reduction ansatz, and we used
this in writing down our results in (41) and (43). The point is that the Bianchi identity will
be satisfied by virtue of the D = 4 equations of motion for the scalar fields and the U (1)
gauge fields being satisfied. (To be precise, the scalar equations of motion in question here
are those of the three dilatons i , in combination with certain non-linear admixtures of
b(4) from A 0 , whose
the three axion equations of motion.) Of course the contribution to F
(3)
precise form we have not been able to determine, does not enter into the discussion of the
b0 = 0.
b0 that identically satisfies d F
Bianchi identity, since it gives a contribution F
(4)
(4)
b(4) = 0 implies the four-dimensional equations of
To see how the Bianchi identity d F
motion for the scalars and the gauge fields, we note from the structure of (41) and (43) that
after acting with d we shall have two distinct classes of terms. First, there will be terms

160

M. Cvetic et al. / Nuclear Physics B 584 (2000) 149170

of the form d2i (4) , where (4) is a 4-form living entirely in the four-dimensional
spacetime. ((4) will comprise terms of the form (4) , and of the form ddYi , etc. Of
course they are all proportional to (4) .) The requirement of the vanishing of these terms
will imply the scalar equations of motion. Secondly, there will be terms of the form
b00 , where (3) is a 3-form
d2i (di 12 gAi(1)) (3) coming from the action of d on F
(4)
living in the four-dimensional spacetime. The vanishing of these terms will imply the fourdimensional equations of motion for the gauge fields.
Let us consider the second type of contribution first, since it is the simpler one. The
b00 , and give
terms of this type come only from d F
(4)
X


2
i
di di gA(1) d |W |2 Ri = 0.
(44)
i

This can immediately be seen to imply precisely the equations of motion for the four U (1)
gauge fields, given in (39).
It remains to check that the terms of the form d2i (4) coming from the Bianchi
identity vanish by virtue of the four-dimensional scalar equations of motion. The kinetic
terms of these scalar equations come from the action of d on the final three lines in (41).
Clearly, we get the combinations of the form

(45)
d 2Y11 dY1 1 Y14 d1 ,
arising (with similar independent expressions involving the (Y2 , 2 ) and (Y3 , 3 ) pairs).
This is a combination of the 1 and the 1 equations of motion. In fact it is




(46)
dd1 + e21 d1 d1 1 d e21 d1 ,
where the first quantity in square brackets is the dilaton equation of motion, and the second
quantity in square brackets is the axion equation of motion.
This particular combination, of the dilaton equation plus an admixture of the axion
equation, is an especially simple one to compare with the scalar equations of motion
coming from the four-dimensional Lagrangian (34). It means that we are looking at the
combination that comes from the following variation of the D = 4 Lagrangian:
4 L4 1 L4 .
L
1
1

(47)

If we define a symbol to denote this specific combination of field variations, i.e.,

1
,

1
1
then we find the great simplification that
12 = Y12 ,
Y

e12 = Y
e12 ,
Y

1 = 0.
b

(48)

(49)

(Of course since we are focusing on the scalars with the index i = 1 at the moment, all of
the scalar quantities with i = 2 or i = 3 labels are invariant under this transformation.) The
1 = 0, leads to an enormous simplification when we vary LKin and
last equation in (49), b
LCS given by (36) and (38). It means that |W |, the Pa , and all the bi are invariant. We need
e1 , and these just vary by the very simple rules given in (49).
only consider Y1 and Y

M. Cvetic et al. / Nuclear Physics B 584 (2000) 149170

161

With these observations, it becomes a relatively straightforward matter to verify that the
b(4) vanish precisely
terms of the form d2i (4) that arise in the Bianchi identity for F
as a consequence of the scalar equations of motion following from (34), to all orders in
scalar fields and gauge field strengths. Note that the contributions to the scalar equations of
motion from the potential V given in (35) arise from the action of the exterior derivative on
the term 2gU (4) in (41). This part of the calculation can be seen quite easily, and can be
examined in isolation from the more complicated contributions from the four-dimensional
gauge fields.
b0 = A 0 in (41) remains undetermined. We know some aspects of
The contribution F
(4)
(3)
it structure, for example that it is of the general from
X


j 
hij 2i d2j 2j d2i di gAi(1) dj gA(1) ,
(50)
A 0(3) =
i6=j

where the functions hij depend on the scalars i and i , and the direction cosines i . At
leading order, these terms will give rise to the linearised ansatz for the axions i . If explicit
expressions for the complete ansatz for the N = 8 SO(8) gauged supergravity embedding
were available, A0(3) could in principle be determined by substituting the expressions for
uij KL and vij KL appearing in (9) and (10) into them. To the extent that such expressions
are implicit in the work of [1], a procedure in principle exists for reading off A0(3) . It is not
clear that attempting such a substitution would be simpler than a brute-force direct attack
on the problem, of the type that has proved successful in previous (simpler) cases [2,3].
2.4. Domain wall solutions and their oxidation
The four-dimensional U (1)4 Lagrangian (34) supports a four-charge AdS black hole
solution [13]. In the extremal limit, the four U (1) gauge fields decouple and the solution
becomes AdS domain wall, supported by the scalar fields only. It is given by [14]
ds4 = (gr)4 (H1 H2 H3 H4 )1/2 dx dx + (H1 H2 H3 H4 )1/2
ei = Y12 = fi ,

i = 0,

dr 2
,
g2 r 2
(51)

where
f1 =

(H3 H4 )1/2
,
(H1 H2 )1/2

f3 =

(H2 H3 )1/2
,
(H1 H4 )1/2

(H2 H4 )1/2
,
(H1 H3 )1/2
`2
Hi = 1 + i2 .
r
f2 =

(52)

This solution can be oxidised back to D = 11 [14], where it acquires the interpretation of
being a continuous ellipsoidal distribution [1419] of M2-branes.
The scalar kinetic terms in the Lagrangian (34) are invariant under global SL(2, R)3
transformations, corresponding to the usual fractional-linear group action on each of the
axion/dilaton pairs. The scalar potential in (34), on the other hand, is invariant only under
the SO(2)3 subgroup transformations

162

M. Cvetic et al. / Nuclear Physics B 584 (2000) 149170

i i0 =

cos i + sin i
,
sin i + cos i

(53)

where i = i + iei . Applying these global transformations to the original domain


walls we obtain new solutions, with
Yi2 = ei =


1 2
fi cos2 i + sin2 i ,
fi

i =

1
2 (fi 1)
.
2
fi cos2 i + sin2 i

(54)

ei are hence given by


The Y
ei2 =
Y

f 2 + 14 (f 1)2 sin2 (2i )


fi (fi2 cos2 i + sin2 i )

(55)

Having obtained the SO(2)3 rotated domain-wall solutions, they can be oxidised back
to D = 11. The eleven-dimensional metric is given by substituting the solution into (28).
These solutions with non-vanishing i no longer simply describe distributed M2-branes.
To see this we note from (50) that with non-vanishing axions the field strength F(4) will
involve components lying purely in the internal S 7 . By contrast, in a distributed M2-brane
solution one has F(4) = d 3 x dH 1 , where H is the harmonic function in the transverse
space. Thus for a distributed M2-brane the field strength F(4) always carries three worldvolume indices.
3. The 2-scalar D = 5 embedding in type IIB
In this section, we consider the embedding of the 2-scalar truncation of D = 5 gauged
supergravity discussed in the introduction, and its embedding in the type IIB theory via an
S 5 reduction. In the early stages of the derivation, we retain all four of the scalar fields of
the truncation discussed in [4].
3.1. The metric reduction ansatz
The set of 42 spin-0 fields in the complete SO(6) gauged N = 8 supergravity in D = 5
[20] are described by a 27-bein V, which transforms under local USp(8) and global E6 .
The truncation to four spin-0 fields is described in [4], in terms of an SL(6, R) SL(2, R)
basis, for which the components of the vielbein are decomposed as (V I J ab , VI ab ). In
terms of this decomposition, the following conjecture for the inverse S 5 metric has been
proposed [4]:
n e
eKLcd ac bd ,
VI J ab V
g mn (x, y) 2/3 g mn (x, y) = 2KImJ KKL

(56)

e is the inverse of the vielbein V, 2 = det(gmn (x, y))/ det(gmn (y)), and gmn (y) is
where V
the undeformed round S 5 metric where the scalar fields are set to zero. The ten-dimensional
metric ansatz will then be
2
= 2/3 ds52 + gmn (x, y) dy m dy n
d s10

= 2/3 ds52 + gmn (x, y) dy m dy n .

(57)

M. Cvetic et al. / Nuclear Physics B 584 (2000) 149170

163

The process of making the 4-scalar truncation in the vielbein V has been described in
detail in [21]. Substituting this into the metric ansatz (56) is a mechanical exercise that is
most conveniently implemented by computer. Since the final result is considerably simpler
than the intermediate stages we shall, without further ado, present the final answer. We find
that the inverse 5-sphere metric s25 g mn m n is given by


2
2
2
2
+ K25
+ K35
+ K45
s25 = X1 cosh 2y2 (cosh 2r sin sinh 2r) K15

2
2
2
2
+ K26
+ K36
+ K46
+ cosh 2y2 (cosh 2r + sin sinh 2r) K16

+ 2 cos sinh 2r sinh 2y2 (K26 K35 K25 K36 + K16 K45 K15 K46 )

 2


2
2
2
+ K34
+ K23
+ K14
+ X2 14 3 cos + 2 cos2 cosh 4r K12

2
2
+ K24
+ 2 cos2 sinh2 2rK12 K34
+ cosh2 2y2 K13

2 sinh2 2y2 K13 K24
2
.
+ X4 K56

(58)

The scalars (X, r, y2 , ) are related to the quantities (, 1 , 2 , ) appearing in [21] by


= X1/2 ,

r = 12 (2 1 ),

y2 = 12 (1 + 2 ),

= 2.

(59)

Note that the D = 5 scalar Lagrangian for this truncation is


L = 2

3
X
(i )2 sinh2 (1 2 )( )2 V ,

i=1

= e 63 /2 ,

and the scalar potential V takes the form [4]


 

V = g 2 X2 1 cos2 sinh2 1 sinh2 2


+ X1 cosh 21 + cosh 22


1 4
X 2 + 2 sin2 2 sin2 cosh 2(1 2 )
+ 16

cosh 41 cosh 42 .

where X

(60)

(61)

At this stage, we impose the further truncation to the 2-scalar subsector that we really
want to consider. This corresponds to setting = 0 and 2 = 0 [21]. It is easily verified
from (60) and (61) that this is a consistent truncation. Thus we shall have r = 12 , and
y2 = 12 , where we now drop the 1 subscript on 1 . The potential (61) reduces to


(62)
V = cosh2 X2 (2 cosh2 ) + 2X1 12 X4 sinh2 .
It is convenient also at this stage to perform a labelling of indices on the Killing vectors
Kij in (58), under which the index values (2, 3, 4) are cycled: 2 3, 3 4 and 4 2.
We now adopt a description of the round 5-sphere that is precisely analogous to the
one that we introduced in Appendix B for S 7 . This time, we shall end up with three

164

M. Cvetic et al. / Nuclear Physics B 584 (2000) 149170

direction cosines i , subject to the condition i i = 1, and three azimuthal angles i .


After manipulations similar to those in Section 2, we arrive at the following expression for
the inverse 5-sphere metric 52 :
s25 = s22 + s23 ,

(63)

where the 2 2 and 3 3 blocks are given by




s22 = cosh2 X1 (1 3 3 1 )2 + (2 3 3 2 )2

+ X2 (1 2 2 1 )2 ,


2 2 
2
2 2 2
s23 = cosh2 X 2
1 1 + 2 2 + X 3 3
2
sinh2 X(1 + 2 ) X2 3 ,

(64)

and
(21 + 22 )X + 23 X2 .

(65)

Note that the 2 2 inverse metric s22 is just equal to the metric for the single-scalar
truncation when = 0, multiplied by a factor of cosh2 . The 3 3 inverse metric is
equal to cosh2 times the = 0 metric, with the correction term appearing in its second
line.
The inverse of the 3 3 block s23 is straightforward to calculate, and we find
d s32 =



sech2 1 2 2
X 1 d1 + 22 d22 + X2 23 d32

2
tanh2 2
1 d1 + 22 d2 23 d3 .
+
2

(66)

Note that the determinant of d s32 is given by (1 2 3 )2 /(3 cosh4 ).


For the 2 2 block, the inversion gives the metric


sech2 1
X d21 + d22 + X2 d23 .
(67)

It is helpful at this stage to reparameterise the direction cosines i , and make


redefinitions of the azimuthal angles (1 , 2 ) as follows:
d s22 =

1 = cos cos 12 ,
1 = 12 ( + ),

2 = cos sin 12 ,

3 = sin ,

2 = 12 ( ).

In fact (, , ) are just the Euler angles on


as
1 + i2 = ei (d + i sin d),

(68)
S3.

One can define left-invariant 1-forms i ,

3 = d + cos d.

(69)

These satisfy d1 = 2 3 , and cyclically. Defining also


c cos ,

s sin ,

(70)

we find that the 5-dimensional internal metric d s52 gmn (x, y) dy m dy n = d s22 + d s32
becomes

M. Cvetic et al. / Nuclear Physics B 584 (2000) 149170

d s52 =



sech2 
X d 2 + 14 X1 c2 12 + 22 + 32 + X2 s 2 d32

2
tanh2 2
c 3 2s 2 d3 ,
+
2
4

165

(71)

where
= Xc2 + X2 s 2 .

(72)

In the absence of the pseudoscalar field , this reduces to the metric ansatz encountered
in the N = 4 gauged SU(2) U (1) supergravity embedding obtained in [22]. In that case,
the scalar field X parameterises inhomogeneous deformations of S 5 viewed as a foliation
of S 3 S 1 surfaces.
With the pseudoscalar non-vanishing, it is advantageous to rewrite the metric (71) as
the sum of squares of just five quantities, by completing the square. After doing this, we
obtain the result
d s52 =

c2 X1

 c2 X 2
2
2

+
1
2
4 3
cosh2
4 cosh2

2
s2
c2 sinh2

d
+

,
3
3
2
2 cosh2
X

d 2 +

(73)

where
X3 c2 + s 2 cosh2 .

(74)

This expression reduces to the one found in [22] if = 0. In that case, the scalar X
parameterises deformations of S 5 corresponding to inhomogeneities of codimension 1 of
the foliation by S 3 S 1 . When the pseudoscalar is included too, the inhomogeneities
remain of codimension 1, but with a slightly more complicated structure. In addition, there
is a sort of twist in the S 3 S 1 product structure of the homogeneous foliating surfaces,
as indicated by the cross-term between the interval d3 on S 1 , and the 1-form 3 on S 3 .
Finally, substituting our result for the internal hatted metric d s52 into (57), we arrive at
the conjectured ten-dimensional metric ansatz for this two-scalar truncation:

X1/2 2
c2 X1
d +
12 + 22
1/2
cosh
4 cosh

2
2
1/2
2
s
c2 sinh2
c X cosh 2
3 + 3/2
3 .
d3
+
4
2
cosh

2
= 1/2 cosh ds52 +
d s10

(75)

3.2. The field-strength ansatz


There does not seem to be any straightforward way to determine the ansatz for the
KaluzaKlein reduction other fields of the ten-dimensional type IIB theory, in this twoscalar reduction. We know that when is taken to be zero, the ansatz must reduce to one
encompassed by the results in [22]. In particular, the remaining scalar field X enters in
the ansatz for the self-dual 5-form, whilst the dilaton, axion and 3-form field strengths of
the type IIB theory vanish when = 0. Since it is a pseudoscalar, the field enters at the

166

M. Cvetic et al. / Nuclear Physics B 584 (2000) 149170

linearised level in the ansatz for the NSNS and RR 2-form potentials A (2) A NS
(2) and
RR

A [23].
(2)

The relevant bosonic equations of motion of the type IIB theory are



1 b1 2
1 b1 2
b2
bMN = 1 H
R
96 MN + 4 F(3) MN 12 F(3) g MN



b2 2 1 F
b2 2
+ 14 F
(3) MN
12 (3) g MN ,
b(5) F
b(3) ,
b(3) = iH
d F
b
b
b(5) = i F

dH
2 (3) F (3) ,

(76)

b(5) = H
b(5),
H

where we have introduced the notation that


RR
A (2) A NS
(2) + iA(2) .

(77)

We are assuming here that the dilaton and axion of the type IIB theory vanish in the
reduction. For this to be consistent with the type IIB equations of motion, it is necessary
that
(3) = 0,
b(3) Fb
F

b(3) F
b(3) = Fb
(3) Fb
(3) .
F

(78)

We shall restrict our discussion from now on to the linearised level.


In the notation that we are using here, the linearised ansatz for pseudoscalars will be
of the form
A (2) = Y(2) ,

(79)

where Y(2) is a complex 2-form spherical harmonic satisfying


dY(2) = iY(2)

(80)

b(5) + G
b(5)
b(5) G
on the unit round 5-sphere. The ansatz for the self-dual 5-form H
b
includes a FreunndRubin term G(5) = 4(5) (we have set the gauge coupling g = 1
here). Substituting into the type IIB equations of motion, one finds that the pseudoscalar
satisfies the linearised equation of motion


(81)
dd + ( 4)(5) Y(2) = 0.
A 2-form harmonic with eigenvalue gives a pseudoscalar with m2 = ( 4). We want
the mass for the 10 and 10 members of the massless multiplet, namely m2 = 3, which
therefore requires = 1 or = 3. In fact, the required harmonics are those with = 3
(there are none with = 1).
There are ten such harmonics on S 5 , which can be written in terms of the Killing spinors.
There are Killing spinors satisfying Da = 2i a . It turns out that the required 2form harmonics are given by the construction
Yab = ab + ,

(82)

where and + are any two Killing spinors of the minus and plus kinds, respectively.
Solving for the Killing spinors, and substituting into (82), we find that one of the ten

M. Cvetic et al. / Nuclear Physics B 584 (2000) 149170

167

harmonics has a structure that is particularly naturally adapted to our parameterisation


of the sphere, namely

(83)
Y(2) = ei3 c d 3 + 12 sc2 1 2 isc2 3 d3 .
One may expect that this harmonic, or a closely related construction, will play a significant
rle in the construction of the reduction ansatz at the full non-linear order, but we have not
yet completed this investigation.
3.3. Oxidation of five-dimensional solutions
Given the conjectured metric reduction ansatz, we can oxidise the metric in any solution
of the two-scalar truncation of five-dimensional maximal gauged supergravity back to a
solution of type IIB supergravity in D = 10. In principle, one can solve the equations of
motion in this two-scalar sector to obtain a supersymmetric domain wall solution, which
has an interpretation as the RG-flow equations on the strongly coupled field theory side,
as discussed in [21]. Unfortunately the equations seem not to allow an explicit solution in
terms of elementary functions.
One simple oxidation that we can perform is to take the D = 5 solution corresponding to
the second (non-trivial) supersymmetric stationary point of the potential. This corresponds
to the stationary point of (62) with [4]
X = 21/3 ,

sinh =

1 .
3

(84)

(The fully-supersymmetric stationary point is at X = 1, = 0.) Substituting into (73), we


find that the internal 5-sphere metric d s52 at this stationary point is given by


3
c2
2c2
2
2
2

2
+
d s5 = 7/3 d 2 +
2
2
2(1 + s 2 ) 1
3 + 5s 2 3

2 
s 2 (3 + 5s 2 )
c2
+

d
.
(85)
3
3
3(1 + s 2 )2
3 + 5s 2
4. Conclusion
In this paper, we have obtained the metric ansatz for two examples of KaluzaKlein
sphere reductions, both of which involve pseudoscalar as well as scalar fields. The first
example is the S 7 reduction of eleven-dimensional supergravity, with a truncation from
N = 8 to the N = 2 theory with U (1)4 gauge fields, three dilatons and three axions.
Among other uses this reduction allows one to study the eleven-dimensional geometries
corresponding to the lifting of the four-dimensional BPS AdS black hole and domain-wall
solutions [13] of gauged supergravity. Our results generalise those obtained previously in
[2], where the problem was studied in the absence of the three axionic scalars.
Our second example is a truncation of five-dimensional maximal gauged supergravity,
to a subsector in which two spin-0 fields are retained, one of which is a scalar, and
the other a pseudoscalar. This truncation retains the fields necessary for describing a

168

M. Cvetic et al. / Nuclear Physics B 584 (2000) 149170

second supersymmetric vacuum in D = 5, with N = 2 supersymmetry and SU(2) U (1)


invariance, in addition to the maximally-supersymmetric one with SO(6) invariance [4].
The metric reduction ansatz that we obtain here allows one to study the ten-dimensional
geometries corresponding to the lifting of solutions of the five-dimensional theory. In
principle, this can include the renormalisation-group flow [21] associated with the second
supersymmetric extremum, although the explicit form of this five-dimensional solution is
not known.

Acknowledgement
We are grateful to Jim Liu, Krzysztof Pilch, Tuan Tran and Nick Warner for discussions.

Appendix A
In this appendix, we present the explicit form of the SO(8) triality rotation that we used
in Section 2.1 in order to simplify the KaluzaKlein metric reduction ansatz:
K12 12 (K12 + K34 + K56 + K78 ),

K13 12 (K13 K24 + K57 K68 ),

K14 12 (K14 + K23 + K58 + K67 ),

K15 12 (K15 K26 + K37 K48 ),

K16 12 (K16 + K25 + K38 + K47 ),

K17 12 (K17 K28 + K35 K46 ),

K18 12 (K18 + K27 + K36 + K45 ),

K23 12 (K23 + K14 K58 K67 ),

K24 12 (K24 K13 + K57 K68 ),

K25 12 (K25 + K16 K38 K47 ),

K26 12 (K26 K15 + K37 K48 ),

K27 12 (K27 + K18 K36 K45 ),

K28 12 (K28 K17 + K35 K46 ),

K34 12 (K34 + K12 K56 K78 ),

K35 12 (K35 + K17 + K28 + K46 ),

K36 12 (K36 + K18 K27 K45 ),

K37 12 (K37 + K15 + K26 + K48 ),

K38 12 (K38 + K16 K25 K47 ),

K45 12 (K45 + K18 K27 K36 ),

K46 12 (K46 + K35 K17 K28 ),

K47 12 (K47 + K16 K25 K38 ),

K48 12 (K48 + K37 K15 K26 ),

K56 12 (K56 + K12 K34 K78 ),

K57 12 (K57 + K13 + K24 + K68 ),

K58 12 (K58 + K14 K23 K67 ),

K67 12 (K67 + K14 K23 K58 ),

K68 12 (K68 + K57 K13 K24 ),

K78 12 (K78 + K12 K34 K56 ).


(86)

Appendix B
In this appendix, we collect some results on the geometry of the 7-sphere. We can
describe S 7 as the unit sphere in R 8 , with 8 real coordinates xI :
xI xI = 1.

(87)

M. Cvetic et al. / Nuclear Physics B 584 (2000) 149170

169

As such, it has a manifest SO(8) symmetry, with 28 Killing vectors KI J given by


KI J = x I

xJ
.
xJ
xI

(88)

We can also describe S 7 as the unit sphere in C 4 , with 4 complex coordinates zi :


z i zi = 1.

(89)

We can relate these complex coordinates to the previous real ones as follows:
z1 = x1 + ix2 ,

z2 = x3 + ix4 ,

z3 = x5 + ix6 ,

z4 = x7 + ix8 .

(90)

We can parameterise these complex coordinates as


z1 = 1 ei1 ,

z2 = 2 ei2 ,

z3 = 3 ei3 ,

z4 = 4 ei4 ,

(91)

where (89) implies that


4
X

2i = 1.

(92)

i=1

These (i , i ) coordinates are precisely the ones used for describing higher-dimensional
rotating black holes in [24], and in the S 7 reduction ansatz obtained in [2].
From the coordinate transformations above, it is straightforward to establish that the real
derivatives /xI that appear in the Killing vectors (88) are given by
sin 1

= cos 1

,
x1
1
1 1

cos 1

= sin 1
+
,
x2
1
1 1

(93)

with analogous expressions involving (2 , 2 ), (3 , 3 ) and (4 , 4 ) for the pairs (x3 , x4 ),


(x5 , x6 ) and (x7 , x8 ), respectively. It is easy to see from this that the four Killing vectors
K12 , K34 , K56 and K78 are simply of the form:
K12 =

,
1

K34 =

,
2

K56 =

,
3

K78 =

.
4

(94)

These are the four commuting U (1) generators. It is convenient to write them as 1 , etc.
We also note that the Killing-vector bilinears in the top 3 lines in (14) are also relatively
simple, when expressed in terms of the i and i coordinates. After some algebra we find,
for example, that
2
2
2
2
+ K14
+ K23
+ K24
= (1 2 2 1 )2 +
K13

22

21

22

2 +
2 1

22

(95)

with analogous results for the other five combinations.

References
[1] B. de Wit, H. Nicolai, The consistency of the S 7 truncation in D = 11 supergravity, Nucl. Phys.
B 281 (1987) 211.

170

M. Cvetic et al. / Nuclear Physics B 584 (2000) 149170

[2] M. Cvetic, M.J. Duff, P. Hoxha, J.T. Liu, H. L, J.X. Lu, R. Martinez-Acosta, C.N. Pope, H. Sati,
T.A. Tran, Embedding AdS black holes in ten and eleven dimensions, Nucl. Phys. B 558 (1999)
96; hep-th/9903214.
[3] M. Cvetic, H. L, C.N. Pope, Four-dimensional N = 4, SO(4) gauged supergravity from D =
11, hep-th/9910252, Nucl. Phys. B, in press.
[4] A. Khavaev, K. Pilch, N.P. Warner, New vacua of gauged N = 8 supergravity in five dimensions,
hep-th/9812035.
[5] M. Cvetic, H. L, C.N. Pope, A. Sadrzadeh, Consistency of KaluzaKlein sphere reductions of
symmetric potentials, hep-th/0002056.
[6] H. Nastase, D. Vaman, P. van Nieuwenhuizen, Consistent nonlinear KK reduction of 11-D
supergravity on AdS7 S 4 and selfduality in odd dimensions, Phys. Lett. B 469 (1999) 96;
hep-th/9905075.
[7] H. Nastase, D. Vaman, P. van Nieuwenhuizen, Consistency of the AdS7 S4 reduction and the
origin of self-duality in odd dimensions, hep-th/9911238.
[8] M. Cvetic, H. L, C.N. Pope, Gauged six-dimensional supergravity from massive type IIA,
Phys. Rev. Lett. 83 (1999) 5226; hep-th/9906221.
[9] H. Nastase, D. Vaman, On the nonlinear KK reductions on spheres of supergravity theories,
hep-th/0002028.
[10] B. de Wit, H. Nicolai, N = 8 supergravity with local SO(8) SU (8) invariance, Phys. Lett.
B 108 (1982) 285.
[11] B. de Wit, H. Nicolai, N = 8 supergravity, Nucl. Phys. B 208 (1982) 323.
[12] B. de Wit, H. Nicolai, N.P. Warner, The embedding of gauged N = 8 supergravity into D = 11
supergravity, Nucl. Phys. B 255 (1985) 29.
[13] M.J. Duff, J.T. Liu, Anti-de Sitter black holes in gauged N = 8 supergravity, Nucl. Phys. B 554
(1999) 237; hep-th/9901149.
[14] M. Cvetic, S. Gubser, H. L, C.N. Pope, Symmetric potentials of gauged supergravities in
diverse dimensions and Coulomb branch of gauge theories, hep-th/9909121, Phys. Rev. D, in
press.
[15] P. Kraus, F. Larsen, S.P. Trivedi, The Coulomb branch of gauge theory from rotating branes,
JHEP 9903 (1999) 003; hep-th/9811120.
[16] D.Z. Freedman, S.S. Gubser, K. Pilch, N.P. Warner, Continuous distributions of D3-branes and
gauged supergravity, hep-th/9906194.
[17] A. Brandhuber, K. Sfetsos, Nonstandard compactification with mass gaps and Newtons Law,
hep-th/9908116.
[18] I. Bakas, K. Sfetsos, States and curves of five-dimensional gauged supergravity, hepth/9909041.
[19] I. Bakas, A. Brandhuber, K. Sfetsos, Domain walls of gauged supergravity, M-branes and
algebraic curves, hep-th/9912132.
[20] M. Gunaydin, L.J. Romans, N.P. Warner, Gauged N = 8 supergravity in five dimensions, Phys.
Lett. B 154 (1985) 268.
[21] D.Z. Freedman, S.S. Gubser, K. Pilch, N.P. Warner, Renormalization group flows from
holography, supersymmetry and a C-theorem, hep-th/9904017.
[22] H. L, C.N. Pope, T.A. Tran, Five-dimensional N = 4, SU(2) U (1) gauged supergravity from
type IIB, Phys. Lett. B 475 (2000) 261; hep-th/9909203.
[23] H.J. Kim, L.J. Romans, P. van Nieuwenhuizen, Mass spectrum of ten-dimensional N = 2
supergravity on S 5 , Phys. Rev. D 32 (1985) 389.
[24] R.C. Myers, M.J. Perry, Black holes in higher-dimensional spacetimes, Ann. Phys. 172 (1986)
304.

Nuclear Physics B 584 (2000) 171196


www.elsevier.nl/locate/npe

On the supersymmetry and gauge structure


of Matrix theory
Y. Kazama , T. Muramatsu 1
Institute of Physics, University of Tokyo, Komaba, Meguro-ku, Tokyo153-8902, Japan
Received 20 March 2000; accepted 26 May 2000

Abstract
Supersymmetric Ward identity for the low energy effective action in the standard background
gauge is derived for arbitrary trajectories of supergravitons in Matrix theory. In our formalism, the
quantum-corrected supersymmetry transformation laws of the supergravitons are directly identified
in closed form, which exhibit an intricate interplay between supersymmetry and gauge (BRST)
symmetry. As an application, we explicitly compute the transformation laws for the source-probe
configuration at 1-loop and confirm that supersymmetry fixes the form of the action completely,
including the normalization, to the lowest order in the derivative expansion. 2000 Elsevier Science
B.V. All rights reserved.
PACS: 11.25.-w; 11.25.Db; 11.30.Db
Keywords: Supersymmetry; Matrix theory; Gauge symmetry; Ward identity

1. Introduction
By now a considerable amount of evidence has been accumulated for the Matrix
theory for M-theory, originally proposed by Banks et al. [1] and later re-interpreted
by Susskind [2] in the framework of discrete light-cone quantization. In particular, just
to mention only the direct comparison with eleven dimensional supergravity, complete
agreement for the multi-graviton scattering (including the recoil effects) at 2-loop [3,4]
and that for the two-body potential between arbitrary fermionic as well as bosonic objects
at 1-loop [5] can be cited as highly nontrivial and remarkable.
Despite such impressive pieces of evidence as well as general supportive arguments
[6,7], the deep reason and the mechanism of the agreement are yet to be fully understood.
Evidently, one of the keys should be the understanding of the structure and the role of
Corresponding author. E-mail: kazama@hep3.c.u-tokyo.ac.jp
1 tetsu@hep1.c.u-tokyo.ac.jp

0550-3213/00/$ see front matter 2000 Elsevier Science B.V. All rights reserved.
PII: S 0 5 5 0 - 3 2 1 3 ( 0 0 ) 0 0 3 6 7 - 9

172

Y. Kazama, T. Muramatsu / Nuclear Physics B 584 (2000) 171196

the symmetries. The well-known symmetries present both in the Matrix theory and in
the supergravity theory are the global Spin(9) invariance, the CPT invariance and the
invariance under 16 supersymmetries. Perhaps less familiar is the generalized conformal
invariance [810], the generalization of the conformal symmetry that plays the major
role in the AdS/CFT correspondence of Maldacena [11]. Besides these symmetries, the
supergravity possesses the general coordinate invariance while the Matrix theory has the
YangMills type gauge symmetry, which must be deeply connected. Finally, although
not yet clearly identified, the agreement of the multi-body scattering amplitudes strongly
suggests that the eleven-dimensional Lorentz invariance is present in a highly nontrivial
manner in the Matrix theory.
Except for the eleven dimensional Lorentz invariance, the above-mentioned symmetries
of the Matrix theory are easily recognized in the original action. However, since
the supergravity interactions between various objects arise only after introducing the
corresponding backgrounds and integrating out the quantum fluctuations around them, the
realizations of these symmetries are in general modified for the effective action of interest.
Besides, being an off-shell quantity, the form of the effective action depends on the choice
of the gauge as well as on the definition of the fields. Further, there is always a vast degree
of ambiguities as one can add total derivatives. For these reasons, the study of how the
symmetries govern the structure of the effective action becomes quite nontrivial.
In this article, we shall focus on the supersymmetry (SUSY), considered to be the
most powerful among the ones listed above. In fact it has been claimed that a large
degree of SUSY, N = 16, present in the theory imposes strong restrictions on the form
of the effective action and leads to a number of non-renormalization theorems [1217].
However, upon close examinations one finds that such assertions in the existing literature
can still be challenged. This is essentially due to the lack of completely off-shell
consideration. As we shall elaborate in some detail in Section 3, to understand precisely
to what extent SUSY is responsible in determining the effective action, one must allow
the background fields to have arbitrary time-dependence. This in turn inevitably leads to
the necessity of examining the gauge (BRST) symmetry, another important ingredient of
Matrix theory, as SUSY and gauge symmetry are known to be intimately intertwined offshell. Such an analysis has not been performed in the past.
What makes the off-shell analysis difficult is that we do not as yet have the off-shell
unconstrained superfield formulation in the case of N = 16 supersymmetry. This prompts
us to resort to the conventional means, namely the Ward identity for the symmetries in
question. In order to be able to check against the available explicit result for the effective
action, one would like to obtain the Ward identity in so-called the background gauge,
in which all the calculations have been performed. Although the derivation of such a
Ward identity is expected to be a text-book matter, this was not to be: one must carefully
disentangle the dependence on the background field of the effective action by making use
of BRST Ward identities. The result is a somewhat complicated Ward identity, in which the
supersymmetry and the BRST symmetry are intertwined. A notable feature of our Ward
identity is that one can read off the effective quantum-corrected SUSY transformations in
closed form and this should serve as a starting point of various truly off-shell investigations.

Y. Kazama, T. Muramatsu / Nuclear Physics B 584 (2000) 171196

173

As a simple application, we compute the transformation laws to the lowest order in


the derivative expansion at 1-loop and analyze the restriction imposed by SUSY on the
effective action for a background with arbitrary time-dependence, to the corresponding
order. We find that at this order the effective action is indeed fully determined by the
requirement of supersymmetry, which agrees with the explicit calculation [18] including
the normalization.
The rest of the paper is organized as follows: in Section 2, we recall the symmetries of the
original action of Matrix theory and the BRST symmetry associated with the background
gauge fixing. Then in Section 3 we make some important remarks on the determination of
the effective action and its symmetries, to emphasize the necessity of completely off-shell
analysis. Having clarified the issue, we proceed in Section 4 to the derivation of the SUSY
Ward identity for the effective action in the background gauge. By carefully separating the
different origins of the dependence on the background field, we obtain the desired Ward
identity together with the closed expressions for the effective SUSY transformation laws.
This constitutes the main result of this work. As an application, we explicitly work out in
Section 5 the SUSY transformation laws to the lowest order in the derivative expansion
at 1-loop and show that the effective action to the corresponding order is completely
determined by the Ward identity. Section 6 is devoted to a short summary and discussions
of future problems.

2. Action and its symmetries


Let us begin by recalling the action of the Matrix theory and its symmetries, which at
the same time serves to set our notations.
The basic action of the Matrix theory can be written as

Z
1
g2
S0 = Tr dt [Dt , Xm ]2 + [Xm , Xn ]2
2
4

i T
g T m
+ [Dt , ] + [Xm , ] ,
(2.1)
2
2
(2.2)
Dt = t igA,
where N N hermitian matrices Xijm (t), Aij (t), ,ij (t) stand for the bosonic, the gauge,
and the fermionic fields, respectively. The middle Latin indices m, n, . . . , running from 1
to 9, denote the spatial directions, while the Greek letters such as , = 1 16 are used
for the SO(9) spinor indices. The 16 16 -matrices m are real symmetric and satisfy
{ m , n } = mn .
To facilitate the quantum computations, it is convenient to define the theory by going to
and the
the Euclidean formulation. Introduce the Euclidean time , the gauge field A,
2
action e
S0 by
it,

A iA,

2 Fermions are not transformed.

e
S0 iS0 .

(2.3)

174

Y. Kazama, T. Muramatsu / Nuclear Physics B 584 (2000) 171196

Then the action and the covariant derivative become



Z
1
g2
e
S0 = Tr d [D , Xm ]2 [Xm , Xn ]2
2
4

1 T
1
T m
+ [D , ] g [Xm , ] ,
2
2

D = ig A.

(2.4)
(2.5)

Besides the obvious Spin(9) symmetry, this action is invariant under the following
transformations:
1. Gauge transformations with a gauge parameter matrix :
A = [D , ],

Xm = ig[, Xm ],

= ig[, ].

(2.6)

2. Supersymmetry transformations with a spinor parameter  :


 Xm = i T m ,
 A =  T ,


g
 = i [D , Xm ] m + [Xm , Xn ] mn ,
2

(2.7)
(2.8)

where mn 12 [ m , n ] is real anti-symmetric. When the system possesses symmetries


other than the supersymmetry, the supersymmetry algebra may close not only on the space
time translation but also on the generators of such additional symmetries. In fact in the
present case it is well-known that it involves the gauge symmetry with a field-dependent

gauge function. For example on A,

[ , ]A = 2i T A + A,

(2.9)


where = 2i  Xm  A ,
T m

(2.10)

and similarly for the other fields. Moreover, for a system with N = 16 supersymmetry,
such as the Matrix theory, formulation in terms of unconstrained superfields is not known
and hence the algebra closes only up to the equations of motion in general. Thus it is
expected that the proper understanding of the supersymmetry of the Matrix theory must
include the analysis of these nontrivial features.
3. Generalized conformal transformations: if we rescale the fields Xm and A by a factor
S0 is invariant
of g, such as Xm Xm /g, and allow g to depend on to a linear order, e
under a generalization of the conformal transformations [8,10]. In particular, the invariance
under the special conformal transformation defined by

K A = 2 A,

K Xm = 2 Xm ,

K =  ,

K g = 3g,

K = 0,
(2.11)

 = an infinitesimal bosonic parameter,


imposes a useful restriction on the form of the effective action.
Besides these well-established symmetries, the remarkable agreement between the 11dimensional supergravity calculations and the 2-loop Matrix theory calculations for multibody scattering processes [3] strongly suggests that the Matrix theory actually possesses
11-dimensional Lorentz symmetry in a highly nontrivial manner.

Y. Kazama, T. Muramatsu / Nuclear Physics B 584 (2000) 171196

175

In this article, we shall focus on how the first two of these symmetries, which
are intimately intertwined, are implemented in the quantum effective action of the
supergravitons. In the M-theory interpretation of the Matrix theory, the coordinates and
the spin degrees of freedom of these supergravitons are represented by the diagonal
backgrounds for Xm and , respectively. We shall denote them by Bm and ,
respectively, and separate them from the quantum parts Ym and as
1
Bm + Ym ,
(2.12)
g
1
(2.13)
= + .
g
As was already emphasized in the introduction and will be further elaborated in the next
section, it is important to take Bm ( ) and ( ) as arbitrary backgrounds, not satisfying
any equations of motion. Only in this way we can unambiguously determine how much
restrictions are imposed by the supersymmetry on the effective action for these background
fields.
To quantize the theory, we need to fix the gauge. Although our derivation of the Ward
identity, to be presented in Section 4, can be readily adapted to any choice of gauge, the
actual computations are extremely cumbersome except in the standard background gauge.
It is specified by the gauge-fixing function of the form


(2.14)
G = A + i B m , Xm .
Xm =

In fact essentially all the existing explicit calculations have been performed in this gauge.
However, as it will become clear, the naive use of this gauge leads to a subtle but
important complication in deriving the correct Ward identity. To avoid this problem, we
em in place of Bm and write the gauge-fixing
will tentatively use a different function B
function as
 m

e , Xm .
e = A + i B
(2.15)
G
em = Bm .
Later at an appropriate stage, we will set B
The corresponding ghost action can be readily obtained by the standard BRST method.
The BRST transformations for the quantum part of the fields are given by
B A = [D , C],

B Ym = ig[Xm , C],

B = ig{C, },
B C = igC ,
2

(2.16)
B C = ib,

B b = 0.

C, C and b are, respectively, the ghost, the anti-ghost and the NakanishiLautrup auxiliary
fields. The background fields are not transformed. Then the combined gauge-ghost action
e
Sgg is generated by

 
Z
1
e
e b .
(2.17)
Sgg = B Tr d C G
i
2
We will henceforth set the gauge parameter to be 1. This leads, after integrating over the
b field, to the familiar gauge-ghost action

176

Y. Kazama, T. Muramatsu / Nuclear Physics B 584 (2000) 171196

 m

1
e , Xm 2
d A + i B
2
Z
 m

[D , C] g C B
e , [Xm , C] .
i Tr d C

e
Sgg = Tr

(2.18)

3. Remarks on the determination of effective action and its symmetries


Before starting the derivation of the off-shell supersymmetry Ward identity, we wish
to make some important remarks on the determination of the effective action and its
symmetries, which point to the necessity of off-shell analysis. Although many of the
remarks will apply for general backgrounds, for clarity of discussions we shall consider the
so-called source-probe configuration. In the case of U (N + 1) gauge group, it is defined
as the situation where a probe supergraviton interacts with N supergravitons situated at the
origin that act as a heavy source. The background fields representing this situation are
Bm = diag(rm , 0, 0, . . . , 0),

= diag( , 0, 0, . . . , 0).

(3.19)

Here, rm represents the coordinate of the probe and its spin content. As usual, the spin
of the source is neglected.
As already emphasized in the introduction, the primary feature of Matrix theory is that
it generates the supergravity interactions among various objects only after (i) introducing
the corresponding backgrounds and (ii) integrating over the quantum fluctuations around
them. Because of this, realizations of the symmetries of the effective theory are in general
modified nontrivially. Besides, being an off-shell quantity, the form of the effective action
is affected by (A) the gauge choice, (B) the (re)definition of the fields, and (C) the freedom
of adding total derivatives. Of course the on-shell S-matrix elements do not depend on these
factors. However, the determination of the full (i.e., quantum-corrected) on-shell condition
itself requires the knowledge of the off-shell effective action. 3 Thus, in order to understand
the symmetry structure of the effective theory fully, it is necessary to perform an off-shell
analysis with (A) (C) properly taken into account.
Now since an exact analysis is practically impossible, one often needs to make some
approximations. In doing so, one must make sure that they are logically consistent for ones
aim. For the present purpose, some of the often used approximations are not appropriate.
For example, the eikonal approximation, where one tries to reconstruct the effective action
from the eikonal phase shift, can be dangerous and misleading. In fact the answer depends
on the form of the effective Lagrangian assumed. As a simple illustration, consider the
1-loop eikonal phase shift [19] for v  1 given by
v3
v5
3 v7
+ 0 10
+ O(v 9 ),
6
b
b
2 b14
rm = vm + bm , v b = 0, bm = impact parameter.

e1e =

3 This was clearly demonstrated in [18], where the agreement between the supergravity and the Matrix theory
calculation was achieved with the recoil corrections.

Y. Kazama, T. Muramatsu / Nuclear Physics B 584 (2000) 171196

177

If one assumes the effective Lagrangian to be of the form L = L(v, r), then the effective
action that reproduces this phase shift is uniquely determined to be


Z
15 v 4
v6
9009 v 8
10
e
+ 0 11
+ O(v ) .
S1 = d
16 r 7
r
4096 r 15
However, even restricting to O(v 4 ), the most general form allowed for the effective action
contains 6 independent structures, after eliminating total-derivative ambiguities:

Z
v4
v 2 (v r)2
(v r)4
e
+
C
S1 = d A 7 + B
r
r9
r 11

2
2
v (a r)
(a r)
a2
+D
+E
+F 5 .
r7
r7
r
On the other hand, there is only one condition, (15/16) = A + (1/7)B + (1/21)C,
required for the correct phase shift, and hence 5 parameters remain undetermined. The
situation at O(v 6 ) is even more striking. Although the eikonal phase shift vanishes, the
explicit computation reveals that there are 5 non-vanishing independent structures present
in the effective action [20].
It should be clear from these illustrations that the only logically consistent procedure, not
affected by the total-derivative ambiguities, is to use off-shell backgrounds with arbitrary
-dependence 4 and to classify terms by derivative expansion according to the order
defined by
1
order = # of + # of fermions.
2

(3.20)

If necessary, one may combine this with the usual loop expansion.
Having emphasized the importance of off-shell considerations, we now make some
related comments on the general arguments on the restrictions imposed by supersymmetry,
often referred to as SUSY non-renormalization theorems. They can be roughly classified
into two categories.
The first type of argument, devised by Paban et al. [12], relies on the closure property of
SUSY transformations. For example, at O(v ) they first make a choice of the definition
R
of the fields so that the action takes the form df (r)v 2 and take the O(v 0 ) SUSY
transformation laws in that basis to be the standard ones without any correction,  r m =
i m ,  = i(/v) . Then demanding that the closure is canonical, namely [ , ] =
2T  , they show that there cannot be a correction to the transformation laws at O(v2 )
and hence the O(v 2 ) effective action is tree-exact. Although the argument is quite simple
and plausible, it is unclear why the closure should be canonical off-shell and further it is
not obvious if the O(v 0 ) transformation laws must be of the standard form in a particular
basis adopted. In general, field redefinitions affect the form of the SUSY transformations
and hence they must be considered as a pair.
4 This was emphasized in the context of generalized conformal symmetry in [21]. Related discussion can also
be found in [18].

178

Y. Kazama, T. Muramatsu / Nuclear Physics B 584 (2000) 171196

The other type of argument is known as the SUSY completion method, which makes
use of the chain of relations produced among terms with different number of s by SUSY
transformations. For example, at O(v 4 ) one expects relations of the form
v4 v3 2 v2 4 v 6 8 .

(3.21)

By showing that the top form, 8 term in this case, is not renormalized beyond 1-loop, one
wishes to infer the non-renormalization of all the other terms in the chain, in particular
the v 4 term. This method appears efficient, but some care is needed in drawing firm
conclusions. One problem is that sometimes the chain starting from the top form stops
at an intermediate stage. Put differently, one may form a super-invariant not containing
An example already occurs at O(v 2 ), where the tree-level expression
Rthe top 2form.
2
2 )) is SUSY-complete without a 4 term. More nontrivial example

d ((v /2g ) + ( /2g


6
is seen at O(v ): although 12 term was shown to vanish [13] at 1-loop, the bosonic
contribution at O(v 6 ) nonetheless exists [20].
An attempt at filling this gap was made in [16]. In this work all the connections in the
chain (3.21) were examined and it was concluded that SUSY is indeed powerful enough to
fix the effective action at this order up to an overall constant. Again one must be cautious
in accepting this conclusion: in this analysis v and were taken to be -independent and
hence the assumed form of the effective Lagrangian was not the most general one allowed
in the proper derivative expansion with arbitrary backgrounds. Later it was recognized [18],
however, that the higher derivative terms neglected in this analysis can actually be absorbed
into the tree-level Lagrangian by a suitable field redefinition, which appeared to resurrect
the validity of the analysis made in [16]. Unfortunately, the problem still persists: by such
a field redefinition the higher derivatives are simply shifted into the SUSY transformation
laws and one must reanalyze the issue with such modifications.
Thus one sees that although the existing analyses are highly plausible they are not airtight. In view of the importance of precise understanding of the role of supersymmetry and
its connection with gauge symmetry, it is desirable to perform an unambiguous off-shell
analysis with arbitrary backgrounds. This motivates us to the study of the Ward identity, to
be described in the next two sections.

4. SUSY Ward identity for the effective action in the background gauge
Having argued the importance of off-shell analysis for arbitrary trajectories, we shall
now derive the SUSY Ward identity for the effective action e in the standard background
gauge, (2.14), used exclusively in the actual computations.
To make use of the well-established method, let us further split the quantum fluctuations
Ym and into two parts, the diagonal and the off-diagonal, in the manner
Ym,ij =

ym,i
bm,ij ,
ij + Y
g

,ij =

,i
b,ij ,
ij +
g

and introduce the sources only for the diagonal fields:

(4.22)

Y. Kazama, T. Muramatsu / Nuclear Physics B 584 (2000) 171196

e
Ss =

179

d (Jm,i ym,i + ,i ,i ).

(4.23)

The Euclidean generating functionals are defined by


Z
e [J, ]),
e ] = D exp(e
Z[J,
Stot ) = exp(W

(4.24)

e
S0 + e
Sgg + e
Ss ,
Stot e

D DADY
D DCDC,

e [J, ] is the one for the connected functions. By making the change of integration
where W
variables corresponding to the supersymmetry transformations, one obtains the primitive
form of the Ward identity
Sgg i + h e
Ss i.
0 = h e

(4.25)

Here and in what follows, hOi for an operator O means


R
e
DOeStot
.
hOi = R
Stot
Dee

(4.26)

We now rewrite this identity (4.25) in terms of the generating functional e, which is
1PI (1-particle-irreducible) with respect to the diagonal fields. Define as usual the classical
fields and e by
ym,i( )

e
W
,
Jm,i ( )

e[y, ] W [J, ]

,i ( )
Z

e
W
,
,i ( )

d (Jm,i ym,i + ,i ,i ).

(4.27)
(4.28)

Then the sources are expressed in terms of e as


Jm,i ( ) =

e
,
ym,i ( )

,i ( ) =

e
.
,i ( )

(4.29)

Therefore, the contribution to the Ward identity from the variation of the source action can
be written as


Z
e
e
h ym,i ( )i +
h ,i ( )i .
(4.30)
Ss i = d
h e
ym,i ( )
,i ( )
Sgg i from the gauge-ghost part, a direct calculation yields
As for the contribution h e
a rather complicated expression, which constitutes an inhomogeneous term in the Ward
identity regarded as a functional integro-differential equation for e. This is undesirable
since what we wish to understand is how the supersymmetry acts on the effective action e.
Fortunately, it was noted long ago [22] in the context of four-dimensional super Yang
Mills theory that one can reexpress such a term in a form similar to (4.30). The first step
is to note that the supersymmetry transformations (2.7) and (2.8) commute with the BRST
transformation (2.16) on all the fields, as can be checked straightforwardly. Therefore,
Sgg i can be written as
starting from (2.17), h e

180

Y. Kazama, T. Muramatsu / Nuclear Physics B 584 (2000) 171196

*


Z
1 e
e
C G
h Sgg i =  B Tr d
i
*


Z
1 e
C G
= B  Tr d
i
where
1
O Tr
i

1
b
2
1
b
2

+
+
= hB O i,

e
d C  G

(4.31)

(4.32)

is a fermionic composite operator. This expression, being an expectation value of a BRSTexact form, vanishes in the ordinary vacuum. However, in the presence of external sources,
it becomes proportional to the sources, and hence to the functional derivatives of e. Let us
collectively denote by and J the basic fields and the corresponding sources, respectively,
and consider the generating functional with a source j for the operator O :
Z
(4.33)
Z[J, j ] = D e(Stot +J +j O ) .
Now make a change of variables corresponding to the BRST transformation. We get
Z

(4.34)
0 = D (1)|| J B j B O e(Stot +J +j O ) ,
where || is 0 (1) if is bosonic (fermionic). By differentiating with respect to j once,
setting j = 0, and then expressing the source J in terms of e, one easily obtains the
following BRST Ward identity:
Z
e
hB ( )O i.
(4.35)
hB O i = d
( )
In this way, we get

Z
e
h Sgg i = d


e
e

hB ym,i ( )O i +
hB ,i ( )O i .
ym,i ( )
,i ( )

(4.36)

Putting all together, we arrive at the following SUSY Ward identity expressed solely in
terms of the derivatives of e:

Z

e
h ym,i ( )i + hB ym,i ( )O i
0 = d
ym,i ( )


e

(4.37)
h ,i ( )i hB ,i ( )O i .

,i ( )
Normally it is now a simple matter to convert this into the desired Ward identity for
the effective action as a functional of the backgrounds B m and : one would rewrite the
derivatives with respect to ym and into those with respect to B m and and then set
ym = = 0. This procedure is indeed valid for the fermions since and always
appear in the combination + in the original action and it is a simple matter to prove
that this gets converted to + in e.

Y. Kazama, T. Muramatsu / Nuclear Physics B 584 (2000) 171196

181

This is not so for the bosonic field. While most of the B m dependence comes from
bm , where Bm and ym appear together, B m in the
the splitting Xm = (1/g)(Bm + ym ) + Y
gauge-ghost sector (in the standard background gauge) is not accompanied with ym . From
the point of view of the Ward identity above, it is an independent extra parameter field
which should be distinguished from the bona fide background field. This is why we chose
em for this field.
to start out with a different symbol B
Now the problem we face is the following. In order to be able to apply the Ward identity
em
to the case of the standard background gauge, we need to express the dependence on B
m
again in the form of the functional derivative of e with respect to B . In other words,
we must disentangle the two different types of B m dependence buried in the standard
background gauge formulation and construct the Ward identity which takes both types
of dependence into account. Although this is a technical rather than a conceptual problem,
it is again a manifestation of the gauge theory nature of the Matrix theory, which has often
been neglected.
em appears only in the gauge-ghost action
The problem can be solved as follows. Since B
and the variation with respect to it commutes with the BRST transformation, we get from
(2.17)
e
Sgg
= B Om,i ( ),
m,i
e
B ( )

(4.38)

where
Om,i =


C j i Ym,ij C ij Ym,j i ,

(4.39)

em . The expectation value of the left-hand side


em,i stands for the diagonal elements B
and B
ii
e as W
e / B
em,i , which is equal
can be expressed in terms of the generating functional W
e
e
to / Bm,i since it is a parameter field. On the other hand the expectation value of the
right-hand side can be treated in exactly the same way as we treated B O . In this way we
get the identity

Z
e
e
0
hB yn,j ( 0 )Om,i ( )i
= d
e
yn,j ( 0 )
Bm,i ( )

e
0

h
(
)O
(
)i
.
(4.40)

+
B ,j
m,i
,j ( 0 )
Now let us replace ym and with Bm and respectively and then set ym = = 0
em
in the previous Ward identity (4.37) and in the relation (4.40) above. In the limit B
Bm , the total variation of e with respect to Bm , which we denote by e/Bm to avoid
confusion, is


e
e
e
=
+
.
(4.41)
em,i ( ) B=B
Bm,i ( )
Bm,i ( ) B
e
Substituting (4.40) we then get

182

Y. Kazama, T. Muramatsu / Nuclear Physics B 584 (2000) 171196

e
=
Bm,i ( )

e
Tnj ;mi (, 0 )
Bn,j ( 0 )
Z
e
hB ,j ( 0 )Om,i ( )i,
d 0
,j ( 0 )
d 0

(4.42)

with
Tnj ;mi (, 0 ) mn ij ( 0 ) hB yn,j ( 0 )Om,i ( )i.

(4.43)

By inverting this relation, we can express the partial variation e/Bm in terms of the total
variation e/Bm :
Z
e
e
1
= d 0 Tmi,nj
(, 0 )
Bm,i ( )
Bn,j ( 0 )
Z
Z
e
1
hB ,k ( 00 )On,j ( 0 )i, (4.44)
(, 0 ) d 00
+ d 0 Tmi,nj
,k ( 00 )
where T 1 is the inverse of T .
Finally, the correct Ward identity in the standard background gauge is obtained by
em Bm is taken, the total
substituting this expression into (4.37). Once the limit B
variation e/Bm can be identified with e/Bm , which denotes the usual functional
derivative of the effective action computed in the standard background gauge (i.e., with
Bm in the gauge-fixing term.) With this understood, the result can be put in the desired
form


Z
e
e
+  ,i ( )
,
(4.45)
0 = d  Bm,i ( )
Bn,j ( )
,i ( )
where the effective SUSY transformation laws are given by
Z

1
( 0 , ) h yn,j ( 0 )i + hB yn,j ( 0 )O i ,
 Bm,i ( ) = d 0 Tmi,nj

(4.46)

 ,i ( ) = h ,i ( )i hB ,i ( )O i
Z
1
( 00 , 0 )hB ,i ( )On,j ( 0 )ih ym,k ( 00 )
d 0 d 00 Tmk,nj
+ B ym,k ( 00 )O i.

(4.47)

(As said before y = = 0 is understood.) This is the main result of this section.
Note the following features:
We have succeeded in putting the Ward identity in the form where the supersymmetry
transformation laws for the effective action are cleanly identified in closed forms.
As expected, the supersymmetry and the gauge (BRST) symmetry are nontrivially
intertwined. Naively, one might expect that the effective transformation laws are obtained
as the expectation values of the original transformation laws (2.7) and (2.8). In our
notation, they are represented by h yn,j i and h ,i i in  Bm,i and  ,i , respectively.
The actual transformation laws, (4.46) and (4.47), are much more complicated. One
can see, however, that the corrections to the naive laws all involve B , i.e., the BRST

Y. Kazama, T. Muramatsu / Nuclear Physics B 584 (2000) 171196

183

transformation. Since the quantization of the system inevitably requires a gauge fixing,
this is a universal feature, not special to the standard background gauge adopted
here.
As has already been remarked, the transformation laws derived above are exact, albeit
somewhat formal at this stage. In particular, there is no inherent distinction between the
tree level contribution and the quantum corrections. Thus it is far from obvious that the
anti-commutator of the effective transformations would close solely on the translation
generator, as it does at the tree level: We have so far not been able to produce a proof.
In the next section, we will carefully examine the structure of this Ward identity at the
1-loop level and draw implications.

5. Explicit calculations and implications


We now compute the effective SUSY transformation laws explicitly and study the
implications of the Ward identity.
5.1. Source-probe situation
Although the Ward identity derived in the previous section is valid for any background,
we shall restrict ourselves to the source-probe situation, since the existing calculations of
the effective action itself, to be compared later, are more complete for this configuration.
The background fields representing this situation for the gauge group U (N + 1) were
already described in (3.19), which we display again for convenience:
Bm = diag(rm , 0, 0, . . . , 0),

= diag( , 0, 0, . . . , 0).

(5.48)

rm represents the coordinate of the probe and its spin content. We shall denote the time
derivative of the coordinate by vm rm .
To facilitate the computations, it is convenient to introduce the following notations: for
any matrix U we define UI U1I , UI UI 1 , where I = 2 N + 1. (The symbol here
does not stand for complex conjugation.) We shall call such index I the off-diagonal matrix
vector index. Then the 11 component of a product of two matrices becomes (U V )11 =
PN+1

I =2 UI VI , which we abbreviate as U V .
With this convention, the basic quantities appearing in the Ward identity take the
following forms:
(5.49)
h ym i = i T m ,


2
g
m
mn
hA Ym A Ym i +
hYm Yn Ym Yn i  ,
h i = i v/ ig 2
2
(5.50)
B ym = ig 2 (Ym C Ym C),

B = ig 2 C C ,

Z

1
C11 + + C ( + /r)
O = i d
g

(5.51)
(5.52)

184

Y. Kazama, T. Muramatsu / Nuclear Physics B 584 (2000) 171196


+ C ( /r) + C I J J I ,
Om = C Ym C Ym .

(5.53)
(5.54)

5.2. One-loop at order two


In this article, we shall perform the calculation of the simplest nontrivial contributions to
the effective transformation laws, namely those which govern the order 2 part of the 1-loop
effective action. Here the order is defined as in (3.20), i.e., the number of derivatives plus
twice the number of fermions. Since the tree action is already of order 2, this means that we
need to compute  rm and  to order 0 at 1-loop, where the transformation parameters
 are considered to be of order 1/2.
Such contributions are further classified according to the number of s involved.
Because rm and are of order 0 and 1/2, respectively, we need to consider  rm to
linear order in , while for  we need 2 contributions as well. In what follows, it
is more convenient to refer only to the number of additional s relative to the tree level
contribution. For instance, a correction to O( 0 ) for  rm refers to a term of the form
f (r).
At 1-loop order, since the terms involving the BRST variation B start only at 1-loop,
the expressions (4.46) and (4.47) for  rm and  simplify to
Z
1
1
m
m
 r ( ) = ( )( ) + d 0 hB ym ( )On ( 0 )i ( n )( 0 )
i
i
+ hB ym ( )O i,
 ( ) = h ( )i hB ( )O i
Z
1
d 0 hB ( )Om ( 0 )i ( m )( 0 ).
i

(5.55)

(5.56)

The expectation values of the composite operators themselves simplify considerably


at this order. It is easy to check that at 1-loop only the 2-point functions contribute.
Moreover, since there are no mixing between the fields with different off-diagonal
matrix vector indices in such propagators, we always have the structure hUI ( )VJ ( 0 )i =
I J hU ( )V ( 0 )i, where U and V can be regarded as single-component fields. I J s then
contract to produce a factor of N , which goes with the coupling g 2 . With these remarks,
the relevant multi-body expectation values become


2
m

A( )Ym ( ) A ( )Ym ( )
h ( )i = i(/v) ( ) + g N


i mn

(5.57)
+ Ym ( )Yn ( ) Ym ( )Yn ( )  ( ),
2
Z
0 )i( + /r) h ( 0 )Ym ( )i
hB yn ( )O i =  g 2 N d 0 hC ( )C(

hC( )C ( 0 )i( /r) h ( 0 )Ym ( )i ,

(5.58)

Y. Kazama, T. Muramatsu / Nuclear Physics B 584 (2000) 171196

hB ( )O i =  g 2 N

Z
g

185

0 )i( + /r) ( 0 ) ( )
d 0 hC ( )C(

hC( )C ( 0 )i( /r) ( 0 ) ( )

( 0 )
b
d 0 C ( )C 11

( 0 ) ,
b
d 0 C ( )C 11


)i
hB yn ( 0 )Om ( )i = ig 2 N Ym ( 0 )Yn ( ) h(C ( 0 )C(



0
+ Ym ( )Yn ( ) h(C( 0 )C ( )i ,
+ g

(5.59)

(5.60)

where stands for the derivative with respect to . We will now evaluate these expressions
to the relevant order by perturbation theory.
5.2.1. Corrections at O( 0 )
Let us begin with the corrections at O( 0 ). First we need to compute the tree-level
propagators. The part of the Lagrangian quadratic in the fields without is of the form

L = Ym ( 2 + r 2 )Ym + A ( 2 + r 2 )A 2ivm A Ym A Ym
1
+ ( + /r)
+ 2
2g
+ i C ( 2 + r 2 )C + i C ( 2 + r 2 )C + i C 11 ( 2 )C11 .
(5.61)
Following the remark already made on the trivial dependence on the off-diagonal matrix
and C, C in L as if they were singlevector indices, we may treat the fields Ym , A,
component. The propagators for the massless fields and C11 , C 11 turn out to be needed
only at O( 2 ), to be discussed in the next subsection.
The simplest is the ghost propagator, which can be directly read off as
)C ( 0 )i = hC ( )C( 0 )i
hC(
0 )i = hC( )C ( 0 )i = ih || 0 i,
= hC ( )C(

(5.62)

where
( 2 + r 2 )1 .
The propagators for the system are given by

| 0 i,
( ) ( 0 ) = h |D


+
| 0 i,
( ) ( 0 ) = h |D

(5.63)

(5.64)
(5.65)

where
D ( /r)1 .

(5.66)

From the relation ( /r)( /r) = ( 2 + r 2 )(1 /v ), we can expand D in powers


of the velocity v( ) as

186

Y. Kazama, T. Muramatsu / Nuclear Physics B 584 (2000) 171196

D = ( /r)(1 /v)1 = ( /r)( /v + ).

(5.67)

To the order of interest, we will only need the v-independent part.


Now as for the Ym A system, we have a mixing term with an arbitrary coefficient
function vm ( ) and it can only be resolved in the derivative expansion. Expanding in
powers of the mixing term, we readily obtain
hY m ( ) Y n ( 0 )i = mn h || 0 i + O(v 2 ),
)A ( 0 )i = h || 0 i + O(v 2 ),
hA(

)Ym ( 0 ) = h |2ivm | 0 i + O(v 3 ).


Ym ( )A ( 0 ) = A(

(5.68)

The only other 2-point functions appearing at this order are h ( )Ym ( 0 )i and
h ( )Ym ( 0 )i, which can be readily computed by inserting the vertices of the form

m
Ym Ym .
(5.69)

The results are

( m ) | 0 i,
( )Ym ( 0 ) = h |D


+
( m ) | 0 i.
( )Ym ( 0 ) = h |D

(5.70)
(5.71)

With this preparation, the calculations of the various expectation values in the Ward
identity can be performed efficiently to the desired order with the use of the so-called
normal ordering method developed in [20]. The essence of this method is to first
rearrange the order of a product of various operators and functions into the standard form
f ( ) m n , using the commutation relations such as [, ] = 2(r v) etc., and
then use BakerCampbellHausdorff and the Gaussian integration formulas to evaluate
it. A useful list of formulas so obtained are collected in the appendix of [23]. Below, we
shall give a sample calculation of this sort and then simply list the results for the needed
expectation values.
As an example, let us consider hB yn ( )O i given in (5.58). Substituting the expressions
for the various propagators already computed, it becomes
hB y n ( )O i = i T g 2 Nh |[( + /r)D + + ( /r)D ] n | i
= 2i T g 2 Nh |( 2 + r 2 ) n | i.

(5.72)

To the order of interest, the normal-ordering is trivial and we get


hB yn ( )O i = 2i Tg 2 Nn h |2 | i = i T n

1 g2 N
.
2 r3

(5.73)

In a similar manner, we obtain the following results:


Z

hB yn ( 0 )Om ( )i = 2g 2 Nnm h 0 || i2 ,
1
1 g2 N
,
d 0 hB ym ( )On ( 0 )i ( n )( 0 ) = im
i
2 r3
hB ( )O i = 0,
hB ym ( )O i = im

g2 N

1
,
2 r3

(5.74)
(5.75)
(5.76)
(5.77)

Y. Kazama, T. Muramatsu / Nuclear Physics B 584 (2000) 171196

N vm
,
A Ym A Ym =
i r3

Ym Yn Ym Yn = 0.

187

(5.78)
(5.79)

Making use of these formulas, we find the effective SUSY transformation laws at O( 0 )
to be


g2 N
(5.80)
 r m = i T m 1 + 3 ,
r


g2 N
(5.81)
 = i/v 1 3 .
r
Already at this order, there are nontrivial corrections to the tree level laws. For  the
entire correction came from the expectation value of the non-linear part of the original
Ym ] m contained in [D , Xm ] m . On
SUSY transformation law, namely from ig[A,
the other hand, the same procedure is not applicable for  rm : if one naively took the
expectation value of the basic transformation law  Xm = i m , one would not get
any corrections to all orders. What we want is the effective SUSY transformation operating
on the effective action e and this can only be obtained through the analysis of the Ward
identity, as we have done. One can see from the calculations outlined above that exactly half
Sgg i and the other half was produced
of the quantum correction for  r m came from h e
from the procedure of taking into account the dependence on the extra parameter field we
em .
originally called B
5.3. Corrections at O( 2 )
Now we move on to the corrections at O( 2 ). Since 2 is of order 1, we need to compute
such contributions only for  . The procedure is entirely similar to the O( 0 ) case but
the calculations are more involved and we relegate the details to the Appendix A. As shown
there, 7 types of diagrams contribute. Two of them, diagrams (B2-a,b), involve genuine 3
point vertices as well as massless propagators given by
1
1
hC11 ( )C 11 ( 0 )i = ih | 2 | 0 i.
(5.82)
h ( ) ( 0 )i = g 2 h | | 0 i,

They are singular in the infrared, but such singularities cancel in the end result. After some
calculations, we find the O( 2 ) corrections to  to be
2

 =

3ig 2 N
rl ( mnl )( nm ) + 2rl ( ml )( m ) 4rl ( m )( ml )
16r 5

+ 4( )(/r ) 4(/r ) .
(5.83)

Although it appears quite complicated, it can be drastically simplified with the use of the
SO(9) Fierz identities [5] described in the Appendix B. The relevant identity is
0 = rl ( mnl )( nm ) + 2rl ( nl )( n ) + 4rl ( n )( nl )
+ 4rl ( )( l ) + 12rl ( l ) .
Applying this to (5.83), we get

(5.84)

188

Y. Kazama, T. Muramatsu / Nuclear Physics B 584 (2000) 171196

3ig 2 N
(/r ) .
(5.85)
r5
Let us summarize the results. To order 2 at 1-loop, the effective supersymmetry
transformations laws are found to be


g2 N
m
T m
(5.86)
 r = i 1 + 3 ,
r


g2 N
3ig 2 N
(/r ) .
(5.87)
 = i(/v) 1 3
r
r5
2

 =

5.3.1. Closure of the effective algebra and field redefinition


It is of interest to examine the closure property of the transformation laws obtained
above. By a simple calculation, one immediately finds
[1 , 2 ]rm = 2rm (2 1 ) + O(g 4 ),
[1 , 2 ] = 2 (2 1 ) + O(g 4 ).

(5.88)
(5.89)

Thus to the 1-loop order, the closure turned out to be precisely canonical.
As we remarked at the end of Section 4, this is not a feature guaranteed by the general
analysis. One way to appreciate this is to consider a field redefinition which makes the
form of  r m to be the same as the one at the tree level. On general grounds, the most
general form of  r m and  , to the order we are considering, are

1
 r m = ( m ) 1 + g 2 F (r) ,
i
 = i(/v) + g 2 G (r, ) .
From (5.90), the desired field redefinition can be read off as

1 + g 2 F (r) .

(5.90)
(5.91)

(5.92)

The transformation law for this new field is then



+ O(g 4 ).
 = i(/v) + g 2 i(/v) F (r) +  F (r) + G (r, )
It is quite nontrivial that the

O(g 2 )

(5.93)

part of this expression vanishes exactly.

5.4. Implication of the Ward identity


Having found the transformation laws, we are now ready to analyze the consequence of
the Ward identity on the structure of e.
Let us write the effective action up to 1-loop as e = e0 + e1 , the subscript denoting the
number of loops. The tree-level action is given by
 2

Z
v

+
.
(5.94)
e0 = d
2g 2 2g 2
As for e1 , it is easy to convince oneself that the most general structure at order 2, up to
total derivatives, is

Y. Kazama, T. Muramatsu / Nuclear Physics B 584 (2000) 171196

e1 =

189

 2
v
(v r)2
T
T /r
d A 3 + B
+C 3 +D 4
5
r
r
r
r


T mn rm vn
r l r k ( lm )( mk )
+E
+F
,
r5
r7

(5.95)

where A F are numerical constants. 5


Now we demand that  e vanish to the order of interest. The simplest way to proceed
is as follows. First, look at the O( 5 ) terms, which can only be produced from the last term
of (5.95) by the tree-level part of the transformation  rm . They read
Z

i
(5.96)
d F 9 r k 2r 2 ( l )( lm ) + 7r l r j ( j )( lm ) ( mk ).
r
It can be checked that the integrand does not vanish by any of the Fierz identities. 6 Thus
we find F = 0. Next, demand that the O( ) part of  e vanish. It is straightforward to
show that this reduces the allowed form of e1 to be
 2

Z
v
3A mn rm vn
,
(5.97)
e1 = d A 3 + (A + N) 3 +
2
r
r
r5
where A remains undetermined. Finally, look at the O( 3 ) part of  e. The contributions
arising from the tree level transformation of e1 are
3iN

rm ( m )( )
r5

3iA
+ 5 2rm ( m )( ) vn ( m )( mn ) rm ( n )( mn )
2r

15iA
+ 7 rm rn vl ( n )( ml ) + ( m )( nl ) .
4r
On the other hand, the 1-loop level transformation applied to e0 produces

(5.98)

3iN
rm ( m )( ),
(5.99)
r5
which cancels the first term of (5.98). The remaining terms in (5.98) are all proportional
to A. Now note that while the four-fermion structures in the second line of (5.98) have
only one free index, contracted with an arbitrary vector rm ( ) or vm ( ), the ones in the
last line carry three free indices. Since the Fierz identities can only relate structures with
the same number of free indices, expressions in these two lines cannot cancel each other.
Moreover, it is easy to check, using the Fierz identities given in the Appendix A, that the
second line does not vanish by itself. This then proves A = 0.
In summary, we have found that the order 2 contribution at 1-loop for the effective action
in the background gauge is completely determined by the requirement of supersymmetry
and takes the form

5 The structure of the form r l r k ( lmn )( mkn ) can be expressed in terms of the last term in (5.95) via
a Fierz identity and hence can be omitted.
6 We have also checked this numerically.

190

Y. Kazama, T. Muramatsu / Nuclear Physics B 584 (2000) 171196

e1 =

Z
d N


.
r3

(5.100)

This indeed agrees, including the overall normalization, with the direct calculation
performed in [18]. It can be easily checked that the normalization is directly linked to
the magnitude of the O( 0 ) quantum corrections in (5.86) and (5.87), which cannot be
determined by the closure property alone.
As we shall discuss in the concluding section, the power of our off-shell Ward identity
can only be fully utilized at the next order in the derivative expansion, where the most
general form of the action unavoidably contains many terms with higher derivatives, such
as rm and . Nevertheless, it is gratifying that already at order 2 it has enabled us to see
explicitly how the supersymmetry and the gauge symmetry intimately work together to
dictate the form of the effective action.

6. Summary and discussions


In this paper, we have derived the exact supersymmetric off-shell Ward identity for
Matrix theory as a step toward answering the old yet important unsettled problem:
to what extent do the symmetries, in particular the supersymmetry, determine the lowenergy effective action of Matrix theory? Our work was motivated by the observation
that the existing analyses are incomplete in that off-shell trajectories with arbitrary timedependence have not been fully considered. An important aspect of our Ward identity is
that it allows the quantum-corrected effective supersymmetry transformation laws to be
directly identified in closed form. They exhibit an intricate interplay with the gauge (BRST)
symmetry of the theory, a feature not properly appreciated previously.
As an application, we computed the explicit form of these transformation laws at 1-loop
to the lowest order in the derivative expansion, and examined if the invariance under them
determines the form of the effective action to the corresponding order. We found that the
answer is affirmative, confirming the earlier result [12]. This is as expected since at this
order the higher derivatives, such as the acceleration etc., can be eliminated from the
effective action by partial integration and the analysis is essentially the same as in the
existing literature.
The full significance of our off-shell Ward identity should become apparent starting
from the next order, i.e., from order 4, where complete elimination of higher derivatives
will no longer be possible. There will be a considerable number of independent structures
allowed in the most general effective action. Even the proper listing of them requires
careful analysis due to the total derivative ambiguities and the existence of nontrivial
Fierz identities. Nonetheless, a preliminary investigation indicates that, with an aid of
computerized calculation, it appears feasible to determine whether SUSY alone is enough
to fix the form of the effective action at order 4 for arbitrary trajectories.
Another important direction into which to extend our present work is to apply our
Ward identity to genuinely multi-body configurations. To find out whether the remarkable

Y. Kazama, T. Muramatsu / Nuclear Physics B 584 (2000) 171196

191

agreement with supergravity in such a situation [3] is due to supersymmetry alone would
certainly deepen our understanding of the Matrix theory further.
At the more formal and structural level, we should mention that a further study should
be made on the issue of the closure property of the effective SUSY transformations. As we
have shown, at the lowest order the closure turned out to be canonical. So far, however, we
have not been able to answer whether this persists at higher orders and loops. An analysis
based on the general closed form expressions for the transformation laws should shed light
on this intriguing question.
We hope to be able to report on these and other related issues in the near future.

Acknowledgement
We would like to express our special gratitude to Y. Okawa for a number of clarifying
discussions and his interest in our work. Y.K. acknowledges the warm hospitality extended
to him by the organizers at the third international symposium on Frontiers of Fundamental
Physics (Hyderabad, India), where a preliminary version of this work was presented. This
work is supported in part by Grant-in-Aid for Scientific Research on Priority Area #707
Supersymmetry and Unified Theory of Elementary Particles, Grant-in-Aid for Scientific
Research No. 09640337, and Grant-in-Aid for International Scientific Research (Joint
Research) No. 10044061, from Japan Ministry of Education, Science and Culture.

Appendix A. Calculations of O( 2 ) terms in 


In this appendix, we exhibit some details of the calculations of O( 2 ) terms in  at
1-loop order.
At this order, what we need to evaluate is (see (5.56))
 ( ) =  A ( ) +  B ( ) +  C ( ),

(A.1)

where
 A ( ) h ( )i,
 B ( ) hB ( )O i,
Z
1
C
 ( ) d 0 hB ( )Om ( 0 )i ( m )( 0 ).
i

(A.2)
(A.3)
(A.4)

Hereafter in this appendix,  will refer only to the O( 2 ) part. Also, we shall omit
the overall factor of N , except in the final expression. The relevant Feynman diagrams are
shown in Fig. 1.
Calculation of  A ( )
The explicit expression is

192

Y. Kazama, T. Muramatsu / Nuclear Physics B 584 (2000) 171196

Fig. 1. Schematic depiction of Feynman diagrams relevant for the calculations in this appendix.
2




m Ym A ( m ) + i g Ym Yn Yn Ym ( mn ) .
A = g 2 AY
2

(A.5)

To compute the expectation values above to O( 2 ), we need to perform the second order
perturbation using the O( ) vertices
Z

A ,
e
=
i
d A
(A.6)
SA

Z

e
(A.7)
SY = d ( m ) Ym Ym .
This generates the diagrams (A1) and (A2) in Fig. 1, for the first and the second term
)Ym ( )i. Using the tree-level
in (A.5), respectively. Consider for example the term hA(
propagators given in (5.64) and (5.68), this can be computed as
Z

)A ( 0 )i ( 0 )( n ( 00 )) h ( 0 ) ( 00 )i
)Ym ( ) = i d 0 d 00 hA(
A(
hYn ( 00 )Ym ( )i
= ih | ( + /r) ( m ) 2 | i.

(A.8)

Performing the normal-ordering, neglecting the terms which generate derivatives, this
becomes


)Ym ( ) = i( /r m )h |3 | i = 3i ( /r m ) = 3i rl ( ml ).
A(
16r 5
16r 5

(A.9)

Y. Kazama, T. Muramatsu / Nuclear Physics B 584 (2000) 171196

193

Likewise, one easily finds that hA ( )Ym ( )i gives exactly the same contribution. The
evaluation of the diagram (A2) proceeds in an entirely similar manner. In this way one
finds
A =


3ig 2
2rl ( ml )( m ) rl ( mnl )( nm ) .
5
16r

(A.10)

Calculation of  B ( )
This receives contributions from two classes of diagrams, (B1) and (B2) in Fig.1.
For (B1), we have
Z

B1
2
0 )i( + /r) ( 0 ) ( )
d 0 hC ( )C(
 ( ) =  g


(A.11)
hC( )C ( 0 )i( /r) ( 0 ) ( ) .
e
Insertions of the vertices e
SA
(A.6) and SY (A.7) twice generate the diagrams (B1-a) and
(B1-b). Again neglecting the derivatives produced in the process of normal ordering we
obtain

3ig 2
2( )(/r ) 2(/r ) 2rl ( m )( ml ) .
(A.12)
5
16r
Now, in distinction to all the other contributions, the one from (B2) involves propagation
of massless fields , C11 and C 11 in the intermediate steps. The original expression is, to
the order of interest,
Z
Z

B2
0

( 0 ) .

b
b
 = g d C ( )C11 ( ) + g d 0 C ( )C 11

B1 =

(A.13)
To extract O( 2 ) contributions, we need to use the following five types of vertices:
Z

m C11 ,
(A.14)
V1 ig d r m C C11 Y m + r m CY
Z
(A.15)
V2 d n ( Y m Y m ),
Z
C A C C A C ,
(A.16)
V3 g d C C 11 A + C C11 A
11
11
Z
),
(A.17)
V4 i d ( A A
Z
1
.
b
(A.18)
V5 2 d
g
First, inserting V1 , V2 and V5 , we get the diagrams of the type (B2-a). This gives the
contribution
Z
1
 B2a = 2ig 2 d 0 d 00 h || 00 ih 00 | | 0 ir n ( 00 )h 00 |( n ) | i( )( 0 ).

(A.19)

194

Y. Kazama, T. Muramatsu / Nuclear Physics B 584 (2000) 171196

Similarly, use of V3 , V4 and V5 generates the diagrams of the type (B2-b), the contribution
of which is worked out to be
 B2b = 2ig 2 h |( )r m ( m ) | i  B2a .

(A.20)

Therefore,
 B2 =  B2a +  B2b = 2ig 2 ( )h |3 | i(/r ) =

3ig 2
2( )(/r ) .
16r 5
(A.21)

Calculation of  C ( )
Finally, consider  C . It takes the form
Z


 C ( ) = g 2 d 0 hC( )C ( 0 )i Ym ( 0 ) ( )

0 )ihYm ( 0 ) ( )i  T m ( 0 ),
+ hC ( )C(

(A.22)

which is represented by the diagram (C). Using the vertex (A.7) and proceeding similarly
to the previous calculations, we obtain
C =


3ig 2
2(/r ) 2rl ( m )( ml ) .
5
16r

(A.23)

Summary
Adding up all the contributions and reinstating the factor of N , the final result is
 =

3ig 2 N
16r 5

rl ( mnl )( nm ) + 2rl ( ml )( m ) 4rl ( m )( ml )



+ 4( )(/r ) 4(/r ) .
(A.24)

Appendix B. SO(9) Fierz identities


In this appendix, we record the SO(9) Fierz identities which are crucial in simplifying
the O( 2 ) part of  at 1-loop.
Adapting the notations of Taylor and Raamsdonk [5], let us introduce the following
quantities for n = 0, 1, 2, 3 (repeated indices are summed):
En = rm ( a1 an m )( an a1 ),
E n = rm ( a1 an )( an a1 m ),
a1 an m

an a1

)(
),
Fn = rm (
a
a
a
a
m
n
n
1
1
)(
),
Fn = rm (
1
P = rm  a1 a2 a3 a4 b1 b2 b3 b4 m ( a1 a2 a3 a4 )( b1 b2 b3 b4 m ),
4!
1
Q = rm  a1 a2 a3 a4 b1 b2 b3 b4 m ( a1 a2 a3 a4 )( b1 b2 b3 b4 m ),
4!

(B.1)
(B.2)
(B.3)
(B.4)
(B.5)
(B.6)

Y. Kazama, T. Muramatsu / Nuclear Physics B 584 (2000) 171196

195

where is an arbitrary spinor. Because = 0 for any symmetric matrix , five of them
actually vanish, namely
F0 = F0 = F1 = F3 = Q = 0.

(B.7)

Since there are nine independent Fierz identities [5], only four structures are independent,
which we take to be F1 , F2 , F2 and F3 . Then the remaining quantities can be expressed in
terms of them as 7
1
1
1
E0 = F1 (F2 + F2 ) F3 ,
(B.8)
16
32
96
1
1
1
(B.9)
E 0 = F1 (F2 + F2 ) + F3 ,
16
32
96
3
1
1
(B.10)
E1 = F1 (F2 F2 ) F3 ,
8
8
48
3
1
1
(B.11)
E 1 = F1 + (F2 F2 ) F3 ,
8
8
48
7
1
1
(B.12)
E2 = F1 (F2 + F2 ) + F3 ,
4
4
24
7
1
1
(B.13)
E 2 = F1 (F2 + F2 ) F3 ,
4
4
24
21
3
3
(B.14)
E3 = F1 + (F2 F2 ) + F3 ,
4
4
8
21
3
3
(B.15)
E 3 = F1 (F2 F2 ) + F3 ,
4
4
8
15
(B.16)
P = (F2 + F2 ).
2
From these relations, one easily finds the identity
0 = F2 2F1 4E 1 + 4E 0 + 12E0 .
Removing , we get
0 = rl ( mnl )( nm ) + 2rl ( nl )( n ) + 4rl ( n )( nl )
+ 4rl ( )( l ) + 12rl ( l ) ,

(B.17)

which was used in Section 5.


References
[1]
[2]
[3]
[4]
[5]
[6]

T. Banks, W. Fischler, S.H. Shenker, L. Susskind, Phys. Rev. D 55 (1997) 5112, hep-th/9610043.
L. Susskind, hep-th/9704080.
Y. Okawa, T. Yoneya, Nucl. Phys. B 538 (1999) 67, hep-th/9806108.
Y. Okawa, T. Yoneya, Nucl. Phys. B 541 (1999) 163, hep-th/9808188.
W. Taylor, M.V. Raamsdonk, JHEP 04 (1999) 013, hep-th/9812239.
A. Sen, Adv. Theor. Math. Phys. 2 (1998) 51, hep-th/9709220.

7 The sign in front of 48E in Eq. (B.20) of [5] should be +.


0

196

[7]
[8]
[9]
[10]
[11]
[12]
[13]
[14]
[15]
[16]
[17]
[18]
[19]
[20]
[21]
[22]
[23]

Y. Kazama, T. Muramatsu / Nuclear Physics B 584 (2000) 171196

N. Seiberg, Phys. Rev. Lett. 79 (1997) 3577, hep-th/9710009.


A. Jevicki, T. Yoneya, Nucl. Phys. B 535 (1998) 335, hep-th/9805069.
A. Jevicki, Y. Kazama, T. Yoneya, Phys. Rev. Lett. 81 (1998) 5072, hep-th/9808039.
A. Jevicki, Y. Kazama, T. Yoneya, Phys. Rev. D 59 (1999) 066001, hep-th/9810146.
J. Maldacena, Adv. Theor. Math. Phys. 2 (1998) 231, hep-th/9711200.
S. Paban, S. Sethi, M. Stern, Nucl. Phys. B 534 (1998) 137, hep-th/9805018.
S. Paban, S. Sethi, M. Stern, JHEP 06 (1998) 012, hep-th/9806028.
D.A. Lowe, JHEP 11 (1998) 009, hep-th/9810075.
S. Sethi, M. Stern, JHEP 06 (1999) 004, hep-th/9903049.
S. Hyun, Y. Kiem, H. Shin, Nucl. Phys. B 558 (1999) 349, hep-th/9903022.
H. Nicolai, J. Plefka, hep-th/0001106.
Y. Okawa, Talk at YITP Workshop, Kyoto, July, 1999.
M.R. Douglas, D. Kabat, P. Pouliot, S.H. Shenker, Nucl. Phys. B 485 (1997) 85, hepth/9608024.
Y. Okawa, Nucl. Phys. B 552 (1999) 447, hep-th/9903025.
H. Hata, S. Moriyama, Phys. Lett. B 452 (1999) 45, hep-th/9901034.
B. de Wit, D.Z. Freedman, Phys. Rev. D 12 (1975) 2286.
H. Hata, S. Moriyama, Phys. Rev. D 60 (1999) 126006, hep-th/9904042.

Nuclear Physics B 584 (2000) 197215


www.elsevier.nl/locate/npe

Deconfinement at the ArgyresDouglas point


in SU(2) gauge theory with broken
N = 2 supersymmetry
A. Gorsky a , A. Vainshtein b , A. Yung c
a Institute of Experimental and Theoretical Physics, Moscow 117259, Russia
b Theoretical Physics Institute, University of Minnesota, Minneapolis, MN 55455, USA
c Petersburg Nuclear Physics Institute, Gatchina, St. Petersburg 188350, Russia

Received 25 April 2000; accepted 6 June 2000

Abstract
We consider chiral condensates in SU(2) gauge theory with broken N = 2 supersymmetry.
The matter sector contains an adjoint multiplet and one fundamental flavor. Matter and gaugino
condensates are determined by integrating out the adjoint field. The only nonperturbative input is the
AffleckDineSeiberg (ADS) superpotential generated by one instanton plus the Konishi anomaly.
These results are consistent with those obtained by the integrating in procedure, including a
reproduction of the SeibergWitten curve from the ADS superpotential. We then calculate monopole,
dyon, and charge condensates using the SeibergWitten approach. We show that the monopole and
charge condensates vanish at the ArgyresDouglas point where the monopole and charge vacua
collide. We interpret this phenomenon as a deconfinement of electric and magnetic charges at the
ArgyresDouglas point. 2000 Elsevier Science B.V. All rights reserved.
PACS: 11.15.-q; 11.30.Pb; 14.80.Hv
Keywords: Supersymmetric gauge theories; chiral condensates; monopole

1. Introduction
The derivation of exact results in N = 1 supersymmetric gauge theories based on low
energy effective superpotentials and holomorphy was pioneered in [13], then new wave
of development was initiated by Seiberg, see [4] for review. Additional input was provided
by the SeibergWitten solution of N = 2 supersymmetric gauge theories with and without
matter [5].
The key feature of the N = 2 theory is the existence of the Coulomb branch where the
vacuum expectation value of the adjoint scalar serves as a modulus [5]. The solution is
described in terms of Riemann surfaces and the Coulomb branch parametrizes the moduli
0550-3213/00/$ see front matter 2000 Elsevier Science B.V. All rights reserved.
PII: S 0 5 5 0 - 3 2 1 3 ( 0 0 ) 0 0 3 4 9 - 7

198

A. Gorsky et al. / Nuclear Physics B 584 (2000) 197215

space of their complex structures. The simplest way to break N = 2 supersymmetry


(SUSY) down to N = 1 amounts to giving a nonvanishing mass to the chiral N = 1
superfield in the adjoint representation. This field is a partner to the gauge fields in the
N = 2 supermultiplet. At small values of the theory is close to its N = 2 counterpart
while at large the adjoint matter decouples and the pure N = 1 theory emerges. The
emerging theory at large is close to supersymmetric QCD (SQCD) but does not coincide
with it. A trace of the massive adjoint remains in the effective theory in the form of
nonrenormalizable quartic terms [6] in the superpotential which are suppressed by 1/.
Although in the N = 1 theory the degeneracy on the Coulomb branch is lifted by the
superpotential, memory of the structure of the Riemann surfaces remains. Namely, the
vanishing of the discriminant of the Riemann surface defines the set of vacua in the
corresponding N = 1 theory [513].
Different vacua are distinguished by the values of chiral condensates, such as the gluino
e
condensate hTr i and the condensate of fundamental matter hQQi.
Generically, the latter
can be found in SQCD using the effective superpotential, while the gluino condensate can
be evaluated using the Konishi anomaly which relates the two condensates (see Ref. [14]
for a review). To obtain the condensate in pure N = 1 YangMills theory one has to start
with the massive SQCD and use the holomorphy to decouple massive matter. Recently,
some points concerning formation of the condensate and identification of the relevant field
configurations were clarified in [1518]
The brane picture provides another approach to the problem. The brane configurations
for N = 1 theories with different matter content are known [19,20] and the recipe
for calculating minima of the superpotentials has been formulated [21]. The key point
concerning the brane configurations is that to break SUSY down to N = 1 one has to
rotate the N = 2 picture. However only configurations which correspond to the vanishing
of the discriminant can be rotated which means that at any value of the adjoint mass
these points remain intact. The superpotentials calculated from brane configurations are
in correspondence with field theory expectations.
In this paper we consider an N = 1 theory with both adjoint and fundamental matter
and limit ourselves to the most tractable case of SU(2) gauge group with one fundamental
flavor and one multiplet in the adjoint representation. Our strategy is as follows: First,
we integrate out the adjoint matter to get SQCD-like effective superpotential for the
fundamental matter. The only nonperturbative input in this effective superpotential is given
by the AffleckDineSeiberg superpotential generated by one instanton [1]. Difference
with pure SQCD is due to the tree level nonrenormalizable term generated by the heavy
adjoint exchange, mentioned above. Similarly to SQCD, the effective superpotential
together with the Konishi relations unambiguously fixes condensates of fundamental and
adjoint matter as well as the gaugino condensates in all three vacua of the theory.
We then compare the condensate of the adjoint matter with points in the u plane
corresponding to the vanishing of the discriminant defined by SeibergWitten solution in
N = 2 theory and find a complete match. Our results for matter and gaugino condensates
are consistent with those obtained by the integrating in method [8,9,22,23] and can be
viewed as an independent confirmation of this method. What is specific to our approach is

A. Gorsky et al. / Nuclear Physics B 584 (2000) 197215

199

that we start from the weak coupling regime where the notion of an effective Lagrangian is
well defined, and then use holomorphy to extend results for chiral condensates into strong
coupling.
We subsequently determine monopole, dyon, and charge condensates following the
SeibergWitten approach, i.e., considering effective superpotentials near singularities on
the Coulomb branch of the N = 2 theory. Again, holomorphy allows us to extend our
results to the domain of the hard N = 2 breaking. This extension includes not only the
mass term of adjoint but also breaking of N = 2 in Yukawa couplings.
Our next step is the study chiral condensates in the ArgyresDouglas (AD) points. These
points were originally introduced in the moduli/parameter space of N = 2 theories as
points where two singularities on the Coulomb branch coalesce [2426]. It is believed
that the theory in the AD point flows in the infrared to a nontrivial superconformal theory.
The notion of the AD point continues to make sense even when the N = 2 theory is broken
to N = 1; in the N = 1 theory it is the point in parameter space where two vacua collide.
In particular, we consider the AD point where the monopole and charge vacua collide at a
particular value of the mass of the fundamental flavor. Our key result is that both monopole
and charge condensates vanish at the AD point. 1 We interpret this as deconfinement of
both electric and magnetic charges at the AD point. It provides evidence that the theory
at the AD point remains superconformal even after strong breaking of N = 2 to N = 1.
Argyres and Douglas conjectured this in their consideration of SU(3) theory [24]
Let us recall that the condensation of monopoles ensures confinement of quarks in
the monopole vacuum [5], while the condensation of charges provides confinement of
monopoles in the charge vacuum. As shown by t Hooft [28] it is impossible for these
two phenomena to coexist. This apparently leads to a paradoxical situation in the AD point
where the monopole and charge vacua collide. Our result resolves this paradox.
This paradox is a part of more general problem: whether there is an uniquely defined
theory at the AD point. Indeed, when two vacua collide the Witten index of the emerging
effective theory at the AD point is fixed, namely there are two bosonic vacuum states. The
question is whether there is any physical quantity which could serve as an order parameter
differentiating these two vacua. The continuity of chiral condensates in the AD point we
find shows that these condensates are not playing this role. The same continuity also leads
to vanishing tension for domain walls interpolating between colliding vacua when we
approach the AD point. We discuss if these domain walls could serve as a signal of two
vacua in the AD point.
The paper is organized as follows. In Section 2 we dwell on the calculation of matter
and gaugino condensates, while monopole, charge and dyon condensates are considered in
Section 3. In Section 4 we briefly discuss a definition of the theory at the AD point and the
related problem of domain walls. Our results are discussed in Section 5.

1 Vanishing of condensates for coalescing vacua was mentioned by Douglas and Shenker [27] in the context of
SU(N ) theories without fundamental matter for N > 3. Note, that it was before the notion of the AD point was
introduced [24].

200

A. Gorsky et al. / Nuclear Physics B 584 (2000) 197215

2. Matter and gaugino condensates


2.1. Effective superpotential and condensates
We consider a N = 1 theory with SU(2) gauge group where the matter sector consists of
the adjoint field = a ( a /2) ( , = 1, 2; a = 1, 2, 3), and two fundamental fields
Qf (f = 1, 2) describing one flavor. The general renormalizable superpotential for this
theory has the form,
W = Tr 2 +

m f
1
Q Q + hfg Qf Qg .
2 f
2

(1)

Here the parameters and m are related to the masses of the adjoint and fundamental
fields, m = /Z , mQ = m/ZQ , by the corresponding Z factors in the kinetic terms.
Having in mind normalization appropriate for the N = 2 case we choose for bare
0
0
= 1/g02 , ZQ
= 1. The matrix of Yukawa couplings hfg is symmetric, and
parameters Z
summation over color indices , = 1, 2 is explicit. Unbroken N = 2 SUSY appears when
= 0 and det h = 1.
To obtain an effective theory similar to SQCD we integrate out the adjoint field
implying that m  mQ . In the classical approximation this integration reduces to the
substitution


1
1

fg

Qf Qg Qf Qg ,
(2)
= h
2
2 2
which follows from W/ = 0. What is the effect of quantum corrections on the effective
superpotential? It is well known from the study of SQCD that perturbative loops do not
contribute and nonperturbative effects are exhausted by the AffleckDineSeiberg (ADS)
superpotential generated by one instanton [1]. The effective superpotential then is
Weff = mV

( det h) 2 2 31
V +
4
4V

(3)

where the gauge and subflavor invariant chiral field V is defined as


1 f
Q Q .
(4)
2 f
The first two terms in Eq. (3) appear on the tree level after substitution (2) into Eq. (1)
while the third nonperturbative one is the ADS superpotential. The scale parameter 1 is
given in terms of the mass of PauliVillars regulator MPV and the bare coupling g0 (plus
the vacuum angle 0 ) as


8 2
3
exp 2 + i0 .
(5)
31 = 4MPV
g0
V=

The coefficient 2 31 /4 in the ADS superpotential is equivalent to 5SQCD in SQCD. The


factor 2 in the coefficient reflects four fermionic zero modes of the adjoint field, see, e.g.,
Ref. [17,29] for details.

A. Gorsky et al. / Nuclear Physics B 584 (2000) 197215

201

The only term in the superpotential (3) which differentiates it from the SQCD case is the
second term which is due to tree level exchange by the adjoint field. At h = 0 it vanishes
and we are back to the known SQCD case with two vacua and a Higgs phase for small m.
When det h is nonvanishing we have three vacua, marked by the vevs of the lowest
component of V ,
v = hV i.
These vevs are roots of the algebraic equation dWeff /dv = 0 which has the form
 
31 2
( det h) v

= 0.
m
2

4 v

(6)

(7)

This equation shows, in particular, that although the second term in the superpotential (3)
seems to be suppressed at large it turns out to be of the same order as the ADS term.
From Eq. (7) it is also clear that the dependence on is given by the scaling v .
To see the dependence on the other parameters let us substitute v by the dimensionless
variable defined by the relation
s
31
.
(8)
v=
4m
Then Eq. (7), when rewritten in terms of ,
1
=0
2
is governed by the dimensionless parameter ,
 
( det h) 1 3/2
.
=
4
m
1

(9)

(10)

We see that the two parameters m and det h enter only as m( det h)2/3 . The dependence
of v on is linear as we discussed above.
The particular dependence of condensate v on the parameters , m and det h follows
from the R symmetries of the theory. Following Seiberg [30] one can consider , m
and det h as background fields and identify nonanomalous R symmetries which prove the
dependence discussed above. Classically, there are three U(1) symmetries in the theory
associated with the three fermion fields (gaugino, adjoint and fundamental fermions). In
the quantum theory one can organize two nonanomalous combinations (a symmetry is
nonanomalous if it does not transform the scale 1 , associated with regulators).
The charges of the fields and parameters of the theory under these two U(1) symmetries
are shown in Table 1. The first of these symmetries UJ (1) is a subgroup of the global
SUR (2) group related to the N = 2 superalgebra [5]. This explains the zero charge of the
coupling h with respect to this symmetry. The symmetry UJ (1) fixes the dependence
of condensates. Namely, it is given by a power of equal to half the UJ (1) charge of
the condensate. In particular, the field V has UJ (1) charge equal to 2 which ensures that
v . Thus, we can use holomorphy to extend results to arbitrary values of .

202

A. Gorsky et al. / Nuclear Physics B 584 (2000) 197215

Table 1
Nonanomalous U(1) symmetries
Fields/parameters

UJ (1) charges
UR (1) charges

0
1

1
1

1
1

1
1

0
4

2
0

0
3

The second nonanomalous symmetry UR (1) is similar to the R symmetry of Ref. [1]
extended to include the adjoint field. As a consequence, for a given chiral field X
d (QJ /2)(QR /4)

hXi = QJ /2 mQR /4 1X

fX ( ),

(11)

where QJ , QR are the UJ (1), UR (1) charges of the field X, dX is its dimension, and fX is
an arbitrary function of the dimensionless parameter defined by Eq. (10). This parameter
is neutral under both U(1)s. Eq. (8) is an example of the general relation (11) with QJ = 2,
QR = 2 and fV = ( )/2.
The important benefit of the consideration above is that in a theory with N = 2 SUSY
strongly broken by large and det h 6= 1 we can still relate chiral condensates with those
in softly broken N = 2 where det h = 1 and is small.
Here is an example. When 0 two roots of Eq. (9) are 1,2 = 1 and the third
one goes to infinity as 3 = 1/ . For two finite roots one can suggest dual interpretations.
Firstly, taking h = 0, one can relate them to two vacua of SQCD in the Higgs phase.
Second, for det h = 1 (which is its N = 2 value) one can make small by taking the
limit of large m. But this limit should bring us to the monopole and dyon vacua of softly
broken N = 2 SYM. The naming of vacua refers to the particle whose mass vanishes in
the corresponding vacuum.
To verify this interesting mapping we need to determine the vev

(12)
u = hU i = Tr 2 ,
which can be accomplished using the set of Konishi anomalies. Generic equation for an
arbitrary matter field Q looks as follows (we are using the notation of the review [14]):
Tr W 2
W
1 2
+ T (R)
D JQ = Q
,
4
Q
8 2

(13)

where T (R) is the Casimir in the matter representation. The left hand side is a total
derivative in superspace so its average over any supersymmetric vacuum vanishes. In our
case this results in two relations for the condensates,


1 fg
1 Tr W 2
m f

Q Q + h Qf Qg +
= 0,
2 f
2 8 2
2


1
Tr W 2
= 0.
(14)
2 Tr 2 + hfg Qf Qg + 2
8 2
2
From the first relation, after the substitution in (2) and comparing with Eq. (7), we find an
expression for gluino condensate [31]

A. Gorsky et al. / Nuclear Physics B 584 (2000) 197215

s=

hTr 2 i
hTr W 2 i 2 31
.
=

=
16 2
16 2
4v

203

(15)
P

This is consistent with the general expression [TG


T (R)]hTr 2 i/16 2 for the
nonperturbative ADS piece of the superpotential (3), see [32]. Combining the two relations
in (14) we can express the condensate u in terms of v,
q



3
2
m31 
1
3 1
3
1
(mv + 3s) =
mv +
=
+
.
(16)
u=
2
2
4 v
4

Now we see that in the limit of large m two vacua = 1 are in perfect correspondence
with u = 20 for the monopole and dyon vacua of N = 2 SYM. Indeed, 40 = m31 is
the correct relation between the scale parameters of the theories.
For the third vacuum at large m the value u = m2 /( det h) corresponds on the Coulomb
branch to the so called charge vacuum, where some fundamental fields become massless.
Moreover, the correspondence with N = 2 results can be demonstrated for the three vacua
at any value of m. To this end we use the relation (16) and Eq. (9) to derive the following
equation for u,
9
27
( det h)u3 m2 u2 ( det h)m31 u + m3 31 + 8 ( det h)2 61 = 0.
8
2

(17)

The three roots of this equation are the vevs of Tr 2 in the corresponding vacua.
How does this look from N = 2 side? The Riemann surface governing the Seiberg
Witten solution is given by the curve [5]
1
1
(18)
y 2 = x 3 ux 2 + 31 mx 61 .
4
64
Singularities of the metric, i.e., the points in the u-plane where the discriminant of the
curve vanishes are defined by two equations, y 2 = 0 and dy 2/dx = 0,
1
1
1
3x 2 2ux + 31 m = 0,
(19)
x 3 ux 2 + 31 mx 61 = 0,
4
64
4
which lead to
9
27
(20)
u3 m2 u2 m31 u + m3 31 + 8 61 = 0.
8
2
We see that this is a particular case of the N = 1 Eq. (17) at det h = 1.
Moreover, when det h is not equal to its N = 2 value (1) Eq. (17) coincides with
Eq. (20) after the rescaling
u = ( det h)1/3 u0 ,

m = ( det h)2/3 m0 ,

v = ( det h)1/3 v 0 .

(21)

This is in agreement with the master parameter which contains the product m3/2 det h
and the nonanomalous U(1) symmetries we discussed above. In other words, breaking of
N = 2 by Yukawa couplings does not influence consideration of the chiral condensates
modulus the rescaling (21).
The consideration above shows that the only nonperturbative input needed to determine
the chiral condensates is provided by the one-instanton ADS superpotential. This means

204

A. Gorsky et al. / Nuclear Physics B 584 (2000) 197215

that any reference to the N = 2 limit is not crucial at all, i.e., in regard to these condensates
the exact SeibergWitten solution of N = 2 is equivalent to the ADS superpotential.
The relations for the condensates we have derived are not new, they were obtained in
[8,9] by the integrating in procedure introduced in [23]. Our approach which is based on
integrating out, plus the Konishi relations, can be viewed as an independent proof of the
integrating in procedure.
What we see as an advantage of our approach it is that, within a certain range of
parameters, the superpotential (3) gives a complete description of the low energy physics.
Indeed, when the mass mV of the field V ,
mV = 2m(2 3 ),

(22)

is much less than the other masses, such as m = g 2 and mW = |g 2 v|1/2 , we are in the
weakly coupled Higgs phase and enjoy full theoretical control. The Konishi relations help
to determine the condensates of heavy fields in this phase. Holomorphy then allows for
continuation of these results for the condensates to strong coupling.
At strong coupling the superpotential (3), like other versions of the Veneziano
Yankielowicz Lagrangians [33,34], does not describe the low energy physics. For example,
it contains no light monopole degrees of freedom near the monopole vacuum point at
small . Moreover, there is no single local superpotential which could describe mutually
nonlocal degrees of freedom which become light in different regions of the moduli space
of the theory. At strong coupling the superpotential (3) is equivalent to the effective
superpotential [8,9] of the integrating in procedure and can be viewed as a shorthand
equation that gives the values of the condensates.
2.2. Matter and gaugino condensates in the limit of large mass
f

Here we summarize the results for matter, v = hQf Q /2i, u = hTr 2 i, and gaugino,
s = hTr 2 /16 2 i, condensates in the limit where the parameter defined by Eq. (10) is
small. This can be achieved in the limit of large m if the Yukawa coupling is fixed, or by
taking det h to be small otherwise. In the charge vacuum:

2m
1 + O 2 ,
( det h)

m2
1 + O 2 ,
uC =
( det h)

31 ( det h)
1 + O 2 .
sC =
8m
vC =

In the monopole and dyon vacua:


s

31
1 + O( ) ,
vM,D =
4m
q

uM,D = 31 m 1 + O( ) ,

(23)

A. Gorsky et al. / Nuclear Physics B 584 (2000) 197215

q

1
sM,D = 31 m 1 + O( ) .
2

205

(24)

The upper sign refers to the monopole vacuum, while the lower one is for the dyon vacuum.
As discussed above we can interpret these vacua also as the two vacua of the Higgs phase
in SQCD. To this end we need to consider the limit of small det h and m  SQCD with
the identification
1
5SQCD = 2 31 .
4

(25)

2.3. Small mass limit


The limit of massless fundamentals m 0 corresponds to . In this limit the
three vacua are related by a Z3 symmetry [5],

1
e2ik/3 1 + O 2/3
(k = 0, 1),
1/3
(2 det h)

3
(k = 0, 1),
u = 21 (2 det h)1/3 e2ik/3 1 + O 2/3
8

1
(k = 0, 1).
s = 21 (2 det h)1/3 e2ik/3 1 + O 2/3
4
v=

(26)

Note that the massless limit exists due to the nonvanishing Yukawa coupling. When h 0
we are back to the runaway vacua of massless SQCD.
2.4. ArgyresDouglas points
When the mass m changes from large to small values we interpolate between the two
quite different structures of vacua shown above. Let us consider this transition when, for
definiteness, det h = 1 and m is real and positive and changes from large to small values.
At large positive m all the vacua are situated at real values of u, from Eqs. (23), (24) we
see that the dyon vacuum is at negative u, the monopole vacuum is at positive u, and the
charge vacuum is also at positive, but much larger, values of u. When m diminishes then at
some point the monopole and charge vacua collide on the real axis of u and subsequently
go more off to complex values producing the Z3 picture at small m.
The point in the parameter manifold where the two vacua coincide is the AD point [24].
In the SU(2) theory these points were studied in [25]. Mutually non-local states, say
charges and monopoles, become massless at these points. On the Coulomb branch of the
N = 2 theory these points correspond to a non-trivial conformal field theory [25]. Here we
study the N = 1 SUSY theory, where N = 2 is broken by the mass term for the adjoint
matter as well as by the difference of the Yukawa coupling from its N = 2 value. Collisions
of two vacua still occur in this theory. In this subsection we find the values of m at which
AD points appear and calculate the values of the condensates at this point. In the next
section we study what happens to the confinement of charges in the monopole point at
non-zero once we approach the AD point.

206

A. Gorsky et al. / Nuclear Physics B 584 (2000) 197215

First, let us work out the AD values of m, generalizing the consideration in [25].
Coalescence of two roots for v means that together with Eq. (7) the derivative of its lefthand-side should also vanish,
 
 3
3 2

( det h) v
1
= 0,
( det h) + 31
= 0.
(27)
m
2

4 v
v
This system is consistent only at three values of m = mAD ,
3
mAD = 1 ( det h)2/3 ,
4

= e2in/3

(n = 0, 1),

(28)

related by Z3 symmetry. The condensates at the AD vacuum are


1
,
( det h)1/3
3
uAD = 1 21 ( det h)1/3 ,
4
1 1
21 ( det h)1/3 .
sAD =
4
vAD =

(29)

3. Dyon condensates
In this section we calculate various dyon condensates at the three vacua of the theory.
As discussed above, holomorphy allows us to find these condensates starting from a
consideration on the Coulomb branch in N = 2 near the singularities associated with a
given massless dyon. Namely, we calculate the monopole condensate near the monopole
point, the charge condensate near the charge point and the dyon (nm , ne ) = (1, 1)
condensate near the point where this dyon is light. Although we start with small values
of the adjoint mass parameter , our results for condensates are exact for any as well as
for any value of det h.
3.1. Monopole condensate
Let us start with calculation of the monopole condensate near the monopole point.
Near this point the effective low energy description of our theory can be given in terms
of N = 2 dual QED [5]. It includes a light monopole hypermultiplet interacting with a
vector (dual) photon multiplet in the same way as electric charges interact with ordinary
photons. Following Seiberg and Witten [5] we write down the effective superpotential in
the following form,

e
(30)
W = 2 MMA
D + U,
where AD is a neutral chiral field (it is a part of the N = 2 dual photon multiplet in the
N = 2 theory) and U = Tr 2 considered as a function of AD . The second term breaks
N = 2 supersymmetry down to N = 1.

A. Gorsky et al. / Nuclear Physics B 584 (2000) 197215

207

e we find that AD = 0, i.e., the


Varying this superpotential with respect to AD , M and M
monopole mass vanishes, and


du
e =
.
(31)
MM

2 daD aD =0
The condition AD = 0 means that the Coulomb branch near the monopole point, where
the monopole mass vanishes, shrinks to a single vacuum state at the singularity while
Eq. (31) determines the value of monopole condensate. Below we consider aD as a function
of u. The value of u, u = uM , at the monopole vacuum was determined in the previous
section.
The non-zero value of the monopole condensate ensures U(1) confinement for charges
via the formation of AbrikosovNielsenOlesen vortices. Let us work out the r.h.s. of
Eq. (31) to determine the and m dependence of the monopole condensate. From the
exact SeibergWitten solution [5], we have
I
2
dx
daD
=
.
(32)
du
8
y(x)

Here for y(x) given by Eq. (18) we use the form


y 2 = (x e0 )(x e )(x e+ ).

(33)

The integration contour in the x plane circles around two branch points e+ and e of
y(x). At the monopole vacuum, when u = uM , two branch points e+ and e coincide,
e+ = e = e and the integral (32) is given by the residue at x = e.

i 2
daD
(uM ) =
.
(34)
du
4 e e0
The value of e e0 (equal at u = uM to (1/2) d2(y 2 )/dx 2 ) is fixed by the equation
d(y 2 )/dx = 0,
r
3
(35)
e e0 = u2M m31 .
4
Substituting this into the expression for the monopole condensate (31) we get finally
1/4



e = 2i u2M 3 m3
.
(36)
MM
1
4
To test the result let us consider first the limit of a large masses m for the fundamental
matter. As in Section 2.1 this limit can be viewed as a RG flow to pure YangMills theory
with the identification
40 = m31 ,

(37)

where 0 is the scale of the N = 2 YangMills theory. In this limit we have uM = 20 .


Then Eq. (36) gives


e = 2 i0 ,
(38)
MM
which coincides with the SeibergWitten result [5]. This ensures monopole condensation
and charge confinement in the monopole point at large m.

208

A. Gorsky et al. / Nuclear Physics B 584 (2000) 197215

Notice, that in the derivation above N = 2 was not broken by the Yukawa coupling, i.e.,
we assume det h = 1. The result, however, can be easily generalized to arbitrary det h by
e is equal to one.
means of U(1) symmetries considered
above. The UR (1) charge of MM

e
Indeed, the coefficient of the 2 MMAD term in the superpotential (30) which is equal
to one in the N = 2 limit remains the same when N = 2 is broken down to N = 1. It
follows from U(1) symmetries together with decoupling of fundamental matter at large m.
As a result we see that the Eq. (36) for the monopole condensate remains valid for arbitrary
det h.
It is instructive to rewrite the result (36) for the monopole condensate in terms of v,



1/2


3 31 1/2
2
e
= i ( det h)v 2 4s
,
(39)
MM = i ( det h)v
v
where we also show a form which uses the gluino condensate s. It follows from the
expression (16) for u and Eq. (7) for v. It is interesting to observe that at m  1 , when
the value of v is large and the nonperturbative term in (39) can be neglected, the monopole

e
condensate reduces to that of the quark, hMMi
iv det h. Another interesting limit is
the SQCD one when h 0. In this limit the nonperturbative term in Eq. (39) dominates,
and the square of the monopole condensate reduces to the gluino condensate.
Now let us address the question: what happens with the monopole condensate when we
reduce m and approach the AD point? The AD point corresponds to a particular value of m
which ensures coalescence of the monopole and charge singularities in the u plane. Near
the monopole point we have condensation of monopoles and confinement of charges while
near the charge point we have condensation of charges and confinement of monopoles. As
shown by t Hooft these two phenomena cannot happen simultaneously [28]. The question
is: what happens when monopole and charge points collide in the u plane?
The monopole condensate at the AD point is given by Eq. (36). When m and u are
substituted by mAD and uAD from Eqs. (28) and (29), we get


e
= 0.
(40)
MM
AD
We see that the monopole condensate goes to zero at the AD point. Our derivation
makes it clear why it happens. At the AD point all three roots of y 2 become degenerate,

e+ = e = e0 , so the monopole condensate which is proportional to e e0 naturally


vanishes.
In the next subsection we calculate the charge condensate in the charge point and show
that it also goes to zero as m approaches its AD value (28). Thus we interpret the AD point
as a deconfinement point for both monopoles and charges.
3.2. Charge and dyon condensates
In this subsection we use the same method to calculate values for the charge and dyon
condensates near the charge and dyon points respectively. We first consider m above its
AD value (28) and then continue our results to values of m below mAD . In particular, in
the limit m = 0 we recover Z3 symmetry.

A. Gorsky et al. / Nuclear Physics B 584 (2000) 197215

209

Let us start with the charge condensate. At = 0, det h = 1 and large m the effective
theory near the charge point

(41)
a = 2m
on the Coulomb branch is N = 2 QED. Here a is the neutral scalar, the partner of photon
in the N = 2 supermultiplet. Half of the degrees of freedom in color doublets become
massless whereas the other half acquire a large mass 2m. The massless fields form one
e+ , Q+ of charged particles in the effective electrodynamics. Once we
hypermultiplet Q
add the mass term for the adjoint matter the effective superpotential near the charge point
becomes
1 e
e+ Q+ + U.
(42)
W= Q
+ Q+ A + mQ
2
Minimizing this superpotential we get condition (41) as well as

e+ Q+ = 2 du
Q
.
da a= 2 m
Now, following the same steps which led us from (31) to (36), we get

1/4


e+ Q+ = 2 u2C 3 m3
det h Q
,
1
4

(43)

(44)

where we include a generalization to arbitrary det h. We choose to consider det h


e
e+ Q+ i because it has the UR (1) charge equal to one, similar to the hMMi
condensate
hQ
considered above. By uC we denote the position of the charge vacuum in the u plane. At
large m uC = m2 /( det h), see Eq. (23), and


e+ Q+ =
Q


2m
1 + O 2 .
( det h)

(45)

Holomorphy allows us to extend the result (44) to arbitrary m and det h. So we can use
Eq. (44) to find the charge condensate at the AD point. Using Eqs. (28) and (29) we see
that the charge condensate vanishes at the AD point in the same manner the monopole
condensate does. As it was mentioned we interpret this as deconfinement for both charges
and monopoles.
As with the monopole condensate, we can also relate the charge condensate with the
quark vev v,

1/2 
1/2

3 31
4s
2
2
e
= v
.
(46)
Q+ Q+ = v
v( det h)
( det h)
This expression differs from the one for the monopole condensate only by a phase factor.
The coincidence of the charge condensate with the quark one at large v, i.e., at weak
coupling, is natural. The difference is due to nonperturbative effects and is similar to the
difference between a 2 /2 and u on the Coulomb branch of the N = 2 theory. At strong
coupling the difference is not small. In particular, the charge condensate vanishes at the
AD point while the quark condensate remains finite.

210

A. Gorsky et al. / Nuclear Physics B 584 (2000) 197215

Now let us work out the dyon condensate. More generally let us introduce the dyon field
Di , i = 1, 2, 3, which stands for charge, monopole and (1, 1) dyon,


(47)
Di = ( det h)1/4 Q+ , M, D .
The arguments of the previous subsection which led us to the result (36) for monopole
ei Di i
condensate give for hD

1/4

ei Di = 2ii u2i 3 m3
,
(48)
D
1
4
where ui is the position of the ith point in the u plane and the i are phase factors.
For the monopole condensate at real values of m larger than the mAD Eq. (36) gives
M = 1,

(49)

while for the charge condensate from Eq. (44) we have


C = i.

(50)

In fact one can fix the charge phase factor by imposing the condition that the charge
condensate should approach the value 2m in the large m limit. For the dyon the phase
factor is
D = i.

(51)

At the particular AD point we have chosen the monopole and charge condensates vanish,
while the dyon condensate remains non-zero, see (48). Below the AD point, condensates
are still given by Eq. (48), but the charge and monopole phase factors can change. 2 The
dyon phase factor (51) does not change when we move through the AD point because the
dyon condensate does not vanish at this point.
In the limit m = 0 we should recover the Z3 -symmetry for the values of condensates.
From Eq. (48) it is clear that the absolute values of all three condensates are equal because
the values of the three roots ui are on the circle in the u plane, see (26). Imposing the
requirement of Z3 symmetry at m = 0 we can fix the unknown phase factors C and M
below the AD point using the value (51) for dyon. This gives
C = i,

M = i.

(52)

3.3. Photino and gaugino condensates


The gaugino condensate hTr 2 i we found in the previous section can be viewed as a
sum of the condensates for charged gauginos and the photino,

(53)
Tr 2 = + + 3 3 .
2
In gauge invariant form the photino condensate can be associated with

(54)
(Tr W )2 .
2 Note that the quantum numbers of the charge and monopole are also transformed, see [35]

A. Gorsky et al. / Nuclear Physics B 584 (2000) 197215

211

We argue here that the photino condensate vanishes so that the gaugino condensate is solely
due to the charged gluino.
Let us start with the Coulomb branch in the N = 2 theory. All gaugino condensates
vanish in N = 2 for a simple reason: 2 is not the lowest component in the corresponding
N = 2 supermultiplet. When the perturbation U which breaks N = 2 is added to
the superpotential the gaugino condensate is proportional to . However, the term U
in the superpotential does not break N = 2 SUSY in the effective QED. Consider, for
example, the monopole vacuum. The corresponding effective superpotential is given by
Eq. (30), where in the expansion of U as function of AD it is sufficient to retain only
linear term. It was shown in [36] that the perturbation linear in AD does not break N = 2
in the effective QED. An immediate consequence of this observation is that the photino
condensate continues to vanish.

4. The ArgyresDouglas point: how well is the theory defined


As discussed in the Introduction, at the AD point we encounter the problem of not having
a uniquely defined vacuum state. Indeed, when the mass parameter m approaches its AD
value mAD we deal with two vacuum states which can be distinguished by values of the
chiral condensates. It is unlikely that the number of states with zero energy will change
when we reach the AD point, it is very similar to the Witten index. However, the continuity
of the chiral condensates we obtained above shows that they are no longer parameters
which differentiate the two states once we reach the AD point.
This does not prove the absence of a relevant order parameter so the quest can be
continued. A natural possibility to consider is a domain wall interpolating between
colliding vacua. In the case of BPS domain walls their tension is given by the central
charge [32],


(55)
Tab = 2 Weff (va ) Weff (vb ) ,
where a, b label the colliding vacua. The central charge here is expressed via values of
exact superpotential (3) in corresponding vacua. The continuity of the condensate v shows
that the domain wall becomes tensionless at the AD point, T (m mAD )3/2 when m
mAD . If such a domain wall were observable at the AD point it could serve as a signal of
two vacua.
We argue, however, that this domain wall is not observable in continuum limit. The
crucial point is that the wall is built out of massless fields, therefore its thickness is infinite
at the AD point. This makes it impossible to observe this tensionless wall in any physical
experiment of a limited spatial scale.
In the conclusion of this section let us review briefly the brane construction of
N = 1 vacua. Gauge theories are realized on brane worldvolumes. Brane configurations
responsible for N = 1 theories were suggested in [19,20] and a derivation of domain
wall tensions from analysis of Riemann surfaces (which is similar to the calculation
of the masses of BPS particles in N = 2 theories) can be found in [21]. The brane

212

A. Gorsky et al. / Nuclear Physics B 584 (2000) 197215

configuration for the N = 1 theory with one flavor and SU(2) gauge group is described
by Riemann surface embedded into three dimensional complex space C 3 parametrized by
three variables t, v and w. The embedding is given by the following equations
v+m=

(w w+ )(w w )
,
w

t = 2 w(w w+ )(w w ),

1
2
w
w+ = (1 )3 ,
(56)
w+ + w + m = 0,
2
with free parameters , m, . The tension of the walls, which have the interpretation of
the M5 branes wrapping three-cycle with the boundaries on the Riemann surface above
can be calculated by integrating the holomorphic three-form over this cycle
Z
T = dv dw d(log t).
(57)
Let us consider the geometry of the brane configuration near the AD point. It was shown
recently [37,38] that the AD point corresponds to a singular CalabiYau 3-manifold which
is resolved if one adds particular perturbation. Since the tension is defined by integration
of the holomorphic 3-form around the resolved singularity the tensionless wall has a
geometrical interpretation as the M5 brane wrapping this vanishing cycle. Actually, the
curves can be considered as fibered over the complex m plane and the AD singularities
correspond to the appearance of vanishing cycles in the fiber in a manner quite similar
to the SeibergWitten solution of N = 2 theories where vanishing cycles correspond to
massless BPS particles.

5. Conclusions
The approach of this work is similar to that used in SQCD. Namely, we integrate out
the adjoint field which leads, in some range of parameters, to an SQCD-like effective
superpotential. This superpotential describes the low energy theory at weak coupling
where we have full theoretical control. The nonperturbative part is given by the ADS
superpotential generated at the one instanton level. The adjoint field shows up only as an
extra (as compared with SQCD) nonrenormalizable term quartic in the fundamental fields.
Results for chiral condensates of matter and gaugino fields are continued into the range
of a small adjoint mass where we find a complete matching with the N = 2 Seiberg
Witten solution. The ArgyresDouglas points introduced in N = 2 theories are shown to
exist in the N = 1 theory as well. Although the bulk of our results for matter and gaugino
condensates overlaps with what is known in the literature we think that our approach
clarifies some aspects of duality in N = 1 theories.
We then analyze monopole, charge and dyon condensates departing from the Coulomb
branch of the N = 2 theory. This resulted in explicit relations between these condensates
and those of the fundamental matter. The most interesting phenomenon occurs at the
AD point: when the monopole and charge vacua collide both the monopole and charge
condensates vanish. We interpret this as a deconfinement of electric and magnetic charges

A. Gorsky et al. / Nuclear Physics B 584 (2000) 197215

213

at the AD point. Vanishing of condensates signals that the theory at this point becomes
superconformal.
In our approach we see straightforwardly that the one-instanton generated ADS
superpotential is the only nonperturbative input needed to fix all chiral condensates. The
general nature of this statement is seen from our derivation which relates polynomial
coefficients in the SeibergWitten curve to the ADS superpotential.
Let us mention a relation to finite-dimensional integrable systems. It was recognized
that N = 2 theories are governed by finite-dimensional integrable systems. The integrable
system responsible for N = 2 SQCD was identified with the nonhomogenious XXX spin
chain [3941]. After perturbation to the N = 1 theory the Hamiltonian of the integrable
system is expected to coincide with the superpotential of corresponding N = 1 theory. This
has been confirmed by direct calculation in the pure N = 2 gauge theory [42] as well in the
theory with a massive adjoint multiplet [43]. It would be very interesting to find a similar
connection between spin chain Hamiltonians and superpotentials in the N = 1 SQCD.
One more point to be clarified is the meaning of the AD point within approach based
on integrability. Since the quark mass is identified as a value of spin [3941] one might
expect that at particular spin values corresponding to the AD mass, the XXX spin chain
would have additional symmetries similar to superconformal ones. We hope to discuss
these points in more details elsewhere.
In this paper we considered only the SU(2) theory with one flavor postponing the generic
Nc , Nf case for a separate publication. The most interesting problem in the generic
situation involves Seiberg IR duality of the electric SU(Nc ) theory with Nf flavors and
the magnetic SU(Nf Nc ) theory. In generic case of nondegenerate fundamental masses
we expect deconfinement at the AD points. A degeneracy in fundamental masses leads to
the appearance of Higgs branches. The approach of the present paper can be applied to this
case as well. However, since Higgs branches do not disappear at the AD points [25] we do
not expect deconfinement to occur in this case [44].

Acknowledgements
Authors are grateful to P. Argyres, A. Hanany, K. Konishi, A. Marshakov, S. Rudaz,
A. Ritz, and M. Shifman for helpful discussions. Part of this work was done when two of
the authors, A.V. and A.Y., participated in the SUSY99 program organized by the Institute
for Theoretical Physics at Santa Barbara. A.V. and A.Y. are thankful to ITP for hospitality
and support from NSF by the grant PHY 94-07194. A.G. and A.Y. thank the Theoretical
Physics Institute at the University of Minnesota where this work was initiated for support.
A.G. thanks J. Ambjorn for hospitality at the Niels Bohr Institute where part of this work
has been done.
The work of A.G. is supported in part by the grant INTAS-97-0103, A.V. is supported
in part by DOE under the grant DE-FG02-94ER40823, and A.Y. is supported in part by
Russian Foundation for Basic Research under the grant 99-02-16576.

214

A. Gorsky et al. / Nuclear Physics B 584 (2000) 197215

References
[1] I. Affleck, M. Dine, N. Seiberg, Phys. Lett. B 137 (1984) 187.
[2] V.A. Novikov, M.A. Shifman, A.I. Vainshtein, V.I. Zakharov, Nucl. Phys. B 229 (1983)
407, Reprinted in: S. Ferrara (Ed.), Supersymmetry, Vol. 1, North Holland/World Scientific,
AmsterdamSingapore, 1987, p. 606.
[3] M. Shifman, A. Vainshtein, Nucl. Phys. B 296 (1988) 445.
[4] K. Intriligator, N. Seiberg, Nucl. Phys. Proc. Suppl. BC 45 (1996) 1, hep-th/9509066.
[5] N. Seiberg, E. Witten, Nucl. Phys. B 426 (1994) 19; Nucl. Phys. B 430 (1994) 485, hepth/9407087; Nucl. Phys. B 431 (1994) 484, hep-th/9408099.
[6] D. Kutasov, A. Schwimmer, N. Seiberg, Nucl. Phys. B 459 (1996) 455, hep-th/9510222.
[7] K. Intriligator, N. Seiberg, Nucl. Phys. B 431 (1994) 551, hep-th/9408155.
[8] S. Elitzur, A. Forge, A. Giveon, E. Rabinovici, Phys. Lett. B 353 (1995) 79, hep-th/9504080;
Nucl. Phys. B 459 (1996) 160, hep-th/9509130.
[9] S. Elitzur, A. Forge, A. Giveon, K. Intriligator, E. Rabinovici, Phys. Lett. B 379 (1996) 121,
hep-th/9603051.
[10] S. Terashima, S. Yang, Phys. Lett. B 391 (1997) 107, hep-th/9607151.
[11] T. Kitao, Phys. Lett. B 402 (1997) 290, hep-th/9611097.
[12] T. Kitao, S. Terashima, S. Yang, Phys. Lett. B 399 (1997) 75, hep-th/9701009.
[13] K. Konishi, H. Terao, Nucl. Phys. B 511 (1998) 264, hep-th/9707005.
[14] M. Shifman, A. Vainshtein, in: M.A. Shifman (Ed.), ITEP Lectures on Particle Physics and
Field Theory, Vol. 2, World Scientific, Singapore, 1999, p. 485, hep-th/9902018.
[15] N.M. Davies, T.J. Hollowood, V.V. Khoze, M.P. Mattis, Nucl. Phys. B 559 (1999) 123, hepth/9905015.
[16] N.M. Davies, V.V. Khoze, JHEP 0001 (2000) 015, hep-th/9911112.
[17] A. Ritz, A. Vainshtein, Nucl. Phys. B 566 (2000) 311, hep-th/9909073.
[18] G. Carlino, K. Konishi, H. Murayama, JHEP 0002 (2000) 004, hep-th/0001036.
[19] K. Hori, H. Ooguri, Y. Oz, Adv. Theor. Math. Phys. 1 (1998) 1, hep-th/9706082.
[20] E. Witten, Nucl. Phys. B 507 (1997) 658, hep-th/9706109.
[21] S. Nam, K. Oh, S. Sin, Phys. Lett. B 416 (1998) 319, hep-th/9707247.
[22] K. Intriligator, R.G. Leigh, N. Seiberg, Phys. Rev. D 50 (1994) 1092, hep-th/9403198.
[23] K. Intriligator, Phys. Lett. B 336 (1994) 409, hep-th/9407106.
[24] P.C. Argyres, M.R. Douglas, Nucl. Phys. B 448 (1995) 93, hep-th/9505062.
[25] P.C. Argyres, M.R. Plesser, N. Seiberg, E. Witten, Nucl. Phys. B 461 (1996) 71, hep-th/9511154.
[26] T. Eguchi, K. Hori, K. Ito, S. Yang, Nucl. Phys. B 471 (1996) 430, hep-th/9603002.
[27] M.R. Douglas, S.H. Shenker, Nucl. Phys. B 447 (1995) 271, hep-th/9503163.
[28] G. t Hooft, Nucl. Phys. B 138 (1978) 1; Nucl. Phys. B 153 (1979) 141.
[29] A. Yung, Nucl. Phys. B 485 (1997) 38, hep-th/9604096.
[30] N. Seiberg, Phys. Rev. D 49 (1994) 6857, hep-th/9402044.
[31] V.A. Novikov, M.A Shifman, A.I. Vainshtein, V.I. Zakharov, Nucl. Phys. B 260 (1985) 157.
[32] G. Dvali, M. Shifman, Phys. Lett. B 396 (1997) 64, hep-th/9612128.
[33] G. Veneziano, S. Yankielowicz, Phys. Lett. B 113 (1982) 231.
[34] G. Veneviano, S. Yankielowicz, T. Taylor, Nucl. Phys. B 218 (1983) 493.
[35] A. Bilal, F. Ferrari, Nucl. Phys. B 516 (1998) 175, hep-th/9706145.
[36] A. Hanany, M. Strassler, A. Zaffaroni, Nucl. Phys. B 513 (1998) 87, hep-th/9707244.
[37] A. Gustavsson, M. Henningson, Phys. Lett. B 463 (1999) 551, hep-th/9906053.
[38] A. Shapere, C. Vafa, BPS structure of ArgyresDouglas superconformal theories, hepth/9910182.
[39] A. Gorsky, A. Marshakov, A. Mironov, A. Morozov, Phys. Lett. B 380 (1996) 75, hepth/9603140.
[40] A. Gorsky, S. Gukov, A. Mironov, Nucl. Phys. B 517 (1998) 409, hep-th/9707120.

A. Gorsky et al. / Nuclear Physics B 584 (2000) 197215

[41]
[42]
[43]
[44]

A. Gorsky, A. Mironov, Nucl. Phys. B 550 (1999) 513, hep-th/9902030.


S. Katz, C. Vafa, Nucl. Phys. B 497 (1997) 196, hep-th/9611090.
N. Dorey, JHEP 9907 (1999) 021, hep-th/9906011.
A. Yung, Nucl. Phys. B 562 (1999) 191, hep-th/9906243.

215

Nuclear Physics B 584 (2000) 216232


www.elsevier.nl/locate/npe

Anomalous dimensions
in N = 4 SYM theory at order g 4
Massimo Bianchi a , Stefano Kovacs b , Giancarlo Rossi a ,
Yassen S. Stanev a,1
a Dipartimento di Fisica, Universit di Roma Tor Vergata, I.N.F.N. Sezione di Roma Tor Vergata,

Via della Ricerca Scientifica, 1, 00173 Roma, Italy


b D.A.M.T.P., University of Cambridge, Wilberforce Road, Cambridge, CB3 0WA, UK

Received 30 March 2000; accepted 19 May 2000

Abstract
We compute four-point correlation functions of scalar composite operators in the N = 4
supercurrent multiplet at order g 4 using the N = 1 superfield formalism. We confirm the
interpretation of short-distance logarithmic behaviours in terms of anomalous dimensions of
unprotected operators exchanged in the intermediate channels and we determine the two-loop
contribution to the anomalous dimension of the N = 4 Konishi supermultiplet. 2000 Elsevier
Science B.V. All rights reserved.
PACS: 11.15.B; 12.60.Jv; 11.25.Hf; 11.10.Gh; 11.10.Hi; 11.15.Bt; 12.38.Bx
Keywords: Anomalous dimension; Superconformal invariance; Supersymmetry; Superdiagrams; AdS/CFT
correspondence; Renormalization

1. Introduction and summary of the results


The conjectured AdS/CFT correspondence [14] has renewed the interest in conformal
field theories (CFTs), in particular N = 4 supersymmetric YangMills theory (SYM)
in D = 4, and has raised the issue of finding to what extent high order perturbative
computations are feasible in the weak coupling regime.
The interest is twofold. On the one hand, given the explicit analytic expressions of
certain four-point amplitudes in the AdS context, one may ask whether it is possible to
recognize some systematic pattern in the perturbative N = 4 field-theoretic results. On the
other hand, insights gained in such finite theories may prove to be useful in theories with
(partial) supersymmetry breaking.
1 On leave of absence from Institute for Nuclear Research and Nuclear Energy, Bulgarian Academy of Sciences,
BG-1784, Sofia, Bulgaria.

0550-3213/00/$ see front matter 2000 Elsevier Science B.V. All rights reserved.
PII: S 0 5 5 0 - 3 2 1 3 ( 0 0 ) 0 0 3 1 2 - 6

M. Bianchi et al. / Nuclear Physics B 584 (2000) 216232

217

In CFT the dependence of two- and three-point functions of scalar primary operators on
the external insertion points is completely fixed by conformal symmetry. For protected
operators non-renormalisation theorems hold that prevent quantum corrections to the
lowest order contribution. Explicit computations have confirmed the validity of these
theorems in the context of the AdS/CFT correspondence [57]. Except for extremal [8
10] and sub-extremal correlators [11], no such non-renormalisation theorems are believed
to hold for higher-point functions.
A particularly interesting class of higher-point functions are the four-point functions of
the lowest chiral primary operators, i.e. dimension-two gauge-invariant scalar composites
in the N = 4 supercurrent multiplet. Perturbative computations involving such operators
have been performed at order g 2 in [1214]. Motivated by the presence of string corrections
to the AdS effective action [1519], instanton computations have been carried out in [14,
2024]. All such computations show logarithmic behaviours at short-distances that allow
an interpretation in terms of anomalous dimensions of unprotected operators in long
supermultiplets exchanged in intermediate channels [14]. Similar analyses have been
performed in the AdS context in the supergravity limit [25]. A detailed comparison of
the two classes of results is problematic because of the different regimes in the (g 2 , N )
parameter space that are explored by the two approaches. Nevertheless, explicit AdS
computations in the scalar sector that are amenable to a perturbative analysis in the
field-theoretic weak coupling regime, such as the one presented here, have been recently
completed in [26] in the supergravity approximation with the help of the results of [27].
In this paper, we compute at order g 4 four-point correlation functions of the lowest
dimension scalar composite operators in the N = 4 supercurrent multiplet using the N =
1 superfield formalism. We confirm the interpretation of the short-distance logarithmic
behaviour in terms of anomalous dimensions of unprotected operators exchanged in
intermediate channels and we extract the two-loop contribution to the anomalous
dimension of the N = 4 Konishi supermultiplet. In order to illustrate our point we will
concentrate in most of our presentation on the simplest of the six independent four-point
functions of chiral primary operators [1214] (see below). Identities that have been proved
to hold both in perturbation theory in [12,13,28] and at the one-instanton level in [14]
should be enough to determine all four-point functions from the one we consider.
A recent interesting paper by Eden, Schubert and Sokatchev has reported on similar
computations at order g 4 from a different vantage point and confirmed the validity of
the above identities [29]. Their approach is based on the less familiar N = 2 harmonic
superspace formalism. It is remarkable that, although we adopt a different approach, we
find the same result for the four-point function on which we focus our attention, up to
an overall constant that was not fixed in [29]. We would like to stress that, contrary
to expectations, the N = 1 superfield approach we have pursued is not prohibitively
complicated and it is of wider applicability. In particular we have in mind applications
to some interesting finite N = 1 superconformal theories that are promising candidates for
realistic theories after soft supersymmetry breaking.
The plan of the paper is as follows. In Section 2, for the sake of completeness, we will
write down the action, propagators and vertices in the N = 1 superfield formalism and

218

M. Bianchi et al. / Nuclear Physics B 584 (2000) 216232

identify the six independent four-point functions of chiral primary operators. In Section 3
we draw the relevant Feynman superdiagrams and sketch the computation for the simplest
possible four-point function. In Section 4 we discuss the final result and interpret the
dominant logarithmic behaviours at short distance in terms of the anomalous dimension
of the lowest dimensional operator, K1 , in the N = 4 Konishi supermultiplet. We confirm
the one-loop results of [14] and extract the two-loop contribution 2 to the anomalous
dimension of K1 . We finally comment on possible extensions of our results to finite N = 1
supersymmetric theories that arise after soft breaking of N = 4 supersymmetry.
2. N = 4 SYM in N = 1 superspace
The field content of N = 4 SYM [30,31] comprises a vector, A , four Weyl spinors, A
(A = 1, 2, 3, 4), and six real scalars, i (i = 1, 2, . . . , 6), all in the adjoint representation
of the gauge group G. Since no off-shell formulation is available that manifestly preserves
N = 4 supersymmetry, in order to compute quantum corrections one has to resort either
to the N = 2 harmonic superspace approach pursued in [12,13,29] or to the more familiar
N = 1 formalism pursued in [14].
In the N = 1 formalism the fundamental fields can be arranged into a vector superfield,
V , and three chiral superfields, I (I = 1, 2, 3). The six real scalars, i , are combined into
three complex fields, namely


1
1
I = I i I +3 ,
(1)
I = I + i I +3 ,
2
2
that are the lowest components of the superfields I and I , respectively. Three of
the Weyl fermions, I , are the spinors of the chiral multiplets. The fourth spinor, =
4 , together with the vector, A , form the vector multiplet. In this formulation only an
SU(3) U (1) subgroup of the original SU(4) R-symmetry group is manifest. I and I
transform in the representations 3 and 3 of the global SU(3) flavour Q-symmetry, while
V is a singlet. Under the axial U (1) R-symmetry the vector A is neutral, the gaugino
has charge +3/2, the spinors of the chiral multiplets I have charge 1/2 and the three
complex scalars I have charge +1.
The (Euclidean) action in the N = 1 superfield formulation reads
Z
 Z

Z
1
d2 W W + h.c. +
d2 d2 egV I egV I
S = 2 tr
d4 x
16
Z

Z



g

d2 I J K I J , K d2 I J K I J , K
,
(2)

3! 2
where W is the chiral superfield-strength of V

1
(3)
W = D 2 e2gV D e2gV .
4g
2 Notice that we call `-loops calculations that are of order g 2` . The authors of Refs. [12,13] and [29] dub
diagrams of this order as (` + 1)-loop calculations.

M. Bianchi et al. / Nuclear Physics B 584 (2000) 216232

219

The trace over the colour indices is defined by


1
tr(T a T b ) = ab ,
2

a, b = 1, . . . , dim(G),

(4)

where T a are the generators in the fundamental representation of the gauge group.
Notice that a gauge fixing term, Sgf , has to be added to the classical action (2). As usual
we decide to take
Z
Z


1
(5)
tr d4 x d2 d2 (D 2 V )( D 2 V ) ,
Sgf =
16
where is a gauge parameter. We will not display here ghost terms since they do not
contribute to the Green functions that we will consider at the order we work. The choice
= 1 greatly simplifies all the computations, as it makes all corrections to the propagators
of the fundamental superfields vanishing at order g 2 [32,35].
Expanding the exponentials egV in (2) gives

Z
g2
a
a
4
2
2
S = d x d d V a Va Ia aI igfabc I V b I c + fabe fecd I V b V c I d
2



ig

abc
I J K
IJK
f
aI bJ cK ( ) + ,
(6)
I J K a b c ( )
3! 2
where fabc are the structure constants of the gauge group and we have displayed only the
terms that are relevant for our subsequent computations.
In the following we will carry on the calculations using N = 1 formalism in coordinate
superspace, that is more suitable for the study of CFTs. The superfield propagators in the
conventions of Eq. (6) are

I J ab ( + 2 ) 1
e ii jj ji j 2 ,
Ia (xi , i , i )bJ (xj , j , j ) =
4 2
xij

(7)


ab (ij )(ij )
,
Va (xi , i , i )Vb (xj , j , j ) = 2
8
xij2

(8)

where
xij = xi xj ,

ij = i j ,

ij = i j .

(9)

In the next section we will describe the calculation of the order g 4 perturbative correction
to the four-point correlation function

(x2 )C 22(x3 )C22


(x4 ) ,
(10)
G(H ) (x1 , x2 , x3 , x4 ) = C 11 (x1 )C11
where the gauge-invariant composite operators

CIJ = tr I J
C I J = tr( I J ),

(11)

are the lowest components of the (anti-)chiral superfields



CIJ = tr I J .
C I J = tr( I J ),

(12)

220

M. Bianchi et al. / Nuclear Physics B 584 (2000) 216232


ij

In turn C I J and CIJ appear in the decomposition of Q20 , the lowest scalar components of
ij
the N = 4 current supermultiplet, under SU(4) SU(3) U (1). The Q20 belong to the
real representation 20 of SU(4) and are defined as


ij
ij
k k .
(13)
Q20 = tr i j
6
ij

The most general four-point function of the Q20


G(Q) (x1 , x2 , x3 , x4 ) = Qi1 j1 (x1 )Qi2 j2 (x2 )Qi3 j3 (x3 )Qi4 j4 (x4 )

(14)

can be expressed as a linear combination of G(H ) , defined in Eq. (10), and the following
five independent four-point functions

(x2 )C 11 (x3 )C11


(x4 ) ,
G(V ) (x1 , x2 , x3 , x4 ) = C 11 (x1 )C11

(x2 )C 22 (x3 )C11


(x4 ) ,
G(1)(x1 , x2 , x3 , x4 ) = C 11 (x1 )C22

(x3 )C22
(x4 ) ,
G(2)(x1 , x2 , x3 , x4 ) = C 11 (x1 )C 22(x2 )C11

(x2 )C11
(x3 )C 11 (x4 ) ,
G(3)(x1 , x2 , x3 , x4 ) = C 11 (x1 )C11

(x3 )C11
(x4 ) .
(15)
G(4)(x1 , x2 , x3 , x4 ) = C 11 (x1 )C 11(x2 )C11
We stress that the above five linearly independent four-point functions have been shown
to be functionally related to (10) both in perturbation theory [12,13,28,29] and nonperturbatively at the one-instanton level [14]. For instance, one finds
G(H ) (x1 , x2 , x3 , x4 ) = sG(V ) (x1 , x2 , x3 , x4 ),

(16)

where s is one of the two independent conformally invariant cross-ratios


r=

2 x2
x12
34
2 x2
x13
24

s=

2 x2
x14
23
2 x2
x13
24

(17)

that can be constructed out of four points.

3. Superdiagrams and calculations


The four-point function G(H ) (x1 , x2 , x3 , x4 ) defined in (10) is the lowest component of
the supercorrelation function

(z2 )C 22 (z3 )C22


(z4 )
(18)
(H ) (z1 , z2 , z3 , z4 ) = C 11 (z1 )C11
viz.


G(H ) (x1 , x2 , x3 , x4 ) = (H ) (z1 , z2 , z3 , z4 ) = =0 ,
i

(19)

where zi = (xi , i , i ) as usual. The N = 1 superfield approach greatly simplifies


the calculations. The choice of external flavours guarantees that there are no connected
supergraphs at tree level. The potential corrections to the propagators at order g 4 would

M. Bianchi et al. / Nuclear Physics B 584 (2000) 216232

221

Fig. 1. Diagrams with only chiral lines.

Fig. 2. Diagrams with two vector lines.

only contribute to disconnected supergraphs 3 since the choice = 1 in the gauge-fixing


term makes all propagator corrections vanish at order g 2 [3236].
The connected superdiagrams contributing to (18) are reported in Figs. 1, 2 and 3. The
numbering of the chiral lines refers to the SU(3) flavours. We have only displayed superdiagrams that do not vanish by colour contractions. There are 4 superdiagrams with only
chiral lines, 21 superdiagrams with one internal vector line and 10 superdiagrams with two
internal vector lines. Notice that none of them involves cubic or quartic pure vector vertices.
This explains why we have not explicitly displayed the corresponding couplings in (6).
The computation of the overall weights, due to combinatorial factors and colour and
flavour contractions, is greatly simplified if fake different coupling constants are introduced
3 Order g 4 corrections to two-point functions of operators belonging to ultra-short supermultiplets vanish [37].

222

M. Bianchi et al. / Nuclear Physics B 584 (2000) 216232

Fig. 3. Diagrams with one vector line.

M. Bianchi et al. / Nuclear Physics B 584 (2000) 216232

223

for each interaction term in the action. Only at the very end one imposes the relations
among the couplings that guarantee N = 4 supersymmetry. As a result one finds 4 for the
superdiagrams with only chiral lines
g4 2 2
N (N 1)(2A1 + A2 + A3 + A4 ),
8
for those with one vector line
Atot =

(20)

g4 2 2
N (N 1)(2B1 + 2B2 + B3 + B4 B5 2B6 + 2B7 + 2B8 + B9
4
+ B10 B11 2B12 + B13 + B14 + B15 + B16

Btot =

+ B17 B18 B19 B20 B21 ),

(21)

and for the superdiagrams with two vector lines


Ctot =

g4 2 2
N (N 1)
2
(2C1 + 2C2 2C3 2C4 + C5 C6 + 2C7 + 2C8 C9 ).

(22)

Note that C10 is zero due to Grassmann integration over the s of the interaction
vertices. In general, the integration over the variables of the interaction vertices is
greatly simplified by our choice of computing the lowest component of the supercorrelation
function (18). This amounts to put to zero the variables of the external insertions.
The final result is a complicated combination of double, triple and quadruple integrals
of convolutions of massless scalar propagators and derivatives thereof.
The general strategy for computing the resulting integrals is the following. Exploiting
the conformal invariance of the full correlation function we send one of the external points
(say x1 ) to infinity after multiplying the correlator by (x12 ) , where is the conformal
dimension of the operator inserted at x1 . Note that this simple prescription only works
for operators (like C) with protected conformal dimension ( = 2 in the case at hand), i.e.,
independent of the coupling constant g 2 . For operators that acquire anomalous dimensions,
one has to carefully subtract divergent terms that behave like powers of log(x12 ) as x1
2 , x 2 and x 2 . The
. In this limit the final answer will depend on the three variables x23
24
34
complete x1 dependence is dictated by conformal invariance and can be recovered after the
2 x 2 x 2 , x 2 x 2 x 2 and x 2 x 2 x 2 .
substitutions x23
23 14 24
24 13
34
34 12
Taking the limit x1 considerably simplifies the computation for two reasons. On
the one hand it lowers the number of propagators to be integrated. On the other hand, before
taking the limit, one can integrate by parts some of the derivatives in the chiral propagators
and make them act on propagators involving the point x1 . This increases the power of x1 in
the denominator making it larger than 4 with the consequence that terms generated in this
way vanish in the limit. This trick also simplifies the integrals. As a result one is left with
only double integrals except for the diagrams B21 and C6 where the remaining integral
4 In what follows we will only display group theory factors relevant for the case G = SU(N ). The generalization
to an arbitrary gauge group amounts to substituting g 2 N with g 2 C2 (A) and N 2 1 with dim(G), C2 (A) being
the quadratic Casimir of the adjoint representation.

224

M. Bianchi et al. / Nuclear Physics B 584 (2000) 216232

is (naively) a triple one. A drawback of our procedure is that the limit x1 breaks
many of the permutation symmetries of the integrand, thus diagrams that were related
by rotations (or reflections) to one another must be separately computed. An important
simplification arises by observing that in the limit x1 certain linear combinations of
diagrams contributing to (21) and (22) actually vanish. Precisely one finds B1 = 0, B9 = 0,
B15 B18 = 0, 2B2 B5 B20 = 0, B3 2B6 + B13 + B17 B19 = 0, C1 C3 = 0,
C2 + C8 = 0, 2C4 C5 + C9 = 0. As already observed C10 = 0 due to the integration
even before taking the limit.
The whole result can be expressed in terms of only two functions
Z
dx5
(23)
g(i, j, k) =
2
2
xi5 xj25 xk5
and

Z
f (i; j, k) =

dx5 dx6
,
2
2 x2 x2
xi5 xj25 xi6
k6 56

(24)

where in the notation used in [12]


f (i; j, k) f (i, j ; i, k).
A useful expression for the function g is [14]
 2
2 
x
x24
2
,
,
g(2, 3, 4) = 2 B 34
2 x2
x23
x23
23

(25)

(26)

where B(r, s) is a box-type integral given by


Z
B(r, s) = d0 d1 d2 (1 0 1 2 )

1
.
(27)
1 2 + r0 1 + s0 2
The result of the integration in (27) can be expressed as a combination of logarithms and
dilogarithms as follows:

 

r +s1 p 2
1
B(r, s) = ln(r) ln(s) ln
p
2




2
2
2Li2
,
(28)
2Li2

1+r s+ p
1r +s + p
with
Li2 (z) =

n
X
z
n=1

n2

and p = 1 + r 2 + s 2 2r 2s 2rs.

The explicit expression of the function f (i; j, k) has been obtained in [38]
 2
2 
4 (2) xij xik
,
,
f (i; j, k) = 2
xj k
xj2k xj2k

(29)

(30)

where (2) (r, s) involves polylogarithms of up to fourth order. For the purpose of the
present investigation, however, we only need the following identities [12,13]
1
(31)
j f (i; j, k) = 4 2 2 g(i, j, k),
xij

M. Bianchi et al. / Nuclear Physics B 584 (2000) 216232

i f (i; j, k) = 4 2

xj2k
2
xij2 xik

g(i, j, k).

225

(32)

The latter identity has to be used in both directions to simplify the integrals appearing in
the diagrams A2 , B7 , B8 and B11 . The triple integrals that appear in the computation of the
diagrams B21 and C6 can be simplified with the help of the relation
Z
2 
2
2
dx5 
2 x24
=
4
(33)
g(3,
4,
5)
g(2, 3, 4) .
3
2
2
x25
x34
Eq. (33) can in turn be proved using the identity

2 

2
2
2
3 g(2, 3, 4) ,
g(2, 3, 4) = x34
2 x24

(34)

which can be verified by resorting to the explicit expression of g(2, 3, 4) given by Eq. (26).
The total contributions of each of the three classes of diagrams of Figs. 1, 2 and 3 is


 
2
g 4 N 2 (N 2 1) 2 
f
(3;
2,
4)
+
f
(2;
3,
4)
+
g(2,
3,
4)
,
(35)
Atot =
2
8(4 2 )8
x34
g 4 N 2 (N 2 1) 1
2
4(4 2)8
4x34

2 2


2
2
g(2, 3, 4) x24
+ x34
x23
4f (3; 2, 4) 4f (4; 2, 3) ,

Btot =

Ctot =


2
g 4 N 2 (N 2 1) 1  2 
x24 g(2, 3, 4) 2f (3; 2, 4) .
2
8
2
2(4 )
4x34

(36)
(37)

Summing the above three contributions yields


g 4 N 2 (N 2 1)
2
x1
4(4 2 )8 x34



1 2
2
2
f (2; 3, 4) + f (3; 2, 4) + f (4; 2, 3) + x24
+ x34
+ x23
[g(2, 3, 4)]2 . (38)
4

G(x2 , x3 , x4 ) lim x14 G(H ) (x1 , x2 , x3 , x4 ) =

As already explained, the actual x1 dependence is recovered by performing in the right


2 x 2 x 2 , x 2 x 2 x 2 and x 2 x 2 x 2 . The
hand side of (38) the substitutions x23
23 14
24
24 13
34
34 12
final result is in perfect agreement with the expression of the corresponding four-point
function of hypermultiplet bilinears computed by means of N = 2 harmonic superspace
techniques in [29].

4. Logarithms and anomalous dimensions


As in previous computations at order g 2 [1214] and at the one-instanton level [14],
the function (38) shows the expected short-distance logarithmic behaviour. Indeed, at short
distances (i.e., in the limit in which pairs of points are taken to coincide) one finds linear
and quadratic logarithms that are not incompatible with the finiteness of N = 4 SYM in
the superconformal phase.

226

M. Bianchi et al. / Nuclear Physics B 584 (2000) 216232

To illustrate this point in a simple fashion we will follow closely [14] and consider as an
example the two-point function of a primary operator of scale dimension


A
,
(39)
O (x)O (y) =
(x y)2
where A is an overall normalisation constant possibly depending on the subtraction
(0)
scale . Now suppose that = + , i.e., the operator under consideration has an
anomalous dimension. In perturbation theory = (g 2 ) is expected to be small and to
admit an expansion in the coupling constant g 2
(g) = 1 g 2 + 2 g 4 + .

(40)
g2

yields
The perturbative expansion of (39) in powers of






a
1 g 2 1 log 2 (x y)2
O (x)O (y) =
(0)
(x y)2



 2
 2


4 1 2
2 2
2
log (x y)
2 log (x y)
+ ,
+g
2 1

(41)

where after renormalisation we have set A = a 2 . Thus, although the exact


expression (39) is conformally invariant, at each order in g 2 Eq. (41) contains powers
of logarithms that seem to even violate scale invariance. Similar considerations apply to
generic n-point Green functions as well.
Assuming the validity of the OPE, a four-point function can be expanded in the schannel in the form

QA (x)QB (y)QC (z)QD (w)


X CCD K (z w, w )


(42)
QA (x)QB (y)OK (w) ,
=

C
D
K
(z w)
K

where K runs over a complete set of primary operators. Descendants are implicitly taken
into account by the presence of derivatives in the Wilson coefficients, Cs. An expansion
like (42) is valid in the other two channels as well. To simplify formulae we assume that
QA , QB , QC , QD are protected operators, i.e., they have no anomalous dimensions. In
(0)
general the operators OK may have anomalous dimensions, K , so that K = K + K ,
(0)
(0)
where K is the tree-level scale dimension. Similarly CI J K = CI J K + I J K , with I J K
the perturbative correction to the OPE coefficients.
The three-point function in the right hand side of Eq. (42) is determined by conformal
invariance to be of the form

QA (x)QB (y)OK (z)


=

(x

y)A +B K (x

CABK (g 2 )
.
z)A B +K (y z)B A +K

(43)

Eq. (43), like the two-point function (41), can be expanded in power series in g 2 giving
rise to logarithmic terms. From these formulae one can extract the corrections to both the
OPE coefficients and the anomalous dimensions of the operators OK . The same procedure

M. Bianchi et al. / Nuclear Physics B 584 (2000) 216232

227

applies to derivatives of the three- and four-point functions as well as to their limits for
x1 .
Let us analyze the function G(H ) in the limit x4 x3 that exposes the s-channel. Only
operators in the singlet of the manifest SU(3) U (1) may be exchanged [14]. Barring the
identity operator, that is clearly not renormalized, the leading contribution comes from the
operator
K1 =

1
: tr( i i ) :
3

(44)
(0)

that has naive dimension = 2 and is the lowest component of the long Konishi
supermultiplet.
In the case under consideration, the relative complexity of the explicit expression
for f (i; j, k) makes it difficult to directly analyse the short distance behaviour of the
function (38). Thus we find it more convenient to compute 2 G(x2 , x3 , x4 ), which can
be expressed in terms of only the much simpler function g(i, j, k). Using Eqs. (31), (32)
and (34), one obtains
 4
g 4 N 2 (N 2 1) X
j g(2, 3, 4) j g(2, 3, 4)
2 G(x2 , x3 , x4 ) =
8(4 2 )8
j =2



1
1
1
(45)
3(4 2 ) 2 2 + 2 2 + 2 2 g(2, 3, 4) .
x23x34 x24 x34 x23x24
The leading behaviour of (45) in the limit x4 x3 is given by

 2 

x34
g 4 N 2 (N 2 1)
1
3 log 2 5 .
2 G(x2 , x3 , x4 )
4 x2
16(4 2 )6
x23
x23
x4 x3
34

(46)

We now compare Eq. (46) with the result of the OPE analysis of the order g 4 contribution
to the function 2 G(x2 , x3 , x4 ). The latter yields
 2 


x34
g 4 (N 2 1)
2
2
2 G(x2 , x3 , x4 )
a0 1 log 2 + a0 (1 + 22) + 2a1 1 ,
2 6 4 2
x23
x4 x3 (4 ) x23 x34
(47)
where the coefficients a0 and a1 are given by
1
a1 = N 2 .
(48)
a0 = (4 2 )2 ,
3
They represent the tree-level contribution of the Konishi scalar K1 to the s-channel
expansion of the function G(H ) and its finite one-loop correction. The coefficients a0 and
a1 are determined by the one-loop analysis performed in [14] that gives
!


2
x34
g 2 (N 2 1) 1
1
log
+
a
a
G(x2 , x3 , x4 )|1-loop
0
1
2 x2
2
2
(4 2 )6 x23
x23
x4 x3
34
 2 


x34
1
g 2 (N 2 1) 1
2
log

1
.
(49)
=
2 x2
2
2
(4 2 )6 x23
x23
34

228

M. Bianchi et al. / Nuclear Physics B 584 (2000) 216232

We stress that the quadratic logarithmic term in G(H ) contributes a linear logarithmic term
to 2 G(H ) in Eq. (47).
Comparison of Eq. (46) with Eq. (47) shows that the coefficient of the logarithmic term
is related to the square of the one-loop anomalous dimension of K1 , and agrees with
3N
the known value 1 = 16
2 [3942]. The constant term corresponds to a combination
of the one- and two-loop anomalous dimensions and the known one-loop correction to
the OPE coefficient. For the g 4 contribution to the anomalous dimension of the Konishi
supermultiplet one obtains
3g 4 N 2
.
(50)
16(4 2)2
The analysis of the other two non-singlet channels (x24 0 and x23 0) is more
subtle. The lowest dimensional operators that can be exchanged in both channels belong to
the 105 and to the 84 representations of the SU(4) R-symmetry. Single- and double-trace
operators of dimension four in the 105 are expected to be protected [7,43,44]. In order to
disentangle the two SU(4) representations it is necessary to consider another four-point
function. We find it convenient to analyse the four-point correlator G(V ) , defined in (15).
One-loop [12,13], one-instanton [14,20] and two-loop [29] computations give sG(V ) =
G(H ) . The OPE analysis of the g 4 contribution to G(V ) , as it is obtained from the above
functional relation, confirms the non-renormalisation of the 105 operators.
The two possible operators of naive dimension four in the 84 mix at one-loop [14]. One
of them, K84 , is a superconformal descendant of K1 and as such has the same anomalous
dimension. A careful OPE analysis, combined with the symmetry of the factor in brackets
in the right hand side of Eq. (38) under the exchange of x2 and x4 , implies that this operator
b84 ,
saturates the logarithmic behaviour in this channel. We thus conclude that the operator D
4
identified in [14] as the combination orthogonal to K84 , is protected also at order g . This
is in agreement with the fact that it belongs to a supermultiplet satisfying a generalized
shortening condition [4553].
2-loop = 2 g 4 =

5. Comments
Let us briefly comment on the bearing of the results of this paper. First of all the very
fact that it was possible to compute in closed form a four-point function of protected
composite operators at order g 4 shows that N = 4 SYM is both non-trivial and calculable.
No off-shell approach is known that preserves N = 4 supersymmetry. At the quantum
level, one has to resort either to the N = 1 superspace approach pursued here or to the
less familiar N = 2 harmonic superspace approach pursued in [29]. The coincidence of all
known results in the two approaches gives independent support to some N = 2 harmonic
superspace identities [54], based on the bonus symmetry proposed in [55], that led to
drastic simplifications in the computations reported in [29].
The introduction of supersymmetric mass-terms gives rise to interesting, in some cases
confining, theories that can be handled with N = 1 superfield techniques. Alternatively
one might consider other finite N = 1 gauge theories some of which are conjectured to be

M. Bianchi et al. / Nuclear Physics B 584 (2000) 216232

229

dual to type IIB superstring on AdS5 S 5 / , where is a discrete subgroup of an SU(3)


subgroup of the SU(4) R-symmetry. These gauge theories govern the dynamics of the light
degrees of freedom of stacks of coincident D3-branes at special orbifold singularities [57
61]. Other N = 1 finite theories can be obtained by perturbing N = 2 theories by massterms. This is the case of D3-branes at generalized conifold singularities [62], whose
supergravity dual replaces S 5 with less trivial Einstein spaces [63,64]. In all such cases the
N = 1 formalism pursued in this paper might allow one to compute correlation functions
of protected composite operators. By OPE one would then extract anomalous dimensions
and couplings of unprotected operators such as those belonging to Konishi-like long
supermultiplets. These are expected to be dual to genuine string excitations and as such
should decouple from the operator algebra in the supergravity limit, that is dual to the large
N and strong t Hooft coupling limit. The fact that the one-loop anomalous dimension of
the Konishi multiplet has been known for some time [3942] to be positive was somewhat
reassuring in this respect. The OPE analysis of the two-loop computations confirm the oneloop result, but at the same time yields a negative two-loop contribution to the anomalous
dimension. This result certainly requires further investigation and some cross-checks [65].
Another related issue is the role of multi-trace operators that are expected to be dual to
multiparticle states in the AdS description. We have shown and confirmed in the present
paper that their mixing with single trace operators is not suppressed in the large N limit at
finite t Hooft coupling. In fact protected operators belonging to supermultiplets that satisfy
generalized shortening conditions [4553] are typically mixtures of single- and multi-trace
operators. One would like to clarify their rle in the operator algebra in relation to the
string exclusion principle" that is expected to drastically reduce the spectrum of AdS
excitations at finite radius", i.e., at finite N [66,67].

Acknowledgements
We would like to thank E. DHoker, S. Ferrara, M. Gunaydin and J.F. Morales for useful
discussions. Ya.S.S. and S.K. would like to thank the Physics Department and I.N.F.N.
Section at Universit di Roma Tor Vergata for hospitality and financial support.
References
[1] J. Maldacena, The large N limit of superconformal field theories and supergravity, Adv. Theor.
Math. Phys. 2 (1998) 231, 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.
[3] E. Witten, Anti De Sitter space and holography, Adv. Theor. Math. Phys. 2 (1998) 253.
[4] O. Aharony, S.S. Gubser, J. Maldacena, H. Ooguri, Y. Oz, Large N field theories, string theory
and gravity, Phys. Rept. 323 (2000) 183, hep-th/9905111.
[5] S. Lee, S. Minwalla, M. Rangamani, N. Seiberg, Three-point functions of chiral operators in
D = 4 N = 4 SYM at large N , Adv. Theor. Math. Phys. 2 (1998) 697, hep-th/9806074.
[6] E. DHoker, D.Z. Freedman, W. Skiba, Field theory tests for correlators in the AdS/CFT
correspondence, Phys. Rev. D 59 (1999) 045008, hep-th/9807098.

230

M. Bianchi et al. / Nuclear Physics B 584 (2000) 216232

[7] W. Skiba, Correlators of short multi-trace operators in N = 4 supersymmetric YangMills


theory, Phys. Rev. D 60 (1999) 105038, hep-th/9907088.
[8] E. DHoker, D.Z. Freedman, S.D. Mathur, A. Matusis, L. Rastelli, Extremal correlators in the
AdS/CFT correspondence, hep-th/9908160.
[9] M. Bianchi, S. Kovacs, Non-renormalisation of extremal correlators in N = 4 SYM theory,
Phys. Lett. B 468 (1999) 102, hep-th/9910016.
[10] B. Eden, P.S. Howe, C. Schubert, E. Sokatchev, P.C. West, Extremal correlators in fourdimensional SCFT, Phys. Lett. B 472 (2000) 323, hep-th/9910150.
[11] J. Erdmenger, M. Perez-Victoria, Non-renormalisation of next-to-extremal correlators in N = 4
SYM and the AdS/CFT correspondence, hep-th/9912250.
[12] B. Eden, P.S. Howe, C. Schubert, E. Sokatchev, P.C. West, Four-point functions in N =
4 supersymmetric YangMills theory at two loops, Nucl. Phys. B 557 (1999) 355, hepth/9811172.
[13] B. Eden, P.S. Howe, C. Schubert, E. Sokatchev, P.C. West, Simplifications of four-point
functions in N = 4 supersymmetric YangMills theory at two loops, Phys. Lett. B 466 (1999)
20, hep-th/9906051.
[14] M. Bianchi, S. Kovacs, G.C. Rossi, Ya.S. Stanev, On the logarithmic behaviour in N = 4 SYM
theory, J. High Energy Phys. 08 (1999) 020, hep-th/9906188.
[15] M.B. Green, M. Gutperle, Effects of D-instantons, Nucl. Phys. B 498 (1997) 195, hepth/9701093.
[16] M.B. Green, M. Gutperle, D-particle bound states and the D-instanton measure, J. High Energy
Phys. 9801 (1998) 005, hep-th/9711107.
[17] M.B. Green, P. Vanhove, D-instantons, strings and M-theory, Phys. Lett. B 408 (1997) 122,
hep-th/9704145.
[18] M.B. Green, M. Gutperle, H. Kwon, Sixteen-fermion and related terms in M-theory on T 2 ,
Phys. Lett. B 421 (1998) 149, hep-th/9710151.
[19] T. Banks, M.B. Green, Non-perturbative effects in AdS5 S 5 string theory and d = 4 SUSY
YangMills, J. High Energy Phys. 05 (1998) 002, hep-th/9804170.
[20] M. Bianchi, M.B. Green, S. Kovacs, G.C. Rossi, Instantons in supersymmetric YangMills and
D-instantons in IIB superstring theory, J. High Energy Phys. 08 (1998) 013, hep-th/9807033.
[21] M. Bianchi, S. Kovacs, YangMills instantons vs. type IIB D-instantons, hep-th/9811060.
[22] N. Dorey, V.V. Khoze, M.P. Mattis, S. Vandoren, YangMills instantons in the large-N limit
and the AdS/CFT correspondence, Phys. Lett. B 442 (1998) 145, hep-th/9808157.
[23] N. Dorey, T.J. Hollowood, V.V. Khoze, M.P. Mattis, S. Vandoren, Multi-instantons and
Maldacenas conjecture, J. High Energy Phys. 06 (1999) 023, hep-th/9810243.
[24] N. Dorey, T.J. Hollowood, V.V. Khoze, M.P. Mattis, S. Vandoren, Multi-instanton calculus and
the AdS/CFT correspondence in N = 4 superconformal field theory, Nucl. Phys. B 552 (1999)
88, hep-th/9901128.
[25] E. DHoker, D.Z. Freedman, S.D. Mathur, A. Matusis, L. Rastelli, Operator product expansion
of N = 4 SYM and the 4-point functions of supergravity, hep-th/9911222.
[26] G. Arutyunov, S. Frolov, Four-point functions of lowest weight CPOs in N = 4 SYM4 in
supergravity approximation, hep-th/0002170.
[27] G. Arutyunov, S. Frolov, Scalar quartic couplings in type IIB supergravity on AdS5 S 5 , hepth/9912210.
[28] B. Eden, P.S. Howe, A. Pickering, E. Sokatchev, P.C. West, Four-point functions in N = 2
superconformal theories, hep-th/0001138.
[29] B. Eden, C. Schubert, E. Sokatchev, Three-loop four-point correlator in N = 4 SYM, hepth/0003096.
[30] L. Brink, J. Scherk, J.H. Schwarz, Supersymmetric YangMills theories, Nucl. Phys. B 121
(1977) 77.

M. Bianchi et al. / Nuclear Physics B 584 (2000) 216232

231

[31] F. Gliozzi, D.I. Olive, J. Scherk, Supersymmetry, supergravity and the dual spinor model, Nucl.
Phys. B 122 (1977) 253.
[32] S. Ferrara, B. Zumino, Supergauge invariant YangMills theories, Nucl. Phys. B 79 (1974) 413.
[33] M. Grisaru, M. Rocek, W. Siegel, Zero value of the three-loop function in N = 4
supersymmetric YangMills theory, Phys. Rev. Lett. 45 (1980) 1063.
[34] W.E. Caswell, D. Zanon, Zero three-loop beta function in the N = 4 supersymmetric Yang
Mills theory, Nucl. Phys. B 182 (1981) 125.
[35] S. Kovacs, A perturbative re-analysis of N = 4 supersymmetric YangMills theory, hepth/9902047.
[36] S. Kovacs, N = 4 supersymmetric YangMills theory and the AdS/SCFT correspondence, PhD
thesis, hep-th/9908171.
[37] S. Penati, A. Santambrogio, D. Zanon, Two-point functions of chiral operators in N = 4 SYM,
J. High Energy Phys. 12 (1999) 006, hep-th/9910197.
[38] N.I. Ussyukina, A.I. Davydychev, Exact results for three and four point ladder diagrams with
an arbitrary number of rings, Phys. Lett. B 298 (1993) 363.
[39] D. Anselmi, D.Z. Freedman, M.T. Grisaru, A.A. Johansen, Universality of the operator product
expansion of SCFT4 , Phys. Lett. B 394 (1997) 329, hep-th/9608125.
[40] D. Anselmi, D.Z. Freedman, M.T. Grisaru, A.A. Johansen, Nonperturbative formulas for central
functions of supersymmetric gauge theories, Nucl. Phys. B 526 (1998) 543, hep-th/9708042.
[41] D. Anselmi, J. Erlich, D.Z. Freedman, A. Johansen, Positivity constraints on anomalies in
supersymmetric gauge theories, Phys. Rev. D 57 (1998) 7570, hep-th/9711035.
[42] D. Anselmi, The N = 4 quantum conformal algebra, Nucl. Phys. B 541 (1999) 369, hepth/9809192.
[43] L. Andrianopoli, S. Ferrara, KK excitations on AdS5 S 5 as N = 4 primary superfields,
Phys. Lett. B 430 (1998) 248, hep-th/9803171.
[44] L. Andrianopoli, S. Ferrara, On short and long SU(2, 2|4) multiplets in the AdS/CFT
correspondence, Lett. Math. Phys. 48 (1999) 145, hep-th/9812067.
[45] V.K. Dobrev, V.B. Petkova, All positive energy unitary irreducible representations of extended
conformal supersymmetry, Phys. Lett. B 162 (1985) 127.
[46] S. Ferrara, talk given at Strings 99.
[47] S. Ferrara, A. Zaffaroni, Superconformal field theories, multiplet shortening and the AdS5 S 5
correspondence, hep-th/9908163.
[48] L. Andrianopoli, S. Ferrara, E. Sokatchev, B. Zupnik, Shortening of primary operators in N extended SCFT4 and harmonic-superspace analyticity, hep-th/9912007.
[49] S. Ferrara, E. Sokatchev, Short representations of SU(2, 2|N ) and harmonic superspace
analyticity, hep-th/9912168.
[50] S. Ferrara, Superspace representations of SU(2, 2, |N ) superalgebras and multiplet shortening,
hep-th/0002141.
[51] M. Gunaydin, D. Minic, M. Zagerman, Novel supermultiplets of SU(2, 2|4) and the AdS5 /CFT 4
duality, Nucl. Phys. B 544 (1999) 737, hep-th/9810226.
[52] M. Gunaydin, D. Minic, M. Zagerman, 4d doubleton conformal theories, CPT and IIB string
on AdS5 S 5 , Nucl. Phys. B 534 (1998) 96, hep-th/9806042; Nucl. Phys. B 538 (1999) 531
(Erratum).
[53] P. Claus, M. Gunaydin, R. Kallosh, J. Rahmfeld, Y. Zunger, Supertwistors as quarks of
SU(2, 2|4), J. High Energy Phys. 05 (1999) 019, hep-th/9905112.
[54] P.S. Howe, C. Schubert, E. Sokatchev, P.C. West, Explicit construction of nilpotent covariants
in N = 4 SYM, hep-th/9910011.
[55] K. Intriligator, Bonus symmetry of N = 4 super YangMills correlation functions via AdS
duality, Nucl. Phys. B 551 (1999) 575, hep-th/9811047.
[56] K. Intriligator, W. Skiba, Bonus symmetry and the operator product expansion of N = 4 super
YangMills, Nucl. Phys. B 559 (1999) 165, hep-th/9905020.

232

M. Bianchi et al. / Nuclear Physics B 584 (2000) 216232

[57] S. Kachru, E. Silverstein, 4-D conformal theories and strings on orbifolds, Phys. Rev. Lett. 80
(1998) 4855, hep-th/9802183.
[58] M. Bershadsky, Z. Kakushadze, C. Vafa, String expansion as large N expansion of gauge
theories, Nucl. Phys. B 523 (1998) 59, hep-th/9803076.
[59] Z. Kakushadze, Gauge theories from orientifolds and large N limit, Nucl. Phys. B 529 (1998)
157, hep-th/9803214.
[60] A. Lawrence, N. Nekrasov, C. Vafa, On conformal field theories in four dimensions, Nucl.
Phys. B 533 (1998) 199, hep-th/9803015.
[61] M. Bershadsky, A. Johansen, Large N limit of orbifold field theories, Nucl. Phys. B 536 (1998)
141, hep-th/9803249.
[62] I. Klebanov, E. Witten, Superconformal field theory on three-branes at a CalabiYau singularity,
Nucl. Phys. B 536 (1998) 199, hep-th/9807080.
[63] A. Ceresole, G. DallAgata, R. Dauria, S. Ferrara, Spectrum of type IIB supergravity on
AdS5 T 11 : predictions on N = 1 SCFTs, Phys. Rev. D 61 (2000) 066001, hep-th/9905226.
[64] A. Ceresole, G. DallAgata, R. Dauria, S. Ferrara, Superconformal field theories from IIB
spectroscopy on AdS5 T 11 , Class. Quant. Grav. 17 (2000) 1017, hep-th/9910066.
[65] M. Bianchi, S. Kovacs, G.C. Rossi and Ya.S. Stanev, in preparation.
[66] J. Maldacena, A. Strominger, AdS3 black holes and a stringy exclusion principle, J. High Energy
Phys. 12 (1998) 005, hep-th/9804085.
[67] A. Giveon, D. Kutasov, N. Seiberg, Comments on string theory on AdS3 , Adv. Theor. Math.
Phys. 2 (1998) 733, hep-th/9806194.

Nuclear Physics B 584 (2000) 233250


www.elsevier.nl/locate/npe

R 4 terms in 11 dimensions and conformal anomaly


of (2, 0) theory
A.A. Tseytlin 1,2
Department of Physics, The Ohio State University, Columbus, OH 43210-1106, USA
Received 16 May 2000; accepted 6 June 2000

Abstract
Using AdS7 /CFT 6 correspondence we compute a subleading O(N) term in the scale anomaly
of (2, 0) theory describing N coincident M5 branes. While the leading O(N 3 ) contribution to the
anomaly is determined by the value of the supergravity action, the O(N) contribution comes from
a particular R 4 term (8-d Euler density invariant) in the 11-dimensional effective action. This R 4
term is argued to be part of the same superinvariant as the P-odd C3 R 4 term known to produce O(N)
contribution to the R-symmetry anomaly of (2, 0) theory. The known results for R-anomaly suggest
that the total scale anomaly extrapolated to N = 1 should be the same as the anomaly of a single
free (2, 0) tensor multiplet. A proposed explanation of this agreement is that the coefficient 4N 3
in the anomaly (which was found previously to be also the ratio of the 2-point and 3-point graviton
correlators in the (2, 0) theory and in the free tensor multiplet theory) is shifted to 4N 3 3N. 2000
Elsevier Science B.V. All rights reserved.

1. Introduction
Two known maximally (2, 0) supersymmetric conformal field theories in 6 dimensions
are the free tensor multiplet theory describing low energy dynamics of a single M5 brane,
and still largely mysterious interacting (2, 0) conformal theory describing N coincident
M5 branes. A way to study the latter theory is provided by its conjectured duality [1]
to M-theory (or, for large N , 11-d supergravity corrected by higher derivative terms) on
AdS7 S 4 background.
Comparison of the 2-point and 3-point correlators of the stress tensor of (2, 0) theory
as predicted by the AdS7 S 4 supergravity [2,3] to those in the free tensor multiplet theory shows [46] that they differ only by the overall coefficient 4N 3 . 3 The remarkable
1 tseytlin@mps.ohio-state.edu
2 Also at Blackett Laboratory, Imperial College, London and Lebedev Physics Institute, Moscow.
3 The same is true also for the correlators of R-symmetry currents [7].

0550-3213/00/$ see front matter 2000 Elsevier Science B.V. All rights reserved.
PII: S 0 5 5 0 - 3 2 1 3 ( 0 0 ) 0 0 3 8 0 - 1

234

A.A. Tseytlin / Nuclear Physics B 584 (2000) 233250

coefficient 4N 3 was originally found in [5] in the comparison of the M5 brane world volume theory and the D = 11 supergravity expressions for the absorption cross-sections of
longitudinally polarized gravitons by N coincident M5 branes. The same coefficient 4N 3
appears also as the ratio of the scale anomalies (or Weyl-invariant parts of conformal anomalies) of the interacting (2, 0) theory [8] and free theory of a single tensor multiplet [9].
The reason why the coefficient 4N 3 was puzzling in [5] was analogy with the d = 4 case:
a similar comparison of the gravitational and world-volume absorption cross-sections in the
case of D3-branes [5,10] led to the ratio N 2 , which is equal to 1 for N = 1. This agreement
in the d = 4 case was later understood [6] as being a consequence of nonrenormalization
of the conformal anomaly and thus of the 2-point stress tensor correlator in N = 4 SYM
theory. The analogy between the d = 4 and d = 6 cases should not, of course, be taken too
seriously, given that the (2, 0) theory should have a different structure than SYM theory,
being an interacting conformal fixed point without a free coupling parameter.
Still, one may expect that anomalies and 2- and 3-point correlators of currents of the
(2, 0) theory may have special protected form, with simple dependence on N , allowing
one to interpolate between N  1 and N = 1 cases.
This was, in fact, observed for the R-symmetry anomaly of the (2, 0) theory [11]:
the anomaly of the (2, 0) theory obtained from the 11-d action containing the standard
supergravity term plus a higher-derivative C3 R 4 term [12] is given by the sum of the
leading supergravity O(N 3 ) and subleading O(N) terms, and for N = 1 is equal to the
R-symmetry anomaly corresponding to the single tensor multiplet [13,14].
Since the conformal and R-symmetry anomalies of the (2, 0) theory should belong to
the same d = 6 supermultiplet [11,15], one should then expect to find a similar O(N)
correction to the O(N 3 ) supergravity contribution [8] to the (2, 0) conformal anomaly.
This O(N) correction should originate from a higher-derivative R 4 term in the 11-d action
which should be a part of the same superinvariant as C3 R 4 term (just like the secondderivative supergravity terms R and C3 F4 F4 are).
Our aim below is to discuss a mechanism of how this may happen. We shall argue that
the 11-d action contains a particular R 4 term, which, upon compactification on S 4 , leads
to a special combination of R 3 terms in the effective 7-d action. These R 3 corrections
produce extra O(N) terms in the conformal anomaly of the boundary (2, 0) conformal
theory. As a result, the coefficient 4N 3 in the ratio of the (2, 0) theory and tensor multiplet
scale anomalies may be shifted to 4N 3 3N . Since the latter is equal to 1 for N = 1, this
would be a resolution of the 4N 3 puzzle.
Since this conclusion is sensitive to numerical values of coefficients in the 11-d low
energy effective action we shall start with a critical review of what is known about the
structure of R 4 terms in type IIA string theory in 10-d and their counterparts in M-theory.
While the type IIB theory effective action contains the same J0 R 4 invariant at the tree
and one-loop levels, the one-loop term in type IIA theory is a combination of two different
R 4 structures. We shall argue that they should be organized into two different N = 2A
superinvariants J0 and I2 (containing P-odd B2 tr R 4 term) in a way different than it was
previously suggested (Section 2). The corresponding tow D = 10 superinvariants lifted
to D = 11 represent the leading R 4 corrections to the 11-d supergravity action (Section 3).

A.A. Tseytlin / Nuclear Physics B 584 (2000) 233250

235

These terms should be supplemented by proper F4 = dC3 dependent terms as required


by supersymmetry and chosen in a specific on-shell scheme not to modify the AdS7 S 4
solution of the D = 11 supergravity. Assuming that, in Section 4 we discuss higher
derivative corrections to the 7-d action of S 4 compactified theory which follow from the
presence of the R 4 terms in D = 11 action. In Section 5 we compute the corresponding
O(N) contributions to the scale anomaly of the (2, 0) theory using the method of [8],
and draw analogy between the total O(N 3 ) + O(N) result and the expression for the Rsymmetry anomaly found in [11].
2. R 4 terms in 10 dimensions
Let us start with a review of the structure of the R 4 terms in the effective actions of type
IIA superstring in 10 dimensions and the corresponding terms in M-theory effective action
in 11 dimensions, paying special attention to explicit values of numerical coefficients.
The relevant terms in the tree + one-loop type IIA string theory effective action can be
written in the form
S = S0 + S1 ,
S0 =

Z
d 10 x

2
210

1
S1 =
2 0

G e2 R

Z
d 10 x

G L1 ,


1
H32 + + b0 03 J0 ,
2 3!

L1 = b1 J0 + b2 K,

(2.1)

(2.2)

where Hmnk = 3[m Bnk] and 4


J0 = J1 + J2 t8 t8 RRRR +

1
10 10 RRRR,
4 2!

(2.3)

J0 = J1 J2 t8 t8 RRRR

1
10 10 RRRR,
4 2!

(2.4)



2
1
K = 10 B2 tr R 4 tr R 2 .
4

(2.5)

In the notation we are using the numerical coefficients are


2
= (2)7 gs2 04 ,
210

b0 =

1
(3),
3 211

b2 = 12b1 =

(2.6)
b1 =

(2)4

1
.
(2)4 3 211

1
,
32 213
(2.7)

4 We use Minkowski notation for the metric and  tensor, so that   = 10!, and upon reduction to 8 spatial
10 10
dimensions mn... mn... 28 8 . For other notation see also [16].

236

A.A. Tseytlin / Nuclear Physics B 584 (2000) 233250

The tree and one-loop coefficients of the well-known J1 = t8 t8 RRRR term 5 can be
determined from the 4-graviton amplitude [1719]. 6
1
The invariant J2 = 42!
10 10 RRRR which will play important role in what follows is
the D = 10 extension of the integrand of the Euler invariant in 8 dimensions
1
J2 = E8 ,
4
1
n
n
D D R 4 = 8![m1 1 . . . m88 ] R m1 m2n1 n2 R m7 m8n7 n8 ,
E8 =
(D 8)!

(2.8)

where correspond to the case of Euclidean or Minkowski signature. 7


The expansion of E8 near flat space (gmn = mn + hmn ) starts with h5 terms (see, e.g.,
[21]), so that its coefficient cannot be directly determined from the on-shell 4-graviton
amplitude. The sigma-model approach implies [2226] that E8 does appear in S0 , i.e., that
(up to usual field redefinition ambiguities) the tree-level type II string R 4 term is indeed
proportional to J0 (2.3).
The structure of the kinematic factor (t8 + 12 8 )(t8 + 12 8 ) in the one-loop type IIA
4-point amplitude with transverse polarisations and momenta suggests [2729] that the
one-loop R 4 terms in D = 10 type IIA theory should be proportional to the oppositesign combination J0 (2.4) of the J1 and J2 terms, and this assumption passes some
compactification tests [28,29].
The presence of the P-odd one-loop term K (2.5) can be established [30] following
similar calculations of anomaly-related terms in the heterotic string [31,32]. Its coefficient
b2 can be fixed by considering compactification to 2 dimensions [30], and its value is in
agreement with the coefficient required by 5-brane anomaly cancellation [12] (see also
below).
The low-energy effective string action should be supersymmetric. 8 Remarkably, the
coefficients in (2.7) are indeed consistent with what is known about the structure of possible
R 4 superinvariants. First, the h4 term in t8 t8 R 4 is the bosonic part of the on-shell linearized
R
superspace invariant [34,35] (i.e., d 16 4 , = + + 4 R + written in terms
of N = 1 or N = 2B [36,37] on-shell superspace superfield ). If one first restricts
consideration to N = 1, D = 10 supersymmetry only, then one can use the classification of
possible bosonic R 4 parts of on-shell non-linear N = 1 superinvariants given in [3840].
A basis of the three independent N = 1 invariants [16,3840] can be chosen as J0 , X1 , X2
5 The more explicit form of this term is J = 24t [tr R 4 1 (tr R 2 )2 ], where R = (R ab ) and t tr R 4
mn
1
8
8
4
tr(16Rmn Rrn Rml Rrl + 8Rmn Rrn Rrl Rml 4Rmn Rmn Rrl Rrl 2Rmn Rrl Rmn Rrl ).
1
6 Note that the total coefficient of the t t RRRR term in S is thus
( (3)/gs2 + 2 /3). The
8 8
(2 )7 3211 0
2
2
2
2
0
4
relative combination (3)/gs + /3 is the same as in [19] (where g = (210 ) (2 ) = 16 7 gs2 ) and in [20],
but our overall normalization of this term is different (by factor 25 compared to [20]).
7 The Euler number in 8 dimensions is =

R 8
1
d x g E8 .
(4 )4 327

8 The string S-matrix is invariant under on-shell supersymmetry, so the leading-order corrections to effective
action evaluated on the supergravity equations of motion should be invariant under the standard supersymmetry
transformations. Since the D = 10 supersymmetry algebra does not close off shell, the full off-shell effective
action should be invariant under deformed supersymmetry transformations (see, e.g., [33]).

A.A. Tseytlin / Nuclear Physics B 584 (2000) 233250

J0 = t8 t8 RRRR +

1
10 10 RRRR,
4 2!

237

(2.9)

1
1
X2 = t8 tr R 2 tr R 2 10 B2 tr R 2 tr R 2 . (2.10)
X1 = t8 tr R 4 10 B2 tr R 4 ,
4
4
One may try to combine these N = 1 invariants to form potential N = 2A superinvariants.
Since t8 t8 R 4 = 24t8 [tr R 4 14 tr R 2 tr R 2 ], one may consider two candidate invariants which
contain combinations of J1 (2.3) or J2 (2.4) with 6K (2.5), i.e.,


1
I1 = 24 X1 X2 = J1 6K
4


2
1
(2.11)
= t8 t8 RRRR 610 B2 tr R 4 tr R 2 ,
4
or


1
I2 = J0 24 X1 X2 = J2 + 6K
4


2
1
1
10 10 RRRR + 610 B2 tr R 4 tr R 2 ,
(2.12)
=
4 2!
4
I1 + I2 = J0 .

(2.13)

The 1-loop term L1 (2.2) with b2 = 12b1 can thus be represented as a combination of
two different R 4 superinvariants [27,28], i.e., as
L1 = b1 J0 + b2 K = b1 (J1 J2 12K) = b1 (J0 + 2I1 ),

(2.14)

L1 = b1 (J0 2I2 ).

(2.15)

or as

The J0 -term should represent a separate N = 2 invariant. 9 A non-trivial question is which


of I1 and I2 can be actually extended to an invariant of N = 2A supersymmetry. 10
We would like to argue that it is I2 and not I1 that is the true N = 2A superinvariant.
Namely, it is the Euler term J2 = 14 E8 and not J1 = t8 t8 RRRR that is the superpartner
of the B2 -dependent term K (2.5). The form of the 1-loop correction L1 that admits a
superextension is then (2.15) and not (2.14). Then the tree + one-loop J0 terms in the type
IIA theory will be exactly the same as in the type IIB theory,


(3) 2
1
J0 ,
+

3
(2)4 3 213 0 gs2
9 In [3840] where non-linear extensions of N = 1 on-shell R 4 superinvariants were constructed the
transformation of the dilaton prefactor was ignored. As a result, one was not able to make a distinction between
J0 terms appearing at the tree and 1-loop levels. It is natural to conjecture that f ()J0 terms should combine
into an N = 2A superinvariant (invariant under deformed supersymmetry). For a discussion of supersymmetry
of e2 R + f ()J0 action in type IIB supergravity theory see [41].
10 Once the dilaton dependence of J terms is taken into account, one will not be able to freely switch between
0
I1 and I2 using (2.13).

238

A.A. Tseytlin / Nuclear Physics B 584 (2000) 233250

with the type IIA theory action containing in addition one extra one-loop contribution
(2.15) proportional to the superinvariant I2 .
Indeed, the weak-field expansions of both E8 and K start with 5-order terms, and the
corresponding 5-point amplitudes should be related by global supersymmetry. At the same
time, it is hard to imagine how the linearized on-shell W 4 N = 2 superspace invariant
corresponding to h4 term in t8 t8 RRRR may have a non-linear extension containing P-odd
term K.
A more serious argument against t8 t8 RRRR being a superpartner of 10 B2 [tr R 4
1
2 2
4 (tr R ) ] is the following. The D = 10 type II supergravity is known to contain a oneloop quadratic 2 UV divergence proportional to t8 t8 RRRR (this can be seen [42] by
taking the field theory limit, 0 0, = fixed, in the one-loop 4-graviton amplitude, cf.
(2.2)). At the same time, the ChernSimons type terms like 10 B2 R 4 cannot appear in the
UV divergent part of one-loop effective action. 11 This can be proved directly by using the
background field method: all one-loop UV divergent terms must be manifestly invariant
under 2-form gauge transformations and as well as diffeomorphisms. Since, e.g., a proper
time cutoff is expected to preserve supersymmetry at the level of one-loop UV divergences,
one concludes that J1 and K cannot be parts of the same superinvariant.
Similar argument can be given in the context of D = 11 theory. The t8 t8 RRRR term
appears [20,27,4345] as a cubic UV divergence (with a particular value of the UV cutoff
being fixed by duality considerations [45]), but 11 C3 R 4 term [12] can have only a finite
coefficient (with a non-perturbative dependence on 11 on dimensional grounds). Thus
(contrary to some previous suggestions in the literature, cf. [20,27,28,46]) these terms
cannot be related by supersymmetry, and the superpartner of the 11 C3 R 4 term should
be the D = 11 analog of J2 = 14 E8 (see Section 3).
Before turning to a detailed discussion of the D = 11 terms, let us add few comments
about the structure of the D = 10 effective action (2.1), (2.2). In addition to the R 4 terms
given explicitly in (2.3) and (2.4), it may contain also other Ricci tensor dependent terms
as well as terms depending on other fields (cf. [47]), for example, terms involving two and
more powers of H3 = dB2 (which were not included in the discussion of superinvariants
in [3840]). The well-known field redefinition ambiguity [18,48] allows one to change the
coefficients of on-shell terms. 12 In particular, the tree-level effective action (2.1) may
contain other Rmn dependent terms in addition to the full curvature contractions present in
J0 (see [2426,3840,49])


1 hkmn
8
hmnk
rsp q
rsp q
Rpmnq Rh R rsk + R
Rpqmn Rh R rsk + O(Rmn ). (2.16)
J0 = 3 2 R
2
The field redefinition ambiguity allows one to choose the action in a specific scheme
where only the Weyl tensor part of the curvature appears in J0 , i.e.,
11 Known examples of induced CS terms have finite coefficients and originate from IR effects (they appear from
1-loop contributions containing 12 massless poles, and thus can be rewritten in a manifestly gauge invariant but

nonlocal form).
12 For example, ignoring other fields, one may use R
4
mn = 0 to simplify the structure of R invariants as the
graviton legs in the string amplitudes they correspond to are on mass shell.

A.A. Tseytlin / Nuclear Physics B 584 (2000) 233250

239



1
J0 J0 = 3 28 C hmnk Cpmnq Ch rsp C q rsk + C hkmn Cpqmn Ch rsp C q rsk . (2.17)
2
That freedom of choice of a special scheme is crucial, in particular, in order to avoid
corrections to certain highly symmetric leading-order solutions, both in 10 and in 11
dimensions (see Section 3). For example, in type IIB theory the (scale of) AdS5 S 5
solution is not modified by the R 4 terms [50] only in the scheme [51] where they have
the form (2.17).

3. R 4 terms in 11 dimensions
Since the invariant I2 in (2.15) contains the P-odd CS type part K, its coefficient
cannot develop dilaton dependence without breaking B2 gauge invariance, i.e., its value
cannot be renormalized from its coupling-independent one-loop value [16]. Taking the
limit gs this term can then be lifted to a corresponding superinvariant in D = 11
theory. Assuming that the coefficient of the J0 invariant (2.3) does not receive higher than
one-loop perturbative string corrections, it can be also lifted [20,2729] to D = 11 (with
its tree-level part giving vanishing contribution). The resulting presence of the t8 t8 R 4 term
in the M-theory effective action is indeed in agreement with what follows directly from the
low-energy expansion of the 4-graviton amplitude in D = 11 supergravity [27,45].
In view of the above discussion, we conclude that the effective action of the D = 11
theory should contain two distinct R 4 superinvariants: (i) J0 with t8 t8 R 4 as its part, and
(ii) I2 which is a sum of the E8 and 11 C3 R 4 structures. With this separation, the coefficient
in front of the J0 term is then in agreement with the 4-graviton amplitude (with the Mtheory cutoff [45]), and the coefficient of the I2 term (its C3 R 4 part) is precisely the one
implied by the M5 brane anomaly cancellation condition [12].
Explicitly, the D = 11 action is then (cf. (2.1), (2.2))
S = S0 + S1 ,
S0 =

1
2
211

d 11 x

g R

d 11 x

g (J0 2I2 ).

Z
S1 = b1 T2


1
1
F42

C
F
F
11 3 4 4 ,
2 4!
6 3! (4!)2

(3.1)

(3.2)

Here Fmnkl = 4[m Cnkl] and the two R 4 superinvariants are (see (2.9), (2.8), (2.11))
1
J0 = t8 t8 RRRR + E8 ,
4

E8 =

1
11 11 RRRR,
3!




1
1
4
2 2
.
I2 = E8 + 211 C3 tr R tr R
4
4

(3.3)

(3.4)

240

A.A. Tseytlin / Nuclear Physics B 584 (2000) 233250

The constant b1 =

1
(2)4 32 213

is the same as in (2.7) and the 10-d and 11-d parameters are

related as follows (T1 and T2 are the string and the membrane tensions) 13
2
11
2
9
2
211
= (2)5 l11
,
10
=
,
2R11

R11 = gs 0 ,
l11 = (2gs )1/3 0 ,

T2 =

1
3
2l11

2
= (2)2/3 211

1/3

(3.5)

T1 =

1
= 2R11 T2 .
2 0

(3.6)

The subleading O(T2 ) term (3.2) in the effective action of 11-d theory may contain
also other O(Rmn ) and O(F4 ) terms. The invariant J0 (supplemented with appropriate
F4 dependent terms) may be considered as a non-linear extension of the linearized
R 4 superinvariant in on-shell D = 11 superspace [52]. The P-even part of the second
superinvariant starting with I2 (3.4) may also include extra O(F4 ) terms. Note that in the
exterior form notation I2 may be written as
I2 e0 e1 e10
2
= 11 e e e R R R R
3

+ 25 3!C3 tr(R R R R)


1
tr(R R) tr(R R) .
4

(3.7)

4. AdS7 S 4 solution and compactification on S 4


The D = 11 supergravity admits the well known AdS7 S 4 solution with F4 flux
N through S 4 [54]. Compactifying on S 4 , one may derive the corresponding d = 7
supergravity action, which gives the O(N 3 ) contribution [8] to the conformal anomaly
in the corresponding boundary conformal (2, 0) theory.
Let us consider how the presence of the R 4 terms in the 11-d effective action S1
(3.2) may influence the existence of the AdS7 S 4 solution and expansion near it.
Using the on-shell superspace description of 11-d supergravity and assuming that all local
higher-order corrections to the equations of motion can be written again in terms of the
basic on-shell supergravity superfield, it was argued in [55] that these corrections cannot
modify the maximally supersymmetric AdS7 S 4 solution. It should be possible to see
explicitly that adding the J0 term in (3.2) (supplemented with F4 dependent terms as
required by supersymmetry 14 and chosen in a special on-shell scheme analogous but
R

13 Note that B and C are canonically normalized, so that the 10-d invariant T
4
2
1 B2 tr(R) in (2.2) goes
R 3
1
1C
4
m
n
m
into the 11-d one T2 C3 tr(R) , where in the form notation B2 = 2 Bmn dx dx , C3 = 3!
mnk dx
R
1
n
k
ab
ab
m
n
4
4
dx dx , R = 2 Rmn dx dx . Thus S1 (3.2) contains T2 C3 tr(R) with the coefficient 4 3! 2 b1 =
1
which is the same as in [12].
(2 )4 326
14 In addition to F dependent terms (which may contain up to 8 powers of F ) there are also F dependent
4
4
4
terms which accompany t8 t8 R 4 part of J0 in the 4-point S-matrix [56] (as suggested by the analysis of tree-level
4-point scattering amplitudes in 11-d supergravity). These derivative terms vanish on AdS7 S 4 background.

A.A. Tseytlin / Nuclear Physics B 584 (2000) 233250

241

not equivalent 15 to (2.17) in 10-d theory) does not change the leading-order AdS7 S 4
solution. One may view J0 as originating from a restricted superspace integral of W 4 ,
where Wabcd (x, ) is the on-shell supergravity superfield [52], which has the structure
W = F4 + + ( R + F4 F4 + DF4 ) + ( stand for products
of gamma matrices). Then J0 (R + F4 F4 )4 and its first, second and third variation over
the metric evaluated on AdS7 S 4 + F4 -flux background (Rmn (F42 )mn , F4 = 0) will
vanish, essentially as in the case of AdS5 S 5 solution of type IIB theory corrected by J0
term [50] (taken in the form (2.17)). 16
The fact that the AdS7 S 4 solution (and, in particular, the radii of its factors) is not
modified by the J0 correction can be also represented as a consequence of the fact that upon
compactification of the 11-d theory on S 4 with F4 flux the J0 term (taken in the special
on-shell scheme) reduces to the Weyl tensor dependent C 4 term (2.17), now defined in
7 dimensions. 17 This term produces an O(N) correction [51] to the leading N 3 term [57]
in the entropy of (2, 0) theory describing multiple M5 branes. As in the AdS5 S 5 case in
type IIB theory, this C 4 term does not, however, modify the expression for the conformal
anomaly of the boundary conformal theory. 18
Let us now discuss the second invariant I2 (3.4) in (3.2). It is easy to see that its P-odd
part 11 C3 [tr R 4 14 (tr R 2 )2 ] does not modify the AdS7 S 4 solution. Upon reduction on
S 4 it leads to O(N) CS terms in d = 7 action [11]. As for the E8 part of I2 , we shall
assume that, as in the case of J0 , there exists an on-shell scheme in which this term,
supplemented with proper F4 -dependent terms, also does not modify the leading-order
AdS7 S 4 solution.
The main point is that upon compactification on S 4 the E8 term in (3.4) should produce
additional R 3 higher-derivative terms in the 7-d effective action which, while not changing
the vacuum solution, will give subleading O(N) corrections to the conformal anomaly of
the boundary CFT. 19
It is known that the C3 R 4 part of I2 (3.4) gives a subleading O(N) correction to the
R-symmetry anomaly of the (2, 0) theory [11,12]. Since the R-symmetry and conformal
anomalies should belong to the same 6-d supermultiplet, it is natural to expect that the
superpartner of the C3 R 4 term, i.e., the E8 term in I2 , should lead to an O(N) correction
15 Note that in contrast to AdS S 5 space with equal radii the 11-d space AdS S 4 space with radii 1 and 1
7
5
2
is not conformally flat.
16 The vanishing of the first variation is equivalent to the vanishing of the first correction to the 11-d supergravity
equations of motion abc DWabcd = 0 due to the supercovariant constancy of W [55]. The argument of [55]
should certainly apply to the first subleading correction to the 11-d supergravity equations of motion coming
from R 4 terms in the action.
17 The tree + one-loop J term in type IIB theory leads to the same C 4 term (2.17) in the 5-d effective action
0
obtained by compactifying the type IIB theory on S 5 with F5 flux.
18 It is important to stress for what follows that in the above discussion we treated J (3.3) as a whole, without
0
splitting it into t8 t8 R 4 and E8 parts. It is only that particular combination of R 4 terms that takes the irreducible
form (2.16) (cf. [49]), and thus should lead only to C 4 terms upon compactification to d = 7. At the same time,
E8 contains reducible curvature contractions like ((Rmnkl )2 )2 + R(Rmnkl )3 + and thus may, in principle,
lead to O(R n ), n < 4, terms upon compactification to d = 7.
19 These terms will give also another O(N ) correction to the entropy of (2, 0) theory, in addition to the one
coming from the J0 term (2.17) found in [51].

242

A.A. Tseytlin / Nuclear Physics B 584 (2000) 233250

to the conformal anomaly of the boundary 6-d theory. This is what we are going to suggest
below.
Since we do not know the F4 (and Rmn ) dependent terms which supplement E8 to a
superinvariant, to determine the terms in the 7-d action that originate from the E8 part of
the invariant I2 in (3.2) we shall use the following heuristic strategy. We shall start with
E8 and compute it in the case when the 11-d space is a direct product, M 11 = M 7 M 4 .
It is easy to see that





(4.1)
E8 M 7 M 4 = 4E2 M 4 E6 M 7 + 12E4 M 4 E4 M 7 ,
where, as in (2.8),

E2n M d

1
d d R n ,
(d 2n)!

d > 2n,

(4.2)

and E2n (M d ) = 0 for d < 2n. In the case when M 4 is a 4-sphere of radius L (RS 4 = 12/L2 )
and M 7 has curvature R we get
 3 25
 32 27

E4 M 7
E8 M 7 S 4 = 2 E6 M 7 +
L
L4
3 26
3 25
(4.3)
= 2 7 7 RRR + 4 7 7 RR.
L
L
A remarkable property of the E8 invariant is that it does not produce a correction to the
cosmological or Einstein term in the 7-d action.
Next, we shall assume that when the same reduction is repeated for the analog of E8
term in a special on-shell scheme (i.e., for E8 supplemented by F4 and Rmn dependent
terms so that it does not produce a modification of the leading-order AdS7 S 4 solution)
then the resulting terms in the 7-d action will be the same as in (4.3) but with the curvature
tensor R of M 7 replaced by its Weyl tensor C part.
In what follows we shall consider only on the E6 (M 7 ) C 3 + term in (4.3) coming
from E8 . The reason is that we shall compute the corresponding contribution to the scale
anomaly of the boundary theory only modulo Rmn -dependent terms, but it is easy to see
that a potential C 2 term in the 7-d action (coming from E4 in (4.3)) can lead only to terms
in the conformal anomaly which vanish when the 6-d boundary space is Ricci flat.
Choosing the normalization in which the radii of AdS7 and S 4 are 1 and L = 1/2 so that
2
2
Vol(S 4 ) = 83 L4 = 6 , and assuming that the value of the quantized F4 flux is N , we get
(see (3.5), (3.6) and [51,57])
N3
2N 3
1
Vol(S 4 )
N3
2N
1
.
=
=
,
=
=
,
T2 =
2
8
5
9
5
2
2
3

3
211 2 L
27
211
R
The relevant [N 3 (R 2) + NC 3 ] terms in the 7-d action are then
Z
Z

N3
N
S (7) = 3 d 7 x g (R + 30) + 2 11 3 d 7 x g E 6 + ,
3
3 2
where the explicit form of the E 6 C 3 correction term is (cf. (2.8))

(4.4)

(4.5)

A.A. Tseytlin / Nuclear Physics B 584 (2000) 233250

243

E 6 = (E6 )Rmn =0 = 7 7 CCC


n

n1
. . . m66 ] C m1 m2n1 n2 C m3 m4n3 n4 C m5 m6n5 n6 = 32(2I1 + I2 ),
= 6![m
1

(4.6)

where I1 and I2 are defined as


I1 = Camnb C mpqn Cp ab q ,

I2 = Cab mn Cmn pq Cpq ab .

(4.7)

As follows from (4.3) the numerical coefficient is


= 1,

(4.8)

but we shall keep it arbitrary, given the uncertainties in the above derivation of the
correction term in (4.5) (for example, the presence of (F4 )2 (Rmnkl )3 terms in I2 would
shift the value of ).

5. Conformal anomaly of (2, 0) theory


Let us now determine the contribution of the C 3 correction term in the 7-d action (4.5)
which originated from the E8 part of the I2 superinvariant in the 11-d action (3.2) to
the conformal anomaly of the d = 6 boundary conformal theory. We shall follow the
same method as used in [8] in computing the leading N 3 term in the anomaly. 20 We
shall compute only the O(N) contribution to the scale anomaly (which is the same as
integrated conformal anomaly, assuming topology of 6-space is trivial) and ignore terms
which depend on Rmn , i.e., concentrate only on the Weyl-invariant non-total derivative C 3
terms (type B part) in the 6-d conformal anomaly.
To obtain the conformal anomaly one is to solve the 7-d equations for the metric (as in
(4.4) we set the radius of AdS7 to be equal to 1)
1
(5.1)
ds 2 = 2 d 2 + 1 gij (x, ) dx i dx j ,
4
evaluate the action on the solution g = g0 (x) + g2 (x) + , and compute its variation
under the Weyl rescaling of the 6-d boundary metric. The anomaly is essentially
determined by the coefficient of the logarithmic divergence produced by the integral over
[8]. In the present case of (4.5) we find (using (4.5), (4.6) and RAdS7 = 42)
 3 Z

Z
Z
N
N
d p
1 d
(7)
6
6
g(x, ) 2 6 3
g (2I1 + I2 ) + .
S = d x
3 3
4
3 2 2

(5.2)
Since [8]
Z



d p
g(x, ) = g0 a0 (x) 3 + a6 (x) ln  + ,
6
4

(5.3)


20 Similar computation of subleading corrections to conformal anomaly of 4-d boundary conformal field theories
(with N < 4 supersymmetry) coming from R 2 curvature terms in 5-d effective action were discussed in [46,58
61].

244

A.A. Tseytlin / Nuclear Physics B 584 (2000) 233250

the anomaly is given by the sum of the O(N 3 ) and O(N) terms 21
3

N
A(2,0) = AN
(2,0) + A(2,0) =

N3
N
2a6 + 2 6 3 (2I1 + I2 + ).
3 3
3 2

(5.4)

Here a6 and I1 , I2 are evaluated for the boundary metric g0 , and dots stand for O(N)
Rmn -dependent and total derivative terms we are ignoring.
3
The result of [8] for the leading-order contribution AN
(2,0) written as a sum of the type A
(Euler), type B (Weyl invariant) and scheme-dependent (covariant total derivative) terms
[62,63] is
3

AN
(2,0) =



4N 3
E6 + 8(12I1 + 3I2 I3 ) + O i J i ,
3
2
5
(4) 3 2

where E6 = 6 6 RRR. The invariants I1 , I2 (4.7) and I3



I3 = Cmnbc 2 C mnbc + O(Rmn ) + O i J i ,

(5.5)

(5.6)

which form the basis of 3 Weyl invariants are the same as used in [9]. They are related to the
1
invariants used in [8,62] as follows: E(6) , I1 , I2 and I3 in [8] are equal to 33 2
11 E6 , I1 , I2
8
1
0
0
i
and 5I3 , I3 = I3 3 (2I1 + I2 ) 12 E6 + O(i J ), in terms of the invariants E6 , I1 , I2
and I3 used in [9] and here. 22
We use this opportunity to point out that the curvature invariant I3 = 5I30 as defined in
[8,62] is not, in fact, covariant under Weyl transformations, contrary to what was assumed
in [8] (this can be easily checked by computing it for the metric of a sphere S 6 : one
finds that while I1 (S 6 ) = I2 (S 6 ) = 0, I30 (S 6 ) 6= 0). The proper third Weyl invariant of type
C 2 C (5.6) was given in [6466] and is equivalent to the Weyl invariant I3 used in [9] and
here. Since I3 of [8] or I30 is a mixture of the true Weyl invariants I1 , I2 , I3 with E6 , the
separation of the leading N 3 Weyl anomaly of the (2, 0) theory [8] into type A and type B
parts was not presented correctly in [8]. The correct separation was given in [9] and is used
here. 23
Note that modulo terms that vanish for Rmn = 0 and total derivative terms, one has the
following relations (cf. (4.6))

I3 = 4I1 I2 + O(Rmn ) + O i J i , (5.7)
E6 = 32(2I1 + I2 ) + O(Rmn ),
3

so that AN
(2,0) vanishes for Rmn = 0, as it should [8].
Eq. (5.5) is to be compared with the expression for the conformal anomaly for the free
(2, 0) tensor multiplet found in [9]:
21 To obtain the O(N ) contribution we evaluate the C 3 term in the 7-d action on the leading-order solution for
2
action in d = 5), separate the C 3 part
the metric (5.1) (see [58] for a similar computation in the case of the Rmnkl
depending on the 6-d metric g0 , and omit other parts that depend on the Ricci tensor of g0 .
a
22 Our curvature tensor R a
bmn = m bn has the opposite sign to that of [8]. Note also that [9] was assuming

Euclidean signature where E6 is defined as 6 6 RRR.


23 Note that when R
0
mn = 0 the two invariants I3 and I3 coincide, up to a covariant total derivative term.
In fact, a separation of the conformal anomaly into type A and type B parts becomes ambiguous on a Ricci flat
background.

A.A. Tseytlin / Nuclear Physics B 584 (2000) 233250

Atens. =




7
1
i
E
+
8(12I
+
3I

I
)
+
O

J
.
6
1
2
3
i
(4)3 32 25 4

245

(5.8)

As was concluded in [9], the Weyl-invariant (type B) parts of the leading (2, 0) theory
anomaly (5.5) and the tensor multiplet anomaly (5.8) have exactly the same form, up to the
overall factor 4N 3 in (5.5).
Since we have found the O(N) correction to the anomaly of the (2, 0) theory in (5.4)
only modulo Rmn -dependent and total derivative terms, we are able to compare only type
B anomalies, or scale anomalies (assuming that the d = 6 space has trivial topology, so
that we can ignore the integral of the Euler term E6 )
Z
Z

Atens. = d 6 x g0 Atens. .
A(2,0) = d 6 x g0 A(2,0),
Using (5.7) to express I3 in terms of I1 and I2 , we find from (5.5), (5.4) and (5.8)
Z

4N 3
N3
A(2,0) =
d 6 x g0 (2I1 + I2 ),
(4)3 32
AN
(2,0) =

N
(4)3 32

1
Atens. =
(4)3 32

d 6x

g0 (2I1 + I2 ),

Z
d 6x

g0 (2I1 + I2 ).

The total scale anomaly of the (2, 0) theory following from (4.5), (5.4) is then
Z

4N 3 N
3
N
+
A
=

d 6 x g0 (2I1 + I2 ).
A(2,0) = AN
(2,0)
(2,0)
3
2
(4) 3

(5.9)

(5.10)

(5.11)

(5.12)

Equivalently,
A(2,0) =

4(N 3 N)
(4)3 33

Z
d 6x

g0 (2I1 + I2 ) + (4 )NAtens. .

(5.13)

Thus if the true value of is 3 instead of the naive value 1 (4.8) which follows directly
from reduction of E8 (4.3), ignoring possible F4 -dependent (F42 R 3 ) terms in the 11-d
superinvariant I2 , then A(2,0) reproduces the scale anomaly (5.11) of a single (2, 0) tensor
multiplet. This N = 1 relation should be expected, given that a similar correspondence is
true for the R-symmetry anomalies [11] (see below). Though we are unable to show that
= 3 does follow from the d = 7 reduction of the 11-d superinvariant I2 containing P-odd
C3 R 4 term, we find it remarkable that the required value of differs from the naive value
1 simply by factor of 3. 24 , 25
24 In the original version of the present paper we mistakenly used the basis of type B invariants including I of
3
[8] instead of the correct invariant of [9] and as a result got the O(N ) term with extra coefficient 3, concluding
that = 1 gives already the desired coefficient 4N 3 3N in (5.12).
25 Note that if we were comparing the full local conformal anomalies evaluated for R
mn = 0 then, since the
N 3 contribution (5.5) vanishes in this case, we would need = 3/4 in order to reproduce the non-zero Rmn = 0
value of the tensor multiplet anomaly (5.8) by the N = 1 limit of the O(N ) term in (5.4).

246

A.A. Tseytlin / Nuclear Physics B 584 (2000) 233250

Making a natural conjecture that the same relation Atens. = (A(2,0) )N=1 should be true
between the full expressions for the conformal anomalies of the (2, 0) theory and tensor
multiplet, one can make a prediction about the complete structure of the O(N) term in the
(2, 0) theory anomaly A(2,0) (5.4) (cf. (5.5), (5.8)) 26
1
32 25





9
3
3
i
4N N E6 + 4N 3N 8(12I1 + 3I2 I3 ) + O i J , (5.14)
4

A(2,0) =

(4)3

or, equivalently,
A(2,0) =


N3 N 
E6 + 8(12I1 + 3I2 I3 ) + O i J i + NAtens. . (5.15)
3
2
3
(4) 3 2

Using (5.7), we can rewrite (5.14) also as


A(2,0) =



N
E6 + O(Rmn ) + O i J i ,
(4)3 3 27

(5.16)

in agreement with the fact that for Rmn = 0 the conformal anomaly of the tensor multiplet
becomes [9,43,44]


1
E6 + O i J i .
Atens. =
3
7
(4) 3 2
It is useful to compare the above expressions (5.12), (5.14) with the previously known
results for the R-symmetry anomalies of the interacting (2, 0) theory and free tensor
multiplet theory. The 1-loop effective action for a free 6-d tensor multiplet in a
background of 6-d Lorentz curvature R and SO(5) R-symmetry gauge field F has local
SO(6) and SO(5) anomalies. They satisfy the descent relations d( ) = I7 , I8 = dI7 ,
with the 8-form anomaly polynomial I8 being [13,14]


2
1
1
tens.
p2 (F ) p2 (R) + p1 (F ) p1 (R) ,
(5.17)
I8 (F, R) =
4
3 24
with (here F 2 F F , etc.)
p1 (F ) =

1 2
tr F ,
2

p2 (F ) =



1
1
tr F 4 tr F 2 tr F 2 ,
4
2

i
F.
(5.18)
F =
2
The corresponding anomalies of the interacting (2, 0) theory describing multiple M5 branes
derived (by assuming that the total M5-brane anomaly + inflow anomaly should cancel)
from the 11-d supergravity action (3.1) with the R 4 correction term (3.2) is [11]
26 The shift of the coefficient of the E term in the conformal anomaly seems to imply a contradiction between
6
our assumption that the R 4 terms in the 11-d action (3.2) do not change the scale of AdS7 S 4 solution (i.e., that
the value of the 7-d action (4.5) evaluated on the AdS7 solution is not changed), and the claim of [59] that the
coefficient of the type A (Euler) term in the anomaly of a generic effective theory is determined only by the value
of the action on the AdS solution.

A.A. Tseytlin / Nuclear Physics B 584 (2000) 233250


(2,0)

I8

(F, R) =

1
3 24

247



2
1 
2N 3 N p2 (F ) Np2 (R) + N p1 (F ) p1 (R) . (5.19)
4

Here the O(N 3 ) term comes [67] from the CS term in supergravity action (3.1) and the
O(N) term [12,14] from the P-odd C3 R 4 part of the superinvariant I2 (3.2), (3.4).
Equivalently,

1
(5.20)
N 3 N p2 (F ) + NI8tens. (F, R).
3
32
Thus for N = 1 the anomaly of the (2, 0) theory is the same as the anomaly of a single
tensor multiplet. This is the same type of a relation we have established above (cf. (5.13))
for the scale anomalies, with the crucial O(N) contribution coming from the P-even
E8 part of the superinvariant I2 (3.4). This is obviously consistent with the fact that Rsymmetry and conformal anomalies should be parts of the same 6-d supermultiplet.
(2,0)

I8

(F, R) =

Acknowledgements
We are grateful to S. Frolov for a collaboration at an initial stage and many useful
discussions. We would like also to acknowledge J. Harvey, P. Howe, K. Intriligator,
R. Metsaev, Yu. Obukhov, H. Osborn, T. Petkou and M. Shifman for helpful discussions
and comments. This work was supported in part by the DOE grant DOE/ER/01545,
EC TMR grant ERBFMRX-CT96-0045, INTAS grant No. 99-0590, NATO grant PST.CLG
974965 and PPARC SPG grant PPA/G/S/1998/00613.
References
[1] J. Maldacena, The large N limit of superconformal field theories and supergravity, Adv. Theor.
Math. Phys. 2 (1998) 231, hep-th/9711200.
[2] H. Liu, A.A. Tseytlin, D = 4 super YangMills, D = 5 gauged supergravity, and D = 4
conformal supergravity, Nucl. Phys. B 533 (1998) 88, hep-th/9804083.
[3] G. Arutyunov, S. Frolov, Three-point Green function of the stress-energy tensor in the AdS/CFT
correspondence, Phys. Rev. D 60 (1999) 026004, hep-th/9901121.
[4] F. Bastianelli, S. Frolov, A.A. Tseytlin, Three-point correlators of stress tensors in maximallysupersymmetric conformal theories in d = 3 and d = 6, Nucl. Phys. B 508 (1997) 245, hepth/9911135.
[5] S.S. Gubser, I.R. Klebanov, A.A. Tseytlin, String theory and classical absorption by threebranes, Nucl. Phys. B 499 (1997) 217, hep-th/9703040.
[6] S.S. Gubser, I.R. Klebanov, Absorption by branes and Schwinger terms in the world volume
theory, Phys. Lett. B 413 (1997) 41, hep-th/9708005.
[7] R. Manvelian, A.C. Petkou, A note on R-currents and trace anomalies in the (2, 0) tensor
multiplet in d = 6 AdS/CFT correspondence, hep-th/0003017.
[8] M. Henningson, K. Skenderis, The holographic Weyl anomaly, JHEP 07 (1998) 023, hepth/9806087.
[9] F. Bastianelli, S. Frolov, A.A. Tseytlin, Conformal anomaly of (2, 0) tensor multiplet in six
dimensions and AdS/CFT correspondence, JHEP 02 (2000) 013, hep-th/0001041.
[10] I.R. Klebanov, World-volume approach to absorption by non-dilatonic branes, Nucl. Phys.
B 496 (1997) 231, hep-th/9702076.

248

A.A. Tseytlin / Nuclear Physics B 584 (2000) 233250

[11] J.A. Harvey, R. Minasian, G. Moore, Non-abelian tensor-multiplet anomalies, JHEP 09 (1998)
004, hep-th/9808060.
[12] M.J. Duff, J.T. Liu, R. Minasian, Eleven-dimensional origin of string/string duality: A one-loop
test, Nucl. Phys. B 452 (1995) 261, hep-th/9506126.
[13] L. Alvarez-Gaume, E. Witten, Gravitational anomalies, Nucl. Phys. B 234 (1984) 269.
[14] E. Witten, Five-brane effective action in M-theory, J. Geom. Phys. 22 (1997) 103, hepth/9610234.
[15] P.S. Howe, G. Sierra, P.K. Townsend, Supersymmetry in six dimensions, Nucl. Phys. B 221
(1983) 331.
[16] A.A. Tseytlin, Heterotictype I superstring duality and low-energy effective actions, Nucl. Phys.
B 467 (1996) 383, hep-th/9512081.
[17] M.B. Green, J.H. Schwarz, Supersymmetrical dual string theory. 2. Vertices and trees, Nucl.
Phys. B 198 (1982) 252.
[18] D.J. Gross, E. Witten, Superstring modifications of Einsteins equations, Nucl. Phys. B 277
(1986) 1.
[19] N. Sakai, Y. Tanii, One loop amplitudes and effective action in superstring theories, Nucl. Phys.
B 287 (1987) 457.
[20] M.B. Green, P. Vanhove, D-instantons, strings and M-theory, Phys. Lett. B 408 (1997) 122,
hep-th/9704145.
[21] B. Zumino, Gravity theories in more than four dimensions, Phys. Rept. 137 (1986) 109.
[22] M.T. Grisaru, A.E. van de Ven, D. Zanon, Four loop beta function for the N = 1 and N = 2
supersymmetric nonlinear sigma model in two dimensions, Phys. Lett. B 173 (1986) 423.
[23] M.T. Grisaru, A.E. van de Ven, D. Zanon, Four-loop divergences for the N = 1 supersymmetric
nonlinear sigma model in two dimensions, Nucl. Phys. B 277 (1986) 409.
[24] M.T. Grisaru, D. Zanon, Sigma model superstring corrections to the EinsteinHilbert action,
Phys. Lett. B 177 (1986) 347.
[25] M.D. Freeman, C.N. Pope, M.F. Sohnius, K.S. Stelle, Higher order sigma model counterterms
and the effective action for superstrings, Phys. Lett. B 178 (1986) 199.
[26] Q. Park, D. Zanon, More on sigma model beta functions and low-energy effective actions, Phys.
Rev. D 35 (1987) 4038.
[27] J.G. Russo, A.A. Tseytlin, One-loop four graviton amplitude in eleven-dimensional supergravity, Nucl. Phys. B 578 (2000) 139, hep-th/9707134.
[28] E. Kiritsis, B. Pioline, On R 4 threshold corrections in type IIB string theory and (p, q) string
instantons, Nucl. Phys. B 508 (1997) 509, hep-th/9707018.
[29] I. Antoniadis, S. Ferrara, R. Minasian, K.S. Narain, R 4 couplings in M- and type II theories on
CalabiYau spaces, Nucl. Phys. B 507 (1997) 571, hep-th/9707013.
[30] C. Vafa, E. Witten, A one-loop test of string duality, Nucl. Phys. B 447 (1995) 261, hepth/9505053.
[31] W. Lerche, B.E. Nilsson, A.N. Schellekens, Heterotic string loop calculation of the anomaly
cancelling term, Nucl. Phys. B 289 (1987) 609.
[32] W. Lerche, B.E. Nilsson, A.N. Schellekens, N.P. Warner, Anomaly cancelling terms from the
elliptic genus, Nucl. Phys. B 299 (1988) 91.
[33] E.A. Bergshoeff, M. de Roo, The quartic effective action of the heterotic string and
supersymmetry, Nucl. Phys. B 328 (1989) 439.
[34] R. Kallosh, Strings and superspace, Phys. Scripta T 15 (1987) 118.
[35] B.E. Nilsson, A.K. Tollsten, Supersymmetrization of (3)R 4 in superstring theories, Phys. Lett.
B 181 (1986) 63.
[36] P.S. Howe, P.C. West, The complete N = 2, D = 10 supergravity, Nucl. Phys. B 238 (1984)
181.
[37] M.B. Green, Interconnections between type II superstrings, M theory and N = 4 YangMills,
hep-th/9903124.

A.A. Tseytlin / Nuclear Physics B 584 (2000) 233250

249

[38] M. de Roo, H. Suelmann, A. Wiedemann, Supersymmetric R 4 actions in ten dimensions, Phys.


Lett. B 280 (1992) 39.
[39] M. de Roo, H. Suelmann, A. Wiedemann, The supersymmetric effective action of the heterotic
string in ten dimensions, Nucl. Phys. B 405 (1993) 326, hep-th/9210099.
[40] J.H. Suelmann, Supersymmetry and string effective actions, PhD Thesis, Groningen, 1994.
[41] M.B. Green, S. Sethi, Supersymmetry constraints on type IIB supergravity, Phys. Rev. D 59
(1999) 046006, hep-th/9808061.
[42] R.R. Metsaev, A.A. Tseytlin, One-loop corrections to string theory effective actions, Nucl. Phys.
B 298 (1988) 109.
[43] E.S. Fradkin, A.A. Tseytlin, Quantum properties of higher-dimensional and dimensionally
reduced supersymmetric theories, Nucl. Phys. B 227 (1983) 252.
[44] E.S. Fradkin, A.A. Tseytlin, Present state of quantum supergravity, in: M.A. Markov et al.
(Eds.), Proc. of Third Seminar on Quantum Gravity, October 2325, 1984, Moscow, World
Scientific, 1985, p. 303.
[45] M.B. Green, M. Gutperle, P. Vanhove, One-loop in eleven dimensions, Phys. Lett. B 409 (1997)
177, hep-th/9706175.
[46] D. Anselmi, A. Kehagias, Subleading corrections and central charges in the AdS/CFT
correspondence, Phys. Lett. B 455 (1999) 155, hep-th/9812092.
[47] A. Kehagias, H. Partouche, The exact quartic effective action for the type IIB superstring, Phys.
Lett. B 422 (1998) 109, hep-th/9710023.
[48] A.A. Tseytlin, Ambiguity in the effective action in string theories, Phys. Lett. B 176 (1986) 92.
[49] R. Myers, Superstring gravity and black holes, Nucl. Phys. B 289 (1987) 701.
[50] T. Banks, M.B. Green, Non-perturbative effects in AdS5 S 5 string theory and d = 4 SUSY
YangMills, JHEP 9805 (1998) 002, hep-th/9804170.
[51] S.S. Gubser, I.R. Klebanov, A.A. Tseytlin, Coupling constant dependence in the thermodynamics of N = 4 supersymmetric YangMills theory, Nucl. Phys. B 534 (1998) 202, hepth/9805156.
[52] E. Cremmer, S. Ferrara, Formulation of eleven-dimensional supergravity in superspace, Phys.
Lett. B 91 (1980) 61.
[53] L. Brink, P. Howe, Eleven-dimensional supergravity on the mass-shell in superspace, Phys. Lett.
B 91 (1980) 384.
[54] K. Pilch, P. van Nieuwenhuizen, P.K. Townsend, Compactification of D = 11 supergravity on
S 4 (or 11 = 7 + 4, too), Nucl. Phys. B 242 (1984) 377.
[55] R. Kallosh, A. Rajaraman, Vacua of M-theory and string theory, Phys. Rev. D 58 (1998) 125003,
hep-th/9805041.
[56] S. Deser, D. Seminara, Tree amplitudes and two-loop counterterms in D = 11 supergravity,
hep-th/0002241.
[57] I.R. Klebanov, A.A. Tseytlin, Entropy of near-extremal black p-branes, Nucl. Phys. B 475
(1996) 164, hep-th/9604089.
[58] M. Blau, K.S. Narain, E. Gava, On subleading contributions to the AdS/CFT trace anomaly,
JHEP 9909 (1999) 018, hep-th/9904179.
[59] C. Imbimbo, A. Schwimmer, S. Theisen, S. Yankelowicz, Diffeomorphisms and holographic
anomalies, Class. Quant. Grav. 17 (2000) 1129, hep-th/9910267.
[60] S. Nojiri, S.D. Odintsov, Weyl anomaly from Weyl gravity, hep-th/9910113.
[61] S. Nojiri, S.D. Odintsov, On the conformal anomaly from higher derivative gravity in AdS/CFT
correspondence, hep-th/9903033.
[62] L. Bonora, P. Pasti, M. Bregola, Weyl cocycles, Class. Quant. Grav. 3 (1986) 635.
[63] S. Deser, A. Schwimmer, Geometric classification of conformal anomalies in arbitrary
dimensions, Phys. Lett. B 309 (1993) 279, hep-th/9302047.
[64] C. Fefferman, C. Graham, Conformal invariants, Asterisque, hors serie 95 (1985).
[65] T. Parker, S. Rosenberg, Invariants of conformal Laplacians, J. Diff. Geom. 25 (1987) 199.

250

A.A. Tseytlin / Nuclear Physics B 584 (2000) 233250

[66] J. Erdmenger, Conformally covariant differential operators: properties and applications, Class.
Quant. Grav. 14 (1997) 2061, hep-th/9704108.
[67] D. Freed, J.A. Harvey, R. Minasian, G. Moore, Gravitational anomaly cancellation for M-theory
fivebranes, Adv. Theor. Math. Phys. 2 (1998) 601, hep-th/9803205.
[68] E.S. Fradkin, A.A. Tseytlin, Conformal anomaly in weyl theory and anomaly free superconformal theories, Phys. Lett. B 134 (1984) 187.
[69] D. Anselmi, D.Z. Freedman, M.T. Grisaru, A.A. Johansen, Nonperturbative formulas for central
functions of supersymmetric gauge theories, Nucl. Phys. B 526 (1998) 543, hep-th/9708042.

Nuclear Physics B 584 (2000) 251283


www.elsevier.nl/locate/npe

The spacetime life of a non-BPS D-particle


Eduardo Eyras a, , Sudhakar Panda b,c,1
a Department of Applied Mathematics and Theoretical Physics, University of Cambridge,

Wilberforce Road, CB3 0WA Cambridge, UK


b The Abdus Salam International Centre for Theoretical Physics Strada Costiera 11,

I-34014, Trieste, Italy


c Mehta Research Institute of Mathematics and Mathematical Physics Chhatnag Road,

Jhoosi Allahabad 211019, India


Received 30 March 2000; accepted 13 June 2000

Abstract
We investigate the classical geometry generated by a stable non-BPS D-particle. We consider the
boundary state of a stable non-BPS D-particle in the covariant formalism in the type IIB theory
orbifolded by (1)FL I4 . We calculate the scattering amplitude between two D-particles in the noncompact and compact orbifold and analyse the short and long distance behaviour. At short distances
we find no force at order v 2 for any radius, and at the critical radius we find a BPS-like behaviour up
to v 4 corrections for long and short distances. Projecting the boundary state on the massless states of
the orbifold closed string spectrum we obtain the large distance behaviour of the classical solution
describing this non-BPS D-particle in the non-compact and compact cases. By using the non-BPS
D-particle as a probe of the background geometry of another non-BPS D-particle, we recover the noforce condition at the critical radius and the v 2 behaviour of the probe. Moreover, assuming that the
no-force persists for the complete geometry we derive part of the classical solution for the non-BPS
D-particle. 2000 Elsevier Science B.V. All rights reserved.
PACS: 11.25.-w; 11.25.Sq
Keywords: Branes; Non-BPS

1. Introduction
BPS D-branes enjoy a double life. On the one hand as a conformal field theory described
by open strings with Dirichlet boundary conditions [13], and on the other hand as
classical solitons of supergravity. 2 For a single D-brane the regimes of validity of these
two descriptions are complementary. The conformal field theory description is valid at
Corresponding author. E-mail: e.eyras@damtp.cam.ac.uk
1 panda@mri.ernet.in
2 See [46] and references therein.

0550-3213/00/$ see front matter 2000 Elsevier Science B.V. All rights reserved.
PII: S 0 5 5 0 - 3 2 1 3 ( 0 0 ) 0 0 3 7 6 - X

252

E. Eyras, S. Panda / Nuclear Physics B 584 (2000) 251283

weak coupling; whereas the supergravity solution corresponds to a strongly interacting


gravitational system, which corresponds to strong coupling. On the other hand, for a large
number of branes N the system can be well described by a classical solution at weak
coupling [7]. The two main properties of the BPS D-branes that assure this consistent
dual behaviour are the fact that BPS-branes preserve a fraction of supersymmetry, hence
they are stable under variations of the string coupling; and the fact that one can consider
a superposition of a large number of D-branes in such a way that they still preserve
a fraction of supersymmetry. These properties make possible, for instance, the entropy
calculation of black-holes using D-branes [8,9]. On the other hand, D-branes can be studied
in more general situations, for which spacetime supersymmetry is not preserved. This is
the case of the non-BPS D-branes [1019]. These branes have an exact conformal field
theory description, whose consistency conditions do not rely on spacetime supersymmetry
[1114,16]. Having such a precise description for the non-BPS D-branes in the conformal
field theory side, it is natural to wonder whether the non-BPS D-branes also enjoy a
double life and have a description in terms of a classical solution of a certain effective
(super)gravity theory.
It has been recently suggested in [20] that the unstable non-BPS branes of type II
theories might have a gravitational counterpart described by a gravitational sphaleron. 3
These are unstable solutions with finite energy which interpolate between two (possibly
distinguishable) vacuum configurations. However, these solutions are unstable and
probably do not remain valid classically. On the other hand, we expect that consistent
classical solutions will be related to non-BPS branes which have some stability properties
similar to the BPS D-branes.
In order to find the appropriate conditions for constructing a classical solution for a nonBPS D-brane we must find out which properties of the D-branes are such that (1) they
assure the validity of a classical supergravity description and (2) they are not necessarily
related to fractional supersymmetry. These properties certainly comprise stability, which
ensures that the state survives at strong coupling. Moreover, stability can be based on
grounds different from supersymmetry, for instance, being the lightest object carrying a
certain quantum number [10,11]. Another key property is the fact that we can superpose
an arbitrary number of parallel D-branes, which is the same as having a no-force condition
at all distances [6,22,23]. This is a consequence of the BPS property, which can exist
independently of supersymmetry. Therefore, it seems natural that in order to find a classical
solution for a non-BPS D-brane, this brane should be stable and enjoy a no-force property.
Stable non-BPS D-branes have been found in different theories [1116,2431]. However,
only non-BPS D-branes in a certain orbifold of type II theories are known to enjoy both
stability and the no-force property [33]. There it was shown that at a particular critical
radius of the compact orbifold the non-BPS branes develop a BoseFermi degeneracy
at one loop. The critical radius is in fact the value beyond which the non-BPS D-brane
becomes unstable and can decay into a pair of BPS D-branes.
3 See [21] and references therein.

E. Eyras, S. Panda / Nuclear Physics B 584 (2000) 251283

253

In this article we study the particular case of a stable non-BPS D-particle in type IIB
string theory 4 orbifolded by (1)FL I4 . This D-particle is a truncated D-brane [30], is
charged electrically under a twisted RR 1-form field and it is the strong-weak coupling
dual of a non-BPS state in the orientifold I4 of type IIB, charged under the U (1)
field of the D5O5 system [11]. The coupling of the non-BPS D-particle to a twisted
RR vector field is the origin of the stability of this non-BPS D-particle [12]. Moreover,
at a critical radius of the compact orbifold this D-particle meets all the requirements
suggesting the existence of a classical solution. We use the technique of the boundary
state in the covariant formalism 5 [3842] to describe the non-BPS D-particle. We analyse
the interaction potential between two non-BPS D-particles in relative motion. Remarkably,
we find no force at order v 2 as for BPS D-branes. At the critical radius, the static force is
moreover vanishing, hence the non-BPS D-particle presents a BPS-like behaviour, up to
v 4 corrections. Unlike for BPS branes, we find no matching between the v4 terms in the
open and closed string description.
Using the boundary state for BPS D-branes, the long distance behaviour of the classical
massless fields generated by the D-brane was computed in 6 [44,45], and it was shown that
the asymptotic behaviour of the corresponding classical solution is precisely recovered.
A BPS Dp-brane is described by a boundary state with two parts, the NSNS part and
the RR part. They generate the asymptotic behaviour of NSNS massless fields (metric
and dilaton), and of the RR massless fields (RR (p + 1)-form potential), respectively.
In this paper we implement the same technique to obtain the long distance behaviour of
the non-BPS D-particle geometry. This is given by the asymptotic form of a metric and
a dilaton propagating in the bulk, and a twisted RR 1-form potential propagating in the
orbifold fixed plane.
One difficulty about recovering the full solution for the non-BPS D-particle from its
asymptotic form is that, in principle, there might be many different possible geometries
with the same asymptotic structure. In order to restrict these geometries we will assume
that the no force condition also takes place when one consider the complete geometry at
the critical radius, as it happens for BPS branes [23]. That is, we take into account the
extra pieces that one would need to add to the asymptotic behaviour in order to recover
the full form of metric. On the other hand, although the BoseFermi degeneracy occurs
at any distance between the non-BPS D-branes [33], at short distances, open strings loops
might spoil this property. In fact, it has been recently proposed in [46] that even at 1-loop in
the open string theory, the no-force is removed in favour of another vacuum configuration
in which the branes attract each other. For these reasons, our assumptions will be only
acceptable for distances much larger than the string scale, which is also the range of validity
of the classical solution for a BPS D-brane [47].
Using the non-BPS D-particle as a probe in the background of another non-BPS
D-particle we recover the no-force behaviour at the critical radius. Moreover, under the
4 The gravity duals of BPS branes in orbifolds have been considered before in [3436].
5 For a recent review see [37].
6 The same results were obtained earlier in [43] using different techniques.

254

E. Eyras, S. Panda / Nuclear Physics B 584 (2000) 251283

assumptions presented above we are able to obtain part of the complete metric, dilaton and
twisted RR 1-form generated by the D-particle. The velocity dependent part of the brane
action multiplies a flat metric, agreeing with the vanishing of the v2 in the interaction
potential. Extending this property for the complete geometry of the non-BPS D-particle
source, we are able to find more properties of the classical geometry. We expect that the
solution is consistent classically at the critical radius only, since it is only there where one
can consider a superposition of a large number of D-particles N .
We find a diagonal metric with SO(5) SO(4) symmetry (in Einstein frame):
2
= g00 (y) dt 2 + gmn (y) dy m dy n + gij (y) dx i dx j ,
ds10,E

where we do the split = (0, m, i) according to the orbifold symmetry, and I4 acts on the
i directions. By using the assumptions mentioned above we are able to find the form of
two of the components of the metric at the critical radius:

 7 a
6

6 T0
2 0 1 1
+

,
2
g00 (y) = 1 +
2a4
|y|3

1a
6

6 T0
2 0 1 1
+

mn .
2
gmn (y) = 1 +
2a4
|y|3
The expressions are given in terms of a parameter a, and some possible extra dependence
in |y|n , n < 3 denoted by which remain to be determined. We find no expression
for gij since, as will be explained in Section 4, the D-particle cannot probe the precise
geometry in these directions. Moreover, the form of the dilaton and twisted RR potential
at the critical radius are found to be:

a
1 1
6 T0
+

,
2 2 0
e = 1 +
2a4
|y|3

 4 a
3
6 Q0 1
(1)
+

1.
C0 = 1 +
3
4a4 |y|
Here T0 is the tension of the D-particle and is related to its charge Q0 by T0 = Q0 2 0 ,
hence at the critical radius the fields above are given in terms of a single function.
This article is organised as follows. In Section 2 we carry out the construction of the
covariant boundary state for the non-BPS D-particle in the orbifold of type IIB generated
by (1)FL I4 , for the non-compact and compact cases. We also evaluate the amplitude
for D-particles in relative motion and analyse the long and short distance behaviour of
the interaction. In Section 3 we evaluate the asymptotic behaviour of the massless fields
excited by the non-BPS D-particle, for the non-compact and compact cases. In Section 4 we
recover the no-force property at the critical radius and derive part of the complete classical
solution, by using the assumptions mentioned above. For completeness, we also include
the derivation of the zero-mode of the twisted RR boundary state in Appendix A, and the
explicit form of the complete covariant boundary state for the D-particle in Appendix B.

E. Eyras, S. Panda / Nuclear Physics B 584 (2000) 251283

255

2. Stable non-BPS D-particle in IIB/(1)FL I4


The stable non-BPS D-particle of type IIB string theory orbifolded by (1)FL I4 was
described in [1012]. It corresponds to the S-dual of a massive particle-like state in the
system of a D5-brane on top of an orientifold 5-plane, which is stable but non-BPS [10].
Its stability is due to the fact that this particle is the lightest state charged under the U (1)
gauge field of the D5O5 worldvolume theory. Similarly, the stability of the D-particle is
due to its charge under a twisted RR 1-form, which can be identified with the U (1) gauge
field in the worldvolume of a NS-5 brane on top of the orbifold fixed plane.
Let us consider first the non-compact theory, i.e., type IIB in Minkowski space
orbifolded by (1)FL I4 . The operator I4 corresponds to a reflection in the directions x i ,
i = 6, 7, 8, 9. In the non-compact case, the orbifold contains a fixed plane at x 6 = x 7 =
x 8 = x 9 = 0, hence the orbifold breaks the SO(1, 9) symmetry down to SO(1, 5) SO(4).
The operator FL is the spacetime fermion number of the left-sector, hence (1)FL changes
the sign of the of the RR groundstate, without having any action on the oscillators. The
closed string spectrum of the orbifold theory consists of an untwisted sector, given by
type IIB states which are invariant under (1)FL I4 , and a twisted sector localised on the
orbifold fixed plane. If we split the coordinates as X = (X , Xi ), where X , = 0, . . . , 5,
is longitudinal to the fixed plane, and Xi , i = 6, . . . , 9, transverse, the oscillators of the
twisted sector have the following modding:


r , r Z + 1/2,
n , n Z,
twisted NS:
ri , r Z + 1/2,
ni , n Z,


n , n Z,
n , n Z,
(2.1)
twisted R:
ri , r Z + 1/2,
ri , r Z + 1/2.
There are 4 fermionic zero-modes in the NS sector in the directions i = 6, 7, 8, 9, and 6
fermionic zero-modes in the R-sector in the directions = 0, 1, . . . , 5. Since the intercept
vanish for both sectors, the twisted groundstate will be given by these zero-modes. The
twisted NSNS sector gives a vector of SO(4), or equivalently 4 scalars under SO(1, 5). On
the other hand, the twisted RR sector gives rise to a vector under SO(1, 5). From the point
of view of the fixed plane, the supersymmetries are generated by two Weyl supercharges
of different chirality. This corresponds to (1, 1) supersymmetry in 6 dimensions. In fact,
the massless twisted sector can be identified with the worldvolume degrees of freedom of
a NS-5 brane of type IIB [11,48,49].
2.1. The D-particle boundary state
In this section we construct the boundary state for the stable non-BPS D-particle in type
IIB orbifolded by (1)FL I4 . This has been carried out in the light-cone gauge in [12]. Here
we use the covariant formalism for the boundary state, which has not been implemented
before for non-BPS D-branes in orbifolds. The D-particle boundary state is made up of an
untwisted NSNS part and a twisted RR part 7 [12]:
7 We use the subindex NS and R in the boundary states for short.

256

E. Eyras, S. Panda / Nuclear Physics B 584 (2000) 251283

|D0i = |D0iNS,U + |D0iR,T .

(2.2)

The NSNS boundary state is defined as the GSO-invariant combination of boundary states
|D0, iNS,U , with = 1, which turns out to be [12]:

1
|D0, +iNS,U |D0, iNS,U ,
2
where the GSO projector in the NSNS sector is given by
|D0iNS,U = PGSO,U |D0, +iNS,U =

(2.3)


1

1 (1)F +G 1 (1)F +G ,
(2.4)
4
with F and G the (worldsheet) fermion and superghost number operators, respectively:
PGSO,U =

F=

m m ,

G=

m=1/2

(m m + m m ),

(2.5)

m=1/2

and , are the superghosts. For the right movers these operators are analogously defined.
The state |D0, iNS,U takes the form [50]:
T0
|D0X i |D0gh i |D0 , iNS |D0sgh, iNS ,
(2.6)
2
where T0 , a constant related to the tension of the D-particle, is to be determined later.
There is a bosonic and a fermionic part, (|D0X i and |D0 , iNS ), and also a ghost and
a superghost part (|D0gh i and |D0sgh , iNS ). The boundary state |D0, iNS,U is very
similar to the NSNS boundary state for type II branes. In fact, the only piece modified
by the orbifold with respect to the type II case is the zero-mode part of |D0X i. For the
bosonic untwisted boundary state the most general conditions invariant under the orbifold
symmetry are
|D0, iNS,U =

X0 | =0 |D0X i = 0,
Xp | =0 |D0X i = y p ,
X | =0 |D0X i = 0,
i

p = 1, . . . , 5,
i = 6, . . . , 9,

from which we can deduce that


|D0X i =

(5)

(q y )
p

(4)

"
(q ) exp
i

(2.7)

X
1
n=1

#
n S n |k = 0i.

(2.8)

Note that the D-particle position in the directions i = 6, . . . , 9 is restricted by the orbifold
symmetry to be on the fixed plane. The other pieces of the untwisted boundary state are
not modified by the orbifold and are the same as for the type II case. These are explicitly
given in Appendix B. Moreover, since there are nine Dirichlet directions for the case of the
D-particle we have 8 S = .
The twisted RR boundary state that we denote by |D0iR,T is defined as the GSO
invariant combination of boundary states |D0, iR,T and is given by [12]:
8 We take the Minkowski metric to be mostly plus.

E. Eyras, S. Panda / Nuclear Physics B 584 (2000) 251283


1
|D0, +iR,T + |D0, iR,T .
2
In this sector the GSO-operator is given by:
|D0iR,T = PGSO,T |D0, +iR,T =

257

(2.9)


1

1 + (1)F +G 1 (1)F +G ,
(2.10)
4
with F and G the (worldsheet) fermion and superghost number operators in the twisted
R-sector, respectively:
PGSO,T =


m
m

(1)r=1/2
(1) = (1)m=1

X
(m m + m m ),
G = 0 0
F

i
r
ij r

,
(2.11)

m=1

and similarly for the right movers. Here (and ) represent the zero-mode parts of the
GSO-projectors, which are explicitly given in Appendix A. The twisted RR boundary
state |D0, iR,T is given by:
Q0
|D0X iT |D0gh i|D0 , iR,T |D0sgh, iR ,
(2.12)
2
where Q0 is a normalisation factor related to the charge density of the brane and to be
determined below. Notice that since this D-particle is non-BPS, Q0 6= T0 , unlike the BPS
D-branes. The (super)ghosts are not affected by the orbifold [51], hence the corresponding
pieces have the same form as for a type II RR boundary state. Since in the twisted RR
sector the bosons have integer modding along the orbifold fixed plane and half-integer
modding along the orbifolded directions, the state (2.12) have zero-modes along the fixed
plane only. Accordingly, the twisted bosonic part takes the form:
"
#

X1
X
1 i

j
(5) p
p

n +
ij n |k = 0i. (2.13)
|D0X iT = (q y ) exp
n n
n n
|D0, iR,T =

n=1

n=1/2

The fermions in the twisted RR sector have integer modding along the fixed plane
and half-integer modding along the orbifolded directions. As a consequence (2.12) have
fermionic zero-modes on the fixed plane only. The fermionic overlap conditions now read:


i S m |D0 , iR,T = 0, m Z,
m
1
j 
(2.14)
ri i S i j r |D0 , iR,T = 0, r Z + .
2
From them we can deduce:
"
#

X
X

i
m m i
m ij m |D0, i(0)
|D0 , iR,T = exp i
R,T .
m=1

m=1/2

(2.15)
|D0, i(0)
R,T

is determined from the zero mode overlap conditions in


The zero-mode part
(2.14) and is constructed upon the twisted RR groundstate:

258

E. Eyras, S. Panda / Nuclear Physics B 584 (2000) 251283

f
|D0, i(0)
R,T = Mab |aiT |biT ,

(2.16)

where a and b are spinor indices of SO(1, 5). The explicit form of Mab and its derivation
are given in Appendix A.
Having constructed the covariant boundary state for the non-BPS D-particle, we
determine next its tension and charge using the openclosed string consistency condition
for boundary states. The interaction between two D-particles separated by a distance y in
the fixed plane is given by the amplitude between two boundary states located at a relative
distance y with respect to each other, with the insertion of a closed string propagator
hD0|D|D0i,
with
0
Da =
4

Z
|z|61

(2.17)

d2 z L0 a L 0 a
z
z
,
|z|2

(2.18)

where a = 1/2 in the untwisted NSNS sector and a = 0 in the twisted RR sector.
Moreover, one needs to take the appropriate moddings in the expression for L0 and L 0 in
the twisted sector. Also care must be taken in computing the matrix elements involving the
zero modes of the superghosts and fermions in the twisted RR sector. This is because, in
general, the superghost zero modes produce infinite number of terms with any superghost
number contribution, hence a regularisation is needed. We use the same regularisation as
(0)
(0)
in [50]. Defining |D0, i(0)
R,T = |D0 , iR,T |D0sgh , iR we give below the regularised
result:
(0)
(0)
R,T hD0, 1 |D0, 2 iR,T

= lim

(0)
(0)
2F0 +2G0
|D0, 2 iR,T
R,T hD0, 1 |

= 41 2 ,1 , (2.19)

where F0 and G0 are the zero-mode parts of the operators F and G, implicitly given in
(2.11). Making a change of variables according to |z| = e` and d2 z = e2` d` d
we find:
9/2
V1 0
2 2 0
hD0|D|D0i =
16

y2

d` `9/2 e 2 0 `



 4 ` )f34 (e` )
f 8 (e` ) f48 (e` )
2
2 0 2 f2 (e
, (2.20)

(Q
(T0 )2 3
)

`
2
0
f14 (e` )f44 (e` )
f18 (e` )
where V1 is the (infinite) length of the D-particle worldline and fi are functions of e`
defined in the usual manner. We can make a worldsheet transformation, ` = 1/ , to
express the above closed string channel result in the open string channel. Note that we
have considered both closed string and open string to have same periodicity in the spatial
direction of the worldsheet. Openclosed string consistency then requires that the result
must be equal to the 1-loop amplitude of the open strings stretched between the two
D-particles. However, care must be taken to allow only states invariant under the orbifold
projection to propagate in the loop [11]. The open string states on a non-BPS D-particle

E. Eyras, S. Panda / Nuclear Physics B 584 (2000) 251283

259

are labelled by ChanPaton factors 9 1 and 1 . States with ChanPaton factor 1 have usual
GSO projection and are even under I4 . On the other hand, states with ChanPaton factor
1 have opposite GSO projection and are odd under I4 . Moreover, the symmetry (1)FL
does not have any action on the open string oscillators. Accordingly, the open string 1-loop
amplitude is given by:
Z
A = 2V1



d
1
TrNSR
1 (1)F +G (1 + I4 )e2 L0
2
4


1
1 + (1)F +G (1 I4 )e2 L0
4


Z
 2 L
d
1
F +G
0
TrNSR
1 (1)
I4 e
,
= 2V1
2
2

(2.21)

where F and G are the (worldsheet) fermion and superghost number operators for the open
string. The tachyon is projected out in the trace and this renders the non-BPS D-particle
stable [12]. We obtain the following 1-loop amplitude:
1/2
1
A = V1 8 2 0
2


Z
f44 (e )f34 (e )
d y 2 0 f38 (e ) f28 (e )
2

e
.
3/2
f14 (e )f24 (e )
f18 (e )

(2.22)

As mentioned above, openclosed string consistency allows us to fix the normalisation of


the boundary state:

Q0 = 8 0 .
(2.23)
T0 = 8( 0 )3/2 7/2 ,
charge.
The tension of the non-BPS D-particle is given by T0 and Q0 is related to its electric
Using that in ten dimensions 10 = 8 7/2gs ( 0 )2 , and that for the orbifold orb = 2 10 ,
we find the mass per unit volume of the D-particle:
M0 =

T0
1
= ,
orb g 2 0

(2.24)

which agrees with the mass of the D-particle found in [11]. In analogy with the BPS
D-branes,
define an electric charge with respect to the (twisted) RR field as
we can
0 = 2 Q0 = 8 2 0 . It is interesting to observe that the tension T0 of this non-BPS
D-particle is the same as that of a BPS D-particle of type IIA theory in ten dimensions. 10
However, their charges are different. In fact, 0 is exactly twice the charge of the BPS
D-particle of type IIA in six dimensions.
9 This can be derived from a D0 anti-D0 system in type IIA following the prescription given in [16,17].

2 times bigger than the type IIA BPS D-particle.


However, in the orbifold
case, there is an extra factor 1/2 in the open string amplitude coming from the projection
operator, such that the 2 factor is compensated.
10 The unstable non-BPS D-particle of type IIB has a tension

260

E. Eyras, S. Panda / Nuclear Physics B 584 (2000) 251283

2.2. The D-particle in the compact orbifold


Let us consider now the boundary state of a non-BPS D-particle in the case of the
compact orbifold, i.e., type IIB on T 4 /(1)FL I4 . The coordinates x i , i = 6, 7, 8, 9,
are periodic and there are 16 fixed planes instead of one, located at x i = 0, Ri . We
consider a non-BPS D-particle located on one of the fixed planes, that we choose x i = 0,
i = 6, 7, 8, 9, for simplicity.
The boundary state is now constructed upon a groundstate with zero winding and zero
momentum. The only piece modified in the boundary state with respect to the non-compact
case is the bosonic part and the overall normalisation. These two modifications will only
affect the NSNS part of the boundary state. The bosonic part is modified to include
KaluzaKlein modes in the directions i = 6, . . . , 9:
"
#
9 X
Y
X1
i ni

i
q

n S n |k = 0, n = 0i. (2.25)
e n Ri exp
|D0X i = (5) q p y p
n
i=6 ni Z

The states

n=1

|ni i = exp(iqni Rnii )|n = 0i

are normalised as

hni |n0i i = i ni n0i ,

(2.26)

where i is the self-dual volume [52] which satisfies


lim i = 2Ri ,

lim i =

Ri

Ri 0

2 0
.
Ri

(2.27)

We consider the new overall normalisation factor as a constant times the normalisation for
the uncompactified case:
T0
N |D0X i |D0gh i |D0 , iNS |D0sgh , iNS .
(2.28)
2
The factor N can be obtained by openclosed string consistency and is such that in the
decompactification limit Ri , we recover the non-compact boundary state [53]. The
amplitude between two non-BPS D-particles in the closed string channel is given by:
|D0, iNS,U =

5/2
V1 0
2 2 0
hD0|D|D0i =
16

Z
0

(
(T0 ) N
2

9
Y
i=6

(Q0 )2

y2

d` `5/2 e 2 0 `

9 X
Y
i=6 ni Z

f24 (e` )f34 (e` )


f14 (e` )f44 (e` )

2 `

ni
Ri

2

f38 (e` ) f48 (e` )


f18 (e` )

Using the worldsheet duality ` = 1/ and the Poisson resummation formula


r
X 2 (mR)2
X 0 ` n 2
2
e 2 R =
R
e 0 `
,
0
`
nZ

one obtains

mZ

(2.29)

(2.30)

E. Eyras, S. Panda / Nuclear Physics B 584 (2000) 251283

5/2
V1 0
hD0|D|D0i =
2 2 0
16
(

261

d y 2 0
e 2
3/2

! 9
2 Y
9
Y X 2 (m R )2 f 8 (e ) f 8 (e )
2 2 2
3
2
(T0 ) N
Ri i
e 0 i i
0
8 (e )

f
1
i=6
i=6 mi Z
)
f 4 (e )f34 (e )
.
(2.31)
(Q0 )2 44
f1 (e )f24 (e )


Openclosed string consistency imposes that this amplitude must be equal to the open
string 1-loop amplitude for the compactified case
Z
A = 2V1




1
d
TrNSR
1 + (1)F I4 e2 L0
2
2

1/2
1
= V1 8 2 0
2

( 9
Y X
i=6 mi Z

d y 2 0
e 2
3/2

2
(mi Ri )2
0

f38 (e ) f28 (e )
f18 (e )

f44 (e )f34 (e )
f14 (e )f24 (e )

)
.

(2.32)
This gives the same value for Q0 as in the uncompactified case, which is expected since
the compactification is done in the transverse directions of the D-particle. This also fixes
the normalisation factor N to be:
!1/2
9
Y
2Ri i
.
(2.33)
N=
i=6

It is easy to check that with this normalisation and using i = 2Ri , in the decompactified
limit, we can recover the untwisted boundary state for the non-compact case. Thus the net
effect of the compactification is a renormalisation of the tension of the D-particle and as
expected, the tension depends on the compactification radii. At this point we can recover
the vanishing of the amplitude at the critical radius [33]. The open strings on the D-particle
have now winding modes:



9 
X
1
wi Ri 2
1
2
+ 0 N
.
(2.34)
M =
0

2
i=6

The groundstate with zero winding is the tachyon and is projected out by the orbifold
symmetry. At a particular critical value of the radii
r
0
, i = 6, 7, 8, 9,
(2.35)
Ri =
2
there are four states, for which only one wi 6= 0, which become massless [16,2429,49].
These are the modes of the tachyon field that at the critical radius correspond to the

262

E. Eyras, S. Panda / Nuclear Physics B 584 (2000) 251283

marginal deformation which takes the D-particle to a bound state of a D-string and an antiD-string [11]. At this critical radius a BoseFermi degeneracy takes place, which translates
to the fact that the 1-loop amplitude vanishes [33]. In the closed string channel this can be
seen by using the relations:
X
2
e`n = f1 (e` )f32 (e` ),
nZ

f4 (e` )f2 (e` )f3 (e` ) =

2.

(2.36)

Moreover, the normalisation factor N at the critical radius becomes


!1/2
9
Y
1
i
.
N=
2 2 0

(2.37)

i=6

Hence we can write the amplitude as


5/2
V1 0
2 2 0
hD0|D|D0i =
16

y2

d` `5/2 e 2 0 `

0
4
`
)
f3 (e
4
4
`
`
f1 (e
)f2 (e
)f44 (e` )

2


T0
8 `
8 `
2 8 `
) f4 (e
) (Q0 ) f2 (e
) ,
f3 (e

2 0

(2.38)

which vanishes, at any distance y, by the abstruse identity and using the expressions for
T0 and Q0 . In the open string channel it is simpler to demonstrate this by using the above
identities.
2.3. Long and short distance interactions
In this section we consider the interaction amplitude for non-BPS D-particles in relative
motion. This will give a velocity dependent potential which we will compare with the
usual BPS case. This amplitude can be obtained by using a boosted boundary state [54] for
the non-BPS D-particle. For definiteness we consider a D-particle moving in the direction
X1 with a velocity v, interacting with another non-BPS D-particle at rest. We consider
first the non-compact orbifold. The boosted boundary state include some v-dependent
modifications that we write explicitly below. In the NSNS untwisted sector, the bosonic
part of the boundary state becomes
!
9
p
Y

p
(q ) q 1 + v q 0
|D0X , vi = 1 v 2
"
exp

i=2

X
n=1

#
1
n S(v) n |k = 0i,
n

where the matrix S(v) is given by

(2.39)

E. Eyras, S. Panda / Nuclear Physics B 584 (2000) 251283

263

1 + v2
2v
,
S10 (v) = S01 (v) =
,
(2.40)
1 v2
1 v2
and S(v) = S for the other components. The other pieces in the NSNS untwisted boundary
state have the usual form except for a substitution of the matrix S by S(v). Similarly in the
RR twisted sector, the boosted bosonic boundary state takes the form:
S00 (v) = S11 (v) =

|D0X , viT =

1 v2
"

5
Y

(q p ) q 1 + v q 0

p=2

exp

X1
t.m.


#

n S(v) n |k = 0, n = 0i,

where with t.m. we indicate that one takes the corresponding twisted moddings of the
twisted sector. For the other pieces we do the same substitution S S(v). Finally, for the
RR zero-mode part there is as well a v-dependent modification:
f
|D0, , vi(0)
R,T = Mab (v)|aiT |biT ,

(2.41)

and the matrix Mab (v) is given in Appendix A. We define the distance between the particles
as r 2 = b2 + t 2 v 2 (1 v 2 )1 , where t is the proper time of the moving particle along which
we also integrate; b is the impact parameter: b2 = y22 + + y52 . For convenience we also
define the following variable
1v
1
ln
.
2i 1 + v
Finally, the cylinder amplitude takes the following form:
u=


2 0 1/2

hD0, v|D|D0i = 8


Z
sin u

d` `

29

(2.42)

r2

dt e 2 0 `

0
6 `
6 `
3 (u, i`)f3 (e
) 4 (u, i`)f4 (e
)
4`2
6
`
1 (u, i`)f1 (e
)

2 (u, i`)f22 (e` )f34 (e` )


,

1 (u, i`)f12 (e` )f44 (e` )


(2.43)

where s (u, i`) are the Jacobi -functions, and we have used the value of the tension and
charge of the D-particle. For v = 0 we recover the amplitude for the static interaction given
in (2.20).
In order to study the interaction we define the interacting potential of the scattering U ,
in the following way:
Z
hD0, v|D|D0i =

dt U(v, r(t)).

(2.44)

We can extract the long range interaction potential taking the limit ` in the integrand
of (2.43), and then performing the integral in `. For slow velocities, we find the following
expansion in powers of v:

264

E. Eyras, S. Panda / Nuclear Physics B 584 (2000) 251283

(2 0 )3
Uclosed '
(4)1/2




1
1
1
(3/2)
(7/2)
1 4 (7/2)
v
8 1 + v2 + v4

+
,
2
24
r7
(2 0 )2 r 3
8
r7

(2.45)

to order v 4 . There is a static force, as expected since the D-particle is non-BPS. Moreover,
the velocity dependent corrections start at order v 2 , as for the potential between two
BPS D-branes of different dimensionality [22,23], and as for the non-BPS D-particle of
type I [55]. This is something we could expect from the fact that the particle breaks all
supersymmetries.
We can analyse this amplitude for very short distances. In this region the open string
description dominates. Using the modular properties of the -functions we can write
(2.43) in the open string channel:

2 0 1/2

hD0, v|D|D0i = i 8


Z
sin u

d
0

12

r2

dt e 2 0

3 (iu, i )f36 (e ) 2 (iu, i )f26 (e )


1 (iu, i )f16 (e )

4 (iu, i )f42 (e )f34 (e )


1 (iu, i )f12 (e )f24 (e )


.

(2.46)

We can derive the interaction potential at short distances by taking the limit in
the integrand and then performing the integral in . Note that the pieces originally from
the NSNS and RR sectors multiply now in (2.46) the same powers of , hence we can
expect that the interaction potential will be very different from the closed string case. This
time we obtain:

2 0 1/2

Uopen = i 8

Z
sin u

r2

d 1/2 e 2 0

2 + 2 cos(2iu ) 8 cos(iu )
,
sin(iu )

(2.47)
which for slow velocities and after integrating in becomes (to order
Uopen =

4
(2 0 )3 (7/2) v 4
r
+
.
2 0
r7
(4)1/2

v 4 ):
(2.48)

Surprisingly, there is no v 2 correction to the potential, as for BPS D-branes. The first term
is a linear atractive force coming from the stretched strings. Moreover, the v4 correction
has precisely the same form as the v 4 term for a scattering of two BPS D-particles. For
BPS branes the v 4 term in the open string channel coincides with the v4 term from the
closed string description [47]. This does not occur in the non-compact orbifold. As we will
see below for the compact case, even though at the critical radius the static potential and
v 2 terms will be suppressed, the matching of the v 4 terms will not occur either. Hence the
BPS-like behaviour will not extend beyond order v 2 .

E. Eyras, S. Panda / Nuclear Physics B 584 (2000) 251283

265

We extend next the analysis of the cylinder amplitude for the case of relative motion
of the D-particles in the compact orbifold. We consider again one non-BPS D-particle
moving in the direction X1 with a velocity v, interacting with another non-BPS D-particle
at rest. The boosted boundary state include the v-dependent modifications on top of the
modifications due to compactification. In the NSNS untwisted sector, the bosonic part of
the boundary state becomes
!
5
p
Y

(q p ) q 1 + v q 0
(2.49)
|D0X , vi = 1 v 2
p=2

9
Y

i
iq ni Rn
i

"

exp

i=6 ni Z

X
1
n=1

#
n S(v) n |k = 0, n = 0i,

(2.50)

where the matrix S(v) is as given before in (2.40). In the RR twisted sector, the boosted
bosonic boundary state and fermionic zero-mode have the same form as for the noncompact case, and the other pieces have the usual form except for a substitution of the
matrix S by S(v). Using the same notation as before, the cylinder amplitude takes the
following form:
r
hD0, v|D|D0i =

2
sin u
2 0
(

Z
d` `
0

9
0 2 Y

4


i=6

Ri1

25

r2

dt e 2 0 `

9
Y

2 `

ni
Ri

2 !

i=6 ni Z

3 (u, i`)f36 (e` ) 4 (u, i`)f46 (e` )

1 (u, i`)f16 (e` )


)
2 (u, i`)f22 (e` )f34 (e` )
.

1 (u, i`)f12 (e` )f44 (e` )

(2.51)

For v = 0 we recover the amplitude for the static interaction given in (2.29). For slow
velocities (u ' (i)1 v) and after integration in `, the long range interaction potential to
order v 4 is given by
!
)
(
 02 Y
9
4 0
v4
1 2 1 4

1
Ri 1
.
(2.52)
1+ v + v
Uclosed ' 3
r
2
6
4
8
i=6

We see that for generic radii the potential has v 2 corrections, as before, which is typical
for potentials between two BPS D-branes of different dimensionality [22,23], and occurs
also for the non-BPS D-particle of type I [55]. On the other hand, at the critical radius
they start at v 4 , as for BPS D-branes [22,56]. Notice that the static and v2 terms of the
potential also vanishes for other values of the radii. However, those radii do not make the
amplitude (2.29) vanish. Moreover, this would require some of the radii to be below the
critical radius, and the D-particle would not be stable.

266

E. Eyras, S. Panda / Nuclear Physics B 584 (2000) 251283

In order to study the short-distance behaviour we write the scattering amplitude in the
open channel:
r
Z
Z
r2
2
12
sin u d
dt e 2 0
hD0, v|D|D0i = i
2
0

(

1
4

9
Y

2
0

(mi Ri )2

!

3 (iu, i )f36 (e ) 2 (iu, i )f26 (e )

i=6 mi Z

4 (iu, i )f42 (e )f34 (e )


1 (iu, i )f12 (e )f24 (e )

1 (iu, i )f16 (e )

(2.53)

To extract the leading contribution at short distances and low velocities we take the limit
in the integrand. This time one must also include contributions with one unit of
winding number mi = 1, which are relevant very close to the critical radius Ri > Rc :
r
Z
r2
2
sin u
d 1/2 e 2 0
Uopen =
2
0

2i sin(iu )
0

1 X 2 R0 i
e
cos(2iu ) 4 cos(iu ) + 1 +
2
9

!
+

(2.54)

i=6

At low velocities the expression simplify. Carrying out the integral in we find:
!
9 q

4
(2 0 )3 (7/2) v 4
1X 2
2 R 2 0 /2
r
+
4
+
. (2.55)
r

Uopen =
i
2 0
4
r7
(4)1/2
i=6

As for the non-compact case we find no v 2 corrections, and the v 4 corrections are again
the same as for the BPS D-particle in ten dimensions. Notice also that the r-dependence of
the v 4 term is as for the uncompactified case. The static part of the potential vanishes when
all the radii are equal to the critical radius. Remarkably, at the critical radius Uopen takes a
very BPS-like form. However, as we announced before, the v 4 terms does not match with
the closed string result.
From this analysis we conclude that the long and short range interactions of the nonBPS D-particle are quite different for generic radii of the orbifold. In the particular case of
the critical radius the static and v 2 terms of the interaction potential are absent in the open
and closed string description. At the critical radius, to order v 2 , we expect an equivalence
between the (super)gravity and worldvolume description of the non-BPS D-particle. At
order v 4 the open strings begin to describe the geometry of non-BPS D-particle very
differently from the closed strings.

3. Spacetime description
As shown in [44], the boundary state of a D-brane encodes the long distance behaviour
of the corresponding classical solution of supergravity. At large distances, this classical

E. Eyras, S. Panda / Nuclear Physics B 584 (2000) 251283

267

solution tends to a flat background configuration. The fluctuations around this background
is exactly reproduced by the boundary state. From the NSNS part one obtains the
asymptotic behaviour of the dilaton and the metric, and from the RR part one obtains
the asymptotic behaviour of the RR form potential under which the brane is charged. In
this section we implement this technique on the stable non-BPS D-particle to obtain the
asymptotic form of the solution. We obtain a metric and a dilaton in the bulk, which depend
on all the coordinates transverse to the D-particle, and whose dependence is the expected
one for a particle in 10 dimensions. On the other hand, we also find a twisted RR 1-form
which is restricted to the fixed plane and whose dependence on the spatial coordinates
on the fixed plane is the correct one for a particle in 6 dimensions. We describe first the
non-compact case.
3.1. The non-compact case
Given a certain massless closed string state |i, normalised as h|i = 2 , one can
define a projection operator hP() | associated to this state, such that hP() |i = [45].
The asymptotic behaviour of the classical field , generated by a D-brane is determined
by computing the overlap of the boundary state with the state |P() i with the insertion of a
closed string propagator [44,45]:
hP() |D|Di.

(3.1)

Since the D-particle has an untwisted NSNS part and a twisted RR part, the D-particle
will excite gravitons g and dilatons in the untwisted sector and a RR 1-form field
C (1) in the twisted sector. At large distances, these fields will be of the form of a fluctuation
around a flat background 11 :

' 10 2 ,
C (1) ' 6 2 C (1) .
(3.2)
g ' + 210 h ,
The fluctuations h , and C (1) , are given by (3.1) using the state |D0i constructed
previously. We consider first the case of the uncompactified orbifold. In the untwisted
sector, only those which are invariant under the orbifold symmetry will appear in the theory.
This implies that string states which depend on the momentum in the orbifolded directions
are only invariant if they are symmetric with respect to
I4 : k i k i ,

i = 6, 7, 8, 9.

In particular, these will include gravitons that propagate off the orbifold fixed plane which
are symmetric in k i .
The projection operators in the untwisted NSNS sector are given in the (1, 1)
picture. For simplicity, we use the following notation for the NSNS groundstate in the
(1, 1) picture:
] 1 .
|ki |k/2i1 |k/2i
The projectors for the dilaton and the metric are given by:
11 We use the same normalisation as in [44].

(3.3)

268

E. Eyras, S. Panda / Nuclear Physics B 584 (2000) 251283


1

hP() | = hk , k i | + hk , k i | 1/2
1/2 ( k ` k ` ),
4 2


1

1/2 + 1/2 1/2


hP(h) | = hk , k i | + hk , k i | 1/2
4
1
hP() | ( k ` k ` ),
2 2

(3.4)

where k = (k , k i ), with k the spatial components of the momentum along the fixed plane
directions. Moreover, ` is such that `2 = 0 and k ` = 1.
In the RR sector, the most natural picture is the asymmetric 12 picture (1/2, 3/2),
where the RR vertex operator couples to the potential instead of the field strength [50,
57]. In analogy to the type II case described in [50] and using a, b as the SO(1, 5) spinor
indices, we write the following quantum state in the (1/2, 3/2) picture associated to
the twisted RR 1-form field 13


1
^
k/2i3/2 ,
|C (1) i = C(1) (C + )ab cos 0 0 + (C )ab sin 0 0 |a, k/2i1/2 |b,
2
whose conjugate state is given by:
1
^
k/2|3/2 ha, k/2|
hC (1) | = C(1) 1/2 hb,
2


(C )ab cos 0 0 + (C + )ab sin 0 0 ,
with
1
(3.5)
= (18 ),
2
the chirality projector in 6 dimensions, where the matrix is defined in Appendix A. This
(1)
state is normalised as hC (1) |C (1) i = C C (1) . The corresponding projector is given by
1
^
k /2|3/2 ha, k /2|
hP(C) | = 1/2 hb,
2


(C )ab cos 0 0 + (C + )ab sin 0 0
such that hP(C) |C (1) i = C(1) . The form of this projector is similar to the case of the type
IIA BPS D-particle, with the difference that the twisted RR groundstate carries spinor
indices of SO(1, 5).
The NSNS (RR) fields have non-zero overlap with the NSNS (RR) boundary state
only. We find the following results for the asymptotic fields in momentum space:
3 V1
(k) = hP() |Da=1/2|D0iNS,U = T0 2 ,
8 k


1
T0 V 1
7 1
,
,
.
.
.
,
diag
.
h (k) = hP(h) |Da=1/2|D0iNS,U =
2 k2
4 4
4
12 See [44] for an explanation about this.
13 For simplicity we omit the subindex T of the twisted groundstate.

(3.6)
(3.7)

E. Eyras, S. Panda / Nuclear Physics B 584 (2000) 251283

269

For the twisted RR field we only have a contribution along the worldline direction of the
D-particle, namely:
Q0 V1
C0(1) (k ) = hP(C) 0 |Da=0 |D0iR,T = 2 .
2 k

(3.8)

Note that C0(1) depends only on the momenta transverse to the D-particle, longitudinal
to the fixed plane. We now make use of the following Fourier transformation in order to
translate the results into position space. For a generic momentum K, with d 1 non-zero
components, we have:
Z
eiKx
V1
=
,
(3.9)
dt d(d1)x
(d 3) d2 |x|(d3) K 2
where x are d 1 spatial coordinates transverse to the D-particle and d2 is the area of
a unit sphere surrounding the D-particle. For the NSNS sector we have d = 9, hence we
find the following asymptotic behaviour for the metric and dilaton:
1
3T0
,

7
|x|
14 2 8


1
T0
7 1
1
diag , , . . . ,
.
h (x) =
14 8 |x|7
4 4
4
(x) =

(3.10)

In the RR sector, since the momentum dependence is restricted to the fixed plane spatial
directions, we have d = 6, and the asymptotic form of the RR 1-form is found to be:
Q0
1
,
C (1) (y) =
3
|y|
3 2 4

(3.11)

where now y indicates the spatial directions in the fixed plane only, and 4 is the volume
of a unit 4-sphere surrounding the particle inside the fixed plane. Note that the power of
|y| is exactly the power expected for a particle in 6 dimensions. We see, however, that the
untwisted fields do see the entire spacetime. Thus, although the D-particle is stuck on the
fixed plane, the associated metric and dilaton background solution in the uncompactified
case will extend also in the orbifolded directions. This is a consequence of the fact that the
D-particle can emit massless untwisted fields off the fixed plane which are symmetrised in
the momenta along the orbifolded directions.
3.2. The compact case
In order to consider a spacetime description for a stable non-BPS D-brane, in a regime
where it remains valid classically, one should make sense of a superposition of them, which
turns out to be possible only in the compact orbifold at the critical radius. In this section
we extend the analysis of the metric and dilaton of the D-particle to the case in which the
space transverse to the orbifold fixed plane is a 4-torus, i.e., type IIB on T 4 /(1)FL I4 .
We derive first the asymptotic form of the metric and dilaton in the compactified case
using an approximation in the background fields, that we will compare with the boundary
state calculation. Consider one of the compact orbifolded directions: x9 x9 + 2R9 .

270

E. Eyras, S. Panda / Nuclear Physics B 584 (2000) 251283

A D-particle sitting at the origin of this S 1 can be seen from the point of view of the
covering space as an infinite array of equally spaced D-particles. Thus we can write:
X
1
1
'
,
(3.12)
7
2
|x|
(r + (x9 2nR9 )2 )7/2
nZ

with |x|2 = r 2 + x92 , and r 2 = x12 + + x82 . We can approximate the sum by an integral
assuming that the distance in the non-compact directions is much larger that the size of the
compact one, i.e., r  R9 . Changing variables we can write
X
nZ

1
1
'
2
2
7/2
(r + (x9 2nR9 ) )
2R9 r 6

Z+
du

1
I5 1
=
,
2
7/2
(1 + u )
2R9 r 6

where we use the notation:


Z
In = d sinn .

(3.13)

(3.14)

Moreover, these integrals satisfy the properties nIn = (n 1)In2 , for n > 2, and
d = 2Id1 Id2 I1 I0 .

(3.15)

If we do this approximation for each of the compactified directions of the orbifold we


obtain
9
Y
1
1
'
I
I
I
I
(2Ri )1 3 ,
2
3
4
5
|x|7
|y|

(3.16)

i=6

with y now indicating the spatial directions in the fixed plane. Within this approximation
we obtain the following asymptotic behaviour for the dilaton and metric for the
compactified case:
9
Y
1
T0
(2Ri )1 3 ,
(y) '
|y|
2 2 4 i=6

h (y) '



9
1
T0 Y
7 1
1
.
(2Ri )1 3 diag , , . . . ,
6 4
4 4
4
|y|

(3.17)

i=6

We rewrite the fluctuations using the 6-dimensional gravitational constant


g ' + 210 h + 26 h ,

' 10 2 6 2 ,
(1)
(1)
C C .
Using the relation 14 between 10 and 6 : 10 = 6

Q9

i=6 (2Ri )

(3.18)
1/2 ,

we obtain

14 We use that for any dimension D, 2 2 = 16 G , and for any lower dimension D d, G
D
Dd = GD /Vd ,
D
with Vd the volume of the compact space.

E. Eyras, S. Panda / Nuclear Physics B 584 (2000) 251283

271

9
Y
T0
1

(y)
'
(2Ri )1/2 3 ,
|y|
2 2 4 i=6

h (y) '



9
1
T0 Y
7 1
1
.
(2Ri )1/2 3 diag , , . . . ,
6 4
4 4
4
|y|

(3.19)

i=6

The asymptotic behaviour of the twisted RR 1-form stays the same as for the non-compact
case (3.11).
We compare now this approximation with the asymptotic behaviour derived from the
boundary state. The closed string fields have an infinite number of KaluzaKlein modes,
since the momentum is quantised in the orbifolded directions. The massless closed string
states excited by the D-particle correspond to the massless KaluzaKlein modes. Using the
same notation as before and using bars for the objects in the compact case, we can write
the projection operators for the massless states as follows:

1 Y 1/2

1/2 ( + )k ` ,
i
hk , ni = 0| 1/2
hP() | =
2 2 i=6
9

hP(h) | =


1 Y 1/2

1/2 + 1/2 1/2


i
hk , ni = 0| 1/2
2
9

i=6


1
hP() | ( + )k ` ,
2 2

(3.20)

where now k ` = 1 and ` ` = 0. Instead of splitting the projectors in terms of


6-dimensional fields, we keep the ten-dimensional notation in order to compare it with the
non-compact results, and with the approximation made previously. As before, the NSNS
fields will have non-zero overlap with the NSNS boundary state. The type of calculation
we perform in this case is similar to the non-compact case, with the difference that there
appear extra normalisation coefficients, and the momentum dependence is only in the
Q 1/2
in the normalisation
spatial directions of the fixed plane. Note that the factors of i i
Q
of the boundary state and the massless states cancel with a factor i i coming from the
normalisation (2.26) in the amplitude. We find the following results:
3T0 Y
V1

(2Ri )1/2 2 ,
(k)
= hP() |Da=1/2 |D0iNS,U =
k
2 2 i
h (k) = hP(h) |Da=1/2|D0iNS,U


1
T0 Y
7 1
1/2 V1
=
(2Ri )
diag , , . . . ,
.
2
2
4 4
4
k

(3.21)

We can make use of the Fourier transformation (3.9) in the NSNS sector to derive the
spacetime behaviour:
Y
1
T0

(2Ri )1/2 3 ,
(y)
=
|y|
2 2 4
i

272

E. Eyras, S. Panda / Nuclear Physics B 584 (2000) 251283

h (y) =



7 1
1
1
T0 Y
(2Ri )1/2 3 diag , , . . . ,
.
6 4
|y|
4 4
4

(3.22)

The twisted RR sector remains the same as for the uncompactified case (3.11). Note
that the asymptotic behaviour derived from the compactified boundary state (3.22) exactly
reproduces the behaviour of the 6-dimensional fluctuations obtained before (3.19) using
the approximation given in (3.16). The asymptotics of the dilaton and the graviton are
now certainly comparable with that of the twisted RR 1-form. In particular, at the critical
radius there is an accidental BoseFermi degeneracy, which as we will see in the next
section, implies a no-force in a brane probe. At the critical radius we can then consider a
large number of branes N , hence the fluctuations take the following form:
1
T0 N

,
(y)
=
2
0
2 2 4 (2 ) |y|3


1
1
T0 N
7 1
,
,
.
.
.
,
diag
,
h (y) =
64 (2 2 0 ) |y|3
4 4
4
Q0 N
1
.
C (1) (y) =
3 2 4 |y|3

(3.23)

4. No-force condition at the critical radius


A composite of branes preserving the BPS property satisfies a no-force condition
between the constituents. This was verified at the level of the effective action and
background geometries for BPS branes in [23]. The no-force property is a consequence
of the vanishing of the one-loop amplitude between the D-branes. On the other hand,
the BoseFermi degeneracy for non-BPS D-branes at the critical radius described in [33]
represents a remarkable example in which the no-force property takes place without being
BPS. In this section we recover the no-force condition using the non-BPS D-particle as
a probe of the geometry of another non-BPS D-particle. Firstly, we consider the effective
action for the non-BPS D-particle.
4.1. The action of the D-particle
Following the prescription given in [58] one can construct the action for the D-particle
in the orbifold of type IIB we are considering. One starts with the action of a non-BPS
D-particle in type IIB, and set to zero all fields which are odd under (1)FL I4 . The

worldvolume theory of this D-particle is given by scalars b1/2|0iNS 1, = 1, . . . , 9,


a tachyon |0iNS 1 , and 16 fermions in the Ramond groundstate of each of the Chan
Paton sectors 1 and 1 . The orbifold (1)FL I4 eliminates the scalars in the directions
i = 6, 7, 8, 9, the tachyon and the fermions with ChanPaton factor 1 [16]. Accordingly,
the non-BPS D-particle is stable and contains 5 physical bosonic degrees of freedom,
describing the motion of the D-particle on the orbifold fixed plane, plus 16 fermionic d.o.f.
If we label the coordinates as before X = (X , Xi ), where = 0, 1, . . . , 5 labels the
orbifold fixed plane directions, we can write the following action for the D-particle:

E. Eyras, S. Panda / Nuclear Physics B 584 (2000) 251283

Skin =

T0
10

q
d e | X X G |,

273

(4.1)

where G is the background string metric. Furthermore, since the D-particle carries
charge under the twisted RR 1-form C (1) , we can include a WZ coupling 15 , which
involves the 1-form potential on the fixed plane:
Z
Q0
(4.2)
X C(1).
SWZ =
26
The couplings of the stable non-BPS D-particle to the background fields given by Eqs. (4.1)
and (4.2) can be also derived by projecting the boundary state of the D-particle onto
the projectors defined previously, analogously to the BPS case considered in [44]. The
coupling to the dilaton is given by
3T0
J() = hP() |D0iNS,U = V1 ,
2 2

(4.3)

whereas the coupling to the metric is given by

J(h) = hP(h) |D0iNS,U = T0 V1 h00 ,

(4.4)

where h is the symmetric and traceless helicity tensor of the graviton. These two results
reproduce the couplings
of (4.1) in Einstein frame (g = e/2 G ) after rescaling the

dilaton as = 10 2 . Finally, the coupling to the twisted RR 1-form is given by:


Q0
(1)

|D0iR,T = V1 C0 ,
J(C) = C(1) hP(C)
2

(4.5)

which reproduces (4.2), after the rescaling C (1) = 6 2 C (1) .


In the compact case, the coupling of the D-particle to the massless closed string fields
can be obtained as above, this time using the compactified boundary state constructed
in Section 2.2. The net result is a change of the factor in front of the kinetic term, which
amounts to the relation of the 6-dimensional gravitational constant with the 10-dimensional
one. The boundary state reproduces the couplings of the following action:
Skin =

Z
9
q
T0 Y
(2Ri )1/2 d e | X X G |.
6

(4.6)

i=6

Thus although the D-particle does not couple to the components Gij (4.1), it does feel
the transverse geometry through the renormalised tension. Moreover, the open strings on
the D-particle carry winding modes, and at the critical radius, the winding modes of the
tachyon become massless. These extra modes appear as new degrees of freedom, i , in the
effective action [49,58]:
Z
q

T0
d e | X X G | 1 G i i .
(4.7)
Skin = 2 0
2 6
However, these extra modes will not play any role in our discussion below.
15 WessZumino couplings for stable non-BPS D-branes have been recently considered in [59].

274

E. Eyras, S. Panda / Nuclear Physics B 584 (2000) 251283

4.2. D-particle at the critical radius


We consider a non-BPS D-particle probe moving in the background of another nonBPS D-particle in the compactified orbifold. We split the embedding coordinates as X =
(X0 , Xm ), m = 1, . . . , 5, and choose the static gauge:
X0 = .

(4.8)

We identify the embedding scalars with the transverse coordinates of the background
geometry and expand the action in derivatives 16 :


q
p
1
m
n

| y y g | = |g00 | 1
0 y 0 y gmn + .
(4.9)
2g00
Moreover, for a D-particle background, only the zeroth component of the twisted RR
1-form is turned on:
X C(1) = C0(1) .

(4.10)

The action can then be approximated by


3
Z
Z
9
e 4
T0 Y
(2Ri )1/2 d
0 y m 0 y n gmn ,
S ' d V(y) +
6
2 |g00 |

(4.11)

i=6

where we have defined the effective static potential V:


3 p
Q0 (1)
T0 Y
(2Ri )1/2 e 4 |g00 | +
C .
6
26 0

V(y) =

(4.12)

i=6

If this potential is not constant or zero for a given background g00 (y), (y), C0(1)(y), there
will be a force term in the field equations for y m . Plugging into this potential the background geometry generated by another D-particle, one can check whether the BoseFermi
degeneracy takes place. Since from the boundary state we obtain the asymptotic form of
the D-particle solution, we rewrite the potential V in terms of the fluctuations (3.18) 17 :


9
Y
3
Q0
(4.13)
V(y) ' T0 (2Ri )1/2 + h 00 + C0(1) .
2 2
2
i=6
We can insert now the asymptotic form of the solution for the compactified case given by
(3.22) and (3.11). The potential then takes the form:
!
9
Y
Q20
2
2
1
(2Ri )
T
V(y) =
4
34 |y|3 0
i=6
!
 0 2 Y
9

4 0
Ri1 1 ,
(4.14)
=
|y|3
2
i=6

16 We use the Einstein frame since we want to use in the action the results given by the boundary state
computation.
17 We neglect the constant parts of the potential, since these would not generate any force.

E. Eyras, S. Panda / Nuclear Physics B 584 (2000) 251283

275

which coincides with the static potential


p derived from the cylinder amplitude in (2.52) and
vanishes for the critical radius (Ri = 0 /2), hence we recover the no-force at the critical
radius using the background geometry generated by the non-BPS D-particles.
According to the interaction potential Uopen derived in Section 2.3, the velocity
corrections start at order v 4 and there is no v 2 corrections to the potential in for any radius,
as happens for BPS branes. From the probe point of view, this translates into the fact that
the metric on the moduli space, multiplying the velocity dependent piece of the action
(4.11), is flat [23]. In the asymptotic limit, for the non-compact orbifold this metric takes
the form:

3

3
e 4
gmn ' mn 1 10 (x) + 10 h00 (x) + 210 hmn (x).
(4.15)

|g00 |
2 2
Using the asymptotic fluctuations given in (3.10), it is easy to see that the metric is flat
(= mn ). In the case of the compact orbifold, this metric factor takes the form:


3
e 4
3

gmn ' mn 1 6 (y)

+ 6 h 00 (y) + 26 h mn (y).
(4.16)
|g00 |
2 2
Substituting the asymptotic fluctuations (3.22), we obtain the same flat metric for any value
of the radii. Thus we recover the behaviour given by the open strings in (2.48) and (2.55).
4.3. The classical geometry of the non-BPS D-particle
In this section we assume that the no-force condition will persist at the full non-linear
level of the field equations. We can then restrict the possible classical geometries by
imposing the no-force at the level of equation (4.12). No-force can occur for V either
constant or zero [23]. However, in our case, the no-force condition at the critical radius
must enforce the relation T0 = ( 2 0 )Q0 , as seen in (2.38). These are in fact the factors of
the NSNS and RR contributions of the potential (4.12), respectively. Hence we conclude
that the classical solution must be such that at the critical radius the following equality
holds
3 p
(1)
(4.17)
e 4 |g00 | = C0 .
We can then deduce part of the form for g00 , , and C (1) :

 7 a
6

6 T 0
2 0 1 1
+

,
2
g00 (y) = 1 +
2a4
|y|3

a
6 T0
1
(2 0 )1 3 + ,
e (y) = 1 +
2a4
|y|

 4 a
3
6 Q0 1
(1)
+
1.
C0 (y) = 1 +
4a4 |y|3

(4.18)

These functions are given in terms of a parameter a yet to be determined, and are such that
asymptotically they become the fluctuations in (3.19) and (3.11), and at the critical radius
equation (4.17) holds. The dots indicate other possible contributions with dependence |y|n ,

276

E. Eyras, S. Panda / Nuclear Physics B 584 (2000) 251283

n < 3, which are subleading in the asymptotic limit (|y| ), but which are relevant
when we come closer to the brane. We expect these extra terms to appear since harmonicity,
which is a direct consequence of supersymmetry, may not be present in the complete
solution.
We can make one further assumption if we consider that the metric factor in the velocity
dependent piece of the action (4.11) remains flat for the complete geometry. We impose
then the following relation:
3

e 4

gmn = mn .
|g00 |

(4.19)

Note that the functions describing the dilaton and the metric must be the same at the critical
radius. Moreover, it is only at this radius where we actually expect to find a consistent
classical geometry. Therefore, we make use of the expressions for g00 (y) and (y) given
above to obtain:

1a
6

6 T0
2 0 1 1
+
mn ,
(4.20)
2
gmn (y) = 1 +
3
2a4
|y|
where the dots represent the same contributions as in (4.18) at the critical radius. Finally,
it seems that gij is out of the reach of the present analysis. A possible way to include it
in the analysis is by a coupling to the extra massless states i in (4.7). On the other hand,
although the asymptotic behaviour of the metric in the non-compact case (3.10) presents
an SO(9) symmetry, this is certainly broken to SO(5) SO(4) by the orbifold when we get
closer to the position of the brane. This is in fact already suggested by the asymptotics in
the compact case (3.19). Accordingly, the form of gij (y) is expected to be different from
gmn (y).

5. Comments
In this paper we have investigated the description of a stable non-BPS D-particle in
terms of a classical solution. We have used the technique of the boundary state to obtain
the asymptotic form of the classical solution for the non-compact and compact versions
of the orbifold. We find a metric and a dilaton propagating in the bulk, and a twisted
RR 1-form propagating in the fixed plane. In the non-compact case the bulk fields have
the dependence expected for a particle in ten dimensions, whereas in the compact case
they have a dependence typical of a particle in six dimensions. The twisted RR 1-form
has in both cases the same asymptotic form with the usual dependence for a particle in
six-dimensions. Using the non-BPS D-particle as a probe in the background of another
non-BPS D-particle, we have recovered the no-force property at the critical radius using
the asymptotic behaviour. Moreover, we have calculated the cylinder amplitude for nonBPS D-particles in relative motion. From it we have extracted the long and short distance
interactions. For generic radii these contain v 2 corrections in the closed string description,
but the open string description presents no v 2 terms for all radii, like for BPS D-branes.
Moreover, the v 4 corrections do not match, unlike for BPS branes. On the other hand, at

E. Eyras, S. Panda / Nuclear Physics B 584 (2000) 251283

277

the critical radius they present a BPS-like behaviour, up to the v4 corrections, which do not
match in the open and closed descriptions.
We have assumed that the no-force property of a brane probe holds for the full
background geometry. This is acceptable for distances much larger than the string scale.
This assumption allows us to derive part of the classical solution, which reproduces the
asymptotic behaviour and the no-force property. On the other hand, we expect that extra
terms may appear in the solution at the classical level. This is due to the fact that the
boundary state only gives information about the next-to-leading term in the asymptotic
limit, hence subleading terms which become relevant at short distances escape from this
analysis. Moreover, the fact that there is a coordinate system in which the metric can be
written in terms of harmonic functions is very much related to residual supersymmetry,
which does not occur in our case. This fact does not permit to find information about
the geometry near the core of the non-BPS D-particle. Moreover, although the asymptotic
behaviour of the metric (3.10) presents an SO(9) symmetry, this is expected to be broken
to SO(5) SO(4) by the orbifold when we get closer to the position of the brane.
The form of the asymptotic behaviour we have found suggests that the classical solution
for the non-BPS D-particle will be a solution to the field equations derived from an action
involving 10-dimensional bulk fields and 6-dimensional matter fields constrained on an
orbifold fixed plane:
Stotal = Sbulk + Splane .

(5.1)

The bulk action involves the metric and the dilaton, and other fields of the massless
untwisted sector. For the case of the D-particle, only the metric and dilaton are relevant,
hence in Einstein frame we have:


Z
p
1
1
10
2
(5.2)
d x |det(g )| R () .
Sbulk = 2
2
210
Since the twisted sector is provided by a type IIB NS-5 brane hidden in the orbifold fixed
plane, we can derive the fixed plane action from the effective action of the type IIB NS-5
brane in Einstein frame:
3
Z
q
e 2
F F ,
d6 y |det(g )|
(5.3)
Splane = m
4
x i =0

where the tilde denotes the restriction of the bulk fields to the position of the fixed plane:
g (y ) = g (y , x i = 0),

) = (y , x i = 0),
(y

(5.4)

and , = 0, 1, . . . , 5 are the indices along the fixed plane. This particular coupling of the
twisted fields with the bulk metric is obtained by expanding the NS-5 brane kinetic term
in powers of the worldvolume fields. The twisted RR 1-form has been identified with
the U (1) gauge field on the NS-5 brane worldvolume F . Here m is a factor related to the
tension of the NS-5 brane. Moreover, the embedding scalars of the NS-5 brane has not
been included, since they correspond to the twisted NSNS sector, to which the non-BPS
D-particle does not couple. It is straightforward to see that the asymptotic fields (3.10) and

278

E. Eyras, S. Panda / Nuclear Physics B 584 (2000) 251283

(3.11) are a solution to the weak field limit of the equations of motion of the action Stotal .
Note that the non-BPS D-particle is not stable below the critical radius, therefore we do
not expect it would appear as a solution of this action reduced to six dimensions.
Finally, the behaviour of the stable non-BPS D-particle at the critical radius suggests
that it probably saturates a BPS type of bound in the effective theory, without being
supersymmetric. This latter property comes from the fact that the particle couples to
a 1-form in the wrong supersymmetric multiplet. The analysis carried out here can be
extended to other stable non-BPS branes. A study of the classical geometry of other stable
non-BPS D-branes will be presented in a future publication [60].

Acknowledgements
We would like to thank E. Bergshoeff and M. de Roo for useful discussions. E.E. is
greatful to M. Gaberdiel, T. Dasgupta, N. Lambert, A. Liccardo, M.J. Perry, B. Stefanski,
P. Townsend and A. Uranga for useful discussions. E.E. has also enjoyed discussions with
C. van der Bruck, P. Vanhove and N. Wyllard. The work of E.E. is supported by the
European Community program Human Potential under the contract HPMF-CT-199900018. This work is also partially supported by the PPARC grant PPA/G/S/1998/00613.

Appendix A
In this appendix we present some details about the zero-mode part of the twisted RR
boundary state and its GSO projection. In order to describe the twisted RR groundstate
we make use of the 8 8 gamma matrices of SO(1, 5):

(A.1)
, = 2 18 .
We define
= 0 1 2 3 4 5 ,

(A.2)

such that { , } = 0 and ( )2 = 18 . Furthermore, there is a conjugation matrix


C = 3 5 0,

(A.3)

such that
{ , C} = 0,

( )T = C C 1 ,

C 1 = C,

C T = C.

(A.4)

The twisted RR groundstate is characterised by left and right spinor indices of SO(1, 5),
and can be constructed from the NS-vacuum by means of spin and twist fields as follows:
z)|0i = |aiT |bi
fT.
lim S a (z)(z)S b (z)(

z,z0

(A.5)

The zero-mode part of the twisted RR boundary state satisfies the following overlap
equations in the twisted sector:

E. Eyras, S. Panda / Nuclear Physics B 584 (2000) 251283



0 i S 0 |D0 , i(0)
R,T = 0,

279

(A.6)

where = 0, 1, . . . , 5 and S is the restriction of the matrix S of the D-particle to the


6-dimensional orbifold fixed plane: S = diag(+1, 1, 1, 1, 1, 1). We define
f T.
|D0 , iR,T = Mab |aiT |bi
(0)

(A.7)

If we define the action of the fermionic zero-modes in the twisted Ramond sector as
f T = 1 ( )a c (18 )b d |ciT |di
f T,
0 |aiT |bi
2

f T = 1 ( )a c ( )b d |ciT |di
fT,
(A.8)
0 |aiT |bi
2
the matrix M must satisfy:

(A.9)
( )T M i S M = 0.
The solution to this equation is given by:
M = C 0

1 + i
.
1 + i

(A.10)

The zero-mode RR boundary state for a non-BPS D-particle moving in a direction m is


given by
f
|D0 , , vi(0)
R,T = Mab (v)|aiT |biT ,

(A.11)

where the matrix Mab (v) is given by


 1 + i
1
.
C 0 + v m
M(v) =
2
1 + i
1v
On the other hand, defining
(0)
f
R,T hD0 , | = T ha|T hb|Nab

(A.12)

(A.13)

and using
1
f ( )a c (18 )b d ,
= T hc|T hd|
2
f = 1 T hc|T hd|
f ( )a c ( )b d
T ha|T hb|
0
2
we can write the matrix N as follows:
1 + i
.
N = C 0
1 i
f
T ha|T hb|
0

(A.14)

(A.15)

The boosted version can be written similarly to (A.12):


 1 + i
1
.
(A.16)
C 0 + v m
N(v) =
2
1 i
1v
In order to implement the GSO projection on the zero-mode of the boundary state, we
define the zero-mode part of the GSO-projector on the twisted RR sector as:

280

E. Eyras, S. Panda / Nuclear Physics B 584 (2000) 251283

= 8 05 00 .

= 805 00 ,

(A.17)

Using this definition and (A.8), the action of and on the twisted RR groundstate is
found to be:
f T = ( )a c |ciT |bi
fT,
|aiT |bi

f T = ( )b d |aiT |di
fT.
|aiT |bi

(A.18)

Finally, combining (A.10) and (A.18) we find:


(0)

(0)
(0)
|D0 , iR,T = |D0 , iR,T .

(0)

|D0 , iR,T = |D0 , iR,T ,

(A.19)

Appendix B
For the sake of completeness we include in this appendix the explicit definitions of the
different parts of the boundary state for the non-BPS D-particle. We include as well the
conjugate states. Before GSO-projection the NSNS boundary state is given by:
|D0, iNS,U =

T0
|D0X i |D0gh i |D0 , iNS |D0sgh, iNS .
2

The bosonic part is:


|D0X i =

(5)

(q y )
p

(4)

(q ) exp
i

X
1
n=1

The ghost part:


|D0gh i = exp

X
n=1

(B.1)
!

n S n |k = 0i.

(B.2)

!

 c0 + c0
f

cn bn bn cn
|1i|1i.
2

The fermionic part:


|D0 , iNS = i exp i

(B.3)

r S r |0i.

(B.4)

r=1/2

The superghost part:

|D0sgh , iNS = exp i

r r r r

!


g
|1i|1i.

(B.5)

r=1/2

The conjugate states are:


hD0X | = hk = 0|

(5)

(q y )
p

(4)

(q ) exp
i

X
1
n=1

f b0 b0 exp
hD0gh | = h2|h2|

X
n=1

NS hD0 , | = i h0| exp

X
r=1/2

bn cn cn bn
!

r S r ,

!


n
,

!
n S n ,

E. Eyras, S. Panda / Nuclear Physics B 584 (2000) 251283

g exp i
NS hD0sgh , | = h1|h1|

r r r r

!


281

(B.6)

r=1/2

For the twisted RR sector we have (before GSO-projection):


Q0
|D0X iT |D0gh i |D0 , iR,T |D0sgh , iR .
2
The bosonic part is given by
 X

1
(5) p
p
n S n |k = 0i
|D0X iT = (q y ) exp
n
t.m.
|D0, iR,T =

(B.7)

(B.8)

where t.m. indicates that the sum is performed according to the twisted moddings of the
(twisted) RR sector given in (2.1). The ghost part is the same as for the NSNS boundary
state. The fermionic part reads:
 X

(0)
n S n |D0 , iR,T ,
(B.9)
|D0 , iR,T = exp i
t.m.
(0)
|D0 , iR,T

and is given in Appendix A. Finally, the superghost part:


!

X

n n n n |D0sgh , i(0)
|D0sgh , iR = exp i
R ,

where

(B.10)

n=1

with the zero-mode given by

^
|D0sgh , iR = ei0 0 |1/2i|3/2i.
(0)

(B.11)

Finally, the conjugate states are:

 X

1
(5) p
p

hD0
|
=
hk
=
0|
(
q

y
)
exp

T
X
n
n ,
n
t.m.
 X

(0)

hD0
,
|
=

hD0
,
|
exp
i

R,T

n
n ,
R,T

(0)
NS hD0sgh , | = R hD0sgh , | exp

t.m.

n n n n

!


(B.12)

n=1
(0)

where R hD0 , | is given in Appendix A and


(0)
^ i0 0 .
R hD0sgh , | = h3/2|h1/2| e

(B.13)

References
[1] J. Dai, M. Leigh, J. Polchinski, New connections between string theories, Phys. Lett. 4 (1989)
2073.

282

E. Eyras, S. Panda / Nuclear Physics B 584 (2000) 251283

[2] M.B. Green, Pointlike states for type IIB superstrings, Phys. Lett. B 329 (1994) 435, hepth/9403040.
[3] J. Polchinski, Dirichlet branes and RamondRamond charges, Phys. Rev. Lett. 75 (1995) 4724,
hep-th/9510017.
[4] G.T. Horowitz, A. Strominger, Black strings and p-branes, Nucl. Phys. B 360 (1991) 197.
[5] A.W. Peet, The Bekenstein formula and string theory, Class. Quant. Grav. 15 (1998) 3291, hepth/9712253.
[6] M.J. Duff, R.R. Khuri, J.X. Lu, String solitons, Phys. Rept. 259 (1995) 213, hep-th/9412184.
[7] J. Maldacena, A. Strominger, Black hole greybody factors and D-brane spectroscopy, Phys.
Rev. D 55 (1997) 861, hep-th/9609026.
[8] A. Strominger, C. Vafa, Microscopic origin of the BekensteinHawking entropy, Phys. Lett.
B 379 (1996) 99, hep-th/9601029.
[9] C.G. Callan, J.M. Maldacena, D-brane approach to black hole quantum mechanics, Nucl. Phys.
B 472 (1996) 591, hep-th/9602043.
[10] A. Sen, Stable non-BPS states in string theory, JHEP 06 (1998) 007, hep-th/9803194.
[11] A. Sen, Stable non-BPS bound states of BPS D-branes, JHEP 08 (1998) 010, hep-th/9805019.
[12] O. Bergman, M.R. Gaberdiel, Stable non-BPS D-particles, Phys. Lett. B 441 (1998) 133, hepth/9806155.
[13] A. Sen, SO(32) spinors of type I and other solitons on braneantibrane pair, JHEP 09 (1998)
023, hep-th/9808141.
[14] A. Sen, Type I D-particle and its interactions, JHEP 10 (1998) 021, hep-th/9809111.
[15] E. Witten, D-branes and K-theory, JHEP 9812 (1998) 019, hep-th/9810188.
[16] A. Sen, BPS D-branes on non-supersymmetric cycles, JHEP 12 (1998) 021, hep-th/9812031.
[17] A. Sen, Non-BPS states and branes in string theory, hep-th/9904207.
[18] A. Lerda, R. Russo, Stable non-BPS states in string theory: a pedagogical review, hepth/9905006.
[19] M. Gaberdiel, Lectures on non-BPS Dirichlet branes presented at the school of the TMRNetwork Quantum Aspects of Gauge Theories, Supersymmetry and Gravity, Torino, 2000.
[20] J.A. Harvey, P. Horava, P. Kraus, D-sphalerons and the topology of string configuration space,
hep-th/0001143.
[21] M.S. Volkov, The tale of gravitational sphaleron, hep-th/0002006.
[22] G. Lifschytz, Comparing D-branes to black-branes, Phys. Lett. B 388 (1996) 720, hepth/9604156.
[23] A.A. Tseytlin, No-force condition and BPS combinations of p-branes in 11 and 10 dimensions,
Nucl. Phys. B 487 (1997) 141, hep-th/9609212.
[24] O. Bergman, M.R. Gaberdiel, Non-BPS states in heterotic type IIA duality, JHEP 03 (1999)
013, hep-th/9901014.
[25] O. Bergman, M.R. Gaberdiel, A non-supersymmetric open string theory and S-duality, Nucl.
Phys. B 499 (1997) 183, hep-th/9701137.
[26] M. Frau, L. Gallot, A. Lerda, P. Strigazzi, Stable non-BPS D-branes in type I string theory,
Nucl. Phys. B 564 (2000) 60, hep-th/9903123.
[27] O. Bergman, M.R. Gaberdiel, Dualities of type 0 strings, JHEP 07 (1999) 022, hep-th/9906055.
[28] Y. Michishita, D0-branes in SO(32) SO(32) open type 0 string theory, Phys. Lett. B 466
(1999) 161, hep-th/9907094.
[29] E. Eyras, Stable D8-branes and tachyon condensation in type 0 open string theory, JHEP 10
(1999) 005, hep-th/9908121.
[30] M.R. Gaberdiel, B. Stefanski Jr., Dirichlet branes on orbifolds, hep-th/9910109.
[31] A. Sagnotti, Some properties of open-string theories, hep-th/9509080.
[32] C. Angelantonj, Non-tachyonic open descendants of the 0B string theory, Phys. Lett. B 444
(1998) 309, hep-th/9810214.

E. Eyras, S. Panda / Nuclear Physics B 584 (2000) 251283

283

[33] M.R. Gaberdiel, A. Sen, Non-supersymmetric D-brane configurations with BoseFermi


degenerate open string spectrum, JHEP 11 (1999) 008, hep-th/9908060.
[34] M. Bertolini, P. Fre, R. Iengo, C.A. Scrucca, Black holes as D3-branes on CalabiYau
threefolds, Phys. Lett. B 431 (1998) 22, hep-th/980309.
[35] I.R. Klebanov, N.A. Nekrasov, Gravity duals of fractional branes and logarithmic RG flow,
hep-th/9911096.
[36] I.R. Klebanov, A.A. Tseytlin, Gravity duals of supersymmetric SU(N) SU(N + M) gauge
theories, hep-th/0002159.
[37] P. Di Vecchia, A. Liccardo, D-branes in string theory, hep-th/9912161, hep-th/9912275.
[38] C. Callan, C. Lovelace, C.R. Nappi, S.A. Yost, String loop corrections to beta functions, Nucl.
Phys. B 288 (1987) 525.
[39] C. Callan, C. Lovelace, C.R. Nappi, S.A. Yost, Adding holes and crosscaps to the superstring,
Nucl. Phys. B 293 (1987) 83.
[40] C. Callan, C. Lovelace, C.R. Nappi, S.A. Yost, Loop corrections to superstring equations of
motion, Nucl. Phys. B 308 (1988) 221.
[41] J. Polchinski, Y. Cai, Consistency of open superstring theories, Nucl. Phys. B 296 (1988) 91.
[42] M.B. Green, P. Wai, The insertion of boundaries in worldsheets, Nucl. Phys. B 431 (1994) 131.
[43] M.R. Garousi, R.C. Myers, Superstring scattering from D-branes, Nucl. Phys. B 475 (1996)
193, hep-th/9603194.
[44] P. Di Vecchia, M.L. Frau, A. Lerda, I. Pesando, R. Russo, S. Sciuto, Classical p-branes from the
boundary state, Nucl. Phys. B 507 (1997) 259, hep-th/9707068.
[45] P. Di Vecchia, M.L. Frau, A. Lerda, A. Liccardo, (F, Dp) bound states from the boundary state,
hep-th/9906214.
[46] N. Lambert, I. Sachs, Non-abelian field theory of stable non-BPS branes, hep-th/0002061.
[47] M.R. Douglas, D. Kabat, P. Pouliot, S.H. Shenker, D-branes and short distances in string theory,
Nucl. Phys. B 485 (1997) 85, hep-th/9608024.
[48] D. Kutasov, Orbifolds and solitons, Phys. Lett. B 383 (1996) 48, hep-th/9512145.
[49] J. Majumder, A. Sen, Blowing up D-branes on non-supersymmetric cycles, JHEP 09 (1999)
004, hep-th/9906109.
[50] M. Bill, P. Di Vecchia, M.L. Frau, A. Lerda, I. Pesando, R. Russo, S. Sciuto, Microscopic
string analysis of the D0D8 brane system and dual RR states, Nucl. Phys. B 526 (1998) 199,
hep-th/9802088.
[51] L. Dixon, D. Friedan, E. Martinec, S. Shenker, The conformal field theory of orbifolds, Nucl.
Phys. B 282 (1987) 13.
[52] A. Giveon, E. Rabinovici, G. Veneziano, Duality in string background space, Nucl. Phys. B 322
(1989) 167.
[53] M. Frau, I. Pesando, S. Sciuto, A. Lerda, R. Russo, Scattering of closed strings from many
D-branes, Phys. Lett. B 400 (1997) 52, hep-th/9702037.
[54] M. Bill, D. Cangemi, P. Di Vecchia, Boundary states for moving D-branes, Phys. Lett. B 400
(1997) 63, hep-th/9701190.
[55] L. Gallot, A. Lerda, P. Strigazzi, Gauge and gravitational interactions of non-BPS D-particles,
hep-th/0001049.
[56] C. Bachas, D-brane dynamics, Phys. Lett. B 374 (1996) 37, hep-th/9511043.
[57] M. Bianchi, G. Pradisi, A. Sagnotti, Toroidal compactification and symmetry breaking in open
string theories, Nucl. Phys. B 376 (1992) 365.
[58] A. Sen, Supersymmetric world-volume action for non-BPS D-branes, JHEP 10 (1999) 008,
hep-th/9909062.
[59] R. Russo, C.A. Scrucca, On the effective action of stable non-BPS branes, hep-th/9912090.
[60] E. Eyras, S. Panda, B. Stefanski, in preparation.

Nuclear Physics B 584 (2000) 284299


www.elsevier.nl/locate/npe

Tachyon couplings on non-BPS D-branes


and DiracBornInfeld action
Mohammad R. Garousi
Department of Physics, University of Birjand, Birjand, Iran
Institute for Studies in Theoretical Physics and Mathematics IPM, P.O. Box 19395-5746, Tehran, Iran
Received 29 March 2000; revised 29 May 2000; accepted 30 May 2000

Abstract
By explicit evaluation of certain disk S-matrix elements in the presence of background B-flux,
we find coupling of two open string tachyons to gauge field, graviton, dilaton or KalbRamond
antisymmetric tensor on the worldvolume of a single non-BPS Dp-brane. We then propose an
extension of the abelian DiracBornInfeld action which naturally reproduces these couplings in
field theory. This action includes non-linearly the dynamics of the tachyon field much like the
other bosonic modes of the non-BPS Dp-brane. On the general grounds of gauge and T-duality
transformations and the symmetrized trace prescription, we then extend the abelian action to nonabelian cases. 2000 Elsevier Science B.V. All rights reserved.
PACS: 11.25.-w

1. Introduction
Recent years have seem dramatic progress in the understanding of non-perturbative
aspects of string theory [15]. With these studies has come the realization that solitonic
extended objects, other than just strings, play an essential role. An important object in
these investigations has been Dirichlet branes [68]. D-branes are non-perturbative states
on which open string can live, and to which various closed strings including Ramond
Ramond states can couple.
Type II string theories have two kind of D-branes, BPS Dp-branes for p even(odd) [68]
and non-BPS Dp-branes for p odd(even) [913] in IIA(IIB) theory. The BPS branes are
stable solitons which break half of the spacetime supersymmetries and their dynamics
are properly described in field theory by DiracBornInfeld(DBI) action [1416] (see also
[17]) and ChernSimons action [1820]. The non-BPS Dp-branes on the other hand suffer
biruni@iran.com

0550-3213/00/$ see front matter 2000 Elsevier Science B.V. All rights reserved.
PII: S 0 5 5 0 - 3 2 1 3 ( 0 0 ) 0 0 3 6 1 - 8

M.R. Garousi / Nuclear Physics B 584 (2000) 284299

285

from open string tachyonic mode whose mass causes the brane in a flat background to be
unstable. However, there are other terms in the tachyon potential which makes it bounded
from below. Consequently, the non-BPS branes decay to minimum of the tachyon potential.
It has been conjectured that at the stationary point of the tachyon potential, the negative
minimum energy of the tachyon potential plus the positive energy of the brane tension
is exactly zero [2125]. Hence, the unstable non-BPS branes in flat spacetime vacuum
should decay to the true vacuum of the theory in which there is no branes. This conjecture
was studied in [2633] by explicit calculation of the tachyon potential using the string field
theory framework.
The dynamics of massless bosonic excitations of non-BPS Dp-branes are suitablely
described by the DBI action in field theory. This action has been generalized to the
supersymmetric form to include the dynamics of massless fermionic modes of the branes
as well [34]. The RR fields of the type II theory have no coupling to the non-BPS D-branes
through the usual ChernSimons action. However, there is a non-vanishing coupling
between the RR field and tachyon on the worldvolume of the branes [35]. The Chern
Simons action hence was modified in [36] to incorporate this coupling. In the present
paper, on the other hand, we are interested in generalizing the DBI action to take into
account the dynamics of the tachyon field. We study this by explicit evaluation of some
nontrivial disk S-matrix elements in the first quantized string theory. From these matrix
elements we conjecture an extension for the DBI action which includes the tachyon field
as well.
An outline of the paper is as follows. We begin in Section 2 by expressing our conjecture
for extension of the DBI action which includes dynamics of the tachyon field. Then we
expand this action around a background B-flux to produce various couplings involving two
tachyons and one gauge field, graviton, dilaton or KalbRamond antisymmetric tensor. In
Section 2.1 we transform the above couplings between commutative fields to their noncommutative counterparts. We do this because the disk S-matrix elements in the presence
of the background B-flux with which we are going to compare our conjectured action
in the subsequent section are corresponding to non-commutative open string fields [37]. In
Section 3, we evaluate the S-matrix elements and check their consistency with the proposed
field theory couplings. In Section 4 we extend our proposed action for describing dynamics
of a single non-BPS Dp-brane to the case of the non-abelian theory of N coincident
branes using the general grounds of the symmetrized trace, and non-abelian gauge and
invariance under T-duality transformations. Appendix contains our conventions and some
useful comments on conformal field theory propagators and vertex operators used in our
calculations.

2. Abelian action
The worldvolume theory of a single non-BPS D-brane in type II theory includes a
massless U(1) vector Aa , a set of massless scalars Xi , describing the transverse oscillations
of the brane, a tachyonic state T and their fermionic partners (see, e.g., [35]). The leading

286

M.R. Garousi / Nuclear Physics B 584 (2000) 284299

order low-energy action for the massless fields corresponds to a dimensional reduction of
a ten-dimensional U(1) YangMills theory. As usual in string theory, there are higher order
0 = `2s corrections, where `s is the string length scale. As long as derivatives of the field
strengths (and second derivatives of the scalars) are small compared to `s , then the action
takes a DiracBornInfeld form [1416]. To take into account the couplings of the massless
open string states with closed strings, the DBI action may be extended naturally to include
massless NeveuSchwarz closed string fields, i.e., the metric, dilaton and KalbRamond
filed. In this case one arrives at the following worldvolume action:
Z
p
S = Tp d p+1 e det(P [Gab + Bab ] + 2 0 Fab .
Here, Fab is the abelian field strength of the worldvolume ordinary gauge field, while the
metric and antisymmetric tensors are the pull-backs of the bulk tensors to the D-brane
worldvolume, e.g.,
X X
a b
= Gab + 2Gi(a b) Xi + Gij a Xi b Xj ,

P [Gab ] = G

(1)

where in the second line above we have used that fact that we are employing static
gauge throughout the paper, i.e., a = Xa for worldvolume and Xi ( a ) for transverse
coordinates.
In order to extend this action to incorporate dynamics of the tachyonic mode as well,
we shall evaluate some disk S-matrix elements in string theory and read from them various
couplings involving the tachyon field. Our results indicate that the tachyon should appear
in the following extension of DBI action:
Z
S = Tp d p+1 e V (T )
p
(2)
det(P [Gab + Bab ] + 2 0 Fab + 2 0 a T b T ),
where the tachyon potential is V (T ) = 1 + 2 0 m2 T 2 /2 + O(T 4 ) and the tachyon mass
is m2 = 1/2 0 . In our conventions the tachyon field is dimensionless. The conjecture
in [2125] is that the tachyon potential is zero at the minimum of the potential, i.e.,
V (T0 ) = 0. Hence, upon tachyon condensation at this point the abelian action (2) of the
non-BPS brane becomes zero.
Note that appearance of the tachyon kinetic term and potential in (2) is similar in form to
the kinetic term and potential of the transverse scalar fields in the non-abelian DBI action
of N coincident BPS D-branes [38]. In this case though the kinetic term appears in the
pull-back of the metric under the square root and the potential for scalar fields multiplies
the square root in the DBI action(see Eq. (22) for T = 0).
We now continue backward, assuming the above action (2) and check its consistency
with some S-matrix elements. To have nontrivial check, we shall evaluate disk amplitudes
describing decay of two tachyons to dilaton, graviton or KalbRamond antisymmetric
tensor on the worldvolume of a single non-BPS Dp-brane with background B-flux.
The amplitude describing the worldvolume coupling of two tachyons to gauge field will

M.R. Garousi / Nuclear Physics B 584 (2000) 284299

287

be evaluated as well. Therefore, we begin by expanding (2) for fluctuations around the
background G = , B = F ab a b , = 0 and T = 0 to extract interactions
expected from the proposed action (2). The fluctuations should be normalized as the
conventional field theory modes which appear in the string vertex operators. As a first
step, we recall that the graviton vertex operator corresponds to string frame metric. Hence,
one should transform the Einstein frame metric G to the string frame metric g via
G = e/2 g . Now with our conventions for string vertex operators (see appendix), the
string mode fluctuations take the form
g = + 2h ,

= 2 ,
B = F ab a b 2b ,
1
,
T=p
2 0 Tp
Aa =

1
p aa ,
2 0 Tp

1
Xi = p i .
Tp
With these normalizations, the pull back of the Einstein frame metric becomes:




1

1 + a i b i + ,
P [Gab ] = ab 1 + + 2P [hab ] +
Tp
2
2
where the dots represents terms with two and more closed string fields.
Now it is straightforward, to expand Eq. (2) using

p
p

1
det(M0 + M) = det(M0 ) 1 + Tr M01 M
2


1
1
1
1 
1 2
+
Tr M0 MM0 M + Tr M0 M
4
8

to produce a vast array of interactions. We are interested in the interactions quadratic


in tachyon, and linear in massless open and closed string fluctuations. The appropriate
Lagrangian are:


1 2 2 1
ab
L2,0 = c m + (VS ) a b ,
2
2

c
1
1 2
m (VA )ab fba 2 + V ab fba V cd c d
L3,0 = p
2
2 Tp 2

V ab fbc V cd d a = 0,

288

M.R. Garousi / Nuclear Physics B 584 (2000) 284299

L2,1 = c




1
V ab (hba bba ) + Tr(V ) 4
2 2


1
1
(VS )ab a b + m2 2

2
2



1
V ab hbc bbc + bc V cd d a ,
2 2

(3)

where the first and the second subscript in above Lagrangians refers to the number of open
and closed string fields, respectively. Here we have dropped some total derivative terms in
L3,0 . In the above Lagrangians, fab = a ab b aa and
p
ab
V ab ( + F )1 ,
(4)
c det(ab + Fab ),
and VS (VA ) is symmetric(antisymmetric) part of the V matrix above. It is important to
note that the antisymmetric matrix VA appears in total derivative terms in the Lagrangian
L3,0 which involves only open string fields.
In the case that the background B-flux is zero, the coupling of KalbRamond field
to tachyon in the second line of L2,1 vanishes, and the graviton and dilaton couplings
reduce to the natural coupling of these fields to the kinetic term of the tachyon field, i.e.,
e Gab a T b T . In that way, there is no nontrivial coupling that confirms the conjectured
action (2) is valid or not. So we continue our discussion for non-vanishing background
B-flux.
The open string fields appearing in the DBI action (2) or (3) are ordinary commutative
fields. Whereas, open string vertex operators in string theory with background B-flux
correspond to non-commutative fields [37]. Hence, in order to compare the couplings in
(3) with corresponding S-matrix elements, one has to transform the commutative fields in
(3) to their non-commutative variables.
2.1. Change of variables
In [37] differential equation for transforming commutative gauge field to its noncommutative counterpart was found to be

1
A a ( ) = cd A c Fad + Fad A c A c d A a d A a A c ,
4

1
Fab ( ) = cd 2Fac Fbd + 2Fbd Fac A c D d Fab + d Fab
4


D d Fab + d Fab A c ,
where the gauge field strength and product were defined to be
Fab = a A b b A a i A a A b + i A b A a
= a A b b A a i[A a , A b ]M ,

i
a b
y=z=x .
f(x) g(x)

= e 2 ab y z f(y)g(z)

(5)

M.R. Garousi / Nuclear Physics B 584 (2000) 284299

289

These differential equations can be integrated perturbatively to find a relation between


ordinary fields appearing in (3) and non-commutative fields corresponding to open string
vertex operators. The result for abelian case that we are interested in is [39]:

1
Aa = A a cd A c 0 Fad A c 0 d A a + O(A 3 ),
2

Fab = Fab cd Fac 0 Fbd A c 0 d Fab + O(A 3 ),

(6)

where the commutative multiplication 0 operates as




sin( 12 ab ya zb )
0

f (y)g(z)

= 1
.
f (x) g(x)
a
b
y=z=x
2 ab y z
In [39] we verified by explicit calculation of S-matrix elements of one massless closed
and two open string states that the transformation (6) is exactly reproduced by perturbative
string theory. Appropriate transformations for scalar fields such as the tachyon can be read
from (6), i.e.,
T = T + cd A c 0 d T + ,



a T = a T i A a , T M + cd Fca 0 b T + A c 0 a d T + ,

(7)

where dots represent terms which involve more than two open string fields. They produce
couplings between more than three fields upon replacing them into (3) in which we are not
interested.
The differential equation (5) expresses infinitesimal variation of linear field, e.g.,

Aa , in terms of infinitesimal variation of the non-commutative parameter, i.e., cd .


Upon integration (6), this transforms the linear commutative gauge field in terms
of nonlinear combination of non-commutative fields. However, we are interested in
transforming quadratic combinations of commutative fields appearing in (3) in terms
of non-commutative fields. Such a transformation, in principle, might be found from a
differential equation alike (5) that expresses infinitesimal variation of the quadratic fields
in terms of infinitesimal variation of non-commutative parameter. Upon integration, that
would produced the desired transformation. In that way, one would find that not only the
fields transform as in (7) but the multiplication rule between fields also undergo appropriate
transformation. We are not going to find such a differential equations here. Instead, we
simply note that the transformation for multiplication rule between two open string fields
can be conjectured from the right hand side of Eq. (6) to be
fg|=0 f 0 g|6=0

(8)

for f and g being any arbitrary open string fields. This transformation rule was confirmed
in [39] by explicit evaluation of S-matrix elements of one massless closed and two open
string states.
Now with the help of Eqs. (7) and (8), one can transform the commutative Lagrangian (3)
to non-commutative counterparts. In doing so, one should first using (8) replace ordinary
multiplication of two tachyons by the 0 multiplication. Then, using (7) the ordinary

290

M.R. Garousi / Nuclear Physics B 584 (2000) 284299

tachyon fields should be transformed to their non-commutative counterparts. The results


are


1
1
ic
p (VS )ab a [a b , ]M + , (9)
L 2,0 = c m2 + (VS )ab a b +
2
2
4 Tp



1
ab

L2,1 = c V (hba bba ) + Tr(V ) 4


2 2


1
1 2 0
ab
0
(VS ) a b + m

2
2




bc
ab
cd
0
(10)
hbc bbc + V d a + ,
V
2 2
where ellipsis represent terms which have more than three fields. Here we have dropped
some total derivative terms which appeared in L2,0 and also replaced 0 between two noncommutative tachyons in (9) with ordinary multiplication rule because the deference is
some total derivative terms. In the Eq. (10) on the other hand, the difference between 0
and ordinary multiplication rules is not just a total derivative terms. Note that upon inserting
the transformation (7) into (3), the antisymmetric matrix (VA )ab appears in Eq. (9) only in
the product terms. 1
It is interesting to note that the symmetric part of the ( + F )1 matrix, i.e., VS , appears
as the metric in (9) and the antisymmetric part appears in the definition of product in the
Moyal bracket. This is consistent with the conclusion reached in [37]. In our discussion,
however, the symmetric part VS appears naturally as a result of expanding the ordinary
DBI action around the background B-flux, and the antisymmetric part VA appears as a
result of transforming commutative fields to non-commutative variables. In the Lagrangian
(10), on the other hand, which involves open and closed string fields, both symmetric and
antisymmetric matrices appear in its different coupling terms.

3. Scattering calculations
The couplings in (9) and (10) should be reproduced by disk S-matrix elements of string
theory if the proposed action (2) is going to be valid. In this section, we work at the string
theory side and evaluate these couplings using the conformal field theory technique. We
begin with the evaluation of the coupling of two tachyons to gauge or scalar fields.
3.1. Openopenopen couplings
In the worldsheet conformal field theory framework, the coupling of two tachyons to a
gauge or scalar field is described properly by the correlation of their corresponding vertex
operators inserted at the boundary of the disk worldsheet, that is
1 Note that our conventions set ab = 4(V )ab .
A

M.R. Garousi / Nuclear Physics B 584 (2000) 284299

Z
A (3 G)

291



dx1 dx2 dx3 : V1 2k1 V T , x1 : :V1 2k2 V T , x2 :


:V0 2k3 V T , x3 : ,

where the vertex operators are given in the appendix. Using the worldsheet conformal
field theory technique, it is not difficult to perform the correlators above and show that the
integrand is invariant under SL(2, R). Gauging this symmetry by fixing the positions of the
vertices at arbitrary points, one finds A(1 , 2 , 3 ) = 0 and
A(1 , 2 , a3 ) =

c sin(l)
p
(k1 VS 3 ),
Tp

(11)

where we have defined l 2k1 V T F V k2 = 2k1 VA k2 . We have also normalized the


p
amplitude by the appropriate coupling factor c/2 Tp . The sin(l) factor above arises
basically from two different phase factors corresponding to two distinct cyclic orderings
of the vertex operators. Each phase factor stems from the second term of the worldsheet
propagator (29). Using the fact that our conventions set ab = 4VAab , it is not difficult to
verify that the S-matrix element (11) is exactly reproduced by the second term in (9). At
the same time, vanishing of A(1 , 2 , 3 ) is consistent with (9).
3.2. Closedopenopen amplitudes
The amplitudes describing decay of two open string tachyons to one massless closed
string NSNS mode is given by the following correlation:
Z



A (3 D) dx1 dx2 d2 z :V0 2k1 V T , x1 : :V0 2k2 V T , x2 :

(p3 D, z3 ): ,
:V1 (p3 , z3 ): :V1
where the closed string vertex operator inserted at the middle and open string vertex
operators at the boundary of the disk worldsheet. Explicit form of the vertex operators
in terms of worldsheet fields are given in the Appendix. Here again using appropriate
worldsheet propagators, one can evaluate the correlations above and show that the
integrand is SL(2, R) invariant. We refer the reader to Refs. [4042] for the details of
the calculations. Gauging the SL(2, R) symmetry by fixing z3 = i and x1 = , one arrives
at

Z
8ik2 V T 3 DV k1
A 22s2 dx (2s + 1) Tr(3 D)
x i

T
8ik1 V 3 DV k2
+
x+i
(x i)sl (x + i)s+l ,
where the integral is taken from to +, and s = p3 VS p3 = 1/2 2k1 VS k2 .
This integral is doable and the result is

292

M.R. Garousi / Nuclear Physics B 584 (2000) 284299


ic
(2s)
a1 (s + l) a2 (s l)
,
(12)
2
(1 s l) (1 s + l)
where a1 and a2 are two kinematic factors depending only on the spacetime momenta and
the closed string polarization tensor
A=

a1 = 4k2 V T 3 DV k1 ,
a2 = (s + l) Tr(3 D) + 4k1 V T 3 DV k2 .
We have also normalized the amplitude (12) at this point by the coupling factor ic/2.
As a check of our calculations, we have inserted the dilaton polarization (31) into (12)
and found that the result is independent of the auxiliary vector ` . The amplitude (12) has
the pole structure at m2open = n/ 0 . 2 This does not have tachyon pole which is consistent
with the fact that coupling of three tachyons is zero. In fact due to the worldsheet fermions
in the tachyon vertex operator, coupling of any odd number of tachyons is zero. Hence,
the worldvolume of the non-BPS Dp-branes has a Z2 symmetry under which the tachyon
changes sign.
3.2.1. Massless poles
Given the general form of the string amplitude in Eq. (12), one can expand this amplitude
as an infinite sum of terms reflecting the infinite tower of open string states that propagate
on the worldvolume of D-brane. In the domain where s 0, the first term of the expansion
representing the exchange of massless string states dominate. In this case the scattering
amplitude (12) reduces to
ic sin(l)
(a1 + a2 ) + ,
(13)
4s
where dots represent contact terms and the infinite series of massive poles. Making the
appropriate explicit choices of polarizations, we find



ic sin(l) l
Tr(D) + 2 4k1 V V k2 + 1 2,
As (1 , 2 , 3 ) =

2
8 2s


ic sin(l) l
Tr(3 D) 4k1 V 3T V k2 + 1 2.
(14)
As (1 , 2 , h3 ) =
4s
2
Here h3 stands for both graviton and KalbRamond antisymmetric tensors. In writing
explicitly the above massless poles, one finds some terms which are proportional to s as
well. We will add these terms which have no contribution to the massless poles of field
theory to the contact terms in (16). The amplitudes (14) should be reproduced in s-channel
of field theory. They can be evaluated in field theory as


b
e3 a a G
ea
e
,
V
A0s (1 , 2 , 3 ) = V
ab a1 2



b
eh3 a a G
ea
e
,
(15)
A0s (1 , 2 , h3 ) = V
V
ab a1 2
A=

where the propagator and the vertices can be read from expansion of (2) in terms of noncommutative fields. They are
2 We explicitly restore 0 here. Otherwise our conventions set 0 = 2.

M.R. Garousi / Nuclear Physics B 584 (2000) 284299

293

i (VS1 )ab
,
c
s
p



a
Tp c 1
a
a
a
e
Tr(D) + 2 p3 VA p3 V V + V V p3 ,
V3 a =
2
2 2


p

eh3 a a = Tp c 1 Tr(3 D)p3 VAa p3 V 3T V a + V a 3T V p3 ,
V
2

ea1 2 a = c sin(l)
p k1 VSa + 1 2.
V
2 Tp
ea
G

ab

In writing the above propagator, we have used the covariant gauge VSab a A b = 0.
Replacing above propagator and vertices into (15), one finds exactly the string massless
poles (14).
3.2.2. Contact terms
Having examined in detail the massless poles of string amplitude, we now extract the
low energy contact terms of the string amplitude (12). Expanding the gamma function
appearing in this amplitude for s 0, one will find




a1 + a2 sin(l)
a1 a2 sin(l)
ic
+
A=
2
2
s
2
l
!

sin(l) X
(2n + 1)l (2n+1) + k 2 O(s, l) .
+ (a1 + a2 )
l
n=1

The factor sin(l)/(l) appears for all the contact terms which is consistent with the
transformation of multiplication rule in (8). The second term of the first line above is the
contact term with minimum number of momentum in which we are interested, that is,
ic sin(l)
(a1 a2 ).
(16)
4l
Inserting appropriate polarization (see appendix) and adding the residue contact terms of
the massless poles (13), one finds



ic sin(l)
s

Tr(D) + 2 4k1 V V k2 + 1 2,
(17)
Ac (1 , 2 , 3 ) =
2
8 2l


ic sin(l)
s
Tr(3 D) 4k1 V 3T V k2 + 1 2,
Ac (1 , 2 , h3 ) =
4l
2
Ac

where again h3 stands for both graviton and KalbRamond antisymmetric tensors. These
contact terms are reproduced exactly by the Lagrangian in (10). The first terms in (17) by
the terms in the first line of (10) and the second terms in (17) by the terms in the second
line of (10). Note that, while the first terms in (17) can be reproduced in field theory by
an action in which the tachyon kinetic term appears linearly like the one proposed in [34],
the second terms in (17) can be reproduced only if the tachyon kinetic term appears nonlinearly in the determinant under the square root in the DBI action, i.e., Eq. (2). This ends
our illustration of consistency between disk S-matrix elements and the proposed action (2).

294

M.R. Garousi / Nuclear Physics B 584 (2000) 284299

4. Non-abelian action
In this section we extend the proposed action (2) for a non-BPS Dp-brane to the case of
N coincident non-BPS Dp-brane where the worldvolume theory involves a U(N ) gauge
theory. Our guiding principle in constructing such a non-abelian action is that the action
should be consistent with the familiar rules of T-duality. This guideline has been recently
employed by Myers [38] to construct non-abelian DBI action. In this way, one should start
with non-abelian action for D9-branes and then use some sort of T-duality transformations
to convert the D9-brane action to non-abelian action for Dp-branes. Therefore, we begin
by extending the abelian action (2) to non-abelian action for non-BPS D9-branes. In this
case there is no scalar field corresponding to the transverse direction of D9-branes. Hence,
the non-abelian action may be constructed from the corresponding abelian case by simply
extending the derivative of open string fields to its non-abelian covariant derivative [43,44],
and a trace over the U(N ) representations [45] (see also [17]), that is,
Z
S = T9 d 10
q


(18)
Tr e V (T ) det(G + B + 2 0 F + 2 0 D T D T ) ,
where G , B and F are the metric, antisymmetric tensor and the non-abelian gauge
field strength, respectively, and
D T =

T
i[A , T ].

This action is still incomplete without a precise prescription for how the gauge trace should
be implemented. We expect that a prescription similar to that given for BornInfeld action
[45] should also be given here. That is, the gauge trace should be completely symmetric
between all non-abelian expression of the form F , D T and individual T appearing in
the tachyon potential V (T ).
Now we generalize (18) to the action appropriate for non-BPS Dp-branes for any p.
To this end, we apply familiar T-duality transformations rules to the non-abelian D9brane action (18). T-duality transformations in i = p + 1, . . . , 9 directions of the D9-brane
worldvolume converts the D9-brane to Dp-brane, the gauge fields in those direction to
e = T . The new scalar fields
A i = Xi /2 0 and leaves unchanged the tachyon, i.e., T
i
X are now transverse coordinates of the new Dp-brane. Under this transformation, the
covariant derivative of tachyon becomes
ei Te = T i [Xi , T ].
D
i
2 0
Using the fact that we are always working in the static gauge, the first term on the right
hand side becomes zero because of the assumption implicit in the T-duality transformations
that all fields must be independent of the coordinates i . Now adding this transformation
to the T-duality transformation of massless fields (see, e.g., [38]), one has complete list of
T-duality transformations for the fields appearing in (18);

M.R. Garousi / Nuclear Physics B 584 (2000) 284299

eab = Eab Eai E ij Ej b ,


E

295

eai = Eaj E j i ,
E

eij = E ij ,
E

Eia = E ij Ej a ,

ei Te = i [Xi , T ],
D
e2 = e2 det(E ij ),
2 0
eai = 1 Da Xi ,
eab = Fab ,
F
F
2 0
i
eia = 1 Da Xi ,
eij =
[Xi , Xj ], F
F
0
2
(2 )
2 0

(19)

where we have defined E G + B . Here E ij denotes the inverse of Eij , i.e.,


E ik Ekj = i j . Under above T-duality transformations the determinant in (18) becomes
!
eaj + Da Xj iDa T [Xj , T ]
eab + 2 0 Fab + 2 0 Da T Db T
E
E
e = det
.
D
eib Db Xi i[Xi , T ]Db T
E

eij i 0 [Xi , Xj ] 1 0 [Xi , T ][Xj , T ]


E
2
2

Manipulating the matrix inside the determinant, one finds






e = det P Eab + Eai (Q1 )ij Ej b + 2 0 Fab + Tab det Qi j det(E ij ), (20)
D
where now the definition of the pull-back above is the extension of (1) in which ordinary
derivative is replaced by its non-abelian covariant derivative. Here the matrices Qi j and
Tab are defined to be
i
1
[Xi , Xk ]Ekj
[Xi , T ][Xk , T ]Ekj ,
2 0
2 0
Tab = 2 0 Da T Db T Da T [Xi , T ](Q1 )ij [Xj , T ]Db T

Qi j = i j

1 i

iDa X

(21)

) j [X , T ]Db T iDa T [X , T ](Q )i Ej b


1
(Q )ij [Xj , T ]Db T iDa T [Xi , T ](Q1 )ij Db Xj .

iEai (Q

In Eqs. (20) and (21), indices are raised and lowered by E ij and Eij , respectively. Now
replacing (20) into T-dual of (18) and using the transformation for dilaton field (19), one
finds the final T-dual action
Z

e
S = Tp d p+1 Tr e V (T )V 0 (T , Xi )
q

(22)
det(P [Eab + Eai (Q1 )ij Ej b ] + 2 0 Fab + Tab ) ,
p
where we have defined V 0 (T , Xi ) = det(Qi j ). This potential term is one for abelian case.
If the tachyon field is set to zero, this action would get to the result of non-abelian action
for N coincident BPS Dp-branes [38]. In this case, the prescription for the gauge trace
is studied in [38]. The trace is completely symmetric between Fab , Da Xi , i[Xi , Xj ] and
individual Xi . The latter non-abelian field stems from non-abelian Taylor expansion of the
closed string fields that appear in the DBI action [46]. Natural extension of this prescription
for the trace in the action (22) is that the trace should be completely symmetric between
all non-abelian expressions of the form Fab , Da Xi , i[Xi , Xj ], Xi , Da T , i[Xi , T ] and
individual T of the tachyon potential.

296

M.R. Garousi / Nuclear Physics B 584 (2000) 284299

Acknowledgements
I would like to acknowledge useful conversation with R.C. Myers. This work was
supported by University of Birjand and IPM.

Appendix A. Perturbative string theory with background field


In perturbative superstring theories, to study scattering amplitude of some external
string states in conformal field theory frame, one usually evaluate correlation function
of their corresponding vertex operators with use of some standard conformal field
theory propagators [47,48]. In trivial flat background one uses an appropriate linear
-model to derive the propagators and define the vertex operators. In nontrivial D-brane
background the vertex operator remain unchanged while the standard propagators need
some modification. Alternatively, one may use a doubling trick to convert the propagators
to standard form and give the modification to the vertex operators [40]. In this appendix
we would like to consider a D-brane with constant gauge field strength/or antisymmetric
KalbRamond field in all directions of the D-brane. The modifications arising from the
appropriate linear -model appear in the following boundary conditions [49]: 3
y Xa iF a b x Xb = 0,
X = 0,
i

for a, b = 0, 1, . . . , p,
for i = p + 1, . . . , 9,

(23)

where Fab are the constant background fields, and these equations are imposed at y = 0.
The worldvolume (orthogonal subspace) indices are raised and lowered by ab (N ij ) and
ab (Nij ), respectively. Now we have to understand the modification of the conformal field
theory propagators arising from these mixed boundary conditions. To this end consider the
following general expression for propagator of X (z, z ) fields:


= log(z w) log(z w)

X (z, z ) X (w, w)
D (z w),
D log(z w)

(24)

where D is a constant matrix. To find this matrix, we impose the boundary condition
(23) on the propagator (24), which yields
ab D ba F ab F a c D bc = 0

(25)

=
for the orthogonal directions, and
for the worldvolume directions,
otherwise. Now Eq. (25) can be solved for D ab , that is
D ij

Dab = 2( F )(1)
ab ab
= 2Vba ab ,

N ij

D ia

=0
(26)
(27)

3 Our notation and conventions follow those established in [40]. So we are working on the upper-half plane with
boundary at y = 0 which means y is normal derivative and x is tangent derivative. And our index conventions
are that lowercase greek indices take values in the entire ten-dimensional spacetime, e.g., , = 0, 1, . . . , 9;
early latin indices take values in the worldvolume, e.g., a, b, c = 0, 1, . . . , p; and middle latin indices take values
in the transverse space, e.g., i, j = p + 1, . . . , 8, 9. Finally, our conventions set `2s = 0 = 2.

M.R. Garousi / Nuclear Physics B 584 (2000) 284299

297

where matrix V is the dual metric that appears in the expansion of DBI action (4). Note
that the D is orthogonal matrix, i.e., D D = .
Using two-dimensional equation of motion, one can write the worldsheet fields in terms
of right- and left-moving components. In terms of these chiral fields, closed NSNS and
open NS vertex operators are


e z), (
z), (
z), p : ,
V NSNS = :VnNS X(z), (z), (z), p : :VmNS X(

e

(x) + (x),
(x) + (x),
k :
V NS = :VnNS X(x) + X(x),
where is super partner of worldsheet field X and is worldsheet superghost field.
The indices n, m refer to the superghost charge of vertex operators, and p and k are closed
and open string momentum, respectively. In order to work with only right-moving fields,
we use the following doubling trick:
e (z) D X (z),
X

(z) D (z),

z) (z).
(

(28)

These replacements in effect extend the right-moving fields to the entire complex plane
and shift modification arising from mixed boundary condition from propagators to vertex
operators. Under these replacement, worldsheet propagator between all right-moving fields
take the standard form [50] except the following boundary propagator:


i
X (x1 ) X (x2 ) = log(x1 x2 ) + F (x1 x2 )
2

(29)

where (x1 x2 ) = 1 (1) if x1 > x2 (x1 < x2 ). Note that the orthogonal property of the
D matrix is an important ingredient for writing the propagators in the standard form. The
vertex operators under transformation (28) becomes


V NSNS = :VnNS X(z), (z), (z), p : :VmNS DX(z), D(z), (z), p : ,

V NS = : VnNS X(x) + DX(x), (x) + D(x), 2(x), k : .
The vertex operator for massless NSNS and NS states and open string tachyon are
V NSNS = (D) :Vn (p, z): , :Vm (pD, z ):
V NS = ( G) :Vn (2k V T , x): ,
V =: Vn (2k V T , x): ,

(30)

where G ab = (ab + D ab )/2 = V ba for gauge field, G ij = (ij D ij )/2 = N ij for scalar
field and G ai = 0 otherwise. The open string vertex operators in (0) and (1) pictures are


V0 (k, x) = X (x) + ik (x) (x) eikX(x),


V1 (k, x) = e(x) (x)eikX(x),

V0 (k, x) = ik (x)eikX(x),
V1 (k, x) = e(x)eikX(x).
The physical conditions for the massless open string and tachyon are

298

M.R. Garousi / Nuclear Physics B 584 (2000) 284299

massless:
tachyon:

k VS k = 0 ,
1
k VS k =
4

k VS = 0,

and for massless closed string are p2 = 0 and p = 0 where is the closed string
polarization which is traceless and symmetric(antisymmetric) for graviton (KalbRamond)
and
1
(31)
= ( ` p ` p ) , `p = 1
8
for the dilaton. Using the fact that D is orthogonal matrix, one finds the following
identities:
G G T = G S ,

(DG T )ab = G ab ,

(DG T )ij = N ij ,

where the G S is symmetric part of the G matrix.


References
[1]
[2]
[3]
[4]
[5]
[6]
[7]
[8]
[9]
[10]
[11]
[12]
[13]
[14]
[15]
[16]
[17]
[18]
[19]
[20]
[21]
[22]
[23]
[24]
[25]
[26]
[27]
[28]
[29]

A. Sen, An introduction to nonperturbative string theory, hep-th/9802051.


C. Vafa, Lectures on strings and dualities, hep-th/9702201.
J. Polchinski, Rev. Mod. Phys. 68 (1996) 1245, hep-th/9607050.
M.J. Duff, Int. J. Mod. Phys. A 11 (1996) 5623, hep-th/9608117.
J.H. Schwarz, Nucl. Phys. Proc. Suppl. B 55 (1997) 1, hep-th/9607201.
J. Polchinski, TASI lectures on D-branes, hep-th/9611050.
J. Polchinski, S. Chaudhuri, C.V. Johnson, Notes on D-branes, hep-th/9602052.
W. Taylor, Lectures on D-branes, gauge theory and M(atrices), hep-th/9801182.
O. Bergman, M.R. Gaberdiel, Phys. Lett. B 441 (1998) 133, hep-th/9806155.
A. Sen, JHEP 9809 (1998) 023, hep-th/9808141.
A. Sen, JHEP 9810 (1998) 021, hep-th/9809111.
A. Sen, JHEP 9812 (1998) 021, hep-th/9812031.
P. Horava, Adv. Theor. Math. Phys. 2 (1999) 1373, hep-th/9812135.
R.G. Leigh, Mod. Phys. Lett. A 4 (1989) 2767.
C.G. Callan, C. Lovelace, C.R. Nappi, S.A. Yost, Nucl. Phys. B 308 (1988) 221.
A. Abouelsaood, C.G. Callen, C.R. Nappi, S.A. Yost, Nucl. Phys. B 280 (1987) 599.
A.A. Tseytlin, BornInfeld action, supersymmetry and string theory, hep-th/9908105.
M.R. Douglas, Branes within branes , hep-th/9512077.
M. Li, Nucl. Phys. B 460 (1996) 351, hep-th/9510161.
M. Green, J.A. Harvey, G. Moore, Class. Quant. Grav. 14 (1997) 47, hep-th/9605033.
A. Sen, JHEP 9808 (1998) 010, hep-th/9805019.
A. Recknagel, V. Schomerus, Nucl. Phys. B 545 (1999) 233, hep-th/9811237.
C.G. Callan, I.R. Klebanov, A.W. Ludwig, J.M. Maldacena, Nucl. Phys. B 422 (1994) 417,
hep-th/9402113.
J. Polchinski, L. Thorlacius, Phys. Rev. D 50 (1994) 622, hep-th/9404008.
A. Sen, Int. J. Mod. Phys. A 14 (1999) 4061, hep-th/9902105.
A. Sen, JHEP 9912 (1999) 027, hep-th/9911116.
A. Sen, B. Zwiebach, Tachyon condensation in string field theory, hep-th/9912249.
V.A. Kostelecky, Samuel, Phys. Lett. B 207 (1988) 169.
W. Taylor, D-brane effective field theory from string field theory, hep-th/0001201.

M.R. Garousi / Nuclear Physics B 584 (2000) 284299

299

[30] N. Moeller, W. Taylor, Level truncation and the tachyon in open bosonic string theory, hepth/0002237.
[31] J.A. Harvey, P. Kraus, D-branes as unstable lumps in bosonic open string field theory, hepth/0002117.
[32] N. Berkovits, The tachyon potential in open NeveuSchwarz string field theory, hepth/0001084.
[33] N. Berkovits, A. Sen, B. Zwiebach, Tachyon condensation in superstring field theory, hepth/0002211.
[34] A. Sen, JHEP 9910 (1999) 008, hep-th/9909062.
[35] A. Sen, Non-BPS States and Branes in String Theory, hep-th/9904207.
[36] M. Billo, B. Craps, F. Roose, JHEP 9906 (1999) 033, hep-th/9905157.
[37] N. Seiberg, E. Witten, JHEP 9909 (1999) 032, hep-th/9908142.
[38] R.C. Myers, JHEP 9912 (1999) 022, hep-th/9910053.
[39] M.R. Garousi, Non-commutative world-volume interactions on D-branes and DiracBorn
Infeld action, hep-th/9909214.
[40] M.R. Garousi, R.C. Myers, Nucl. Phys. B 475 (1996) 193, hep-th/9603194.
[41] A. Hashimoto, I.R. Klebanov, Phys. Lett. B 381 (1996) 437, hep-th/9604065.
[42] A. Hashimoto, I.R. Klebanov, Nucl. Phys. Proc. Suppl. B 55 (1997) 118, hep-th/9611214.
[43] H. Dorn, Nucl. Phys. B 494 (1997) 105, hep-th/9612120.
[44] C.M. Hull, JHEP 9810 (1998) 011, hep-th/9711179.
[45] A.A. Tseytlin, Nucl. Phys. B 501 (1997) 41, hep-th/9701125.
[46] M.R. Garousi, R.C. Myers, Nucl. Phys. B 542 (1999) 73, hep-th/9809100.
[47] M.E. Peskin, Introduction to string and superstring theory II, in: H. Haber et al. (Eds.), From
the planck scale to the weak scale: toward a theory of the universe, Proc. of TASIs86, World
Scientific, 1987.
[48] V.A. Kostelecky, O. Lechtenfeld, W. Lerche, S. Samuel, S. Watamura, Nucl. Phys. B 288 (1987)
173.
[49] J. Dai, R.G. Leigh, J. Polchinski, Mod. Phys. Lett. A 4 (1989) 2073.
[50] M.R. Garousi, JHEP 9812 (1998) 008, hep-th/9805078.

Nuclear Physics B 584 (2000) 300312


www.elsevier.nl/locate/npe

Tachyon condensation and brane descent relations


in p-adic string theory
Debashis Ghoshal , Ashoke Sen
Mehta Research Institute of Mathematics and Mathematical Physics, Chhatnag Road, Jhusi,
Allahabad 211019, India
Received 11 May 2000; accepted 6 June 2000

Abstract
It has been conjectured that an extremum of the tachyon potential of a bosonic D-brane represents
the vacuum without any D-brane, and that various tachyonic lump solutions represent D-branes of
lower dimension. We show that the tree level effective action of p-adic string theory, the expression
for which is known exactly, provides an explicit realisation of these conjectures. 2000 Elsevier
Science B.V. All rights reserved.
PACS: 11.10.Lm; 11.25.Sq; 11.27.+d
Keywords: p-adic string theory; D-branes; Solitons

1. Introduction and summary


The world-volume theory on the D-brane of a bosonic string theory contains a tachyonic
mode. It has been conjectured that the tachyon potential has a non-trivial extremum where
the potential energy of the tachyon exactly cancels the tension of the D-brane, and that
this configuration represents the closed string vacuum without any D-brane [1]. It has been
further conjectured that various tachyonic lump solutions on the D-brane world-volume
represent D-branes of lower dimensions [14]. These conjectures and their generalisations
to superstring theories [59] have been tested by various methods [24,8,1023]. However,
since the exact effective action for the tachyon field is not known, there is no direct proof
of these conjectures.
In this paper we point out that in the p-adic string theory introduced in [2433] (see
[34] for a review) we can explicitly check these conjectures. It should be emphasised that
although the p-adic string is an exotic object, the spacetime it describes is the familiar
Corresponding author.

E-mail addresses: ghoshal@mri.ernet.in (D. Ghoshal), ashoke.sen@cern.ch, sen@mri.ernet.in (A. Sen).


0550-3213/00/$ see front matter 2000 Elsevier Science B.V. All rights reserved.
PII: S 0 5 5 0 - 3 2 1 3 ( 0 0 ) 0 0 3 7 7 - 1

D. Ghoshal, A. Sen / Nuclear Physics B 584 (2000) 300312

301

one. 1 In the p-adic open string theory, which in modern language can be regarded as the
world-volume theory of a space-filling D-brane, the exact classical action of the tachyon
field and various solutions of the equations of motion are known [26]. Among the known
non-trivial solutions is a translationally invariant solution with the property that it is a
local minimum of the potential, and that the propagator of the tachyon field describing
fluctuations around this background has no physical pole. Thus this configuration has no
physical open string excitations, and is naturally identified with the vacuum without a
D-brane. The exact tachyon equation of motion of the p-adic string theory also has classical
lump solutions for all codimension > 1, which approach the vacuum solution far away
from the core of the soliton. If the original open string theory is formulated in (d 1, 1)dimensional spacetime, 2 then such a lump solution of codimension (d q 1) describes
a solitonic q-brane. We show that the world-volume theory on the solitonic q-brane agrees
with the expected world-volume theory on a Dirichlet q-brane in the p-adic string theory,
to the extent that we can compare them with the present knowledge. This provides strong
evidence that these lump solutions can be identified as lower-dimensional D-branes.
The paper is organised as follows. In Section 2 we summarise the exact effective action
of the tachyon in the p-adic string theory, the known solutions of the equation of motion
derived from the action and their properties. In Section 3 we analyse the world-volume
theory of the solitonic q-brane, and in Section 4 we compare this with the world-volume
theory of a Dirichlet q-brane. Section 5 contains some comments on further extension of
this work, and ends with speculation on its possible application to the study of tachyon
condensation in ordinary bosonic string theory.

2. Solitonic q-branes of p-adic string theory


In Ref. [24] p-adic string theory was defined as follows. Consider the expressions for
various amplitudes in ordinary bosonic open string theory, written as integrals over the
boundary of the world-sheet which is the real line R. Now replace the integrals over R by
integrals over the p-adic field Qp with appropriate measure, and the norms of the functions
in the integrand by the p-adic norms. These rules were subsequently derived from a local
action defined on the world-sheet of the p-adic string [34,37,38]. Using p-adic analysis,
it is possible to compute N tachyon amplitudes at tree-level for all N > 3.
This leads to an exact action for the open string tachyon in d-dimensional p-adic string
theory. This action is given in Ref. [26]


Z
Z
1 p2
1 12
1
d
d
p+1

d x p 2 +
,
(2.1)
S = d xL= 2
2
p+1
g p1
where 2 denotes the d-dimensional Laplacian, is the tachyon field (after a rescaling and
a shift), g is the open string coupling constant, and p is an arbitrary prime number. We are
1 A different type of p-adic string was considered in [35,36].
2 There is as yet no compelling reason for a critical dimension in p-adic string theory, but the so called adelic

formula [25,34] relating the product of four tachyon amplitudes in p-adic strings for all primes p to that in the
bosonic string suggests that they all have the same critical dimension d = 26.

302

D. Ghoshal, A. Sen / Nuclear Physics B 584 (2000) 300312

using metric with signature (, +, +, . . . , +). If we denote by (2p0 )1 the tension of


the p-adic string as defined in Refs. [37,38], then our choice of units correspond to [37,38]
ln p
.
(2.2)
p0 =
2
We have added a constant term to the Lagrangian density L so that it vanishes at = 0.
Fig. 1 shows the qualitative features of the tachyon potential for different values of p.
The equation of motion derived from this action is,
1

p 2 2 = p .

(2.3)

Different known solutions of this equation are as follows [26]:


The configuration = 1 is the original vacuum around which we quantised the
string. 3 We shall call this the D-(d 1)-brane solution. The energy density associated
with this configuration, which can be identified as the tension Td1 of the D-(d 1)-brane
configuration, is given by
1 p2
.
(2.4)
2g 2 p + 1
The configuration = 0 denotes a configuration around which there is no perturbative
physical excitation. We shall identify this with the vacuum configuration. By definition we
have taken the energy density of this vacuum to be zero.
The configuration:




(2.5)
(x) = f x q+1 f x q+2 f x d1 F (dq1) x q+1 , . . . , x d1 ,
Td1 = L( = 1) =

with
f () p

1
2(p1)



1p1 2
,
exp
2 p ln p

(2.6)

describes a soliton solution with energy density localised around the hyperplane x q+1 =
= x d1 = 0. This follows from the identity:
p
1 2
(2.7)
p 2 f () = f () .

Fig. 1. The effective tachyon potential for the p-adic string.


3 For p 6= 2, there is also an equivalent solution corresponding to = 1. Since the action is symmetric under
, we shall restrict our analysis to solutions with positive .

D. Ghoshal, A. Sen / Nuclear Physics B 584 (2000) 300312

303

We shall call (2.5), with f as in (2.6), the solitonic q-brane solution. Let us denote by
x = (x q+1 , . . . , x d1 ) the coordinates transverse to the brane and by xk = (x 0 , . . . , x q )
those tangential to it. The energy density per unit q-volume of this brane, which can be
identified as its tension Tq , is given by
Z
1 p2
,
(2.8)
Tq = d dq1 x L( = F (dq1) (x )) = 2
2gq p + 1
where,

gq = g

p2 1

(dq1)/4

2 p2p/(2p1) ln p

(2.9)

From Eqs. (2.4), (2.8) and (2.9) we see that the ratio of the tension of a q-brane to a
(q 1)-brane is
2p

 1 p 2
Tq
2 p p1 ln p 2
p 1
1
q .
=
=
(2.10)
p
Tq1
p2 1
p p1 2 0
p

In the above equation we have used dimensional analysis and (2.2) to restore factors of
p0 . Note that the ratio (2.10) is independent of q. This is a feature of the D-branes in
ordinary bosonic string theory, and suggests that the solitonic q-branes of p-adic string
theory should have interpretation as D-branes. This also suggests that the self-dual radius
Rsd of the p-adic string theory, where the tension 2Rsd Tq of a wrapped q-brane is equal
to the tension Tq1 of a (q 1)-brane, is given by
pp/(p1) q 0
p .
Rsd = p
p2 1

(2.11)

Note that as p , this approaches the formula for the self-dual radius in ordinary
bosonic string theory.

3. World-volume theory on the solitonic q-branes


Let us now consider a configuration of the type
(x) = F (dq1) (x )(xk ),

(3.1)

with F (dq1) (x ) as defined in (2.5), (2.6). For = 1 this describes the solitonic q-brane.
Fluctuations of around 1 denote fluctuations of localised on the soliton; thus (xk )
can be regarded as one of the fields on its world-volume. We shall call this the tachyon
field on the solitonic q-brane world-volume. 4 Substituting (3.1) into (2.3) and using (2.7)
we get
p 2 2k = p ,
1

4 In the linearised approximation this tachyonic mode was discussed in Ref. [39].

(3.2)

304

D. Ghoshal, A. Sen / Nuclear Physics B 584 (2000) 300312

where 2k denotes the (q + 1)-dimensional Laplacian involving the world-volume


coordinates xk of the q-brane. The action involving can be obtained by substituting
(3.1) into (2.1):

Sq () = S = F (dq1)(x )(xk )


Z
1
1
1 p2
1
p+1 ,
d q+1 xk p 2 2k +
= 2
(3.3)
gq p 1
2
p+1
where gq has been defined in Eq. (2.9).
Note that a solution of (3.2) gives an exact solution of the full equation of motion (2.3).
Thus Eq. (3.2) describes the dynamics of the mode on the q-brane world-volume exactly.
This does not mean that there are no other modes on the q-brane world-volume; rather
what this implies is that it is possible to obtain a consistent truncation of the world-volume
theory of the q-brane by setting all the modes except to zero. In terms of scattering
amplitudes this means that the tree level S-matrix on the q-brane world-volume, involving
only external tachyon states, can be calculated exactly from the action (3.3).
Of the various other (infinite number of ) modes living on the q-brane world-volume
are the (d q 1) massless modes i associated with translations of the brane in the
(d q 1) directions x transverse to the brane. Inclusion of these modes correspond to
deformation of of the form
(x) = F (dq1) (x ) (xk ) + x i F (dq1) (x ) i (xk ) + .

(3.4)

Substituting this in Eq. (2.3), and comparing the coefficients of the independent functions
(F (x ))p and (F (x ))p1 x i F (x ) on both sides, we get the following equations of

motion:

1
p 2 2k = p + O 2 ,

1
(3.5)
p 2 2k i = p1 i + O 2 .
The above equations can be derived from the effective action:

Z
1
1 p2
1
1
i
q+1
p+1
xk p 2 2k +
d
Sq (, ) = 2
2
p+1
gq p 1




1 i 1 2k i 1 p1 i i
3
2
C p

+O .
2
2

(3.6)

C is a normalisation constant whose value is not important to this order, as it can be


changed by rescaling i .
= 1 corresponds to the solitonic q-brane solution. For computing amplitudes
involving the world-volume fields on the q-brane, we define the shifted field and rescaled
fields i through the relation
gq
gq
,
i =
(3.7)
i,
=1+
p
pC
and expand the action (3.6) in powers of and i . This gives

D. Ghoshal, A. Sen / Nuclear Physics B 584 (2000) 300312

p
Sq =
p1

305



gq p+1
1 p
p
1 1 2k 1
2
1+
xk p
+ 2
2
2
gq p + 1
p
gq
2gq




gq p1 i i
1 i 1 2k i 1
3
2
1+
p
+
+ O . (3.8)
2
2
p


Z
d

q+1

There is no linear term in in action (3.8) reflecting the fact that = 0 is a solution of the
equation of motion. The momentum space and the i propagators computed from this
action are given by:
1 (k) = i
1 i j (k) = i

1
p1
,
1 2
p p 2 k 1 1

1
p1
ij .
p p 12 k 2 1

(3.9)

The residues at the poles in the and the i propagators (at k 2 = 2 and k 2 = 0,
respectively) have the same values. This will help us compare the amplitudes involving
s and i s in the external legs.

4. Comparison with the world-volume theory on a Dirichlet q-brane


We shall now compare the world-volume action on the solitonic q-brane with that on
the D-q-brane. We are immediately faced with the question whether it is possible to have
Dirichlet branes in p-adic string theory. Fortunately, a world-sheet approach to p-adic
string has been developed in Refs. [37,38]. According to these authors, this world-sheet
is a so-called BruhatTits tree a Bethe lattice with p + 1 nearest neighbours the
boundary of which is the field of p-adic numbers Qp . The generalisation of the Polyakov
action is the lattice discretisation of the action for free scalar fields corresponding to the
target space coordinates. Now one can either choose Neumann or Dirichlet boundary
conditions as in the case of ordinary strings. It was shown in [37,38] that Neumann
boundary conditions leads to the tachyon amplitudes postulated in [24,26].
While this may be the proper way to define D-branes in p-adic string theory, we shall
content ourselves with the continuation of the relevant formul from ordinary bosonic
string theory. Thus, for our purposes the world-volume theory of a p-adic D-q-brane
is defined by taking the expressions for various amplitudes for an ordinary D-q-brane,
written as integrals over world-sheet coordinates of the appropriate vertex operators, and
then replacing the integrals over real line by integrals over the p-adic field, with all the
norms appearing in the integrand replaced by p-adic norms as in Ref. [26]. In principle
one should be able to derive these rules from the world-sheet description in Refs. [37,38].
For amplitudes involving the external tachyons, described by the vertex operators of the
type eikXk with momentum k restricted to lie along the world-volume of the D-q-brane, the
computation of the amplitude is identical to the one described in Ref. [26]. Thus, following
the analysis there, these S-matrix elements can be obtained from an effective action of the
form:

306

D. Ghoshal, A. Sen / Nuclear Physics B 584 (2000) 300312

1 p2
b
Sq () = 2
gq p 1



1
1
1
p+1 ,
d q+1 xk p 2 2k +
2
p+1

(4.1)

where the tachyon field is shifted so that = 1 describes the D-q-brane, and gq denotes
the coupling constant which characterises the strength of the interaction in the worldvolume theory of the D-q-brane. Comparing this with (3.3) we see that the world-volume
actions for the tachyon fields on the solitonic q-brane and the Dirichlet q-brane agree
exactly if we choose:
gq = gq .

(4.2)

At present there is no independent derivation of g q in terms of g, and hence we cannot


verify Eq. (4.2) independently. 5 But assuming (4.2) to be true, we have a complete
agreement between the world-volume theories involving the tachyon fields on the D-qbrane and the solitonic q-brane.
Next we shall compare the amplitude h i j i on the solitonic q-brane and the D-qbrane. First let us compute this on the solitonic q-brane using the action (3.8). The four
Feynmann diagrams contributing to it have been shown in Fig. 2. These can be easily
evaluated, and the answer is:


p1
1
1
2 p2
+
+ k k +1
Aij (k1 , k2 , k3 , k4 ) = ij gq
k
k
+1
1
2
1
3
p
p
p
1 p
1

1
,
(4.3)
+ k k +1
p 1 4 1
with the contribution to the four terms in the right hand side of (4.3) coming from the
Feynman diagrams (a), (b), (c) and (d), respectively, in Fig. 2. In deriving (4.3) we have
used the mass-shell conditions
k12 = k42 = 2,

k22 = k32 = 0 .

(4.4)

Fig. 2. The Feynmann diagrams contributing to the amplitude h i on the solitonic q-brane.
A dashed line denotes the propagator and a solid line denotes the propagator.
5 This is related to the problem of computing the tension of the D-q-brane independently.

D. Ghoshal, A. Sen / Nuclear Physics B 584 (2000) 300312

307

Let us now evaluate the same amplitude on the D-q-brane. The vertex operator associated
i ikXk
e
. Inserting the i and j
with the mode i on the D-q-brane is given by 6 X
vertex operators carrying momenta k2 and k3 at 0 and 1, respectively, and the two vertex
operators carrying momenta k1 and k4 at x and , respectively, we can express the amplitude as:
Z
dx |x|k1 k2 |1 x|k1 k3 .
(4.5)
A ij (k1 , k2 , k3 , k4 ) = ij gq2
Qp

Here | | denotes the p-adic norm and integral over x is over the p-adic field. This is
precisely the integral evaluated in [26]. Using the identity
k1 k2 + k1 k3 + 2 = k1 k4 ,

(4.6)

we can express this amplitude as



1
1
p2 p1
+
+
A ij (k1 , k2 , k3 , k4 ) = ij (gq )2
p
p
pk1 k2 +1 1 pk1 k3 +1 1

1
.
+ k k +1
p 1 4 1

(4.7)

This agrees precisely with Eq. (4.3) for gq = gq .


In fact, it is possible to give a general argument showing that an amplitude with two
external fields and N external fields for arbitrary N , computed from the action (3.8),
agrees with the corresponding amplitude on a Dirichlet q-brane. To see this, let us consider
the situation where we start with the action (2.1) with g replaced by
another coupling
constant g,
and compactify 7 (d q 1) directions on circles of radii 1/ 2. Let ui denote
the compact coordinates and z the non-compact ones, and consider an expansion of the
field of the form:
s dq1


C X i
e
e(z) +
(z) 2 cos( 2 ui ) + .
(4.8)
(x) =
p
i=1

We have restricted to be even under ui ui for each i; this gives a consistent


truncation of the theory at the tree level. The dots stand for higher momentum modes
which will not be required for our analysis. Substituting this into (2.1) (with g replaced
by g)
we get the action:



Z
1 e 1 2z e
2 dq1
1 ep+1
1 p2
q+1
2

z p

+
d
2
g p 1
2
p+1
2



1 i 1 2z i 1 p1 i i
e e
p 2 e

e
+ O(e
C e
3) + .
2
2
(4.9)
6 The notation X is schematic, as care is needed to define the correct vertex operator that corresponds to the
analogous one for ordinary string.
7 Alternatively, we can work with the uncompactified theory, but just examine those modes of which carry

either 0 or 2 units of momentum in (d q 1) of the directions.

308

D. Ghoshal, A. Sen / Nuclear Physics B 584 (2000) 300312

If we identify
gq2

 dq1
2
= g
,
2
2

(4.10)

e, e
i and
this action looks identical to the one in (3.6) with the fields , i replaced by
the identification xk z. In particular we can define the analogues of Eqs. (3.7)
e=1+

gq e

,
p

gq
e
i =

ei ,
pC

(4.11)

e = 1 by expanding (4.9) in a
and compute the S-matrix elements around the vacuum
i
power series in e
and
e . Similarity of (4.9) and (3.6) (and hence (3.8)) shows that the Smatrix elements computed from the action (4.9) around the = 1 background are identical
e = 1 background. In particular the S-matrix
to those computed from (3.8) around the
ej , and an arbitrary number of e
quanta for (4.9) is identical to
element involving a
ei , a
i
j
the S-matrix element involving , and an arbitrary number of quanta in (3.8). 8
On the other hand, the S-matrix elements computed from (4.9) have direct string theory
interpretation, as the action is obtained by compactifying a p-adic string theory [41]. In
ej e
N i is given in terms of correlation functions of
ei ,
ej and
particular the amplitude he
i
Ne
vertex operators on the upper half plane.
operator for e
is proportional
Thevertex
ikZ
i
i
ikZ
, whereas that for
e is given by 2 cos( 2 U )e
. Comparing this with the
to e
corresponding computation for the D-q-brane we see that the e
vertex operator is identical
to the vertex operator with Xk replaced by Z. The i vertex operator on the D-q-brane,
i ikXk
e
, looks different from the
ei vertex operator even after we identify Xk
given by X
with Z. However
ifwe note that
half plane the two point
on the
boundary of the upper

j
i
(x1 )X (x2 )i are identical,
functions h 2 cos( 2 U i (x1 )) 2 cos( 2 U j (x2 ))i and hX
2
both being equal to ij |x1 x2 | , we can conclude that these particular amplitudes in the
compactified string theory are indeed identical to those on the D-q-brane.
To summarise, we have shown that the amplitudes h i j N i computed from (3.8) are
identical to the corresponding amplitudes in the compactified string theory, which in turn
are identical to the corresponding ones on the D-q-brane. This establishes the desired
result. In presenting this argument we have not been careful about the overall normalisation
factors, but the equality already established for the amplitudes h N i and h i j i in the
two theories guarantees that the overall normalisation factors also agree in the two theories.
This provides strong evidence that the solitonic q-branes of the p-adic string theory
should be identified with Dirichlet q-branes.
It will be interesting to systematically extend this comparison to S-matrix elements
involving more than two external i states, and also to S-matrix elements involving higher
level states. It is not easy to establish this in all generality, however we can consider a
subset of the massive modes on the q-brane and show the agreement between the S-matrix
elements on the solitonic q-brane and D-q-brane with at most two of these states on the
external leg.
8 Since the similarity of (3.6) and (4.9) holds only to quadratic order in (e
), we can only make this claim for
two or less external (e
) particles.

D. Ghoshal, A. Sen / Nuclear Physics B 584 (2000) 300312

309

We start with the solitonic q-brane, and consider a generalisation of the expansion (3.4):
(x) = F (dq1)(x )(xk ) +

dq1
X

X0

r=1 {i1 ,...ir }

x1

x ir F (dq1)(x ) i1 ir (xk ) + ,

(4.12)
P
where 0 above denotes sum over those indices {i1 , ir } for which no two in the set are
equal. In this case
p
i1 x ir F (dq1)(x )
x

p1
= pr F (dq1) (x )
x i1 x ir F (dq1) (x ).
(4.13)

Substituting (4.12) into the equation of motion (2.3), and using Eq. (4.13) we get

1
p 2 2k = p + O 2 ,

1
p 2 2k i1 ir = p1r p1 i1 ir + O 2 .

(4.14)

The action involving these fields is given by



Z
1
1 p2
1
1
q+1
p+1
xk p 2 2k +
d
Sq (, ) = 2
gq p 1
2
p+1
dq1
X X0  1
1

Cr i1 ir p 2 2k +r1 i1 ir
2
r=1 {i1 ,...ir }




1
p1 i1 ir i1 ir +O 3 .
2

(4.15)

Cr is a normalisation constant which can be absorbed into the definition of i1 ir . From


this we see that the mass-shell constraint for i1 ir , in the = 1 background, is
k 2 = 2(1 r) .

(4.16)

In order to compare this with the world-volume theory on the D-q-brane, we need to
first identify the vertex operator corresponding to the mode i1 ir . We take this to be
i1
ir ikXk
X
e
.
Vi1 ir = X

(4.17)

This describes a physical state satisfying the same mass shell constraint as Eq. (4.16) as
long as the indices in the set {i1 , . . . ir } are all different.
We shall now compare the S-matrix elements computed from the action (4.15) with
two or less external -legs to that computed directly in the D-q-brane. For two external
tachyons and two external we can do this explicitly and verify that it agrees with
the corresponding computation on the D-q-brane. The computation is identical to the
one discussed earlier. For arbitrary number of external tachyon legs, one can generalise
the argument given for external i -legs. The key ingredient of this argument is that the

310

D. Ghoshal, A. Sen / Nuclear Physics B 584 (2000) 300312


i0

i0

i1
ir ikXk
r ik Xk
1
two point function of X
X
e
and X
is

r X e i for the D-q-brane


1 ), . . . , cos( 2 U ir )e ikZ
2)
cos(
2
U
identical
to
that
between
the
vertex
operators
(

0
0
0
and ( 2)r cos( 2 U i1 ), . . . , cos( 2 U ir )eik Z of the compactified string theory.

5. Comments
We have shown that one can get a consistent truncation of the world-volume theory
on a D-q-brane in p-adic string theory by keeping only the tachyonic mode. Thus by
examining the tree level tachyon amplitudes in the world-volume theory we shall not
discover the existence of the other modes. This suggests that there may be other (massless
and massive) modes living on the world-volume of the space-filling D-(d 1)-brane
as well, inspite of the fact that there are no poles in the tachyon S-matrix elements
corresponding to these states. Indeed, Ref. [40] attempted to generalise the p-adic string
amplitudes to external vector states. If these modes are present they will give rise to new
degrees of freedom on the solitonic q-brane, and will have to be taken into account in
comparing the world-volume theory on the D-q-brane with that on the solitonic q-brane.
It will also be of interest to compute the tension of a Dirichlet q-brane in the padic string theory independently, and compare with Eq. (2.8) describing the tension of a
solitonic q-brane. This will require careful analysis of the cylinder amplitude, and a proper
understanding of the closed string sector of the theory.
It has been shown in Ref. [26] that it is possible to assign ChanPaton factors to
the open string states of a p-padic string theory. This shows the existence of multiple D(d 1)-branes. Furthermore if there are massless gauge fields in the spectrum of open
strings in the p-adic string theory, and if there is a T-duality transformation relating the
D-(d 1)-brane to D-q-brane, then by switching on Wilson lines corresponding to the
gauge fields followed by a T-duality transformation, we can produce static configuration of
D-q-branes separated in space. It will be interesting to examine if the equation of motion
(2.3) admit such solutions. Ideas developed in Ref. [42] may be useful in this context.
It is natural to ask if this analysis has any relevance to the ordinary bosonic string
theory. Firstly, we would like to point out that even if the tachyon potential in the p-adic
string theory is totally unrelated to that in the ordinary bosonic string theory, it can be
regarded as a toy model which nicely illustrates the features expected of the full bosonic
string field theory action. Besides, there is evidence of close relationship between tachyon
amplitudes in the p-adic and ordinary bosonic string theory [25,37,38]. Thus one might
hope that the full tachyon effective action in bosonic string field theory is related in some
way to the tachyon effective action in p-adic string theory. In this direction, we cannot
resist the temptation to point out some apparent similarities between the equation of motion
(2.3) and that in the open bosonic string field theory [43]. To lowest order in the level
truncation scheme [13,14], the tachyon equation of motion in open bosonic string field
theory may be written as [20]
i
h

0
2,
(5.1)
0 2 + 1 ec 2 2 = g

D. Ghoshal, A. Sen / Nuclear Physics B 584 (2000) 300312

311

where c = ln(33 /42 ), g is open string coupling constant after suitable normalisation, and
0
is related to the original tachyon field T by a field redefinition T = ec 2/2 + g 1 ,
so that = 0 is the vacuum without any D-brane, and = 1/g denotes the D-brane. If
we drop the first and the third terms on the left hand side of Eq. (5.1) by hand, then this
equation, after suitable rescaling of x and , reduces to Eq. (2.3) for p = 2. Of course there
is no justification for dropping these terms, so we shall not pursue this matter any further;
but it is not inconceivable that some exact relation between bosonic string field theory and
p-adic string theory will be discovered in the future.

Acknowledgement
We would like to thank Sunil Mukhi for collaboration in the initial stages and for many
useful discussions.

References
[1] A. Sen, Descent relations among bosonic D-branes, Int. J. Mod. Phys. A 14 (1999) 4061, hepth/9902105.
[2] A. Recknagel, V. Schomerus, Boundary deformation theory and moduli spaces of D-branes,
Nucl. Phys. B 545 (1999 ) 233, hep-th/9811237.
[3] C.G. Callan, I.R. Klebanov, A.W. Ludwig, J.M. Maldacena, Exact solution of a boundary
conformal field theory, Nucl. Phys. B 422 (1994) 417, hep-th/9402113.
[4] J. Polchinski, L. Thorlacius, Free fermion representation of a boundary conformal field theory,
Phys. Rev. D 50 (1994) 622, hep-th/9404008.
[5] A. Sen, Stable non-BPS bound states of BPS D-branes, JHEP 9808 (1998) 010, hep-th/9805019.
[6] A. Sen, Tachyon condensation on the brane antibrane system, JHEP 9808 (1998) 012, hepth/9805170.
[7] E. Witten, D-branes and K-theory, JHEP 9812 (1998) 019, hep-th/9810188.
[8] A. Sen, BPS D-branes on non-supersymmetric cycles, JHEP 9812 (1998) 021, hep-th/9812031.
[9] P. Horava, Type IIA D-branes, K-theory, and matrix theory, Adv. Theor. Math. Phys. 2 (1999)
1373, hep-th/9812135.
[10] A. Sen, SO(32) spinors of type I and other solitons on brane-antibrane pair, JHEP 9809 (1998)
023, hep-th/9808141.
[11] M. Frau, L. Gallot, A. Lerda, P. Strigazzi, Stable non-BPS D-branes in type I string theory,
Nucl. Phys. B 564 (2000) 60, hep-th/9903123.
[12] J. Majumder, A. Sen, Vortex pair creation on braneantibrane pair via marginal deformation,
hep-th/0003124.
[13] V.A. Kostelecky, S. Samuel, The static tachyon potential in the open bosonic string theory, Phys.
Lett. B 207 (1988) 169.
[14] V.A. Kostelecky, R. Potting, Expectation values Lorentz invariance, and CPT in the open
bosonic string, Phys. Lett. B 381 (1996) 89, hep-th/9605088.
[15] A. Sen, B. Zwiebach, Tachyon condensation in string field theory, hep-th/9912249.
[16] N. Berkovits, The tachyon potential in open NeveuSchwarz string field theory, hepth/0001084.
[17] N. Berkovits, A. Sen, B. Zwiebach, Tachyon condensation in superstring field theory, hepth/0002211.

312

D. Ghoshal, A. Sen / Nuclear Physics B 584 (2000) 300312

[18] N. Moeller, W. Taylor, Level truncation and the tachyon in open bosonic string field theory,
hep-th/0002237.
[19] J.A. Harvey, P. Kraus, D-branes as unstable lumps in bosonic open string field theory, hepth/0002117.
[20] R. De Mello Koch, A. Jevicki, M. Mihailescu, R. Tatar, Lumps and p-branes in open string field
theory, hep-th/0003031.
[21] P. Fendley, H. Saleur, N.P. Warner, Exact solution of a massless scalar field with a relevant
boundary interaction, Nucl. Phys. B 430 (1994) 577, hep-th/9406125.
[22] J.A. Harvey, D. Kutasov, E.J. Martinec, On the relevance of tachyons, hep-th/0003101.
[23] A. Recknagel, D. Roggenkamp, V. Schomerus, On relevant boundary perturbations of unitary
minimal models, hep-th/0003110.
[24] P.G.O. Freund, M. Olson, Nonarchimedean strings, Phys. Lett. B 199 (1987) 186.
[25] P.G.O. Freund, E. Witten, Adelic string amplitudes, Phys. Lett. B 199 (1987) 191.
[26] L. Brekke, P.G.O. Freund, M. Olson, E. Witten, Non-archimedean string dynamics, Nucl.
Phys. B 302 (1988) 365.
[27] P.H. Frampton, Y. Okada, The p-adic string N point function, Phys. Rev. Lett. 60 (1988) 484.
[28] P.H. Frampton, Y. Okada, Effective scalar field theory of p-adic string, Phys. Rev. D 37 (1988)
3077.
[29] P.H. Frampton, Y. Okada, M.R. Ubriaco, On adelic formulas for the p-adic string, Phys.
Lett. B 213 (1988) 260.
[30] B.L. Spokoiny, Quantum geometry of non-archimedean particles and strings, Phys. Lett. B 208
(1988) 401.
[31] G. Parisi, On p-adic functional integrals, Mod. Phys. Lett. A 3 (1988) 639.
[32] R.B. Zhang, Lagrangian formulation of open and closed p-adic strings, Phys. Lett. B 209 (1988)
229.
[33] Z. Hlousek, D. Spector, p-adic string theory, Ann. Phys. 189 (1989) 370.
[34] L. Brekke, P.G.O. Freund, p-adic numbers in physics, Phys. Rep. 133 (1993) 1, and references
therein.
[35] I.V. Volovich, p-adic string, Class. Quant. Grav. 4 (1987) L83.
[36] B. Grossman, p-adic strings, the Weyl conjectures and anomalies, Phys. Lett. B 197 (1987) 101.
[37] A.V. Zabrodin, Non-archimedean strings and BruhatTits trees, Commun. Math. Phys. 123
(1989) 463.
[38] L.O. Chekhov, A.D. Mironov, A.V. Zabrodin, Multiloop calculations in p-adic string theory
and BruhatTits trees, Commun. Math. Phys. 125 (1989) 675.
[39] P.H. Frampton, H. Nishino, Stability analysis of p-adic string solitons, Phys. Lett. B 242 (1990)
354.
[40] A.V. Marshakov, A.V. Zabrodin, New p-adic string amplitudes, Mod. Phys. Lett. A 5 (1990)
265.
[41] L.O. Chekhov, Y.M. Zinoviev, p-Adic string compactified on a torus, Commun. Math.
Phys. 130 (1990) 130.
[42] R. Gopakumar, S. Minwalla, A. Strominger, Noncommutative solitons, hep-th/0003160.
[43] E. Witten, Noncommutative geometry and string field theory, Nucl. Phys. B 268 (1986) 253.

Nuclear Physics B 584 (2000) 313328


www.elsevier.nl/locate/npe

A three three-brane universe:


new phenomenology for the new millennium?
Ian I. Kogan a,1 , Stavros Mouslopoulos a,2 , Antonios Papazoglou a,3 ,
Graham G. Ross a,4 , Jos Santiago b,5
a Theoretical Physics, Department of Physics, Oxford University, 1 Keble Road, Oxford, OX1 3NP, UK
b Department of Theoretical Physics, Granada University, Avenue Fuente Nueva, Granada, 18071, Spain

Received 14 February 2000; revised 7 April 2000; accepted 17 April 2000

Abstract
We consider an extension of the RandallSundrum model with three parallel 3-branes in a
5-dimensional spacetime. This new construction, apart from providing a solution to the Planck
hierarchy problem, has the advantage that the SM fields are confined on a positive tension brane. The
study of the phenomenology of this model reveals an anomalous first KK state which is generally
much lighter than the remaining tower and also much more strongly coupled to matter. Bounds on
the parameter space of the model can be placed by comparison of specific processes with the SM
background as well as by the latest Cavendish experiments. The model suggests a further exotic
possibility if one drops the requirement of solving the hierarchy problem. In this case gravity may
result from the exchange of the ordinary graviton plus an ultralight KK state and modifications of
gravity may occur at both small and extremely large scales. 2000 Elsevier Science B.V. All rights
reserved.
PACS: 11.10.Kk; 04.50.+h; 11.27.+d

1. Introduction
Recently, there has been considerable interest in theories in which the SM fields are
localized on a 3-brane in a higher dimensional spacetime. Depending on the dimensionality
and the particular form of the geometry of this space, the long standing (Planck)
hierarchy problem 6 can find alternative resolutions. Furthermore, these models make
1 i.kogan@physics.ox.ac.uk
2 s.mouslopoulos@physics.ox.ac.uk
3 a.papazoglou@physics.ox.ac.uk
4 g.ross@physics.ox.ac.uk
5 jsantiag@ugr.es
6 I.e., the hierarchy between the Planck scale and the electroweak scale. There remains the problem of the

hierarchy between the weak and gauge unification scales [6].


0550-3213/00/$ see front matter 2000 Elsevier Science B.V. All rights reserved.
PII: S 0 5 5 0 - 3 2 1 3 ( 0 0 ) 0 0 2 4 1 - 8

314

I.I. Kogan et al. / Nuclear Physics B 584 (2000) 313328

dramatical phenomenological predictions which can be directly confronted with current


and future accelerator experiments as well as cosmological observations. The idea that a
multidimensional universe may not be of KK type, but rather a low dimensional (3 + 1 in
the case of our universe) brane dates back to early eighties when independently Akama
[1] and Rubakov and Shapochnikov [2] suggested models of our universe as topological
defects and later Visser [3] and Squires [4] described how particles can be gravitationally
trapped on the brane. The possibility that gravity be trapped on a brane was suggested by
Gogberashvili [5].
Antoniadis, Arkani-Hamed, Dimopoulos and Dvali [79] proposed that we live on a
3-brane in a 3 + 1 + n space with fully factorizable geometry. The higher dimensional
2 = M n+2 V , where
Planck scale M is then related to the 4D Planck scale by MPl
n
Vn is the compactification volume. Taking the size of the new n dimensions to be
sufficiently large and identifying the (4 + n)-Planck scale with the electroweak scale, an
hierarchy between the electroweak and the Planck scale is introduced. Experimental and
astrophysical constraints demand that n > 2, M & 30 TeV and allow new dimensions even
of submillimeter size. However, a new hierarchy must now be explained, namely the ratio
of the large compactification radius to the electroweak scale. Even for six extra dimensions
this must be greater than 105 .
In the light of this new problem, Randall and Sundrum proposed [10] an alternative
scenario where they assumed one extra dimension along which the geometry is nonfactorizable. Their construction consists of two parallel 3-branes sitting on the fixed points
of an S 1 /Z2 orbifold (see Fig. 1). The 5D spacetime is essentially a slice of AdS5 and
the tensions of the two 3-branes are chosen so that the 4D spacetime appears flat. This
last requirement forces the one of the two branes to have negative tension. An exponential
warp factor in the metric then generates a difference of the mass scales between the two
branes that could be O(1015 GeV) although the size of the orbifold is only of the order
of Planck length. Assuming that the fundamental mass scale on the positive brane is of
the order of MPl we can readily get a mass scale on the negative brane of the order the
electroweak scale, thus solving the hierarchy problem. In this the compactification radius
need only be some 35 times larger than the Planck length. The phenomenology of this
model has been extensively explored in Refs. [11,12]. The KK tower of spin-2 graviton

Fig. 1. RandallSundrum model with two branes at the orbifold S 1 /Z2 fixed points. Here and further
a brane with positive cosmological constant is called positive or + brane and a brane with negative
cosmological constant is called negative or brane.

I.I. Kogan et al. / Nuclear Physics B 584 (2000) 313328

315

resonances starts from the TeV scale with TeV spacing giving rise to characteristic signals
in high energy colliders.
Several models have been constructed since then [1315] that extend the original RS
model to multibrane configurations, parallel or intersecting, with a single or different
cosmological constants between them. A generic characteristic of these models is the
presence of negative tension branes which are necessary for the branes to be flat. Interesting
cases where the branes are not flat or the extra dimension(s) is not compact have also been
considered in Refs. [1620]. The phenomenology of these models is generally complicated
and little has been written about them.
For a viable theory the brane world must reproduce correct gravity and cosmology
of our universe. In the RS picture the negative cosmological constant of the bulk is used
to cancel the cosmological constant or tension on the brane. On a brane with a positive
tension (as for example in the single brane scenario [18]), gravity is effectively confined
to the brane by the steep warp factor generated by the tension dominating the brane. Of
course in realistic cosmologies the energy density of the universe must be dominated by
matter. An important observation about the cosmology of the brane world was made in
Refs. [2124]. These papers showed that the Hubble parameter H governing the expansion
of the scale factor on the brane has a different behavior than derived from the usual 4-d
Friedmann equations. In particular, the Hubble parameter is proportional to the energy

density on the brane instead of the familiar dependence H .


In the papers [2529] (see also [3035]) it was shown that for the RandallSundrum
construction the full energy density is a sum of a vacuum energy density, i.e., brane
tension br and a matter energy density m and the correct expression for Hubble constant
squared can be obtained by cancelling the leading (br + m )2 term with the term 2B

coming from the negative bulk cosmological constant B so that H br m . From


this picture one can immediately see that in these simplest models life on a negative
brane is impossible we either have normal matter with positive energy density, but
imaginary Hubble constant, or real Hubble constant, but negative energy density, i.e.,
antigravity. Because we are living in the universe with a real Hubble constant and without
any noticeable effects of antigravity the negative brane as a model of our world must be
excluded in the original RS model. This statement is not true for more general models with
nontrivial bulk stressenergy tensor [36,37] in which it was shown that one may have real
Hubble constant without antigravity. However it puts extra constraints on parameters of the
model and we need to make some fine tuning to avoid potential cosmological problems.
The purpose of this paper is to formulate a new model of a brane world in which we
live on a positive brane thus avoiding these cosmological problems. The model consists of
two positive branes located at the fixed points of a S1 /Z2 orbifold with one negative brane
which can move freely in between (see Fig. 2). It is easy to see that the two-brane RS model
is nothing but the limiting case of our three-brane model, when the negative brane hits one
of the positive branes. The model has three parameters, the bulk curvature, the warp factor
and the x factor which effectively measures the distance between one of the positive
branes and the negative brane. The RS model corresponds to the limiting case x = 0.

316

I.I. Kogan et al. / Nuclear Physics B 584 (2000) 313328

Fig. 2. + + model with two + branes at


the fixed points and moving branes. In
the limiting case when x 0 we have a RS
configuration.

Fig. 3. + + model which is a RS configuration on Figure 1 with an insertion of


a moving + brane. This configuration can
be obtained from a configuration with only
branes at the orbifold points and several
moving + branes.

Let us note that so far it is unclear what are the selection rules on branes at orbifold fixed
points and moving branes. In a heterotic M-theory for example [38] one can think about
boundaries of eleven-dimensional space after compactification on a suitable CalabiYau
manifold as negative branes where a fundamental five-brane in M-theory after wrapping
on 2-cycles may play the role of a + brane. Moving a + brane towards one of the
branes we can transform it into a + brane, so one can get a configuration of the type
+ + or with more than one moving + brane the only constraint is the that the
sum of all charges is zero. Let us note that each moving brane on an orbifold must be
counted with a double charge because of it mirror image. 7 It is unclear if one may have
moving branes. At the same time it is necessary to say that so far the full string/Mtheory description of these multibrane configurations is missing and we can not exclude a
possibility that there are moving negative branes.
In this paper we shall discuss in details the + + model of Fig. 2. One of the
striking predictions of our model is the fact that the first KK mode can be very light
and strongly coupled compared to the rest of the KK states, so that the phenomenology
is determined mostly by it. An unusual possibility arises if we relax the requirement that
the Planck hierarchy problem is solved. In this case this light mode can be so light that
the corresponding wavelength can be by order of 1% of the observable size of the universe
while the second KK mode is in submillimeter region. Surprisingly enough this situation
is not excluded experimentally! Thus one may have bi-gravity in all experimentally
analyzed regions gravitational attraction is due to an exchange of two particles the
massless graviton and ultralight first KK mode. Only at scales larger than 1026 cm will
the first KK mode decouple leading to a much smaller gravitational coupling beyond this
length scale.
7 We are grateful to Andre Lucas for interesting discussion on this subject.

I.I. Kogan et al. / Nuclear Physics B 584 (2000) 313328

317

The reason the anomalously light KK mode exists is due to the fact that with more
that one + brane there will be a bound state on each of them when they have infinite
separation. At finite distances there is a mixing between the two localized states. One
superposition is the true ground state while the other configuration has non-zero mass, but
the gap may be very small it is given by a tunneling factor. This effect takes place not
only in + + model but in + + and other models with more than two + branes (in
which case there may be more than one light mode). In this paper we consider only the
+ + configuration which has the interesting bi-gravity possibility. Other models will
be considered in separate publications.
The organization of the paper is as follows. In the next section we construct the model
and discuss the spectrum of KK excitations. In Section 3 the first and subsequent KK modes
are considered in more detail and their masses and couplings as a function of the warp
factor and the parameter x are discussed using both analytical and numerical methods. In
Section 4 we discuss the phenomenology when the first KK mode has a mass in the meVTeV region. In Section 5 we discuss the unusual bi-gravity scenario when the Compton
wavelength of the first KK mode is by order of 1026 cm while the Compton wavelength of
a second KK mode is less than 1 mm. In conclusion, we discuss the possible generalization
of our model and questions for future investigation.

2. The 3-brane model


Our model consists of three parallel 3-branes in an AdS5 space with cosmological
constant < 0. The 5th dimension has the geometry of an orbifold and the branes are
located at L0 = 0, L1 and L2 where L0 and L2 are the orbifold fixed points (see Fig. 2).
Firstly we consider the branes having no matter on them in order to find a suitable vacuum
solution. The action of this setup is:
ZL2

Z
S=

d x

dy


X
G + 2M 3 R

L2

p
b(i) ,
d 4 x Vi G

(1)

i y=L
i

b is the induced metric on the branes and Vi their tensions. The notation is the
where G
same as in Ref. [10]. The Einstein equations that arise from this action are:
!
p
X
b(i) (i)
1
1
G
b
G M N (y Li ) . (2)
Vi
GMN +
RMN GMN R =
2
4M 3
G
i
(i)

At this point we demand that our metric respects 4D Poincar invariance. The metric
ansatz with this property is the following:
ds 2 = e2 (y) dx dx + dy 2.

(3)

Here the warp function (y) is essentially a conformal factor that rescales the 4D
component of the metric. A straightforward calculation gives us the following differential
equations for (y):

318

I.I. Kogan et al. / Nuclear Physics B 584 (2000) 313328

( 0 )2 = k 2 ,
X Vi
(y Li ),
00 =
12M 3
i
p
where k = /(24M 3 ) is a measure of the curvature of the bulk.
The solution of these equations consistent with the orbifold geometry is precisely:



(y) = k L1 |y| L1

(4)
(5)

(6)

with the requirement that the brane tensions are tuned to V0 = /k > 0, V1 = /k < 0,
V2 = /k > 0. If we consider massless fluctuations of this vacuum metric as in Ref. [10]
and then integrate over the 5th dimension, we find the 4D Planck mass is given by

M3 
1 2e2kL1 + e2k(2L1L2) .
(7)
k
The above formula tells us that for large enough kL1 and k(2L1 L2 ) the three mass
scales MPl , M, k can be taken to be of the same order. Thus we take k O(M) in order
not to introduce a new hierarchy, with the additional restriction k < M so that the bulk
curvature is small compared to the 5D Planck scale so that we can trust our solution.
Furthermore, if we put matter on the third brane all the physical masses m on the third
brane will be related to the mass parameters m0 of the fundamental 5D theory by the
conformal (warp) factor
2
=
MPl

m = e (L2 ) m0 = ek(2L1 L2 ) m0 .

(8)

Thus we can assume that the third brane is our universe and get a solution of the Planck
hierarchy problem arranging ek(2L1L2 ) to be of O(1015 ), i.e., 2L1 L2 35k 1 . In
1
this case all the parameters of the model L1
1 , L2 and k are of the order of Plank scale.
To determine the phenomenology of the model we need to know the KK spectrum that
follows from the dimensional reduction. This is determined by considering the (linear)
fluctuations of the metric of the form:


2
(9)
ds 2 = e2 (y) + 3/2 h (x, y) dx dx + dy 2 .
M
Here we have ignored the dilaton mode that could be used to stabilize the brane positions
L1 and L2 as discussed in Refs. [37,3941]. While the dilaton may lead to additional
observable phenomena its inclusion should not change the phenomenology of the graviton
KK modes.
We expand the field h (x, y) in graviton and KK states plane waves:
h (x, y) =

(n)
h(n)
(y),
(x)

(10)

n=0
(n)

(n)

(n)

where ( m2n )h = 0 and fix the gauge as h = h = 0. The function (n) (y)
will obey a second order differential equation which after a change of variables reduces to
an ordinary Schrdinger equation:


2
1 2
b(n) (z) = mn
b(n) (z)
(11)
z + V (z)
2
2

I.I. Kogan et al. / Nuclear Physics B 584 (2000) 313328


15k 2
3k 

(z) + (z z2 ) (z z1 ) (z + z1 ) .
2
2g(z)
8[g(z)]
The new variables and wavefunction in the above equation are defined as:


y [L1 , L2 ],
2ekL1 e2kL1 ky 1 /k,

eky 1/k,
y [0, L1 ],
z

ky

1 /k,
y [L1 , 0],
e

kL
2kL
+ky
1
1
1 /k, y [L2 , L1 ],
2e e
V (z) =

with

b(n) (z) (n) (y)e/2,

319

(12)

(13)

(14)

and the function g(z) as g(z) k{z1 ||z| z1 |} + 1, where z1 = z(L1 ).


This is a quantum mechanical problem with -function potentials of different weight
and an extra 1/g 2 smoothing term (due to the AdS geometry) that gives the potential a
double volcano form. The change of variables has been chosen so that there are no first
derivative terms in the differential equation.
An interesting characteristic of this potential is that it always gives rise to a (massless)
zero mode which reflects the fact that Lorentz invariance is preserved in 4D spacetime. It
is given by
A
b(0) =
.
(15)

[g(z)]3/2
The normalization factor A is determined by the requirement
Zz2
 (0) 2
b (z) = 1,
dz
0

chosen so that we get the standard form of the FierzPauli Lagrangian.


In the specific case where L1 = L2 /2 (and with zero hierarchy) the potential and thus
the zero modes wavefunction is symmetric with respect to the second brane. When the
second brane moves towards the third one the wavefunction has a minimum on the second
brane but different heights on the other two branes, the difference generating the hierarchy
between the first and the third brane.
For the KK modes the solution is given in terms of Bessel functions. For y lying in the
regions A [0, L1 ] and B [L1 , L2 ], we have:
 


  r
 
 
g(z)
mn
mn
A2
A1
(n) A
b
g(z) +
g(z) .
(16)
J2
Y2

=
B1
B2
B
k
k
k
The boundary conditions (one for the continuity of the wavefunction at z1 and three for
the discontinuity of its first derivative at 0, z1 , z2 ) result in a 4 4 homogeneous linear
system which, in order to have a non-trivial solution, leads to the vanishing determinant:





J1 m
Y1 mk
0
0

k



m
m

Y1 k g(z2 )
0
0
J1 k g(z2 )


(17)
J m g(z ) Y m g(z ) J m g(z ) Y m g(z ) = 0,
1 k
1
1 k
1
1 k
1
1 k
1





J2 m g(z1 ) Y2 m g(z1 ) J2 m g(z1 ) Y2 m g(z1 )
k

320

I.I. Kogan et al. / Nuclear Physics B 584 (2000) 313328

where we have suppressed the subscript n on the masses mn .


This is essentially the mass quantization condition which gives the spectrum of the KK
states. For each mass we can then determine the wave function with normalization
Zz2
 (n) 2
b (z) = 1.
dz
0

From the form of the potential we can immediately deduce that there is a second bound
state, the first KK state. In the symmetric case, L1 = L2 /2, this is simply given by reversing
the sign of the graviton wave function for y > L1 (it has one zero at L1 ). When the second
brane moves towards the third this symmetry is lost and the first KK wave function has a
very small value on the first brane, a large value on the third and a zero very close to the
first brane.
The interaction of the KK states to the SM particles is found as
p in Ref. [42] by expanding
R
b G,
b SM fields) with
the minimal gravitational coupling of the SM Lagrangian d4 x GL(
respect to the metric. After the rescaling due to the warp factor we get:
Lint =

g(z2 )3/2 X (n)


b (z2 )h(n)

(x)T (x)
M 3/2
n>0

X
b(n) (z2 )g(z2 )3/2 (n)
A (0)
h
(x)T
(x)

h (x)T (x)

M 3/2
M 3/2

(18)

n>0

with T the energy momentum tensor of the SM Lagrangian. Thus the coupling
suppression of the zero and KK modes to matter is respectively:
A
1
= 3/2 ,
c0 M
b(n) (z2 )g(z2 )3/2

1
=
.
cn
M 3/2

(19)
(20)

For the zero mode the normalization constant A is M 3/2 /MPl which gives the Newtonian
gravitational coupling suppression c0 = MPl .

3. The first and subsequent KK modes: masses and coupling constants


As discussed above the KK spectrum has a special first mode which for all x significantly
different from unity has very different behaviour compared to the other KK states. In the
case x  1 we may obtain a reliable approximation to its mass by using the first terms
of the Bessel power series. From now on we shall use a convenient choice of parameters:
the measure of the curvature of the bulk k, the separation of the second and third brane
x = k(L2 L1 ) and the hierarchy factor w = ek(2L1L2 ) . The first KK mode has mass
given by
m1 = 2kwe2x .

(21)

I.I. Kogan et al. / Nuclear Physics B 584 (2000) 313328

321

The normalization integral can also then be done analytically and the coupling
suppression is found to be independent of x and equal to
c1 = wMPl .

(22)

The reason for this is readily understood because, being dominantly a bound state of the
volcano potential on our brane, it is largely localized on it.
The masses of the other KK states in the above region are found to depend in a different
way on the parameter x. Numerically we find out that the mass of the second state and the
spacing 1m between the subsequent states have the form:
m2 4kwex ,
1m kwe

(23)

(24)

where is a number between 1 and 2. The spacing only approaches a constant for high
enough levels when the Bessel arguments become much greater than one. Again solving
numerically, we find the couplings of the higher modes are suppressed relative to the lowest
mode by a factor proportional to ex . To illustrate the difference between the first and second
modes at large x, we consider the case w = e35 , x = 5 and k = 1017 GeV. The masses
and coupling suppressions of the first two modes are
m1 5.7 MeV

with

c1 630 GeV,

m2 1.6 GeV

with

c2 52 500 GeV.

(25)

When the second brane approaches the third brane, the above approximations break
down and the first mode is not so different from the others. Its mass rises above 100 GeV
and the coupling is no longer constant but has a small dependence on x. In the extreme
case where the positions of the second and the third brane coincide, the positive brane
disappears and we obtain the original RandallSundrum model. In this case the spectrum
starts from several hundreds of GeV up to some TeV depending on the choice of k. It has
large spacing between the KK states and TeV coupling suppression.

4. Phenomenology
In this section we will present a brief discussion of the phenomenology of the KK modes
to be expected in colliders, concentrating on the simple process e+ e + . The
analysis is readily generalized to include q q,
gg initial and final states. A more complete
discussion will appear elsewhere.
Using the Feynman rules of Ref. [42] the contribution of the KK modes to e+ e
+
is given by
2
s 3
D(s)
(26)
1280
where D(s) is the sum over the propagators multiplied by the appropriate coupling
suppressions:

e + e + =

322

I.I. Kogan et al. / Nuclear Physics B 584 (2000) 313328

Fig. 4. e+ e + .

D(s) =

X
n>0

1/cn2
s m2n + in mn

(27)

and s is the C.M. energy of e+ e .


Note that the bad high energy behaviour of this cross section is expected since we
are working with an effective low energy non-renormalizable theory of gravity. Our
effective theory is valid up to an energy scale Ms (which is O(TeV)), which acts as an
ultraviolet cutoff. The theory that is valid above this scale is supposed to give a consistent
description of quantum gravity. Since this is unknown we are only able to determine
the contributions of the KK states with masses less than this scale. This means that the
summation in the previous formula should stop at the KK mode with mass near the cutoff.
The decay rates of the KK states that are present in the above formula are given by:
n =

m3n
,
cn2

(28)

39
0.039 (in the case that the KK
where is a dimensionless constant that is between 320
is light enough, i.e., smaller than 0.5 MeV, that can decay only to massless gauge bosons
71
0.094 (in the case where the KK is heavy enough that can decay
and neutrinos) and 240
to all SM particles).
The details of computation of the total cross section depends on the KK spectrum. In
the case that x is small, x . 5, we have a widely spaced discrete spectrum (from the point
of view of TeV physics) close to the one of the RandallSundrum case with cross section
at a KK resonances of the form res s 3 /m8n . To be definite we will take k = 1017 GeV,
w = e35 . In the limit of very low x (x . 0.1) the mass of the first mode is above current
experimental energies (m1 200 GeV) and the phenomenology is that discussed in [11].
For the rest of the discrete spectrum region (0.1 . x . 5) however there are always KK
resonances in the range of energies of collider experiments. Due to the uncertainty in the
collision energy the narrow KK peak will be reduced as shown in Fig. 5. In this case the
contribution of the KK modes gives an excess of 1000% over the SM contribution which
would have been seen either by direct scanning if the resonance is near the energy at which
the experiments actually run or by means of the process e+ e + which scans a
continuum of energies below the center of mass energy of the experiment. Thus this range
of k, w, and x is already excluded. Of course if k is raised the KK modes become heavier
and there will be a value for which the lightest KK mode is above the experimental limits.

I.I. Kogan et al. / Nuclear Physics B 584 (2000) 313328

323

Fig. 5. e+ e + Resonance of the third mode for k = 1017 GeV, w = e35 and x = 1. The
higher curve is the theoretical curve, the smaller one is the result of the smoothing by the convolution
of the theoretical curve and the experimental resolution of the beam. The SM background lies below
them.

However as k decreases it reintroduces the hierarchy problem which the warp factor is
designed to eliminate.
For values of x greater than x 5 the spacing in the spectrum is so small that we can
safely consider it to be continuous. In this case we substitute in D(s) the sum for n > 2 by
an integral over the mass of the KK excitations, i.e.,
1/c12

1
+
D(s)KK
s m21 + i1 m1 1m c2

ZMs
dm
m2

1
,
s m2 + i

(29)

where the value of the integral is i/(2 s) with the principal value negligible in the

region of interest ( s  Ms ) and we have considered constant coupling suppression c for


the modes with n > 2. The first state is singled out because of its different coupling. In fact
at these values of x the biggest contribution comes from this state (the coupling of the rest
being very small).
For w e35 the contribution of the KK tower is negligible compared with the SM
background. However if we allow the value of w to decrease, the coupling of the first

state increases enough to give noticeable contribution to the cross section ( 30% at s =
189 GeV for w e36 ). Thus we can exclude in this continuum limit values of w e36
and below. A quantitative estimate is given in Fig. 6 where three curves are shown, the
lowest one corresponds to the SM alone, the second takes into account also the tower of
KK in the case of w = e35 and the last one is also with the KK tower but this time with
w = e36 which as can be seen from the plot is already excluded.
To summarize, from the process e+ e + we can exclude a window of the form
0.1 . x . 5 for the preferred value of w (this window corresponds to a value of the mass
of the first KK mode between 0.3 GeV and 200 GeV). Both the the very low x limit
and the continuum limit are allowed. Finally, we can also exclude values of the hierarchy
greater than e36 except for the 0 6 x . 0.1 region.

324

I.I. Kogan et al. / Nuclear Physics B 584 (2000) 313328

Fig. 6. e+ e + Cross section in the continuum limit for k = 1017 GeV and x = 11. The
upper curve is for w = e36 , the medium one for w = e35 and the lower one is the SM contribution
(one loop calculation). The resonance we see is the Z peak.

We have considered as an example the process e+ e + but it may be possible


to obtain similar or even stricter bounds could be obtained from other processes like
e+ e + missing energy, for example. We shall present the detailed cross-section of
the process e+ e + light KK mode in another publication, let us only mention here
that by dimensional analysis the ratio of this cross-section to two-photon electronpositron
annihilation
s
s
(e+ e + light KK mode)
2 102 2 ,
(e+ e + )
c1
c1

(30)

and so we see that this process may be phenomenologically very important.


A further bound on the parameters of our model can be put from the Cavendish
experiments. The fact that gravity is Newtonian at least up to millimeter distances implies
that the corrections to gravitational law due to the presence of the KK states must be
negligible for such distances. The gravitational potential is the Newton law plus a Yukawa
potential due to the exchange of the KK massive particles (in the Newtonian limit):
!
X MPl 2
1 M1 M2
1+
emn r .
(31)
V (r) = 2
cn
MPl r
n>0
The contribution to the above sum of the second and higher modes is negligible
compared with the one of the first KK state, because they have larger masses and coupling
suppressions. Thus, the condition for the corrections of the Newton law to be small for
millimeter scale distances is:


ln w
1
[GeV] .
(32)
x < 15 ln
2
kw
The logarithm for any reasonable choice of the parameters k, w gives a contribution
O(1), so we can safely say that x 15 is the maximum brane separation allowed.

I.I. Kogan et al. / Nuclear Physics B 584 (2000) 313328

325

Fig. 7. Excluded regions for k = 1017 GeV and w = e35 .

Combining the above results the allowed region of the parameters of our model for
k = 1017 GeV and w = e35 are shown in Fig. 7.

5. Bi-gravity
Equations (21) and (24) show that, for large x, the lightest KK mode splits off from the
remaining tower. This leads to an exotic possibility in which the lightest KK mode is the
dominant source of Newtonian gravity!
Cavendish experiments and astronomical observations studying the motions of distant
galaxies have put Newtonian gravity to test from submillimeter distances up to distances
that correspond to 1% of the size of observable universe, searching for violations of the
weak equivalence principle and inverse square law. In the context of the graviton KK
modes discussed above this constrains m < 1031 eV or m > 104 eV. Our exotic scheme
corresponds to the choice m1 1031 eV and m2 > 104 eV. In this case, for length scales
less than 1026 cm gravity is generated by the exchange of both the massless graviton and
the first KK mode. This implies (taking into account the different coupling suppressions of
the massless graviton and the first KK state), that the gravitational coupling as we measure
it is related with the parameters of our model as:


1
1
1
1
=
H MPl wM.
(33)
1
+

2
2
2
w
(wM)2
MPl M
We see that the mass scale on our brane, wM, is now the Planck scale so, although the
warp factor, w, may still be small (i.e., the fundamental scale M  MPlanck ), we do not
now solve even the Planck hierarchy problem. However our example does illustrate how
gravity may be quite different from the form that is usually assumed.
Using equations (33) and (21) and assuming as before that k M, we find that m1 =
2ke2x MPlanck e2x . For m1 = 1031 eV we have m2 102 eV. This comfortably
satisfies the bound m > 104 eV coming from Cavendish experiments. Here we should
note that since the coupling of the second mode is always smaller than the one of the
first mode, the phenomenology of the continuum of the KK states is similar to the case of
Refs. [7,8] and thus does not conflict to experiment.
According to this picture deviations from Newtons law will appear in the submillimeter
regime as the Yukawa corrections of the second and higher KK states become important.
Also the presence of the ultralight first KK state will give deviations from Newtons
law as we probe cosmological scales (of the order of the observable universe). The

326

I.I. Kogan et al. / Nuclear Physics B 584 (2000) 313328

Fig. 8. Exclusion regions for the bi-gravity case and correlation of the first two KK states.

phenomenological signature of this scenario is that gravitational interactions will appear to


become weaker by the factor w for distances larger than 1026 cm!

6. Conclusions
In this paper we discussed a model of a three 3-brane universe with two positive and
one intermediate negative tension brane. Our world is confined to a positive tension brane
which makes this construction the minimal realistic model of the RS class. Due to the
presence of two positive branes there are now two bound states, one associated with the
graviton and one with the first KK mode. Compared to the remaining tower of KK states,
the latter is has relatively small mass and large coupling. Bounds on the parameter space of
this model were placed by comparison with data from collider and Cavendish experiments.
However, the phenomenology of the model needs further investigation and it may be that
missing energy processes can put stricter bounds on the parameter space.
We also explored the possibility of bi-gravity in which observable gravity is due to the
exchange of both the ordinary graviton and the first ultralight KK state. The novel feature
of this description is that gravity is modified at both large and small scales. In particular at
large scales the strength of the gravitational force will be reduced by the warp factor. It is
clearly of interest to explore the cosmological consequences of such a scheme.

Acknowledgements
S.M., A.P. and J.S. would like to thank Peter B. Renton and Peter Richardson for
useful discussions. All authors would like to thank Pavel Kogan for the first three figures
and Subir Sarkar for informative discussions. J.S. would like to thank the Department
of Theoretical Physics, Oxford University, for a kind hospitality during the course of
this project. S.M.s work is supported by the Hellenic State Scholarship Foundation
(IKY) No. 8117781027. A.P.s work is supported by the Hellenic State Scholarship
Foundation (IKY) No. 8017711802. J.S.s work is supported by M.E.C., CICYT (Spain,
AEN 96-1672) and Junta de Andaluca (FQM 101). The work of I.I.K. and G.G.R is
supported in part by PPARC rolling grant PPA/G/O/1998/00567, the EC TMR grant
FMRX-CT-96-0090 and by the INTAS grant RFBR-950567.

I.I. Kogan et al. / Nuclear Physics B 584 (2000) 313328

327

Note added in proof


We would like to thank Prof. Akama for sending manuscript of his paper before it
became available from xxx archive.
References
[1] K. Akama, in: K. Kikkawa, N. Nakanishi, H. Nariai (Eds.), Gauge Theory and Gravitation,
Proceedings of the International Symposium, Nara, Japan, 1982, Springer-Verlag, 1983, p. 267;
hep-th/0001113.
[2] V.A. Rubakov, M.E. Shaposhnikov, Phys. Lett. B 125 (1983) 136.
[3] M. Visser, Phys. Lett. B 159 (1985) 22.
[4] E.J. Squires, Phys. Lett. B 167 (1985) 286.
[5] M. Gogberashvili, hep-ph/9904383.
[6] D. Ghilencea, G.G. Ross, hep-ph/9908369.
[7] N. Arkani-Hamed, S. Dimopoulos, G. Dvali, Phys. Lett. B 429 (1998) 263.
[8] N. Arkani-Hamed, S. Dimopoulos, G. Dvali, Phys. Rev. D 59 (1999) 086004.
[9] I. Antoniadis, N. Arkani-Hamed, S. Dimopoulos, G. Dvali, Phys. Lett. B 436 (1998) 257.
[10] L. Randall, R. Sundrum, Phys. Rev. Lett. 83 (1999) 3370.
[11] H. Davoudiasl, J.L. Hewett, T.G. Rizzo, hep-ph/9909255.
[12] S. Chang, M. Yamaguchi, hep-ph/9909523.
[13] I. Oda, hep-th/9908104; hep-th/9909048.
[14] H. Hatanaka, M. Sakamoto, M. Tachibana, K. Takenaga, hep-th/9909076.
[15] C. Cski, Y. Shirman, Phys. Rev. D 61 (2000) 024008.
[16] T. Li, hep-th/991123.
[17] N. Kaloper, Phys. Rev. D 60 (1999) 123506.
[18] L. Randall, R. Sundrum, hep-th/9906064.
[19] J. Lykken, L. Randall, hep-th/9908076.
[20] N. Arkani-Hamed, S. Dimopoulos, G. Dvali, N. Kaloper, hep-th/9907209.
[21] A. Lukas, B.A. Ovrut, K.S. Stelle, D. Waldram, Phys. Rev. D 59 (1999) 086001.
[22] A. Lukas, B.A. Ovrut, D. Waldram, Phys. Rev. D 60 (1999) 086001; hep-th/9902071.
[23] N. Kaloper, A. Linde, Phys. Rev. D 59 (1999) 101303.
[24] P. Bintruy, C. Deffayet, D. Langlois, hep-th/9905012.
[25] T. Nihei, Phys. Lett. B 465 (1999) 81.
[26] N. Kaloper, Phys. Rev. D 60 (1999) 123506.
[27] C. Cski, M. Graesser, C. Kolda, J. Terning, Phys. Lett. B 462 (1999) 34.
[28] J.M. Cline, C. Grojean, G. Servant, Phys. Rev. Lett. 83 (1999) 4245.
[29] H.B. Kim, H.D. Kim, hep-th/9909053.
[30] C. Grojean, J. Cline, G. Servant, hep-th/9910081.
[31] P. Kraus, hep-th/9910149.
[32] A. Kehagias, E. Kiritsis, JHEP 9911 (1999) 022.
[33] T. Shiromizu, K. Maeda, M. Sasaki, gr-qc/9910076.
[34] E. Flanagan, S.H.H. Tye, I. Wasserman, hep-ph/9910498.
[35] D. Ida, gr-qc/9912002.
[36] P. Kanti, I.I. Kogan, K.A. Olive, M. Pospelov, Phys. Lett. B 468 (1999) 31; hep-ph/9912266.
[37] C. Cski, M. Graesser, L. Randall, J. Terning, hep-ph/9911406.
[38] P. Horava, E. Witten, Nucl. Phys. B 460 (1996) 506; Nucl. Phys. B 475 (1996) 94.
[39] W.D. Goldberger, M.B. Wise, Phys. Rev. D 60 (1999) 107505; Phys. Rev. Lett. 83 (1999) 4922;
hep-ph/9911457.

328

I.I. Kogan et al. / Nuclear Physics B 584 (2000) 313328

[40] O. De Wolfe, D.Z. Freedman, S.S. Gubser, A. Karch, hep-th/9909134.


[41] M.A. Luty, R. Sundrum, hep-th/9910202.
[42] T. Han, J.D. Lykken, R.J. Zhang, Phys. Rev. D 59 (1999) 105006.

Nuclear Physics B 584 (2000) 329358


www.elsevier.nl/locate/npe

Constraints on effective Lagrangian of D-branes


from non-commutative gauge theory
Yuji Okawa a, , Seiji Terashima b,1
a Department of Physics, Graduate School of Science, Osaka University, Toyonaka, Osaka 560-0043, Japan
b Department of Physics, Faculty of Science, University of Tokyo, Tokyo 113-0033, Japan

Received 29 February 2000; revised 25 May 2000; accepted 30 May 2000

Abstract
It was argued that there are two different descriptions of the effective Lagrangian of gauge fields
on D-branes by non-commutative gauge theory and by ordinary gauge theory in the presence of
a constant B field background. In the case of bosonic string theory, however, it was found in the
previous works that the two descriptions are incompatible under the field redefinition which relates
the non-commutative gauge field to the ordinary one found by Seiberg and Witten. In this paper
we resolve this puzzle to observe the necessity of gauge-invariant but B-dependent correction terms
involving metric in the field redefinition which have not been considered before. With the problem
resolved, we establish a systematic method under the 0 expansion to derive the constraints on the
effective Lagrangian imposed by the compatibility of the two descriptions where the form of the field
redefinition is not assumed. 2000 Elsevier Science B.V. All rights reserved.
PACS: 11.25.-w; 11.15.-q
Keywords: String theory; D-branes; Non-commutative geometry

1. Introduction
In the light of the recent developments in superstring and M-theory brought by introducing D-branes, it would be impossible to underestimate the importance of understanding the
dynamics of collective coordinates of D-branes, such as scalar fields on the worldvolume
of the D-branes representing their transverse positions and gauge fields describing internal
degrees of freedom. In some situations or limits, the effective Lagrangian describing such
collective coordinates is approximated or even supposed to be exactly described by the dimensional reduction of super YangMills theory from ten dimensions to the worldvolume
Corresponding author. E-mail: okawa@het.phys.sci.osaka-u.ac.jp; tel.: +81-6-6850-5733; fax: +81-6-68505741
1 seiji@hep-th.phys.s.u-tokyo.ac.jp

0550-3213/00/$ see front matter 2000 Elsevier Science B.V. All rights reserved.
PII: S 0 5 5 0 - 3 2 1 3 ( 0 0 ) 0 0 3 6 0 - 6

330

Y. Okawa, S. Terashima / Nuclear Physics B 584 (2000) 329358

dimensions of the D-branes [1,2]. For example, the matrix model of M-theory [3] is based
on the description of D0-branes in terms of super YangMills theory and the most typical
case of the AdS/CFT correspondence [4], namely the correspondence between AdS5 and
the four-dimensional super YangMills theory, is based on that of D3-branes.
Since the perturbative interactions between D-branes and those between D-branes and
elementary excitations of strings are completely defined by the open string sigma model
with the Dirichlet boundary condition, it is in principle possible to calculate systematic
corrections of the effective Lagrangian to the YangMills theory. For example, if we want
to obtain the effective Lagrangian of gauge fields on D-branes, we should calculate the
S-matrix of the scattering processes of the gauge fields on D-branes in string theory,
then construct the effective Lagrangian such that it reproduces the S-matrix correctly.
Another way to calculate the effective Lagrangian is to calculate the beta function of the
open string sigma model with Dirichlet boundary condition and to look for a Lagrangian
whose equation of motion coincides with the condition that the beta function vanishes. The
resulting Lagrangian is believed to coincide with the one obtained from the string S-matrix
at least for tree-level processes. However, the complexity of the calculation will necessarily
increase if we proceed to higher orders in the expansion with respect to 0 and the string
coupling constant gs in the S-matrix approach and to higher loops in the beta function
approach so that it would be helpful if other complementary approaches to the effective
Lagrangian are available.
Recently it is argued that the effective Lagrangian of the gauge fields on D-branes
is described by non-commutative gauge theory [511] in the presence of a constant
background field of the Neveu-SchwarzNeveu-Schwarz two-form gauge field which is
usually referred to as B field. It is also possible to describe it in terms of ordinary
gauge theory, however, the B-dependence in the two descriptions is totally different and
it turned out that it is possible to constrain the form of the effective Lagrangian by the
compatibility of the two descriptions. Actually, it was shown in [12] that the Dirac
BornInfeld (DBI) Lagrangian [1315] 2 satisfies the compatibility in the approximation
of neglecting derivatives of field strength and its particular form was essential for the
compatibility. It is impossible to derive the DBI Lagrangian from the gauge invariance
alone so that this shows that the requirement of the compatibility does provide us with
information on the dynamics of the gauge fields.
The proof of the equivalence of the two descriptions for the DBI Lagrangian in [12]
was beautiful, however it is not clear how we can obtain the constraints for other terms
imposed by the compatibility in a systematic way so as to study how powerful and useful
this approach will be. This is our basic motivation of the present paper and we will present
a method to obtain the constraints systematically in the 0 expansion.
Actually a method towards this goal was developed to some extent in [17] where
the problem of whether it is possible to include two-derivative corrections 3 to the DBI
2 For a recent review of the DiracBornInfeld theory see [16] and references therein.
3 By n-derivative corrections to the DBI Lagrangian, we mean terms with n derivatives acting on field strengths

(not on gauge fields).

Y. Okawa, S. Terashima / Nuclear Physics B 584 (2000) 329358

331

Lagrangian satisfying the compatibility was discussed and the most general form of the
two-derivative corrections up to the quartic order of field strength, F 4 , in the 0 expansion
was derived. However there was a puzzle that the form of the two-derivative terms which
is consistent with the compatibility disagreed with the effective Lagrangian derived from
bosonic string theory although it was consistent with superstring theory. Does this mean
that the equivalence of the two descriptions in the presence of a constant B field fails in
bosonic string theory? In the light of the argument in [12], we do not think that it is the
case. It is most likely that we had made too strong assumptions so that we only obtained
a limited class of Lagrangians which excludes that of bosonic string theory. If it is the
case, the methods which are currently available such as the one in [17] do not fulfill our
purpose to derive the constraints correctly. Furthermore the problem may not be limited
to the case of bosonic string theory. Without resolving the puzzle, it would be dangerous
to develop discussions based on the equivalence of the two descriptions. Therefore we
have to reconsider the assumptions which have been made and find out the correct set
of the assumptions from which we should derive the constraints using the problem of
whether the puzzle in bosonic strings is resolved as a touchstone of the validity of our
approach.
It will turn out that the assumption which is not satisfied in bosonic strings is the one
on the form of the field redefinition which relates the ordinary gauge field to the noncommutative one. The field redefinition which preserves the gauge equivalence relation
found in [12] and further discussed in [18] should be modified in general and suffered
from gauge-invariant but B-dependent correction terms involving metric. In particular,
our result will show that such terms must exist in the case of bosonic string theory.
We will argue that the form of the field redefinition should not be assumed as input
when constraining the form of the effective Lagrangian and can be rather regarded as a
consequence of the compatibility of the two descriptions. This argument is essential in
resolving the puzzle in the case of bosonic string theory as we will see. We could jump
into the problem of two-derivative correction terms to resolve the puzzle, however, we will
first determine the F 4 terms which coincide with those in the DBI Lagrangian correctly
without assuming the form of the field redefinition in order to show that the idea presented
in this paper is useful to constrain the effective Lagrangian. We then apply it to twoderivative corrections to resolve the puzzle and the generalization to other cases would
be straightforward.
The organization of this paper is as follows. In Section 2, we first review the two
descriptions of the effective Lagrangian of the gauge fields on D-branes in the presence
of a constant B field, namely, the one in terms of ordinary gauge theory and the one by
non-commutative gauge theory, to clarify what we assume when deriving the constraints.
We then derive the F 4 terms in the DBI Lagrangian without assuming the form of the
field redefinition which relates the ordinary gauge field to the non-commutative one in
Section 3. We extend our consideration to two-derivative corrections to the DBI Lagrangian
in Section 4 where the discrepancy in the case of bosonic string theory is resolved by
generalizing the form of the field redefinition. Section 5 is devoted to conclusions and
discussions.

332

Y. Okawa, S. Terashima / Nuclear Physics B 584 (2000) 329358

2. Review of the two descriptions in the presence of B


Let us first review the two descriptions of the effective Lagrangian of D-branes in the
presence of a constant B field background Bij . In this paper, we concentrate on the effective
Lagrangian of a gauge field on a single D-brane in flat spacetime, with constant metric
gij , for simplicity.
The worldsheet action describing this system is
Z

1
d 2 gij a x i a x j 2i 0 Bij  ab a x i b x j
S=
0
4

Z
Z
i
1
2
i a j
d gij a x x
d Bij x i x j ,
(2.1)
=
4 0
2

where is the string worldsheet with Euclidean signature and is its boundary.
A background gauge field couples to the string worldsheet by adding
Z
(2.2)
Sint = i d Ai (x) x i

to the action (2.1). Comparing (2.1) and (2.2), we see that a constant B field can be replaced
by the gauge field
1
Ai = Bij x j ,
2
whose field strength is Fij = Bij . Thus we conclude that there exists a definition of a gauge
field in the effective Lagrangian such that the effective Lagrangian depends on B and F
only in the combination B + F when we turn on a constant B field. This gauge field is an
ordinary one, namely, the gauge transformations and its field strength are defined by
Ai = i ,
Fij = i Aj j Ai ,
Fij = 0.

(2.3)
(2.4)
(2.5)

This is the first description of the effective Lagrangian in terms of ordinary gauge theory.
To derive the second description in terms of non-commutative gauge theory, let us
examine the propagator in (2.1). In the presence of a constant B field, the boundary
condition of open strings is modified and is no longer the Neumann one along the D-brane.
Thus the propagator in the sigma model is also modified so as to satisfy the new boundary
condition. The explicit form of the propagator evaluated at boundary points is [1315]


i
x i ( )x j ( 0 ) = 0 (G1 )ij log( 0 )2 + ij ( 0 ),
2

(2.6)

where the worldsheet is mapped to the upper half plane, and 0 are points on the boundary
and
Gij = gij (2 0 )2 (Bg 1 B)ij ,

(2.7)

Y. Okawa, S. Terashima / Nuclear Physics B 584 (2000) 329358

333

ij
1
1
B
= (2 )
.
(2.8)
g + 2 0 B g 2 0 B
There are two important modifications here. The first one is the coefficient in front of the
log term is no longer the metric (g 1 )ij . The second one is the appearance of the term
proportional to the step function ( ) which is 1 or 1 for positive or negative .
Now consider the -dependence of correlation functions of open string vertex operators
which are given by
+
* k
Y

n
Pn x(n ), 2 x(n ), . . . eip x(n )
0 2

ij

n=1

G,

!
i X n ij m
= exp
p pj (n m )
2 n>m i
+
* k
Y
 ipn x(n )
2
Pn x(n ), x(n ), . . . e

n=1

(2.9)

G,=0

where Pn s are polynomials in derivatives of x and x are coordinates along the D-brane.
Since the second term in the propagator does not contribute to contractions of derivatives
of x, the -dependent part can be factorized as the right-hand side of (2.9). The string
S-matrix can be obtained from these correlation functions by putting external fields on
shell and integrating over the s. Therefore, the S-matrix and the effective Lagrangian
constructed from it have a structure inherited from this form.
So we can see how the effective Lagrangian is modified when we turn on the constant
B field. To distinguish the gauge field in this description from that in the preceding one,
The Lagrangian L is
let us rename it to A and denote the Lagrangian in terms of A as L.
constructed from the one L in the absence of B as follows.
First, the metric which appears when contracting Lorentz indices is modified to Gij
instead of gij corresponding to the modification in the propagator. Secondly, since the
coupling constant can depend on B, let us denote the coupling constant in the presence of
B as Gs . Finally, let us go on to the most important modification related to the appearance
of the -dependent factor
!
i X n ij m
p pj (n m )
(2.10)
exp
2 n>m i
in (2.9). It corresponds to modifying the ordinary product of functions to the associative
but non-commutative product defined by





i
,
(2.11)
f (x) g(x) = exp ij i j f (x + )g(x + )
2
= =0
in the momentum-space representation. Now the B-dependence of the effective Lagrangian
gij by Gij ,
in this description can be obtained through the following replacements: A by A,
gs by Gs and ordinary multiplication by the product. Corresponding to the modification
of the product, the gauge transformations and the definition of field strength are also
modified as follows:

334

Y. Okawa, S. Terashima / Nuclear Physics B 584 (2000) 329358

A i = i + i A i i A i ,

(2.12)

bij = i A j j A i i A i A j + i A j A i ,
F
bij = i F
bij i F
bij .

(2.13)
(2.14)

We have seen that there are two different effective Lagrangians of the gauge field on the
D-brane which reproduce the S-matrix of string theory in the presence of a constant B.
What we have learned from the action (2.1) and the interaction (2.2) can be summarized as
follows.
1. There exists a definition of a gauge field Ai such that the Lagrangian in terms of it
respects the ordinary gauge invariance and it depends on B only in the combination
B + F.
2. There exists a definition of a gauge field A i such that the Lagrangian in terms of it
respects the non-commutative gauge invariance and it depends on B only through
Gij , Gs and ij in the non-commutative product.
(2.15)
These are our fundamental assumptions and we will consider constraints on the form of
the effective Lagrangian imposed by the compatibility of them in what follows.
It is not surprising that there are different descriptions of the effective Lagrangian since
the S-matrix is unchanged under field redefinitions in the effective Lagrangian so that the
construction of the effective Lagrangian from the S-matrix elements is always subject to
an ambiguity originated in the field redefinitions. Thus we do not expect that the two gauge
fields Ai and A i coincide: they would be related by a field redefinition. Usually we consider
field redefinitions of the form
Ai Ai + fi (, F ),
where fi (, F ) denotes an arbitrary gauge-invariant expression made of Fij , k Fij ,
k l Fij , and so on. The field redefinitions of this kind preserve the ordinary gauge
invariance. However they will not work in this case because the gauge transformation of
A i is different from that of Ai . The field redefinition which relates A i to Ai must preserve
the gauge equivalence relation, namely it satisfies

+ A),

= A(A
A(A)
+ A(A)

(2.16)

Whether there exists a field redefinition which satisfies (2.16)


with infinitesimal and .
is a nontrivial question, however, a perturbative solution with respect to was found by
Seiberg and Witten [12]. Its explicit form for the rank-one case is given by 4
1
A i = Ai kl Ak (l Ai + Fli ) + O( 2 ),
(2.17)
2
1
(2.18)
= + kl k Al + O( 2 ).
2
However we should emphasize here that we do not assume the explicit form of the field
redefinition which relates A i to Ai when we derive constraints on the form of the effective
4 Solutions to the gauge equivalence relation were further discussed in [18].

Y. Okawa, S. Terashima / Nuclear Physics B 584 (2000) 329358

335

Lagrangian in the present paper. What we assume is the two assumptions (2.15) alone.
This is an important difference from the previous works such as [12] or [17]. The form of
the field redefinition is rather regarded as a consequence of the compatibility of the two
descriptions in terms of ordinary and non-commutative gauge theories as we will see in the
next section.
Before proceeding, we should make a comment on the relation between our assumptions (2.15) and regularization schemes in the sigma model. We mentioned the ambiguity
related to field redefinitions in constructing effective Lagrangian from S-matrix elements.
In the case of string theory, we can also understand the origin of the ambiguity in the point
of view of the sigma model to be coming from degrees of freedom to choose different regularization schemes as was discussed in [12]. We arrived at the assumptions (2.15) from
the properties of (2.1) and (2.2) at classical level. However it is necessary to regularize
the theory to define composite operators such as (2.2) at quantum level. The description in
terms of the ordinary gauge field Ai will be derived from a PauliVillars type regularization while the description in terms of the non-commutative gauge field A i will be derived
from a point-splitting type regularization. However if we take the simple point-splitting
regularization discussed in [12] in which we cut out the region | 0 | < and take the
limit 0, the non-commutative gauge transformation suffers from 0 corrections before
taking the zero slope limit. Therefore it is not clear whether there is an appropriate regularization corresponding to the non-commutative gauge field A i in the second assumption
of (2.15) where no zero slope limit is taken. In this sense, we regard (2.15) as assumptions although we can argue that they are plausible in the following way. If the effective
action before turning on B is invariant under the ordinary gauge transformation and the
B-dependence can be made only through Gij , Gs and ij , the action after turning on B is
automatically invariant under the non-commutative gauge transformation (2.12) at least for
the case where the rank of the gauge group is greater than one. The case with the rank-one
gauge theory may be slightly subtle but it would be naturally expected that it holds in this
case as well. At any rate, our basic standpoint is that the effective Lagrangian we discuss
in this paper is constructed so as to reproduce the S-matrix elements correctly and it is
not necessary to consider its relation to the background field in the sigma model in what
follows.
3. Determination of F 4 terms revisited
3.1. Determination without assuming the form of the field redefinition
Let us now proceed to see how the F 4 terms in the DBI Lagrangian are determined by
the assumptions (2.15) although the form of the field redefinition is not assumed. Since we
study the effective Lagrangian in the 0 expansion, we present the following formulas for
convenience which will be used repeatedly:
(G1 )ij = (g 1 )ij + (2 0 )2 (g 1 Bg 1 Bg 1 )ij + O( 0 4 ),
0 2

= (2 ) (g
ij

Bg

1 ij

04

) + O( ),

(3.1)
(3.2)

336

Y. Okawa, S. Terashima / Nuclear Physics B 584 (2000) 329358

i
f g = fg (2 0 )2 (g 1 Bg 1 )kl k f l g + O( 0 4 ).
(3.3)
2
The lowest order term of the effective Lagrangian of a gauge field on a D-brane in the 0
expansion is the F 2 term:


det g  1 ij
(g ) Fj k (g 1 )kl Fli + O( 0 )
L(F ) =
gs


det g 
Fij Fj i + O( 0 )
=
gs


det g 
(3.4)
Tr F 2 + O( 0 ) .

gs
Here we omitted a possible overall factor including an appropriate power of 0 . Since the
discussions presented in this paper do not depend on the dimension of spacetime on which
the gauge theory is defined, namely, the dimension of worldvolume of the D-brane, if we
want to supply the overall factor, we only need to multiply an appropriate power of 0 to
the Lagrangian to make the action dimensionless and a numerical constant which depends
on the convention. In the second line of (3.4), we made g 1 implicit as
Ai Bi (g 1 )ij Ai Bj .

(3.5)

Since Lorentz indices in most of the expressions in what follows are contracted with respect
to the metric gij , we will adopt this convention together with
2 (g 1 )ij i j ,

(3.6)

to simplify the expressions unless the other metric Gij is explicitly used. And Tr denotes
the trace over Lorentz indices as can be seen from the third line of (3.4).
Now the assumptions (2.15) imply that we can describe the system in two different ways
when we turn on B as follows:


det g 
(3.7)
Tr(B + F )2 + O( 0 ) ,
L(B + F ) =
gs


det G  1 ij b

bli + O( 0 ) .
b
(3.8)
(G ) Fj k (G1 )kl F
L(F ) =
Gs
In the case of higher-rank gauge theory, it follows from the comparison between (3.7)
and (3.8) when B vanishes that
Gs = gs + O( 0 ),
A i = Ai + O( 0 ).

(3.9)
(3.10)

In the rank-one case, on the other hand, we can only determine the normalizations of Gs
and A i as
Gs = tgs + O( 0 ),

A i = t Ai + O( 0 ),

(3.11)
(3.12)

from the consideration at the lowest order in 0 alone since there is no interaction in the
F 2 term. The normalizations of Ai and A i and hence that of Gs are already determined

Y. Okawa, S. Terashima / Nuclear Physics B 584 (2000) 329358

337

by (2.15) since if we rescale Ai or A i then the B-dependence does not take the combination
bij does not take the
B + F for the description in terms of Ai and the field strength F
form (2.13) anymore as for the description using A i . Therefore we can in principle
determine the constant t from the assumptions (2.15). However the calculation for the
determination of t is slightly messy so that we will defer it to Appendix A and proceed
assuming t = 1 in this section for the sake of brevity which will be justified in Appendix A.
F
b ) coincide at the lowest order in 0 , which
Let us first check that L(B + F ) and L(
is necessary to be consistent with (2.15). In general, the Lagrangian L on the noncommutative side reduces to the one L on the commutative side at the lowest order in 0 .
In this case,
bj k (G1 )kl F
bli = (i A j j A i )(j A i i A j ) + O( 0 2 ).
(G1 )ij F

(3.13)

reduces to
up to total
What is less trivial is the question whether Tr(B +
2
derivative, namely, whether Tr F satisfies the initial term condition defined by
F )2

Tr F 2

f (B + F ) = f (F ) + total derivative,

(3.14)

in [17], which is the condition for a term to be qualified as an initial term of a consistent
Lagrangian in the 0 expansion. It is verified that Tr F 2 satisfies this condition as follows:
Tr(B + F )2
= Tr F 2 + 2 Tr BF + Tr B 2
= Tr F 2 + total derivative + const.

(3.15)

The F 4 terms in the DBI Lagrangian are determined by the consideration at the next
order terms in the 0 expansion of (3.8), which are given by

det G 1 ij b
bli
(G ) Fj k (G1 )kl F
Gs

det g 
=
(i Aj j A i )(j A i i A j ) 4(2 0 )2 Bkl k A i l A j j A i
Gs
+ 2(2 0 )2 (B 2 )ij (j A k k A j )(k A i i A k )

12 (2 0 )2 Tr B 2 (i A j j A i )(j A i i A j ) + O( 0 4 ) .

(3.16)

What is important here is the existence of the second term on the right-hand side of (3.16).
It gives a non-vanishing contribution to the three-point scattering amplitude of the gauge
fields. More precisely, if we represent the asymptotic fields in N -point scattering as
asym a

Ai

(x) = ia eik

(k a )2 = 0,

a x

a = 1, 2, . . . , N,

a k a = 0,

N
X

(3.17)
kia = 0,

(3.18)

a=1

the second term on the right-hand side of (3.16) gives a contribution to the three-point
amplitude of order O(B, 3 , k 3 ). It can be easily shown that no other term can produce the
contribution of this form on the non-commutative side. Therefore this contribution cannot

338

Y. Okawa, S. Terashima / Nuclear Physics B 584 (2000) 329358

be canceled and must be reproduced from the Lagrangian L(B + F ) on the commutative
side.
There are two terms on the commutative side which can produce the O(B, 3 , k 3 )
contribution to the three-point amplitude. They are Tr(B + F )4 and [Tr(B + F )2 ]2 :


Tr(B + F )4 = Tr F 4 + 4 Tr BF 3 + O(B 2 ),
2
Tr(B + F )2 = (Tr F 2 )2 + 4 Tr BF Tr F 2 + O(B 2 ).

(3.19)
(3.20)

There are several terms in Tr BF 3 and Tr BF Tr F 2 when we expand them as Fij = i Aj


j Ai , but some of them which contain 2 Ai or i Ai do not contribute to the S-matrix of the
three-point scattering because of the on-shell conditions k 2 = 0 and k = 0. Moreover, it
will be useful to observe that terms of the form f i g i h in general do not contribute to the
S-matrix of three-point scattering where f , g and h are massless fields or their derivatives.
This follows from the fact k 1 k 2 = k 2 k 3 = k 3 k 1 = 0 which can be easily seen as
0 = (k 3 )2 = (k 1 + k 2 )2 = 2k 1 k 2 ,

(3.21)

where we used k 1 + k 2 + k 3 = 0 and (k a )2 = 0. Another way to see this is to rewrite


f i g i h as follows:
1
1
f i g i h = ( 2 fgh f 2 gh fg 2 h) + 2 (fgh) i (i fgh).
2
2

(3.22)

Having been equipped with this formula, we can extract the part which contributes to the
S-matrix from Tr BF 3 and Tr BF Tr F 2 as follows:
Tr BF 3 = Bij Fj k Fkl Fli
= 2Bij j Ak k Al l Ai 2Bij j Ak k Al i Al
+ terms with 2 A + total derivative,

(3.23)

Tr BF Tr F = Bij Fj i Fkl Flk


2

= 4Bij j Ai k Al l Ak
= 8Bij Ai k Al j l Ak + total derivative
= 8Bij l Ai k Al j Ak + a term with l Al + total derivative.

(3.24)

To summarize, we have found that all the terms which contribute to the S-matrix of order
O(B, 3 , k 3 ). On the non-commutative side, there was only one source,
b G1 F
b ) 4(2 0 )2 Bkl k A i l A j j A i ,
Tr(G1 F
while there were two on the commutative side:
Tr BF 3 2Bij j Ak k Al l Ai 2Bij i Al j Ak k Al ,
Tr BF Tr F 2 8Bij j Ak k Al l Ai .
It is not difficult to show that the contributions to the S-matrix from Bij j Ak k Al l Ai
and Bij i Al j Ak k Al are non-vanishing and linearly independent. Thus the conclusion

Y. Okawa, S. Terashima / Nuclear Physics B 584 (2000) 329358

339

derived from (2.15) is that to reproduce the contribution to the S-matrix from the
F
b ), the following terms must exist in the Lagrangian L(B + F ):
Lagrangian L(
1
(3.25)
2(2 0 )2 Tr BF 3 (2 0 )2 Tr BF Tr F 2 .
2
We can uniquely construct the Lagrangian L(F ) such that L(B + F ) generates the
terms (3.25), which is given by




1
det g
2
0 2 1
4
2 2
Tr F (Tr F )
Tr F + (2 )
L(F ) =
gs
2
8

(3.26)
+ O( 0 4 ) + derivative corrections .
This coincides with the 0 expansion of the DBI Lagrangian for a single Dp-brane,
p
1
det(g + 2 0 F ),
(3.27)
LDBI (F ) =
p
0
(p+1)/2
gs (2) ( )
up to an overall factor and an additive constant. Thus we have succeeded in determining
the F 4 terms in the DBI Lagrangian from the assumptions (2.15) without referring to the
explicit form of the field redefinition which relates A i to Ai . We will derive its form in the
next subsection.
3.2. Field redefinition
We have seen that the two effective Lagrangians,
L(B + F )





1
det g
2
0 2 1
4
2 2
Tr(B + F ) Tr(B + F )
Tr(B + F ) + (2 )
=
gs
2
8

+ O( 0 4 ) + derivative corrections ,
and
F
b) =
L(



det G
1 b
1 b
0 2 1
b )4
Tr(G1 F
Tr(G F G F ) + (2 )
arbitrary
Gs
2



1
1 b 2 2
04
Tr(G F ) arbitrary + O( ) + derivative corrections ,
8

(3.28)

(3.29)

produce the same contribution to the S-matrix of order O(B, 3 , k 3 ). Here we added the
F
b ) which were required by the existence of the corresponding F 4 terms
b4 terms to L(
F
in L(B + F ) and the subscripts arbitrary there mean that the ordering of the four field
strengths in each term is arbitrary. Since the product is non-commutative, we have to
specify the ordering of field strengths as in the case of the ordinary YangMills theory.
However, there is no principle in determining the ordering for the rank-one case and
we leave it arbitrary for now. The fact that two effective Lagrangians produce the same
contribution to the S-matrix does not mean that the two must coincide at off-shell level but

340

Y. Okawa, S. Terashima / Nuclear Physics B 584 (2000) 329358

implies that the fields in the two effective Lagrangians can be related by a field redefinition.
Let us see this explicitly for the case in hand. By expanding the O(0 2 ) terms in L(B + F ),
we have

2
1
1
(2 0 )2 Tr(B + F )4 (2 0 )2 Tr(B + F )2
2
8
1
1
0 2
4
= (2 ) Tr F (2 0 )2 (Tr F 2 )2
2
8
1
+ 2(2 0 )2 Tr BF 3 (2 0 )2 Tr BF Tr F 2
2
1
+ 2(2 0 )2 Tr B 2 F 2 (2 0 )2 Tr B 2 Tr F 2 + total derivative + const.,(3.30)
4
where we used the fact that


1
0 2
2
2
(3.31)
(2 ) Tr(BF ) (Tr BF ) = total derivative.
2
Obviously the O(B) and O(B 2 ) parts of (3.30) do not coincide with those of (3.16) if we
assume A i = Ai . Let us first consider the difference in the O(B) part:
1
1L 2(2 0 )2 Tr BF 3 (2 0 )2 Tr BF Tr F 2
2

4(2 0 )2 Bkl k Ai l Aj j Ai
= 2(2 0 )2 Bkl Flj Fj i Fik + 2(2 0 )2 Bkl Ak l Fij Fj i
+ 2(2 0 )2 Bkl k Ai l Aj Fj i + total derivative
= 2(2 0 )2 Bkl Fj i (Flj Fik + Ak l Fij + k Ai l Aj ) + total derivative.

(3.32)

This must be reduced to the field redefinition which relates A i to Ai . We can make it
manifest by noting the fact that


Bkl Fj i i Ak (l Aj + Flj )


= Bkl Fj i i Ak (l Aj + Flj ) + Ak (l i Aj + i Flj )


= Bkl Fj i (Fik + k Ai )(Flj + l Aj ) + Ak 12 l Fij + i Flj
= Bkl Fj i (Flj Fik + Ak l Fij + k Ai l Aj ),
where we used the facts that
1
Fj i i Flj = Fj i l Fij ,
2
and that
Bkl Fj i (Fik l Aj + k Ai Flj ) = 0.

(3.33)

(3.34)

(3.35)

Then the difference 1L can be rewritten using (3.33) as




1L = 2(2 0 )2 Bkl Fj i i Ak (l Aj + Flj ) + total derivative
= 2(2 0 )2 Bkl i Fij Ak (l Aj + Flj ) + total derivative.

(3.36)

Y. Okawa, S. Terashima / Nuclear Physics B 584 (2000) 329358

341

The fact that 1L does not contribute to the S-matrix and can be reduced to the field
redefinition of A i is now manifest in this form since 1L is proportional to i Fij and hence
vanishes using the equation of motion. If we write
A i = Ai + (2 0 )2 1Ai + O( 0 4 ),

(3.37)

it obeys that
(i A j j A i )(j A i i A j ) = Fij Fj i + 4(2 0 )2 i Fij 1Aj + O( 0 4 ).

(3.38)

Thus the appropriate field redefinition is determined by solving the equation


4(2 0 )2 i Fij 1Aj = 1L.

(3.39)

The solution is given by


1
(3.40)
1Ai = Bkl Ak (l Ai + Fli ),
2
up to gauge transformations, and the relation between A i and Ai is
1
(3.41)
A i = Ai + (2 0 )2 Bkl Ak (l Ai + Fli ) + O( 0 4 ).
2
This precisely coincides with the field redefinition (2.17) found by Seiberg and Witten [12]
if we express in terms of B. This was expected since we assumed in (2.15) the
ordinary gauge invariance in the description in terms of Ai and the non-commutative gauge
invariance in the description using A i so that the gauge equivalence relation (2.16) must be
satisfied. Our result is therefore consistent with the previous works. However it is important
to note that this form of the field redefinition should be regarded as a consequence of
the assumptions (2.15) in our approach. We did not have to know the form of the field
redefinition in the determination of the F 4 terms and the form of the field redefinition was
determined from the difference between the two effective Lagrangians at off-shell level.
The O(B 2 ) part of the difference between (3.16) and (3.30),
1
(2 0 )2 Tr B 2 Tr F 2 ,
(3.42)
4
is proportional to the F 2 term so that it can be absorbed into the definition of Gs as follows:


1
0 2
2
04
(3.43)
Gs = gs 1 (2 ) Tr B + O( ) .
4
Here it is also possible to take care of the difference (3.42) by a field redefinition of A i just
as in the case of the difference in the O(B) part and we cannot determine how we should
treat (3.42) from the consideration at the order 0 2 . However since the normalizations of
Ai and A i are already determined by (2.15) as we mentioned below (3.12), the ambiguity
must be fixed by the consideration at higher orders. We will determine the O(0 2 ) part of
Gs in Appendix B from the consideration at order 0 4 , which justifies (3.43).
We have demonstrated how to constrain the effective Lagrangian of gauge fields on
D-branes from the assumptions (2.15) for the F 4 terms in the DBI Lagrangian. We should
now proceed to the reconsideration of the constraints on the two-derivative corrections
to the DBI Lagrangian where the discrepancy was found in the case of bosonic string
theory [17].

342

Y. Okawa, S. Terashima / Nuclear Physics B 584 (2000) 329358

4. Constraints on two-derivative corrections


4.1. O( 0 ) terms
The two-derivative corrections to the DBI Lagrangian can first appear at order 0
compared with the F 2 term. Let us first survey possible terms at this order in both ordinary
and non-commutative gauge theories.
In ordinary gauge theory, Lagrangians are made of field strength and its derivatives. At
order 0 , terms of the forms F F , F 2 F and F 3 are possible. However since the F 2 F
terms can be transformed to the F F terms using the integration by parts and F 3 terms
vanish for the rank-one case, it is sufficient to consider the F F terms. There are three
different ways to contract Lorentz indices:
T1 i Fik j Fj k ,

T2 j Fik i Fj k ,

T3 k Fij k Fj i .

(4.1)

Using the Bianchi identity, the term T3 reduces to T2 ,


T3 = 2T2 ,

(4.2)

and the two remaining terms T1 and T2 coincide up to total derivative:


T1 = Fik i j Fj k + total derivative,

(4.3)

T2 = Fik j i Fj k + total derivative.

(4.4)

Thus any term at order 0 can be transformed to T1 .


The story is slightly different in non-commutative gauge theory. The building blocks
b and its covariant
of Lagrangians in non-commutative gauge theory are field strength F
derivatives defined by
bj k + i F
bj k A i .
bj k = i F
bj k i A i F
bi F
D

(4.5)

bF
bD
bF
b, F
bD
b 2F
bD
b2F
b and F
b3 are possible. The F
b terms
At order 0 , terms of the forms D
b
b
b
b
can be transformed to the D F D F terms using the integration by parts as in the case of
b3 terms no longer vanish even for the rank-one case. Thus
ordinary gauge theory, but F
there are four terms at order 0 : 5
bi F
bj F
bik D
bj k ,
Tb1 D
bij F
bj k F
bki ,
Tb4 i F

b2 D
bj F
bi F
bik D
bj k ,
T

bk F
bk F
bij D
bj i ,
Tb3 D
(4.6)

b3 term by i to make it Hermitian. Using the Bianchi identity,


where we multiplied the F
b
b
the term T3 reduces to T2 as before,
Tb3 = 2Tb2 ,

(4.7)

but the terms Tb1 and Tb2 do not coincide up to total derivative since
5 Lorentz indices on the non-commutative side should be regarded as being contracted using G although we
ij
will not write it explicitly in this subsection contrary to the conventions (3.5) and (3.6).

Y. Okawa, S. Terashima / Nuclear Physics B 584 (2000) 329358

bik D
bi D
bj F
bj k + total derivative,
Tb1 = F
bik D
bj D
bi F
bj k + total derivative,
Tb2 = F

343

(4.8)
(4.9)

bj no longer commute. The remaining three terms Tb1 , T


b2 and Tb4 are not
bi and D
where D
independent which can be seen as follows:
b2 = F
bik [D
bi , D
bj ]F
bj k + total derivative
Tb1 T
bij F
bj k + i F
bj k F
bij ) + total derivative
bik (i F
= F
= 2Tb4 + total derivative.

(4.10)

b4 } as a basis of O( 0 ) terms in non-commutative gauge theory.


We will choose {Tb1 , T
The origin of the extra term Tb4 can be interpreted as an ambiguity in constructing noncommutative gauge theory from ordinary gauge theory for the rank-one case. This can be
seen manifestly if we rewrite Tb4 as
ib
bij [D
bj k F
bki F
bki F
bj k ) = 1 F
bk , D
bi ]F
bj k ,
(4.11)
Tb4 = F
ij (F
2
2
which precisely corresponds to the ambiguity of the ordering of covariant derivatives when
we construct a non-commutative counterpart of the term Fij k i Fj k . This is characteristic
of the rank-one theory and there is no such ambiguity in higher-rank cases where the
ordering of field strengths or covariant derivatives is already determined in ordinary Yang
Mills theory.
We have found bases of O( 0 ) terms for both ordinary and non-commutative gauge
theories. We will next consider the properties of the bases with respect to their behavior
under field redefinitions and to the relation to our assumptions (2.15).
For ordinary gauge theory our basis consists of T1 alone. It is possible to absorb T1 into
the F 2 term by a field redefinition which is given by
A i = Ai + a(2 0 ) j Fj i + O( 0 2 ),
ej i = Fij Fj i + 4a(2 0 ) i Fij k Fkj + total derivative + O( 0 2 ).
eij F
F

(4.12)
(4.13)

It is important to notice that this field redefinition has the following property:
e)ij = (B + F )ij + a(2 0 ) 2 (B + F )ij + O( 0 2 ).
(B + F

(4.14)

This implies that if the effective Lagrangian in terms of A i depends on B only in the
form of B + F , the Lagrangian in terms of Ai also depends on B only in the combination
B + F , namely, both Ai and A i satisfy the first assumption of (2.15). As can be seen from
this example, the first assumption of (2.15) does not determine the definition of the gauge
field uniquely. For instance, field redefinitions of the form
A i = Ai + fi (F, 2 F, . . .),

(4.15)

where field strengths in fi are accompanied by at least one derivative, do not spoil the first
assumption of (2.15). Since the term T1 satisfies the initial term condition (3.14) because
of the fact that i (B + F )j k = i Fj k for a constant B, we can proceed allowing a finite
T1 term to be present in the Lagrangian without restricting the definition of the gauge field

344

Y. Okawa, S. Terashima / Nuclear Physics B 584 (2000) 329358

further. However we will take an alternative approach that we choose a definition of the
gauge field in terms of which the T1 term vanishes in the Lagrangian among the ones which
satisfy the first assumption of (2.15) for convenience.
For non-commutative gauge theory our basis consists of Tb1 and Tb4 . As in the case of
b2 term by a field redefinition
b1 can be absorbed into the F
ordinary gauge theory, the term T
given by
e
b
F
ij

bj i + O( 0 2 ),
bj F
A i = A i + a(2 0 )D
e =F
b
bij F
bj i + 4a(2 0 )D
bi F
bk F
bij D
bkj
F
ji
+ total derivative + O( 0 2 ).

(4.16)
(4.17)

This field redefinition preserves the second assumption of (2.15) so that we can select
a definition of the non-commutative gauge field satisfying (2.15) such that the term Tb1
vanishes in the Lagrangian. With this convention and the one for the ordinary gauge field
we mentioned in the last paragraph, there is no O( 0 ) term in L(B + F ) and only the Tb4
F
b ) at order 0 , which implies that
term exists in L(
A i = Ai + O( 0 2 ),

(4.18)

namely, no O( 0 ) part in the field redefinition.


On the other hand the term Tb4 cannot be redefined away and it gives a non-vanishing
contribution to the S-matrix at O(B) as we will see shortly. It would be rather trivial that
the existence of Tb4 in the effective Lagrangian is consistent with our assumptions (2.15)
b4 is
for the rank-one case since it vanishes in the commutative limit. Incidentally, the term T
3
consistent for higher-rank cases as well since its commutative counterpart i tr Tr F , where
tr denotes the trace over color indices, satisfies the initial term condition (3.14), which can
be shown as follows:


i
i tr Tr(B + F )3 = tr(B + F )ij (B + F )j k , (B + F )ki
2
i
i
= tr Fij [Fj k , Fki ] + Bij tr[Fj k , Fki ]
2
2
(4.19)
= i tr Tr F 3 .
4.2. Constraints on two-derivative corrections
b2 term produces a non-vanishing contribution to
In Section 3.1, we have shown that the F
the S-matrix of order O(B, 3 , k 3 ) and that the F 4 terms are determined by the requirement
that the Lagrangian L(B + F ) should reproduce the contribution. Having understood that
b4 is possible at order 0 , let us develop a similar discussion for two-derivative
the term T
corrections.
b4 is evaluated in the 0 expansion as follows:
The term T
1
bij n F
bj k m F
bki + O( 0 4 ).
(4.20)
Tb4 = (2 0 )2 Bnm F
2
We can extract the part which gives a non-vanishing contribution to the three-point
amplitude using the formula (3.22). The result is

Y. Okawa, S. Terashima / Nuclear Physics B 584 (2000) 329358

345

Tb4 = (2 0 )2 Bnm i A j n j A k m k A i
+ terms with 2 A + total derivative + O( 0 4 ).

(4.21)

The first term on the right-hand side of (4.21) provides a non-vanishing contribution to the
S-matrix of order O(B, 3 , k 5 ).
On the commutative side, only terms of the form O( 2 F 4 ) can produce the same
form of the contribution after replacing F with B + F . Any term of order O( 2 F 4 )
can be transformed to the following form using the integration by parts and the Bianchi
identity [19]:
L=

7
X

bi Ji ,

(4.22)

i=1

where
J1 = n Fij n Fj i Fkl Flk ,

J2 = n Fij n Fj k Fkl Fli ,

J3 = Fni Fim n Fkl m Flk ,

J4 = n Fni m Fim Fkl Flk ,

J5 = n Fni m Fij Fj k Fkm ,

J6 = 2 Fij Fj i Fkl Flk ,

J7 = 2 Fij Fj k Fkl Fli ,

2 Fij = i k Fkj j k Fki .

(4.23)

The terms J4 , J5 , J6 and J7 contain the part j Fj i so that they do not contribute to the
S-matrix. This holds after replacing F with B + F since the part j Fj i remains intact in the
replacement. Thus we do not need to consider these terms in the search for the term which
b4 . On the other hand, the first three coefficients
reproduces the contribution from the term T
b1 , b2 and b3 in this basis do not change under field redefinition and unambiguous [20].
Therefore our goal is to answer the question whether these coefficients are constrained by
our assumptions (2.15).
Let us denote the O(B n ) part of Ji with F replaced by B + F as Ji (B n ) following [17].
Explicit expressions of Ji (B) and Ji (B 2 ) for i = 1, 2, 3 are given by
J1 (B) = 2 n Fij n Fj i Bkl Flk ,

J1 (B 2 ) = n Fij n Fj i Bkl Blk ,

J2 (B) = 2Bij Fj k n Fkl n Fli ,

J2 (B 2 ) = n Fij n Fj k Bkl Bli ,

J3 (B) = 2Bni Fim n Fkl m Flk ,

J3 (B 2 ) = Bni Bim n Fkl m Flk .

(4.24)

It is easily seen that the values of J1 (B 2 ), J2 (B 2 ) and J3 (B 2 ) vanish if they are evaluated at
on-shell configurations (3.17) satisfying (3.18). We can also show that the terms J1 (B) and
J2 (B) do not contribute to the S-matrix using the formula (3.22). Therefore the term J3 (B)
is the only one which contributes to the S-matrix of order O(B, 3 , k 5 ) on the commutative
side, which can be rewritten using (3.22) as follows:
J3 (B) = 4Bni i Am n k Al m l Ak + terms with 2 A + total derivative
= 4Bni i l Am n k Al m Ak
+ a term with l Al + terms with 2 A + total derivative
= 4Bnm i Aj n j Ak m k Ai
+ a term with A + terms with 2 A + total derivative.

(4.25)

346

Y. Okawa, S. Terashima / Nuclear Physics B 584 (2000) 329358

The non-vanishing contribution to the S-matrix from J3 (B) takes the same form as that
of Tb4 (4.21) so that it is possible to reproduce the S-matrix from Tb4 by J3 (B) with the
following normalization factor:
1
Tb4 (2 0 )2 J3 (B).
4

(4.26)

In addition to J3 , we can add the terms J1 and J2 to the effective Lagrangian without
violating the assumptions (2.15) since J1 (B) and J2 (B) do not contribute to the S-matrix
at the order we are discussing. In general, if a term f (F ) satisfies the condition that
f (B + F ) = f (F ) + total derivative using the equation of motion,

(4.27)

we can add the term to the effective Lagrangian without violating the assumptions (2.15)
at the same order of 0 as f (F ). We will call (4.27) the on-shell initial term condition.
Following this terminology, we can say that the terms J1 and J2 do not satisfy the initial
term condition (3.14) but satisfy the on-shell initial term condition (4.27).
To summarize, the coefficients in front of J1 and J2 are not constrained by the
assumptions (2.15) since J1 and J2 satisfy the on-shell initial term condition. The
coefficient in front of J3 is correlated with that in front of Tb4 on the non-commutative
b4 was arbitrary
side following the relation (4.26). However, the coefficient in front of T
as we discussed in the preceding subsection so that the coefficient in front of J3 is also
arbitrary. Thus our conclusion is that two-derivative corrections of the form O( 2 F 4 ) are
not constrained at all by the assumptions (2.15) at this order.
This result may seem discouraging in view of our motivation to obtain constraints on
the effective Lagrangian. However we do not expect that it holds at higher-order terms
in the 0 expansion because of the following argument. In general it would become more
difficult to satisfy the on-shell initial term condition when the number of field strengths
minus the number of derivatives increases in the term under consideration. If we note
that the existence of the solutions to the on-shell initial term conditions was essential
to our conclusion that there is no constraint on the O( 2 F 4 ) terms, we can reasonably
expect severe constraints on such higher-order terms. We admit, however, that the approach
presented in this paper will not be practical in deriving the constraints on the higher-order
terms and we need more efficient methods. As an example of promising methods we can
refer to the one discussed in [21]. We will get back to this point after discussing the issue
on field redefinitions.
There is another comment on our result regarding the relation between the coefficients
b4 and J3 (4.26). This provides no information on the effective Lagrangian for
in front of T
the rank-one case since Tb4 vanishes in the commutative limit. However if we succeed in
extending our consideration to higher-rank cases, it might be possible to obtain a prediction
on a relation between the coefficient in front of the F 3 term and coefficients in O(D 2 F 4 )
terms.
We should now clarify the relation between the result in this paper and that in [17].
The most general form of O( 2 F 4 ) terms was derived in [17] from the requirement that
F
b ) coincide up to total derivative under the assumption that the field
L(B + F ) and L(

Y. Okawa, S. Terashima / Nuclear Physics B 584 (2000) 329358

347

redefinition is given by (2.17). The result was that the terms J1 , J2 and J3 must appear in
the combination that
1
(4.28)
J1 + 2J2 + J3 .
4
This was inconsistent with the O( 2 F 4 ) terms in bosonic string theory derived from the
string four-point amplitude [19] or from the two-loop beta function in the open string sigma
model [20] which are proportional to 6
1
J1 2J2 + J3 .
4

(4.29)

The conclusion in this paper that no constraint is imposed on O( 2 F 4 ) terms is trivially


consistent with (4.29) and the difference between this conclusion and that in [17] implies
that the relation between the two gauge fields A i and Ai in (2.15) does not in general
take the form of (2.17) assumed in [17]. In particular, the discrepancy between (4.28)
and (4.29) shows that it is indeed the case for bosonic string theory. We will construct a
field redefinition which is relevant to bosonic string theory in the next subsection.
4.3. Corrections to the field redefinition
We presented the on-shell initial term condition (4.27) as a necessary condition for a
term to be added to the effective Lagrangian without violating the assumptions (2.15) in
the preceding subsection. The relation between A i and Ai must be in general modified if
we add a term which satisfies the on-shell initial term condition (4.27) but does not satisfy
the initial term condition (3.14). As we have seen, the terms J1 and J2 are examples of
such terms since J1 (B), J1 (B 2 ), J2 (B) and J2 (B 2 ) are not total derivative although values
of them vanish when evaluated at configurations satisfying the on-shell conditions (3.18).
The terms J4 , J5 , J6 and J7 also satisfy the on-shell initial term condition, however, they
are less interesting than J1 and J2 since they do not contribute to the S-matrix. An explicit
form of the required field redefinition which relates A i to Ai when we add a term which
satisfies the on-shell initial term condition to the effective Lagrangian can be derived in the
same way as we did in Section 3.2 but we will not do that for completely general cases. It
would be sufficient to demonstrate it for some examples including the one which is relevant
to bosonic string theory since the generalization is straightforward.
Let us first consider a case where only J2 exists in the O( 0 3 ) part. In particular, the
F
b ) because of the relation (4.26).
absence of J3 means that Tb4 is not allowed to exist in L(
F
b ) under our convention that Tb1 should be redefined
Thus there are no O( 0 ) terms in L(
away. This simplifies the discussion since the O( 0 2 ) part in the field redefinition (3.41),
which is necessary to satisfy the assumptions (2.15) as we have seen in the preceding
section, does not affect O( 0 3 ) terms under consideration if there are no O( 0 ) terms in
F
F
b ) cannot generate B-dependent terms of order
b ). Furthermore, the O( 0 2 ) part of L(
L(
0
3
which is manifest under our convention (4.18). Therefore the terms J2 (B) and J2 (B 2 ),
6 This expression is slightly different from (4) in [20] but one of the authors was informed of a misprint in (4)
of [20]: the last coefficient b3 should have sign +.

348

Y. Okawa, S. Terashima / Nuclear Physics B 584 (2000) 329358

which are necessary to realize the B-dependence of the form B + F when we add J2 , must
b 2 term by the O( 0 3 ) part of the field redefinition of A i . Its explicit
be generated from the F
form is easily derived if we rewrite J2 (B) and J2 (B 2 ) as follows:
J2 (B) = 2 n Fni j (FBF )j i + total derivative,
J2 (B ) = n Fni j (B F + FB )j i + total derivative.
2

(4.30)
(4.31)

It follows from a similar argument to the one used to determine the form (3.41) that the
field redefinition
1
A i = Ai + (2 0 )2 Bkl Ak (l Ai + Fli )
2
1
(4.32)
c2 (2 0 )3 j (2FBF + B 2 F + FB 2 )j i + O( 0 4 )
4
b2 term. To summarize, the two
generates c2 (2 0 )3 (J2 (B) + J2 (B 2 )) from the F
Lagrangians,
L(B + F )




2
1
det g
1
Tr(B + F )2 + (2 0 )2 Tr(B + F )4 Tr(B + F )2
=
gs
2
8


+ c2 (2 0 )3 n (B + F )ij n (B + F )j k (B + F )kl (B + F )li + O( 0 4 ) , (4.33)

and
F
b)
L(


det G
b G1 F
b)
Tr(G1 F
=
Gs



1
b )4arbitrary 1 Tr(G1 F
b )2 2
+ (2 0 )2 Tr(G1 F
arbitrary
2
8


04
b
b
b
b
b
b
+ c2 (2 ) (Dn Fij Dn Fj k Fkl Fli )G, arbitrary + O( ) , (4.34)
0 3

bF
bs and F
bs in the O( 0 3 ) term contracted using Gij as
with an arbitrary ordering of D
indicated by the subscript, are related by the field redefinition (4.32).
This example shows that 0 corrections of O(B) to the field redefinition (3.41) are in
general possible. Since
Bij =

1
(gg)ij + O( 2 ),
(2 0 )2

(4.35)

this does not take the form of (2.17). Therefore it would be helpful to confirm that (4.32)
preserves the gauge equivalence relation (2.16). Let us decompose the field redefinition
(4.32) as follows:
A i A i Ai ,
where

(4.36)

Y. Okawa, S. Terashima / Nuclear Physics B 584 (2000) 329358

349

1
eli ) + O( 0 4 ),
A i = A i + (2 0 )2 Bkl A k (l A i + F
(4.37)
2
1
(4.38)
A i = Ai c2 (2 0 )3 j (2FBF + B 2 F + FB 2 )j i + O( 0 4 ).
4
By the first part (4.37), the non-commutative gauge field A i is mapped to an ordinary
gauge field A i which respects the ordinary gauge invariance while A i is mapped to another
ordinary gauge field Ai by the second part (4.38) since the difference between A i and
Ai is gauge invariant although it depends on B. This shows that (4.32) preserves the
gauge equivalence relation (2.16). In general, the field redefinition (3.41) maps a noncommutative gauge field to an ordinary gauge field but the B-dependence of the effective
Lagrangian in terms of the resulting gauge field, A i in this example, does not take the form
of B + F . Therefore further B-dependent redefinition like (4.38) is necessary to map it to
the gauge field which satisfies the first assumption of (2.15).
The form of the field redefinition (4.32) does not belong to the class of solutions to the
gauge equivalence relation (2.16) found in [18]. However there is no contradiction since
it was assumed in [18] that Lorentz indices in a mapping from Ai to A i are contracted
among derivatives of the gauge field and ij alone while (g 1 )ij is used in our case (4.32)
although it is implicit under our convention (3.5). 7
Now the extension to cases where other Ji s except J3 exist in the effective Lagrangian
would be straightforward. However if J3 exists the story becomes slightly complicated
F
b ) which accompanies J3 following the relation (4.26).
b4 in L(
because of the presence of T
We have to consider the effects of the O( 0 2 ) part in the field redefinition (3.41) when it
b4 . Here it is convenient to utilize the results of [17]. Let us review
acts on the O( 0 ) term T
them briefly.
It was shown in [17] that the two Lagrangians,





det G b
1

b
(4.39)
T3 + (2 0 )2 J1 + 2J2 + J3 + O( 0 4 ) ,
L1 (F ) =
Gs
4
and
b) =
L 2 (F





1
1
det G b
0 2
04
T1 + (2 ) J5 J6 + J7 + O( ) ,
Gs
8
2

(4.40)

F
b ) = L(B + F ) up to total derivative under the field redefinition (3.41) with
satisfy L(
the definition of Gs (3.43). Here Ji s are the non-commutative counterparts of Ji s with
an arbitrary ordering of the fields. We presented the Lagrangians on the non-commutative
side because we can uniquely construct their commutative counterparts while the other
F
b ), suffers from the ambiguity in the rank-one case discussed
direction, L(B + F ) L(
in Section 4.1. A linear combination of the two Lagrangians is expressed in our basis
{Tb1 , Tb4 } as follows:





b4 + a(2 0 )2 J5 1 J6 + 1 J7
b1 + b T
b ) = det G a T
L(F
Gs
8
2
7 We would like to thank I. Kishimoto for clarifying this point.

350

Y. Okawa, S. Terashima / Nuclear Physics B 584 (2000) 329358




1
1
1
b(2 0 )2 J1 + 2J2 + J3 + 2J5 J6 + J7 + O( 0 4 ) .
4
4
4

(4.41)

It was further argued in [17] that (4.41) is the most general form of two-derivative
F
b ) = L(B + F ) up to
corrections up to this order in the 0 expansion which satisfy L(
total derivative under the field redefinition (3.41) with the definition of Gs (3.43). 8 To see
that it is the case, it is helpful to notice that if there is another Lagrangian


0
b ) = det G a T
b4
b1 + b T
L (F
Gs

F
b ) + O( 0 4 ) , (4.42)
+ O( 0 2 ) terms different from those of L(
b ) = L0 (B + F ) up to total derivative under the field redefiniwhich also satisfies L 0 (F
tion (3.41) with the definition of Gs (3.43), then the difference L0 (F ) L(F ) must be
a solution of the form O( 2 F 4 ) to the initial term condition (3.14). Thus the question
whether (4.41) is the most general form reduces to the one whether there are solutions of
the form O( 2 F 4 ) to the initial term condition (3.14). Regarding the latter question, it was
shown [17] that the condition that O( 2 F 4 ) terms must be proportional to the combination
that
1
1
(4.43)
F (F ) J1 + 2J2 + J3 + 2J5 J6 + J7 ,
4
4
is necessary to satisfy the initial term condition (3.14). It is difficult to see whether
F (F ) satisfies the initial term condition by a direct calculation, however, we can obtain
the answer by an indirect argument in the following way. From (4.20) and the fact
F
b ) = L(B + F ) up to total derivative under the field
that (4.41) with a = 0 satisfies L(
redefinition (3.41) with the definition of Gs (3.43), it follows that
1
1
1
(4.44)
F (B + F ) = Bnm Fij n Fj k m Fki F (F ) + total derivative,
4
2
4
which implies that F (F ) does not satisfy the initial term condition (3.14). Now that the
only remaining possibility was denied, the statement that there is no solution of the form
O( 2 F 4 ) to the initial term condition (3.14) was shown and this implies that (4.41) is the
most general form of two-derivative corrections up to this order in the 0 expansion which
F
b ) = L(B + F ) up to total derivative under the field redefinition (3.41) with the
satisfy L(
definition of Gs (3.43).
This result provides us with a good starting point for the case where J3 is non-vanishing.
Namely, the Lagrangian


F
b G1 F
b ) + b(2 0 )Tb4
b ) = det G Tr(G1 F
L(
Gs



1
b )4arbitrary 1 Tr(G1 F
b )2 2
+ (2 0 )2 Tr(G1 F
arbitrary
2
8
8 The argument for proving this statement developed in Section 3 of [17] was incorrect as explained in the note
added at the end of hep-th/9909132 v2.

Y. Okawa, S. Terashima / Nuclear Physics B 584 (2000) 329358

351




1
1
1
b(2 0 )3 J1 + 2J2 + J3 + 2J5 J6 + J7 + O( 0 4 ) , (4.45)
4
4
4
F
b ) are related by the field redefinition (3.41). If we
and L(B + F ) constructed from L(
want to change the coefficients in font of Ji s except J3 , we should modify the form of
the field redefinition at order 0 3 appropriately as in the preceding example where only J2
exists.
As an interesting example of such cases, let us derive the form of the field redefinition
which is relevant to bosonic string theory. As we mentioned in the preceding subsection,
the coefficients in front of J1 , J2 and J3 calculated in bosonic string theory are proportional
to (4.29) [19,20]. This corresponds to adding b(2 0 )3 J2 to (4.45) so that the form of the
field redefinition is modified to
1
A i = Ai + (2 0 )2 Bkl Ak (l Ai + Fli )
2
1
b(2 0 )3 j (2FBF + B 2 F + FB 2 )j i + O( 0 4 ).
4

(4.46)

If we further change the coefficients in front of J4 , J5 , J6 and J7 which do not affect the
S-matrix, the form of the field redefinition (4.46) itself is modified correspondingly.
However we cannot make the O( 0 3 ) terms vanish since (4.29) does not take the
general form (4.41) in the absence of the O( 0 3 ) terms. Thus the corrections to the field
redefinition (3.41) are not only possible in principle but also realized actually in bosonic
string theory. For superstring theory, it was found that the coefficients in front of J1 , J2
and J3 vanish [19] so that corrections to the field redefinition (3.41) at order 0 3 are not
required. However there is no general argument that it persists to higher orders in the 0
expansion. We should keep such possibility of corrections in mind when we use properties
of the field redefinition which relates the non-commutative gauge field to the ordinary
one. In particular, it would be important to note that corrections of O(B) O( ) modify
) introduced in [12] for more general descriptions of the
the differential equation of A(
system in terms of non-commutative gauge theory.

5. Conclusions and discussions


We considered the constraints on the effective Lagrangian of the gauge field on a
single D-brane in flat spacetime imposed by the compatibility of the description by non F
b ) with that by ordinary gauge theory L(B + F ) in the
commutative gauge theory L(
presence of a constant B field background. We presented a systematic method under the 0
expansion to derive the constraints based on the assumptions (2.15) alone without assuming
the form of the field redefinition which relates the non-commutative gauge field A i to the
ordinary one Ai .
By applying this method to two-derivative corrections to the DBI Lagrangian, we
established the equivalence of the two descriptions for a larger class of Lagrangians.
In particular it contains the effective Lagrangian of bosonic string theory and thus the

352

Y. Okawa, S. Terashima / Nuclear Physics B 584 (2000) 329358

puzzle in bosonic strings found in the previous works was resolved and the equivalence in
this case was first made consistent.
In resolving the puzzle it was essential to observe that the gauge-invariant but
B-dependent corrections to the field redefinition (2.17) are in general necessary for the
compatibility. They were not considered previously because it was assumed that the metric
gij does not appear in the field redefinition which relates A i to Ai , however we showed that
they must exist for the case of bosonic string theory. It should be emphasized that it was
crucial to reach this observation that we did not assume the form of the field redefinition
when we derive the constraints.
It is sometimes said that the form of the field redefinition which relates A i to Ai can be
determined by solving the differential equation derived from the gauge equivalence relation
in [12] up to the ambiguities found in [18]. However our result clearly shows that it is no
longer the case if we allow the metric gij to appear in the field redefinition as in the case of
bosonic string theory. Even the form of the differential equation itself can be modified and
the form of the field redefinition is hardly constrained by the gauge equivalence relation
without the assumption. In the superstring case the field redefinition of the form (2.17) can
be consistent up to two derivatives [21] and may not be corrected. However, as far as we
are aware of, there is no argument which justifies the assumption in the superstring case
that the metric gij does not appear in the field redefinition which relates A i to Ai .
We believe that we have elucidated the mechanism to constrain the effective Lagrangian
of the gauge fields on D-branes using non-commutative gauge theory. Since we presented
a systematic method to obtain the constraints in the 0 expansion, it is in principle possible to calculate the general form of the effective Lagrangian which satisfies the assumptions (2.15) up to an arbitrary order in 0 . Furthermore our study implies that the number
of the free parameters in the general form is equal to or less than the number of solutions to the initial term condition (3.14) plus that of nontrivial solutions to the on-shell
initial term condition (4.27). Therefore if there are no solutions to these conditions in the
two-derivative terms at higher order in 0 , for example, this implies that the form of the
two-derivative terms is in principle determined uniquely by the requirement of the compatibility up to the parameters in front of J1 , J2 and J3 . However we admit that the approach
adopted in the present paper is not practically useful to proceed to the higher orders to
obtain the constraints or the explicit form of the terms as we mentioned in Section 4.2.
Regarding this issue, a general method to construct 2n-derivative terms to all orders in 0
which satisfy the compatibility of the two descriptions in the approximation of neglecting
(2n + 2)-derivative terms when the field redefinition takes the form of (2.17) was presented in [21]. It is therefore necessary to extend the method to apply it to more general
cases where there are corrections to the field redefinition of the form (2.17) such as the
case of bosonic string theory. If we could succeed in such generalization, it would be
expected that it will provide us with a new powerful method to study the dynamics of
D-branes. For developments in this direction, the simplified SeibergWitten map considered in [22,23] and [24] may be useful because of its geometric nature although we should
clarify its meaning for our approach. See also a related work [25]. How to construct actions
which are invariant under the simplified map was recently discussed in [26]. In addition,

Y. Okawa, S. Terashima / Nuclear Physics B 584 (2000) 329358

353

it is interesting to combine our approach with consideration of supersymmetry and string


dualities. It will probably provide us with further constraints.
We only considered the constraints on the effective Lagrangian at the lowest order in
the expansion with respect to the string coupling constant gs in this paper. There seems
to be no crucial obstruction to the extension of our approach to higher orders in gs
although some modifications may be required. An issue related to this kind of extension
was discussed in [27]. Furthermore, although the assumptions (2.15) were derived from
the action of the sigma model, they are not related to the expansions with respect to 0
and gs once extracted. It would be interesting if we could obtain some non-perturbative
information on the dynamics of D-branes from them. Of course it might be the case that
there are limitations of the description in terms of non-commutative gauge theory at nonperturbative level and it is important to investigate them.
Another important extension of our approach is to consider higher-rank gauge theory. It
would be interesting if we could obtain some insight into the non-Abelian generalization
of the DBI Lagrangian [28]. Although we foresee possible complication originated in its
non-Abelian nature which exists even on the side of ordinary gauge theory, it will be worth
investigating in view of the various important developments which have been made by the
super YangMills theory in the description of multi-body systems of D-branes.

Acknowledgements
We would like to thank K. Hashimoto, H. Kajiura and T. Kawano for useful communications on non-commutative gauge theory. Y.O. also thanks N. Ohta, M. Sato and
T. Yoneya for valuable discussions and comments. This works of Y.O. and S.T. were
supported in part by the Japan Society for the Promotion of Science under the Postdoctoral
Research Programs No. 11-01732 and No. 11-08864, respectively.
Appendix A. Determination of the O( 0 0 ) part of Gs
In this appendix, we determine the O( 0 0 ) part of Gs , namely the constant t in (3.11), by
the assumptions (2.15). Since the normalizations of Ai and A i do not coincide when t 6= 1
as can be seen from (3.12), we should be careful when evaluating the S-matrix. A safer
approach is to make calculations after rescaling both fields such that their normalizations
i:
coincide. We denote the normalized fields by Ai and A
Ai
Ai = ,
gs

i = Ai .
A
Gs

(A.1)

If we define the field strengths of the normalized fields as


Fij i Aj j Ai ,
p
p
j + i Gs A
i,
j j A
i i Gs A
i A
j A
b
Fij i A

(A.2)
(A.3)

F
b ) (3.8) can be rewritten as follows:
the effective Lagrangians L(B + F ) (3.7) and L(

354

Y. Okawa, S. Terashima / Nuclear Physics B 584 (2000) 329358

L(B + F ) =

p
p

B
det g Tr + F
gs

2

+ O( 0 )

det g Tr F2 + total derivative + O( 0 ),

F
b ) = det G Tr(G1b
L(
F G1b
F ) + O( 0 ).
=

(A.4)
(A.5)

i coincide at the lowest


It is clear from these expressions that the normalized fields Ai and A
0
order in :
i = Ai + O( 0 ).
A

(A.6)

Following the procedure presented in Section 3.1, the F 4 terms can be determined in
this case as well. The evaluation of the Lagrangian on the non-commutative side in the 0
i by
expansion is given in terms of the normalized field A

det G Tr(G1b
F G1b
F)
p
p


i )(j A
i i A
j ) 4(2 0 )2 Gs Bkl k A
i l A
j j A
i
= det g (i Aj j A
k k A
j )(k A
i i A
k)
+ 2(2 0 )2 (B 2 )ij (j A

1
j j A
i )(j A
i i A
j ) + O( 0 4 ) .
(2 0 )2 Tr B 2 (i A
2
The relevant terms on the commutative side are


4
p
B
det g
4
Tr(B + F ) = det g gs Tr + F
gs
gs
p



4
3
= det g gs Tr F + 4 gs Tr BF + O(B 2 ) ,

2 2
 
p

det g 
B
2 2
Tr(B + F ) = det g gs Tr + F
gs
gs
p



2 2
= det g gs (Tr F ) + 4 gs Tr BF Tr F2 + O(B 2 ) .

(A.7)

(A.8)

(A.9)

F
b ) and L(B + F ) should produce the same
The requirement that both Lagrangians L(
3
3
S-matrix of order O(B, , k ) determines the form of the Lagrangian L(B + F ) in a
completely parallel way to Section 3.1. The result is as follows:
L(B + F )

 
2
4

p
p
B
B
1
0 2
Tr + F
= det g Tr + F + (2 ) Gs gs
gs
2
gs
 
2 2 

B
1
Tr + F

+ O( 0 4 ) + derivative corrections
8
gs




det g
2
0 2 1
Tr(B + F )4
Tr(B + F ) + t (2 )
=
gs
2



1
2 2
04
Tr(B + F )
(A.10)
+ O( ) + derivative corrections ,
8
where we used t defined by (3.11). The O(B) part of this Lagrangian coincides with that
of (A.7) up to field redefinition since both produce the same S-matrix, but it may not the

Y. Okawa, S. Terashima / Nuclear Physics B 584 (2000) 329358

355

case for the O(B 2 ) part. The O( 0 2 ) part of the Lagrangian (A.10) expanded with respect
to B is given by

 
4
2 2 

p
p
B
B
1
1
0 2
Tr + F
Tr + F
det g (2 ) Gs gs
2
gs
8
gs



p

1
1
Tr F4 (Tr F2 )2
= det g (2 0 )2 t gs
2
8



1
+ t gs 2 Tr BF3 Tr BF Tr F2
2




1
(A.11)
+ t 2 Tr B 2 F2 Tr B 2 Tr F2 + total derivative + const. .
4
The O(B 2 ) part does not coincide with that of (A.7). The difference in the Tr B 2 Tr F 2 term
can be absorbed by field redefinition as we explained in Section 3.3 so that it is irrelevant
to the determination of t, whereas the values of the Tr B 2 F 2 term evaluated at on-shell
configurations satisfying (3.18) do not vanish so that if there is a difference in the term it
cannot be redefined away. Therefore the Tr B 2 F 2 terms in (A.7) and (A.11) must coincide,
which determines the value of t. The result is
t = 1.

(A.12)

Appendix B. Determination of the O( 0 2 ) part of Gs


As we discussed in the last part of Section 3, the consideration at order 0 2 is not
sufficient to determine the O( 0 2 ) part of Gs and that of the field redefinition which relates
A i to Ai uniquely but allows the following ambiguity:


c1
(2 0 )2 Tr B 2 + O( 0 4 ) ,
(B.1)
Gs = gs 1 +
4
1
c
(B.2)
A i = Ai + (2 0 )2 Bkl Ak (l Ai + Fli ) + (2 0 )2 Tr B 2 Ai + O( 0 4 ),
2
8
where c is an undetermined constant. In this appendix, we determine the value of c by the
consideration at order 0 4 .
We should first note that whatever ordering of the field strengths we choose, the
b 4 terms of (3.29) does not affect O( 0 4 ) terms,
product between the field strengths in the F
namely,
b )4arbitrary = (2 0 )2 Tr(G1 F
bG1 F
bG1 F
bG1 F
b ) + O( 0 6 ),
(2 0 )2 Tr(G1 F
0 2

(2 )


b )2 2
Tr(G1 F
arbitrary

0 2

= (2 ) Tr(G

1 b 1 b

FG

F)

2

(B.3)
06

+ O( ),

(B.4)

where the product between the field strengths on the right-hand sides of these expressions
b4 terms in (3.29) are evaluated in the 0
is the ordinary one, not the product. Now the F
expansion using (3.1), (3.3), (B.1) and (B.2) as follows:

356

Y. Okawa, S. Terashima / Nuclear Physics B 584 (2000) 329358




1
det G 1
0 2
1 b 4
0 2
1 b 2 2
(2 ) Tr(G F )arbitrary (2 ) Tr(G F ) arbitrary
Gs
2
8


1
det g 1
(2 0 )2 Tr F 4 (2 0 )2 (Tr F 2 )2
=
gs
2
8


1
1
1
0 4
5
4
3
2
2 2
Tr BF (Tr F )
+ (2 ) 2 Tr BF Tr BF Tr F Tr BF Tr F +
4
2
16

c1
Tr B 2 Tr F 4
+ (2 0 )4 2 Tr B 2 F 4 +
8

1
1c
2 2
2
2
2 2
Tr B (Tr F )
Tr B F Tr F +
2
32

(B.5)
+ total derivative + O( 0 6 ) .
Since no other terms can produce O(BF 5 ) terms, (B.5) gives the complete form of them
on the non-commutative side. On the other hand, there are three sources for O(BF 5 ) terms
on the commutative side, which are
Tr(B + F )6 = Tr F 6 + 6 Tr BF 5 + O(B 2 ),

(B.6)

Tr(B + F )2 Tr(B + F )4 = Tr F 2 Tr F 4


Tr(B + F )


2 3

+ 2 Tr BF Tr F 4 + 4 Tr F 2 Tr BF 3 + O(B 2 ),
= (Tr F 2 )3 + 6 Tr BF (Tr F 2 )2 + O(B 2 ).

(B.7)
(B.8)

By comparison, we can see that the O(BF 5 ) terms in (B.5) are reproduced by the following
terms in L(B + F ):



3
1
1
(2 0 )4 det g 1
Tr(B + F )6 Tr(B + F )2 Tr(B + F )4 +
Tr(B + F )2 .
gs
3
8
96
(B.9)
These are precisely the terms needed to take the form of the DBI Lagrangian (3.27) under
our normalization convention. To show that this is the unique structure of the F 6 terms
consistent with the assumptions (2.15), we must verify that no solution to the on-shell
initial term condition (4.27) is possible in the F 6 terms. However, even if there exist such
solutions, although we believe that there is none, the resulting ambiguity does not affect the
determination of the O( 0 2 ) part of Gs since solutions to the on-shell initial term condition
by definition do not contribute to the B-dependent part of the S-matrix. Thus the argument
which has been made so far is sufficient for the determination.
Now let us compare the O(B 2 ) part of (B.9),


1
1
1
(2 0 )4 det g
Tr B 2 (Tr F 2 )2
2 Tr B 2 F 4 Tr B 2 Tr F 4 Tr F 2 Tr B 2 F 2 +
gs
8
2
32
+ 2 Tr BFBF 3 + Tr BF 2 BF 2 Tr BF Tr BF 3

1
1
2
2
2
Tr F Tr BFBF + Tr F (Tr BF ) ,
4
8

(B.10)

Y. Okawa, S. Terashima / Nuclear Physics B 584 (2000) 329358

357

with the corresponding one on the non-commutative side. In addition to the O(B 2 ) part
b2 term:
of (B.5) at order 0 4 , there is the following contribution from the F

(2 0 )4 det g 
Tr BF 2 BF 2 + 2Ak Bkl l Fij (FBF )j i + Ak Bkl l Fij An Bnm m Fj i
gs

+ 2Bkl Bnm An (m Ai + Fmi ) l Aj k Fj i + O(B 3 ) + total derivative .
(B.11)
By the comparison with respect to the terms Tr B 2 Tr F 4 and Tr B 2 (Tr F 2 )2 which
obviously contribute to the S-matrix, the value of the constant c is determined. The result
is
c = 0.

(B.12)

To complete the argument of the O(B 2 ) part at order 0 4 , it is necessary to verify that the
difference,


(2 0 )4 det g
2 Tr BFBF 3 Tr BF Tr BF 3
gs
1
1
Tr F 2 Tr BFBF + Tr F 2 (Tr BF )2
4
8
2Ak Bkl l Fij (FBF )j i Ak Bkl l Fij An Bnm m Fj i

2Bkl Bnm An (m Ai + Fmi ) l Aj k Fj i ,

(B.13)

does not contribute to the S-matrix and is absorbed by a field redefinition of A i at order 0 4 .
This is an interesting problem itself since it is related to the O( 2 ) part of (2.17). However
it is easily seen that it is irrelevant to the determination of the constant c at any rate because
none of the terms in (B.13) have the structure of the contraction Tr B 2 = Bij Bj i which was
relevant to the determination.
References
[1]
[2]
[3]
[4]
[5]
[6]
[7]
[8]
[9]
[10]
[11]
[12]
[13]
[14]

E. Witten, Nucl. Phys. B 460 (1996) 335.


M.R. Douglas, D. Kabat, P. Pouliot, S.H. Shenker, Nucl. Phys. B 485 (1997) 85.
T. Banks, W. Fischler, S.H. Shenker, L. Susskind, Phys. Rev. D 55 (1997) 5112.
J.M. Maldacena, Adv. Theor. Math. Phys. 2 (1998) 231.
A. Connes, M.R. Douglas, A. Schwarz, J. High Energy Phys. 2 (1998) 003.
M.R. Douglas, C. Hull, J. High Energy Phys. 2 (1998) 008.
Y.E. Cheung, M. Krogh, Nucl. Phys. B 528 (1998) 185.
T. Kawano, K. Okuyama, Phys. Lett. B 433 (1998) 29.
C. Chu, P. Ho, Nucl. Phys. B 550 (1999) 151; Nucl. Phys. B 568 (2000) 447.
F. Ardalan, H. Arfaei, M.M. Sheikh-Jabbari, J. High Energy Phys. 2 (1999) 016; Nucl.
Phys. B 576 (2000) 578.
V. Schomerus, J. High Energy Phys. 6 (1999) 030.
N. Seiberg, E. Witten, J. High Energy Phys. 9 (1999) 032.
E.S. Fradkin, A.A. Tseytlin, Phys. Lett. B 163 (1985) 123.
C.G. Callan, C. Lovelace, C.R. Nappi, S.A. Yost, Nucl. Phys. B 288 (1987) 525.

358

[15]
[16]
[17]
[18]
[19]
[20]
[21]
[22]
[23]
[24]
[25]
[26]
[27]
[28]

Y. Okawa, S. Terashima / Nuclear Physics B 584 (2000) 329358

A. Abouelsaood, C.G. Callan, C.R. Nappi, S.A. Yost, Nucl. Phys. B 280 (1987) 599.
A.A. Tseytlin, hep-th/9908105.
Y. Okawa, Nucl. Phys. B 566 (2000) 348.
T. Asakawa, I. Kishimoto, J. High Energy Phys. 11 (1999) 024.
O.D. Andreev, A.A. Tseytlin, Nucl. Phys. B 311 (1988) 205.
O.D. Andreev, A.A. Tseytlin, Mod. Phys. Lett. A 3 (1988) 1349.
S. Terashima, J. High Energy Phys. 2 (2000) 029.
L. Cornalba, R. Schiappa, hep-th/9907211.
L. Cornalba, hep-th/9909081.
N. Ishibashi, hep-th/9909176.
K. Okuyama, J. High Energy Phys. 3 (2000) 016.
L. Cornalba, hep-th/9912293.
O. Andreev, Phys. Lett. B 481 (2000) 125.
A.A. Tseytlin, Nucl. Phys. B 501 (1997) 41.

Nuclear Physics B 584 (2000) 359386


www.elsevier.nl/locate/npe

General properties of the self-tuning domain wall


approach to the cosmological constant problem
Csaba Cski a,1 , Joshua Erlich a , Christophe Grojean b,c, ,
Timothy J. Hollowood a,d
a Theory Division T-8, Los Alamos National Laboratory, Los Alamos, NM 87545, USA
b Department of Physics, University of California, Berkeley, CA 94720, USA
c Theoretical Physics Group, Lawrence Berkeley National Laboratory, Berkeley, CA 94720, USA
d Department of Physics, University of Wales, Swansea, SA2 8PP, UK

Received 1 May 2000; accepted 20 June 2000

Abstract
We study the dynamics of brane worlds coupled to a scalar field and gravity, and find that selftuning of the cosmological constant is generic in theories with at most two branes or a single brane
with orbifold boundary conditions. We demonstrate that singularities are generic in the self-tuned
solutions compatible with localized gravity on the brane: we show that localized gravity with an
infinitely large extra dimension is only consistent with particular fine-tuned values of the brane
tension. The number of allowed brane tension values is related to the number of negative stationary
points of the scalar bulk potential and, in the case of an oscillatory potential, the brane tension for
which gravity is localized without singularities is quantized. We also examine a resolution of the
singularities, and find that fine-tuning is generically reintroduced at the singularities in order to retain
a static solution. However, we speculate that the presence of additional fields may restore self-tuning.
2000 Elsevier Science B.V. All rights reserved.
PACS: 04.50.+h; 04.65.+e; 11.10.Kl
Keywords: Brane universe; Localized gravity; Self-tuning; Cosmological constant

1. Introduction
There has been renewed interest in KaluzaKlein theories over the past two years,
mainly due to the realization that localization of matter [1,2] and localization of gravity [3]
may drastically change the commonly assumed properties of such models. These theories
Corresponding author. E-mail: grojean@thsrv.lbl.gov
1 J. Robert Oppenheimer fellow.

0550-3213/00/$ see front matter 2000 Elsevier Science B.V. All rights reserved.
PII: S 0 5 5 0 - 3 2 1 3 ( 0 0 ) 0 0 3 9 0 - 4

360

C. Cski et al. / Nuclear Physics B 584 (2000) 359386

clearly open new approaches to the cosmological constant problem 2 , since using extra
dimensions the no-go theorem of Weinberg for adjustment mechanisms ([4], for some
recent proposals for a 4D adjustment mechanism, see also [5,6]) may be circumvented.
A particularly simple scenario has been recently suggested in [7,8] to at least improve on
the cosmological constant problem (see also [9] for an earlier mechanism involving extra
dimensions), by proposing a mechanism for the cancellation of all order Standard Model
(SM) loop contributions and the leading (tree-level) gravity contribution to the effective
4D cosmological constant,
O(M 4 ) + O(Tbr ) + O(Tbr2 M 4 ) +
| {z }
| {z } |
{z
}

tree-level

SM loops

(1.1)

higher order

where M is the fundamental scale of gravity, the Planck scale in the bulk for instance. The
mechanism is based on a single 3-brane (to which the SM fields are localized) embedded
into a 4+1 dimensional spacetime. The essential new ingredient is a bulk scalar field
, which is coupled to the brane tension Tbr (assumed to be the full loop-corrected SM
vacuum energy density). The authors of [7,8] then showed that one can find static solutions
to the classical equations of motion for a vanishing bulk potential of the scalar field, but for
arbitrary values of the brane tension Tbr . However, all the solutions found in [7,8] which
localize gravity involve a naked spacetime singularity, which has been interpreted as the
boundary of the extra dimension (see also [10,11] for a discussion on singularity in a braneworld context). Since the bulk is effectively compactified, we have to worry whether the
size of the extra dimension is compatible with our phenomenological knowledge of gravity
which has been tested up to millimeter distances. The proper distance from the brane to the
2/3
is the 5D Planck scale, which
singularity is found to be yc 52 Tbr1 where M5 = 5
1
is related to the 4D Planck scale, M4 = 4 , by Tbr = 42 /54 . If the tension, associated to
the vacuum energy of the SM fields, is of the order of the electroweak scale, Tbr TeV4 ,
we naturally obtain 3
M5 108 GeV and yc 1 mm

(1.2)

which is phenomenologically safe. Finally, while the SM contribution to the 4D


cosmological constant would be of the order of 1064 M44 , the self-tuning mechanism
cancels this term. However, the next possible term is of the order of 1084 M44 , which is
still forty orders of magnitude too large. It is worth noticing that the self-tuning mechanism
eliminates twenty orders of magnitude and thus should be considered on equal footing to
the RandallSundrum solution to the hierarchy problem.
The necessity of singularities in the bulk is here only dictated by the phenomenological
requirement of localized gravity. However, it complements the singularity theorem of
2 The cosmological constant problem concerns the brane cosmological constant that governs the expansion of
our universe and it has to be distinguished to the vacuum energy density, or tension, on the brane: in particular it
is possible, for a non-vanishing tension, to find solutions to the Einstein equations with a flat Minkowski metric
on the brane in which case the brane cosmological constant is zero.
3 The fact that the electroweak scale on the brane is five orders of magnitude below the fundamental scale, M ,
5
is the gauge hierarchy problem in this model and a mechanism, such as low energy supersymmetry for instance,
is needed to cancel the quadratic divergences which should bring Tbr near M54 .

C. Cski et al. / Nuclear Physics B 584 (2000) 359386

361

Hawking and Penrose under which generic initial conditions lead to singular solutions
of Einsteins equations [12].
Several works [1315] have studied the self-tuning mechanism. In this paper, we
examine the general properties of the solutions to the coupled 5D scalar-gravity system
in detail. First, we investigate the solutions in the presence of a general bulk potential
for the scalar field. The motivation to include a scalar bulk potential are twofold: (i) to
give a mass to the scalar field and thus evade the cosmological problems associated to a
massless scalar field coupled to gravity and/or those associated to an unstabilized extradimension; (ii) to overcome the singularity problem by considering a more general bulk
geometry. We explain that the self-tuning behavior is expected to occur for a generic bulk
potential. However, in the generic solutions (except for a vanishing bulk potential) selftuning does have the more restricted interpretation that there is at least a finite region for
the brane tension for which static solutions can be found (but not necessarily for all values
of Tbr ). More precisely, Standard Model corrections are expected to occur at the weak
scale, whereas self-tuning works up to the larger 5D Planck scale.
After presenting our general results and methods on the solutions in the presence of a
generic bulk potential we ask the following question: are there bulk potentials such that the
self-tuning solutions avoid naked spacetime singularities for a range of values of the brane
tension in such a way, that gravity is localized to the brane (so as to reproduce 4D gravity
for the observer on the brane). We show that for the 5D scalar-gravity system the only
possibility for such solutions is when spacetime asymptotes to five-dimensional anti-de
Sitter space (AdS5 ) far from the brane (therefore producing solutions of the sort considered
in [18]). After a careful analysis we show that such solutions are always fine-tuned; that
is, they occur only for particular isolated values of the brane tension Tbr . We show that the
number of allowed brane tensions is related to the number of negative stationary points
of the bulk potential, which appears as a maximal index. In the case of an oscillatory
bulk potential, we thus obtain a quantization of the brane tension as in a brane world
construction from supergravity [19,20].
The paper is organized as follows: in Section 2 we present our method for searching
for solutions of the coupled scalar-gravity system using the superpotential formalism
of [18,2128]. The advantage of this method is that it results in first order ordinary
differential equations, and all the degrees of freedom are transparent without having to
make a particular ansatz for the solution. In Section 3 we present a class of exactly solvable
systems, given by an exponential bulk potential, which includes all the models presented
in [7,8]. We analyze these models in detail and find that all of these solutions necessarily
involve a naked singularity or lead to an infinite Planck scale on the brane. We also find
that, by relaxing an ansatz for solutions made in [8], the exponential bulk potential is in fact
self-tuning. In Section 4 we compare a perturbative and a numerical method to solve the
equations for the most general bulk potentials, and present an example for both methods
using the exactly solvable cases of Section 3. In Section 5 we present a no-go theorem
for self-tuning branes that would localize gravity without singularities. In Section 6 we
comment on the effect of resolution of the singularities on self-tuning. We conclude in
Section 7.

362

C. Cski et al. / Nuclear Physics B 584 (2000) 359386

2. Self-tuning and scalar fields


In this section we demonstrate that the 4D cosmological constant of branes coupled to
scalar fields is generically self-tuned to zero, as in [7,8]. We begin with the action for a
brane coupled to gravity and a real scalar field (such dilatonic domain walls have been
extensively studied from a supersymmetric point of view [2931], a new ingredient here is
the localization of SM fields) 4 ,


Z

1 p
D2
D
M
M + V []
S = d x 2 |G| R
D1
2D
Z
p
D2
f [] T ,
(2.1)
dD1 x |g|
2
(D 1)D
where GMN is the D-dimensional metric and g is the pullback of the metric onto the
flat domain wall at x D1 y = 0. For now we will not be concerned with the origin
of the coupling f [] to the brane, and allow it to be arbitrary; T is the non-canonically
normalized brane tension and includes standard model vacuum contributions. We look for
static solutions with the metric ansatz 5 ,
2
+ dy 2.
ds 2 = e2A(y)/(D1)dxD1

(2.2)

The unconventional normalization of the action (2.1) and the warp factor in (2.2) will be
convenient in what follows.
The equations of motion which follow from the action (2.1) with the ansatz (2.2) are
[18,28],
A00 (y) = 0 (y)2 + f [(y)] T (y),
0

A (y) = (y) V [(y)],


1 V [] f []
+
T (y),
00 (y) A0 (y) 0 (y) =
2

(2.3)
(2.4)
(2.5)

where the primes denote derivatives with respect to y. If (y) is monotonic in the bulk then
the bulk equations of motion can be written in a first order form [18,2126,28] introducing
an auxiliary field W [],


W [] 2
W []2 ,
(2.6)
V [] =

W [(y)]
,
(2.7)
0 (y) =
(y)
(2.8)
A0 (y) = W [(y)].
4 Our conventions correspond to a mostly positive Lorentzian signature ( + +) and the definition of the
curvature in terms of the metric is such that a Euclidean sphere has positive curvature. Bulk indices will be
denoted by capital Latin indices and brane indices by Greek indices.
5 We will not look for non-flat solution on the brane such as de Sitter or anti-de Sitter 4D configurations. When
they exist simultaneously with flat Minkowskian solutions, it is a dynamical question to know which solution is
preferred by stability. A nice feature of the case with vanishing bulk potential is that the flat solutions are the only
4D maximally symmetric solutions [7].

C. Cski et al. / Nuclear Physics B 584 (2000) 359386

363

Because of the relation (2.6), W [] will be called superpotential even though no


supersymmetry is involved. We will use the first order formalism in order to construct
solutions in the bulk on both sides of the brane: W , and A on the right (+) and
left () hand side. But the equations of motion (2.6)(2.8) must then be supplemented by
boundary conditions at the brane. The boundary conditions due to the delta-function terms
in (2.3) and (2.5) can be written,

f [(y)]
W [(y)]
0
=T
,
1 = 1
(y)
(y) y=0

(2.9)
1A0 = 1W [(y)] = T f [(y)] y=0 ,
where 1F indicates the jump of a discontinuous function F at y = 0, F (0+ ) F (0 ).
In addition, (y) and A(y) must be continuous across the boundary. Hence, there are four
boundary conditions at the brane.
A count of free parameters in the solutions to the equations of motion immediately
demonstrates that given f [(y)], the tension T is generically not fine tuned. If we do not
impose orbifold boundary conditions in addition to those above, then there are naively six
free parameters: one from the solution on each side of the brane of each of equations (2.6)
(2.8). However, one overall constant shift in A(y) is not relevant because A(y) enters into
the equations of motion only through its derivatives. That leaves five free parameters and
four boundary conditions. There is generically a (finite) line of solutions for a region of
scalar-brane couplings f [] T .
Once again, this is what we mean by self-tuning: Given an arbitrary scalar-brane
coupling f [] (possibly satisfying some constraints), there is a range of brane tensions
T such that static solutions exist. In the generic case, as argued above, there is in fact a
continuous set of static solutions for a given boundary condition specified by f [] and T .
Furthermore, if f [] = f 0 [] then the range of T often extends to infinity. The reason is
that if W []  V [] asymptotically when W is large, then W 0 ' W there, and T can be
chosen arbitrarily large such that (f T , f 0 T ) (2W, 2W 0 ).
Orbifold boundary conditions are more constraining. The additional constraints from the
orbifold condition are A(y) = A(y) and (y) = (y), which implies that W+ [] =
W [], where W [] are the solutions for W on the two sides of the brane. However,
continuity of A(y) and (y) is then guaranteed. Hence, in this case there are two free
parameters (there would be three but a constant shift in A is not relevant) and two boundary
conditions, and there is generically a solution for a region of couplings f [] T . Thus, there
is generically self-tuning in this case, as well.
In the absence of orbifold type boundary conditions, if there are N branes then there are
3N + 3 1 = 3N + 2 free parameters and 4N constraints, leaving 2 N free parameters
in the solution for a given set of boundary conditions. Hence, there can be up to two branes
without fine-tuning. With orbifold boundary conditions, where image branes are included
in N and there is assumed to be a brane at the orbifold fixed point (which contributes 1 to
N ), there are 3(N 1)/2 + 3 1 = (3N + 1)/2 free parameters and 4(N 1)/2 + 2 = 2N
constraints, leaving (1 N)/2 free parameters in the solution. Hence, only if there is a
single brane at the orbifold fixed point will self-tuning occur. If the space is compactified

364

C. Cski et al. / Nuclear Physics B 584 (2000) 359386

on a circle with orbifold boundary conditions, then a parameter count for the case of branes
at the two orbifold fixed points demonstrates that a fine-tuning is necessary in this case,
as well [18]. Namely, there are two free parameters in the solution but four boundary
conditions. In what follows we will concentrate on the case of a single brane coupled
to a scalar field.

3. Integrable bulk potentials


In this section we study some special cases which were also partially discussed in [7,
8]. Our discussion of the exact solutions with a vanishing bulk potential is similar, but
we will not restrict ourselves to the ansatz A0 [] 0 made in [8]. In agreement with the
parameter count in the previous section we will find that there is no fine-tuning in these
theories, although some of the exact solutions exhibit non-generic behavior.
Let us first find the exactly solvable models. The challenge is finding a class of solutions
to the nonlinear equation (2.6) for a given V []. If (in a region where 6 V < 0) we write
W [] and W 0 [] as 7


1p
1
,
(3.1)
W=
V [] w[] +
2
w[]


1p
1
V [] w[]
,
(3.2)
W0 =
2
w[]
then (2.6) is immediately satisfied for any prepotential w[]. The consistency of (3.1) and
(3.2) then translates into a differential equation for w[]:
V 0 2 + 1

.
(3.3)
2V 2 1
Exact solutions for w[] can be found when (3.3) is separable, i.e., when V 0 []/V [] is a
constant.
We distinguish three cases:
a vanishing bulk potential: V [] = 0;
a negative bulk cosmological constant: V [] = < 0;
an exponential bulk potential: V 0 []/V [] = const 6= 0.
0 =

3.1. Vanishing bulk potential


This case has been extensively studied by Refs. [7,8] but it is a worthwhile exercise to
repeat the discussion in terms of a superpotential. The equation for the superpotential can
be solved without introducing a prepotential. Indeed, Eq. (2.6) becomes simply
W 0 []2 = W []2 ,

(3.4)

6 The case of a positive bulk potential can be studied in a very similar way up to some changes of sign in Eqs.
(3.1), (3.2).
7 From now, we will denote by a prime a derivative of V , f , W or with respect to ; or a derivative of or
A with respect to y.

C. Cski et al. / Nuclear Physics B 584 (2000) 359386

365

with two branches of solutions,


W [] = ce .

(3.5)

where c is a constant of integration and  is a sign, both of them can take different values on
the two sides of the brane (the constants of integration relative to the right (left) hand side
of the brane will be denoted with a + () subscript). With this form of the superpotential,
Eqs. (2.7), (2.8) can be easily solved as
(y) =  ln(d cy),

(3.6)

A(y) = ln(d cy) + e,

(3.7)

where d and e are some constants of integration that can also differ on the sides of the
branes. Moreover, to make sense Eq. (3.6) need: d+ > 0 and d > 0. Thus the continuity
requires
(0) 0 = + ln d+ =  ln d ,

(3.8)

A(0) A0 = e+ ln d+ = e ln d ,

(3.9)

while the jump equations are


+ c+  c

= f 0 [0 ] T ,
d+
d
c
c+

= f [0 ] T .
d+ d

(3.10)
(3.11)

From the expression


of the warp factor, we conclude that the Planck scale on the brane,
2
2 R
= D
dy e(D3)A/(D1), is finite iff singularities are encountered on both sides
D1
of the brane, i.e., c+ > 0 and c < 0:
If +  = 1: the consistency of the jump equations requires that f 0 [0 ] = f [0 ] and
then it is possible to find solutions with or without singularities for any value of the
brane tension but the singular solutions correspond to f [0 ] T > 0;
If +  = 1: the solutions with singularities exist only if f [0 ] T > 0 and 1 <
f 0 [0 ]/f [0 ] < 1 but do not require any fine-tuning.
If we choose f [] = Ce for the case  = + =  then it is clear that the boundary
conditions can be satisfied for any T . There are several important comments to make about
this case, which is quite non-generic. First of all, the fact that a specific form had to be
chosen for f [] is a result of two non-generic features of the model: First, there is a
symmetry + const. As a result of this symmetry, one of the free parameters, namely
0 , does not appear on the left hand side of the boundary conditions for the derivatives 0
and A0 . In addition, it turned out that the left hand sides of the two boundary conditions had
the same form up to a sign. As a result, only f 0 []/f [] is relevant, and given one solution
(f []T , f 0 []T ), there is an infinite set of solutions with the same f [] and arbitrary
T . The fact that T is completely arbitrary is most likely unique to the case of vanishing
bulk potential. But the fact that given f, f 0 O(M5 ), self-tuning occurs for T to within
O(M5 )  O(MEW ) (where O(MEW ) is the expected Standard Model contribution to the
tension) is generic, and is what we mean by self-tuning.

366

C. Cski et al. / Nuclear Physics B 584 (2000) 359386

Even given the constraint on f [] from the shift symmetry in in this case, as explained
in [7,8] the required exponential form of f [] might be natural from a stringy perspective
where is interpreted as the dilaton.
Furthermore, as pointed out in [8], the solutions with singularities on either side of the
brane must be chosen in order to have a finite gravitational coupling (assuming that the
spacetime can be cutoff at the singularity in a consistent way). This feature will turn out
to be generic except in theories which admit solitonic solutions for fine-tuned boundary
conditions, i.e., for special discrete values of the brane tension.
3.2. Bulk cosmological constant
Even though the case of a (negative) bulk cosmological constant can be studied without
introducing a prepotential, we will present our method on this rather simple example before
proceeding to the somewhat more complicated case of an exponential bulk potential in
the next subsection. We will denote by V [] = < 0 the cosmological constant. The
equations of motion are:
d
= ,
d
d 1
=
(1 2 ) ,
dy 2
dA 1
=
(1 + 2 ) .
dy
2

(3.12)
(3.13)
(3.14)

The differential equation for the prepotential can be easily integrated,


= c e ,

(3.15)

where c is a constant of integration. Plugging this expression for the prepotential into the
remaining equations of motion, we obtain,

(y) = ln c (y) ,
(3.16)




2




(3.17)
A(y) = ln (y) + ln 1 (y) + a,
where a is a constant of integration and another constant of integration, yc , also appears in
the expression of the function (y) defined, on the two different branches of solutions, by

(y yc ) or (y) = coth
(y yc ) .
(3.18)
(y) = tanh
2
2
The nature of the solution depends on the sign of yc on each side of the brane. On the right
(left) hand side, a positive (negative) value of yc will correspond to a solution involving
a singularity at a finite proper distance from the brane, y = yc . Near this singularity, the
warp factor, exp(2A/(D 1)), goes to zero so the singularity appears as an horizon in
the bulk. Conversely, if yc+ , the value of yc on the right hand side of the brane, is negative
(respectively, yc > 0), then the transverse dimension will be infinitely large; however, near
infinity, the warp factor blows up and the Planck scale on the brane diverges, ruining any
phenomenological relevance.

C. Cski et al. / Nuclear Physics B 584 (2000) 359386

367

The values of the constants of integration are constrained by the continuity and jump

equations (with = tanh yc /2),


0 ln(c+ + ) = ln(c ),
2
| + a+
A0 ln |+ | + ln |1 +


2
2
1 + + 1 +
1

(3.21)


2
2
1 +
1
1

= f 0 [0 ] T .
2
+

(3.22)

(3.20)

= f [0 ] T ,

(3.19)
2
= ln | | + ln |1
| + a ,

A solution to these equations can be found, provided that f 0 [0 ] 6= f [0 ], for any value
of the brane tension such that
4
.
(3.23)
T2 >
f [0 ]2 f 0 [0 ]2
Moreover, singularities will exist on both sides of the brane iff
f [0 ] T > 0 and

1 <

f 0 [0 ]
< 1,
f [0 ]

(3.24)

just as in the case of a vanishing bulk potential.


It is interesting to look at the Z2 symmetric solution for which there are the additional
constraints:
yc+ = yc , c+ = c and a+ = a .

(3.25)

The continuity conditions are automatically satisfied but the consistency of the jump
equations requires a fine-tuning,
4
.
(3.26)
T2
As already discussed in the case with a vanishing bulk potential, this fine-tuning is a
consequence of the translational symmetry, + const, in the theory. As before,
because of this shift symmetry we lose the appearance of a free parameter to adjust in
the jump equations, which leads to a more restricted set of boundary conditions than for
a generic bulk potential V []. The fine-tuning (3.26) is precisely the one appearing in the
RandallSundrum model when the scalar coupling to the brane is a constant:
f [0 ]2 f 0 [0 ]2 =

bk =

D1 2 2
T ,
8(D 2) D br

(3.27)

where bk and Tbr are the canonically normalized quantities [19,20],


bk =

D2
D2
and Tbr =
f [0 ]T .
2
2
2(D 1)D
(D 1)D

(3.28)

The solution constructed by Randall and Sundrum that localizes gravity with an
infinitely large extra dimension corresponds to a limit of the singular solution where the
singularities are pushed to infinity, i.e., = 1. The jump equations then require

(3.29)
f 0 [0 ] = 0 and f [0 ] T = 2 ,

368

C. Cski et al. / Nuclear Physics B 584 (2000) 359386

i.e., the coupling between the brane and the scalar field vanishes and the canonically
normalized brane tension, Tbr , is fine-tuned to (3.27). In this limit, the expressions for
the scalar field and the warp factor simply become

(3.30)
= 0 and A = |y|.
3.3. Exponential bulk potential
The last case that can be solved analytically with our method involves an exponential
potential for the scalar field in the bulk. We will concentrate on negative potential while
the case of positive potential requires some minimal changes. So the bulk potential will be
parametrized by two real numbers a and b:
V [] = a 2e2b .

(3.31)

On each side of the brane, the equations of motion are simply:


2 + 1
d
= b 2
,
d
1
d 1 b
= ae (1 2 ),
dy 2
dA 1 b
= ae (1 + 2 ).
dy
2

(3.32)
(3.33)
(3.34)

The sign of a is not fixed and can be chosen independently on the two sides of the brane,
as we will discuss. The first differential equation can be easily solved to express in terms
of :

b2 /(b2 1)
,
(3.35)
eb = ebc ||b/(1+b) 1 + b (1 b)2
where c is a constant of integration. This last result can be used to obtain a parametric
representation of A and y as functions of :
2
y() = ebc
a

Z
0

Z
A() = A0

d
0

||
b/(1+b) |1 + b (1 b) 2|
d
(1 + b (1 b) 2 )
(1 + 2 )
.
(1 + b (1 b) 2 )

b2 /(b2 1)

(3.36)

(3.37)

On the two sides of the brane, the parameter a can differ by a sign while the initial bound of
integration, 0 , and the constant of integration, c, can take any values compatible with the
continuity conditions. A Z2 symmetric solution will correspond to two different choices of
sign for a but the same values for 0 and c.
Different kinds of solutions can be obtained depending on the value of b and of the range
of integration for the variable :

C. Cski et al. / Nuclear Physics B 584 (2000) 359386

369

b < 1: in that case, dy/d has an integrable singularity at = + but a nonintegrable singularity at = 0 while the singularities of dA/d are both nonintegrable. So we can find solution with or without bulk singularity at finite proper
distance.
Solutions without singularity will be given by (3.36), (3.37) where on the right
(left) hand side of the brane, a has to be chosen positive (negative) and 0 is a
negative initial bound of integration. The parameter will range from 0 to 0 .
It is easy to find the asymptotic behavior of this solution for large |y|, i.e., near
0 :
A ln |y|,
|y|

(3.38)

from where it becomes evident that the singularity free solutions do not localize
gravity because the Planck scale on the brane diverges.
Solutions with singularity will still be given by Eqs. (3.36), (3.37) with a positive
(negative) parameter a on the right (left) hand side of the brane. And the range of
integration goes from a positive initial value, 0 , to +. In that case, y reaches a
finite value yc while the warp factor goes like:
A ln |y yc |,
yyc

(3.39)

which indicates that the singularity is a horizon where the metric on the brane
vanishes. The behavior of the warp factor near the singularity insures that the
2
2 R
= D
dy e(D3)A/(D1), is finite.
Planck scale on the brane, D1
1 < b < 1: in that case the singularities of dy/d at = , =

(1 + b)/(1 b) and = 0 are integrable while those appearing in dA/d are


non-integrable. All the solution are singular with a horizon or a curvature singularity.

Fig. 1. Shapes of the warp factor for singular and non-singular solutions to the equations of motion
with an exponential bulk potential. We have drawn Z2 symmetric solutions for different values of the
tension on the brane: when the tension increases, the horizon becomes closer and closer to the brane
for the singular solutions while the warp factor, e2A/(D1) , blows up faster and faster for the the
singularity-free solutions. In both cases, the jump conditions require that the scalar coupling to the
brane satisfies: |(df/d)/f | < 1 on the brane.

370

C. Cski et al. / Nuclear Physics B 584 (2000) 359386

1 < b: this case is quite similar to the first case because the singularity of dy/d at
= 0 is integrable but the singularity at = + is non-integrable while the two
singularities of dA/d are both non-integrable. So we can construct solutions with or
without singularity.
Solutions without singularity will be given by (3.36), (3.37) where on the right
(left) hand side of the brane, a has to be chosen negative (positive) and 0 is a
positive initial bound of integration. The parameter will range from 0 to +.
It is easy to find the asymptotic behavior of this solution for large |y|, i.e., near
+:
A ln |y|,
|y|

(3.40)

from where, once again, it becomes evident that the singularity free solution does
not localize gravity.
Solutions with singularity will still be given by Eqs. (3.36), (3.37) with a negative
(positive) parameter a on the right (left) hand side of the brane. And the range of
integration goes from a negative initial value, 0 , to 0 . In that case, y reaches a
finite value yc while the warp factor goes like:
A ln |y yc |,
yyc

(3.41)

which indicates that the singularity is a horizon where the metric on the brane
vanishes.
b = 1: in these two cases, we can construct singularity free solutions with a blowing
up warp factor as well as singular solutions with a horizon at a finite proper distance.
Fig. 1 illustrates the different types of solutions.
The main result of the analysis of these integrable bulk potentials is that we find two
kinds of solutions: (i) solutions with a horizon in the bulk at a finite proper distance from
the brane and with a finite lower dimensional Planck scale; (ii) solutions without singularity
at a finite proper distance but associated to a bulk geometry that decouples gravity on the
brane. Both kinds of solutions do not require any fine-tuning and can be constructed for a
range of brane tension, T . However it seems impossible to find a singularity free solution
that localizes gravity on the brane, unless the brane tension is fine-tuned as in the Randall
Sundrum model. This point surely deserves further scrutiny which the next sections will
be devoted to.

4. Perturbative and numerical methods


In this section we give two methods for finding approximate solutions to the coupled
equations of motion in the bulk that satisfy the boundary conditions at the brane. First, we
present a perturbative method, which we then show to break down for the interesting case
of perturbing around a solution with localized gravity. Then we give a systematic numerical
method for solving the equations.

C. Cski et al. / Nuclear Physics B 584 (2000) 359386

371

4.1. A perturbative method


As emphasized in the previous section, the key to finding the self-tuned solutions is to
utilize the fact that one can transform the superpotential without changing the bulk potential
for the scalar field. Thus, one has an additional degree of freedom if one picks a different
superpotential function on the two sides of the brane, W+ [] and W []. Let us now
assume that we have found a static solution to the equations of motion 0 (y) and A0 (y),
and assume that we have chosen one of the integration constants such that A0 (0) = 0.
This is a solution for a particular value T of the brane tension. In order to find the solution
obtained by perturbing the brane tension as T T + T , we first need to find how one can
change the superpotential W around W0 [] so as to leave the bulk potential unchanged:
V [] = W00 []2 W0 []2 ,

(4.1)

which should be invariant under W W + W . Linearizing (4.1) around W0 we obtain a


differential equation for W of the form
W 0 W0
= 0,
W
W0
which is solved by
W [] = C exp

(4.2)
Z

W0 []
d,
W00 []

(4.3)

where the arbitrary constant C yields the extra degree of freedom needed to find a
solution for any value of T . The superpotential variation can be expressed in terms of
the unperturbed background solution using the equations of motion d0 /W00 [0 ] = dy and
W0 [0 ] = dA0 /dy, we obtain
Z
dA0
= CeA0 (y).
(4.4)
W [(y)] = C exp dy
dy
From (4.3) we can see that the change in the derivative of the superpotential is given by
W 0 [(y)] = C

A00 (y) A0 (y)


e
.
00 (y)

(4.5)

In order to satisfy the jump equations (2.9) for the perturbed tension T + T , we need to
choose C differently on the two sides of the brane. The values of C are then given by
C =

0 (0)
f 0 [0 ] f [0 ]A00 (0)/0

1(A00 /00 )

T ,

(4.6)

where 1 (A00 /00 ) denotes the jump in A00 /00 at y = 0. Once C is determined from (4.6),
we can simply integrate the equations
0 (y) = W 0 [0 (y)],

A0 (y) = W [0 (y)]

(4.7)

to obtain the perturbed solutions 0 (y) + (y) and A0 (y) + A(y). This method always
results in a perturbed solution. However, in the most interesting case, when the unperturbed
solution asymptotes to AdS space thereby localizing gravity to the brane (that is for A(y)

372

C. Cski et al. / Nuclear Physics B 584 (2000) 359386

(D 1)|y|/RAdS for large values of y), one can easily see that the perturbative method
presented here always breaks down. This can be seen by inspecting (4.4), which shows
that in this case W [(y)] e(D1)|y|/RAdS . Therefore the perturbed values of and A
grow exponentially, and thus the linearized approximation breaks down. We will examine
the case of localized gravity in detail in Section 5. But first we give a numerical method
that can be used for any choice of the bulk potential.
4.2. A numerical method
Next we present a method for solving the system (2.6)(2.8) numerically for any brane
tension and potential in the bulk. Thus the input functions are the bulk potential V [], the
brane tension T , the coupling of the scalar to the brane determined by the function f [],
and in addition we can pick the value of the scalar field at the brane (0) 0 arbitrarily.
In order to find a numerical solution to these equations, one has to first make sure that the
boundary conditions that one imposes do satisfy the jump equations (2.9). Our strategy is
the following: we first determine the superpotential functions W+ [] and W [] to the left
and the right of the brane numerically such that the boundary conditions arising from the
coupling to the brane are satisfied. This can be done by noting that once 0 is fixed, the
jump equations are just given by
W+ W = f0 T ,

W+0 W0 = f00 T ,

(4.8)

where W refers to the values of the superpotential functions to the right and left of the
brane at 0 , f0 = f [0 ], etc. In addition, the superpotential functions must be such that
they reproduce the correct value of the bulk potential at the brane:
W+0 W+2 = V0 ,
2

W0 W2 = V0 ,
2

(4.9)

where V0 = V [0 ]. Eqs. (4.8) and (4.9) together are enough to determine the values of both
W and W0 at the branes. They are given by the expressions:
s
1
1 0
4V0
,
W = f0 T + f0 T 2 + 2
0
2
2
f0 f02
s
1 0
1
4V0
0
.
(4.10)
W = f0 T + f0 T 2 + 2
0
2
2
f f2
0

Due to the quadratic nature of Eqs. (4.8) and (4.9) there is a second solution, where the
signs in front of the square roots in (4.10) are both simultaneously flipped. Once the value
of W and W0 are fixed, one can numerically integrate the equation 8
q
W0 [] = sgn(W0 ) V [] + W []2
(4.11)
8 We will see in Section 5 that, unless the brane tension is fine-tuned, the superpotential, W [], will be a
0 [] will keep the sign of W 0 .
monotonic function of and thus W

C. Cski et al. / Nuclear Physics B 584 (2000) 359386

373

to obtain the superpotential functions to the left and the right of the brane that satisfy all
boundary conditions. Once W [] are numerically known, we can simply integrate the
equations
0 (y) = W 0 [(y)],

A0 (y) = W [(y)]

(4.12)

to the left and the right of the brane to obtain the numerical solutions for (y) and A(y).
We will show an example for this below for the case when the exact solution is known, and
compare the two results.
4.3. An example for the perturbative method
In this section we test how well the perturbative method described in Section 4.1 works.
We will compare the analytic solution of the model with a vanishing bulk potential to
the perturbed solution around a different analytic solution. The main conclusions are as
expected: the perturbative method works well far from the singularities. However it gets
worse as we approach the singularity itself, and does not capture the essential feature of the
self-tuning solutions: whereas in the self-tuning mechanism, for a fixed value of the scalar
field on the brane, the place of the singularity adjusts itself with respect to the value of
the brane tension, here the singularity of the perturbed solution remains at the same place
where the singularity of the unperturbed solution was. Therefore, we conclude that this
method is not very efficient in capturing the basic properties of the self-tuning solutions.
The example we consider is the vanishing bulk potential discussed in Section 3.1, with
the choice + = 1,  = 1, and we choose f [] = e/2 , 0 = 1, and for the unperturbed
solution we choose T = 1, while we pick T = 0.1 for the perturbed solution. The analytic
solution is given by (3.6), (3.7), with
d = e0 ,

3
c+ = T e0 /2 ,
4

1
c = T e30 /2 ,
4

e = 0 ,

(4.13)

Fig. 2. The exact solution versus the perturbed solution for the case of a vanishing bulk potential. The
curve that blows up at smaller values of y corresponds to the exact solution. The initial unperturbed
solution is given by the dashed curve. The singularity of the perturbed solution appears at the same
distance from the brane as the singularity of the initial solution; the shift of the singularity with a
variation of the tension is missing.

374

C. Cski et al. / Nuclear Physics B 584 (2000) 359386

where A has been normalized to zero on the brane. The perturbed solution from (4.7) is
given by




d + c+ y
d c y
T
T
ln
ln
,
(y) =
,
(4.14)
+ (y) =
T
d+
T
d
where has been normalized to zero on the brane. The perturbed solution obtained for
T = 1, T = 0.1 compared to the exact solution for T = 1.1 can be seen in Fig. 2. As
mentioned above, the perturbative solution nicely follows the exact solution away from
the singularity, but deviates from it close to the singularity, in particular the place of
the singularity is incorrectly predicted to coincide with the singularity of the unperturbed
solution.
4.4. An example for the numerical method
We have shown above, that the perturbative method in general does not do a good job in
finding the solutions, since it becomes unreliable close to the singularities. However, the
numerical solution should not have these problems. Indeed, we analyze the same example
as above (the case with vanishing bulk potential) using the numerical method, and find that
the exact and numerical curves are virtually indistinguishable. Therefore, we suggest that
in order to analyze potentials for which no exact solutions can be found, one should use
the numerical method rather than the method based on perturbations.
We are looking for a numerical solution to the case analyzed perturbatively above, that
is vanishing bulk potential, f [] = e/2 , 0 = 1, T = 1 and + = 1,  = 1. From Eqs.
(3.4) and (3.8) we find the starting values of the superpotential to the left and the right of
the brane:
3
W+ = e0 /2 T ,
4

1
W = e0 /2 T .
4

(4.15)

Numerically integrating the equation


W0 [] = W []

(4.16)

with the boundary conditions W [0 ] = W one obtains the numerical values for W ().
Finally, the values for (y) can be obtained by numerically inverting the integral
Z
y=

d
W 0 ()

(4.17)

to the left and right of the brane. The numerical solution obtained this way overlayed on
the exact solution of Section 2 can be seen in Fig. 3. One can see that the two curves are
virtually indistinguishable, suggesting that the numerical method works very well around
the singularities, and should be the preferred method of looking for solution in the absence
of exact solutions.

C. Cski et al. / Nuclear Physics B 584 (2000) 359386

375

Fig. 3. The numerical solution overlayed on the exact solution for the case of a vanishing bulk
potential. The fact that the two curves are indistinguishable shows that the numerical method works
extremely well even close to the singularities.

5. Localized gravity without singularities


5.1. A no-go theorem
We would like to reexamine in this section the count of free parameters versus finetuning parameters needed to preserve a static Poincar invariance on the brane, with the
restriction that the solution corresponds to an infinitely large extra-dimension (without
singularities) in the bulk and localizes gravity on the brane.
Consider the general D-dimensional background preserving PoincarD1
2
ds 2 = e2A(y)/(D1)dxD1
+ dy 2.

(5.1)

The graviton zero-mode is localized on the brane 9 precisely when the effective Planck
scale is finite on the brane [28]. In terms of the warp factor, this condition is:
Z
1
1
=
(5.2)
dy e(D3)A(y)/(D1) < .
2
2
D1
D
It is convenient to introduce a transverse coordinate z for which the bulk metric is
conformally flat:

2
+ dz2 ,
(5.3)
ds 2 = 2 (z) dxD1
where the conformal factor, , is related to the warp factor, e2A(y(z))/(D1), by the two
identities:
(z) = eA(y)/(D1) and 2 (z) dz2 = dy 2.

(5.4)

Then the condition (5.2) is equivalent to having a massless normalizable bound state, which
is interpreted as the graviton on the brane:
Z
dz |0 |2 < .
(5.5)
0 (D2)/2 with
9 We do not consider the recently proposed possibility that gravity might be quasi-localized to the brane [33
36], since in those models the A00 > 0 condition is not satisfied, therefore it is not possible to generate those
backgrounds from a single scalar field.

376

C. Cski et al. / Nuclear Physics B 584 (2000) 359386

Let us assume that the behavior of 0 at infinity is a power law:


0 z .
z

(5.6)

The localization of gravity (5.5) then requires: > 1/2.


In our study, the value of the parameter is constrained by the fact that the background
is created by a scalar field coupled to gravity. From the equations of motion, we easily
deduce that A has to satisfy: d2 A/dy 2 > 0, which translates, in the z coordinate, in a lower
bound on the value of :
D2
.
(5.7)
>
2
Furthermore, it is worth noticing that an upper bound on comes by the requirement
10
of a geometry without singularity at a finite
R proper distance. Indeed the proper distance
from the brane to infinity is given: l = dz that diverges iff:
D2
.
(5.8)
2
So the only background for the scalar field coupled to gravity that localizes gravity
without singularity (with an infinitely large extra dimension) is asymptotic, at infinity,
to the horizon of an anti-de Sitter space, as in the RS model, and corresponds to =
(D 2)/2. In that case, the warp factor is exponentially decreasing with the proper distance
to the brane:
6

A (D 1)|y|/RAdS.
|y|

(5.9)

The aim of this section is to show that such a background necessarily requires a finetuning between the brane and the bulk. This is not to say that there is no self-tuning in these
models, only that the nonsingular solutions require fine-tuning.
The previous asymptotic behavior has a nice interpretation in terms of the superpotential,
W []. According to the equations of motion, the fact that A is asymptotically linear means
that becomes constant and we will denote by c and c+ the asymptotic values of at
y = and y = +, respectively.
The equation
dW
=
,
y
d

(5.10)

is similar to an RGE with W 0 playing the role of the -function. In order for to approach
a constant (fixed point) at infinity, the -function dW/d must have zeroes. In other words,
in order to extend the range of the transverse coordinate from y = to y = +, the
values c and c+ at infinity must be some roots of dW/d:


dW
dW
= 0 and
=0.
(5.11)
d c
d c+
10 For a power law conformal factor, the curvature always vanishes at infinity. However quadratic invariants such
as RM1 M2 M3 M4 R M1 M2 M3 M4 will be singular at infinity as soon as > (D 2)/2.

C. Cski et al. / Nuclear Physics B 584 (2000) 359386

377

Furthermore, at infinity, A has to be linearly increasing in |y|; otherwise the conformal


infinity of AdS would be reached without localized gravity [19,20]. Given that
A
= W [],
y

(5.12)

we conclude that c and c+ must satisfy,


W [c ] < 0 and W [c+ ] > 0.

(5.13)

Finally, c and c+ must be dynamically reached at y = and y = +, which


according to (5.10) is possible iff:


d2 W
d2 W
>
0
and
< 0,
(5.14)
d 2 c
d 2 c+
or at least a similar condition for the first non-vanishing higher order derivatives at c and
c+ .
Pictorially, the previous conditions are summarized in Fig. 4.
Finally, the last equation of motion that relates the superpotential, W , to the scalar
potential in the bulk, V , partially fixes the possible values of c and c+ . Indeed, this
differential equation evaluated at infinity gives:
V [c ] = W [c ]2 < 0 and V [c+ ] = W [c+ ]2 < 0.

(5.15)

while a differentiation with respect to gives:






dW d2 W
dV
dV
dV
=2

W
;
thus
=
0
and
= 0.
d
d d 2
d c
d c+

(5.16)

More information on W can be obtained by considering higher order derivatives of


the differential equation between W and V . Indeed it is easy to prove by induction the
following relation:

Fig. 4. Asymptotic behaviors of W [] leading to a singularity free bulk geometry localizing gravity
0

on the brane. The absence of singularities is equivalent to the conditions: W c = W c = 0; the


00
00 +
dynamics of the equations of motion require W c > 0 and W c < 0, while the localization of

+
gravity requires Wc < 0 and Wc > 0.

378

C. Cski et al. / Nuclear Physics B 584 (2000) 359386

(n)



n
X

n1
=
2
W (k) W (nk+2) W (nk) ,
k1

(5.17)

k=1

where V (n) denotes the nth order derivative of V and similarly for W (n) ; in addition, we

will denote by W (n) c the values of W (n) at = c . At the second order, by evaluating

(5.17) at = c , we obtain a quadratic equation for W (2) c :


2

2W (2) c 2Wc W (2) c V (2) c = 0.

(5.18)

The superpotential will be real-valued provided that:

V (2) c > Vc /2,

(5.19)

and then there are four different branches at each asymptotic point:
q
(5.20)
Wc = 1 Vc ,
q
p

2V (2) c Vc
Vc
(2)
+ 2
,
(5.21)
W c = 1
2
2
with 1 = 2 = 1.
However, the compatibility between the gravity localization requirement (5.13) and the
dynamics of the differential equation (5.14) can be fulfilled iff

V (2) c > 0,

(5.22)

and only one out of the four branches is retained:


c :
c+ :

Wc = Vc ,
Wc+ =

Vc+ ,

q
p

2V (2) c Vc
Vc
(2)
+
,
W c =
2
2
q
p
+
2V (2) c Vc+
Vc+
(2) +

.
W c =
2
2

(5.23)
(5.24)

At higher order, the relation (5.17) becomes linear in W (n) c and allows to compute

W (n) c recursively in terms of lower derivatives:



P
n1
(k)
(nk+2) W (nk) +2(n1)W (2) W (n2)
V (n) c n1
c
c
c
c
k=3 2(k1)W c W
(n)

.
(5.25)
W c =
(2)

2 nW

Wc

Of course, for this expression to make sense for any integer n > 2, it is important that

no solution to the equation nW (2) c Wc = 0 can be found. However, the roots of this
equation would satisfy:



2
1
V (2) c
=
1
,
(5.26)
n
n
Vc
which is incompatible with the physical requirements (5.15) and (5.22) of localized gravity.
At this stage, it is worth noticing that the uniqueness of a superpotential W reaching
a given asymptotic point follows from our requirements of a localized gravity without

C. Cski et al. / Nuclear Physics B 584 (2000) 359386

379

singularity. Otherwise, a scalar field V in the bulk could be constructed such that the

equation nW (2) c Wc = 0 has some solution in which case W (n) c would not be
determined, leading to a continuum of solutions.
Finally, the whole expression of W reaching the asymptotic points c can be uniquely
reconstructed from its derivatives through a Taylor expansion:
y <0:
y >0:

W [] =
W+ [] =

X
1 (n)
W c ( c )n ,
n!

(5.27)

1 (n) +
W c ( c+ )n .
n!

(5.28)

n=0

X
n=0

From the expression of the superpotential, we can solve the equations of motion on the two
sides of the brane. And the jump conditions are given by:
W+ [0 ] W [0 ] W+0 [0 ] W0 [0 ]
=
= T.
f [0 ]
f 0 [0 ]

(5.29)

The first equation will fix the value, 0 , of the scalar field on the brane while the second one
fine-tunes the value of the brane tension. Generically, we will obtain only discrete values of
0 . Indeed, if there exist a continuum interval of solutions for 0 , then the jump equation
becomes a differential equation that can be integrated on this interval and we obtain that
f [] has to be proportional to W+ [] W [], but in that case there is only one possible
value for the brane tension. 11 In all the other cases, given two asymptotic points, c , we
will obtain discrete solutions for 0 and T . Moreover, using different critical asymptotic
points, we usually find different values of the brane tension and the number of values of
T that allows an infinitely large extra dimension with localized gravity is related to the
number of critical points of the bulk potential V satisfying (5.15), (5.16) and (5.22):
nT 6 (3nC 2)nS ,

(5.30)

where nT stands for the number of values of T such that gravity is localized without a
singularity, while nC is the number of critical points of V as defined above. The multiplicity
factor comes from the fact that the scalar field can asymptote either the same critical point
or two adjacent ones on the two sides of the brane. The upper bound has been semiquantitatively corrected by taking into account the average number, nS , of solutions to
the jump equation for 0 . It may also happen that some values of T are degenerate. We
will explicitly describe an example in the next subsection.
It is worth noticing that in the case of an oscillatory bulk potential, we will obtain an
infinite number of discrete values for the brane tension that is as quantized.
To complete the proof of the no-go theorem presented above, we need to show that there
is no loophole in the above argument due to the fact that we have used the superpotential
11 Whereas the non-canonically normalized brane tension, T , is fine-tuned, the physical brane tension, T
br
f [0 ]T , is not fine-tuned since the value of 0 can vary continuously. Whether this is a solution to the
cosmological constant problem or not, beside the fine-tuning requires on f [], depends whether SM loops will
modify T or Tbr . This issue deserves further analysis in future work.

380

C. Cski et al. / Nuclear Physics B 584 (2000) 359386

formalism. The subtlety that one might worry about is that in the proof above we have
implicitly assumed that is monotonic, by writing the second order equations (2.3), (2.4)
in terms of first order equations involving W . In particular, if is not monotonic (that is if
0 = 0 at a finite value of y) one does not have a globally defined superpotential function
W (), but instead one must define separate superpotential functions Wi for the regions
between yi and yi+1 , where 0 (yi ) = 0 (yi+1 ) = 0. For these superpotentials that are not
globally defined it is then possible to have W 0 ( ) = 0 without satisfying V 0 ( ) = 0.
However, it is impossible to continue the solution beyond . This by itself however may
not be a problem, as long as at one is smoothly switching over to another branch of
W . Of course this switch-over can only happen at a point where 0 = 0, since otherwise
is monotonic, and one can solve the equations in terms of the superpotential, which is
well-defined around . Next we show that such possibilities do not get around the nogo theorem presented above. The reason is that in order to have localized gravity with an
infinitely large extra dimension, we need 0 0, 00 0 for |y| , and therefore we
find from the second order equation (2.5) that V 0 0. Thus the critical points at infinity
must belong to an ordinary branch described above, where W can be continued at both
sides of c . However, once we are on an ordinary branch which can be globally defined,
all the critical points will actually happen at V 0 = 0. Since the only possibility for switching
over to another branch is at W 0 = 0, and at those points V 0 = 0, one can never switch off
the ordinary branch, and therefore one can not circumvent the no-go theorem by gluing
non-monotonic s together.
5.2. A numerical example
In this section we demonstrate the ideas of the previous sections in a numerical example.
The bulk potential in this example (Fig. 5) is
V () = 6 + 11 4 7 2 + 1,

(5.31)

which is generated by the superpotential W [] = 3 .


As in some of our previous illustrative examples the potential (5.31) is unbounded from
below, and so the theory might be unstable to quantum fluctuations. However, we will
only be concerned with static solutions and for our purposes any instabilities will not be
relevant.
As emphasized
in Fig. 5, the bulk potential (5.31) has two negative stationary points at

= 1/ 3. According to our no-go theorem, there


are isolated superpotentials solving
the equations of motion with critical points at 1/ 3. Here these superpotentials are very
simple because from (5.25) only a finite number of derivatives are non-vanishing thus the
superpotentials are polynomial and take the form

(5.32)
W [] = 3 reaches c = 1/ 3 on the negative (positive) branch,

W [] = 3 + reaches c = 1/ 3 on the positive (negative) branch. (5.33)


Depending on which critical point we want to asymptote at infinity, we can construct four
types of solutions that localizes gravity

C. Cski et al. / Nuclear Physics B 584 (2000) 359386

381

Fig. 5. The scalar bulk potential (5.31). The physically interesting region, near the two negative
stationary points, is emphasized.

For (c+ = 1/ 3, c = 1/ 3 ): the solution is




1
(y) = tanh 3 y yc ,
3

 2

1
tanh2 3 y yc + ln cosh 3 y yc + a ,
A(y) =
18
9

(5.34)
(5.35)

where yc and a are four constants of integration to be determined by the continuity


and jump conditions. The continuity conditions imply that yc+ = yc and a+ = a .
The jump equations will depend on the precise form of f []. Except in the degenerate
case where f [] W+ [] W [] that has been discussed in footnote 11, we will
generically obtain a finite number of discrete values for 0 . For instance in the case
of an exponential coupling, f [] = a eb , the values of 0 will satisfy
302 1
0 (02 1)

= b,

(5.36)

which admits three solutions whatever the value of b is. And thus there exist three
values for the brane
tension that

will lead to a localized


gravity;
For (c+ = 1/ 3, c = 1/ 3 ) or (c+ = 1/ 3, c = 1/ 3 ): in those cases,
there is no discontinuity
in the
superpotential and the brane tension has to vanish;

For (c+ = 1/ 3, c = 1/ 3 ): this case is analogous to the first one and, for an
exponential coupling, three values for the brane tension are possible and they are just
the opposite of the ones obtained in the first case.
Besides the previous fine-tuned solutions with an infinitely large extra dimension and
localized gravity, all other solutions will either decouple gravity on the brane or involve
a horizon at a finite distance. We would like now to examine numerically the self-tuning
of these singular solutions. To do that we scan for solutions to (2.6) by fixing the value of
W [] at = 0. The solutions are plotted in Fig. 6. Note that in agreement with the no-go
theorem of the previous section there is a unique solution in this branch of solutions with
stationary points at the positions of the minima of V []. As a result of the uniqueness of

382

C. Cski et al. / Nuclear Physics B 584 (2000) 359386

Fig. 6. Solutions for W [], (y) and A(y) in the bulk potential 5.31. There is a unique, fine-tuned,
regular solution which approximates the RandallSundrum solution for large |y|. All other solutions
are singular.

the solitonic solution, there are no solutions (in this branch) with 0 < |W [0]| < W [ ] '
0.385, where is the value of the field at the position of the local minimum of V [], in this
case ' 0.577. Also in agreement with the results of previous sections, the solitonic
solution is the unique regular solution for (y) and A(y), which in this case approximates
the RandallSundrum solution for large |y|.
In order to determine the range of boundary conditions at the brane which could be
satisfied, we note that except for the solitonic solution, all other solutions span an infinite
range in . As noted in Section 4.2 the boundary conditions can be expressed in terms
of W [] and W 0 [] at the boundary. Hence, if there is a space-filling range of solutions
(W [], W 0 []) then no fine-tuning is necessary in order to have a static solution with
arbitrary boundary conditions in that range. In other words, a given (2W [0 ], 2W 0 [0 ]) can
be equated with a certain (orbifold) boundary condition (f [0 ] T , f 0 [0 ] T ). Then, given
one such (f [0 ] T , f 0 [0 ] T ) there is a continuous set of solutions around that point. That
means that for a given (f [0 ], f 0 [0 ]) in a certain range of f 0 /f there is a range of tensions
T for which there is a solution; hence the theory is self-tuning. Fig. 7 is a parametric plot
of W [] versus W 0 [] for several solutions of W []. It is intended to illustrate the fact that
there is a space-filling region for W > 0.385. Given any boundary condition parametrized
by f [] and T , T can be rescaled by an amount given by the intersection of the set
of solutions (2W, 2W 0 ) with a line through the origin and (f T , f 0 T ). As is generically
expected, one can see from Fig. 7 that if f, f 0 are O(1) = O(M5 ), then T can be rescaled
by O(M5 ) without eliminating a solution. Hence we have demonstrated the self-tuning of
this model. However, the caveat is that because of the isolated solitonic solution, there is
a fine-tuned region near W, W 0 O(1). If f is O(1) at some matching point 0 , and T

C. Cski et al. / Nuclear Physics B 584 (2000) 359386

383

Fig. 7. Parametric plot of several solutions for (W [], W 0 []), which specify the boundary conditions
at the wall.

is O(MEW /M5 )  1 then a fine tuning is reintroduced because of the uniqueness of the
solitonic solution. There is still a large region of parameter space where the theory is selftuning, but whether that region is natural or not requires exploration. This phenomenon
is a result of the existence of solitonic solutions, and is non-generic. Furthermore, if we
do not require orbifold boundary conditions, then once again self-tuning is natural. Note
also that asymptotically the parametric plots of W vs. W 0 include the line W = W 0 . This
is generic in any region where the solutions satisfy W []  V []. This also implies that
for any solution f f 0 , there is a very large range for T for which there are solutions.
This behavior for large values of W is common, and extends the range of T over which
self-tuning occurs possibly to if f [] = f 0 [], which may be natural from a stringy
perspective [7,8].

6. Resolution of singularities and fine-tuning


In this section we reconsider the resolution of singularities 12 as proposed in [16,17,
32]. As we have seen, the case of zero bulk potential, which is the case studied in [32], is
quite non-generic. There is a shift symmetry in the scalar field which makes the boundary
conditions more constraining, and two of the boundary conditions turn out to have the same
form. Hence, we study the generic case, and propose a new resolution of the singularity
which may restore self-tuning.
The idea of [16,17,32] is to add a brane at each of the singularities such that the equations
of motion, or boundary conditions, are satisfied there, as well. As pointed out in [1317,
32] 13 the singularities contribute to the effective 4D energy density when integrated over
the extra dimension, and the contribution of the singularities is essential for vanishing of
the 4D cosmological constant.
12 As they stand in our solutions, Einsteins equations are not satisfied at the singularity but require a singular
stress-energy tensor located at the singularity that may correspond to a brane: the introduction of this brane, for
instance, is what we mean by resolution of the singularity.
13 See also Refs. [1315] for a discussion on vanishing 4D effective vacuum energy.

384

C. Cski et al. / Nuclear Physics B 584 (2000) 359386

However, addition of branes at the singularities adds new boundary conditions. In order
to answer the question of whether or not the theory is self-tuning we must better understand
the continuation past the singularities. In [32] it is proposed that the spacetime is either
periodically continued or cut-off at the singularity. In the case that the spacetime is cut-off
at the singularities on each side, there are generically two new boundary conditions at each
singularity, but no additional free parameters. Hence, without orbifold boundary conditions
there are 3 + 3 1 = 5 free parameters and 4 + 2 + 2 = 8 boundary conditions, and the
system is overconstrained. If one imposes orbifold boundary conditions, then there are
only half as many additional boundary conditions (2 + 2 = 4 boundary conditions in all),
but only 3 1 = 2 free parameters. In either case the system is overconstrained. Hence, a
fine-tuning is required. Although in the absence of a bulk potential the situation is modified
because of the non-generic features mentioned above, in that case a fine-tuning at each of
the singularities is required, as well.
Although the details of any continuation of the spacetime beyond the singularity
will depend on quantum gravitational dynamics, we propose another scenario which
does slightly better than the cut-off scenario, although fine-tuning will still be required.
The singularities are generically horizons (cf. Section 3.3) because the warp factor
e2A(y)/(D1) vanishes there. Hence, light does not cross the horizon and we can imagine
a scenario in which there are bulk fields living beyond the singularity, out of causal contact
with our universe. The Penrose diagram for this scenario is illustrated in Fig. 8. However,
the bulk fields across the horizon would provide an additional three free parameters which
one might hope would help avoid fine-tuning. More precisely, following the counting in
Section 2, if there are orbifold boundary conditions, then there are 3 + 3 1 = 5 free
parameters from the bulk fields on both sides of the singularity, but 2 + 4 = 6 boundary
conditions. Hence, although the system is less constrained than the case in which the space
time is cut-off at the singularity, one fine-tuning is still required.
One might suspect that additional fields would add additional degrees of freedom which
could help in self-tuning. For example, the addition of a scalar field with second derivatives
in its equation of motion would contribute two additional free parameters on each side

Fig. 8. Penrose diagram for causally disconnected regions glued together at the resolved singularity.

C. Cski et al. / Nuclear Physics B 584 (2000) 359386

385

of a brane, but only two boundary conditions (continuity and change in the derivative
at the brane). Hence in a system of branes there are net, at least naively, two additional
free parameters from the extra scalar field, which could be used to restore self-tuning at
the singularities or perhaps even produce nonsingular solutions. However, a more detailed
analysis is required in this case.

7. Conclusions
We have studied brane worlds coupled to a scalar field and have found that self-tuning is
a generic feature of these models. In these models the dynamics of the scalar field provides
additional degrees of freedom, which generically alleviates the need for fine-tuning of static
solutions. We have reexamined the exactly solvable models, two of which were studied
previously [7,8], and have found that those case are more constrained than the generic case
because of a shift symmetry in the scalar field in these models. Still, these theories are
self-tuning, including the case of an exponential potential. Whereas in [8] a fine-tuning
was necessary in this case due to a particular ansatz for the scalar and graviton fields,
we showed that the more general solution is not fine-tuned, in agreement with a counting
of free parameters in these models. We demonstrated that singularities in the self-tuned
solutions are generic if gravity is to be localized, and we presented a no-go theorem to
this effect. We provided perturbative and numerical techniques in order to calculate the
self-tuned solutions, and we illustrated the major points of the paper via several numerical
examples. Finally, we pointed out that the fine-tuning that is required in order to resolve
the singularities in the spirit of [16,17,32] for the case of zero bulk potential is generic.

Acknowledgements
We are extremely grateful to Chris Kolda for collaboration at early stages of this project.
We thank Nima Arkani-Hamed, Tanmoy Bhattacharya, Pierre Bintruy, Martin Schmaltz
and Anupam Singh for useful discussions. C.C. thanks the Theory Group at Berkeley
for hospitality while this work was initiated. C.C. is a J. Robert Oppenheimer fellow
at the Los Alamos National Laboratory. C.C., J.E. and T.H. are supported by the US
Department of energy under contract W-7405-ENG-36. C.G. is supported in part by the
US Department of energy under Contract DE-AC03-76SF00098 and in part by the National
Science Foundation under grant PHY-95-14797.
References
[1] N. Arkani-Hamed, S. Dimopoulos, G. Dvali, Phys. Lett. B 429 (1998) 263, hep-ph/9803315.
[2] I. Antoniadis, N. Arkani-Hamed, S. Dimopoulos, G. Dvali, Phys. Lett. B 436 (1998) 257, hepph/9804398.
[3] L. Randall, R. Sundrum, Phys. Rev. Lett. 83 (1999) 3370, hep-ph/9905221; Phys. Rev. Lett. 83
(1999) 4690, hep-th/9906064.

386

[4]
[5]
[6]
[7]
[8]
[9]
[10]
[11]
[12]
[13]
[14]
[15]
[16]
[17]
[18]
[19]
[20]
[21]
[22]
[23]
[24]
[25]
[26]
[27]
[28]
[29]
[30]
[31]
[32]
[33]
[34]
[35]
[36]

C. Cski et al. / Nuclear Physics B 584 (2000) 359386

S. Weinberg, Rev. Mod. Phys. 61 (1989) 1.


V.A. Rubakov, Phys. Rev. D 61 (2000) 061501, hep-ph/9911305.
A. Hebecker, C. Wetterich, hep-ph/0003287.
N. Arkani-Hamed, S. Dimopoulos, N. Kaloper, R. Sundrum, Phys. Lett. B 480 (2000) 193,
hep-th/0001197.
S. Kachru, M. Schulz, E. Silverstein, hep-th/0001206; hep-th/0002121.
V.A. Rubakov, M.E. Shaposhnikov, Phys. Lett. B 125 (1983) 139.
S.S. Gubser, hep-th/0002160.
C. Gmez, B. Janssen, P. Silva, JHEP 0004 (2000) 027, hep-th/0003002.
S.W. Hawking, R. Penrose, Proc. Roy. Soc. Lond. A 314 (1970) 529.
D. Youm, hep-th/0002147.
J. Chen, M.A. Luty, E. Ponton, hep-th/0003067.
I. Low, A. Zee, hep-th/0004124.
S.P. de Alwis, hep-th/0002174.
S.P. de Alwis, A.T. Flournoy, N. Irges, hep-th/0004125.
O. DeWolfe, D.Z. Freedman, S.S. Gubser, A. Karch, hep-th/9909134.
C. Grojean, J. Cline, G. Servant, Nucl. Phys. B 578 (2000) 259, hep-th/9910081.
C. Grojean, Phys. Lett. B 479 (2000) 273, hep-th/0002130.
A. Brandhuber, K. Sfetsos, JHEP 9910 (1999) 013, hep-th/9908116.
I. Bakas, K. Sfetsos, Nucl. Phys. B 573 (2000) 768, hep-th/9909041.
K. Behrndt, M. Cvetic, hep-th/9909058; Phys. Rev. D 61 (2000) 101901, hep-th/0001159.
K. Skenderis, P.K. Townsend, Phys. Lett. B 468 (1999) 46, hep-th/9909070.
A. Chamblin, G.W. Gibbons, Phys. Rev. Lett. 84 (2000) 1090, hep-th/9909130.
R. Kallosh, A. Linde, JHEP 0002 (2000) 005, hep-th/0001071.
M. Gremm, Phys. Lett. B 478 (2000) 434, hep-th/9912060; hep-th/0002040.
C. Cski, J. Erlich, T.J. Hollowood, Y. Shirman, hep-th/0001033, to be published in Nucl.
Phys B.
M. Cvetic, Phys. Rev. Lett. 71 (1993) 815, hep-th/9304062.
M. Cvetic, Phys. Lett. B 341 (1994) 160, hep-th/9402089.
M. Cvetic, H.H. Soleng, Phys. Rev. D 51 (1995) 5768, hep-th/9411170.
S. Frste, Z. Lalak, S. Lavignac, H.P. Nilles, Phys. Lett. B 481 (2000) 360, hep-th/0002164.
R. Gregory, V.A. Rubakov, S.M. Sibiryakov, Phys. Rev. Lett. 84 (2000) 5928, hep-th/0002072.
C. Cski, J. Erlich, T.J. Hollowood, hep-th/0002161, to be published in Nucl. Phys B.
G. Dvali, G. Gabadadze, M. Porrati, hep-th/0002190.
E. Witten, hep-ph/0002297.

Nuclear Physics B 584 (2000) 387414


www.elsevier.nl/locate/npe

The interaction of Dirac particles with non-abelian


gauge fields and gravity bound states
Felix Finster a, , Joel Smoller b,1,3 , Shing-Tung Yau c,2,4
a Max Planck Institute for Mathematics in the Sciences, Inselstr. 22-26, 04103 Leipzig, Germany
b Mathematics Department, The University of Michigan, Ann Arbor, MI 48109, USA
c Mathematics Department, Harvard University, Cambridge, MA 02138, USA

Received 1 February 2000; revised 29 March 2000; accepted 13 June 2000

Abstract
We consider a spherically symmetric, static system of a Dirac particle interacting with classical
gravity and an SU(2) YangMills field. The corresponding EinsteinDiracYangMills equations are
derived. Using numerical methods, we find different types of soliton-like solutions of these equations
and discuss their properties. Some of these solutions are stable even for arbitrarily weak gravitational
coupling. 2000 Elsevier Science B.V. All rights reserved.
PACS: 03.65.Pm; 04.20.Jb; 11.15.Kc
Keywords: Relativistic quantum mechanics; Dirac equation; Relativity; YangMills field

1. Introduction
The coupling of gravity to other classical force fields and to quantum mechanical
particles has led to many interesting solutions of Einsteins equations and has given some
insight into the nature of the nonlinear interactions. The first such examples are the BartnikMcKinnon (BM) solutions of the EinsteinYangMills (EYM) equations [1]. For these
solutions, the repulsive YangMills force compensates the attractive gravitational force;
unfortunately, these solutions are unstable [2]. If, on the other hand, one considers quantum
mechanical Dirac particles, the gravitational attraction is balanced by the repulsion due to
the Heisenberg Uncertainty Principle, and this leads to stable bound states of the resulting
EinsteinDirac (ED) system [3]. However, a pure ED system is somewhat artificial,
Corresponding author. E-mail: felix.finster@mis.mpg.de
1 smoller@umich.edu
2 yau@math.harvard.edu
3 Research supported in part by the NSF, Grant No. DMS-G-9802370.
4 Research supported in part by the NSF, Grant No. 33-585-7510-2-30.

0550-3213/00/$ see front matter 2000 Elsevier Science B.V. All rights reserved.
PII: S 0 5 5 0 - 3 2 1 3 ( 0 0 ) 0 0 3 7 0 - 9

388

F. Finster et al. / Nuclear Physics B 584 (2000) 387414

because physical Dirac particles also exhibit electroweak and strong interactions, which
in all realistic situations are much stronger than gravity. Thus the question arises if Dirac
particles in a gravitational field still form bound states if an additional strong coupling to
a non-abelian YangMills (YM) field is taken into account (the case of an abelian gauge
field was considered in [4]). Related questions are, do the BM solutions become stable if
one includes Dirac particles, and which physical effects does the nonlinear coupling in the
EinsteinDiracYangMills (EDYM) equations lead to?
In order to address these questions, we consider here a static, spherically symmetric
EDYM system of one Dirac particle in a gravitational field and an SU(2) YangMills
field. In this system, the spinors couple only to the magnetic component of the YM field,
and we thus obtain a consistent ansatz by setting the electric component identically equal
to zero. In the limit of weak coupling of the spinors, our system goes over to the EYM
system [1]. In contrast to the two-particle singlet state studied in [3], we consider here only
one Dirac particle (this becomes possible because the inclusion of the SU(2) YangMills
field changes the representation of the rotation group on the spinors; see Section 2). Thus
one cannot recover exactly the ED system [3], but the limit of a weak YangMills field
yields equations which are closely related to the ED equations of the two-particle singlet.
By numerically seeking bound states of the EDYM system, we find a surprisingly rich
solution structure. First of all, we find solutions where the Dirac particle is bound by the
gravitational attraction, and where the Dirac particle also generates a YM field. Stable
solutions of this type exist also in the physically realistic situation of weak gravitational
and strong YM coupling. This result shows that the magnetic component of the YM field,
which usually has a repulsive effect (like, e.g., for the BM solutions) cannot prevent
Dirac particles from forming stable bound states. We also find other types of solutions
where the binding comes about through the nonlinear interaction of all three fields. These
solutions have stable and unstable branches, whereby the stable solutions are found for
weak gravitational coupling, provided that the YM coupling is sufficiently strong, but not
too strong. Finally, we study the relation between these solutions and the BM solutions.
We find one-parameter families of solutions which join the BM ground state with our new
solutions. This shows that the BM ground state can be made stable by the inclusion of
a Dirac particle, but only if the coupling to the spinors is sufficiently strong. The first
excited BM state, on the other hand, cannot be joined with our new solutions. Namely,
perturbing this state by an additional Dirac particle yields a separate unstable branch of
EDYM solutions.
The plan of the paper is as follows. In Section 2 we derive the SU(2)EDYM equations.
In Section 3 we obtain a limiting system constructed by letting the gravitational constant
tend to zero and letting the YM coupling constant tend to infinity. We find numerical
solutions of this system and discuss their properties. In the last section, we consider
solutions of the full EDYM equations, obtained by tracing the solutions of our limiting
system and the BM solutions while continuously varying the coupling constants.

F. Finster et al. / Nuclear Physics B 584 (2000) 387414

389

2. Derivation of the equations


The general EDYM equations are obtained by variation over Lorentzian metrics gij , YM
connections A, and Dirac wave functions , of the action

Z 
 p
1
1
ij
R + (G m)
Tr Fij F
det g d 4 x,
(2.1)
S=
16
16e2
where R is scalar curvature, G is the Dirac operator (which depends on the gravitational
and YM fields), Fij is the YM field tensor, and where the trace is taken over the YM
index. The gravitational and YM coupling constants are denoted by and e, respectively.
The appearance of the factor e2 in (2.1) requires a brief explanation. In contrast to the
usual form of the gauge-covariant derivative Dj = j ieAj , we use here the convention
Dj = j iAj (this makes it possible to work with the particularly convenient form of
the gauge potentials used in [1]). Our convention is obtained from the usual one by the
transformation Aj e1 Aj . Under this transformation, the field strength tensor behaves
like Fij e1 Fij , and this gives rise to the factor e2 in (2.1).
In this paper, we shall study a particular EDYM system, which is obtained as follows.
First of all, we consider a spherically symmetric, static metric in polar coordinates,
ds 2 =

1
1
dr 2 r 2 d 2 r 2 sin2 d 2 ,
dt 2
2
T (r)
A(r)

(2.2)

with positive functions A and T . The Einstein tensor corresponding to this metric is given
in [3]. Moreover, as in [1], we restrict attention to the magnetic component of an SU(2)
YangMills field and choose for the YM potential the ansatz
A = w(r) 1 d + (cos 3 + w(r) sin 2 ) d

(2.3)

with a real function w, where E = 12 E is the standard basis of su(2) (E are the Pauli
1
ia
matrices). The YM current j and energy-momentum tensor Tji = 4e
2 Tr(F Fj a
1 ab
i
4 F Fab j ) are computed to be


A 00 A0 T 2AT 0 0 w(1 w2 )
1
w
2w
j=
4e2
2r
4r 2 T
2r 4



+ 2 csc()
,
1



1
4
2
2 2
02
(1

w
)

Aw

,
T00 =
4e2
r4
r2


1
4
2
2 2
02
(1

w
)
+
Aw

,
T11 =
4e2
r4
r2


1
2
2 2
(1

w
)
,
T22 = T33 =
4e2 r 4

(2.4)
(2.5)
(2.6)

and all other components vanish. When coupled to the Dirac spinors, the YM potential (2.3)
has the disadvantage that it depends on and , in a way which makes it difficult to

390

F. Finster et al. / Nuclear Physics B 584 (2000) 387414

separate variables in the Dirac equation. To remedy this, we perform the SU(2) gauge
transformation Aj U Aj U 1 + iU (j U 1 ) with


i 2
.
U (, ) = exp(i 1 ) exp(i( + ) 3 ) exp
2
The resulting YM potential is
A = (w 1)r sin ( d d),

(2.7)

where we use the following polar linear combinations of the matrices,


r = 1 cos + 2 sin cos + 3 sin sin ,
1
= ( 1 sin + 2 cos cos + 3 cos sin ),
r
1
( 2 sin + 3 cos ).
(2.8)
=
r sin
By minimally coupling the SU(2) potential (2.7) to the Dirac operator in the gravitational
field [3, Eq. (2.23)], we obtain the Dirac operator



i
i T0
t
r
A
+ i + i
G = iT t + i Ar + ( A 1)
r
2
T
2i
+ (w 1)(E E r r ) r ,
(2.9)
r
where ( j )j =t,r,, are, in analogy to (2.8), the Dirac matrices of Minkowski space in polar
coordinates, where we work in the Dirac representation,




1 0
0 E
,
E =
.
(2.10)
t =
0 1
E
0
Notice that the Dirac operator (2.9) acts on 8-component wave functions; this is because the
additional YM index doubles the number of components compared to usual Dirac spinors.
More precisely, it is convenient to regard the wave functions as sections of
C8 = C2up/down C2large/small C2YM ,

(2.11)

where C2up/down describes the two spin orientations, C2large/small corresponds to the upper
and lower components of the Dirac spinor (i.e., usual Dirac spinors are sections of C4 =
C2up/down C2large/small), and C2YM is acted upon by the SU(2) gauge group. For clarity, we
shall refer to the factors in (2.11) by separate indices, i.e., we write a wave function
as ( ua ),u,a=1,2 , where , u, and a label the components of C2up/down, C2large/small,
and C2YM , respectively. Thus the operators E act on the index a, the spin operators SE are
given by SE = 1 E acting on the Greek indices, and t coincides with the matrix t =
2

diag(1, 1) acting on the index u, i.e.,


 ua

,
if u = 1,
t ua
=

ua
, if u = 2.

It is apparent in (2.9) that the Dirac operator commutes with the three operators
E + SE + E,
JE = L

(2.12)

F. Finster et al. / Nuclear Physics B 584 (2000) 387414

391

E is angular momentum. It is very convenient to regard the operators JE as the total


where L
angular momentum operators of the system. Since the total angular momentum operators
are the infinitesimal generators of rotations (as explained for angular momentum in [5,
26]), we can then interpret (2.12) as saying that the inclusion of the YM field influences
the representation of the rotation group on the spinors. The Dirac operator is invariant under
this group representation, because the operators JE commute with G; this makes spherical
symmetry of the Dirac operator manifest.
E to
Since (2.12) coincides with the formula for the addition of angular momentum L
1
E
E
two spin- 2 -operators S and E , it is clear that the operators J have integer eigenvalues.
Thus we can expect that the operator J 2 has a nontrivial kernel. In this case, the simplest
spherically symmetric ansatz for the Dirac particles would be to take one Dirac particle
whose wave function is in the kernel of J 2 . We now work out this ansatz in detail, whereby
we consider JE as operators on the spinors a (, ) on S 2 (i.e., the a are sections
of C4 = C2up/down C2YM ). Adding the two spin operators SE and E , we can decompose
C2up/down C2YM into the direct sum of one state of total spin zero and three states of
total spin one (see [5, 31]; these states are usually called the singlet and triplet states,
E according to the standard
respectively). By subsequently adding the angular momentum L
rules for the addition of angular momentum [5, 31], one sees that the operator J 2 has
indeed a nontrivial kernel. More precisely, the kernel of J 2 is two-dimensional, spanned
by two vectors 0 and 1 with angular momentum zero and one, respectively. The state
0 is (up to a phase) uniquely characterized by the conditions
E 0 = 0 = (SE + E)0
L

and k0 kS 2 = 1.

(2.13)

Using (2.13), we can write 1 as


1 = 2S r 0 = 2 r 0 .
Namely, representing

Sr

and

(2.14)
in the form

S = xEsE and = xE E,
r

E xE , and SE (see
and using the standard commutation relations between the components of L,
[5, 26, 54]), we obtain that
E (E
E 0 + 2[SE + E, (E
E 0
x S)]
x S)]
JE1 = 2[JE, S r ]0 = 2[L,
E 0 + 2i xE S
E 0=0
= 2i xE S

(2.15)

R3 ),

and
(where is the wedge product in
Z
k1 k2S 2 = h2S r 0 , 2S r 0 i d = k0 k2S 2 = 1.
S2

One can verify directly that 1 has angular momentum one; namely
E L,
E S r ]0 = 2i L(E
E x S)
E 0
L2 1 = 2L2 S r 0 = 2L[
E (E
E 0 = 4(E
E 0 = l(l + 1)1
= 2i[L,
x S)]
x S)
with l = 1. Furthermore, using the fact that (2S r )2 = 1 = (2 r )2 and S 2 =
obtain that

3
4

= 2 , we

392

F. Finster et al. / Nuclear Physics B 584 (2000) 387414

0 = 2S r 1 = 2 r 1 ,
(2.13)
E 0 = 3 0 ,
SEE0 = SE S
4
 (2.15) 1
1 E
1
2
E
(S + E ) S 2 2 1 =
= (L2 S 2 2 )1 = 1 ,
S E1 =
2
2
4
E 0 = 0,
SE L

E 2 S 2 L2 1 = 1 ( 2 S 2 L2 )1 = 1 .
E 1 = 1 (SE + L)
SE L
2
2

(2.16)
(2.17)
(2.18)
(2.19)
(2.20)

Finally, it is useful to observe that (cf. [6, Eq. (3.3)])


2
E
S + S = S r (SEL).
r

(2.21)

Using the relations (2.13)(2.21), one can easily compute the Dirac operator (2.9) on the
invariant subspace J 2 = 0. It turns out that we obtain a consistent ansatz for the Dirac wave
function by setting


ua
it T (r)
(r)0a (, )u,1 + i(r)1a (, )u,2
(t, r, , ) = e
(2.22)

r
with real functions and , where > 0 is the energy of the Dirac particle, and .,. is the
Kronecker delta. For this ansatz, the Dirac equation reduces to the system of ODEs
0 w
(2.23)
A = (m + T ),
r
0
w
(2.24)
A = (m + T ) .
r
The Dirac current j and Dirac energy-momentum tensor Tj k = Re( G(j Dk) ) corresponding to the ansatz (2.22) are obtained by a straightforward computation similar to that
in [3]. The result is



2T
1
2
+ csc()
,
j = 3

r
T 2 2
( + 2 ),
r2
T 2
T
mT
T11 = 2 ( 2 + 2 ) + 2 3 w + 2 ( 2 2 ),
r
r
r
T
2
3
T2 = T3 = 3 w,
r
T00 =

and all other components vanish. The normalization condition for the spinors is (as in [3]),
Z
T
( 2 + 2 ) = 1.
A

(2.25)

By substituting the formulas for the YM current and energymomentum tensor into the
Einstein and YM equations, we obtain the following system of ODEs,

F. Finster et al. / Nuclear Physics B 584 (2000) 387414

rA0 = 1 A

393

(1 w2 )
2
2T 2 ( 2 + 2 ) 2 Aw02 ,
2
2
e
r
e

(2.26)

(1 w2 )
T0
= 1 + A + 2
+ 2mT ( 2 2 ) 2T 2 ( 2 + 2 )
T
e
r2
2
T
+ 4 w 2 Aw02 ,
r
e
A0 T 2AT 0 0
w.
r 2 Aw00 = (1 w2 )w + e2 rT r 2
2T

2rA

(2.27)
(2.28)

The Einstein equations are (2.26) and (2.27), whereas (2.28) is the YM equation. Our
EDYM system is given by the five ODEs (2.23), (2.24), (2.26)(2.28), together with the
normalization condition (2.25).
We are interested here in bound states of the Dirac particles. Thus we want to find
particle-like solutions of our EDYM system, i.e., solutions which are smooth and tend to
the vacuum solution as r . According to the explicit formulas (2.4)(2.6), the energymomentum tensor of the YM field is regular at r = 0 only when |w(0)| = 1 and w0 (0) = 0.
Using the remaining gauge freedom, we can assume that w(0) = 1, and thus

w(r) = 1 r 2 + O(r 3 )
2

(2.29)

with a real parameter . Using this result, a local Taylor expansion of the Einstein and
Dirac equations around r = 0 yields (just as in [3]) that
(r) = 1 r + O(r 3 ),
A(r) = 1 + O(r 2 ),

1
(r) = (T0 m)1 r 2 + O(r 3 ),
2
T (r) = T0 + O(r 2 )

(2.30)
(2.31)

with two parameters 1 and T0 > 0. Using linearity of the Dirac equation, we can always
assume that 1 > 0. Furthermore, we demand that our solution has finite ADM mass,
r
(1 A(r)) < ,
(2.32)
:= lim
r 2
and goes asymptotically to the vacuum solution,
lim T (r) = 1,

lim (w(r), w0 (r)) = (1, 0)

lim ((r), (r)) = (0, 0).

(2.33)

3. The reciprocal coupling limit


Under all realistic conditions, the coupling of Dirac particles to the YM field (describing
the weak or strong interactions) is much stronger than the coupling to the gravitational
field. Thus we are particularly interested in the case of weak gravitational coupling. In
preparation, it is instructive to briefly consider the case without gravitation. In this limit,
the Dirac equations read

394

F. Finster et al. / Nuclear Physics B 584 (2000) 387414

w
(m + ),
(3.1)
r
w
(3.2)
0 = (m + ) .
r
For large r, these equations go over to a linear system of ODEs with constant coefficients,
and the sign of m determines whether the solutions of these equations behave
oscillatory or exponentially. The normalization condition (2.25) excludes the oscillatory
case (as in [6, Section 5]) and thus m > 0. In the case m = 0, the -equation
is independent of , and the boundary conditions (2.30) imply that 0. As a
consequence, (2.28) reduces to the homogeneous YM equation
0 =

r 2 w00 = (1 w2 )w.

(3.3)

It is well-known [7] that the only solution to this equation satisfying the boundary
conditions (2.29),(2.33) is the trivial solution w 1. But then the -equation simplifies
to
1
0 = ,
r
whose solution = 1 r violates the normalization condition (2.25). In the case m
> 0, on the other hand, the local Taylor expansion (2.30) yields that the (, )-curve lies for
small r in the fourth quadrant, i.e., (r) < 0 < (r) for small r. Using the Dirac equations
(3.1), (3.2), one sees that the fourth quadrant is an invariant region, and thus (r) < 0
< (r) for all r. But in the fourth quadrant, both (r) and (r) are increasing for large r
(as one sees in (3.1), (3.2) taking into account that w/r 0 for r ), and thus the
normalization condition (2.25) will again be violated.
These considerations show that the gravitational field is essential for the formation of
bound states. Nevertheless, for arbitrarily weak gravitational coupling, we can hope to find
bound states. It is even conceivable that these bound state solutions might have a welldefined limit when the gravitational coupling tends to zero, if we let the YM coupling go
to infinity at the same time. Our idea is that this limiting case might yield a system of
equations which is simpler than the full EDYM system, and can thus serve as a physically
interesting starting point for the analysis of the coupled interactions described by the
EDYM equations. Expressed in dimensionless quantities, we shall thus consider the limits
m2 0

and e2 .

(3.4)

Let us determine how the quantities of our EDYM system should behave in this limit.
Since we are considering weak gravitational coupling, it is clear that the metric will be
close to the Minkowski metric, i.e., A 1 and T 1. Furthermore, the YM potential w
should have a finite limit. Similar to our flat space consideration at the beginning of this
section, one sees that the normalization condition (2.25) can be satisfied only if the function
mT (r) changes sign, and thus m (but both m and may go to zero or infinity in the
limit (3.4)). Putting this information together, we conclude that the Dirac equations (2.23)
and (2.24) have a meaningful limit only under the assumptions that converges and that

m(r) (r),

m2 (T (r) 1) ,

m( m) E

(3.5)

F. Finster et al. / Nuclear Physics B 584 (2000) 387414

395

and a real parameter E. Multiplying (2.24) with m and taking


with two real functions ,
the limits (3.5) as well as A, T 1, the Dirac equations become
w

(3.6)
0 = 2,
r
w

(3.7)
0 = (E + ) .
r
We next consider the YM equation (2.28). The last term in (2.28) drops out in the limit
of weak gravitational coupling (3.4). The second summand converges only under the
assumption that
e2
q
(3.8)
m
with q a real parameter, playing the role of an effective coupling constant. Together
with (3.4), this implies that m . The YM equations thus have the limit

r 2 w00 = (1 w2 )w + qr .

(3.9)

In order to get a well-defined and non-trivial limit of the Einstein equations (2.26), (2.27),
we need to assume that the parameter m3 has a finite, non-zero limit. Since this parameter
has the dimension of inverse length, we can arrange by a scaling of our coordinates that
m3 1.

(3.10)

We differentiate the T -equation (2.27) with respect to r and substitute (2.26). Taking the
limits (3.5) and (3.10), a straightforward calculation yields the equation
r 2 1 = 2 ,

(3.11)

where 1 = r 2 r (r 2 r ) is the radial Laplacian in Euclidean R3 . Indeed, this equation


can be regarded as Newtons equation with the Newtonian potential . Thus our limiting
case (3.11) for the gravitational field corresponds to taking the Newtonian limit. Finally,
the normalization condition (2.25) reduces to
Z
(r)2 dr = 1.
(3.12)
0

The boundary conditions (2.29)(2.33) are transformed into

w(r) = 1 r 2 + O(r 3 ),
2
(r) = 1 r + O(r 3 ),
(r) = 0 + O(r ),
2

lim w(r) = 1,

= O(r 3 ),
(r)
lim (r) <

(3.13)
(3.14)
(3.15)

with the three parameters , 1 , and 0 . We point out that the limiting system contains only
one coupling constant q. According to (3.8) and (3.10), q is in dimensionless form given
by
e2 m2 q.

(3.16)

Hence in dimensionless quantities, our limit (3.4) describes the situation where the
gravitational coupling goes to zero, while the YM coupling constant goes to infinity like

396

F. Finster et al. / Nuclear Physics B 584 (2000) 387414

e2 (m2 )1 . Therefore, we call our limiting case the reciprocal coupling limit. The
reciprocal coupling system is given by the equations (3.6), (3.7), (3.9), and (3.11) together
with the normalization condition (3.12) and the boundary conditions (3.13)(3.15).
According to (3.5), the parameter E coincides up to a scaling factor with m, and
thus has the interpretation as the (properly scaled) energy of the Dirac particle. As in
Newtonian mechanics, the potential is determined only up to a constant R; namely,
the reciprocal limit equations are invariant under the transformation
+ ,

E E .

(3.17)

Let us consider how the ADM mass behaves in the reciprocal coupling limit. First of all,
we can write the quotient /m as
1
r

= lim
(1 A(r)) =
m r 2m
m

Z

r
(1 A(r))
2

0
dr.

After substituting the A-equation (2.26), we can take the limits (3.4) and (3.5) and obtain
that

(3.12)

(r)2 dr = 1.

(3.18)

Thus the ADM mass coincides with the rest mass of the Dirac particle; this shows that the
total binding energy B := m goes to zero in our limit. Indeed, the behavior of the total
binding energy can be described in more detail as follows. For a solution of the full EDYM
system, we can write the binding energy using the normalization condition (2.25) as
Z
B=

r
(1 A)
2

0


T
dr.
m( 2 + 2 )
A

We again substitute the A-equation (2.26) and obtain


Z
B=
0

1
T
1 (1 w2 )2
02
2
2
(
+
Aw
+
(wT
A

m)
+

)
dr.

2e2
r2
e2
A

(3.19)

According to (3.16), it is obvious that the first two summands in (3.19) have a finite limit
after dividing by m2 . In order to treat the last summand, we first multiply the T -equation
(2.27) with m2 and take the limits (3.4), (3.5), (3.10),
m2 (A 1) 2r 0 .
Using again (3.10), (3.5), and m, T 1, we obtain that

1
1
1
(T A m) = 3 m( A 1)T + 3 m(T m)
2
m
m
m
r 0 + ( + E).

F. Finster et al. / Nuclear Physics B 584 (2000) 387414

397

From this we conclude that the binding energy (3.19) divided by m2 has a meaningful
limit; more precisely
b := B
B
m2

Z


1 (1 w2 )2 1 02
2
0
+ w + (E + + r ) dr.
2q
r2
q

(3.20)

We now describe our method for constructing numerical solutions of our reciprocal limit
system. Since it is difficult to take into account the integral condition (3.12) in the numerics,
we discard this condition for the construction of the solution; it will be taken care of later
via a rescaling technique (see (3.34), (3.35)). This rescaling method requires only that the
normalization integral be finite,
Z
0 < :=

(r)2 dr < .

(3.21)

According to (3.6), (3.7) and (3.13), (3.15), the behavior of the Dirac spinors at infinity is
either oscillatory or exponential. As a consequence, the normalization integral in (3.21) will
be finite only if (r) tends to zero for r . Furthermore, we can use the transformation
(3.17) to set (0) = 0. Hence in the first construction step, we want to find solutions of the
modified system
w

(3.22)
0 = 2,
r
w

(3.23)
0 = (E + ) ,
r

(3.24)
r 2 w00 = (1 w2 )w + qr ,
r 2 1(r) = 2

(3.25)

with the following conditions at the origin,


w(r) = 1 r 2 + O(r 3 ),

(r) = 1 r + O(r 2 )

(3.26)

(r) = O(r ),

(r) = O(r ),

(3.27)

together with the conditions at infinity


lim w(r) = 1,

lim (r) = 0,

| | := | lim (r)| < .


r

(3.28)
(3.29)

For any given value of the coupling constant q, we thus have two free parameters and 1
to characterize the solutions near the origin r = 0. Each solution has a unique extension to
larger values of r. Asymptotically for r , we must satisfy the two conditions (3.28).
Thus we have as many free parameters as asymptotic conditions, and we therefore expect
for fixed q a discrete set of solutions satisfying (3.26), (3.27), and (3.28). For each solution,
we must then verify that the conditions (3.21) and (3.29) are also satisfied.
For the construction of numerical solutions, we enhanced the technique used in [3,4] to
a two-parameter shooting method. Since two-parameter problems are considerably more

398

F. Finster et al. / Nuclear Physics B 584 (2000) 387414

difficult than one-parameter problems, we describe the method in some detail. For clarity,
we first consider the simplified situation where (r) and w(r) have prescribed boundary
values for a given finite r = r1 . In this case, one can apply the standard multi-parameter
shooting method as, e.g., described in [8]. More precisely, one can for given initial data
compute (r1 ) and (w(r1 ), w0 (r1 )) numerically, compare with the prescribed boundary
conditions, and iteratively adjust the two free parameters and 1 until the boundary
conditions are satisfied to sufficient accuracy. In our case, the situation is more difficult
because we have boundary values not for finite r = r1 , but for r = . In order to deal
with this problem, we first choose a finite r1 . Using an ansatz for the asymptotic form of
w, ) at infinity, we approximately compute (r1 ) and (w(r1 ), w0 (r1 ))
the solution (, ,
and derive conditions between these functions. Taking these conditions as the boundary
conditions at r = r1 , we can apply the two-parameter shooting method on the finite interval
(0, r1 ] as described above. The so-obtained solution on (0, r1 ] gives, in combination with
the asymptotic formulas on (r1 , ), an approximate solution for all r > 0. Since our
asymptotic description becomes precise only in the limit r1 , we must, in order
to attain the desired accuracy, choose r1 sufficiently large. In order to ensure that r1 is
appropriately increased during the computation, we modified the two-parameter shooting
method in such a way that both the initial data and r1 are adjusted in each iteration step. The
iteration is stopped when the numerics has stabilized and the accuracy no longer improves.
This modified shooting method was implemented in the M ATHEMATICA programming
language using the standard ODE solver with a working precision of 16 digits. The initial
data is adjusted in the iteration with a secant method, and the step size for incrementing r1
is determined from the relative error of the numerical solution at the upper boundary r1 .
After the iteration has been stopped and a numerical solution has been found, our program
slightly changes the initial data and searches for a nearby solution. In this way, we can
automatically trace a one-parameter family of solutions. Finally, we explain our method for
describing the asymptotic behavior of the solutions at infinity. According to the asymptotics
of the solutions of the ED and EYM equations [1,3], we can expect that the spinors and
will decay exponentially fast at infinity, whereas the potentials (r) and w(r) for r
will behave like rational functions. Therefore it is a reasonable asymptotic approximation
to set and to zero for r > r1 . In this approximation, the potential is harmonic
according to (3.25). The YM equation (3.24), on the other hand, reduces to the vacuum
YM equation (3.3). In the new variable u = log r, this equation becomes autonomous;
namely [7]
u2 w u w = (1 w2 )w.

(3.30)

This autonomous equation allows us to derive boundary conditions for w as follows. We


set
x =w

and y = u w.

(3.31)

Then the YM orbits in the (x, y) plane are described by the following differential equation,
y 0 (x) =

u2 w (3.30)
(1 x 2 )x
.
= 1
u w
y

(3.32)

F. Finster et al. / Nuclear Physics B 584 (2000) 387414

399

According to the boundary conditions (3.28) and the differential equation (3.30), the
variables x and y must behave in the limit r like either x 1, y & 0 or x
1, y % 0. In both of these cases, there is a unique YM orbit y(x), which can be easily
calculated numerically by integrating (3.32). By transforming (3.31) back to the variable r,
we obtain the following mixed boundary conditions for w(r) at r = r1 ,
w0 (r1 ) =

1
y(w(r1 )).
r1

(3.33)

We next describe our rescaling method needed to arrange the normalization condition (3.12). Suppose that we have a solution of the modified system (3.22)(3.29) with
finite normalization integral, (3.21). A direct calculation shows that the transformed functions
(r)

= 2 (2 r),
4

(r)
=


(2 r) ,

= 4 (2 r),
(r)
2

w(r)

= w(

r)

(3.34)
(3.35)

solve our original reciprocal limit system (3.6), (3.7), (3.9), (3.11), and (3.12) with
boundary conditions (3.13)(3.15), if one replaces the energy E and coupling constant q
by
e = 4 (E + )
E

and q = 4 q.

(3.36)

We point out that only the rescaled solutions (3.34), (3.35) and rescaled parameters (3.36)
have a physical meaning. Therefore, we will in what follows consider only the rescaled
tilde solutions; for ease in notation, the tilde will be omitted.
In the remainder of this section, we describe our numerical solutions of the reciprocal
limit equations. Just as in the case for the ED and EYM equations [1,3], there are solutions
for different rotation numbers of the spinors in the (, )-plane and for the YM potential in
the (w, w0 )-plane. For simplicity, we restricted attention to solutions with rotation number
zero for the spinors (as for the ground state solutions in [3]). For the YM potential, we
consider only the cases where the (w, w0 )-curve either makes a half rotation joining the
points (1, 0) and (1, 0), or makes a full rotation, ending at its starting point (1, 0).
A typical example for a solution of each type is shown in Figs. 1 and 2. Because of the
similarity of the YM potential to the BM ground state and the BM first excited state, we
refer to these two types in what follows as the ground states and the first excited states,
respectively. Notice that the curves in the (w, w0 )-plane are not plotted all the way to their
rest points at (1, 0) or (0, 1), respectively. The reason is that we plot only the numerical
solution on the interval (0, r1 ]. One sees that the spinors have become practically zero
for r = r1 , and it is thus an admissible approximation to smoothly join the (w, w0 )-curve
with a vacuum YM solution by using the boundary conditions (3.33). We first discuss the
ground state solutions. In Fig. 3, the main characteristics of the solutions are plotted versus
the coupling constant q. As explained above, E has the interpretation of the (appropriately
scaled) energy of the Dirac particle. Since E is negative, the Dirac particle has gained
b (3.20), gives the total binding energy,
energy by forming the bound state. The parameter B,
b is
i.e., the amount of energy which is set free when the binding is broken up. Since B

400

F. Finster et al. / Nuclear Physics B 584 (2000) 387414

Fig. 1. Reciprocal coupling limit: the ground state for q = 8.49811, E = 0.377446.

Fig. 2. Reciprocal coupling limit: the first excited state for q = 6.96132, E = 0.227762.

negative, we can expect that solutions of the full EDYM system, which are close to the
solutions of the reciprocal coupling equations, should be stable. Finally, rw and r are
the characteristic length scales of the solutions; more precisely, rw is the radius where w
changes sign, and r is the radius where has its maximum,
w(rw ) = 0 and 0 (r ) = 0.

(3.37)

The characteristic radii are interesting because they give information about the size of
the solutions as functions of r; i.e., they tell whether the fields are spread out in space,
or whether they are localized close to the origin. It is also worth considering both radii
because rw and r can behave quite differently (cf. Fig. 3).
The plots in Fig. 3 have a turning point at q 8.49. Similar to the situation described
for the spiral in [3], this is a bifurcation point which comes about as a consequence of our
rescalings. One branch of solutions can be extended up to q 11.6. For solutions close to
this end point, the potential w leaves the interval [1, 1] as shown in Fig. 4. Since rw and
r both go to zero in this limit, the spinors and YM field are both confined to a smaller
and smaller neighborhood of the origin. At the same time, the energy of the Dirac particle
and the binding energy become infinite. The other branch of solutions ends near q = 8.95.
For solutions near this end point, the (w, w0 )-curve comes very close to the origin before
running into the rest point at (1, 0), see Fig. 5. This makes the numerics rather delicate,
and we therefore have not yet analyzed this regime in much detail. It is interesting that
r is bounded near this end point, whereas rw seems to become infinite. This shows that,
while the Dirac particle stays in a bounded region of space, the YM field becomes more
and more spread out.
For the first excited state, the energy spectrum and characteristic radii are shown in
Fig. 6. Since in general w never equals zero for the first excited state, we define rw via the
minimum of w, i.e.,
w0 (rw ) = 0 and 0 (r ) = 0.

(3.38)

F. Finster et al. / Nuclear Physics B 584 (2000) 387414

401

Fig. 3. Reciprocal coupling limit: the energy spectrum and characteristic radii for the ground state.

Fig. 4. Reciprocal coupling limit: the ground state for q = 11.5838, E = 512161.

Fig. 5. Reciprocal coupling limit: the ground state for q = 8.76701, E = 0.334974.

In contrast to the ground state, the solutions can now be extended up to q = 0. In this
regime, the YM potential stays close to w = 1; see Fig. 7. The solutions have a bifurcation
point at q 9.866. The branch coming out at the bifurcation point for larger values of E
is difficult to study numerically because the (w, w0 )-curve comes close to the origin, see
Fig. 8.
It is interesting that for the ground state solutions in Fig. 3, the parameter q stays
bounded away from zero, whereas the plots for the first excited state in Fig. 6 could
be extended up to q = 0. Let us consider how this can be understood directly from the

402

F. Finster et al. / Nuclear Physics B 584 (2000) 387414

Fig. 6. Reciprocal coupling limit: the energy spectrum and characteristic radii for the first excited
state.

Fig. 7. Reciprocal coupling limit: The first excited state for q = 0.285092, E = 0.16431.

Fig. 8. Reciprocal coupling limit: the first excited state for q = 9.75946, E = 0.405186.

equations. The parameter q enters only the YM equation (3.9). In the limit q 0, this
equation goes over to the vacuum YM equation, which has only the trivial solution w 1.
Hence if we assume that the spinors have a finite limit for q 0, then w(r) must go
uniformly in r to one. This shows that the solutions can be regular for q 0 only if
w satisfies the boundary condition limr w(r) = +1. In particular, our ground state
solutions cannot be regular in this limit. We next consider the limit q 0 for the first
excited state in more detail. Since w converges uniformly in r to one, the reciprocal limit

F. Finster et al. / Nuclear Physics B 584 (2000) 387414

403

equations (3.6), (3.7), and (3.11) go over to the DiracNewton equations


1

0 = 2,
r

(3.39)

1
0 = (E + ) ,
r
r 2 1(r) = 2 .

(3.40)
(3.41)

These equations are obtained by taking the nonrelativistic limit of the ED equations [3],
and according to the results obtained in that paper, the equations (3.39)(3.41) together
with the normalization integral (3.12) have a countable number of solutions, characterized
by the rotation number of the spinors (called the ground state, the first excited state, etc.).
) corresponding to solutions of the reciprocal
We thus expect that the functions (, ,
limit equations should for q 0, go over to a solution of the DiracNewton equations.
The behavior of the YM potential w can now be analyzed in more detail by taking the
) of the DiracNewton equations as a given inhomogeneity in the YM
solution (, ,
equation (3.9) and performing a perturbation calculation for small q. More precisely, the
ansatz w(r) = 1 + q u(r) to first order in q, leads to the linear equation

r 2 u00 = 2u + r ,
which can be solved by integration. Fixing the integration constants with our boundary
conditions u(0) = u() = 0 and u0 (0) = 0, we obtain the unique solution
Zr
u(r) = r

2
0

ds
s4

Zs

2
dt r
t(t)(t)
3

1
dt.
(t)(t)
t2

(3.42)

This consideration shows that for q 0, the rotation number of w is uniquely determined
by the rotation number of the spinors. Furthermore, one sees that in the limit q 0, the
Dirac wave function is determined by the DiracNewton equations (3.39)(3.41). Thus
only the gravitational attraction is responsible for the formation of the bound state, whereas
the YM field has no influence on the spinors.

4. Solutions of the EDYM equations


In this section, we shall construct numerical solutions of the full EDYM equations
and discuss their properties. Our method is to first find special solutions which are small
perturbations of either the BM solutions [1] or solutions to the reciprocal limit equations of
the previous section. We then trace these solutions while gradually changing the coupling
constants. This yields one-parameter families of solutions which can be extended even to
regions in parameter space where the solutions are far from all of the known limiting cases.
In order to simplify the connection between the EDYM equations and the reciprocal
limit equations of Section 3, it is useful to introduce a parameter > 0 in such a way that
the reciprocal limit equations are obtained when 0. To this end, we parametrize the
EDYM equations in terms of the new variables (, q, E) as follows,

404

F. Finster et al. / Nuclear Physics B 584 (2000) 387414

q
,

= + qE.
q

= (q) 2 ,

e2 =

1
m= ,
q

Since the EDYM equations involve three dimensionless parameters (namely m2 , /m,
and e2 ), introducing (, q, E) is merely a transformation to new independent parameters,
prescribing at the same time the gravitational constant (this means that we give up the
freedom to rescale r by fixing our length scale). In the limit 0, both q and E go over
to the corresponding parameters of the reciprocal limit system (see (3.8) and (3.5)). Also,
it is easy to check that the limits (3.4), (3.10), and (3.16) are satisfied if we let 0 and
keep (q, E) fixed. The parameters and q can be written in dimensionless form as
m2
,
q = m2 e 2 .
(4.1)
e2
Thus describes the relative strength of gravity versus the YM interaction, whereas q is the
product of the gravitational and YM coupling constants. Up to a scale factor, E = m.
Since is the relativistic energy and m the rest mass of the Dirac particle, E can, exactly
as in the previous section, be interpreted as the energy of the Dirac particle. Finally, we
also describe the binding energy by a parameter which corresponds to our notation for the
reciprocal limit system (3.20) and set
b = m .
B
q
=

For the construction of numerical solutions, we use a two-parameter shooting algorithm


combined with a rescaling method. Since this technique is quite similar to that described
for the reciprocal limit equations in the previous section, we shall merely outline our
procedure. In the first step of the construction, we consider the EDYM equations (2.23),
(2.24) and (2.26)(2.28) with the side conditions
Z
T
0 < := ( 2 + 2 ) dr < ,
A
2

(4.2)

0 < := lim T (r) < ,


r

lim w(r) = 1,

(4.3)
(4.4)

together with the following expansions near r = 0,


(r) = 1 r + O(r 3 ),

(r) = O(r 2 )

A(r) = 1 + O(r 2 ),

T (r) = 1 + O(r 2 ).

For fixed and q, we thus have the two parameters 1 and E to characterize a solution
of this modified EDYM system near the origin r = 0. On the other hand, we must satisfy
two conditions at infinity; namely, w must converge to 1, (4.4), and the spinors must go
asymptotically to zero in order for the normalization integral to be finite (4.2). Hence, we
can apply a two-parameter shooting method as described in Section 3. In order to have

F. Finster et al. / Nuclear Physics B 584 (2000) 387414

405

optimal boundary conditions at the upper end point r = r1 , we again match this with the
solution of the autonomous vacuum YM equation (see (3.32) and (3.33)). The shooting
method was again implemented in M ATHEMATICA, using an accuracy of 32 digits. For
each solution constructed in this way, we verify that (4.3) is satisfied and that the ADM
mass is finite (2.32). Once we have found a solution of the modified equations, we rescale
the solution according to

= 2 (2 r),
(r)
(4.5)
(r)

= 2 (2 r),
2
1
2
e

T (r) = T ( r),
(4.6)
A(r) = A( r),
and transform the parameters (m, , , e2 ) as follows,
m
= 2 m,

= 2 ,

(4.7)

= ,

e = e .

(4.8)

2 2

A straightforward calculation shows that the so-rescaled solution satisfies the EDYM
equations (2.23), (2.24) and (2.26)(2.28) together with the original side conditions (2.25)
and (2.30)(2.33). The parameters (, q, E) transform under the rescalings as
= ,

q = 4 q,

e = 4 (E + ( 1)m).
E

(4.9)

In the limit 0, these transformations coincide with the rescalings of the reciprocal
limit equations (3.34)(3.36). However, we remark that for the ED equations [3] a much
different rescaling is used. Namely, in order to get a better correspondence to the reciprocal
limit equations, we here scale the gravitational constant , whereas in [3] is fixed to be
1 throughout. Clearly, only the rescaled solutions have physical significance. Therefore, in
what follows we will consider only the rescaled solutions and again omit the tilde.
In a realistic physical situation, the gravitational coupling is very weak, whereas the YM
coupling constant is of order one, i.e., m2  1 and e2 1. Hence, according to (4.1), we
will here only investigate the parameter range  1, and we are particularly interested
in the situation for small q. In the limit 0, there are known solutions of our EDYM
system, namely the BM solutions (which more precisely are solutions in the limits 0
and e2 / 1), and the solutions of the reciprocal limit system constructed in Section 3.
We take these special solutions as the starting point for the numerics. By varying the
parameters and q and tracing the solutions with our shooting and rescaling methods, we
obtain a two-parameter family of solutions. In order to reduce the computational workload,
we did not step systematically through the two-parameter space, but always kept one
parameter fixed while varying the other parameter. Since remains unchanged under the
rescaling (see (4.9)), it is most convenient to construct one-parameter families of solutions
for different, fixed values of .
We now describe the solutions we found. Exactly as for the reciprocal limit system in
Section 3, we restricted attention to solutions with rotation number zero for the spinors
and a rotation angle of or 2 for the YM potential. We again refer to these types of
solutions as the ground state and the first excited state, respectively. For the ground
state solutions, the energy spectrum and the characteristic radii are in Figs. 9 and 10 plotted
for different values of the parameter (the characteristic radii are again defined by (3.37)).

406

F. Finster et al. / Nuclear Physics B 584 (2000) 387414

Fig. 9. The energy spectrum and characteristic radii for the EDYM ground state and = 0 (A),
= 0.0003933 (B), = 0.005322 (C), and = 0.02067 (D).

Fig. 10. Details of the energy spectrum and characteristic radii for the EDYM ground state and
= 0 (A), = 0.0003933 (B), = 0.005322 (C), and = 0.02067 (D).

F. Finster et al. / Nuclear Physics B 584 (2000) 387414

407

Fig. 11. The EDYM ground state for q = 8.8373, E = 0.3367, and = 0.0003933.

The curves A for = 0 coincide with the plots for the reciprocal limit system in Fig. 3.
For small values of the parameter , there are solutions for the EDYM equations which
are close to the solutions of the reciprocal limit equations (compare the curves A and B in
Fig. 9). In this parameter regime, the EDYM solutions look typically as shown in Fig. 11.
The metric functions A and T are both close to one; thus the gravitational interaction is
weak, in agreement with our considerations after (3.4). The spinors and the YM potential
look very similar to the solution of the reciprocal limit equations in Fig. 5. We conclude that
the reciprocal limit system of Section 3 indeed describes a significant limiting case of the
EDYM equations. However, one also sees that even for small , not all the solutions of the
EDYM equations are close to the reciprocal limit solutions. More precisely, curve B leaves
the vicinity of curve A at q 10 (see Fig. 10). If one follows curve B after it branches off
from curve A, the parameter q first increases up to a turning point, and then decreases to
q = 0. If gets large, the solution curves no longer come so close to the reciprocal limit
solutions (see curves C and D). The maximum of q decreases (see curve C) and finally
disappears (see curve D). Fig. 12 shows a typical solution for small q. We note that in this
parameter region, the metric functions A and T are not near one; this explains why the
reciprocal limit equations are no longer a good approximation. Indeed, the potentials w,
A, and T now resemble a BM solution of the EYM equations [1], and the spinors look like
the solution of the Dirac equation in the BM background. Hence q 0 corresponds to the
limit of weakly coupled spinors; i.e., spinors in a fixed BM background. Notice that the
characteristic radii go to zero and the energies go to infinity in the limit q 0 (see Fig. 9).
This can be understood from our rescalings. Namely, for the (unscaled) solutions of our
modified EDYM system, the BM solutions are easily obtained by taking the limit 1 0
(in which the spinors go uniformly in r to zero). In this limit, the normalization integral
(4.2) tends to zero, and thus the rescalings (4.5)(4.9) lead to a singular behavior of the
rescaled solutions for q 0. To summarize, there is a one-parameter family of solutions

408

F. Finster et al. / Nuclear Physics B 584 (2000) 387414

Fig. 12. The EDYM ground state for q = 0.005881, E = 17867.5, and = 0.005322.

(obtained by continuously changing the coupling constants), connecting the BM solutions


to our reciprocal limit solutions
We remark that our plots of the curves B have a small gap at q 8.7. The reason is that
in this region the numerics become unstable, and could not be carried out with our methods.
But we were able to construct two branches of solutions which approach the problematic
region from both sides. We suspect that the instability of the numerics is merely an artifact
of our rescaling method, but it might well be an indication for a possible bifurcation point
in this region. For the other curves C and D, we analyzed only the branch of solutions
which extends towards smaller values of q.
For the first excited state, the energy spectrum and characteristic radii are plotted in
Figs. 13 and 14 (the characteristic radii are again defined by (3.38)). The curves A for
= 0 correspond to the solutions of the reciprocal limit equations in Fig. 6. In contrast to
the situation for the ground state, the solutions for small are all close to the reciprocal
limit solutions (compare the curves A and B). Fig. 15 shows a typical solution for large q;
one sees that the spinors and YM potential look similar to those in Fig. 8. The form of
the energy spectrum and the characteristic radii gradually change when is increased; for
example, the cusp in the (q, rw )-plot becomes smooth (see curve D). It is interesting that
for q 0, the curves converge independent of to a single limit point (see Fig. 13). This
limit point was already described at the end of Section 3 as the case when the spinors
form a bound state due to their gravitational attraction, and the spinors generate a YM field
(see (3.39)(3.42)). This picture is in agreement with our numerics, since the spinors and
metric functions, for a solution near this limit point, look similar to the ED solutions [3] in
the Newtonian limit, and w 1 (see Fig. 16). The fact that this limit point is independent
of follows, because as explained at the end of Section 3, for q 0, the YM equation
decouples from the ED equations. For clarity, we point out that it would not be correct
to say that the gravitational interaction dominates the YM interaction in the limit q 0.

F. Finster et al. / Nuclear Physics B 584 (2000) 387414

409

Fig. 13. The energy spectrum and characteristic radii for the first excited EDYM state and = 0 (A),
= 0.0003933 (B), = 0.005322 (C), and = 0.02067 (D).

Fig. 14. Details of the energy spectrum and characteristic radii for the first excited EDYM state and
= 0 (A), = 0.0003933 (B), = 0.005322 (C), and = 0.02067 (D).

410

F. Finster et al. / Nuclear Physics B 584 (2000) 387414

Fig. 15. The first excited EDYM state for q = 9.548, E = 0.4153, and = 0.0003933.

Fig. 16. The first excited EDYM state for q = 0.09013, E = 0.1634, and = 0.02067.

Namely, according to (4.1), the ratio of the gravitational and YM coupling constants is
kept fixed, and thus q 0 corresponds to the limit where both coupling constants go to
zero at the same rate. Nevertheless, the YM field has for q 0 no influence on the energy
spectrum and the characteristic radii.
A main qualitative difference between the ground state and the first excited state is that
for the first excited state, we could not continuously join the solutions of the reciprocal
limit equations with a BM solution. In order to see how this comes about, we did numerical
calculations starting with a Dirac particle in the BM background (similar to that shown in
Fig. 17) and gradually increased the coupling of the spinors to gravity and to the YM field.

F. Finster et al. / Nuclear Physics B 584 (2000) 387414

411

Fig. 17. The deformation of the first excited BM state for q = 0.00047737, E = 367616, and
= 0.02067.

Fig. 18. The energy spectrum and characteristic radii for the deformation of the first excited BM state
for = 0.005322 (C) and = 0.02067 (D).

For these deformations of the first excited BM state, the curves of the energy spectrum
and the characteristic radii have spirals, whose size and shape drastically changes when
is increased, see Figs. 18 and 19. In the parameter regime where the energy plots spiral
around, the spinors have self-intersections similar as observed for the ED solutions [3],
see Fig. 20.
We now discuss the stability of our solutions. The relevant parameter for the stability

412

F. Finster et al. / Nuclear Physics B 584 (2000) 387414

Fig. 19. The energy spectrum and characteristic radii for the deformation of the first excited BM state
for = 0.0003933 (B), = 0.005322 (C), and = 0.02067 (D).

Fig. 20. The deformation of the first excited BM state for q = 0.0005528, E = 729921, and
= 0.0003933.

b Namely, if B
b is negative and smaller than the total
analysis is the total binding energy B.
energies of all other states, then energy is needed to break up the binding or to make a
transition to any of the other states, and therefore for physical reasons the solution must be
stable. Clearly, this energy argument does not provide a rigorous stability proof, and it also
cannot replace the numerical analysis of linear stability (like, e.g., in [2] or [3, Section 8]),
but it gives a strong indication for stability and is therefore commonly used (see, e.g., [9]).

F. Finster et al. / Nuclear Physics B 584 (2000) 387414

413

Fig. 21. Detailed plots of the total binding energy of the EDYM ground state for = 0.0003933 (B)
and = 0.005322 (C).

Let us first apply this energy argument to the ground state solutions of Figs. 9 and 10.
One sees that the total energy becomes negative for large q. For the curves B and C, this
region is plotted in more detail in Fig. 21. For the solutions on branch b, the total binding
energy is minimal, and thus this branch is stable. Applying Conley index methods with q
as the bifurcation parameter (see [10]), we obtain, as in [3], that the two other branches
a and c are unstable. Indeed, the instability of branch c follows also from the continuity
of the Conley index and the fact that in the limit q 0, this branch goes over to the
ground state BM solution which is known to be unstable [2]. When is increased (see
curve D, Fig. 9 and 10), only one branch of solutions remains, which comprises the BM
solutions as a limiting case and is therefore unstable. More precisely, the one-parameter
family has in this case no bifurcation points, and in the limit q 0 the solutions tend to
an unstable BM solution. Thus using Conley index techniques, it follows that the entire
one-parameter family is unstable. We conclude that for small , there is a stable branch
of ground state solutions for which q lies in a finite interval away from q = 0; all other
ground state solutions are unstable. For the stability of the first excited state, we consider
the plots of Figs. 13 and 14. Since for q 0, the spinors and metric functions go over to
the Newtonian limit of the ED solutions, we conclude from [3] that the branch of solutions
starting at q = 0 should be stable. This is in agreement with our above energy argument,
b is negative, and is smaller than the total
because on this branch the total binding energy B
binding energy of the second branch of solutions, which comes out of the bifurcation point
located at the maximum of q. Again, Conley index theory yields that this second branch
is unstable. For the deformations of the first excited BM state, the total binding energy is
positive (see Fig. 19 and 20), and hence these solutions should be unstable. Indeed, for the
branch of solutions which extends up to q = 0 (i.e., before the first bifurcation point), this
also follows from the continuity of the Conley index and the instability of the first excited
BM solution.
References
[1] R. Bartnik, J. McKinnon, Particlelike solutions of the EinsteinYangMills equations, Phys.
Rev. Lett. 61 (1988) 141144.
[2] N. Straumann, Z. Zhou, Instability of the BartnikMcKinnon solution, Phys. Lett. B 237 (1990)
353356.

414

F. Finster et al. / Nuclear Physics B 584 (2000) 387414

[3] F. Finster, J. Smoller, S.-T. Yau, Particlelike solutions of the EinsteinDirac equations, Phys.
Rev. D 59 (1999) 104020, gr-qc/9801079.
[4] F. Finster, J. Smoller, S.-T. Yau, Particlelike solutions of the EinsteinDiracMaxwell equations,
Phys. Lett. A 259 (1999) 431436, gr-qc/9802012.
[5] L.D. Landau, E.M. Lifshitz, Quantum Mechanics, Pergamon Press, 1977.
[6] F. Finster, J. Smoller, S.-T. Yau, Non-existence of time-periodic solutions of the Dirac equation
in a ReissnerNordstrm black hole background, J. Math. Phys. 41 (2000) 21732194, grqc/9805050.
[7] C.N. Yang, T.T. Wu, Some solutions of the classical isotopic gauge field equations, in: H. Mark,
S. Fernbach (Eds.), Properties of Matter Under Unusual Conditions, Wiley Interscience, New
York, 1969, pp. 349354.
[8] J. Stoer, R. Bulirsch, Numerische Mathematik 2, 3rd edn., Springer, 1990.
[9] T.D. Lee, Mini-soliton stars, Phys. Rev. D 25 (1987) 36403657.
[10] J. Smoller, Shock Waves and Reaction-Diffusion Equations, 2nd edn., Springer, 1994.

Nuclear Physics B 584 (2000) 415435


www.elsevier.nl/locate/npe

Brane surgery: energy conditions, traversable


wormholes, and voids
Carlos Barcel 1 , Matt Visser
Physics Department, Washington University, Saint Louis, MO 63130-4899, USA
Received 5 April 2000; revised 10 May 2000; accepted 6 June 2000

Abstract
Branes are ubiquitous elements of any low-energy limit of string theory. We point out that negative
tension branes violate all the standard energy conditions of the higher-dimensional spacetime they
are embedded in; this opens the door to very peculiar solutions of the higher-dimensional Einstein
equations. Building upon the (3 + 1)-dimensional implementation of fundamental string theory, we
illustrate the possibilities by considering a toy model consisting of a (2 + 1)-dimensional brane
propagating through our observable (3 + 1)-dimensional universe. Developing a notion of brane
surgery, based on the IsraelLanczosSen thin shell formalism of general relativity, we analyze
the dynamics and find traversable wormholes, closed baby universes, voids (holes in the spacetime
manifold), and an evasion (not a violation) of both the singularity theorems and the positive mass
theorem. These features appear generic to any brane model that permits negative tension branes:
This includes the RandallSundrum models and their variants. 2000 Elsevier Science B.V. All
rights reserved.
PACS: 04.60.Ds; 04.62.+v; 98.80.Hw
Keywords: Branes; Brane surgery; Energy conditions; Wormholes; Voids

1. Introduction
Branes, ubiquitous elements of any low-energy limit of string theory, have recently
attracted much attention as essential ingredients of the semi-phenomenological Randall
Sundrum models [1,2]. These models have been used to both ameliorate the hierarchy
problem [1] and to explore the possibility of exotic KaluzaKlein theories with their
infinitely large extra dimensions [2]. Essential ingredients in these RS models are the
existence of both positive and negative tension branes.
Corresponding author. E-mail: visser@kiwi.wustl.edu; Homepage: http://www.physics.wustl.edu/ visser
1 E-mail: carlos@hbar.wustl.edu; Homepage: http://www.physics.wustl.edu/ carlos

0550-3213/00/$ see front matter 2000 Elsevier Science B.V. All rights reserved.
PII: S 0 5 5 0 - 3 2 1 3 ( 0 0 ) 0 0 3 7 9 - 5

416

C. Barcel, M. Visser / Nuclear Physics B 584 (2000) 415435

Now a brane tension is normally thought of as being completely equivalent to an internal


cosmological constant, and from the point of view of physics constrained to the brane this
is certainly correct. However, from the higher-dimensional point of view (that is, as seen
from the embedding space) this is not correct: For a (p + 1)-brane embedded in (n + 1)
dimensions a brane tension leads to the stress energy
!
np
X

np a


( ) = D g
na na np (a ),
(1.1)
T = D ginduced
a=1

where the sum runs over the np normals to the brane, and the a are suitable Gaussian
normal coordinates. Contracting with a higher-dimensional null vector, k , we see
" np
#
X
2 np a

np a

( ) = D
( ).
(1.2)
na k
T k k = D ginduced k k
a=1

If the brane tension is negative, D < 0, and the null vector is even slightly orthogonal to
the brane, then on the brane
T k k < 0.

(1.3)

That is, the embedding-space null energy condition (NEC) is violated. In fact, integrating
across the brane, even the averaged null energy condition (ANEC) is violated. (Ipso facto,
all the energy conditions are violated.) This is a classical violation of the energy conditions,
which we shall soon see is even more profound than the classical violations due to nonminimally coupled scalar fields [3].
In a recent series of papers [46] we have made a critical assessment of the current
status of the energy conditions, finding a variety of both classical and quantum violations
of the energy conditions. We now see that uncontrolled violations of the energy conditions
are also a fundamental and intrinsic part of any brane-based low-energy approximation to
fundamental string theory. Among the possible consequences of these energy condition
violations we mention the occurrence of traversable wormholes (violations of topological
censorship), possible violations of the singularity theorems (more properly, evasions of the
singularity theorems), and even the possibility of negative asymptotic mass.
A particular example of this sort of phenomenon occurs in the (finite size) Randall
Sundrum models, where one has two parallel branes (our universe plus a hidden brane)
of equal but opposite brane tension. One or the other of these branes (depending on
whether one is considering the RS1 or RS2 model) violates the (4 + 1)-dimensional energy
conditions and exhibits the flare out behaviour reminiscent of a traversable wormhole [7].
That these branes do not quite represent traversable wormholes in the usual sense [8,9]
follows from the fact that the throat is an entire flat (3 + 1) Minkowski space, instead
of the more usual R 1 S d1 . Furthermore, in the infinite-size version of the Randall
Sundrum (RS2) model, where the hidden sector has been pushed out to hyperspatial
infinity, our universe is itself represented by a positive-tension (3 + 1)-brane, which does
not violate any (4 + 1)-dimensional energy conditions. The energy-condition violating
brane has in this particular model been pushed out to hyperspatial infinity and discarded. Be

C. Barcel, M. Visser / Nuclear Physics B 584 (2000) 415435

417

that as it may, the occurrence of negative tension branes in modern semi-phenomenological


models is generic, and a feel for some of the peculiar geometries they can engender is
essential to developing any deep understanding of the physics.
In this particular paper we shall for illustrative purposes choose a particularly simple
model: We work with a (3 + 1)-dimensional bulk, which contains a (2 + 1)-dimensional
brane (of either positive or negative brane tension). We choose this particular model
because it is sufficiently close to reality to make the points we wish to make as
forcefully as possible, and because it arises naturally in certain types of fundamental
string theory. While it is most often the case that fundamental string theories (or their
various offspring: membrane models, M-theory, etc.) are formulated in either (9 + 1) or
(10 + 1) dimensions, 2 this is not absolutely necessary: There is an entire industry based on
formulating string theories directly in (3 + 1) dimensions, with the price that has to be paid
being the inclusion of extra (1 + 1)-dimensional quantum fields propagating on the worldsheet [1014]. 3 Now even in such a (3 + 1)-dimensional incarnation of string theory, open
strings will terminate on D-branes (Dirichlet branes), and an effective theory involving the
(3+1)-dimensional bulk plus (2+1)-, (1+1)-, and (0+1)-dimensional D-branes (domain
walls, cosmic strings, and soliton-like particles) can be contemplated as a lowenergy approximation. 4 While D-branes are perhaps the most straightforward examples of
membrane-like solitons in string theory, they do come with additional technical baggage:
the most elementary implementation of D-branes occurs in bosonic string theories [15],
but often D-branes are associated with specific implementations of supersymmetric string
theories [16] and carry various types of RamondRamond or NeveuSchwarz charge.
There are in addition other types of brane-like configurations that sometimes arise in
fundamental string theory such as non-dynamical orientifold planes [16], which generate
gravitational fields corresponding to negative tensions, but which do not themselves exhibit
2 In many specific cases the actual implementation is directly in terms of a Euclidean-signature 10- or
11-dimensional spacetime; with the underlying Lorentzian-signature reality hidden under several layers of
scaffolding.
3 Consider for example the bosonic string, which is most often viewed as a (1 + 1)-dimensional world sheet
propagating in (25 + 1) dimensions: There is a trivial re-interpretation in which the bosonic string propagates
in (3 + 1) dimensions and there are 22 free scalar fields propagating on the world-sheet. These 22 scalar fields
are there just to soak up the conformal anomaly and make the theory manageable. If these scalar fields are
now constrained by appropriate identifications the re-interpretation is less trivial it is an example of the fact
that compactifications of some of the dimensions of the higher-dimensional embedding spacetime that the world
sheet propagates through can be traded off for a lower-dimensional uncompactified embedding spacetime plus
interacting fields on the world-sheet. When this procedure is applied to superstrings the technical details are
considerably more complex, but the basic result still holds.
4 More traditional string theorists who absolutely insist on working directly in the higher-dimensional
embedding space can view the current calculations as a particular toy model in which only selected sub-sectors of
the grand total degrees of freedom are excited. Additionally, it should be borne in mind that many of the generic
features of the analysis presented in this paper will extend mutatis mutandis to embedding spaces and branes of
higher dimensionality. You do not want the bulk to have fewer than (3 + 1) dimensions since then bulk gravity
is either completely or almost trivial. You do not want the bulk to have more than (10 + 1) dimensions since the
model is then difficult to interpret in terms of fundamental string theory. For technical reasons (to be able to use
the thin-shell formalism) you want the brane to be of co-dimension 1, so if the bulk is (n + 1)-dimensional the
brane should be ([n 1] + 1)-dimensional. Within these dimensional limitations, the qualitative features of this
paper are generic.

418

C. Barcel, M. Visser / Nuclear Physics B 584 (2000) 415435

internal dynamics. We will not delve further into this bestiary, but will instead content
ourselves with the observation that the low-energy limit of fundamental string theory
(of whatever persuasion) generically leads to an effective theory containing brane-like
excitations.
This overall picture is actually very similar to the notion of extended topological defects
arising from symmetry breaking in point-particle field theories: There are many semiphenomenological GUT-based point particle field theories that naturally contain domain
walls, cosmic strings, and/or solitons. The key difference here is that point particle field
theories inevitably lead to positive brane tensions, with negative brane tensions being energetically disfavoured (they correspond to an unnatural form of symmetry breaking that
forces one to the top of the potential). The key difference in brane-based models is that
there is no longer any particular barrier to negative brane tension in fact negative brane
tensions are ubiquitous, now being so commonly used as to almost not require explicit
mention. An exhaustive list of papers using negative tension branes would by now be impractical. Among many instances of their use (apart from [1] and [2]) we mention: [1726].
Within the model we have chosen, we demonstrate that negative tension branes lead to
traversable wormholes in some cases to stable traversable wormholes. (Positive tension
branes quite naturally lead to closed baby universes; these are not FLRW universes, and are
not suitable for cosmology, but are perhaps of interest in their own right.) We also explore
the possibility of viewing the brane as an actual physical boundary of spacetime, with the
region on the other side of the brane being null and void.
The basic tools used are the idea of Schwarzschild surgery as developed in [27] (see
also the more detailed presentation in [9]), which we first extend to brane surgery,
specialize to ReissnerNordstrmde Sitter surgery, and then use to present an analysis
of both static and dynamic spherically-symmetric (2 + 1)-dimensional branes in a (3 + 1)dimensional ReissnerNordstrmde Sitter background geometry. 5 We find both stable
and unstable traversable wormhole solutions, stable and unstable baby universes, and stable
and unstable voids.

2. Brane surgery
We start by considering a rather general static spherically symmetric geometry (not the
most general, but quite sufficient for our purposes)
ds 2 = F (r) dt 2 +

dr 2
+ r 2 d22 .
F (r)

(2.4)

To build the class of geometries we are interested in, we start by taking two copies of this
geometry, truncating them at some time-dependent radius a(t), and sewing the resulting
geometries together along the boundary a(t). The result is a manifold without boundary
5 As we shall soon see, brane surgery is essentially a specific implementation of the IsraelLanczosSen
junction conditions of general relativity; as such it has been used implicitly in many brane-related papers (see, for
example, [1,2,2830]); the key difference in the present paper is in the details and in the questions we address.

C. Barcel, M. Visser / Nuclear Physics B 584 (2000) 415435

419

that has a kink in the geometry at a(t). If we sew together the two external regions
r (a(t), ), then the result is a wormhole spacetime with two asymptotic regions. On
the other hand, if we sew together the two internal regions r (0, a(t)), then the result is a
closed baby universe.
At the kink a(t) the spacetime geometry is continuous, but the radial derivative (and
hence the affine connexion) has a step-function discontinuity. The Riemann tensor in this
situation has a delta-function contribution at a(t), and this geometry can be analyzed using
the IsraelLanczosSen thin shell formalism of general relativity [3235]. The relevant
specific implementation of the thin-shell formalism can be developed by extending the
formalism of [27] and [9]. Because of its relative simplicity we shall start with the static
case a = constant.
2.1. Brane statics
The unit normal vector to the sphere a = constant is (depending on whether one is
considering inward or outward normals)


 p

1

n = 0,
, 0, 0 .
(2.5)
n = 0, F (a), 0, 0 ,
F (a)
The extrinsic curvature (second fundamental form) can be written in terms of the normal
derivative
1 g
1 g
1p
g
= n
.
(2.6)
=
F (a)
K =

2
2
x
2
r
If we go to an orthonormal basis, the relevant components are 6
1p
gt t t t
1p
F (r) 1
g =
Ktt =
F (r)
F (r)
2
r
2
r F (r)


F
(r)
1
,
= F (r)1/2
2
r r=a


1p
g
1p
r 2 1
F (r)
g =
F (r)
F (r)
=
.
K =
2
r
2
r r 2
r r=a

(2.7)

(2.8)

The discontinuity in the extrinsic curvature is related to the jump in the normal derivative
of the metric as one crosses the brane
+

K
.
= K

(2.9)

In general, one could take the geometry on the two sides of the brane to be different
[F + (r) 6= F (r)], but in the interests of clarity the present models will all be taken to
have a Z2 symmetry under interchange of the two bulk regions. 7 Under these conditions
6 The use of an orthonormal basis makes it particularly easy to phase the calculation in terms of the physical
density and physical pressure.
7 Remember that we have already decided to take the range of the r coordinate to be either two copies of
(a(t), ), corresponding to a wormhole; or two copies of (0, a(t)), corresponding to a baby universe. Then Z2
symmetry corresponds to F + (r) = F (r), with a kink in the geometry at r = a(t). Our normal vectors do not
flip sign as we cross the brane.

420

C. Barcel, M. Visser / Nuclear Physics B 584 (2000) 415435

tt = F (r)

1/2


F (r)
,
r r=a


F (r)
= 2
r

(2.10)

(2.11)

r=a

The upper sign refers to a wormhole geometry where the two exterior regions have been
sewn together (discarding the two interior regions), while the lower sign is relevant if one
has kept the two interior bulk regions.
The thin-shell formalism of general relativity [3235] relates the discontinuity in
extrinsic curvature to the energy density and tension localized on the junction: 8


1
1 p
= =
F (r) ,
(2.12)
4
2r
r=a

1  p
1
r F (r) .
= [ tt] =
8
4r r
r=a

(2.13)

If the material located in the junction is a clean brane (a brane in its ground state, without
extra trapped matter in the form of stringy excitations), then its equation of state is =
and the condition for a static brane configuration (either a wormhole or baby universe
geometry) is simply


p
 p

r F (r)
=
= 2 F (r)
r=a
r
r=a


F (r)

= 0.
(2.14)
r
r 2 r=a
Thus we have a very simple result: static wormholes (baby universes) correspond to
extrema of the function F (r)/r 2 , though at this stage we have not yet made any assertions
about stability or dynamics. The only difference between wormholes and baby universes
is that for wormholes the brane tension must be negative, whereas for baby universes it is
positive.
It is instructive to note that the locations of these static brane solutions correspond to
circular photon orbits in the original spacetime (and this is true for arbitrary F (r)). That
is: at these static brane solutions any particle that is emitted form the brane, which
then follows null geodesics (of the bulk spacetime), and which initially has no radial
momentum, will just skim along the brane; never moving off into the bulk. (Note that
this is a purely kinematic effect that occurs over and above any trapping due to stringy
interactions between the brane and excited string states.)
This may easily be verified by considering the photon orbits for arbitrary F (r). The
time-translation and rotational Killing vectors lead to conserved quantities
8 The numerical coefficients appearing herein are dimension-dependent (because of the implicit trace over the
Ricci tensor and extrinsic curvature hidden in the Einstein equations).

C. Barcel, M. Visser / Nuclear Physics B 584 (2000) 415435

, k = 
t

gt t

, k = `

421

dt
= 
d

dt
= .
d

(2.15)

d
=`
d

r2

d
= `.
d

(2.16)

Inserting this back into the condition that the photon momentum be a null vector, (k, k) = 0,
we see
 2
F (r)`2
dr
+
=  2.
(2.17)
d
r2
Now is an arbitrary affine parameter, so we can reparameterize ` and define
b = /` to see that photon orbits are described by the equation
 2
F (r)
dr
+ 2 = b2 .
(2.18)
d
r
The circular photon orbits (and at this stage we make no claims about stable versus unstable
circular photon orbits) are, as claimed, at extrema of the function F (r)/r 2 (which coincide
with the location of the static brane configurations).
2.2. Brane dynamics
If now the brane is allowed to move radially a a(t), we start the analysis by first
parameterizing the motion in terms of proper time along a curve of fixed and . That is:
the brane sweeps out a world-volume

(2.19)
X (, , ) = t ( ), a( ), , .
The 4-velocity of the (, ) element of the brane can then be defined as


dt da
, , 0, 0 .
V =
d d

(2.20)

Using the normalization condition and the assumed form of the metric, and defining a =
da/d ,
!
p

 p
a
F (a) + a 2

, a,
0, 0 ,
V = F (a) + a 2 ,
, 0, 0 .
(2.21)
V =
F (a)
F (a)
The unit normal vector to the sphere a( ) is


a p

2
, F (a) + a , 0, 0 ,
n =
F (a)

p
n = a,

!
F (a) + a 2
, 0, 0 . (2.22)
F (a)

The extrinsic curvature can still be written in terms of the normal derivative
1 g
.
K = n
2
x
If we go to an orthonormal basis, the component is easily evaluated [9,27]

(2.23)

422

C. Barcel, M. Visser / Nuclear Physics B 584 (2000) 415435

p
1p
g
F (a) + a 2
2
g =
.
(2.24)
K =
F (a) + a
2
r
a
The component is a little messier, but generalizing the calculation of [27] or [9] (which
amounts to calculating the four-acceleration of the brane) quickly leads to 9


1
dp
1
dF (r)
+ 2a =
K = p
F (a) + a 2 .
(2.25)
2 F (a) + a 2
da
da
Applying the thin-shell formalism now gives:
1 p
F (a) + a 2 ,
(2.26)
=
2a

1 d p
a F (r) + a 2 .
(2.27)
=
4a da
These equations can easily be seen to be compatible with the conservation of the stress
energy localized on the brane

d 2
d
a2 =
a .
(2.28)
d
d
So as usual, two of these three equations are independent, and the third is redundant.
From the above we see that traversable wormhole solutions, corresponding to the
minus sign above, require negative brane tension (and so positive internal pressure and
negative internal energy density). This is in complete agreement with [31] where it was
demonstrated that even for dynamical wormholes there must be violations of the null
energy condition at (or near) the throat.
If the material located in the junction is again assumed to be a clean brane ( = )
then all the dynamics can be reduced to a single equation 10
a 2 + F (a) = (2 )2 a 2 .

(2.29)

This single dynamical equation applies equally well to both wormholes and baby universes,
the that shows up in the brane Einstein equations quietly goes away upon squaring
thus for questions of dynamics and stability these surgically constructed baby universes
and wormholes can be dealt with simultaneously the only difference lies in question of
whether the brane tension is positive or negative.
Note that we could re-write this dynamical equation as


d ln(a) 2 F (a)
+ 2 = (2 )2 .
(2.30)
d
a
From this it is clear that static solutions must be located at extrema of the function F (a)/a 2 ,
in agreement with the static analysis.
9 We do not repeat the details here since this calculation is now standard textbook fare [9], pp. 182183. If one
wishes to avoid the need for this particular calculation one can instead work backwards from the conservation of
stress-energy, together with the already-calculated expression for K , to deduce an expression for K . But if
you choose this route you lose the opportunity to make a consistency check.
10 And if the brane is not clean in this sense one only needs to keep track of one additional piece of information
the on-brane conservation equation (2.28).

C. Barcel, M. Visser / Nuclear Physics B 584 (2000) 415435

423

In the next section we shall make use of this general formalism by specializing F (r) to
the ReissnerNordstrmde Sitter form. We shall then exhibit some explicit solutions to
these brane equations of motion, and perform the relevant stability analysis.

3. ReissnerNordstrmde Sitter surgery


For the ReissnerNordstrmde Sitter geometry
2M Q2 2
+ 2 r .
r
r
3
It is then most instructive to write the dynamical equation in the form


d ln(a) 2
+ V (a) = E,
d
F (r) = 1

(3.31)

(3.32)

with a potential
1
2M Q2
F (a)
= 2 3 + 4 ,
2
a
a
a
a
3
and an energy
V (a) =

(3.33)

E = +(2 )2 .

(3.34)

The extrema of this potential are easily located, their positions are independent of and
occur at
s


3M
3M 2

2Q2 .
(3.35)
r =
2
2
(As promised, these are indeed the locations of the circular photon orbits of the Reissner
Nordstrmde Sitter geometry; note that the cosmological constant does not affect the
location of these circular photon orbits.) The value of this potential at these extrema is
somewhat tedious to calculate, we find

V (M, Q, ) V r (M, Q)




3/2

9 M 2 27 M 4
M
3M 2
1
2
+

2Q
.

= 2 1
4Q
2 Q2
8 Q4
4Q6
2
3
(3.36)
Though it may not be obvious, the Q 0 limit formally exists and is given by
V (M, Q 0, ) ,

V+ (M, Q 0, )

.
2
27M
3

(3.37)

The behaviour of the potential V (a) is qualitatively:


V (a) + as a 0 (Q 6= 0);
V (a) /3 as a +;
There is at most one local minimum (V located at r ) and one local maximum (V+
located at r+ ).

424

C. Barcel, M. Visser / Nuclear Physics B 584 (2000) 415435

Two figures, where we have plotted V (a) for two special cases, are provided in the
discussion below. When looking for a stable brane solution we want to satisfy the
following:
1. We want the local minimum to exist, and the brane to be located in its basin of
attraction.
2. The energy must be at least equal to V (to even get a solution), and should not
exceed V+ (to avoid having the solution escape from the local well located around
r ).
3. We also do not want (at least for now) the brane to fall inside (or even touch) any
horizon the original ReissnerNordstrmde Sitter geometry might have for two
reasons:
(a) Because if it does fall inside (or even touch) an event horizon the wormhole
geometry is operationally indistinguishable from a ReissnerNordstrmde Sitter black hole and therefore not particularly interesting (but see the discussion
regarding singularity avoidance later in this paper) whereas the baby universe
geometry is for Q = 0 doomed to a brief and unhappy life, and for Q 6= 0 is just
plain weird.
(b) For technical reasons (r is now timelike and t spacelike) a few key minus signs
flip at intermediate steps of the calculation, more on this later.
These physical constraints now imply:

1. To get a local minimum we need M > 8/9Q.


2. To then trap the solution, to make it one of bounded excursion, we need
V (M, Q, ) 6 +(2 )2 6 V+ (M, Q, ).

(3.38)

3. Horizon avoidance requires F (a) > 0 over the entire range of motion; this implies
V (a) =

F (a)
>0
a2

V (M, Q, ) > 0.

(3.39)

In view of this the horizon avoidance condition might more properly be called horizon
elimination horizons can be avoided if and only if the inner and outer horizons are
actually eliminated. (We could however still have a cosmological horizon at very
large distances, this cosmological horizon is never reached if the bounded excursion
constraint is satisfied.) We can also explicitly separate out the cosmological constant
to write the horizon elimination condition as
< 3V (M, Q, 0),

(3.40)

which makes it clear that a powerful enough negative (bulk) cosmological constant
is guaranteed to eliminate all the event horizons from the geometry.
That these constraints can simultaneously be satisfied (at least in certain parameter
regimes) can now be verified by inspection. The best way to proceed is to sub-divide the
discussion into several special cases.

C. Barcel, M. Visser / Nuclear Physics B 584 (2000) 415435

425

3.1. M > |Q| = 0


There is one maxima (at a = 3M) and no minimum. There are no stable solutions,
though the arbitrarily advanced civilization posited by Morris and Thorne [8] might
like to try to artificially maintain the unstable static solution at a = 3M. (This solution
is unstable to both collapse and explosion.)
Adding Q 6= 0 provides a hard core to the potential so that collapse is avoided (modulo
the horizon crossing issue which must be dealt with separately).
3.2. M > |Q| 6= 0
There are now both a local maximum (at r+ < 3M) and a global minimum (at r > 0).
The potential is plotted in Fig. 1. Stable solutions exist (both static stable solutions and
stable solutions of bounded excursion), but since V < 0 ( = 0) at the global minimum
horizon avoidance requires
< 3V (M, Q, 0) < 0.

(3.41)

That is, stable traversable wormhole or baby universe solutions exist only if the bulk is
anti-de Sitter space (adS(3+1)) with a strong enough negative cosmological constant.

Fig. 1. Sketch of the potential V (a) for M > |Q| and = 0. Adding a cosmological constant merely
moves the entire curve up or down: the lower horizontal line represents /3, and for sufficiently
large and negative the inner and outer horizons (which are given by the intersection of this horizontal
line with the = 0 potential) are guaranteed to be eliminated. The upper horizontal line represents
/3 + (2 )2 , and its intersection with the = 0 potential gives the turning points of the motion.
If inner and outer horizons exist they lie between the inner and outer turning points.

426

C. Barcel, M. Visser / Nuclear Physics B 584 (2000) 415435

Indeed, if you consider the original geometry prior to brane surgery and extend it down
to r = 0 then for this choice of parameters (because of the large negative cosmological
constant) you encounter a naked singularity. For the stable wormhole geometries based on
this brane prescription this is not a problem since the naked singularity was in the part of
the spacetime that you threw away in setting up the brane construction. (The baby universe
models on the other hand, while stable, explicitly do contain naked singularities.) 11
A particularly simple sub-class of these solutions occurs when the bulk cosmological
constant is tuned to a special value in terms of the brane tension. This is the analog of
the RandallSundrum fine tuning [1,2] and corresponds to a zero effective cosmological
constant, in the sense that the brane equation of motion can be rearranged and reinterpreted
as being governed by E = 0 and
effective = + 3(2 )2 .

(3.42)

If this effective is now tuned to zero


= 3(2 )2 < 3V (M, Q, 0) < 0.

(3.43)

3.3. M = |Q|
There are still both a local maximum (at r+ = 2M) and a global minimum (at r =
M). Stable solutions exist. Since now V ( 0) = 0 at the global minimum horizon
avoidance requires anti de Sitter space with an arbitrarily weak cosmological constant.
(And again this is an example of horizon elimination.)
3.4. M

8/9 |Q|, |Q|

There are still both a local maximum (at r+ < 2M) and a global minimum (at r > M).
The potential is plotted in Fig. 2. Stable solutions exist. Since now V ( 0) > 0 at the
global minimum, horizon avoidance can be achieved with zero cosmological constant in
the bulk. For instance, picking
= 0,

(2 )2 = V (M, Q, 0),

(3.44)

yields the stable static solution at r . This is perhaps the most physical of these
traversable wormholes in that it resides in an asymptotically flat spacetime.
11 This is part of a general pattern: The stable (or even merely static) brane configurations that do not
possess naked singularities in the bulk region are the wormhole configurations with negative brane tension.
This observation also applies to the other sub-cases discussed below. This is compatible with the discussion
of Chamblin, Perry, and Reall [36] who discovered qualitatively similar behaviour for (8 + 1)-dimensional branes
in a (9 + 1)-dimensional bulk. Specifically, they found that static (8 + 1)-dimensional brane configurations with
positive brane tension led to naked singularities in the bulk, and that eliminating the naked singularities forced
the adoption of negative brane tension (and implicitly a wormhole configuration). This observation also serves
to buttress our previous comments to the effect that the qualitative features of the calculations presented in this
paper are generic, and are not just limited to (2 + 1) branes in (3 + 1) dimensions.

C. Barcel, M. Visser / Nuclear Physics B 584 (2000) 415435

427

Fig. 2. Sketch of the potential V (a) for M in the critical range ( 8/9 |Q|, |Q|), and = 0. Adding
a cosmological constant merely moves the entire curve up or down. In this case, even for = 0, we
see a stable minimum at r with no event horizons. For small positive a cosmological horizon
will form at very large radius, but this is of no immediate concern because of the barrier at r+ . If
becomes too large however, > V (M, Q, 0), inner and outer horizons will reappear between
the inner and outer turning points.

3.5. M =

8/9 |Q|

The maximum and minimum merge into a single point of inflection (at r = 3M/2).
There are no stable solutions. All the solutions exhibit runaway to large radius.
3.6. M

8/9 |Q|, 0

There is not even a point of inflection: the potential is monotonic decreasing. There are
no stable solutions.
3.7. M = 0, Q 6= 0
There is not even a point of inflection: the potential is monotonic decreasing. There are
no stable solutions.

428

C. Barcel, M. Visser / Nuclear Physics B 584 (2000) 415435

3.8. M < 0
Letting the central mass M go negative is not helpful M < 0 helps stabilize against
collapse, but actually destroys the possibility of stable solutions because the location of
extrema r is pushed to unphysical nominally negative values of the radius.
3.9. Baby bangs?
The fact that so many of these baby universe models are unstable to explosion is
intriguing, and potentially of phenomenological interest. While these particular babyuniverse models are not suitable cosmologies for our own universe, we believe that more
realistic scenarios can be developed.
3.10. Singularity avoidance?
We have so far sought to implement horizon avoidance in our models: we have sought
conditions that would prevent the brane from falling through or even touching any horizon
that might be present in the underlying pre-surgery spacetime. Suppose we now relax that
constraint. The best way to analyze the situation is to note that inside the horizon (more
precisely between the outer horizon and the inner horizon) the pre-surgery metric can be
written in the form
ds 2 = +|F (r)| dt 2

dr 2
+ r 2 d22 .
|F (r)|

(3.45)

The calculation of the four-velocity, normal, extrinsic curvatures, and their discontinuities
can be repeated, with the result that in this region [F (r) < 0]
!
p
2 |F (a)|
a

, a,
0, 0 ,
V =
|F (a)|


q
a
, a 2 |F (a)|, 0, 0 ,
(3.46)
n =
|F (a)|
and

q
1
a 2 |F (a)|.
2a
After rearrangement this leads to the same dynamical equation as before


d ln(a) 2 F (a)
+ 2 = (2 )2 .
d
a
=

(3.47)

(3.48)

So that all of our previous arguments can be extended inside the event horizon.
A few key observations:
The two turning points occur at F (a)/a 2 = (2 )2 > 0. Thus F (a) > 0 at the turning

6=
points. So if there are horizons present (that is, if F (a) = 0 has solutions rhorizon
r ), and one is in the potential well near r , then one turning point will be outside the
outer horizon, and the second turning point will be inside the inner horizon.

C. Barcel, M. Visser / Nuclear Physics B 584 (2000) 415435

429

Even though the brane oscillation will take finite proper time this corresponds to
infinite t-parameter time when the brane re-emerges from the outer horizon it will
emerge from a past outer horizon of a future incarnation of the universe; the brane
will not re-emerge into our own universe. (For simplicity you may wish to set = 0
and consider the Penrose diagram of the maximally extended ReissnerNordstrm
geometry as presented, for instance, on page 158 of Hawking and Ellis [37]. A partial
Penrose diagram for ReissnerNordstrmde Sitter may be found in [38,39]. See also
Fig. 3.)
Operationally, from our asymptotically flat region, once the brane passes the horizon the geometry will be indistinguishable from an ordinary ReissnerNordstrm
de Sitter black hole.
The original pre-surgery spacetime has two asymptotic regions, two outer horizons,
and two inner horizons, which are then repeated an infinite number of times in the
maximal analytic extension. If the brane starts out in the rightmost asymptotic region
and falls through the right (future) outer horizon, then you can quickly convince
yourself that it must pass through the left inner horizon (twice, once on the way in,

Fig. 3. Sketch of the Penrose diagram for the maximally extended ReissnerNordstrm geometry
when M > |Q| ( = 0). A timelike (2 + 1)-brane [spacelike normal] will oscillate between the
turning points r+ and r , but each oscillation will take infinite coordinate time even if it takes finite
proper time. For wormhole solutions keep the right half of the diagram, make two copies, and sew
them together along the brane. For baby universe geometries keep the left half of the diagram, make
two copies and sew them up along the brane.

430

C. Barcel, M. Visser / Nuclear Physics B 584 (2000) 415435

and once again on the rebound) before moving back out through the right (past) outer
horizon back into the (next incarnation of) the right asymptotic region. (See Fig. 3.)
The wormhole geometry based on this brane surgery is an explicit example of
partial evasion of the usual singularity theorems [37]. (We say evasion, not violation,
because the presence of the negative tension brane vitiates the usual hypotheses
used in proving the singularity theorems.) The wormhole geometry certainly has
trapped surfaces once the brane falls inside the horizon, but by construction there
is no left curvature singularity. (The right curvature singularity is still there, and
the right inner horizon is still a Cauchy horizon.) 12 Note that this is a idealized
statement appropriate to clean wormhole universes containing only a few test
particles of matter: in any more realistic model where the universe contains a finite
amount of radiation, inner event horizons are typically unstable to a violent blue
shift instability, and are typically converted by back-reaction effects to some sort of
curvature singularity [38,39]. This process however, lies far beyond the scope of the
usual singularity theorems.
If you wish to eliminate both left and right singularities a more drastic fix is called for:
You will need to use a (3 + 0)-dimensional brane, something you might call an instantonbrane because it represents a spacelike hypersurface through the spacetime at early
times theres nothing there, the brane switches on for an instant, and then its gone again.
The simplest example of such a instanton-brane is to place one at r , the static minimum
of the potential V (a). 13 If there are event horizons then this minima will be inside the
event horizon (between inner and outer horizons) and a hypersurface placed at r will
be spacelike. Placing the instanton-brane at this location will eliminate both singularities
and both inner horizons you are left with two asymptotic regions and two (outer) event
horizons, infinitely repeated.
More generally one could think of an instanton-brane described by a location a(`),
where ` is now proper length along the brane (and the notion of dynamics is somewhat
obscure). The spacelike tangent and timelike normal are now (outside the horizon)
!
p
(da/d`)2 F (a) da

, , 0, 0 ,
V =
F (a)
d`



q
1 da
2
, (da/d`) F (a), 0, 0 ,
n =
F (a) d`
and a brief computation yields
q
1
(da/d`)2 F (a).
=
2a

(3.49)

(3.50)

12 If you think of the ReissnerNordstrmde Sitter geometry as arising from gravitational collapse of an
electrically charged star, then it is the left curvature singularity (which is eliminated by the present construction)
that would arise from the central density of the star growing to infinity. The right curvature singularity (which is
unaffected by the present construction) has a totally different genesis as it arises in a matter-free region due to
gravitational focussing of the electromagnetic field.
13 Although this is a static minimum of the usual V (a) it is in the present context not stable. This arises because
for a spacelike shell the overall sign of the potential flips.

C. Barcel, M. Visser / Nuclear Physics B 584 (2000) 415435

This can be rearranged to give




d ln(a) 2 F (a)
2 = (2 )2 .
d`
a

431

(3.51)

So the net result is that for an instanton-brane the sign of the potential has flipped, but that
of the brane contribution to the energy has not. (And exactly the same result continues to
hold inside the horizon, a few intermediate signs flip, but thats all.)
In summary: certain varieties of brane wormhole provide explicit evasions (either partial
or complete) of the usual singularity theorems.

4. Voids: the brane as a spacetime boundary


A somewhat unusual feature of brane physics is that the brane could also be viewed
as an actual physical boundary to spacetime, with the other side of the brane being
null and void. In general relativity as it is normally formulated the notion of an actual
physical boundary to spacetime (that is, an accessible boundary reachable at finite distance)
is anathema. The reason that spacetime boundaries are so thoroughly deprecated in general
relativity is that they become highly artificial special places in the manifold where some
sort of boundary condition has to be placed on the physics by an act of black magic.
Without such a postulated boundary condition all predictability is lost, and the theory is
not physically acceptable. Since there is no physically justifiable reason for picking any
one particular type of boundary condition (Dirichlet, Neumann, Robin, or something more
complicated), the attitude in standard general relativity has been to exclude boundaries, by
appealing to the cosmic censor whenever possible and by hand if necessary.
The key difference when a brane is used as a boundary is that now there is a specific
and well-defined boundary condition for the physics: D-branes (D for Dirichlet) are
defined as the loci on which the fundamental open strings end (and satisfy Dirichlet-type
boundary conditions). D-branes are therefore capable (at least in principle) of providing
both a physical boundary and a plausible boundary condition for spacetime. For Neveu
Schwarz branes the boundary conditions imposed on the fundamental string states are more
complicated, but they still (at least in principle) provide physical boundary conditions on
the spacetime.
When it comes to specific calculations, this may however not be the best mental picture
to have in mind after all, how would you try to calculate the Riemann tensor for
the edge of spacetime? And what would happen to the Einstein equations at the edge?
There is a specific trick that clarifies the situation: Take the manifold with brane boundary
and make a second copy, then sew the two manifolds together along their respective
brane boundaries, creating a single manifold without boundary that contains a brane, and
exhibits a Z2 symmetry on reflection around the brane. Because this new manifold is a
perfectly reasonable no-boundary manifold containing a brane, the gravitational field can
be analyzed using the usual thin-shell formalism of general relativity [3235]: The metric
is continuous, the connection exhibits a step-function discontinuity, and the Riemann

432

C. Barcel, M. Visser / Nuclear Physics B 584 (2000) 415435

curvature a delta-function at the brane. The dynamics of the brane can then be investigated
in this Z2 -doubled manifold, and once the dynamical equations and their solutions have
been investigated the second surplus copy of spacetime can quietly be forgotten.
In particular, all the calculations we have performed for the spherically symmetric
wormholes of this paper apply equally well to spherically symmetric holes in spacetime
(not black holes, actual voids in the manifold), with the edge of the hole being a brane
we deduce the existence of a large class of stable void solutions, and an equally large class
of unstable voids that either collapse to form black holes, or explode to engulf the entire
universe.
Equally well, the baby universes of the preceding section can, under this new physical
interpretation of the relevant mathematics, be used to investigate finite volume universes
with boundary. The bulk of the physical universe now lies in the range r (0, a), and the
edge of the universe is located at a. Again, we deduce the existence of a large class of
stable baby universes with boundary, and an equally large class of unstable baby universes
that either collapse to singularity, or explode to provide arbitrarily large universes. Note
that these particular exploding universes are not FLRW universes, and are not suitable
cosmologies for our own universe. Nevertheless, this notion of using a brane as an actual
physical boundary of spacetime is an issue of general applicability, and we hope to return
to this topic in future publications.

5. Discussion
The main point of this paper is that in the brane picture there is nothing wrong with
the notion of a negative brane tension, and that once branes of this type are allowed to
contribute to the stress-energy, the class of solutions is greatly enhanced, now including
many quite peculiar beasts not normally considered to be part of standard general relativity.
As specific examples, the energy condition violations caused by negative tension branes
allow one to construct classical traversable wormholes, at least some of which (as we
have seen) are actually dynamically stable. Now for spherically symmetric wormholes
of the type considered in this paper, attempting to cross from one universe to the other
requires the traveller to cross the brane, a process which is likely to prove disruptive of
the travellers internal structure, well being, and overall health. This problem, or rather
the no-brane analog of this problem, was already considered by Morris and Thorne in
their pioneering work on traversable wormholes [8]. A possible resolution comes from the
fact that spherical symmetry is a considerable idealization: One of the present authors has
demonstrated that if one uses negative tension cosmic strings instead of negative tension
domain walls, then it is possible to construct traversable wormhole spacetimes that do not
possess spherical symmetry, and contain perfectly reasonable paths from one asymptotic
region to the other that do not involve personal encounters with any form of exotic
matter [40]. (See also the extensive discussion in [9].) In a brane context this means we
should consider the possibility of a negative tension (1 + 1)-dimensional brane in (3 + 1)dimensional spacetime.

C. Barcel, M. Visser / Nuclear Physics B 584 (2000) 415435

433

Now the peculiarities attendant on widespread violations of the energy conditions are not
limited to violations of topological censorship; as we have seen there is also the possibility
of violating (evading) the singularity theorem. If this is not enough, then it should be
borne in mind that without some form of energy condition we do not have a positive mass
theorem. (Looking out into our own universe, we do have a positive mass observation, but
it would be nice to be able to deduce this from general principles.) A discussion of some
of the peculiarities attendant on negative asymptotic mass can be found in the early work
of Bondi [41], and a possible observational signal (particular types of caustics in the light
curves due to gravitational lensing) has been pointed out by Cramer et al. [42]. Finally,
energy condition violations are also the sine qua non for the Alcubierre warp drive [43].
In summary, all of these somewhat peculiar geometries, which were investigated within
the general relativity community more with a view to understanding the limitations of
general relativity (and more specifically, of semiclassical general relativity) than in the
expectation that they actually exist in reality, are now seen to automatically be part and
parcel of the brane models currently being considered as semi-phenomenological models
of empirical reality.

Acknowledgements
The research of CB was supported by the Spanish Ministry of Education and Culture
(MEC). MV was supported by the US Department of Energy. We wish to thank Harvey
Reall and Sumit Das for their comments and interest.

References
[1] L. Randall, R. Sundrum, A large mass hierarchy from a small extra dimension, Phys. Rev.
Lett. 83 (1999) 3370, hep-ph/9905221.
[2] L. Randall, R. Sundrum, An alternative to compactification, Phys. Rev. Lett. 83 (1999) 4690,
hep-th/9906064.
[3] E.E. Flanagan, R.M. Wald, Phys. Rev. D 54 (1996) 6233.
[4] C. Barcelo, M. Visser, Traversable wormholes from massless conformally coupled scalar fields,
Phys. Lett. B 466 (1999) 127, gr-qc/9908029.
[5] M. Visser, C. Barcelo, Energy conditions and their cosmological implications, gr-qc/0001099.
[6] C. Barcelo, M. Visser, Scalar fields, energy conditions, and traversable wormholes, grqc/0003025.
[7] L.A. Anchordoqui, S.E. Perez Bergliaffa, The world is not enough, gr-qc/0001019.
[8] M.S. Morris, K.S. Thorne, Wormholes in spacetime and their use for interstellar travel: A tool
for teaching general relativity, Am. J. Phys. 56 (1988) 395.
[9] M. Visser, Lorentzian Wormholes: From Einstein to Hawking, American Institute of Physics,
Woodbury, 1995.
[10] H. Kawai, D.C. Lewellen, S.H. Tye, Construction of four-dimensional fermionic string models,
Phys. Rev. Lett. 57 (1986) 1832.
[11] H. Kawai, D.C. Lewellen, S.H. Tye, Construction of fermionic string models in four dimensions,
Nucl. Phys. B 288 (1987) 1.

434

C. Barcel, M. Visser / Nuclear Physics B 584 (2000) 415435

[12] H. Kawai, D.C. Lewellen, S.H. Tye, Four-dimensional type II strings and their extensions: typeIII strings, Phys. Lett. B 191 (1987) 63.
[13] H. Kawai, D.C. Lewellen, S.H. Tye, Construction of four-dimensional fermionic string models
with a generalized supercurrent, Int. J. Mod. Phys. A 3 (1988) 279.
[14] S. Chaudhuri, H. Kawai, S.H. Tye, Chiral bosonic supercurrents and four-dimensional fermionic
string models, Nucl. Phys. B 322 (1989) 373.
[15] J. Polchinski, An Introduction to the Bosonic String, String Theory, Vol. 1, Cambridge
University Press, England, 1998.
[16] J. Polchinski, Superstring Theory and Beyond, String Theory, Vol. 2, Cambridge University
Press, England, 1998.
[17] C. Csaki, M. Graesser, L. Randall, J. Terning, Cosmology of brane models with radion
stabilization, hep-ph/9911406.
[18] C. Csaki, J. Erlich, T.J. Hollowood, J. Terning, Holographic RG and cosmology in theories with
quasi-localized gravity, hep-th/0003076.
[19] C. Csaki, J. Erlich, T.J. Hollowood, Y. Shirman, Universal aspects of gravity localized on thick
branes, hep-th/0001033.
[20] J. Cline, C. Grojean, G. Servant, Inflating intersecting branes and remarks on the hierarchy
problem, Phys. Lett. B 472 (2000) 302, hep-ph/9909496.
[21] C. Grojean, J. Cline, G. Servant, Supergravity inspired warped compactifications and effective
cosmological constants, hep-th/9910081.
[22] J.M. Cline, C. Grojean, G. Servant, Cosmological expansion in the presence of extra
dimensions, Phys. Rev. Lett. 83 (1999) 4245, hep-ph/9906523.
[23] J.M. Cline, Cosmological expansion in the RandallSundrum warped compactification, hepph/0001285, Cosmo99 Proceedings.
[24] I.I. Kogan, S. Mouslopoulos, A. Papazoglou, G.G. Ross, J. Santiago, A three three-brane
universe: new phenomenology for the new millennium?, hep-ph/9912552.
[25] T. Li, Classification of 5-dimensional spacetime with parallel 3-branes, hep-th/9912182.
[26] P. Kanti, I.I. Kogan, K.A. Olive, M. Pospelov, Single-brane cosmological solutions with a stable
compact extra dimension, hep-ph/9912266.
[27] M. Visser, Traversable wormholes from surgically modified Schwarzschild spacetimes, Nucl.
Phys. B 328 (1989) 203.
[28] N. Arkani-Hamed, S. Dimopoulos, G. Dvali, N. Kaloper, Infinitely large new dimensions, Phys.
Rev. Lett. 84 (2000) 586, hep-th/9907209.
[29] Z. Chacko, A.E. Nelson, A solution to the hierarchy problem with an infinitely large extra
dimension and moduli stabilization, hep-th/9912186.
[30] C. Grojean, T self-dual transverse space and gravity trapping, hep-th/0002130.
[31] D. Hochberg, M. Visser, The null energy condition in dynamic wormholes, Phys. Rev. Lett. 81
(1998) 746, gr-qc/9802048.
[32] W. Israel, Singular hypersurfaces and thin shells in general relativity, Nuovo Cimento B 44 (10)
(1966) 114; Nuovo Cimento B 48 (10) (1967) 463463 (Errata).
[33] K. Lanczos, Untersuching ber flchenhafte verteiliung der materie in der Einsteinschen
gravitationstheorie, 1922, unpublished.
[34] K. Lanczos, Flchenhafte verteiliung der materie in der Einsteinschen gravitationstheorie, Ann.
Phys. (Leipzig) 74 (1924) 518540.
[35] N. Sen, ber dei grenzbedingungen des schwerefeldes an unstetig keitsflchen, Ann. Phys.
(Leipzig) 73 (1924) 365396.
[36] A. Chamblin, M.J. Perry, H.S. Reall, Non-BPS D8-branes and dynamic domain walls in massive
IIA supergravities, JHEP 9909 (1999) 014, hep-th/9908047.
[37] S.W. Hawking, G.F.R. Ellis, The Large Scale Structure of Spacetime, Cambridge University
Press, England, 1973.

C. Barcel, M. Visser / Nuclear Physics B 584 (2000) 415435

435

[38] E. Poisson, Black-hole interiors and strong cosmic censorship, gr-qc/9709022; in: L.M. Burko,
A. Ori (Eds.), The Internal Structure of Black Holes and Spacetime Singularities, Institute of
Physics Press, Bristol, 1997.
[39] E.E. Flanagan, Quantum mechanical instabilities of Cauchy horizons in two dimensions: a
modified form of the blueshift instability mechanism, gr-qc/9711066; in: L.M. Burko, A. Ori
(Eds.), The Internal Structure of Black Holes and Spacetime Singularities, Institute of Physics
Press, Bristol, 1997.
[40] M. Visser, Traversable wormholes: some simple examples, Phys. Rev. D 39 (1989) 3182.
[41] H. Bondi, Negative mass in general relativity, Rev. Mod. Phys. 29 (1957) 423.
[42] J.G. Cramer, R.L. Forward, M.S. Morris, M. Visser, G. Benford, G.A. Landis, Natural
wormholes as gravitational lenses, Phys. Rev. D 51 (1995) 3117, astro-ph/9409051.
[43] M. Alcubierre, The warp drive: hyper-fast travel within general relativity, Class. Quantum
Grav. 11 (1994) L73.

Nuclear Physics B 584 (2000) 436458


www.elsevier.nl/locate/npe

p-Brane black holes as stability islands


K.A. Bronnikov , V.N. Melnikov 1
Centre for Gravitation and Fundamental Metrology, VNIIMS, 3-1 M. Ulyanovoy St., Moscow 117313, Russia
Institute of Gravitation and Cosmology, RUDN, 6 Miklukho-Maklaya St., Moscow 117198, Russia
Received 25 February 2000; accepted 18 April 2000

Abstract
In multidimensional gravity with an arbitrary number of internal Ricci-flat factor spaces,
interacting with electric and magnetic p-branes, spherically symmetric configurations are considered.
It is shown that all single-brane black hole solutions are stable under spherically symmetric
perturbations, whereas similar solutions possessing naked singularities turn out to be catastrophically
unstable. The black hole stability conclusion is extended to some classes of configurations with
intersecting branes. These results do not depend on the particular composition of the D-dimensional
spacetime, on the number of dilatonic scalar fields a and on the values of their coupling constants
sa . Some examples from 11-dimensional supergravity are considered. 2000 Elsevier Science B.V.
All rights reserved.
PACS: 04.20.Jb; 04.50.+h; 04.70.Bw
Keywords: Multidimensional gravity; Unified theories; Spherical symmetry; Stability

1. Introduction
In this paper we continue our studies of multidimensional gravitational models based on
D-dimensional Einstein equations with fields of antisymmetric forms of arbitrary rank
(see [13] and references therein) as some low-energy limit of a future unified model
(M-, F - or other type). Our main interest here will be in the stability properties of
multidimensional black-hole (BH) and non-BH solutions with nonzero fields of forms,
associated with charged p-branes. There exist a large number of such solutions in arbitrary
dimensions see, e.g., [2,49] and references therein. They are important in connection
with studies of processes at early stages of the Universe, counts of micro-states in BH
thermodynamics and now especially due to new developments in M-theory [10] related to
Corresponding author. K.A. Bronnikov, RGS, 3-1 M. Ulyanovoy St., Moscow 117313, Russia; E-mail:
kb@rgs.mccme.ru
1 melnikov@rgs.phys.msu.su

0550-3213/00/$ see front matter 2000 Elsevier Science B.V. All rights reserved.
PII: S 0 5 5 0 - 3 2 1 3 ( 0 0 ) 0 0 2 4 0 - 6

K.A. Bronnikov, V.N. Melnikov / Nuclear Physics B 584 (2000) 436458

437

the AdS/CFT correspondence [11]. For recent reviews of this rapidly developing field, see,
e.g., [12,13].
BH stability studies have a long history, of which we will only mention (more or less
arbitrarily) some milestones, concerning spherically symmetric backgrounds. Regge and
Wheeler [14] considered the stability of the Schwarzschild spacetime and developed
the formalism of spherical harmonics for metric perturbations. Vishveshwara [15] finally
proved the linear stability of Schwarzschild BHs; Moncrief [16] did the same for Reissner
Nordstrm ones. BHs with a conformally coupled scalar field were shown to be unstable
under spherically symmetric perturbations [17], as well as minimally coupled scalar field
configurations in general relativity possessing naked singularities [18]. The monopole
degree of freedom is present there due to the scalar field; it was argued that monopole
perturbations were most likely to be unstable due to the absence of centrifugal terms in the
effective potentials; catastrophic instabilities were indeed found and it was unnecessary to
study other multipoles. On the other hand, coloured BHs, containing non-Abelian gauge
fields, were shown to be, in general, unstable due to their sphaleronic degrees of freedom
see [19] and references therein. A recent overview of 4-dimensional perturbation studies
may be found in Ref. [20].
For BHs in multidimensional theories of gravity the situation is more complex since,
on the one hand, there emerge new effective scalar fields (extra-dimension scale factors,
sometimes called moduli fields) in the external spacetime, and, on the other, instabilities
may be caused by waves in extra dimensions. Instabilities of the latter kind were indeed
found by Gregory and Laflamme [21,22] for a limited class of neutral and charged black
strings and branes, having a constant internal space scale factor. Furthermore, it was
argued that compactification on a sufficiently small length scale should prevent the onset
of instability, and, moreover, that extremal black branes are stable [23]. It was concluded
that only very light BHs, whose horizon radii have the same order of magnitude as their
extra dimensions, manifest this form of instability.
It is therefore of interest to inquire whether or not there are other forms of instability,
maybe more dangerous, on more general backgrounds, containing nontrivial internal
space structures and/or several dilatonic scalars and brane charges. As was previously the
case with backgrounds containing effective scalar fields, it is natural to consider first the
simplest, monopole perturbations.
Earlier we analyzed the stability of static, spherically symmetric solutions to the
EinsteinMaxwell scalar equations with a dilatonic type coupling between scalar and
electromagnetic fields in D-dimensional gravity [24,25]. It was proved there that only BH
configurations were stable under linear spherically symmetric perturbations, while nonBH solutions turned out to be catastrophically unstable. A similar result was obtained
for dilatonic BHs with the inclusion of the GaussBonnet curvature term due to one-loop
quantum corrections [26]. We will now show that in the simplest case of a single charged
black brane the solution is stable under linear spherically symmetric perturbations, whereas
single-brane solutions with naked singularities are unstable. So the results of [24,25] are
generalized.

438

K.A. Bronnikov, V.N. Melnikov / Nuclear Physics B 584 (2000) 436458

We also present a tentative consideration of multi-brane BHs and conclude that in


cases when the perturbation equations decouple, the stability conclusion is also valid. Two
classes of such systems are indicated, both characterized by certain relations among brane
charges, such that, in terms of Section 3, the constituent vectors YE s form a single block of
a block-orthogonal system (BOS) single-block BHs for short. Namely, the stability is
proved for arbitrary two-brane single-block BHs and multi-brane single-block BHs with
mutually orthogonal vectors YE s (see the details in Section 6.2). For many single-block
configurations which do not belong to these classes, the stability can be proved as well,
but their properties require individual studies; see an example in Appendix A, Eqs. (A.4)
(A.7). There are, however, numerous multi-brane BHs for which decoupling is impossible
and one may expect that some of them show a new type of instability connected with mode
interaction; a study of these systems is in progress.
The paper is organized as follows. Section 2 describes the general features of the field
model to be considered. Section 3 presents some known static solutions, including BHs,
on the basis of the target space V connected with dimensional reduction. In Section 4 a
truncated target space V, more appropriate for treating the perturbations, is introduced,
and wave equations for perturbations are derived. In Section 5 the stability properties of
single-brane configurations are deduced, while in Section 6 the stability of some multibrane BHs under spherically symmetric perturbations is established. The appendix gives
some examples from 11-dimensional supergravity.
The word stable throughout the paper means stable under linear spherically
symmetric perturbations.

2. The model
Our starting point is, as in Refs. [18], the model action for D-dimensional gravity with
several scalar dilatonic fields a and antisymmetric ns -forms Fs :
(
)
Z
X s
p
D
MN
a
b
2sa a 2
e
Fs ,
(1)
S = d z |g| R[g] ab g M N
ns !
sS

in a pseudo-Riemannian manifold M = Ru M0 Mn , with factor space dimensions


di , i = 0, . . . , n; R is the scalar curvature. We will assume M to be spherically symmetric,
so that the metric is
0

2
= gMN dzM dzN = e2 du2 +
dsD

n
X

e2 dsi2

i=0
0

= e2 du2 + e2 d 2 e2 dt 2 +

n
X

e2 dsi2 .

(2)

i=2

Here u is a radial coordinate ranging in Ru R; ds02 = d 2 is the metric on a unit d0 dimensional sphere M0 = S d0 ; t M1 Rt is time; the metrics g i = dsi2 of the extra
factor spaces (i > 2) are assumed to be Ricci-flat and can have arbitrary signatures i =

K.A. Bronnikov, V.N. Melnikov / Nuclear Physics B 584 (2000) 436458

439

M ...M

sign g i ; |g| = |det gMN | and similarly for subspaces; Fs2 = Fs, M1 ...Mns Fs 1 ns ; sa are
coupling constants; s = 1 (to be specified later); s S, a A, where S and A are some
i
finite sets. The scale factors e and the scalars a are assumed to depend on u and t
only.
The F -forms should be also compatible with spherical symmetry. A given F -form
may have several essentially (non-permutatively) different components; such a situation
is sometimes called composite p-branes. 2 For convenience, we will nevertheless treat
essentially different components of the same F -form as individual (elementary) F forms. A reformulation to the composite ansatz, if needed, is straightforward.
Each ns -form F = dA [M1 AM2 ...Mns ] dzM1 dzMns is then associated with a certain
subset I = {i1 , . . . , ik } (i1 < < ik ) of the set of numbers labelling the factor spaces:
{i} = I0 = {0, . . . , n}. The forms Fs are naturally classified as electric (F eI ) and magnetic
(FmI ) ones. By definition, the potential AI of an electric form F eI carries the coordinate
indices of the subspaces Mi , i I , and is u-dependent (since only a radial component
of the field may be nonzero). A magnetic form FmI is built as a form dual to a possible
electric one, and its nonzero components carry coordinate indices of the subspaces Mi , i
def

I = I0 \ I , One can write:


nmI = rank FmI = D rank F eI = d( I ),
(3)
n eI = rank F eI = d(I ) + 1,
P
where d(I ) = iI di are the dimensions of the subspaces MI = Mi1 Mik . The
index s will be used to jointly describe the two types of forms, so that [2,4]
S = {s} = { eIs } {mIs }.

(4)

We will make some more assumptions to assure that all F -forms behave like genuine
electric or magnetic fields in the physical subspace Mphys = Ru Rt M0 , namely:
(i)

1 Is , s

(the subspaces MIs contain the time axis Rt );

(5)

(ii)

0
/ Is , s

(the branes only live in extra dimensions);

(6)

(iii)

Ttt (Fs ) > 0,

(the energy density is positive).

(7)

By (i), the so-called quasiscalar forms [2,4] (forms with 1


/ Is , behaving as effective
scalar or pseudoscalar fields in Mphys ) are excluded. The reason for adopting (i) is that our
interest here is mostly in BHs which do not admit nonzero quasiscalar forms (the no-hair
theorem for brane systems [3]).
Assumption (iii) holds if all extra dimensions are spacelike (i = 1, i > 2) and in (1) all
s = 1. In more general models, with arbitrary i , (iii) holds if
def Y
mIs = ( I s ),
(I ) =
i .
(8)
eIs = (Is ),
iI

We will consider static configurations and their small (linear) time-dependent perturbations. It turns out, however, that under the above assumptions the Maxwell-like field
2 There is an exception: two components, having only one non-coinciding index, cannot coexist since in this
N , while
case there emerge nonzero off-block-diagonal components of the energymomentum tensor (EMT) TM
the Einstein tensor in the l.h.s. of the Einstein equations is block-diagonal. See more details in Ref. [1].

440

K.A. Bronnikov, V.N. Melnikov / Nuclear Physics B 584 (2000) 436458

equations for Fs may be integrated in a general form for their arbitrary dependence on u
and t. Indeed, for an electric m-form Fs (s = eI , m = d(Is ) + 1) the field equations due
to (1)
 

p
u
a
(9)
Fsut M3 ...Mm |g| e2sa = 0
t
are easily integrated to give
p
0
a
|gI |
Fsut M3 ...Mm = Qs e 0 2sa M3 ...Md(I )
1 2
a
F = (I )Q2s e2 ( I )2sa ,
(10)

m! s
Q
where ... and ... are Levi-Civita symbols, |gI | = iI |g i |, and Qs = const are charges.
In a similar way, for a magnetic m-form Fs (s = mI , m = d( I s )), the field equations and
the Bianchi identities dFs = 0 lead to
p
1 2
a
Fs = ( I )Q2s e2 ( I )+2sa . (11)
Fs,M1 ...Md( I ) = Qs M1 ...Md( I ) |gI |
m!
We use the notations
n
X
X
dj j (u, t),
(I ) =
di i (u, t).
(12)
i =
j =i

iI

Evidently, the expressions (10) and (11) differ only in the signs before sa and the
signature-dependent prefactors . Due to (8), their energy-momentum tensors (EMTs)
coincide up to the replacement sa sa , and their further treatment is quite identical.
In what follows we therefore mostly speak of electric forms, but the results are easily
reformulated for any sets of electric and magnetic forms. We also assume that all Qs 6= 0.
3. Static systems
3.1. The target space V
Under the above assumptions, the system is well described using the so-called model
representation [1], to be briefly outlined here as applied to static, spherically symmetric
systems. This formulation can be derived by reducing the action (1) to the (d0 + 1)dimensional space Ru M0 .
As in [27] and many later papers, we choose the harmonic u coordinate ( M M u = 0),
such that
n
X
di i .
(13)
0 (u) = 0 (u)
i=0

Due to (6), the combination


integrated giving
e

0 0

1
2
1 + 2

= (d0 1)s(k, u),

of the Einstein equations has a Liouville form and is


1
k sinh ku, h > 0,
s(k, u) = u,
h = 0,
1
k sin ku, h < 0,
def

(14)

K.A. Bronnikov, V.N. Melnikov / Nuclear Physics B 584 (2000) 436458

441

where k is an integration constant. With (14) the D-dimensional line element may be
def

written in the form ( d = d0 1)




n
X
e21 /d
du2
1
i
2
=
+ d 2 e2 dt 2 +
e2 dsi2 .
dsD
2
[ ds(k, u)]2/d [ ds(k, u)]
i=2

(15)

The u coordinate is defined for 0 < u < umax where u = 0 corresponds to spatial infinity
while umax may be finite or infinite depending on the form of a particular solution.
The remaining set of unknowns i (u), a (u) (i = 1, . . . , n, a A) can be treated
as a real-valued vector function x A (u) (so that {A} = {1, . . . , n} A) in an (n + |A|)dimensional vector space V (target space). The field equations for x A can be derived from
the Toda-like Lagrangian
X
Q2s e2ys (u)
L = GAB xuA xuB +
s

n
X
i=1

di (ui )2 +

2
1,
u

d0 1

+ ab ua ub +

Q2s e2ys (u)

(16)

(the subscript u means d/du), with the energy constraint


X
d0
k 2 sign k.
Q2s e2ys =
E = GAB xuA xuB
d
0 1
s
The nondegenerate symmetric matrix


di dj /d + di ij
0
(GAB ) =
0
ab

(17)

(18)

defines a positive-definite metric in V; the functions ys (u) are scalar products:


ys = (Is ) sa a Ys,A x A ,

(Ys,A ) = (di iIs , sa ),

(19)

where iI = 1 if i I and iI = 0 otherwise. The contravariant components and scalar


products of the vectors YE s are found using the matrix GAB inverse to GAB :
 ij


/di 1/(D 2) 0
AB
=
,
G
0
ab



d(Is )
, sa ,
(20)
Ys A = iIs
D2
d(Is )d(Is 0 )
+ as as 0 .
(21)
Ys,A Ys 0 A YE s YE s 0 = d(Is Is 0 )
D2
The equations of motion in terms of YE s read
X
Q2s YsA e2ys (u) .
(22)
x A =
s

3.2. Exact solutions: orthogonal systems (OS)


The integrability of the Toda-like system (16) depends on the set of vectors YE s . In many
cases general or special solutions to Eqs. (22) are known. Here we will mention the simplest

442

K.A. Bronnikov, V.N. Melnikov / Nuclear Physics B 584 (2000) 436458

case of integrability: a general solution is available if all YE s are mutually orthogonal in


V [2], that is,


Ys2 = d(I ) 1 d(I )/(D 2) + 2s > 0,
(23)
YE s YE s 0 = ss 0 Ys2 ,
P
where 2s = a 2sa . Then the functions ys (u) obey the decoupled Liouville equations
ys,uu = Q2s Ys2 e2ys , whence
e2ys (u) = Q2s Ys2 s 2 (hs , u + us ),

(24)

where hs and us are integration constants and the function s(. , .) has been defined in (14).
For the sought functions x A (u) and the conserved energy E we then obtain:
x A (u) =
E=

X
s

X Ys A
Ys2

ys (u) + cA u + cA ,

s
2
hs sign hs
Ys2

+ cE 2 =

d0 2
k sign k,
d0 1

(25)
(26)

where the vectors of integration constants cE and cE are orthogonal to all YE s : cA Ys,A =
cA Ys,A = 0, or
ci di iIs ca sa = 0,

ci di iIs ca sa = 0.

(27)

3.3. Exact solutions: block-orthogonal systems (BOS)


The above OS solutions are general for input parameters (D, di , YE s ) satisfying Eq. (23):
there is an independent charge attached to each (elementary) F -form. One can, however,
obtain special solutions for more general sets of input parameters, under less restrictive
conditions than (23). Namely, assuming that some of the functions ys (u) (19) coincide,
one obtains the so-called BOS solutions [4], where the number of independent charges
coincides with the number of different functions ys (u).
Indeed, suppose [4] that the set S splits into several non-intersecting non-empty subsets,
[
S ,
|S | = m(),
(28)
S=

such that the vectors YE () (() S ) form mutually orthogonal subspaces V V:


YE () YE (0 ) = 0,

6= 0 .

(29)

Then the corresponding result from [4] can be formulated as follows:


Proposition 1. Let, for each fixed , all YE V be linearly independent, and let there be
P
a vector YE = S a YE with a > 0 such that
def
YE YE = Y2 = YE 2 ,

S .

Then one has the following solution to the equations of motion (22), (17):

(30)

K.A. Bronnikov, V.N. Melnikov / Nuclear Physics B 584 (2000) 436458

xA =

X Y A

Y2

y (u) + cA u + cA ,

e2y = q Y2 s 2 (h , u + u ),

443

(31)
def

q =

Q2 ,

(32)

E=

X h2 sign h
d0

+ cE 2 =
k 2 sign k,
2
d

1
Y
0

(33)

where h , u , cA and cA are integration constants; cA and cA are constrained by the


orthogonality relations (27) (so that the vectors cE and cE are orthogonal to each individual
vector YE s V).
Eqs. (30) form a set of linear algebraic equations with respect to the charge factors
P
a = Q2 /q , satisfying the condition S a = 1. A solution to (30) for given YE
can contain some a < 0; according to [4], this would mean that such a p-brane is
quasiscalar, violating the assumption (5). Solutions with such branes are possible but
are rejected here since they do not lead to black holes. Furthermore, if a solution to (30)
gives a = 0 for some S , this means that the block cannot contain such a p-brane,
and then the consideration may be repeated without it. 3
The function y (u) is equal to y() (u) = Y(),A x A , which is, due to (30), the same
for all S . The BOS solution generalizes the OS one, (24), (25): the latter is restored
when each block contains a single F -form.
Both kinds of solutions are asymptotically flat, and it is natural to normalize the
functions ys (u) and y (u) by the condition ys (0) = 0 or y (0) = 0, so that the constants
us and u are directly related to the charges.
Other solutions to the equations of motion are known, connected with Toda chains and
Lie algebras [57,9].
3.4. Black-hole solutions
Black holes (BHs) are distinguished among other spherically symmetric solutions by
the existence of horizons instead of singularities in the physical d0 + 2 dimension space
Mphys ; the extra dimensions and scalar fields are also required to be well-behaved on the
horizon to provide regularity of the D-dimensional metric. Thus BHs are described by the
above solutions under certain constraints upon the integration constants. Namely, for the

i
1
BOS solution (30)(33), requiring that all the scale factors e (except e = |gt t | which
should tend to zero) and scalars a tend to finite limits as u umax , we get [2]:
X
cA = k
Y2 Y A k1A ,
(34)
h = k > 0, ,
s
3 Geometrically, the vector YE solving Eqs. (30) is the altitude of the pyramid formed by the vectors YE ,

S with a common origin. The condition a > 0 means that this altitude is located inside the pyramid, while
a = 0 means that the altitude belongs to one of its faces.

444

K.A. Bronnikov, V.N. Melnikov / Nuclear Physics B 584 (2000) 436458

where A = 1 corresponds to i = 1 (time). The constraint (26) then holds automatically. The
value u = umax = corresponds to the horizon. The same condition for the OS solution
(24)(27) is obtained by replacing 7 s.
Under the asymptotic conditions a 0, i 0 as u 0, after the transformation
e2ku = 1

2k
dr d

def

d = d0 1

(35)

the metric (15) for BHs and the corresponding scalar fields may be written as
2
dsD

Y

"
HA

def

A =



2k Y 2/Y2
H
dt 1
dr d
#

 X
n
Y Ai
dr 2
2
2
2

+ r d +
+
dsi
H ,
1 2k/(dr d )

i=2
2

2 X a d(I ) OS 2 d(Is )
,
= 2
D2
Y2
Ys D 2
S

def

Ai =

2 X
2
OS
a iI = 2 iIs ,
Y2
Ys

(36)

a =

X 1
X sa
X
OS
ln H
a a =
ln Hs ,
2
Y
Ys2

(37)

OS

where = means equal for OS, with 7 s, and H are harmonic functions in R+ S d0 :
H (r) = 1 + p /(dr d ),

def

p =

k 2 + q Y2 k.

(38)

The subfamily (34), (36)(38) exhausts all BOS BH solutions; the OS ones are obtained
in the special case of each block S consisting of a single element s.
The above relations describe the so-called non-extremal BHs. Extremal ones, corresponding to minimum black hole mass for given charges (the so-called BPS limit), are
obtained in the limit k 0. The same solutions follow directly from (31)(33) under the
conditions h = k = cA = 0. For k = 0, the solution is defined in the whole range r > 0,
while r = 0 in many cases corresponds to a naked singularity rather than an event horizon, so that we no more deal with a black hole. However, in many other important cases
r = 0 is an event horizon of extremal ReissnerNordstrm type, with an AdS near-horizon
geometry; some examples are mentioned in the appendix.
Other families of solutions, mentioned at the end of the previous section, also contain
BH subfamilies. The most general BH solutions are considered in Ref. [9].

K.A. Bronnikov, V.N. Melnikov / Nuclear Physics B 584 (2000) 436458

445

4. Perturbation equations
4.1. Truncated target space V
Consider now nonstatic spherically symmetric configurations corresponding to the
action (1) with the metric (2) and all field variables depending on u and t. As before,
we are dealing with true electric and magnetic forms Fs , so that their Is 3 1, or
Is = 1 Js ,

Js {2, . . . , n}.

(39)

As in Refs. [24,25], it is helpful to pass to the Einstein frame in the physical (d0 + 2)dimensional spacetime Mphys = Ru M0 M1 . The action (1) is then rewritten in terms
of the metric g , the d0 + 2-dimensional part of gMN , and is transformed to the Einstein
frame in Mphys with the metric
g = e22 /d0 g .

(40)

The electric ns -forms are re-parametrized as follows:


Fut M3 ...Mns = F ut ,
s

1 2 1
a
F = F F e2 (Js ) = Q2s e2sa ,
ns ! s
2s s

(41)

where the indices M3 , . . . , Mns belong to Js ; here and henceforth the indices , are
raised and lowered using the metric g ; in the last equality the solution (10) and the
positive energy assumption (7) are taken into account.
The action (1) is written in terms of g and F as follows (up to a constant prefactor,
s
connected with the volume of extra dimensions, and a subtracted total divergence):

Z
n
X
p
 
2

1
dd0 +2 z |g| R g (2 )2
di i ab a , b
S=
d0
i=2

Mphys

1X
a
F F e22 /d0 2 (Js )+2sa

s
s
2

Z
=
Mphys

sS



 
 1X
2Zs,K x K
d0 +2
K
L
d
z |g| R g HKL x , x
F F e
(42)
s
s
2
p

sS

where (f, g) = g f g, (f )2 = (f, f ); the non-degenerate symmetric matrix




di dj /d0 + di ij
0
(43)
(HKL ) =
0
ab
defines a positive-definite metric in the vector space V (truncated target space) parametrized by the variables (x K ) = ( 2 , . . . , n ; a ); the constant vectors Z s V are characterized by the components 4
4 We will use the indices K, L for quantities specified in V to distinguish them from those in V where the
indices A, B are used; vectors in V are marked with overbars, those in V by arrows. Scalar products are written
as YE ZE = GAB Y A Z B (as before) and Y Z = HKL Y K Z L .

446

K.A. Bronnikov, V.N. Melnikov / Nuclear Physics B 584 (2000) 436458



di
(Zs,K ) = di iJs , sa ,
d0




d(Is )
K
KL
, sa ,
Zs = H Zs,L = iJs
D2
where the matrix (H KL ) is inverse to (HKL ):
 ij

/di 1/(D 2) 0
KL
.
(H ) =
0
ab

(44)

(45)

The truncated target space V may be considered as the hyperplane x 1 = 2 /d0 in V, with
the metric HKL induced by GAB (18). The components H KL turn out to be the same as
GAB for i 6= 1; the components Ys A and Zs K coincide in the same manner. It is easy to
find that for vectors whose Is satisfy (39) (which is always the case in the present paper),
d0 1
,
YE s YE s 0 = Z s Z s 0 +
d0

(46)

whence it follows that, first, when different YE s are mutually orthogonal in V, the
corresponding Z s are never mutually orthogonal in V; second, for any YE s whose Is 3 1
one has
d0 1
.
(47)
YE s YE s Ys2 >
d0
4.2. Wave equations
The action (42) may be used to obtain the equations governing small spherically
symmetric perturbations of static solutions. The metric (40) in Mphys will be written in
the form
dsE2 = g dz dz = e2 du2 + e2 d 2 e2 dt 2 ,

(48)

where E stands for the Einstein frame and (u, t), (u, t), (u, t) are connected with
the corresponding quantities from (2) as follows:
= 0 + 2 /d0 ,

= 0 + 2 /d0 ,

= 1 + 2 /d0 .

(49)

Since the field equations for the F -forms have been integrated see (10) and (41), the
remaining unknowns are , , and x K (that is, i , i > 2, and a ). In what follows we
will write
(u, t) = (u) + (u, t),
where is a small perturbation, and similarly for other unknowns. We accordingly
preserve only terms linear in and similar quantities and in time derivatives. The field
equations may be written in the form
X
 
Zs K Q2s e2d0 +2Zs x ,
(50)
 g xK =
s

1
def
R = T T = Te ,
d0

(51)

K.A. Bronnikov, V.N. Melnikov / Nuclear Physics B 584 (2000) 436458

447

 
where  g = g is the DAlembert operator, while for the nonzero components
of the EMT T corresponding to (42) one has (no summing in )
2
HKL xuK xuL diag(0, 1, 0, . . . , 0)
Te
= e


X
1
1
1
1
2 2+2 +2Zs x
Qs e
diag 1 , , , . . . ,
,

d0
d0
d0
d0
s

Teut = HKL xuK xtL ,

(52)

where xu = u x and xt = t x; the first and second places under the symbol diag belong
to t and u, respectively.
As in our previous papers on stability, we use the coordinate freedom in the perturbed
spacetime and put
0

(53)

but preserve the harmonic u coordinate condition in the unperturbed (static) spacetime. 5
Then Eqs. (50) and (51) give
X
K
b
= 2
Q2s e2 +2Zs x Zs K Zs,Lx L ,
(54)
L x K + xuK (u u ) 2xuu
s

d0 t = x u x t ,

(55)

d0 u (u u ) = 2x uu x 2uu ,

(56)

where
def
b
L = e2d0 t t + uu ;

(57)

(55) follows from the (ut ) component of (51) and (56) from one of the angular components
of (51); we have also used the equations valid for static systems, in particular, (22), where,
according to the definitions of YE and Z, ys (u) = YE s xE = + Z s x. Integrating (55) in t and
omitting the emerging arbitrary function of u (since we neglect static perturbations), we
obtain
d0 = x u x.

(58)

Substituting from (58) and u u from (56) into (54), we finally arrive at the set of
wave equations for the dynamical degrees of freedom in our system, represented by x K :

 X
1 d xuK xL,u
K
K
L
K
b
P L=
Q2s e2ys (u) ZsK Zs,L. (59)
+
Lx = 2P L x ,
d0 du
u
s
The stability problem is now reduced to a boundary-value problem for x K (u, t).
Namely, if there exists a nontrivial solution to Eqs. (59) satisfying some physically
reasonable conditions at the ends of the range of u, such that |x K | (at least some of them)
grow unboundedly with t, then the static system is unstable. Otherwise it is stable in the
linear approximation.
5 This coordinate is harmonic for both metrics (2) and (48); in the latter the coordinate condition has the form
= d0 + .

448

K.A. Bronnikov, V.N. Melnikov / Nuclear Physics B 584 (2000) 436458

The condition at infinity, u = 0, is evident: the perturbations must vanish,


dx K 0 as u 0.

(60)

It is less evident at u = umax since some of the background static solutions are singular
there. As in Refs. [18,25] and others, dealing with minimally coupled or dilatonic scalar
fields, we will use the minimal requirement providing the validity of the perturbation
scheme, namely
K K
x /x < .
(61)
When the background is regular, this condition requires that the perturbation should be
regular as well.

5. Stability properties of single-brane solutions


5.1. Decoupling cases
Eqs. (59) in general do not decouple. Even in the simplest case when there is only one
antisymmetric form F (that is, one p-brane), so that Eqs. (24)(27) yield the general static
solution to the field equations, Eqs. (59) contain various linear combinations of x K with
u-dependent coefficients.
There is, however, an important case when Eqs. (59) do decouple for any configuration
of M with the metric (2), namely, the single-brane solution (24)(27) under the condition
that the vector c = (cK ) is parallel to Z in V: 6
cK = BY K /Y 2 ,

B = const

(62)

(here and henceforth in this section we omit the index s since, by our assumption, it takes
only one value). This condition is automatically valid for the case of utmost interest, BHs
with one p-brane (a black p-brane), which, by (34), corresponds to B = k = h > 0.
Due to the collinearity condition (62) and the constraint (26), the constants are now
connected by the relation

(63)
N 0 h2 sign h B 2 = k 2 sign k B 2
with N 0 = Y 2 (d0 1)/d0 < 1. It turns out that, besides BHs, the condition (62) is satisfied
for some singular solutions whose behaviour is quite generic for the system under study:
1. k > 0, h > 0, such that umax = and a singularity at the centre of symmetry is
attractive at least in terms of the metric (48), e 0;
2. h < 0, so that the solution behaviour is determined by the function s(h, u + u1 ) =
h1 sin h(u + u1 ) in (24) where u1 = const (0, /|h|). In this case the central
singularity is repulsive, e , of ReissnerNordstrm type.
Due to (62), Eqs. (59) take the form
6 Curiously, they are parallel in V although the corresponding vectors cE and YE are mutually orthogonal in the
surrounding target space V. For a clear picture, imagine two vectors in 3-dimensional space whose projections
onto a plane lie on the same ray.

K.A. Bronnikov, V.N. Melnikov / Nuclear Physics B 584 (2000) 436458

  2

1 fu
b
fuu (Zx),
Lx K = 2Z K
d 0 u u

def

f (u) =

y + Bu
Y2

with y(u) determined by (24); the area function has the form




1
1
ln (d0 1)s(k, u) + 0 f (u) Bu + const ,
=
d0 1
N

449

(64)

(65)

where the value of the constant is inessential.


Since V is an l-dimensional Euclidean space (l = n 1 + |A|), there are l 1 linearly
independent vectors Z such that Z Z = 0. Therefore the set of wave equations (64)
decouples into one equation for Zx and l 1 equations for different Z x:
  2

1 fu
b
U (u) = 2Z 2
fuu ;
(66)
L (Zx) = U (u)(Zx),
d 0 u u
b
(67)
L (Z x) = 0.
The static nature of the background solution makes it possible to separate the variables:
0

Z x = 0 (u) e t ,

Zx = (u) et ,

so that Eqs. (66) and (67) lead to




uu = e2d0 2 + U (u) ,
0
uu

=e

2d0

02

(68)

(69)

(70)

The existence of an admissible solution of any of these equations with a real value of
or 0 would mean that the perturbation can grow exponentially with time, hence the
instability.
It is hard to solve Eqs. (69), (70) in their full range but it is rather easy to assess the
asymptotic behaviour of their solutions near u = 0 and umax , and this will be sufficient for
making stability conclusions.
In particular, for u 0, which corresponds to spatial infinity, one has
U (u) 0 and

ed0 c0 ud0 /d ,

c0 = d d0 /d .

(71)

The general asymptotic form of solutions to (69) and (70) at small u for all cases under
study may be written as follows:



(72)
(or 0 ) = ud0 /(2d) c1 exp c0 d u1/d + c2 exp c0 d u1/d ,
with c1 , c2 = const. The boundary condition (60) then requires that in (72) c1 = 0, and it
remains to look at the other end of the u range, u umax .
5.2. Instability of naked singularities
Consider Case 1 of the previous subsection, a scalar type singularity. As u , the
relevant functions of the static solution behave as follows:
u N1 1
(k B)(h B),
(73)
y = hu + O(1),
2
Y d0 h + k

450

K.A. Bronnikov, V.N. Melnikov / Nuclear Physics B 584 (2000) 436458

where N1 = d0 Y 2 /d > 1. Since, due to (63), in the present case


|B| > k |B| > h and |B| < k |B| < h,
one sees that e(u) 0 exponentially. The same happens to U (u), therefore the asymptotic
0 = 0, so that
form of (69) or (70) is simply uu = 0 or uu
(or 0 ) = c3 u + c4

(74)
xK

also behave as
with constants c3 and c4 . On the other hand, the background functions
const u as u , therefore the second boundary condition (61) is satisfied for any
solution (74), including the one joining the solution (72) with c1 = 0 at small u. We
conclude that there are growing modes of perturbations for any , hence the singular
solution is catastrophically unstable.
In Case 2, h < 0, we have umax = /|h| u1 < and the relevant functions in the
static solution approach umax in the following way:

y(u) ln |h|1u ,

1
ln 1u ,
d0 Y 2

xK


ZK
ln |h|1u ,
2
Y
(75)

where 1u = umax u. One can make sure that U (u) does not affect the asymptotic
behaviour of solutions to Eq. (69) as u umax as compared with that of Eq. (70), and
for both one can write:
(or 0 ) = c5 + c6 1u,

(76)

while the condition (61) only requires |x K / ln 1u| < . Thus the solution satisfies (61)
for any choice of the constants c5 , c6 , and, as in Case 1, this leads to the instability of the
background singular solution.
5.3. Stability of black holes
In the BH case it is again hard to solve Eqs. (69), (70), but for our purpose it is sufficient
to note that, to realize an instability, a solution should begin with a zero value at u = 0 and
tend to a finite limit as u . This is evidently impossible for a solution to (70) since
0 / 0 > 0. We conclude that at least the 0 modes of BH perturbations are stable. The
uu
same reasoning works for the mode provided U (u) > 0 for all u > 0. Let us pass to
the variable R = r d /d in the expression for U in (66), so that 0 < u < corresponds to
> R > 2k:



0
0
N 0 R2
2Z 2 pR(R 2k)
0 4kR + 2k(p + p ) + pp
2k
+
p
1

+
pN
,
U=
Y 2 (R + p)2
(R + p0 )2
(R + p0 )2
def 1 d0 1
def
< 1,
p0 = p(1 N 0 ),
(77)
N0 = 2
d0
Y
where we have used the explicit form of the single-brane BH solution (24)(34) and the
substitution (35) with R = r d /d, replacing Ys Y , ps p, us u1 . Note that N 0 < 1

K.A. Bronnikov, V.N. Melnikov / Nuclear Physics B 584 (2000) 436458

451

due to (47), so that, in particular, p0 > 0. The expression (77) is manifestly positive for >
R > 2k, therefore the mode also does not lead to an instability. Thus linear stability of all
single-brane BH solutions under spherically symmetric perturbations has been established.
Our consideration did not apply to extremal BHs since in this case the behaviour of the
background functions x K (u) is generically singular as u (R = 1/u 0):

X ZK 
u
s
K
ln 1 +
,
(78)
x =
Y2
us
s
and so there is no reason to require |x K | < . In some cases it is regular (see Section 3.2
and examples in Appendix A). One can see, however, that again, as u , Eqs. (69) and
(70) for a single-brane extremal BH take the form uu = 0; the linearly growing solution
is discarded since it grows faster than x K in (78), so we are left with a constant and have
to require | < | for both regular and singular backgrounds. Then the same reasoning
with uu / > 0 makes us conclude that such allowed solutions with > 0 do not exist
and extremal BHs are stable as well. Indeed, an explicit form of U (u) is

2
00 2 
2Z 2
0 u1 + N u
1+N
,
(79)
U (u) = 2
Y (u + u1 )2
(u1 + N 00 u)2
where N 0 < 1 was defined in (77) and N 00 = 1 N 0 . The reasoning works since U > 0 for
all u > 0 and U 0 as u .
We can now formulate the following result, to be used in the further consideration:
Proposition 2. If a decoupled linear perturbation mode of a static, spherically
symmetric BH solution obeys the equation
b
L = U (u)

(80)

with U (u) > 0 (including the case U 0), this mode is stable.

6. Some black holes with multiple branes


6.1. Two-brane black holes
We have seen that one-brane singular background solutions are catastrophically
unstable; we would not like to treat more complex singular solutions since there is no
reason to believe that interaction of modes can prevent the instability. We instead consider
some multi-brane BH solutions for which the perturbation equations decouple and show
that they are stable.
Suppose there is a BH background solution (34)(38) with two branes, so that s takes
two values, s = 1, 2. The solution is characterized by two charges Qs , two vectors YE s V
and their counterparts Z s V, which we assume to be non-collinear (if they are collinear,
the consideration simplifies and the result is the same as for a single brane).

452

K.A. Bronnikov, V.N. Melnikov / Nuclear Physics B 584 (2000) 436458

The matrix P K L in the perturbation equations (59) may be written in the form
X
X
Q2s e2ys (u) ZsK Zs,L +
ZsK Zs 0 ,L fss 0 (u),
PKL =
s

with

(81)

ss 0



1
(ys,u + k)(ys 0 ,u + k)
.
f (u) =
d 0 u
Ys2 Ys20
u

(82)

ss 0

Just as in the one-brane case, one easily separates the transversal degrees of freedom:
for vectors Z V such that Z Z s = 0 (they fill a (dim V 2)-dimensional plane), the
function = Z x obeys the wave equation (80) with U 0.
However, Eqs. (59) with the matrix (81) in general do not decouple. An exception is the
special case when the two functions ys coincide,


(83)
y1 = y2 = y(u) = k sinh(ku1 )/ sinh k(u + u1 )
although the vectors YE s are different; in the BOS terminology (Section 3.3) the two vectors
YE s form a block and in our case this single block exhausts the whole system. If we suppose
for simplicity that the norms of YE s coincide, Y12 = Y22 = Y 2 , then the charges coincide as
well, Q21 = Q22 = Q2 , and one obtains for the two modes = ( Z 1 Z 2 )x:


b
(84)
L + = 2Z 2 (1 + cos ) Q2 e2y + 2F + ,
2
2
2y
b = 2Z Q e (1 cos ) ,
(85)
L
where Z 2 = Y 2 (d0 1)/d0 (see (46), is the angle between the vectors Z 1 and Z 2 in
V and F is the function (82) with Y12 = Y22 and y1 = y2 = y.
A direct substitution of y and u into (84) shows, as before, that the coefficient by
x+ at the r.h.s. is non-negative; for x in (85) this is manifestly so. Hence the previous
reasoning works and we conclude that such BHs (including extremal ones) with two branes
are stable.
The case Y12 6= Y22 is covered in the next section.
6.2. Single-block black holes
A natural question arises, whether or not the stability conclusion of the previous section
extends to an arbitrary multi-brane BH described by a single function y(u), in other words,
to any single-block BH. Note that any set of linearly independent vectors YE s may be treated
as a BOS-block, hence a special static solution of this kind (and hence a BH solution)
may always be obtained; the only restriction is a > 0 for the charge factors obeying the
consistency conditions (30).
Consider such a system: let there be a BOS BH solution with m linearly independent
vectors YE s V, s S = S , and the charge factors as satisfy (30). The following relations
are valid:
X
X
as YE s ,
YE s YE = Y2 s,
as = 1.
(86)
YE =
s

K.A. Bronnikov, V.N. Melnikov / Nuclear Physics B 584 (2000) 436458

453

It is easy to see that, due to (46), similar relations hold for the corresponding vectors Z s
V:
X
Z =
as Z s s,
Z s Z = Z 2 .
(87)
s

For certainty we suppose that Z s are linearly independent; if they are not, the consideration
is slightly modified without changing the results.
The wave equations (59) take the form
X
1b K
L x = q e2y(u)
as Zs K ZsL x L + F (u)Z K Z L x L ,
2
s


1 (yu + k)2
,
(88)
F (u) = 4
Y
d 0 u
u
P
E x
where q = q = s Q2s and y(u) is given by (83). As before, the perturbations Z
E s , whose dimension is dim V m >
E belong to the plane V orthogonal to all Z
(where Z
0) are decoupled and obey Eq. (67), giving no unstable modes.
Multiplying (88) by Z , one obtains a decoupled equation for = Z x:

b
U (u) = q e2y + F Z2 ,
(89)
L = 2U (u) ,
where Z2 = Z 2 . Since, as is directly verified, U (u) > 0, this mode is also stable.
The remaining (m 1) degrees of freedom may be described in terms of the vectors
W s = Z s Z and the functions s = W s x, such that
X
X
W s Z = 0,
as W s = 0,
as s = 0.
(90)
s

Using (87) and (90), one obtains the following m equations, coupled due to (90), for
(m 1) independent variables:
b
L s = 2q e2y

m
X

Kss 0 s 0 ,

Kss 0 = as 0 W s W s 0 .

(91)

s 0 =1

Excluding one of the unknowns, say, m , by virtue of (90), we arrive at a determined set of
wave equations for s = s m , s = 1, . . . , m 1:
!
m1
m1
X
X
1
2y
0
b
Fss 0 s ,
Fss 0 = as 0 W s 0 W s +
as 00 W s 00 .
(92)
L s = 2q e
am 00
0
s =1

s =1

This is a good way of studying specific models. In the general case, however, the situation
looks more transparent if we consider, instead, an auxiliary system with m independent
unknowns, described by Eqs. (91) where W m is slightly shifted from its true value by
some 1W , so that all W s become linearly independent; the relation among s in (90) is
then cancelled as well. Our system is restored when 1W 0.
For the auxiliary system the matrix (W s W s 0 ) is symmetric and positive-definite; if all
as are equal, the same is true for the matrix of coefficients in (91), (Kss 0 ), hence there
is a similarity transformation bringing it to a diagonal form with its positive eigenvalues

454

K.A. Bronnikov, V.N. Melnikov / Nuclear Physics B 584 (2000) 436458

along the diagonal. Such a transformation applied to Eqs. (91) decouples them into m
separate wave equations like (89), with some positive function replacing U (u). In the limit
1W 0, the worst thing that can happen is that some of the eigenvalues tend to zero,
L = 0 which, as we know, does not lead
giving for some combinations of s the equation b
to an instability. One can assert by continuity that this picture is generic, at least for as
close enough to being equal, and stability is again concluded according to Proposition 2.
On the contrary, when the numbers am > 0 are different, one cannot guarantee that the
non-symmetric matrix Kss 0 is similar to a diagonal one [28]. A failure in its diagonalization
can be connected with the occurrence of a pair (or pairs) of complex roots s of the
characteristic equation det |Kss 0 ss 0 | = 0. In this case there is at least one pair of
coupled perturbations for which a special investigation is necessary. An inspection of
the characteristic equation shows that the matrix Kss 0 cannot have negative eigenvalues,
therefore a separate unstable mode cannot occur and the only possible instability can be
connected with coupling between modes.
In particular, in an arbitrary OS BH solution there is a subfamily where all ys (u) coincide
(i.e., the constants us are the same for all s), so that the branes form a BOS block, and it
turns out that all as are also equal, as well as the squared charges Q2s . The above reasoning
shows that such solutions are stable.
If rank(Kss 0 ) < m 1, that is, there are additional linear dependences among Z s , then
L = 0, and for the
some combinations of s decouple leading to equations of the form b
remaining modes the above discussion can be repeated with slight modifications.
This is what can be said about the general case of single-block BOS BH solutions. If
there is a block of only two branes (m = 2), one can make a common stability conclusion
generalizing the one made in Section 6.1. Indeed, for = a1 1 + a2 2 there is Eq. (89),
whereas for = 1 2 one obtains
b
L = 2q e2y a1 (1 a1 )(Z 1 Z 2 )2 .

(93)

In the special case Z12 = Z22 = Z 2 one recovers (84), (85).


For m > 3 one has to study specific models individually.

7. Concluding remarks
We have shown that all static single-brane BH solutions with the metric (2) are stable
under linear spherically symmetric perturbations, whereas non-BH solutions possessing
naked singularities of different types are unstable. Very probably other singular solutions,
for which perturbation equations do not decouple, are unstable as well, since, as known
from vibration theory, coupling between modes can hardly stabilize them. On the contrary,
coupled modes can be unstable even when single ones are stable. It is therefore of interest
to study the stability properties of more complex BH solutions; this work is in progress.
We have also shown that the BH stability conclusion can be extended to some BHs
with multiple intersecting p-branes, namely, for the BOS case, characterized by a single
function y(u) (i.e., all ys coincide). It turns out that for such backgrounds the wave

K.A. Bronnikov, V.N. Melnikov / Nuclear Physics B 584 (2000) 436458

455

equations for perturbations also generically decouple and the absence of unstable modes
can be proved. Though, such a general proof is available only in two cases: (i) twobrane BH solutions (m = 2) and (ii) equal-charge subfamilies of arbitrary OS solutions.
Nontrivial brane systems with m > 2 should be studied individually to see whether or not
the corresponding matrix (Kss 0 ) in (91) can be diagonalized over the field of real numbers.
If yes, the solution is stable, otherwise a further study of coupled modes is necessary.
It should be stressed that a BOS-block solution exists for an arbitrary set of linearly
independent vectors YE s . In particular, if any multi-brane static BH solution for a certain set
of input parameters with independent vectors YE s is known, e.g., any OS or BOS solution
(see Section 3), then the additional requirement that all the functions ys coincide selects
from it a special BOS-block solution, for which a stability study can be performed as
described above. The only restriction is the requirement a () > 0 for the charge factors
obeying Eqs. (30).
Some technical points are worth mentioning. First, in gravitational stability studies it is
sometimes rather hard to separate real physical perturbations from purely gauge degrees
of freedom. We avoid this problem by obtaining the set of wave equations (59) where the
number of equations is precisely the number of dynamical degrees of freedom, represented
by scalars in the physical spacetime Mphys . Due to the latter circumstance, one more
complication is avoided: when dealing with vector and tensor perturbations of BHs, one
has to take into account the apparent singularity of the metric on the horizon; to properly
formulate the boundary conditions, it is then necessary to pass to Kruskal-like coordinates;
to be admissible, the perturbations are required to be finite on the future horizon [15,21,22].
In our case the perturbations are scalars, so the finiteness requirement can be imposed in
any coordinates. The choice of gauge only remains important for making the treatment
more transparent.
To conclude, we would like to emphasize that our consideration did not depend on the
number and dimensions of the factor spaces in the original spacetime M, on the number
of scalar fields a and on the particular values of their coupling constants sa .

Acknowledgement
V.M. is grateful to Department of Mathematics, University of the Aegean, Greece,
and to Department of Physics, Nara University, Japan, for their hospitality during his
stay there in October and NovemberDecember 1999, respectively. K.B. acknowledges
the hospitality of the colleagues of DFis-UFES, Vitria, ES, Brazil during his stay there
in NovemberDecember 1999. The work was supported in part by the Russian Basic
Research Foundation and by the Russian Ministry of Science and Technologies.

Appendix A
Consider, for illustration, some solutions of D = 11 supergravity, representing the lowenergy limit of M-theory, as examples of systems to which our stability results apply.

456

K.A. Bronnikov, V.N. Melnikov / Nuclear Physics B 584 (2000) 436458

The action (1) for this theory does not contain scalar fields ( a = sa = 0) and the
only F -form is of rank 4, whose various nontrivial components Fs (elementary F -forms
according to Section 2, to be called simply F -forms) are associated with electric 2-branes
[for which d(Is ) = 3] and magnetic 5-branes [such that d(Is ) = 6] (see [12] and references
therein). The orthogonality conditions (23) are satisfied if the following intersection rules
hold:
3 3 = 1,

3 6 = 2,

6 6 = 4.

(A.1)

(the notations are evident); for all F -forms Ys2 = 2.


We will designate the branes by figures labelling their world volume coordinates
(covered by Is ), beginning with 1 which corresponds to the time axis. Thus, e.g., (123)
is an electric 2-brane whose world volume includes the time axis M1 = Rt and two extra
dimensions. The number of dimensions where branes can be located is D 1 d0 =
10 d0 .
1. Single-brane BH solutions are described by (36), (38) where all sums and products
in s consist of a single term. These solutions are well known; the metric (2) can be
presented as



dr 2
1 2k/(dr d ) 2
2
dt +
+ r 2 d 2
= H d(I )/9
ds11
H
1 2k/(dr d )

2
2
+ dsoff
,
(A.2)
+ H 1 dson
p
2 and ds 2
where H = H (r) = 1 + p/(dr d ), p = k 2 + 2Q2 k, d = d0 1; dson
off
are the on-brane and off-brane extra-dimension line elements, respectively; the
dimension d0 of the sphere M0 varies from 2 to 7 for d(I ) = 3 (an electric brane) and
from 2 to 4 for d(I ) = 6 (a magnetic brane). In particular, the cases of maximum d0 ,
when off-brane extra dimensions are absent, correspond in the extremal near-horizon
limits to the famous structures AdS4 S 7 (electric) and AdS7 S 4 (magnetic). All
these solutions are stable under linear spherically symmetric perturbations.
2. Some examples of orthogonal systems (OS), whose stability in the general case is yet
to be studied, are:
(i)

(123), (145), (167) 3 electric branes; d0 = 2 or 3.

(ii)

(123), (124567) 1 electric and 1 magnetic branes; d0 = 2 or 3.

(iii)

(123), (145), (124678), (135678); d0 = 2.

The metrics are easily found from (36) with D 2 = 9, Ys2 = 2 and the equalities
OS

marked = .
The systems (i) with d0 = 3 and (iii) are remarkable in that their extremal limits have
regular horizons and the near-horizon geometries are, respectively, AdS2 S 3 T 6 and
AdS2 S 2 T 7 if the remaining extra dimensions are compactified on tori.
When the orthogonal systems form BOS-blocks (i.e., in the special case of equal charges
and a unique function y(u)), the solutions are stable according to Section 6.2.

K.A. Bronnikov, V.N. Melnikov / Nuclear Physics B 584 (2000) 436458

457

3. All two-brane BOS-block BHs are stable according to Section 6.1, for instance,
(i)

(123), (123456);

(ii)

(123), (145678);

YE 1 YE 2 = 1.
YE 1 YE 2 = 1.

The norms are equal (Ys2 = 2), and the angle is 60 in case (i) and 120 in case (ii).
4. Many seemingly possible three-brane blocks turn out to be forbidden due to a zero
value of a certain as (see Section 3.3). Consider, e.g.,

2 0
1

(123), (145), (123678);
YE s YE s 0 = 0 2 1 .
(A.3)
1 1 2
One easily finds from (30) that (a1 , a2 , a3 ) = (0, 1/2, 1/2), so one of the charges
should be zero, which means that such a system cannot exist.
5. The following is an example of a single-block BH whose stability can be established
by an individual study as described in Section 6.2:

2 0 1

YE s YE s 0 = 0 2 0 .
(A.4)
(123), (145), (123467); d0 = 2 or 3,
1 0 2
From (30) it follows
(a1 , a2 , a3 ) = (2/7, 3/7, 2/7),

YE =

3
X

as YE s ,

YE 2 = 6/7.

(A.5)

s=1

Suppose for certainty d0 = 3. Then, in agreement with (46),

4/3 2/3 1/3



4
Z s Z s 0 = 2/3 4/3 2/3 ,
Z2 = .
21
1/3 2/3 4/3

(A.6)

The next stage is to separate the perturbation Z x, after which the remaining two
degrees of freedom obey Eqs. (92) of the form


2
X
2/7
0
2y
b
Fss 0 s 0 ,
Fss 0 =
.
(A.7)
L s = 2q e
2/7 6/7
0
s =1

The characteristic equation det |Fss 0 ss 0 | = 0 has the form (2/7 )(6/7 ) = 0,
and, according to Proposition 2, the positivity of its roots proves the stability of the
background configuration.
References
[1]
[2]
[3]
[4]
[5]

V.D. Ivashchuk, V.N. Melnikov, Class. Quant. Grav. 14 (1997) 3001; hep-th/9705036.
K.A. Bronnikov, V.D. Ivashchuk, V.N. Melnikov, Grav. Cosmol. 3 (1997) 203; gr-qc/9710054.
K.A. Bronnikov, J. Math. Phys. 40 (1999) 924; gr-qc/9806102.
K.A. Bronnikov, Grav. Cosmol. 4 (1998) 49; hep-th/9710207.
V.D. Ivashchuk, V.N. Melnikov, Class. Quant. Grav. 16 (1998) 849; hep-th/9802121.

458

K.A. Bronnikov, V.N. Melnikov / Nuclear Physics B 584 (2000) 436458

[6] V.R. Gavrilov, V.N. Melnikov, in: Proc. Int. Sem. on Mathematical Cosmology, Potsdam,
Current Topics in Mathematical Cosmology, WS, Singapore, April 1998, p. 310.
[7] V.D. Ivashchuk, V.N. Melnikov, in: S. Cotsakis, G. Gibbons (Eds.), Mathematical and Quantum
Aspects of Relativity and Cosmology, Lecture Notes in Physics, Vol. 537, Springer, Berlin,
2000 (Proc. 2nd Samos Meeting, 1998); gr-qc/9901001.
[8] S. Cotsakis, V.D. Ivashchuk, V.N. Melnikov, Grav. Cosmol. 5 (1999) 52.
[9] V.D. Ivashchuk, V.N. Melnikov, Black hole p-brane solutions for general intersection rules,
hep-th/9910041, submitted to Clas. Quant. Grav.
[10] E. Witten, Nucl. Phys. B 443 (1995) 85.
[11] J. Maldacena, Adv. Theor. Math. Phys. 2 (1998) 231.
[12] J.L. Petersen, Introduction to the Maldacena conjecture on AdS/CFT, hep-th/9902131.
[13] E. Kiritsis, Dualities and instantons in string theory, hep-th/9906018.
[14] T. Regge, J.A. Wheeler, Phys. Rev. 108 (1957) 1063.
[15] C.V. Vishveshwara, Phys. Rev. D 1 (1970) 2870.
[16] V. Moncrief, Phys. Rev. D 10 (1974) 1057.
[17] K.A. Bronnikov, Yu.N. Kireyev, Phys. Lett. A 67 (1978) 95.
[18] K.A. Bronnikov, A.V. Khodunov, Gen. Rel. Grav. 11 (1979) 13.
[19] P. Kanti, E. Winstanley, Do stringy corrections stabilize coloured black holes?, gr-qc/9910069.
[20] K.D. Kokkotas, B.G. Schmidt, Quasi-normal modes of stars and black holes, gr-qc/9909058.
[21] R. Gregory, R. Laflamme, Phys. Rev. Lett. 70 (1993) 2387; hep-th/9301052.
[22] R. Gregory, R. Laflamme, Nucl. Phys. B 428 (1994) 399; hep-th/9494071.
[23] R. Gregory, R. Laflamme, Phys. Rev. D 51 (1995) 7007; hep-th/9410050.
[24] K.A. Bronnikov, Izv. Vuzov, Fizika 1 (1992) 106110.
[25] K.A. Bronnikov, V.N. Melnikov, Ann. Phys. (NY) 239 (1995) 40.
[26] P. Kanti, N.E. Mavromatis, J. Rizos, K. Tamvakis, E. Winstanley, Phys. Rev. D 57 (1998) 6255.
[27] K.A. Bronnikov, Acta Phys. Pol. B 4 (1973) 251.
[28] F.R. Gantmacher, Matrix Theory, Nauka, Moscow, 1988, in Russian.

Nuclear Physics B 584 (2000) 459479


www.elsevier.nl/locate/npe

Large-angle Bhabha scattering and luminosity at


flavour factories
C.M. Carloni Calame a , C. Lunardini b , G. Montagna a, , O. Nicrosini c ,
F. Piccinini c
a Dipartimento di Fisica Nucleare e Teorica, Universit di Pavia, and INFN, Sezione di Pavia,

via A.Bassi 6, 27100, Pavia, Italy


b SISSA-ISAS, via Beirut 2-4, 34100, Trieste, Italy and INFN, Sezione di Trieste,

via Valerio 2, 34127, Trieste, Italy


c INFN, Sezione di Pavia, and Dipartimento di Fisica Nucleare e Teorica, Universit di Pavia,

via A. Bassi 6, 27100, Pavia, Italy


Received 6 April 2000; revised 26 May 2000; accepted 5 June 2000

Abstract
The luminosity determination of electronpositron colliders operating in the region of low-lying
hadronic resonances (Ecm ' 110 GeV), such as BEPC/BES, DA8NE, KEKB, PEP-II and VEPP2M, requires the precision calculation of the Bhabha process at large scattering angles. In order to
achieve a theoretical accuracy at a few 0.1% level, the inclusion of radiative corrections is mandatory.
The phenomenologically relevant effect of QED corrections is taken into account in the framework
of the Parton Shower (PS) method, which is employed both for cross section calculation and event
generation. To test the reliability of the approach, a benchmark calculation, including exact O()
corrections and higher-order leading logarithmic contributions, is developed as well and compared
in detail with the PS predictions. The effect of O() next-to-leading and higher-order leading
corrections is investigated in the presence of realistic event selections for the Bhabha process at
the factories. A new Monte Carlo generator for data analysis (BABAYAGA) is presented, with an
estimated accuracy of 0.5%. Possible developments aiming at improving its precision and range of
applicability are discussed. 2000 Elsevier Science B.V. All rights reserved.
PACS: 02.70.Lq; 12.20.Ds; 13.10.+q
Keywords: Electronpositron collision; Flavour factories; Bhabha scattering; Radiative corrections; Parton
Shower; Monte Carlo

1. Introduction
The accurate determination of the machine luminosity is a fundamental ingredient for
the successful accomplishment of the physics programme of electronpositron (e+ e )
Corresponding author. E-mail: guido.montagna@pv.infn.it; tel: +39-0382-507742; fax: +39-0382-526938

0550-3213/00/$ see front matter 2000 Elsevier Science B.V. All rights reserved.
PII: S 0 5 5 0 - 3 2 1 3 ( 0 0 ) 0 0 3 5 6 - 4

460

C.M. Carloni Calame et al. / Nuclear Physics B 584 (2000) 459479

colliders operating in the region of the low-lying hadronic resonances, such as BEPC/BES
(Beijing), DA8NE (Frascati), VEPP-2M (Novosibirsk), as well as for the BELLE and
BABAR experiments around the at KEKB (KEK) and PEP-II (SLAC). In particular,
the precise measurement of the hadronic cross section at the factories requires a
luminosity determination with a total relative error better than 1% [14]. As well known,
the luminosity L of e+ e colliders can be precisely derived via the relation L = N/th ,
where N and th are the number of events and the theoretical cross section of a given
reference reaction, respectively. In order to make the total (experimental and theoretical)
luminosity error as small as possible, the cross section th of the reference process should
be large, in order to keep the statistical uncertainty small, and calculable with high
theoretical accuracy.
At e+ e machines operating in the energy range 110 GeV, the best candidate fulfilling
the above criteria is the process e+ e e+ e (Bhabha scattering) detected at large
scattering angles, say in the angular range 20 6 6 160 . For example, at the factory
DA8NE the KLOE detector can be used to detect such events [4], where the Bhabha
scattering cross section is significant, being of the order of 104 nb at a center of mass

(c.m.) energy around the resonance ( s ' 1 GeV).


Therefore, on theoretical side, precision calculations of the large-angle Bhabha (LABH)
cross section are demanded, with a theoretical accuracy at a few 0.1% level. This implies
to include in the calculation all the phenomenologically relevant radiative corrections, in
particular the large effects due to photonic radiation. Furthermore, such effects should be
implemented and accurately simulated in event generators, which are strongly demanded
by the experimental analysis.
At present, the status of the theoretical predictions and generators of interest for the
LABH process at low-energy e+ e machines can be summarized as follows. An exact
O() generator, based on the calculation of Ref. [5] and modified to match DA8NE
characteristics, is used in Monte Carlo (MC) studies by the KLOE collaboration [4,6].
An independent O() generator is also in use by the CMD-2 and SND experiments at
VEPP-2M [7]. In both the programs used in such MC simulations the effect of higherorder corrections is not taken into account. A semi-analytical calculation of the cross
section for large-angle QED processes, i.e., e+ e + , e+ e , , below 3 GeV
was performed in Ref. [8]. It includes exact O() plus leading logarithmic (LL) higherorder corrections. This formulation is available as a computer code described in Ref. [9].
The recently developed Bhabha generator BHWIDE [10], based on the YennieFrautschi
Suura approach for the treatment of QED radiation, appears in the list of simulation tools
presently under consideration by the BABAR collaboration at the B factory PEP-II [11].
A QED Parton Shower (PS) algorithm, which is employed in the present study, is adopted
in Refs. [12,13] for the computation of radiative corrections to LABH scattering. These
calculations, however, are optimized to high-energy LABH and differ, as it will be
discussed, in some aspects from the present implementation of the PS model. A complete
inventory of existing calculations and programs, for both small- and large-angle Bhabha
scattering, used at very high-energy e+ e colliders can be found in Ref. [14].

C.M. Carloni Calame et al. / Nuclear Physics B 584 (2000) 459479

461

The paper is organized as follows. In Section 2 the theoretical formulation, which is


based upon a QED realization of the PS method to account for radiative corrections
due to photon emission, is described. The steps and kinematics of the algorithm are
reviewed. Section 3 is devoted to the description of a new, original PS generator for the
simulation of the LABH process and based upon the formulation previously discussed.
A first sample of numerical results from the PS Bhabha generator is also given, with
particular emphasis on the simulation of the Bhabha process at the factories in the
presence of realistic event selections (ES). In the following sections tests of the reliability
of the PS approach are shown and commented, both at the level of integrated cross sections
and differential distributions. To this end, the calculation of the exact O() cross section
and its matching with higher-order LL corrections is addressed in Section 4. This is meant
as a benchmark calculation, developed in order to check the physical + technical precision
of the PS generator. Detailed comparisons between the PS predictions and the results of
the benchmark computation are given in Section 5. Conclusions, open issues and possible
developments are discussed in Section 6.

2. Theoretical approach and the Parton Shower method


In order to approach the aimed theoretical accuracy, the calculation of the QED corrected
Bhabha scattering cross section and the relative event generation is performed according
to the master formula: 1
Z
Z


(s) = dx dx+ dy dy+ dlab D x , Q2 D x+ , Q2

 d0
(x x+ s, cm )J (x , x+ , lab )(cuts),
D y , Q2 D y+ , Q2
dcm

(1)

which is based on the factorization theorems of (universal) infrared and collinear


singularities. Eq. (1) can be worked out within a QED PS algorithm for the calculation
of the electron Structure Function (SF) D(x, Q2 ), both for initial-state radiation (ISR)
and final-state radiation (FSR). In Eq. (1) d0 /d is the Born-like differential cross
section relevant for centre c.m. energy between 110 GeV, including the photonic sand t-channel diagrams and their interference, and the contributions due to exchange of
vector resonances, such as , J / and . Following the standard procedure described in
Ref. [16], the contribution of hadronic resonances is taken into account in terms of their
effective couplings to the electron. At c.m. energy around 1 GeV, the total contribution
amounts to 0.1 (0.3)% for 20 (50) 6 6 160 (130 ), where is the electron
scattering angle. For higher energies, as in the case of BELLE and BABAR experiments
around the , the effect of the resonance to the Bhabha cross section is of the same
order. It is worth noticing that for the low-energy colliders the LABH cross section is
largely dominated by t-channel photon exchange; hence, its leading dynamics is quite
similar to small-angle Bhabha (SABH) at LEP1/SLC and LABH at LEP2 and higher
1 See, for example, Ref. [15].

462

C.M. Carloni Calame et al. / Nuclear Physics B 584 (2000) 459479

energies [14]. In the hard-scattering cross section, the relevant correction due to vacuum
polarization is taken into account as well, by adopting the parameterization of Ref. [17].
In the evaluation of the hadronic contribution to the vacuum polarization, the Euclidean
value of the momentum in the photon propagator has been used as the appropriate scale

for time-like momenta [18,19]. The effect of the running coupling constant at s = M is
to enhance the cross section by 2 (2.5)% for 20 (50 ) 6 6 160 (130). The factor
J (x , x+ , lab ) in Eq. (1) accounts for the boost from the c.m. to the lab frame due to
emission of unbalanced ISR from the electron and positron legs, while (cuts) represents
cuts implementation.
The basic ingredient of Eq. (1) is the electron SF D(x, Q2 ), which represents the
probability density of finding inside a parent electron an electron with momentum
fraction x and virtuality Q2 . It can be explicitly obtained in QED by solving the
DokshitzerGribovLipatovAltarelliParisi (DGLAP) evolution equation [2022] in the
non-singlet channel:


D x, Q2 =
Q
2
Q2

Z1



x 2
dy
P+ (y)D
,Q ,
y
y

(2)

where P+ (x) is the regularized e e + splitting function


1 + x2
(1 x)
P+ (x) =
1x

Z1
dt P (t).

(3)

Q2

entering the SF will be, in general, dependent on the


Notice that the energy scale
specific process under study. No exact analytical solution of Eq. (2) is known in the
literature. After the efforts undertaken in the last decade for the program of precision
physics at LEP/SLC, the theoretical situation can be summarized as follows (see for
instance Ref. [23] and references therein):
(i) Approximate (accurate) analytical solutions in the collinear limit [2432]. This
kind of SFs is implemented in most of the programs for data analysis at
LEP/SLC [23];
(ii) Exact numerical solution, obtained via a numerical calculation of the Mellin
transform of the SF. This can be considered as a benchmark solution, but it is
unusable from the practical point of view for implementation in computational
tools;
(iii) Exact MC solution, obtained by means of the PS approach, which is particularly
powerful for exclusive event generation [3342]. 2 The PS method, originally
developed and widely applied in perturbative QCD [3339], has been recently
introduced in QED [4042] as a convenient framework to compute photonic
radiative corrections in e+ e collisions. In fact, exponentiation of soft photons and
the contribution of multiple emission of hard collinear photons can be automatically
accounted for.
2 A review of the PS method in QCD and relative QCD event generators can be found in Ref. [39].

C.M. Carloni Calame et al. / Nuclear Physics B 584 (2000) 459479

463

Let us summarize the basics of the PS method. The starting point of the PS approach is the
Sudakov form factor [43]:
"
#
Zs1 0 Zx+

ds
dz P (z) ,
(4)
(s1 , s2 ) = exp
2
s0
s2

which represents the probability that an electron evolves from virtuality s2 to virtuality
s1 with no emission of photons of energy fraction greater than  = 1 x+ , where  is an
infrared regulator. In terms of the factor (s1 , s2 ), the DGLAP equation can be written in
iterative form as:
D(x, s) = (s, m2 )(1 x)
Zs
Zx+
 0
0 ds
0
2
(s , m )
dy P (y)(x y)
+ s, s
s0
2
0

m2

Zs
+

ds 0
(s, s ) 0
s
0

m2

2

Zx+

s0

ds 00
(s 00 , m2 )
s 00

m2
Zx+

dx2 P (x1 )P (x2 )(x x1 x2 ) + ,

dx1
0

(s 0 , s 00 )

(5)

= s is understood, s = 4E 2 being the total c.m. energy. Eq. (5) suggests the
steps to compute D(x, s) by means of a MC algorithm [3341]:
(i) Set initial values for the electron virtuality and momentum fraction: K 2 = m2e and
x = 1 (with me electron mass).
(ii) Choose a random number in [0, 1].
(iii) If < (s, K 2 ) then stop: no photon is radiated.
(iv) Else if > (s, K 2 ), a photon is emitted: calculate the value K 02 of the electron
virtuality after the branching as the solution of the equation = (K 02 , K 2 ).
(v) Choose the residual momentum fraction z of the electron in [0, x+ ], according to
P (z).
(vi) Replace x by zx and K 2 by K 02 . Go back to step 2.
In this way the emission of a shower of photons by an electron is simulated, and the x
distribution of the PS event sample reproduces D(x, Q2 ). Such a distribution, obtained
from a 105 event sample at Q2 = s = (190)2 GeV2 , is compared in Fig. 1 with a numerical
solution of Eq. (2), normalized to the same number of events. The agreement is excellent.
A further test of the algorithm is the comparison between the exact analytical Mellin
moments D(N) of the SF and the corresponding ones calculated in the PS scheme. This
comparison is shown in Fig. 2, where it can be seen that the PS simulation (markers) for
N = 1, 2, 10, 50, 200 well agrees, within the statistical errors, with the analytical moments
(solid line). The results shown in Fig. 1 and Fig. 2 have been obtained with the value  =
109 for the infrared cut-off entering Eq. (4). However, independence of the PS predictions
for the QED corrected cross section from  variation, on a wide range of  values (from

where here Q2

464

C.M. Carloni Calame et al. / Nuclear Physics B 584 (2000) 459479

Fig. 1. x distribution of the electron SF at s = 190 GeV. Solid line: numerical solution of
DGLAP equation, by means of numerical inversion of Mellin transform. Histogram: result of the
PS algorithm.

1016 to 104 ), has been successfully checked, as shown in Fig. 3 for the LABH cross
section at the factory. In the present implementation of the PS model  is taken constant,
in order to avoid loss of accuracy in the determination of the absolute value of the electron
SF, as first pointed out in Refs. [40,41].
An up to O() PS algorithm has been developed as well. It allows to calculate the
corrected cross section of Eq. (1) up to O(). Such a calculation is strongly required
for fully consistent comparisons between the PS predictions and an exact perturbative
calculation. The steps required for the O() PS can be obtained by using Eq. (5) and
expanding the product D(x , Q2 )D(x+ , Q2 )D(y+ , Q2 )D(y , Q2 ) present in Eq. (1) up
to O(). It is easy to see that:


D(x , s)D(x+ , s)D(y+ , s)D(y , s) O()


= 4 s, m2 O() (1 x+ )(1 x )(1 y+ )(1 y )
s 

ln
(1 x )(1 y+ )(1 y )P (x+ )
+
2 m2
+ (1 x+ )(1 y+ )(1 y )P (x )
+ (1 x+ )(1 x )(1 y+ )P (y )


+ (1 x+ )(1 x )(1 y )P (y+ ) ,

(6)

where [ 4 (s, m2 )]O() is the product of four Sudakov form factors expanded up to O().
The steps for O() PS are therefore the following:

C.M. Carloni Calame et al. / Nuclear Physics B 584 (2000) 459479

465

Fig. 2. Comparison for the Mellin moments of the electron SF at s = 190 GeV. Solid line: exact
analytical moments. Markers: results of the PS algorithm for N = 1, 2, 10, 50, 200.

(i) Set initial values for the fermions virtuality and momentum fractions: K 2 = m2e and
x+ = x = y+ = y = 1.
(ii) Choose a random number in [0, 1].
(iii) If < [ 4 (s, K 2 )]O() then stop: no photon is radiated.
(iv) Else if > [ 4 (s, K 2 )]O() , a photon is emitted: calculate the value K 02 of
the fermion virtuality after the branching as the solution of the equation =
[ 4 (K 02 , m2 )]O() .
(v) Choose randomly the fermion which has emitted the photon.
(vi) Choose the residual momentum fraction z of the fermion in [0, 1 ], according
to P (z) and replace x+ , x , y+ or y with z, according to the particle which has
radiated.
(vii) Stop the algorithm.
The PS algorithm, both in the all order and O() implementation, offers the possibility to
go naturally beyond the strictly collinear treatment of the electron evolution, by generating
the transverse momentum p of electrons and photons at each branching. To this end,
a definite role for the variable x must be chosen. In the QED PS model of Refs. [40,41] x
is understood as the longitudinal momentum fraction of the electron after photon emission.
In the present PS simulation, x is chosen as the energy fraction of the electron after photon
emission (hereafter denoted as E scheme), in agreement with the known perturbative
results for the photon spectrum in QED [4446] and previous interpretation in perturbative
QCD (as, for example, in the paper by MarchesiniWebber in Refs. [3339]). It is worth

466

C.M. Carloni Calame et al. / Nuclear Physics B 584 (2000) 459479

Fig. 3. QED corrected Bhabha cross section as a function of the infrared regulator , at the peak
of the resonance. Cuts used are given in the text and correspond to a realistic event selection at
DA8NE.

noticing that the two prescriptions are coincident in the collinear limit. In the E scheme,
the kinematics of the branching process e(p) e0 (p0 ) + (q) can be written as:
E pz ),
p = (E, 0,
p0 = (zE, pE , pz0 ),


q = (1 z)E, pE , qz ,

(7)

with built-in energy and transverse momentum conservation. After having generated the
variables k 2 , k 0 2 and z by the PS algorithm, the on-shell conditions p2 = k 2 , p0 2 = k 0 2 ,
q 2 = 0, together with the longitudinal momentum conservation, allow to obtain complete
event reconstruction as follows:
k2
,
2E
(1 z)k 2 + k 0 2
,
pz0 = zE
2E
zk 2 k 0 2
,
qz = (1 z)E
2E

2
2
= (1 z) zk 2 k 0 ,
p
pz = E

(8)

2 /E 2  1. An alternative procedure, followed in the


at first order in k 2 /E 2  1, p
literature [42], consists in generating a photon p according to the leading pole behaviour

C.M. Carloni Calame et al. / Nuclear Physics B 584 (2000) 459479

467

Fig. 4. Comparison between the QED corrected LABH cross section and the Born one, in the c.m.
energy range from 1 GeV to 10.5 GeV. Solid circles: Born approximation; solid triangles: QED
corrected cross section. Cuts are given in the text.

1/p q. As a cross-check of the results obtained by means of Eq. (8), the method of
Ref. [42] has been also employed in the present implementation of the PS model, finding
agreement between the procedures for the calculation of the QED corrected Bhabha cross
section.

3. Bhabha generator and first sample of phenomenological results


In the spirit of the PS approach described above, a new MC event generator
(BABAYAGA) for simulation of the LABH process at e+ e flavour factories has been
developed. It is a generator of unweighted events, giving as output the QED corrected
cross section and the momenta of the final-state particles. Both ISR and FSR are simulated.
In principle, an arbitrary number of photons can be generated, including their p . In
the standard version, BABAYAGA records the momenta of electron and positron, and of
the most energetic and next-to-most energetic photons generated by the electromagnetic
shower. The possibility of an up to O() calculation of Eq. (1) is included as an option, in
order to compare it with the exact O() perturbative results (see Section 5). Also for the
O() branch, particles momenta are reconstructed.
As a first example, in Fig. 4, the PS corrected LABH cross section is compared with the
Born-like cross section, as a function of the c.m. energy. An energy threshold of 0.8 Ebeam

468

C.M. Carloni Calame et al. / Nuclear Physics B 584 (2000) 459479

for both electron and positron is required, the acceptance cuts are 20 6 6 160 and an
acollinearity cut of 5 is imposed. In such a situation, the correction due to QED radiation
is of the order of 10% at the factories and of the order of 20% at the B factories, pointing
out the need of a careful treatment of photonic corrections in simulation tools of the LABH
process at flavour factories.
A further sample of phenomenological results obtained by means of the BABAYAGA
program is shown in Figs. 57. The parameters and selection criteria adopted in the analysis
are very similar to those considered in previous MC simulations [4,6] and correspond to

realistic data taking at DA8NE and VEPP-2M, for c.m. energy at the peak, i.e., s =

=
1.019 GeV. The energy cut imposed on the final-state electron and positron is Emin

0.4 GeV; two different angular acceptances of 20 6 6 160 and 50 6 6 130


are considered, with (maximum) acollinearity cut allowed to vary in the range max =
5 , 10 , 15 , 20, 25 . The energy cut refers to the energy of the bare electron and positron,
corresponding to a so-called bare ES (see for instance Ref. [14]), which is not far from
realistic due to the presence of magnetic fields as for DA8NE and CMD-2 experiment at
VEPP-2M. The tight acollinearity cut max = 5 is generally introduced in MC simulations
in order to single out quasi-elastic Bhabha events [4]. However, it is worth noticing that
this acollinearity cut, in association with the high energy thresholds on bare electrons and
positrons, defines a rather severe set of constraints, which can be expected to emphasize
the effects of QED radiative corrections, especially from the final-state. This conjecture is
confirmed by the numerical results of Fig. 5, where the (relative) effect, with respect to the
Born cross section, of ISR only (open circles) is compared with the whole effect of ISR and
FSR (solid circles). Actually, it can be seen that the photon radiation produces a lowering of
the integrated cross section of the order of 10% and that half of this effect has to be ascribed
to FSR, showing the need of including FSR for realistic simulations. In this simulation,
the Q2 entering the electron SF is fixed to be Q2 = s as typical virtuality in initial- and
final-state shower. The relevance of FSR in such ES is further illustrated in Fig. 6 for the
electron energy, electron scattering angle, electron transverse momentum distribution and
acollinearity distribution as well. It can be noticed that the electron energy and p are
significantly affected by FSR, while this is not the case for the electron scattering angle
and acollinearity, which are actually largely dominated, as expected, by the x distribution
of ISR [15]. In the comparison shown in Fig. 6 the numbers of generated events for ISR
only and ISR+FSR are consistently normalized to the same luminosity.
A further illustration of the potential of BABAYAGA in event generation is given
in Fig. 7, showing the energy, polar angle and transverse momentum of the most
energetic
photon, as well as the missing mass distribution of the event, defined as
p
(p+ + p q+ q k)2 , where p (q ) are the initial(final) electron/positron momenta and k is the momentum of the most energetic photon. A minimum energy cut of
5 MeV is imposed as visibility criterion. It can be seen that the expected physical features of photon radiation are correctly reproduced by the PS generator, noticeably the soft

peaking behaviour of E , p and missing mass distribution as well as the initial- and the
final-state collinear peaks, which are clearly visible in the cos distribution.

C.M. Carloni Calame et al. / Nuclear Physics B 584 (2000) 459479

469

Fig. 5. Relative effect of ISR and ISR + FSR on the integrated Bhabha cross section at the
factories. Open circles: ISR only; solid circles: ISR + FSR. Cuts used are given in the text.

4. Benchmark calculation
In order to test the (physical + technical) precision of the PS approach and corresponding Bhabha generator, an exact O() perturbative calculation has been addressed, by computing the up to O() corrected cross section as follows
()
()
= S+V
(E < k0 ) + H() (E > k0 , cuts).
exact

(9)

()
is obtained by integration over
In the above equation the O() soft + virtual part S+V
the electron angle of the soft+virtual differential cross section, which is explicitly given
by [47,48]


(),i
(10)
dS+V = d0 1 + 2 (e + int ) ln k0 /E + CFi .

In Eqs. (9) and (10), k0 stands for a (small) photon energy softhard (fictitious) separator,
i is an index for the photonic contributions to the Bhabha Born matrix element (i =
| (s)|2 ,| (t)|2 , (s) (t)), e = 2/(ln(s/m2e ) 1) is the leading collinear factor for
initial- and final-state radiation, int = 2/ ln(t/u) is the leading angular factor for
initialfinal state interference. Furthermore, in Eq. (10) CFi = CFi () are the soft + virtual
K-factors for the three QED channels, including box/interference finite terms [47,48]. The
hard bremsstrahlung contribution H() is included, in the photon phase-space region above
the energy cut k0 , via the e+ e e+ e matrix element calculated in Ref. [49]. Its
contribution is computed numerically in the presence of experimental cuts by means of
the MC method with standard importance sampling technique, in order to handle collinear

470

C.M. Carloni Calame et al. / Nuclear Physics B 584 (2000) 459479

Fig. 6. Effect of FSR on Bhabha differential distributions at the factories: electron energy, electron
scattering angle, electron transverse momentum and acollinearity. Dashed histograms: ISR only.
Solid histograms: ISR + FSR. Cuts are given in the text.

and infrared photon singularities. Needless to say, the independence of Eq. (9) from the
fictitious infrared cut-off k0 was successfully checked with high numerical precision.
The comparison between the exact O() calculation and the O() predictions of the PS
generator is of interest because it allows to evaluate the size of the O() next-to-leading
order (NLO) corrections, which are missing in the PS. Moreover, this comparison can be
a useful guideline to improve the agreement between perturbative and LL PS results, for
example, by properly choosing the virtuality Q2 in the electron SF in such a way that O()
NLO terms are effectively reabsorbed into the LL contributions. It can be noticed, in fact,
that by choosing the scale Q2 as Q2 = st/u in the collinear logs ln(Q2 /m2 ) generated by
the PS method, then one has that ln(Q2 /m2 ) ln(s/m2 ) + ln(t/u) [50]. If this procedure
is applied both to ISR and FSR by choosing as maximum virtuality of the electromagnetic
shower Q2 = st/u, it is possible to keep under control, besides the large logarithms from
ISR and FSR, also the leading angular contribution from initialfinal state interference.
In order to estimate the precision of the PS approach with all order corrections, higherorder LL terms must be added to the exact O() cross section in the benchmark calculation.
The general algorithm recently proposed in Ref. [51], and there applied to the highprecision computation of the SABH cross section, can be advocated, by writing the
benchmark cross section as (in the so-called additive form) [51]
()
()
()
LL
+ exact
,
= LL

(11)

C.M. Carloni Calame et al. / Nuclear Physics B 584 (2000) 459479

471

Fig. 7. Distributions of the most-energetic photon in the LABH process at the factories: energy,
polar angle, transverse momentum and missing mass. Cuts are given in the text.
()

()

where LL is the all-order LL cross section as given by Eq. (1), LL is the up to O()
()
truncation of the LL cross section, exact is the exact perturbative O() cross section of
Eq. (9). Collinear SFs as given in Ref. [24] are used in the calculation of the LL cross
sections. Adopting such a procedure, exact O() corrections are simply matched with
LL higher-orders in the collinear approximation. Therefore, the cross section of Eq. (11)
includes exact O() + O( n Ln ), with n > 2, leading corrections.
A more accurate factorized form, as motivated and discussed in detail in Ref. [51], can
be supplied, by computing the cross section as
() 
F ' 1 + CNL LL ,
()

()

CNL

()
()

exact LL
NL ,
0
0

(12)

()

where CNL is the whole NLO content of Eq. (9). In such a way, the bulk of the most
important second-order NLO corrections, i.e., the O( 2 L) terms, are added to the content
of Eq. (11) [51]. 3
The procedure here shortly sketched works efficiently for cross section calculation (not
unweighted event generation) and allows, by comparison with the full PS predictions,
to estimate the overall precision of the PS approach. Furthermore, it is possible to get
3 Actually, in Ref. [51] a more refined treatment of the factorized cross section is given, and it is shown that
Eq. (12) is an approximation of the full treatment at the 0.1% level, which is sufficient for the present study.

472

C.M. Carloni Calame et al. / Nuclear Physics B 584 (2000) 459479

a measure of O( n Ln ) and O( 2 L) corrections. The benchmark calculation is available


in the form of MC integrator program (LABSPV), which is a suitable modification of the
SABSPV code [52] from SABH to low-energy LABH process.

5. Tests of the approach and further numerical results


As already remarked in the previous section, the detailed comparison between the exact
O() cross section of Eq. (9) and the up to O() expansion of the PS results is a valuable
tool to establish the physical precision of the PS approach, since the registered difference
is due to the NLO corrections left over in the pure LL predictions of the PS scheme. The
relative difference between Eq. (9) and the up to O() PS cross section is shown in Fig. 8
for the angular acceptances 20 6 6 160 and 50 6 6 130 and for two different
choices of the Q2 scale in the PS, i.e., for Q2 = st/u and Q2 = 0.75 st/u. The relative
deviations shown in the figure are plotted as functions of the acollinearity cut. For the
natural choice of the scale Q2 = st/u which allows to completely reproduce the LL
structure of the exact O() calculation, the difference is an unambiguous evaluation of
the NLO corrections. Such contributions, as a priori expected, are important in view of
the required theoretical precision, especially at the acollinearity value max = 5 , being
of the order of several 0.1% (see Fig. 8). However, it is worth noticing that a simple ad
hoc variation of the Q2 scale from Q2 = st/u to Q2 = 0.75 st/u significantly reduces
the difference, which goes from more than the 0.51% level down to about 0.10.3% at
max = 5 , depending on the angular set up. In general, with the adjusted scale Q2 =
0.75 st/u the difference between the exact O() calculation and the PS predictions is
within 0.5%, which is below the theoretical accuracy required for the calculation of the
LABH process in typical event selections at the -factories. Notice that improving further
the agreement between the PS and the benchmark calculation with a more refined choice
of the Q2 scale would be sensible only in the presence of a more stringent requirement on
the theoretical error and, in any case, a further refining down to 0.10.2% level would
mean entering the regime of the theoretical uncertainty of the benchmark calculation
itself. This naive example illustrates how, for a given set of cuts, the level of agreement
can be substantially improved by a simple redefinition of the maximum virtuality of the
electromagnetic shower. Going beyond this simple procedure would require a merging
between perturbative calculation and PS scheme, which is beyond the scope of the present
work.
In order to assess the reliability of the PS approach in the presence of higher-order
corrections, the comparison between the benchmark and PS calculation has been extended
to the computation of the exact O() plus higher-orders cross section. The relative
difference between Eq. (11) and the full PS cross section is shown in Fig. 9, still as
functions of the acollinearity cut. According to what discussed about the O() comparison,
the scale in the PS is, for definiteness, fixed at Q2 = 0.75 st/u. Relative differences
contained within 0.5% are still observed, confirming the equivalent implementation of
higher-order LL corrections in the PS and benchmark calculation. This difference between

C.M. Carloni Calame et al. / Nuclear Physics B 584 (2000) 459479

473

Fig. 8. Relative differences for the LABH process at the factories between the exact O() cross
section (LABSPV) and the up to O() PS one (BABAYAGA), as functions of the acollinearity
cut and for two choices of the Q2 scale in the electron SF. On the left the acceptance region
20 6 6 160 is considered, while on the right the acceptance region is 50 6 6 130 . Other
cuts are specified in the text.

Fig. 9. The same as Fig. 8 for the O() plus higher orders Bhabha cross section at the factories.
Cuts are given in the text.

BABAYAGA and LABSPV in the presence of higher-order corrections can be considered


as an estimate of the physical precision of the modified PS approach for the cross section
calculation for realistic ES at the factories.
In addition to the evaluation of the O() NLO corrections, for an assessment of
the theoretical precision, it is important to evaluate the impact of the higher-order LL
contributions. The size of LL O( n Ln ), with n > 2, corrections can be derived, as already

474

C.M. Carloni Calame et al. / Nuclear Physics B 584 (2000) 459479

Fig. 10. Effect of higher-order O( n Ln ) (n > 2) and O( 2 L) corrections on the integrated Bhabha
cross section at the factories. On the left the acceptance region 20 6 6 160 is considered,
while on the right the acceptance region is 50 6 6 130 . Other cuts are specified in the text.

remarked, by comparing in the benchmark calculation the results of Eqs. (9) and (11) or,
equivalently, in the PS scheme the full all-order predictions with the corresponding up to
O() truncation. Furthermore, the difference between Eqs. (11) and (12) in the benchmark
calculation is able to provide an indication of the size of O( 2 L) corrections, thus yielding
an estimate of the most important second-order NLO corrections left over in the PS method.
Such an analysis of higher-order effects leads to the results shown in Fig. 10, where the
relative effect of the above higher-order corrections is shown for the natural scale Q2 =
st/u. It can be clearly seen that LL O( n Ln ) corrections are unavoidable in view of the
expected theoretical accuracy. In fact, their contribution is at 12% level at max = 5
and of the order of some 0.1% for larger acollinearity cuts. The impact of O(2 L) is,
instead, negligible, being well below the 0.1% level, in agreement with the estimate given
in Ref. [8].
Having established that the inclusion of higher-order contributions does not alter, as
expected, the results of the O() comparison, further O() tests have been performed
at the level of exclusive differential distributions. The results are illustrated in Figs. 11
13. In these plots, the histograms represent the distributions of the events generated by
means of the up to O() PS generator BABAYAGA with scale Q2 = 0.75 st/u, while
the markers correspond to the predictions of the benchmark calculation LABSPV. A few
comments are in order here. Since the benchmark calculation is not suited for unweighted
event generation, the number of events corresponding to the markers have been obtained by
calculating, as a first step, the integrated cross section over each bin and next by multiplying
it for a reference luminosity calculated as L = NPS /PS , where NPS and PS are the number
of events and the cross section obtained with the PS generator, respectively. The set of
cuts used for the analysis of the distributions corresponds to the angular range 20 160.
Generally, as it can be seen, the level of agreement is quite satisfactory. In particular,

C.M. Carloni Calame et al. / Nuclear Physics B 584 (2000) 459479

475

Fig. 11. Electron energy distribution for the LABH process at the factories. Markers: exact O()
via LABSPV. Histograms: PS prediction via BABAYAGA with scale Q2 = 0.75 st/u.

Fig. 12. Electron angle distribution for the LABH process at the factories. Markers: exact O()
via LABSPV. Histograms: PS prediction via BABAYAGA with scale Q2 = 0.75 st/u.

the electron energy distribution simulated by BABAYAGA, well agrees with the exact
LABSPV calculation, as shown in Fig. 11. Because the electron energy distribution is
driven in the present implementation of the PS method by the x distribution of the SF,

476

C.M. Carloni Calame et al. / Nuclear Physics B 584 (2000) 459479

Fig. 13. Acollinearity angle distribution for the LABH process at the factories. Markers: exact
O() via LABSPV. Histograms: PS prediction via BABAYAGA with scale Q2 = 0.75 st/u.

the agreement observed between exact and PS predictions reinforces, a posteriori, the
interpretation of the x variable as residual fraction energy as the most natural one in QED
PS models.
Also the PS description of the electron angular variables is in nice agreement with the
exact calculation, as shown by the electron scattering angle and acollinearity distribution of
Figs. 1213. Some disagreement is seen in the first acollinearity bins, where still acceptable
differences at a few per cent level are registered. The satisfactory PS predictions for the
electron angles can be understood since the distributions of these variables are dominated,
beyond the tree-level approximation, by the effect of the longitudinal boost due to the
emission of unbalanced electron and positron ISR, which, in turn, is governed by the x
distribution of the electron SF.

6. Conclusions and perspectives


The LABH process is used at present and B factories as the reference reaction to
determine the machine luminosity. In order to reach a total accuracy at a few 0.1% level,
precision calculations of the LABH scattering cross section and distributions become more
and more urgent. In particular, the relevant effects due to photon emission have to be
kept under control and accurately simulated in the computational tools required by the
experimental analysis.
Along this direction, an original calculation of the LABH process has been addressed.
It is based on a realization of the PS method in QED to account for photonic
corrections due to ISR, FSR and initialfinal-state interference. The approach adopted

C.M. Carloni Calame et al. / Nuclear Physics B 584 (2000) 459479

477

allows the calculation of integrated cross sections as well as event generation, including
reconstruction of photon p . A new MC event generator (BABAYAGA) has been
developed and is available for a full experimental simulation of the LABH process
at flavour factories. The program has been used to provide several numerical results
of phenomenological interest, thus showing the potential of BABAYAGA in physics
analysis. For example, it has been shown, both at the level of cross section and interesting
distributions, that the effects due to radiation from final-state electrons have to be carefully
taken into account in the presence of realistic ES at the factories DA8NE and VEPP2M, since the impact of FSR can be as large as the one due to ISR.
With the aim of checking the overall (physical + technical) precision of the PS approach
and relative generator, a benchmark calculation has been carried out as well. It relies upon
the exact O() calculation of the LABH cross section supplemented with higher-order LL
corrections in the collinear approximation. The benchmark computation is also available
in the form of computer code (LABSPV), which is a MC integrator allowing for precise
cross section calculations, with an estimated precision at 0.1% level.
By comparing the predictions of the PS generator BABAYAGA and of the benchmark
calculation LABSPV for the up to O() cross section with typical cuts, it turns out that
the contribution of the O() NLO corrections is, as expected, important in the light of the
required theoretical precision, being at the 0.51% level for typical ES at the factories.
However, it has been also shown that the scale Q2 entering in the PS LL calculation can be
simply adjusted in order to agree within 0.5% with the benchmark calculation. An energy
scale like Q2 st/u reveals to be the best choice that effectively reabsorbs O() NLO
terms into LL PS contributions. This conclusion, which holds for the integrated cross
section, has been proved to be generally valid also for the most interesting differential
distributions.
With the scale choice Q2 = st/u, the effect of higher-order O( n Ln ) LL corrections
has been evaluated and found to be at the 12% level, while the role of O( 2 L)
corrections is marginal, below 0.1% accuracy. Therefore, one of the main conclusions of
the present analysis is that, given the size of the radiative corrections discussed above,
theoretical predictions for the LABH process at flavour factories aiming at a few 0.1%
precision must include the contribution of both O() NLO terms and O( n Ln ) leading
logarithmic contributions. From the whole of the present analysis, it turns out that the
present physical precision of BABAYAGA generator is 0.5% for typical ES at the
factories. If particularly stringent requirements of theoretical accuracy would be in the
future necessary, the PS approach as presently implemented in BABAYAGA should be
updated by means of an appropriate merging of the exact O() matrix element with the
exclusive photon exponentiation realized by the PS algorithm. Some proposals addressing
such an issue are already available in the literature [5358], in order to obtain sensible
QCD phenomenological predictions for experiments at the TEVATRON and LHC.
A second possible development of the present work concerns application of the
PS scheme to other phenomenological studies of strong interest at e+ e flavour
factories. In fact, since the PS is a very general method to compute photonic radiative
corrections, the same formulation here applied to the LABH process could be employed

478

C.M. Carloni Calame et al. / Nuclear Physics B 584 (2000) 459479

to evaluate radiative corrections to other large-angle interesting QED processes, such


as e+ e + (n ), (n ), and also to hadronic final states, in particular for
e+ e hadrons and e+ e hadrons + , the latter of great interest for an energy
scan of the hadronic cross section below the nominal c.m. energy [5962].
A further foreseen development is a phenomenological analysis of QED processes at the
B factories, along the lines followed in the present study.
Such possible developments are by now under consideration.

Acknowledgements
The authors are very grateful to A. Bukin, G. Cabibbo, S. Eidelman, V.N. Ivanchenko,
F. Jegerlehner, V.A. Khoze, G. Kukartsev, J. Lee-Franzini, G. Pancheri, M. Piccolo,
I. Peruzzi, Z. Silagadze and G. Venanzoni for useful discussions, remarks and interest
in their work. The authors acknowledge partial support from the EEC-TMR Program,
Contract N. CT98-0169. C.M. Carloni Calame and C. Lunardini wish to thank the INFN,
Sezione di Pavia, for the use of computer facilities.
References
[1]
[2]
[3]
[4]
[5]
[6]
[7]
[8]
[9]
[10]
[11]
[12]
[13]
[14]

[15]
[16]
[17]
[18]

L. Maiani, G. Pancheri, N. Paver (Eds.), The DA8NE Physics Handbook, 1992.


L. Maiani, G. Pancheri, N. Paver (Eds.), The Second DA8NE Physics Handbook, 1995.
G. Pancheri, Acta Phys. Polonica 30 (1999) 2243.
G. Cabibbo, G. Venanzoni, Measuring the DA8NE luminosity with large angle Bhabha
scattering, KLOE MEMO n. 168, November 9, 1998.
F.A. Berends, R. Kleiss, Nucl. Phys. B 228 (1983) 537.
E. Drago, G. Venanzoni, A Bhabha generator for DA8NE including radiative corrections and
resonance, INFN/AE-97/48, 1997.
S. Eidelman, V.N. Ivanchenko, private communication.
A.B. Arbuzov et al., Sov. Phys. JETP 10 (1997) 001.
A.B. Arbuzov, LABSMC: Monte Carlo event generator for large-angle Bhabha scattering, hepph/9907298.
S. Jadach, W. Paczek, B.F.L. Ward, Phys. Lett. B 390 (1997) 298.
See, for example, the home page of the BABAR Event Generator Group, http://www.slac.
stanford.edu/BFROOT/www/Physics/Tools/generators/.
J. Fujimoto et al., Prog. Theor. Phys. 91 (1994) 333.
H. Anlauf et al., Comput. Phys. Commun. 79 (1994) 466.
S. Jadach, O. Nicrosini et al., Event generators for Bhabha scattering, in: G. Altarelli,
T. Sjstrand, F. Zwirner (Eds.), Physics at LEP2, CERN Report 96-01, Vol. 2, CERN, Geneva,
1996, p. 229.
G. Montagna, O. Nicrosini, F. Piccinini, Phys. Rev. D 48 (1993) 1021, and references therein.
M. Greco, G. Montagna, O. Nicrosini, F. Piccinini, in: L. Maiani, G. Pancheri, N. Paver (Eds.),
The Second DA8NE Physics Handbook, Vol. II, 1995, p. 629.
S. Eidelman, F. Jegerlehner, Z. Phys. C 67 (1995) 585.
F. Jegerlehner, private communication and DESY-99-007, hep-ph/9901386, in: J. Sola (Ed.),
Proc. of the IVth International Symposium on Radiative Corrections, RADCOR98, Barcelona,
Spain, 1998.

C.M. Carloni Calame et al. / Nuclear Physics B 584 (2000) 459479

[19]
[20]
[21]
[22]
[23]
[24]
[25]
[26]
[27]
[28]
[29]
[30]
[31]
[32]
[33]
[34]
[35]
[36]
[37]
[38]
[39]
[40]
[41]
[42]
[43]
[44]
[45]
[46]
[47]
[48]
[49]
[50]
[51]
[52]

[53]
[54]
[55]
[56]
[57]
[58]
[59]
[60]
[61]
[62]

479

J.G. Krner, A.A. Pivovarov, K. Schilcher, Eur. Phys. J. C 9 (1999) 551.


V.N. Gribov, L.N. Lipatov, Sov. J. Nucl. Phys. 15 (1972) 298.
G. Altarelli, G. Parisi, Nucl. Phys. B 126 (1977) 298.
Y.L. Dokshitzer, Sov. Phys. JETP 46 (1977) 641.
G. Montagna, O. Nicrosini, F. Piccinini, Riv. Nuovo Cimento 21 (1998) 1, hep-ph/9802302.
E.A. Kuraev, V.S. Fadin, Sov. J. Nucl. Phys. 41 (1985) 466.
G. Altarelli, G. Martinelli, in: J. Ellis, R. Peccei (Eds.), Physics at LEP, CERN Report 86-02,
Vol. 1, CERN, Geneva, 1986, p. 47.
O. Nicrosini, L. Trentadue, Phys. Lett. B 196 (1987) 551; Z. Phys. C 39 (1988) 479.
F.A. Berends, G. Burgers, W.L. van Neerven, Nucl. Phys. B 297 (1988) 429.
M. Skrzypek, S. Jadach, Z. Phys. C 49 (1991) 577.
S. Jadach, M. Skrzypek, B.F.L. Ward, Phys. Lett. B 157 (1991) 173.
M. Cacciari et al., Europhys. Lett. 17 (1992) 123.
G. Montagna, O. Nicrosini, F. Piccinini, Phys. Lett. B 406 (1997) 243.
A.B. Arbuzov, Phys. Lett. B 470 (1999) 252.
R. Odorico, Nucl. Phys. B 172 (1980) 157; Phys. Lett. B 102 (1981) 341.
P. Mazzanti, R. Odorico, Phys. Lett. B 95 (1980) 133.
G. Marchesini, B.R. Webber, Nucl. Phys. B 238 (1984) 1.
T. Sjstrand, Phys. Lett. B 310 (1985) 321.
K. Kato, T. Munehisa, Phys. Rev. D 39 (1989) 156.
K. Kato, T. Munehisa, H. Tanaka, Z. Phys. C 54 (1992) 397.
R.K. Ellis, W.J. Stirling, B.R. Webber, QCD and Collider Physics, Cambridge Univ. Press, 1996.
J. Fujimoto, T. Munehisa, Y. Shimizu, Prog. Theor. Phys. 90 (1993) 177.
Y. Kurihara, J. Fujimoto, T. Munehisa, Y. Shimizu, Prog. Theor. Phys. 96 (1996) 1223;
Prog. Theor. Phys. 95 (1996) 375.
H. Anlauf et al., Comput. Phys. Commun. 70 (1992) 97.
V.V. Sudakov, Sov. Phys. JETP 3 (1956) 65.
G. Bonneau, F. Martin, Nucl. Phys. B 27 (1971) 381.
V.N. Baier, V.S. Fadin, V.A. Khoze, Nucl. Phys. B 65 (1973) 381.
F.A. Berends, R. Kleiss, Nucl. Phys. B 260 (1985) 32.
M. Caffo, E. Remiddi et al., Bhabha scattering, in: G. Altarelli, R. Kleiss, C. Verzegnassi (Eds.),
Z Physics at LEP1, CERN Report 89-08, Vol. 1, CERN, Geneva, 1989, p. 171.
M. Greco, Riv. Nuovo Cimento 11 (1988) 1.
F.A. Berends et al., Nucl. Phys. B 202 (1981) 63.
M. Greco, O. Nicrosini, Phys. Lett. B 240 (1990) 219.
G. Montagna, O. Nicrosini, F. Piccinini, Phys. Lett. B 385 (1996) 348.
M. Cacciari, G. Montagna, G. Montagna, O. Nicrosini, F. Piccinini, in: D. Bardin, W. Hollik,
G. Passarino (Eds.), Report of the Working Group on Precision Calculations for the Z
Resonance, CERN Report 95-03, CERN, Geneva, 1995, p. 389; Comput. Phys. Commun. 90
(1995) 301.
F. Krauss, R. Kuhn, G. Soff, Acta Phys. Polonica 30 (1999) 3875.
G. Miu, T. Sjstrand, Phys. Lett. B 449 (1999) 313.
G. Corcella, M. Seymour, Phys. Lett. B 442 (1998) 417.
J. Andr, T. Sjstrand, Phys. Rev. D 57 (1998) 5767.
H. Baer, M.H. Reno, Phys. Rev. D 44 (1991) 3375.
M.H. Seymour, Comput. Phys. Commun. 90 (1991) 95.
S. Spagnolo, Eur. Phys. J. C 6 (1999) 637.
S. Binner, J.H. Kuhn, K. Melnikov, Phys. Lett. B 459 (1999) 279.
M. Benayoun et al., Mod. Phys. Lett. A 14 (1999) 2605.
A.B. Arbuzov et al., J. High Energy Phys. 12 (1998) 009.

Nuclear Physics B 584 (2000) 480508


www.elsevier.nl/locate/npe

Warped compactifications in M and F theory


Brian R. Greene a,1 , Koenraad Schalm b,2 , Gary Shiu c,
a Department of Physics and Mathematics, Columbia University, New York, NY 10027, USA
b NIKHEF Theory Group, P.O. Box 41882, Amsterdam 1009DB, The Netherlands
c C.N. Yang Institute for Theoretical Physics, State University of New York,

Stony Brook, NY 11794-3840, USA


Received 18 April 2000; revised 6 June 2000; accepted 16 June 2000

Abstract
We study M and F theory compactifications on CalabiYau four-folds in the presence of non-trivial
background flux. The geometry is warped and belongs to the class of p-brane metrics. We solve for
the explicit warp factor in the orbifold limit of these compactifications, compare our results to some
of the more familiar recently studied warped scenarios, and discuss the effects on the low-energy
theory. As the warp factor is generated solely by backreaction, we may use topological arguments
to determine the massless spectrum. We perform the computation for the case where the four-fold
equals K3 K3. 2000 Elsevier Science B.V. All rights reserved.

1. Introduction
Recently, there has been renewed interest in the idea that our world may be a domain
wall in a higher-dimensional spacetime [1]. This scenario, often known as the Brane
World scenario [210], is motivated by the existence of solitonic objects in string theory,
such as D-branes, in which gauge and matter fields are localized on lower-dimensional
subspaces. Indeed, if the Standard Model fields are localized on a brane, whereas gravity
propagates in a higher-dimensional spacetime, a number of phenomenological issues have
to be revisited. A rather interesting 5-dimensional example was suggested in [11] where
gravity seems to be localized on a 4-dimensional brane, even though the 5th dimension
has an infinite extent. A key ingredient is their setup is that the ambient spacetime is
warped; the localization of gravity is due to an exponentially damped warp factor. Several
attempts have been made to find domain wall solutions in 5D gauged supergravity that
may perhaps localize gravity [1215]. A warped metric can also be used to generate
Corresponding author. E-mail: shiu@insti.physics.sunysb.edu
1 greene@phys.columbia.edu
2 kschalm@nikhef.nl

0550-3213/00/$ see front matter 2000 Elsevier Science B.V. All rights reserved.
PII: S 0 5 5 0 - 3 2 1 3 ( 0 0 ) 0 0 4 0 0 - 4

B.R. Greene et al. / Nuclear Physics B 584 (2000) 480508

481

the hierarchy between the electroweak and the four-dimensional Planck scale [11]. The
possible embedding of such a scenario in string theory was discussed in [16].
In fact, warped compactifications are a rather logical generalization of compactifications
of the product type M4 K, and were considered some time ago in the context of
supergravity [17,18] and heterotic string theory [19,20]. In perturbative string theory or
supergravity there are no branes and in these scenarios there are therefore no gauge and
matter fields which are a priori localized in some lower-dimensional subspace. In strongly
coupled heterotic string theory, on the other hand, there are two end of the universe
branes at the boundaries of the non-perturbative 11th direction on which gauge and
matter fields are localized [2,3,8,9]. In general, the four-form field strength G is non-zero,
and as a result, spacetime is warped. However, the long wavelength expansion in elevendimensional supergravity is valid provided that R11 6 RCY / 2/3 where is the eleventh
dimensional gravitational coupling, and RCY is the size of the CalabiYau on which the
heterotic string theory is compactified. One cannot make the extra dimension transverse to
the branes too large, or else the effective field theory breaks down.
F theory [21,22] provides another powerful tool in studying non-perturbative string
vacua (see, e.g., [2325], for four-dimensional F theory compactifications). In compactifying M and F theory to lower dimensions, anomaly cancellation (cancellation of tadpoles)
often requires one to introduce branes [28] (or alternatively, to turn on some background
flux [26,27]). All p-brane supergravity solutions are examples of warped metrics due to
the backreaction of the branes. The recent developments give us motivation to better understand these warped compactifications in a variety of circumstances. And while the consistency requirements for M theory compactifications on CalabiYau four-folds (and their
lifting to F theory) have been studied in some detail more recently [2630], the form of
the background geometry has not been determined explicitly in the presence of non-zero
background flux (see, however [51]). A purpose of the present paper is to take some modest
steps in this direction.
In the first part of this paper, we will determine the warp factor for some M/F theory
compactifications with background flux and analyze its behavior. To solve the equation for
the warp factor explicitly, we consider orbifold limits of CalabiYau spaces. In the F theory
limit, the dilaton will also be taken to be constant over the base (the 7-brane charges are
locally cancelled). It is, however, likely that the generic features we find here will also hold
for general CalabiYau manifolds. We find that the background flux contributes to the warp
factor in the same manner and with the same scale as p-branes (both are 0 suppressed as
they are due to backreaction). The resulting metrics are thus generalizations of the p-brane
ones with a contribution to the harmonic function from the background flux. In particular
the warp factor always contains a constant part. This is to some extent expected as the
contribution of the background flux to the energy momentum tensor scales inversely with
the compactification volume. In the large volume limit, the contribution is negligible, and
we should recover the regular p-brane metric. By way of comparison, the contribution
of a cosmological constant (as in [11]) does not scale with the compactification volume.
In the proposal of [16], however, one restricts attention to the near-horizon region where
the constant part of the warp factor is unimportant and the effective geometry changes.

482

B.R. Greene et al. / Nuclear Physics B 584 (2000) 480508

The compactification data as well as the undecoupled gravitional interactions are encoded
in a hypothetical Planck brane at the edge of the AdS throat.
In the second part of this paper we will address some aspects of the computation of
(the closed string part of) the low-energy spectrum for compactifications with background
flux. Generic string compactifications suffer from an abundance of massless fields. The
presence of background flux, however, will lift a number of these [27]. A priori, the number
of massless fields are determined by the moduli of the CalabiYau four-fold, subject to a
set of topological consistency conditions on the non-zero background flux. On the other
hand, a KaluzaKlein reduction of an arbitrary warped metric generates extra mass terms
due to the warp factors. Generically the massless spectrum is no longer determined purely
from the dimensions of spaces of harmonic forms on the internal space. However, the
powers of the warp factors that appear in this type of warped compactifications, which are
determined by supersymmetry, balance each other out, so one may indeed use topological
information to determine the massless spectrum. Moreover the non-trivial parts of the warp
factor are generated solely by the gravitational backreaction, and in the weak coupling limit
spacetime reduces back to a product form. Using this information we recover explicitly the
superpotential for the complex structure moduli conjectured in [30].
The organization of the paper is as follows. In Section 2, we review the results of M and
F theory compactifications on CalabiYau 4-folds in the presence of background fluxes. In
Section 3, we determine the warp factor for orbifold examples, in which the anomaly is
cancelled by a combination of background flux (from the untwisted and twisted sectors)
and M2/D3-branes. In Section 4, we comment on the shape of the graviton wavefunction
in this type of warped compactification. For completeness, we include in the appendix
a parallel discussion on the graviton wavefunction for Heterotic-M theory on a Calabi
Yau 3-fold. In Section 5, we present a general discussion on determining the number of
complex and Kahler moduli in vacua with background flux. For illustrative purpose, we
consider in Section 6 compactifications of M theory with background flux which satisfy
the field equations for an orbifold limit of K3 K3. The conditions on the background
flux are greatly simplified if we choose it to be a product of (1, 1) forms of each K3. For
this example we will determine the spectrum using the explicit topological data and we
will discuss how to solve more general cases. We end with some comments in Section 7.

2. M and F theory vacua with background flux


Warped compactifications of perturbative string theories, or rather supergravities,
preserving minimal supersymmetry in four dimensions were considered in [19,20]. The
supersymmetry requirements are highly restrictive; it was found that for type II theories
there are no non-trivial solutions if the four non-compact dimensions are Minkowski [20].
For type I/heterotic theory there is a solution if the internal space has nonzero torsion: i.e.,
the vacuum expectation value of the three-form NSNS field strength hH NS i is non-zero.
In this case the warp factor equals the dilaton, whose profile is determined by the vacuum
expectation value hH NS i.

B.R. Greene et al. / Nuclear Physics B 584 (2000) 480508

483

It has since been discovered that the M theory effective action has at the first subleading
order in the derivative expansion a topological term [32,33]
Z
(1)
C X8 (R),
where




1
1
4
2 2
tr R tr R
.
X8 =
8 4!
4

(2)

The inclusion of such higher derivative terms in the low energy supergravities allows for
new solutions preserving minimal supersymmetry, i.e. four supercharges, which are of the
warped kind [26,27]. (See also [2830].) Let us give a brief review.
On compactifications down to three dimensions, the eight-form X8 (R) can take on a
background expectation value. This will act as a source term for the three-form field C.
To maintain a solution to the field equations one needs to introduce M2-brane sources or
non-trivial G-flux; G is the four-form which locally equals dC. Either the M2-branes or
the non-trivial G-flux or both induce in turn a warping of the metric [26]

ds 2 = e(y) dx dx + e 2 (y)ga b dy a d y b .
1

(3)

The requirement that minimal supersymmetry (four supercharges, N = 2 in d = 3) is


preserved determines the relative weights of the warp-factors, as is known from p-brane
solutions [34]. In addition, supersymmetry demands that ga b is the metric of a CalabiYau
four-fold. The four-form G must furthermore obey
Gabcd = 0 = Gabcd ,

Ga =  a e 2 .
3

g cd Ga bc
d = 0,

(4)

The first two conditions mean that G is a (2,2) form on CY 4 which is self-dual or,
equivalently, primitive with respect to the Kahler form J of the CY 4 [27]
G = G

J G = 0.

(5)

Dirac quantization requires G to be an element of integer cohomology. The last equation


of (4) says that the only nonvanishing part of the three-form C is
3

C =  e 2 .

(6)

By definition the antisymmetric tensor is taken with respect to the unwarped metric. 3
The field equation for C determines the warp factor [26,34]
 X

n
1
3/2
) = CY 4 X8 G G
8 (y yj ).
(7)
2CY 4 (e
2
j =1

The factor G G has its origins in the well known C G G supergravity interaction and
we have introduced n M2-branes at arbitrary points. This number n is determined by the
3 The three-form thus equals the induced volume on the three-dimensional space C
=
p
Levi-Civita as is again familiar from p-branes. For the Dirac-delta function in
det |gMN zM zN | 
R
curved space, we use the convention that dx g(x) (x xi ) = 1.

484

B.R. Greene et al. / Nuclear Physics B 584 (2000) 480508

consistency requirement the absence of tadpoles or anomalies that the integral over
the right-hand side of (7) vanish,
Z
Z
1
CY 4
=n+
X8 =
G G.
(8)
24
2
CY 4

CY 4

Lift to F theory
Such compactifications can be lifted to F theory if the CY 4 is elliptically fibered. Fourdimensional Lorentz-invariance requires that the four-form G has one leg on the toroidal
fiber and three on the base B [27]. We will limit our attention to the case where the G-flux
is not localized in the fiber of the elliptic fibration, but is constant. In that case the M theory
solution can straightforwardly be lifted to F theory with [27]

ds 2 = e 4 dx 2 + e 4 gaBb dy a d y b ,
3

H NS = B ,
H RR = B ,
+
=  e 2 ,
D
3

(9)

R
R
where H (1,2)(B) with ( ) B = CY 4 G G Z+ . The metric and fourform potential are again that of the D3-brane solution. Formally the warp factor is still
given by (7), but one has to be careful with the dependence on the internal directions. We
will solve for the warp factor for a four-fold of the form K3 T 4 /Z2 where the fiber
belongs to T 4 /Z2 . In this case, the warp factor can equivalently be determined from the
IIB field equation for D + ,
d dD + =


1 X
tr(R R) 2 z1 zi1 H NS H RR
16
4

i=1

n
X


2 z1 zj1 4 ( j ).

(10)

j =1

Here, i labels the fixed points of T 2 /Z2 and j labels the n D3-branes. Tadpole cancellation
requires that this number equals
Z
K3K3
= 24.
(11)
H NS H RR + n =
24
B

Substituting the background expectation value for D + we find the F theory analogue of (7)

2B (e

3/2

) = B


1 X
tr(R R) 2 z1 zi1 H NS H RR
16
4

i=1

n
X
j =1


2 z1 zj1 4 ( j ),

(12)

B.R. Greene et al. / Nuclear Physics B 584 (2000) 480508

485

which determines the shape of the warp factor. In the orientifold limit, the first term of the
expression contains the contribution of the orientifold planes [37].
We will solve the warp-factor equations (7) and (12) for an orbifold limit of the Calabi
Yau. Specifically we will concentrate on the case K3 K3 = T 4 /Z2 T 4 /Z2 . This allows
us to take the F theory limit in the end, for which we will then attempt to compute the
massless four-dimensional spectrum.
Let us note from the outset that the warp factor, as a solution to (7), (12) is determined
only up to an integration constant whose value is fixed by the boundary conditions of the
warp factor. If the fluxes are sufficiently localized, the metric (3) should be approximately
flat away from such special points and the constant can be fixed to 1. We will discuss the
effect of a constant flux in the next section.
The dilaton
In the previously known warped compactifications of (heterotic) string theory, the warp
factor turned out to be equivalent to the dilaton [19,20]. Let us briefly recall why this is not
the case for the type IIB dilaton when we lift the above M theory vacua to F theory [27].
Weyl rescaling between the Einstein and the string frame can therefore alter the form of
the warp factors if a non-constant dilaton profile is consistent with the field equations.
By compactifying M theory in the eleventh and ninth-direction to IIA on S1 and
T-dualizing on the latter one finds that the IIB dilaton is given by a ratio of the
compactification radii [35,36]
exp (IIB ) =

R11
.
R9

(13)

Since in the background solution for an elliptically fibered CY 4 ,




2
(y) dw22 ,
ds s = e dx 2 + e/2 ga b dy a d y b + R92 (y) dw12 + R11

(14)

the warp factor is common to both R11 and R9 , the IIB dilaton is independent hereof. In
general, though, the dilaton will be a function of the internal dimensions y. In F theory,
the 7-branes are sources of the complex field = a + ie where a is the RR axion. For
example, near the location of a 7-brane at z = zi (z is the complex coordinate transverse to
the 7-branes),
1
log(z zi ).
2i
In the Einstein frame the metric is then warped to [21,22,39]
Y
Y
gz z = 2 2 2 (z zi )1/12 (z z i )1/12 .

(15)

(16)

In F theory, the D7-branes are not mutually local, but in the orientifold limit, which we will
consider, the charges of the D-branes are locally cancelled against the orientifold planes
(which for non-zero gs , are bound states of some (p, q) 7-branes). In that case the complex
field becomes a constant, and the metric (16) reduces to a flat one (except for conical

486

B.R. Greene et al. / Nuclear Physics B 584 (2000) 480508

singularities at the locations of the orientifold planes). In the general case where the 7brane charges are not locally cancelled, the calculation of the effective four-dimensional
Planck scale would involve an integral over a rather non-trivial function of the internal
manifold (due to (15), (16)). Another complication is that the metric for the 37 brane
system where the 7-brane charges are not locally cancelled has not yet been solved; for
progress in this direction see [40,41].

3. The warp factor


M theory
The warp factor is to be determined from the equation of motion (7) for the gauge field
C . To find its explicit form, we have to invert the Laplacian on the compact internal
space. This can be done with the help of Greens functions. On a compact space, or
equivalently when the Laplacian has non-trivial zero modes, the inversion is only defined
on the subset of functions orthogonal to these zero modes, i.e., on those functions not
belonging to the kernel of the Laplacian (see appendix for a brief review). For our purposes,
it is relevant to know that on a compact space the scalar Greens function obeys,
1
,
(17)
Vol
where d (x xi ) is the d-dimensional Dirac delta-function, and Vol is the volume of the
compact space.
On an orbifold, internal fluxes fall into two categories: those localized at the fixed points
and those wrapped over periods of the underlying torus. The equation (7) determining the
warp factor thus has two different kinds of contributions: the part of the flux term G G
which consists of a constant flux corresponding to untwisted cycles and that part which
is built from twisted cycles, proportional to Dirac-delta functions at the fixed points. The
constant fluxes are naturally expressed as an integer divided by the volume (after the Hodge
dual has been taken). In our case, the orbifold limit of K3 K3, we take the G-flux to be
constant on one of the K3s, so that we may take the F theory limit later. Then the flux
term on the r.h.s. of (7) can be written as
r1 X
r1 r2
+
mp 4 (x xp ),
(18)
(G G) =
V1 V2 V1 x
2G(x, xi ) = d (x xi )

where ri , Vi is, respectively, the total untwisted flux and volume on each K3 and mp is the
total twisted flux located at each fixed point xp of the second K3.
In the orbifold limit the curvature is also localized at the fixed points. On K3 K3 the
total contribution to the warp factor from the curvature is
Z
K3K3
= 24.
(19)
X8 =
24
The (T 4 /Z2 )2 limit has 162 fixed points. Each contributes of course equally to the
curvature and the first term in (7) may thus be written as

B.R. Greene et al. / Nuclear Physics B 584 (2000) 480508

X8 =

24 X 4
(x xp ) 4 (z zp ),
162 x ,z
p

487

(20)

where zp are the fixed points of the first K3.


Combining this information the equation (7) determining the warp factor reduces to

r1 r2
r1 X
3 X 4
(x xp ) 4 (z zp )

mp 4 (x xp )
2 e3/2 =
32 x ,z
2V1 V2 2V1 x
p

n
X

4 (x xi ) 4 (z zi )

i=1



1
3 X 4
(x xp ) 4 (z zp )
32 x ,z
V1 V2
p

n 
X

1
V1 V2
i=1


1
r1 X
mp 4 (x xp )
.

2V1 x
V2

4 (x xi ) 4 (z zi )

(21)

Here we have made use of the fact that tadpole cancellation implies that 48 = r1 r2 +
P
r1 mp + 2n.
The warp factor is now easily expressed in terms of Greens functions
 X (8)

3 X (8)
G x, (xp , zp )
G x, (xi , zi )
e3/2 = c0 +
32 x ,z
xi ,zi
p p
r1 X
mp G(4) (x, xp ).
(22)

2V1 x
p

Note that the completely constant flux term, proportional to r2 , does not contribute to the
warp factor in any obvious way. The Greens functions G(8) on K3 K3 and G(4) on a
single K3 can be constructed in the orbifold limit by the method of images. For example,
G(4) is given by
G(4) (x, xi ) =

E x xEi ) + e i p(E
E x +xEi )
X ei p(E
pE

pE 2

(23)

The momenta pE are quantized in units of the inverse radii of the T 4 . The second term is due
to the Z2 image. In the above sum, the zero momentum mode (p1 = p2 = p3 = p4 = 0) is
excluded.
The integration constant c0 is determined by the boundary conditions. We already
explained why, if all fluxes are localized (r2 = 0 in our example above) this constant is
fixed to unity. In that case far away from the points where flux (energy density) is localized,
spacetime should be approximately flat. In the case of constant G-flux over the whole
internal manifold, we can determine c0 from the curvature in the internal directions by
looking at the Einstein equation,

488

B.R. Greene et al. / Nuclear Physics B 584 (2000) 480508

RMN =



1
3
gMN G G ,
GM GN
2
D2

(24)

where
GMABC GN ABC
GMNRS GMNRS
,
GM GN =
.
(25)
4!
3!
We have ignored contributions to the stress-tensor from the curvature and the M2-branes
as they are localized. Substituting the background expectation values (and noting that G is
self-dual in the internal dimensions)
GG=

C =  e3/2 ,

e 6= 0
Ga bc
d G
a bcd

eG
e = r/VCY ,
G

(26)

directions, 4

one finds for the Ricci curvature in the internal




2
1
gmn
e3/2 e e
m ln e3/2 n ln e3/2
Rmn =
ln e3/2 + gmn
GG ,
2
3
6

(27)

with gmn the metric on the CY 4 . In the decompactification limit the explicit G-flux term
vanishes as it scales inversely with the volume. The warp factor does not depend explicitly
on the constant flux and we should therefore recover the regular p-brane metric for which
c0 = 1. (Strictly speaking we have only shown that limVCY c0 (V ) = 1.) The above
argument is rather general, and Ris not restricted
to orbifold cases where the curvatures are
R
localized. This is again because X8 and GG are quantized (to a finite number), and so
their contribution to the
R curvature is small when the size of the CalabiYau is large.
R internal
In other words, both X8 and G G, unlike the cosmological constant, are not extensive
quantities. By way of comparison, a non-zero c0 means that the decompactification limit
of the present scenario has significantly different properties than those studied in [11]. For
instance, one would not obtain a finite Planck scale in this limit.
An issue regarding the solution (22) is that due to the opposite sign of the Greens
functions corresponding to the curvature induced charge, the metric has a naked singularity
at the fixed points of the orbifold: the conical orbifold singularity. This is, however, an
artifact of the long-range supergravity approximation. The solution may roughly only be
trusted as long as the distance to any special point is larger than the Planck length or string
scale. The naked singularity is expected to be cured by stringy effects. Examples of this
have recently been discussed in [49,50]. Since these effects will only modify the metric
close to the singularity, they are unlikely to change the arguments above.
F theory
For F theory the solution is similar. We denote V6 as the volume of the 6 compactified
dimensions and V2 the volume of the two dimensions transverse to the 7-branes. The
twisted fluxes are proportional to 4 (x xp ) since we consider only fluxes that are not
localized around a singular fiber of the elliptic fibration: constant on one of the K3s. The
flux factor can thus be written as
4 To find this answer one needs to use that the field equations require that G
em G
en = (gmn /2)G
e G.
e This is the
combined constraint of primitivity and G  H (2,2) expressed in components.

B.R. Greene et al. / Nuclear Physics B 584 (2000) 480508

(H NS H RR ) =

489

r
1 X
+
mp 4 (x xp ).
V6 V2 x

(28)

The curvature contribution to the equation for the warp factor is due to the (unit charge)
D3-branes or O3-planes (charge O ). In the orbifold limit the curvature is completely
localized at the fixed points and thus gives rise to a six-dimensional delta-function. This
corresponds to the fact that we are considering the limit that the D7-brane charges are
locally cancelled and we are left with only O3-plane charge at the fixed points. Explicitly
3

2e 2 =

n
 X
O 6 (x xO ) H NS H RR
6 (x xi ).

xO

(29)

i=1

Substituting (28) we can rewrite this as


X
3
r
O 4 (x xO ) 2 (z zO )
2e 2 =
V6
x
O

n
X
1 X
mp 4 (x xp )
6 (x xi )
V2 x
i=1
p


X
1
O 4 (x xO ) 2 (z zO )
=
V6
xO


1
1 X
mp 4 (x xp )

V2 x
V4


n 
X
1
6 (x xi )
,

V6

(30)

i=1

where we have made use of the fact that anomaly cancellation implies r = n +
O ). In terms of Greens functions the warp factor is:
3

e 2 = c0

n
X
X
1 X
mp G4 (x, xp ) +
O G6 (x, xO )
G6 (x, xi ).
V2 x
x
p

xp (mp

(31)

i=1

For the more general case where the 7-brane charges are not locally cancelled, Eq. (10)
(which is written in the string frame), should receive contributions from O3- and O7-planes
and D3- and D7-branes plus an appropriate contribution from the dilaton which is no longer
constant. The net effect is that, in the Einstein frame, the warp factor is given by an equation
of the form
X
 X
ij
g g j e3/2 =
O 6 (x xO ) H NS H RR
6 (x xi ),
n

xO

(32)

i=1

where the metric gij takes into account the backreaction of the 7-branes according to (16).
The exact solution to this equation is not known, although it can be solved approximately
[40,41].

490

B.R. Greene et al. / Nuclear Physics B 584 (2000) 480508

4. The shape of gravity


With the warped metric (9) in F theory, the tree-level four-dimensional Planck scale is
given in terms of the 10-dimensional Planck scale by
Z
q
8
e2 e3/2 g B ,
(33)
M42 = M10
B

whereas the tree-level gauge coupling on a three-brane at y = yi has no warp factor


contribution and is given as usual by
1
2 (y )
gYM
i

= e2(yi ) .

(34)

As the non-trivial terms in the warp factor are solely due to backreaction, the usual relation
between the four-dimensional Newton constant and the three-brane gauge coupling still
holds at leading order.
At subleading order, the warp factor might come into play. However, an explicit
calculation shows that this is not the case. First, let us consider the case where the dilaton
is constant. Note that the power of the warp factor that appears in (33) is exactly equal to
the one in the field equation (12). Since the Greens function may be written as
G(y, yi ) =

X f (y)f (yi )
6=0

(35)

where f are orthonormal eigenfunctions of the Laplacian with eigenvalues , the fourdimensional Planck mass can be expressed as
8 2
e
M42 = M10
Z

Z
q
q
q
X 1 Z
6 0
6
B (y 0 )f (y 0 )
B f (y) . (36)

d 6 y gB
y
g
d
y
g
d

2
6=0

The quantity equals the charge density on the r.h.s. of (12),


0

(y ) = B

!
4
n
X
1 X
2 0
NS
RR
tr(R R) (y yi ) H H
6 (y 0 yj ). (37)

16
i=1

j =1

Because the zero mode is explicitly excluded in the Greens function, the second term in
(36) vanishes. This bears out our expectations. The warp factor does not change the relation
between the four- and ten-dimensional Planck scale. 5
In [11], an exponentially decaying wavefunction of the graviton was used to generate
the hierarchy between the electroweak scale and the four-dimensional Planck scale. The
exponential decay is due to the fact that the wavefunction of the massless graviton is
dressed by some powers of the warp factor. By way of comparison, let us also deduce
5 If the dilaton is not constant (taking into account the 7-branes), the net effect is the modification of the field
equation to (32) where the metric is expressed in the Einstein frame. The above argument based on the properties
of the Greens function should also hold in this case.

B.R. Greene et al. / Nuclear Physics B 584 (2000) 480508

491

the shape of the graviton wavefunction in the present setup by linearizing fluctuations
about the background warped metric in Section 3. We will see that although the massless
spectrum is indeed unaffected, the masses of the KK modes will get dressed by powers
of the warp factor. A few comments are in order. If the compactification is of the order
of string scale, we expect, a priori, that the wavefunctions of the KK modes are no longer
given by linearizing Einstein gravity since stringy effects may become important. However,
in compactifying Heterotic-M theory on a CalabiYau three-fold [8,9] (the G-flux in
this case is generically non-zero), there is a regime in which the theory is effectively
five-dimensional, the wavefunctions of some of the lower-lying KK modes (in the fifth
dimension) can still be obtained by linearizing Einstein gravity. The analysis for HeteroticM theory on CY 3 is very similar, so for completeness, we have included a derivation of the
KK spectrum in the appendix.
For a metric of the form
ds 2 = e2A(y) dx dx + e2B(y)gab dy a dy b ,

(38)

the gravitational fluctuations are given by + h where h is small compared


with .
From Einsteins equations, if we choose the gauge h = 0, the linear fluctuations
can be shown to satisfy the covariant wave equation (see the discussion in the next section
or, e.g., [42]):

1

M gg MN N h = 0,
g

(39)

where the indices M, N = 0, 1, . . . , 9 are raised and lowered with the warped background
metric (in Einstein frame). For the warped metric in M and F theory that we considered
(i.e., A(y) is proportional to B(y)), this reduces to

(40)
e2A(y)2Mink + 2g h = 0.
Expanding h (x, y) = (y)h (x), with 2x h (x) = m2 h (x), we have
2g (y) = m2 e2A(y)(y).

(41)

Hence the masses of the KK modes depend on the warp factor, which in turn depends on
the locations of the branes.
For the massless graviton (i.e., m2 = 0), (41) always admit the solution (y) = constant.
The wavefunction for the graviton, however, should be properly normalized:
Z

S d 10 x g h h +
Z
Z
p

(42)
= dy e2A(y) g(y) 2 (y) d 4 x h (x) h (x) + ,
where hatted indices are with respect to the unwarped metric. Therefore, the properly
normalized wavefunction is
p
1/2


B 1/4
(y) = e e3/4 det gab
,
(43)
(y) = e2A(y) g(y)

492

B.R. Greene et al. / Nuclear Physics B 584 (2000) 480508

which is the square root of the integrand in (33), as expected. For the orbifold examples
B ]1/4 are constant, with e 3/2 given
that we discuss in Section 3, the dilaton and [det gab
in terms of the Greens functions, hence
#1/2
"
n
X
X
1 X
4
6
6
mp G (x, xp ) +
O G (x, xO )
G (x, xi )
. (44)
(y) c0
V2 x
x
p

i=1

An alternative view of these warped compactifications was suggested in [16]. In


extending the AdS/CFT correspondence to the full string theory (i.e., without taking the
scaling limits as in [31]), the closed string degrees of freedom including gravity are no
longer decoupled from the effective theory on the worldvolume of the branes. To account
for the closed string degrees of freedom, one introduces into the AdS supergravity a
hypothetical Planck brane with dynamical degrees of freedom representing these closed
string modes. This hypothetical Planck brane is placed at the edge of the AdS throat
created by the branes and effectively cuts off the radial AdS coordinate. In the AdS/CFT
correspondence, distances from the brane correspond to energy scales in the worldvolume
theory. Therefore, the Planck brane serves as an UV cutoff and quantum gravity effects
become important as we get closer to the Planck brane.
In the case that the spacetime transverse to the branes is compactified, the information
about the compactification geometry would then be encoded in vacuum expectation values
of the excitations of the Planck brane. In principle one could derive this set-up by
integrating out coordinate shells of constant warp factor (momentum shells on the worldvolume) extending beyond the throat of the AdS near horizon region.

5. The low energy spectrum


The low energy spectrum of warped compactifications is naively different from that of
product space compactifications where one may use topological information of the internal
manifold to determine the massless spectrum. For example the equation of motion of a
scalar field,
1

M g g MN N (z) = 0,
g
when reduced on a generic warped metric of the form
 2A(y)

e
g (x)
0
gMN (z) =
0
gab (y)

(45)

(46)

yields

e2A 2g + 2g + d g ab a Ab (z) = 0.

(47)

Here d denotes the dimension of the internal manifold; D is the dimension of the ambient

one finds that the low energy modes descending


space. Redefining (z) = edA/2(z)
from this field are determined by

B.R. Greene et al. / Nuclear Physics B 584 (2000) 480508

493




dA
d
d2
= 0.
e2A 2g + e 2 2g 2g A (A)2 (z)
(48)
2
4
The last two terms act as additional mass terms for the dimensionally reduced field. The
massless modes are those in which the eigenvalue of the internal Laplacian cancels the
terms descending from the warp factor. Naively one thus loses the power of topological
arguments to determine the massless spectrum.
The particular warped solutions known in string theory/supergravity, the p-brane
metrics, belong to a special class, however. The internal manifold itself is also multiplied
by a warp factor which precisely compensates for the warp factor of the external space. For
example, in the scalar field above, the effect of an extra warp factor on the internal space,

 2A(y)
e
g (x)
0
,
(49)
gMN =
0
e2bA gab (y)
changes the field equation for the low-energy modes to


(50)
e2A 2g + 2g + d b(D d 2) g ab a Ab (z) = 0.
This internal warp factor can cancel the additional mass terms in (48) if b is chosen
appropriately: b = d/(D d 2). This is in fact exactly the combination one finds for
p-brane metrics in supergravity. Here, the physical reason is that the warp factor is solely
due to the backreaction of the vacuum configuration. Indeed one can explicitly see from
the solution that the non-constant terms in the warp factor are suppressed by powers of the
gravitational coupling constant. In the limit where this vanishes the warp factor is trivial
and the space is of the product form M4 CY 4 . This is another argument why also in the
case with background fluxes, the integration constant equals unity. The backreaction alters
the geometry only locally, but does not change the topology of the vacuum manifold. The
number of massless closed string sector fields (which correspond to marginal deformations
of the vacuum manifold) is still determined topologically. In fact, the balancing of
the warp factors is essentially the reason why when we first quantize open strings with
Dirichlet-boundary conditions, the gravitational backreaction of the D-branes does not
change the corresponding massless closed string spectrum.
The argument that in balanced warp metrics the warp factor is to a large extent
inconsequential as regards to the massless part of the low-energy spectrum holds
irrespective of whether G-flux is present or not. The introduction of the latter does
complicate the determination of the massless spectrum as in the presence of non-trivial
background expectation values of matter fields the mass matrix of low energy modes
is generically off-diagonal in terms of the original fields. Fortunately the fact that the
allowed background fluxes are subject to topological conditions G should be primitive
and a (2, 2) form in addition to the topological interpretation of the moduli of the
CalabiYau four-fold ought to allow one to also use topological methods to determine
the massless spectrum with G-flux [27]. In general these low-energy modes could be a
linear combination of the original fields, but their number can be determined by looking at
the topological constraints.
To be specific, those Kahler moduli which spoil the primitivity condition J G = 0
are lifted as well as those complex structure moduli which fail to keep G a (2, 2) class.

494

B.R. Greene et al. / Nuclear Physics B 584 (2000) 480508

In addition those Wilson lines which are not orthogonal to G: C G 6= 0 are lifted as
well [27]. In principle one has hereby computed the massless spectrum. In practice, the
first and the last constraint are readily solved but the counting of those complex structure
deformations which keep G a (2, 2) class is more involved. Consider a deformation of the
complex structure. This is given by a coordinate transformation which is not holomorphic
zi y i (z, z ).

(51)

Infinitesimally the mixed and the pure deformations of the metric under an arbitrary
coordinate transformation are
c
c
c
ga b = ga c b y c + gbc
a y + y c ga b + y c ga b ,
c
gab = ga c b y c + gbc a y c gc(a
b) .

(52)

This is just the well known fact that non-holomorphic coordinate changes correspond
g hc with the constant antito pure type metric deformations. By contracting g ca
ab
b
holomorphic (0, n) form one recovers the (1, n 1) forms that are in one-to-one
correspondence with complex structure deformations. Under such a transformation a (2, 2)
form transforms infinitesimally as

b
d
a
c
G Ga bc
d dz d z dz d z

G + K (3,1) + K (1,3)

(53)

with
 e b d e
a c
c a
K (1,3) = Ga bc
d e e e e dz d z d z d z .

(54)

For those complex structure deformations which keep G a (2, 2) class K (1,3) must vanish.
This constraint can be expressed as follows. The complex structure deformations ba are
e H (0,2)(2 T )
representatives of the cohomology H (0,1)(T ). Define the form G
eab = G ef ab .
G
cdef
cd

(55)

The constraint that K (3,1) must vanish can then be expressed by requiring that the triple
intersection numbers
Z

[a
ebc d]
G
(56)
abcd (0,1)
(0,2)
I (0,1) (4,0) = 0
vanish for all basis elements I of H (0,1)(T ). The latter are constructed from basis
elements I of H (3,1)(CY4 ) as in (55). Eq. (56) thus represents that the natural inner
product of K (1,3) with an arbitrary element I vanish. There is however no general analysis
of this condition if the (2, 2) form G is also required to be integral. One therefore has to
solve (56) on a case by case basis. Below we will do so for the simple example where the
CY 4 = K3 K3.
The deformations that do not respect the constraints should give rise to mass terms
when we perform a KK reduction. By our previous reasoning we can to first order ignore
the effects of the warp factor. Then, we are on a product space and the KK reduction is
straightforward, with the caveat that the field equations are now not satisfied. One finds

B.R. Greene et al. / Nuclear Physics B 584 (2000) 480508

495

that the kinetic term for C is responsible for mass terms for the pure and mixed type metric
fluctuations (numerical coefficients are suppressed)
Z
g e bc
a b
cd b c
m2 h2 hab Ga bc
d Gbc h + h Gabed Gb c h

dk e
dk e
a bc
a bc
+ Ga bc
+ Ga bc
.
d h hk Ge
d h hk Ge

(57)

In the first and the third term one recognizes the square of K (3,1) in (54). For massless
deformations it must vanish.
One may compare this with the conjectures made in [30]. There it was argued that the
complex structure deformations have a superpotential
Z
(58)
WC = G,
where is the holomorphic (4, 0) form. For the calculation of the mass terms involving
the complex structure deformations, the relevant parts of the first and second deformation
of the four-form are
()abcn abcd g d m hm n + ,

2 abn f abcd g cm hm n g d e hef + .
The non-trivial quadratic fluctuations in the superpotential are thus given by
Z
Z
n d
g hm hc abdn Gabcm ,
2 W = 2 G

(59)

(60)

where we have used the fact that the background value of G(2,2) is selfdual. Reducing to
N = 0 components and noting that the W 2 |=0 term does not contribute, we find the mass
term for the pure deformations
Z
(61)
m2 h2zz hdc abdn Gabcm Gnpq r rq sm hsp .
Using the relation
abcd a b cd abcd a b cd ,
the mass term reduces to
Z
m2 h2zz hsd Gq r nd Gnpq r hsp + hdq Gsr nd Gnpq r hsp ,

(62)

(63)

which is of the same form as the first and third mass terms from the Lagrangian in (57).
For the Kahler deformations the authors of [30] conjectured the superpotential
Z
(64)
WK = K K G,
where K is the complexified Kahler form K = J +iB. The reduction to N = 0 components
in this case is more involved. This will be reported elsewhere.

496

B.R. Greene et al. / Nuclear Physics B 584 (2000) 480508

6. An example: K3 K3
In this section, we will compute the spectrum for the simple example where the CY 4
equals K3 K3. We will first determine the number of complex and Kahler moduli by
explicitly constructing the complete parameter space of solutions to the G-flux constraints
on the orbifold limit of K3 K3. This allows us to simply count the number of moduli
by hand. At the end we will compare it to the topological determination of allowed
deformations outlined above.
Vanishing fluxes
Before proceeding let us briefly recall for comparison the spectrum of M theory on
a K3 K3 without G-flux. In this case there are h(3,1) + h(2,1) N = 1, d = 4 chiral
multiplets, corresponding to deformations of the complex structure and Wilson lines of the
three-form respectively, and h(1,1) N = 1, d = 4 vector fields, corresponding to Kahler
deformations. For CY 4 = K3 K3 the Hodge diamond equals
h0,0
h1,0
h2,0
h3,0
h4,0

h1,1
h2,1

h3,1

1
h0,1

h0,2
h1,2

h2,2

h0,3
h1,3

0
2
0

h0,4

0
40

0
40

404

(65)

2
0

0
40

The K3 K3 compactifaction is non-minimal in that it preserves four more supercharges


than the minimal four required by N = 1, d = 4 supersymmetry. Indeed we see that the
spectrum is given by 40 N = 2, d = 4 vector multiplets.
Taking one of K3s to be elliptically fibered we can lift this M theory compactification
to F theory on K3 K3, which is related to a number of other compactifications by a chain
of dualities [28,43,44]. Since F theory on K3 is heterotic on T 2 , compactifying both sides
on another K3 gives heterotic on K3 T 2 . On the other hand, heterotic on K3 is F theory
on a CalabiYau threefold CY 3 . As a result, F theory on K3 K3 is dual to F theory on
CY 3 T 2 .
The orientifold dual of the above F theory compactification can be found via Sens
map [45]. We take the orbifold limit K3 = T 4 /Z2 for each K3. Let (z1 , z2 , z3 , z4 ) be
the complex coordinates of T 8 , with z4 being the fiber coordinate. The orientifold dual is
given by type IIB on T 6 /{R1 ((1)FL R2 )} where is the worldsheet parity operation,
R1 and R2 act as follows:
R1 z1,2 = z1,2 ,
R2 z1,2 = z1,2 ,

R1 z3 = z3 ,

(66)

R2 z3 = z3 .

(67)

The resulting orientifold dual is therefore the GimonPolchinski model [4648] further
compactified on a T 2 and then T-dualized in the T 2 directions, so that instead of 5-branes
and 9-branes, we have 3-branes and 7-branes. At a special point of the moduli space where
is constant, i.e., four D7-branes are placed on top on each O7-plane, the gauge group
from the 7-branes is U (4)4 . In addition, there are two N = 2 hypermultiplets in the 6
representation of each U (4) from the 77 open strings. The 3-branes give rise to additional

B.R. Greene et al. / Nuclear Physics B 584 (2000) 480508

497

gauge symmetries. In the case of maximal gauge symmetries, i.e., the 3-branes are sitting
on top of each other, the gauge group from the 3-branes is U (16). The 33 open strings
also give rise to two N = 2 hypermultiplets in the 120 representations of U (16). Finally,
there is a hypermultiplet in the bi-fundamental representation (16, 4) of U (16) U (4) if
the 3-branes sit on one of the four groups of 7-branes. The orientifold model at a generic
point of the moduli space can be obtained from the above by Higgsing.
6.1. G-flux conditions
Now we turn on background fluxes. The field-equations demand that we seek an integer
(2, 2) form on K3 K3 = T 4 /Z2 T 4 /Z2 which is primitive, i.e.,
J G = 0.

(68)

We will choose G of the form


G = 1 2 ,

(69)

where i H (1,1)(T 4 /Z2 ) H (2)(T 4 /Z2 , Z). This guarantees that G  H (2,2)(T 4 /Z2 , Z)
though it is not necessary. This simplifies the solutions to the quantization conditions
(as compared to the general case [27]). We could have chosen i H (2,0)(T 4 /Z2 , Z) or
H (0,2)(T 4 /Z2 , Z) but these forms correspond to the class of the fiber and its Hodge dual
and are no longer normalizable when we shrink the volume of the fiber in the F theory limit
[27].
The complex structure is inherited from the tori and the condition (68) requires that each
i is primitive with respect to Ji [27]. Thus our task is reduced to finding primitive (1, 1)
forms on T 4 /Z2 . The cohomology H (1,1)(K3 = T 4 /Z2 ) has dimension 20, but these can
be subdivided in four untwisted (1, 1) forms inherited from the T 4 and sixteen twisted
ones. As any integer form is an integer combination of the basis forms we can consider the
two situations separately.
Constant fluxes
Consider untwisted forms first. These are inherited from the T 4 and are constant on the
orbifold. The intersection matrix of two-forms on K3 = T 4 /Z2 is block-diagonal with in
the upper left-hand corner minus the Cartan-matrix of E8 E8 and in the lower right-hand
corner three times the Pauli matrix 1 . The latter blocks equal the intersection-matrix of
two-forms on T 4 and thus correspond to the untwisted forms. Hence we only have to check
that the periods of the untwisted forms are integer over the untwisted cycles. This means
that we have reduced our problem to finding the set of integer (1, 1) forms D on T 4 .
For each T 2 (with volume normalized to 1), we make the standard choice for the periods
Z
Z
i
i
dz = j ,
dzi = i ji , no sum on i,
(70)
j

where xi and yi are the x and y cycles; i, j = 1, 2. The (1, 1) forms xi and yi dual to
these cycles are

498

B.R. Greene et al. / Nuclear Physics B 584 (2000) 480508

xi = dy i ,

yi = dx i ,

with the volume element


Z
dx i dy i = 1, no sum on i.

(71)

(72)

Ti

The Kahler form is given by


Ji = dzi d z i ,

(73)

and equals (i i ) times the volume form dx i dy i .


Take the Kahler form J for the T 4 to be the sum of J1 and J2 . The general form of (1, 1)
forms obeying the primitivity condition is then
J D = 0 D = A dz1 d z 2 + A d z 1 dz2 + iB(dz1 d z 1 dz2 d z 2 ),

(74)

where A is complex and B is real. The requirement that D H (1,1)(T 4 , Z)


H (1,1)(T 4 /Z2 , Z) demands that
A + A = n,
2 A + 2 A = m,
1 A + 1 A = p,

(75)

1 2 A + 1 2 A = q,
where n, m, p, q Z. In addition
iB(1 1 ) = v,
iB(2 2 ) = w,

(76)

where v, w Z. These are six equations with in principle three unknowns: Re(A), Im(A),
B; the system is overconstrained and has no solutions.
However, if one relaxes the requirement that all the i are free parameters, namely one
allows the complex structure 2 of one torus to be determined in terms of the other, and
in addition requires that the ratio of the imaginary parts of 1 and 2 be a rational number,
then there is a solution. This poses three real constraints on the moduli of the T 4 . The first
constraint can be seen by rewriting the integrality conditions as
A = n A,
m n2
p n1
=
,
A=
(1 1 ) (2 2 )
n1 p
1 1
=
,
2 2 m n2
1 2 A + 1 2 A = q.

(77)

Note, however, that we have imposed both the primitivity and integrality conditions in
deriving these constraints. Moreover, in (74) we are only considering (1, 1) classes. Hence,
the three constraints determine the loci in the combined space of complex and Kahler

B.R. Greene et al. / Nuclear Physics B 584 (2000) 480508

499

structures on K3 where one may find primitive integral forms which are purely (1, 1).
The number of constraints therefore correspond to the total number of moduli, complex
structure plus Kahler, which are lifted. Complex structure deformations must form complex
pairs. The total number of moduli which become massive is therefore 1 (complex valued)
complex structure deformation and 1 real Kahler deformation. For instance, one can see
this explicitly by generalizing the choice of the Kahler class to J1 + aJ2 and then noting
that the analogue of Eq. (76) fixes the value of a in terms of the complex structures 1
and 2 .
In the end we are interested in the H (2,2)(T 4 , Z) flux
Z

D D = 2 |A|2 + B 2 (1 1 )(2 2 ),
(78)
T4

which should be an integer. This is proportional to the number of M2/D3-branes we will


have to introduce. Substituting (76) and the second equation of (77) in (78) we find that
Z
D D = 2vw 2(p n1 )(m n2 ).
(79)
T4

Suppose v or w vanishes; as the last factor should be a negative semi-definite integer


(i H + in (78)), let us make the Ansatz that


r
r
m
2 =

(80)
m n2 =
(p n1 )
n
n(p n1 )
is indeed a solution to (77). Here r Z+ /2. Substituting this relation into the four equations
(77) we find that our Ansatz is a solution with r = qn + mp. Note that mp > qn. Hence
the flux equals
Z
D D = 2vw 2(p n1 )(m n2 ) = 2(qn mp vw).
(81)
T4

Note that if D is also an element of H (1,0)(T 2 , Z) H (0,1)(T 2 , Z) H (1,1)(T 4 , Z), i.e.,


if
R n = n1 n2 ; m = n1 m2 ; p = m1 n2 and q = m1 m2 for integer n1 , n2 , m1 , m2 then the flux
T 4 D D = 2(m1 m2 n1 n2 n1 m2 m1 n2 ) vanishes, if B = 0.
Twisted fluxes
We will restrict ourselves to twisted forms whose dual cycles are localized at the fixed
points. These are two-spheres S 2 ' CP1 shrunk to a point. Denoting them as B i with
i = 1, . . . , 16 running over the fixed points, the general form of a twisted flux D is
D = ci B i .

(82)

are a linear combination of the generating forms


of
Bi =
The
ani V n with ani integer. As we are limiting our attention to twisted fluxes we may restrict the
V n to those generating minus the E8 E8 Cartan matrix I nm . This also means that D is
automatically primitive as the Kahler form J consists purely of untwisted forms.
Bi

Vn

H (1,1)(T 4 /Z2 , Z):

500

B.R. Greene et al. / Nuclear Physics B 584 (2000) 480508

The condition that also D H (1,1)(T 4 /Z2 , Z) means that


Z
i mn
D = ci am
I = pn ,

(83)

Vn

with pn Z for all n. The cycles B i have an intersection matrix [19]


Z
j
B i B j = 2 ij = ani I nm am .

(84)

This means that the coefficients ci are all multiples of 1/2,


ci =

pn ani
.
2

(85)

R
The four-form flux of interest to us T 4 /Z2 D D equals
Z
X
1 m
D D = 2
ci ci = pn Inm
p ,
T 4 /Z2

as one can show that


X
i
1
ani am
= 2Inm
.

(86)

(87)

The inverse of the Cartan matrix has integer entries because the E8 lattice is even and selfdual, and the diagonal entries are even. The r.h.s. of (86) is therefore an integer as expected.
R
Note again that D D is negative semidefinite just as in the case of constant fluxes.
Note that the Bi or more precisely arbitrary half-integer combinations thereof do not
generate H (1,1)(T 4 /Z2 , Z) [19], but a larger group. For integer cohomology the ci are
subject to the additional constraint
X
i i
am
c = pn I vnm .
(88)
i

We are therefore not missing any integer forms by limiting our attention to fluxes of the
form (82).
In this case those Kahler deformations J = n V n where V n H (1,1)(T 4 /Z2 , Z) are
twisted fluxes that do not preserve the primitivity condition are frozen. Those deformations
such that
Z
i
= 0, n
(89)
J D = n I nm ci am
do survive. The integrated primitivity condition is sufficient to guarantee the local one [19].
R
In fact we can determine the contribution to D D of each fixed point separately. As
the intersection matrix of the cycles B i is proportional to the unit matrix, the contribution
of the ith cycle is just given by the ith term in Eq. (86) [19]. Since the orientifold planes
are also localized at the fixed points twisted fluxes change the effective O-plane charge.

B.R. Greene et al. / Nuclear Physics B 584 (2000) 480508

501

6.2. The spectrum


In the previous section, we explicitly constructed the parameter space of solutions for
pure (1, 1) forms on K3 that are primitive and integral. For constant flux this showed
that the dimension of the space of complex structures and the dimension of the space of
Kahler forms are each reduced by one. In this section we will compare these results with a
topological deformation analysis.
Suppose one is given a complex structure and Kahler form on K3 for which it is possible
to find a primitive integral (1, 1) flux i . The i H (1,1)(K3) H 2 (K3, Z) define a single
direction in H (1,1)(K3) H 2 (K3, Z). The Kahler deformations which are preserved are
those which are orthogonal to i in the sense that
Z
(90)
J i = 0.
Thus there is always exactly one Kahler deformation in each K3 that is lifted.
As for the complex structure deformations, on a K3 they will cause a (1, 1) form to
become a mixture of a (1, 1) and a (0, 2) + (2, 0) form; the latter being complex conjugates.
Since H (0,2) is 1-dimensional, a 19-dimensional subspace of (1, 1) forms is preserved.
These allowed deformations will not spoil integrality, but could spoil primitivity. One can,
however, always find a compensating Kahler deformation which restores primitivity. This
is illustrated in the constant flux example by our earlier explicit calculation.
Since on K3 K3, the (2, 2) classes come from (1, 1) (1, 1) and (2, 0) (0, 2), we
find that there is a (38 = 19 + 19)-dimensional subspace of preserved complex structure
deformations on K3 K3.
An alternative way to see this embodies the symmetry between complex structure
deformations and Kahler deformations. Consider in analogy with the superpotential (58)
of [30] the intersection
Z
(91)
i i
for each K3. i is now the holomorphic (2,0) form under a complex structure deformation
this potential changes to
Z
(92)
i i .
We see that that deformation which is parallel to the i flux is lifted.
These constraints thus change the number of chiral and vector multiplets by one in each
K3. As a result, a total of two N = 2, d = 4 vector multiplets are lifted due to the G-flux.
The spectrum (in addition to that from the branes) thus consists of 38 N = 2, d = 4 U (1)
vector multiplets, coupled to gravity.

7. Discussion
In this paper, we have taken some modest steps towards understanding warped
compactifications in M and F theory on CalabiYau four-folds in the presence of

502

B.R. Greene et al. / Nuclear Physics B 584 (2000) 480508

non-trivial background flux. The introduction of background flux can have interesting
consequences for low energy physics and this subject is certainly worthy of further study.
Detailed investigations require the explicit form of the warp factor, and to facilitate its
determination, we considered orbifold limits of these compactifications. The contribution
of the background flux is a backreaction effect similar to explicit p-brane sources and
because the energy density associated with the G-flux is inversely proportional to the
volume, the leading term in the warp factor is always a constant. This contrasts sharply
to what has been found in a number of other warped scenarios. In the orbifold limit the
background fluxes are either constant or localized. The constant fluxes allow the warp
factor to be consistently solved in terms of Greens functions on the internal space. The
twisted fluxes act as sources for the Greens functions.
The extensive nature of the background flux suggests that its introduction will not have
drastic consequences; indeed we find that the usual relation between the four- and tendimensional Planck scales is recovered. This, again contrasts with the scenarios involving
a cosmological constant [11], in which gravity is localized. The shape of the graviton
wavefunction is sharply peaked at the location of the branes but this is just the source
divergence of a Greens function. In a similar analysis for Heterotic-M theory on a Calabi
Yau three-fold, discussed in an appendix, we also derived the wavefunctions of the KK
modes. In our M/F theory setup, the higher KK modes can also be obtained by linearizing
gravity if the compactification size is large. Interestingly, aside from the compactification
size, at subleading order the masses of the KK modes appear to depend on the location of
the branes.
In the view of [16], in which the renormalization flow of the effective worldvolume
theory is correlated with a translation in one of the real internal directions, the
compactification geometry is encoded by a hypothetical Planck brane at the throat of the
near-horizon AdS created by the branes. The shape of the warp factor should determine the
characteristics of this Planck brane.
Finally we discussed the low energy spectrum of these warped compactifications.
Supersymmetry or the argument that the warp factor is solely due to backreaction
guarantees that topological arguments may still be used to determine the massless
spectrum. The presence of background fluxes can lift a number of moduli, but the rules
to determine them are of (pseudo) topological nature as well. For K3 K3 we found that
two N = 2 vector multiplets were lifted. In addition to the complex and Kahler moduli
that arise in these compactifications, the low energy degrees of freedom include gauge
and matter fields on the branes. For example, in the K3 K3 case, there are generically
3 and 7 branes on which gauge and matter fields are localized. We recalled in Section 6
the spectrum from the branes when the background flux is zero. The determination of the
spectrum from the brane sector of F theory compactifications in the presence of background
three-form NSNS and RR fluxes is an interesting direction to pursue in the future.

B.R. Greene et al. / Nuclear Physics B 584 (2000) 480508

503

Acknowledgements
We would like to thank Jan de Boer, Claude LeBrun, Lennaert Huiszoon, Zurab Kakushadze, David Morrison, Horatiu Nastase, Peter van Nieuwenhuizen and Savdeep Sethi
for discussions and correspondence. The research of G.S. is partially supported by the
NSF grant PHY-97-22101, and that of B.R.G. is partially supported by the DOE grant
DE-FG02-92ER40699B. K.S. acknowledges the hospitality of the C.N. Yang ITP and the
Spinoza Institute.

Appendix A. Greens functions on a compact space


On a compact space, or equivalently when the Laplacian has non-trivial zero modes, the
inversion of the Laplacian is only defined on the subset of functions orthogonal to these
zero modes. One can see this explicitly by constructing the Greens function in terms of
the complete set of eigenfunctions of the Laplacian. These eigenfunctions must obey the
same boundary conditions as the function on which the Laplacian acts. Recalling that the
Laplacian has non-positive definite eigenvalues we can denote an orthonormal set by m (x)
with
2m (x) = m2 m (x).

(A.1)

The Greens function is then


X m (x)m (x 0 )
.
G(x, x 0 ) =
m2

(A.2)

m6=0

All the zero modes of the Laplacian must be omitted in the sum on the r.h.s. This is a
generic feature of all Greens functions. When one writes
2G(x, x 0 ) = 1 = D (x x 0 ),

(A.3)

it is implicitly understood that the -function or identity-operator is only defined in the


space of functions orthogonal to the zero modes of the Laplacian. This is not necessarily
the same as the Dirac-delta function. On a circle of radius R for instance, there is exactly
one non-trivial zero mode, the constant function. The Greens function equals
in(xx 0 )

1 Xe R
G(x, x ) =
2R
n2 /R 2
0

(A.4)

n6=0

and therefore obeys 6


2G(x, x 0 ) = x2 G(x, x 0 ) =

1 X in(xx 0 )
1
,
e R = (x x 0 )
2R
2R

(A.5)

n6=0

rather than
6 Recall that we use the convention that the Dirac-delta function in curved space obeys

(x xi ) = f (xi ).

dx g(x)f (x)

504

B.R. Greene et al. / Nuclear Physics B 584 (2000) 480508

2G(x, x 0 ) = (x x 0 ).

(A.6)

The former combination (x x 0 ) (1/2R) is obviously the delta-function in the space


of functions orthogonal to the zero mode. In the infinite volume limit we are left with the
conventional Dirac-delta function.
The above is a reflection of the fact that for differential forms, the Greens function can
be used as a projector onto the space of non-harmonic forms, see, e.g., [38]. An arbitrary
p-form can be decomposed into an harmonic p-form and a non-harmonic part as
= + 2G.

(A.7)

The combination 2G is equivalent to the projector


Z
X
n
R
n .
2G = 1
( n n )
n

(A.8)

Here the n form a basis of harmonic p-forms. For scalars on a compact manifold there is
only one harmonic form, the constant, and for 0-forms the above combination reduces to
2G(x, x 0 ) = D (x x 0 )

1
.
Vol

(A.9)

For one dimension this equals the r.h.s. of (A.5)


Particularly in the case of tori, one occasionally imposes the boundary conditions
signifying the compactness of the space by introducing image charges on the covering
space. The torus is considered as the quotient Rd /d of Rd by the lattice d . The Greens
function on Rd /d is then a sum of regular Rd Greens functions obeying
D
(x x 0 )
2GRd (x, x 0 ) = Dirac

such that GRd /d and its derivative is periodic, i.e.,


X
GRd (x, x 0 + nE eE ).
GRd /d (x, x 0 ) =

(A.10)

(A.11)

nE

Here eE are the lattice vectors generating d .


The Laplacian acting on GRd /d (x, x 0 ) now gives
X
D
Dirac
(x x 0 + nE eE).
2GRd /d (x, x 0 ) =

(A.12)

nE

The only relevant part of the r.h.s, however, is the term with n = 0 as both x, x 0 belong to
a single fundamental region x i [0, ei ). The delta functions for other values of nE never
contribute.
If one chooses to solve for the warp factor using such a Greens function, one would
ignore the contribution from any constant terms on the r.h.s. This is evident from the
preceding discussion. However, a consistency condition is required, namely, the total
integral on the r.h.s. of the warp factor equation of motion must vanish. This just says
that the total flux on the compact space is zero or that the anomaly cancels.

B.R. Greene et al. / Nuclear Physics B 584 (2000) 480508

505

Appendix B. Heterotic-M theory on CY 3


The effective five-dimensional theory by compactifying Heterotic-M theory on a Calabi
Yau 3-fold was derived in [8,9]. Here, we will follow their discussion and notation. We will
consider the simplest case, in which there are no 5-branes, and the low energy degrees of
freedom include only gravity and the hypermultiplet containing the breathing mode V of
the CalabiYau (the universal solution in [8,9]). With the standard embedding of the spin
connection in one of the E8 s, the effective action becomes [8,9]


Z

1 2 55 0 2 1 2 2
2
g R + V g V + V
25 S5 =
2
3
M5
Z
Z

g V 1 2 2
g V 1 ,
(B.1)
+2 2
M4(1)

M4(2)

where 5 is the five-dimensional gravitational coupling. The constant is given by


 2/3 Z
Z

tr R R, v =
gCY3 ,
(B.2)
=
8 2 v 4
CY 3

CY 3

where is the Kahler form on the CY 3 . The metric is warped to


ds 2 = a 2 (y) dx dx + b2 (y) dy 2.

(B.3)

The solution to the field equations is:


a = a0 H 1/2,
b = b0 H ,
2

2
|y| + c0 c|y| + c0 ,
H=
3

V = b0 H 3 ,
where a0 , b0 and c0 are integration constants.
The Einstein equation linearized in the fluctuations give:


 0


b0
1
c
a
1 00
0
4 b h 3cb0 H 3
(y) (y ) h
h
2
2
a
r
H
 
1 b0 2 3
=
H h, .
2 a0

(B.4)

(B.5)

Let us first consider the equation in the bulk. It is easy to see that the above equation
can be written as the covariant wave Eq. (39). As in Section 4, we expand h (x, y) =
h (x)(y) with 2x h (x) = m2 h (x). The properly normalized wavefunction is
p
1/2

(y) = ab1/2(y).
(B.6)
(y) = e2A(y) g(y)
For massless graviton,
(y) ab 1/2 H 3/2.
Therefore, the wavefunction (y) vanishes at the singularity where H (y) is zero.

(B.7)

506

B.R. Greene et al. / Nuclear Physics B 584 (2000) 480508

The compactification scale of the M theory direction is usually taken to be slightly larger
than that of the CalabiYau (from gauge and gravitational unification [2,3]). Therefore,
there is a regime in which the theory is five-dimensional, and the wavefunction of some of
the low-lying massive KK gravitons can still be described by the above wave equation



c
1 00
(y) (y ) (y)
(y) 3cb0H 3
2
H
 2
b0
1
H 3 (y).
(B.8)
= m2
2
a0
Let us rewrite the above equation such that m2 becomes the eigenvalue of a
Schrdinger equation. Define a new variable u = (2/5)c1 (b0 /a0 )H 5/2 and hence du =
(b0 /a0 )H 3/2dy. In terms of this variable, the linearized equation becomes

 


1
1
3c a0
1
(u

(u

u
)


H 5/2 3a0 c
1
2
2
4 b0
H 1/2 b0 H 5/2
1
(B.9)
= m2 ,
2
where the dot denotes the derivative with respect to u. The locations u1 and u2 are defined
by u = ui when y = 0 and , respectively.
It
Finally, to eliminate the first derivative term in the above equation. Define = H s .
is easy to see that first order derivatives of do not appear if s = 3/4. The function
satisfies the following equation

1
1 d2

(u) + V(u)(u)
= m2 (u),
2
2 du
2

(B.10)

where

 

 


a0
15
c
21 a0 2 c2
2
(u u1 ) (u u2 )

3b0H
V(u) =
32 b0 H 5
b0 H 5/2
4
 



2
5c a0 4/5 15
21

3b

(u u1 ) (u u2 ) .
=
0
2
200u
5u
2 b0
4

The potential is not of the volcano type [11]. In particular, we have seen that gravity is
not localized.
We have seen that the wavefunction of the massless graviton can be deduced without
having to solve (B.10) since it is clear from (41) that = constant is a solution. The
massive modes can be found by solving (41), or equivalently (B.10). Let us focus on
the latter, as it allows us to compare our solution with that of [11]. Let = u1/2 f , the
Schrdinger equation can be written as


2
f
1
2 f
2
+ v
+v
f (v) = 0,
(B.11)
v
v 2
v
25
where v = m(u + u0 ). The solution is simply


f = k1 J1/5 m(u + u0 ) + k2 J1/5 m(u + u0 ) ,

(B.12)

B.R. Greene et al. / Nuclear Physics B 584 (2000) 480508

507

where k1 , k2 are constants which can be determined by the matching conditions at the
delta function sources. The boundary conditions at the two delta sources also imply that
the masses m of the KK modes are quantized in units of 1/.
References
[1]
[2]
[3]
[4]
[5]
[6]
[7]
[8]
[9]
[10]
[11]
[12]
[13]
[14]
[15]
[16]
[17]
[18]
[19]
[20]
[21]
[22]
[23]
[24]
[25]
[26]
[27]
[28]
[29]
[30]
[31]
[32]
[33]
[34]
[35]
[36]
[37]
[38]
[39]
[40]

V.A. Rubakov, M.E. Shaposhnikov, Phys. Lett. B 125 (1985) 372; Phys. Lett. B 125 (1983) 139.
P. Horava, E. Witten, Nucl. Phys. B 460 (1996) 506; Nucl. Phys. B 475 (1996) 94.
E. Witten, Nucl. Phys. B 471 (1996) 135.
N. Arkani-Hamed, S. Dimopoulos, G. Dvali, Phys. Lett. B 429 (1998) 263.
I. Antoniadis, N. Arkani-Hamed, S. Dimopoulos, G. Dvali, Phys. Lett. B 436 (1998) 257.
G. Shiu, S.-H.H. Tye, Phys. Rev. D 58 (1998) 106007.
Z. Kakushadze, S.-H.H. Tye, Nucl. Phys. B 548 (1999) 180.
A. Lukas, B.A. Ovrut, K.S. Stelle, D. Waldram, Phys. Rev. D 59 (1999) 086001.
B.A. Ovrut, hep-th/9905115 and references therein.
K.R. Dienes, E. Dudas, T. Gherghetta, Phys. Lett. B 436 (1998) 55; Nucl. Phys. B 537 (1999)
47, hep-ph/9807522.
L. Randall, R. Sundrum, Phys. Rev. Lett. 83 (1999) 3370; Phys. Rev. Lett. 83 (1999) 4690.
K. Behrndt, M. Cvetic, hep-th/9909058, hep-th/0001159.
M. Cvetic, H. Lu, C.N. Pope, hep-th/0001002.
R. Kallosh, A. Linde, hep-th/0001071.
M. Wijnholt, S. Zhukov, hep-th/9912002.
H. Verlinde, hep-th/9906182.
P. van Nieuwenhuizen, in: B. de Wit, R. Stora (Eds.), Relativity, groups and topology II, NorthHolland, Amsterdam, 1983.
P. van Nieuwenhuizen, N. Warner, Comm. Math. Phys. 99 (1985) 141.
A. Strominger, Nucl. Phys. B 274 (1986) 253.
B. de Wit, D.J. Smit, N.D. Hari Dass, Nucl. Phys. B 283 (1987) 165.
C. Vafa, Nucl. Phys. B 469 (1996) 403.
D.R. Morrison, C. Vafa, Nucl. Phys. B 473 (1996) 74; Nucl. Phys. B 476 (1996) 437.
R. Friedman, J. Morgan, E. Witten, Comm. Math. Phys. 187 (1997) 679.
M. Bershadsky, A. Johansen, T. Pantev, V. Sadov, Nucl. Phys. B 505 (1997) 165.
M. Bershadsky, T.-M. Chiang, B.R. Greene, A. Johansen, C.I. Lazaroiu, Nucl. Phys. B 527
(1998) 531.
K. Becker, M. Becker, Nucl. Phys. B 477 (1996) 155, hep-th/9605053.
K. Dasgupta, G. Rajesh, S. Sethi, JHEP 9908 (1999) 023, hep-th/9908088.
S. Sethi, C. Vafa, E. Witten, Nucl. Phys. B 480 (1996) 213.
E. Witten, Nucl. Phys. B 471 (1996) 195, hep-th/9603150.
S. Gukov, C. Vafa, E. Witten, hep-th/9906070.
J. Maldacena, Adv. Theor. Math. Phys. 2 (1998) 231.
C. Vafa, E. Witten, Nucl. Phys. B 447 (1995) 261, hep-th/9505053.
M.J. Duff, J.T. Liu, R. Minasian, Nucl. Phys. B 452 (1995) 261, hep-th/9506126.
M.J. Duff, R.R. Khuri, J.X. Lu, Phys. Rept. 259 (1995) 213, hep-th/9412184.
E. Witten, Nucl. Phys. B 443 (1995) 85, hep-th/9503124.
P.S. Aspinwall, Nucl. Phys. Proc. Suppl. 46 (1996) 30, hep-th/9508154.
K. Dasgupta, D.P. Jatkar, S. Mukhi, Nucl. Phys. B 523 (1998) 465.
T. Eguchi, P.B. Gilkey, A.J. Hanson, Phys. Rept. 66 (1980) 213.
B.R. Greene, A. Shapere, C. Vafa, S.T. Yau, Nucl. Phys. B 337 (1990) 1.
O. Aharony, A. Fayyazuddin, J. Maldacena, JHEP 9807 (1998) 013.

508

[41]
[42]
[43]
[44]
[45]
[46]
[47]
[48]
[49]
[50]
[51]

B.R. Greene et al. / Nuclear Physics B 584 (2000) 480508

A. Kehagias, Phys. Lett. B 435 (1998) 337.


C. Csaki, J. Erlich, T. Hollowood, Y. Shirman, hep-th/0001033.
M. Bershadsky, V. Sadov, Nucl. Phys. B 510 (1998) 232.
Z. Kakushadze, G. Shiu, S.-H.H. Tye, Nucl. Phys. B 533 (1998) 25.
A. Sen, Phys. Rev. D 55 (1997) 7345, hep-th/9709159.
E.G. Gimon, J. Polchinski, Phys. Rev. D 54 (1996) 1667.
G. Pradisi, A. Sagnotti, Phys. Lett. B 216 (1989) 59.
M. Bianchi, A. Sagnotti, Phys. Lett. B 247 (1990) 517.
C.V. Johnson, J. Polchinski, A. Peet, Phys. Rev. D 61 (2000) 086001.
L. Jarv, C.V. Johnson, hep-th/0002244.
C.S. Chan, P.L. Paul, H. Verlinde, hep-th/0003236.

Nuclear Physics B 584 (2000) 511527


www.elsevier.nl/locate/npe

Locality bound for effective four-dimensional


action of domain-wall fermion
Yoshio Kikukawa
Department of Physics, Nagoya University, Nagoya 464-8602, Japan
Received 25 January 2000; revised 10 May 2000; accepted 16 May 2000

Abstract
We discuss locality in the domain-wall QCD through the effective four-dimensional Dirac operator
which is defined by the transfer matrix of the five-dimensional Wilson fermion. We first derive an
integral representation for the effective operator, using the inverse five-dimensional WilsonDirac
operator with the anti-periodic boundary condition in the fifth direction. Exponential bounds are
obtained from it for gauge fields with small lattice field strength. 2000 Elsevier Science B.V. All
rights reserved.

1. Introduction
Locality properties of Neubergers lattice Dirac operator [1], which is derived from the
overlap formalism [2] and satisfies the GinspargWilson relation [3], has been examined
by Hernndes, Jansen and Lscher [4]. For a certain class of gauge fields with small lattice
field strength, exponential bounds have been proved rigorously on the kernels of the Dirac
operator and its differentiations with respect to the gauge field. These properties assure that
the index theorem holds true on the lattice [2,5]. The index theorem implies the topological
properties of chiral anomalies. It plays the crucial role in the recent construction of lattice
chiral gauge theories [611]. Numerical studies of the locality of Neubergers lattice Dirac
operator are also found in [4,12].
The purpose of this paper is to argue the locality of the low energy effective action of the
domain-wall fermion [1329]. It has been known that the partition function of the domainwall fermion, in the anti-periodic subtraction scheme [18], reduces to a single determinant
of an effective four-dimensional Dirac operator [30],


1
N e
(N)
1 + 5 tanh a5 H
=
.
(1.1)
Deff
2a
2
kikukawa@eken.phys.nagoya-u.ac.jp

0550-3213/00/$ see front matter 2000 Elsevier Science B.V. All rights reserved.
PII: S 0 5 5 0 - 3 2 1 3 ( 0 0 ) 0 0 2 9 7 - 2

512

Y. Kikukawa / Nuclear Physics B 584 (2000) 511527

Here N and a5 denote the lattice size and the lattice spacing of the fifth dimension,
e is defined through the transfer matrix of the five-dimensional Wilson
respectively. H
fermion

1
1
C
B
e
B
,
(1.2)
T = ea5 H =

1
1
B +C C
C
B
B
where 1

1
+ ,
2

m0
a
.
B = 1 + a5
2
a
C = a5

(1.3)
(1.4)

e2 > 0. The effective Dirac operator Eq. (1.1)


The limit N is defined well as long as H
e,
then reduces to Neubergers lattice Dirac operator using H


e
1
H
1 + 5 p
,
(1.5)
Deff =
2a
e2
H
and turns out to satisfy the GinspargWilson relation.
Moreover, the propagator of the light fermion field which is introduced by Furman and
Shamir [15],
q(x) = L (x, 1) + R (x, N),

q(x)

= L (x, 1) + R (x, N),

can be expressed in terms of the effective Dirac operator [31]:




a5 1 (N) 1
Deff
(x, y) .
hq(x)q(y)i

= 4
a a
The anomalous term in the axial WardTakahashi identity
 


 
 

2
N
N
N
N
(N)

R x,
L x, + 1
L x, + 1 R x,
X (x) =
a5
2
2
2
2
can also be expressed with it:

(N) 
a 4 X(N) (x) = tr 5 2 1 aDeff
(x, x)


1
1 (N) 1
5
tr Deff
(x, x).
e
a
cosh2 N2 a5 H

(1.6)

(1.7)

(1.8)

(1.9)

In the limit N , this reduces to the chiral anomaly associated with the exact chiral
symmetry [3238],
a 4 hX(x)i = tr 5 2(1 aDeff )(x, x).

(1.10)

It would have the topological properties, if the effective Dirac operator is local and depends
smoothly on the gauge fields.
1 In this expression, the positivity of B is required for the transfer matrix to be defined consistently. It is assured
when 0 < (a5 /a)m0 < 1 . It is also assumed that N is even.

Y. Kikukawa / Nuclear Physics B 584 (2000) 511527

513

In view of this direct relation, 2 it seems reasonable to argue locality in the domain-wall
fermion approach through the locality properties of the effective Dirac operators Eq. (1.1)
and Eq. (1.5). It is expected that a similar exponential bound could be established under
certain conditions, like the result obtained by Hernndez, Jansen and Lscher [4]. In our
e. Then the
case, though, the hermitian WilsonDirac operator should be replaced with H
use of the Legendre polynomials [4] does not lead to the expansion in terms of operators
with finite ranges.
For our purpose, we first derive an integral representation for the effective Dirac operator.
The inverse square root in Neubergers lattice Dirac operator can be written by the integral:


Z
1
H
dp
1
1
1 + 5
= + 5
2
2
2
2
i
p
+
D
H
5
w

m0  5 .
a

(1.11)

Corresponding to this, we can show that the effective Dirac operator admits the following
representation [31]:

1

1
(N)
= 1 PR a5 D 5w NN PL PL a5 D 5w 11 PR
aDeff

1

1
PR a5 D 5w N1 PR PL a5 D 5w 1N PL ,
(1.12)
where D 5w is the five-dimensional WilsonDirac operator with the anti-periodic boundary
condition in the fifth dimension. Its inverse may be expressed as


a5 D 5w

1
st

1 X
eip(st )
N p i5 sin p + 1 cos p + a5 Dw

m0  .
a

(1.13)

1
The summation is taken over the discrete momenta p = 2
N (k 2 ) (k = 1, 2, . . . , N) and
in the limit N it reduces to the continuous integral.
From this representation, it is rather clear that the effective Dirac operator can be defined
consistently if the five-dimensional WilsonDirac operator with the anti-periodic boundary
condition is not singular and invertible for all N . In this respect, we should note that the
lower bound on the square of the five-dimensional WilsonDirac operator is related closely
to that on the square of the four-dimensional WilsonDirac operator [4,42], because the
gauge field is four-dimensional. In fact, the same lower bound can be set for the class of
gauge fields with small lattice field strength.
Given the positive lower bound on the square of the five-dimensional WilsonDirac
operator, it is possible to formulate a series expansion in terms of the five-dimensional
WilsonDirac operator, using the generating function of the Chebycheff polynomials [43].
2 In Eq. (1.7) the propagator of the boundary variables turns out to be chiral invariant in the limit N due
to the negative contact term. In this respect, it is important to note the contribution of massive modes (including
the PauliVillars modes), which takes account of the chiral anomaly in Eqs. (1.8) and (1.9) and fills the gap
to the
P GinspargWilson fermion [31]. Following [32,39,40], we may also introduce an heavy auxiliary field

+ (x),

so that
a 4 x (x) (x) and redefine the boundary variables as q 0 (x) = q(x) + (x), q 0 (x) = q(x)
the contact term in the propagator is removed. It is also pointed out in [41] that a certain modification of the
action of the domain-wall fermion leads directly to the propagator of the boundary variables which satisfies the
GinspargWilson relation.

514

Y. Kikukawa / Nuclear Physics B 584 (2000) 511527

The exponential bounds on the effective Dirac operator and its differentiations can be
established from it.
We may also discuss the GinspargWilson relation of the effective Dirac operator
through this integral representation. We will see that this reduces to the question concerning
the property of the five-dimensional Dirac operator under the chiral transformation
introduced by Furman and Shamir [15].
Another interesting aspect of the effective Dirac operator is its behavior in case with
the singular gauge configuration for which isolated eigenvalues of the hermitian Wilson
Dirac operator collapse to zero. In [4], it has been proved rigorously that Neubergers
Dirac operator in terms of the hermitian WilsonDirac operator remains local even with
such singular gauge configurations. We will argue that it is also true for the effective Dirac
operator Eq. (1.5).
The approach to the chiral symmetry limit from a finite N is the most important issue
for the practical implementation of exact chiral symmetry using the domain-wall fermion
[19,25,26,28,4448]. In this respect, our result of the locality and exact chiral symmetry of
the effective four-dimensional action is restricted for the class of gauge fields with small
lattice field strength. It is a nonperturbative result and it gives a sufficient condition for
that the exact chiral symmetry based on the GinspargWilson relation can be implemented
using the domain-wall fermion. But it does not assure that it would work practically in
the numerical simulations using the standard Wilsons gauge action. Our result, however,
presents an explicit method to connect the locality and chiral symmetry properties of the
domain-wall fermion to the spectrum of the four-dimensional WilsonDirac operator. We
hope that such a method would be useful in order to study the above practical issue.
This paper is organized as follows. In Section 2, we briefly review the domain-wall
fermion in order to fix our notation. In Section 3, we describe how to derive the integral
representation for the effective Dirac operator. We also discuss how the GinspargWilson
relation for the effective Dirac operator follows in this integral representation. In Section 4,
we discuss the positivity of the five-dimensional WilsonDirac operator with the antiperiodic boundary condition. With this result, we consider exponential bounds for the
effective Dirac operator and its differentiations in Section 5. We also discuss the locality in
the case with the singular gauge configurations. In Section 6, we summarize our result and
give some discussions concerning the issue of the approach to the chiral symmetry limit.

2. Domain-wall fermion in the anti-periodic subtraction scheme


In this section, we review the domain-wall fermion [1315] and fix our notation. The
domain-wall fermion is defined by the five-dimensional WilsonDirac fermion with the
Dirichlet boundary condition.
SDW =

N
X
t =1

a4

X
x

(x,
t)D5w (x, t),

(2.1)

Y. Kikukawa / Nuclear Physics B 584 (2000) 511527

515


1
+ st + PL Mst + PR Mst .
(2.2)
2
We assume the lattice size of the fifth dimension N is even. For N = 6, the mass matrix
reads

B 1 0 0 0 0
0 B 1 0 0 0

1
0 0 B 1 0 0
(2.3)
Mst =

,
a5 0 0 0 B 1 0

0 0 0 0 B 1
0 0 0 0 0 B
D5w =

where B is defined by


m0
a

.
B = 1 + a5
2
a

(2.4)

The chiral transformation is introduced as vector-like one so that the symmetry breaking is
minimized [15]:
N
,
(2.5)
2
N
(2.6)
(x, t) = +(x, t), t > + 1.
2
Accordingly, the anomalous term is given by
 


 
 

2
N
N
N
N
R x,
L x, + 1 . (2.7)
L x, + 1 R x,
X(N) (x) =
a5
2
2
2
2
(x, t) = (x, t),

t6

The partition function of the domain-wall fermion may be defined with the subtraction
of the PauliVillars fields, which is subject to the anti-periodic boundary condition in the
fifth dimension. 3
det D5w
ZDW =
,
(2.8)
det D 5w
where

1
D 5w = + st + PL M st + PR M st ,
2

B 1 0 0 0 0
0 B 1 0 0 0

1
0
0
B
1
0
0

M st =

, N = 6.
a5 0 0 0 B 1 0

0 0 0 0 B 1
1 0 0 0 0 B

(2.9)

(2.10)

3 For this subtraction to work consistently, we should require the positivity of D


5w : D D 5w > 0. We will
5w
see later that this requirement also assures the locality and the GinspargWilson relation of the effective Dirac
operator.

516

Y. Kikukawa / Nuclear Physics B 584 (2000) 511527

For later convenience, we perform a chirally asymmetric parity transformation in the


fifth dimension:
(x, t) = (PR + PL P )t s 0 (x, s),

(x,
t) = 0 (x, s)(PR P + PL )st ,
where

0
0

0
Pst =
0

0
1

0
0
0
0
1
0

0
0
0
1
0
0

0
0
1
0
0
0

0
1
0
0
0
0

1
0

0
,
0

0
0

(2.11)
(2.12)

N = 6.

(2.13)

Accordingly, the five-dimensional Dirac operators are transformed as follows:


0
D5w
= (PR P + PL )D5w (PR + PL P )

1
= + Pst + MstH ,
2
0
= (PR P + PL )D 5w (PR + PL P )
D 5w

1
= + Pst + M stH ,
2

where, for N = 6,

0
0

1
0

H
Mst = Mst P = P Mst =

a5 0

1
B

0
0

1
0
M st =

a5 0

1
B

0
0
0
1
B
0

0
0
1
B
0
0

0
1
B
0
0
0

1
B
0
0
0
0

0
0
0
1
B
0

(2.14)
(2.15)
(2.16)
(2.17)

0
0
1
B
0
0

B
0

0
.
0

0
1

0
1
B
0
0
0

1
B
0
0
0
0

B
0

0
.
0

0
0

(2.18)

(2.19)

In this basis, the chiral transformation adopted by Shamir and Furman [15] can be
expressed as follows:
s0 (x) = (5 )st t0 (x),
where 5 is given (for N = 6) by

(2.20)

Y. Kikukawa / Nuclear Physics B 584 (2000) 511527

517

5 0
0 0 0 0
0 0 0 0 0
5

0 5 0 0 0
0
(5 )st =
,
0
0
0 5 0 0

0
0
0 0 5 0
0
0
0 0 0 5

N = 6.

(2.21)

0
0
and D 5w
satisfy the following
With this definition of the chiral transformation, D5w
identities, respectively,

2
5 s N t N ,
2
2
a5


2
2
0
0
+ D 5w
5 st =
5 s N t N + 5 sN t N .
5 D 5w
2
2
a5
a5
0
0
+ D5w
5
5 D5w

st

(2.22)
(2.23)

The chiral symmetry breaking occurs at t = N/2 in the five-dimensional Dirac operator
for the domain-wall fermion. On the other hand, it occurs both at t = N/2 and at t = N for
the PauliVillars field, because of the anti-periodic boundary condition.

3. Effective four-dimensional Dirac operator


3.1. An integral representation of the effective four-dimensional Dirac operator
The functional determinant of the domain-wall fermion, in the anti-periodic subtraction
scheme, reduces to a single determinant of a four-dimensional Dirac operator,
det D5w
(N)
= det aDeff .
det D 5w

(3.1)

In this section, we reproduce this result and derive an integral representation for the
effective four-dimensional Dirac operator.
We may write the partition function as follows:
0
 0 1 
det D5w det D5w
0
=
= det D5w
D 5w
.
0
det D 5w det D 5w

(3.2)

Then, we note a simple relation between two five-dimensional WilsonDirac operators:


0
0
= D 5w

D5w

1
sN Nt .
a5

(3.3)

This relation implies that


 0 1
 0 1
1
0
= st sN D 5w
.
D 5w
D5w
1t
a5

(3.4)

Since this matrix is lower triangle in the lattice indices of the fifth dimension, we can easily
see that its determinant reduces to a single four-dimensional determinant:

518

Y. Kikukawa / Nuclear Physics B 584 (2000) 511527



 0 1 
1  0 1
0
det D5w
D 5w
= det 1
D5w NN .
a5
From this result, we may set
1  0 1
(N)
D
aDeff = 1
a5 5w NN

1

1
1
PR D 5w NN PL + PL D 5w 11 PR
=1
a5

1

1 
+ PR D 5w
PR + PL D 5w
PL .
N1

(3.5)

(3.6)

(3.7)

1N

Thus the effective four-dimensional Dirac operator can be expressed in terms of the inverse
of the five-dimensional WilsonDirac operator with the anti-periodic boundary condition.
Since the gauge field is four-dimensional, the inverse of this five-dimensional Wilson
Dirac operator may be expressed as follows:



1

1 X ip(st )
m0 1
e
,
(3.8)
i5 sin p + 1 cos p + a5 Dw
a5 D 5w st =
N p
a
where the summation is taken over the discrete momenta p =
and Dw is the four-dimensional WilsonDirac operator

X 1
 a
+ .
Dw =
2
2

2
N (k

12 ) (k = 1, 2, . . . , N)
(3.9)

Then the effective Dirac operator may be expressed as follows:


1 X
1
(N)
 PL
aDeff = 1 PR
N p i5 sin p + 1 cos p + a5 Dw ma0
1 X
1
 PR
PL
N p i5 sin p + 1 cos p + a5 Dw ma0
+ PR

1 X
eip
N p i5 sin p + 1 cos p + a5 Dw

m0  PR
a

+ PL

1 X
e+ip
N p i5 sin p + 1 cos p + a5 Dw

m0  PL .
a

(3.10)

In the limit N , the summation over the discrete momentum reduces to the continuous
integral:
1
N

X
p= 2
N

(k 21 )

Z
H

dp
.
2

(3.11)

Note that, since we do not use the transfer matrix in this derivation, this expression
could hold true even if B is not positive definite and the transfer matrix is not defined
consistently. m0 can be chosen as any value within m0 [0, 2] (when a5 = a), as long as
the five-dimensional WilsonDirac operator with the anti-periodic boundary condition in
the fifth dimension is not singular and invertible.

Y. Kikukawa / Nuclear Physics B 584 (2000) 511527

519

3.2. The GinspargWilson relation


Next we discuss the GinspargWilson relation for the effective Dirac operator in the
integral representation. As we have seen in the previous subsection, the effective Dirac
operator, Deff , is defined by
aDeff = 1

1  0 1
,
D
a5 5w NN

N = .

(3.12)

If it would satisfies the GinspargWilson relation


5 Deff + Deff 5 = 2aDeff 5 Deff ,

(3.13)

then the following identity must hold true in the limit of N :


 0 1
 0 1
2  0 1  0 1
+ D 5w
=
5 D5w NN ,
D
5 D 5w
NN
NN 5
a5 5w NN

N = .

(3.14)

0
We may compare this identity with Eq. (2.23) which express the chiral property of D 5w
under the chiral transformation introduced by Furman and Shamir Eq. (2.20). The latter
we may write

 0 1  0 1
2  0 1  0 1
+ D 5w
5 st =
D5w s N 5 D5w N t
5 D 5w
2
2
a5
2  0 1  0 1
+
D5w sN 5 D5w Nt .
a5

(3.15)

Setting s = t = N , we obtain
 0 1
 0 1
2  0 1  0 1
+ D 5w
=
D5w N N 5 D5w N N
5 D 5w
NN
NN 5
2
2
a5
2  0 1  0 1
+
D5w NN 5 D5w NN .
a5

(3.16)

Then we can see that Eq. (3.14) is equivalent to the following condition in the limit of
N :
 0 1
(3.17)
D 5w N N = 0, N .
2

As we will see below, this condition is fulfilled as long as the five-dimensional Wilson
Dirac operator with the anti-periodic boundary condition is not singular and invertible.
Then we obtain Eq. (3.14) and Eq. (3.13), the GinspargWilson relation for the effective
Dirac operator.

4. Positivity of the square of the five-dimensional WilsonDirac operator with


anti-periodic boundary condition
From Eqs. (3.10) and its N limit, we see that for the effective four-dimensional
Dirac operator to be defined consistently, it is required that the five-dimensional Wilson
Dirac operator with the anti-periodic boundary condition should be non-singular and

520

Y. Kikukawa / Nuclear Physics B 584 (2000) 511527

invertible. In this section, we examine the positivity of the five-dimensional WilsonDirac


operator square.
To examine this requirement, we evaluate the square of the five-dimensional Wilson
Dirac operator. Setting a5 = a for simplicity, we have



i5 sin p + 1 cos p + (aDw m0 ) i5 sin p + 1 cos p + (aDw m0 )


a2
2

(4.1)
= 4 sin (p/2) 1 m0 + (aDw m0 ) (aDw m0 ).
2

D5w
For m0 = 1, the first term is positive semi-definite and then the positivity of a 2 D 5w
is entirely determined by the positivity of the four-dimensional WilsonDirac operator
square, (aDw 1) (aDw 1). According to the result of [4], if the plaquette variables
U (p) are uniformly bounded as

k1 U (p)k < ,

(4.2)

we obtain



i5 sin p + 1 cos p + (aDw 1) i5 sin p + 1 cos p + (aDw 1)


 2



a
2


4
sin

(p/2)

1)
(aD

1)
+
(aD
=

w
w


2
> 1 30.

(4.3)

For the generic value of m0 [0, 2], we also obtain [42,49,50]





i5 sin p + 1 cos p + (aDw m0 ) i5 sin p + 1 cos p + (aDw m0 )
2

1
(4.4)
> (1 30) 2 |1 m0 | , if 1 30 > |1 m0 |2 .
Recently, it has been shown
by Neuberger [42] that the constant 30 in the above bounds
can be improved to 6(2 + 2).
From these considerations, we may assume the positive lower and upper bounds of
the square of the five-dimensional WilsonDirac operator with the anti-periodic boundary
condition as



(4.5)
0 < 6 4 sin2 (p/2)B + (aDw m0 ) (aDw m0 ) 6 ,
under the following condition,
k1 U (p)k < ,

<


1 |1 m0 |2 .
6(2 + 2)
1

(4.6)

5. Expansion with Chebycheff polynomials and an exponential bound


Given the bounds on the square of the five-dimensional WilsonDirac operator with the
anti-periodic boundary condition, we will derive in this section exponential bounds on the
effective four-dimensional Dirac operator and its differentiations with respect to the gauge
field. We will also obtain an exponential bound on the inverse of the five-dimensional

Y. Kikukawa / Nuclear Physics B 584 (2000) 511527

521

WilsonDirac operator which is needed to prove Eq. (3.17) and the GinspargWilson
relation.
5.1. Locality bounds
From Eqs. (3.10) and its N limit, we see that the locality property of the effective
Dirac operator of the domain-wall fermion is determined by the locality properties of the
following operators in the integral representation:
1 X +ip ip
1
(5.1)
1, e , e
I (N) =
N p
i5 sin p + 1 cos p + (aDw m0 )
and
Z

1
dp  +ip ip
1, e , e
.
2
i5 sin p + 1 cos p + (aDw m0 )

I=

(5.2)

The integrand can be written as


 +ip ip
1
1, e , e
i5 sin p + 1 cos p + (aDw m0 )
1
=
2
4 sin (p/2)B + (aDw m0 ) (aDw m0 )


1 PR eip PL eip + (aDw m0 ) 1, e+ip , eip .

(5.3)

From this expression it is clear that the operators in the numerator are local and bounded.
Then we may omit these operators in the following considerations.
We can obtain an expansion of the integrand using the generating function of the
Chebycheff polynomials [43]

X
1
=
t k Uk (z),
2
1 2tz + t

kUk (z)k 6 Uk (1) = k.

(5.4)

k=0

Following [4], we set

t = e ,

cosh =

+
,

(5.5)

and
z=

+ 2{4 sin2 (p/2)B + (aDw m0 ) (aDw m0 )}


.

(5.6)

Then we obtain
1
4 sin2 (p/2)B + (aDw m0 ) (aDw m0 )

4t

k=0

t k Uk (z).

(5.7)

This defines an expansion in terms of the square of the WilsonDirac operator and B
with only nearest-neighbor and next-to-nearest-neighbor couplings. In order to contribute

522

Y. Kikukawa / Nuclear Physics B 584 (2000) 511527

to the kernel of the operator Eq. (5.7) between two lattice sites x and y of the lattice
distance d(x, y) = |x y|, the order of the polynomials Uk (x) in the expansion must be
greater than d(x, y)/2a:
d(x, y)
.
(5.8)
2a
Then for the given distance d(x, y), the series expansion Eq. (5.7) can be arranged as
follows:
 X


4t
exp d(x, y)
t k Uk+d/2a (z).
(5.9)
2a

k>

k=0

Noting the bound on the polynomials, kUk (z)k 6 k, we obtain






1


4 sin2 (p/2)B + (aD m ) (aD m ) (x, y)
w
0
w
0
 X


4t
exp d(x, y)
t k kUk+d/2a (z)k
6
2a

k=0




t
1 d(x, y)

4t
+
d(x,
y)
.
exp

6
2a
(1 t)2
1 t 2a

(5.10)

Since the summation over the momentum in I (N) (the integration in I ) is normalized to
unity, the above bounds implies the exponential bound for the integrals, I (N) and I .
As for the differentiations of the effective Dirac operator, we can also derive the
exponential bounds, following [4]. We consider the differentiations of the Chebycheff
expansion, Eq. (5.7). We first introduce an integral representation for the Chebycheff
polynomials:
I
d k1
1

,
(5.11)
Uk (z) =
2
2 2z + 1
where the integration is defined along a circle in the complex plane centered at the origin.
The radius r of the circle should be strictly less than 1 to avoid the singularities of the
integrand. The denominator of the integrand can be factorized according to
2 2z + 1 = ( u )( u),
Since, u u = 1, it is clear that
2

2z + 1 1 6 (1 r)2 .

u = z + i(1 r)1/2 .

(5.12)

(5.13)

If we denote the differentiation of z with respect to the gauge fields as z , then we obtain


U k (z) 6 2kzkr k (1 r)4 .
(5.14)
We may now adjust the radius r so that the factor r k (1 r)4 is minimized. We obtain


U k (z) 6 constant kzk(1 + k)4 .
(5.15)
With these bounds, we can see that the differentiated series Eq. (5.7) is also exponentially
convergent with the same exponent as the original series. By similar estimations, we can

Y. Kikukawa / Nuclear Physics B 584 (2000) 511527

523

see that this is also true for higher-order differentiations (each differentiation give rise to
an additional factor of (1 + k)2 in the bound on the Chebycheff polynomials).
5.2. Exponential bounds in the fifth-direction and the GinspargWilson relation
We next consider the exponential bound on the inverse of the five-dimensional Wilson
Dirac operator which is necessary to prove Eq. (3.17) and the GinspargWilson relation.
From the above derivation of the exponential bounds for the summation and integral
Eqs. (5.1) and (5.2), we can see that the same bound holds true for the inverse of the
five-dimensional WilsonDirac operator, D 5w , itself:


a D 5w

1
st

1 X
e+ip(st )
.
N p i5 sin p + 1 cos p + (aDw m0 )

(5.16)

In fact, we can obtain






 2

a D D 5w 1 (x, s; y, t) 6 C exp d5 (x, s; y, t) ,
5w
2a
where d5 (x, s; y, t) = |x y| + min(|s t|, N |s t|) and


t
1 d5 (x, s; y, t)
4t
+
.
C=
2a
(1 t)2
1 t
From this bound, it follows immediately that
 0 1
= 0.
lim D 5w
NN
N

(5.17)

(5.18)

(5.19)

This completes the proof of the GinspargWilson relation under the condition on the
plaquette variables Eq. (4.6). Note again that this proof does not refer to the transfer matrix
and it applies for any value of m0 [0, 2].
5.3. Singular case
In this subsection, we examine locality of the effective Dirac operator Eq. (1.5) with the
singular gauge configuration for which isolated eigenvalues of the hermitian WilsonDirac
operator collapse to zero:
H 0 (x) = 0 (x),

' 0,

(5.20)

where



m0
.
H = 5 Dw
a

(5.21)

We will argue that the effective Dirac operator remains local even with such singular gauge
configurations.
For this purpose, however, the integral representation and the Chebycheff expansion in
terms of the five-dimensional WilsonDirac operator, considered so far, does not seem
to be useful. When the isolated near-zero mode occurs in the four-dimensional hermitian

524

Y. Kikukawa / Nuclear Physics B 584 (2000) 511527

WilsonDirac operator, it is associated with many modes with small fifth momenta in the
spectrum of the five-dimensional WilsonDirac operator (with the anti-periodic boundary
condition). This means that the continuum spectrum would collapse to zero in the fivedimensional WilsonDirac operator in the limit N . Then the separation of the effect
of the near-zero mode does not seem easy in this representation (cf. [4]). Therefore, in this
e.
section, we use the formula for the effective action in terms of the transfer matrix and H
In [4], it has been proved rigorously that the contribution of the near-zero-mode to
Neubergers Dirac operator,

H

,
(5.22)
H 2 near-zero
remains local. This result can be understood from the localization properties of the
eigenvectors of the near-zero modes. In fact, it is well localized with exponentially
decaying tails [4,44].
e is related to H by the formula
Since H
e

ea5 H + e+a5 H 2 = a52H

1
H,
B

(5.23)

e is the exact zero mode of H [2,39,44,46]. Therefore, the


the exact zero mode of H
e to the effective Dirac operator is identical to the
contribution of the zero mode of H
contribution of the zero mode of H to Neubergers Dirac operator,


e
H
H
p
=
= lim sign() 0 (x)0 (y),
(5.24)
2
0
2
e
H
zero
H zero
and it remains local, according to the result of [4]. It is expected that this localization
e.
properties persist also for the contribution of the near-zero modes of H
It is desirable to make the above argument rigorous. We will leave this issue for future
study.

6. Discussion
We have argued locality in the domain-wall fermion approach through its effective
four-dimensional Dirac operator. As expected, all the properties proved rigorously for
Neubergers Dirac operator holds true for the effective Dirac operator. In particular, we
have shown explicitly that the locality properties of the domain-wall fermion depends
crucially on the spectrum of the four-dimensional WilsonDirac operator, which is closely
related to that of the five-dimensional WilsonDirac operator (with the anti-periodic
boundary condition). Then we can see that the bound for the plaquette variables leads
to the locality bound for the effective Dirac operator. We have also shown that the effective
Dirac operator satisfies the GinspargWilson relation with the same bound for plaquette
variables.
The approach to the chiral symmetry limit from a finite N is the most important issue
for the practical implementation of exact chiral symmetry using the domain-wall fermion

Y. Kikukawa / Nuclear Physics B 584 (2000) 511527

525

[19,25,26,28,4648]. In order to examine the effect of the finite N , the explicit breaking
term in the axial WardTakahashi identity has been measured, among other physical
quantities. This breaking term can be written by the correlation function between the
middle and the boundary of the fifth dimension [31]:


1

1 (N) 1
1
0

D
(x, y).
(6.1)
a5 D5w N ,N =
e
2
a eff
2 cosh N2 a5 H
From the point of view of the effective four-dimensional action, this effect may be
examined through the breaking term in the GinspargWilson relation. 4 As we have seen
in the Section 5.2, it is given by the similar correlation function defined through the fivedimensional WilsonDirac operator with the anti-periodic boundary condition:


0
a5 D 5w

N
2 ,N

=
=

1
e
2 cosh N2 a5 H
N
(eip ) 2
1 X
N p i5 sin p + 1 cos p + a5 Dw

m0  .
a

(6.2)

It is the smallest eigenvalue of square of the four-dimensional WilsonDirac operator


which determines the behavior of the breaking term in the limit N . When the
isolated near-zero mode occurs in the four-dimensional hermitian WilsonDirac operator,
it is associated with many modes with small fifth momenta in the spectrum of the fivedimensional WilsonDirac operator (with the anti-periodic boundary condition). This
means that the continuum spectrum tends to collapse to zero and the lower bound for
this continuum spectrum determines the rate of the exponential decay in the limit N .
Therefore, it would be important to examine the behavior of the near-zero modes as done
in [4], also in the context of the domain-wall fermion, as suggested in [25,26]. In view

D5w tends to be
of the data obtained in [4] and also in [51], lowest eigenvalues of a 2 D5w
4
6
small and of order O(10 ) to O(10 ) for = 5.85. Then it would happen that a quite
large N (of order hundreds) would be required to achieve good chiral behaviors.
It is also desirable to clarify the nature of the distribution of small eigenvalues of the fourdimensional WilsonDirac operator with a negative mass, for the gauge field configurations
used in the current simulations [45]. The phase structure of the domain-wall fermion
is under investigation in relation to the possible parity-flavor breaking phase, which is
expected to give qualitative understanding how the chiral symmetry limit is approached in
the limit N [52].

Acknowledgements
The intensive discussions at the Summer Institute 99 at Yamanashi, Japan, were very
suggestive and useful to complete this work. The author would like to thank S. Aoki,
T. Onogi, Y. Taniguchi, T. Izubuchi, K. Nagai, J. Noaki, K. Nagao, N. Ukita, and H. So
4 Numerical study of the GinspargWilson relation of the effective Dirac operator is found in [41].

526

Y. Kikukawa / Nuclear Physics B 584 (2000) 511527

for enlightening discussions. He is grateful to D.A. Adams and H. Neuberger for the
correspondences concerning the bound of the hermitian WilsonDirac operator. The author
would like to thank T.-W. Chiu for the kind hospitality at Chiral 99 in Taipei. The author
is also grateful to M. Lscher and A. Borii for the comments on the first version of this
article. This work is supported in part by Grant-in-Aid for Scientific Research of Ministry
of Education (#10740116).
References
[1] H. Neuberger, Phys. Lett. B 417 (1998) 141; Phys. Lett. B 427 (1998) 353.
[2] R. Narayanan, H. Neuberger, Nucl. Phys. B 412 (1994) 574; Phys. Rev. Lett. 71 (1993) 3251;
Nucl. Phys. B 443 (1995) 305.
[3] P.H. Ginsparg, K.G. Wilson, Phys. Rev. D 25 (1982) 2649.
[4] P. Hernndes, K. Jansen, M. Lscher, Nucl. Phys. B 552 (1999) 363378, hep-lat/9808010.
[5] P. Hasenfratz, V. Laliena, F. Niedermayer, Phys. Lett. B 427 (1998) 125.
[6] M. Lscher, Nucl. Phys. B 538 (1999) 515.
[7] M. Lscher, Nucl. Phys. B 549 (1999) 295.
[8] M. Lscher, Weyl fermions on the lattice and the non-abelian gauge anomaly, hep-lat/9904009.
[9] T. Fujiwara, H. Suzuki, K. Wu, Non-commutative differential calculus and the axial anomaly in
abelian lattice gauge theories, Phys. Lett. B 463 (1999) 63, hep-lat/9906015.
[10] D.H. Adams, A topological aspect of the non-abelian anomaly for Weyl fermions on the lattice,
hep-lat/9910036.
[11] M. Lscher, Plenary talk at the International Symposium on Lattice Field Theory, Pisa, Italy,
June 29July 3, 1999.
[12] A. Borii, Phys. Lett. B 453 (1999) 46; hep-lat/9910045.
[13] D.B. Kaplan, Phys. Lett. B 288 (1992) 342.
[14] Y. Shamir, Nucl. Phys. B 406 (1993) 90.
[15] V. Furman, Y. Shamir, Nucl. Phys. B 439 (1995) 54.
[16] T. Blum, A. Soni, Phys. Rev. D 56 (1997) 174; Phys. Rev. Lett. 79 (1997) 3595.
[17] T. Blum, A. Soni, M. Wingate, Phys. Rev. D 60 (1999) 114507.
[18] P. Vranas, Phys. Rev. D 57 (1998) 1415.
[19] P. Chen et al. (Columbia Univ. Collaboration), Phys. Rev. D 59 (1999) 054508; Nucl. Phys.
(Proc. Suppl.) 73 (1999) 204; Nucl. Phys. (Proc. Suppl.) 73 (1999) 207; Nucl. Phys. (Proc.
Suppl.) 73 (1999) 456; hep-lat/9812011; hep-lat/9903024; hep-lat/9909140; hep-lat/9911002.
[20] S. Aoki, Y. Taniguchi, Phys. Rev. D 59 (1999) 054510; Nucl. Phys. (Proc. Suppl.) 63 (1998)
290.
[21] S. Aoki, T. Izubuchi, Y. Kuramashi, Y. Taniguchi, Phys. Rev. D 59 (1999) 094505; Phys. Rev.
D 60 (1999) 114504.
[22] S. Aoki, Y. Taniguchi, Phys. Rev. D 59 (1999) 094506.
[23] S. Aoki, T. Izubuchi, J. Noaki, Y. Kuramashi, Y. Taniguchi, hep-lat/9909049.
[24] J. Noaki, Y. Taniguchi, Scaling property of domain-wall QCD in perturbation theory, heplat/9906030.
[25] S. Aoki, T. Izubuchi, Y. Kuramashi, Y. Taniguchi, hep-lat/9909154.
[26] A. Ali Khan et al. (CP-PACS Collaboration), hep-lat/9909049.
[27] T. Blum et al. (RIKEN-BNL-Columbia Collaboration), A first study of epsilon/epsilon on the
lattice using domain wall fermions, hep-lat/9908025.
[28] T. Blum et al. (RIKEN-BNL-Columbia Collaboration), hep-lat/9909093; hep-lat/9909101; heplat/9909117; hep-lat/9909382.
[29] T.-W. Chiu, hep-lat/9912005.

Y. Kikukawa / Nuclear Physics B 584 (2000) 511527

[30]
[31]
[32]
[33]
[34]
[35]
[36]
[37]
[38]
[39]
[40]
[41]
[42]
[43]
[44]
[45]
[46]
[47]
[48]
[49]
[50]
[51]
[52]

527

H. Neuberger, Phys. Rev. D 57 (1998) 5417.


Y. Kikukawa, T. Noguchi, hep-lat/9902022.
M. Lscher, Phys. Lett. B 428 (1998) 342.
Y. Kikukawa, A. Yamada, Phys. Lett. B 448 (1999) 265.
D.A. Adams, hep-lat/9812019.
H. Suzuki, Prog. Theor. Phys. 102 (1999) 141.
K. Fujikawa, Nucl. Phys. B 546 (1999) 480.
T.-W. Chiu, Phys. Lett. B 445 (1999) 371.
T.-W. Chiu, T.-H. Hsieh, hep-lat/9901011.
M. Lscher, private communication.
H. Neuberger, Phys. Rev. Lett. 81 (1998) 4060; Int. J. Mod. Phys. C 10 (1999) 1051; Phys. Rev.
D 60 (1999) 065006; hep-lat/9909043.
A. Borii, hep-lat/9909057; hep-lat/9912040.
H. Neuberger, hep-lat/9911004.
B. Bunk, hep-lat/9805030.
R.G. Edwards, U.M. Heller, R. Narayanan, Nucl. Phys. B 535 (1998) 403.
R.G. Edwards, U.M. Heller, R. Narayanan, Phys. Rev. D 60 (1999) 034502.
Y. Shamir, Phys. Rev. D 59 (1999) 054506.
R.C. Brower, B. Svetitsky, hep-lat/9912019.
Y. Shamir, hep-lat/9912027.
D.H. Adams, Contribution to Chiral 99, Workshop on Chiral Gauge Theories, September
1318, 1999, Taipei, hep-lat/0001014.
M. Lscher, unpublished note.
P. Hernandez, K. Jansen, L. Lellouch, A numerical treatment of Neubergers lattice Dirac
operator, hep-lat/0001008.
S. Aoki, Y. Kikukawa, Y. Nakayama, T. Onogi, M. Sugiura, in preparation.

Nuclear Physics B 584 (2000) 528542


www.elsevier.nl/locate/npe

Improved gauge actions on anisotropic lattices I.


Study of fundamental parameters in the weak
coupling limit
S. Sakai a , T. Saito b, , A. Nakamura c
a Faculty of Education, Yamagata University, Yamagata, 990-8560, Japan
b Faculty of Science, Hiroshima University, Higashi-Hiroshima, 739-8526, Japan
c Research Institute of Information Sciences and Education, Hiroshima University,

Higashi-Hiroshima, 739-8521, Japan


Received 29 February 2000; revised 25 May 2000; accepted 7 June 2000

Abstract
On anisotropic lattices with the anisotropy = a /a the following basic parameters are
calculated by perturbative method: (1) the renormalization of the gauge coupling in spatial and
temporal directions, g and g , (2) the parameter, (3) the ratio of the renormalized and bare
anisotropy = /B and (4) the derivatives of the coupling constants with respect to , g2 / and
g2 / . We employ the improved gauge actions which consist of plaquette and six-link rectangular
loops, c0 P (1 1) + c1 P (1 2) . This class of actions covers Symanzik, Iwasaki and DBW2
actions. The ratio shows an impressive behavior as a function of c1 , i.e., > 1 for the standard
Wilson and Symanzik actions, while < 1 for Iwasaki and DBW2 actions. This is confirmed nonperturbatively by numerical simulations in weak coupling regions. The derivatives g2 / and
g2 / also change sign as c1 increases. For Iwasaki and DBW2 actions they become opposite
sign to those for standard and Symanzik actions. However, their sum is independent of the type of
actions due to Karschs sum rule. 2000 Elsevier Science B.V. All rights reserved.

1. Introduction
Anisotropic lattices allow us to carry out numerical simulations with the fine temporal
resolution while keeping the spatial lattice spacing coarse, i.e., a < a , where a and a
are lattice spacing in the temporal and spatial directions, respectively. This is especially
important for QCD Monte Carlo simulation at finite temperature and heavy particle
spectroscopy. There have been many such calculations like glue thermodynamics [1],
hadron masses at the finite temperature [2], glueballs [3] and heavy quark spectra [4].
Corresponding author. E-mail: tsaito@hirohe.hepl.hiroshima-u.ac.jp

0550-3213/00/$ see front matter 2000 Elsevier Science B.V. All rights reserved.
PII: S 0 5 5 0 - 3 2 1 3 ( 0 0 ) 0 0 3 8 3 - 7

S. Sakai et al. / Nuclear Physics B 584 (2000) 528542

529

The anisotropic lattice may become an important tool for the calculation of the transport
coefficients of the quarkgluon plasma [57] and for the determination of spectral
functions at finite temperature [8,9]. In numerical simulations on the anisotropic lattice we
need the information upon the renormalization of anisotropy which is given by = /B ,
where is the renormalized anisotropy, a /a , and B is the bare one. Karsch has first
studied anisotropic lattices perturbatively with the standard plaquette action and obtained
together with anisotropy coefficients which are defined by the derivatives of the spatial
and temporal gauge couplings with respect to , i.e., g2 / and g2 / , and the QCD
parameter [10]. In Ref. [11] was determined non-perturbatively by analyzing Wilson
loops in numerical simulations.
Anisotropic lattices play an essential role in the analysis of thermodynamics of QCD. To
get the thermodynamics quantities like the internal energy and the pressure from numerical
simulations, one needs to know the anisotropy coefficients.
On the anisotropic lattice, we can change the temperature by changing not only N but
also a . This allows us to adjust the temperature continuously with fixed spatial volume.
Recently it has been recognized that improved actions are effective to get reliable results
from lattice QCD simulations on the relatively coarse lattice; lattice artifacts due to the
discretization are expected to be much less. Therefore, it is important to employ improved
actions in the anisotropic lattice calculations, since the lattice is rather coarse in spatial
direction in current numerical simulations. Garca Prez and van Baal pursued first this
direction, i.e., they have determined the one-loop correction to the anisotropy for the square
Symanzik improved action [12].
In this article we study the improved actions which consist of plaquette and six-link
rectangular loops,
X

(1)
c0 P (1 1) + c1 P (1 2) ,
S
where c0 and c1 satisfy the relation c0 + 8c1 = 1. This class of actions covers tree level
1
) [13,14], Iwasaki action
Symanzik action without the tadpole improvement (c1 = 12
(c1 = 0.331) [15] and DBW2 action (c1 = 1.4088) [16]. They are most widely used
in simulations of recent days. For the class of actions which consists of planar loops the
anisotropic lattice can be formulated in the same way as for the standard plaquette action.
This may not be the case for improved actions which include non-planar loops in three or
four dimensions.
In the following, we will calculate the parameter, = /B and the anisotropy
coefficients, in weak coupling regions mainly by perturbative calculations.
In Section 2 we briefly review the formulation of the anisotropic lattice with the
improved actions and summarize formulae which will be used in this paper. In Section 3,
we outline the background field method and discuss the removal of the infrared divergence.
In Section 4 we present results of the perturbative calculations. The c1 dependence of
and anisotropy coefficients are studied in detail. Since the behavior of is very important
for practical use, we will study it by the numerical simulations in Section 5. Section 6 is
devoted to concluding remarks.

530

S. Sakai et al. / Nuclear Physics B 584 (2000) 528542

2. Anisotropic lattice with improved gauge actions


In case of improved actions that consist of plaquette and rectangular loops, the
anisotropic lattice may be formulated in the same way as for the standard plaquette
action [10]. The action takes the form
XX
XX
Pij +
P4i ,
(2)
Sg =
x i>j

i6=4

where P are plaquette and six-link rectangular loop operators in plane,


P = c0 P (1 1) + c1 P (1 2) ,

(3)

which are constructed of the link variable Un, , and


=

2Nc
1 ,
2
g (a, )

2Nc
.
2
g (a, )

(4)

Here g and g are the coupling constants in the spatial and temporal directions,
respectively. The action can be also written with the bare anisotropy parameter B as
!
XX
1 XX
Pij + B
P4i ,
(5)
Sg =
B x
x
i>j

i6=4

p
where = 2Nc /g2 = and /B = g2 /g2 .
The weak coupling limit of the anisotropic lattice is fully discussed in Ref. [10].
Therefore, we will summarize only equations which are necessary in the following studies.
In the continuum limit a 0 (g 0), the lattice spacing and the coupling g are
related with each other by the scale parameter through the renormalization group
relation,

b /(2b2)

(6)
a ( ) = b0 g2 1 0 exp 1/ 2b0 g2 ,

where b0 and b1 are the universal first two coefficients of -function,




11Nc
34 Nc 2
,
b1 =
.
b0 =
3 16 2
48 2

(7)

We calculate the effective action using the background field method [21] up to one-loop
order,


 X 2 3
1
1
2
C ( ) + O g
Fij a a
Seff =
4 g2 ( )
i,j


 X 2

1
1
2

C
+
(
)
+
O
g
(8)
Fi4 + F4i2 a3 a .

2
4 g ( )
i6=4

Effective actions with different value of the anisotropy parameter correspond to different
regularization scheme, but they should have the same continuum limit and we require
6=1
=1
1Seff = Seff Seff = 0. Then the relations are obtained,

S. Sakai et al. / Nuclear Physics B 584 (2000) 528542

1
g2 ( )

1
g 2 (1)



+ C ( ) C (1) + O g 2 ,



1
1
2
=
+
C
(
)

C
(1)
+
O
g
.

g2 ( ) g 2 (1)

531

(9)

(10)

In the following, the deviation of the one-loop quantum correction from the isotropic case
is often employed and written as 1C ( ) = C ( ) C (1), 1C ( ) = C ( ) C (1).
Perturbatively, all fundamental parameters on the anisotropic lattice are given in terms
of C ( ) and C ( ). The parameter on the anisotropic lattice is given by


1C ( ) + 1C ( )
( )
= exp
.
(11)
(1)
4b0
The quantum correction for the anisotropy parameter = /B is written as
 2 1/2

g
Nc

1 ( ) + O 2 ,
=
=1+
(, )
2
B

g
1 ( ) = C ( ) C ( ),

(12)
(13)

where = 2Nc /g 2 , and 1 ( ) is the quantum correction from the one-loop calculation.
Anisotropy coefficients are given by the derivative of the C ( ) and C ( ) with respect to
,
g2 ( ) C ( )
=
,

g2 ( ) C ( )
=
.

(14)

They play an important role in QCD thermodynamics [10,11,17,18] as will be discussed in


Section 4.3.

3. Perturbative calculation of C and C


3.1. Background field method
We calculate C ( ) and C ( ) in one-loop order, by applying the background field
method. The background field method on the lattice is well known [1921] and therefore
here we will only outline the method of the calculation and stress the points related to
the anisotropic lattice. The gauge field is decomposed into a quantum field and a
background one B which satisfies the classical equation of motion,
(0)
,
Un, = eig a (n) Un,

(0)
Un,
= eia B (n) .

A gauge fixing term is introduced as,


2
X X
3
(0)
Tr
D (n) .
Sg.f. = a a
n

(15)

(16)

532

S. Sakai et al. / Nuclear Physics B 584 (2000) 528542

Here

1
(0)
(0)
(n + )Un,
(n) ,
Un,
a

1
(0)
(0)
D (0)
(n )Un,
(n) .
Un,
=
a
D(0) =

The FaddeevPopov term resulting from the gauge fixing is


XX 


Tr D(0) (n) D(0) (n) .
SF.P. = 2a3 a
n

(17)

(18)

The total action becomes


Stot ( , , B ) = Sg ( , B ) + Sg.f. ( , B ) + SF.P. (, B ),

(19)

where Sg is the gauge action constructed from plaquette and six-link rectangular loops.
It is invariant under the following gauge transformation,
(0)
(0)

Un, V (n)Un, V (n + ),
(20)
(n) V (n) (n)V (n),

(n) V (n)(n)V (n),


(

D (n) V (n)D (n)V (n),


D (n) V (n)D (n)V (n),

(21)

where V is an element of SU (Nc ). For the calculation of the effective action in oneloop order we expand the total action up to second order in and B , which we
(2)
. With the help of CampbellHausdorffs formula and a relation V exp(i) =
denote as Stot
exp(iV V 1 )V , we split the total action into a classical action and a bilinear term of
and ,
(2)

Stot = S(B ) + Stot ( , , B ).

(22)

The term linear in B is missing because of the equation of motion for the background
field.
Thanks to the gauge invariance of the background field, it is sufficient to calculate the
coefficients of p p B B to obtain the effective action. For the calculation of this term,
we have applied the method explained in the appendix of Ref. [22], which makes the
calculation much simpler.
0
(2)
(2)
into several parts, i.e., Stot
= S0 + S0 + +
It is convenient to separate the action Stot
0
0
0
S6 + SF.P. . S0 , S0 , . . . , and S6 are symbolically expressed as follows [22]

2
S3 : [B, ][B, ],

S00 : f ,

2,

:
(1)
S
S
4 : f [B, [B, ]],
0
S5 : [[B, ], ]W,
S1 : f [B, ],
(23)

2
2
S1 : 1[B, ],
S6 : f W ,

S6 : W [, W ],
S2 : [, ]W,

S. Sakai et al. / Nuclear Physics B 584 (2000) 528542

533

where B and represent B and , respectively, and W and f are the field strength
tensors of background and quantum fields, respectively. 1 is a lattice derivative when
0
we set U (0) = 1 in Eq. (17). S0 and S0 are the free part of the action, which defines
0
gluon propagators. S1 , S1 and S2 terms correspond to three-point diagram from which
0
we construct the one-loop self-energy. S3 to S6 contribute to the tadpole self-energy. Here
0
0
S0 and S1 result from gauge fixing terms. The FaddeevPopov term is the same as the
previous calculations except that the anisotropy parameter is included [1921].
3.2. Effects of anisotropy parameters
By the integration over the quantum fields, we obtain the effective action in one-loop
order. We carry out the integration in the momentum space. Fourier transform of the gauge
and FaddeevPopov fields are defined as
Z
a (n + 1/2) =

a (ka) exp(i(n + 1/2)ka)

Z
a (n) =

a (ka) exp(inka)

Y d(k a )

Y d(k a )

(24)

(25)

We also define Fourier transformation of the classical field in a similar manner.


By this Fourier transformation, the anisotropy parameters are factorized in the action
(2)
.
Stot
(2)

(2)

Stot, ( ) = X Stot, (1)

(26)

(2)
(1) are already given by Iwasaki and Sakai [22] on the isotropic lattice and X are
Stot,
defined as
1 1 1
, , ,
1 1 1
, , ,

(27)
X =
1 1 1 .
, , ,

, , , 3
In this way the perturbative calculation of the anisotropic lattice becomes very systematic
and transparent.
For example, the free part of the anisotropic improved action S0 is given by
Z
1 X
(ka)G (ka),
(28)
S0 =
2 ,
k

534

S. Sakai et al. / Nuclear Physics B 584 (2000) 528542


1

Gii = qil kl2 + ki2 + qi4 k42 ,

3 2

2
=
G

44 q4l kl + k4 ,
1
Gij = {1 qij }ki kj ,

G4j = {1 q4j }k4kj ,

Gj 4 = G4j ,

k = 2 sin 2 k a ,

q = 1 c1 k2 + k2

q = 0,

( 6= ),

(29)

(30)

R
where Rthe 3 term in G44 results from the gauge fixing term and k stands for
Q

=1 dk a /2 . Note that off-diagonal elements of G vanish for the Wilson


action.
Propagators D are defined by

i
0
0
(31)
(ka)j (k a) = ij (2)4 (4)(ka + k a)D (ka),
and they are obtained by solving the equations,
G D = .

(32)

The FaddeevPopov propagator is given by


DF.P. (ka) =

ki 2 + 2 k4 2

(33)
0

The explicit forms of the S1 to S6 terms in Eq. (23) are obtained by setting c2 = c3 = 0
in formulae of Ref. [22] and by introducing the anisotropy factors as shown in Eqs. (26)
and (27).
(2)
( , , B ) over a and Faddeev
The effective action is obtained by integrating Stot
Popov fields a ,
Z
(2)
S(a B )
(34)
D(a )D(a )eStot (a ,a ,a B ) = eSeff (a B ) .
e
2 in S ,
Then C ( ) and C ( ) in Eq. (8) are obtained as coefficients of Fij2 and Fi4
eff
respectively.

3.3. Infrared divergence


We will discuss here a subtle point concerning the cancellation of the infrared
divergence. The contributions for C and C from the self-energy type diagram of the term
h(S1 + S2 )2 i have the infrared divergence. However in the difference given by Eqs. (9) and
(10) they are canceled. But numerically the calculation of the divergent integral is very
delicate problem. In the numerical evaluation, we discretize the momentum integration

S. Sakai et al. / Nuclear Physics B 584 (2000) 528542

535

d4 k, and the divergence comes from the segment including k = 0. We should not take
the difference of integration with different directly, because their measures of segment
are different for isotropic and anisotropic lattices. We use the following method for the
calculation of the Eqs. (9) and (10),


Imp.
Imp.
Imp.
C ( ) C (1) = C ( ) CStand. ( ) + CStand. ( ) CStand. (1)

Imp.
(35)
+ CStand. (1) C (1) ,
where C Imp. and C Stand. are the coefficients with improved and standard actions,
respectively.
Imp.
In the first term of r.h.s. of Eq. (35), C ( ) and CStand. ( ) have the same infrared
divergence and they are canceled exactly by each other. The second term of Eq. (35) can
be calculated by the analytic integration of the 4th component of the loop momentum [10].
The results do not include the infrared divergence and the numerical integration is stable.
The divergence in the last term has been already calculated in Ref. [22]. We have checked
that our calculations for the second and the third terms of Eq. (35) coincide with those for
Wilson and Symanzik case respectively given in Ref. [10] and Ref. [22]. 1
Imp.
Imp.
In this way the difference C ( ) C (1) is calculated in numerically stable
Imp.
Imp.
manner. Similar calculation has been done for C ( ) C (1).
Table 1
C and C for Symanzik, Iwasaki and DBW2 actions. Here 1C = C ( ) C (1) and 1C =
C ( ) C (1)
Symanzik action

Iwasaki action

DBW2 action

1C

1C

1C

1C

1C

1C

SU(2)

1.0
1.1
1.5
2.0
3.0
4.0
5.0
6.0

0.00000
0.00541
0.02072
0.03180
0.04304
0.04862
0.05168
0.05314

0.00000
0.00092
0.00173
0.00085
0.00103
0.00159
0.00059
0.00162

0.00000
0.00201
0.00908
0.01564
0.02406
0.02934
0.03340
0.03725

0.00000
0.00614
0.02327
0.03612
0.05016
0.05613
0.05692
0.05444

0.00000
0.01124
0.04875
0.08400
0.13384
0.16932
0.19695
0.21970

0.00000
0.01464
0.05150
0.07510
0.09537
0.09766
0.09055
0.08000

SU(3)

1.0
1.1
1.5
2.0
3.0
4.0
5.0
6.0

0.00000
0.00930
0.03602
0.05571
0.07596
0.08614
0.09182
0.09462

0.00000
0.00253
0.00701
0.00829
0.00856
0.00951
0.01216
0.01625

0.00000
0.00427
0.01852
0.03119
0.04693
0.05644
0.06350
0.07012

0.00000
0.01045
0.03967
0.06184
0.08679
0.09843
0.10163
0.09953

0.00000
0.02087
0.08983
0.15389
0.24348
0.30658
0.35533
0.39532

0.00000
0.02584
0.09206
0.13609
0.17729
0.18776
0.18221
0.17034

1 Although the values of C Stand. ( ) C Stand. (1) were not given in Ref. [10], we have calculated them from
Table and formulae in Ref. [10].

536

S. Sakai et al. / Nuclear Physics B 584 (2000) 528542

4. Results of one-loop calculation


Values of C ( ) C (1) and C ( ) C (1) are given in Table 1 for Symanzik, Iwasaki
and DBW2. is varied from 1 to 6, since these anisotropy parameters are often used in
Monte Carlo simulations on anisotropic lattices.
4.1. The ratio
When we calculate physical quantities, we must take into account the variation of the
scale a due to ( ). In weak coupling regions it is given by Eq. (11). The ratio is
calculated as a product of three factors,
Imp. ( )
Imp. ( )
Stand. ( ) Stand. (1)
=

.
Imp. (1) Stand. ( ) Stand. (1)
Imp. (1)

(36)

Table 2
parameter ratio for improved actions
Imp. ( )/Imp. (1)

SU(2)

Symanzik

Iwasaki

DBW2

1.0
1.1
1.5
2.0
3.0
4.0
5.0
6.0

1.00000
0.97614
0.90282
0.84653
0.78873
0.76309
0.75471
0.75778

1.00000
0.97803
0.92641
0.89562
0.86890
0.86566
0.88107
0.91159

1.00000
0.98187
0.98529
1.04907
1.23013
1.47071
1.77320
2.12140

Imp. ( )/Imp. (1)

SU(3)

Symanzik

Iwasaki

DBW2

1.00000
0.97599
0.90111
0.84350
0.78512
0.75955
0.75133
0.75483

1.00000
0.97809
0.92691
0.89583
0.86670
0.86010
0.87209
0.89981

1.00000
0.98234
0.99204
1.06597
1.26812
1.53177
1.86138
2.24211

Fig. 1. The ratio of parameter for SU(3).

S. Sakai et al. / Nuclear Physics B 584 (2000) 528542

537

The numerical results are given in the Table 2 for Symanzik, Iwasaki and DBW2 and are
shown in Fig. 1.
4.2. The parameter
The parameter is given by Eqs. (12) and (13). Because of the gauge invariance we can
see C (1) = C (1) at the isotropic case, and therefore these coefficients do not appear in
the definition of . But to cancel the infrared divergence we calculate Eq. (13) as 1 ( ) =
(C ( ) C (1)) (C ( ) C (1)). We plot 1 as a function of c1 in Fig. 2.
The parameter 1 changes sign around c1 = 0.18 0.19. This means that in weak
coupling regions, there is no renormalization for the anisotropy parameter for this action.
Table 3
The 1 for improved actions

SU(2)

Symanzik

Iwasaki

DBW2

1.0
1.1
1.5
2.0
3.0
4.0
5.0
6.0

0.00000
0.00633
0.02246
0.03266
0.04201
0.04702
0.05109
0.05476

0.00000
0.00816
0.03235
0.05177
0.07423
0.08548
0.09032
0.09170

0.00000
0.02588
0.10025
0.15911
0.22921
0.26698
0.28751
0.29971

SU(3)

Symanzik

Iwasaki

DBW2

0.00000
0.01183
0.04303
0.06400
0.08452
0.09565
0.10398
0.11088

0.00000
0.01472
0.05820
0.09304
0.13372
0.15487
0.16513
0.16966

0.00000
0.04671
0.18190
0.28998
0.42077
0.49434
0.53755
0.56567

Fig. 2. The 1 as a function of c1 for SU(3).

538

S. Sakai et al. / Nuclear Physics B 584 (2000) 528542

The interesting point is that the dependence of the with Iwasaki and DBW2 actions
is opposite to those with standard and Symanzik actions; As decreases decreases for
Iwasaki and DBW2 action, while it increases for standard and Symanzik actions. This is a
new feature and we shall confirm it non-perturbatively in the next section.
4.3. Anisotropy coefficients
The anisotropy coefficients, which are the derivatives of spatial and temporal gauge
couplings with respect to the anisotropy , are calculated as,



g2
Imp.
=
C ( ) CStand. ( ) +
CStand. ( ) CStand. (1) .

(37)

In this manner, we are free from the infrared divergence, and the numerical evaluation is
stable as in Section 3.3.
Table 4
The anisotropy coefficients for improved actions. The results are checked by estimating the derivative
numerically by using the results with = 1.05, 1.1
SU(2)
g2 / | =1
Symanzik

0.058406

SU(3)
g2 | =1
0.01196

g2 / | =1

g2 / | =1

0.10025

0.030365

Iwasaki

0.020569

0.067009

0.044305

0.113965

DBW2

0.117103

0.163543

0.21793

0.287592

Fig. 3. Anisotropy coefficients with the standard and improved actions for SU(3). Dotted line
represents the r.h.s. of Eq. (38), which is about 0.0697 for SU(3) case. The cross symbol stands
for the l.h.s. of Eq. (38).

S. Sakai et al. / Nuclear Physics B 584 (2000) 528542

539

From the invariance of the string tension on the isotropic and anisotropic lattice, Karsch
has derived the following sum rule [10],
g2 g2 11Nc
+
=
( = 1, a 0).
(38)

48 2
The same arguments are applied to improved actions. We show results in Table 4 and Fig. 3.
This sum rule is satisfied quite well.
An interesting point is that individual terms g2 / and g2 / also change sign
as c1 increases. For Iwasaki and DBW2 actions, g2 / becomes positive while
it is negative for standard and Symanzik actions. These anisotropy coefficients have
a contribution to QCD thermodynamics. For example, Okamoto et al. [24] used our
perturbation results to study the energy and pressure with Iwasaki action and had no
negative pressure problem, contrary to Wilson action case [25].

5. Numerical results in weak coupling regions


In the previous section we have found by the perturbative calculation that the ratio of
the renormalized and bare anisotropy, , becomes less than one for Iwasaki and DBW2
actions.
Since the is important in QCD simulations on the anisotropic lattice, we will study its
behavior further by numerical simulations.
Numerically the parameter is calculated from the relation [11,26],
= /B .

(39)

The anisotropy B appears in the action given by Eq. (5) while the renormalized anisotropy
is defined by
= a /a .

(40)

For the probe of the scale in the space and temperature direction, we use the lattice potential
in these directions, which is defined by


W (l, t)
,
(41)
V (B , l, t) = ln
W (l + 1, t)
where W (l, t) is the Wilson loop of the size l t in the temporal plane. Similar formula
holds for the potential in space direction. We fix = 2, and calculate the ratio at a few B
points,
V (B , l, r)
.
(42)
V (B , l, t)
Then we search for the point R = 1 by interpolating B [11,26]. In Ref. [27], an extensive
study is done for the determination of B .
The simulations are done on the 123 24 lattice. Numerical results with large region
( > 10) together with perturbative ones are shown in Fig. 4. They agree with each other
at large region. The parameter decreases as decreases for Iwasaki and DBW2 while
it increases for standard and Symanzik actions.
R(B , l, r) =

540

S. Sakai et al. / Nuclear Physics B 584 (2000) 528542

Fig. 4. Perturbative and non-perturbative results of as a function of . Data for Wilson actions are
taken from Ref. [26]

6. Concluding remarks
We have calculated the QCD scale parameter , which is shown in Table 2 and Fig. 1.
Ratios ( )/(1) for standard, Symanzik and Iwasaki actions have very similar behavior;
they are slightly less than one, but the behavior of ratio of the DBW2 is quite different
from those of other actions. DBW2 action is expected to be very near to the renormalized
trajectory [16], and may have a special feature. It may be interesting to study the c1
dependence of the parameters between Iwasaki and DBW2 actions in more detail.
The parameters and anisotropy coefficients, the derivatives of the coupling constants
with respect to the anisotropy parameter are calculated in one-loop order for improved
actions. We have found that the 1 in the Eq. (13) changes sign for c1 ' 0.18 0.19
as shown in Fig. 2. The 1 ( ) is positive for the standard plaquette and Symanzik actions
while it becomes negative for Iwasaki and DBW2 actions. This behavior is confirmed nonperturbatively by the numerical simulations in weak coupling regions as shown in Fig. 4.
We have also found that the anisotropy coefficients change sign as c1 increases. For
Iwasaki and DBW2 actions, g2 / is positive while for standard and Symanzik actions
it is negative. These may be good properties for the thermodynamics using those improved
actions [24].
We have found that obtained by the perturbation calculations is close to one in
the region c1 ' 0.18 0.19 for = 1 6. 2 A natural question is whether this
is true also at intermediate and strong coupling regions. Parts of the results on the
parameter are reported at Lattice99 at Pisa [23]. The study of the lattice spacing a on
the anisotropic lattice in intermediate coupling regions has been started. Detailed results
2 The value is sligtly shifted from that in Ref. [23]. After Lattice99, we have done an extensive study of the
numerical calculation of the loop integral, so that we can safely extrapolate to the continuum integral.

S. Sakai et al. / Nuclear Physics B 584 (2000) 528542

541

will be reported in the forthcoming paper. Moreover, with these fundamental properties
of the improved actions on anisotropic lattices, we are going to start simulations of heavy
quark spectroscopy, transport coefficients of quark gluon plasma etc. on these lattices.

Acknowledgements
We are grateful to R.V. Gavai for the discussion for thermodynamics quantities. This
work is supported by the Grant-in-Aide for Scientific Research by Monbusho, Japan
(No. 10640272). In this work, QCD Monte Carlo simulations have been done on SX-4
at RCNP, Osaka University, and on VPP500 at KEK (High Energy Accelerator Research
Organization) and Tsukuba University.

Appendix A
In this appendix we show all the data used in Eq. (35). In Table 5 the data of the first
term in Eq. (35) are summarized, and in Table 6 we report the data for the standard action.
The values of the third terms in Eq. (35) are given by isotropy cases in Table 5.
Table 5
Imp.
C and C for Symanzik, Iwasaki and DBW2 actions. Here C ( ) = C ( ) CStand. ( ) and
Imp.
C ( ) = C ( ) CStand. ( )
Symanzik action

Iwasaki action

DBW2 action

C ( )

C ( )

C ( )

C ( )

C ( )

C ( )

SU(2)

1.0
1.1
1.5
2.0
3.0
4.0
5.0
6.0

0.13173
0.13690
0.15149
0.16130
0.16908
0.17135
0.17218
0.17292

0.13173
0.12672
0.11355
0.10393
0.09261
0.08643
0.08359
0.08300

0.32668
0.33928
0.37625
0.40371
0.43115
0.44427
0.45223
0.45828

0.32668
0.31461
0.28349
0.26190
0.23844
0.22685
0.22221
0.22188

0.57505
0.59687
0.66429
0.72042
0.78929
0.83262
0.86414
0.88909

0.57505
0.55447
0.50362
0.47128
0.44160
0.43368
0.43694
0.44468

SU(3)

1.0
1.1
1.5
2.0
3.0
4.0
5.0
6.0

0.23211
0.24156
0.26836
0.28649
0.30099
0.30524
0.30674
0.30799

0.23211
0.22293
0.19832
0.17988
0.15771
0.14509
0.13848
0.13586

0.57349
0.59652
0.66429
0.71477
0.76527
0.78921
0.80344
0.81412

0.57349
0.55132
0.49302
0.45112
0.40373
0.37852
0.36606
0.36144

0.99987
1.03950
1.16198
1.26386
1.38820
1.46573
1.52166
1.56570

0.99987
0.96232
0.86701
0.80326
0.73962
0.71558
0.71187
0.71702

542

S. Sakai et al. / Nuclear Physics B 584 (2000) 528542

Table 6
C and C for the standard action. 1C ( ) = CStand. ( ) CStand. (1) and 1C ( ) = CStand. ( )
CStand. (1). We have evaluated these data from Table and formulae in Ref. [10]

SU(2)

1C ( )

1C ( )

1.0
1.1
1.5
2.0
3.0
4.0
5.0
6.0

0.00000
0.01058
0.04048
0.06137
0.08039
0.08824
0.09213
0.09433

0.00000
0.00593
0.01991
0.02865
0.03807
0.04370
0.04754
0.05035

SU(3)

1C ( )

1C ( )

0.00000
0.01875
0.07227
0.11009
0.14485
0.15928
0.16645
0.17051

0.00000
0.01171
0.04079
0.06051
0.08296
0.09652
0.10578
0.11250

References
[1] T. Hashimoto, A. Nakamura, I.O. Stamatescu, Nucl. Phys. B 400 (1993) 267.
[2] QCD-TARO Collaboration, Ph. de Forcrand et al., Nucl. Phys. B (Proc. Suppl.) 8384 (2000)
411.
[3] C. Morningstar, M. Peardon, Phys. Rev. D 56 (1997) 4043.
[4] T.R. Klassen, Nucl. Phys. B 73 (1999) 918.
[5] F. Karsch, H.W. Wyld, Phys. Rev. D 35, 2518.
[6] A. Nakamura, S. Sakai, K. Amemiya, Nucl. Phys. B 42 (1996) 432.
[7] A. Nakamura, T. Saito, S. Sakai, Nucl. Phys. B 63 (1998) 424.
[8] QCD-TARO Collaboration, Ph. de Forcrand et al., Nucl. Phys. B 63AC (1998) 460.
[9] Y. Nakahara, M. Asakawa, T. Hatsuda, Phys. Rev. D 60, 091503.
[10] F. Karsch, Nucl. Phys. B 205 [FS5] (1982) 285.
[11] G. Burgers, F. Karsch, A. Nakamura, I.O. Stamatescu, Nucl. Phys. B 304 (1988) 587.
[12] M. Garca Prez, P. van Baal, Phys. Lett. B 392 (1997) 163.
[13] K. Symanzik, Nucl. Phys. B 226 (1983) 187.
[14] M. Lscher, P. Weisz, Phys. Lett. B 158 (1985) 250.
[15] Y. Iwasaki, Report UTHEP-118, Univ. of Tsukuba, 1983; Nucl. Phys. B 258 (1985) 141.
[16] QCD-TARO Collaboration, Ph. de Forcrand et al., Nucl. Phys. B 577 (2000) 263.
[17] J. Engels, F. Karsch, H. Satz, I. Montvay, Nucl. Phys. B 205 [FS5] (1982) 545.
[18] S. Ejiri, Y. Iwasaki, K. Kanaya, Phys. Rev. D 58 (1998) 094505.
[19] A. Hasenfratz, P. Hasenfratz, Phys. Lett. B 93 (1980) 165.
[20] A. Hasenfratz, P. Hasenfratz, Nucl. Phys. B 193 (1981) 210.
[21] R. Dashen, D.J. Gross, Phys. Rev. D 23 (1981) 2340.
[22] Y. Iwasaki, S. Sakai, Nucl. Phys. B 248 (1984) 441.
[23] S. Sakai, A. Nakamura, T. Saito, hep-lat/0001004.
[24] M. Okamoto et al., CP-PACS Collaboration, Phys. Rev. D 60, 094510.
[25] B. Svetitsky, F. Fucito, Phys. Lett. B 131 (1983) 165.
[26] T.R. Klassen, Nucl. Phys. B 533 (1998) 557.
[27] J. Engels, F. Karsch, T. Scheideler, Nucl. Phys. B 564 (2000) 303.

Nuclear Physics B 584 (2000) 545588


www.elsevier.nl/locate/npe

D1D5 system and noncommutative geometry


Andrei Mikhailov ,1
California Institute of Technology, Pasadena, CA 91125, USA
Department of Physics, Princeton University, Princeton, NJ 08544, USA
Received 14 December 1999; revised 29 May 2000; accepted 30 May 2000

Abstract
Supergravity on AdS3 S 3 T4 has a dual description as a conformal sigma-model with the
target space being the moduli space of instantons on the noncommutative torus. We derive the precise
relation between the parameters of this noncommutative torus and the parameters of the near-horizon
geometry. We show that the low energy dynamics of the system of D1D5-branes wrapped on the
torus of finite size is described in terms of the noncommutative geometry. As a byproduct, we give
a prediction on the dependence of the moduli space of instantons on the noncommutative T4 on the
metric and the noncommutativity parameter. We give a compelling evidence that the moduli space of
stringy instantons on R4 with the B-field does not receive 0 -corrections. We also study the relation
between the D1D5 sigma-model instantons and the supergravity instantons. 2000 Elsevier Science
B.V. All rights reserved.
PACS: 11.25.-w; 11.30.Pb; 04.65.+e; 03.50.-z
Keywords: Ads/CFT correspondence; Attractors; Instantons; Noncommutative geometry

1. Introduction
The formalism of noncommutative geometry [1] is useful in string theory when one
studies certain special points in the moduli space of compactifications. In string theory,
small size manifolds are related to finite size manifolds by T-duality. However, the required
T-duality transformation usually does not have a nice limit when the size of the manifold
goes to zero. The form of the T-duality which brings the background with the small size
manifold to the background with the finite size manifold depends very irregularly on the
moduli of the small size manifold, such as the metric and the B-field. Therefore, one
could naively expect that the physics also does not have a regular limit. But it turns out
that there exists a smooth description of physics at very small distances in terms of the
andrei@viper.princeton.edu
1 On leave from the Institute of Theoretical and Experimental Physics, 117259, Bol. Cheremushkinskaya, 25,
Moscow, Russia.

0550-3213/00/$ see front matter 2000 Elsevier Science B.V. All rights reserved.
PII: S 0 5 5 0 - 3 2 1 3 ( 0 0 ) 0 0 3 6 5 - 5

546

A. Mikhailov / Nuclear Physics B 584 (2000) 545588

noncommutative geometry. This amazing fact was discovered in string perturbation theory
in [2] and in the study of the Matrix theory in [3]. The relation between the noncommutative
geometry and the string theory has been rigorously derived and made more precise in
[47]. The authors of [8] have found the zero slope limit in which the noncommutative
YangMills theory becomes a valid description of the open string sector of the string
theory, established the relation between the conventional and the noncommutative gauge
fields and found the stringy interpretation of the Morita equivalence [39]. Unfortunately,
all the successful applications of noncommutative geometry to string theory have so far
been limited to flat backgrounds, although the natural prescription for constructing the
noncommutative algebra from the background with the B-field is known in mathematical
literature [9,10].
Putting a field theory on the noncommutative space effectively introduces the nonlocality. This non-locality should presumably lead to softening some of the field theory
divergences (as discussed, for example, in [11]). One very important example of
divergences affected by the noncommutativity is the divergences in perturbation theory;
the study of the effects of noncommutativity in Feynman diagrams has been initiated in
[12]. The other example, directly related to our paper, is the resolution of singularities in
the instanton moduli spaces. It was pointed out in [13] that introducing noncommutativity
for the field theory on R4 is equivalent from the point of view of the instanton moduli
space to considering the ADHM construction with the nonzero value of the hyper-Khler
moment map. (In other words, it introduces the FayetIlliopoulos terms in the ADHM
sigma-model.) This discovery has an important application in the study of the string theory
on anti-de Sitter space. It was conjectured in [14], that this theory has a dual description
in terms of the conformal field theory. More precisely, the type IIB string theory on
AdS3 S 3 X, where X is K3 or T4 , is dual to the conformal sigma-model whose target
space is the moduli space of instantons on X. This proposal, suggested in [14] and further
studied in [1517], is valid at the specific subspace of the moduli space. On this subspace,
the fluxes of the NSNS B-field through the two-cycles of X are zero. It was found in [18],
that the non-zero fluxes of B correspond to the non-zero value of the moment map in the
ADHM construction. Therefore, the string theory on AdS3 S 3 X with the nonzero Bfield corresponds to the conformal sigma-model on the moduli space of instantons on the
noncommutative X.
The purpose of this paper is to fix the precise relation between the parameters of the
noncommutative SYM theory and the parameters of the near-horizon geometry. We will
proceed in the following way. Wrapping Q4 D5-branes on X, Qi2 D3-branes on the ith
two-cycle of X, and adding Q0 D-strings unwrapped, we end up having a black string in
the (5 + 1)-dimensional flat space. The background AdS3 S 3 X arises in type IIB as
the near-horizon limit of this black string. The corresponding supergravity solution looks
in a neighborhood of a given point in the (5 + 1)-dimensional space as R5+1 X, but
the moduli of X depend on the distance from the black string. For such a solution, let
us consider the geometry of X near the horizon (which we denote Xh ) and X far away
from the string (we denote it X ). Given X , one can find Xh , using the supergravity
equations of motion. It turns out that Xh cannot be arbitrary K3 (or T4 ); the moduli of Xh

A. Mikhailov / Nuclear Physics B 584 (2000) 545588

547

should satisfy certain attractor conditions [19]. On the other hand, the moduli of X can
be arbitrary. Therefore, the correspondence X 7 Xh is not one-to-one; there is a whole
family of X flowing to a given Xh .
The CFT dual to the near-horizon supergravity can be found using the procedure
suggested by Maldacena [14]. We look at the worldsheet theory of the corresponding
system of branes. In the low energy limit, it becomes a conformal theory. According to
the Maldacena conjecture, this is the conformal theory dual to AdS supergravity. The
worldsheet effective theory feels the flat background, the one which would be there without
branes. The geometry of this background is R1+5 X . Therefore the moduli of the
conformal theory depend on X . However, as we have already mentioned there are
many possible X which give the same near-horizon Xh . The low energy theories on
the worldvolume of the black string should be equivalent for all the backgrounds X
flowing to the same Xh . This is a necessary condition for AdS/CFT correspondence to
be self-consistent. It was argued in [14], that the equivalence of the conformal theories
corresponding to different X flowing to the same Xh follows from supersymmetry.
We want to ask the following question: given the family of flat backgrounds X
flowing to the same near-horizon background, can we find a canonical representative,
the one which gives the simplest description of the D1D5 worldsheet theory? At first
sight, the most appealing possibility would be to find X of a very large size. If we
could do that, we would get the description of our CFT as the sigma model with the
target space being the moduli space of YangMills instantons. But it turns out that most
of the attractors Xh cannot be obtained as the near-horizon limit of the solutions with
very large X . (In fact, there is an upper limit on the volumes of X flowing to a given
Xh .) What we can do instead is to find a representative with very small X . For such
backgrounds the low-energy effective action is formulated rather explicitly in terms of
the noncommutative geometry. We can use the results of [8] relating the shape of X
and the B-field fluxes to the parameters of the noncommutative manifold which we will
. According to [8], the effective low-energy theory for the bound state of D1,
g
denote X
R1+1 . Therefore, the effective
g
D3 and D5 is supersymmetric YangMills theory on X
low-energy theory for the black string is (1 + 1)-dimensional sigma-model with the target
. (In fact, there is a subtlety here which we
g
space the moduli space of instantons on X
will discuss later.) This gives the dual description of the near-horizon supergravity.
Let us mention an interesting consequence of this correspondence. It turns out that even
if we restrict ourselves to the small-size X , there is still the whole family of them flowing
to the same Xh . This gives us the family of noncommutative manifolds having the same
moduli space of instantons. Indeed, the corresponding sigma-models should be equivalent;
the equivalence of (4, 4) sigma-models implies that they have the same target spaces,
modulo some discrete symmetries; but the equivalence classes turn out to be connected,
therefore the possibility of discrete symmetries is excluded. It has been proven in [8] that
instantons on R4 depend only on the self-dual part of the noncommutativity parameter. Our
method allows us to formulate the analogue of this statement for instantons on T4 .
The relation between the near-horizon geometry and the moduli of the dual CFT has
been studied in great detail in the paper by Dijkgraaf [20]. The precise correspondence

548

A. Mikhailov / Nuclear Physics B 584 (2000) 545588

between the parameters of the attractor and the moduli space of the dual CFT has been
given in that paper. Combined with our paper, this gives the description of the geometry of
the moduli space of instantons on the noncommutative torus.
Another application of our formulas is the study of the low-energy dynamics of
D-branes wrapped on the torus of the finite size (of the order ls ). The most straightforward
way of doing it would consist of two steps: (1) study the low-energy worldvolume theory
of D5-branes wrapped on T4 (with the numbers of D1 and D3 specifying the topological
sector of this theory); (2) compactify this low energy theory on T4 , to get the low energy
theory of the black string. However, we think that performing the first step is very hard for
the finite size torus. We do not know how to construct the low-energy theory describing the
slow degrees of freedom of N > 1 branes of finite size in the nontrivial topological sector.
(The nonabelian BornInfeld theory [32] is valid when the covariant derivatives of the
field strength are small, and this condition is not satisfied in our case.) But we can bypass
the first step and go directly to the low energy dynamics of the black string. Indeed, our
considerations show that it is equivalent to the low energy dynamics of branes wrapped
on some very small torus. Therefore, the low energy Lagrangian for the black string
can always be represented as the sigma-model on the moduli space of noncommutative
instantons.
There is a subtlety in the relation between the instanton sigma-model and the lowenergy supergravity on AdS3 . The statement that the low-energy dynamics of the D1D5
system is described by the sigma-model on the moduli space of instantons follows from
the dimensional reduction of the (1 + 5)-dimensional classical super YangMills theory on
the noncommutative torus. Strictly speaking, we can treat this (1 + 5)-dimensional theory
classically only if it is weakly coupled on the compactification scale. On the other hand,
classical AdS supergravity is valid only if the radius of curvature of AdS space is large
enough. It turns out that these two conditions are incompatible: we cannot trust dimensional
reduction in the regime where AdS supergravity is valid. However, we conjecture that
the shape of the target space of the sigma-model describing the dynamics of the D1D5
system is independent of the string coupling constant. We give two arguments confirming
this conjecture. The first argument is based on considering the dependence of the sigmamodel for the D1D5 system on the parameters of the background. Our considerations in
Section 3 will imply that the structure of equivalence classes of backgrounds giving the
same sigma-model does not depend on gstr . Although this fact does not necessarily imply
that the structure of the target space does not dependent of gstr , we think that it supports the
conjecture. The second argument is based on the relation between the worldsheet instantons
of the D1D5 sigma-model and the D-instantons of the six-dimensional supergravity. An
example of this relation was given in [25]. Instantons in toroidally compactified string
theories where considered in [2224]. We argue that under certain conditions there is a
correspondence between these two kinds of instantons, and the action of the worldsheet
instantons is equal to the action of the supergravity instantons. The action of the worldsheet
instantons is related to the hyper-Khler period map of the target space. And the action of
the supergravity instantons can be found when gstr is small from supergravity. The formula
agrees with the period map for the target space conjectured in [20] and does not contain N .

A. Mikhailov / Nuclear Physics B 584 (2000) 545588

549

This suggests that the target space for small gstr N and large gstr N should have the same
period map. Therefore the hyper-Khler structure of the target space is the same in the
regime when we can use dimensional reduction and in the regime when we can trust
supergravity.
It was claimed in [21] that the moduli space of the twisted little string theories on
T3 is the moduli space of noncommutative instantons on T4 . This result was derived by
relating the theory on NS5 brane to the D2D6 system, with the D6-brane wrapped on
the four-torus. Strictly speaking, D2D6 system on T4 is described by the noncommutative
YangMills theory only when this four-torus is very small. However, the situation with
the D2D6 system should be similar to what happens with D1D5: namely, the moduli
space does not depend on some combination of the background fields, and there is a small
torus in each equivalence class. Therefore, the moduli space is always the moduli space of
noncommutative instantons.
The paper is organized as follows. In the second section, we give a brief review of
the attractor equations and explain when the supergravity approximation can be trusted.
In the third section, we give the argument for independence of the sigma-model on
certain combinations of the background fields, based on supersymmetry. This section is
auxiliary, and it is not necessary for understanding the rest of the paper. Our arguments
based on supersymmetry should be closely related to the argument given in [30], but
we have tried to make them more precise. In the fourth section, we derive the relation
between the asymptotic background and the near-horizon (attractive) background. This
section is somewhat technical. The main formula is (56), giving the condition for the two
backgrounds to flow to the same attractor. The formula (63) expresses the near-horizon
RamondRamond fields in terms of the RamondRamond fields at infinity. In the fifth
section, we review the correspondence between the string theory on the small torus and the
YangMills theory on the noncommutative torus. Combining these results with the results
from the fourth section, we construct the flows on the moduli space of the noncommutative
tori, which leave the moduli space of instantons invariant. In Section 5.2 we make an
observation which suggests that the moduli space of instantons on noncommutative R4
does not receive 0 -corrections. In the sixth section we discuss the relation between the
worldsheet instantons and the supergravity instantons. In Appendix A, we review the
correspondence between the moduli of the type IIB compactification and the points of
Grassmanian.

2. Attractor equation
2.1. Supergravity solution
The dependence of the moduli of K3 or T 4 on the distance from the horizon is given by
the attractor equation. This equation can be found in many papers, for example in [26],
although in somewhat implicit form. In this section, we will briefly review the attractor
equations and present them in the form most convenient for our purposes.

550

A. Mikhailov / Nuclear Physics B 584 (2000) 545588

We will concentrate on the case X = K3. First of all, we want to review some properties
of the chiral N = 4, D = 6 supergravity. We will follow the original paper [27] but use
slightly different language. This theory describes an interaction of N = 4b gravity multiplet
with n tensor multiplets. For the compactification of type IIB on K3, we have n = 21.
The gravity multiplet contains a graviton, four left-handed gravitini and five antisymmetric
tensors with selfdual field strength. The tensor multiplet consists of an antisymmetric
tensor with antiselfdual field strength, four right-handed symplectic Majorana spinors and
five scalars. To describe the interacting theory, we think of the scalars as parametrizing the
coset
SO(5, 21)
.
Gr(5, 5 + 21) =
SO(5) SO(21)
We represent this coset as the manifold of five-dimensional positive planes in the space
L = R5+21 .
For a given five-plane W L, we denote P+ the projector on this plane, and P the
projector on the orthogonal plane. We denote the (1 + 5)-dimensional Minkowski space M.
The field strengths of the gravity multiplet and the tensor multiplets may be combined in a
vector H 3 M L, with the constraint
H = (P+ P )H.

(1)
1/2
S M

be the
Instead of writing P+ H and P H , we will often write H+ and H . Let
3/2
bundle of antichiral spinors on M, and S+ M the bundle of RaritaSchwinger fields (i.e.,
chiral spinors with vector indices). Also, we denote W+ the tautological bundle on the
Grassmanian Gr(5, 5 + 21), that is the bundle whose fiber over the point represented by
the plane W is the plane W itself. Similarly, W will denote the bundle whose fiber is
W . The gravitino and the fermions from the tensor multiplets are the sections of
the following bundles:
 
 3/2
S+ M S(W+ ) R ,


 1/2
(2)
S M S(W+ ) R W .
Here S(W+ ) is the spinor bundle associated with the Spin(5) vector bundle 2 W+ (we
will call its sections internal spinors). The subindices R mean that some Majorana
3/2
conditions are imposed. Let us explain the Majorana conditions for . (S+ M S(W+ ))R
P
3/2
is generated by the expressions of the form a sa a , where sa S+ M, a S(W+ ),
with the reality condition
X
X
sa a =
C(1) saT C(2) aT .
a

We have denoted C(1) and C(2) the charge conjugation matrices for the space-time spinors
and the internal spinors, respectively. The Majorana condition for is defined in the same
way. The supersymmetry generators  are spacetime and internal spinors:
2 The moduli space of scalars is topologically trivial (the global coordinates may be found, for example, in
[28]). Therefore, there is no difference between the Spin(5) bundles and the SO(5) bundles.

A. Mikhailov / Nuclear Physics B 584 (2000) 545588

 1/2
 
 S+ M S(W+ ) R .

551

(3)

The supersymmetry transformations for the gravitino and are:


1
 = D  + 0 (H+ )( ),
4
1
1
(4)
 = ( )  0 H .
2
12
Here 0 is the asymptotic value of the six-dimensional string coupling constant, 0

gstr / VK3 (we will not need the precise expression for it). In the transformation law for ,
the vector H+ W+ acts on S(W+ ) index of  as gamma-matrices act on spinors. D is
a covariant derivative, which includes the natural connection on W+ . In the expression for
, denotes the point in the moduli space of scalars. Its derivative acts on  as
follows. There is a natural isomorphism between the tangent bundle to Gr(5, 5 + 21) and
W W+ . Indeed, a tangent vector to Gr(5, 5 + 21) is an infinitesimal rotation W of the
plane W , and it is completely specified by saying what is w W for each vector w W .
. Since L is equipped with a
This gives us a map from W to W , or a section of W W+

metric, we can identify W+ with W+ . This means, that we can think of d as an element of
1/2
W W+ 1 M. Then, to act on  ([S+ M S(W+ )]R ) by W W+ , we
use the action of W+ on S(W+ ) (gamma-matrices act on spinors) and tensor multiplication
by W . After acting by the spacetime gamma-matrix we get a section of the bundle
1/2
(S M S(W+ ))R W , which is where lives.
We want to find a supersymmetric background, corresponding to our black string. This
means that for some parameters , we should have  = 0 and  = 0. For our solution,
there will be eight linearly independent  satisfying these equations. We will try the
following ansatz for the metric:

(5)
ds 2 = e2U (r)(dt 2 dx 2 ) e2U (r) dr 2 + r 2 d32 .
Let us introduce the vielbein {et , ex , er , e1 , e2 , e3 }, so that
ds 2 = (et )2 (ex )2 (er )2

3
X

(ei )2 .

i=1

We use the following ansatz for the three-form field strength:


e3U
ij k Z,
r3
e3U
(6)
Ht xr = 3 Z .
r
The vector Z should be integer-valued, Z 5,21 L, because the fluxes of H through S 3
are integers. This vector should be identified with the charge of the black string. We look
for supersymmetry with the parameter  which depends only on r. Also, we assume that
the scalars depend only on r. First, let us write explicitly the transformations of various
components of . We will denote the covariant derivative on W+ . We do not need to
know explicitly what is the spin connection in W+ , only the spacetime spin-connection
will be important for us. We have the following conditions for = 0 (with U 0 = dU/dr):
Hij k =

552

A. Mikhailov / Nuclear Physics B 584 (2000) 545588

1
1 e3U
t = U 0 eU t r  + 0 3 x r Z+  = 0,
2
2 r
1 0 U x r
1 e3U
x = U e  0 3 t r Z+  = 0,
2
2 r
1 e3U t x
r = r  + 0 3 Z+  = 0,
2 r
1 0 U i r
1 e3U
i = U e  + 0 3 ij k j k Z+  = 0.
2
4 r

(7)

Let us denote
b+ = p Z+ .
Z
||Z+ ||2
Then, it follows from the first equation in (7), that
q
e2U
2 ||Z+ ||2 ,
r 3 0
b+ x t  = 0.
Z

U0 =

(8)

The second row in (7) follows from (8), and the third gives the equation for the radial
dependence of :
1 e3U
r (r) = 0 3 (r).
2 r

(9)

b+ will be covariantly
b+ x t ) = 0 because for our solution Z
This is consistent with (Z
b+ (r) = 0). The last equation in (7) follows from (8) if one takes into
constant (that is, r Z
account that  is chiral.
Now we turn to the variation of the fermions from tensor multiplets. We have:
e3U
1
= eU r (r )  3 0 1 2 3 Z  = 0.
2
r

(10)

Taking into account the chirality condition 123  = t xr  and Eqs. (8) one can see that
= 0 if satisfies the following equation:
d = 2dU

1
Z Z+ .
||Z+ ||2

(11)

This equation has a very clear geometrical meaning. Remember that we represent tangent
vectors to the moduli space of scalars as linear maps from W to W , giving the variations
of vectors in W . Suppose that we took a vector w W which is orthogonal to Z. Then
(Z+ w) = 0 and (11) tells us that the variation of w is zero. Therefore, the subspace
W Z W remains constant. We can represent W = (W Z )RZ+ . When r changes
(W Z ) remains constant and the one-dimensional subspace RZ+ gets rotated in the
plane RZ+ RZ :

A. Mikhailov / Nuclear Physics B 584 (2000) 545588

553

In particular, we see that the near-horizon background is represented by (W Z ) RZ.


This is a qualitative picture. Now we want to write the explicit formula for (r). Let
b+ =
us introduce the hyperbolic angle between the vector Z and the plane W . Let Z
2
1/2
2
1/2
b
Z+ and Z = (||Z || )
Z . Then,
(||Z+ || )
b
b = cosh Z
b+ + sinh Z
Z

(12)

(this is the definition of ) and


b
b+ = 2 tanh dU Z
dZ

(13)

b+ W+ ). Combining these two equations, we


(this is d from (11) applied to the vector Z
have
d = 2 tanh dU.

(14)

This equation together with the second equation from (8) determines the dependence of U
and on r:
q
e2U
U 0 = 3 02 ||Z||2 cosh ,
r
q
e2U
(15)
0 = 2 3 02 ||Z||2 sinh .
r
Adding and subtracting these two equations, we get:
q
e2U
d
(2U ) = 2 02 ||Z||2
.
(16)
dr
r3
The result of integration depends on two constants, c+ and c :
p
p

 

0 ||Z||2 1/2
0 ||Z||2 1/2
2U
= c+ +
,
e
c +
r2
r2
p


0 ||Z||2 + c r 2 1/2
p
.
e =
0 ||Z||2 + c+ r 2

(17)

554

A. Mikhailov / Nuclear Physics B 584 (2000) 545588

We want U = 0 at spatial infinity, therefore c+ c = 1. The ratio c+ /c is related to the


asymptotic value of moduli. This gives:
c = e .

(18)

Two examples
(1) The system of Q0 D-strings and Q4 D5-branes with B = 0. (The corresponding
supergravity solution has been found in [29].) The plane W is generated by three vectors
orthogonal to Z and the vector (1, 0, V ). In this situation only the volume of K3 changes
with r, the shape remains fixed. We have:
s
s
s
s
!
!


Q4
Q0
Q4
Q0
1
, V = cosh
,
,
+ sinh
.
(19)
Q0
Q4
Q0
Q4
V
Therefore,
e

(r)

s
=

s
Q4 V (r)
,
Q0

c+ =

Q4 V
.
Q0

(20)

Now, (17) gives

r

 


Q4 1/2
2
Q0 1/2

,
1+
e2U (r) = 1 + 2V 0 2
0
V r2
r
q
1 + V2 0 Qr 20
.
V (r) = V

1 + 2V 0 Qr 24

(21)

(2) Although we will not consider adding symmetric branes and fundamental strings in
this paper, we will now give the corresponding solution as an example. Consider a system
of q0 fundamental strings and q4 NS5-branes wrapped on X. Using the definition (12) of
and formula (125) from Appendix A, we have the equation for (r):

r
r

r 
r 
q4
q0
q4
q0
1
,
,
= cosh (r)
+ sinh (r)
.
(22)
(r),
(r)
q0
q4
q0
q4
q
This gives c+ = 10 qq40 and the following dependence of the metric and the coupling
constant on r:


 

2q4 1/2
2q0 1/2
,
1 + 02 2
e2U (r) = 1 + 2
r
r

2
2
2 r + 2q4
.
(23)
(r) = 0

r 2 + 202 q0
Lagrangean interpretation
The attractor equations have a Lagrangean interpretation. Indeed, let us introduce the
time = 1/r 2 . The attractor equations (15) imply that:

A. Mikhailov / Nuclear Physics B 584 (2000) 545588

1
d 2U
= 02 ||Z||2 e4U (cosh2 + sinh2 ),
2
d 2
d 2
= 2e4U 02 ||Z||2 sinh cosh .
d 2
These second order equations may be derived from the following Lagrangian:
 


dU 2 02 ||Z||2 4U
1 d 2
e (cosh2 + sinh2 ).
+
+
L=
4 d
d
4

555

(24)

(25)

The potential energy is


1
Epot = 02 e4U (||P+ Z||2 ||P Z||2 ).
4

(26)

The expression ||P+ Z||2 ||P Z||2 = 2||P+ Z||2 ||Z||2 is, up to the coefficient and the
constant term, the square of the tension of the black string. At the attracting point, that is
when P+ Z = Z, this tension reaches its minimum. Also, we can see from our solution (17),
that e4U is zero at the horizon. Therefore, the motion in , corresponding to the geometry
of the black string, may be considered as climbing up the potential. The initial velocity
should be adjusted in such a way that getting to the top requires infinite time.
In what follows, we will not need the full solution for U (r) and W (r). The only thing
important for us is that W changes with r in such a way that W (r) Z remains constant,
and also that the near-horizon W is W h = (W Z ) RZ.
2.2. When can we use the supergravity approximation?
We want to be able to use supergravity equations until we reach small enough values
of r, where the moduli are close to their fixed values. The low energy effective action
does not receive any quantum corrections because of the supersymmetry. Therefore, the
supergravity approximation is valid when we can neglect higher derivative corrections. We
will assume that the near-horizon string coupling is weak, therefore the main source of
higher derivative corrections is the string perturbation theory. These corrections behave
like a positive power of lstr /r where r is the typical radius of curvature. Therefore, the
condition that supergravity is applicable is that the size of the AdS region is much larger
then lstr . Let us explain what we mean by the size of the AdS region. The metric for the
D1D5 solution contains factors of the form
p
p



e ||Z||2
e ||Z||2
1 + 0
.
1 + 0
r2
r2
We are in AdS regime when we can approximate this expression as 02 ||Z||2/r 4 . Assuming
> 0, we see that the AdS regime starts at
q
2
= 0 e ||Z||2 .
(27)
r 2 < rcr
The condition for supergravity to be valid is that rcr  lstr . Now let us estimate lstr in the
near-horizon region. It is convenient to start with estimating the size of the core of the
D-string. For the typical attractor, all the moduli of the near-horizon torus are of the order

556

A. Mikhailov / Nuclear Physics B 584 (2000) 545588

one, therefore all the black D-strings (for example, D1, D3 wrapped on a two-cycle and
D5 wrapped on T4 ) have approximately the same string length and the same size of the
core. For the typical D-string in the typical background the size of the core is
2
2
rcore,
D gstr lstr .

(28)
p
As a typical D-string, let us choose the string with the charge Z/ ||Z||2 . This charge is
usually not integer, but we need only the supergravity solution. It has the metric with the
characteristic factor
p

p


1
1
||Z||2
||Z||2

+
+
1 + 0 e
. (29)
1 + 0 e
rE2
(Er rE0 )2
rE2
(Er rE0 )2
One can see from this expression that the size of the core for the D-string at the coordinate
distance r0 from the origin is of the order
q
2
2

r
||Z||2 .
(30)
rcore,
D
0
This equation together with (28) implies
2
(r0 ) p
lstr

r02
||Z||2gstr (r0 )

Since = const. for the D-string, we have


r

gstr
V (r0 )
g

.
gstr (r0 ) =
V str
V
This implies that for r0 rcr we have:
p
2
||Z||2
gstr
rcr

.
2 (r )
lstr
V
cr

(31)

(32)

(33)

(Remember that we measure all the lengths for the torus in string units.) The right-hand
side of this expression has a clear physical meaning. It is the t Hooft coupling constant for
f4 on the scale of the size of T
f4 . (Here
the (1 + 5)-dimensional super YangMills on R1,1 T
4
f
4
T is the noncommutative torus dual to the torus T at infinity as explained in Section 5.)
Our estimate (33) tells us that in the regime when we can trust the supergravity solution
in the AdS region, the six-dimensional noncommutative YangMills describing the D1D5
system moduli poise necessarily strongly coupled on the compactification scale. Therefore,
we cannot trust the classical dimensional reduction. We conjecture that the shape of the
target space of the sigma-model on R1,1 which is the dimensional reduction of this sixdimensional theory actually does not depend on the six-dimensional coupling constant.
We will give some evidence confirming this conjecture in Section 6.

3. Argument based on supersymmetry


In this section we will prove using supersymmetry that two backgrounds flowing to the
same attractor give the same low energy effective theory on the worldvolume of the black

A. Mikhailov / Nuclear Physics B 584 (2000) 545588

557

string. Our argument is very close to the one used in [30] to prove independence of the
hypermultiplet metric on the scalars in the vector multiplet.
We will study some auxiliary brane configuration. It consists of the system of N parallel
black strings, all having the same charge Q. This system has a Coulomb branch. Let us
consider the submanifold of this Coulomb branch, where N 1 black strings sit at one
point (the origin), and one is moving around:

We will call that black string which is moving around the single black string. We want
to look at the worldsheet theory of this single black string. The low-energy worldsheet
c of this sigma-model. It can be
theory is some sigma-model. Consider the target space M
4
parametrized by the position in the transversal R plus the internal degrees of freedom,
which specify the instanton
on T4 . Notice that the size of the moduli space of instantons

4
distances
on T is of the order V /gstr and does not depend on N , while the characteristic p
in the transversal R4 are growing with N . (From our solution, we see that R 2 N ||Z||2 .)
Therefore, in the limit N , we can consider the internal degrees of freedom as fast,
and the motion along R4 as slow. Moreover, in this limit the single black string sitting
at some point x R4 does not feel that the background changes with the distance from
the origin. Therefore, the configuration space of the internal degrees of freedom should be
approximately the same hyper-Khler manifold as for the black string in flat space. In other
c as a bundle with the base R4 and the fiber
words, in the limit N we can think of M
the moduli space M(x) of internal degrees of freedom of D1D5 wrapped on T4 (x):
M 4
c
R .
M

(34)

The dependence of T4 (x) on the point x R4 (and therefore the dependence of M(x)
on x) is dictated by the supergravity solution for N 1 black strings sitting at the origin.
Our arguments will go as follows. We will prove, using supersymmetry, that although the
moduli of the torus T4 depend on the distance from the origin, the fiber M does not. But
in the limit N , the fiber M(x0 ) over the given point x0 R4 is the same as the target
space M0 for the D1D5 in the flat background with the torus T4 (x) T4 (x0 ). Therefore,
independence of M(x0 ) on x0 implies that the target space for D1D5 is the same for all
the backgrounds T4 = T4 (x0 ), x0 R4 . Since T4 (x0 ) depends only on the distance r of
x0 from the origin, we had proven that the moduli space of D1D5 is the same for the
backgrounds from the given one-parametric family T4 (r), r [0, ]. In particular, it is

558

A. Mikhailov / Nuclear Physics B 584 (2000) 545588

the same for T4 (r = 0) and T4 (r = ). This implies, that all the backgrounds at infinity
r = which flow to the same near-horizon T4 (r = 0) give the same moduli space.
Let us proceed with the proof. We want to understand what kind of restrictions the
c It is useful to start with
supersymmetry imposes on the geometry of our target space M.
considering the flat sigma-model, which describes the dynamics of D1D5 in flat space
(without N 1 strings at the origin). In this case, we would have:
c 0 = R4 M 0 ,
M
where M0 does not depend on the point of the base. Both R4 and M0 are hyper-Khler,
and therefore we get the system with (4, 4) supersymmetry. In fact, there are two hyperKhler structures on R4 : one with the self-dual hyper-Khler forms, and the other with antiself-dual. Closer examination of the supersymmetry transformations shows that they are
different in the left- and right-moving sectors: the left-moving (4, 0) supersymmetry uses
self-dual complex structures on R4 , and the right-moving one uses anti-self-dual complex
structures.
When we introduce N 1 black strings at the origin the flat target space gets corrected;
c should
for example, the metric on R4 is not flat anymore. Nevertheless, our manifold M
still support a sigma-model with (4, 4) supersymmetry. This follows from the fact that
our single black string preserves the same supersymmetry as those N 1 black strings
which we have added. Therefore, introducing these N 1 black strings does not break
any more supersymmetry. The (4, 4) supersymmetry implies the existence of six complex
structures, three for the left-moving sector, and three for the right-moving [31]. The sigmamodel action can include torsion, which is an antisymmetric tensor b with db 6= 0. The
Lagrangian is:
L = g (dX dX ) + 2ga (dX dXa ) + gab (d a d b )
+ b dX dX + 2ba dX d a + bab d a d b .

(35)

Here we denote X the coordinates in R4 , and a the coordinates in the fiber (the moduli
space of instantons). The complex structures should be covariantly constant, but with
respect to the modified connection, with the torsion T = db added. The sign of the torsion
is opposite for the left and the right connections. We will write this condition explicitly for
the corresponding Khler forms:
I
I
I
T
T
= 0

(36)

(the index I = 1, 2, 3 distinguishes between the three different Khler forms, and the sign
distinguishes between the right and the left sectors). In our situation, we know something
about these six complex structures from comparison to the flat case. We know that in the
vicinity of a given fiber, the Khler forms are:
I
I
dX dX + (0)ab
d a d b + small corrections.
I = (0)

(37)

A. Mikhailov / Nuclear Physics B 584 (2000) 545588

559

I
Here (0)
are the basic self-dual (with plus) or anti-self-dual (with minus) forms on R4 ,

I
are the Khler forms on flat M0 . The corrections are small, because locally our
and (0)ab
background is almost flat. 3
We also have some information about the torsion. First of all, the torsion is proportional
to the background RamondRamond fields. Indeed, the corresponding term in the black
string effective action violates parity. In the brane picture, parity turns a black string into
an anti-black string. It is the same as changing the sign of the RamondRamond fields.
Therefore, the torsion should be zero if the RamondRamond fields are turned off. Also,
we know that in flat case the torsion is zero. 4 Now we want to see what happens to the
torsion when we put N 1 black strings at the origin in R4 . First of all, the components of
the torsion along R4 become nonzero. For example, D5-branes from the pile at the origin
create B RR , which couples to D1 from the single black string. Closer examination shows
that Tij k N||Z||2 ij k and the other components in R4 are zero. Also, we should admit
the possibility that the component Tabr is not zero any more. Indeed, the RamondRamond
fields along T4 change when we move in the radial direction. This implies that theta-angles
on M may change when we move in R4 . This precisely means that Tabr 6= 0. On the other
hand, we expect that Tabc is still zero, or at least very small when N . Indeed, we have
seen that Tabc is zero when we put our black string in flat background. The non-flatness of
the background can be locally described by the gradient of the background fields, which
we can schematically denote ( denotes the background fields). From the rotational
symmetry, and assuming the analytic dependence of the torsion on the background fields,
we estimate:

Tabc g .

(38)

In other words, it is of the second order in the gradient of the background fields. Therefore
we will neglect it. Also, the other components of the torsion (for example, Tjia , where i
and j are indices from the tangent space to S 3 ) are zero because of the rotational symmetry
of R4 .
Now we want to consider the variation of the period map (the integrals of the Khler
forms over the two-cycles in M), when we move in the radial direction. Let us consider
two points, one at the distance r1 from the origin, and the other at the distance r2 (both
3 The left Khler forms on M coincide with the right Khler forms, therefore it might seem strange that
0
I + and I . The reason why we distinguish them
we are using the superscript to distinguish between (0)ab
(0)ab
is the following. The forms I should be covariantly constant (with modified connection). In particular, they
should be covariantly constant along the radial direction in R4 . Although the three-dimensional space generated
I+
I
and (0)ab
coincide, we do not a priori exclude the possibility that the covariantly constant bases in
by (0)ab
these spaces are different in left and write sectors.
4 Indeed, the components of the torsion along R4 are not allowed in flat case because of SO(4) invariance.
The components of the torsion along M0 , Tabc , are also zero for the following reason. In the flat case, the only
RamondRamond fields are the constant fields along the torus. We know how they couple to our D1D5 system
from supergravity. They couple to the topological numbers. Therefore, the corresponding contribution to the black
string effective action cannot change under the small deformations of the fields. In other words, these constant
RamondRamond fields generate theta-angles in our sigma-model, but not a torsion.

560

A. Mikhailov / Nuclear Physics B 584 (2000) 545588

on the same line with the origin). Consider the cycle c(r1 ) in M at the first point, and the
same cycle c(r2 ) in M at the second point. We have, according to the Stokes theorem,
Z
Z
Z

I
I
I

=
dx dx dx T[ ]
.
(39)
c(r2 )

c(r1 )

cI[r1 ,r2 ]

The leading term on the right hand side is


Z
Z
I
c
d a d b Tr[a
b]c
.
dr

(40)

c(r)

We know that the space of periods over the fiber of I + coincides with that of I , modulo
small corrections. On the other hand, we see from (40) that the change in the periods of I +
and the periods of I as we move along r is opposite. The only possibility to reconcile
c I is either zero or some combination AI J . In any
these two observations is that Tr[a
J
b]c
case, it follows that the space of periods does not depend on r.
The condition
I
c
b]c
= linear combination of J
Tr[a

(41)

is a restriction on the torsion. This restriction implies that the hyper-Khler structure of the
c is zero in other words,
fiber does not depend on r. It is automatically satisfied when Tra
when the theta-angles on M(x) do not depend on the position in x R4 . We believe that
this is what actually happens:
b
= 0.
Tra

(42)

Indeed, our analysis of supergravity solution shows that of the eight RamondRamond
fields on the torus, there are seven linear combinations which do not depend on r.
It is natural that the seven worldsheet theta-angles depend precisely on these seven
combinations of the RamondRamond fields. We will check it in the case when the size
of the torus is very small at the end of Section 5, using the methods of noncommutative
geometry.

4. Near horizon geometry


In this section we will study the correspondence between the background fields far away
from the black string (we will call them asymptotic background) and the background
fields at the horizon. The answer is given by the formulas (52) and (54). We will explain
under which conditions two different asymptotic backgrounds flow to the same nearhorizon background (Eqs. (56)). We will show that for each background there exists a
background with the small size torus which flows to the same near-horizon geometry. To
simplify the discussion, we will first turn off the RamondRamond fluxes and study only
the perturbative moduli, parametrized by Gr(4, 4 + 20). We will include the Ramond
Ramond fields at the end of this section. The correspondence between the Ramond
Ramond fluxes at infinity and at the horizon is given by (63).

A. Mikhailov / Nuclear Physics B 584 (2000) 545588

561

4.1. Moduli of string perturbation theory


We will use the correspondence between the string theory backgrounds and the
planes W , which was first discussed in [33] and then made more precise in [34] and [20].
The plane W is generated by the vectors vE and v 0 , given by:
vE = (0, ,
E B ),
E


v 0 = 1, B, V 12 B B .

(43)

Here
E = (1 , 2 , 3 ) are the three Khler forms, V is the volume and B is the B-field.
Now suppose that we are given the charge vector Z:
Z = (Q4 , Q2 , Q0 ).

(44)

According to the previous section, the near-horizon geometry corresponds to the plane
W h = (W Z ) RZ. Let us describe this W h explicitly.
We begin with introducing a useful notation. The space 2+ T4 of self-dual forms on our
torus is generated by i . Let us consider the two-dimensional subspace




(45)
C = 2+ T4 (Q2 Q4 B) = 0 .
We have denoted this space C for the following reason. As explained in [33], the choice
of the two-dimensional subspace in H 2 (X, R) (where X is K3 or T4 ) is equivalent to the
choice of the complex structure up to a sign (to fix a sign, we have to orient the two-plane).
Therefore, the charge vector Z gives our torus a preferred complex structure modulo sign,
C . The orthogonal space to C in 2+ X is one-dimensional. Therefore, there are two
ways to choose the Khler form, which differ by a sign. The choice of the sign of the
Khler form also fixes the sign of the complex structure. (In the case of torus this may
be understood as follows: we need the value of the Khler form on any tangent bivector
to any holomorphic surface to be positive.) Our preferred complex structure has a clear
geometrical meaning in the case when B = 0. In this case, the bound state of D1D5 may
be described as an instanton configuration on U (Q4 ) gauge theory with the topological
charges (Q2 , Q0 ). The corresponding field strength is of the type (1, 1) in the complex
structure selected by the condition (Q2 C ) = 0. Our definition (45) is the generalization
of this complex structure for B 6= 0.
Now we describe W Z . This space contains a two-dimensional subspace



(46)
WC = 0, , (B ) C .
Since W Z is three-dimensional, we need one more vector. Choose R 2+ T4 , which
satisfies two conditions:
(1)

R C ,

||R ||2 = 1.

The two-form V R is the Khler form up to sign. The vector




1
0
vR = 1, B + sR , V B B sB R ,
2
(2)

(47)

(48)

562

A. Mikhailov / Nuclear Physics B 584 (2000) 545588

where
s=

1 ||Q2 Q4 B||2 ||Z||2 2Q24 V


2Q4
((Q2 Q4 B) R )

(49)

belongs to W Z . Moreover, it is orthogonal to WC . Therefore, W Z is generated by


0 . The near-horizon W h is generated by
the two-dimensional space WC and the vector vR
0 and Z:
this W Z and Z. Let us introduce the following linear combinations of vR
h
=q
vR

v0h =

0 Z
Q4 vR

||Z||2 + Q24 (2V + s 2 )

0 + Q (2V + s 2 )Z
||Z||2 vR
4

||Z||2 + Q24 (2V + s 2 )

(50)

h is of the form
They are useful, because they are orthogonal to each other and WC , vR
(0, , ) and ||vR ||2 = 1, and v0h is of the form (1, , ). It follows from the relation (43)
between the background fields and the plane W that the near-horizon Khler form, B-field
h and v h in the following way:
and volume are related to the components of vR
0

h
h
vR = 0, R , ,

(51)
v0h = 1, B h , ,

||v0h ||2 = 2V h .
The near-horizon background as a function of the background at infinity
This gives the following expressions for the near-horizon background fields in terms of
the asymptotic background fields:
hC = C ,
Q4 (B + sR ) Q2
h
,
=q
R
||Z||2 + Q24 (2V + s 2 )
Bh =
2V h =

||Z||2(B + sR ) + (2V + s 2 )Q4 Q2


,
||Z||2 + Q24 (2V + s 2 )

(52)

||Z||2 (2V + s 2 )
.
||Z||2 + Q24 (2V + s 2 )

The sign of the square root is not fixed (as we have explained before, the choice of the sign
of the Khler form corresponds to the choice of the sign of the complex structure). One
can see that the near-horizon background fields satisfy the attractor conditions:
q
h
.
(53)
Q4 B h Q2 = ||Z||2 2Q24 V h R
Which backgrounds at infinity flow to the given near-horizon background?
Let us now invert (52) and find the backgrounds flowing to the given attractor.
The condition (53) tells us that the near-horizon B-field is expressed in terms of the

A. Mikhailov / Nuclear Physics B 584 (2000) 545588

563

near-horizon volume and Khler form. Therefore, it is natural to parametrize attractors


h
, hC ). There is a twenty-parameter family of backgrounds flowing to
by (V h , R
h
h
h
(V , R , C ). They may be parametrized by a real number s and the Khler form R .
The B-field, the volume and the complex structure are determined from the following
equations:
C = hC ,
B + sR =
V+

Q2
||Z||2
q
+
h ,
Q4 Q ||Z||2 2Q2 V h R
4
4

(54)

||Z||2
s2
=
V h.
2
||Z||2 2Q24 V h

From the last of these relations we get a restriction on V :


V6

||Z||2
V h.
||Z||2 2Q24 V h

(55)

Therefore, for any given attractor there is an upper limit on the volumes of the possible tori
at infinity.
When are two backgrounds at infinity equivalent?
We will call two backgrounds equivalent if they flow to the same background at the
horizon. Our equivalence classes are 20-dimensional subspaces in Gr(4, 4 + 20) in the
case of K3, or 4-dimensional subspaces in Gr(4, 4 + 4) in the case of T4 . Indeed, given a
space W Z , we need to specify a single line in R4,20 /(W Z ), in order to specify W .
Therefore, the equivalence class is itself a Grassmanian manifold, Gr(1, 1 +20). The space
of equivalence classes is also a Grassmanian, Gr(3, 3 + 20). It follows from (52) or (54)
that for the background with parameters B 0 , 0i to be equivalent to the background B, ,
the following conditions should be satisfied:
(1) Complex structures are the same,
0
= B + sR ,
(2) B 0 + s 0 R

(3) V 0 +

(s 0 )2

(56)

s2

=V + .
2
2
(A priori, the complex structures are the same modulo sign. But since the equivalence
classes are connected, it is natural to choose the sign to be the same.) The third condition
follows from the first two and the definition (49) of s, therefore we have in total 40 + 20 =
60 conditions for X = K3, or 8 + 4 = 12 for T4 . (This is the dimension of Gr(3, 3 + 20)
or Gr(3, 3 + 4), respectively.)
Two examples
(1) As a consistency check, let us see what happens when V  1. In this case, the nearhorizon plane W h is generated by the vectors:

564

A. Mikhailov / Nuclear Physics B 584 (2000) 545588


h
vC
= (0, C , 0),

h
vR
' (0, R , 0),

v0h ' (1, Q2 /Q4 , Q0 /Q4 ).

(57)

This plane does not depend on B. Therefore, for the large volume torus the moduli space
of instantons does not depend on B. This is what we expect. Indeed, in this limit we expect
that the system is described by the modified six-dimensional YangMills action:
Z
1
(58)
tr (F B)2 d 6 x.
SB = const. + 0
gstr
The corresponding equations of motion do not depend on B, and the moduli space of
solutions is the same as for B = 0.
(2) Now consider the equivalence of backgrounds with very small torus, V = V 0 = 0.
Then, we have simply s = s 0 . This allows us to rewrite the equivalence conditions in
somewhat simpler form. Namely, the two small tori are equivalent if and only if:
(1) They have the same complex structure,
(2) They have the same two-form
(R (Q2 Q4 B))
(Q2 Q4 B).
R 2
||Q2 Q4 B||2 ||Z||2

(59)

Each equivalence class contains small tori


Given the torus of the finite size, it is easy to find the corresponding family of small
tori. Let us denote the background fields for the finite size torus (B, R , C ), and the
0
, 0C ). From the last of equations (56), we
background fields for the small tori (B 0 , R
0
express s for the family of small tori in terms of the volume and s for our finite size torus:
p
(60)
s 0 = 2V + s 2 .
0
, which is constrained to be orthogonal
Our family of small tori can be parametrized by R
0
0
to C = C . The field B is found from the second equation in (56) which tells us that
0
is equal to B + sR . (If s 0 is related to s by the last of equations (56) and
B 0 + s 0 R
0
) satisfy the second of (56), and s is related to (R , B, V ) by the formula (49),
(B 0 , R
0 , B 0 , V 0 = 0) by the same formula (49).) We see that, indeed,
0
then s is also related to (R
there are small tori in each equivalence class. This allows us to describe the dynamics of
branes wrapped on a finite torus in terms of branes wrapped on a small torus. The latter is
related to the noncommutative geometry, as reviewed in the next section.

4.2. Turning on the RamondRamond fields


Now we want to include the RamondRamond fluxes. We will use the notation for the
vectors in R5,21 which is explained at the end of Appendix A. Notice that turning on
the RamondRamond fields does not change the near-horizon values of the perturbative
moduli, calculated in the previous subsection. Indeed, the four-plane specifying the
perturbative moduli is given by (W v )/v where v = [(0, 0, 0), (0, 1)]. But v Z ,
which means that the reduction to v /v commutes with the operation W (W Z )
RZ used to compute the near-horizon moduli. We want to answer the question: what

A. Mikhailov / Nuclear Physics B 584 (2000) 545588

565

are the near-horizon RR fluxes, given the RR fluxes at infinity? Consider the plane W
corresponding to the background at infinity. It is generated by five vectors the fifth of which
is:


(61)
v 5 = (C0 , C2 , C4 ), (1, 2 12 ||C||2) .
Let us denote W4 = W v , v = [(0, 0, 0), (0, 1)]. In particular W4 is the plane generated
by the first four vectors v 0 , . . . , v 3 . Consider the vector
vh5 = v 5

(v 5 Z)
PW Z.
||PW4 Z||2 4

(62)

We claim that vh5 = [(C0h , C2h , C4h ), (1, h2 12 ||C||2)], where Cih and h are the nearhorizon values of the RamondRamond fluxes and the six-dimensional string coupling
constant. Let us prove it. The subspace W4 W consists of the vectors of the form:
[(, , ), (0, )]. We can characterize v 5 as the vector of the form [(, , ), (1, )], with
the conditions v 5 W and v 5 W4 . The vector vh5 defined in (62) belongs to W Z .
Since W Z is constant, v5 also belongs to the near-horizon five-plane W h . Because
of the attractor condition, W h contains Z. Since we do not consider fivebranes and
fundamental strings, the last two components of Z are zero. Therefore Z belongs to the
near-horizon four-plane, Z W4h . Also, notice that W4 Z is three-dimensional. Now
it follows that the near-horizon W4h is generated by Z and W4 Z . Besides being
orthogonal to Z, our vector vh5 is also orthogonal to W4 Z (indeed, v 5 is orthogonal to
W4 and PW4 Z is orthogonal to W4 Z ). Therefore, vh5 is orthogonal to W4h . Also, it is
clearly of the form [(, , ), (1, )]. Because of our characterization of v 5 , these properties
imply that the vector vh5 defined in (62) is, indeed, the near-horizon v 5 .
Straightforward computation using (62) gives (when V = 0):
C0h = C0 + t,
C2h = C2 tB,
t
C4h = C4 + ||B||2 ,
2
where we have denoted:
Q0 C0 + Q4 C4 + (Q2 C2 )
.
t=
Q0 + (Q2 B) 12 Q4 ||B||2

(63)

(64)

One can check that the total flux of the near-horizon RR fields through our branes is zero:
Q0 C0h + Q4 C4h + (Q2 C2h ) = 0.

(65)

This is the attractor condition for the RamondRamond fields. The relation (63) between
the asymptotic and the near-horizon values of the RamondRamond fields may be rewritten
in terms of C = C0 + C2 + C4 :
C h = C + t eB .

(66)

Notice that the expressions (63) for Cih do not involve R (when V = 0). It has been shown
in [20] that the near-horizon values of the RamondRamond fluxes correspond to the fluxes

566

A. Mikhailov / Nuclear Physics B 584 (2000) 545588

of the B-fields in the sigma-model on the moduli space of instantons. In other words, they
are the theta-angles of the instanton sigma-model. When B = 0, (63) tells us that these
theta-angles depend only on C2 and C4 . Intuitively this is what one would expect. Indeed,
after T-duality C4 becomes C00 the new RR zero-form, and C2 becomes (C20 ) . Our Q4
becomes instanton number, and Q0 becomes the rank of the gauge group. The interaction
with the RR fluxes for the (1 + 5)-dimensional theory is of the form:
Z
Z

(67)
C2 tr F F + C4 tr F F F ,
C0 tr F +
R2

T4 R2

where F is the six-dimensional field strength. Since these expressions are topological,
they give theta-angles after the dimensional reduction to 1 + 1 dimensions. 5 Notice
that the RamondRamond zero-form C0 couples only to the U (1) gauge field, which
decouples from the other fields and does not participate in the dual description of the AdS
supergravity.
We explain the meaning of (63) when B 6= 0 in the next section.

5. Relation to the noncommutative torus


5.1. Small torus and noncommutative torus (a very brief review)
There is a beautiful relation between the noncommutative geometry and the compactification of the string theory on the torus of very small size. We will not review it here,
since it is thoroughly explained in [2,3,8]. But we will give a very brief description of this
correspondence, in order to fix our notations. Consider the compactification of the type
IIA string theory on the small torus T4 with the metric gij and the B-field Bij . Consider
N D0-branes in this background. Suppose that the size of the torus is much less than the
string length. Consider first the case Bij = 0. Making T-duality in all the four directions of
the torus, we get the theory of N D4-branes on the dual torus (which has very large size).
According to [35], the low energy worldsheet theory for these D4-branes is the N = 4
supersymmetric YangMills theory. Now let us turn on some very small B-field. (In our
notations, B-field is dimensionless; for the string worldsheet wrapping the cycle (ij ) of
the torus, the B-field gives the phase factor e2iBij in the path integral.) Since B is small,
we expect that it does not considerably change the dynamics of D0 -branes. The effect of
the small B-field is some small correction to N = 4 SYM on T4 . It turns out, that this
correction is precisely turning on the noncommutativity parameter. Moreover, this result
holds for arbitrary Bij , not necessarily small.
The precise correspondence goes as follows. The noncommutative torus is defined in
terms of the algebra of functions on it (see [38] and references therein for details). This
is the noncommutative deformation of the algebra of functions on the usual torus. If we
5 The map from the worldsheet R2 to the moduli space of instantons on T4 implies the choice of the U (N )bundle over T4 R2 . The theta-angles assign a phase to such a map, which is the linear combination of the Chern
classes specified in (67).

A. Mikhailov / Nuclear Physics B 584 (2000) 545588

567

denote i the coordinates on the torus, then the algebra of functions is generated by eij for
j = 1, . . . , 4. The algebra of functions on the noncommutative torus may also be thought
of as generated by eij , but instead of eij eik = eik eij we have:
eij eik = e2ijk eik eij .
The bivector ij is called the noncommutativity parameter. This defines the noncommutative torus as a manifold. To define the YangMills functional, we have also to specify
the metric. It is done in the following way. One introduces a four-dimensional abelian algebra, acting on the algebra of functions as derivations. This algebra is the analogue of the
algebra of constant vector fields on the commutative torus. It remains abelian in the noncommutative case. The metric on the noncommutative torus is defined as the metric Gij on
this algebra (as on the linear space). We have the following correspondence between the
parameters of the string theory torus T4 and the parameters of the noncommutative torus
e
T4 :
ij = Bij ,

Gij = (g 1 )ij .

(68)

The relation between these formulas and the effective open string metric/noncommutativity
parameter of [8] is the following. The effective open string metric and noncommutativity
parameter are given by Eq. (2.5) of [8]:
ij
ij


1
1
,
ij =
.
(69)
Gij =
g+B S
g+B A
Suppose that we are given the background with the metric gij and the B-field Bij . First,
we do the T-duality g + B 1/(g + B). Then, we use (69) to get G = 1/g and = B.
The vector bundle over the noncommutative torus is defined as a module E over
the algebra of functions. The gauge field is defined as the set of covariant derivatives
1 , . . . , d acting in E in a way that agrees with the action of the algebra of functions
(see [3] for details). The curvature Fij = [i , j ] is an element of End E. To understand
the correspondence between the numbers of branes and the topological numbers of the
noncommutative YangMills theory we have to review the noncommutative analogue of
F
the Chern classes. In conventional (commutative) geometry, we can think of tr e 2 i as
a differential form; its cohomology class is the Chern character of our bundle. In the
F
noncommutative case we cannot think of tr e 2 i as a differential form, in particular because
the operation of integration cannot be separated from the operation of taking the trace.
We will think of it in the following way. Given the noncommutative torus Td , consider
a commutative torus U (1)d , consisting of the automorphisms of the algebra eij eik =
e2ijk eik eij of the following form:
eij eij eij ,

j [0, 2].

(70)

Consider a trivial bundle over this torus with the fiber being our module E (the infinitedimensional space). Then, Fij = [i , j ] is a two-form on this commutative torus with
F

values in End E. Let us consider tr e 2 i as a (non-homogeneous) constant form on U (1)d .

568

A. Mikhailov / Nuclear Physics B 584 (2000) 545588

This form by itself does not have any integrality properties. However, it is known from
mathematical literature [40] that the following form on U (1)d :
F

= e() tr e 2 i

(71)

is an integer form. It is the noncommutative version of the Chern character for bundles on a
torus. The correspondence between the numbers of branes and the topological numbers of
the noncommutative YangMills theory goes as follows. For a two-form = 12 ij dx i
dx j , let us denote the two-form on the dual torus with the components ( )ij =
1 ij kl
kl . Then, for d = 4 we have the following relation between the Chern character
2
and the numbers of branes:
= 0 + 2 + 4 vol,
0 = Q0 ,

4 = Q4 ,

2 = Q
2.

(72)

Indeed, to fix the relation between the components of and the charge Z, we can consider
the case B = 0. After T-duality, C = C0 + C2 + C4 goes to
C 0 = C4 + C2 C0

(73)

(it follows from (125) and (126); T-duality acts on 4,4 as (x, v, y) (y, v , x)). We use
the coupling to the RamondRamond fields to find :
Z
(74)
C 0 = Q0 C0 + (Q2 C2 ) + Q4 C4 .
This implies (72). In particular, we have the following relations:
1
tr 1 = Q0 + (Q2 B) Q4 ||B||2 ,
2
F

= 2 4 = Q2 + Q4 B .
(75)
tr
2i
Also notice that for the geometrically dual torus, the Khler forms are related to the Khler
forms of the original torus as follows:
i 7 (i ) .
In our situation, we have type IIB Q0 D1, Q2 D3 and Q4 D5-branes. In this case, we get
T4 , where R1,1 is commutative. (We may
the five-dimensional U (Q0 ) theory on R1,1 e
consider this as a particular case of the six-dimensional noncommutative YangMills, with
the noncommutativity bivector vanishing when one of its indices is in the tangent
space to R1,1 .) Dimensional reduction of this theory on the four-torus gives the sigmamodel with the target space the moduli space of instantons on e
T4 . We will not discuss
the details of the dimensional reduction in this paper. Let us explain why the resulting
metric on the moduli space of noncommutative instantons is hyper-Khler. The argument
goes precisely as for the conventional (commutative) instantons [41]. Consider the infinitedimensional space of all connections i in the module E over the algebra of functions on
e
T4 . This space is endowed with the flat metric ds 2 = tr di d i and the sphere worth of
T4 . There is an
Khler forms [] =  ij kl tr ij dk dl , being a Kahler form on e

A. Mikhailov / Nuclear Physics B 584 (2000) 545588

569

infinite-dimensional group of gauge transformations i gi g 1 preserving the metric


and the complex structures. Choosing a particular Kahler form [] as a symplectic form,
we can think of the gauge transformation with the infinitesimal parameter as generated by
the Hamiltonian H [] =  ij kl ij tr Fkl . The anti-self-duality condition may be written
as the condition that for any Kahler form
H [] = ( + ) tr ,
where + is the constant two-form on the right hand side of the instanton equations F + =
+ [37]. Therefore, the moduli space of the anti-self-dual connections is the hyper-Khler
reduction of the space of all connections. According to Theorem 3.2 in [42], the natural
metric on the hyper-Khler quotient is hyper-Khler.
As an example of how this correspondence works, let us compute the tension of the
black string in the limit of the small size torus. From the solution (17), we see that the
tension is:
q
1
V ||Z+ ||2 .
(76)
T=
gstr
At small V , we have:
(Z v 0 )2
' ||(Q2 Q4 B)+ ||2
||v 0 ||2
(||Z||2 ||Q2 Q4 B||2 )2
+
8Q24 V
1
+ (||Z||2 ||Q2 Q4 B||2 )
2
(we have neglected the terms of the order V ). Then, the mass formula (76) gives:

e

1
V
||Z||2 ||Q2 Q4 B||2
T= 2
2Q
2gYM
4



||(Q2 Q4 B)+ ||2
1
e
+ O(V ) ,
Q4 1 + 2
||Z||2 ||Q2 Q4 B||2
||Z+ ||2 = (Z vE)2 +

(77)

(78)

e = (V )1 . The first term is leading


where is the sign of ||Z||2 ||Q2 Q4 B||2 and V
when V 0. The second term after identifications (68) and (72) is proportional to the
action of the noncommutative instanton, found in [37].
Now we are in a position to give a formula relating the near-horizon geometry to the
noncommutative torus. First, we look at which small size tori at infinity flow to the given
torus at the horizon. We use (54) with V = 0. The third of relations (54) expresses s in
terms of V h . Substituting this s to the second equation, we get the relation between B and
R . We have:
C = hC ,

p
||Z||2
2V h ||Z||2
Q2
h
q
+
R q
R .
B=
Q4 Q ||Z||2 2Q2 V h
2 2Q2 V h
||Z||
4
4
4

(79)

570

A. Mikhailov / Nuclear Physics B 584 (2000) 545588

These relations determine the family of small tori, which flow to our attractor. This family
is parametrized by R . Each of these small tori determines the noncommutative torus,
which is geometrically dual to it, with the noncommutativity parameter = B. Therefore,
to the given attractor we associate the family of noncommutative tori. This family is
characterized by the fixed complex structure and the following relation between the Khler
form and the noncommutativity parameter:
p

||||2
2V h ||||2
h
2
+ q
R q
R .
(80)
=
4
2
2
||||2 2 V h
||||2 2 V h
4

(Here we have used instead of Z.) The formula (80) gives the answer to the question:
which noncommutative torus corresponds to the AdS background with parameters
h , V h , B h )? We see that the answer is not unique. It follows that all the tori from
(hC , R
the family (80) have the same moduli space of noncommutative instantons.
Since the parameters of the attractor are unambiguously determined by the parameters
of the small torus (B, R , C ), we can rewrite the equations (80) for the family of
noncommutative tori so that it does not contain the parameters of the attractor. Indeed,
rewriting (59) in terms of and (, i ), we get:
(1) C = const. (modulo phase),

(R (2 4 ))
2 4 = const.
(2) R 2

2
2
||2 4 || ||||

(81)

Different constants give different families of noncommutative tori. The form 2 4


is expressed in terms of tr F in (75). It would be interesting to prove in noncommutative
geometry, that all the noncommutative tori from the family (81) have the same moduli
space of instantons with instanton number .
Notice that Eqs. (81) imply that the self-dual part of the form
2 4
||2 4 ||2 ||||2

(82)

T4 of selfis a covariantly constant section of the restriction of the universal bundle 2+e
dual forms on the family of equivalent tori.
Notice that for small noncommutativity and 2 = 0, the second equation (81) implies
that: (1) R is constant and (2) the self-dual part of is constant. This is what we
intuitively expect. Indeed, in the limit of almost commutative torus there are large regions
in the moduli space, where some instantons become very small. But very small instantons
on zero volume T4 is almost the same as instantons on R4 . According to [8], the moduli
space of instantons on R4 depends only on the self-dual part of the noncommutativity
parameter. Some properties of supersymmetric configurations on R4 are discussed in the
next subsection.
5.2. Supersymmetric configurations on R 4
It was conjectured in [13,43] that the supersymmetric configurations on R4 with nonzero
B-field are noncommutative instantons. This has been proven in [8] in the zero-slope limit,

A. Mikhailov / Nuclear Physics B 584 (2000) 545588

571

that is when 0 0. We want to show that this is true beyond the zero slope limit. The
zero slope limit may be characterized by:
( 0 )2 g ik g j l Bij Bkl '

1
 1.


(83)

1
Gik Gj l ij kl
( 0 )2

(84)

Notice that
( 0 )2 g ik g j l Bij Bkl =

(this follows from the formula (69) for the effective open string metric and noncommutativity parameter).
We will argue that the moduli space of supersymmetric configurations depends only
on certain combination of metric and B-field, and that for the background with generic
g and B we can always find the background which gives the same moduli space of
supersymmetric configurations and satisfies the zero slope condition. Let us prove it.
Suppose that we are given the generic metric and B-field on R4 , not necessarily satisfying
(83). The moduli space of supersymmetric configurations should not change significantly
if we replace our R4 by sufficiently large torus. (We will make this statement more
precise below.) This torus will have a very large volume and a very large B-field; in
fact, for the generic background the B-field will scale like the square root of the volume.
The target space of the sigma-model describing the low energy dynamics of the D1D5
system wrapped on this torus is the moduli space of supersymmetric configurations on this
torus. For finite values of the charges, it has a region corresponding to supersymmetric
configurations on R4 . Indeed, for a very large torus we expect to have such field
configurations that the field strength is localized in the small region of the torus which
may be thought of as R4 with a nontrivial B-field. We will call such field configurations
localized. Since we have gij ' Bij , these field configurations do not satisfy the condition
of the zero slope limit. We want to argue, however, that our large torus is equivalent to the
torus with |gij |  |Bij |, for which the localized supersymmetric configurations do satisfy
the zero slope condition. Let us show it. It is very convenient to do first the T-duality
g + B 1/(g + B) which gives us the torus with small gij and Bij . The metric and the
B-field are, still, of the same order of magnitude. In other words, we have:
||B||2 V  1,

Q0 ' Q4

(85)

(and we put Q2 = 0). Although the volume is small, it is finite. As we have argued in the
previous section, we can always find an equivalent torus with zero volume. The Khler
0 and the B-field B 0 for that equivalent torus satisfy:
form R
0
= B + sR ,
B 0 + s 0 R

where
s0 =

2V + s 2 ,

s'

(86)
Q0
1
Q4 (B R )

(87)

(we have used an equation (49) for s in the regime (85)). Since s  1 and V  1 we have:
s 0 = s + O(V /s).

(88)

572

A. Mikhailov / Nuclear Physics B 584 (2000) 545588

Then, (86) is satisfied by:


0
= R ,
R

B 0 = B + O(V /s) ' B

(89)

(the correction for B is much smaller then B itself). We see that the torus with small V and
B satisfying (85) is equivalent to the zero volume torus, which has the same shape and the
same (small) B-field. (Of course, we can take any other zero volume torus with the same
shape and the same self-dual part of the B-field.) In terms of the T-dual torus (gij , Bij )
with large gij and Bij (the one we have started with) this means that it is equivalent to
eij . However, the relation between the shape of
the zero size torus with large B-field B
e and B is
this equivalent torus and the shape of gij , as well as the relation between B
more complicated. In fact, it is very simple in terms of the effective open string metric
and the noncommutativity parameter. Namely, the noncommutative torus (Gij , ij ) with
small ij and Gik Gj l ij kl ' ( 0 )2 is equivalent to the very large noncommutative
torus (Gij = ) with the same shape and the same noncommutativity parameter. The
corresponding moduli space is just the moduli space of instantons on large noncommutative
T4 and the region in the moduli space corresponding to localized field configurations is the
moduli space of instantons on the noncommutative R4 .
Let us summarize what we have. We have started with the torus of the large size
and large B-field. We have argued that there are localized field configurations, which
are the same as supersymmetric configurations on R4 with the B-field. On the other
hand, we have proven that the moduli space of supersymmetric configurations on this
torus is the same as the moduli space of supersymmetric configurations on the small size
torus with some other B-field (which is also large). This moduli space also has a region
corresponding to localized configurations. It is a natural conjecture (we did not prove it)
that the localized supersymmetric field configurations on the large torus correspond to the
localized configurations on the equivalent small torus. But the localized configurations on
the small torus are described by the zero slope limit, and therefore they are instantons on
the noncommutative R4 .
This shows that the moduli space of supersymmetric configurations on R4 does not
depend on 0 .
5.3. Wilson lines
The moduli space of anti-self-dual connections on T4 has an obvious U (1)4 symmetry:
i i + iai ,

(90)

where ai are real numbers. This symmetry does not have fixed points. Therefore, the
moduli space has a structure of a bundle with the fiber four-torus T 4 . In the commutative
case (when B = 0) we could identify this T4 with the dual to the torus on which our Yang
Mills fields live. In the noncommutative case we expect more subtle relation between these
tori. One way to see it is to notice that the periodic identifications ai ai + 2 result
from the gauge transformations, which are modified for the noncommutative torus. The
gauge transformation with the parameter e2ii would not just change ai ai + 2i,

A. Mikhailov / Nuclear Physics B 584 (2000) 545588

573

but will instead add some nontrivial function of 1 and 2 to i . Shifting ai by constant
will require more complicated gauge transformation, and the periodicity properties of the
Wilson lines will be changed. We expect that the torus T 4 parametrizing the Wilson lines
will be different from the geometrically dual to T4 . Notice that in the presence of the
B-field the shape of the torus T4 depends on the representative in the equivalence class
of backgrounds. Therefore, if T 4 was just geometrically dual to T4 , we would get nonequivalent sigma-models for equivalent backgrounds. In fact, we know from Section 4
that C and B + sR are invariants of the equivalence relation. Therefore, it is natural
to conjecture that the hyper-Khler structure on the moduli space of flat connections in
the bundle
with the given (Q4 , Q2 , Q0 ) is specified by the forms Re C , Im C and
p
eR ||2 where
eR is given by the second line in (81). We do not know how to
eR / ||

prove it in noncommutative geometry.


5.4. Possible singularities in the moduli space
If the instanton moduli space for one of the tori from the family (81) has singularities,
then all the tori from this family have singular instanton moduli spaces. Singularities of
the instanton moduli spaces are related in string theory to the possibility for the system of
branes to split into two or more subsystems preserving the BPS condition. The conditions
for the possibility of the decay of the black string have been worked out in [36]. The answer
is the following. For the black string with the charge Z in the background W to be able
to decay into two strings, one with the charge Z1 and the other with the charge Z Z1 , it
is necessary that the projection (Z1 )+ of Z1 on W and the projection Z+ of Z on W are
collinear vectors. This means, that W Z1 = W Z . In other words, W Z should be
orthogonal to Z1 . Therefore, singularity or nonsingularity depends only on Z W . But
our definition of equivalent tori is precisely that they have the same Z W .
We get the simplest example of the singular moduli space, if we take Q2 = 0, B antiself-dual, and turn off the RamondRamond fields. In this situation, (B ) = 0 (this is
the definition of B being anti-self-dual), and the plane W Z is generated by the three
vectors [(0, i , 0), (0, 0)] together with the fourth vector [(0, 0, 0), (1, 2)]. We see that
the whole lattice 1,1 generated by the integer vectors of the form [(, 0, ), (0, 0)] is
orthogonal to W Z . Therefore, our (Q4 , Q0 ) system may decay into (Q04 , Q00 ) and
(Q4 Q04 , Q0 Q00 ), provided that Q04 Q00 > 0 and (Q4 Q04 )(Q0 Q00 ) > 0. The
corresponding singularity in the moduli space of instantons is due to small instantons.
Notice that if we turn on the generic RamondRamond fields, then the separation of
branes becomes impossible. Therefore, although the moduli space of instantons remains
geometrically singular, the conformal sigma-model is nonsingular. The reason is, of course,
that we have turned on theta-angles.
It is not true that an arbitrary background with singular instanton moduli space is related
to this example by dualities. Indeed, it can happen that the sublattice orthogonal to W Z
is not 1,1 (and has dimension higher then two). Then, the corresponding singularity
cannot be related to the small instantons by the chain of dualities.

574

A. Mikhailov / Nuclear Physics B 584 (2000) 545588

5.5. Theta-angles
The target space M of the D1D5 sigma-model has a nontrivial second cohomology
group. There is a natural map Z from the cohomology lattice of T4 to the second
cohomology group of the moduli space of noncommutative instantons with the charge Z.
Let us remind how this map is constructed. Given a cohomology class H 2 (T4 , Z), we
want to describe how to compute Z () on a cycle M. We describe the embedding
M as a connection on a vector bundle Eb over T4 . Then, the value of Z () on
b . (This is the noncommutative generalization of
is the highest component of (E)
R
b
T4 ch(E) .)
Although Z is naturally defined over Z we can consider it over R. Consider the fluxes
of the RamondRamond field through the cycles of T4 as an element of H (T4 , Z). We
want to show that the theta-terms in the D1D5 effective action are the values of Z on the
near-horizon RamondRamond fields:
Theta-angles = Z (C h ).

(91)

In particular, the theta terms do not depend on the choice of the representative in the
equivalence class of the backgrounds.
To derive this formula, we will use the following trick. We first consider the case of
Euclidean D-branes wrapped on T6 . We will take this T6 to be a product of T2 and a
T4 . In this case, we know that the Ramond
small T4 , and go to the dual noncommutative e
Ramond fields couple to the winding numbers of various D-branes on this six-torus:
SRR = C 0 T2 e
T4 ,

(92)

2
e4
where T2 e
T4 is the integer cohomology class of T T specifying the number of
branes wrapped on various cycles. Then we look at (92) from the perspective of the
noncommutative geometry. We can explicitly express T2 e
T4 in terms of the field strength
T4 using the Eliotts formula (71):
of the YangMills field on T2 e



1
0

F
=
C

e
tr
exp
SRR = C 0 T2 e
T4
2i




1
1
F = A0 tr exp
F
(93)
= (e C)0 tr exp
2i
2i

(the forms A and C are defined in Appendix A, and C 0 is given by (73)). Now we want to
T4 by making T2 very large, much larger then e
T4 . The low energy theory is
pass to R2 e
a dimensional reduction of the six-dimensional YangMills on e
T4 . This theory lives on an
ordinary, commutative two-torus. However, it does remember, in a way, that it was obtained
from the reduction on the noncommutative manifold; it turns out that the integral along this
commutative two-torus of tr F is not integer. This follows from the formula
1

tr e 2 i F = e

(94)

derived in [40]. For the worldsheet instanton has two indices along T2 and two or four
indices along e
T4 . Its contraction with the noncommutativity parameter gives nontrivial

A. Mikhailov / Nuclear Physics B 584 (2000) 545588

575

(and noninteger) flux of tr F along the two-torus. This flux contributes to the formula (93)
for the coupling to the RR fields.
R
The appearance of noninteger tr Fzz may seem strange and we want to make a
comment about it. Let us consider the following example. Take our noncommutative torus
T4 to be a product of two noncommutative tori T2(1) T2(2) , with the noncommutativity
R
parameters 1 and 2 . Let us consider an instanton with T2 T2 F F 6= 0 (the first T2
(1)

in the product T2 T2(1) is the commutative torus). In the adiabatic limit ( i.e., when the
size of T2 is much larger then the size of T2(1) ) this configuration may be thought of as
wrapping the large (and commutative) T2 on the nontrivial two-cycle in the moduli space
of flat connections in T2(1) (see [44]). Naively it seems that we can just put Az and Az
equal to zero, which would clearly give Fzz = 0, and no flux. However, there is a subtlety
here. In fact, we have to consider flat connections modulo the gauge transformations.
And it is not true that for our map from T2 to the moduli space of flat connections on
T2(1) we can globally choose the representative in the gauge equivalence class smoothly
depending on the point in T2 . Therefore, we should really cover our T2 with patches
and construct separately the family of flat connections on T2(1) over each patch, and then
glue them together by the appropriate gauge transformations. These gauge transformations
can make it impossible to choose Az = 0 and Az = 0. In the case if the small torus is
noncommutative, we have additional complications due to the fact that our fields A are
not really gauge fields, but noncommutative gauge fields. That is, they have non-standard
gauge transformations. In fact, non-commutative gauge fields can be expressed in terms
of the ordinary gauge fields, as explained in [8]. The formula to the first order in the
noncommutativity parameter is:

1
A i = Ai kl Ak (l Ai + Fli ) + (l Ai + Fli )Ak .
4
The corresponding noncommutative field strength is, in our case:
1 ij h
2 Fzi Fz j + Fz j Fzi Ai (j Fzz + j Fzz )
Fc
zz = Fzz +
4
i
+ (j Fzz + j Fzz )Ai .

(95)

(96)

bzz cannot be taken to be zero. The noncommutative instanton


This expression shows that F
R
R
is the deformation of the configuration with F F = 8 2 ninst and T2 F = 0. In this
bzz 6= 0:
topological sector, we can take Fzz = 0. However, F
Z
Z
bzz = 1 ij Fzi Fz j ' 1 ninst,
trF
(97)
2
T4

T4

which is in agreement with (94).


Now let us see what happens when we replace T2 by R2 , which gives us our original
configuration of D1D5 wrapped on T4 with two common noncompact directions. In this
case, the flux of tr F through R2 is not determined by the topology. The configuration
with zero tr Fzz minimizes the action. Indeed, we can add to the gauge field Az the piece

576

A. Mikhailov / Nuclear Physics B 584 (2000) 545588

constant in T4 : Az Az + a(z, z ), Az Az + a(z,


z ), a C. The only modification to
the action will be (remember that tr includes integration along e
T4 ):
Z

Z
2
d 2 z trFz
z
T2

d 2 z tr(F da)2zz .
T2

Varying with respect to a, we get (da)zz tr 1 = tr Fzz . This means that the configurations
with minimal action have trFzz = 0. In particular, the flux of tr F through R2 is zero. This
implies, that the coupling to the RamondRamond fields becomes different from (93).
In fact, putting trFzz 0 in (93) is equivalent to replacing the RamondRamond fields
(C0 , C2 , C4 ) by their near-horizon value (C0h , C2h , C4h ) given by the formula (63). Let us
prove it. We will consider the formula (93) valid for the theory on the six-torus, but put the
diagonal part of the components of F along T2 equal to zero. In other words, replace F
R
R
with F ( tr11 T2 tr F ) where is the fundamental cohomology class of T2 , T2 = 1.
We will get:
Z

R




1
2 tr F

0
F T

d z (e C) tr exp
2i
tr 1




 R
Z
F
2 tr F
C 0 d 2 z e tr exp
= exp T
2i tr 1
2i
T2

 R
T2 tr F
C 0 T2 e
= exp
T4
2i tr 1

R
0
0
T2 tr F

= C T2 e
4
2
4
e
T
T T
2i tr 1
Z
0
(C
F

)
4
T
.
tr
= C 0 T2 e
T4
tr 1
2i
2

T2

(98)

T2

Let us compare this expression with what we would get by replacing C C h in (92):
Z



1
F
e (C te ) tr exp
2i
Z
F
.

t
tr
= C 0 T2 e
T4
2i

T2

0

(99)

T2

Substituting t from (64) and taking into account (75) we see that (98) is equal to (99).
In other words, the coupling of the sigma-model to the RamondRamond fields is given
h
by the formula (92) with integer T2 e
T4 but C replaced with C . This means that the
sigma-model on R2 couples to the RamondRamond fields only through their near-horizon
values.

A. Mikhailov / Nuclear Physics B 584 (2000) 545588

577

6. Sigma-model instantons and supergravity instantons


In this section we want to discuss the supergravity meaning of the period map. The
period map associates to each hyper-Khler manifold the space of integrals (periods) over
its two-cycles of the three Khler forms. Given the period map of the target space, one can
immediately compute the action of any sigma-model instanton. Indeed, the corresponding
map f : R2 M from the worldsheet to the target space should be holomorphic in some
complex structure and the action is given by
v
!2
u 3 Z
uX
u
f I .
(100)
S =t
I =1

R2

But the integrals over the worldsheet of the pullbacks of the Khler forms is given in terms
of the period map and the topology of the instanton. More exactly, let us introduce the
cohomology lattice H 2 (M, Z). There is a natural nondegenerate pairing on this lattice
(the construction is reviewed in [20]). The topology of the sigma-model instanton can
be specified by giving a vector z H 2 (M, Q), so that the integral of p H 2 (M, Z)
over the worldsheet equals (z p). The period map gives us a three-plane W3 -model
R H 2 (M, Z), and the action of the instanton is
q
(101)
S = ||PW -model z||2 .
3

We want to interpret this formula in supergravity. Our sigma-model arises on the


worldvolume of the D1D5 system wrapped on T4 . Let us consider the D1D5 system
with the worldvolume not R2 , but some compact Riemann surface (for example, S 2 ).
Of course, this system will not be BPS and therefore will presumably collapse and decay
into gravitons and other fields of six-dimensional supergravity. But let us consider the case
when we have nontrivial instanton charge on . This instanton charge will couple to the
RamondRamond background fields, and therefore it cannot decay into the states from the
perturbative string spectrum. In fact it decays into some perturbative state plus one of the
D-instantons of the six-dimensional supergravity, such as type IIB D-instanton, or type
IIB Euclidean D-string wrapped on the two-cycle in T4 , or type IIB Euclidean D3-brane
wrapped on T4 , or a combination of them. Therefore, we can think of D-instantons as
black strings with compact worldsheet wrapping a nontrivial cycle in the target space.
Do we expect the action of these supergravity instantons to be equal to the action
of the corresponding instantons in our sigma-model? For generic backgrounds, we do
not. One reason is that we expect the sigma-model description to break down for small
enough worldsheets. But even if we forget about the possible breakdown of the low energy
description, there is still a reason why the action of supergravity instantons is different. As
we have explained in Section 5.3, the sigma-model instanton on R2 usually has nonzero
abelian gauge field strength on R2 , determined from minimizing the action. In generic
situation, the flux of this gauge field through R2 is non-integer. If we want to compactify
R2 to S 2 or any other Riemann surface, we have to change the two-dimensional abelian
gauge field so that the flux is integer. This will increase the action. Therefore, we expect

578

A. Mikhailov / Nuclear Physics B 584 (2000) 545588

that in the generic background, the actions of the supergravity instantons are different from
the actions of the sigma-model instantons. However, for some special backgrounds and
instanton charges, the flux of the abelian field happens to be zero. The condition for it is that
the coupling of the instanton to (C0 , C2 , C4 ) is the same as the coupling to (C0h , C2h , C4h ).
In other words, the charge of the supergravity instanton, defined so that the phase factor
is e2i(C ) , satisfies:
(PW4 Z ) = 0.

(102)

We will see that this condition has a natural interpretation in supergravity. We will argue
that if this condition is satisfied (together with some other conditions) then the action of
the supergravity instanton is equal to the action of the sigma-model instanton.
6.1. Can supergravity instantons be absorbed by branes?
We will argue that under certain conditions (including (102)) there exists the solution
of the Euclidean six-dimensional supergravity, corresponding to the black string with the
worldsheet R2 and the D-instanton at some distance from it. This solution preserves four
real supercharges. It has moduli corresponding to the motion of the D-instanton in the
space transverse to the black string. The action does not depend on the position of the
D-instanton. It is possible that such a D-instanton can approach our black string and
dissolve into it, becoming a sigma-model instanton. We conjecture that the action is not
changed in this process. For example, we may consider the black string which is obtained
by wrapping N > 1 type IIB D3-branes on some two-cycle of the torus. If we turn off
the B-field on the torus, then introducing a D1 -brane will leave eight supersymmetries
unbroken. The D1 -brane can move in the directions transversal to N D3, or it can become
an instanton on the worldvolume of N D3. From the point of view of the YangMills
theory on the worldvolume of the threebranes, we can describe this process as follows.
The D1 -brane on top of N D3-branes may be thought of as point-like instanton. This
point-like instanton corresponds to the special point in the moduli space of instantons of
SU(N) theory. Moving in the moduli space, we deform it to the finite size instanton, which
can be described as a classical supersymmetric field configuration on the worldsheet.
We should stress, however, that it is not always true that the supersymmetric configuration of the supergravity instanton and the D-brane can be represented as a classical
instanton on the worldvolume of the D-brane. For example, the configuration consisting of
a single D3 brane and a D-instanton on top of it is supersymmetric; nevertheless, there are
no finite size instantons in the U (1) gauge theory living on the worldvolume of the single
D3-brane. For our configuration consisting of the black string and the instanton, we do
not know whether the instanton can really be represented by the smooth worldsheet field
configuration. 6 We can probably say that there is always an ideal instanton represented
6 Supersymmetric solutions of YangMills equations on six-dimensional manifolds have been studied in [45
48] and references therein. For these solutions, the field strength should belong to su(3) su(4) = so(6)
modulo constant. (The variation of fermion with nonzero constant field strength can be compensated by the
transformation, as explained in [8].)

A. Mikhailov / Nuclear Physics B 584 (2000) 545588

579

by the singular fields, but the precise meaning of this statement is not very clear.
To prove the connection between the supergravity instantons and the sigma-model
instantons, we should answer two questions. The first question is whether the supergravity
instanton sitting on top of the brane can be deformed to the smooth supersymmetric
solution of the six-dimensional YangMills equations. In other words, whether the
corresponding singular field configuration is just a point on the boundary space of smooth
configurations. The second question is whether this smooth six-dimensional solution (if it
exists) has a good limit when we shrink T4 to zero size; if yes, then it should correspond
to the instanton of the D1D5 sigma-model.
We did not prove that the answers to these two questions are positive. We want to give
an example which shows that at least the first assumption is true for some instantons.
Some supersymmetric solutions of the six-dimensional YangMills theory on CalabiYau
threefold were obtained (in the language of complex geometry) in [49]. Our example is
analogous to the solutions found in that paper.
Consider type IIB compactified on a six-torus T6 = T2 T4 . Introduce the following
brane configuration. Take N D3-branes wrapped on T4 and one D3-brane wrapped on
T2 where H2 (T4 , Z), ||||2 > 0. The four-torus T4 with a given metric admits
a family of complex structures parametrized by S 2 . We can choose one of these complex
structures so that the cohomology class Poincar-dual to (which we will call []) is of
the type (1, 1). Then, there are holomorphic line bundles on T4 whose Chern class is [].
Let us choose one of them and call it L. The other bundles with the same Chern class may
be obtained from this one by shifts in T4 . It is proven in Chapter 2.6 of [50] that
||||2
.
2
Zeroes of the global sections of L are holomorphic curves representing . Fixing one
particular bundle corresponds to fixing two out of (||||2 /2) + 1 complex moduli of
such holomorphic curves. Let us parametrize the global sections of L by the vectors v
2
C|||| /2 = H 0 (M, L). We will denote the corresponding sections [v](x1 , x2 ), where x1
and x2 are the local complex coordinates on T4 . Our configuration of D3-branes can be
represented by the following equation in T6 = T2 T4 :
dim H 0 (T4 , L) =

def

F0 (x1 , x2 , x, y) = [v0 ](x1 , x2 )(a0 + a1 x + a2 y + + aN x n/2 ) = 0.

(103)

described by the Weierstrass equation = + ax + b. If


Here (x, y) is the point of
N is odd, then the last term is x (N3)/2 y. This equation represents a union of N D3-branes
wrapping T4 and the single D3 brane wrapping T2 . It preserves 18 of supersymmetries.
One can think of this equation as specifying a single complex surface in T6 , but then this
surface is singular. To make it nonsingular, we consider a deformation of (103) specified
by the choice of

(104)
[u0 , . . . , uN ] P CN+1 H 0 (T4 , L) .
T2

This deformation is:


F [u0 , . . . , uN ](x1 , x2 , x, y)

y2

x3

580

A. Mikhailov / Nuclear Physics B 584 (2000) 545588

= [u0 ](x1, x2 ) + [u1 ](x1 , x2 )x + [u2 ](x1 , x2 )y + = 0.

(105)

In particular, F [a0 v0 , . . . , aN vN ] = F0 . Unlike the initial equation (103), it represents a


smooth four-surface X[u0 , . . . , uN ] in T6 . This smooth four-surface belongs to the same
homology class N[T4 ] + [T2 ] as the original singular surface (103). Let us now assume
that N is even and is divisible by two. Then, the Chern class of the normal bundle to X
is divisible by two, and X is a spin manifold. Therefore, we can wrap a Euclidean threebrane on it [51,52]. Let us take n > 1 such D3-branes on top of each other and put k SU (n)
instantons on their worldvolume (the corresponding smooth field configuration exists for
large enough k [53]).
After making T-duality in both directions of T2 , we get a system of nN D5-branes with
nontrivial field configuration on it. Since the initial field configuration was nonsingular, we
should get nonsingular configuration after T-duality. The brane charge of this configuration
is


N||||2
[T4 ] + n[T2 ] .
Nn[T6 ] + k
8
The term N||||2 /8 appears because X has nontrivial topology, and therefore our D3
brane couples to axion even if we turn off the world volume gauge fields. Namely, the
D-instanton charge induced by the nontrivial topology of X is
Z
1
1
p1 (T X)

[X] =
[X] [X] [X] = N||||2 .
A(T X)[X] =
24
24
8
T6

This example demonstrates that at least some configurations exist as smooth solutions
on the brane worldvolume. However, the configuration which we considered is obviously
not the most generic one. The most serious restriction in this example is that
Z
tr F F F = 0
T6

(this is what allowed us to reduce dimension from six to four by doing T-duality). Also, we
did not prove that turning on the B-field will not destroy our configuration.
In the remaining part of this section we will just assume that supergravity instantons
correspond to smooth field configurations and are well described by the D1D5 sigmamodel, and see what this assumption implies.
6.2. Classification of instantons in six-dimensional supergravity
The simplest instanton to consider is the D1 -instanton, which is the dimensional
reduction of the D-instanton in ten-dimensional type IIB. Its action is
SD1 = C0

i
.
gstr

(106)

The real part of this action is the phase associated to the instanton. It is equal to the
expectation value of the axion field. In the classical low energy supergravity we have

A. Mikhailov / Nuclear Physics B 584 (2000) 545588

581

a symmetry C0 C0 + const. This symmetry is a subgroup R1 of SO(5, 5, R). But the


higher derivative corrections to the low energy action do not have such a symmetry. The
reason is precisely the contribution of the D1 -instantons, which carry a phase depending
on C0 . Now we want to consider more general instantons. Notice that the subgroup R1
SO(5, 5, R) consisting of the shifts of axion can be represented in terms of the light-like
vector v R 5,5 , and the orthogonal vector v /v. Indeed, let us consider the
following v and :
v = [(0, 0, 0), (0, 1)],
= [(1, 0, 0), (0, 0)].

(107)

One can check that the one-parametric group of transformations Tv, (t) SO(5, 5, R)
acting on vectors x R 5,5 in the following way:



t
2
(108)
Tv, (t)x = x + t ( x)v (v x) + |||| v
2
is precisely the group of shifts of C0 . We can consider arbitrary pairs (v, ) 5,5 , such
that v is lightlike and is orthogonal to it. Any pair (v, ) specifies the R-subgroup of
SO(5, 5, R), and there is a corresponding instanton breaking this symmetry. To find the
action of this instanton, we have to rewrite (106) in terms of v and :
s
||PW v ||2
( PW v)

i
.
(109)
S=
||PW v||2
||PW v||2
Invariance under U-duality group SO(5, 5, Z) implies that (109) gives the action of the sixdimensional instanton for an arbitrary pair (v, ). Although the D1 -instanton we have
started with was somewhat special because the corresponding vector was a null-vector,
the expression (109) is valid for all D-instantons. Indeed, let us consider the configuration
consisting of Q3 Euclidean D3-branes wrapped on T4 , and Q1 D1 -instantons. If the
B-field on the torus is anti-self-dual, this action for this configuration is the sum of
the action of Q1 D1 -instantons and Q3 Euclidean D3-branes. This is in agreement
with (109). The backgrounds with a generic B-field may be obtained from this one by
the rotation by the element of SO(4, 4, R) preserving the vector = (Q3 , 0, Q1 ). The
formula (109) is invariant under such rotations. This means that (109) is valid for an
arbitrary of the form (Q3 , 0, Q1 ). But any other primitive vector in the lattice 4,4
is equivalent to one of these.
We will actually consider only D-instantons, therefore all that we need is (101) for v =
[(0, 0, 0), (0, 1)]. In this case, ||PW v||2 = 2 , and (109) becomes:
q
i
||PW4 ||2 .
(110)
S = (C )

(We denote W4 = W v .)
6.3. When do supergravity instanton and black string preserve supersymmetry?
So far we have discussed instantons in flat six-dimensional space. Now let us introduce
the black string with some charge vector Z. In the presence of this black string,

582

A. Mikhailov / Nuclear Physics B 584 (2000) 545588

the group SO(5, 5, R) of symmetries of the classical supergravity is broken to the subgroup
SO(4, 5, R) preserving the vector Z. In particular, the subgroup of shifts Tv, defined in
(108) preserves Z if and only if (Z ) = 0 (we are considering only D black strings,
that is (Z v) = 0). Therefore, if (Z ) = 0, it makes sense to consider the instantons
charged under this symmetry. The condition (Z ) = 0 by itself does not guarantee that
the instanton in the presence of the black string is a supersymmetric configuration. For
example, let us consider the D1 -instanton in the background of the D3-brane wrapped on
the two-cycle of the torus. If there is a nonzero flux of the B-field through this two-cycle,
then this configuration breaks all the supersymmetry. The condition for this configuration
to be supersymmetric is that the flux of B through the two-cycle on which we wrap our
D3-brane is zero. This means that
(PW4 Z) = 0.

(111)

In general case, the condition (111) is a necessary condition for the instanton and the
black string to be a supersymmetric configuration. One can see it in the following way.
Let us consider the instanton in the presence of the large number N of coinciding black
strings, each having charge Z. At each point of the transverse space, the instanton feels
approximately flat background. Therefore, its action is given approximately by (110) but
with W depending on the distance from the black string, according to the attractor equation.
The conditions for the action (110) to be constant when W changes according to the
attractor equation with given Z are precisely that (PW4 Z ) = 0 and (Z ) = 0. We will
prove in Appendix B that the configuration consisting of the black string and the instanton
preserves supersymmetry if the following conditions are satisfied:
(1)
(Z ) = 0,
(2) (Z PW4 ) = 0,
(3)
||||2 > 0.

(112)

6.4. Supergravity instantons and period map


When the condition (111) is satisfied, we can rewrite ||PW4 ||2 as follows:
||PW4 ||2 = ||PW4 Z ||2 .

(113)

It makes sense to think of D-instantons classically when their action is very


p large, that
is gstr  1. However, nothing prevents us from taking simultaneously gstr ||Z||2  1.
Therefore, there is a regime where we can trust both the formula for the action of the
D-instanton and the near-horizon AdS supergravity.
How can D-instantons help us to prove that the geometry of the target space does not
receive string loop corrections? We know that the target space is a hyper-Khler manifold
with the second cohomology group isomorphic to 4,4 Z . The period map can be
described as a choice of the positive-definite three-plane W3 -model Z (R 4,4 ). It
was conjectured in [20] that
W3 -model = Z W4 .

(114)

A. Mikhailov / Nuclear Physics B 584 (2000) 545588

583

We want to check this formula by comparing the actions of the sigma-model instantons
to the actions of the supergravity instantons. We have argued that for some special
backgrounds the sigma-model instantons are related to the D-instantons in supergravity.
The actions of these supergravity instantons are given by (110). We expect the action
of the supergravity instanton with the charge Z to be equal to the action of the
corresponding sigma-model instanton when ( PW4 Z) = 0. In other words, when
is orthogonal to the projection of Z+ = PW4 Z on Z . Let L R 4,4 be the sixdimensional space of all vectors in Z which are orthogonal to PW4 Z:
L = (R 4,4 ) Z (PW4 Z) .

(115)

Then, our condition on the instanton charge is just: L. For generic background there
are no such integer vectors which satisfy this condition. However, there is a dense set
of such backgrounds that we can find L. Moreover, for any vector v L we can find
an integer vector V 4,4 and a number R such that V v. This means (assuming
the smooth dependence of W3 -model on the background fields) that the action formula for
the supergravity instantons gives us the angle between the plane W3 -model and an arbitrary
vector in L. Namely, given v L, the action formula (110) implies that:
||PW -model v||2 = ||PW4 v||2 .
3

(116)

For v Z W4 we have ||PW -model v||2 = ||v||2 . This implies (114). Therefore, the action
3
formula for supergravity instantons is in agreement with the conjectured period map. In
particular, we see that the periods do not depend on the string coupling constant.
We should notice, however, that the validity of this argument depends on the assumption
made at the end of Section 6.1.

Acknowledgements
I would like to thank Prof. E. Witten for many interesting and helpful discussions.
Conversations with M. Krogh and N. Nekrasov were very useful. This work has been
supported in part by the Russian grant for support of scientific schools No. 96-15-96455.
Appendix A. Moduli of K3 or T4 as points of Grassmanian
Here we will review the correspondence between the moduli and the points of the
Grassmanian manifold, following mostly [33]. We will concentrate on the case of K3,
the case of T4 is similar.
Let us first turn off the RR fields and consider weak coupling limit. There is a one to one
correspondence between the points of the moduli space and the positive 4-planes in R4,20 .
The moduli include the metric and the B-field. First, let us describe the shape of K3. Let
R4,20 be a positive 4-plane. Choose w 4,20 a primitive light-like vector. Notice
that
(w 4,20 )/w ' 3,19.

(117)

584

A. Mikhailov / Nuclear Physics B 584 (2000) 545588

e = w and = p(
e) where p : w w /w
We can consider the three-plane
is the projection. The 3-plane corresponds, via the period map, to the hyper-Khler
structure of our K3. Now we want to specify the volume and the B-field. To do this, let us
choose another primitive light-like vector, w , so that (w w ) = 1. We represent as an
orthogonal direct sum:
e
e RB,
=

e
e
B

(118)

with
e = w + w + B
B

(119)

(B w) = (B w ) = 0.

(120)

and

We identify B as the B-field, and as V 12 (B B). We have an equivalence B B + ,


with H 2 (K3, Z). Such a shift of B can be accounted for as an ambiguity in the choice
of w . Indeed, let us identify the cohomology classes of K3 with the vectors in the lattice
3,19 which is the sublattice in 4,20 , orthogonal to both w and w . Given such a vector ,
let us consider the following automorphism of the lattice:
1
w 7 (w )0 = w w( ),
2
7 0 = + ( )w,

(121)

w w,
where is any vector in 3,19 . Then the new w and w are orthogonal to the new 3,19 .
If we also change B and as follows:
B 7 (B + w(B )) + ( + w( )),

1
7 + ||B||2 ||B + ||2
2

(122)

e coincides with the old B,


e therefore all that we have done was to pick
then the new B

different w . The transformation laws (122) suggest that B should be identified with the
B-field, and should be identified as follows:
1
= V (B B).
2

(123)

As far as we know, this formula for was first obtained in [34]. Following [20], we use
the notation (x, v, y) for the vector xw + yw + v, where x, y R and v R3,19 .
Now let us include the RamondRamond fields and the coupling constant. We enlarge
our lattice from 4,20 to 5,21 by adding 1,1 . We will denote the vectors from R5,21 in
the following way: [(x, v, y), (x 0 , y 0 )] where (x, v, y) is the vector from R4,20 and (x 0 , y 0 )
is the vector from R1,1 . The scalar product is:


[(x, v, y), (x 0 , y 0 )] 2 = 2xy + ||v||2 + 2x 0 y 0 .

A. Mikhailov / Nuclear Physics B 584 (2000) 545588

585

Now we are ready to describe the correspondence between the moduli and the points of
Grassmanian. The plane W is generated by the following five vectors. Four vectors are of
the form


(124)
E = v , (0, (C v ))
with v i = (0, i , (B i )) and v 0 = (1, B, V 12 (B B)), and the fifth vector is

 
1
||C||2
.
E 5 = C, 1, 2
2

(125)

Let us explain what is and C. We define C R4,20 as follows:


C = (C0 , C2 , C4 ),

(126)

where C0 , C2 , C4 are related to the RamondRamond fluxes A in the following way:


C = C0 + C2 + C4 = (A0 + A2 + A4 )eB .

(127)

Here we consider A0 , C0 zero-forms, A2 , C2 and B two-forms and A4 , C4 four-forms.


The coupling of the fields on the brane world volume to the RamondRamond fluxes is:


Z
1
F B .
(128)
A tr exp
2i
The formula (127) can be verified by considering the transformations of C under the
element of the T-duality group which shifts the B-field. For the large volume torus (or
K3) we have the following relation between the numbers of branes (Q4 , Q2 , Q0 ) and the
parameters of the (1 + 5)-dimensional YangMills:


1
F .
(129)
(Q4 , Q2 , Q0 ) = tr exp
2i
(For the large volume torus, the YangMills theory lives on D5-branes, and the rank of
the gauge group is Q4 . The sign of Q0 may be understood in the following way. Let us
consider the D1D5 system wrapped on K3. On one hand, to get BPS configuration we
should have Q0 Q4 >R0. On the other hand, D-strings should become anti-self-dual gauge
1
F )2 < 0.)
fields on K3, so that tr( 2i
The six-dimensional string coupling constant is related to the ten-dimensional gstr :
gstr
(130)
= .
V
This is invariant under T-duality.

Appendix B. An instanton in the presence of a string: an example


Here we want to prove that the configuration consisting of the black string with the
charge Z and the D-instanton with the charge vector preserves 18 of supersymmetry, if
( Z) = 0, ( PW4 Z) = 0 and ||||2 > 0. We will consider the configuration consisting
of the black string with Q2 = 0 (only D1 and D5, no D3), and the instanton obtained by

586

A. Mikhailov / Nuclear Physics B 584 (2000) 545588

wrapping the D1-string on the holomorphic cycle in T4 . It is enough to consider this


configuration, because any other pair of orthogonal vectors and Z can be brought to this
one by SO(4, 4, R) transformations (we do not need this transformations to be defined over
Z, because we want just to understand whether supersymmetry is preserved, and this is the
property of the classical supergravity solution).
Let us temporarily turn off the B-field. The Euclidean Dp-brane stretched along
the coordinates y 1 , . . . , y p+1 preserves the combination of supercharges, satisfying the
condition:
1 p+1 L = iR .

(131)

For example, for the D1 -instanton, L = iR [54]. Let us denote x 0 and x 5 coordinates
parallel to the black string, and x 1 , . . . , x 4 coordinates in the torus. The supersymmetries
preserved by the D1/D5 system are:
0 5 L = iR ,
0 5 1 4 L = iR .

(132)

Suppose that the cycle on which we wrap our D-string to get an instanton is holomorphic
in some complex structure of the torus. The corresponding Khler form R can act
on spinors (being contracted with gamma-matrices). We will denote the corresponding
operator R . The supersymmetry preserved by the instanton is selected by the condition
that L is chiral in the tangent space to T4 and R is:
R L = iR .

(133)

This equation, together with (132) implies that L is chiral in the tangent space of T4 and
also subject to the condition R L = 0 5 L . Then, R is expressed in terms of L by one
of the equations (132). We see that our background preserves 18 of the supersymmetry.
Our background was quite generic except that we have put B = 0. To turn on the B-field,
we use the subgroup SO(3, 3, R) SO(4, 4, R) which preserves both Z = (Q4 , 0, Q0 ) and
= (0, , 0). This groups contains rotations in the plane generated by (Q4 , 0, Q0 ) and
(0, , 0), where H 2 (T4 , R) is orthogonal to . One can see that infinitesimal rotations
in such a plane transform our background with zero B field to the background with the
infinitesimal B-field B satisfying the condition (B ) = 0 and otherwise arbitrary. This
means, that there is one and only one condition for the background with nonzero B to be
connected to the background with zero B by the rotation orthogonal to both Z and . This
condition is (Z PW4 ) = 0.
References
[1] A. Connes, Noncommutative Geometry, Academic Press, 1994.
[2] M. Douglas, C. Hull, D-branes and noncommutative geometry, JHEP 9802 (1998) 008, hepth/9711165.
[3] A. Connes, M. Douglas, A. Schwarz, Noncommutative geometry and Matrix theory: compactification on tori, JHEP 9802 (1998) 003, hep-th/9711162.

A. Mikhailov / Nuclear Physics B 584 (2000) 545588

587

[4] Y.-K.E. Cheung, M. Krogh, Noncommutative geometry from 0-branes in a background B field,
Nucl. Phys. B 528 (1998) 185.
[5] C.-S. Chu, P.-M. Ho, Noncommutative open string and D-brane, Nucl. Phys. B 550 (1999)
151, hep-th/9812219; Constrained quantization of open string in background B field and
noncommutative D-brane, Nucl. Phys. B 568 (2000) 447456, hep-th/9906192.
[6] V. Schomerus, D-Branes and deformation quantization, JHEP 9906 (1999) 030, hepth/9903205.
[7] F. Ardalan, H. Arfaei, M.M. Sheikh-Jabbari, Mixed branes and (M)atrix theory on noncommutative torus, hep-th/9803067; Noncommutative geometry from strings and branes, JHEP 9902
(1999) 016, hep-th/9810072; Dirac quantization of open strings and noncommutativity in
branes, hep-th/9906161.
[8] N. Seiberg, E. Witten, String theory and noncommutative geometry, JHEP 9909 (1999) 032,
hep-th/9908142.
[9] M. Kontsevich, Deformation quantization of Poisson manifolds, q-alg/9709040.
[10] A.S. Cattaneo, G. Felder, A path integral approach to the Kontsevich quantization formula,
math.QA/9902090.
[11] M. Douglas, Two lectures on D-geometry and noncommutative geometry, hep-th/9901146.
[12] T. Filk, Divergences in a Field theory on a quantum space, Phys. Lett. B 376 (1996) 53.
[13] N. Nekrasov, A. Schwarz, Instantons on noncommutative R4 and (2, 0) superconformal sixdimensional theory, Commun. Math. Phys. 198 (1998) 689703, hep-th/9802068.
[14] J. Maldacena, The large N limit of superconformal Field theories and supergravity, Adv. Theor.
Math. Phys. 2 (1998) 231252, hep-th/9711200.
[15] A. Giveon, D. Kutasov, N. Seiberg, Comments on string theory on AdS3 , Adv. Theor. Math.
Phys. 2 (1998) 733780, hep-th/9806194.
[16] J. de Boer, H. Oogiri, H. Robins, J. Tannenhauser, String theory on AdS3 , JHEP 9812 (1998)
026, hep-th/9812046.
[17] D. Kutasov, N. Seiberg, More comments on string theory on AdS3 , JHEP 9904 (1999) 008,
hep-th/9903219.
[18] O. Aharony, M. Berkooz, N. Seiberg, Light-cone description of (2, 0) superconformal theories
in six dimensions, Adv. Theor. Math. Phys. 2 (1998) 119153, hep-th/9712117.
[19] S. Ferrara, R. Kallosh, A. Strominger, N = 2 extremal black holes, Phys. Rev. D 52 (1995),
hep-th/9508072.
[20] R. Dijkgraaf, Instanton strings and hyper-Khler geometry, Nucl. Phys. B 543 (1999) 545571,
hep-th/9810210.
[21] Y.-K.E. Cheung, O.J. Ganor, M. Krogh, A.Yu. Mikhailov, Instantons on non-commutative
T4 from twisted (2, 0) and little string theories, Nucl. Phys. B 564 (2000) 259284, hepth/9812172.
[22] B. Pioline, D-effects in toroidally compactified type II string theory, hep-th/9712155.
[23] B. Pioline, E. Kiritsis, U-duality and D-brane conbinatorics, Phys. Lett. B 418 (1998) 6169,
hep-th/9710078.
[24] O.J. Ganor, Two conjectures on gauge theories, gravity, and infinite-dimensional KacMoody
groups, hep-th/9903110.
[25] I.I. Kogan, G. Luzon, D-instantons on the boundary, Nucl. Phys. B 539 (1999) 121134, hepth/9806197.
[26] R. DAuria, P. Fre, BPS black holes in supergravity, hep-th/9812160.
[27] L.J. Romans, Self-duality for interacting fields, Nucl. Phys. B 276 (1986) 7192.
[28] L. Andrianopoli, R. DAuria, S. Ferrara, M.A. Lledo, Horizon geometry, duality and fixed
scalars in six dimensions, Nucl. Phys. B 528 (1998) 218228, hep-th/9802147.
[29] C.G. Callan, J.M. Maldacena, D-brane approach to black hole quantum mechanics, Nucl.
Phys. B 472 (1996) 591610, hep-th/9602043.

588

A. Mikhailov / Nuclear Physics B 584 (2000) 545588

[30] J.M. Maldacena, D-branes and near extremal black holes at low energies, Phys. Rev. D 55 (1997)
76457650, hep-th/9611125.
[31] S.J. Gates, C.M. Hull, M. Rocek, Twisted multiplets and new supersymmetric non-linear
-models, Nucl. Phys. B 248 (1984) 157186.
[32] A.A. Tseytlin, On non-Abelian generalization of BornInfeld action in string theory, Nucl.
Phys. B 501 (1997) 4152, hep-th/9701125.
[33] P. Aspinwall, K3 surfaces and string duality, hep-th/9611137.
[34] S. Ramgoolam, D. Waldram, Zero-branes on a compact orbifold, JHEP 9807 (1998) 009, hepth/9805191.
[35] E. Witten, Bound states of strings and p-branes, Nucl. Phys. B 460 (1996) 335350.
[36] N. Seiberg, E. Witten, The D1/D5 system and singular CFT, JHEP 9904 (1999) 017, hepth/9903224.
[37] A. Astashkevich, N. Nekrasov, A. Schwarz, On noncommutative Nahm transform, hepth/9810147.
[38] A. Konechny, A. Schwarz, BPS states on noncommutative tori and duality, Nucl. Phys. B 550
(1999) 561584, hep-th/9811159.
[39] M.A. Rieffel, A.S. Schwarz, Morita equivalence of multidimensional noncommutative tori, Int.
J. Math. 10 (1999) 289299, math.QA/9803057.
[40] C.A. Eliott, On the K-theory of the C -algebras generated by a projective representation of a
torsion-free discrete abelian group, in: Operator Algebras and Group Representations, Pitman,
London, 1984, pp. 157184.
[41] S.K. Donaldson, P.B. Kronheimer, The geometry of four-manifolds, Clarendon Press, Oxford
Univ. Press, New York, 1990.
[42] N.J. Hitchin, A. Karlhede, U. Lindstrm, M. Rocek, Hyper-Khler metrics and supersymmetry,
Comm. Math. Phys. 108 (1987) 535589.
[43] M. Berkooz, Nonlocal field theories and the noncommutative torus, Phys. Lett. B 430 (1998)
237, hep-th/9802069.
[44] M. Bershadsky, A. Johansen, V. Sadov, C. Vafa, Topological reduction of 4D SYM to 2D
-model, Nucl. Phys. B 448 (1995) 166, hep-th/9501096.
[45] E. Corrigan, C. Devchand, D.B. Fairlie, J. Nuyts, First-order equations for gauge fields in spaces
of dimension greater then four, Nucl. Phys. B 214 (1983) 452464.
[46] R.S. Ward, Completely solvable gauge-field equations in dimension greater than four, Nucl.
Phys. B 236 (1984) 381396.
[47] B.S. Acharya, J.M. Figueroa-OFarrill, B. Spence, M. OLoughlin, Euclidean D-branes and
higher-dimensional gauge theory, Nucl. Phys. B 514 (1998) 583602, hep-th/9707118.
[48] M. Blau, G. Thompson, Euclidean SYM theories by time reduction and special holonomy
manifolds, Phys. Lett. B 415 (1997) 242252, hep-th/9706225.
[49] R. Friedman, J. Morgan, E. Witten, Vector bundles and F theory, Commun. Math. Phys. 187
(1997) 679743, hep-th/9701162.
[50] P. Griffiths, J. Harris, Principles of Algebraic Geometry.
[51] E. Witten, D-branes and K-theory, JHEP 9812 (1998) 019, hep-th/9810188.
[52] D.S. Freed, E. Witten, Anomalies in string theory with D-branes, hep-th/9907189.
[53] C.H. Taubes, Self-dual connections on 4-manifolds with indefinite intersection matrix, J. Diff.
Geom. 19 (1984) 517560.
[54] M. Green, M. Gutperle, Effects of D-instantons, Nucl. Phys. B 498 (1997) 195227, hepth/9701093.

Nuclear Physics B 584 (2000) 589614


www.elsevier.nl/locate/npe

Instantons and Gribov copies in the maximally


Abelian gauge
F. Bruckmann

a, ,

T. Heinzl a,1 , A. Wipf a , T. Tok b,1

a Friedrich-Schiller-Universitt Jena, Theoretisch-Physikalisches Institut, Max-Wien-Platz 1,

D-07743 Jena, Germany


b Universitt Tbingen, Institut fr Theoretische Physik, Auf der Morgenstelle 14,

D-72076 Tbingen, Germany


Received 1 February 2000; revised 30 May 2000; accepted 5 June 2000

Abstract
We calculate the FaddeevPopov operator corresponding to the maximally Abelian gauge for
gauge group SU(N). Specializing to SU(2) we look for explicit zero modes of this operator. Within
an illuminating toy model (YangMills mechanics) the problem can be completely solved and
understood. In the field theory case we are able to find an analytic expression for a normalizable zero
mode in the background of a single t Hooft instanton. Accordingly, such an instanton corresponds
to a horizon configuration in the maximally Abelian gauge. Possible physical implications are
discussed. 2000 Elsevier Science B.V. All rights reserved.

1. Introduction
The configuration space A of gauge theories is a bigger-than-real-life-space [1]. This
is due to the fact that the action of the gauge group G relates physically equivalent
configurations along the gauge orbits. Therefore, this action has to be divided out. In
principle, this division leads to the physical configuration space, Aphys = A/G. In practice,
however, this division is not easily performed. The most efficient method to do so is gauge
fixing, where a subset of A is identified with Aphys. This subset is characterized by choosing
some condition on the gauge potentials A of the form [A] = 0. Prominent examples are
the covariant gauge, cov = A , or the axial gauge, ax = n A. One hopes that this
condition satisfies both the requirements of existence and uniqueness. Existence means
that the hypersurface : = 0 intersects every orbit, while uniqueness requires that it
does so once and only once. It has first been shown by Gribov that the latter requirement
Corresponding author. E-mail: bruckmann@tpi.uni-jena.de
1 Supported by DFG.

0550-3213/00/$ see front matter 2000 Elsevier Science B.V. All rights reserved.
PII: S 0 5 5 0 - 3 2 1 3 ( 0 0 ) 0 0 3 5 5 - 2

590

F. Bruckmann et al. / Nuclear Physics B 584 (2000) 589614

cannot be satisfied for non-Abelian gauge theories in the covariant and Coulomb gauge [2].
Shortly afterwards, Gribovs observation has been proven for a large class of continuous
gauge fixings [3]. In the physics community, the lack of uniqueness has become known
as the Gribov problem. This just paraphrases the difficulty in constructing the physical
configuration space which, by definition, is void of any (residual) gauge (or Gribov) copies.
In order to analyse this issue it has turned out useful to describe the gauge fixing not
simply by a condition = 0. Instead, in order to study the global aspects of the problem,
one formulates the gauge fixing procedure in terms of a variational principle [48]. To
this end one tries to define an action functional F in such a way that the associated
classical trajectories are nowhere parallel to the orbits so that their union defines a gauge
fixing hypersurface. By this construction one completely suppresses fluctuations in gauge
directions which in the unfixed formulation do not cost energy (or action) and thus make
the path integral ill-defined. Of course, by conservation of difficulties, one cannot avoid the
Gribov problem this way.
The variational approach to gauge fixing has mainly been studied for background type
gauges like the Coulomb gauge, where one can indeed construct a functional F [A; U ] with
the following generic properties: the critical points of F along the orbits generated by U
are the potentials A satisfying the Coulomb gauge condition, i Ai = 0. The Hessian of F
at these points is the FaddeevPopov operator FP. The Gribov region 0 is defined as the
set of transverse gauge fields for which det FP is positive. This is the set of relative minima
of F . Its boundary 0 is the Gribov horizon, where, accordingly, det FP = 0 because
the lowest eigenvalue of FP changes sign. It has been shown that 0 A is convex [7,9,
10]. This is basically due to the linearity of FP in A [11]. Contrary to early expectations
the Gribov region still contains Gribov copies [7,12,13]. Only if one restricts to the set
of absolute minima and performs certain boundary identifications, one ends up with the
physical configuration space (also called the fundamental modular domain) [11].
As stated above, the appearance of horizon configurations A 0 implies that the
gauge is not uniquely fixed; in other words, there are gauge fixing degeneracies. Somewhat
symbolically, this can be shown as follows. Let A have the infinitesimal gauge
variation A = D[A] , D denoting the covariant derivative. To check whether the gauge
transform A + A also satisfies the gauge condition, one calculates
A
N[A] D[A] .
(1.1)
[A + A] = [A] +
A
Here we have used that [A] = 0 and defined the normal N to the gauge fixing
hypersurface . Now, if A + A is also in we see that the FP operator,
FP[A] N[A] D[A] ,

(1.2)

must have a zero mode given by the infinitesimal gauge transformation . In this case,
there are two gauge equivalent fields A and A + A on and A is a horizon configuration.
From (1.2) one infers that there are two generic reasons for this to happen. First,
can be a zero mode already of D[A]. As the latter can be viewed as the velocity of a
fictitious motion along the orbits, its vanishing (on ) corresponds to a fixed point under
the action of the gauge group. In this case, the configuration A is called reducible [3,14].

F. Bruckmann et al. / Nuclear Physics B 584 (2000) 589614

591

Obviously, these are always horizon configurations. One might speculate whether there
is a gauge fixing such that reducible configurations are the only horizon configurations
[15]. The second possibility for det FP to vanish is that orbit velocity D and normal N
are orthogonal, which means that a particular orbit is tangent to . This is what usually
happens for background type gauges like the Coulomb gauge where N is constant, i.e.,
independent of A.
In general, it is very hard to explicitly find horizon configurations. For this reason, one
has to concentrate on rather simple and/or symmetric gauge potentials. Again, the case
best studied is the Coulomb gauge. It is known that there are Gribov copies of the classical
vacuum A = 0 [11,16,17]. An even simpler example of Gribov copies is provided by
constant Abelian gauge fields on the torus (the torons) [11,18]. Configurations with a radial
symmetry have been discussed in the original work of Gribov [2]. An explicit example with
axial symmetry has been given by Henyey [11,19].
On the lattice, the detection of Gribov copies has been reported for the first time in
[20]. It turns out that some of these copies are lattice artifacts while others survive in
the continuum limit [21]. In a sense, therefore, the Gribov problem becomes even more
pronounced upon gauge fixing on the lattice. This is of particular relevance for the lattice
studies of the dual superconductor hypothesis of confinement [2224], where one mainly
uses (a lattice version [25]) of t Hoofts maximally Abelian gauge (MAG) [6]. In order
to extract physical results within this approach one clearly has to control the influence of
Gribov copies. Finding the critical points of the lattice gauge fixing functional is similar to
a spin glass problem due to the high degree of degeneracy. The difficulties in numerically
determining the absolute maximum 2 of the lattice functional lead to an inaccuracy in
observables of the order of 10% [26,27].
The Gribov problem for the MAG so far has not been discussed in the continuum. The
purpose of this paper is to (at least partly) fill this gap. The MAG and its defining functional
will be reviewed in Section 2. The Hessian of this functional is the FP operator which is
calculated for gauge group SU (N). In Section 3 we specialize to SU (2) and give general
arguments showing the existence of a Gribov horizon. To provide some intuition, Section 4
introduces a simple toy model for which the FP operator and determinant can be calculated
exactly. The presence of Gribov copies is shown explicitly. Finally, in Section 5, we return
to field theory and calculate the FP operator in the background of a single instanton (in the
singular gauge). Again, we find an analytic expression for a normalizable zero mode which
shows that the single instanton is a horizon configuration in the MAG. Some technical
issues are discussed in Appendices A to C.

2. The maximally Abelian gauge


As explained in Appendix A, we decompose the gauge potential A into diagonal (Ak
and off-diagonal (A H ) components, A = Ak + A . The MAG is then defined
by minimizing the following functional

Hk )

2 The maximization of the lattice functional corresponds to a minimization of the continuum functional.

592

F. Bruckmann et al. / Nuclear Physics B 584 (2000) 589614


 2
F [A; U ] UA .

(2.1)

F is thus a functional of both the gauge field A and the gauge transformation U SU (N).
Via the parametrization

(2.2)
U () = exp(i) = exp i a T a , su(N) ,
F equivalently can be viewed as depending on the argument of U . The action of U on
A is
A = U 1 A U + iU 1 U.

(2.3)

With F of (2.1) we are thus minimizing the charged component A along its orbit, which,
roughly speaking, amounts to maximizing the Abelian or neutral component Ak . Hence
the name maximally Abelian gauge.
The YangMills norm in (2.1) is the same as in the YangMills action and induced by
the scalar product (A.6),
Z
(2.4)
kAk2 hA, Ai d d x tr A2 .
Note that our conventions are such that this norm is positive for hermitian gauge fields A
with values in su(N). The norm (2.4) can be viewed as the distance (squared) between A
and the zero configuration A = 0. As the space A of gauge potentials is affine, the norm is
gauge invariant in the following sense,


(2.5)
kA Bk = UA UB .
If the configuration B is kept fixed, however, the norm ceases to be gauge invariant and
explicitly depends on U or . The same is thus true for F which accordingly changes along
the orbit of A unless there is some (residual) invariance. For the functional (2.1) such an
invariance can indeed be found. Let V = exp(i k ) be an Abelian gauge transformation
and consider

 2
 2
(2.6)
F [A; V ] = VA = V 1 Ak V + V 1 A V + iV 1 dV .
As V is Abelian, the first and last terms on the r.h.s. of (2.6) vanish due to the projection
on H , and we are left with

 2
(2.7)
F [A; V ] = V 1 A V .
At this point it is crucial to note that V 1 A V is in H ,



tr Hi V 1 A V = tr V Hi V 1 A = tr Hi A = 0.

(2.8)

Therefore, we can write for (2.6),


2 2

F [A; V ] = V 1 A V = A = F [A; 1].

(2.9)

This immediately leads to the following Abelian invariance of F ,






F [A; V U ] = F UA; V = F UA; 1 = F [A; U ].

(2.10)

F. Bruckmann et al. / Nuclear Physics B 584 (2000) 589614

593

Note that our notation is such that U acts prior to V , i.e.,


A = (U V )1 A U V + i(U V )1 d(U V ).

VU

(2.11)

Roughly speaking, the Abelian invariance implies that F can be thought of as some kind
of mexican hat with the residual symmetry corresponding to (Abelian gauge) rotations
around its symmetry axis. Accordingly, the Hessian of F will have trivial zero modes
associated with the constant directions of F .
We are interested in the behavior of F [A; U ] around the point U = 1, i.e., = 0, on the
orbit of A. To this end we Taylor expand F as
1

F [A; U ] F [A; ] = F [A; 0] + F 0 [A; 0], + , F 00 [A; 0] +O 3 . (2.12)


2
In order to do so we need the gauge transform UA as a power series in . Using a standard
formula for the derivative of the exponential [11,28], one finds
exp(i ad ) 1
(D )
i ad
i
 i2 h
i
, [, D ] + .
(2.13)
= A + D + , D +
2
3!
Not surprisingly, the covariant derivative D = i ad(A ) with ad(A) B [A, B]
appears at this stage. Inserting (2.13) into (2.1) we obtain
2


+ 2 A , (D ) + (D ) , (D )
F [A; ] = A

+i A
(2.14)
, [, D ] + .
A = A +

In the following we are going to evaluate this expression term by term. This requires some
preparations. We will need the commutator identity,
hA, [B, C]i = hB, [C, A]i = hC, [A, B]i,

(2.15)

which follows straightforwardly from the definition of the scalar product. The latter
equation shows that both the operator ad(A) and the covariant derivative D[A] are antihermitean,
h, ad(A)i = had(A), i,

(2.16)

h, D[A]i = hD[A], i.

(2.17)
F 0,

The last two identities allow for an evaluation of the first derivative


A , (D ) = A
, D = D A , = D A , ,
k

(2.18)

with D i ad A . We thus have, to first order in ,


2



2 D k A , + O 2 .
F [A; ] = A

(2.19)

Note that to this order, F does not depend on the Cartan component k . We immediately
read off the critical points defining the MAG,

Dk A
D A = 0.

(2.20)

594

F. Bruckmann et al. / Nuclear Physics B 584 (2000) 589614

The second derivative requires considerably more efforts. We relegate the explicit
calculations to Appendix B, where we obtain for the Taylor expansion of F [A; ],
2



2 D k A , + ih , ad D A k
F [A; ] = A





k
, Dk Dk + ad2 A
i ad A QD i ad D A

(2.21)
+ O 3 .
Here we have defined a projection Q onto the complement H of the Cartan subalgebra
such that Q = . The term in (2.21) depending on k may seem somewhat strange but
is actually necessary to guarantee the Abelian invariance (2.10). It vanishes on the gauge
fixing hypersurface defined by (2.20).
From (2.21) we can easily read off the FaddeevPopov operator which is the Hessian of
F evaluated on (i.e., at the critical points),



(2.22)
FP = Q Dk Dk + ad2 A
i ad A QD Q.
In effect we have performed a saddle point approximation to the functional F [A; ]. The
equation of motion is the gauge fixing condition, and the fluctuation operator is the FP
operator. In this approximation the functional on reads
2

+ h, FPi + O 3 .
(2.23)
F [A; ] = A

As stated in the introduction, it is in general rather difficult (in a continuum formulation)


to find explicit examples of Gribov copies. The MAG is no exception from this rule.
The nontrivial task is to find normalizable zero modes of FP given by (2.22) which
is a complicated partial differential operator. We are, however, encouraged by lattice
calculations, in which such copies have been detected numerically, for the first time in
[29] and with refined techniques in [26,27]. One should keep in mind, though, that some
(if not all) of these copies can be lattice artifacts which do not survive in the continuum
limit. To study the possible appearance of Gribov copies in the continuum we have to
perform several simplifications. The first one will be to consider the case of gauge group
SU (2).

3. The FP operator for SU(2) general considerations


For SU(2), the gauge fixing condition (2.20) of the MAG can be rewritten in terms of

the gauge field components A3 Hk and A


H ,


1
2
(3.1)
iA3 A
= 0, A A iA .
The fact that these are only two requirements already implies (by counting of degrees of
freedom) that there remains a residual gauge freedom corresponding to a one-dimensional
subgroup which can only be U (1). Superficially, the gauge fixing looks like a background
gauge which would actually be true if the neutral component A3 were independent of
the charged one, A
. As these, however, are two components of one and the same
configuration they are not independent, and the gauge fixing condition is quadratic,

F. Bruckmann et al. / Nuclear Physics B 584 (2000) 589614

595

i.e., nonlinear in A . This makes life somewhat complicated. For instance, in order to
prove perturbative renormalizability within a BRST approach, one has to introduce fourghost terms [30,31]. The standard FP treatment thus is not sufficient. In a path integral
formulation, these ghost interactions regularize the usual bilinear FP ghost term in the
presence of possible zero modes [32]. For the MAG, such zero modes have never been
constructed explicitly. One objective of this paper is to remedy this situation.
The FP operator for SU (2) simplifies considerably as the last term in (2.22) vanishes.
One is thus left with the following sum of two operators,

(3.2)
FP = Q Dk Dk + ad2 A
Q.
Using the notation (A.8), FP can be viewed as a 3 3 matrix in color space. The operator
Q projects onto the two directions perpendicular to the z-axis so that the third row and
column of FP vanish identically. The associated trivial zero mode corresponds to the
residual U (1) gauge freedom which remains unfixed by the MAG. Explicitly, one has for
the nonvanishing entries of FP,
a b



= a b 2 A3 A3  a b A3 + 2A3 ,
(3.3)
Dk Dk


b
a

= a b Ac Ac Aa Ab .
(3.4)
ad2 A

Summing these two terms leads to the representation of FP given in equation (12) of [33]. 3
k k
Being (the negative of) a Laplacian, the operator D D is nonnegative. The same is

true for ad(A


) ad(A ) as will be shown in what follows. We define the hermitean matrix
C via
 
(3.5)
A , iC,
and calculate, using (2.16),





, Q ad A
ad A Q = ad A , ad A
= hiC, iCi = hC, Ci > 0.

(3.6)
(3.7)

One can as well use the representations (3.3), (3.4) and the CauchySchwarz inequality to
end up with the same result. The SU (2) FP operator from (3.2) is thus the difference of two
positive semidefinite operators which we abbreviate for the time being as
FP = A B,

A, B > 0.

(3.8)

The inequality denotes the fact that A and B have nonnegative spectrum. The identity (3.8)
already suggests that if B is sufficiently large, FP will develop a vanishing eigenvalue.
Let us make this statement slightly more rigorous. To this end we modify an argument used
in [14,34] for background type gauges.
First of all we note that together with the configuration (Ak , A ) also the scaled
configuration (Ak , A ), with some (positive real) parameter, will be in the MAG. The
associated FP operator is


(3.9)
FP Ak , A FP() = A 2 B.
3 Note, however, that in this reference the gauge potentials are defined as being anti-hermitean.

596

F. Bruckmann et al. / Nuclear Physics B 584 (2000) 589614

Fig. 1. Qualitative behavior of the lowest eigenvalue of FP as a function of the flow parameter .
The parameter value h corresponds to a horizon configuration.

Let us denote the lowest eigenvalue and the associated eigenfunction of FP() by E0 ()
and 0 (), respectively,
FP() 0 () = E0 () 0 ().

(3.10)

From (3.8) one must have E0 (0) > 0. If we turn on , a straightforward application of the
HellmannFeynman theorem leads to

E0 () = 2 0 (), B 0 () 6 0,

(3.11)

whence the function E0 () has negative slope. In addition, it has to be concave [35] 4 so
that, for sufficiently large, there will be a zero-mode at some value, say h (see Fig. 1).
In a way we have thus determined a path within the MAG fixing hypersurface that leads
us from the interior of the Gribov region ( = 0) to its boundary ( = h ).
As a result we can state that generically there have to be Gribov copies within the MAG
if the non-diagonal components A of the gauge fields become sufficiently large.

4. A toy model
In order to have an illustration of the somewhat abstract notions of the preceding sections
we will analyse an example with a finite number of degrees of freedom [36]. To this end
we employ a Hamiltonian formulation in d = 2 + 1 and consider only gauge potentials
A which are spatially constant. Renaming Aai = xia , i = 1, 2, a = 1, 2, 3, the Lagrangian
becomes
4 It is exactly for this reason that the second order perturbation theory correction to any groundstate is always
negative.

F. Bruckmann et al. / Nuclear Physics B 584 (2000) 589614

597

2
1 ab b 2 1 a
D0 xi xi  abc Ac0 xib .
(4.1)
2
2
One way of arriving at this Lagrangian is by gauging a free particle Lagrangian L0 =
xia xia /2 via minimal substitution, i.e., by replacing the ordinary time derivative 0 with the
covariant derivative D0ab . To keep things as simple as possible, we have not introduced
any (YangMills type) interaction; we are anyhow only interested in the kinematics of the
problem.
Defining the canonical momenta pia = D0ab xib , the Lagrangian (4.1) can be recast in first
order form
1
(4.2)
L = pia xia pia pia + Aa0 Ga ,
2
where we have introduced the operator Ga leading to Gausss law
L=

Ga  abc xib pic Diab pib = 0.

(4.3)

Obviously, Ga is the total angular momentum of two point particles in R3 (= color


isospace) with position vectors x 1 and x 2 . Gauge transformations are thus SO(3) rotations
of these vectors which do not change their relative orientation (i.e., the angle in-between
them). This is illustrated in Fig. 2.
As usual we will work in the Weyl gauge, A0 = 0, so that Gausss law has to be imposed
by hand, and, after quantization, holds upon acting on physical states. Once the Weyl
gauge has been chosen, there still is the freedom of performing time independent gauge
transformations. This will be (partially) fixed using the MAG. For the case at hand, there
are several equivalent ways of formulating the latter.
To avoid writing too many indices we denote x 1 x = (x, y, z), x 2 X = (X, Y, Z).
An arbitrary vector A will be decomposed according to

Fig. 2. An isospace (gauge) rotation (by an angle ) in the toy model, transforming the configuration
(x1 , x2 ) (x10 , x20 ). The lengths of the vectors and the angle in-between them are in-variant.

598

F. Bruckmann et al. / Nuclear Physics B 584 (2000) 589614

Ak Az ez ,

(4.4)

A Ax ex + Ay ey ,

(4.5)

which represents the decomposition into Cartan (= z) component and its complement. The
MAG condition then reads explicitly
b c
xi = 0,
a Diab (xik )xi =  abc xik

(4.6)

or, in components,
1 = yz Y Z = 0,
2 = xz + XZ = 0,

(4.7)

= 0.
3

The last condition is just an empty tautology so that there are in fact only two gauge
conditions. 5 Of course, this just corresponds to the fact that the gauge rotations generated
by G3 (the rotations around the z-axis) remain unfixed (cf. the remark after (3.1)).
The MAG conditions (4.7) can be easily visualized. The projections x and X have to
be collinear with their magnitudes being related through
|z|x = |Z|X .

(4.8)

The MAG is thus obtained by rotating the configuration (x, X) in such a way that both
vectors are as close to the z-axis as possible. This is achieved as shown in Fig. 3. x and X
are the diagonals of two rectangles with sides |z|, x and |Z|, X , respectively. If the areas

Fig. 3. The MAG condition in the toy model. The areas A and a have to be the same. We have
arbitrarily chosen x and X to lie in the yz-plane. The residual U (1) gauge freedom corresponds to
rotations around the z-axis.
5 In Diracs terminology [37], 3 is strongly zero and thus does not contribute in the calculation of any Poisson
bracket.

F. Bruckmann et al. / Nuclear Physics B 584 (2000) 589614

599

a and A of the rectangles coincide, a = A, the configuration is in the MAG. Algebraically,


the notion of being close to the z-axis is measured by the function
2
2
+ X
.
F (x, X) x

(4.9)

One can easily show that the conditions (4.6) or (4.7) minimize F and thus make the
nondiagonal components of x and X as small as possible. We mention in passing that the
trivial solution of (4.7) given by z = Z = 0 corresponds to a maximum of F so that we can
always assume z or Z 6= 0 (except for the zero-configuration representing the origin).
It is obvious from Fig. 3 that rotations around the z-axis leave both F and the MAG
condition invariant and thus correspond to a residual U (1) gauge freedom. As expected,
this situation is reflected in the FP operator,


(4.10)
FP = a , Gb =0 ,
which, in matrix notation, can be written as

2
xy + XY
0
z + Z2 y 2 Y 2
FP =
xy + XY
z2 + Z 2 x 2 X 2 0 .
0
0
0

(4.11)

The zero entries in the third row and column are a trivial consequence of the residual U (1)
and correspond to the action of the Q-projection in (2.22). The eigenvalues of FP are found
to be
E3 = 0,

(4.12)

E+ = z + Z ,
2

E = z + Z
2

2
2

2
x

(4.13)
2
X
.

(4.14)

Let us concentrate on the eigenvalues E which are not related to the residual Abelian
gauge freedom. Configurations where one of these vanishes are located on the Gribov
horizon and reflect some non-trivial residual gauge freedom different from the U (1)
above. A particular (in some sense trivial) class of horizon configurations consists in the
reducible configurations as discussed in the introduction. These have a higher symmetry
than generic configurations (a nontrivial stabilizer or isotropy group). In other words, they
are fixed points under the action of (a subgroup of) the gauge group. Technically, they
show up by inducing zero modes of the Laplacian 1ab = Diac Dicb (see Appendix C).
Within our example, the reducible configurations are readily identified [36,38] by simple
symmetry considerations. The origin is invariant under the whole action of SO(3) while
configurations with x and X collinear are invariant under rotations around their common
direction which clearly corresponds to a U (1). This is nicely reflected in the spectrum of
FP. At the origin, both E vanish, while a collinear configuration can always be rotated
in the z-axis so that its stabilizer coincides with the standard residual U (1) corresponding
to E3 = 0. This U(1) stabilizer is thus hidden in the residual U (1). Fixing the latter by
demanding, e.g., x = X = 0, does, however, not affect configurations collinear along the
z-axis so that these will induce zero modes of FP even after residual gauge fixing [36].

600

F. Bruckmann et al. / Nuclear Physics B 584 (2000) 589614

2 + X2
Fig. 4. Behavior of the eigenvalues of FP in the toy model as a function of the magnitude x

2
2
of the nondiagonal components. A zero mode arises when x + X = z2 + Z 2 .

There is a remaining possibility for a vanishing eigenvalue. While E+ is always positive,


2 + X 2 . This happens for configurations where x and X are
E vanishes if z2 + Z 2 = x

of the same length and orthogonal to each other. Elementary trigonometry implies that
in this case the two areas a and A are always the same, irrespective of the location of
the configuration relative to the z-axis. Thus, there is an additional residual U (1) gauge
freedom for such exceptional configurations. This can be nicely illustrated in terms of a
2 + X 2 (see Fig. 4). We thus have found an explicit
spectral flow as a function of x

realization of the general results of Section 3, in particular of Fig. 1.

5. The FP operator in an instanton background


The natural question arising at this point is the following: is there a way of extending the
results of the toy model to the realistic field theory case? The answer given in this section
will be affirmative.
Our motivation stems from the observation made by Brower et al. [39] that the single
t Hooft instanton both in the singular and regular gauge satisfies the MAG condition
(2.20). For the instanton in the singular gauge 6 (or singular instanton, for short) given
by
sing

a
A (x) = 2

2
x
a /2,
2
2
x x + 2

(5.1)

with denoting the instanton size, the MAG fixing functional F is finite, while for the
instanton in the regular gauge,
6 We use the conventions of [40].

F. Bruckmann et al. / Nuclear Physics B 584 (2000) 589614


reg

a
A (x) = 2

x
a /2,
x2 + 2

601

(5.2)
sing

it diverges. The two configurations A


transformation

reg

and A

are related through the gauge

g(x) = x4 + i x a a ,

(5.3)

where x = x /r, r = (x 2 )1/2 denoting the modulus of the Euclidean position x. If we


adopt the point of view that we have to take the minima of F to define the Gribov region
sing
reg
0 of the MAG then A is located in 0 while A is not. This is corroborated by
the quoted work of Brower et al. [39] which, when translated into our language, amounts
sing
to the following. One numerically constructs a path (R) connecting A with
reg
A . Along this path 7 (beginning at the singular instanton) the MAG functional F is
monotonically rising. The configurations A (R) along the path are determined by applying
a (singular) gauge transformation  which takes the singular instanton to A (R), i.e.,
sing
A (R) = A . Hence (R) is a path both within and the single instanton orbit.
Accordingly, there must be an infinitesimal gauge transformation of the singular instanton
that does not leave and thus must be a zero mode of FP [Asing]. In what follows we will
try to explicitly determine this zero mode.
The first step of this program consists in the calculation of the FP operator in the
background of a singular instanton. Plugging (5.1) into (3.3) and (3.4) one obtains the
result

FPa b = a b 2 + 2  a b a(r)(x21 x1 2 + x3 4 x4 3 ).

(5.4)

We have discarded the vanishing third row and column (resulting from the action of Q)
and introduced the (singular) instanton profile function,


1
1
2 /r 2
=
2

.
(5.5)
a(r) = 2 2
r + 2
r 2 r 2 + 2
We are looking for normalizable zero modes of the FP operator,
FP = 0,

h, i < ,

(5.6)

where (x) now is a two-component vector (field) living in the complement of the Cartan
subalgebra. Solving Eq. (5.6) for the zero mode is basically an exercise in group theory as
will become clear in a moment. If we define the generators of four-dimensional Euclidean
rotations as
L = i(x x ),

, = 1, . . . , 4,

the FP operator can be written in 2 2 matrix notation as




2
2i a(r)(L12 L34 )
.
FP =
2
2i a(r)(L12 L34 )

(5.7)

(5.8)

7 In [39] the parameter R is the radius of a monopole loop associated with the configuration A (R) located on

sing
reg
somewhere in-between A = A (R = 0) and A = A (R = ).

602

F. Bruckmann et al. / Nuclear Physics B 584 (2000) 589614

It is straightforward to check that the L indeed satisfy the Lie algebra of SO(4).
In analogy with the Lorentz group one introduces the angular momentum and boost
generators
1
Li ij k Lj k ,
2
Ki Li4 ,

(5.9)
(5.10)

and their linear combinations,


1
i i
x ,
(5.11)
Mi (Li Ki ) =
2
2
1
i i
x .
(5.12)
Ni (Li + Ki ) =
2
2
These can be viewed as the anti-self-dual and self-dual parts of L , if duality is
understood as the exchange of L and K. The operators Mi and Ni generate two independent
SU (2) subgroups with Casimirs M 2 and N 2 having eigenvalues M(M + 1) and N(N + 1),
respectively [41]. It is important to note that M and N will in general be half-integer,
M, N {0, 1/2, 1, . . .}.

(5.13)

This fact is well known from the algebraic treatment of the hydrogen atom which has a
hidden dynamical O(4) symmetry (see, e.g., [42]). In addition, as FP is a 22 matrix, it
can be expanded in terms of Pauli matrices, so that altogether we find the rather compact
result,
FP = 21 + 4a(r)M3 2 .

(5.14)

Plugging this into (5.6) results in a four-dimensional Schrdinger equation with spin having
a high degree of symmetry. A complete set of commuting observables is given by the
Casimirs M 2 and N 2 , their projections M3 and N3 (with eigenvalues m and n) and the
Pauli matrix 2 (eigenvalues s = 1). Replacing 2 by its eigenvalue and rewriting the
Laplacian in terms of the radial coordinate r we are left with

3
2
(5.15)
FP(s) r2 r + 2 M2 + N2 + 4a(r)M3 s.
r
r
This is indeed a 4d radially symmetric Hamiltonian. Upon closer inspection, the Casimir
term turns out to become even simpler. Using the representations (5.7), (5.9) and (5.10)
one finds that
N2 M2 = L K = 0,

(5.16)

so that FP finally becomes


3
4
(5.17)
FP(s) = r2 r + 2 M2 + 4a(r)M3 s.
r
r
The eigenfunctions of FP will therefore depend on the quantum numbers M
{0, 1/2, 1, . . .}, m, n {M, M + 1, . . . , M} and s = 1. Choosing the coordinates
x = r (cos cos 12 , cos sin 12 , sin cos 34 , sin sin 34 ),

(5.18)

F. Bruckmann et al. / Nuclear Physics B 584 (2000) 589614

603

with 0 6 6 /2, 0 6 12 , 34 6 2 , the eigenfunctions can be written as follows,


= fMm (r) hMmn ( ) ymn (12 ) zmn (34 ) s .
The s are the eigenspinors of 2 ,


1
1
=
.
2 i

(5.19)

(5.20)

The Schrdinger equation factorizes accordingly. Introducing the dimensionless variable


R = r/ and defining a function g(R) via
f (R) g(R)/R 3/2 ,
(we omit the subscripts of f ) the radial equation for the zero mode becomes


4M(M + 1) + 3/4
8ms
2
+ 2
g(R) = 0.
R +
R2
R (1 + R 2 )

(5.21)

(5.22)

We are looking for a normalizable zero mode, or, in other words, a bound state with
vanishing energy. For this we need an attractive potential. We thus must have ms < 0, and
we choose s = 1, m > 0 in what follows. The bound state equation (5.22) thus becomes


4M(M + 1) 8m + 3/4
8m
2
+
g(R) = 0.
(5.23)
R +
R2
1 + R2
This equation has already been obtained by Brower et al. [39] in the stability analysis of
their monopole solutions. These authors, however, have overlooked the fact that M is halfinteger which is crucial for obtaining the correct solution (see below). In addition they
approximated the profile function a(r) by 1/r 2 (in the limit of small monopole loops). We
will instead solve (5.23) exactly. The latter is an effective one-dimensional Schrdinger
equation with a Hamiltonian
HR R2 + V1 (R) + V2 (R).

(5.24)

The second potential term, V2 , is always positive (for m > 0). Only the first term, V1 has
a chance of becoming negative leading to attraction. As m is bounded by M, the Casimir
term M(M + 1) in (5.23) will always win for large M. We thus should make M as small
and m as large as possible, implying m = M. From (5.23), there is exactly one solution for
M which makes V1 negative, namely M = 1/2 = m. We have explicitly checked that for
M > 1/2 there is no bound state solution. 8 The associated potential V1 + V2 is plotted in
Fig. 5. For M = 1/2, the normalizable solution of (5.23) is given by




1
(5.25)
g(R) = R 1 (1 + R 2 ) ln 1 + 2 .
R
Close to the origin, f (R) = g(R)/R 3/2 behaves as
f (R) =


1
(1 + 2 ln R) R(1 2 ln R) + O R 2 ,
R

(5.26)

8 The claim of Brower et al. [39], that attraction occurs for m = 1 with the ground state having M = 1, thus
cannot be substantiated.

604

F. Bruckmann et al. / Nuclear Physics B 584 (2000) 589614

Fig. 5. The bound-state potential V1 + V2 as a function of R = r/ for the attractive case (quantum
numbers M = m = 1/2).

while asymptotically it drops as 1/R 3 . Both types of behavior are sufficient to make f (or
) normalizable. The radial wave function f (R) and the associated probability distribution
p(R) = R 3 f 2 (R) are shown in Fig. 6. From this figure it is obvious that f has no nodes
and therefore corresponds to the ground state in the sector with M = 1/2 (cf. the analogous
reasoning in [11]).
The degeneracy of the solution is found as follows. FP does not depend on N3 , therefore
n can arbitrarily be chosen as an half-integer from {M, M + 1, . . . , M}, i.e., for M =
1/2, one has n = 1/2. Furthermore, FP is invariant under (m, s) (m, s), so that,
altogether, there is a four-fold degeneracy. In terms of abstract states |M, m, n, si the zero
modes are linear combinations of the four degenerate basis states |1/2, 1/2, 1/2, i and
|1/2, 1/2, 1/2, +i.
To explicitly determine the zero mode, we still have to find the functions hMmn , ymn and
zmn for M = 1/2. ymn and zmn are eigenfunctions of the operators L3 and K3 so that their
product becomes an eigenfunction of the two operators



1
i

,
(5.27)
M3 = (L3 K3 ) =
2
2 12 34


1
i

+
,
(5.28)
N3 = (L3 + K3 ) =
2
2 12 34
according to
M3 ymn zmn = m ymn zmn ,

(5.29)

N3 ymn zmn = n ymn zmn .

(5.30)

F. Bruckmann et al. / Nuclear Physics B 584 (2000) 589614

605

Fig. 6. The radial wave function f (R) of the zero mode and the associated probability distribution
p(R). While f diverges at the origin, p vanishes due to the radial measure factor R 3 .

Explicitly, one finds


ymn (12 ) = ei(m+n)12 ,

(5.31)

zmn (34 ) = ei(mn)34 .

(5.32)

The function hMmn ( ) satisfies the differential equation





(m + n)2 (m n)2
1
sin 2
+ 4M(M + 1)

hMmn ( ) = 0. (5.33)
sin 2

cos2
sin2
For M = 1/2, we can circumvent solving this equation by considering only the two
extremal states in a multiplet with m = M, which obey
M |M, M, ni = 0.

(5.34)

The associated differential equation is much simpler than (5.33) and straightforwardly
solved in terms of the functions
hM,M,n ( ) = cosMn sinM+n ,
hM,M,n ( ) = sinMn cosM+n .

(5.35)

Direct application to m = 1/2 finally yields the four degenerate zero modes for M = 1/2
(using the notation mns ),

606

F. Bruckmann et al. / Nuclear Physics B 584 (2000) 589614

1/2,1/2,+(x)
1/2,1/2,+(x)
1/2,1/2,(x)
1/2,1/2,(x)

=
=
=
=

cf (r) cos ei12 +


cf (r) sin ei34 +
cf (r) cos ei12
cf (r) sin ei34

1 ,
4 ,
2 ,
3 ,

(5.36)

where c denotes a normalization constant which will be determined in a moment. To this


end we rewrite the measure
d 4 x = r 3 dr cos sin d d12 d34 ,
and calculate the integral ( denoting any of the basic zero modes)


Z
2 !
2
1+
= 1.
d 4 x (x) (x) = c2 4
6
3

(5.37)

(5.38)

This determines the normalization c. Any zero mode of FP satisfying (5.6) must be a
linear combination of the four basis modes (5.36). For the following considerations it is
convenient to introduce the real basis,


1
c f (r) x4
(3 4 ) =
,
1
2i
2 r  x3 
1
c f (r) x3
,
2 (3 + 4 ) =
2
2 r  x4 
(5.39)
1
c f (r) x2
(1 2 ) =
,
3
2i
2 r  x1 
1
c f (r) x1
,
4 (1 + 2 ) =
2
x2
2 r
which, upon using the properties of t Hoofts symbols [40], can be compactly written as
c
a
x .
(5.40)
a (x) = f (r)
2
A general linear combination thus assumes the form

2
a
a
n a F (r) n
x ,
(5.41)
(x)
c
where n is a constant four vector, and F (r) f (r)/r.
At this point it is due time to ask for the physics associated with the zero mode a . From
sing
the argument given in the introduction, it is clear that with the singular instanton A also
sing
its infinitesimal neighbor, A + D , is in the MAG. However, applying the finite gauge
a

transformation  = exp(i a /2) to the singular instanton leads to a configuration that is


no longer in the MAG. This is at variance with the solution R found by Brower et al. [39]
which yields a monopole configuration within the MAG.
One might speculate that the zero mode (5.41) is induced by some spacetime symmetry
of the instanton. Recall that a configuration A is called symmetric under a symmetry
operation R if R translates A along its orbit, i.e., RA = A. In other words, the action
of R is compensated by the gauge transformation 1 . It should, however, be noted
that a symmetry of a configuration does not necessarily imply a symmetry of the MAG

F. Bruckmann et al. / Nuclear Physics B 584 (2000) 589614

607

functional as the latter is not gauge invariant. If, on the other hand, there is such a
symmetry, then, by Goldstones argument, zero modes of the Hessian must be present (see
Section 2). In covariant background gauges, for instance, there indeed appears a whole
S 4 of gauge equivalent configurations induced by an SO(5) symmetry transformation of
the instanton [43]. Let us, therefore, have a closer look at this idea in the context of
the MAG. A subclass of the SO(5) symmetry is given by the combination of conformal
transformations and translations,

(5.42)
x = r 2 + 2 n + 2n x x .
Its action on the instanton can be gauge compensated by the infinitesimal change D ,
where the algebra element is given by [44]
a
x .
a 2n

(5.43)

If we extend our zero mode (5.41) by adding the appropriate color 3-component, a a,
(which is always possible as FP is blind against this component), we note that it is a
deformation of (5.43): the constant prefactor 2 gets replaced by the profile function
F (r). The spacetime dependence of the zero mode (5.41) is thus much more complicated
than that of the compensating algebra element (5.43). Interestingly, there is also a nonnormalizable zero mode of FP that interpolates between the two. It is given by changing
the profile function F (r) to
r 2 + 2
,
(5.44)
r2
which goes to a constant to large r. Asymptotically, this zero mode thus approaches the
conformal mode (5.43) (up to a factor 2). For small r, the two zero modes coincide as
both F and G go like 1/r 2 . The situation is depicted in Fig. 7.
We thus conclude that our zero mode (5.41) is not generated by any of the spacetime
symmetries of the instanton. It is maybe not too surprising that this result differs from the
one obtained in [43]. After all, the MAG represents a non-linear gauge fixing implying a
G(r) =

Fig. 7. Lines of constant modulus of the zero modes as a function of |x| = (x a x a )1/2 and x4 . (a)
Normalizable zero mode with profile function F (r); (b) non-normalizable zero mode with profile
function G(r). The modulus increases in half-integer steps from 0.5 to 3. Full lines correspond to
integer values. The conformal mode (5.43) would be given by straight vertical lines.

608

F. Bruckmann et al. / Nuclear Physics B 584 (2000) 589614

curved gauge fixing hypersurface. Background type gauges, on the other hand, are linear in
the gauge potential A. They thus lead to gauge fixing hyperplanes with constant normals.
The residual gauge symmetries in both cases will thus in general be distinct. Nevertheless,
a common feature of both gauges is the fact that the singular instanton is located on the
associated Gribov horizons. For the MAG, this has been unambiguously shown by the
explicit construction of the zero mode (5.41).

6. Discussion
Among the different Abelian gauges used for the lattice study of the dual superconductor
hypothesis, the MAG is the one that has been analysed in greatest detail. In this paper we
have tried to supplement these achievements by analytic investigations. As the gauge fixing
is nonlinear, this requires some effort. We have calculated the FP operator for general gauge
group SU (N). The result is fairly complicated; considerable simplifications only seem to
arise for gauge group SU (2). For this particular case we were able to show by quite general
reasoning that there must be Gribov copies. This finding was confirmed both for a simple
toy model and the full field theory. In the latter case it turns out that the singular instanton
is a horizon configuration in the MAG. The associated zero modes of the FP operator have
explicitly been constructed.
Let us finally discuss some possible physical consequences of our findings. The
two pronounced manifestations of the QCD vacuum are confinement and spontaneous
breakdown of chiral symmetry. As stated above, the MAG is well suited for studying
the former by checking dual superconductivity which is believed to be due to monopole
condensation [2224]. On the lattice, condensation of monopoles has been confirmed
for various Abelian gauges [4548]. The monopole vacuum, however, does not provide
a straightforward explanation of chiral symmetry breaking which is due to instantons
rather than monopoles [40,49,50]. It is thus of conceptual importance to relate these
two complementary pictures of the vacuum to each other. In computer experiments
correlations between instantons and monopoles have indeed been detected [39,5158].
The dynamical origin of these correlations, however, remains unclear, despite considerable
efforts to investigate this problem analytically, in the MAG [39,5860], the Polyakov
gauge [6165] and other Abelian gauges [66,67]. For the MAG, the situation is as follows.
There are basically three known solutions which represent finite transformations from the
singular instanton Asing into another MAG configuration. These are (i) the transformation
(5.3) to the regular gauge instanton Areg , (ii) the hedgehog transformation of Chernodub
and Gubarev [59], and (iii) the family of solutions {A(R)} given by Brower et al. [39],
interpolating between Asing and Areg . Of these solutions only (ii) and (iii) induce magnetic
monopoles. Solution (ii) leads to an infinite Dirac string, solution (iii) to a monopole
loop of radius R. The associated MAG functional F diverges in cases (i) and (ii). In
case (iii) it is finite, however such that F [A(R)] > F [A(0)] F [Asing]. As a result one
concludes that the instanton in the singular gauge defines the global minimum of F along
the single instanton orbit. In other words, the MAG functional does not support monopoles

F. Bruckmann et al. / Nuclear Physics B 584 (2000) 589614

609

associated with single instantons as these configurations give rise to a larger value of F .
This is actually consistent with lattice results. In [26,27,29] it was observed that the number
of monopoles decreases the better the MAG is fixed, i.e., the closer one approaches the
absolute maximum of the lattice MAG functional. Due to monopole dominance, the string
tension also becomes smaller. This effect might well be caused by the suppression of
monopole loops associated with single instantons.
In favor of the instantonmonopole correlation, Brower et al. argue that a possible
zero mode of FP can be interpreted as a kinematical instability of the singular instanton
against monopole formation. In the limit of small monopole loops, R  , their solution
(Eq. (31) in [39]) indeed is a zero mode of FP. It goes like sin 2 sin cos , and thus,
upon comparing with (5.35), is seen to correspond to M = 1. Therefore, from our general
analysis in the preceding section, it is not normalizable and thus should be discarded from
the stability analysis. It is probably not too surprising that singular gauge transformations
like the ones found in [39] lead to zero modes with diverging norm.
The physical interpretation of the normalizable M = 1/2 zero mode given in (5.41)
is not completely clear. We have seen that it is not due to any of the known space
time symmetries of the instanton. Contrary to our expectations, it also has nothing
to do with the solution of Brower et al. In particular, it does not induce monopole
singularities. Furthermore, as stated in the last section, the finite transformation even leads
out of the MAG. All this confirms the result that in the MAG single instantons are not
correlated with monopoles. One is thus left with a possible correlation between multiinstanton configurations and monopoles. Numerically, this has been observed [39,5155].
In particular, the instantonanti-instanton (IA) system seems to be physically interesting.
In this case one finds that both I and A are surrounded by a single monopole loop if the
IA distance is large. Below a critical distance, however, the two loops merge into a single
one [39,54] which can be viewed as a kinematical precursor to monopole percolation. Of
course, an analytic treatment of multi-instanton systems is quite involved, but maybe not
hopeless. In this respect let us just mention Rossis old construction of the BPS monopole
in terms of an infinite number of instantons aligned along the time axis [68]. We have
performed some preliminary investigations of the IA system which show that the simple
sum ansatz, AI A = AI + AA is not in the MAG. The ansatz suggested by Yung [69],
however, does fulfill the differential MAG conditions (3.1), though the MAG functional
probably diverges. Further work in this direction is surely necessary.

Acknowledgements
The authors thank M. Mller-Preussker for enlightening discussions and D. Hansen for
a careful reading of the manuscript. T.H. and T.T. acknowledge support under DFG grant
WI-777/3-2 and RE 856/4-1, respectively. Part of the research of T.T. was performed during
his stay at the Institute of Theoretical Physics in Jena.

610

F. Bruckmann et al. / Nuclear Physics B 584 (2000) 589614

Appendix A. Notations and conventions


The generators of SU (N) are hermitean matrices denoted by T a with normalization
tr(T a T b ) = ab /2. Any gauge field A = Aa T a is decomposed into a component Ak in the
Cartan subalgebra Hk su(N) and a component A in the complement, H , such that
su(N) = Hk H and
A = A k + A = A i H i + A E .

(A.1)

The different generators obey the commutation relations [70]


[Hi , Hj ] = 0,

i = 1, . . . , r,

[Hi , E ] = i E ,
[E , E ] = N E+ ,

(A.2)
(A.3)

+ 6= 0,

[E , E ] = i Hi .

(A.4)
(A.5)

The rank of the Lie algebra is denoted by r, the i are the roots, and N is a normalization
the value of which is not important for us. For SU (2), which has only two roots , the
third commutator (A.4) becomes obsolete, and the situation simplifies considerably.
The decomposition (A.1) is orthogonal with respect to the scalar product
Z
(A.6)
hA, Bi d d x tr AB,
where A and B denote some arbitrary Lie algebra valued L2 functions. Thus we have

k
(A.7)
A , B = 0.
We will also use an alternative notation [33] where we simply divide the N 2 1 generators
T a into neutral and charged ones by means of their superscripts, namely T a =
(T a0 , T a ) with T a0 Hk and T a H . A gauge field thus is decomposed as
A = Aa T a = Aa0 T a0 + Aa T a .

(A.8)

The superscripts a0 and a take on r and N 2 1 r values, respectively. For SU (2), for
example, we have a0 = 3 and a {1, 2}, while for SU (3), a0 {3, 8} etc.

Appendix B. The second derivative of the MAG functional


In this appendix we calculate the second derivative of the MAG functional given by the
last two terms in (2.14). First we evaluate (D ) ,







k
i A
(D ) = i Ak , i A
,
,



k
= D i A
, .
This yields for the square term in (2.14),

(B.1)

F. Bruckmann et al. / Nuclear Physics B 584 (2000) 589614

611






(D ) 2 = , D QD + 2i , A , D k + k , A , A , k




k
(B.2)
+ 2i , D A
, ,
where we have made use of the Leibniz rule,
D [B, C] = [D B, C] + [B, D C],

(B.3)

Cartan complement, QA = A . Note

that the last term


and defined a projection Q onto the
2
in (B.2) vanishes at the critical points (2.20). The second term of order in (2.14) is





k
i , A
i A
, [, D ] = i , A , D
, D




k
k
i , D A
i , A
,
, D



k
(B.4)
i k , A
, D .
Adding (B.2) and (B.4) we see that the terms which mix and D k cancel. The O( 2 )
term in F thus becomes




k
+ i , D A
F (2) [A; ] , D QD i , A
, D
,

 k 


k
i k , A
(B.5)
+ k , A
, A ,
, D .
The two terms bilinear in k add up to zero whence expression (B.5) simplifies to





k
+ i , D A
F (2) [A; ] = , D QD i A
, D
,





k
F (2) A; + i , D A
, .

(B.6)

Introducing P = 1 Q, the terms quadratic in assume the following form,





F (2) A; = , Dk Q D + i ad A
PD .

(B.7)

With the projections




PD = iP A
, ,



QD = Dk iQ A
, ,
k

(B.8)
(B.9)

and the identity h , QD Ai = h , D Ai, (B.7) becomes










i A
F (2) A; = , Dk Dk i Dk A
,
, QD
 

.
+ A
, A ,

(B.10)

This is the result used in (2.21).

Appendix C. The Laplacian of the toy model


Using matrix notation, the Laplacian 1ab = Diac Dicb of the toy model is given by

2
2
2
2
xy + XY
xz + XZ

y z Y Z

2
2
2
2
1=
. (C.1)
yz + Y Z
yx + Y X
x z X Z

2
2
2
2
zx + ZX
zy + ZY
x y X Y

612

F. Bruckmann et al. / Nuclear Physics B 584 (2000) 589614

Denoting r |x| and R |X|, the determinant of the Laplacian becomes




det 1 = r 2 + R 2 (x X)2 r 2 + R 2 d 2 6 0.

(C.2)

As is appropriate for a Laplacian, 1 is a negative-semidefinite operator. It has zero modes


for reducible configurations only [14], which for the case at hand are given by the zero
configuration, x = X = 0 (with the full group SO(3) as its stabilizer), and the collinear
configurations, x = X, with U (1) stabilizer.
The eigenvalues of 1 are given by
(C.3)
E0 = r 2 + R 2 ,
 1p
1 2
2
(r 2 R 2 )2 + 4(x X)2 .
(C.4)
E = r + R
2
2
It is reassuring to note that the eigenvalues and, accordingly, the determinant only depend
on the gauge invariant scalar products r 2 , R 2 and x X. At the origin, which has the
largest stabilizer, all eigenvalues vanish. For collinear configurations with x X = rR,
the eigenvalues are E = 0 and E0 = E+ = r 2 + R 2 , so that there is a zero mode and
the first excited state is degenerate. For the horizon configurations of the MAG, having
x X = 0, r = R , one finds E = 2 and E0 = 2 2 . Thus, the groundstate becomes
degenerate. The latter fact corresponds to the gauge fixing degeneracies of the Laplacian
gauge [7173] as particularly discussed in [72]. As the MAG and the Laplacian gauge
coincide for constant gauge fields, the degeneracies have to be the same, and, indeed, they
are.
References
[1]
[2]
[3]
[4]
[5]
[6]
[7]

[8]
[9]
[10]
[11]
[12]
[13]
[14]
[15]

K. Kuchar, Phys. Rev. D 34 (1986) 3031.


V. Gribov, Nucl. Phys. B 139 (1978) 1.
I.M. Singer, Commun. Math. Phys. 60 (1978) 7.
L.G. Yaffe, Nucl. Phys. B 151 (1979) 247.
K. Wilson, in: G. t Hooft et al. (Eds.), Recent Developments in Gauge Theories, Plenum, New
York, 1980, p. 363.
G. t Hooft, Nucl. Phys. B 190 (1981) 455.
M.A. Semenov-Tyan-Shanskii, V.A. Franke, Zapiski Nauchnykh Seminarov Leningradskogo
Otdeleniya Matematicheskogo Instituta im. V.A. Steklov AN SSR, Scientific Notes of the
Leningrad Branch of the Mathematical Institute V.A. Steklov 120 (1982) 159, Translation:
Plenum, New York, 1986, p. 1999.
W.I. Weisberger, in: A.H. Guth et al. (Eds.), Asymptotic Realms of Physics, MIT Press,
Cambridge, MA, 1983.
D. Zwanziger, Phys. Lett. B 114 (1982) 337.
D. Zwanziger, Nucl. Phys. B 209 (1982) 336.
P. van Baal, Nucl. Phys. B 369 (1992) 259.
G. DellAntonio, D. Zwanziger, in: P.H. Damgaard et al. (Eds.), Probabilistic Methods in
Quantum Field Theory and Quantum Gravity, Plenum, New York, 1990, p. 107.
G. DellAntonio, D. Zwanziger, Commun. Math. Phys. 138 (1991) 291.
O. Babelon, C.M. Viallet, Commun. Math. Phys. 81 (1981) 515.
T. Heinzl, in press.

F. Bruckmann et al. / Nuclear Physics B 584 (2000) 589614

[16]
[17]
[18]
[19]
[20]
[21]
[22]
[23]
[24]
[25]
[26]
[27]
[28]
[29]
[30]
[31]
[32]
[33]
[34]
[35]
[36]
[37]
[38]
[39]
[40]
[41]
[42]
[43]
[44]
[45]
[46]
[47]
[48]
[49]
[50]

[51]
[52]
[53]
[54]
[55]
[56]
[57]

613

R. Jackiw, I. Muzinich, C. Rebbi, Phys. Rev. D 17 (1978) 1576.


C.M. Bender, T. Eguchi, H. Pagels, Phys. Rev. D 17 (1978) 1086.
J. Koller, P. van Baal, Nucl. Phys. B 302 (1988) 1.
F.S. Henyey, Phys. Rev. D 20 (1979) 1460.
P. de Forcrand, Nucl. Phys. B (Proc. Suppl.) 20 (1991) 194.
P. de Forcrand, J.E. Hetrick, Nucl. Phys. B (Proc. Suppl.) 42 (1995) 861.
G. Parisi, Phys. Rev. D 11 (1975) 970.
S. Mandelstam, Phys. Rep. 23 (1976) 245.
G. t Hooft, in: A. Zichichi (Ed.), High Energy Physics, Proceedings of the EPS International
Conference, Palermo 1975, Editrice Compositori, Bologna, 1976.
A.S. Kronfeld, M.L. Laursen, G. Schierholz, U.-J. Wiese, Phys. Lett. B 198 (1987) 516.
G.S. Bali, V. Bornyakov, M. Mller-Preussker, K. Schilling, Phys. Rev. D 54 (1996) 2863, heplat/9603012.
A. Hart, M. Teper, Phys. Rev. D 55 (1997) 3756, hep-lat/9606007.
H. Rothe, Lattice Gauge Theory, World Scientific, Singapore, 1992, Chapter 14.
S. Hioki et al., Phys. Lett. B 271 (1991) 201.
H. Min, T. Lee, P. Pac, Phys. Rev. D 32 (1985) 440.
K.-I. Kondo, T. Shinohara, hep-th/0004158.
M. Schaden, hep-th/9909011.
M. Quandt, H. Reinhardt, Int. J. Mod. Phys. A 13 (1998) 4049, hep-th/9707185.
M. Creutz, I.J. Muzinich, T.N. Tudron, Phys. Rev. D 19 (1979) 531.
A. Galindo, P. Pascual, Quantum Mechanics II, Texts and Monographs in Physics, Springer,
Berlin, 1991, p. 127.
T. Pause, T. Heinzl, Nucl. Phys. B 524 (1998) 695, hep-th/9801169.
P.A.M. Dirac, Lectures on Quantum Mechanics, Benjamin, New York, 1964.
J.M. Arms, J.E. Marsden, V. Moncrief, Commun. Math. Phys. 78 (1981) 455.
R.C. Brower, K.N. Orginos, C.-I. Tan, Phys. Rev. D 55 (1997) 6313, hep-th/9610101.
T. Schfer, E.V. Shuryak, Rev. Mod. Phys. 70 (1998) 323, hep-ph/9610451.
P. Ramond, Field Theory A Modern Primer, Addison Wesley, Reading, 1990, p. 8 f.
E. Merzbacher, Quantum Mechanics, Wiley, New York, 1998, Chapter 12.
L. Baulieu, A. Rozenberg, M. Schaden, Phys. Rev. D 54 (1996) 7825, hep-th/9607147.
R. Jackiw, Diverse Topics in Theoretical and Mathematical Physics, World Scientific,
Singapore, 1995, Paper III.3.
L. Del Debbio, A. Di Giacomo, G. Paffuti, P. Pieri, Phys. Lett. B 355 (1995) 255.
N. Nakamura et al., Nucl. Phys. B (Proc. Suppl.) 53 (1997) 512, hep-lat/9608004.
M.N. Chernodub, M.I. Polikarpov, A.I. Veselov, Phys. Lett. B 399 (1997) 267, hep-lat/9610007.
A. Di Giacomo, B. Lucini, L. Montesi, G. Paffuti, Phys. Rev. D 61 (2000) 034503, heplat/9906024.
G. t Hooft, Phys. Rev. D 14 (1976) 3432.
D. Diakonov, in: A. Di Giacomo, D. Diakonov (Eds.), Selected Topics in Nonperturbative QCD,
Proceedings International School of Physics Enrico Fermi, Course CXXX, Varenna, Italy,
1995, IOS Press, Amsterdam, 1996, hep-ph/9602375.
M. Gckeler, A.S. Kronfeld, M.L. Laursen, G. Schierholz, U.-J. Wiese, Phys. Lett. 233 (1989)
192.
H. Suganuma, K. Itakura, H. Toki, O. Miyamura, hep-ph/9512347.
V. Bornyakov, G. Schierholz, Phys. Lett. B 384 (1996) 190, hep-lat/9605019.
A. Hart, M. Teper, Phys. Lett. B 371 (1996) 261, hep-lat/9511016.
S. Thurner, M. Feurstein, H. Markum, W. Sakuler, Phys. Rev. D 54 (1996) 3457.
H. Suganuma, S. Sasaki, H. Ichie, F. Araki, O. Miyamura, Nucl. Phys. B (Proc. Suppl.) 53
(1997) 528, hep-lat/9609033.
M. Fukushima et al., Phys. Lett. B 399 (1997) 141, hep-lat/9608084.

614

F. Bruckmann et al. / Nuclear Physics B 584 (2000) 589614

[58] S. Sasaki, O. Miyamura, Phys. Rev. D 59 (1999) 094507, hep-lat/9811029.


[59] M.N. Chernodub, F.V. Gubarev, JETP Lett. 62 (1995) 100, hep-th/9506026.
[60] R.C. Brower, K.N. Orginos, C.-I. Tan, Nucl. Phys. B (Proc. Suppl.) 53 (1997) 488, heplat/9608012.
[61] H. Reinhardt, Nucl. Phys. B 503 (1997) 505, hep-th/9702049.
[62] O. Jahn, F. Lenz, Phys. Rev. D 58 (1998) 085006, hep-th/9803177.
[63] C. Ford, U.G. Mitreuter, T. Tok, A. Wipf, J.M. Pawlowski, Ann. Phys. (NY) 269 (1998) 26,
hep-th/9802191.
[64] C. Ford, T. Tok, A. Wipf, Nucl. Phys. B 548 (1999) 585, hep-th/9809209.
[65] C. Ford, T. Tok, A. Wipf, Phys. Lett. B 456 (1999) 155, hep-th/9811248.
[66] E.-M. Ilgenfritz, S. Thurner, H. Markum, M. Mller-Preussker, Phys. Rev. D 61 (2000) 054501,
hep-lat/9904010.
[67] O. Jahn, hep-th/9909004.
[68] P. Rossi, Nucl. Phys. B 149 (1979) 170.
[69] A.V. Yung, Nucl. Phys. B 297 (1988) 47.
[70] B. Wybourne, Classical Groups for Physicists, Wiley, New York, 1974.
[71] G. Schierholz, J. Seixas, M. Teper, Phys. Lett. B 157 (1985) 209.
[72] J.V. Vink, U.-J. Wiese, Phys. Lett. B 289 (1992) 12, hep-lat/9206006.
[73] A.J. van der Sijs, Nucl. Phys. B (Proc. Suppl.) 53 (1997) 535, hep-lat/9608041.

Nuclear Physics B 584 (2000) 615640


www.elsevier.nl/locate/npe

Equations of motion for massive spin 2 field


coupled to gravity
I.L. Buchbinder a,b , D.M. Gitman a , V.A. Krykhtin c , V.D. Pershin d,
a Instituto de Fsica, Universidade de So Paulo, P.O. Box 66318, 05315-970, So Paulo, SP, Brasil
b Department of Theoretical Physics, Tomsk State Pedagogical University, Tomsk 634041, Russia
c Department of Theoretical and Experimental Physics, Tomsk Polytechnical University, Tomsk 634050, Russia
d Department of Theoretical Physics, Tomsk State University, Tomsk 634050, Russia

Received 5 November 1999; revised 2 May 2000; accepted 20 June 2000

Abstract
We investigate the problems of consistency and causality for the equations of motion describing
massive spin two field in external gravitational and massless scalar dilaton fields in arbitrary
spacetime dimension. From the field theoretical point of view we consider a general classical action
with non-minimal couplings and find gravitational and dilaton background on which this action
describes a theory consistent with the flat space limit. In the case of pure gravitational background all
field components propagate causally. We show also that the massive spin two field can be consistently
described in arbitrary background by means of the lagrangian representing an infinite series in the
inverse mass. Within string theory we obtain equations of motion for the massive spin two field
coupled to gravity from the requirement of quantum Weyl invariance of the corresponding twodimensional sigma-model. In the lowest order in 0 we demonstrate that these effective equations of
motion coincide with consistent equations derived in field theory. 2000 Elsevier Science B.V. All
rights reserved.
PACS: 04.40.-b; 11.10.Ef; 11.25.Db
Keywords: Higher spin fields; Strings

1. Introduction
The purpose of this paper is twofold first, to describe consistent theory of massive
spin 2 field interacting with external gravity from the point of view of classical field theory
and, second, to derive effective equations of motion for this field from string theory.
Corresponding author. E-mail: pershin@ic.tsu.ru

E-mail addresses: ilb@mail.tomsknet.ru (I.L. Buchbinder), gitman@fma.if.usp.br (D.M. Gitman),


krykhtin@phys.dfe.tpu.edu.ru (V.A. Krykhtin).
0550-3213/00/$ see front matter 2000 Elsevier Science B.V. All rights reserved.
PII: S 0 5 5 0 - 3 2 1 3 ( 0 0 ) 0 0 3 8 9 - 8

616

I.L. Buchbinder et al. / Nuclear Physics B 584 (2000) 615640

The problem of consistent description of higher spin fields interaction has a long history
but is still far from the complete resolution. Within framework of standard field theory
it is possible to construct a classical action for the higher spin fields only on specific
curved spacetime manifolds [16]. 1 For example, massive integer spins are described
by symmetric tensors of corresponding ranks and one can write down the equations of
motion and find classical actions for them but only in special spacetimes (e.g., in Ricci flat
spaces). It means that in such an approach gravity field does not feel the presence of higher
spins matter through an energymomentum tensor. So a consistent classical action for the
system of dynamical gravity and a higher massive field is still unknown and there are some
indications that perhaps it does not exist at all.
One of these indications comes from considering a KaluzaKlein decomposition of
Einstein gravity in D-dimensional spacetime into gravity plus infinite tower of massive
second rank tensor fields in (D 1)-dimensional world, masses being proportional to
inverse compactification radius. Then the resulting four-dimensional theory of spin 2 fields
interacting with gravity and with each other should be consistent as one started from the
ordinary Einstein theory and just considers it on a specific manifold. But as was shown in
[12], it is impossible to reduce this theory consistently to a finite number of spin 2 fields,
i.e., consistency can be achieved only if the whole infinite tower of higher massive fields
are present in the theory.
The main problem in higher spin fields theories is that introduction of interaction
may excite new unphysical degrees of freedom which lead in general to appearance of
negative norm states in the Hilbert space and to violation of causality. In free theories
these unphysical degrees are absent due to transversality and tracelessness conditions.
Lagrangian description of these conditions requires auxiliary fields vanishing on equations
of motion [1315]. In arbitrary interacting theories the auxiliary fields may become
dynamical and to make them vanishing again one has to impose some additional
restrictions on the kind of interaction.
As a first step towards a full dynamical theory of interacting higher spin fields one
usually tries to describe a single higher spin massive field in an external electromagnetic
or gravitational background [3,5,6,1619]. In this case one way to achieve consistency
is to impose appropriate restrictions on the external background field. For example, it is
well known that consistent theories of higher massive spin fields can be easily built in
the spacetimes of constant curvature. In the Section 2 we describe in detail the theory of
massive spin 2 field in curved spacetime and show how these restrictions on the external
gravitational background arise. As a result we will arrive to a one-parameter family of
lagrangian theories which describe consistent propagation of the spin 2 field in arbitrary
Einstein spacetime.
There exists another possible way to achieve consistency by constructing equations
for higher massive fields in form of infinite series in inverse mass (or, equivalently, in
curvature). A recent attempt in this direction was undertaken in [6] where, however, only
1 For description of higher spin massless fields on specific background see, e.g., [79], the case of a collection
of massless spin 2 fields was investigated in [10,11].

I.L. Buchbinder et al. / Nuclear Physics B 584 (2000) 615640

617

theories on symmetric Einstein manifolds were considered and consistent equations were
derived in the simplest approximation linear in curvature.
In this paper we demonstrate that consistent equations for the spin 2 massive field can, in
principle, be constructed as infinite series in inverse mass square in arbitrary gravitational
background. These kinds of infinite series arise naturally in string theory which represent
another approach for consistent description of higher spins interaction.
String theory contains an infinite number of massive fields with various spins interacting
with each other and with a finite number of massless fields. Unfortunately, there are
arguments that in string theory a general coordinate invariant effective field action
reproducing the correct S-matrix both for massless and massive string states does not exist
[20]. The full effective action for all string fields is not general coordinate invariant and
general covariance arises only as an approximate symmetry in effective action for massless
fields once all the massive fields are integrated out. That effective action for massive fields
cannot be covariant follows, for example, from the fact that terms cubic in massive fields
can contain only flat metric and there is no terms of higher powers [20].
Influence of massive string modes is negligible at low energies but they become
important, for instance, in string cosmology [21] (for a recent review see, e.g., [22]) so
it would be desirable to have consistent equations describing interaction of the massive
string fields with gravity.
And indeed, there exists a possibility to derive from the string theory some covariant
parts of equations for massive higher spins fields interacting with background gravity. The
aim of our paper is to show explicitly how this procedure works using as an example
dynamics of the second rank tensor from the first massive level of open bosonic string.
A convenient method of deriving effective field equations of motion from the string
theory is provided by the -model approach [2333]. Within this approach a string
interacting with background fields is described by a two-dimensional field theory and
effective equations of motion arise from the requirement of quantum Weyl invariance.
Perturbative derivation of these equations is well suited for massless string modes
because the corresponding two-dimensional theory is renormalizable and loop
expansion
corresponds to expansion of string effective action in powers of string length 0 .
Inclusion of interaction with massive modes [3446] makes the theory non-renormalizable but this fact does not represent a problem since in string theory one considers
the whole infinite set of massive fields. Infinite number of counterterms needed for
cancellation of divergences generating by a specific massive field in classical action leads
to renormalization of an infinite number of massive fields. The only property of the theory
crucial for possibility of derivation of perturbative information is that number of massive
fields giving contributions to renormalization of the given field should be finite. As was
shown in [3739] string theory does fulfill this requirement. To calculate -function for
any massive field it is sufficient to find divergences coming only from a finite number of
other massive fields and so it is possible to derive effective equations of motion for any
background fields in any order in 0 .
This procedure have serious limitation it does not allow to derive non-perturbative
contributions to the string effective action. It is well known that correct terms cubic in

618

I.L. Buchbinder et al. / Nuclear Physics B 584 (2000) 615640

massive fields (quadratic terms in equations of motion) cannot be found within perturbative
renormalization of -model. For example, perturbative -function of tachyon field is linear
in all orders in 0 and interacting tachyonic terms are due to non-perturbative (from the
point of view of two-dimensional field theory) effects [4749].
To find these non-perturbative contribution one should use the method exact renormalization group (ERG) [5055]. It is ideally suited for string theory because, first of all, its
formulation does not require any a priori defined perturbative scheme, and secondly, the
equations of exact renormalization group are quadratic in interaction terms thus resembling
the cubic structure of exact string field action. But ERG method also has a serious disadvantage as it requires explicit separation of classical action into free and interaction parts
and thus leads to non-covariant equations for background fields. From the general point
of view this fact does not contradict general properties of string theory. We have already
mentioned that exact effective action describing both massless and massive string modes
should be non-covariant. But non-covariance of the ERG method makes it rather difficult
to establish relations between string fields equations and ordinary field theory.
So in this paper we derive covariant equations of motion for massive string fields
interacting with gravity by means of ordinary perturbative analysis of quantum Weyl
invariance condition in the corresponding -model. Of course, perturbatively we can obtain
equations only linear in massive fields. It was noted long ago [20] that one can make such a
field redefinition in the string effective action that terms quadratic in massive fields (linear
terms in equations of motion) acquire dependence on arbitrary higher powers of massless
fields and so may be covariant. In this paper we explicitly obtain these interaction terms
in the lowest in 0 approximation. We do not get terms quadratic in massive fields which
should be non-covariant and arise only non-perturbatively.
As a model for our calculations we use bosonic open string theory interacting with
background fields of the massless and the first massive levels. First massive level contains
symmetric second rank tensor and so this model provides the simplest example of massive
higher spin field interacting with gravity. In order to obtain equations of motion for
string fields we build effective action for the corresponding two-dimensional theory,
perform renormalization of background fields and composite operators and construct the
renormalized operator of energymomentum tensor trace.
This rather standard scheme was first developed for calculations in closed string theory
with massless background fields [24,31,32] and then was generalized for the open string
theory [56,5860] and for strings in massive fields [3739]. The new feature appearing
in open string theory is that the corresponding two-dimensional sigma-model represents a
quantum field theory on a manifold with a boundary. To construct quantum effective action
in such a theory we use generalization of SchwingerDe Witt method for manifolds with
boundaries developed in the series of papers [61].
Making perturbative calculations we restrict ourselves to string world sheets with
topology of a disk. The resulting equations of motion for graviton will not contain
dependence on massive fields from the open string spectrum because these fields interact
only with the boundary of world sheet and so can not influence the local physics in the bulk.
For example, in the case of graviton and massive fields from the open string spectrum

I.L. Buchbinder et al. / Nuclear Physics B 584 (2000) 615640

619

one expects that equations of motion for the graviton should look like ordinary vacuum
Einstein equations without any matter. Of course, one can obtain contributions from open
string background fields to the right hand side of Einstein equations for gravity but it would
demand considering of world sheets of higher genus [57,59].
The organization of the paper is as follows. In the Section 2 we describe the most
general consistent equations of motion for massive spin 2 field on a specific class of curved
manifolds from the point of view of ordinary field theory and generalize them for the
interaction with scalar dilaton field in the Section 3. In the next section we show how the
scheme can be generalized to the case of arbitrary gravitational background by means of
either non-lagrangian equations or equations representing infinite series in inverse mass.
Section 5 contains description of the string model that we use for derivations of effective
equations of motion of string massive fields. We calculate divergences of the theory in the
lowest order, carry out renormalization of background fields and composite operators and
construct the renormalized operator of the energymomentum tensor trace. Requirement of
quantum Weyl invariance leads to effective string fields equations of motion. We compare
with each other equations of motion derived in two previous sections and show that string
theory gives (at least in the lowest order) consistent equations for the spin 2 massive field
interacting with gravity. Conclusion contains summary of the results.

2. Massive spin 2 field coupled to gravity in field theory


In this section we give detailed analysis of the lagrangian formulation for the free spin 2
massive field and then generalize it for the presence of external gravitational field. The
first requirement one should impose on such a theory is preservation of the same number
of degrees of freedom and constraints as in the flat theory. Another important feature of
any consistent relativistic theory is causality, which means that the equations of motion
should not describe superluminal propagation [16,17] (see also [18] for a review). The
result obtained in this section is a lagrangian which depends on one arbitrary dimensionless
parameter of non-minimal coupling and describes consistent propagation of the spin 2
massive field in arbitrary Einstein spacetime, i.e., it contains the same number of lagrangian
constraints as in the flat spacetime and does not violate causality.
In the flat spacetime the massive spin 2 field is described (as follows from the analysis
of irreducible representations of 4-dimensional Poincar group) by symmetric transversal
and traceless tensor of the second rank H satisfying mass-shell condition:
( 2 m2 )H = 0,

H = 0,

H = 0.

(1)

In higher-dimensional spacetimes Poincar algebras have more than two Casimir operators
and so there are several different spins for D > 4. Talking about spin 2 massive field in
arbitrary dimension we will mean, as usual, that this field by definition satisfies the same
equations (1) as in D = 4. After dimensional reduction to D = 4 such a field will describe
massive spin two representation of D = 4 Poincar algebra plus infinite tower of Kaluza
Klein descendants.

620

I.L. Buchbinder et al. / Nuclear Physics B 584 (2000) 615640

The most convenient approach to building interacting field theories is a lagrangian one
and, in fact, the general point of view is that any consistent equations of motion should
follow from some classical action. In the case of the free massive spin 2 field it is well
known that all the equations (1) can be derived from the FierzPauli action [13]:

Z
1
1
1
1
S = d D x H H H H H H + H H
4
4
2
2

m2
m2 2

H H +
H ,

(2)
4
4
where H = H .
The general scheme of calculating the complete set of constraints in lagrangian
formalism [63] is equivalent to the DiracBergmann procedure in hamiltonian formalism
and in the case of second class constraints (which is relevant for massive higher spin fields)
consists in the following steps. If in a theory of some set of fields A (x), A = 1, . . . , N the
original lagrangian equations of motion define only r < N of the second time derivatives
(accelerations) A then one can build N r primary constraints, i.e., linear combinations
of the equations of motion that does not contain accelerations. Requirement of conservation
in time of the primary constraints either define some of the missing accelerations or lead to
new (secondary) constraints. Then one demands conservation of the secondary constraints
and so on, until all the accelerations are defined and the procedure closes up.
Applying this procedure to the action (2) one can see that the equations of motion
E = 2 H 2 H + H + H H H
m2 H + m2 H = 0

(3)

contain D primary constraints (expressions without second time derivatives H ):


E00 = 1Hii i j Hij m2 Hii 0(1) 0,
E0i = 1H0i + i H kk k H ki i k H0k m

(4)
2

(1)
H0i i

0.

(5)

The remaining equations of motion Eij = 0 allow to define the accelerations H ij in terms
of H and H . The accelerations H 00 , H 0i cannot be expressed from the equations
directly.
Conditions of conservation of the primary constraints in time E 0 0 lead to D
secondary constraints. On-shell they are equivalent to
(2) = E = m2 H m2 H 0.

(6)

Conservation of i(2) defines D 1 accelerations H 0i and conservation of 0(2) gives


another one constraint. It is convenient to choose it in the covariant form by adding suitable
terms proportional to the equations of motion:
(3) = E +

m2
D1
E = H m4
0.
D2
D2

(7)

Conservation of (3) gives one more constraint on initial values


(4) = H 00 + H kk = H 0

(8)

I.L. Buchbinder et al. / Nuclear Physics B 584 (2000) 615640

621

and from the conservation of this last constraint the acceleration H 00 is defined.
Altogether there are 2D + 2 constraints on the initial values of H and H . The theory
contains the same local dynamical degrees of freedom as the system (1) and describes
traceless and transverse symmetric tensor field of the second rank.
Now if we want to construct a theory of massive spin 2 field on a curved manifold first
of all we should provide the same number of propagating degrees of freedom as in the flat
case. It means that new equations of motion E should lead to exactly 2D + 2 constraints
and in the flat spacetime limit these constraints should reduce to their flat counterparts. In
addition to consistency with the flat space limit any field theory should possess one more
crucial property connecting with causal propagation [1618]. In general case interaction
with external fields changes light cones describing propagation of spin 2 massive field [1]
so causality may give another restrictions on the theory.
Generalizing (2) to curved spacetime we should substitute all derivatives for the
covariant ones and also we can add non-minimal terms containing curvature tensor with
some dimensionless coefficients in front of them. As a result, the most general action for
massive spin 2 field in curved spacetime quadratic in derivatives and consistent with the
flat limit should have the form [1]:

Z

1
1
1
S = d D x G H H H H H H
4
4
2
a
a2
1
1
+ H H + RH H + RH 2
2
2
2
a3
a4
a5
+ R
H H + R H H + R H H
2
2
2

2
2
m 2
m
H H +
H ,
(9)

4
4

where a1 , . . . , a5 are so far arbitrary dimensionless coefficients, R = ,


R = R .
Equations of motion
E = 2 H G 2 H + H + G H H
H + 2a1 RH + 2a2 G RH + 2a3 R H
+ a4R H + a4 R H + a5 R H + a5 G R H
m2 H + m2 H G 0
contain second time derivatives of H in the following way:

E00 = Gmn G00 G00 Gmn + G00 G0m G0n 0 0 Hmn + O(0 ),

E0i = G0i G00 Gmn + G0i G0m G0n G0m in 0 0 Hmn + O(0 ),

Eij = G00 im jn Gij G00 Gmn + Gij G0m G0n 0 0 Hmn + O(0 ).

(10)

(11)

So we see that accelerations H 00 and H 0i again (as in the flat case) do not enter the
equations of motion while accelerations H ij can be expressed through H , H and their
spatial derivatives.

622

I.L. Buchbinder et al. / Nuclear Physics B 584 (2000) 615640

There are D linear combinations of the equations of motion which do not contain second
time derivatives and so represent primary constraints of the theory:
(1) = E 0 = G00 E0 + G0j Ej .

(12)

Now one should calculate time derivatives of these constraints and define secondary ones.
(1)
In order to do this in a covariant form we can add to the time derivative of any linear
combination of equations of motion and primary constraints. So we choose the secondary
constraints in the following way:
(1)
i
(1)
i
+ i
E 0
i
E
(2) = E = (1) + i E i + 0


2
2

= 2a1 R m H + 2a2 R + m H + 2a3 R H

+ a4 R H + (a4 2)R H + a5 R H




a4
a5

H R + 2a2 +
H R
+ (a5 + 1)R H + 2a1 +
2
2


+ H (2a3 + a5 + 1) R + (a4 2a3 2) R .

(13)

At the next step conservation of these D secondary constraints should lead to one new
constraint and to expressions for D 1 accelerations H 0i . This means that the constraints
(13) should contain the first time derivatives H 0 through the matrix with the rank D 1:
0 = A H 00 + B j H 0j + ,
(2)

= Ci H 00 + Di j H 0j + ,


A B j



= D 1.

rank rank
Ci Di j
(2)

(14)
(15)

In the flat spacetime we had the matrix




0 0


b =

.
0 m2 ij

(16)

In the curved case the explicit form of this matrix elements in the constraints (13) is:
A = RG00 (2a1 + 2a2) + R 00 (a4 + a5 ) + R 0 0 G00 (a4 + a5 1),
B j = m2 G0j + RG0j (2a1 + 4a2 ) + 2a3 R 0j 0 0 + R j 0 G00 (a4 2)
+ R 0j (a4 + 2a5 ) + R 0 0 G0j (a4 + 2a5 ),
Ci = R 0 i G00 (a4 + a5 1),
j

Di j = m2 G00 i + 2a1 RG00 i + 2a3 R 0j i 0 + a4 R 00 i


+ (a4 2)R j i G00 + (a4 + 2a5)Ri0 G0j .

(17)

At this stage the restrictions that consistency imposes on the non-minimal couplings and
on the external gravitational field reduce to the requirements that the above matrix elements
b = 0 while det Di j 6= 0.
give det

I.L. Buchbinder et al. / Nuclear Physics B 584 (2000) 615640

623

One way to fulfill these requirements is to impose the following restriction on the
external gravitational fields:
1
(18)
R = G R.
D
It means that one considers only Einstein spacetimes [67] representing solutions of vacuum
Einstein equations with cosmological constant. In these spacetimes the scalar curvature R
is constant as follows from the Bianchi identity R = 12 R but the Weyl tensor part
of the curvature tensor can be arbitrary.
If the Einstein equation (18) for external gravity is fulfilled the coefficients a4 , a5 in the
b takes the form:
lagrangian (9) are absent and the matrix


RG00 (2a1 + 2a2 1/D) RG0j (2a1 + 4a2) + 2a3 R 0j 0 0 + m2 G0j



.
b =



0j 0
00 j
2 00 j

0
2a3 R i + RG i (2a1 2/D) m G i
(19)
The simplest way to make the rank of this matrix to be equal to D 1 is provided by the
following choice of the coefficients:


1
1
a3 = 0,
2R a1
(20)
2a1 + 2a2 = 0,
m2 6= 0.
D
D
As a result, we have one-parameter family of theories:

1 2
,
a2 =
,
a3 = 0,
a4 = 0,
a5 = 0,
D
2D
1
2(1 )
R + m2 6= 0
R = G R,
D
D
with an arbitrary real number.
The action in this case takes the form

Z

1
1
1
S = d D x G H H H H H H
4
4
2

1 2
1
RH H +
RH 2
+ H H +
2
2D
4D

m2
m2 2
H H +
H

4
4
and the corresponding equations of motion are
a1 =

(21)

(22)

E = 2 H G 2 H + H + G H H
2
1 2
RH G m2 H + m2 H G = 0.
H + RH +
D
D
(23)
The secondary constraints built out of them are


 2 2(1 )
(2)

R
= E = H H m +
D

(24)

624

I.L. Buchbinder et al. / Nuclear Physics B 584 (2000) 615640

b looks like
and the matrix


0j
2(1 )
0 G

2
b = m +
R

0 G00ij
D

(25)

Just like in the flat case, in this theory the conditions i(2) 0 define the accelerations H 0i
and the condition 0(2) 0 after excluding H 0i gives a new constraint, i.e., the acceleration
H 00 is not defined at this stage.
To define the new constraint in a covariant form we use the following linear combination
of (2) , equations of motion, primary and secondary constraints:
m2
2(1 )
G E + E +
RG E
D2
D(D 2)



1
2(1 )
D + 2(1 D)
R + m2
R + m2 (D 1) 0.
=H
D2
D
D

(3) =

(26)

Requirement of its conservation leads to one more constraint


(3) H

(4) = H 0.

(27)

The last acceleration H 00 is expressed from the condition (4) 0.


Using the constraints for simplifying the equations of motion we see that the original
equations are equivalent to the following system:
2( 1)
RH m2 H = 0,
D
H = 0,
H = 0,

2 H + 2R H +
H = 0,
i

0i

MAB B + = 0,

0i

, = 0, . . . , D 1

ij

i j H0

(28)

2R H
G 0 i H G 0 i H G i 0 H G
2( 1)
RH 0 + m2 H 0 = 0.

D
The last expression represents D primary constraints.
For any values of the theory describes the same number of degrees of freedom as
in the flat case the symmetric, covariantly transverse and traceless tensor. D primary
constraints guarantees conservation of the transversality conditions in time.
Now we turn to the discussion of causality for the spin 2 field equations. Analogous
problem with dynamical gravity was investigated in [1] (see also [1618] for causality
problem in electromagnetic background). In general, when one has a system of differential
equations for a set of fields B (to be specific, let us say about second order equations)
00

(29)

the following definitions are used. A characteristic matrix is the matrix function of D
arguments n built out of the coefficients at the second derivatives in the equations:
MAB (n) = MAB n n . A characteristic equation is det MAB (n) = 0. A characteristic
surface is the surface S(x) = const. where S(x) = n .
If for any ni (i = 1, . . . , D 1) all solutions of the characteristic equation n0 (ni ) are real
then the system of differential equations is called hyperbolic and describes propagation of

I.L. Buchbinder et al. / Nuclear Physics B 584 (2000) 615640

625

some wave processes. The hyperbolic system is called causal if there is no timelike vectors
among solutions n of the characteristic equations. Such a system describes propagation
with a velocity not exceeding the speed of light. If there exist timelike solutions for n
then the corresponding characteristic surfaces are spacelike which violates causality.
In the flat spacetime the equations for the spin 2 field (1) lead to the characteristic
equation
det M(n) = (n2 )D(D+1)/2 ,

(30)

which has 2 multiply degenerate roots:


q
n0 = n2i .
n20 + n2i = 0,

(31)

The solutions for n are real and null hence the equations are hyperbolic and causal.
Now consider the curved spacetime generalization. If we tried to use the equations of
motion in the original lagrangian form (23) then the characteristic matrix
M (n) = () () n2 G G n2 + G n n + G n n
( n) n ( n) n

(32)

would be degenerate. This fact can be seen from the relation


n M (n) 0,

(33)

which means that any symmetric tensor of the form n( t) (with t an arbitrary vector)
represents a null vector for the matrix M(n) and therefore det M = 0.
After having used the constraints we obtain the equations of motion written in the
form (28) and the characteristic matrix becomes non-degenerate:
M (n) = n2 ,

n2 = G n n .

(34)

The characteristic cones remains the same as in the flat case. At any point x0 we can choose
locally G (x0 ) = and then

(35)
n2 x = n20 + n2i .
0

Just like in the flat case the equations are hyperbolic and causal.
So we demonstrated that in Einstein spacetimes spin 2 massive field can be consistently
described by a one-parameter family of theories (22). For any value of the parameter
the corresponding equations describe the correct number of degrees of freedom which
propagate causally. Our lagrangian for the spin 2 field in curved spacetime is the most
general known so far, in all previous works only the theories with specific values of the
parameter were considered [3,5].
The next natural step would consist in building a theory describing dynamics of both
gravity and massive spin 2 field. In such a theory in addition to dynamical equations for
the massive spin 2 field one would have dynamical equations for gravity with the energy
momentum tensor constructed out of spin 2 field components. The analysis of consistency
then changes and one needs to have correct number of constraints and causality for both
fields interacting with each other [1].

626

I.L. Buchbinder et al. / Nuclear Physics B 584 (2000) 615640

The only known consistent system of a higher spin field interacting with dynamical
gravity is the theory of massless helicity 3/2 field, i.e., supergravity [64,65] (see also
the book [66]). In that case consistency with dynamical gravity requires four-fermion
interaction. If a consistent description of spin 2 field interacting with dynamical gravity
exists it may also require some non-trivial modification of the lagrangian. At least, it is
known that lagrangians quadratic in spin 2 field do not provide such a consistency [1]. In
the Section 4 we will describe a possible way of consistent description of the spin 2 field
on arbitrary gravitational background which is given by representation of the lagrangian in
the form of infinite series in inverse mass.

3. Coupling to background scalar field


Now we investigate a possibility to generalize the above analysis for the case of spin 2
massive field interacting not only with background gravity but also with a scalar dilaton
field. This set of fields arises naturally in string theory which contains dilaton field (x) as
one of its massless excitations.
Writing a general action similar to (22) for this system one should take into account
all possible new terms with derivatives of (x) and also containing arbitrary factors f ()
without derivatives of the dilaton fields. For example, string effective action can contain
in various terms the factors ek , k = const. We will consider here the class of actions for
the field H where all these factors can be absorbed to the metric G by a conformal
rescaling.
The most general action of this type is
S = SG + S ,

(36)

where SG is the general action without dependence on scalar field (9) and S can contain
(up to total derivatives) ten new terms:

Z

c1
c2
c3
H H + H H + H H
S = d D x G
2
2
2
c5
c6
c4

+ H H + H H + H H
2
2
2
c8
c9
c7


+ H H + H H + H H ()2
2
2
2

c10 2
2
H () .
+
(37)
2
We consider as a background field on the same footing with the metric G .
What values of coupling parameters c1 , . . . , c10 are permissible and how does the
condition on the background (18) change in the presence of ? To answer these questions
one should repeat the analysis of the previous section for the action (36) calculating all
constraints of the theory.
The equations of motion are:

I.L. Buchbinder et al. / Nuclear Physics B 584 (2000) 615640

627

E = 2 H G 2 H + H + G H H H
+ 2a1RH + 2a2 G RH + 2a3R H + a4 R H + a4 R H

c3 c6
H + H
+ a5R H + a5 G R H +
2
 c5 c4

c 6 c3

H + H +
H + H
+
2
2

c3 + c6

H + H
+ (c4 c5 )G H
2
c1 H 2 c2 G H 2 c4 H c5 G H

+ c7 H + H + c8 H + c8 G H
+ 2c9H ()2 + 2c10G H ()2 m2 H + m2 H G = 0.

(38)

Introducing of background dilaton does not change the second derivatives terms in
the equations and so (1) = E 0 0 are again primary constraints. Conditions of their
conservation should give D secondary constraints:

1
(2)
0 = E0 = (c3 c6 + c4 c5 )G00 k G0k H 00 + 2Gkj H 0j + ,
2

1
i(2) = Ei = (c3 c6 + c4 c5 )G00 i G00 H 00 + 2G0j H 0j + . (39)
2
Hence we should impose the restriction
c 3 c6 + c4 c5 = 0

(40)

in order to cancel second time derivatives in (2) . Note that without dilaton couplings these
expressions do not contain second derivatives at all and represent constraints for any values
of non-minimal couplings with gravity (12).
We will restrict ourselves to even simpler particular class of the theories with
c 3 = c6 ,

c4 = c5 .

(41)

Then the secondary constraints contain first time derivatives of H0 with the following
coefficients (14):


A = G00 2(a1 + a2 )R (c1 + c2 ) 2 + 2(c9 + c10 )()2
+ (a4 + a5 )R 00 (c3 + c4 ) 0 0 + (c7 + c8 ) 0 0


+ G00 (a4 + a5 1)R 0 0 (c3 + c4 )0 0 + (c7 + c8 )0 0 ,

B j = 2a3 R 0j 0 0 + G0j m2 + (2a1 + 4a2 )R (c1 + 2c2 ) 2

+ (2c9 + 4c10 )()2
+ (a4 + 2a5)R 0j (c3 + 2c4 ) 0 j + (c7 + 2c8 ) 0 j


+ G0j (a4 + 2a5 )R 0 0 (c3 + 2c4 )0 0 + (c7 + 2c8 )0 0


+ G00 (a4 2)R j 0 c3 0 i + c7 0 i ,


Ci = G00 (a4 + a5 1)R 0 i (c3 + c4 )i 0 + (c7 + c8 )i 0 ,

628

I.L. Buchbinder et al. / Nuclear Physics B 584 (2000) 615640


j
Di j = 2a3 R 0j i 0 + G00 i m2 + 2a1R c1 2 + 2c9 ()2

j
+ i a4 R 00 c3 0 0 + 0 0


+ G00 (a4 2)R j i c3 i j + c7 i j


+ G0j (a4 + 2a5 )Ri0 (c3 + 2c4 )i 0 + (c7 + 2c8 )i 0 .

(42)

b (15) with the elements (42) to be


The simplest way to make the rank of the matrix
equal to D 1 consists in the choice
R =

1
RG ,
D
c1 = c2 ,

2a1 + 2a2
c3 = c4 ,

1
= 0,
a3 = a4 = a5 = 0,
D
c7 = c8 ,
c9 = c10

(43)

supplemented by (41).
In this case consistent action for the spin 2 field interacting with background
gravitational and scalar fields contains five arbitrary constant parameters:

Z

1
1
1
S = d D x G H H H H H H
4
4
2

1 2
1
1
RH H +
RH 2 + H H
+ H H +
2
2D
4D
2
1
2
2

H H + H H H H
2
2
2
2
3
2
H H + H H + H H
2
2
2
3
4

4
H H + ()2 H H ()2 H 2
2
2
2

m2
m2 2
H
H H +
(44)
4
4
and conservation conditions for the secondary constraints (2) = E can be used for
building one new constraint:

(3) = E + 3 2 E


2(1 )
1
R (3 + 24 )()2 + (1 + 2 ) 2
G E
+ m2 +
D
D2


= H 22R + 23 23


22
R 23 2
+ H H 2(1 + 2 ) 2 +
D

(45)
2(3 + 44 ) + H f () + Hf ().
Here f and f are functions of the scalar field and its derivatives, we will not write
the explicit form of them here. The important fact about this constraint is that it does not
contain the time derivative of the field component H00 . Hence, the conservation condition
(3) 0 does not contain the acceleration H 00 and leads to another one constraint (4)

I.L. Buchbinder et al. / Nuclear Physics B 584 (2000) 615640

629

providing thus the correct total number of constraints in the theory. The acceleration H 00
is defined only from the condition (4) 0. For any values of the coupling parameters the
theory (44) describes correct number of degrees of freedom.
Causality in presence of dilaton may be violated even in the flat spacetime. It follows
(2)
from the fact that in general the constraints and (3) cannot be solved algebraically
with respect to the trace H and the longitudinal part H and used for cancelling the
corresponding terms in the equations of motion. As a result, the characteristic matrix differs
significantly from its flat counterpart and the characteristic equation may possess new nontrivial solutions spoiling causality for some values of the parameters , . We postpone a
detailed study of the causality in the presence of dilaton for a separate publication.
In the rest of the paper we will be considering only pure gravitational background when
the scalar field is absent.

4. Consistent equations in arbitrary background


In previous sections we analyzed a possibility of consistent description of the spin 2 field
on arbitrary Einstein manifold. Now we will describe another possibility which allows to
remove any restrictions on the external gravitational background by means of considering
a lagrangian in the form of infinite series in inverse mass m. Existence of dimensionful
mass parameter m in the theory let us construct a lagrangian with terms of arbitrary orders
in curvature multiplied by the corresponding powers of 1/m2 :

Z

1
1
1
SH = d D x G H H H H H H
4
4
2
a
a
a
1
1
2
3
+ H H + RH H + RH 2 + R H H
2
2
2
2
a4
a5
+ R H H + R H H
2
2
1
+ 2 (RH H + RH H + RRH H )
m

1
+ 4 RRH H + RRH H
m
 
1
+ RRH H + RRH H + RRRH H ) + O
m6

m2 2
m2

H H +
H .
(46)

4
4
Actions of this kind are expected to arise naturally in string theory where the role of mass
parameter is played by string tension m2 = 1/ 0 and perturbation theory in 0 will give for
background fields effective actions of the form (46). 2
2 In fact, string theory should lead to even more general effective actions than (46) since in the higher 0
corrections higher derivatives of all background fields should appear.

630

I.L. Buchbinder et al. / Nuclear Physics B 584 (2000) 615640

Possibility of constructing consistent equations for massive higher spin fields as series
in curvature was recently studied in [6] where such equations were derived in particular
case of symmetrical Einstein spaces in linear in curvature order.
Here we will demonstrate that requirement of consistency with the flat spacetime limit
can be fulfilled perturbatively in 1/m2 for arbitrary gravitational background at least in the
lowest order. We will use the same general scheme of calculating lagrangian constraints
as in the previous sections. The only difference is that each condition will be considered
perturbatively and can be solved separately in each order in 1/m2 .
Primary constraints in the theory described by the action (46) should be given by the
equations E 0 0. Requirement of absence of second time derivatives in these equations
will give some restrictions on coefficients in higher orders in 1/m2 , for example, in terms
like RH H .
Secondary constraints in the lowest order in 1/m2 were already calculated in the
Section 2 (13). Consistency with the flat spacetime limit requires existence of one
additional constraint among conservation conditions of these secondary constraints. So
we should calculate the rank of the matrix


A B j




b
(47)
=
Ci Di j
with the elements in the lowest order given by (17). The advantage of having a theory in the
form of infinite series consists in the possibility to calculate the determinant of the above
matrix perturbatively in 1/m2 . Assuming that the lower right subdeterminant of the matrix
is not zero (it is not zero in the flat case) one has
b = (A BD 1 C) det D,
det

det D 6= 0.

(48)

Converting the matrix D perturbatively


 
1
1
j
1
D = 2 00 i + O
m G
m4
we get
A BD 1 C = RG00 2(a1 + a2 ) + R 00 (2a4 + 2a5 1) + O

(49)



1
.
m2

(50)

So consistency with the flat limit imposes at this order in m2 two conditions on the five nonminimal couplings in the lagrangian (46) and we are left with a three parameters family of
theories:
1
1
3
1
(51)
a1 = , a2 = , a3 = , a4 = 2 , a5 = 2 .
2
2
2
2
The action (46) then takes the form:

Z

1
1
1
D
SH = d x G H H H H H H
4
4
2
1
1
1 22
1
R H H
+ H H + RH H RH 2 +
2
4
4
4

I.L. Buchbinder et al. / Nuclear Physics B 584 (2000) 615640

631

2
3
m2
R H H + R H H
H H
2
2
4
 
m2 2
1
H +O
+
.
(52)
4
m2
b is equal to D 1 and one can construct from the
In this case the rank of the matrix
conservation conditions for the secondary constraints
+

b H 0 +
0 (2) = 0 E =

(53)

one linear combination which does not contain acceleration H 0 :


 

 

1 00
1
1 0i
1
(2)
(2)
00
0i
0 0 + G 2 R + O
0 i
G 2R +O
4
m
m
m
m4
 
1 0
1
0

= G 0 E 2 R 0 E + O
.
(54)
m
m4
After excluding the accelerations H ij by means of equations of motion this expression will
represent a new constraint restricting the trace of the field H . To make it covariant we can
add spatial derivatives of the secondary constraints i E :
 
1
1
(3) E 2 R E + O
m
m4

= (m2 1 R) 2 H H

+ 2 R 2 H + H H H
+ 3 R H + (1 + 22 + 23 ) R ( H H )



2
1
21 + 2 R H H + (1 + )H 2 R
+
2
2



1
3
R
+ H (1 + 2 + 3 ) 2 R + + 1 2
2
2

 
1
.
(55)
3 R R + 3 R R + O
m2
Derivatives of the field H enter this expression in such a way that it does not contain
the accelerations H 00 , H 0 and the velocity H 00 . This velocity does not appear in (55)
after excluding the accelerations H ij because the equations of motion do not contain H 00
either. It means that just like in the flat case the conservation condition (3) 0 leads to
another new constraints (4) and the last acceleration H 00 is defined from (4) 0. The
total number of constraints coincides with that in the flat spacetime.
The constraints (3) and (2) can be solved perturbatively in 1/m2 with respect to the
trace and the longitudinal part of H :
 
 
1
1
(2)

H
+
O
(3) H + O
,

(56)

2
m
m2
and used for simplifying the form of the original equations of motion. It is convenient to
take these equations in the following linear combination which does not contain the term
m2 H :

632

I.L. Buchbinder et al. / Nuclear Physics B 584 (2000) 615640

1
G G E
D1
= 2 H + H H H

1
G H 2 H m2 H + 1 RH + (3 + 2)R H
+
D1



1
2
+ 2 R H + R H + 2 R H
G RH

2
D1
 
1
2 3 1
G R H + O
.
(57)
+
D1
m2
Adding to these equations a suitable combination of the constraints (56) we obtain that in
the lowest order in 1/m2 the spin 2 massive field is described by the conditions
E

2 H m2 H + 1 RH + (3 + 2)R H


 1 
 2 3 1
1
+ 2 R H + R H +
G R H + O 2 = 0,

2
D1
m
 
 
1
1
= 0,
H + O
=0
(58)
H +O
m2
m2
and also by the D primary constraints E 0 . We see that even in this lowest order in m2 not
all non-minimal terms in the equations are arbitrary. Consistency with the flat limit leaves
only three arbitrary parameters while the number of different non-minimal terms in the
equations is four.
However, if gravitational field is also subject to some dynamical equations of the form
R = O(1/m2 ) then the system (58) contains only one non-minimal coupling in the
lowest order
 
1
= 0,
2 H m2 H + (3 + 2)R H + O
m2
 
 
1
1

H
+
O
=
0,

= 0,
H +O

2
m
m2
 
1
=0
(59)
R + O
m2
and is consistent for any its value.
Requirement of causality does not impose any restrictions on the couplings in this order.
The characteristic matrix of (58) is non-degenerate, second derivatives enter in the same
way as in the flat spacetime, and hence the light cones of the field H described by (58)
are the same as in the flat case. Propagation is causal for any values of 1 , 2 , 3 . In higher
orders in 1/m2 situation becomes more complicated and we expect that requirement of
causality may give additional restrictions on the non-minimal couplings.
Concluding this section we would like to stress once more that the theory (52) admits any
gravitational background and so no inconsistencies arise if one treats gravity as dynamical
field satisfying Einstein equations with the energymomentum tensor for the field H .
The action for the system of interacting gravitational field and massive spin 2 field and the
Einstein equations for it are:

I.L. Buchbinder et al. / Nuclear Physics B 584 (2000) 615640

S = SE + SH ,

SE =

1
D2

1
H
,
R G R = D2 T
2

Z
dDx

633

G R,

1 SH
H
T
=
G G

(60)

with SH given by (52). However, making the metric dynamical we change the structure
of the second derivatives by means of nonminimal terms RH H which can spoil causal
propagation of both metric and massive spin 2 field [1]. This will impose extra restrictions
on the parameters of the theory. Also, one can consider additional requirements the theory
should fulfill, e.g., tree level unitarity of graviton massive spin 2 field interaction [4].

5. Open string theory in background of massive spin 2 field


In this section we will consider sigma-model description of an open string interacting
with two background fields massless graviton G and second rank symmetric tensor
field H from the first massive level of the open string spectrum. We will show that
effective equations of motion for these fields are of the form (59) and explicitly calculate
the coefficient 3 in these equations in the lowest order in 0 .
Classical action has the form
Z
Z
1
1
2
ab

d z g g a x b x G +
e dt H x x .
S = S0 + SI =
4 0
2 0
M

(61)
Here , = 0, . . . , D 1; a, b = 0, 1 and we introduced the notation x = dx /e dt. The
first term S0 is an integral over two-dimensional string world sheet M with metric gab
and the second SI represents a one-dimensional integral over its boundary with einbein e.
We work in euclidian signature and restrict ourselves to flat world sheets with straight
boundaries. It means that both two-dimensional scalar curvature and extrinsic curvature of
the world sheet boundary vanish and we can always choose such coordinates that gab =
ab , e = 1.
Theory has two dimensionful parameters. 0 isfundamental string length squared,
D-dimensional coordinates x have dimension 0 . Another parameter carries
dimension of inverse length in two-dimensional field theory (61) and plays the role
of renormalization scale. It is introduced in (61) to make the background field H
dimensionless. In fact, power of is responsible for the number of massive level to which
a background field belongs because one expects that open string interacts with a field from
nth massive level through the term
Z
n+1
e dt x 1 x n+1 H1 ...n+1 (x).
n ( 0 ) 2
M

The action (61) is non-renormalizable from the point of view of two-dimensional


quantum field theory. Inclusion of interaction with any massive background produces in
each loop an infinite number of divergencies and so requires an infinite number of different

634

I.L. Buchbinder et al. / Nuclear Physics B 584 (2000) 615640

massive fields in the action. But massive modes from the nth massive level give vertices
proportional to n and so they cannot contribute to renormalization of fields from lower
levels. Of course, this argument assumes that we treat the theory perturbatively defining
propagator for X only by the term with graviton in (61). Now we will use such a scheme
to carry out renormalization of (61) dropping all the terms O(2 ).
Varying (61) one gets classical equations of motion with boundary conditions:


(G)a x b x = 0,
g ab Da b x g ab a b x +


2
1
(62)
G n x M Dt2 x H + x x H H H = 0,

where n = na a , na is unit inward normal vector to the world sheet boundary and

Dt2 x = x + (G)x x is the derivative covariant with respect to D-dimensional


diffeomorphisms.
(1)
we need to expand (61) up
To calculate divergencies of one-loop effective action div
to the second order in normal coordinates
1
(63)
x = x + + O( 3 )
2
around a solution x of classical equations of motion (62):


1
1
V
(1) = Tr ln S0 +

2
2 0
1
1

TrV G0 + O(2 ),
(64)
= Tr ln S0 +
2
2
2 S0
1
2 SI
V
=
.
S0 = ,


2 0

Here Green function G0 of the operator S0 is defined as


2 0 S0 G0 = .

The terms O(2 ) in (64) contribute to renormalization of the second and higher massive
levels only and will be omitted from here on.
Explicit calculations give the following expressions for V and S0

1
G g ab Da Db + g ab a x b x R (z, z0 )
2 0
1
M (z)G na Da (z, z0 ),

2 0
2
X
(k)
V (z)(Dt )k (z, z0 ),
V (z, z0 ) = M (z)

S0 (z, z0 ) =

k=0
(0)
= 2 H Dt2 x
V


+ x x H 2 H 2HR ,

(1)
= 2x ( H H H ),
V
(2)
= 2H .
V

(65)

I.L. Buchbinder et al. / Nuclear Physics B 584 (2000) 615640

635

Here delta-function of the boundary M (z) is defined as


Z
Z
p

2
M (z)V (z) g(z) d z =
V
e(t) dt.
M

(66)

zM

Using dimensional regularization and following the procedure [61,62] one gets divergences of the Green function and its derivatives:
Z
div

d 2+ z g g ab a x b x R ,
Tr ln S0; =
2
M

div

,
G0 M =


div
Dt G0 M = 0,

div

ab
g a x b x R ,
Dt2 G0 M =
2
M

(67)
(68)
(69)

As a result, divergent part of the one loop effective action has the form
Z

(1)
=
d 2+ z g g ab a x b x R
div
4

M
1

Z
dt e(t)x x 2 H 2R H + R H

(70)

which leads to one-loop renormalization of the background fields:


0
R ,

0 2

H 2R ( H) + R H
(71)
H = H +

with circles denoting bare values of the fields. We would like to stress once more that
higher massive levels do not influence the renormalization of any given field from the lower
massive levels and so the result (71) represents the full answer for perturbative one-loop
renormalization of G and H .
Now to impose the condition of Weyl invariance of the theory at the quantum level we
should calculate the trace of energymomentum tensor in d = 2 + dimension:

G = G

T (z) = gab (z)

ab
S
1
=
g (z)a x b x G
H x x M (z)
0
gab (z) 8
4 0
(72)

performing one-loop renormalization of the composite operators g ab (z)a x b x G and


H x x .
Divergencies in vacuum expectations values of the operators
R
Dx eS[x] H x x

R
(73)
hH x x i =
Dx eS[x]

636

I.L. Buchbinder et al. / Nuclear Physics B 584 (2000) 615640

are calculated by standard background field method with the use of normal coordinates
(63) as quantum fields.
In the one loop approximation it is enough to expand the composite operator and the
action in (73) up the to second order in normal coordinates and we neglect the terms more
than linear in H as giving only contributions to renormalization of higher massive levels:



H (x)
+ h i0 x x
+ R H (x)

x x H = x x H (x)
2

+ Dt 0 2x H (x).

(74)
+ Dt Dt 0 H (x)
Here

(z) (z )


D e 2 S0; [x]
(z) (z0 )

= 2G0 (z, z0 )
R
1
2 S0; [x]

D e

(75)

is Green function for the fields .


Using (69) one gets:




2
1

x x H = x x H + R H 2 H

g ab a x b x R H + (fin).
(76)

Renormalized composite operator should have finite expectation value and so the
renormalization should have the form (we use minimal subtraction scheme):




2
1
x x H 0 = x x H + R H 2 H



x x R H + (fin),
(77)

where (. . .)0 and [. . .] stands for bare and renormalized operators, respectively. Expressing

bare values of background field H in terms of the renormalized ones (71) we finally see
that in the lowest order in 0 the operator does not receive any renormalization:




(78)
x x H 0 = x x H .
After the same but more tedious calculations we get the following renormalization of
another composite operator:





0
ab

ab

g a x b x G 0 = g a x b x G R

0
1+
 00

H M (z)
+


+ Dt2 x H 4 H M (z)
+ x x H 4 ( H) + 2 2 H


2R H M (z)
and build the renormalized operator of the energymomentum tensor trace:

(79)

I.L. Buchbinder et al. / Nuclear Physics B 584 (2000) 615640


 2


(0)
(1)
8[T ] = g ab a x b x E
(x) + M (z) x x E
(x)

 2 (2)  1 00
1
+ M (z) Dt x E (x) + M (z)[E (3)(x)],

637

(80)

where
(0)
(x) = R + O( 0 ),
E
(1)
(x) = 2 H H H
E
1
1
R H + H 0 H + O( 0 ),
2

E(2) (x) = H 4 H + O( 0 ),

E (3) (x) = H + O( 0 ).

(81)

Terms of order O( 0 ) arise from the higher loops contributions.


The requirement of quantum Weyl invariance tells that all E(x) in (81) should vanish
and so they are interpreted as effective equations of motion for background fields. They
contain vacuum Einstein equation for graviton (in the lowest order in 0 ), curved spacetime
generalization of the mass shell condition for the field H with the mass m2 = ( 0 )1 and
D + 1 additional constraints on the values of this fields and its first derivatives. Taking into
account these constraints and the Einstein equation we can write our final equations arising
from the Weyl invariance of string theory in the form:
2 H + R H
H + O( 0 ) = 0,

1
H + O( 0 ) = 0,
0
H + O( 0 ) = 0,

R + O( 0 ) = 0.

(82)

They coincide with the equations found in the previous section (59) with the value of nonminimal coupling 3 = 1.
In fact, Einstein equations should not be vacuum ones but contain dependence on the
H . Our calculations could not produce
field H through its energymomentum tensor T
this dependence because such dependence is expected to arise only if one takes into account
string world sheets with non-trivial topology and renormalizes new divergencies arising
from string loops contribution [57,59].
H

R + O( 0 ) = T

1
T H ,
D2

(83)

H can
where explicit form of the lowest contributions to the energymomentum tensor T
be determined only from sigma-model on world sheets with topology of annulus.
In order to determine whether the equations (82) can be deduced from an effective
lagrangian (and to find this lagrangian) one would need two-loop calculations in the string
sigma-model. Two-loop contributions to the Weyl anomaly coefficients E (i) are necessary
because the effective equations of motion (81), (82) are not the equations directly following
from a lagrangian but some combinations of them similar to (58). In order to reverse the

638

I.L. Buchbinder et al. / Nuclear Physics B 584 (2000) 615640

procedure of passing from the original lagrangian equations to (58) one would need the
next to leading contributions in the conditions for H and H (82).

6. Conclusion

Let us summarize the obtained results. We investigated the problem of consistency of


the equations of motion for spin 2 massive field in curved spacetime and found that two
different description of this field are possible. First, for gravitational background satisfying
vacuum Einstein equations (18) one can build an action leading to consistent equations
including the tracelessness and transversality conditions. Of course, such gravitational
backgrounds include all popular vacuum solutions of Einstein equations such as constant
curvature spacetimes, Schwarzschild solution, plane waves etc. It would be interesting to
investigate properties of the massive spin 2 field dynamics on these specific exact solutions.
Another possibility (naturally arising in string theory) consists in building the theory as
perturbation series in inverse mass. In the lowest order no problems of consistency with
the flat space limit and causality arise and equations of motion have the form (58).
Then we calculated the equations for the massive spin 2 background field arising in
sigma model approach to string theory from the condition of quantum Weyl invariance
in the lowest order in 0 . The explicit form of the derived equations (82) appears to be
a particular case of the general equations in field theory (58). We expect that in general
in each order in 0 the situation remains the same and it is possible to construct the
part of string effective action quadratic in massive background field which should lead
to generalized mass-shell, tracelessness and transversality conditions.
To determine this part of the bosonic string effective action completely one should also
consider other massless background field including the dilaton (x) and antisymmetric
tensor B (x). Inclusion of dilaton will require investigation of strings with curved world
sheets and with non-vanishing extrinsic curvature on the boundary which complicates
the sigma-model calculations. Interaction with massless antisymmetric tensor will be
especially interesting in presence of a D-brane because such a system is a source of noncommutative geometry in string theory (see [68] and references therein). In the limit when
components of B along the brane are large there should arise some non-commutative
counterpart of the spin two massive field theory and we hope to derive its explicit form in
the future.
Also it would be interesting to repeat our analysis in the case of closed string which
contains the fourth rank tensor at the lowest massive level. From the field theoretical
point of view investigation of such higher spin fields interacting with gravity will require
to generalize analysis made in the Section 2 to the case of arbitrary spin fields whose
dynamics is governed by more complex lagrangians with auxiliary fields [14]. We leave
this investigation for the future work.

I.L. Buchbinder et al. / Nuclear Physics B 584 (2000) 615640

639

Acknowledgements
I.L.B. is grateful to S. Kuzenko, H. Osborn, B. Ovrut, A. Tseytlin and G. Veneziano for
useful discussions of some aspects of this work. The work of I.L.B., V.A.K. and V.D.P. was
supported by GRACENAS grant, project 97-6.2-34 and RFBR grant, project 99-02-16617;
the work of I.L.B. and V.D.P. was supported by RFBR-DFG grant, project 99-02-04022
and INTAS grant N 991-590. I.L.B. is grateful to FAPESP and D.M.G. is grateful to CNPq
for support of the research.
References
[1]
[2]
[3]
[4]
[5]
[6]
[7]
[8]
[9]
[10]
[11]
[12]
[13]
[14]
[15]
[16]
[17]
[18]
[19]
[20]
[21]
[22]
[23]
[24]
[25]
[26]
[27]
[28]
[29]
[30]
[31]
[32]
[33]
[34]
[35]
[36]
[37]
[38]

C. Aragone, S. Deser, Nuovo Cimento A 3 (1971) 709, Nuovo Cimento B 57 (1980) 33.
A. Higuchi, Nucl. Phys. B 282 (1987) 397.
I. Bengtsson, J. Math. Phys. 36 (1995) 5805.
A. Cucchieri, M. Porrati, S. Deser, Phys. Rev. D 51 (1995) 4543.
A. Hindawi, B.A. Ovrut, D. Waldram, Phys. Rev. D 53 (1996) 5583.
S.M. Klishevich, Class. Quant. Grav. 16 (1999) 2915.
E.S. Fradkin, M.A. Vasilev, Nucl. Phys. B 291 (1987) 141.
M.A. Vasilev, Phys. Lett. B 243 (1990) 378.
M.A. Vasiliev, Int. J. Mod. Phys. D 5 (1996) 763.
C. Cutler, R.M. Wald, Class. Quant. Grav. 4 (1987) 1267.
R.M. Wald, Class. Quant. Grav. 4 (1987) 1279.
M.J. Duff, C.N. Pope, K.S. Stelle, Phys. Lett. B 223 (1989 ) 386.
M. Fierz, W. Pauli, Proc. R. Soc. A 173 (1939) 211.
S.J. Chang, Phys. Rev. 161 (1967) 1308.
L.P.S. Singh, C.R. Hagen, Phys. Rev. D 9 (1974) 898.
G. Velo, D. Zwanziger, Phys. Rev. 188 (1969) 2218.
G. Velo, Nucl. Phys. B 43 (1972) 389.
D. Zwanziger, Lecture Notes in Physics 73 (1978) 143.
S. Deser, V. Pascalutsa, A. Waldron, Massive spin 3/2 electrodynamics, hep-th/0003011.
A.A. Tseytlin, Phys. Lett. B 185 (1987) 59.
M. Maggiore, Nucl. Phys. B 525 (1998) 413.
G. Veneziano, String cosmology: the pre-big bang scenario, hep-th/0002094.
C. Lovelace, Phys. Lett. B 135 (1984) 75.
C. Callan, D. Friedan, E. Martinec, M. Perry, Nucl. Phys. B 262 (1985) 593.
A. Sen, Phys. Rev. Lett. 55 (1985) 1846.
E.S. Fradkin, A.A. Tseytlin, Phys. Lett. B 158 (1985) 316; Nucl. Phys. B 261 (1985) 1.
E.S. Fradkin, A.A. Tseytlin, Phys. Lett. B 163 (1985) 123.
A.A. Tseytlin, Nucl. Phys. B 276 (1986) 391.
A.A. Tseytlin, Phys. Lett. B 178 (1986) 34.
C.M. Hull, P.K. Townsend, Nucl. Phys. B 274 (1986) 349.
A.A. Tseytlin, Nucl. Phys. B 294 (1987) 383.
H. Osborn, Nucl. Phys. B 294 (1987) 595.
G. Curci, G. Paffuti, Nucl. Phys. B 286 (1987) 399.
J.M.F. Labastida, M.A.H. Vozmediano, Nucl. Phys. B 312 (1989) 308.
J.-C. Lee, Nucl. Phys. B 336 (1990) 222.
S. Frste, Ann. Phys. 1 (1992) 98.
I.L. Buchbinder, E.S. Fradkin, S.L. Lyakhovich, V.D. Pershin, Phys. Lett. B 304 (1993) 239.
I.L. Buchbinder, V.A. Krykhtin, V.D. Pershin, Phys. Lett. B 348 (1995) 63.

640

I.L. Buchbinder et al. / Nuclear Physics B 584 (2000) 615640

[39] I.L. Buchbinder, Nucl. Phys. (Proc. Suppl.) B 49 (1996) 133.


[40] I.L. Buchbinder, V.D. Pershin, G.V. Toder, Mod. Phys. Lett. A 11 (1996) 1589; Class. Quant.
Grav. 14 (1997) 589; Nucl. Phys. (Proc. Suppl.) B 57 (1997) 280.
[41] B. Sathiapalan, Nucl. Phys. B 326 (1989) 376; Nucl. Phys. B 415 (1994) 332; Mod. Phys. Lett.
A 9 (1994) 1681; Mod. Phys. Lett. A 10 (1995) 1565; Mod. Phys. Lett. A 11 (1996) 571.
[42] K. Bardakci, L.M. Bernardo, Nucl. Phys. B 505 (1997) 463.
[43] K. Bardakci, Nucl. Phys. B 524 (1998) 545.
[44] A. Wilkins, Mod. Phys. Lett. A 13 (1998) 1289.
[45] S.M. Klishevich, Int. J. Mod. Phys. A 15 (2000) 395.
[46] I. Giannakis, J.T. Liu, M. Porrati, Phys. Rev. D 59 (1999) 104013, hep-th/9809142.
[47] S.R. Das, B. Sathiapalan, Phys. Rev. Lett. 56 (1986) 2654.
[48] C. Itoi, Y. Watabiki, Phys. Lett. B 198 (1987) 486.
[49] A.A. Tseytlin, Phys. Lett. B 264 (1991) 311.
[50] T. Banks, E. Martinec, Nucl. Phys. B 294 (1987) 733.
[51] R. Brustein, D. Nemeschansky, S. Yankielowicz, Nucl. Phys. B 301 (1988) 224.
[52] J. Hughes, J. Liu, J. Polchinski, Nucl. Phys. B 316 (1989) 15.
[53] A.A. Tseytlin, Int. J. Mod. Phys. A 4 (1989) 4249.
[54] U. Ellwanger, J. Fuchs, Nucl. Phys. B 312 (1989) 95.
[55] U. Ellwanger, J. Schnittger, Int. J. Mod. Phys. A 7 (1992) 3389; Int. J. Mod. Phys. A 9 (1994)
1821.
[56] H. Dorn, H.-J. Otto, Z. Phys. C 32 (1986) 599.
[57] W. Fischler, L. Susskind, Phys. Lett. B 171 (1986) 383; Phys. Lett. B 173 (1986) 262.
[58] A. Abouelsaood, C.S. Callan, C.R. Nappi, S.A. Yost, Nucl. Phys. B 280 (1987) 599.
[59] C.S. Callan, C. Lovelace, C.R. Nappi, S.A. Yost, Nucl. Phys. B 288 (1987) 525.
[60] K. Behrndt, H. Dorn, Phys. Lett. B 266 (1991) 59; Int. J. Mod. Phys. A 7 (1992) 1375.
[61] D.M. McAvity, H. Osborn, Class. Quant. Grav. 8 (1991) 603; Class. Quant. Grav. 8 (1991) 1445;
Nucl. Phys. B 394 (1993) 728.
[62] D.M. McAvity, Class. Quant. Grav. 9 (1992) 1983.
[63] D.M. Gitman, I.V. Tyutin, Quantization of Fields with Constraints, Springer-Verlag, 1990.
[64] D.Z. Freedman, P. van Nieuwenhuizen, S. Ferrara, Phys. Rev. D 13 (1976) 3214.
[65] S. Deser, B. Zumino, Phys. Lett. B 62 (1976) 335.
[66] I.L. Buchbinder, S.M. Kuzenko, Ideas and Methods of Supersymmetry and Supergravity, I,
OP Pub., Bristol, Philadelphia, 1995, 1998.
[67] A.Z. Petrov, Einstein Spaces, Pergamon, 1969.
[68] N. Seiberg, E. Witten, JHEP 9909 (1999) 032.

Nuclear Physics B 584 (2000) 641655


www.elsevier.nl/locate/npe

YangMills integrals for orthogonal,


symplectic and exceptional groups
Werner Krauth a, , Matthias Staudacher b,1,2
a CNRS-Laboratoire de Physique Statistique, Ecole Normale Suprieure 24, rue Lhomond,

F-75231 Paris Cedex 05, France


b Albert-Einstein-Institut, Max-Planck-Institut fr Gravitationsphysik, Am Mhlenberg 1,

D-14476 Golm, Germany


Received 9 May 2000; accepted 7 June 2000

Abstract
We apply numerical and analytic techniques to the study of YangMills integrals with orthogonal,
symplectic and exceptional gauge symmetries. The main focus is on the supersymmetric integrals,
which correspond essentially to the bulk part of the Witten index for susy quantum mechanical
gauge theory. We evaluate these integrals for D = 4 and group rank up to three, using Monte
Carlo methods. Our results are at variance with previous findings. We further compute the integrals
with the deformation technique of Moore, Nekrasov and Shatashvili, which we adapt to the groups
under study. Excellent agreement with all our numerical calculations is obtained. We also discuss
the convergence properties of the purely bosonic integrals. 2000 Elsevier Science B.V. All rights
reserved.
PACS: 11.10.Kk; 11.15.Pg; 11.15.-q; 02.70.-c

1. Introduction
Recent attempts to describe D-branes through effective actions have revealed the
existence of a new class of gauge-invariant matrix models. These models are related to
(1)-branes. They differ from the classic systems of random matrices, which have been
extensively studied ever since Wigners and Dysons work in the 1950s. The models
consist in matrix integrals of D non-linearly coupled matrices. They are obtained by
complete dimensional reduction of D-dimensional euclidean continuum (susy) YangMills
theory to zero dimensions, and we term them quite generally YangMills integrals.
Corresponding author. E-mail: krauth@physique.ens.fr; tel.: (+33) 1 44 32 35 00; fax: (+33) 1 44 32 34 33
1 matthias@aei-potsdam.mpg.de
2 Supported in part by EU Contract FMRX-CT96-0012.

0550-3213/00/$ see front matter 2000 Elsevier Science B.V. All rights reserved.
PII: S 0 5 5 0 - 3 2 1 3 ( 0 0 ) 0 0 3 8 2 - 5

642

W. Krauth, M. Staudacher / Nuclear Physics B 584 (2000) 641655

The supersymmetric D = 4, 6, 10 integrals with SU(N) symmetry have already found


several important applications. They are relevant to the calculation of the Witten index of
quantum mechanical gauge theory [13], and to multi-instanton calculations [4,5] of fourdimensional SU() susy conformal gauge theory. The D = 10 integrals are furthermore
the crucial ingredient in the so-called IKKT model [6], which possibly provides a nonperturbative definition of IIB superstring theory. Finally, it remains to be seen whether
YangMills integrals contain information on the full, unreduced field theory through the
EguchiKawai mechanism [7] as the size of the matrices gets large. Some very interesting
recent considerations along these lines can be found in [8].
YangMills integrals are ordinary, not functional integrals. Despite this tremendous
simplification, no systematic analytic tools for their investigation are known to date. We
have developed [912] accurate and reliable Monte Carlo methods which allow to study
the new matrix models as long as the dimension of the gauge group is not too large. We
have found, e.g., that supersymmetry is generically not necessary for the existence of the
integrals [10]. We also computed their asymptotic eigenvalue distributions, which we found
to qualitatively differ between the susy and bosonic case as the size of the gauge group gets
large [11]. For related, complementary studies see [13].
To date all existing studies have focused on the case where the gauge group is SU(N). In
the present paper we generalize to the cases of all other semi-simple compact Lie groups
of rank r 6 3. These are, (besides the already known cases SU(2), SU(3), SU(4)) the
groups SO(3), SO(4), SO(5), SO(6), SO(7), Sp(2), Sp(4), Sp(6), and G2 , for which we
compute the susy D = 4 partition functions. We were motivated in part by a recent paper
of Kac and Smilga [14] which presented conjectures about the values of the bulk part of
the Witten index (and therefore for the corresponding integrals). Intriguingly, our results
are at variance with their predictions in most cases, indicating that the index calculations
for these groups are even more subtle than the corresponding considerations for SU(N),
where the approach of [14] agrees with the known values.
Moore et al. [15] recently employed sophisticated deformation techniques to evaluate
the SU(N) susy bulk index for all N and D = 4, 6, 10. The method apparently leads to
the correct result for all SU(N) [9,10]. Below, we adapt the technique to the more general
groups, and we find again excellent agreement with the Monte Carlo calculation. This
further indicates that the deformation method is indeed reliable.

2. YangMills integrals for semi-simple compact gauge groups


For a general semi-simple compact Lie group G we define supersymmetric or bosonic
YangMills integrals as
! N
!
Z dim(G)
D
Y Y
Y
dXA
N
A

d
ZD,G :=
2
A=1
=1
=1




1
1
Tr[X , X ][X , X ] + 2 Tr X , ,
(1)
exp
4g 2
2g

W. Krauth, M. Staudacher / Nuclear Physics B 584 (2000) 641655

643

where dim(G) is the dimension of the Lie group and the D bosonic matrices X = XA T A
and the N fermionic matrices = A TA are anti-hermitean and take values in the
fundamental representation of the Lie algebra Lie(G), whose generators we denote by
T A . The integral Eq. (1) depends on the gauge coupling constant in a trivial fashion, as
we can immediately scale out g. Nevertheless, there is a natural convention for fixing g:
for an orthogonal set of generators we should pick g according to their normalization:
Tr T A T B = g 2 AB . This convention is imposed by the index calculations of the next
section. In the bosonic case we simply drop the fermionic variables.
The supersymmetric integrals can formally be defined in D = 3, 4, 6, 10, which
corresponds to N = 2, 4, 8, 16 real supersymmetries. We are not aware of a mathematically
rigorous investigation of their convergence properties. However, our numerical studies
indicate that they are absolutely convergent in D = 4, 6, 10 (but not in D = 3) for all
semi-simple compact gauge groups. The convergence properties of the bosonic (N = 0)
integrals are discussed in Section 6. The variables A are real Grassmann-valued and can
be integrated out, leading to a bosonic integral with very special measure:


Z dim(G)
D
Y Y
dXA
1
N
exp
Tr[X , X ][X , X ] PD,G (X).
(2)
ZD,G =
4g 2
2
A=1 =1
PD,G (X) is a homogeneous Pfaffian polynomial of degree 12 N dim(G) given by
AB

PD,G = Pf MD,G , with MD,G = if ABC XC ,

(3)

are defined through the real Lie algebra


=
where the structure

ABC
C
T . Explicit expressions for the Gamma matrices and further details on
f
PD,G (X) may be found in [9].
constants 3

[T A , T B ]

3. Group volumes and bulk indices


It is well known that the susy YangMills integrals Eq. (2) naturally appear when one
computes the Witten index of quantum mechanical gauge theory (i.e., the reduction of the
field theory to one, as opposed to zero, dimension) by the heat kernel method. For the
details of the method we refer to [13]. Specifically, the integrals are related to the bulk
part indD
0 (G) of the index as
1 N
F H
=
Z
.
(4)
indD
0 (G) = lim Tr(1) e
0
FG D,G
The total Witten index indD (G) is then the sum of this bulk part and a boundary
D
D
D
contribution indD
1 (G): ind (G) = ind0 (G) + ind1 (G). The constant FG relating the
D
bulk index ind0 and the YangMills integral is independent of D and can be interpreted as


1
G
Volume
FG =
,
(5)
1
ZG
(2) 2 dim(G)
3 Note that in [9] we used hermitean generators but defined the structure constants through [T A , T B ] =
if ABC T C , so Eq. (3) remains valid.

644

W. Krauth, M. Staudacher / Nuclear Physics B 584 (2000) 641655

i.e., essentially the volume of the true gauge group, which turns out to be the quotient
group G/ZG , with ZG the center group of G. In practice, great care has to be taken in
using the relation Eq. (5), as the volume depends on the choice of the local metric on the
group manifold. For the present purposes we simply adapted our method for computing
FSU(N) (see [10]) to the relevant gauge groups. An invariant average over the group allows
to project onto gauge invariant states and to derive Eq. (4) from the quantum mechanical
path integral. In the ultralocal limit, the quantum mechanics of D 1 matrices turns into
an integral over D matrices. Then, this integration is over the anti-hermitean generators of
the group
dim(G)
A
1 Y dXD
.
DU
FG
2
A=1

(6)

The normalized Haar measure DU on the group elements U G simplifies significantly


if we restrict attention to the Cartan subgroup of G. A beautiful result of Weyl allows
to explicitly write down the restricted measure. If we parametrize the Cartan torus T by
angles 6 1 6 , . . . , 6 r 6 , where r = rank(G), the measure reads
!
r
Y
di
1
|G |2 ,
(7)
DU DT =
|WG |
2
i=1

where |WG | is the order of the Weyl group WG of G, and |G |2 the squared modulus of
the Weyl denominator:
Y i

i
e 2 (,) e 2 (,) .
(8)
G =
>0

Here the product is over the set of positive roots of the Lie algebra Lie(G). In the vicinity
of the identity in G the angles i are small and we can approximate the measure Eq. (7) by
!
r
Y
di Y
1
[(, )]2 .
(9)
|WG |
2
i=1

>0

Now restricting the flat measure on Lie(G) on the right-hand side of Eq. (6) to the Cartan
modes i we get
!
dim(G)
r
Y
Y dXA
FG |ZG | Y
D

di
[(, )]2 .
(10)
r
(2) |WG |
2
A=1

i=1

>0

An important subtlety is that we needed to multiply the measure Eq. (10) by an additional
factor |ZG | of the order of the center group ZG of G, as the averaging over the group
manifold localizes on |ZG | points. Finally, the constant FG in Eq. (10) is fixed by noting
that the flat measure on Lie(G) is normalized with respect to Gaussian integration
!


Z dim(G)
Y dXA
1 X A 2
D

XD
exp
= 1.
(11)
2
2
A=1

The Gaussian integration of the right-hand side of Eq. (10) leads to Selberg-type integrals,
see, e.g., [16]. Explicit details on how to implement the above procedure for the groups

W. Krauth, M. Staudacher / Nuclear Physics B 584 (2000) 641655

645

under study can be found in the appendices. With our conventions for the normalization
of the generators (SO(N): Tr T A T B = 2 AB and Sp(2N), G2 : Tr T A T B = 12 AB ) one
finds
N

1
2
,
FSO(N) =
N(N5) QN
2CN 2 4
j =1 (j/2)

(12)

where C2N = 2 and C2N+1 = 1, as well as


1 22N + 2 2
FSp(2N) = QN
2 j =1 (2j )
2

and finally
FG 2

(13)

36864 3
.
=
5

(14)

4. Deformation method
In [15] it was suggested that the original susy YangMills integrals Eq. (1) may be
vastly simplified by a deformation technique. It consists in adding a number of terms to
the action which break the number of supersymmetries from N = 2, 4, 8, 16 to N = 1.
Keeping one of the supersymmetries means that the partition function is protected and
should not change under the deformation. This gives the correct result 4 for SU(N) and
D = 4, 6, 10. The final outcome is a much simpler integral involving only a single Liealgebra valued matrix. The remaining integral is still invariant under the gauge group, and
one can therefore pass from the full algebra to the Cartan subalgebra degrees of freedom.
This was derived in [15] in detail for SU(N) but should carry over immediately to other
gauge groups. For D = 4 (N = 4) one finds, in the notation of the previous section (here
the product is over all roots of Lie(G))
!
Z Y
r
|ZG | 1
dxi Y [ 12 (x, ) 12 (x, )]
D=4
.
(15)
ind0 (G) =
|WG | E r
2i [ 12 (x, ) 12 (x, ) E]
i=1

This r-dimensional integral (r = rank(G)) is divergent. There are divergences due to the
poles of the denominator of Eq. (15), as well as at infinity, where the integrand tends
to one. These divergences are present since the method starts from the partition sums
Eq. (1) with Minkowski signature, i.e., with the Wick-rotated versions of our integrals.
The Minkowski integrals are divergent without a prescription. The poles are regulated by
giving an imaginary part to the parameter E. The singularity at infinity is regulated by
interpreting the integrals as contour integrals. 5 It would be interesting to complete the
arguments by demonstrating that the Wick-rotation leads to precisely these prescriptions.
4 For unclear reasons it fails for D = 3.
5 This interpretation furthermore necessitates the inclusion of the factors of i in the measure of Eq. (15) which

would not be present in an ordinary integral over a real Lie algebra.

646

W. Krauth, M. Staudacher / Nuclear Physics B 584 (2000) 641655

Very encouraging signs for the consistency of this method are that the final result neither
depends on the location of the parameter E nor on whether the contours are closed in the
upper or the lower half plane (it is important though that all r contours are closed in the
same way). For D = 6, 10 expressions very similar to Eq. (15) can be found in [15].
We now present the explicit form of the contour integrals Eq. (15) for the groups studied
in the present work (see appendices for details)
(SO(2N + 1))
indD=4
0
I Y
N
N
N
Y
(xi2 xj2 )2
xi2
1
dxi Y
1
,
= N
2 N! E N
2i
[(xi xj )2 E 2 ][(xi + xj )2 E 2 ]
xi2 E 2
i=1

i<j

i=1

(16)
(SO(2N))
indD=4
0
=

1
2N1 N! E N
2

I Y
N
N
(xi2 xj2 )2
dxi Y
,
2i
[(xi xj )2 E 2 ][(xi + xj )2 E 2 ]

(17)

i<j

i=1

(Sp(2N))
indD=4
0
I Y
N
N
N
Y
(xi2 xj2 )2
xi2
1
dxi Y
2
,
= N
N
2
2
2
2
2
2 N! E
2i
[(xi xj ) E ][(xi + xj ) E ]
xi (E/2)2
i=1

i<j

i=1

(18)
(G2 )
indD=4
0
I
(x1 x2 )2 (x1 + x2 )2 x12 x22
1 1
dx1 dx2
=
12 E 2
2i 2i [(x1 x2 )2 E 2 ][(x1 + x2 )2 E 2 ][x12 E 2 ][x22 E 2 ]
(2x1 + x2 )2 (x1 + 2x2 )2
.
(19)
[(2x1 + x2 )2 E 2 ][(x1 + 2x2)2 E 2 ]
They are easily evaluated for low rank, and we present the results in Table 1. We
highlighted the cases which were not already indirectly known due to the standard lowrank isomorphisms so(3) = sp(2) = su(2), so(4) = su(2) su(2), so(6) = su(4). Note,
however, that these identities, as well as the final semi-simple Lie algebra isomorphism
so(5) = sp(4), constitute non-trivial consistency checks on the expression Eq. (15), as the
precise form of the corresponding contour integrals is different in all these cases. It would
be interesting to compute Eqs. (16)(19) for arbitrary rank, as has been done in the case
of SU(N) in [15]. We also checked that the analogous, more complicated D = 6 contour
integrals lead to the same bulk indices, as one expects.
For the groups not related by an isomorphism to SU(N) the rational numbers in Table 1
differ from the ones proposed in [14]. I would be important to understand why. We also
do not see how the arguments of Section 8 of [15], which seemed to furnish a shortcut
explanation of the SU(N) results, could be adapted to reproduce the numbers highlighted
in Table 1.
We next turn to numerical verification of these proposed bulk indices.

W. Krauth, M. Staudacher / Nuclear Physics B 584 (2000) 641655

647

Table 1
D = 4 and D = 6 bulk indices for the orthogonal,
symplectic and exceptional groups of rank 6 3
Group

Rank

indD=4,6
0

SO(3)
SO(4)
SO(5)
SO(6)
SO(7)

1
2
2
3
3

1/4
1/16
9/64
1/16
25/256

Sp(2)
Sp(4)
Sp(6)

1
2
3

1/4
9/64
51/512

G2

151/864

5. Monte Carlo evaluation of YangMills integrals


As in previous works [9,10], we evaluate the YangMills integrals using Monte Carlo
methods. Both the Pfaffian polynomial PD,G and the action S = 4g1 2 Tr[X , X ][X , X ]
in Eq. (2) are homogeneous functions of the XA
(  
 
1
P XA 2 N dim(G) P XA ,
A
A
X X (; A)
 
 
S XA 4 S XA .

(20)

D
) = (d , R), with d = D dim(G) the total
We introduce polar coordinates (X11 , . . . , Xdim(G)
dimension of the integral. As an example, P(, 1) and S(d , 1) denote the value of the
D
) on
Pfaffian polynomial and the action, respectively, for a configuration (X11 , . . . , Xdim(G)
the surface of the d-dimensional unit hyper-sphere, with polar coordinates (d , R = 1).
Eq. (20) allows us to to perform the R-integration analytically for each value of d , and
to express the YangMills integral as an expectation value over these angular variables:
R
Dd zG (d )
N
R
,
(21)
ZD,G =
Dd

with
zG (d ) = 2

d
d/21 4

N
8


dim(G)


d

P(, 1)

.
(22)
d N
[S(d , 1)] 4 + 8 dim(G)
As discussed previously, the integrand zG (d ) is too singular to be obtained by direct
sampling of random points on the surface of the unit hyper sphere. Slightly modifying our
procedure from [10], we therefore write
 R
R

Dd z z1 1 1 Dd z1 z2 1 1
N
R
R
ZD,G =
Dd z
Dd z1
R

Dd z2 z2 1
R

.
(23)
Dd z2

648

W. Krauth, M. Staudacher / Nuclear Physics B 584 (2000) 641655

Each of the terms [ ] in Eq. (23) is computed in a separate run. For example, the second
quotient in Eq. (23):
R
Dd z1 z2 1
R
= hz2 1 i1 ,
(24)
Dd z1
is simply the average of z2 1 for points d on the unit hyper-sphere distributed according
to the probability distribution (d ) z1 (d ). (As it stands, Eq. (24) is immediately
applicable to D = 4, where the integrand z is positive semi-definite [8,9]. In the general
case, we have to sample with |z1 (d )| (cf. [9]).)
We sample angular variables d according to z1 with a Metropolis Markov-chain
method, which we now explain: at each iteration of the procedure, two distinct indices
(A1 , 1 ) and (A2 , 2 ), and an angle 0 < < 2 are chosen randomly. An unbiased trial
move d d0 is then constructed by modifying solely the coordinates XA11 and XA22 :
 A1 0 q


 A1 
X1
X1
A1 2
A2 2 sin()
= X1 + X2

,
(25)
cos()
XA22
XA22
all other elements of {XA } remaining unchanged. The trial move Eq. (25) preserves the
norm R of the vector {XA }, i.e., keeps the configuration on the surface of the unit sphere.
Furthermore, it is unbiased (the probability to propose d d0 is the same as for the
reverse move).
Finally, the move (for the example in Eq. (24)) is accepted according to the Metropolis
acceptance probability


z1 ( 0 )
0
.
(26)
P ( ) = min 1,
z 1 ()
Empirically, we found the values 1 = 0.95, 2 = 0.6 to be appropriate. Each of the
averages in Eq. (23) was computed within between a few hours and more than a thousand
hours of computer time (on a work station array), corresponding to a maximum of 5 109
samples. Results are presented in the Table 2.
N =4 by the corresponding group volume factors
Dividing the Monte Carlo results for ZD=4,G
(cf. Eqs. (12), (13), and (14)) we arrive at our numerical predictions for the bulk indices
(G), which we compare in Table 3 to the proposed analytical values.
indD=4
0
Agreement between the Monte Carlo results and theory is excellent, both in cases where
rigorous results are known (SO(3), SO(4), Sp(2)) and where the deformation technique
was applied. Among the latter cases, we again indicate in bold type new values, which had
not been obtained before.

6. Bosonic convergence for orthogonal, symplectic and exceptional integrals


Our qualitative Monte Carlo method (cf. [911]) allows us to determine the convergence
N =0 . We have found the following:
properties of the bosonic YangMills integrals ZD,G

W. Krauth, M. Staudacher / Nuclear Physics B 584 (2000) 641655

649

Table 2
Direct evaluation of YangMills integrals
Group
G

Monte Carlo result


N =4
ZD=4,G

SO(3)
SO(4)
SO(5)
SO(6)
SO(7)

1.255 0.003
0.197 0.004
0.589 0.004
0.0407 0.0007
0.0169 0.0003

Sp(2)
Sp(4)
Sp(6)

1.253 0.001
18.65 0.2
279.2 9.7

G2

6943 120

Exact

1.2533 . . .
0.1963 . . .
0.589 . . .
0.04101 . . .
0.01708 . . .
1.2533 . . .
18.849 . . .
285.59 . . .
7011.4 . . .

Table 3
Monte Carlo results for the D = 4 bulk index

N =0
ZD,SO(N=3,4)

Group
G

Monte Carlo
indD=4
(G)
0

Exact
(Table 1)

SO(3)
SO(4)
SO(5)
SO(6)
SO(7)

0.2503 0.0006
0.0627 0.0013
0.1406 0.001
0.0620 0.001
0.0966 0.0017

0.25 (1/4)
0.0625 (1/16)
0.1406 (9/64)
0.0625 (1/16)
0.0976 (25/256)

Sp(2)
Sp(4)
Sp(6)

0.2500 0.0002
0.139 0.0015
0.0973 0.003

0.25 (1/4)
0.1406 (9/64)
0.0996 (51/512)

G2

0.173 0.003

0.1747 (151/864)

< , for D > 5,


N =0
ZD,Sp(2)

N =0

ZD,SO(N>5)

N
=0
ZD,Sp(4,6,...) < , for D > 3.

Z N =0

(27)

D,G2

All other bosonic integrals diverge. We thus obtain conditions which are fully consistent
with the group isomorphisms discussed in the appendix. Let us note that we also performed
the same qualitative computations for the susy integrals, as an important check of the
convergence of the underlying Markov chains during the simulation.

650

W. Krauth, M. Staudacher / Nuclear Physics B 584 (2000) 641655

7. Conclusions and outlook


In this paper we provided further evidence that YangMills integrals encode surprisingly
rich and subtle structures, which may prove to have important bearings on gauge and string
theory.
The chief result of the present paper was to demonstrate that YangMills integrals, as
well as the methods to study them, can be naturally generalized from the previously studied
special unitary symmetries to other gauge symmetries. We numerically evaluated the
partition functions for all semi-simple gauge groups of rank r 6 3 and compared the results
to conjectured exact values, which were obtained by a generalization of certain contour
integrals derived from a supersymmetric deformation procedure. The connection between
the YangMills integrals and the bulk indices is provided by the group volumes, that we
computed explicitly. We provided details on these very subtle calculations. Agreement
between the approaches is perfect within the tight error margins left by our Monte Carlo
technique.
It would be very interesting to gain a simpler understanding of the rational numbers
collected in Table 1, although it is already evident that the bulk indices of the groups in
question are more complicated than those of the special unitary case. In particular it would
be nice to find general formulas for arbitrary rank.
In the present paper we have focused on D = 4 since this is the case where our numerical
approach is most accurate. One should clearly study the dimensions D = 6 and, especially,
D = 10 as well. It is straightforward, if more involved, to work out the predictions of the
deformation method for these cases, at least for low rank gauge groups. In [14] exact values
for the total (bulk plus boundary) Witten index indD (G) were proposed. In D = 4, 6 one
should have indD=4,6(G) = 0 while for D = 10 it is argued to be a positive integer which,
for groups other than SU(N), can be larger than one. It is important to check the arguments
by computing the index from the path integral. The results in the present paper indicate
that the bulk contributions indD
0 (G) are correctly reproduced by the deformation method;
however, we still lack a reliable method for computing the boundary terms indD
1 (G).
Since the deformation method of [15] successfully reproduces the partition functions,
it is natural to ask whether it can be extended to calculate correlation functions of the
ensembles Eq. (1), such as the quantities studied numerically (so far only for SU(N))
in [8,11,12]. This would likely lead to new insights both in string theory [6] and gauge
theory [7,8].
Numerically, it might be interesting to compare SU(N), SO(N) and Sp(2N) for large
values of N , as in the standard large N limit of t Hooft these groups are expected
to lead to identical results.

Acknowledgements
We thank H. Nicolai, H. Samtleben, G. Schrder and especially A. Smilga for useful
discussions. This work was supported in part by the EU under Contract FMRX-CT960012.

W. Krauth, M. Staudacher / Nuclear Physics B 584 (2000) 641655

651

Appendix A. Details and conventions for SO(N )


The Lie algebra so(N) has 12 N(N 1) generators which we choose to be the following
standard anti-symmetric matrices T pq (i.e., the index A becomes a double index pq)
p q

p q

(T pq )j k = j k k j ,
where p < q (p, q = 1, 2, . . . , N). The Lie algebra reads then
X
f pq,rs,t uT t u .
[T pq , T rs ] = qr T ps qs T pr pr T qs + ps T qr =
t <u

= 2 pq,rs and the structure


In this basis the generators are normalized as
1
pq,rs,t
u
pq
rs
= 2 Tr(T [T , T t u ]). The gauge potentials are
constants are given through f
P
pq pq
then X = p<q X T and the SO(N) YangMills integrals in these conventions read
Tr T pq T rs

N
ZD,SO(N)

1
2 N(N1)

p<q



D
pq
Y
1
dX
exp Tr[X , X ][X , X ] PD,N (X),

8
2
=1

(A.1)

where PD,N is the Pfaffian as defined in Eq. (3). These conventions are such that, in view
of the isomorphism so(3) = su(2) we have for all N (i.e., N = 0, 4, 8, 16)
N
N
N
= ZD,SU(2)
= ZD,Sp(2)
,
ZD,SO(3)

(A.2)

where the symplectic case is discussed in Appendix B. One checks that in these
normalizations the isomorphism so(4) = so(3) so(3) results in
2
3
1
N
N
= 2 2 (D 2 N ) ZD,SO(3)
,
(A.3)
ZD,SO(4)
while the isomorphism so(6) = su(4) leads to
N
N
= 2 4 (D 2 N ) ZD,SU(4)
,
ZD,SO(6)
15

(A.4)

where the SU(4) integral is defined as in [9]. Finally, the isomorphism so(5) = sp(4)
translates into
N
N
= 2 2 (D 2 N ) ZD,Sp(4)
,
ZD,SO(5)
5

(A.5)

where the Sp(4) integral is defined in Appendix B. One verifies that these isomorphisms
are in perfect agreement with the results of the present paper as well as with [10].
We next provide the details necessary for verifying the group volume factor FSO(N) of
Eq. (12). The natural Cartan subalgebra is spanned by the generators T 12 , T 34 , . . . . The
corresponding maximal compact tori are given for SO(2N + 1) by the (2N + 1) (2N + 1)
matrix

rot 1

rot 2

.
..
(A.6)
T =
,

rot N
1

652

W. Krauth, M. Staudacher / Nuclear Physics B 584 (2000) 641655

while for SO(2N) one has the 2N 2N matrix


rot

T =

rot 2
..

(A.7)

rot N
Here rot i are the 2 2 rotation matrices


cos i
sin i
rot i =
sin i cos i

(A.8)

and matrix elements with no entries are zero. The corresponding reduced, normalized Haar
measure on SO(2N + 1) (i.e., Eq. (7)) reads

 N


 
2 N
N
i + j Y 2 i
22N Y di Y 2 i j
sin
sin
sin2
(A.9)
DT = N
2 N!
2
2
2
2
i=1

i<j

i=1

while for SO(2N) one has






N
N
22N(N1) Y di Y 2 i j
2 i + j
sin
sin
.
DT = N1
2
N!
2
2
2
i=1

(A.10)

i<j

Eqs. (A.9), (A.10) may also be used to work out the detailed form of the contour integrals
Eqs. (16), (17) of Section 4: one simply expands the Haar measure around i 0.
Finally we recall the center groups of SO(N): one has ZSO(2N+1) = {1} and ZSO(2N) =
{1, 1} and therefore |ZSO(2N+1) | = 1 and |ZSO(2N) | = 2.

Appendix B. Details and conventions for Sp(2N )


The Lie algebra sp(2N) has 2N 2 + N generators which we choose as follows. Define
the N N matrices E pq (p, q, j, k = 1, 2, . . . , N )
p q

(E pq )j k = j k .
Then we define the N N matrix generators T a,pq for p < q by
 pq

E E qp
0
1
,
T 0,pq =
0
E pq E qp
2 2


0
i(E pq + E qp )
1
1,pq
,
=
T
0
2 2 i(E pq + E qp )


0
(E pq + E qp )
1
,
T 2,pq =
0
2 2 (E pq + E qp )


i(E pq + E qp )
0
1
T 3,pq =
pq
0
i(E + E qp )
2 2
and the remaining generators are (p = 1, . . . , N )

W. Krauth, M. Staudacher / Nuclear Physics B 584 (2000) 641655

653



1
0
iE pp
,
0
2 iE pp


1
0
E pp
,
T 2,pp =
0
2 E pp


1 iE pp
0
T 3,pp =
.
0
iE pp
2
T 1,pp =

In this basis the generators are normalized as Tr T A T B = 12 AB and the structure


constants are given through f ABC = 2 Tr T A [T B , T C ] where A is the multi-index
(a, pq). The Cartan subalgebra is spanned by the generators T 3,pp with p = 1, . . . , N .
The corresponding maximal compact torus is given by the matrix

i1
e
.

..

i
e N

(B.1)
T =
.
i

e 1

..

.
eiN
The corresponding normalized Haar measure reads

 N


2 N
N
i + j Y 2
22N Y di Y 2 i j
sin
sin i .
sin2
DT = N
2 N!
2
2
2
i=1

i<j

(B.2)

i=1

The center of Sp(2N) is the group ZSp(2N) = {1, 1} and thus |ZSp(2N) | = 2.
Appendix C. Details and conventions for G2
The Lie algebra G2 has 14 generators. For the fundamental representation we choose
them as the following explicit 7 7 matrices (see, e.g., [17]). Define
p q

p q

(Xp,q )j k = j k k j ,

p q
p q
(Y p,q )j k = i j k + k j .

Then
1
1
T 1 = (X1,2 + X3,4 ) + (X5,6 + X6,7 ) ,
2 6
2 3
1
1
2
1,2
3,4
5,6
6,7
T = (Y + Y ) + (Y + Y ) ,
2 6
2 3
1
T 3 = (X1,7 X4,5 ) ,
2 2
1
T 4 = (Y 1,7 + Y 4,5 ) ,
2 2
1
1
T 5 = (X1,6 + X4,6 ) + (X2,7 X3,5 ) ,
2 3
2 6

654

W. Krauth, M. Staudacher / Nuclear Physics B 584 (2000) 641655

1
1
T 6 = (Y 1,6 Y 4,6 ) + (Y 2,7 + Y 3,5 ) ,
2 3
2 6
1
T 7 = (X1,3 + X2,4 ) ,
2 2
1
8
1,3
2,4
T = (Y + Y ) ,
2 2
1
1
9
1,5
4,7
T = (X + X ) + (X2,6 X3,6 ) ,
2 6
2 3
1
1
T 10 = (Y 1,5 Y 4,7 ) + (Y 2,6 + Y 3,6 ) ,
2 6
2 3
1
T 11 = (X2,5 X3,7 ) ,
2 2
1
T 12 = (Y 2,5 + Y 3,7 ) .
2 2

(C.1)

Finally, the matrices T 13 and T 14 are diagonal matrices with elements:




i
i
i
i
i
i
T = diag ; ; ; ; ; 0; ,
2 6 2 6 2 6 2 6
6
6


i
i
i
i
T 14 = diag ; ; ; ; 0; 0; 0 .
2 2 2 2 2 2 2 2
13

(C.2)
(C.3)

In this basis the generators are normalized as Tr T A T B = 12 AB and the structure


constants are given through f ABC = 2 Tr T A [T B , T C ]. The Cartan subalgebra is spanned
by the generators T 13 and T 14 . The corresponding maximal compact torus is given by the
matrix

T =

ei1
ei2

ei2
ei1
ei(1 +2 )
1

(C.4)

ei(1 +2 )
The normalized Haar measure on G2 with respect to this torus reads
DT =



 
 


2
1 2
1 + 2
212 d1 d2 2 1
sin
sin2
sin2
sin2
12 2 2
2
2
2
2




21 + 2
1 + 22
sin2
.
sin2
2
2

The center of G2 is trivial: ZG2 = {1} and thus |ZG2 | = 1.

(C.5)

W. Krauth, M. Staudacher / Nuclear Physics B 584 (2000) 641655

655

References
[1] A.V. Smilga, Witten index calculation in supersymmetric gauge theory, Yad. Fiz. 42 (1985)
728; Nucl. Phys. B 266 (1986) 45; Calculation of the witten index in extended supersymmetric
YangMills theory, Yad. Fiz. 43 (1986 ) 215 (in russian).
[2] P. Yi, Witten index and threshold bound states of D-branes, Nucl. Phys. B 505 (1997) 307,
hep-th/9704098.
[3] S. Sethi, M. Stern, D-brane bound state redux, Commun. Math. Phys. 194 (1998) 675, hepth/9705046.
[4] N. Dorey, T.J. Hollowood, V.V. Khoze, M.P. Mattis, S. Vandoren, Multi-instantons and
Maldacenas conjecture, JHEP 9906 (1999) 023, hep-th/9810243; Multi-instanton calculus and
the AdS/CFT correspondence in N = 4 superconformal field theory, Nucl. Phys. B 552 (1999)
88, hep-th/9901128.
[5] T.J. Hollowood, V.V. Khoze, M.P. Mattis, Instantons in N = 4 Sp(N) and SO(N) theories and
the AdS/CFT correspondence, hep-th/9910118.
[6] N. Ishibashi, H. Kawai, Y. Kitazawa, A. Tsuchiya, A large N reduced model as superstring,
Nucl. Phys. B 498 (1997) 467, hep-th/9612115.
[7] T. Eguchi, H. Kawai, Reduction of dynamical degrees of freedom in the large N gauge theory,
Phys. Rev. Lett. 48 (1982) 1063.
[8] J. Ambjrn, K.N. Anagnostopoulos, W. Bietenholz, T. Hotta, J. Nishimura, Large N dynamics
of dimensionally reduced 4D SU(N) super YangMills theory, hep-th/0003208.
[9] W. Krauth, H. Nicolai, M. Staudacher, Monte Carlo approach to M-theory, Phys. Lett. B 431
(1998) 31, hep-th/9803117.
[10] W. Krauth, M. Staudacher, Finite YangMills integrals, Phys. Lett. B 435 (1998) 350, hepth/9804199.
[11] W. Krauth, M. Staudacher, Eigenvalue distributions in YangMills integrals, Phys. Lett. B 453
(1999) 253, hep-th/9902113.
[12] W. Krauth, J. Plefka, M. Staudacher, YangMills integrals, Class. Quant. Grav. 17 (2000) 1171,
hep-th/9911170.
[13] T. Hotta, J. Nishimura, A. Tsuchiya, Dynamical aspects of large N reduced models, Nucl. Phys.
B 545 (1999) 543, hep-th/9811220.
[14] V.G. Kac, A.V. Smilga, Normalized vacuum states in N = 4 supersymmetric YangMills
quantum mechanics with any gauge group, hep-th/9908096.
[15] G. Moore, N. Nekrasov, S. Shatashvili, D-particle bound states and generalized instantons,
Commun. Math. Phys. 209 (2000) 77, hep-th/9803265.
[16] M.L. Mehta, Random Matrices, 2nd edn., Academic Press, 1991.
[17] I. Pesando, Exact results for the supersymmetric G2 gauge theories, Mod. Phys. Lett. A 10
(1995) 1871, hep-th/9506139.

Nuclear Physics B 584 [PM] (2000) 659681


www.elsevier.nl/locate/npe

Target space duality I: general theory


Orlando Alvarez
Department of Physics, University of Miami, P.O. Box 248046, Coral Gables, FL 33124, USA
Received 10 April 2000; accepted 22 May 2000

Abstract
We develop a systematic framework for studying target space duality at the classical level. We
e arises because of the existence of a very
show that target space duality between manifolds M and M
e and admits a double fibration.
special symplectic manifold. This manifold locally looks like M M
We analyze the local geometric requirements necessary for target space duality and prove that both
manifolds must admit flat orthogonal connections. We show how abelian duality, nonabelian duality
and PoissonLie duality are all special cases of a more general framework. As an example we exhibit
e = Rn . 2000 Elsevier Science B.V. All rights reserved.
new (nonlinear) dualities in the case M = M
PACS: 11.25.-w; 03.50.-z; 02.40.-k
Keywords: Duality; Strings; Geometry

1. Introduction
The (1 + 1)-dimensional sigma model describes the motion of a string on a manifold.
The sigma model is specified by giving a triplet of data (M, g, B), where M is the target
n-dimensional manifold, g is a metric on M, and B is a 2-form on M. The Lagrangian for
this model is
 x i x j
x i x j 
x i x j
1

+ Bij (x)
(1.1)
L = gij (x)
2



with canonical momentum density
L
= gij x j + Bij x 0j ,
(1.2)
x i
where an overdot denotes the time derivative (/ ) and a prime denotes the space
derivative (/ ) on the worldsheet. What is remarkable is that it possible for two
e g,
e to describe the same physics.
completely different sigma models, (M, g, B) and (M,
B),
i =

This work was supported in part by National Science Foundation grant PHY-9870101.

oalvarez@miami.edu

0550-3213/00/$ see front matter 2000 Elsevier Science B.V. All rights reserved.
PII: S 0 5 5 0 - 3 2 1 3 ( 0 0 ) 0 0 3 1 4 - X

660

O. Alvarez / Nuclear Physics B 584 [PM] (2000) 659681

By this we mean that there is a canonical transformation between the space of paths
e that preserves the respective Hamiltonians. This
on M and the corresponding one on M
phenomenon is known as target space duality.
This is the first of two articles where we develop a systematic framework for studying
target space duality at the classical level. We do not consider quantum aspects of target
space duality nor do we consider examples involving mirror symmetry. Most of our
considerations are local but phrased in a manner that is amenable to globalization. We
analyze the local geometric requirements necessary for target space duality. The study
of target space duality has developed by discovering a succession of more and more
complicated examples (see below). We show that the known examples of abelian duality,
nonabelian duality and PoissonLie duality are all derivable as special cases of the
framework. We show that target space duality boils down to the study of some very special
symplectic manifolds that allow the reduction of the structure group of the frame bundle
to SO(n). In article I we develop the general theory and apply it so some very simple
examples. In article II [1] we systematically apply the theory to a variety of scenarios and
we reproduce nonabelian duality and PoissonLie duality. The theory is applied to other
geometric situations that lead us deep into unknown questions in Lie algebra theory. We
try to make article I self contained. References to equations and sections in article II are
preceded by II, e.g., (II-8.3).
What is the value in developing a general framework for studying classical target space
duality? The framework may say something about the what is string theory. We believe
that there is some parameter space that describes string theory. For special values of the
parameters we get the familiar Type I, Type II-A, Type II-B, etc., theories and that these
are related by various dualities. If we can get a handle on the class of symplectic manifolds
that lead to target space duality we may be able to get a better idea about the parameter
space of string theory.
The simplest target space duality is abelian duality. Here a theory with target space S 1
or R is dual to a theory with target space S 1 or R. For a comprehensive review and history
of abelian duality look in [2]. It should also be mentioned that it has been known for a long
time, see, e.g., [3], that the abelian duality transformation is a canonical transformation.
A first attempt to generalize abelian duality to groups led to the pseudochiral model of
Zakharov and Mikhailov [4] as a dual to the nonlinear sigma model. Nappi [5] showed
that these models were not equivalent at the quantum level. The correct dual model was
first found by Fridling and Jevicki [6] and Fradkin and Tseytlin [7] using path integral
methods. String theory motivated a renewed interest in abelian and nonabelian duality [8
15]. It was shown that the duality transformation was canonical [16,17] and these ideas
were generalized in a variety of ways [1822]. The form of the generating functions for
duality transformation gave hints that nonabelian duality was associated with the geometry
of the cotangent bundle of the group.
The most intricate target space duality discovered thus far is the PoissonLie duality of
Klimcik and Severa [2325]. In this example we see a very nontrivial geometrical structure
playing a central role. A PoissonLie group G is a Lie group with a Poisson bracket that
is compatible with the group multiplication law. Drinfeld [26] showed that PoissonLie

O. Alvarez / Nuclear Physics B 584 [PM] (2000) 659681

661

groups are determined by a Lie bialgebra gD = g g where g is the Lie algebra of G


e See Appendix II-B.1. The two Lie algebras are
and g is the Lie algebra of a Lie group G,
coupled together in a very symmetric way. A Lie group GD with Lie algebra gD is called
e is also a PoissonLie group. By using
a Drinfeld double. It should be pointed out that G
a clever argument, Klimcik and Several discovered that if the metric g and B field on a
PoissonLie group G was of a special form then there would be a corresponding metric g
e
e Their observations follow from the symmetric way that G and
and B-field
on the group G.
e
G enter into the Drinfeld double GD . They showed that that by writing down a first order
e by taking
sigma model on GD they could derive either the model on G or the model on G
an appropriate slice. Here one explicitly sees that the the target manifold and the target dual
manifold are carefully glued together into a larger space. They are totally explicit about
the metric and B field. Klimcik and Severa discuss PoissonLie duality as a canonical
transformation transformation in [23] and show that it is related to the symplectic structure
on the Drinfeld double [27]. In [24] they explicitly mention that the generating function
for PoissonLie duality may be obtained from the symplectic structure on the Drinfeld
double. A different approach was pursued by Sfetsos [28,29] who also wrote down the
duality transformation, verified that it was a canonical transformation, and constructed the
generating function for the canonical transformation, see also [30].
At the time of the work by Klimcik and Severa, the author had been working on a
program to develop a general theory of target space duality, see [31]. In that article I
advocated the use of generating functions of the type (2.2) because they would lead to

)
that preserved the quadratic
a linear relationship 1 between (dx/d, ) and (dx/d,
nature of the sigma model Hamiltonians. I discussed the geometry which was involved
and explained the role played in this geometry by the hamiltonian density H and the
momentum density P. Explicit formulas relating the geometries of the two manifolds were
not given in that article for the following reason. The formulation I had at the time involved
variables (x, p) where essentially = dp/d . This gave a certain symmetry to some of
the equations but at a major price. The B field gauge symmetry B B + dA became
a nonlocal symmetry in (x, p) space and the gauge symmetry was no longer manifest.
Only for special choices of A was the gauge transformation local. The formulas I had
derived respected the special gauge transformations but I could not verify general gauge
invariance. Sfetsos [28] exploited some of the geometric constraints I had proposed and
he was able to explicitly construct the duality transformation for PoissonLie duality.
Sfetsos work is very interesting. He conjectures the form of the duality transformation
e g,
e from the work of Klimcik and
and he knows the geometric data (M, g, B) and (M,
B)
Severa. He now uses this information and certain integrability constraints to explicitly work
out the generating function for the canonical transformation. Sfetsos computation may be
reinterpreted as the construction of a known symplectic structure [24,27,34] on the Drinfeld
double, see Section II-3.
In this article I present a general theory for target space duality that is manifestly gauge
invariant with respect to B field gauge transformations. I consider what could be called
1 For nonpolynomial generating functions look at [32,33].

662

O. Alvarez / Nuclear Physics B 584 [PM] (2000) 659681

irreducible duality where there are no spectator fields. All the fields participate actively
in the duality transformation. I show that the duality transformation arises because of
e and
the existence of a special symplectic manifold P that locally looks like M M
admits a double fibration. The duality transformation exists only when there exists a
compatible confluence of several distinct geometric structures associated to the manifold
P : an O(2n) structure related to the hamiltonian density (3.1), an O(n, n) structure related
to the momentum density (3.2), an O(n) O(n) structure associated with the sigma model
metrics, and a Sp(2n) structure related to the symplectic form. This is why these symplectic
manifolds are very special and rare. I develop the general theory and then show how the
known examples of abelian duality, nonabelian duality and PoissonLie duality follow.
The general theory indicates that there are probably many more examples. For example, in
Section 8.3 I write down families of nonlinear duality transformations that map a theory
with target space Rn into one with target space Rn . I also investigate a variety of scenarios
and pose open mathematical questions deeply related to the theory of Lie algebras.
This work differs from the work of Sfetsos [28] in a variety of ways. There are two
types of constraints on the canonical transformation: algebraic constraints having to do
with quadratic form of the Hamiltonian density and differential constraints having to do
integrability conditions. Sfetsos writes these down but in a way that is neither geometric
nor gauge invariant. He applies them to PoissonLie duality and derives the generating
function. Sfetsos formulation does not exploit the fact that there are natural geometric
structures associated to these equations. This is what I was trying to do in [31] but failed
due to a bad choice of variables (x, p) leading to an absence of manifest B field gauge
invariance. The formulation presented here uses the variables (x, ) and is manifestly
gauge invariant. In Section II-2.2.2 I give a geometric interpretation of B field gauge
invariance. In this article I work in terms of adapted frame fields. In this way, the formalism
has an immediate interpretation in terms of H -structures on the bundle of frames. In fact
the discussion presented in Section II-4.1 is done in a sub-bundle of the bundle of frames.
The framework developed in this work allows one to attack a variety of interesting
e We show
questions. Are there any interesting restrictions on the manifolds M and M?
e
in Section 6 that the manifolds M and M have to admit flat orthogonal connections. We
know for any manifold M there always exists a natural symplectic manifold P = T M, the
cotangent bundle. We can ask what type of dualities arises from the standard symplectic
structure on the cotangent bundle? We show that this can only happen if M is a Lie group,
see Section II-2.2.1. This formalism allows general question to be asked. For example there
are a series of PDEs that have to be solved to determine the duality transformations. These
PDEs depend on some functions. If these functions are zero then one gets abelian duality,
if some are made nonzero then you get nonabelian duality, etc. This is a framework that
can be used for a systematic study of duality. It opens up the possibility to study dualities
involving parallelizable manifolds that are not Lie groups such as S 7 or sub-bundles of the
frame bundle. This work indicates that duality is a very rich geometrical framework ripe
for study and we have only scratched the surface.

O. Alvarez / Nuclear Physics B 584 [PM] (2000) 659681

663

2. The symplectic structure


We review briefly the notion of a generating function in canonical transformations
because our methods introduce a secondary symplectic structure into the formulation of
target space duality and it is important to understand the difference between the two.
Assume you have symplectic manifolds, P and Pe, with respective symplectic forms
e Pe.
e:P P
and .
Consider P Pe with standard projections : P Pe P and
You can make P Pe into a symplectic manifold by choosing as symplectic form  =
e . By definition, a canonical or symplectic transformation f : P Pe satisfies

e will be symplectic
f = . We describe f by its graph f P Pe. It is clear f : P P

if and only if |f = 0. Locally we have = d and = d . Thus we see that is


a closed 1-form on f . Consequently there exists locally a function F : f R such
that = dF . This function F is called the generating function for the symplectic
transformation. The reason is that if in local Darboux coordinates we have that = p dq
and = p dq then we have that F is locally a function of only q and q,
p = F / q and
p = F /q. We can now use the inverse function theorem to construct the map from
(q, p) to (q,
p).
Note that dim f = 2n and therefore F is a function of 2n variables.
Had we chosen = q dp then we would have that F is a function of q and p. In this
case it is worthwhile to observe F = qp
generates the identity transformation. We mention
this because the identity transformation is not in the class of transformations generated by
functions of q and q.

All this generalizes to field theory. We discuss only the case of (1 + 1) dimensions. Let
P (M) be the path space of M. By this we mean the set of maps { : N M}, where
N can be R, S 1 or [0, ] depending on whether we are discussing infinite strings, closed
strings or open strings. Most of the discussion in this article is local and so we do not
specify N . In the case of a sigma model with target space M, the basic configuration space
is P (M) with associated phase space P (T M). If (x, ) are coordinates on T M then the
symplectic structure on P (T M) is given by
Z
( ) x( )d.
(2.1)
In what follows we are interested in looking for canonical transformations between a sigma
e of the same dimensionality. We
model with target space M and one with target space M
e g,
e
say that a sigma model with geometrical data (M, g, B) is dual to a sigma model (M,
B)

e that preserves the


if there exists a canonical transformation F : P (T M) P (T M)
e = H, where the Hamiltonian density is given by (3.1).
hamiltonian densities, F H
In the case of abelian duality where the target space is a circle you can choose the
generating function to be
Z
dx
F [x, x]
= x d.
d
This leads to the standard duality relations ( ) = dx/d

and (
) = dx/d.
The nonabelian duality relations follow from the following natural choice [20,22] for
generating function. Assume the target space is a simple connected compact Lie group

664

O. Alvarez / Nuclear Physics B 584 [PM] (2000) 659681

G with Lie algebra g. The dual manifold is the Lie algebra with an unusual metric. The
generating function is very natural:


Z
e = Tr Xg
e 1 dg d,
F [g, X]
d
e is a Lie algebra valued field.
where X
We now consider a class of generating functions for target space duality that leads to
a linear relationship [31] between (dx/d, ( )) and the corresponding variables on the
e choose locally a 1-form = i (x, x)
dx i + i (x, x)
dx i . We can
dual space. On M M
e by
define a natural generating function on P (M M)


Z
Z
 dx i
 dx i
+ i x( ), x(
d. (2.2)
)
)
F [x( ), x(
)] = =
i x( ), x(
d
d
We only consider target space duality that arises from this type of canonical transformation.
e with compact support which
Let v be a vector field along the path (x( ), x(
)) MR M
R
represents a deformation of the path. Note that v F = Lv = v d. In the previous
formula Lv = v d + dv is the Lie derivative with respect to v. Since v has compact
support, the exact term can be neglected. Thus the variation of F is determined by the exact
2-form = d:
Z
(2.3)
v F = v .
We use to construct the duality transformation. If x and x are respectively local
e then
coordinates on M and M
1
1
dx i dx j + mij (x, x)
dx i dx j + lij (x, x)
dx i dx j ,
(2.4)
= lij (x, x)
2
2
m are used to
lij = lj i and lij = lj i . The three n n matrix functions l, l,
where l:
construct the canonical transformation on the infinite dimensional phase space. A brief
calculation shows that the canonical transformations are
dx j
dx j
+ lij (x, x)
,

d
d
dx j
dx j
+ lij (x, x)
.

i ( ) = mij (x, x)
d
d

i ( ) = mj i (x, x)

(2.5)
(2.6)

e requires
The invertibility of the canonical transformation between P (T M) and P (T M)
m to be an invertible matrix. This implies that is of maximal rank, i.e., a symplectic
e
form 2 on M M.
It is important to recognize that there are two very different symplectic structures in this
problem. The first one is the standard symplectic structure on phase space P (T M) given
e given by arises from the class of generating functions
by (2.1). The second one on M M
2 It is possible for to be symplectic and have m = 0 but this will not define an invertible canonical
e For example, if M and M
e are symplectic manifolds with
transformation between P (T M) and P (T M).
respective symplectic forms and then choose = .

O. Alvarez / Nuclear Physics B 584 [PM] (2000) 659681

665

e and the horizontal


Fig. 1. A bifibration where the vertical fibers are diffeomorphic to a cover of M
ones are diffeomorphic to a cover of M.

(2.2) we are considering. The generating function arguments are local and suggest that the
e may be generalized to a symplectic manifold P which
symplectic structure on M M
e
e you have natural cartesian projections
contains M M. In the cartesian product M M
e
e
e
e
c : M M M and c : M M M. The product structure can be generalized by the
introduction of the concept of a bifibration. A 2n-dimensional manifold P is said to be a
e and projections : P M
bifibration if there exists n-dimensional manifolds M and M
e
e
e
and : P M such that the respective fibers are diffeomorphic to coverings spaces of M
e
and M and they are also transverse. This means that if p P then ker |p ker |p =
e are the differential maps of the projections. Note that the cartesian
Tp P , where and
e is an example of a bifibration. If the product projection
product manifold P = M M
e A covering space example is given
e is injective 3 then P = M M.
e :P M M

2
1
ix
e
e : (x, x)
7 e and
7 eix .
by P = R and M = M = S with : (x, x)
We introduce the following terminology illustrated in Fig. 1. At a point p P we have a
splitting of the tangent space Tp P = Hp Vp where the horizontal tangent space Hp is
e, and the vertical tangent space Vp is tangent to the fiber of .
tangent to the fiber of
A symplectic form is said to be bifibration compatible if for every p P one has the
following nondegeneracy conditions:
1. Given Y Vp , if for all X Hp one has (X, Y ) = 0 then Y = 0.
2. Given X Hp , if for all Y Vp one has (X, Y ) = 0 then X = 0.
This is the coordinate independent way of stating that the matrix mij is invertible. The
reader can verify that the symplectic form given in footnote 2 fails the above. Our abstract
scenario is a bifibration 4 with a bifibration compatible symplectic form. We will refer
to such a manifold as a special symplectic bifibration. We believe that the formulation
of canonical transformations in path space in terms of is probably more fundamental
than the use of a generating function. This is probably analogous to the ascendant role the
symplectic 2-form has taken in symplectic geometry because of global issues. The 2-form
3 The definition of a fiber bundle implies that
e is surjective.
4 It may be possible to generalize from a bifibration to a bifoliation. It is not clear to the author what is the most

general formulation.

666

O. Alvarez / Nuclear Physics B 584 [PM] (2000) 659681

may play a role in the quantum aspects of duality maybe in some geometric quantization
type of framework.
A scenario for how a natural generating function of type (2.2) might arise is the
following. For any M, the cotangent bundle T M is a symplectic manifold. Firstly, one
has to investigate whether the cotangent bundle admits a second fibration transverse to the
defining one. Secondly, it may be necessary to deform the original symplectic structure.
In the case of a Lie group G, the cotangent bundle is trivial and is thus a product T G =
G g where we have used the metric on G to identify the Lie algebra g with its vector
space dual g .

3. Hamiltonian structure
The discussion in the Section 2 is general and makes no reference to the Hamiltonian.
The Hamiltonian only played an indirect role because we chose a class of canonical
transformations which are linear with respect to dx/d and ( ) in anticipation of future
application to the nonlinear sigma model. The nonlinear sigma model has target space a
riemannian manifold M with metric g and a 2-form field B. The Hamiltonian density and
the momentum density are respectively given by



dx i dx j
1
dx k
dx l
1
,
(3.1)
j Bj l
+ gij (x)
H = g ij (x) i Bik
2
d
d
2
d d
dx i
.
(3.2)
d
We are interested whether we can find a canonical transformation with generating function
of type (2.2) which will map the Hamiltonian density and momentum density into that
e metric
of another sigma model (the dual sigma model) characterized by target space M,
e
tensor g and 2-form B.
It winds up that working in coordinates is not the best way of attacking the problem. It is
best to use moving frames la Cartan and Chern. Let ( 1 , . . . , n ) be a local orthonormal
coframe 5 for M. The Cartan structural equations are
P = i ( )

d i = ij j ,
1
dij = ik kj + Rij kl k l ,
2
where ij = j i is the riemannian connection. 6 Next we define dx/d in the
orthonormal frame to be x by requiring that i = x i d . If is now the canonical
momentum density in the orthonormal frame then in this frame (3.1) and (3.2) become
5 Because we will be working in orthonormal frames we do not distinguish an upper index from a lower index
in a tensor.
6 The riemannian connection is the unique torsion free metric compatible connection. A metric compatible
connection will also be referred to as an orthogonal connection. In general an orthogonal connection can have
torsion.

O. Alvarez / Nuclear Physics B 584 [PM] (2000) 659681


 1
1
i Bik x k i Bil x l + x i x i ,
2
2

P = i x i = i Bij x j x i .

H=

667

(3.3)
(3.4)

In this coframe we can write (2.4) as


1
1
i j + mij (x, x)
i j + lij (x, x)
i j .
(3.5)
= lij (x, x)
2
2
We use the same letters l, m, l but the meaning above is different from (2.4). In this notation
equations (2.5) and (2.6) become
x j + lij (x, x)x
j,
i ( ) = mj i (x, x)
j + lij (x, x)
x j .
i ( ) = mij (x, x)x

(3.6)
(3.7)

In matrix notation the above may be written as


 


 t 
m 0
x
l I
x
=
.

m 0
l I
Rewrite the above in the form
 



 t
n I
x
x
m 0
,
ex = m 0
B
Bx
n I
where
n = l B,
e
n = l B.

(3.8)
(3.9)

The rewriting above is closely related to (A.2), see below. This equation is not very
interesting in this form but it becomes much more interesting when rewritten as
  t


1 

m 0
n I
x
x
ex = n I
B
Bx
m 0



x
(mt )1
(mt )1 n
.
(3.10)
=
n(m
t )1 n + m n(m
t )1
Bx
Notice that equation (3.10) gives us a linear transformation between (x , Bx )
ex ). The preservation of the Hamiltonian density means that this linear
and (x , B
transformation must be in O(2n). If in addition you want to preserve the momentum density
then this transformation must be in OQ (n, n), the group of 2n 2n matrices isomorphic to
O(n, n) which preserves the quadratic form


0 In
.
(3.11)
Q=
In 0
In the formula above, In is the n n identity matrix. Properties of OQ (n, n) and
its relation with O(2n) are reviewed in Appendix A. They key observation 7 is that
7 I do not understand geometrically why automatically induces this pseudo-orthogonal matrix.

668

O. Alvarez / Nuclear Physics B 584 [PM] (2000) 659681

the matrix appearing in (3.10) is automatically in OQ (n, n) which means that our
canonical transformation automatically preserves the canonical momentum density (3.4).
As previously mentioned to preserve the hamiltonian density (3.3) is it necessary that the
matrix above also be in O(2n). Thus the matrix


(mt )1
(mt )1 n
(3.12)
n(m
t )1 n + m n(m
t )1
must be in O(2n) OQ (n, n), a compact group locally isomorphic to O(n) O(n), see
Appendix A. Using the equations in the appendix we learn that the condition that (3.12) be
in the intersection O(2n) OQ (n, n) is that
mmt = I n 2 ,

(3.13)

m m=I n ,

(3.14)

mn = nm.

(3.15)

We can now simplify (3.12) to




(mt )1
(mt )1 n
.
(mt )1
(mt )1 n

(3.16)

To better understand the above is is worthwhile using the conjugation operation (A.4)
and switch the quadratic from from Q to


I
0
.
0 +I
Under this conjugation operation (3.10) becomes


ex )
x ( B
ex )
x + ( B



x ( Bx )
(mt )1 (I + n)
0
.
=
0
(mt )1 (I n)
x + ( Bx )
This leads to the pair of equations



ex + x = +T+ ( Bx ) + x ,
B



ex x = T ( Bx ) x ,
B

(3.17)

(3.18)
(3.19)

where
T = (mt )1 (I n) O(n).

(3.20)

An equivalent way of writing the above is m = T (I n). Also note that T+ and T
are not independent. They are related by T1 T+ = (I + n)1 (I n) which is the Cayley
transform of n. It is often convenient to think that (3.5) is determined by two orthogonal
matrices T O(n) with
n = (T+ + T )1 (T+ T ).

(3.21)

O. Alvarez / Nuclear Physics B 584 [PM] (2000) 659681

669

4. Gauge invariance
It is well known that the sigma model (M, g, B) has a gauge invariance given by B
B + dA where A is a 1-form on M. We can manifest these
R gaugeR transformations within the
class (3.20) of canonical transformation by considering ( + A) which transforms
appropriately. An observation and a change of viewpoint will give us a manifestly
gauge invariant formulation. Notice that both the left hand side and right hand side of
Eq. (3.17) is manifestly gauge invariant. This suggests that m, n, n may be gauge invariant.
Looking at (3.8) and (3.9) and incorporating the remark about how we implement gauge
invariance we see that n and n are gauge invariant quantities, i.e., the gauge transformations
e This suggest that instead of
are implemented by shifting l, l respectively by dA and dA.
working with it may be worthwhile to work with defined by
1
1
i j + mij (x, x)
i j + n ij (x, x)
i j ,
= nij (x, x)
2
2
where is not closed but satisfies
e,
d = H H

(4.1)

(4.2)

e. We have now
e = dB.
e More correctly one has d = H
e H
where H = dB and H
achieved a gauge invariant formulation.

5. The geometry of P
e is it best to
To gain further insight into relations between the geometry of M and M
e We can use the freedom of
work in P which you may think of it locally being M M.
e
working in P to simplify results and then project back to either M or M.
There are two closely related ways of simplifying the geometry. One way is to work
in the bundle of orthonormal frames. The other is to adapt the orthonormal frames to the
problem at hand similar to the way one uses Darboux frames to study surfaces in classical
differential geometry. The former gives a global formulation but the latter is more familiar
to physicists hence we choose the latter. All our computations will be local and can be
patched together to define global objects.
The first thing to observe is that the existence of the double fibration allows us to
e and
naturally define a riemannian metric on P by pulling back the metrics on M and M
declaring that the fibers are orthogonal to each other. In a similar fashion we pullback local
coframes and get local coframes on P . These orthonormal coframes satisfy the Cartan
structural equations
d i = ij j ,
d i = ij j ,
1
dij = ik kj + Rij kl k l ,
2
1 e k l
d ij = ik kj + R
ij kl .
2

(5.1)
(5.2)
(5.3)
(5.4)

670

O. Alvarez / Nuclear Physics B 584 [PM] (2000) 659681

Once we begin working on P then we have the freedom to independently rotate and
at each point. Once we do this these coframes will no longer be pullbacks but this doesnt
matter because it does not change the metric on each fiber. We are going to exploit this
e in a way similar to the way the intrinsic
freedom to relate the geometry of M to that of M
curvature of a submanifold is related to the total curvature of the space and the curvature
of the normal bundle. Note that with these choices there is a natural group of O(n) O(n)
gauge transformations on the tangent bundle of P which is compatible with the metric
structure and the bifibration.

6. Constraints from the algebraic structure of


First we derive various constraints that follow from the algebraic constraints on
imposed by the preservation of H and P. Eqs. (3.14) and (3.15) tell us that
m = T (I + n)

and n = T nT t

(6.1)

where T O(n). Since T connects a to a we see that its covariant differential is


given by
dTij + ik Tkj + j k Tik = +Tij k k Teij k k ,

(6.2)

where the components of the covariant differential in the M direction is +Tij k and in the
e direction is Teij k . The negative sign is introduced for future convenience. Notice that
M
eij k are tensors defined on P whose existence is guaranteed by the existence of
Tij k and T
the tensor Tij on P .
We now invoke a symmetry breaking mechanism to reduce the structure group of
gauge transformations from O(n) O(n) to O(n). At each point in P we can rotate
e R 1 where
(or ) and make T = I because under these gauge transformations T RT
e O(n) O(n). The isotropy group of T = I is the diagonal O(n). This is no
(R, R)
different than giving a scalar field a vacuum expectation value to break the symmetry.
This symmetry breaking leads to an identification at each point of P of the vertical and
horizontal tangent spaces. This does not tell us that the metrics are the same but allows us
to identify an orthonormal frame in one with an orthonormal frame in the other. Let us be a
bit more precise and abstract on the reduction of the structure group and the identification
of the vertical and horizontal tangent spaces. We already mentioned that at p P one
has Tp P = Hp Vp . The tensor m(p) may be viewed as an element of Vp Hp . Because
there is a metric on Vp we can reinterpret m as giving us an invertible linear transformation
m
: Hp Vp . We also have a metric on Hp and thus we can study the orbit of m(p)
under the action of O(n) O(n). Our previous discussion shows that a canonical form
for m(p) may be taken to be m(p) = I + n(p) with isotropy group being the diagonal
O(n). If (e1 , . . . , en ) is an orthonormal basis at Hp and (e1 , . . . , en ) is the corresponding

orthonormal basis at Vp then they are related by m(p)e


i = ej (j i + nj i (p)).
From now on we assume we have adapted our coframes such that T = I and
mij = ij + nij ,

(6.3)

O. Alvarez / Nuclear Physics B 584 [PM] (2000) 659681

nij = n ij.

671

(6.4)

In this frame, simplifies to


1
1
= i i + nij i j nij i j nij i j .
2
2
The duality equations are particularly simple now and they are given by

ex + x = ( Bx ) + x ,
B



ex x = T ( Bx ) x ,
B

(6.5)

(6.6)
(6.7)

where the orthogonal matrix T is the Cayley transform of n:


I +n
.
(6.8)
I n
The matrix T is not arbitrary because there are constraints on nij as we will see later
on. Without constraints on T there are interesting solutions to (6.6) and (6.7) which map
spaces of constant positive curvature into spaces of negative constant curvature or more
generally dual symmetric spaces. 8
We can now exploit equation (6.2) to relate the connections in the adapted coframing.
Inserting T = I into the above leads to
T =

ij ij = +Tij k k Teij k k .

(6.9)

Thus we see that in the reduction of the structure group we have generated torsion and
that this torsion satisfies Tij k = Tj ik and Teij k = Tej ik . We now define an orthogonal
connection on our adapted frames by
ij = ij + Tij k k = ij + Teij k k .

(6.10)

e by
First we define the components of the covariant derivatives of T and T
dTij k + ( T )ij k = Tij0 kl l + Tij00kl l ,

(6.11)

e)ij k = Teij0 kl l + Teij00kl l .


dTeij k + ( T

(6.12)

In the above ( T ) and ( Te) are abbreviations for standard expressions. We have chosen
to use the connections and rather than in the definition of the covariant derivative
for the following reasons: if Tij k is the pullback of a tensor on M then Tij00kl = 0; if Teij k is
e then Te0 = 0. A notational remark is that a primed tensor
the pullback of a tensor on M
ij kl
denoted the covariant derivative in the M direction and a doubly primed tensor denotes the
e direction. Doubly primed does not mean second derivative.
covariant derivative in the M
The curvature of this connection may be computed by either using the expression
involving or the one involving .
A straightforward computation of the curvature matrix
2-form
ij = dij + ik kj
in these two ways leads to the following expressions
8 O. Alvarez, unpublished.

(6.13)

672

O. Alvarez / Nuclear Physics B 584 [PM] (2000) 659681

ij = Tij00lm l m


1
+ Rij lm Tij0 lm Tij0 ml + (Tikl Tkj m Tikm Tkj l ) l m ,
2
and
ij = Teij0 lm l m


1e
eij00lm Teij00ml + Teikl Tekj m Teikm Tekj l l m .
+ R
ij lm T
2
Comparing these two expression we learn that the curvature two form matrix is given by
ij = dij + ik kj = Tij00lm l m .

(6.14)

The following constraints must also hold



Rij lm Tij0 lm Tij0 ml + (Tikl Tkj m Tikm Tkj l ) = 0,


eikl T
ekj m Teikm T
ekj l = 0,
eij lm Teij00lm Teij00ml + T
R

(6.16)

+ Teij0 ml = 0.

(6.17)

Tij00lm

(6.15)

Form (6.14) is reminiscent of a Khler manifold where the curvature is of type dz dz


and there are no dz dz or dz dz components. The absence of these many curvature
components is due to the reduction of the structure group from O(2n) to O(n) at the
expense of generating torsion.
There are a variety of equivalent ways of interpreting the above. The most geometric is
to observe that ij defines a connection on P and thus a connection when restricted to any
e1 (x)
be a horizontal fiber. Notice that along this
of the fibers. For example, let Mx =
fiber = 0 and thus ij = 0. Since Mx is isometric to M we have found a flat orthogonal
connection (generally with torsion) on M. Note that this is true for all horizontal fibers.
One can make a similar statement about the vertical fibers. We have our first major result.
e respectively, admit flat
Target space duality requires that the manifolds M and M,
e
orthogonal connections. The connection ij is flat when restricted to either M or M.
At a more algebraic level equations (6.15) and (6.16) are the standard equations for
parallelizing the curvature by torsion. A manifold M is said to be parallelizable if
the tangent bundle is a product bundle T M = M Rn . This means that you can
globally choose a frame on M. The existence of a flat connection on a manifold does
not imply parallelizability. The reason is that in a non-simply connected manifold there
is an obstruction to globally choosing a frame if there is holonomy. If the manifold
is simply connected and the connection is flat then it is parallelizable. Finally we
observe that if a manifold is parallelizable then there are an infinite number of other
possible parallelizations. 9 Assume we have an orthogonal parallelization, i.e., a choice
of orthonormal frame at each point. Given any other orthogonal parallelization we can
always make a rotation point by point so that both frames agree at the point. Thus the
space of all orthogonal parallelizations is given by the set of maps from M to O(n).
9 I would like to thank I.M. Singer for the ensuing argument.

O. Alvarez / Nuclear Physics B 584 [PM] (2000) 659681

673

e the tensor Tij k on the respective


Note that given two distinct points x1 , x2 M,
horizontal fibers Mx1 and Mx2 do not have to be the same. There are many flat orthogonal
connections on M as can be seen by a variant parallelizability argument. In fact you could
e
in principle have a multiparameter family parametrized by M.
There is a special case of interest when Tij k is the pullback of a tensor on M. In this case
e0 = 0.
a previous remark tells us that Tij00kl = 0 and consequently by (6.17) we also have T
ij kl
e This means that the same torsion
Therefore Teij k is also the pullback of a tensor on M.
tensors make the connection flat on all the fibers. Note that in this case ij = 0 and the
orthogonal connection ij is a flat connection on P .
e and
If Tij k is the pullback of a tensor on M then Teij k is the pullback of a tensor on M
ij = 0. In this case ij is a flat connection on P .

7. Simple examples
e introduces relations among H, H
e, Tij k and Teij k . First we
The equation d = H H
point out some facts.
7.1. The case of nij = 0
As a warmup we study the case where nij = 0. In this case = i i and we compute
d by using the Cartan structural equations (5.1), (5.2) and the condition which follows
from the reduction of the symmetry group (6.9). A brief computation yields
eij k i j k .
d = Tkij i j k T
e vanish. Next we see that Tkij = Tkj i and Teij k =
First we learn that the 3-forms H and H
e
Tikj . We remind the reader that a tensor Sij k which is skew symmetric under i j
eij k = 0. It follows
and symmetric under j k is zero. Thus we conclude that Tij k = T
eij kl = 0. Since the Riemannian curvatures
from equations (6.15) and (6.16) that Rij kl = R
e are manifolds with universal cover Rn . There are no other
vanish we know that M and M
possibilities if nij = 0. For example, you can have M = Tk Rnk . This is the case of
abelian duality. Other potential singular cases of interest are orbifolds or cones which are
flat but have holonomy due to the presence of singularities.
7.2. The case of a Lie group
We verify that the standard nonabelian duality results are reproducible in this formalism.
We present a schematic discussion here because the Lie group example is a special case of
a more general result presented in Section II-2.2.1. Let G a compact simple Lie group
with Lie algebra g. Let (ei , . . . , en ) is an orthonormal basis for g with respect to the
Killing form. The structure constants fij k are defined by [ei , ej ] = fkij ek . In this case

674

O. Alvarez / Nuclear Physics B 584 [PM] (2000) 659681

the structure constants are totally antisymmetric. Let i be the associated MaurerCartan
forms satisfying the MaurerCartan equations
1
(7.1)
d i = fij k j k .
2
Because of the Killing form we can identify the Lie algebra g with its vector space dual g .
We choose P to be the cotangent bundle T G which is a product bundle T G = G g =
G g. If (p1 , . . . , pn ) are the standard coordinates on the cotangent bundle with respect
to the orthonormal frame then the we take in (2.2) to be = pi i , the canonical 1-form
on T G. Therefore = d is the standard symplectic form on T G given by
1
= dpi i pi fij k j k .
2

(7.2)

By looking at reference [22] one can see that the orthonormal coframe ( 1 , . . . , n ) on the
fiber g is given by dpj = i (ij + fkij pk ). This suggests that mij = (ij + fkij pk ) and
that in this basis the symmetry breaking is manifest and thus nij = fkij pk . Thus we expect
that is given by
1
1
(7.3)
= fkij pk i j + (ij + fkij pk ) i j fkij pk i j .
2
2
e because the modification of going from the closed form to
Note that d = H
involved a term of the type nij i j . To verify this we observe that i = dpj m1 and
ji

thus nij i j only depends on p and dp, therefore, its exterior derivative can only be of
type dp dp dp . In fact, 12 fkij pk i j is the standard representation for
e
the 2-form B.
If we write d i = 1 fij k j k , then a straightforward exercise shows that
2

fij k = (mj m fmkl mkm fmj l )m1


li .
By using (B.1) one can compute ij . It is now an algebraic exercise to compute
parallelizing torsions Tij k and Teij k .

8. The case of a general connection


8.1. General theory
e
We already saw that the connection ij on P gives a flat connection on both M and M,
e to be target space duals of each other. We are going to
a necessary condition for M and M
take the following approach. Assume we are given a ij on P , how do we determine nij?
We will derive PDEs that nij must satisfy. If there exist solutions to these PDEs then we
e For it is
automatically have a duality between the sigma model on M and the one on M.
worthwhile to rewrite the Cartan structural equations in terms of ij :
1
d i = ij j fij k j k ,
2

(8.1)

O. Alvarez / Nuclear Physics B 584 [PM] (2000) 659681

1
d i = ij j fij k j k ,
2
dij = ik kj Tij00lm l m ,

675

(8.2)
(8.3)

where fij k = fikj , fij k = fikj and Tij00kl = Tj00ikl . The structure functions fij k and fij k
eij k by
are related to Tij k and T
1
Tij k = (fij k fj ik fkij ),
(8.4)
2

eikj , Teij k = 1 fij k fj ik fkij .
(8.5)
fij k = Teij k T
2
We define the components n0ij k , n00ij k , fij0 kl , fij00kl , fij0 kl , fij00kl of the covariant derivatives of
nij , fij k , fij k with respect to the connection ij by
fij k = Tij k Tikj ,

dnij + ik nkj + j k nik = n0ij k k + n00ij k k ,

(8.6)

dfij k + il flj k + j l filk + kl fij l = fij0 kl l + fij00kl l ,

(8.7)

dfij k + il flj k + j l filk + kl fij l = fij0 kl l + fij00kl l .

(8.8)

There are several important constraints which follow from d2 = d2 = 0:



fij0 kl + fmj k fiml j k l = 0,
fij00kl
k l

= Tij00kl


fij00kl + fmj k fiml j = 0,
fij0 kl

00
Tikj
l,

Tij00lk

00
Tiklj

(8.9)
(8.10)


.

(8.11)
(8.12)

Note that Tij00kl = 0 if and only if fij00kl = fij0 kl = 0, i.e., fij k and fij k are, respectively,
pullbacks in accord with a previous remark. The d2 ij = 0 constraints are not used in this
report and will not be given.
To derive the PDE satisfied by nij we compute d :
e
d = H H
1
1
= n0ij k i j k + fij k nil j k l
2
2
1
n00ij k i j k n0ij k i k j
2
1
1
+ fij k j k i fij k nil j k l
2
2
1
00
i
j
k
+ nij k n0ij k k i j
2
1 i
1
j
k
fij k fij k nil l j k
2
2
1
1
n00ij k i j k + fij k nil j k l .
2
2
If we write the closed 3-forms in components as

(8.13)

676

O. Alvarez / Nuclear Physics B 584 [PM] (2000) 659681

H=

1
Hij k i j k ,
3!

eij k i j k ,
e= 1 H
H
3!

(8.14)

eij k are totally skew symmetric then we immediately see that


where Hij k and H
n0ij k + n0j ki + n0kij = Hij k + (flij nlk + flj k nli + flki nlj ),

eij k + flij nlk + flj k nli + flki nlj ,
n00ij k + n00j ki + n00kij = +H

n0kij n0kj i n00ij k = (fkij nlk flij ) = mkl flij ,


n0ij k + n00kij n00kj i = + fkij + nlk flij = flij mlk .

(8.15)
(8.16)
(8.17)
(8.18)

The number of linearly independent equations above is 13 n(n 1)(2n 1). The best way
to see this is that if we define i = ( i i ) then the term containing nij in is basically
j
nij +i + . If the components of the covariant derivatives of nij in this basis are n
ij k

k
k
i
i
then d(nij +i +i ) n+
ij k + + + + nij k + + . The stuff in ellipsis does
+
not involves derivatives of nij . Since nij k is linearly independent of n
ij k we see that the
1
1
number of equations we get is 3! n(n 1)(n 2) + n 2 n(n 1). The first remark we
make is that the PDEs given by (8.6) generally make an overdetermined system if n > 1.
The reason is that there are 13 n(n 1)(2n 1) equations for 12 n(n 1) functions nij . This
means that for a solution to exist integrability conditions arising from d2 nij = 0 must be
satisfied.
V
Let tij k = tj ik be a tensor in ( 2 V ) V for some n dimensional vector space V
V
with inner product. The vector space ( 2 V ) V has an orthogonal decomposition into
V2
V3
V ) ((
V ) V )mixed where the latter are the tensors of mixed symmetry under the
(
permutation group. The orthogonal projectors A (antisymmetrization) and M (mixed) that
V
V
respectively project onto 3 V and (( 2 V ) V )mixed are
j

1
(At)ij k = (tij k + tj ki + tkij ),
3
1
(Mt)ij k = (2tij k tj ki tkij ).
3

(8.19)
(8.20)

A detailed analysis (see below) of Eqs. (8.15)(8.18) shows that they determine An0 , An00
and M(n0 + n00 ). These equations do not provide information about M(n0 n00 ).
To solve the equations above it is best on introduce the following auxiliary tensors:
Vij k = Hij k (flij nlk + flj k nli + flki nlj ),

eij k + flij nlk + flj k nli + flki nlj ,
eij k = H
V
Wij k = (fkij nlk flij ) = mkl flij ,

eij k = fkij + nlk flij = flij mlk .
W

(8.21)
(8.22)
(8.23)
(8.24)

e are totally
They are all skew symmetric under the interchange i j and V , V
antisymmetric. Given a value for nij , these tensor are determined by the geometric data
which specifies the sigma models. This data is not independent because these tensors are
linearly related due to the right hand sides of (8.15)(8.18).

O. Alvarez / Nuclear Physics B 584 [PM] (2000) 659681

677

A little algebra shows that


e +V
e.
n0 + n00 = W V = W

(8.25)

All the content of (8.15)(8.18) is contained in (8.15), (8.16) and (8.25). These equations
e, W, W
e . Immediate conclusions are that
place constraints on V , V
e =V +V
e,
W +W
e ) = 0,
M(W + W
1e
2
,
AW = V + V
3
3
e.
e = 1V + 2V
AW
3
3

(8.26)
(8.27)
(8.28)
(8.29)

In deriving the last two equation we used (8.15), (8.16) and applied the A operator to (8.25).
The equations above imply linear algebraic relations among the data that defines the sigma
models. They tell us that there exists a tensor Uij k of mixed symmetry, i.e., Uij k = Uj ik
and AU = 0 such that
1e
2
,
W = +U + V + V
3
3
e.
e = U + 1 V + 2 V
W
3
3

(8.30)
(8.31)

Collating all our information we can now write down the 13 n(n 1)(2n 1) first order
linear PDEs that determine nij :
1
An0 = V ,
3
1e
00
An = + V
,
3
M(n0 + n00 ) = +U.

(8.32)
(8.33)
(8.34)

There is no equation for M(n0 n00 ). It is worthwhile to note that


1
1
n0 = U + M(n0 n00 )
2
2
1
1
n00 = U M(n0 n00 ) +
2
2

1
V,
3
1e
V.
3

(8.35)
(8.36)

You can envision using this formalism in four basic scenarios.


e g,
e are dual to each other. This
1. Test to see if two sigma models (M, g, B) and (M,
B)
entails the construction of the symplectic manifold P .
2. Given a sigma model (M, g, B) and a symplectic manifold P , naturally associated
e g,
e
with M, can you construct the dual sigma model (M,
B)?
3. Given a symplectic manifold P that admits a bifibration, attempt to construct dual
sigma models.
4. Find all symplectic manifolds P that admit dual sigma models.

678

O. Alvarez / Nuclear Physics B 584 [PM] (2000) 659681

8.2. Covariantly constant nij


Here we show that the assumption of covariantly constant nij leads to a flat connection
on P . Assume that in our adapted coframes the nij are covariantly constant with respect
to the connection, i.e., n0ij k = n00ij k = 0. In this case it is immediate from (8.17) and
e = 0.
(8.18) that fij k = fij k = 0. Subsequently we see from (8.15) and (8.16) that H = H

From (8.10) we see that Tij00kl = 0 and thus the curvature vanishes, ij = 0. We are mostly
interested in local properties so we might as well assume P is parallelizable. We can use
parallel transport with respect to this connection to get a global framing. In this special
framing the connection coefficients vanish and thus we can make the substitution ij =
0 in all the equations in Section 8.1. Note that the orthonormal coframes satisfy d i =
e are manifolds with cover Rn . Following up on remarks made
d i = 0 and thus M and M
in Section 7.1 we see that this is the case of abelian duality but with constant nij in the
eij .
adapted frames corresponding to constant Bij and B
8.3. Case of fij k = 0

e has cover Rn ? Note that by (8.5)


What is the most general manifold M whose dual M
00
0
we have that Teij k = 0 and thus Tij lm = Teij ml = 0. This means that the curvature (6.14) of
the connection ij vanishes. Again using the remarks just made we can choose a parallel
framing such that ij = 0. Since fij k = 0 we have that d i = 0 and thus locally there
exists functions x i such that i = dx i . We also have that d i = 12 fij k j k . Previous
arguments also tell us that fij k is the pullback of a tensor on M. From (8.22) we see that
e=H
e and from (8.24) we have that W
e = 0. Eq. (8.31) tells us that U = 0 and V =
V
e
e
2V = 2H . Inserting into (8.30) we find that
eij k .
Wij k = mkl flij = (kl + nkl )flij = H

(8.37)

e = 0 then fij k = 0 and M is also a


An elementary consequence of this equation is that if H
manifold with cover Rn . Inserting the above into (8.21) we find that
eij k = Hij k (fij k + fj ki + fkij ).
H

(8.38)

e and the right hand side is the pullback


The left hand side is the pullback of a tensor on M
eij k are constants.
of a tensor on M thus each side must be constant. We have learned that H
We assume that nij are not constant (see Section 8.2). From (8.37) we expect the fij k not
to be constant. Let ij k be a tensor of mixed symmetry then from (8.35) and (8.36) we see
that




2e
1e
m
(8.39)
dnkl = klm Hklm + klm + Hklm m .
3
3
e has cover
The answer to the question, What is the most general manifold M whose dual M
Rn ? and the construction of the duality transformation is given by the general solution to
the following system of exterior differential equations:
1e i
j
(kl + nkl ) d l = H
kij ,
2

(8.40)

O. Alvarez / Nuclear Physics B 584 [PM] (2000) 659681

d i = 0,




2e
1e
m
dnkl = klm Hklm + klm + Hklm m .
3
3

679

(8.41)
(8.42)

e = 0. From (8.37) we have that fij k = 0.


As an example consider the special case of H
We are now asking, What is the most general duality transformation between manifolds
with cover Rn ? The equations above tell us that there exists functions x i and x j such that
i = dx i and j = dx j . Equation (8.42) becomes dnkl = klm dy m where y i = x i x i .
We learn that nij is a function of y only. Since the tensor has mixed symmetry we see
that d(nij (y) dy i dy j ) = 0 and thus we conclude that locally there exists functions ri of
the independent variables y j such that

1
nij (y) dy i dy j = d ri (y) dy i .
(8.43)
2
We now have all the information required to construct the duality transformation. The
duality transformations are given by
+ x = + x ,
[ x ]
x = T (x x)
with T (y) = (I + n(y))(I n(y))1 . By taking the sum and difference of the equations
above one gets ODEs that can be solved for (x(
), (
)) given (x( ), ( )).

Acknowledgements
I would like to thank O. Babelon, L. Baulieu, T. Curtright, L.A. Ferreira, D. Freed,
S. Kaliman, C.-H. Liu, R. Nepomechie, N. Reshetikhin, J. Snchez Guilln, N. Wallach
and P. Windey for discussions on a variety of topics. I would also like to thank Jack Lee for
his M ATHEMATICA package R ICCI that was used to perform some of the computations.
I am particularly thankful to R. Bryant and I.M. Singer for patiently answering my many
questions about differential geometry. This work was supported in part by National Science
Foundation grant PHY-9870101.

Appendix A. Facts about orthogonal groups


We collate some basic properties of orthogonal groups in this section. An orthogonal
matrix in O(n, n) may be written in terms of n n blocks as


A B
,
C D
where At A + C t C = I , B t B + D t D = I and B t A + D t C = 0. A matrix in the Lie algebra
so(n, n) may be written as


a b
,
c d

680

O. Alvarez / Nuclear Physics B 584 [PM] (2000) 659681

where a = a t , d = d t and c = bt .
The group OQ (n, n) GL(2n, R) is defined to be the linear transformations, which
leave
 0 I 
n
Q=
In 0
invariant. A matrix in OQ (n, n) may be written in n n blocks as


W X
,
(A.1)
Y Z
where W t Z + Y t X = I , W t Y + Y t W = 0 and Xt Z + Z t X = 0. It is very important to
observe that matrices of the form


I 0
,
(A.2)
Y I
where Y = Y t are in OQ (n, n). A matrix in the Lie algebra soQ(n, n) may be written as


w x
,
(A.3)
y z
where z = wt , y = y t and x = x t .
Conjugating by the orthogonal matrix


1 In In
T=
2 In In
leads one to the observation that




In 0
0 In
1
T =
.
T
In 0
0
In

(A.4)

Therefore, the group OQ (n, n) is isomorphic to O(n, n).


We will also need to identify the compact group O(2n) OQ (n, n). To do this we observe
that at the Lie algebra level so(2n) soQ(n, n) is given by matrices of the form


a b
,
b a
where a t = a and bt = b. Conjugating by T one sees that




ab
0
a b
.
T
T 1 =
0
a+b
b a
Therefore, the intersection so(2n) soQ(n, n) is conjugate to so(n) so(n).
Appendix B. On the structure functions
In a local orthonormal coframe one has the Cartan structural equations (5.1). Locally
one can always write d i = 12 fij k j k for some structure functions fij k which are
skew symmetric in j k. If we write the riemannian connection as ij = ij k k then
1
ij k = (fkij fij k + fj ik ).
2

(B.1)

O. Alvarez / Nuclear Physics B 584 [PM] (2000) 659681

681

This allows us to reconstruct all the local Riemannian geometry of the manifold in terms
of the structure functions.
References
[1]
[2]
[3]
[4]
[5]
[6]
[7]
[8]
[9]
[10]
[11]
[12]
[13]
[14]
[15]
[16]
[17]
[18]
[19]
[20]
[21]
[22]
[23]
[24]
[25]
[26]
[27]
[28]
[29]
[30]
[31]

[32]
[33]
[34]

O. Alvarez, Nucl. Phys. B 584 (2000) 682, next article in this issue.
A. Giveon, M. Porrati, E. Rabinovici, Phys. Rep. 244 (1994) 77202, hep-th/9401139.
A. Giveon, E. Rabinovici, G. Veneziano, Nucl. Phys. B 322 (1989) 167.
V.E. Zakharov, A.V. Mikhailov, Sov. Phys. JETP 47 (1978) 10171027.
C.R. Nappi, Phys. Rev. D 21 (1980) 418.
B.E. Fridling, A. Jevicki, Phys. Lett. B 134 (1984) 70.
E.S. Fradkin, A.A. Tseytlin, Ann. Phys. 162 (1985) 31.
E.B. Kiritsis, Mod. Phys. Lett. A 6 (1991) 28712880.
M. Rocek, E. Verlinde, Nucl. Phys. B 373 (1992) 630646, hep-th/9110053.
A. Giveon, M. Rocek, Nucl. Phys. B 380 (1992) 128146, hep-th/9112070.
X.C. de la Ossa, F. Quevedo, Nucl. Phys. B 403 (1993) 377394, hep-th/9210021.
M. Gasperini, R. Ricci, G. Veneziano, Phys. Lett. B 319 (1993) 438444, hep-th/9308112.
A. Giveon, M. Rocek, hep-th/9406178.
A. Giveon, M. Rocek, Nucl. Phys. B 421 (1994) 173190, hep-th/9308154.
A. Giveon, E. Kiritsis, Nucl. Phys. B 411 (1994) 487508, hep-th/9303016.
T. Curtright, C. Zachos, Phys. Rev. D 49 (1994) 54085421, hep-th/9401006.
E. Alvarez, L. Alvarez-Gaume, Y. Lozano, Phys. Lett. B 336 (1994) 183189, hep-th/9406206.
E. Alvarez, L. Alvarez-Gaume, Y. Lozano, Nucl. Phys. B 424 (1994) 155183, hep-th/9403155.
E. Alvarez, L. Alvarez-Gaume, J.L.F. Barbon, Y. Lozano, Nucl. Phys. B 415 (1994) 71100,
hep-th/9309039.
Y. Lozano, Phys. Lett. B 355 (1995) 165170, hep-th/9503045.
Y. Lozano, Mod. Phys. Lett. A 11 (1996) 28932914, hep-th/9610024.
O. Alvarez, C.H. Liu, Commun. Math. Phys. 179 (1996) 185, hep-th/9503226.
C. Klimcik, P. Severa, Phys. Lett. B 351 (1995) 455462, hep-th/9502122.
C. Klimcik, P. Severa, Phys. Lett. B 372 (1996) 6571, hep-th/9512040.
C. Klimcik, P. Severa, Phys. Lett. B 376 (1996) 8289, hep-th/9512124.
V.I. Drinfeld, Dokl. Akad. Nauk SSSR 268 (2) (1983) 798820.
A.Y. Alekseev, A.Z. Malkin, Commun. Math. Phys. 162 (1994) 147174, hep-th/9303038.
K. Sfetsos, Nucl. Phys. B 517 (1998) 549566, hep-th/9710163.
K. Sfetsos, Nucl. Phys. Proc. Suppl. B 56 (1997) 302, hep-th/9611199.
A. Stern, Phys. Lett. B 450 (1999) 141, hep-th/9811256.
O. Alvarez, in: L. Baulieu, V. Kozakov, M. Picco, P. Windey (Eds.), Low Dimensional
Applications of Quantum Field Theory, Plenum Press, 1997, pp. 118; hep-th/9511024,
Lectures presented at the Cargse Summer School in July 1995.
C. Klimcik, P. Severa, Phys. Lett. B 381 (1996) 5661, hep-th/9602162.
K. Sfetsos, Nucl. Phys. B 561 (1999) 316, hep-th/9904188.
D. Alekseevsky, J. Grabowski, G. Marmo, P.W. Michor, J. Geom. Phys. 26 (1998) 340379,
math.DG/9801028.

Nuclear Physics B 584 [PM] (2000) 682704


www.elsevier.nl/locate/npe

Target space duality II: applications


Orlando Alvarez 1
Department of Physics, University of Miami, P.O. Box 248046, Coral Gables, FL 33124, USA
Received 10 April 2000; accepted 22 May 2000

Abstract
We apply the framework developed in Target space duality I: general theory. We show that both
nonabelian duality and PoissonLie duality are examples of the general theory. We propose how
the formalism leads to a systematic study of duality by studying few scenarios that lead to open
questions in the theory of Lie algebras. We present evidence that there are probably new examples of
irreducible target space duality. 2000 Elsevier Science B.V. All rights reserved.
PACS: 11.25.-w; 03.50.-z; 02.40.-k
Keywords: Duality; Strings; Geometry

1. Introduction
In this article we consider some applications of the general theory derived in article I [1].
We show that nonabelian duality [211] and PoissonLie duality [1215] are special cases
of the general theory. In fact we show that nonabelian duality is a special case of a more
general situation. The spirit of this paper is that there are natural geometric scenarios
that need to be explored. We explore a few of the easier ones and see that they lead to
open mathematical questions in the theory of Lie algebras. For example, Lie bialgebras
and generalizations, and R-matrices naturally appear in this framework. References to
equations and sections in article I are preceded by I, e.g., (I-8.3).

2. Examples with a flat connection


2.1. General remarks
To best understand how to use the equations given in the Section I-8.1 it is best to do
a few examples. The examples we will consider in Section 2 assume that the connection

This work was supported in part by National Science Foundation grant PHY-9870101.

1 oalvarez@miami.edu

0550-3213/00/$ see front matter 2000 Elsevier Science B.V. All rights reserved.
PII: S 0 5 5 0 - 3 2 1 3 ( 0 0 ) 0 0 3 1 3 - 8

O. Alvarez / Nuclear Physics B 584 [PM] (2000) 682704

683

ij is flat. We are mostly interested in local properties so we might as well assume P is


parallelizable. We can use parallel transport with respect to this connection to get a global
framing. In this special framing the connection coefficients vanish and thus we can make
the substitution ij = 0 in all the equations in Section I-8.1. A previous remark tells us that
e Consequently
in this case fij k and fij k are pullbacks respectively of tensors on M and M.
0
l
00
l
00
0
we have that dfij k = fij kl and dfij k = fij kl , i.e., fij kl = fij kl = 0. Because we are
e are
interested in mostly local considerations we might as well assume both M and M
parallelizable.
e
2.2. The tensor nij is the pullback of a tensor on M
Here we assume that the connection is flat. In Section I-8.2 we considered what
happened if nij was covariantly constant in this section we relax this condition to one
where nij only depends on a natural subset of the variables. We assume that nij is the
e This means that
pullback of a tensor on M.
dnij = n00ij k k

or equivalently n0ij k = 0.

(2.1)

e We
An equivalent formulation is that 12 nij i j is the pullback of a 2-form on M.
immediately see from (I-8.17) and (I-8.18) that
n00ij k = mkl flij ,
n00kij

n00kj i

(2.2)

= mj l flki mil flkj = flij mlk .

(2.3)

A brief computation shows that d(mj i j ) = 0 and therefore we can locally find n functions
p1 , . . . , pn such that
dpi = mj i j .

(2.4)

Note that since mij is invertible the differentials {dp1 , . . . , dpn } are linearly independent.
e and they
The functions p1 , . . . , pn are the pullbacks of functions locally defined on M
e We immediately see that
define a local coordinate system on M.
dnij = fkij dpk .

(2.5)
0 l
fkij
l

0
fkij
l

=
dpk immediately tells us that
=0
The integrability condition 0 =
0
00
since {1, . . . , n , dp1 , . . . , dpn } are linearly independent. Since fij kl = fij kl = 0 we see
that fij k are constants and thus M is diffeomorphic to a Lie group G. The i are the
pullbacks by of the left invariant MaurerCartan forms on G. The next question is
whether the pullback metric i i is invariant under the action of the group. Choose a G
left-invariant vector field X along M, i.e., X i = 0. A brief computation shows that
d2 nij

LX ( i i ) = Xk fij k ( i j + j i ).

(2.6)

The metric on M is G-invariant if and only if fij k = fj ik , i.e., the structure constants are
totally antisymmetric, a standard result.
We can integrate (2.5) to obtain
nij = n0ij + fkij pk ,

(2.7)

684

O. Alvarez / Nuclear Physics B 584 [PM] (2000) 682704

where n0ij are constants. Define the constant tensor m0 by m0ij = ij + n0ij . Eq. (2.3)
immediately gives us an expression for fij k in terms of fij k and pl . It is a straightforward
computation to verify that fij k satisfies the integrability conditions for d l = 12 flij i
j . In general fij k are not constants and are thus not generically MaurerCartan forms for
a Lie group.
To determine H we use (I-8.15). Note that our hypotheses imply that the left hand side
is automatically zero. After using the Jacobi identity satisfied by the fij k we find that
Hij k = flij n0lk + flj k n0li + flki n0lj .

(2.8)

If we define a left invariant 2-form by n0 = 12 n0ij i j then the above equation is H =


e:
dn0 which tells us that H is cohomologically trivial. We use (I-8.16) to determine H

eij k = fij k + fj ki + fkij + Hij k (fij k + fj ki + fkij ).
(2.9)
H
2.2.1. Cotangent bundle duality
We will see in this section that cotangent bundle duality is a special case of what we
have been discussing in Section 2.2. In particular this corresponds to what is often called
nonabelian duality. The cotangent bundle has a natural symplectic structure and thus we
automatically have a candidate symplectic manifold P for free. Assume the manifold M
is parallelizable. This means that the cotangent bundle T M is trivial, i.e., it is a product
e will be taken to be the Cartesian
space, P = T M = M Rn . The projections and
e = Rn . If i is an orthonormal coframe on M then the canonical
projections and therefore M
1-form on T M may be written as = pi i . The pi are coordinates along the fibers of .
e (diffeomorphic to M) given by dpi = 0, i = 1, . . . , n, are the same as the
The fibers of
fibers given by i = 0, i = 1, . . . , n, therefore, there must exist an invertible matrix defined
by functions m
bij such that
bij i .
dpj = m

(2.10)

The canonical symplectic form is given by = d = dpj j 12 pk fkij i j . In going


from to the term is not changed and so we immediately learn that m
bij = mij =
ij + nij . Taking the exterior derivative of (2.10) leads to n0ij k = 0 and (2.3). Thus nij
is the pullback of a tensor on Rn and we are back to our general discussion given in
Section 2.2. This is an example of what is often called nonabelian duality which has
generating function given by (I-2.2) with = pk k . We know that M is a Lie group G,
e = g . The metric on g is immediately computable from (2.10)
T G = G g and thus M
e = g = Rn there is an
since mij = ij + n0ij + pk fkij . It is worth remarking that because M
e which does not leave the metric invariant. Said differently
abelian Lie group action on M
the dp are the MaurerCartan forms for the abelian Lie group g . This statement is made
in anticipation of our discussion about PoissonLie duality in Section 3.
Nonabelian duality usually refers to the case with n0ij = 0 where (2.8) tells us that
e from (2.9) directly or it was already remarked in Section I-7.2
Hij k = 0. We can compute H
e
that Bij is easily determined.

O. Alvarez / Nuclear Physics B 584 [PM] (2000) 682704

685

2.2.2. Cotangent bundle duality and gauge invariance


In this section we revisit cotangent bundle duality and try to understand geometrically
the role of the B field gauge transformation. We assume the manifold M has a trivial
cotangent bundle T M = M Rn with canonical 1-form = pi i and symplectic form
= d. We modify the discussion of Section 2.2.1 by demanding that the bifibration not
be given by the Cartesian projections. We want the vertical fibers to be original ones so we
e will
e : T M M
have : (x, p) T M 7 x M. On the other hand the projection
not be the canonical projection. The fibers of this projection are slanted relative to the
e are the integral manifolds
ec . Note that the fibers of
fibers of the Cartesian projection
i
of the Pfaffian equations = 0. From general principles we know that
bj i j + uj i j
dpi = m

(2.11)

e are slanted
for some functions m
b, u. Geometrically this is the statement that the fibers of
e
relative to the fibers of c . As before the structure of the symplectic form leads to the result
that m
bij = mij . The integrability condition d2 pi = 0 leads to the following equations
n00ij k n00ikj = flj k mli ,

(2.12)

u0j ik

= flj k uli ,

(2.13)

= n0kij .

(2.14)

u0kij
u00j ik

Comparing (2.12) with (I-8.18) we see that n0ij k = 0 and thus nij is the pullback of a tensor
e The general discussion of Section 2.2 tells us that there exists function p i in T M
on M.
such that dp i = mj i j . Geometrically this is just the statement that the fibers is given by
p i = constant. Note that (2.14) tells us that u00ij k = 0 and thus uij is the pullback of a tensor
on M. Eq. (2.11) tells us that d(pi p i ) = uj i j and thus we see that pi = p i + ki where
ki are pullbacks of functionsR on M, i.e., ki = ki (x). Thus we we see that theRcanonical
transformation generated by pi i is gauge equivalent to the one generated by p i i and
corresponds to a different choice of fibration.
e naturally be a Lie group?
2.2.3. Can M
Is it possible for the dual manifold to be naturally a Lie group under the assumptions
underlying the discussions in Section 2.2? By naturally we mean that the i are the
e Solving (2.3) for fij k generally leads to a nonMaurerCartan forms for a Lie group G.
e g,
e
constant solution. There is a possibility that the solution may be constant, i.e., (M,
B)
is naturally a Lie group. Unfortunately, we will see that both g and g must be abelian and
so there are no new interesting examples. Inserting (2.7) into (2.3) and using the linear
independence of the dpi leads to two equations:
m0j l f l ki m0il f l kj = fl ij m0lk ,
f m il f l j k = f m il fl j k .

(2.15)
(2.16)

To obtain the latter equation we used the Jacobi identity satisfied by fij k . The indices
have also been set at their natural (co)variances for future convenience. The fi j k satisfy
the Jacobi identity if they satisfy the two equations above because of our remark about

686

O. Alvarez / Nuclear Physics B 584 [PM] (2000) 682704

integrability after (2.7). You should think of M as a Lie group G with Lie algebra g
e Let adX (Y ) = [X, Y ] be the adjoint action of g. Let [, ] be the
and similarly for M.
e eY
e
e e
Lie bracket on g with adjoint action denoted by ad
X = [X, Y ] . The reduction of the
structure group of P to O(n) meant that we could identify the horizontal tangent space
with the vertical tangent space and thus we can identify g with g . We should think of a
single vector space V with two Lie brackets giving us two Lie algebras g = (V , [, ]) and
g = (V , [, ] ). In this notation (2.16) becomes
e Y ) = 0.
adX (adY ad

(2.17)

There are some immediate important consequences of this equation. Let d be the vector
e Y Z for all Y, Z. The subspace d is contained in the
subspace of g spanned by adY Z ad
e then the center z is a nontrivial abelian ideal
center z of g by (2.17). If d 6= 0, i.e., ad 6= ad,
e and
in the Lie algebra g and thus g is not semisimple. If d = 0 then we have that ad = ad
we will show that g is abelian when we incorporate (2.15) into our reasoning. Note that if
z = 0 then d = 0.
e i.e., d = 0. Eq. (2.15) may be rewritten as
Let us study the implications of ad = ad,

(2.18)
2fkij = (fij k + fj ki + fkij ) + n0il flj k + n0j l flki + n0kl flij .
Notice that the right hand side is totally antisymmetric under permutations of i, j, k. This
means that fkij is totally antisymmetric and thus the metric i i is a bi-invariant
positive definite metric on G. The proposition proven in Appendix A tells us that there
is a decomposition into ideals g = k a where k is a compact semisimple Lie algebra and
a is an abelian Lie algebra. The next part of the argument only involves the compact ideal
k and without any loss we assume a = 0 for the moment. Using the antisymmetry of the
structure constants the above may be rewritten as
fij k = n0il flj k + n0j l flki + n0kl flij .
Using the remark made immediately after (2.8) we see that 3!1 fij k i j k = dn0 .
The closed 3-form 3!1 fij k i j k is exact and this contradicts H 3 (k) 6= 0 as discussed
e then we conclude that g = a is abelian and we are back in
in Appendix A. If ad = ad
familiar territory.
The next we consider the general case by exploiting the observation that d z. Since
we have a positive definite metric on g there is an orthogonal direct sum decomposition
g = z z . The orthogonal complement z is not generally an ideal because the metric
is not necessarily (ad g)-invariant. Nevertheless we can choose an orthonormal basis {e }
for z and an orthonormal basis {ea } for z . Greek indices are associated with z, indices in
{a, b, c, d} are associated with z ; and indices in {i, j, k, l} run from 1, . . . , dim g and are
associated with all of g. The Lie algebra g is given by
[e , ej ] = 0,
[ea , eb ] = f

(2.19)
ab ec

+f

ab e .

(2.20)

Note that h = g/z is a Lie algebra since z is an ideal. It follows that the structure constants
of h are f c ab . Also, g is a central extension of h with extension cocycle f ab . It is well

O. Alvarez / Nuclear Physics B 584 [PM] (2000) 682704

687

known that the cocycle is trivial if f ab = t c f c ab corresponding to a redefinition of the


basis given by ea ea + t a e . The Lie algebra g must have the form below because the
two Lie brackets are the same modulo the center z:
[e , ej ] = f j e ,

[ea , eb ] = f

ab ec

(2.21)

+f

ab e .

(2.22)

In Eq. (2.15) choose k = then the left hand side vanishes and the equation becomes
0 = fl ij m0l = fd ij m0d + f ij m0 . Using the different choices for (i, j ) we find

= 0,
f j = fd j m0d (m0 )1


d
0
0
1
.
(2.23)
f ab = f ab md (m )
These equations tell us that g is a central extension of h with a trivial cocycle. Next choose
i = in (2.15) with result 0 = m0l f l kj . Choosing k = a and j = b leads to

(2.24)
f ab = (m0 )1 m0c f c ab = 0
which tells us that the cocycle f ab is also trivial. Finally choose (i, j, k) = (a, b, c) in
(2.15) and substitute (2.23) and (2.24) for f ab and f ab to obtain




m0bd m0b (m0 )1 m0d f d ca m0ad m0a (m0 )1 m0d f d cb


= f d ab m0dc m0d (m0 )1 m0 c .
(2.25)
Next we observe that since m0 = + n we conclude that ((m0 )1 ) = s + a
where s is symmetric and positive definite and a is skew. In particular note that
m0b ((m0 )1 ) m0d = s m0 b m0d + a m0 b m0d where the first term is symmetric and
positive definite and the second is skew. Using this we immediately see that

m0bd m0b (m0 )1 m0d = Sbd + Nbd
where Sbd is symmetric and positive definite and Nbd is skew. We can use Sab as a second
metric on g and use it to raise and lower the indices in (2.25) leading to

(2.26)
2fcab = (fabc + fbca + fcab ) + Nad f d bc + Nbd f d ca + Ncd f d ab .
This tells us that fabc is totally antisymmetric and therefore Sab is an invariant metric on h.
The same chain of arguments used after (2.18) tell us that h is abelian which implies that
f ab = 0 and f ab = 0 concluding that g and g are abelian Lie algebras.
2.2.4. Connection with R-matrices
Eq. (2.15) is closely related to the theory of R-matrices developed by Semenov-TianShansky [16] which is different than the one developed by Drinfeld [17]. The discussion
here suggests that PoissonLie groups may play a role in duality, see Section 3. We observe
that (2.15) may be rewritten as
0 1
0 1
0 1
0
(m0 )1
j l fkil (m )kl fj il = (m )j l (m )km fnlm mni .

(2.27)

Next we show that this equation describes a potential double Lie algebra structure on g
la Semenov-Tian-Shansky (B.11). Assume the Lie algebra g admits an invariant metric K.

688

O. Alvarez / Nuclear Physics B 584 [PM] (2000) 682704

For example if the Lie algebra is semisimple then K may be taken to be the Killing metric.
Let us use the indices a, b, c, d, e to denote generic components in a generic basis. In
terms of a basis (e1 , . . . , en ) the structure constants are given by [ea , eb ] = f c ab ec . For the
moment it is best to forget about the orthonormal basis we were previously using before
because the metric K may not be related to the previous metric. The components of the
invariant metric are given by Kab = K(ea , eb ). We will use K to identify g with g , i.e.,
raise and lower indices. The tensor with components K ab is the inverse of the invariant
metric, i.e., the induced metric on g . The statement that K is g-invariant is equivalent to
fabc being totally antisymmetric. With these assumptions equation (B.13) may be rewritten
as



0
0
0
(2.28)
R d e K ea f b cd R d e K eb f a cd = K aa K bb Kcc0 f c a 0 b0 R .
The indices of R and f have their natural (co)variances.
We are now going to compare (2.28) with (2.27). Remember that (2.27) is valid in an
orthonormal basis with respect to a specific metric on g which may not be related to the
invariant metric K. What we will have to do is express (2.28) in the orthonormal basis
but this will be simple because we adjusted all the (co)variances correctly in the equation
above. The indices i, j, k, l, m, n refer to our orthonormal basis. We raise and lower indices
using the Kronecker delta tensor. The only potential confusion is that we have to be careful
and remember that Kab becomes Kij and the inverse invariant metric K ab becomes Kij1 .
The correspondence is made by choosing R (in our orthonormal basis) to be defined by
1
(m0 )1
ij = Rj l Kli .

(2.29)

The R-matrices we are considering are invertible because both m0 and K are always
invertible. This is different that in the Drinfeld case where the R-matrix is skew adjoint
and may not be invertible. If you think of K as a map K : g g then the equation above
is (m0 )t = KR 1 : g g which suggests that m0 should be interpreted as a map m0 :
g g. Comparing the right hand sides of (2.27) and (2.28) we see that
film Rlj Rmk = Ril (flj k )R .

(2.30)

If Rij satisfies the modified YangBaxter equation (B.10) then (f i j k )R are the structure
constants of a Lie algebra gR associated with a Lie group GR . Let (1 , . . . , n ) be the left
invariant MaurerCartan forms for GR :
dl = 12 (flj k )R j k .
Going to a different basis given by i = Ril l we see that
di = 12 film l m .
We conclude that if R is a solution of the modified classical YangBaxter equation then we
e
can construct the Lie algebra g with structure constant fij k and an associated Lie group G.
In our situation we also have to impose a second equation (2.16) and this is very
restrictive. Our in-depth analysis that eliminated all except the abelian case. Logically,

O. Alvarez / Nuclear Physics B 584 [PM] (2000) 682704

689

there is the possibility of non-trivial solutions by the use of R-matrices. The reason is that
using (2.30) and (B.11) we can rewrite (2.16) as

0 = f m il f l j k R j p R k q f r nq R n p R l r f r pn R n q R l r .
The stuff between the parentheses is the left hand side of modified classical YangBaxter
equation (B.12). If R satisfies the modified classical YangBaxter equation the above
becomes 0 = cf m il f l pq . The solution to this equation is c = 0 or adX adY = 0. We
know from our general analysis that it is unnecessary to proceed along these lines.
2.3. Symmetric duality
In the previous sections we had = pi i for some functions pi on P . A consequence
was that the pi were good coordinates on the fibers of and these fibers were also
Lagrangean submanifolds of . There was a certain asymmetry in the way fibers of
e were treated. In this section we consider a more symmetric situation. To motivate
and of
the ensuing presentation let us go to the case of M = Rn and P = T M where we have the
symplectic form = dpi dq i . Up to constant 1-forms the most general antiderivative is
= (1 u)pi dq i uq i dpi where u R is a constant.
Let U P be a neighborhood, we can always write
= qi i q i i ,

(2.31)

where q i and qi are local functions on U . We now make some special assumptions about
e such that
the functions q and q.
Assume there exist matrix valued functions E and E
dq i = Eij j ,
eij .
dqj = i E

(2.32)
(2.33)

These functions are not arbitrary and must satisfy a variety of constraints. For example, the
e are invertible then the
ranks are constrained by Eq. (2.41) below. For example, if E and E
1
functions (q, q)
are independent in the sense that the map (q , . . . , q n , q1 , . . . , qn ) : U
R2n is of rank 2n. Consequently the fibers of are locally described by (q = constant)
e will not be invertible. In this
e by (q = constant). In general, E or E
and the fibers of
case not all the functions will be independent and (q = constant) will define a family of
e
manifolds each diffeomorphic to M.
As before we use the prime and double prime notation to denote derivatives in the
appropriate directions. The equation d2 q i = 0 tells us that
Eij00 k = 0,
Eij0 k

0
Eikj

= Eil flj k .

(2.34)
(2.35)

Likewise the equation d2 q = 0 tells us that


eij0 k = 0,
E

(2.36)

00
e00
e
ekj
E
l Elj k = fikl Eij .

(2.37)

690

O. Alvarez / Nuclear Physics B 584 [PM] (2000) 682704

Note that you can solve (2.35) for Eij0 k as a function of Eil flj k and you can solve (2.37)
eij .
e00 and as function of fikl E
for E
ij k

Next we compute = d:

eij + Eij i j 1 qi fij k j k + 1 q i fij k j k .
= E
2
2

(2.38)

e are Lagrangean submanifolds of . This is very


In general, neither the fibers of or of
different from the cotangent bundle cases previously discussed. Next we make we write E
e as
and E
Eij = ij + ij ,
eij = ij + ij ,
E

(2.39)
(2.40)

where and are symmetric, and and are antisymmetric. Comparing to (I-6.5) we see
that
ij = ij + ij

and nij = ij + ij .

(2.41)

e cannot be arbitrary and must satisfy conditions given


The matrix valued functions E and E
0
e0 = E 0 = 0 + 0 . Since nij =
by the above. We remark that nij k = m0ij k = Eij0 k + E
ij k
ij k
ij k
ij k
nj i we immediately see that
ij0 k = 0 and Eij0 k = n0ij k = ij0 k .

(2.42)

Also Eij00 k = 0 and thus we see that ij00 k = 0 and ij00 k = 0. This means that ij is constant.
e00 = n00 = 00 and 0 = 0.
Similarly we conclude that ij is constant, E
ij k
ij k
ij k
ij k
We can take the results given above and insert into (I-8.18) and (I-8.17) to obtain
Eij0 k = n0ij k = ij0 k = flij Elk ,
ekl flij .
eij00 k = n00ij k = ij00 k = +E
E

(2.43)
(2.44)

We now insert the above into (2.35) and (2.37) to obtain the basic equations
ej l flik = flj k E
eli ,
ekl flij E
E
flij Elk flik Elj = Eil flj k .

(2.45)
(2.46)

Note that
dEij = Eij0 k k = flij Elk k = flij dq l ,

(2.47)

eij00 k k
eij = E
dE

(2.48)

ekl flij = flij dq ,


=E
k

0 k and df = f00 k .
where we used (2.32) and (2.33). We know that dflij = flij
lij
k
lij k
Therefore, by taking the exterior derivatives of (2.47) and (2.48) we learn that

eml (dflij ) = 0,
E

dflij Elm = 0.

(2.49)
(2.50)

To make progress we have to make some assumptions. The simplest assumption is that
Eij = 0. In this case we are back to the discussion given in Section 2.2.1. In this paragraph

O. Alvarez / Nuclear Physics B 584 [PM] (2000) 682704

691

e g . You can
we use the notation that a Lie group G via nonabelian duality is dual to G
generalize cotangent bundle duality along the following lines. Consider matrices with with
the same 2 2 block form



 (1)
0
0
0
E
e
,
E=
E=
e(2) .
0
0
0 E
e = G(1) G
e(2) .
e(1) G(2) and M
This will lead to a manifold M = G
e
Next we look at the case where Eij is invertible. Eq. (2.49) tells us that fij k are constants
and thus M is naturally a Lie group G with structure constants fij k . We can immediately
integrate (2.48) obtaining
eij = ij + 0 + fkij q k ,
E
ij

(2.51)

where 0 is a constant tensor. With this information we can use (2.45) to determine fij k .
It is straightforward to verify that fij k satisfy the integrability conditions for (I-8.2) with
ij = 0. All we have to do is to find a tensor Eij that satisfies (2.46) and (2.47). Note that
(2.41) tells us that ij = ij ij . This case merits further analysis.
eij being invertible we also impose that Eij is invertible then fij k are constant
If besides E
e is naturally a Lie group G.
e We can integrate (2.47) to obtain
(see (2.49)) and M
Eij = ij + ij0 + flij q l ,

(2.52)

where 0 is a constant tensor. It is convenient to define


E0 = + 0,
e0 = + 0 .
E

(2.53)
(2.54)

We can insert (2.51) and (2.52) into (2.45) and (2.46) and expand both sides in powers of
q and q to obtain:
e0 ,
e0 f l kj = fl ij E
e0 f l ki E
E
jl
il
lk
f

0
ki Elj

il f j k
fl kj Eli0

=f

il f j k ,
0 l
= Ekl f ij ,
fm il fl j k = fm il f l j k ,

(2.55)
(2.56)
(2.57)
(2.58)

where the Jacobi identity was used to simplify the above. We are now in a situation very
similar to that in Section 2.2.3. The difference is that the relevant metrics are now ij and
ij . The difficulty arises in that we have lost positive definiteness of the metrics. The only
constraint is that + = I . If either or is definite then an analysis along the lines of
Section 2.2.3 leads to the conclusion that g and g are abelian. If both are indefinite or both
are singular then the analysis previously provided breaks down. This situation also merits
further study.

3. PoissonLie duality
Here we discuss a beautiful example of scenario I-3 described at the end of Section I-8.1
where we are given a special symplectic bifibration and we have to construct the metrics

692

O. Alvarez / Nuclear Physics B 584 [PM] (2000) 682704

e The Drinfeld double Lie group is an example of


and antisymmetric tensors on M and M.
a special symplectic bifibration. The metrics and antisymmetric tensors constructed in this
manner correspond to the PoissonLie duality of Klimcik and Severa [12]. The explicit
duality transformation was obtained by Klimcik and Severa [12,13], and by Sfetsos [15].
We can use our general framework to determine both the metrics and antisymmetric tensors
by making educated guesses. The Drinfeld double GD is a Lie group whose Lie algebra
e with
gD is a Lie bialgebra, see Appendix B.1. The bifibration is by Lie groups G and G

respective Lie algebras g and g g . The Lie algebras are related by gD = g g = g g .


If {Ta } is a basis for g and {Tea } is the associated dual basis for g then
[Ta , Tb ] = C c ab Tc ,


ea bc Tc C b ac T
ec ,
Ta , Teb = C
 a b
ec ab Tec ,
Te , Te = C
ea bc are respectively the structure constants for G and G.
e The two
where C c ab and C
sets of structure constants must satisfy compatibility condition (B.4). To write down
the symplectic structure in a convenient way we introduce some notation slightly different than the one given in [12]. Let g G then the adjoint representation on g is
given by gTb g 1 = Ta a a b (g). One also has g Tea g 1 = a a b (g 1 )(Teb + bc (g)Tc ) where
e then g Teb g 1 = Tea a a b (g)
and gT
a g 1 =
ab = ba . Similarly, one has that if g G
ebc (g)
e is antisymmetric. It is worthwhile to note that if g =
Tec ), where
a a b (g 1 )(Tb +
aT
x
ab
c
ab
a
e
eab . Drinfeld shows that the bivecthen (g) = x Cc + O(x 2) and similarly for
e
ab
tor Ta Tb on G defines a Poisson bracket that is compatible with the group multiplication law [17,18]. A PoissonLie group is a Lie group with a Poisson structure which is compatible with the group operation. Thus we have that G is a PoissonLie group. Note that the
eb makes G
e a PoissonLie
eab Tea T
Poisson bivector is degenerate. Similarly the bivector
group. Klimcik and Severa discovered that the sigma model defined on the PoissonLie
e hence the name
group G is dual to the sigma model defined on the PoissonLie group G
PoissonLie duality. To exhibit the duality transformation we write down the symplectic
structure on GD . It was already noted by Klimcik and Severa [12,13] that the symplectic
structure on the Drinfeld double determines the canonical transformation. Here we see how
this fits inside our framework. We remark that having a symplectic structure on GD is not
sufficient. The symplectic structure has to be very special. We note that the Drinfeld double
is not a PoissonLie group in the Poisson structure associated with the symplectic structure
since the Poisson bivector would be nondegenerate. In the (perfect) Drinfeld double every
e
element k GD can uniquely be written as k = gu or k = vh where g, h G and u, v G.
The inverse function theorem shows that you can choose h and u as local coordinates on
GD near the identity. Let = h1 dh and = u1 du be respectively the left invariant
e The symplectic structure [19,20] may be written as
MaurerCartan forms on G and G.
2 = (du u1 )a (g 1 dg)a + (v 1 dv)a (dh h1 )a ,



e 1 )(h1 ) 1 b a
= a b + c cb (h1 ) I (u



ecb (u1 ) a I (h1 )(u
e 1 ) 1 b a .
+ b + c

(3.1)

O. Alvarez / Nuclear Physics B 584 [PM] (2000) 682704

693

Using we can construct the duality transformations. All we have to verify is that we get
e
metrics and antisymmetric tensor fields on G and G.
Klimcik and Severa show that the symmetric and antisymmetric parts of the rank two
e = t (F + )
e 1 are, respectively, the metrics
tensors E = t (F 1 + )1 and E
e The coefficients Fab are
and antisymmetric tensors for the dual sigma models on G and G.
constants. By an appropriate choice of basis for g one can always choose F = I + b where
b is antisymmetric. For pedagogical reasons we first look at the special case where b = 0
which is analogous to choosing n0 = 0 in (2.7). By looking at (3.1) we make an educated
e
guess and conjecture that our orthonormal bases should be given by in G and in G
where
= (I + )

e .
and = (I + )

(3.2)

We can verify that this is in agreement with the Klimcik and Severa data by noting that
e = t (I )
e . The symmetric parts on E and E
e in this basis
E = t (I ) and E
t
t

are respectively and and thus we see that we have orthonormal bases on G
e Likewise we see that the components of the antisymmetric tensors in this basis are
and G.
e because
e, respectively. Note that ( ) is well defined on G (G)
given by and
e
e
( ) is defined on G (G).
We are now ready to verify that there is a duality transformation. Postulate that frames
given by and are the orthonormal ones we need. Rewrite (3.1) in the orthonormal frame
where you find


e I
e 1 (I + ),
(3.3)
m= I


1
e ,
e I
e
I +
(3.4)
l = I
1
e I
e (I + ),
(3.5)
l = (I )
using the notation in (I-3.5). A brief computation shows that m = I + n where n is
antisymmetric and given by
n=

X


k=1

k=0

k

k 

X


e
e k.
e k

k=0

In this frame m is already in normal form and we can proceed. Note that n = n as
follows from (I-3.15). Using (I-3.8) and (I-3.9) we see that B = l m + I = and
e = l + m I =
e are quantities which
e. The important result here is that B and B
B
e and thus we have constructed the PoissonLie duality of
respectively live on G and G
Klimcik and Severa for the special case b = 0.
The general solution for arbitrary b is given by choosing the orthonormal frames to be
given by

e .
= (I + F ) and = F +
(3.6)

694

O. Alvarez / Nuclear Physics B 584 [PM] (2000) 682704

e = (F t )
e with the components of the
We have E = t (F F t F ) and E
t
e = b
e in this basis. The
antisymmetric tensor fields given by B = b F F and B
components of the symplectic form in this basis are


e I
e 1 (I + F ),
(3.7)
m = Ft

1

t
e I
e
e ,
F +
(3.8)
l = F


1
e I
e (I + F ).
(3.9)
l = I Ft
A brief computation shows that m = I + n where n is antisymmetric and given by
n = b +

X


e
F t

k

 
e kF

k=1

X


e
e k.
e k F

F t

k=0

k=0

e = l + m I =
Using (I-3.8) and (I-3.9) we see that B = l m + I = b F t F and B
e. We have reproduced the results of Klimcik and Severa, and Sfetsos.
b
4. Infinitesimally homogeneous nij
4.1. General theory
In this section we address the question, What if nij is the same everywhere? We
will see that this is a much weaker condition than saying n0 = n00 = 0 which we already
studied in Section I-8.2 and lead to abelian duality. We will show that P is a homogeneous
space under certain assumptions. First we have to address the question of what does
same everywhere mean. The best way to do this is to exploit some ideas developed
by Singer [21] for the study of homogeneous spaces. It is convenient to work in the bundle
F (P ) of the adapted orthogonal frames we have been using. This bundle has structure
group O(n) and it admits a global coframing given by ( i , j , kl ) where ( i , j ) are the
canonical 1-forms on the frame bundle and ij is an O(n) connection. Remember that the
MaurerCartan form on O(n) is the restriction of to a fiber of F (P ). The relationships
e P are encapsulated in the Cartan structural equations for
among the geometries of M, M,
F (P ):
d i = ij j 12 fij k j k ,
d i = ij j 12 fij k j k ,
dij = ik kj Tij00lm l m ,

(4.1)
(4.2)
(4.3)

where fij k = fikj , fij k = fikj and Tij00kl = Tj00ikl . Here f, f, T 00 are all functions on
F (P ). Note that there is torsion arising from the reduction of the structure group. The
ideal generated by { 1 , . . . , n } is a differential ideal with integral submanifolds being
the restriction of F (P ) to the fibers of . The ideal generated by { 1 , . . . , n } is a

O. Alvarez / Nuclear Physics B 584 [PM] (2000) 682704

695

differential ideal with integral submanifolds being the restriction of F (P ) to the fibers
e. The degenerate quadratic forms i i and i i on F (P ) are, respectively,
of
e The pullback of the Riemannian connection on
pullbacks of the metrics on M and M.
e Said differently,
the frame bundle of M is schematically + f and likewise for M.
when restricting ij to a horizontal fiber you get an orthogonal connection on the fiber
with torsion, etc. We remind the reader that a tensor in P is a collection of functions on
F (P ) which transform linearly under the action of O(n) on F (P ), i.e., as you change
frames the tensor transforms appropriately. For future use the frame dual to the coframe
( i , j , kl ) will be denoted by (ei , ej , Ekl ). The horizontal vector fields with respect to
are spanned by {eA } = {ei , ej }.
We are now ready to define the statement nij is the same everywhere. Pick a point
b F (P ). If we go to a rotated frame Rb, R O(n), then nij (b) becomes nij (Rb) =
Rik Rj l nkl (b). Notice that as we move along the fiber going through b we will get the
full orbit of nij (b) under O(n). Thus to make sense of nij is the same everywhere we
should not really talk about nij but about the invariants of antisymmetric tensors under the
orthogonal group. We should be thinking in terms of the space of orbits of antisymmetric
tensors under O(n). In the frame bundle, the functions nij define a map n : F (P )
V
V2 n
(R ). We say that P is n-homogeneous if the image of the map n : F (P ) 2 (Rn ) is
a single O(n)-orbit. Said differently, you get the same 2 2 block diagonalization of nij
at each point of P . This is a weaker condition than covariantly constant n. If P is simply
connected then a covariantly constant nij is determined by parallel transporting nij from a
reference point. The value of n at the reference point determines n everywhere.
The condition that P be n-homogeneous is not strong enough for us. This leads to the
notion of infinitesimally n-homogeneous where not only is n the same everywhere but
also the first (N + 1) covariant derivatives of n. Let n denote the covariant derivative of
n, 2 n the second covariant derivative of n, etc. Consider the map
s = (n, n, 2 n, . . . , s n) : F (P ) RJs ,
where Js is an integer we do not compute. We say that P is infinitesimally n-homogeneous
if image of the map N+1 is a single O(n)-orbit. The integer N is determined inductively
as follows.
First we do a rough argument and afterwards we state Singers result. Look at 0 = n :
V
F (P ) 2 (Rn ) and pick a point n0 in the orbit. Consider B = {b F (P ) : n(b) = n0 }.
Note that B is a sub-bundle of F (P ) because n(F (P )) is a single orbit. If K00 O(n) is
the isotropy group of n0 then the action of K00 on a point b B leaves you in B. Thus B
is a principal sub-bundle of F (P ) with structure group K00 . The choice of n0 has broken
the symmetry group to K00 . Now let us be precise about Singers result. Pick a b0 F (P ).
There exists a principal sub-bundle B0 F (P ) containing b0 such that n is constant on
B0 and the structure group K0 O(n) of B0 is the connected component of the isotropy
group of n(b0 ). Note that for a generic orbit, K0 will be a maximal torus of O(n).
Next we invoke n to reduce the symmetry group some more. We use 1 and apply
Singers theorem to it. There exists a principal sub-bundle B1 B0 containing b0 such
that (n, n) is constant on B1 and the structure group K1 K0 of B1 is the connected

696

O. Alvarez / Nuclear Physics B 584 [PM] (2000) 682704

component of the isotropy group of (n(b0 ), n(b0 )). We continue the procedure by
looking at 2 , 3 , . . ., and finding a sequence of principal sub-bundles F (P ) B0
B1 BN BN+1 with respective structure groups O(n) K0 K1 KN
KN+1 . Since O(n) is finite dimensional there exists an integer N such that the chain of
groups satisfies the property that K0 6= K1 6= 6= KN1 6= KN = KN+1 . In fact Singer
establishes that BN = BN+1 , henceforth denoted by H , is a principal bundle with structure
group K = KN . Our arguments show that N+1 = (n, n, . . . , N+1 n) is constant on H .
Note that structure group K is the connected component of the isotropy group of N (b0 ).
The chain of groups tells us that N 6 12 n(n 1) and later on we will see that we also
require N > 1.
Next we show that H is a Lie group and conclude that P = H /K is a homogeneous
space. The strategy is to write down the Cartan structural equations for the principal bundle
H and show that they are actually the MaurerCartan equations for a group. Pick a point
b F (P ) and observe that d N is tangent to the orbit because N (F (P )) is a single orbit.
The orbit is generated by the action of O(n) therefore for X Tb F (P ) there exists a linear
map : Tb F (P ) so(n) such that d N (X) = (X) N (b). Use the standard metric on
SO(n) to write an orthogonal decomposition so(n) = k k , where k is the Lie algebra
of K and k is its orthogonal complement. Let b0 H then we observe that if , 0 are
such that for X Tb0 F (P ) you have d N (X) = (X) N (b0 ) = 0 (X) N (b0 ) then
0 (X) (X) k. At b0 H you can uniquely specify by requiring that (X) k .
We will make this choice. Note that we allow X to be in the full tangent space Tb0 F (P ). In
summary, for b0 H F (P ) there exists a unique linear transformation : Tb0 F (P )
k such that
d N (X) = (X) N (b0 ).

(4.4)

The definition of the covariant derivative is


d N (X) = (X) N (b0 )


+ (X n)(b0 ), X (n)(b0 ), X ( 2 n)(b0 ), . . . , X ( N n)(b0 ) .

(4.5)

Under the decomposition so(n) = k k we have = k + k . Upon restriction to H ,

k is a K-connection on the principal bundle H and k will become torsion. Since K


is the isotropy group of N (b0 ) we conclude that k N (b0 ) = 0. If we restrict (4.5) to
H F (P ) and choose X Tb0 H then d N (X) = 0 because N is constant on H . Thus
for X Tb0 H we have that

k (X) N (b0 )


= (X n)(b0 ), X (n)(b0 ), X ( 2 n)(b0 ), . . . , X ( N n)(b0 ) .
k

(4.6)

If we think of the above as a series of linear equations for (X) then it is easy to see
that if a solution exists then in must be unique. Next we show that the solution exists. To
do this we observe that the covariant derivative (with connection ) of N in direction eA
is given by d N (eA ) = (eA ) N (b0 ), see (4.4), (4.5). Thus we have

(eA ) N (b0 ) = (X n)(b0 ), X (n)(b0 ), X ( 2 n)(b0 ), . . . , X ( N n)(b0 ) .

O. Alvarez / Nuclear Physics B 584 [PM] (2000) 682704

697

Comparing this with (4.6) and using the uniqueness of the solution we conclude that the

torsion k (eA ) = (eA ). Also note that the right hand side of (4.6) is constant on H and

thus the k (eA ) must be constant on H by uniqueness. On restriction to H we have

k k
k k
,
+ kij
ijk = kij

(4.7)

k and k are constant on H .


where kij
kij
We are now almost ready to write down the Cartan structural equations for H . First we
observe that certain functions are constant. If we let F denote fij k or fij k then equations
(I-8.17) and (I-8.18) may schematically be written (I + n)F = n. By differentiating we
learn that s F is a function of (n, n, . . . , s+1 n) only. If N > 1 then fij k , fij k , Tij00kl in
(4.1), (4.2) and (4.3) are constant on H since N+1 is constant on H . We require N > 1.
The first Cartan structural equations (4.1) and (4.2) become
k j
k ,
d i = ijk j 12 cij k j k + kij
k

d i = ijk j 12 cij k j k + kij j k ,


k

(4.8)
(4.9)

where c, c,
, are constant and cij k = cikj , cij k = cikj . Think of the ij indices of
Tij00lm as taking values in the so(n) Lie algebra and denote the projection of Tij00lm onto k
by Kijk lm . Note that Kijk lm is constant on H . The second Cartan structural equation (4.3)
may be written as


k
k
k
k k
k
k k
kj
ik
kj
ik
kj
dijk = ik

k
k k
ik
kj
+ Kijk lm l m ,
(4.10)

where you substitute (4.7) for k in the above. The important lesson is that H has a
coframing given by (, , k ) and that the structural equations (4.8), (4.9) and (4.10) only
involve constants and thus are the MaurerCartan equations for a Lie group. We have
shown that if P is infinitesimally n-homogeneous with N > 1 then P is a homogeneous
space H /K where the Lie group H is a sub-bundle of the frame bundle F (P ).
4.2. The case of K = {e}
This is the situation where the residual symmetry group K is broken all the way down
to the identity group {e}. In this case P = H , the symplectic manifold P is a Lie group,
and k = so(n). The MaurerCartan equations for P are
d i = 12 cij k j k + kij j k ,
d i = 12 cij k j k + kij j k ,

(4.11)
(4.12)

where cij k = cikj , cij k = cikj . Also kij and kij are antisymmetric under i j for
arbitrary i, j reflecting that k = so(n). Note that generates a differential ideal and thus
e with fibers isomorphic to a Lie group G with structure
e:P M
= 0 defines a fibration
constants cij k . Likewise, generates a differential ideal and thus = 0 defines a fibration
e with structure constants cij k . We also
: P M with fibers isomorphic to a Lie group G

698

O. Alvarez / Nuclear Physics B 584 [PM] (2000) 682704

e are Lie subgroups of P . If = 0 then G is a normal subgroup of P .


remark that G and G
e is a normal subgroup of P .
If = 0 then G
Let (ei , ej ) be the basis dual to ( i , j ). The MaurerCartan equations may be
reformulated as the Lie algebra relations
[ei , ej ] = ckij ek ,

(4.13)

[ei , ej ] = ckij ek ,

(4.14)

[ei , ej ] = ikj ek j ki ek .

(4.15)

All the statements made in the previous paragraph also follow from the above.
Consider the left invariant vector fields X = Xi ei , Y = Y i ei . Note that LX i =
Xk kij j . Let ij (X) = Xk kij then using the identity [LX , LY ] i = L[X,Y ] i you obtain
[ (X), (Y )] = ([X, Y ]). Thus we have a Lie algebra representation : g so(n). This
means we have a representation of G by real orthogonal n n matrices. Likewise, : g
e by real orthogonal
so(n) is a Lie algebra representation and we have a representation of G
matrices. This does not mean that G is a compact group if 6= 0. A comment made
in Appendix A tells us that g/(ker ) is a Lie algebra of the form compact semisimple
+ abelian. We know nothing at all about the ideal ker so we cannot make a stronger
e
statement about the structure of g. Similar remarks apply to G.
A Drinfeld double GD admits the following geometric characterization:
1. It is a Lie group of dimension 2n with a bi-invariant quadratic form of type (n, n).
2. It is a bifibration with the property that the fibers are isotropic submanifolds of GD .
e
The fibers are also isomorphic to Lie groups G and G.
If we apply the above to our situation by requiring that the quadratic form i i + i i
be bi-invariant then we learn that P = H is a Drinfeld double, ij k = cij k , ij k = cij k ,
kij = ikj and kij = ikj . It follows that both cij k and cij k are totally antisymmetric
and thus the quadratic forms i i and i i give bi-invariant positive definite metrics
e respectively. Thus G and G
e are of the type compact semisimple + abelian.
on G and G,
The symplectic structure on P (if it exists) appears to be different than the standard
symplectic structure on the Drinfeld double, see (3.1). In the examples I am familiar, if
e constructed via R-matrices is neither simple nor
G is simple and compact the its dual G
compact. I do not know what is known in this more general case compact semisimple +
abelian.
Returning to the general case we remark that the equations d2 = 0 and d2 = 0 leads
to the conclusion that cij k and cij k are respectively the structure constants for Lie groups
e : g so(n) and : g so(n) are Lie algebra representations and
G and G,
kj m j il lj m j ik + cij k mj l cij l mj k + cj kl mij = 0,

(4.16)

kj m j il lj m j ik + cij k mj l cij l mj k + cj kl mij = 0.

(4.17)

These equations are generalizations of the corresponding equations (B.4) in the bialgebra
case. These equations follow just from the structure equations for the group P = H . There
e
are additional constraints which follow from duality considerations such as d = H H
which lead to

O. Alvarez / Nuclear Physics B 584 [PM] (2000) 682704

cij k nil + cikl nij + cilj nik = +Hj kl ,


ej kl ,
cij k nil + cikl nij + cilj nik = H

699

(4.18)
(4.19)

mj l kli mkl j li + nj l ilk nkl ilj = mil clj k ,

(4.20)

mj l kli mkl j li + nj l ilk nkl ilj = +clj k mli .

(4.21)

We do not know if there are non-trivial solutions to these equations.

Acknowledgements
I would like to thank O. Babelon, L. Baulieu, T. Curtright, L.A. Ferreira, D. Freed,
S. Kaliman, C.-H. Liu, R. Nepomechie, N. Reshetikhin, J. Snchez Guilln, N. Wallach
and P. Windey for discussions on a variety of topics. I would also like to thank Jack Lee for
his M ATHEMATICA package R ICCI that was used to perform some of the computations.
I am particularly thankful to R. Bryant and I.M. Singer for patiently answering my many
questions about differential geometry. This work was supported in part by National Science
Foundation grant PHY-9870101.

Appendix A. Some Lie groups facts


For clarification purposes we make some remarks about the left and right actions on a
Lie group. For notational simplicity we restrict to matrix Lie groups. The identity element
will be denoted by I . Let G be a Lie group. Let a G then the left and right actions on
G are respectively defined by La g = ag and Ra g = ga for g G. We take g, the Lie
algebra of G, to be the left invariant vector fields and we identify it with the tangent space
at the identity TI G. The left invariant MaurerCartan forms are = g 1 dg. They satisfy
the MaurerCartan equations d = . Pick a basis (e1 , . . . , en ) of left invariant vector
fields for g with bracket relations [ei , ej ] = f k ij ek . If the dual basis of left invariant forms
is ( 1 , . . . , n ) then d i = 12 f i j k j k .
Naively you would expect that if X is a left invariant vector field then LX = 0 since
is left invariant. In fact a brief computation shows that LX i = (Xk f i kj ) j which is the
adjoint action. The answer to this conundrum is that the left invariant vector fields generate
the right group action. The easiest way to see this is to use old fashioned differentials.
Let v TI G then the left invariant vector field X at g which is v at the identity is given
by X = gv. The infinitesimal action of this vector field at g is given by g g + X =
g + gv = g(I + v) gev which is the right action of the group. Thus we see that
g 1 dg ev (g 1 dg)ev which is the adjoint action in accordance with the Lie derivative
computation. Take any metric hij at the identity then hij i j is a left invariant metric
on G. In general this metric is not invariant under the right action of the group. The right
invariance condition is hil f l j k + hj l f l ik = 0 which means that the structure constants are
totally antisymmetric if the indices are lowered using hij . In such a situation the metric is
bi-invariant.

700

O. Alvarez / Nuclear Physics B 584 [PM] (2000) 682704

Assume you have a Lie algebra g with an invariant positive definite metric then you
have an orthogonal decomposition into ideals g = k a where k is a compact semisimple
Lie algebra and a is abelian. The proof is straightforward and involves putting together
a variety of observations. The invariance of the inner product (, ) is the statement that
the adjoint representation adX Y = [X, Y ] is skew adjoint with respect to the metric:
(adX Y, Z)+(Y, adX Z) = 0. The skew adjointness immediately leads to the decomposition
of g into irreducible pieces. It is an elementary exercise in linear algebra to show that if h is
a non-trivial ideal in g then its orthogonal complement h is also an ideal. Since g = hh
we conclude that by continuing this process the Lie algebra decomposes into irreducible
ideals g = g1 g2 gN . Let us collate all the abelian subalgebras together into a
and rewrite the decomposition as g = k1 k2 kM a, where each kj is a simple
lie algebra. Since this is a decomposition into ideals we have that a is the center of the Lie
algebra. Let (, )j be the restriction of the invariant inner product to kj . An application of
Schurs lemma tells us that an invariant bilinear form on a simple Lie algebra is a multiple
of the Killing form. Thus we conclude that (, )j = Kj (, ), where is a non-zero scalar
and Kj is the Killing form 2 on kj . Since (, ) is positive definite, the Killing form Kj
must be definite and this is only possible if the Lie algebra kj is of compact type. This
concludes the proof of the proposition in the opening sentence.
Closely related to the above is the following. If a Lie algebra g has a faithful
representation : g so(n) by skew adjoint matrices then g = k a where k is a compact
semisimple Lie algebra and a is abelian. The proof follows from the observation that
because the representation is faithful we can think of g as a matrix Lie subalgebra of so(n).
We know that so(n) has a positive definite invariant metric so restriction to g induces
a positive definite invariant metric on g. We now use the proposition from the previous
paragraph.
We remark that if the representation in the previous paragraph is not faithful then
(ker ) g is a nontrivial ideal in g. The Lie algebra g/(ker ) is of the form k a but we
can say nothing about the Lie algebra (ker ).
It is easy to see that the space of all invariant positive definite metrics on a Lie algebra
is a convex set. In a simple Lie algebra g, the third cohomology group H 3 (g) is one
dimensional and generated by the three form (X, Y, Z) = K(X, [Y, Z]) where K is the
killing form which may be written in terms of the structure constants as ij k = Kil f l j k . If
h is an invariant metric then h(X, [Y, Z]) is also a closed three form and by the convexity
of the space of positive definite invariant metrics it must be in the same cohomology class
as .

Appendix B. A primer on classical R-matrices


There are two main nonequivalent approaches to classical R-matrices. The more familiar
one is due to Drinfeld and based on the study of Lie bialgebras [17]. The other due to
2 The sign of the Killing form is chosen such that it is positive on a compact simple Lie algebra.

O. Alvarez / Nuclear Physics B 584 [PM] (2000) 682704

701

Semenov-Tian-Shansky is based on double Lie algebras [16] is the one directly related
to our work. Here we discuss the interconnections between these two approaches. For an
introduction to R-matrices and PoissonLie groups see the book by Chari and Pressley [22]
or the article [20].
B.1. The Drinfeld approach
Drinfeld bases his approach on the notion of a Lie bialgebra. We begin with a down to
earth approach. Assume you have a Lie algebra g with basis (e1 , . . . , en ) and Lie bracket
relations [ea , eb ] = f c ab ec . If X, Y g then the adjoint action by X is given by adX Y =
[X, Y ]. The adjoint action extends naturally to tensor products of g. Let g be the vector
space dual with corresponding basis (1 , . . . , n ). The Lie algebra g acts on g via the
coadjoint representation adea b = f b ac c . In general g is not a Lie algebra but there is
a natural Lie algebra structure on g g given by
[ea , eb ] = f c ab ec ,
[ea , b ] = f b ac c ,
[a , b ] = 0.
The most famous example is combining g = so(3) and its Lie algebra dual g R3 into
the Lie algebra of the euclidean group E(3). The situation becomes much more interesting
when g is a Lie algebra in its own rights [a , b ] = fc ab c . You observe that g acts
via its coadjoint action on its dual (g ) g. Thus one can consider the following more
symmetric structure which takes into account the respective coadjoint actions
[ea , eb ] = f c ab ec ,
[ea , b ] = fa bc ec f b ac c ,

(B.2)

[a , b ] = fc ab c .

(B.3)

(B.1)

This will be a Lie algebra if the following conditions are satisfied:


f e ab fe cd = fb ed f c ae + fb ce f d ae fa ed f c be fa ce f d be .

(B.4)

According to Drinfeld a Lie bialgebra is a Lie algebra with Lie brackets (B.1), (B.2) and
(B.3).
Drinfeld gives a more abstract formulation which is more suitable for studying the
abstract properties of a bialgebra and seeing the origins of classical R-matrices. Assume
V
you have a Lie algebra g and a cobracket : g 2 g. Drinfeld requires that the
cobracket defines a Lie algebra on g . The structure constants on g are related to the
cobracket by (ea ) = 12 fa bc eb ec . Compatibility condition (B.4) is incorporated via a
cohomological argument. The complex in question is
^2 ^2
^2
^2
g g
g
g
g .
The coboundary operator (differential) is given by



( )(X, Y ) = adX (Y ) adY (X) [X, Y ] .

(B.5)

702

O. Alvarez / Nuclear Physics B 584 [PM] (2000) 682704

The condition (B.4) that glues the Lie algebras into a Lie bialgebra is seen to be equivalent
to the cocycle condition = 0. In this language, a Lie bialgebra is a Lie algebra g along
with a cobracket such that g a Lie algebra and the cobracket is a 1-cocycle.
V
A Lie bialgebra is exact if the cocycle is exact. This means that there exists a r 2 g
such that (X) = r = adX r. A computation shows that r defines a bialgebra structure if
V
and only if the Schouten bracket [[r, r]] is ad(g)-invariant: [[r, r]] ( 3 g)inv . The Schouten
bracket is defined by
[[W X, Y Z]] = [W, Y ] X Z [W, Z] X Y
[X, Y ] W Z + [X, Z] W Y.
V3
g)inv is called the modified classical YangBaxter equation
The condition [[r, r]] (
(MCYBE) and [[r, r]] = 0 is called the classical YangBaxter equation (CYBE).
V
If g is semisimple then the Whitehead lemma states that H 1 (g, 2 g) = 0 and thus the
cocycle is always a coboundary = r. Thus in this case we need to understand the
V
set of all r 2 g which satisfy MCYBE.
V
If g is simple then ( 3 g)inv is one dimensional and is generated by the three index
tensor obtained by raising two indices on the structure constants using the Killing metric.
If we call this object BK then MCYBE becomes [[r, r]] = aBK for some constant a.
Let us work in a basis. If r = 12 r ab ea eb . Then (ea ) = adea r = 12 r bc adea (eb ec ) =
1 dc b
bd c
2 (r f ad + r f ad )eb . This tells us that
fa bc = r dc f b ad + r bd f c ad .

(B.6)

Note that fa = fa because r is skew.


It is possible to generalize the above by allowing the cocycle (now called C) to be in
g g. If you write C = C ab ea eb and you let (X) = adX C. In this case you find you
get a Lie bialgebra if the following two conditions are satisfied:
1. (C ab + C ba )ea eb , the symmetric part of C, is ad(g)-invariant,
2. [[C, C]] [C 12 , C 13 ] = [C 12 , C 23 ] + [C 13 , C 23 ] g g g is ad(g)-invariant.
We use standard quantum group notation where for example C 13 = C ab ea I eb , etc.
The bracket [[, ]] above reduces to the Schouten bracket if C is skew symmetric. We
remark that the equation [[C, C]] = 0 is also called the classical YangBaxter equation.
C or r are called classical R-matrices (by Drinfeld). The modified classical YangBaxter
equation is [[C, C]] (g g g)inv .
A brief computation shows that
bc

cb

fa bc = C dc f b ad + C bd f c ad .
Let us write C = 12 s ab (ea eb + eb ea ) + 12 r ab ea eb where s
antisymmetric. Since s ab is an ad(g)-invariant tensor we have that
fa bc = r dc f b ad + r bd f c ad .

(B.7)
is symmetric and r is
(B.8)

only depends on the antisymmetric part of C. Note


Thus the Lie algebra structure on
that fa bc will be skew under b c as required. The effect of the symmetric part sab is
to change the ad g-invariant term in the right hand side of the MCYBE. Eq. (B.7) may be
interpreted as giving a Lie algebra homomorphism C : g g.

O. Alvarez / Nuclear Physics B 584 [PM] (2000) 682704

703

B.2. Semenov-Tian-Shansky approach


Semenov-Tian-Shanskys approach is directly influenced by classical integrable models
where he needs that a single Lie algebra admits two different Lie brackets. Let g be a Lie
algebra and let R : g g be a linear transformation (not necessarily invertible). Define a
skew operation [, ]R by
[X, Y ]R = [RX, Y ] + [X, RY ].

(B.9)

If [, ]R is a Lie bracket then R is called a classical R-matrix by Semenov-Tian-Shansky.


The Jacobi identity for [, ]R may be written as


X, [RY, RZ] R [Y, Z]R + cyclic permutations = 0.
A Lie algebra g with two Lie brackets [, ] and [, ]R is called a double Lie algebra by
Semenov-Tian-Shansky. The equation is [RY, RZ] R([Y, Z]R ) = 0 is also called the
CYBE. The equation

(B.10)
[RY, RZ] R [Y, Z]R = c[Y, Z],
where c is a constant is also called the MCYBE. Solutions to either of these satisfy the
Jacobi identity displayed above.
In a basis we have that Rea = eb R b a , [ea , eb ]R = (f c ab )R ec and consequently (B.9)
becomes [ea , eb ]R = (R d a f c db + R d b f c ad )ec . Thus the new structure constants are
(f c ab )R = R d a f c db + R d b f c ad .

(B.11)

The Semenov-Tian-Shansky version of the MCYBE (B.10) is


f c de R d a R e b f e db R d a R c e f e ad R d b R c e = cf c ab .

(B.12)

B.3. Relating Drinfeld and Semenov-Tian-Shansky


To relate the Semenov-Tian-Shansky approach and the Drinfeld approach one needs
an ad(g)-invariant metric on g. If g is semisimple then one can take the ad(g)-invariant
metric to be the Killing metric. The ad(g)-invariant metric is used to identify g with g . By
lowering indices fabc is completely antisymmetric (due to the ad(g)-invariance). We wish
to identify fR with f. Note that by rearranging indices we have
(fc ab )R = R da fcd b + R db fc a d
= R da f b cd R db f a cd .

(B.13)

=
then we have related fR to f. Said differently
If we are in a situation where
R : g g is a skew-adjoint operator with respect to the invariant metric. In fact there
is a theorem [16] which states that if g has an ad(g)-invariant metric and if R : g g
is skew-adjoint then the double Lie algebra is isomorphic to a Lie bialgebra and all the
structures in the Semenov-Tian-Shansky approach (CYBE, MYBE) go into the structures
in the Drinfeld approach (CYBE, MYBE). The isomorphism is given by thinking of the
metric as giving a map g g , i.e., lowering/raising indices. We are in a different situation
because not all our R-matrices are skew adjoint.
R ab

r ab

704

O. Alvarez / Nuclear Physics B 584 [PM] (2000) 682704

References
[1]
[2]
[3]
[4]
[5]
[6]
[7]
[8]
[9]
[10]
[11]
[12]
[13]
[14]
[15]
[16]
[17]
[18]
[19]
[20]
[21]
[22]

O. Alvarez, Nucl. Phys. B 584 (2000) 659, preceding article in this issue.
B.E. Fridling, A. Jevicki, Phys. Lett. B 134 (1984) 70.
E.S. Fradkin, A.A. Tseytlin, Ann. Phys. 162 (1985) 31.
E.B. Kiritsis, Mod. Phys. Lett. A 6 (1991) 2871.
M. Rocek, E. Verlinde, Nucl. Phys. B 373 (1992) 630, hep-th/9110053.
A. Giveon, M. Rocek, Nucl. Phys. B 380 (1992) 128, hep-th/9112070.
X.C. de la Ossa, F. Quevedo, Nucl. Phys. B 403 (1993) 377, hep-th/9210021.
M. Gasperini, R. Ricci, G. Veneziano, Phys. Lett. B 319 (1993) 438, hep-th/9308112.
A. Giveon, M. Rocek, hep-th/9406178.
A. Giveon, M. Rocek, Nucl. Phys. B 421 (1994) 173, hep-th/9308154.
A. Giveon, E. Kiritsis, Nucl. Phys. B 411 (1994) 487, hep-th/9303016.
C. Klimcik, P. Severa, Phys. Lett. B 351 (1995) 455, hep-th/9502122.
C. Klimcik, P. Severa, Phys. Lett. B 372 (1996) 65, hep-th/9512040.
C. Klimcik, P. Severa, Phys. Lett. B 376 (1996) 82, hep-th/9512124.
K. Sfetsos, Nucl. Phys. B 517 (1998) 549, hep-th/9710163.
M.A. Semenov-Tian-Shansky, Func. Anal. Appl. 17 (4) (1983) 17.
V.I. Drinfeld, Dokl. Akad. Nauk SSSR 268 (2) (1983) 798.
M.A. Semenov-Tian-Shansky, RIMS 21 (1985) 1737.
A.Y. Alekseev, A.Z. Malkin, Commun. Math. Phys. 162 (1994) 147, hep-th/9303038.
D. Alekseevsky et al., J. Geom. Physics 26 (1998) 340, math.DG/9801028.
I.M. Singer, Commun. Pure Appl. Math. XIII (1960) 685.
V. Chari, A. Pressley, A Guide to Quantum Groups, Cambridge University Press, 1994.

Nuclear Physics B 584 [PM] (2000) 705718


www.elsevier.nl/locate/npe

Crosscaps, boundaries and T-duality


L.R. Huiszoon, A.N. Schellekens
NIKHEF Theory Group,
P.O. Box 41882, 1009 DB Amsterdam, Netherlands
Received 4 May 2000; revised 22 May 2000

Abstract
Open descendants with boundaries and crosscaps of non-trivial automorphism type are studied.
We focus on the case where the bulk symmetry is broken to a Z2 orbifold subalgebra. By requiring
positivity and integrality for the open sector, we derive a unique crosscap of automorphism type
g Z2 and a corresponding g-twisted Klein bottle for a charge conjugation invariant. As a specific
example, we use T-duality to construct the descendants of the true diagonal invariant with symmetry
preserving crosscaps and boundaries. 2000 Elsevier Science B.V. All rights reserved.
PACS: 11.25.Hf

1. Introduction
The general prescription to construct open unoriented strings from closed oriented
ones is known as the method of open descendants [13]. It is not limited to circle
compactifications and orbifolds of circles, as orientifolds are. In short, one has to find a
set of crosscap and boundary coefficients from which one can calculate the Klein bottle,
annulus and Mbius strip partition function that together with the torus generate the full
spectrum of the open unoriented string. Various consistency conditions [49] constrain
these coefficients. However, most of these constraints require an explicit knowledge of
OPE-coefficients and duality (fusing and braiding) matrices, which are only known in a
limited number of cases. There is however one very powerful consistency condition that
can be applied to all conformal field theories; the partition functions in the open sector have
to generate positive and integral state multiplicities. In all known cases [10], this condition
determines the crosscap coefficients uniquely once the boundary coefficients are known.
Historically, much progress on open descendants was made after discoveries in
conformal field theory on surfaces with boundaries. Cardy [4] derived the boundary
coefficients that describe symmetry preserving boundary conditions in case that the closed
strings are described by a charge conjugation invariant. Sagnotti and collaborators [11,12]
found a formula for the crosscap coefficient, thus completing the open descendants for the
0550-3213/00/$ see front matter 2000 Elsevier Science B.V. All rights reserved.
PII: S 0 5 5 0 - 3 2 1 3 ( 0 0 ) 0 0 3 2 0 - 5

706

L.R. Huiszoon, A.N. Schellekens / Nuclear Physics B 584 [PM] (2000) 705718

C-diagonal case. The boundary coefficients for some non-charge conjugation invariants
were derived in [13]. In [10,14] we constructed open descendants for these theories by
deriving the unique crosscap coefficients that, together with the boundary coefficients,
satisfy positivity and integrality of the open sector partition function.
Due to the work of Fuchs and Schweigert [13,1518], the boundary coefficients for
boundary conditions that leave an orbifold subalgebra of the full (bulk) symmetry
invariant are now known. In this letter, we will complete the construction of open
descendants for this case. The organization of this letter is as follows: In Section 2,
we briefly review the relation between Z2 orbifolds and simple current extensions. In
Section 3, the open descendants of an integer spin invariant in the orbifold theory are
presented. We derive two inequivalent Klein bottles for the integer spin invariant. These
results are interpreted in terms of the extended theory in Section 4. We will show that Tduality appears in an elegant way. As a specific example of our results, we construct the
descendants of the true diagonal invariant with symmetry preserving boundary conditions.
All technicalities are confined in the appendix.

2. Orbifolds and simple currents


In this letter we will consider the following situation. We start with a theory with a chiral
algebra A and an order two, integer spin simple current J . We can then extend A by J , and
obtain a new theory, whose chiral algebra we will denote as AE . The fields of the A-theory
can be divided into three classes: uncharged, non-fixed fields (labeled i0 ); charged fields
(labeled i1 ) and fixed fields (labeled f ), which are always uncharged. The fields of the AE
theory are then labeled by the uncharged, length two orbits (formed by a pair of fields i0
and J i0 ), plus two fields for each fixed point f . Choosing an orbit representative we denote
the former as [i0], and the latter as [f, ], where is a Z2 character. 1
Conversely, the chiral algebra AE has an order two automorphism, denoted and there
are two kinds of irreducible highest weight representations [19]: symmetric fields [i0 ] are
invariant under the action of and non-symmetric fields [f, ] transform as [f, ] =
[f, ]. This automorphism has the special property that it leaves the Virasoro algebra,
and in particular the conformal weights fixed.
Such automorphisms are of interest in open string constructions, since one may consider
boundaries and crosscaps that are -twisted in the sense of [16]. Such boundaries or
crosscaps do not preserve the full chiral algebra, but only the sub-algebra A. For string
consistency this must include the Virasoro algebra and in particular L0 , which explains
why one must require that conformal weights are preserved exactly. A well-known example
of such an automorphism is charge conjugation. The symmetric fields are then the selfconjugate fields, and the non-symmetric fields are the complex fields. Another example
is the permutation automorphism of a tensor product of two identical conformal field
1 Denoting the Z elements as (1, g), the characters are explicitly : (1) = (g) = 1, and : (1) =
2
0
0
0
1
1
1, 1 (g) = 1. We will furthermore use the convention that , without arguments, denotes the value of the
character on g, which is precisely what distinguishes the two characters.

L.R. Huiszoon, A.N. Schellekens / Nuclear Physics B 584 [PM] (2000) 705718

707

theories [20]. The point is now that any such automorphism can be described most easily
in terms of the -orbifold theory of the AE -theory, and that orbifold theory is precisely the
A-theory. 2
Our approach is thus to take the orbifold theory A as the starting point, and extend
in the closed sector by the current J to AE . Then we find corresponding boundaries and
crosscaps, but without insisting that they also have the full symmetry AE . From the point
of view of the AE -theory this goes by the name of broken bulk symmetry, whereas from
the point of view of the A theory the term extended bulk symmetry would seem more
appropriate.

3. Open descendants of integer spin invariants


In this section, we will focus on the A theory. Consider the following modular invariant
torus partition function [21] 3
X
X
X

ZJ =
ZijJ i j =
(i0 + J i0 ) i0c + J i0c + 2
f f c ,
(1)
ij

i0 ,Rep

where i denotes a generic field in the orbifold and i the corresponding (Virasoro)
character. The first sum in the last expression is over all representatives of integer charge
orbits.
We will proceed as follows. We first give the boundary coefficients, which were
presented in [1618]. From these, we will derive two crosscap coefficients. We will find
that one of them satisfies all open and closed string positivity and integrality conditions.
The other only satisfies those conditions if the boundary conditions of [1618] are
modified. In the next section, we will interpret these results in terms of the AE theory.
Three sets of labels have to be distinguished in the following. The transverse channel
labels belong to fields propagating in the bulk. They are the fields that according to the
torus partition function are paired with their charge conjugate, with a multiplicity given
by the order of the untwisted stabilizer. In this case, the latter equals the stabilizer Sm , the
group of simple currents that fix m. Hence, the transverse channel labels are the chargeless
fields m (i.e., i0 or f ) with a multiplicity label m which is the character of Sm . Note
that the multiplicity label is trivial if m = i0 . The boundary labels distinguish different
boundaries. They were determined in [1618] by considering the classifying algebra. The
result is that the boundary labels are in one-to-one correspondence with the orbits, but
with an extra multiplicity: [, ] = [i], [f, ], where [i] = [i0 ], [i1]. The third kind of
label that occurs in the following is simply the primary field label of the A-theory, which
will be denoted as i. The relevant quantities appearing in the positivity conditions are the
boundary coefficients Bma , where m is a generic transverse channel label (m, m ) and a
2 Although our notation is similar to [1619], there is one important difference: the chiral algebras A and A of
these authors are denoted respectively AE and A in the present letter. This is because the orbifold theory plays
the most prominent rle in our work, and we want to avoid excessive occurrences of bars in formulas.
3 This invariant is a product of an integer spin invariant and charge conjugation. The charge conjugate of a field
i is denoted by i c .

708

L.R. Huiszoon, A.N. Schellekens / Nuclear Physics B 584 [PM] (2000) 705718

a generic boundary label [, ] and the crosscap coefficients m . In terms of these, the
direct annulus, Mbius and Klein bottle are, respectively,
X
X
X
S im Bma Bmb ;
M ia =
P im Bma m ;
Ki =
S im m m , (2)
Aiab =
m

since the A-characters on which S act do not


with the understanding that
depend on . Our conventions and normalizations are as in our previous papers [30]

P
=
and [14]. In particular, the reflection coefficients Rma = Bma Sm0 satisfy m Rma Rmb
P
= .
ab and a Rma Rna
mn
The boundary coefficients are
P

J
|G|
J m (J ) (J )Sm
,
(3)
B(m,m )[, ] =

|Sm | |S |
S0m
S i[f,]

S if ,

where |Sm | is the dimension of the stabilizer of m, and |G| = 2 is the dimension of the
simple current group. The sum is over all currents in the intersection Sm S . The matrix
S 0 S is the usual S-matrix of the A theory and 4 S J S is the fixed point resolution
matrix for the current J . These matrices are explicitly known for WZW-models [23,24]
and extended WZW-models [25]. This result (3) is obtained from [16], apart from the
normalization, which we have adapted to our conventions.
The direct annulus can be computed from the boundary coefficients using (2):
X

(4)
N ij k + N Jjik i ,
A[j ][k] =
i

A[j ][g, 0 ] =

N ijg i ,

(5)

A[f,][g, 0 ] =


1X i
N fg + 0 N ifg i ,
2

(6)

where N are the fusion coefficients of the orbifold theory and


i
=
N fg

X Sf m Sgm S

im

S0m

(7)

Note that the annuli have the following property due to (monodromy) charge conservation.
When the charges of the boundary labels are equal, only untwisted sector fields contribute
to the sums. When the boundary labels have a different charge, i.e., in mixed annuli,
only twisted sector fields contribute. Furthermore all characters appear in AE linear
combinations i + J i , although these combinations are AE -characters only if i has zero
charge.
Let us now turn to the crosscap coefficients. They can be derived in a similar way as was
done in [10]. That is, we have to require that the Mbius strip satisfies the positivity and
integrality relation

i
i
i
i

M
(8)
[, ] 6 A[, ][, ] and M[, ] = A[, ][, ] mod 2.
4 Note that in [14] S is defined differently: it is related to S J by a phase.

L.R. Huiszoon, A.N. Schellekens / Nuclear Physics B 584 [PM] (2000) 705718

709

If we choose the special boundary [0], this implies


i
= 1 0i + 2 0J i ,
M[0]

where 1 and 2 are signs. Inverting the relation between M and the crosscap coefficients
gives
s
X
|Sm | 1 P0m + 2 PJ m
(m,m ) =

.
(9)
|G|
S0m

Note that for fixed points only the sum over m is determined, not the coefficients
separately. The solution involves therefore a set of unknown quantities m :


1
1 P0m + 2 PJ m
m

+ m
(m,m ) =
.
(10)
|G||Sm |
S0m
S0m
Intuitively one expects m to vanish because there is only one crosscap; indeed, such
corrections occur for Bma only for boundary labels that occur in pairs originating from a
fixed point. We will find that all positivity and integrality conditions are satisfied if m = 0.
Introducing non-zero m s leads in most cases to violations of these conditions in the closed
and/or open channel or to complex Mbius coefficients. We can show that m = 0 if P0m 6=
0 or PJ m 6= 0 and also that all m must vanish if S J is purely real or imaginary (as it is in
most cases), but we cannot rule out m in all imaginable cases. We will therefore assume
from now on that it vanishes. Then the crosscap coefficients are fixed up to two signs 1
and 2 .
Requiring positivity and integrality for a boundary label [f, ] fixes the relative sign
(see Appendix A), and the overall sign is in any case never fixed by CFT considerations.
Up to this overall sign, the result is
1
P0m + PJ m
+
=

,
(m,
m)
|Sm ||G|
S0m

(11)

where  eihJ is a sign. The meaning of the superscript + becomes clear later. Now
we can compute the direct Klein bottle
X
(Yi00 + Yi0J )i .
(12)
K ++ =
i,QJ (i)=0

Since this Klein bottle was derived using open sector positivity constraints (in fact, just a
few of them), it is perhaps surprising that it satisfies the positivity and integrality condition
for the closed sector (see Appendix A.1 for details).
Let us assume for the sake of definiteness that AE has complex representations,
with conjugation corresponding to [f, ]c = [f, ]. (This amounts to taking = C,
i.e., charge conjugation. All of the following holds in more general situations.) As we
will explain in the next section, the invariant (1) can either be interpreted as a charge
conjugation invariant or a diagonal invariant for the extension AE . We will also see that the
Klein bottle K ++ is a standard Klein bottle for the charge conjugation invariant; it projects
on world-sheet parity invariant states. However, K ++ is a twisted Klein bottle for the
diagonal invariant, which means that it projects on invariant states for this invariant

710

L.R. Huiszoon, A.N. Schellekens / Nuclear Physics B 584 [PM] (2000) 705718

(see Section 4.1). Recall that the crosscap we derived is unique. 5 So in order to find a
standard Klein bottle projection for the diagonal invariant, we either have to change the
boundary coefficients or the positivity and integrality condition of the open sector. Suppose
for the moment that we keep the boundary conditions (3) fixed. Instead of (8), we require
a -twisted positivity and integrality condition

i
i
i
i

M
(13)
[, ] 6 A[, ][, ] and M[, ] = A[, ][, ] mod 2.
It is easy to see that we can now derive a unique crosscap given by
1
P0m PJ m

,
(m,
m)
|Sm ||G|
S0m
and corresponding Klein bottle
X
(Yi00 Yi0J )i ,
K =

(14)

(15)

i,QJ (i)=0

which is inequivalent to K ++ but also satisfies positivity and integrality of the closed
sector. This Klein bottle has the desired property that it is a standard Klein bottle for the
diagonal invariant (see Section 4.1).
Alternatively, one may leave the positivity and integrality conditions (8) unchanged,
but modify the boundary conditions simply by replacing S J by iS J in (3). This flips the
sign in the annulus (6) for two fixed point boundary labels and leads straightforwardly
to the crosscap (14). This is reminiscent of what happens if one chooses different Klein
bottle projections in the Cardy case, as in [11,30]. If one leaves the boundary coefficients
unchanged, one may encounter contributions like 12 (N a2 + N b2 )0 in the open string
partition function, where 0 is the identity character and N the CP factors. Changing the
appropriate boundary conditions by a factor i changes this to Na Nb 0 . For Na = Nb the
latter can be interpreted in term of a U (Na ) gauge group, whereas the former (even though
for Na = Nb it is numerically equal) does not seem to allow a gauge group interpretation.
Therefore we think changing the boundary coefficients is the correct interpretation. It is
not clear to us whether this affects the analysis of [1618], in which (3) is derived from the
sewing constraint for the bulk-bulk-boundary correlator.
Finally, we display the Mbius strip amplitudes. In the direct channel (open string loop
channel) they are
X

(16)
Yj 0 i Yj J i i ,
M[j] =
i

M[f,]


1X
=
Yf 0 i Yf J i i ,
2

(17)

where the refer to the crosscap (11 or 14) that appear in the transverse Mbius strip. By
(monodromy) charge conservation and Eq. (A.7) of the appendix, only chargeless fields
5 A possible non-uniqueness of the crosscap due to the in the crosscap coefficient (10) cannot provide us a
m
standard Klein bottle for the diagonal invariant: in case of = C, S J is purely imaginary so all m must vanish.

L.R. Huiszoon, A.N. Schellekens / Nuclear Physics B 584 [PM] (2000) 705718

711

contribute. In the appendix, we show that M + and M satisfy the positivity and integrality
conditions (8) and (13) respectively for all boundary labels.
One is tempted to consider, as the notation might suggest, the introduction of Chan
Paton factors M for the crosscap coefficients in the same way as normal CP-factors
are introduced for boundary labels. However, it is not hard to show that only twisted states
(Q(i) 6= 0) contribute to mixed Klein bottles K + . Since these states do not occur in the
torus partition function, the requirement of positivity and integrality of the closed sector,
i.e., Eq. (A.9), forces us to put one of the crosscap CP-factors to zero. From now on, we
will switch to a more economical notation by defining K ++ K + and K K .
Note that the boundary and crosscap coefficients that appeared in this section are very
similar to the coefficients of the descendants of order two half-integer spin invariants [14].
In that case however, the two different sets of crosscap coefficients have a different origin;
they correspond to simple current Klein bottles [30].

4. Open descendants and T-duality


As already stressed in Section 2, the orbifold theory A has an integer spin simple current
J by which we can extend the chiral algebra. The result of this extension is simply AE .
The characters of the AE theory are related to those of the orbifold as follows:
[i0 ] = i0 + J i0 ,

[f,] = [f,] = f .

(18)

The invariant (1) can therefore be interpreted as a charge conjugation invariant Z c of the
AE theory or as an invariant that is the product of charge conjugation and , denoted by
Z c :
X
X
CI J I J ,
Z c =
CI,J I J ,
(19)
Zc =
IJ

IJ

AE

theory. These invariants are known to be T-dual. By


where I, J are generic fields of the
T-duality, we simply mean a one-sided transformation that acts on closed string states
as
T : |I, J i |I, J i,

(20)

which is a duality for every automorphism that preserves the conformal weights exactly.
Note that T -duality acts not only on the ground states but also on the currents in AE /A.
This implies in particular that the definition of the Ishibashi [4,15,26] states flips: AE symmetric Ishibashi states are turned into Ishibashi states of automorphism type [15,26]
and vice-versa. This is a direct consequence of the fact that in one of the chiral algebras J
is replaced by (J ).
From the point of view of the A theory T-duality is trivial, since the automorphism
is defined only after resolution of the fixed points. Hence all boundary and crosscap states
and all amplitudes are invariant under T-duality. T-Duality becomes non-trivial only once
we interpret the result in terms of the AE -theory.

712

L.R. Huiszoon, A.N. Schellekens / Nuclear Physics B 584 [PM] (2000) 705718

A second clue to the automorphism type of crosscaps and boundaries is the expression
one obtains in terms of A characters for the direct Klein bottle, annulus and Mbius strip.
If these amplitudes are obtained from two AE symmetry-preserving boundaries/crosscaps,
they must be expressible in terms of AE characters; conversely if they cannot be expressed
in terms of AE characters, at least one of the boundary/crosscap states must be symmetry
breaking. Since the amplitudes are T-duality invariant, but the automorphism type flips, it
follows that also amplitudes obtained with two boundaries/crosscap of automorphism type
must be expressible in terms of AE characters, since this is the case in the T-dual theory.
To summarize, an amplitude can be written in terms of characters of the extension if and
only if the two boundaries/crosscaps are of the same automorphism type.
In the previous section we observed that in the mixed annuli only twisted sector fields
(Q(i) 6= 0) contribute. On the other hand, both Klein bottles K + and K as well the other
annuli can be expressed in terms of AE characters. It is not hard to show that also the

and M[i+1 ] cannot be written in terms of characters of


mixed Mbius strips M[i0 ] , M[f,]
the extension.
4.1. Twisted Klein bottles

Before we discuss T-duality for the Klein bottles, we have to introduce g-twisted Klein
bottles. When g is a symmetry of a closed string theory that commutes with world-sheet
parity , we can divide out [27] the group (1, g). At the level of partition functions, we
have to add a (halved) g-twisted Klein bottle to the (halved) torus. Only eigenstates of g
appear in the g-twisted Klein bottle, that is, the Klein bottle coefficients satisfy a g-twisted
positivity and integrality condition
|KI | = ZI,gI

(21)

with the modular invariant Z. What kind of Klein bottles did we find in the last section?
From Appendix A.1, we know

+
K = CI,I .
K = CI I ,
(22)
I
I
When we compare with Eq. (21), we come to the following conclusion: K + is an untwisted
Klein bottle from the point of view of the charge conjugation invariant Z c and a -twisted
Klein bottle for the invariant Z c . For K it is the other way around: it is an -twisted
Klein bottle from the point of view of the charge conjugation invariant and an untwisted
Klein bottle for the invariant Z c . Note that the operations and are T-dual.
4.2. Symmetry breaking crosscaps
We have observed above that K + is the standard Klein bottle for the charge conjugation
invariant. Indeed, from Eq. (A.3) of the appendix, it follows that the corresponding cross
cap coefficient is just that of the C-diagonal case I+ = P0I / S0I . This coefficient is
generically non-vanishing for all I , so the corresponding automorphism type g has to
satisfy ZI,gI c = 1 for all I . Therefore g = 1 and thus + must have trivial automorphism

L.R. Huiszoon, A.N. Schellekens / Nuclear Physics B 584 [PM] (2000) 705718

713

type from the point of view of the charge conjugation invariant Z c . Hence it has
automorphism type for the invariant Z c .
Since only one of the Mbius amplitudes M + and M for a given boundary label can
be written in terms of characters of the AE theory, it follows that + and must have
opposite automorphism types. So I is trivial for Z c and of automorphism type for
the invariant Z c . An important check on this interpretation is the fact that the AE -Ishibashi
states [f, ] are not present in the theory if one uses the Z c modular invariant. Hence the
AE crosscap should vanish for the corresponding transverse channel labels. Indeed, one
can show that I vanishes identically for fixed points.
4.3. Symmetry breaking boundaries
Given the symmetry properties of the crosscaps and those of the Mbius strip, one can
now read off those of the boundary coefficients. In agreement with [16] we find that from
the point of view of a charge conjugation invariant Z c , the boundary coefficients (3) have
the following automorphism types. When the charge of the boundary label is zero, the
boundaries leave the AE algebra invariant; they are of trivial automorphism type. Charged
boundaries [i1 ] are of automorphism type ; they only leave the orbifold subalgebra
A invariant. From the point of view of Z c , the automorphism types of the boundary
conditions are reversed: the chargeless boundaries break the symmetry, whereas the
charged boundaries do not. This is of course nothing but a reformulation of the wellknown fact [27] that Dirichlet (automorphism type ) and Neumann (trivial automorphism)
boundary conditions are interchanged under T-duality. A similar check can be made as
in the last paragraph of the previous section. If we use the Z c modular invariant, the
symmetric boundary coefficients B(f,)[i1 ] should vanish if f is a fixed point. This is
indeed true, as a consequence of the fact that S vanishes between fixed points and charged
fields.
4.4. Open descendants of diagonal invariants
Let us first conclude: there are two types of inequivalent open descendants for the charge
conjugation invariant Z c . In the first one, we project on invariant states with the Klein
bottle K + . Via the channel transformation, this leads to crosscaps that preserve the full bulk
symmetry. The open sector has two kinds of boundary conditions; chargeless boundary
conditions have trivial automorphism and charged boundaries have automorphism type .
A second descendant can be constructed when we project on invariant states with a
-twisted Klein bottle K that satisfies a twisted positivity and integrality condition. The
corresponding crosscaps have automorphism type . The open sector has again two kinds
of boundary conditions; chargeless boundary conditions with trivial automorphism and
charged boundaries with automorphism type . In order to satisfy positivity and integrality
of the open sector, some boundary coefficients differ by a factor i relative to those of the
untwisted Klein bottle projection as explained in the previous sector.

714

L.R. Huiszoon, A.N. Schellekens / Nuclear Physics B 584 [PM] (2000) 705718

We could equally well have started with the invariant Z c . The automorphism types
of boundaries and crosscaps are opposite to those of the charge conjugation invariant.
T-duality relates both invariants and also the corresponding open descendants.
As a specific example, take = C, i.e., charge conjugation. We can now construct the
open descendants of a diagonal invariant ZI J = I J with crosscaps and boundaries of
trivial automorphism type. By T-duality, this is equivalent to a charge conjugation invariant
with crosscaps and boundaries of automorphism type C. So we have to take the Klein
bottle K and put the CP-factors of the chargeless boundaries to zero. The standard Cardy
case [4,11], i.e., a charge conjugation invariant with trivial crosscaps and boundaries, can
be obtained in a similar way: take K + and put the CP-factors of the charged boundaries to
zero.

Acknowledgements
We would like to thank N. Sousa, J. Fuchs and C. Schweigert for useful discussions.
L.H. would like to thank the Samenwerkingsverband Mathematische Fysica for financial
support.

Appendix A. Positivity and integrality


Let us first relate the fusion and Y-fusion coefficients of the AE theory to those of the
orbifold A. Let us first be a bit more general, and allow the simple current group to be G.
We denote a generic field of the AE theory by [i, i ]. Fields in the A theory are denoted
by i, j . We will not add the superscript E to quantities of the AE theory, since the indices
attached to these quantities make the formulas unambiguous. The S-matrix of the extension
is given by [23,24]
|G| X
(A.1)
i (J )j (J ) SijJ .
S[i,i ][j,j ] =
|Si ||Sj |
This gives the fusion coefficients via the Verlinde formula

X X S[m,m ][j1 ,j1 ] S[m,m ][j2 ,j2 ] S[m,


[j3 ,j ]
m ][j3 ,j3 ]
.
N[j1 ,j ][j2 ,j ] 3 =
1
2
S[m,m ][0]

(A.2)

[m] m

The P-matrix of the extension is [14]


|G| X
i (J )j (J ) PijJ ,
P[i,i ][j,j ] =
|Si ||Sj |

(A.3)

where the sum is over the intersection Si Sj and where


1 X i[hi hKi ] J
e
PKi,j ,
PijJ =
|G|

(A.4)

and where the sum is now over all K and where

P J = T S J T 2S J T .

(A.5)

L.R. Huiszoon, A.N. Schellekens / Nuclear Physics B 584 [PM] (2000) 705718

715

The Y-fusion coefficients are given by [12]


[j3 ,j ]
Y[j1 ,j ][j2 ,j ] 3
1
2

X X S[m,m ][j1 ,j1 ] P[m,m ][j2 ,j2 ] P[m,


m ][j3 ,j

S[m,m ][0]

[m] m

(A.6)

Recall [29,30] that Yi00 is the FrobeniusSchur indicator of a field i in a conformal


field theory; its value is (minus) one for (pseudo) real fields and zero for complex fields.
Furthermore, the tensor Y is integral and satisfies a positivity and integrality relation
with the fusion coefficients [20,29,30]:
j
j
j
Y 6 N j ,
Yi0 = Nii mod 2,
(A.7)
ii
i0
which plays a crucial rle in all proofs that follow.
A.1. Positivity and integrality of the closed sector
In this subsection, we prove that the Klein bottle coefficients, given by
Ki = Yi00 Yi0J ,
satisfy

K 6 Z J ,
i

ii

 = eihJ ,

Ki = ZiiJ mod 2.

(A.8)

(A.9)

From the torus (1) we see that there are three kinds of fields that appear on the diagonal:
self-conjugate fixed points f = f c , self-conjugate i0 = i0c and fields that satisfy i0 = J i0c .
We first concentrate on the fixed points. From (A.2) and (A.6) we find

1
1
N0ff N 0ff ,
(A.10)
Y[f,][0][0] = (Yf 00 + Yf 0J ).
2
2
We can distinguish three situations:
[f, ]c = [f, ]. So Y[f,][0][0] = 1 which implies Yf 00 = Yf 0J = 1. The
corresponding Klein bottles (A.8) therefore satisfy

+
K = 0,
K = 2,
(A.11)
f
f
N[0][f,][f,] =

which satisfies (A.9) since f = f c .


[f, ]c = [f, ]. So N[0][f,][f,] = 1 which implies N0ff = N 0ff = 1 and f
is self-conjugate. Furthermore, since Y[f,][0][0] = 0 we have Yf 00 = Yf 0J = 1.
The corresponding Klein bottles (A.8) therefore satisfy

+
K = 2,
K = 0,
(A.12)
f
f
which satisfies (A.9) as well.
[f, ]c 6= [f, ] and [f, ]c 6= [f, ]. In this case f 6= f c and both Klein bottles
vanish in agreement with (A.9).
Now we turn to the fields i0 . Eqs. (A.2) and (A.6) give
N[0][i0 ][i0 ] = N0i0 i0 + NJ i0 i0 ,

Y[i0 ][0][0] = Yi0 00 + Yi0 0J .

(A.13)

716

L.R. Huiszoon, A.N. Schellekens / Nuclear Physics B 584 [PM] (2000) 705718

There are now two different cases:


[i0 ]c = [i0 ]. So N[0][i0 ][i0 ] = 1 and either i0 = i0c or i0 = J i0c . When i0c = i0 , Yi0 00 = 1
and Yf 0J = 0 and when i0 = J i0c it is the other way around. In any case

K = 1.
(A.14)
i0
[i0 ]c 6= [i0 ]. So N[0][i0 ][i0 ] = 0 which implies i0 6= i0c and i0 6= J i0c . Both Klein bottles
vanish for these fields, in agreement with (A.9).
So the Klein bottle of Section 3 satisfies positivity and integrality. In Section 4, we regard
the Klein bottles as projections for the theory described by AE . Note that we have to
be careful in case of fixed points. Since one fixed point of the orbifold theory resolves
into two fields [f, ] of the extension, the same happens for the corresponding Klein
bottle coefficients; the coefficient Kf = 2 splits into two coefficients K[f,] = 1
and K[f,] = 1. We will assume that [f, ] and [f, ] have the same Klein bottle
coefficient, so that a coefficient Kf = 0 in the orbifold theory cannot split in a K[f,] = 1
and K[f,] = 1, for instance. This is required by the Klein bottle constraint [11,30],
which forbids [f, ] and [f, ] to have opposite Klein bottle coefficients when [f, ]c =
[f, ], a situation that occurs generically.
From the above analysis, it follows that the Klein bottles satisfy

+
K = CI,I ,
K = CI I ,
(A.15)
I
I
where I is a generic field in the AE theory.
A.2. Positivity and integrality of the open sector
In this section, we prove that the two pairs of Mbius and annulus coefficients from
Section 3 satisfy


M
(A.16)
[, ]i 6 A[, ][, ]i and M[, ]i = A[, ][, ]i mod 2,
for all boundary labels [, ] and all fields i. The annuli A are not defined explicitly
in the main text. By A+ we denote the annulus that corresponds to the boundary
coefficient (3) and by A the annulus of the modified boundary coefficient. It differs from
A+ by a relative minus sign when both boundary labels are fixed points.
For the boundary labels [i] = [i0 ], [i1 ], Eq. (A.16) follows immediately from equation (A.7). For the fixed point boundary labels, we have to prove

1
1
|Yf 0i Yf J i | 6 Niff N iff ,
(A.17)
2
2

1
1
(Yf 0i Yf J i ) = Niff N iff mod 2,
(A.18)
2
2
where i = i0 , g. In order to prove these relations, it is convenient to do a similar trick as
was done in [14]. So we first tensor the A theory with a theory A that has an order two
integer spin simple current J and fixed points f. Let us denote the fields of the tensor
The (Y -)fusion rules of this theory are simply
theory At by I = (i, i).
t

N(i,

= Nij k Ni jk ,
i)(j,
j)(k,k)

Y(i,

= Yij k Yi jk .
i)(j,
j)(k,
k)

(A.19)

L.R. Huiszoon, A.N. Schellekens / Nuclear Physics B 584 [PM] (2000) 705718

717

The tensor theory has an order two integer spin simple current (J, J) by which we can
extend to a theory with chiral algebra AE,t . The fields in this extension are the chargeless
+ (J i, Ji)]
with QJ (i) = Q (i)
and
orbits of the tensor theory, denoted by [I0 ] = [(i, i)
J
the resolved fixed points [f, ]. The (Y -)fusion rules for this extension are related to those
of the tensor theory as in equation (A.2) and (A.6). With the use of (A.19), we can then
relate the coefficients of AE,t and A. Consider

E,t
E,t
E,t
E,t

Y
Y[f,][0][I
(A.20)
[f,][0][I ] 6 N[f,][f,][I ] ,
] = N[f,][f,][I ] mod 2,
where i = i0 , g. In
which holds by Eq. (A.7). In this equation, F = (f, f) and I = (i, 0),

terms of quantities of the A and A theory, the above conditions become



1
1
Yf 0i Yf0 0 + J,J Yf J i YfJ0 6 Nff i N ff0 + N ff i N ff0 ,
(A.21)
2
2


 1
1
Yf 0i Yf0 0 + J,J Yf J i YfJ0 = Nff i N ff0 + N ff i N ff0 mod 2,
(A.22)
2
2
where J,J = ei[hJ +hJ ] . Now we use for the A theory B2 level 2k. This theory has an order
= 1 and Y
= Y
= N
=
two, spin k simple current and a fixed point with N
ff0

f0 0

f0 J

ff0

(1)k . So Eq. (A.17) with the (plus) minus sign follows by taking k (even) odd.

References
[1] A. Sagnotti, in: G. Mack et al. (Eds.), Non-Perturbative Quantum Field Theory, Pergamon Press,
1988, p. 521.
[2] M. Bianchi, A. Sagnotti, On the systematics of open-string theories, Phys. Lett. B 247 (1990)
517.
[3] M. Bianchi, A. Sagnotti, Twist symmetry and open-string Wilson lines, Nucl. Phys. B 361
(1991) 519.
[4] J. Cardy, Boundary conditions, fusion rules and the Verlinde formula, Nucl. Phys. B 324 (1989)
581.
[5] D. Lewellen, Sewing constraints for conformal field theories on surfaces with boundaries, Nucl.
Phys. B 372 (1992) 654.
[6] D. Fioravanti, G. Pradisi, A. Sagnotti, Sewing constraints and non-orientable open strings, Phys.
Lett. B 321 (1994) 349.
[7] G. Pradisi, A. Sagnotti, Ya.S. Stanev, The open descendants of non-diagonal SU(2) WZW
models, hep-th/9506014; Phys. Lett. B 356 (1995) 230.
[8] G. Pradisi, A. Sagnotti, Ya.S. Stanev, Completeness conditions for boundary operators in 2D
conformal field theory, hep-th/9603097; Phys. Lett. B 381 (1996) 97.
[9] R. Behrend, P. Pearce, V. Petkova, J.-B. Zuber, On the classification of bulk and boundary
conformal field theories, hep-th/9809097; Phys. Lett. B 444 (1998) 163.
[10] A.N. Schellekens, Open strings, simple currents and fixed points, hep-th/0001198.
[11] A. Sagnotti, Ya.S. Stanev, Open descendants in conformal field theory, hep-th/9605042;
Fortsch. Phys. 44 (1996) 585; Nucl. Phys. Proc. Suppl. B 55 (1997) 200.
[12] G. Pradisi, A. Sagnotti, Ya.S. Stanev, Planar duality in SU(2) WZW models, hep-th/9503207;
Phys. Lett. B 354 (1995) 279.
[13] J. Fuchs, C. Schweigert, A classifying algebra for boundary conditions, Phys. Lett. B 414 (1997)
251.

718

L.R. Huiszoon, A.N. Schellekens / Nuclear Physics B 584 [PM] (2000) 705718

[14] L.R. Huiszoon, A.N. Schellekens, N. Sousa, Open descendants of non-diagonal invariants, hepth/9911229, to appear in Nucl. Phys.
[15] J. Fuchs, C. Schweigert, Branes: from free fields to general backgrounds, hep-th/9712257, Nucl.
Phys. B 530 (1998) 99.
[16] J. Fuchs, C. Schweigert, Orbifold analysis of broken bulk symmetries, hep-th/9811211, Phys.
Lett. B 447 (1999) 266.
[17] J. Fuchs, C. Schweigert, Symmetry breaking boundaries 1, hep-th/9902132; Nucl. Phys. B 558
(1999) 419.
[18] J. Fuchs, C. Schweigert, Symmetry breaking boundaries 2, hep-th/9908025.
[19] L. Birke, J. Fuchs, C. Schweigert, Symmetry breaking boundary conditions and WZW orbifolds,
hep-th/9905038.
[20] L. Borisov, M.B. Halpern, C. Schweigert, Systematic approach to cyclic orbifolds, hepth/9701061; Int. J. Mod. Phys. A 13 (1998) 125.
[21] A.N. Schellekens, S. Yankielowicz, Simple currents, modular invariants and fixed points, Int. J.
Mod. Phys. A 5 (15) (1990) 2903.
[22] K. Intriligator, Bonus symmetry in conformal field theory, Nucl. Phys. B 332 (1990) 541.
[23] A.N. Schellekens, S. Yankielowicz, Field identification fixed points in the coset construction,
Nucl. Phys. B 334 (1990) 67.
[24] J. Fuchs, A.N. Schellekens, C. Schweigert, A matrix S for all simple current extensions, Nucl.
Phys. B 473 (1996) 323.
[25] A.N. Schellekens, Fixed point resolution in extended WZW models, Nucl. Phys. B 558 [PM]
(1999) 484.
[26] A. Recknagel, V. Schomerus, D-branes in Gepner models, hep-th/9712186; Nucl. Phys. B 531
(1998) 185.
[27] J. Polchinski, Tasi lectures on D-branes, hep-th/9611050.
[28] E. Verlinde, Fusion rules and modular transformations in 2D conformal field theory, Nucl.
Phys. B 300 (1988) 360.
[29] P. Bantay, FrobeniusSchur indicator in conformal field theory, Phys. Lett. B 394 (1997) 87.
[30] L.R. Huiszoon, A.N. Schellekens, N. Sousa, Klein bottles and simple currents, hep-th/9909114;
Phys. Let. B 470 (1999) 95.

Nuclear Physics B 584 [PM] (2000) 719748


www.elsevier.nl/locate/npe

Conformal tension in string theories and M-theory


Manuel Barros a , Angel Ferrndez b , Pascual Lucas b,
a Departamento de Geometra y Topologa, Universidad de Granada, 18071 Granada, Spain
b Departamento de Matemticas, Universidad de Murcia, 30100 Espinardo, Murcia, Spain

Received 25 January 2000; accepted 26 May 2000

Dedicated to Professor A.M. Naveia, el maestro, on the occasion of his 60th birthday

Abstract
This paper deals with string theories and M-theories on backgrounds of the form AdS M,
M being a compact principal U (1)-bundle. These configurations are the natural settings to study Hopf
T-dualities (Duff et al., Nucl. Phys. B 544 (1999) 145), and so to define duality chains connecting
different string theories and M-theories. There is an increasing great interest in studying those
properties (physical or geometrical) which are preserved along the duality chains. For example,
it is known that Hopf T-dualities preserve the black hole entropies (Duff et al., Nucl. Phys. B
544 (1999) 145). In this paper we consider a two-parameter family of actions which constitutes a
natural variation of the conformal total tension action (also known as WillmoreChen functional in
differential geometry). Then, we show that the existence of wide families of solutions (in particular
compact solutions) for the corresponding motion equations is preserved along those duality chains.
In particular, we exhibit ample classes of WillmoreChen submanifolds with a reasonable degree of
symmetry in a wide variety of conformal string theories and conformal M-theories, that in addition
are solutions of a second variational problem known as the area-volume isoperimetric problem.
These are good reasons to refer those submanifolds as the best worlds one can find in a conformal
universe. The method we use to obtain this invariant under Hopf T-dualities is based on the principle
of symmetric criticality. However, it is used in a two-fold sense. First to break symmetry and so
to reduce variables. Second to gain rigidity in direct approaches to integrate the EulerLagrange
equations. The existence of generalized elastic curves is also important in the explicit exhibition
of those configurations. The relationship between solutions and elasticae can be regarded as a
holographic property. 2000 Elsevier Science B.V. All rights reserved.
PACS: 02.40.Ky; 03.40.Dz; 04.50.+h; 04.65.+e; 11.10.Kk; 11.17.+y; 11.30.-j
Keywords: Anti-de Sitter space; String theories; M-theories; T-dualities; Conformal total tension actions;
Generalized elasticae

Corresponding author. E-mail: plucas@um.es; tel.: +34-68-364182

E-mail addresses: mbarros@ugr.es (M. Barros), aferr@um.es (A. Ferrndez).


0550-3213/00/$ see front matter 2000 Elsevier Science B.V. All rights reserved.
PII: S 0 5 5 0 - 3 2 1 3 ( 0 0 ) 0 0 3 5 9 - X

720

M. Barros et al. / Nuclear Physics B 584 [PM] (2000) 719748

1. Introduction and set up


There is an increasing interest in the interplay between bulk and boundary dynamics.
The two main directions being explored presently are:
(1) The conjecture of Maldacena [30], that concerns with the string theory or Mtheory on certain backgrounds of the form AdSp MDp . Here, AdSp is the antide Sitter space of dimension p and MDp is a compact space with dimension D p.
Now, D is 10 or 11 depending on whether we are doing string theory or M-theory,
respectively. This conjecture postulates that the quantum string or the M-theory
on this background is mathematically equivalent to an ordinary but conformally
invariant quantum field theory in a space time of dimension p 1 which plays the
part of boundary of AdSp . This relationship is known as AdS/CFT correspondence.
(2) The holographic hypothesis (see [24] and references therein) states that all
information of a theory in the bulk of a bounded region is available, in some sense,
on the boundary of the region. However, the above some sense means that, as far
as we know, there are at least four possible non-equivalent definitions of what
holography may mean. For instance, the AdS/CFT correspondence can be viewed
as an example of the holographic hypothesis [41].
Consequently, the AdS/CFT correspondence has revived a great interest in gauged extended supergravities which arise as the massless sector of the KaluzaKlein compactifications of D = 11 supergravity, such as AdS4 S7 and AdS7 S4 , as well as in type
IIB supergravity, such as AdS5 S5 (see, for example, [12,14,31]). In [14], Duff et al.
considered that S5 is a U (1) bundle over CP2 (by means of the usual Hopf map) to define an unconventional type of T-duality. This is applied to construct the duality chain
[n = 4 YangMills][type IIB string on AdS5 S5 ][type IIA string on AdS5 CP2
S1 ][M-theory on AdS5 CP2 T 2 ]. This kind of duality can be extended to more general contexts including backgrounds as AdS M, where M are Einstein spaces that are
not necessarily round spheres but still U (1) fibrations. Notice that these spaces naturally
appear as the near horizon geometries of supermembranes with fewer Killing spinors [16].
Hopf T-duality drastically changes the topology of the compactification space S5 by
untwisting it to CP2 S1 . However, it preserves the black hole entropies. The purpose
of this paper is to show an interesting property which remains invariant along the Hopf
T-duality chain. In fact, we prove the existence of rational one-parameter families of
WillmoreChen submanifolds (i.e., critical points of the conformal total tension action) at
any stage of the duality chain. This is done in Sections 4 and 5. The case of type IIB theory
on the background AdS5 S5 is particularly interesting. The high rigidity of the round five
sphere allows us to get a wide variety of solutions, having three quite different families of
them. The first one contains solutions which are obtained via a standard use of the principle
of symmetric criticality [36], when lifting elasticae in the complex projective plane CP2
by the usual Hopf mapping. In particular, many properties of these solutions are reflected,
by a kind of holographic principle, in those elasticae (for example, they have constant
mean curvature if and only if they come from elasticae which have constant curvature
in CP2 ). Other two classes of configurations are obtained by exploiting the principle of

M. Barros et al. / Nuclear Physics B 584 [PM] (2000) 719748

721

symmetric criticality in a different approach, which is used to gain rigidity when integrating
the corresponding motion equations. Therefore, we show that minimal flat tori are always
solutions, which seems reasonable and however it is not an immediate result. This allows
us to give a second family of explicit solutions by considering results of [26,27]. We still
have a third family by using the Chen finite type theory of submanifolds [3,11]. Namely,
we get explicit solutions constructed in the corresponding Euclidean space through the
eigenfunctions of the Laplacian associated with exactly two different eigenvalues. Of
course these three classes of solutions can be projected down to any 5-dimensional lens
space.
The above stated results are extended in Section 6 to other configurations of the form
AdS5 M5 , where M5 is a principal U (1)-bundle over a compact 4-dimensional manifold
N which admits an Einstein metric with positive scalar curvature. These U (1)-bundles are
classified by the cohomology group H 2 (N, Z). Therefore, as an illustration we consider
the Stiefel manifold regarded as the principal U (1)-bundle on the complex quadric N =
S2 S2 associated with a suitable multiple of its first Chern class.
In Section 7, we investigate the above invariant along the Hopf T-duality chains
generated by type IIB theories on backgrounds carrying both NSNS and RR electric and
magnetic 3-form charges, and whose near horizon geometry contains AdS3 S3 . This case
is richer than those we have considered above where only RR charges were supported. In
this case, we show the existence of ample classes of solutions along any duality chain. In
particular, we prove:
The conformal structure associated with the string-frame in the near horizon limit of
any non dilatonic black hole in D = 5 and D = 4 admits a rational one-parameter class
of 4-dimensional WillmoreChen configurations that have constant mean curvature in the
original string metric. However, the degree of symmetry of these solutions is preserved only
when NSNS charges appear while that is decreased when RR charges are carried.
All these results are preceded by a symmetry breaking method which is developed
through two general settings in Section 2. The algorithm we use is based on the principle of
symmetric criticality, rather in a formulation of this principle due to Palais [36]. Certainly
this method can be applied to other backgrounds different from those considered in this
paper, and also it is open to be extended to other contexts. For example, we do not know
if the statement of Theorem 4 works with no assumption on the flatness of the concerned
gauge potential. The paper is completed with Sections 3 and 8. The former is dedicated
to exploit the nice geometry of AdS3 in order to obtain wide classes of elastic helices
(in particular closed elastic helices, see Theorem 12) which generate solutions of the
motion equations in any configuration of the form AdS3 M (Corollaries 14 and 15).
In Section 8, we show how our results can be extended to higher-dimensional theories,
including F-theory.
Now we give here a generic formulation of our method.
Let I (Q, (L, ds 2 )) be the space of immersions of a compact manifold Q in a pseudoRiemannian manifold (L, ds 2 ). Define a real two-parameter family of actions


War : I (Q, (L, ds 2 )) R : a, r R ,

722

M. Barros et al. / Nuclear Physics B 584 [PM] (2000) 719748

by

Z
War () =

2 ae

r/2

dv,

for all I (Q, (L, ds 2 )), where and e stand for the mean curvature and the extrinsic
scalar curvature functions of , respectively. Also dv is the volume element of (ds 2 )
on Q. It should be noticed that when a = 1 and r is the dimension of Q, then we get
the WillmoreChen functional, which provides a variational problem of great interest
in differential geometry, due in part to its invariance under conformal changes of the
surrounding metric ds 2 , also known as the conformal total tension action. Moreover,
if r = 2, we have the Willmore functional which formally coincides with the Canham
Helfrich bending energy of fluid membranes and lipid vesicles [8,23], and amazingly also
with the Polyakov extrinsic action in the bosonic string theory [37]. It should be also
observed that if the dimension of Q is one (i.e., we are talking about immersed curves),
then e vanishes identically and is nothing but the curvature function of the curve in
family of actions reduces to a one-parameter one of elastic
(L, ds 2 ). In this case the above
R
energy functionals, Fr = r ds, acting on the space of closed curves in (L, ds 2 ).
Now, all these actions naturally appear from the own WillmoreChen functional. This
may be showed via an interesting argument, involving the principle of symmetric criticality
[36], as well as the KaluzaKlein inverse mechanism, which allows us to obtain the
above actions by a reduction of symmetry process from the WillmoreChen action in the
conformal KaluzaKlein ansatz.

2. Symmetry breaking phase transitions


Let be a principal connection on a principal fibre bundle P (M, G), G being an
m-dimensional, compact Lie group endowed with a bi-invariant metric d 2 . We denote
the spaces of pseudo-Riemannian metrics on M and P , respectively. For
by M and M
be the mapping defined by
{1, +1}, let : M M
(g) = (g) + (d 2 ),
where : P M stands for the projection map of the principal fibre bundle.
is called a KaluzaKlein metric on P if it belongs
A pseudo-Riemannian metric g M
to the image of . These metrics are also known as bundle like metrics and they are
the natural ones working on a unified theory, collecting the gravitation g with the gauge
potential , in the sense of KaluzaKlein.
It is obvious that : (P , g)
(M, g) is a pseudo-Riemannian submersion whose leaves
are the fibres, and so they are diffeomorphic to the structure group G. It is also evident that
the natural action of G on P is carried out by isometries of (P , g).

Let V be the vertical distribution of the pseudo-Riemannian submersion and H,


the horizontal one, its complementary g-orthogonal

distribution. Note that while V


is involutive, whose leaves are the fibres, H is not integrable, in general. Let N be

M. Barros et al. / Nuclear Physics B 584 [PM] (2000) 719748

723

a q-dimensional submanifold in (M, g), then one can define a distribution D along 1 (N)
p , so D is an (m + q)by D p = (dp )1 (T(p)
N), for any p P . Notice that Vp D
dimensional distribution which is integrable. Furthermore N = 1 (N) is a G-invariant
submanifold of dimension m + q. The converse is also true. Assume N is an (m + q)dimensional submanifold of (P , G), which is G-invariant. Then N is foliated by the fibres
and so the horizontal vectors tangent to N define a q-dimensional distribution in N . Notice
that this is not involutive though it projects down to an integrable distribution D in (M, g).
Hence, N = 1 (N), where N is a leaf of D. Summarizing we have:
Lemma 1. Let N be a q-dimensional submanifold of (M, g), then N = 1 (N) is an
(m + q)-dimensional submanifold of (P , g)
which is G-invariant. Conversely, if N is a
d-dimensional submanifold of (P , g)
which is G-invariant, then N = 1 (N), where N is
a (d m)-dimensional submanifold of (M, g).

From now on, we will denote by overbars the lifts of corresponding objects on the base.
In particular, X will denote the horizontal lift of a vector field X on M.
Proposition 2. Let and be the second fundamental forms on N in (M, g) and N =
respectively. Denote by and the corresponding mean curvature
1 (N) in (P , g),
functions, then we have
1. (V , W ) = 0, for vertical vector fields V and W .
Y ) = (X, Y ).
2. (X,
q2
2
3. 2 = (m+q)
2 .
V ) = 0.
4. If is flat, then (X,

Proof. To show the first claim, just notice that the fibres are totally geodesic not only in
(P , g),
but also in N , with respect to the g-induced

metric. To prove the second statement

observe that (X, Y ) is horizontal. The third one is now clear. The last assumption is also
evident because the ONeill invariant, which measures the obstruction to the integrability
of the horizontal distribution, vanishes identically if is assumed to be flat. 2
Proposition 3. Let e and e be the extrinsic scalar curvature functions of N in (M, g)
respectively. If is flat, then we have
and N = 1 (N) in (P , g),
q(q 1)
e .
e =
(m + q)(m + q 1)
Proof. Let {Va ; 1 6 a 6 m} be the fundamental vector fields in P associated with a frame
of unit left-invariant vector fields in (G, d 2 ). Let {Xi ; 1 6 i 6 q} be a local orthonormal
frame on (N, g 0 ), g 0 being the g-induced metric on N . Then we have

724

M. Barros et al. / Nuclear Physics B 584 [PM] (2000) 719748

1
e =
(m + q)(m + q 1)

"

q
X

X i , X j )
K 0 (X i , X j ) K(

i,j =1
q X
m
X

X i , Va )
K 0 (X i , Va ) K(

i=1 a=1
m
X

b , Va )
K 0 (Vb , Va ) K(V

#


a,b=1

and N with respect


where K and K 0 stand for the sectional curvature functions of (P , g)
to the g-induced

metric. Now, we use Proposition 2 jointly with the Gauss equations of N


in (P , g)
and N in (M, g), respectively, to get the result. 2
= {
For a compact, q-dimensional manifold N , we write IG (N G, (P , g))
is a
I (N G, (P , g))
: is G-invariant}. Since G is compact, then IG (N G, (P , g))
submanifold of I (N G, (P , g)),
which according to Lemma 1 can be identified with
if and only if there exists I (N, (M, g))
I (N, (M, g)). That is, IG (N G, (P , g))

such that (N
G) = 1 ((N)). Then we use Propositions 2 and 3 to obtain
r

q

vol(G, d 2 )Wbr (),


(1)
War () =
m+q
where b = (q1)(m+q)
q(m+q1) a.
ar
the sets of critical points
Now, given (a, r) R2 we denote by ar , ar and G
and IG (N G, (P , g)),

of the functional War on I (N, (M, g)), I (N G, (P , g))


respectively. In other words, they are the sets of configurations which are solutions of the
and IG (N
motion equations for the War -dynamics on I (N, (M, g)), I (N G, (P , g))
G, (P , g)),
respectively. Then the principle of symmetric criticality [36] assures us that
ar
= ar IG (N G, (P , g)).

(2)

which do not break


Said otherwise, to obtain critical points of War on I (N G, (P , g))
the natural G-symmetry of the problem, we only need to compute War restricted to
Since we have already computed this restriction (see (1)), then we
IG (N G, (P , g)).
obtain the following result, which can be regarded as a criterion for reduction of variables
(in the sense of Palais) in the variational problems associated with the functionals War .
Theorem 4. ar if and only if br , with b =

(q1)(m+q)
q(m+q1) a.

Remark 5. Roughly speaking, the last result reduces the search for symmetric solutions of
to that for solutions of the Wbr -dynamics in
the War -dynamics in a unified theory (P , g)
its gravitatory component (M, g), a and b being related as above. It should be noticed,
in addition, that the best worlds to live in a conformal unified universe (P , [g]),
[g]

denoting the conformal structure associated with g,


are the WillmoreChen submanifolds
which preserve the internal symmetry. We use the term best in a two-fold sense. First,
because they preserve the above mentioned symmetry, and secondly because they support

M. Barros et al. / Nuclear Physics B 584 [PM] (2000) 719748

725

the smallest global tension possible from the surrounding universe, and consequently
they must be WillmoreChen submanifolds. For these worlds, we obtain the following
gravitatory characterization.
Corollary 6. 1,m+q , that is, is a G-invariant WillmoreChen submanifold in
(P , [g])
if and only if b,m+q , with b = (q1)(m+q)
q(m+q1) .
It should be pointed out that the constancy of the mean curvature (tension) is preserved
in this symmetry breaking phenomenon, which can be viewed as an interesting holographic
property when we reflect that constancy in the gravitatory component. Furthermore, this
provides another reason to use the term best to name those solutions, since they are also
solutions of the well known area-volume isoperimetric problem.
Next we exhibit a second framework where one can reduce symmetry. The background
is globally a product and so it could be considered as a particular case of the above one, for
example when M is simply connected. However, the metric (or the conformal structure)
involves a warping function and the fibre part is provided by a compact homogeneous
space. Therefore it can be regarded as a complementary of the first stated setting.
To describe this new situation let (M1 , g1 ) and (M2 , g2 ) be a compact homogeneous
space with group of isometries H and a pseudo-Riemannian manifold, respectively. We
recall that M = M1 f M2 is endowed with the metric g = g1 + f 2 g2 (with the obvious
meaning), where f is a positive smooth function on M1 . It is usually called the warped
product with warping function f (see [6] for details).
Let S be a d-dimensional submanifold in M2 with second fundamental form h. Then
N = M1 f S is an H -invariant (n1 + d)-dimensional submanifold of M, n1 being the
dimension of N1 . Moreover, every H -invariant submanifold of M is obtained in this way.
e of M is foliated
To do clear this claim, just notice that every H -invariant submanifold N
e} and so it projects down, via the second projection, in a
by leaves {M1 {p}; p N
e n1 )-dimensional submanifold of M2 .
(dim N
The volume form of the g-induced metric on M1 f S is given by
dv = f dv1 ds,
where dv1 and ds stand for the volume forms of (M1 , g1 ) and S with the g2 -induced metric,
respectively. Now we are going to compute the second fundamental form of M1 f S
in M.
Proposition 7. The following statements hold:
1. (X, Y ) = 0, for X and Y tangent to M1 ;
2. (X, V ) = 0, for X tangent to M1 and V tangent to S;
3. (V , W ) = h(V , W ), for V and W tangent to S.
Proof. The first one follows from the fact that the leaves are totally geodesic not only in M,
but also in M1 f S. Let be any vector field normal to M1 f S in M, then it can be
viewed as a normal one to S in M2 . On the other hand, since each fibre {p} M2 is totally

726

M. Barros et al. / Nuclear Physics B 584 [PM] (2000) 719748

umbilical in M, then its second fundamental form satisfies b(V , ) = 0. Consequently


g( (X, V ), ) = 0, which shows the second statement. The last one follows from [33]. 2
Theorem 8. If S is compact, then M1 f S is in br if and only if S is in cr , where
1+d)
c = b (d1)(n
d(n1 +d1) .
Proof. Let H and H2 be the mean curvature vector fields of M1 f S in M and S in M2 ,
respectively. We use Proposition 2 to see that
H=

1
d
H2 ,
n1 + d f 2

and so the corresponding mean curvature functions are related by


2 =

d2
1 2
.
2
(n1 + d) f 2 2

On the other hand, the extrinsic scalar curvature functions and 2 of M1 f S in M and
S in M2 , respectively, are related by
=

1
d(d 1)
2 .
(n1 + d)(n1 + d 1) f 2

Now the restriction of Wbr to the space of H -invariant submanifolds, M1 f S, in M is


given by
!
r Z

d
1
dv1
Wbr (M1 f S) =
n1 + d
f r1
M1


Z 
(d 1)(n1 + d) r/2
2
ds.
2 b

d(n1 + d 1)
S

Once more the principle of symmetric criticality allows us to obtain the statement.

The next result is obtained when S is chosen to be a curve.


Corollary 9. Let be a closed curve in M2 . Then M1 f is in br if and only if is a
critical point of the elastic energy functional F given by
Z
r
F ( ) = r ds,

acting on closed curves in (M2 , g2 ), where denotes the curvature function of .


Remark 10. Closed curves which are critical points of F r are called generalized elasticae
[4]. In particular, those which are critical points of F 2 are called free elasticae [28].

M. Barros et al. / Nuclear Physics B 584 [PM] (2000) 719748

727

3. Some AdS-geometry
The role of the anti-de Sitter (AdS) geometry in the high energy physics increased due
in part to both the Maldacena conjecture and the holographic hypothesis. Furthermore,
AdS-geometry plays a very important role in the theory of higher spin gauge fields where
iterations contain negative powers of the cosmological constant [19]. That theory may be
considered [43] as a candidate for a more symmetric phase of string theory. The group
manifold case AdS3 is of special interest in many subjects (see [38] for some of them,
including the study of conserved currents of arbitrary spin built from massless scalar and
spinor fields in AdS3 ). Next, we study another interesting property emanating from the
geometry of AdS which also corroborates the importance of AdS3 . In fact, we go to
classify all generalized elasticae with constant curvature in AdS, for arbitrary dimension.
To do that, one first uses a standard argument, which involves several integrations by parts,
to compute the EulerLagrange equations associated with the elastic energy functional
F r . An early consequence obtained from these equations is that any generalized elasticae
in AdS must lie fully in some AdS3 totally geodesic in AdS. We can also read from those
equations that a generalized elastica with constant curvature in AdS3 also has constant
torsion and so it is a helix in AdS3 . This is a nice reason to study the geometry of helices
in AdS3 .
The 4-dimensional pseudo-Euclidean space with index 2, R42 , can be identified with
2
C = {z = (z1 , z2 ): z1 , z2 C} endowed with the usual inner product hz, wi = Re(z1 w 1
z2 w 2 ). The 3-dimensional anti-de Siter space is the hyperquadric AdS3 = {z R42 :
hz, zi = 1}, and the induced metric defines a Lorentzian structure, with constant
sectional curvature 1, on AdS3 . The circle of radius one S1 , regarded as the set of unit
complex numbers, acts naturally (multiplication coordinate to coordinate) on AdS3 . The
space of orbits, under this action, can be identified with the hyperbolic 2-plane H2 (4)
of Gaussian curvature 4. The natural projection : AdS3 H2 (4) gives a semiRiemannian submersion.
A global unit timelike
vector field V can be defined on AdS3 by putting Vz = iz, for all
z AdS3 (as usual i = 1 ). The V flow is made up by fibres, which are unit circles with
negative definite metric. We will use the standard notation and terminology of [33], relative
to semi-Riemannian submersions. In particular, one has the splitting Tz = Vz Hz , z
AdS3 , where Tz is the tangent 3-space to AdS3 in z, Vz = span(Vz ) is the vertical line and
Hz is the horizontal subspace (iHz = Hz ). Recall that Vz = ker(dz ) and dz restricted
to Hz gives an isometry between Hz and the tangent plane to H2 (4) at (z). Overbars
will denote the horizontal lifts of corresponding objects on H2 (4). The semi-Riemannian
connections and of AdS3 and H2 (4), respectively, satisfy

(3)
X Y = X Y + go (J X, Y ) V ,

(4)
V = V X = i X,
X

V V = 0,

(5)

here J and go denote the standard complex structure and metric of H2 (4), respectively.
Notice that the third equation gives the geodesic character of the fibres in AdS3 .

728

M. Barros et al. / Nuclear Physics B 584 [PM] (2000) 719748

The mapping : AdS3 H2 (4) is also a principal fibre bundle over H2 (4) with
structure group S1 (a circle bundle). We define a connection on this bundle by assigning to
each z AdS3 the horizontal 2-plane Hz . The Lie algebra u(1) of S1 = U (1) is identified
with R, so V is the fundamental vector field 1 corresponding to 1 u(1).
We denote by and the connection 1-form and the curvature 2-form of this
connection, respectively. It is well known that there is a unique R-valued 2-form on
H2 (4) such that = (). We also put dA to denote the canonical volume form on
H2 (4), in particular dA(X, J X) = 1 for any unit vector field X on H2 (4). It is clear
i X)
and so we can use the structure equation, the horizontality of
that (X, J X) = (X,
and i X and the formula (3) to obtain
X,

i X)
= d(X,
i X)
= [X,
i X]
= 2(V ) = 2,
(X,
and consequently
= 2 dA.

(6)

Let : [0, L] H2 (4) be an immersed curve with length L > 0. We always assume that
is parametrized by the arclength. The complete lift C = 1 () will be called the Hopf
tube associated with and it can be parametrized as follows. We start from a horizontal

lift : [0, L] AdS3 of and then we get all the horizontal lifts of by acting S1 over .
Therefore we have : [0, L] R AdS3 with

(s, t) = eit (s).


It is not difficult to see that C is a Lorentzian flat surface which is isometric to [0, L]
S1 (where the second factor is endowed with its negative definite standard metric). In
particular, if is closed, then C is a Lorentzian flat torus (the Hopf torus associated
with ). It will be embedded in AdS3 if so is in H2 (4), and its isometry type depends
not only on L but also on the area A > 0 in H2 (4) enclosed by .
Theorem 11. Let be a closed immersed curve in H2 (4) of length L and enclosing an
area A. The corresponding Hopf torus C is isometric to L2 / , where is the lattice in
the Lorentzian plane L2 = R21 , generated by (0, 2) and (L, 2A).
Proof. Let be any horizontal lift of and : L2 C AdS3 the semi-Riemannian

The lines parallel to the t axis in L2 are mapped by


covering defined by (s, t) = eit (s).
onto the fibres of , while the lines parallel to the s axis in L2 are mapped by onto the
horizontal lifts of . These curves are not closed because of the holonomy of the involved
connection, which was defined above. However the non closeness of the horizontal lifts of
closed curves is measured just for the curvature as follows (we will apply, without major
details, a well known argument which is nicely exposed in [21, Vol. II, p. 293]). There
exists [, ) such that (L)

= ei (0) for any horizontal lift. The whole group of


deck transformations
of is so generated by the translations (0, 2) and (L, ). Finally we
R
have = c , where c is any 2-chain on H2 (4) with boundary c = . In particular,
from (6), we get = 2A and the proof finishes. 2

M. Barros et al. / Nuclear Physics B 584 [PM] (2000) 719748

729

From now on, we assume that is an arclength parametrized curve with constant
curvature in H2 (4). Then C = 1 () is a Lorentzian flat surface with constant
mean curvature. Moreover it admits an obvious parametrization (s, t) by means of fibres
(s constant) and horizontal lifts of (t constant). Let be a non-null geodesic of C , that
is determined from its slope g (which is measured with respect to ). It is not difficult to
see that is a helix in AdS3 , with curvature and torsion given, respectively, by
( + 2g)
,
1 g2
(1 + g + g 2 )
,
=
1 g2

(7)
(8)

where = 1 represents the causal character of .


We also have a converse of this fact, namely, given any helix in AdS3 with curvature
and torsion , then it can be regarded as a geodesic in a certain Hopf tube of AdS3 . Indeed,
just consider the Hopf tube C = 1 (), where is a curve in H2 (4) with constant
2
2 +1)
, here denotes the causal character of , and then we choose a
curvature = (

geodesic in C with slope g = +


.
We suppose that is closed, that is, it is a geodesic circle of a certain radius r > 0
in H2 (4). Then its curvature is = 2 coth 2r (notice we choose orientation to get
negative values for curvature). The length of is L = sinh 2r and the enclosed area in
H2 (4) is A = 2 (cosh 2r 1). As we already know the Hopf torus C = 1 () comes
from a lattice in L2 which is generated by (0, 2) and (L, 2A). Now a geodesic (s)
of C = 1 () is closed if and only if there exists so > 0 such that 1 ((so )) .
Consequently


A
2
q+
,
(9)
g=
L

where q is a rational number.


The slope of closed helices can be also written in terms of as follows
p
1
g = q 2 4 ,
2

(10)

where q Q {0}.
The EulerLagrange equation for helices in AdS3 of curvature > 0 and torsion 6= 0
being critical points of F r is
(r 1) 2 + r 2 r = 0,

(11)

that is, in the (, ) plane of helices in AdS3 , the action F r has exactly one ellipse of
critical points. To determine the closed helices in AdS3 , which are in the above ellipse, we
use the discussion made above. In particular the EulerLagrange equation can be written
in terms of and the slope g as follows

(12)
2rg 3 + r 2 + 4(2r 1) g 2 + 2(3r 2)g + (r 1) 2 = 0.

730

M. Barros et al. / Nuclear Physics B 584 [PM] (2000) 719748

The following theorem shows the existence of wide families of generalized elasticae in
AdS3 for arbitrary degrees.
Theorem 12. For any non-zero rational number q and an arbitrary natural number r,
there exists a closed helix qr in AdS3 which is a generalized elastica in AdS3 , i.e.,
a critical point of F r .
Proof. We manipulate Eqs. (10) and (12) to see that a closed helix in AdS3 is a critical
point of F r if and only if, regarded as a geodesic of rational slope q in a Hopf torus on
a geodesic circle in H2 (4) with curvature , then both parameters give a zero of the
following function


p
4r 2
2
2
2
.
(13)
F (, q) = r(4q + 1) 4 4q
r
It is not difficult to see that for any non-zero rational number q, there exists a real
number (, 2) such that F (, q) = 0. We choose a geodesic circle in H2 (4)
A
with curvature and a geodesic in C = 1 () whose slope is g = 2
L (q + ), where L
and A are, respectively, the length of and the enclosed area by in H2 (4). Certainly
is a closed helix in AdS3 and its curvature and torsion satisfy the EulerLagrange equation
associated with F r . 2
From now on we will denote by qr the set of closed helices qr in AdS3 obtained in
the last theorem. Notice that these sets give the complete moduli space (up to motions in
AdS3 ) of generalized elasticae with constant curvature in AdS3 .
Remark 13. It is not difficult to deduce from F (, q) = 0 that the relationship between
q and gives the following property. Every q 6= 0 occurs for exactly one , while each
1
determines exactly two values of q, except when 2 = 4r2
r (which corresponds to q = 2
1
1
or q = 2 ). The product of these two values of q is always 4 , therefore when one of the
them is rational the other one must also be rational. Thus the corresponding Hopf tori in
AdS3 has transverse foliations by closed generalized elastic helices.
Now we can combine Corollary 9 with Theorem 12 to obtain a more general existence
result for critical points of the two-parameter family of functionals {War }. However, we
will change a little bit the notation (namely the order in products) to agree with the classical
one used in Physics. The setting can be described as follows. Let (M, g) be a compact
homogeneous space with group of isometries H . On the product space AdS M, we
consider the Lorentzian metric f 2 go + g, where go is the canonical metric on AdS, say
for instance with constant curvature 1, and f is a positive smooth function on M, which
works as a warping function on the above product. We consider the action War acting
on I (Q, (AdS M, f 2 go + g)), here dim Q = dim M + 1, and use the above mentioned
results to get the complete classification of H -invariant critical points of War which have
constant mean curvature according to the following statement.

M. Barros et al. / Nuclear Physics B 584 [PM] (2000) 719748

731

Corollary 14. The class of H -invariant critical points of War which have constant
mean curvature in (AdS M, f 2 go + g) is {qr M: qr qr }. In particular, if dim
M = r 1, the above family gives all the H -invariant, WillmoreChen submanifolds with
dimension r in (AdS M, [f 2 go + g]) that have constant mean curvature relative to
f 2 go + g.
As a first conclusion, we get a rational one-parameter family of solutions of the motion
equations associated with the actions War , at any stage of the Hopf T-duality chain. These
solutions emerge from closed generalized elastic helices in AdS3 and can be regarded as
solitons corresponding to extended dynamical objects obtained when the compactification
space propagates in the target space by describing closed helicoidal orbits. We can also
study, of course, non-closed helicoidal motions. In this case a real one-parameter class of
non congruent solutions may be obtained and the above closeness property is obtained
in terms of the rationality nature of the parameter. However we are interested in closed
solutions which are narrowly related with the AdS-geometry.
Corollary 15. For any couple of real numbers a, r, and any non-zero rational number q
we have
1. qr S5 belongs to ar in the type IIB string on AdS5 S5 .
2. qr CP2 S1 belongs to ar in the type IIA string on AdS5 CP2 S1 .
3. qr CP2 T 2 belongs to ar in the M-theory on AdS5 CP2 T 2 .
Furthermore, these configurations always lie in a codimension two, totally geodesic
submanifold of the corresponding background. Also they are either SO(6)-invariant,
(SU(3) U (1))-invariant or (SU(3) U (1) U (1))-invariant depending on whether we
are in type IIB, type IIA strings or M-theory, respectively.

4. WillmoreChen submanifolds in type IIA string theories and M-theory


The Hopf T-duality has the effect of untwisting S5 to CP2 S1 . This corresponds with
type IIB configurations carrying strictly RR electric and magnetic 5-form charges [14]. In
our case, the compactification spaces are CP2 S1 for type IIA string and CP2 T 2 for
M-theory, and they can be treated according to the settings we have considered in Section 2.
In both cases we naturally break symmetry to study generalized elasticae in the complex
projective plane CP2 endowed with its usual FubiniStudy metric go , with holomorphic
sectional curvature 4.
For a curve in CP2 , one can consider the angle between the complex tangent
plane span{ 0 (s), J 0 (s)} and the osculating plane of , J standing for the usual complex
structure of CP2 . A curve is said to be of constant slant if the angle is constant along .
In [1] the first author gave the complete classification of curves with constant slant in CP2
R
which are critical points of the elastic energy functional ( 2 + )2 ds, where is a certain
constant. The argument used there can be adapted now to get the completeRclassification of
critical points with constant slant for elastic energy functionals of the form ( 2 + )r/2 ds,

732

M. Barros et al. / Nuclear Physics B 584 [PM] (2000) 719748

where r is any real number, in particular generalized elasticae with constant slant in CP2
(see also [5] for another related problem). In the next discussion, the term generalized
elastica will be used to name a critical point of the above types of functionals. Before to
explain the main points in that classification, where the parameters and r are referred as
the potentials, we will exhibit the argument for an arbitrary because it will be used later.
1. First of all, notice that the standard Frenet equations of curves in CP2 are useful, for
example, in defining the concept of helix. However, to study generalized elasticae in CP2
one needs a different reference frame along curves in CP2 which involves the complex
structure J of CP2 . One way to describe this frame is to begin by lifting horizontally the
curve (s) in CP2 , via the usual Hopf mapping, to a curve Y (s) in S5 . The unit tangent
vector field T (s) = 0 (s) lifts to T (s) = Y 0 (s). Now, we may choose a vector field U (s)
along (s) such that its horizontal lifting U (s) gives the third component in a special
unitary frame (s) = {Y (s), T (s), U (s)} in C3 . In other words, (s) is a lift of the curve
(s) to a curve in SU(3). This curve satisfies a natural differential equation which projects
down to CP2 and gives the natural equations of the new frame {T (s), J T (s), U (s), J U (s)}
along (s) (see [1,5] for more details).
2. By computing the EulerLagrange equations for generalized elasticae in terms of the
new frame, one can see that each generalized elastica with constant slant in CP2 is a helix.
3. The main point in the geometrical integration of the motion equations for helices
is the following: every generalized elastic helix in CP2 is the image, under the natural
projection, of a one-parameter subgroup of SU(3).
4. In this framework one can obtain the moduli space, up to congruences in CP2 ,
of generalized elasticae with constant slant in CP2 . This space consists of a real threeparameter family of helices in CP2 where two parameters in this family can be chosen
to be the potentials. Now the closedness characterization for these curves can be obtained
in terms of a rationality condition of the third parameter. Therefore, for any couple of real
: q
numbers r and , the potentials, we obtain a rational one-parameter family r = {rq
Er Q} of generalized elastic closed helices in CP2 , here Er is a certain subset of rational
numbers determined in terms of the potentials.
The above argument, which shows the existence (in particular) of generalized elasticae
in CP2 , can be combined with the methods of breaking symmetry in the motion equations
associated with the action War that was given in Section 2. It allows us to obtain the
following result of existence of solutions in the type IIA string on AdS5 CP2 S1 .
0 S1 : q E 0 } is a rational oneFor any couple of real numbers a, r the class {rq
r
parameter family of U (1)-invariant tori which belong to ar in the type IIA string on
AdS5 CP2 S1 .
In particular, if r = 2, we obtain Willmore tori as solutions of the conformal motion
equations associated with the corresponding action. Of course we can exploit, once more,
the breaking symmetry process to obtain solutions relative to metrics on AdS5 CP2 S1
which are given as double warping metrics coming from a couple of warping functions on
the circle.
The above solutions can be Hopf T-dualized to obtain solutions not only in the M-theory
on AdS5 CP2 T 2 , but also in the M-theory on any AdS5 P , where P is a circle

M. Barros et al. / Nuclear Physics B 584 [PM] (2000) 719748

733

bundle on CP2 S1 endowed with a KaluzaKlein ansatz associated with a principal


flat connection. To be precise, let be a real number such that / is not a rational
number, the map : Z S1 given by (k) = eik defines a monomorphism between
(Z, +) and S1 C regarded as a multiplicative group. Let U = CP2 R be the universal
covering of CP2 S1 . Certainly U can be regarded as a principal Z-bundle on CP2 S1 ,
which admits an obvious principal flat connection, say . The transition functions of this
bundle can be extended, via , to S1 -valued functions and they can be used as transition
functions to define a principal S1 -bundle, say P , on CP2 S1 . Moreover, one can extend
to a principal flat connection, also called , on the whole S1 -bundle whose holonomy
subbundle is isomorphic to U (CP2 S1 , Z). Notice that when is chosen to be one,
then P1 is nothing but the direct product CP2 T 2 on which the usual M-theory is Hopf
T-dualized. If we call : P CP2 S1 the projection of the above fibration, then we
0 S1 ): 0 0 } is a rational one-parameter family of (U (1) U (1))have { 1 (rq
rq
r
invariant submanifolds which are solutions of the motion equations associated with the
action War in the M-theory on AdS5 P .

5. A zoo of solutions in the type IIB string theory


Throughout this section M will be a principal fibre H -bundle, H being a d-dimensional
compact Lie group, over a certain pseudo-Riemannian manifold (M 0 , g 0 ). Let p : M M 0
be the projection and let be a principal connection on this principal fibre H -bundle. We
denote by da 2 the bi-invariant metric on H , so that the KaluzaKlein metric writes down
g = p (g 0 ) + (da 2 ).
As above, we can determine the manifold made up of (d + 1)-dimensional, H -invariant,
compact submanifolds in (M, g). It can be identified with the set of complete lifts of
closed curves immersed in M 0 . Therefore, to compute the critical points of Wbr in (M, g)
which are H -invariant, we use again the principle of symmetric criticality and compute
the restriction of Wbr to the above submanifold. To do that, let and e be the mean
curvature and the extrinsic scalar curvature functions of p1 ( ) in (M, g), respectively.
We also denote by S and S 0 the Ricci curvatures of (M, g) and (M 0 , g 0 ), respectively.
Finally, assume that (s) is arclength parametrized, denote by its curvature function
in (M 0 , g 0 ) and let 0 (s) be the horizontal lift to (M, g) of its unit tangent 0 (s). In this
setting, we have (see [1])
2 =

1
( 2 p),
(d + 1)2

e =


1
S 0 ( 0 ) p S( 0 ) .
d(d + 1)

On the unit tangent bundle U M 0 of (M 0 , g 0 ), let : U M 0 R be defined by


( 0 ) p =


b(d + 1) 0 0
S ( ) p S( 0 ) .
d

734

M. Barros et al. / Nuclear Physics B 584 [PM] (2000) 719748

Then
Wbr p

vol(H, da 2)
( ) =
(d + 1)r

2 + ( 0 )

r/2

ds.

Consequently we have:
Theorem 16. p1 ( ) is a critical point of Wbr in (M, g) if and only if is a critical point
of the elastic energy functional F defined by
Z
r/2
ds,
2 + ( 0 )
F ( ) =

on the space of closed curves in (M 0 , g 0 ).


The most interesting situation in Theorem 16 occurs when the potential is constant.
In this case, recall that we used the term generalized elastica, with potentials (, r/2), to
refer to critical points of F .
A sufficient condition to guarantee the constancy of is to assume that both (M, g) and
(M 0 , g 0 ) are Einstein, and then must be a YangMills connection. Therefore, if and 0
denote the cosmological constants of (M, g) and (M 0 , g 0 ), respectively, we find
b(d + 1) 0
( ).
d
In particular > 0 (see [1]).
However that condition is not necessary. In fact, suppose that both (M, g) and (M 0 , g 0 )
are Einstein giving a constant potential . Now, we deformate the metric g by changing the
relative scales of the base and the fibres. To be precise, we define a one-parameter family
of metrics on M by putting {gt = p (g 0 ) + t (da 2), t > 0}. This gives a one-parameter
family of Riemannian submersions with totally geodesic fibres, all of them having the
same horizontal distribution associated with . This is nothing but the canonical variation
of the starting Riemannian submersion. Since we are assuming that g1 = g is Einstein,
then there is at most one more Einstein metric in {gt } (see [6]). However, if t denotes the
corresponding potential, it is not difficult to see that t = t and so it is constant for any t.
The existence of one-parameter families of generalized elastic closed helices, for
arbitrary potentials, in S3 and CP2 , respectively, has been established in [1] and in the
last section, respectively.
=

Example 17. Let be the usual Hopf map from the 5-sphere S5 over the complex
projective plane CP2 . Then for any t > 0 and any pair of real numbers b, r, there exists a
rational one parameter family of U (1)-invariant flat tori with constant mean curvature in
(S5 , gt ) which are critical points of Wbr . Here g1 = g is the canonical metric of curvature
one on S5 .
Example 18. Let be the usual quaternion Hopf map from the 7-sphere S7 over the
quaternion projective line S4 . Then for any t > 0 and any pair of real numbers b, r, there

M. Barros et al. / Nuclear Physics B 584 [PM] (2000) 719748

735

exists a rational one parameter family of SO(4)-invariant 4-dimensional submanifolds with


constant mean curvature in (S7 , gt ) which are critical points of Wbr . Here g1 = g is the
canonical metric of curvature one on S7 .
We have just obtained a rational one-parameter family of flat tori with constant mean
curvature in the round 5-sphere, which are critical points of Wbr for arbitrary b and r.
These tori appeared connected, via the Hopf map, with the existence of generalized elastic
closed helices in the complex projective plane. The chief point to obtain this variational
reduction of variables was provided by the Palais principle of symmetric criticality.
Now, we are going to exhibit a new method to get flat tori in S2n+1 which are not
obtained as lifts of closed curves in the complex projective space CPn , but still are critical
points for Wbr .
Let T be a compact genus one surface and let (S2n+1 , g0 ) be the round sphere of radius
one. For any pair of real numbers b and r we have Wbr : I (T , S2n+1 ) R given now by
Z
Wbr () = ( 2 + b bG)r/2 dv,
T

where is the mean curvature of , G and dv being the Gaussian curvature and the volume
element of (T , (g0 )), respectively.
The computation of the first variation of Wbr is not easy in general. Certainly, the case
r = 2 (the Willmore functional corresponds with r = 2 and b = 1) is the simplest one.
R
In fact, using the GaussBonnet formula, the functional is reduced to T ( 2 + b) dv. In
this case, the first variation formula was computed in [45], and from there one sees that
the minimal immersions are automatically solutions of the corresponding EulerLagrange
equations. Moreover, it is not clear, in general, that the minimal immersions are critical
points of Wbr . We will overcome these obstacles by using again the Palais principle
of symmetric criticality. Contrarily to those occasions where we used the principle for
reduction of variables, now we will use it to improve rigidity. Notice that the highest
rigidity for a metric on a compact genus one surface means flatness.
To clarify that idea, let K be a 2-dimensional compact subgroup of SO(2n + 2) regarded
as the isometry group of (S2n+1 , g0 ). It is obvious that K acts naturally on I (T , S2n+1 ).
Let be a K-invariant immersion, then (T , (g0 )) has a subgroup of isometries of
dimension two. We apply a well known classical argument to see that K acts transitively
on (T , (g0 )). This homogeneity implies constant Gaussian curvature and so (T , (g0 ))
is a flat torus. Consequently, the submanifold of K-invariant immersions IK (T , S2n+1 ) is
made up of flat tori. It is not difficult to see that each isometric immersion from a flat
torus in (S2n+1 , g0 ) can be viewed as an orbit associated with a 2-dimensional, compact
subgroup of SO(2n + 2) (see [26,27] for minimal flat tori in the 5-sphere). Then we apply
the principle of symmetric criticality to obtain the following.
Proposition 19. IK (T , S2n+1 ) is a critical point of Wbr if and only if it is a critical
point of Wbr , but restricted to the submanifold IK (T , S2n+1 ).

736

M. Barros et al. / Nuclear Physics B 584 [PM] (2000) 719748

Now we can compute the first variation of this restriction, that is, we take variations of
IK (T , S2n+1 ) in IK (T , S2n+1 ). We use a standard argument, which involves some
integrations by parts, to obtain the following EulerLagrange equation
) + 4 2 H 2(r 2)bH = 0,
r1H r A(H

(14)

where H is the mean curvature vector field, 1 denotes the Laplacian associated with the
normal connection and A is the Simons operator [40].
A first consequence makes mention to the minimality.
Corollary 20. Every minimal flat tori in (S2n+1 , g0 ) is a critical point of Wbr , for
arbitrary b and r.
To illustrate that, we consider the following one-parameter family of minimal flat tori in
(S5 , g0 ) (see [26]). Let q (0, 1] be a rational number and consider, in the Euclidean plane
R2 , the lattice generated by




2 q2
2
2
p
p
.
,
,0
and
2
2
2
q 4q
q 2(4 q )
We define y : R2 C3 by
y(s, t) = p

1
4 q2

q
i(s+t) 
eis , eit , 2 q 2 e q .

(15)

It is not difficult to see that this gives an isometric immersion. Furthermore, it induces an
isometric immersion x from the flat torus T = R2 / in (S5 , g0 ). Since the coordinate
functions of x in C3 are eigenfunctions of the Laplacian of T , associated with the
eigenvalue 2, x is minimal in (S5 , g0 ). Notice that the case q = 1 gives the so called
equilateral flat torus, because it comes from an equilateral lattice.
Next we give some explicit examples of non-minimal flat tori in (S5 , g0 ) which are
critical points of the functional Wbr . Given three real numbers c, d and e, with c, d > 0,
we consider the lattice in R2 generated by (2c, 2e) and (0, 2d). Choose n, m, n
= nc
satisfy 6= and 6= 0. We can also assume
Z {0} such that = nc me
cd and
2 C3 by
>
1.
Then
we
define
y
:
R
that m
d
 


 
q
t is
t is

e , p sin
e ,
1 p2 ei s
,
(16)
y(s, t) = p cos
p
p
where p = d/m. It is easy to see that y defines an isometric immersion if and only if
p2 2 + (1 p2 ) 2 = 1. Furthermore, in this case it induces an isometric immersion x
from the flat torus T = R2 / in (S5 , g0 ). Some interesting properties of these immersions
are collected in the following.
1. The center of mass of x in C3 coincides with the center of S5 , in this sense we say
that it is of mass symmetric. Notice that minimal submanifolds in a round sphere are
always mass symmetric.
2. The immersion x is not minimal in (S5 , g0 ). In fact, it is constructed in C3 by
using eigenfunctions of the Laplacian of T associated with two different eigenvalues,

M. Barros et al. / Nuclear Physics B 584 [PM] (2000) 719748

737

namely 2 + p12 and 2 . Therefore, we say that it is of 2-type in the sense of Chen
(see [11]).
3. These immersions have non-zero constant mean curvature.
4. They are not Hopf tori. That is, they are not invariant under the natural U (1)-action
on S5 to obtain CP2 as space of orbits.
The following result shows the existence of an ample family of non-minimal flat tori in
(S5 , g0 ) which are critical points of Wbr for arbitrary b and r > 0.
Proposition 21. For any pair of real numbers b and r, there exist infinitely many nonminimal flat tori in the family (16) which are critical points of Wbr in (S5 , g0 ).
Proof. Let A be the shape operator of x : T (S5 , g0 ) in the unit direction defined by H .
Let = 2 + p12 and = 2 be the two eigenvalues of the Laplacian of T which are
involved in the 2-type nature of x in C3 . Now, a straightforward but long computation
allows us to see that x is a critical point of Wbr in (S5 , g0 ), that is, a solution of (14), if
and only if the following equation holds
2r|A|2 4 2 = r( + 2) + 2(2 r)b.

(17)

A messy computation shows that (17) is equivalent to


(2r 1)p4 4 + (4p2 2rp2 r 2)p2 2 + M(b, r, p) = 0,

(18)

where

M(b, r, p) = r 1 4p(r p rp2 + p3 ) + p2 (p2 1) 2(2 r)b 2r .
The equation (18) can be regarded as a biquadratic one in , so if D denotes its
discriminant, it is easy to see that limp1 D = r 2 . Then given (b, r) there exists an open
subset I in (0, 1) such that for all p I one can get solutions of (18). Now it is clear that
we can determine flat tori in the family (16) with these parameters, and the proof finishes.
Along this section, we have obtained three explicit families of flat tori in (S5 , g0 ) that
are critical points of Wbr for arbitrary b and r. Let us recall them.
1. The family C1 was obtained by lifting, via the usual Hopf map : S5 CP2 ,
a rational one-parameter family of generalized elastic closed helices in CP2 with
suitable potentials. Therefore, the flat tori in C1 have non-zero constant mean
curvature in (S5 , g0 ).
2. The family C2 is made up of minimal flat tori in (S5 , g0 ). Explicit parametrizations
for tori in C2 were given by Kenmotsu [26]. This is also a rational one-parameter
family of tori.
3. The family C3 (see (16) and Proposition 21) consists of a multi-indexed family of nonzero constant mean curvature, flat tori in (S5 , g0 ) which are constructed in C3 using
eigenfunctions of their Laplacians associated with exactly two different eigenvalues.

738

M. Barros et al. / Nuclear Physics B 584 [PM] (2000) 719748

6. Further backgrounds
The above considered AdS5 S5 solution can be extended to any other configuration of
the form AdS5 M5 , where M5 is any 5-dimensional compact Einstein space with positive
scalar curvature. An interesting class of such solutions is provided by choosing M5 to be a
principal U (1)-bundle over a compact 4-dimensional manifold N . Now, the U (1)-bundles
over such a N are classified by the cohomology group H 2 (N, Z).
For example, let N be the Grassmannian of oriented 2-planes in R4 , viewed as the
complex quadric Q2 = S2 S2 with its natural Einstein metric g. By choosing 13 c1 (N)
H 2 (N, Z), where c1 (N) denotes the first Chern class of N , we obtain a principal U (1)bundle M = T1 S3 on N which coincides with the unit tangent bundle of the round 3-sphere
and admits a natural Einstein metric (Stiefel manifold) g.
This example can be regarded as
a special one of an integer two-parameter family of U (1)-bundles, {M(p, q): p, q Z},
over S2 S2 , where the integers p and q are the winding numbers of the fibres over both
S2 factors in the complex quadric. Natural Einstein metrics, with positive scalar curvature,
can be obtained on each M(p, q) [14]. Observe that one can consider p and q to be
relatively prime, otherwise if a = gcd(p, q), then M(p, q) = M(p/a, q/a)/Zp is a lens
space. Notice also that T1 S3 = M(1, 1) is diffeomorphic to S2 S3 , however it is not a
product either as a homogeneous space or as an Einstein manifold.
Finally remark that the situation of greatest interest comes out when the Einstein space
M(p, q) admits Killing spinors, which automatically implies that the AdS5 M(p, q)
solution preserves some supersymmetries. However it only occurs when p = q = 1, i.e.,
for the Stiefel manifold [14]. We may follow steps analogous to those described for the
five sphere [14], to reduce the AdS5 T1 S3 solution of the type IIB theory to D = 9, and
perform a T-duality transformation. Upon oxidation back to D = 10 type IIA theory, we
have a solution on AdS5 Q2 S1 . This can be oxidised further to D = 11 M-theory,
giving a solution not only on AdS5 Q2 T 2 , but also on any background of the form
AdS5 P , where P is any principal U (1)-bundle over Q2 S1 , which admits a principal
flat connection and it is endowed with the corresponding KaluzaKlein antsaz, I (P ) being
the corresponding group of isometries. Consequently, we obtain the following duality
chain [n = 4 YangMills] [type IIB string on AdS5 T1 S3 ] [type IIA string on
AdS5 Q2 S1 ] [M-theory on AdS5 P ].
According to the reduction of symmetry program, we can obtain examples of solutions,
for the motion equations associated with the War -dynamics, in any step of the above
constructed duality chain which reduces to generalized elasticae in the complex quadric, as
well as in a AdS3 totally geodesic in AdS5 . In the latter case, we can use those rational oneparameter families of generalized elastic helices in AdS3 which were obtained in Section 3.
Therefore we have a result analogous to Corollary 15.
1. {qr T1 S3 : q Q {0}} is a rational one-parameter family of (SO(4) U (1))invariant solutions of the War -dynamic in the type IIB theory on AdS5 T1 S3 .
2. {qr Q2 S1 : q Q {0}} is a rational one-parameter family of (SO(4) U (1))invariant solutions of the War -dynamic in the type IIA theory on AdS5 Q2 S1 .

M. Barros et al. / Nuclear Physics B 584 [PM] (2000) 719748

739

3. {qr P : q Q {0}} is a rational one-parameter family of I (P )-invariant solutions


of the War -dynamic in the M-theory on AdS5 P .
In the former case, we can emulate an argument given in [1] to construct real one-parameter
families of generalized elastic closed helices in the complex quadric for any pair of given
potentials. These helices appear as geodesic in certain flat tori embedded in Q2 , so that a
simple argument allows us to get a closeness condition in terms of a rationality condition
for the involved parameter (see the appendix of [1]). Using now the breaking of symmetry
program, we can obtain holographic solutions for the War -motion equations at any stage
of the above described T-duality chain. These solutions are similar to those obtained in
Sections 4 and 5 where CP2 played the part of Q2 here.
Certainly, spherical compactifications of supergravity are maximally supersymmetric
and therefore the boundary superconformal field theory, SCFT, has sixteen supercharges.
However, the conjecture of Maldacena is believed to be true for any supersymmetry. This
is a good reason to investigate SCFT with less than sixteen supercharges [17,18,20,22,25,
29,34]. This reduction of supersymmetry is obtained by orbifolding the space transverse
to the boundary. Hence, the AdS part of the geometry remains intact while the Sn part of
the geometry is orbifolded and depending on the orbifold one obtains distinct CFT with
different amounts of supersymmetry. An interesting way to do it is by considering odd
dimensional transverse spheres and then to regard them as Hopf fibrations, i.e., principal
U (1)-bundles over complex projective spaces. Now, we can break supersymmetry either
by reducing over U (1)-fibre (this Hopf reduction has been already widely used along this
paper) or by considering multiple windings of the U (1)-fibre over the base space. In this
second setting we do not reduce dimension and arrive to the lens spaces.
Let S2n1 Cn be the (2n 1)-dimensional sphere of radius one, i.e.,
(
)
n
X
|zj |2 = 1 .
S2n1 = z = (z1 , . . . , zn ) Cn : |z|2 =
j =1

be a primitive rth root of unity and {s1 , . . . , sn }


For any natural number r, let =
integers which are relatively prime to r. We define an action of Zr = {1, , 2, . . . , r1 }
on S2n1 by

(z1 , . . . , zn ) = s1 z1 , . . . , sn zn .
e2i/r

The orbit space is denoted by L(r, s1 , . . . , sn ) and it will be called a lens space. The natural
projection p : S2n1 L(r, s1 , . . . , sn ) gives the universal covering of this space. Hence
Zr is not only the fundamental group of L(r, s1 , . . . , sn ), but also the deck transformation
group of this covering space. The classical case appears when n = 2 and L(r, 1, s) is
usually denoted by L(r, s). In particular, L(2, 1) is just the real projective space RP 3 .
Let {k` } and 0 be the transition functions of S3 (L(r, s), Zr ) and the connection
1-form of its canonical flat principal connection, respectively. For any compact Lie group
G endowed with a bi-invariant metric d 2 , we choose an arbitrary closed geodesic
through the identity of G, say (t) = exp(tA), where A g. We define a monomorphism
: Zr G by identifying Zr with the group of primitive rth roots of unity and then
using that the exponential mapping defines an isomorphism between S1 and . We may

740

M. Barros et al. / Nuclear Physics B 584 [PM] (2000) 719748

then extend {k` } via to obtain a set of G-valuated functions which can be used to
construct a principal fibre bundle P (L(r, s), G). Furthermore, is extended to get a
monomorphism : S3 P which maps 0 into a flat connection on P . Summing up,
we have obtained the following result.
Proposition 22. Let G be a compact Lie group. Then there exists a G-principal fibre
bundle P (L(r, s), G) over the lens space L(r, s) which admits a principal flat connection
with holonomy subbundle isomorphic to the 3-sphere, that is, S3 (L(r, s), Zr ).
This proposition is also true if G is not compact and the construction can also be
generalized to lens spaces of higher dimensions.
Let (L5 , g0 ) be a five dimensional lens space. Then we can use the natural covering
mapping from S5 over L5 to project C1 , C2 and C3 (the three families of solutions obtained
in the above section on the round five sphere). We also denote by C1 , C2 and C3 the projected
families in (L5 , g0 ). Then we have,
Corollary 23. Let G be an m-dimensional compact Lie group endowed with a bi-invariant
metric d 2 . Let : P L5 be a principal fibre G-bundle, endowed with a principal flat
connection . Let [g0 ] be the KaluzaKlein conformal class on P associated with g0 =
(g0 ) + (d 2 ). Then there exist infinitely many (m + 2)-dimensional, G-invariant,
WillmoreChen submanifolds in (P , [g0 ]), which have constant mean curvature in (P , g0 ).
This family includes the three subfamilies obtained by lifting, via p, the families of flat tori
C1 , C2 and C3 .
The above result can be obviously extended to any action War , not necessarily to that
giving the WillmoreChen functional. Furthermore it can be applied to a wide variety of
contexts. For example, suppose the Lie group G is chosen to be U (1) = S1 , so that P is a
principal U (1)-bundle on L5 and then we have:
The War -dynamic in the M-theory on AdS5 P has infinitely many U (1)-invariant
compact solutions with dimension three. This class includes the three subfamilies obtained
when lifting, via p : P L5 , the families of flat tori C1 , C2 and C3 on L5 .
It should be noticed that these solutions have constant mean curvature (tension) in the
original KaluzaKlein metric on AdS5 P .
Remark 24. We can construct examples of 2-type (in the sense of Chen) flat tori in (S7 , g0 )
which are critical points of Wbr for arbitrary b and r. Moreover, we can use an argument
similar to that used in [3] to show that 2-type compact surfaces which are solutions of
the Wbr -dynamic in any round sphere (Sm , g0 ) are actually flat tori lying fully in either
(S5 , g0 ) or (S7 , g0 ).

M. Barros et al. / Nuclear Physics B 584 [PM] (2000) 719748

741

7. Dyonic strings and non-dilatonic black holes


General spaces of the type AdS N , where N are Einstein spaces, necessarily spheres,
emerge naturally in supergravity as the near horizon geometries of supermembranes
with fewer Killing spinors and whose boundary conformal field theories have so less
supersymmetry. In particular, the space AdS3 S3 appears as the near horizon geometry
of the self-dual string or, more generally, the dyonic string (see [15] and references therein
for details). In that paper, truncated six-dimensional type IIB and type IIA Lagrangians are
obtained and their T-duality transformations were explicitly computed. This construction
is necessary if one wishes to consider solutions carrying both NSNS and RR electric and
magnetic 3-form charges and whose near horizon geometry contains AdS3 S3 .
All the non-dilatonic black holes in D = 5 and D = 4 were listed there and their near
horizon limits, when they are oxidised to D = 6, were obtained. Since all these near
horizon limits can be obtained by Hopf T-duality on AdS3 S3 (actually, Hopf T-duality
relates not only near horizon limits but also their associated full solutions), we can combine
Corollary 9 with Theorem 12 to obtain an existence result of WillmoreChen submanifolds
which can be regarded as a Hopf T-duality invariant. In fact, it is known that Hopf dualities
preserve the area of the horizons and hence they also preserve the black hole entropies.
Now we have
Corollary 25. The conformal structure associated with the string-frame in the near
horizon limit of any non-dilatonic black hole in D = 5 and D = 4 admits a rational one
parameter family of four dimensional WillmoreChen submanifolds which have constant
mean curvature in the original string metric. Moreover, in the case that there are only NS
NS charges, the invariance of these WillmoreChen submanifolds is preserved, while this
is decreased when RR charges are carried.
Remark 26. Dilatons and axions are constant in the solutions and for simplicity we have
considered them to be zero. Otherwise we could have started from original solutions
AdS3 S3 /Zn , for the type IIB low-energy effective action. It should be pointed out that
the string metric of any near horizon limit is always homogeneous and have constant scalar
curvature. For example, by considering that dilatons and axions are zero and applying Hopf
T-duality on the U (1)-fibres of S3 , the following possibilities could appear:
1. There are only RR charges, then S3 is untwisted to S2 S1 .
2. There are only NSNS charges, then S3 becomes a cyclic lens space S3 /Zp with its
round metric and p being the magnetic NSNS charge.
3. In the generic case, with both NSNS and RR charges, S3 not only becomes S3 /Zp ,
but it is also squashed, with a squashing parameter that is related to the values of the
charges. In other words, the metric on S3 /Zp is covered by a metric on S3 which may
be realized as a distance sphere in the complex projective plane or its symmetric dual
(the complex hyperbolic plane) according to the squashing parameter is less than or
greater than 1, respectively. The squashing parameter equal to 1 corresponds with the
round metric.

742

M. Barros et al. / Nuclear Physics B 584 [PM] (2000) 719748

Most of the near horizon limits are not simply connected. Only those solutions with
dilatons and axions being zero of the type IIB Lagrangian are simply connected. Since
Corollary 9 holds for any Wbr , one can obtain an analogous to Corollary 25 for any
functional Wbr . Consequently, we combine this argument with the first reduction of
symmetry phase to obtain the following result.
Corollary 27. Let AdS3 N be any near horizon limit and choose a monomorphism from
the fundamental group 1 (N) in a compact Lie group endowed with a bi-invariant metric
d 2 . Then
1. There exists a principal fibre G-bundle : P AdS3 N which admits a principal
flat connection .
2. There exists a rational one parameter family of WillmoreChen submanifolds in
(P , [g]),
where g = (g) + (d 2 ), g being the string metric, which have nonzero constant mean curvature in (P , g),
and are (G H )-invariant, where H is the
group of isometries of N .
This result works, in particular, along the oxidation process of the involved metrics to
D = 10 and D = 11. Consequently, we have ample families of solutions of War -dynamics
(in particular equivariant WillmoreChen submanifolds) in the oxidised D = 10 metrics
AdS3 S3 T 4 and AdS3 S3 K3 of type IIB string theory. Also in the duals
AdS3 S2 S1 T 4 , AdS3 S2 S1 K3, AdS3 S3 T 4 , AdS3 S3 K3,
AdS3 (S3 /Zp ) T 4 and AdS3 (S3 /Zp ) K3 type IIA theories. Finally the result
applies to the oxidised D = 11 metrics AdS3 S3 T 5 and AdS3 S3 K3 S1 of
M-theory.

8. Higher-dimensional theories
The type IIA string can be obtained by compactifying the D = 11 supermembrane on
a circle. An obvious question is whether type IIB string also admits a higher-dimensional
explanation. The appearance of MajoranaWeyl spinors and self-dual tensors in both the
twelve-dimensional and type IIB theories supplied evidence in favour of a corresponding
and natural conjecture posed in [7].
Despite of all the objections one might raise to a world with two time dimensions,
associated with the idea of a (2, 2) object moving in a (10, 2) spacetime, it has been revived
in the context of the F-theory [42]. This involves type IIB compactification where the
axion and dilaton from RR sector are allowed to vary on the internal manifold. In general,
given a manifold M that has the structure of a fiber bundle , with fiber T 2 = S1 S1 , on
some manifold L, then [F-theory on M] [type IIB theory on L]. Of course, the most
conservative point of view is that the twelfth dimension is merely a mathematical artifact
and so the F-theory should be considered as a clever way of compactifying the type IIB
string [39]. However, time will tell [13].
Consequently in this section we will investigate the following general framework. Let
M be an F -bundle associated with a certain principal H -bundle Q(L, H ), where H is

M. Barros et al. / Nuclear Physics B 584 [PM] (2000) 719748

743

a compact Lie group and the fibre F is a homogeneous space with dimension d. Let
p : M L be the projection map, a gauge potential on Q(L, H ) and da 2 an H -invariant
metric on F . In this setting, given any metric g 0 on L, one can find a unique metric g
on M such that p : (M, g) (L, g 0 ) is a Riemannian submersion with totally geodesic
fibres isometric to (F, da 2 ) and horizontal distribution associated with [44]. As above,
the manifold of the (d + 1)-dimensional, H -invariant compact submanifolds in (M, g) can
be identified with the set of complete lifts of closed curves in L. Once more, the search for
symmetric configurations which are solutions of the Wbr -dynamics on (M, g) is reduced
to that of solutions of the dynamics associated with the restriction of the action Wbr to
the space of symmetric submanifolds. This restriction has been computed in [2] and is
formally similar to that obtained in Section 5, that is
Z
 vol(F, da 2 )
r/2
ds,
2 + ( 0 )
Wbr p1 ( ) =
r
(d + 1)

where denotes de curvature of in (L, g 0 ) and is defined on the unit tangent bundle of
(L, g 0 ) to measure, up to a constant involving b, the difference between the Ricci curvature
of (L, g 0 ) in a direction and the Ricci curvature of (M, g) in the corresponding horizontal
direction. Therefore we formally have the same Theorem 16, though in a more general
context
Theorem 28. p1 ( ) belongs to br in (M, g) if and only if is a critical point of the
elastic energy functional
Z
r/2
ds,
2 + ( 0 )
F ( ) =

defined on closed curves in (L, g 0 ).


As above, most of the important applications of this result occur when the potential is
constant. In this case we will use again the term of generalized elastica to name a critical
point of F . An obvious sufficient condition to guarantee this constancy of is to assume
that both (M, g) and (L, g 0 ) are Einstein. In this case, if and 0 denote the corresponding
0
cosmological constants then = b(d+1)
d ( ). This condition is not sufficient as the
squashing method shows.
Although the list of examples satisfying that condition is too large, we have chosen some
of them as an illustration.
First example
It is known that the field equations of d = 11 supergravity describe a 4-dimensional
spacetime with negative cosmological constant and a compact 7-dimensional Einstein
space (M, g) with positive scalar curvature. Many of these spaces can be regarded
as principal fibre H -bundles over certain spaces (L, g 0 ) and the Einstein metric g is
obtained from the KaluzaKlein mechanism, i.e., g = p1 (g 0 ) + (da 2 ). Consequently,

744

M. Barros et al. / Nuclear Physics B 584 [PM] (2000) 719748

Theorem 28 can be directly applied to most of the stable vacuum states in the Freund
Rubin spontaneous compactification of d = 11 supergravity.
On the other hand, if we squash an Einstein metric (M, g) by scaling the size of
fibres, we still obtain Riemannian submersions providing constant potentials. Therefore,
if we consider the canonical variation of the quaternion Hopf fibration {p : (S7 , gt )
(S4 , g 0 ); t > 0}, we can find exactly two values of t, namely t = 1 (the round metric)
and t = 1/5 (the squashed metric) providing Einstein metrics. It is known that generalized
elasticae in a round 4-sphere yield in a totally geodesic 3-sphere, as well as the existence,
for any pair of potentials r, , of a rational one parameter family of closed helices, r, =
{j }, in a round 3-sphere, which are generalized elasticae [2]. Therefore, for any t > 0,
we obtain a rational one parameter family {p1 ( ); r, } of critical points of Wbr in
(S7 , gt ). This family of critical points can be projected down to the lens space S7 /Z` =
L7 (`, 1, 1, 1), which is also obtained in the list of solutions for d = 11 supergravity given
(1)
in [10], as SU(4)U
SU(3)U (1) and ` is the number of times that a simple loop in the U (1) of the
denominator winds around the U (1) in the numerator. We will also denote by g the round
metric on S7 /Z` and use the already recalled way to generate the principal fibre bundles
admitting a flat connection to obtain
Corollary 29. Let G be any compact Lie group endowed with a bi-invariant metric d 2
and a monomorphism from Z` in G. Then
1. There exists a principal fibre G-bundle : P S7 /Z` which admits a principal flat
connection .
2. There exists a rational one parameter family of WillmoreChen submanifolds in
the conformal KaluzaKlein structure [g],
g = (g) + (d 2 ) on P which are
(G SU(2))-invariant and have non zero constant mean curvature in (P , g).

Second example
(1)
In [9], the spaces N abc = SU(3)U
U (1)U (1) were studied, where a, b and c Z characterize the
embedding of U (1) U (1) in SU(3) U (1). These spaces can be viewed as principal fibre
SU(2)-bundles on CP2 . Using the KaluzaKlein inverse mechanism and the squashing
method, one can see that these spaces, except when 3a = b, admit exactly two different
Einstein metrics which make the above fibration a Riemannian submersion with totally
geodesic fibres and the base CP2 endowed with the FubiniStudy metric (see also [35]). We
denote by (N abc , g) the above CastellaniRomans Einstein Riemannian manifolds. On the
other hand, for any couple of constants r, one can find a rational one parameter family of
closed helices in CP2 which are generalized elasticae [2]. Therefore, for arbitrary b, r one
can get rational one parameter families of critical points for Wbr in each (N abc , g) which
are SU(2)-invariant and have non-zero constant mean curvature. To avoid Riemannian
product in the next result, we will consider r 6= 0 (otherwise, N ab0 is simply connected)
and the embedding of U (1) U (1) in SU(3) U (1) is carried out by mapping a simple
loop of the U (1) in the denominator to wind ` times around the U (1) in the numerator. In
this case the fundamental group of N abc is Z` and we have

M. Barros et al. / Nuclear Physics B 584 [PM] (2000) 719748

745

Corollary 30. Let G be any compact Lie group with a bi-invariant metric d 2 and a
monomorphism from Z` in G. Then
1. There exists a principal fibre G-bundle : P N abc which admits a principal flat
connection .
2. There exists a rational one parameter family of WillmoreChen submanifolds in
(P , [g]),
g = (g) + (d 2 ), which are (G SU(2))-invariant and have non-zero
constant mean curvature.
Remark 31. We can obtain similar results in other spaces giving solutions for d = 11
5
3 U (1)
supergravity, such as the spaces of Witten M abc = SUS
(1)U (1) , which can be viewed as
circle bundles on CP2 S2 .
Since our result works for associated bundles (not necessarily principal), it can be
applied, for instance, to certain Penrose twistor spaces. Namely, to the twistor spaces of
S4 and CP2 to obtain examples of immersions in br [2].
The method that we have developed in this paper can be applied to other backgrounds
different from those considered here. It is also open to be extended to other contexts.

9. Conclusions
In this paper we have considered a two-parameter class of actions, {War : a, r R},
defined on the space of immersions of a given smooth manifold Q in a pseudo-Riemannian
manifold (L, ds 2 ). This constitutes a natural variation of the conformal total tension
action, also known as the WillmoreChen functional, whose importance is due in part
to its invariance under conformal changes in the surrounding metric ds 2 . This class also
includes the popular Willmore action, so as the CanhamHelfrich bending energy of fluid
membranes and lipid vesicles, as well as the Polyakov extrinsic action in the bosonic
string theory. Roughly speaking, given a universe (L, ds 2 ) and a Lie group G, which acts
on L through isometries of (L, ds 2 ), the best worlds to live in this universe are those
submanifolds which satisfy the following properties:
1. They are G-invariant configurations. That means that they have a natural degree of
a priori stated G-symmetry.
2. They are solutions of the isoperimetric area-volume problem, in particular they must
have constant mean curvature (tension) in (L, ds 2 ).
3. They are extremes for some tension action War and so solutions of the corresponding
motion equations.
Along this section, we will use the term Gar -configuration to name those submanifolds
in (L, ds 2 ) which are G-invariant, critical points of some War and have constant mean
curvature in (L, ds 2 ).
The existence of Gar -configurations is investigated in string theories, M-theory and
F-theory (even in higher-dimensional theories) on backgrounds of the form AdS M,
where M is some principal U (1)-bundle.

746

M. Barros et al. / Nuclear Physics B 584 [PM] (2000) 719748

Recall that string theory, emerging as a candidate for the unification of the fundamental
forces in the nature, has a main objection. In fact, there are five different, but consistent,
ten-dimensional string theories which are all distinct in their perturbative spectra. The
understanding of the problems derivated from this ambiguity has undergone a great
improvement with the appearance of dualities. String dualities are, in some sense, the
statements one has to relate all five different perturbative superstrings. It is tempting then to
imagine that they are the expansion of a single and more powerful eleven dimensional, nonperturbative, unified theory (known as M-theory and that contains the D = 11 supergravity
as low energy limit) around five different sets of perturbative variables. A very important
ingredient to account for these dualities are those properties (encoded in the physics or
in the geometry of the theory) which remain invariant along the duality chains. When
the compactifying space M is a principal U (1)-bundle, one can define a natural kind of
T-duality [14,15]. These Hopf T-dualities relate different black holes preserving entropies
[15]. Therefore, the black holes entropies provide a nice invariant along duality chains.
In this paper we have shown that:
The existence of wide families of Gar -configurations, for arbitrary a, r and suitable
choices of G, is also an invariant along any duality chain.
However, this general statement is actually a series of results that were obtained as
conclusions from a general program of breaking symmetry which was developed in
Section 2. The method exhibited there is mainly based on a formulation due to Palais
[36] of the so-called principle of symmetric criticality. This method allows one to break
the G-symmetry by reducing the number of variables in the study of War -variational
problems. Furthermore, the constancy of the mean curvature function remains invariant
through this process.
To conclude, we summarize the main points in the above series of invariants. The first
natural problem, that we solved, corroborates once more the important role that AdSgeometry plays in these theories and specially the group manifold AdS3 . It corresponds
with the case where the transverse space (M, g) is a G-homogeneous one. In AdSp
MDp , we consider the metric f 2 go + g, where f is any smooth positive function on M.
Given a curve in AdS, one can evolve the transverse space M through to generate the
tube T = M with the metric f 2 dt 2 + g. It is obvious that T is G-invariant and so
it seems natural to ask whether T is a Gar -configuration. The method we exhibited here
allows us to reduce this problem to one for in the AdS part. Using the nice geometry of
AdS, we are able to get not only the characterization of T to be a Gar -configuration, but
also the complete classification of these solutions. This provides us the moduli space of
Gar -configurations with dimension D p + 1. These moduli spaces must be understood,
up to isometries of f 2 go + g, except in the case of the WillmoreChen action, where we
can relax to conformal transformations of f 2 go + g. In this context, we can also obtain
Gar -configurations with dimension greater than D p + 1, however we do not know the
moduli spaces and this can be regarded as an open problem.
Other solutions in this paper are directly obtained in the transverse space. For example,
if we start from type IIB theory on AdS5 M5 , where M5 is a principal U (1)-bundle
over, say B, then we have the duality chain [type IIB theory on AdS5 M5 ] [type IIA

M. Barros et al. / Nuclear Physics B 584 [PM] (2000) 719748

747

theory on AdS5 B S1 ] [M-theory on AdS5 B T 2 ]. Then we use our method


of breaking symmetry to reduce the search for Gar -configurations, where G is now an
internal gauge group, in these backgrounds to that for certain elastic curves in the base
space B. Giving explicit examples of elasticae, for example, in CP2 or in the complex
quadric Q2 = S2 S2 , we get wide families of Gar -configurations along the above duality
chain.
The method also works on theories that carry both NSNS and RR electric and
magnetic charges. Therefore, we showed the existence of ample families of solutions in
the near horizon limit of any non-dilatonic black hole in D = 5 and D = 4. The degree
of symmetry of these configurations is preserved only when NSNS charges appear, while
that is decreased if RR charges are carried. Higher-dimensional theories, such as F-theory,
are also investigated in relation with the existence of Gar -configurations

Acknowledgements
This research has been partially supported by DGICYT grant PB97-0784 and Fundacin
Sneca (C.A.R.M.) grant PB/5/FS/97.
References
[1]
[2]
[3]
[4]
[5]
[6]
[7]
[8]
[9]
[10]
[11]
[12]
[13]
[14]
[15]
[16]
[17]
[18]
[19]
[20]
[21]

M. Barros, Nucl. Phys. B 535 (1998) 531.


M. Barros, WillmoreChen branes and Hopf T-duality, Class. Quantum Grav. 17 (2000) 1979.
M. Barros, B.Y. Chen, J. Math. Soc. Jpn. 39 (1987) 627.
M. Barros, A. Ferrndez, P. Lucas, M.A. Meroo, Trans. Amer. Math. Soc. 352 (2000) 3015.
M. Barros, O.J. Garay, D.A. Singer, Thoku Math. J. 51 (1999) 177.
A.L. Besse, Einstein Manifolds, Springer-Verlag, 1987.
M. Blencove, M.J. Duff, Nucl. Phys. B 310 (1988) 387.
P.B. Canham, J. Theor. Biol. 26 (1970) 61.
L. Castellani, L.J. Romans, Nucl. Phys. B 238 (1984) 683.
L. Castellani, L.J. Romans, N.P. Warner, Nucl. Phys. B 241 (1984) 429.
B.Y. Chen, Total Mean Curvature and Submanifolds of Finite Type, World Scientific, 1984.
M.J. Duff, Int. J. Mod. Phys. A 14 (1999) 815.
M.J. Duff, Supermembranes, hep-th/9611203.
M.J. Duff, H. L, C.N. Pope, Nucl. Phys. B 532 (1998) 181.
M.J. Duff, H. L, C.N. Pope, Nucl. Phys. B 544 (1999) 145.
M.J. Duff, H. L, C.N. Pope, E. Sezgin, Phys. Lett. B 371 (1996) 206.
S. Ferrara, A. Zaffaroni, Phys. Lett. B 431 (1998) 49.
S. Ferrara, A. Kehagias, H. Partouche, A. Zaffarono, Phys. Lett. B 431 (1998 ) 42.
E.S. Fradkin, M.A. Vasiliev, Phys. Lett. B 189 (1987) 89; Nucl. Phys. B 298 (1987) 141.
J. Gomis, Phys. Lett. B 435 (1998) 299.
W. Greub, S. Halperin, R. Vanstone, Connections, Curvature and Cohomology, Academic Press,
1972, 1973, 1976, 3 Vols.
[22] E. Halyo, Supergravity on AdS5/4 Hopf fibrations and conformal field theories, Modern Phys.
Lett. A 15 (2000) 397.
[23] W. Helfrich, Z. Naturf. C 28 (1973) 693.
[24] V. Husain, S. Jaimungal, Phys. Rev. D 60 (1999) 061501.

748

[25]
[26]
[27]
[28]
[29]
[30]
[31]
[32]
[33]
[34]
[35]
[36]
[37]
[38]
[39]
[40]
[41]
[42]
[43]
[44]
[45]

M. Barros et al. / Nuclear Physics B 584 [PM] (2000) 719748

S. Kachru, E. Silverstein, Phys. Rev. Lett. 80 (1998) 4855.


K. Kenmotsu, Thoku Math. J. 27 (1975) 83.
K. Kenmotsu, J. Math. Soc. Jpn. 28 (1976) 182.
J. Langer, D.A. Singer, J. Diff. Geom. 20 (1984) 1.
A. Lawrence, N. Nekrasov, C. Vafa, Nucl. Phys. B 533 (1998) 199.
J. Maldacena, Adv. Theor. Math. Phys. 2 (1998) 231.
R.R. Metsaev, A.A. Tseytlin, Nucl. Phys. B 533 (1998) 109.
B.E.W. Nilsson, C.N. Pope, Class. Quantum Grav. 1 (1984) 499.
B. ONeill, Semi-Riemannian Geometry, Academic Press, New York, 1983.
Y. Oz, J. Terning, Nucl. Phys. B 532 (1998) 163.
D.N. Page, C.N. Pope, Phys. Lett. B 147 (1984) 55.
R.S. Palais, Comm. Math. Phys. 69 (1979) 19.
A.M. Polyakov, Nucl. Phys. B 268 (1986) 406.
S.F. Prokuskin, M.A. Vasiliev, Phys. Lett. B 464 (1999) 53.
A. Sen, Nucl. Phys. B 475 (1996) 562.
J. Simons, Ann. Math. 88 (1968) 62.
L. Susskind, E. Witten, The holographic bound in Anti-de Sitter space, hep-th/9805114.
C. Vafa, Nucl. Phys. B 469 (1996) 403.
M.A. Vasiliev, Int. J. Mod. Phys. D 5 (1996) 763.
J. Vilma, J. Diff. Geom. 4 (1970) 73.
J.L. Weiner, Indiana Univ. Math. J. 27 (1978) 19.

Nuclear Physics B 584 [PM] (2000) 749783


www.elsevier.nl/locate/npe

Nonabelian duality and solvable large N lattice


systems
Andrey Yu. Dubin
ITEP, B. Cheremushkinskaya 25, Moscow 117259, Russia
Received 5 November 1999; revised 29 February 2000; accepted 6 June 2000

Abstract
We introduce the basics of the nonabelian duality transformation of SU(N) or U (N) vector-field
models defined on a lattice. The dual degrees of freedom are certain species of the integer-valued
fields complemented by the symmetric groups n S(n) variables. While the former parametrize
relevant irreducible representations, the latter play the role of the Lagrange multipliers facilitating
the fusion rules involved. As an application, I construct a novel solvable family of SU(N) D-matrix
systems graded by the rank 1 6 k 6 (D 1) of the manifest [U (N)]k conjugation-symmetry. Their
large N solvability is due to a hidden invariance (explicit in the dual formulation) which allows
for a mapping onto the recently proposed eigenvalue-models [8] with the largest k = D symmetry.
Extending [8], we reconstruct a D-dimensional gauge theory with the large N free energy given
(modulo the volume factor) by the free energy of a given proposed 1 6 k 6 (D 1) D-matrix
system. It is emphasized that the developed formalism provides with the basis for higher-dimensional
generalizations of the GrossTaylor stringy representation of strongly coupled 2d gauge theories.
2000 Elsevier Science B.V. All rights reserved.
PACS: 11.15.Ha; 11.15.Pg; 11.15.Tk
Keywords: Lattice; YangMills; Duality; Solvability

1. Introduction
A duality of the D = 4 continuum YangMills gauge system to a kind of string theory
remains to be one of a few intuitive guiding principles to attack nonperturbative dynamics
of the strong interactions. Among the circumstantial evidences, the central role is played by
the Wilsons D > 2 string-like representation [1] of the strong-coupling (SC) series but in a
lattice cousin of the continuum YMD theory. As it is well known, this particular expansion
(running in terms of the inverse powers of the bare coupling constant) can not be directly
extended into the weak-coupling phase relevant for the continuum limit. Nevertheless, we
E-mail: dubin@vxitep.itep.ru; tel.: +7-095-129-9674; fax: +7-095-883-9601

0550-3213/00/$ see front matter 2000 Elsevier Science B.V. All rights reserved.
PII: S 0 5 5 0 - 3 2 1 3 ( 0 0 ) 0 0 3 5 1 - 5

750

A.Yu. Dubin / Nuclear Physics B 584 [PM] (2000) 749783

believe that (properly chosen) lattice YMD systems hide a stringy pattern relevant for the
D > 2 continuum gauge theories at least in the regime being the continuum counterpart of
the lattice SC phase. To this aim, one is to consider a continuum D > 2 YMD model with
a finite ultraviolet cut off UV and sufficiently large coupling constant(s).
To support this idea, we refer to the well-studied D = 2 case, where Gross and Taylor
proposed an elegant stringy representation [2,3] of the large N SC series in the continuum
SU(N) gauge system on an arbitrary 2d surface. Recall that a continuum YM2 can be
directly reproduced through the corresponding lattice gauge model with the action defined
via the associated self-reproducing plaquette-factor [4,5]. As a result, the pattern of the
proposed D = 2 representation is the same both in the SC regime of a given continuum
YM2 theory and in the SC phase of the corresponding self-reproducing lattice model. The
only considerable distinction is that in the latter case one would deal with the discretized
surfaces rather than with the 2d manifolds. The crucial point is that, as far as the built in
topological data is concerned, this difference does not matter. As a result, in both instances
the appropriate SC series can be reinterpreted in terms of statistics of all admissible
branched coverings (associated to the base-surface) described canonically in terms of the
symmetric groups elements.
The key-ingredient of the above D = 2 construction is the so-called SchurWeyl
complementarity (see, e.g., [6] for a review) between certain structures in the Lie and in the
symmetric groups, with the latter being suitable for defining the relevant topological data.
Altogether, for YM2 it fulfills the role of a bridge between the symmetry and topology that
fits in the construction of the gauge string representation. Unfortunately, the proposed in
[2,3] technology can not be directly extended to D > 3. The purpose of the present paper is
to develop the basics of an approach which, among other things, renders accessible higherdimensional generalizations of the GrossTaylor pattern.
One of the central elements of our approach is the nonabelian duality transformation
which can be considered as a natural extension of the SchurWeyl duality. Complementary,
it should be viewed as a realization of the long sought nonabelian version of the abelian
transformation well known in the context of the pure U (1) lattice system (see [14,15] for
a review). This generalization can be compared, in particular, with the recent conjecture
of Polyakov [7]. He advocated that the lattice abelian transformation encodes the
N = 1 string-like pattern which might be generalized (in a yet unknown way) for the U (N)
continuum gauge theory with an arbitrary N .
The gauge string construction, we keep in mind, is facilitated by the formalism which is
to synthesize both the nonabelian duality (i.e., symmetry) and the topology of the branched
coverings. We find it appropriate to introduce the former ingredient in a simpler setting
that avoids entanglement with the topology. For this purpose, we find simpler applications
of the duality transformation. As a warm-up, it is natural to rederive (see Appendix D)
the dual representation for the partition function of the two-dimensional gauge theory on
an arbitrary 2d manifold. (It was previously obtained in [5,23] by a different method and
used by Gross and Taylor [2,3] as the starting point.) As a more nontrivial application, we
construct a family of solvable large N SU(N) matrix systems which are not tractable by
other methods.

A.Yu. Dubin / Nuclear Physics B 584 [PM] (2000) 749783

751

In the general setting, the considered duality transformation provides with the appropriate mathematical framework to operate with the results of the multiple U (N) or SU(N)
er of an arbitrary D-matrix model
integrations like those defining the partition function X
X
X
 p

(s)
eA({wq }) Re q=1 tr(Uq Uq . . . Uq ) ,
(1.1)
eSr ({U }) =
n(+)Z

>0 {w (s) }
q

n(+)

-sum runs over all cyclic-symmetrized words wq(s) [q q . . . q ](s)


where the {wq(s)}n(+)

(made of the 2 D different -, 1 -letters) of lengths nk which can be composed


Pp

(Conventionally,
from the total number n(+)

of the {U } factors: q=1 nq = n(+).


Re[ ] denotes the real part.) Extending the abelian transformation [14,15], we propose the
following dual variables. These are the relevant -species of the integer-valued sets {i (),
i = 1, . . . , N}, (labeling the associated U (N) or SU(N) representations (irreps) R )
complemented by the elements of the symmetric groups {S(n )}.
On the one hand, the latter elements (together with some of the -species of R ) are
encoded in the 1-link integration formula [18,19] that in Section 2 will be manifestly
rewritten in terms of the dual variables. On the other hand, the remaining species of R
enter certain S(n ) operators which provide with the parametrization of the actions
(s)
(1.1) alternative to the one in terms of the words wq . Algebraically, the S(n )-valued
degrees of freedom (being composed into the characters of the tensor representation
associated to the common enveloping space S(n )) act as the Lagrange multipliers.
They facilitate the nonabelian fusion-rule constraints for the complementary integer-valued
fields {()} entering the construction within the canonical S(n )-valued Young projectors
(N)
PR , R Yn .
N
Complementary, the pivotal role of the n S(n)-variables in fact foreshadows a tight
relation to the D > 2 Gauge String construction generalizing the GrossTaylor 2d stringy
pattern [2,3]. Actually, the D > 3 nature of the string amplitudes is encoded [22] in the
outer-product structure of certain combination of the S(n )-blocks embedded to act in
the common enveloping space of the S(n+ ) = S(n ) algebra. The technique, dealing
with such compositions, is naturally inherited from the formalism developed in the present
analysis given in Sections 2 and 3.
Employing the nonabelian duality we design a mapping between the recently proposed
solvable D-matrix eigenvalue-theories [8] and a novel class of the D-matrix models which
apparently are not of the eigenvalue-type. There are a few reasons why the latter models are
worth studying by themselves. First, it provides with a rare example of solvable multimatrix
models nontrivially depending on the nondiagonal components of the SU(N) or U (N)
matrices involved. As we will see, the mechanism behind their solvability is different from
that in the popular systems computable owing to the built in ItzyksonZuber integral
[9]: the KazakovMigdal model [1012] together with the conventional and conformal
multimatrix systems (see, e.g., [13] for a review). Second, generalizing the prescription
of [8], the new models can be viewed as the large N reduction of the associated DeLD
dimensional lattice gauge theories. In other words, the large N partition function (PF) X
D
of certain gauge theory (defined on a cubic lattice of D-volume L ) can be reproduced

752

A.Yu. Dubin / Nuclear Physics B 584 [PM] (2000) 749783

er )LD ,
eLD = lim (X
lim X

(1.2)

er of the corresponding D-matrix model with the reduced spacetime


through the PF X
dependence. The latter noneigenvalue models are computable owing to the claimed duality
to the basic SU(N) eigenvalue-family [8]
e

(2)

Sr ({U })

S (2) ({R })

D(D1)/2
Y

{R }



R (U )R (U ) 2 ,

(1.3)

=1

formulated in terms of the eigenvalues of the D-matrices U SU(N), or to its minor


modification
(1)

eSr

({U })

eS

(1) ({R })

{R }

R (U+ )R (U+ )

{}

D
Y

R (U ),

(1.4)

=1

both being defined on a single D-cube with periodic boundary conditions: {} =


{1, . . . , D(D 1)/2}. The relevant sums run over all SU(N) irreducible representations
(N)
, {}, {}, and in the case of (1.4) it is postulated that the numbers
(irreps) R Yn()
P
of boxes n() in the associated Young tableau are constrained by n() = D1
6= n().
er(m)
The key-advantage of the eigenvalue-systems (1.3), (1.4) is that their large N PF X
is explicitly computed [8] employing the saddle-point (SP) method applied to the irreps
{R }. To construct the purported solvable noneigenvalue deformations of (1.3), (1.4), we
er(m) of the latter systems in terms [8] of the D-products
first rewrite the PF X
"
#m
X
(D1)
(m)
S({R
})/m
D

er = lim
e
=1 L
(1.5)
lim X
N

{R }

R |{R }

of the generalized LittlewoodRichardson (GLR) coefficients of (D 1)th order


Z


(D1)
eSU(N) R (U
e+ ) D1 R (U
e ) Z>0 ,
LR |{R } = d U
6=

(1.6)

which encode the fusion rules of the SU(N) characters. I assert that the reduction of the PF
to the GLR generating functional (1.5) takes place in a larger, apparently noneigenvalue
variety of D-matrix models like (1.1).
The representation (1.1) is not particularly helpful to distinguish the GLR computable
variety, we are interested in, from the generic D-matrix system. The only explicit general
structure is that the family (1.1) forms a natural hierarchy graded by the rank k = D, . . . , 1,
of the [U (N)]k conjugation-invariance

k
: U() g+ U() g , = 1, . . . , k, g U (N),
(1.7)
U (N)
where U factors are recollected into a set of -families {()}. The deep reason for the
GLR solvability of the k < D noneigenvalue-systems (i.e., for their duality to the k = D
models (1.3), (1.4)) is a specific hidden symmetry. The latter becomes manifest only in
the dual representation for the PF of a k 6 D subvariety of (1.1) (including (1.3), (1.4)).
In the dual reformulation, the GLR coefficients in the PF (1.5) can be referred (owing to

A.Yu. Dubin / Nuclear Physics B 584 [PM] (2000) 749783

753

the SchurWeyl complementarity [16,17]) to the simplest available fusion-pattern of the


S(n )-valued Young idempotents (YI) CR PR . It is the nonabelian duality which, as
we will see, allows to reveal that among (1.1) there are k < D systems with the same (as
in the k = D case (1.3), (1.4)) GLR pattern (1.5) of the underlying YI fusion-rules.
Finally, the organization of the paper is as following. The details of the exact duality
transformation, applied first to the eigenvalue-models (1.3), (1.4), are discussed at length
in in Section 2. In particular, for the latter models we rederive the GLR form (1.5)
er(m) directly in the framework of the dual representation. Building on this
of the PF X
formalism, in Section 3 we construct the GLR computable k 6 D 1 subvariety of the
D-matrix systems (1.1). Among the deformations, we select a k = 1 noneigenvaluefamily (see Eqs. (3.11), (3.13)) specifically suitable for the discussion of the continuum
limit (CL) in the corresponding D-dimensional induced gauge theories. The algorithm
to reconstruct the latter theories is formulated in Section 4. To address the issue of the
CL, in Section 5 we first transform the associated large N GLR functional (1.5) into the
1-matrix representation [8]. Then we prove that the simple criterion imposed on the latter
1-matrix system (formulated in [8] for the k = D family (1.3), (1.4)) is valid in the case of
the selected k = 1 variety as well. For this purpose, following [8] we demonstrate that the

link-variables in the corresponding induced lattice gauge theory are localized {U (z) 1}
(modulo the relevant symmetries) which is tantamount to the regime of the CL.
Our conclusion briefly summarizes the results obtained. Also we announce the pattern
of the D > 3 gauge string representation of the strongly coupled YMD theories developed
in [22]. Those features of the latter are emphasized which are novel compared to the earlier
stringy pattern [24,25] of the D > 3 SC expansion in the large N lattice gauge theories.
A few appendices contain relevant technical details of the derivations used in the main text.

2. The dual form of D-matrix models


It is instructive to introduce the concept of the nonabelian Duality extending the abelian
construction [14,15] (dating back to the classical paper due to Kramers and Wannier who
discovered the selfduality of the 2d Ising model). For a brief review of the latter, take the
most relevant for our analysis option of the U (1) lattice gauge theory. Its D > 2 partition
function is defined as the multiple link-integral
Np Nl Z
X
Y
Y
U (1)
eU (1) =
Z(U (p)),
Z(U ) =
n (U )Q(n),
(2.1)
dUl
X
Np ,Nl
p=1 l=1

nZ

of the product composed of the standard U (1) plaquette-factors (e.g., for the Villain
Polyakov action Q(n) = exp[g 2 n2 /2]). In Eq. (2.1),
Z
Z
U (1)
U (1)
il
=e ,
dUl
dl /2,
Ul

= ei[5](p)

is the holonomy around the pth elementary plaquette of the basewhile U (p)
lattice {p; l} consisting of Nl links and Np plaquettes.

754

A.Yu. Dubin / Nuclear Physics B 584 [PM] (2000) 749783

In the abelian case, the crucial simplification arises due to the fact that all U (1) irreps
(labelled by the integers n Z) are one-dimensional, while the U (1) characters n (U ) =
U n = ein form the so-called character-group [14,15], n1 (U )n2 (U ) = n1 +n2 (U ),
isomorphic
to Z. Combining it with the simple structure the generic U (1) 1-link integral
R
n (U + )m = [n, m], one easily derives the dual form of the partition function
dU
U
U (1)
(2.1). It is represented by the constrained multiple sum over the integer-valued variables
(assigned to the plaquettes)
" 2(D1)
! #
Np
Nl
Y
X
Y
X
U
(1)
e
Q(n(p))

n(pl ) , 0 ,
(2.2)
X
Np ,Nl =
p=1 n(p)Z

p l =1

l=1

with the weight Q(n) being introduced in Eq. (2.1). In Eq. (2.2) the sum in the argument of
the Kronecker delta-function, running over the 2(D 1) plaquettes pl which share a given
link l in common, represents the relevant U (1) fusion-rule algebra. Remark also that the
relevant U (1) (and more generally U (N) or SU(N)) D-matrix models (1.1) are defined on
the reduced base-lattice with the topology of a D-dimensional cube with periodic boundary
conditions.
As we will see, in the nonabelian D-matrix models, the fusion rules can be imposed
combining Kronecker delta-functions with the trace (i.e., character) of the appropriate
element of certain symmetric groups algebra in the tensor representation. To derive the
nonabelian counterpart of (2.2), we first introduce the dual representation of the U (N)
1-link integral making manifest the latter structure.
2.1. The dual form of the U(N) measure
Recall that the functional measure in nonabelian lattice vector-field theories, considered
as a distribution, can be defined specifying all the moments of this distribution. On a
given base-lattice, these moments are specified defining at each link the set of generic
p ...q
1-link integrals M G (n, m)j11...lmm
Z
Z
p
p
q
q
{p n }
{q m }
(2.3)
dU(U )j11 (U )jnn (U + )l11 (U + )lmm dUD(U ){j n } D(U + ){l m }
composed from the N N matrices (U )jkk , (U + )lkk in the (anti)fundamental representation
of the Lie group G in question. The crucial observation is that M G (n, m) can be dually
reformulated in terms of the S(n)-valued variables. In particular, in the G = U (N) case
M U (N) (n, m) reads
X
{q n }
p ...q
{p n }
D 1 (1)
D(){l n }
(2.4)
M U (N) (n, m)j11...lmm = [n, m]
n
{j n }
p

S(n)

generalizing the U (1) pattern. The derivation of the identity (2.4) is given in Appendix B,
and here we simply explain the meaning of its building blocks. The factor D( ) stands
for the canonical tensor representation of a given S(n)-algebra element deduced (by
linearity) from the representation [16,17] of a S(n)-group element
{i n }

D( ){j n } = j1 (1) j2 (2) jn(n) ,

: k (k), k = 1, . . . , n,

(2.5)

A.Yu. Dubin / Nuclear Physics B 584 [PM] (2000) 749783

755

while the introduced in (2.4) operator (1)


S(n) can be viwed as belonging to the family
n
X
dR (n! dim R/dR )m CR , m Z,
(2.6)
(m)
n =
(N)

RYn

which is expressed in terms of the dimensions dR and dim R of the S(n)- and chiral U (N)irreps R Yn(N) , respectively, with the Young tableaus Yn(N) containing not more than N
rows. As for the operator CR S(n) in Eq. (2.6), it denotes the canonical Young idempotent
proportional PR = dR CR to the Young projector PR [16,17] (see Appendix A for more
details)
dR X
R ( ) , R Yn ,
(2.7)
PR =
n!
S(n)

where R ( ) is the corresponding character. Summarizing, one observes that the S(n)representation (2.4) of the 1-link integral (2.3) establishes a remarkable duality extending
the SchurWeyl complementarity (see Appendix A for a review of the latter). Considered
as the operators acting on |pin |qin -space, the left- and right-hand sides of Eq. (2.4)
belong to the complementary structures: the Lie group ring and the symmetric group
algebra (being augmented by the integer-valued fields parametrizing irreps R) respectively.
Next, in contradistinction to the U (N) case, the SU(N) 1-link integral (2.3) does not
vanish provided that n = m mod N [20] which makes it generically more complex. In the
context of the D-matrix systems, the important simplification arises because (without loss
of generality) the SU(N) action (1.1) can be restricted to be invariant under the D copies
of the transformations
[U (1)]D : U t U ,

t U (1),

(2.8)

taking values in U (1) rather than in the center-subgroup T = ZN of SU(N). In other words,
the total amounts n () of the U factors in each trace-product of (1.1) is constrained
to be equal for each . As a result, the nondiagonal moments M SU(N) (n, m), n 6= m, do
er , while in the remaining diagonal integrals M SU(N) (n, n) the
not contribute into the PF X
SU(N) link-variables can be substituted by the U (N) = [SU(N) U (1)]/ZN ones
M SU(N) (n, n) = M U (N) (n, n),

n Z>0 ,

(2.9)

as it is proven in Appendix B.
2.2. The dual parametrization of D-matrix actions.
Complementary to the reformulation (2.4) of the measure, a generic D-matrix action
(1.1) can be rewritten in a more concise synthetic form combining both the normal Uji variables and the dual degrees of freedom. As a result, integrating out {U } with the help of
er of (1.1) can be expressed in terms of the dual variables
the S(n)-formula (2.4), the PF X
only.
It is appropriate to recall first the synthetic representation of the SU(N) group characters
(see, e.g., [6])

756

A.Yu. Dubin / Nuclear Physics B 584 [PM] (2000) 749783

R (U ) =





1 X
R ( ) Trn D( )U n = Trn D(CR )U n ,
n!

(2.10)

S(n)

(N)

associated to the set of irreps R Yn with the Young tableaus containing a given number
n of boxes. Eq. (2.10) simply rewrites the Frobenius formula [20] in terms the conventional
algebraic notations
N
X



Trn D( )U n =

i1 i2 ...in =1

Ui1 (1) Ui2 (2) Uin (n) =

n
Y


tr(U k )

 pk

(2.11)

k=1

where [1p1 2p2 npn ], i.e., belongs to the S(n) conjugacy class [ ] defined by the
P
associated partition of n: nk=1 k pk = n. Note also that the complete set of R (U ) is
expressed (see, e.g., [20]) with the help of (U )ij -factors, while (U + )ij is not engaged.
At this step, it can be helpful to make a cross-check of the technology applying the
proposed representation (2.4) (combined with Eqs. (2.9) and (2.10)) to some known
nontrivial system. For this purpose we refer to Appendix D, where the dual form of the
partition function of the SU(N) gauge theory on an arbitrary 2d manifold is rederived
matching with the result [5,23] previously obtained by a different method.
Next, let us generalize the 1-cycle construction (2.11) for the case of the D-matrix
action (1.1), To this aim, observe first that topologically each individual trace tr(U k ) in
Eq. (2.11) can be visualized by the k-fold winding of a path around a single base-cycle
associated to the U -factor. More generally, one is to consider a single closed q-loop
tr(Uq Uq Uq ). One notes that the latter trace can be visualized now by a path wrapped,
(s)

according to the structure of the associated word wq , around the D independent -cycles
of the base-lattice. In turn, any (C(2m)-cyclic symmetrized) word . . . (constrained for
L
simplicity to be invariant under (2.8)) of a length 2m = 2 D
=1 m can be reproduced
#
"
D
O

m
+ m
(U )
(U )
,
(2.12)
tr(U U U ) = Tr2m D({m} )
=1

with the help of the equivalence class [{m} ] of the 2m-cycle permutations {m}
C(2m) defined modulo certain conjugations (immaterial for our present discussion). More
explicitly, the structure of the r.h.s. of Eq. (2.12) adopts the pattern (2.11) to the presence
of the 2D different U , U+ basis-factors


i(m )

(U1 )jm
(U1 )j(1)
1

k(m +1)

(U1+ )lm

+1

k(2m ) 

(U1+ )l2m


k(2m) 
(UD+ )l2m
,

(2.13)

where the mapping n (n), n = 1, . . . , 2m, defines the S(2m) permutation {m} .
Next, Eq. (2.12) by the same token represents a generic product of traces (with 2m =
n(+))

entering the [U (1)]D -invariant D-matrix action (1.1). For this purpose, one is to
choose such ({wk(s) }) S(2m) that can be decomposed into the ordered product =
p( )
P k=1 cnk of the nk -cycle permutations cnk reproducing the trace-product originally
(s)
. Summarizing, the exponent of the D-matrix action (1.1) can
parametrized by {wk }n(+)

be reformulated in the following synthetic form

A.Yu. Dubin / Nuclear Physics B 584 [PM] (2000) 749783

Sr ({U })

Re

"
X

{n }

"
{n } ( ) Tr4n+

D
O
(U )n (U+ )n
D( )

757

##
, (2.14)

=1

P
where n Z>0 , 2n+ = D
=1 n , and S(4n+ ).
As our attention is restricted to the solvable deformations of (1.3), (1.4), we impose extra
condition that {n } ( )is functionally parametrized by a set {R Yn(N)
} of the relevant
SU(N) irreps
{n } ( ) =

eS({R }) {n } ( |{R }),

{R }

n =

D1
X

n .

(2.15)

6=

As a result, the S(4n+ )-algebra elements (defining the D-matrix action (2.14), (2.15))
X
X
{n()} ( ) =
eS({R }) 4n+ ({R }),
(2.16)
4n+ =
{R }

S(4n(+))

can be resummed in terms of the alternative set of the (properly normalized) operators
4n+ ({R }) S(4n+ ), {}, {}, where it is convenient to separate out a weightfactor eS({R }) to be specified below.
Summarizing, in the nonabelian case the dual representation introduces the extended
(compared to (2.2)) set of the dual variables: the integer-valued {()}-fields parametrizing
N
the relevant irreps R are complemented by the elements of the n S(n)-algebra. In the
particular case of the SU(N) GLR generating functionals (1.3), (1.4), the pertinent dual
degrees of freedom fit the pattern
)
(
O
 N
D(D1)/2  N
D
S(n), {} Z /S(N)
Z /S(N)
,
(2.17)
n

For a given {}, {}, each {()}-sector is composed of the sets of N nonnegative
integers {} = {1 > 2 > > N > 0} (and N = 0 for {}). The latter enter the
scene through the relevant Young idempotents CR .
Let us also remark the advantage of choosing the SU(N) option of the [U (1)]D invariant D-matrix models (1.1), where the SU(N) link-variables can be extended
(according to (2.9)) to the U (N) ones. In this way, we combine the simpler structure of the
U (N) 1-link integral (2.4) with the more compact pattern (2.10) of the SU(N) characters
(implicitly entering the action through the operators (2.16)).
2.3. D-matrix amplitudes v.s. Tr 4n(+) -characters
Let us now put together the dual pattern (2.4) of the measure and the synthetic
representation (2.14) of the D-matrix action. As a result, one can to express the D-matrix
er as the weighted sum of the master-integrals
PF X
Z
Tr4n(+) [D(A4n+ )] =

D
Y

Tr4n(+) D(4n+ )D({U U+ })
dU ,
=1

(2.18)

758

A.Yu. Dubin / Nuclear Physics B 584 [PM] (2000) 749783

represented as the characters of the corresponding master-elements A4n+ which belong to


the tensor representation of the S(4n(+))-algebra. In Eq. (2.18) we have defined (following
Eq. (2.14))
D
 O

(U )n() (U+ )n() ,
D {U U+ )}

(2.19)

=1

where 4n+ S(4n(+)) is introduced in Eq. (2.16) so that the block (2.18) is associated
in Eq. (2.14) to the subset of the terms summed up for the particular partition {n n()}
of a given 2n+ 2n(+).
To derive the master-element A4n+ , in Eq. (2.18) the result of the D different U integrations (2.4) is to be represented as an operator embedded to act in the same
enveloping S(4n(+))-space where both 4n+ and the complementary block (2.19) (being
considered as the operator) act. Let us first specify a S(4n(+))-basis suitable to accomplish
our program. As it is reviewed in Appendix A, each individual matrix U or U+ can be
viewed as the operator acting on the associated elementary subspace
N
X

b |i ()i =
U

j ()

(U )i() |j ()i,

j ()=1

b+ |i+ ()i. Thus, a given realization of the S(4n(+))-basis is to be


and similarly for U
constructed as a properly ordered outer product of the elementary building blocks |i ()i.
In particular, the ordering of the {U }-factors in Eqs. (2.18), (2.13) is associated to the
following basis
|I4n(+) i =

D
O

|I2n() i,

(+) ()
|I2n() i = In() In() ,

(2.20)

|In() i = |i ()in() |i ()in() ,

(2.21)

=1

O ()
() D1
I
I
=
n()
n() ,

()

6=

P
where 2n(+) = D
=1 n(). As for the S(4n(+))-group operators (represented by
Eq. (2.5)), they are postulated to act as the corresponding permutations of the elementary
subspaces |i ()i (see Appendix A for more details).
Returning to the S(4n(+)) representation of the D different 1-link integrations, it is
more effective to employ the alternative S(2n)-reformulation of the S(n) S(n) formula
(2.4) (see Appendix B). In the |I2n i = |In(+) i|In() i basis (where |In() i = |i in matches
()
i of Eq. (2.21)) it reads
with |In()
Z
jn+1 ...j2n
j ...j
{j 2n }
(1)
1 [n] )){i 2n } ,
(2.22)
dU D(U )i11...inn D(U + )in+1
...i2n = D(2n (2n)(n
where 1 [n] denotes the unity-permutation of the S(n) group, n S(n) is defined by
P
Eq. (2.6), while (2n) = S(n) ( 1 ), i.e.,
X
j1 ...jn
jn+1 ...j2n
{j 2n }
D( 1 )i1 ...in D()in+1
(2.23)
D( (2n)){i 2n } =
...i2n S(2n).
(1)

S(n)

A.Yu. Dubin / Nuclear Physics B 584 [PM] (2000) 749783

759

Let us restore the -labels, i.e., n n(). For a given link , the left and the right
(+)
i,
S(n())-sub-blocks of (2n()) in Eq. (2.23) act, respectively, on the chiral, |In()
()

and the antichiral, |In() i, S(n())-subspaces of |I2n() i entering Eq. (2.20). The same
(1)
1 [n()] )
convention is used for the S(n())-sub-blocks in the direct product (
n()

entering Eq. (2.22). The remaining S(2n())-operator 2n() , being considered in the
alternatively ordered basis |I2n() i for each |I2n() i-subsector
n()
,
(2.24)
|I2n() i |I2n() i = |i+ ()i |i ()i
n() ), takes the simple form of the outer product of
(with |i ()in() = D1
6= |i ()i
the 2-cycle permutations c2 C(2)

2n() = (c2 )n() S(2n()),

c2 : {12} {21},

(2.25)

with each c2 S(2) acting on the elementary sector |i+ ()i |i ()i. Combining all the
pieces together, the master-element A4n+ is supposed to be constructed as the composition
of the involved (into Eqs. (2.16) and (2.22)) S(n())-valued operators embedded to act in
the common enveloping S(4n(+))-space.
Let us apply this algorithm in order to rederive the GLR pattern (1.5) directly evaluating
the master-integrals (2.18) associated to the eigenvalue-models (1.3) or (1.4). Given the
basis (2.20), the corresponding to (1.4) operator (2.16) can be rewritten (after identification
S (1) ({R }) = S({R })) in terms of
(1)
({R }) =
4n
+

D
O

(1)
K2n()
,


O

(1)
D1
K2n()
= CR
6= CR ,

(2.26)

=1
(2)

while the substitution CR [D1


6= CR ] reproduces K2n() of the option (1.3). By
()
i
definition, each CR -factor in Eq. (2.26) is postulated to act on the corresponding |In()

(+)
i subspace.
subspace of (2.21), while each CR -factor acts on the associated |In()
Actually, a preliminary variant of the GLR pattern (1.5) can be deduced from (2.26)
already at this step. For this purpose, one is to combine the the peculiar structure of the
(m)
dual 4n+ ({R })- operator (2.26) with the invariance of any D-matrix action (2.14)(2.16)
with respect to the substitution of 4n+ by its twisted partner
X
[+ ]1 4n+ [+ ], = D
(2.27)
4n+
=1 (),
{ ()}
()

where () S(n()) is postulated to act on the corresponding |In() i-subspace of the


alternative S(4n(+)) basis
()

0
(+) O ()

D ()

I
I
I
(2.28)
4n(+) = I2n(+)
2n(+) ,
2n(+) = =1 In() ,
()
(m)
i being defined by Eq. (2.21). Performing the substitution (2.27) for 4n
({R })
with |In()
+
defined by (2.26), we finally employ the fusion rules of the Young idempotents (see
Appendix D)

M (p)
X [ (p CRk ) 1 ]
k=1
=
LR+ |{Rk } CR+ ,
(n+ )!

S(n+ )

R+ Yn+

n+ =

p
X
k=1

nk ,

(2.29)

760

A.Yu. Dubin / Nuclear Physics B 584 [PM] (2000) 749783

to arrive at the structure which foreshadows the D-products (1.5) of the GLR coefficients.
As for the invariance (2.27), it is a consequence of the basic commutativity
[D( ), U n ] = 0,

S(n)

(see Appendix A). The latter ensures that the ordered product (2.19) (associated to the
|I4n(+) i-basis (2.20)) is equal to its twisted counterpart
D
O
X

+ () ()

1 



(U )n (U+ )n + () () ,

(2.30)

=1 { ()}

where () S(n()). Being rewritten in the |I 0 4n(+) i basis (2.28), it matches with
(2.27). Remark also that, compared to Eq. (1.6), the decomposition (2.29) involves a larger
set of the GLR coefficients of pth order: the involved irreps R are parametrized by the
S(n ), rather than SU(N), Young tableaus Yn (see Appendices A and D for the relevant
details).
(m)
Returning to the derivation of the master-element A4n+ associated to the eigenvaluemodels (1.3), (1.4), one is to put together Eqs. (2.18), (2.22) and (2.26). Prior to the twisting
(2.27), it results in
(1)

(1)
A4n+

= D
=1 2n()

(2n()) dR K2n()
,
n!
dim R

(2.31)

N D1
(1)
([D1
[6= CR ]) reproduces A(2)
while the substitution K2n()
6= CR ]PR
4n+
corresponding to the option (1.3). To derive the GLR functional (1.5). we first substitute
(m)
({R }) by its (+ )-twisted partner (2.27). Combining (2.29) with
the operator 4n
+
the identity which in the |I2n i-basis reads
(2n) [+ ] = [ 1 [n] ] (2n) [+ 1 [n] ],
inside the character (2.18) one can substitute
(2n())
(1)
A 4n+ = D
=1 2n()
n()!

(1)
A4n+

(1)
A 4n+

S(n),
where

(D1)
X  LR |{R } CR
(N)
R Yn()

dim R

(2.32)


1 [n()] .

(2.33)

(2)

Similarly, for the option A4n+ one reproduces the product of the GLR coefficients which
matches (after some auxiliary trick discussed in Section 5.1) with the m = 2 case of (1.5).
To ensure that the remaining factors in (2.33) conspire to reproduce exactly the correct
GLR functional (1.5), one is to use first the defining property of 2n which in the basis
|I2n i = |In(+) i |In() i assumes the form




(2.34)
Tr2n D (+ ) 2n = Trn D(+ ) , S(n),
where acts on the corresponding |In() i-subspace. Combining (2.34) with the completeness condition
R ()
1 X
R1 ()R2 ( 1 ) = R1 ,R2 1
(2.35)
n!
dR1
S(n)

A.Yu. Dubin / Nuclear Physics B 584 [PM] (2000) 749783

761

and the projection-formula, which tells that Trn [D(PR )] (where PR = dR CR and R
Yn(N) ) is nonzero only for R Yn when
Trn [D(PR )] = dim RR ( ),

R Yn(N) ,

(2.36)

we finally rederive (1.5).

3. GLR-computable D-matrix PFs


Now we are in a position to formulate the concept of the hidden symmetry inherent in
a subvariety of the [U (1)]D -invariant D-matrix systems (1.1) graded by the rank k =
1, . . . , D of the conjugation-invariance (1.7). This subvariety is specific because its fusion
rule algebra is resolved in terms of the GLR coefficients (1.6) (see Eqs. (3.11), (3.13) for
the main results used in the analysis of Section 5).
The idea is to induce the k < D GLR solvable systems generalizing the k = D operator
(m)
({R }) (defined for m = 1 in Eq. (2.26)) in such a way so that the following basic
4n
+
property of the latter operator is preserved. Namely, the invariance of a generic Dmatrix action under the twisting (2.27) of 4n+ must be transformed into that under the
(m)
complementary twisting (akin to (2.29)) of the {CR }-factors entering 4n+ ({R }) in the
master-integral (2.18). One easily observes that the required symmetry (with respect to the
switching of the (+ )-twist (2.30) from the {U }- to the {CR }-block) holds true
provided in Eq. (2.18) 4n+ is chosen in the following form
X
(m)
eE({R }) 4n+ ({R }) 4n+ ({R }),
(3.1)
4n+ =
{R }

(m)
({R }) is determined by Eq. (5.4), E({R }) is some numerical weight-factor,
where 4n
+
while {}, {} and m = 1, 2. As for 4n+ S(2n+ ), it should commute with any
0 i-basis (2.28))
element of the group-product (in what follows we employ the |I4n
+
 O D

 D
=1 S(n ) ,
(3.2)
=1 S(n )

which is inherent in the twisting (2.30). The latter condition can be concisely formalized
by the pattern
X
1
[+ ]1 M4n+ ({R }) [+ ],
(3.3)
4n+ =
2
[n(e
S)]
e
{ S}

N
where e
S e
S is any subgroup of S(2n+ ) S(2n+ ) which includes (3.2). As for M4n+ ,
a priori it may be a generic element of the S(4n+ ) algebra (consistent with the convergence
S) denotes the number of
of the final summation (2.14) over {n } [Z>0 ]D ). Finally, n(e
the elements in e
S.
S S(2n+ ), the resulting
Depending on the choice of M4n+ and the admissible subgroup e
(via (3.1) and (2.18)) D-matrix action (1.1) is endowed with the conjugation-symmetry
(1.7) of a different rank k. In what follows, we concentrate on the simplest case when
e
S) = (2n+ )! . It results in the k = 1 subvariety of the models (1.1) which
S = S(2n+ ), n(e
N

762

A.Yu. Dubin / Nuclear Physics B 584 [PM] (2000) 749783

can be viewed as the noneigenvalue deformations of the two basic k = D systems (1.3),
(1.4).
To develope an intuition in what is going on, let us obtain the explicit form of the
associated D-matrix actions with the arguments restricted to the subspace of the coinciding
link-variables: U = U, = 1, . . . , D. In this case, there appears the larger symmetry
under the switching of the extended (S(2n+ ) S(2n+ ))-twisting (which generalizes
(2.30))
X


(+ )1 (U )2n+ (U + )2n+ (+ ),
(3.4)
{ S(2n+ )}

onto the complementary {CR }-block of (3.1). To be more specific, consider the m = 2
option of (3.1) and choose the simplest separable form of M4n+
(1)

(2)

(1,2)

M4n+ = M2n+ ({R }) M2n+ ({R }),

M2n+ S(2n+ ),

(3.5)

where {}, {}. Taking into account the identity


!
 1
ND
D1
D
X Y
X
=1 {PR [6= CR ]}
(D1)
(D)
=
LR |{R } LR+ |{R } CR+ ,
(2n
)!
+

R+ =1
P
(where S(2n+ ), R+ Y2n+ ) and summing up R Yn PR = 1, one obtains
e

(2)
e
Sr ({U })

E({R })

2 h
 (q)  Y
L(D)
=
R
(p)
p=1

(q) 
R


(p) 
(p)
2 (p) M
(p) U
Y
2n+ R+
R+
dR (p)

p=1

(q)
{R }

(p)

R+ |{R }

(D1)
D
=1 L (p)

R |{R }

(3.6)

(3.7)

i

(3.8)

where U (2) U + , U (1) U and +, {}, {}.


Next, the analysis of the continuum limit (in the associated induced gauge theory) will
require the knowledge of the explicit relation between the two weights: E({R }) (entering
e
e
eSr ({U }) eSr ({U }) |{U =U } ) and S({R }) involved into the GLR computable D-matrix
partition function (1.5). To derive a transparent example of such a relation, we concentrate
on the deformations of (1.4) and specify the operator M4n+ further in the form generalizing
(3.5)
 (1)

e F2n+ ) M
e(2) (4n+ ) ,
(3.9)
M4n+ ({R }) = (M
2n+
2n+
(2n+ )!
where (4n+ ) S(4n+ ) is defined by Eq. (2.23) (with each S(2n+ )-subblock acting on
()
i-subspace of (2.28)). The auxiliary factor
the corresponding |I2n
+
X
eE({R },R+ )+E({R }) PR+ .
(3.10)
F2n+ =
(N)
+

R+ Y2n

is introduced to trade E({R }) in the final amplitudes for its R+ dependent counterpart
(and for simplicity we consider {R }-independent weights).

A.Yu. Dubin / Nuclear Physics B 584 [PM] (2000) 749783

763

To begin with, employing (3.6) and the identities listed in the end of the previous section,
one easily obtains for the associated to (3.9) deformation of the eigenvalue-action (1.4)
(considered for the coinciding arguments)

e(2)
e (1) M
X

M


(1)
R
+
2n
2n
e
(q)
+
+
eE Q R
|R+ (U )|2 ,
(3.11)
eSr ({U }) =
3
d
R+
(q)
{R }

2
 (q)  Y
L(D)
=
Q R
p=1

(p)
R+ |{R }

(D1)
D
=1 L (2)

R |{R }

(3.12)

er associated to (3.11). We refer the


It is not difficult to evaluate the partition function X
reader to Appendix E for the details and now simply write the final result

e(2) dim R+
e (1) M
X

M
R
+
2n
2n
+
+
e })
er =
eE({R },R+ ) Q({R
,
(3.13)
X
3
D
d

dim
R
=1
R
{R }
+
(D1) 
D
e }) = L(D)
(3.14)
Q({R
R+ |{R } =1 LR |{R } ,
where the sum runs over the SU(N) irreps {R }, +, {}, {}.
In conclusion, one observes that both in Eqs. (3.13), (3.11) there appears the same
3
e (1) M
e(2)
factor K(R+ ) = R+ (M
2n+ 2n+ )/dR+ violating the invariance under the Z2 -conjugation:
R R , where {}, {}, +. To retain this auxiliary symmetry (and make
contact with S({R }) in (5.7) of Section 5), we redefine
e

eE({R },R+ ) K(R+ ) = eE({R },R+ ) = eS({R })

D
=1 dim R
dim R+

(3.15)

e }, R+ ) is Z2 -invariant.
postulating that E({R
4. Mapping onto the SU (N ) gauge theory
In [8] we have developed the algorithm that associates to the k = D eigenvalue models
like (1.3), (1.4) the D-dimensional induced lattice gauge theory in such a way that the
correspondence (1.2) (between the the PFs) holds true. Our present purpose is to generalize
[8] and induce, preserving (1.2), gauge theories from the generic k 6 D D-matrix systems
(1.1) invariant under (1.7) and (2.8) (including those belonging to the GLR computable
variety defined via Eqs. (3.1), (3.3)).
To begin with, taking (1.3) as an example, let us briefly review the the original
prescription [8] which consists of the two steps. First, one employs the large N saddlepoint method to prove that the SU(N) system (1.3) is reduced (eliminating the spacetime
dependence) from the following D-dimensional eigenvalue-system. The latter is defined
associating to each site x (of LD lattice) the factor
Y
X
eS
R (U (x))R (U (x + ))R (U+ (x + ))R (U+ (x)), (4.1)
{R }

{}

764

A.Yu. Dubin / Nuclear Physics B 584 [PM] (2000) 749783

where S S({R }) and R R (x)). Observe that the correspondence between (1.3)
and (4.1) implies the particular choice of the mapping {U } {U (z)} between the linker and X
eLD , respectively.
variables entering the PFs X
The constructed in this way intermediate system (4.1) is invariant under the local
[U (N)]D conjugation-symmetry
U (z) h+
(z) U (z) h (z),

h U (N),

= 1, . . . , D,

(4.2)

combined with the reduced gauge symmetry with respect to the center T of the Lie group
U (z) H + (z) U (z) H (z + ),

H (z) T ,

T = ZN ,

(4.3)

complemented by the global [ZN ]D -invariance


[T ]D : U (z) t U (z),

t T = ZN .

(4.4)

The latter two symmetries substantiate consistency of the second step: the gauge theory is
induced from the system (4.1) through the gauge transformation
U (z) G+ (z) U (z) G(z + ),

= 1, . . . , D,

(4.5)

introducing the auxiliary SU(N) scalar field G(z) assigned to the lattice sites. Integration
R
over G(z) with the Haar measure (normalized by dG(z) = 1) results [8] in the associated
effective theory with the manifestly gauge-invariant action e
Seff ({U (z)}).
The pairing between the local conjugation-symmetry (4.2) and its global k = D
counterpart (1.7) is crucial for maintaining the large N correspondence (1.2) between (4.1)
and its reduced partner (1.3). Upon a reflection, the k < D case of the noneigenvalue Dmatrix models (1.1)(3.1) can be easily reduced to the previous k = D option: one is to map
the former models onto the associated effective D-matrix eigenvalue-theories. Afterwards,
the considered above algorithm becomes applicable provided a minor modification of the
pattern (4.1) of the associated intermediate D-dimensional eigenvalue-theory.
To fulfil this program, let us start with the construction of the effective eigenvalue-theory
associated to a given k < D model (1.1) which for simplicity is restricted to be invariant
under (2.8), Additionally, the total amounts n of the U - (or, equally, U+ -) factors in each
trace product of (1.1) are constrained by
n =

D1
X

n ,

n Z>0 .

(4.6)

6=

To begin with, in the SU(N) action (1.1) one is to employ (2.9). Having decomposed the U (N) link-matrices: U = diag[ei() ] + , we then integrate over
U (N)/[U (1)]N , = 1, . . . , D. One observes that the integrations over the right-cosets
can be extended, d d( Te ) dW to those over W spanning the full
U (N) group-manifold (introducing the auxiliary matrix Te [U (1)]N ). Complementary, the remaining integrals over T diag[ei() ] [U (1)]N by the same token can
e to those over U
e U (N) (introducing
be promoted, dT d( T + ) d U
N
U (N)/[U (1)] ).

A.Yu. Dubin / Nuclear Physics B 584 [PM] (2000) 749783

765

Altogether, in terms of this extended set of the variables, the PF of a generic D-matrix
model (1.1) can be rewritten
Z
Z
D
D
Y
Y
e })
e })
Seff ({U
Seff ({U
e
e
e
e
d U =
e
dU
(4.7)
Xr =
=1

U (N)

=1

SU(N)

as the PF of the associated effective eigenvalue-theory manifestly invariant under the k = D


conjugation-symmetry (1.7)
" Z
#
D
Y
+
e
e
e }) = ln
eSr ({W U W })
dW ,
(4.8)
Seff ({U
=1

U (N)

where U (N) dU = 1 is assumed. To return from (4.7) to the original representation, one
absorbs by the opposite shift W W + of the W -variables and then employs the
commutativity Te diag[ei() ] Te+ = diag[ei()]. As for the maximal k = D symmetry
(1.7) of (4.8), it follows from the possibility to reabsorb the g -rotations in the same way
as we have done for .
To reconstruct an intermediate D-dimensional eigenvalue-system associated to the
resulting effective SU(N) eigenvalue-theory (4.8), let us first represent the action of the
latter in the form
" D
#
X X
Y
(q)
e })
e
Sr ({U
A({R })
+
e
e
=
e
Re
R () (U )R (+) (U ) ,
(4.9)
e

{n } {R (q) Y (N) }

=1

(q)

(q)

where A A({R }) and the associated to {R } numbers of boxes {n (q)} satisfy (owing
to (2.8)) n (1) = n (2) n , . As we demonstrate in Appendix F, the algorithm of [8]
is not directly applicable to the more general class (4.9) of models. Instead, the pattern (4.9)
can be reduced (eliminating the spacetime dependence) from the following D-dimensional
eigenvalue-systems. The latter are defined associating to each site x the factor (where
(+)
()
U U+ , U U )
#
" D
YY
X X
(q)

A({R })
(q)
(4.10)
e
Re
R (q) U (x) ,
{n } {R (q) Y (N) }

=1 q=

which provides with the mapping {U } {U (z)} alternative to the one encoded in the
eLD of (4.10) is
pairing between (F.1) and (F.2). By the same token as in [8], the PF X
er of (4.9) through the large N correspondence (1.2).
related to that X
Finally, application of the mapping (4.5) (together with the subsequent integration over
G(z)) converts the D-dimensional eigenvalue-system (4.10) into a gauge theory. In sum,
this prescription provides with the algorithm which induces the SU(N) gauge model from
a generic [U (1)]D -invariant D-matrix system (1.1) including the k 6 D family (3.1),
(3.3) with the GLR computable PF. The subtlety is that, in contradistinction to (4.1), the
intermediate eigenvalue-system (4.10) in addition is invariant under the (finite N ) local
[ZN ]D symmetry
[ZN ]D : U (z) t (z)U (z),

t (z) ZN ,

(4.11)

766

A.Yu. Dubin / Nuclear Physics B 584 [PM] (2000) 749783

much larger than the ZN gauge invariance (4.3). In turn, symmetry (4.11) is present in
the induced via (4.5) gauge theory that is known to set zero the average of any Wilson loop
WC (U ) = tr(U (x)U (x + ) U (x )) provided the corresponding to the contour C
(minimal) area does not vanish. Therefore, it calls for a modification of the prescription to
get rid of the unwanted invariance (4.11) keeping (4.3) intact. We refer to Appendix F for
the algorithm which, synthesizing the patterns (4.1) and (4.10), circumvents this problem
so that the final mapping (4.5) allows to induce the theory invariant under the conventional
SU(N) gauge symmetry but not under (4.11).

5. Continuum limit of the induced theories


The analysis of the continuum limit in the lattice gauge theories, induced from the GLR
computable k < D models (1.1)(4.6), essentially follows the route employed in [8] for
gauge theories induced directly from the k = D models like (1.3). In what follows we
briefly sketch the major steps with the emphasis on a few novel details.
The idea is to take advantage of the fact that the required infinite correlation length
in a lattice gauge theory (with continuous link-variables like U (z) SU(N) or U (N))
is supposed to be ensured by the following effect. Namely, the link-variables U (z)
should be dynamically localized (modulo (4.4) and the gauge transformations) in the
The gaugeinfinitesimal vicinity, scaling as O(N (0) ), of the group-unity element 1.
invariant representation of this condition implies the existence of some O(N 0 ) functional
g 2 N g 2 ({gk })N 0 (of the relevant coupling constants {gk }) so that



1
(5.1)
lim lim htr[U (pl)]i 1 O(g 2 N),
N g 2 N0 N
where U (pl) = U (x)U (x + )U+ (x + )U+ (x) stands for the holonomy around an
elementary plaquette in an arbitrary -plane. Let us fix the maximal tree gauge [20]
putting U (z) = 1 on a largest possible tree (made of the links) which by definition
does not contain nontrivial 1-cycles. Then, introducing the quantum fluctuations Aab
(z) =
i ln[Uab (z)], the required localization can be formulated in the large N limit in the form
lim

2
2
lim h[Aab
(z)] i O(g ) mod (4.4),

N g 2 N0

a, b = 1, . . . , N.

(5.2)

Following [8], we intend to prove that in the induced theory the constraint (5.2) is
fulfilled if in the associated D-matrix model (defined for definiteness by Eqs. (3.9), (3.15))
the condition

2
i O(g 2 ) mod (2.8),
(5.3)
lim lim h Aab

N g 2 N0

ab
(where Aab
= i ln[U ]), is valid for any given a, b = 1, . . . , N . In turn, representing
er(m) as the (mD)th power of the
the large N D-matrix partition function (PF) limN X

effective 1-matrix SU(N) theory formulated in terms of irreps R


"
#mD
X00
(m)
S (m) (R|D)
e
e
,
lim Xr = lim
N

(5.4)

A.Yu. Dubin / Nuclear Physics B 584 [PM] (2000) 749783

767

we will demonstrate that the condition (5.3) can be reformulated as the following constraint
on the saddle-point (SP) values of the j Z fields canonically parametrizing the irreps R
(0)
in the effective 1-matrix system (5.4). Namely, the SP values (0)
i = N i should approach
infinity according to the complementary scaling-condition
(0)
(0)


1 
1 1 
2 i O gN
2
,
(5.5)
lim lim i O N/ gN
N g 2 N0

with the functional g({g


k }) (which enters (5.1), (5.3)) tending to zero.
er(m) ] O(N 2 ) and the
As a clarifying remark, the reduction (5.4) implies [8] that ln[X
(m)
(m)
er ) is invariant under the group-product
weight S ({R }) in Eq. (1.5) (defining X
S(D) S(D(D 1)/2) Z2 ,

Z2 : R R ,

(5.6)

combining the separate permutations within the two sets ({} and {}) of the irrep-indices
{}, {} together with the simultaneous conjugation of all the involved into Eq. (1.5)
P
irreps R . Also, the sum R 00 in (5.4) is in fact constrained [8] by the condition that both
must be nonnegative multiples of (D 1) (where n(R) is the number of
n(R) and n(R)
(N)
associated to R).
boxes in the Young tableau Yn(R)
5.1. The effective N 1-matrix theory
To justify the above statements, we first derive the 1-matrix representation (5.4) of the
large N PF (1.5) associated to the D-matrix model (1.1) specified by (3.9), (3.15). Upon a
reflection, the pattern (3.14) suggests to start with a little bit more specific m = 1 form of
(1.5)
X
(D)
(D1) 
er =
eS({R }) LR+ |{R } D
(5.7)
X
=1 LR |{R } ,
{R }

where each sum over R Yn(N)


, {}, {}, +, runs over the -species of the SU(N)
irreps. As for the weight-factor S({R }), it is defined in (3.15) being invariant under
the group-product (5.6), where Z2 symmetry is extended for = +, {}, {}. Next,
integrating out in Eq. (5.7) the auxiliary R+ variable, one brings it into the required
m = 1 form (1.5) with the identification
X
eS(R+ ,{R },{R }) L(D)
(5.8)
eS({R },{R }) =
R+ |{R } .
R+

Returning to the reduction of (1.5) to (5.4), it is built on the localization of the large N
(0)
} of the corresponding saddlesummations over {R }{R } on the solution {R(0) }{R
point equations. We refer to [8] for the discussion of these equations, and now simply
assert the properties of the solution in the case when the constraints (5.6) are additionally
imposed. To be more specific, we select the option when the effective 1-matrix system in
Eq. (5.4) is reduced to the simplest solvable class of the SU(N) or U (N) models with
S(R|D) being defined as
( M
"N 
 # pk )
2n X
0 X
X
Y
N 1 k
S({})
q
= |dim R({})| exp
gr ({p})
,
(5.9)
i
e
2
n=1 r Y2n

k=1 i=1

768

A.Yu. Dubin / Nuclear Physics B 584 [PM] (2000) 749783

where q > 0, and in the SU(N) case the set {gr({p}) } is supposed to maintain invariance of
S({}) under the translations i i + m. The latter is to match with the fact that U (N)
P
U (N)
(constrained by N1
irreps are labelled by a set of N integers i
i=1 (i (N 1)/2) =
N mod N )
{U (N) } = {1 + N > > N1 + N > N } [ZN /S(N)]

(5.10)

generated from the SU(N)-set of N 1 nonnegative integers {SU(N) } = {1 > 2 > >
P
N1 > 0} by the extra integer number N > 0 or N < 0. As for the sum r in the
exponent of (5.9), it runs over the irreps r r ({p}) Y2n of the even symmetric group
P
S(2n) (labelled by the partititions {p} = [1p1 2p2 2np2n ] of 2n: 2n
k=1 kpk = 2n) with
(0)
},
n 6 M0 Z>1 . Under these conditions, the saddle-point SU(N)-set {R(0) } {R
R(0) = R (0) = R (0) ,

(0)
R
= R2(0) = R 2(0) ,

(5.11)

is supposed to be unique, {}{} independent, respectively, and selfdual. Consequently,


the generating functional (1.5) is equivalent in the large N limit to the reduced system
resulting after the identification
,
R R2 Yn(N)
2

R R Yn(N) ,

(5.12)

with the remaining summations over R, R2 being localized on the same saddle-point
values (5.11).
The effective action for R, R2 , resulting from the reduction (5.12), contains (according
to (1.5)) the Dth power of L(D1)
D1 . To simplify this expression further, one can employ
R|R2

the ( = D case of the) identity [8]


lim

Z Y
p X

=1 R

Sp ({R })

"Z

= lim

p X
Y

#
e

Sp ({R })/

(5.13)

=1 R
(0)

(0)

valid provided that > 0, the saddle-point values of both eSp ({R }) and eSp ({R })/
are unique and positive, while the corresponding free energy is O(N 2 ). Summarizing, it
defines the following one-matrix representation of the large N family (1.5)
"Z
#mD
X
(D) (R,R )/mD
S
+
D1
2
er = lim
e
R (U )[R2 (U )]
, (5.14)
dU
lim X
N

R,R2

where the weight S (D) (R, R2 ) is deduced from S({R }, {R }) of (1.5) through dimensional reduction (5.12). Next, owing to the invariance of (1.6) (in (1.5)) under (2.8), the
er is invariant under the substitution (2.9) of the SU(N) linkSU(N) partition function X
variables by the U (N) = [SU(N) U (1)]/ZN ones. Therefore, the sum in (5.14) over
SU(N) irreps R is effectively constrained by the Z2 -invariant pair of the U (N) conditions
(nontrivial in D > 2)
(D1)
R|R2D1

6= 0 n(R) = n(R2 )(D 1),

= n(R 2 )(D 1).


n(R)

(5.15)

A.Yu. Dubin / Nuclear Physics B 584 [PM] (2000) 749783

769

Here the integers n(R ), n(R ) Z>0 denote the number of boxes in the Z2 -invariant pair
of the Young tableaus corresponding to (5.12):
N
N
X
 X
ni =
(i N + i),
n R({}) =
i=1

(5.16)

i=1

while Z2 conjugation R R reads: {i } {Ni+1 + }, where U (N) = (N 1) and


SU(N) = 1 in the U (N) and SU(N) cases, respectively.
Finally, in order to recast Eq. (5.14) into the form of Eq. (5.4), let us first introduce
X
(D)
(D1)
eS (R,R2 )/mD L D1 .
(5.17)
eH (R|D) =
R|R2

R2

By the same token as in [8], the Z2 -invariant pair of the D > 2 SU(N) conditions (5.15) is
(D)
the only constraint defining the whole eH (R|D) -family induced from the eS (R,R2 )/mD variety via (5.17). It suggests to factorize the latter constraints out
X
n(R),[D1]k n(R),[D1]
(5.18)
eH (R|D) = eS(R|D)

k ,

k,kZ

so that, for a fixed D, in (5.4) any residual R-valued function eS(R|D) (consistent with

er ] O(N 2 ) and with Z2 -invariance S(R|D) = S(R|D))


can be induced
the scaling ln[X
(D) (R,R )/mD
S
2
. The two periodic
through (5.17) provided the judicious adjustment of e
Kronecker delta-functions are supposed to be defined [8] via certain -regularization of
the Poisson resummation formula (with explicit form of the latter being immaterial for the
present discussion).
Actually, more detailed analysis [8] shows that the constraints (5.15) are irrelevant in the
proper large N limit. To ensure the latter N regime, in the selected models (5.9) the
er ] O(N 2 ), |j | O(N), can be adjusted [8] provided for each
required scaling ln[X
r({p}) Y2n the following pattern is valid
j = j /N,

gr({p}) = br({p}) N r ,

r({p}) = 2 2n

2n
X

pk ,

(5.19)

k=1

where it is postulated that br({p}) O(N 0 ).


localization
5.2. The {(htr[U (pl)]i/N) 1}
Now we are in a position to prove that, in the gauge theories induced via (4.5) from the
k = 1 system (1.1) (specified by (3.9), (3.15)), the required localization (5.1) of U (z) is
predetermined by the scaling-condition (5.5) for the {} fields entering (5.4), (5.9).
To begin with, we recall that the constraint (5.5) is dynamically fulfilled [8] provided in
(5.9) all coupling constants (rescaled according to (5.19)) tend to zero: {br({p}) 0}. In the
simplest case of the U (N) action (5.9) with M0 = 1, the functional g 2 is to be introduced
by
"
 #
N 
X
X
N 1 2
q
|dim R({})| exp gr0
(5.20)
, g 2 /2 = gr0 ,
i
2
R({})

i=1

770

A.Yu. Dubin / Nuclear Physics B 584 [PM] (2000) 749783

where gr0 O(1/N) and r0 = [21 ]. In a general case (5.9), (5.19), one is to choose b2k =
lim sup [|br |] as the largest |br({p})| in each k-subset of r Y2k . Then, N g 2 is equated with
1

the lim sup [(b2k ) k ] found among all k 6 M0 (provided the associated leading terms, by
themselves, ensure the convergence of the {}-series in (5.9)).
Next, building on the results of [8], one might expect that the scaling (5.5) results in the
complementary localization (5.3) of the link-variables {U } in the D-matrix systems (1.1)
specified by (3.9), (3.15). This assertion, in particular, employs that the action (of the WC
perturbation theory) in the latter system evidently contains (owing to (3.11)) the quadratic
in Aab
term. In turn, the patterns of the involved mappings (4.7) and (4.5) suggest that (5.3)
indeed entails the required localization (5.1) in the gauge theory induced from (1.1)(3.9),
(3.15).
To substantiate these statements by an explicit computation, we consider the large N
WC asymptotics g 2 N 0 of the properly normalized partition function (PF)
Z Y


(in)

dU (z) exp S({U (z)}) S({1})


(5.21)
XLD =
{,z}

associated to a (induced) gauge theory on a cubic lattice with LD sites. In [8] it is shown
that the localization (5.1) generically results in the power-like large N WC asymptotics

1 (D1)N 2 LD
(in)
= C gN
2
, C > 0,
(5.22)
lim lim XLD (g)
N g 2 N0

R
where the dU = 1 normalization of the Haar measure is used, and C is a model dependent
constant. Thus, our aim is to prove that (5.22) is valid in the specific case of the gauge
theories induced from the D-matrix models (1.1)(3.9), (3.15) constrained by (5.5).
For this purpose one first observes that, according to the mapping (4.5), the factor

S({
1})
e(a)D of the auxiliary D-dimensional
can be rewritten as the partition function (PF) X
e
L
model. The latter is deduced from the intermediate D-dimensional eigenvalue-system
(induced on LD lattice via the decomposition (F.5), (F.6)) in the following way. Namely,
in the plaquatte-factor defining the latter eigenvalue-system, one is to substitute the linkvariables U (z) by the composite field
U (z) G+ (z)G(z + )

(5.23)

as it is predetermined by the pattern (4.5). Consequently, the properly normalized PF of


the induced gauge theory can be represented as the ratio [8]
(in)
(a)
XL
D = XLD /XLD ,
(a)

(5.24)

where XLD and XLD are the PFs (both normalized akin to (5.21)) associated to the
intermediate D-dimensional eigenvalue-system and the auxiliary model defined through
(5.23), respectively.
Next, the correspondence (1.2) allows to express the large N limit of XLD as the LD th
power of the PF Xr of the k = 1 D-matrix model (1.1)(3.9), (3.15). As we prove in
Appendix G for the particular case of the latter model, the {}-localization (5.5) indeed
results in the power-like asymptotics

A.Yu. Dubin / Nuclear Physics B 584 [PM] (2000) 749783

lim


1 DN 2
lim Xr ({gr }) = B gN
2
,

N g 2 N0

B > 0,

771

(5.25)

R
e(1)
where dU = 1 is implied. Being combined with the pattern (3.11))(3.15) of eSr ({U }) ,
localization (5.3) of the fluctuations Aab
it ensures the complementary {U 1}
. Similarly
to [8], the pattern (5.25) is tantamount to the following large N scaling-law (for each
particular j )
lim

lim hj2 ()i O(g 2 N) mod (2.8),

N g 2 N0

(5.26)

where j (), j = 1, . . . , N, are the eigenvalues of U = diag[ei() ] + entering the


effective eigenvalue-theory (4.7). In turn, by the same token as in [8], it ensures that in the
auxiliary model (5.23) the SU(N) field G(z) is localized (modulo (4.4)) in the vicinity of
1 so that

1 D 2
(a)
e 2 L N , B
e > 0.
(5.27)
lim lim XL
D ({gr }) = B gN
N g 2 N0

Summarizing, we reproduce the purported asymptotics (5.22) of the PF of the gauge theory
induced from the k = 1 GLR computable D-matrix model (1.1)(3.9), (3.15).
Finally, we note also that the pattern of the large N phase transitions in the gauge theories
induced from (1.1)(3.9), (3.15) can be analysed in essentially the same way as it is done
for the theories [8] induced from the eigenvalue-models (1.3), (1.4).

6. Conclusions
In this paper we have introduced the basic concepts of the nonabelian duality
transformation and applied them constructing a novel family of solvable D-matrix
models (defined via (3.3)) graded by the rank 1 6 k 6 D 1 of the manifest [U (N)]k
conjugation-symmetry (1.7). The key-ingredients of the transformation are the dual
representation (2.4) of the U (N) 1-link integral and the synthetic form (2.14) of a generic
D-matrix SU(N) system (1.1). Combining these ingredients together, the partition function
of any matrix theory (1.1) can be rewritten in terms of the Tr4n+ characters. The latter are
the traces of the different S(n ) tensors (represented via (2.5)) which, being composed
of the dual variables, are embedded into the enveloping S(4n+ ) space. The dual set
consists of the integer-valued {} fields (parametrizing via (2.15) relevant irreps {R }) that
are complemented by the n S(n)-valued degrees of freedom (facilitating the fusion-rule
algebra of the Young idempotents CR involved).
So far, the available solvable D-matrix models of the 1 6 k 6 D 1 type are mainly
associated to the situations [1012,21] where an application of the ItzyksonZuber formula
[9] transforms the model into some k = D eigenvalue-theory of q (hermitean or unitary)
matrices. The proposed nonabelian duality suggests the alternative mechanism of the
solvability realized for the subclass (3.3) of (1.1). Here the underlying reason is the hidden
symmetry of the action which becomes manifest after reformulation in terms of the dual
variables. It is this somewhat unconventional symmetry which predetermines that the

772

A.Yu. Dubin / Nuclear Physics B 584 [PM] (2000) 749783

involved Young idempotents satisfy the simplest pattern (2.29) of the fusion-rules encoded
by the GLR coefficients (1.6). As a result, the associated 1 6 k 6 D 1 systems (1.1),
(3.3) can be mapped onto the k = D D-matrix eigenvalue-models (1.3), (1.4) solvable [8]
in the limit N .
Generalizing [8], we reconstruct the lattice gauge theory with the large N free energy
er ] of
eLD ] equal (modulo the LD -volume factor (1.2)) to the free energy ln[X
ln[X
a given 1 6 k 6 D 1 system (1.1) invariant under (2.8). To clarify the issue of the
continuum limit in thus induced gauge systems, we choose the judiciously constructed
k = 1 family (3.9) constrained by (3.15) where the 1-matrix model of the representation
(5.4) is selected in the simplest form (5.9). Given this choice, we prove that the proposed
sufficient
in [8] scaling-condition (5.5) on (5.4)) does ensure the localization {U (z) 1}
for a nontrivial continuum limit.
Combining the techniques developed in Sections 2 and 3 with the methods of algebraic
topology, we will extend [22] the D = 2 stringy construction of Gross and Taylor [2,3]
for the case of the D > 3 lattice gauge models [5] (which are renormgroup invariant in
D = 2) in the strong coupling (SC) phase. In particular, the D > 3 free energy of the
latter systems can be represented as the following statistics of the gauge strings. The lattice
e )] of a given connected string worldsheet M(T
e ) (with the support on a 2d
weight w[M(T
cell-complex T ) of the total area A and genus h reads

e )|{bk }],
e )] = eA0 ({bk }) N 22h J [M(T
w[M(T

e T,
:M

(6.1)

where the dependence of the string tension 0 ({bk }) on the relevant YM coupling constants
e )|{bk }] (equal to unity for a nonselfintersecting
{bk } is derived. As for the factor J [M(T
e given a particular choice among the gauge models [5], it is sensitive only to
surface M),
e
the topology (but not to the geometry) of selfintersections of M.
e T of a (discretized) 2d
The latter topology is described in terms of the mapping : M
e onto a 2d cell-complex given by a subspace T of the 2d skeleton TD of the Dsurface M
dimensional base-lattice LD . The maps locally comply with the standard definition of the
e T anywhere except for a set of isolated points on T . At this set, movable
immersion M
or nonmovable singularities are allowed of the same type (i.e., the branch points or/and
the collapsed, to a point, subsurfaces) as for the mappings in the D = 2 proposal of Gross
e )|{bk }], it is collected from that
and Taylor [2,3]. As for the {bk }-dependence of J [M(T
of the elementary weights assigned to the admissible (movable) singularities of the map (6.1). Actually, the weight (6.1) is invariant under the continuous group (rather than
the discrete one as one could expect from the lattice formulation) of the area-preserving
homeomorphisms so that Eq. (6.1) can be readily used to define the corresponding smooth
gauge string. The latter features sharply distinguish our D > 3 proposal from the earlier
variant of the large N D > 3 SC lattice string [24,25] (that also combines, but in a
different way, the Wilsons stringy form [1] of the SC series with the power of the large N
expansion). In particular, the pattern (6.1) does not encode singular lattice geometry of T
that is built into the curvature-defects inherent in the earlier formulation [24,25].

A.Yu. Dubin / Nuclear Physics B 584 [PM] (2000) 749783

773

Acknowledgements
This project was started when the author was the NATO/NSERC Fellow at University of
British Columbia, and I would like to thank all the staff and especially Gordon Semenoff
for hospitality.

Appendix A. SU (N )/

n S(n)

complementarity

To facilitate the nonabelian duality transformation on the lattice, one needs a piece of
the formalism which we now focus on. Let us start with the basic facts about the action of
the GL(N) (which can be further restricted to U (N) or SU(N)) and the symmetric groups
on the tensor spaces associated to the structures introduced in the Section 2.
Recall that GL(N) is the group of the automorphisms of nondegenerate (det V 6= 0)
j
complex N N matrices Vi . Given any basis {|iii, i = 1, 2, . . . , N} on a N -dimensional
vector space XN , the fundamental matrix representation of GL(N) is defined canonically
P
j
n = |i i |i i |i i
b|ii = N
as V
1
2
n
j =1 Vi |j i. Given XN , one introduces the basis |ii
n
n
according to
for the direct product space XN . The elements of GL(N) act on XN
n
n
b n |ip i = D(V ){jn } |j in , where conventionally D(V ){jn } =
the standard rule V
p=1

{i

{i

(V )i11 (V )i22 (V )inn V n and the sum over {jk } is implied.


Next, the representation theory proves (see, e.g., [16,17]) that the symmetric group
S(n) is the most general group of transformations commuting with the elements V n of
n
. The elements of the S(n)-group are represented by the linear
the GL(N) on the XN
tranformation
j

{j n }

|iin = |i 1 (1)i |i 1 (2)i |i 1 (n) i = D( ){i n } |j in ,

(A.1)

{j n }

where D( ){i n } is given by Eq. (2.5) (where ji denotes the N -dimensional Kronecker
delta-function), with the basic property being the commutativity
[D( ), D(V )] = 0,

S(n), V G.

(A.2)

The representation of the S(n)-algebra elements is deduced from the group-representation


(2.5) by linearity. Remark that the unity element 1 [n] of the S(n) group is represented by
the trivial permutation (A.1) 1 [n] : (k) = k, k = 1, . . . , n.
The central operation, we will employ, is the decomposion of the direct product space
V n (or, equally, (V + )n ) into the irreps of the Lie group or, dually, into the irreps of
S(n). Recall that in the S(n) group the irreps are labelled [16,17] by a set of k nonnegative,
nonincreasing integers {ni ; n1 > n2 > > nk > 0} constrained by the single condition
Lk
i=1 ni = n and visualized as the Young tableau Yn with n boxes. The U (N) or GL(N)
irreps can be parametrized in a similar fashion by a set of N integers {ni ; n1 > n2 > >
nN } that is related to the alternative classification (5.10) identifying ni = i N + i. When
all ni , i = 1, . . . , N, are nonnegative (nonpositive), the associated U (N)-characters are
expressed in terms of the V - (V + -) tensors only. The corresponding (anti)chiral sector
of U (N)-irreps can be visualized [16,17] by the U (N) Young tableaus Yn(N) containing

774

A.Yu. Dubin / Nuclear Physics B 584 [PM] (2000) 749783

P
n= N
i=1 ni boxes distributed in not more than N rows. In the SU(N) case, the complete
set of irreps is labelled by the SU(N) Young tableaus Yn(N) containing not more than N 1
rows, i.e., nN = 0.
The required decomposition of V n (or, equivalently, (V + )n ) can be canonically
generated by the Yn(N) -subset of the Young projectors {PR } (which, being defined by
Eq. (2.7), belong to the center of the S(n) algebra: [PR , ] = 0, S(n), R Yn .
The GL(N) (or U (N)) tensor product V n is mapped [16,17] by an admissible projector
eR or equivalently onto the dR
PR , R Yn(N) onto the dim R copies of the S(n) irrep T
copies of the Lie group irrep TR
D(PR )V n = (TeR )dim R = (TR )dR = TeR TR ,

R Yn(N) ,

(A.3)

where dR , dim R stand, respectively, for the dimensions of S(n)- and the chiral GL(N)(or, equivalently, U (N)-) irreps R, respectively. When V is further restricted to SU(N),
one obtains in this way all the irreducible representations. For V U (N), the space V n
(N)
contains irreps included into the single chiral sector of U (N)-irreps R Yn .
Combining equation (A.3) with the defining properties of {PR }
X
PR = 1,
(A.4)
PR1 PR2 = R1 ,R2 PR1 ,
RYn

for V U (N) or GL(N) we deduce one of the central results of the representation theory,
the so-called SchurWeyl duality (see, e.g., [6])
X
eR TR ,
(A.5)
T
V n =
RYn(N)

which formalizes the complementary roles of the Lie and symmetric groups.
Employing Eqs. (A.3), (A.5), the formulas relevant for the duality transformation
can be represented in the concise algebraic form. First, taking trace of the SchurWeyl
decomposition (A.5) and using the completeness condition (A.4) for {PR }, one deduces
(see, e.g., [6]) the second Frobenius formula
X
R ( )R (V ),
(A.6)
[ ] (V ) = Trn [D( )V n ] =
(N)

RYn

where the sum runs over the chiral U (N) irreps R. Taking into account Eq. (A.3), we
obtain the useful modification of (A.6): Trn [D(PR )D( )V n ] = R ( )R (V ). Similarily,
the first Frobenius formula (2.10) for R (V ) can be obtained multiplying Eq. (A.5) by
D( ) and taking the trace as previously.

Appendix B. The dual form of the 1-link integral


In this appendix, we derive the dual representation (2.4) of the 1-link integral (2.3). To
p ...q
compute M G (n, m)j11...lmm of Eq. (2.3) for a particular Lie group G, the starting idea [18
20] is to differentiate the simple generating function

A.Yu. Dubin / Nuclear Physics B 584 [PM] (2000) 749783

Z
G
Fn,m
(A, B) =

775

dV (Tr[AV ])n (Tr[BV + ])m

(B.1)

with respect to A, B G.
B.1. The S(n)-form of the U (N) formula
In the U (N)-case the invariance (2.8) of the integral (B.1) under the U (1)-subgroup
of U (N) = (SU(N) U (1))/ZN ensures that it doesnt vanish only for n = m. For
Fn,n (A, B), A, B U (N), one derives [1820]
U (N)
(A, B) =
Fn,n

dR2
R (AB),
dim R
(N)

(B.2)

RYn

where the second Frobenius formula (A.6) and the standard orthogonality condition of the characters have been employed. Applying to Fn,n (A, B) the operator
Q
j
l
(n!)2 nk=1 2 /Apkk Bqkk (where A, B in the r.h.s. of Eq. (B.2) can be extended [20]
to be GL(N)-valued), one obtains for the U (N) 1-link integral (2.3)
1 X dR2
(n!)2
dim R
(N)
RYn

X
, S(n)

...q

(1) (2)
(n)
(1)
(2)
(n)
D(CR )j(1)
j(2) ...j(n) l (1) l (2) l (n) .

(B.3)

Introducing = 1 S(n), we arrive at the concise representation given by the


(1)
being defined by Eqs. (2.5) and (2.6),
Eq. (2.4), with the S(n)-tensors D() and n
(1)
the sum over the U (N) G-irreps R Yn(N) is restricted
respectively. Note that, in n
(N)
to the single chiral sector parametrized by the U (N) Young tableaus Yn (containing not
more than N rows). Let us remark also that Eq. (2.4) is complementary to already existing
representations [2,3,1820] of (2.3) which are not suitable for our present purposes.
B.2. The S(2n)-form of the U(N) formula
The integration formula (2.4) can be represented in terms of the elements of the S(2n)algebra enveloping S(n) S(n). Employing the ordered link-basis |I2n i = |In(+) i|In() i
()
(with |In i = |i in akin to (2.21)), one rewrites Eq. (2.4) in the S(2n)-form of
(+)
()
Eqs. (2.22), (2.23). By construction, the chiral, |In i, and the antichiral, |In i, sectors
are associated, respectively, to the first and to the second S(n)-sub-blocks of (2.22), (2.23).
As for the operator 2n S(2n), comparing (2.4) and (2.22) one deduces for its explicit
form
{j 2n }

1
2
in+2
i2nn i1n+1 i2n+2 in2n
D(2n ){i 2n } = in+1

(B.4)

which in turn can be concisely represented in the alternative ordered S(2n)-basis |I2n i =
(|i+ i|i i)n as the outer product
D(2n ) = D((c2 )n ) S(2n)

(B.5)

of the 2-cycle permutations c2 C2 , c2 : {12} {21}, where each individual c2 S(2) acts
on the (|i+ i|i i)-subspace of |I2n i.

776

A.Yu. Dubin / Nuclear Physics B 584 [PM] (2000) 749783

Finally, we mention the commutation-rules





[2n (2n)] = (2n), (m)


n 1[n] = 0,

(B.6)

where in the second case the above |I2n i-basis is employed. In particular, it makes the
relative order of the operators in the product (2.22) immaterial. Also, let us include the
following useful identities


(2n) En(+) En() = En() En(+) (2n),


(B.7)
2n En(+) En() = En() En(+) 2n ,
()

valid for En

S(n) where again the |I2n i-basis is implied.


SU (N )

B.3. V SU(N ) V U (N ) extension for Fn,n

(A, B)
SU(N)

Let us prove that the diagonal SU(N) moments Fn,n (A, B), defined by Eq. (B.1),
are invariant under the substitution of the SU(N) link-variables by the U (N) = [SU(N)
U (1)]/ZN ones

(B.8)
V SU(N) V SU(N) V U (1) /ZN = V U (N) , dV SU(N) dV U (N) ,
that is tantamount to Eq. (2.9). Actually, the identity (2.9) for n < N directly follows from
equivalence [20] of any polynomial U (N) representation with n < N to the corresponding
SU(N) one. But for n > N one needs a more refined consideration (valid for any n) we
now focus on.
First, we note that the extra U (1)-factor in Eq. (B.8) matches with the auxiliary
SU(N)
(A, B) (while
U (1)-invariance (2.8) which is implicit in the diagonal moments Fn,n
nondiagonal n 6= m SU(N) integrals (2.3) are only ZN -invariant). To promote (within
SU(N)
Fn,n (A, B)) the ZN center-subgroup (2.8) into U (1), one is to multiply each V SU(N)
by the auxiliary factor V U (1) and then integrate dV U (1) with the U (1) Haar measure
normalized to unity. The resulting pattern is to be confronted with the factorized
representation of the U (N) measure
Z
Z
Z
dV U (N)
dV U (1)
dV SU(N)
(B.9)
U (N)

U (1)

SU(N)

that will be derived below. Identifying the U (1) sector in Eq. (B.9) with the averaging over
SU(N)
the auxiliary U (1)-transformation (that leaves diagonal integrals Fn,n (A, B) invariant),
we justify the required formula (2.9).
To obtain Eq. (B.9), recall first that the explicit expressions [20] for the U (N) and
SU(N) measures (after decomposition V = diag[ei ] + ) are given, respectively, by
Eq. (G.3) and
Z
dV

SU(N)

where + =

PN

k=1 k ,

Z+Y
N

k=1

dk (2)

(+ ) |1({p })|2 d ,
2

and (2) () = 2

nZ (

(B.10)

2n) is the periodic -function.

A.Yu. Dubin / Nuclear Physics B 584 [PM] (2000) 749783

777

Next, the factorized form (B.9) of the U (N) measure (G.3) is predetermined by the
decomposition V U (N) = (V U (N) /det[V U (N) ]1/N ) det[V U (N) ]1/N that allows to identify

 SU(N) 
, ,
(B.11)
V U (N) /det[V U (N) ]1/N
= V SU(N) k

U(N)

1/N
= ei+ /N ,
(B.12)
V U (1)
= det V U (N)
U (N)

U (1) V U (1) ( /N),


+
+ R/N [, +] mod 2. As a result, in Eq. (B.9) V
RwhereU
+
(1)
dV
= d(+ /N)/2 , while
U (N)

kSU(N) = kU (N)

+
N

SU(N)
+
=

N
X

kSU(N) = 0 mod 2.

(B.13)

k=1

In order to convert the decomposition (B.11), (B.12) into (B.9), one is to change
variables in Eq. (G.3) going over from {kU (N) } to the overcomplete set {kSU(N) ; k =
U (N)
U (N)
SU(N)
}) = 1({k
}). Second,
1, . . . , N} (+ /N) entering Eq. (B.13). First, 1({k
SU(N)
SU(N)
= 0 mod 2 for the {k
}-set can be imposed by the (2) the constraint +
function (reminiscent of the SU(N) measure (B.10)), and no additional Jacobian is needed.
Summarizing, we arrive at the identity

 U (N) 
 SU(N) 
U (N)
, = dV U (1) + /N dV SU(N) k
, ,
(B.14)
dV U (N) k

U (N)
U (N) 

+ +
mod 2, k = 1, . . . , N,
(B.15)
kSU(N) + +
N
N
R
R
where dV U (N) and dV SU(N) are defined by Eqs. (G.3) and (B.10), respectively.
SU(N)
Evaluating the diagonal SU(N) 1-link integral Fn,n (A, B), one observes that the overall
U (N)
SU(N)
}-variables does not affect the result. Altogether,
(+ /N)-shift (B.15) of the {k
it transforms the equality (B.14) into the announced decomposition (B.9) of the U (N)
measure.
Finally, the invariance (2.8) of the diagonal SU(N) 1-link integral (B.1) under the
extension of the measure

 SU(N) 
 SU(N) 
U (N)
, V U (1) + /N V SU(N) k
,
(B.16)
V SU(N) k
SU(N)

U (N)

allows
R + to identify Fn,n (A, B) with the Fn,n (A, B). For this purpose, one is to integrate
d(+ /N)/2 that trades (B.10) for the factorized pattern (B.9) of the U (N) measure.
This completes the proof of the identity (2.9).

Appendix C. Fusion rules of Young idempotents


In this appendix we derive the identities relating Young idempotent CR+ S(n+ ), n+ =
Pp
p
k=1 nk , with the corresponding direct product k=1 CRk . Let us start with the following
observation. The outer product of two idempotents CR1 CR2 S(n1 ) S(n2 ), being
embedded into S(n1 + n2 ), ceases to be a composition of the S(n1 + n2 ) Young
idempotents defined as

778

A.Yu. Dubin / Nuclear Physics B 584 [PM] (2000) 749783

([ ]) =

S(n+ )

Rn+ CRn+ ,

(C.1)

(N)
R+ Yn+

where ([t t 1 ]) = ([ ]), for t S(n1 + n2 ). Indeed, (for Rk Ynk ) the product
(CR2 CR1 )V (n1 +n2 ) , being a Lie group representation
(N)

 p


p
Trn+ k=1 CRk V n+ = k=1 Rk (V ),

Rk Yn(N)
,
k

n+ =

p
X

nk ,

(C.2)

k=1

does not generate a S(n1 + n2 )-representation. To preserve the SchurWeyl duality (A.3),
we introduce the twisted deformation
CR1 CR2

1
(n1 + n2 )!



(CR1 CR2 ) 1 .

(C.3)

S(n1 +n2 )

By construction, the twisted product (C.3) commutes with t S(n1 + n2 ) and thus
belongs to the center (C.1) of the S(n1 + n2 )-algebra. More generally, given V U (N)
we arrive at the pattern (2.29) of the fusion rules of the S(n )-valued Young idempotents.
By the same token, one arrives at the inverse of the identity (2.29)
M

CR+ =

{Rk Ynk }

(p)
L{Rk }|R+

X [ (p CRk ) 1 ]
k=1
,
(n+ )!

R+ Yn+ ,

(C.4)

S(n+ )

expressed, strictly speaking, in terms of some other set of the GLR coefficients
(p)
(p)
LR1 ...Rp |R+ L{Rk }|R+ of pth order.
Next, in the framework of the duality transformation the fusion rules (2.29) and (C.4) are
realized in the tensor representation (2.5) and enter only inside the associated traces Trm .
Consequently, according to the second Frobenius formula (A.6), it effectively eliminates
the contribution of those S(n )-irreps R Yn which do not correspond to a U (N) irrep
(i.e., do not belong to Yn(N)
). Therefore, we confine our attention to the GLR coefficients
(p)

(p)

LR+ |{Rk } , L{Rk }|R+ with all R Yn(N)


. Upon a reflection, this subset coincides with the
coefficients in the associated via (C.2) decomposition of the U (N) characters (with the
irreps restricted to the chiral sector), i.e., assume the integral form
Z
(p)
(p)
p
dV R+ (V + ) k=1 Rk (V ),
(C.5)
LR+ |{Rk } = L{Rk }|R+ =
U (N)

Pp
where n+ = k=1 nk . To prove this statement, one first applies both sides of (2.29)
(or (C.4)) to V n+ U (N) and then takes the overall trace Trn+ . Employing the
commutativity (A.2), we get rid of the sum over S(n+ ) and use the identity (C.2) which
(p)
altogether results in (C.5). Finally, let us remark that a generic coefficient LR+ |{Rk } , Rk
(N)

Ynk , can be represented as a combination of the elementary, p = 2 Littlewood


Richardson (LR) coefficients L(2)
R+ |R1 R2 [16,17] (entering the two-character fusion rules).

A.Yu. Dubin / Nuclear Physics B 584 [PM] (2000) 749783

779

Appendix D. The YM2 theory on a 2d manifold


Let us aplly the dual representation (2.4) of the 1-link integral to the computation of the
eG of the continuous gauge theory on a 2d manifold of genus G
partition function (PF) X
and the total area A. As it is known [5,23], the latter can be directly reproduced by the
following lattice model with the action generalizing the HeatKernal one
!
Z Y
G
G
X
Y
g 2 C2 (R)
+
+

A
eG =
2
dim R e
dVk dWk R
Vi Wi Vi Wi ,
(D.1)
X
R

k=1

i=1

where R (U ) and C2 (R) are the character and the second Casimir eigenvalue both
associated to a given irreducible representation (irrep) R of the Lie group in question.
Most effectively, the 1-link integration formula (2.4) can be applied in the SU(N) case,
being combined with Eqs. (2.9) and (2.10). First, one is to employ Eq. (2.10) to rewrite
(N)
the character in Eq. (D.1) as the product of Trn [D(CR )(Vi Wi Vi+ Wi+ )n ], R Yn . As
the expression (D.1) is invariant under (2.8), the SU(N) integrations over Vi , Wi can be
traded for the U (N) ones according to the identity (2.9). Applying Eq. (2.4) 2G times and
contracting the relevant tensor indices in compliance with the pattern of (D.1), we obtain
eG =
X

+ X
X

(dim R)22G e

g 2 C2 (R)
A
2

B(R),

(D.2)

n=0 RY (N)

"

B(R) = Trn

G
2
Y
CR Y X
[k , k ]
D
dim R
k=1 {k ,k }

q=1 {R (q) Y (N) }


k

dR (q) PR (q)
k

n!

!#
,

(D.3)

where the 2G species of -twists are denoted by k , k , k = 1, . . . , G, and the


corresponding {k , k }-sums in (D.3) run over the S(n) group elements (with [i , i ]
i i i1 i1 ). Taking into account that Trn [CR ] = dim RR ( )/dR , B(R) is reduced
(due to the orthonormality (A.4) of PR = dR CR ) to
!
 2G Y
G
G
X
Y
dR

R
[k , k ] /dR = 1,
(D.4)
B(R) =
n!
i=1 i ,i S(n)

k=1

that is equal to unity owing to the identity derived in [2,3] (see also [6]). Summarizing, we
eG previously deduced in
recover the dual representation of the SU(N) partition function X
[5,23] by a different method.

Appendix E. The partition function (3.13)


er associated to (3.11). Similarly to Eq. (2.33), the
Let us evaluate the partition function X
master-integral (2.18) associated to (3.9) is expressed in terms of the following S(4n+ )algebra element A 4n+
(
)
X
X


W4n+
(2)
E e (1)
1
e
e
[+ ]
[+ ]
M2n+ M
,
(E.1)
2n+ 4n+
([2n+ ]!)2
{R }

{ }

780

A.Yu. Dubin / Nuclear Physics B 584 [PM] (2000) 749783

"
W4n+ =

PR+

D
=1

CR

L(D1)
R |{R }

dim R



(4n+ ) D (2n )
=1
, (E.2)
1 [n ]
[2n+ ]!
[2n ]!

where the second sum runs over S(2n+ ) and 4n+ = D


=1 2n . To simplify
D
Eq. (E.1) further, we first note that the factor =1 (2n )/[2n ]! in the (+ )twisted operator W4n+ can be substituted by the 1 [4n+ ] -unity employing the proper change
of the variables. Indeed, let (4n+ ) and (2n ) are defined by Eq. (2.23) in terms of
S(2n+ ) and S(n ) elements, respectively. Making (for a given ) the two shifts
D
1
1
[D
=1 ], [=1 ], one eliminates the dependence of (E.1) on { }.
Second, the resulting form of Eq. (E.1) can be shown to be invariant under the
substitution of [PR+ D
=1 CR ] by
X

[PR+ D
=1 CR ]
[2n+ ]!

S(2n+ )

1 = LR+ |{R } CR+


(D)

(E.3)

made in the chiral S(2n+ )-block of the combination in the first rectangular brakets of
(E.2). To this aim, one is to perform (for a given ) the following composition of the shifts
and the conjugation: + + 1 , 1 , 1 . Altogether, the tensordependent part of Eq. (E.1) can be rewritten as
X

(4n+ )
e(2) 4n+
e(1) M
(+ )
(+ )1 ,
(E.4)
CR+ M
2n+
2n+
([2n+ ]!)3
{ }

where S(2n+ ). Employing (2.36) together with (2.34) and introducing + =


e (2) ]+ instead of + , one transforms Eq. (E.4) into
[1 M
2n+

1 1 e (2) e (1)
1
dim R+ X R+ + + { M
2n+ M2n+ }
,
(E.5)
dR+
([2n+ ]!)3
, + ,

where the sums over , + , run over the S(2n+ )-group elements. Finally, combining
P
1 1 )/(n!)2 = 1/d (see [2,3]) together with
R
{, S(n)} R (
X R (Bn )
X Bn 1
=
PR ,
n!
dR

S(n)

Bn S(n),

(E.6)

RYn

(see Appendix D), and with R1 (PR2 ) = R1 ,R2 R1 ( ), we finally arrive at (3.13), (3.14).
Appendix F. The D-dimensional eigenvalue-theory
To associate to (4.9) an intermediate D-dimensional eigenvalue-theory devoid of the
unwanted local symmetry (4.11), let us first compare the pattern (4.9) with the one (where
U(+) U+ , U() U )
"D(D1)/2
#
X
Y Y
X
(q)


A({R })
(q)
(q)
e
Re
R (q) U R (q) U
,
(F.1)
{n } {R (q) Y (N) }

=1

q=

A.Yu. Dubin / Nuclear Physics B 584 [PM] (2000) 749783

781

that generalizes (1.3) still remaining compatible with the original algorithm [8]. The Ddimensional eigenvalue-system, corresponding via (1.2) to (F.1), can be deduced from
(q)
(4.1) substituting each -block of the characters by the more general block (where R =
(q)
R (x))




(F.2)
R (1) U (x) R (1) U (x + ) R (2) U+ (x + ) R (2) U+ (x) ,

(q)
S({R (x)}).

Evidently, (F.2) does not support


with the overall weight being given by S
(4.11).
One observes that in D > 3 the pattern (4.9) of the effective eigenvalue-theory, according
to Frobenius formula (2.10), (2.11), generically can not be reproduced in terms of (F.1)
(resulting from (F.2) after the large N SP reduction of the spacetime dependence of
(q)
U (z), R (x)). To circumvent this problem, we propose the following synthetic algorithm
defining the mapping {U } {U (z)} for the link-variables of (4.9). First, one is to use
the Frobenius formula (2.10), (2.11) expanding the characters in (4.9) in terms of the trace
e U )
products. In the resulting sum, consider a particular term (substituting U

D Y
Y
=1

tr U )k

p () Y
k

tr U+

k p+ ()
k

kpk () = n ,

(F.3)

containing total amount n of the U (or, equally, U+ ) factors. Second, let us separate
(in a way specified below) the product (F.3) into two blocks splitting the partititions

[1p1 () 2p2 () ] according to pk () = lk () + fk () so that


X
X
(1)
(2)
klk () = m(1)
kfk () = m(2)
(F.4)
,
, m + m = n .
k

Given (F.4), perform (with the help of the second Frobenius formula (A.6)) multiple
Fourier expansion of the first, {lk ()}-block
X
Y
Y
k l q ()
(q)

k
tr U(q)
=
eB1 ({R })
R (q) U(q) .
(F.5)
{,k,q}

{,q}

(q)

{R }

Let the link-variables U in (F.5) be mapped onto U (z) of the intermediate D-dimensional
(1)
eigenvalue-system in compliance with the pattern (4.10). We postulate that the set {m },
PD
(1)
minimizing the function =1 [m ]2 , is constrained by the following condition. There
should exist a set {m Z>0 } of D(D 1)/2 integers so that the second {fk ()} block
can be represented in the form
X
Y
Y
f q ()


tr U(q) )k k
=
eB2
R (q) U(q) R (q) U(q) ,
(F.6)
{,k,q}

(q)

{R }
(q)

{,q}

(where B2 B2 ({R })) that matches with the pattern (F.1). In Eq. (F.6), it is supposed
P
(q)
(N)
(1)
that R Ym with m = D1
6= m . As a result, the latter condition evidently allows
to map {U } in (F.6) onto {U (z)} according to the pattern (F.2).
Upon a reflection, the above prescription remains essentially ambiguous. To fix the
freedom of choosing the {m Z>0 } set, we impose extra constraint that the latter

782

A.Yu. Dubin / Nuclear Physics B 584 [PM] (2000) 749783

P
integers minimize {} [m n ]2 (where n enters (4.6)). In case if there remains
(accidental) residual ambiguity, one is to symmetrize over all the admissible options.
Summarizing, we have formulated the algorithm to map a given [U (1)]D -invariant Dmatrix model (1.1) (constrained by (4.6)) onto the intermediate D-dimensional eigenvaluesystem which is invariant under (4.3), (4.2) but not under (4.11).

Appendix G. The WC asymptotics of Xr ({gr })


Let us demonstrate that in the k = 1 D-matrix model (1.1)(3.9), (3.15) the large N WC
asymptotics (5.25) is indeed valid. To begin with, one readily obtains (from (3.11), (3.15))
for the action of the latter model

e(1) ({1})

eSr

eE dim2 R+

2
Y
p=1

(q)
{R }

L(D)

(p)
R+ |{R }

(D1)
D
=1 L (2)

R |{R }


,

(G.1)

e is defined by Eq. (5.8). Thus, the asymptotics of the ratio (5.24) is predetermined
where E
by the g 2 N -scaling of the factor

 (D)
D
(2) 1
(G.2)
L
(1) dim R+ =1 dim R
R+ |{R }

(2)

responsible for the mismatch between (3.13) and (3.11) (where we have used that R in
(3.12) can be identified with R of Eq. (3.14)). Next, recall that according to Eq. (5.12), the
irrep R in Eq. (5.4) represents the irreps {R } entering the GLR fusion-rules (1.5). Owing
to the pattern (3.11), (3.13), (3.15) of the involved GLR fusion rules, the scaling-condition
(5.5) is valid for the characteristic values of all species {(0) ()}, {}, {}, +,
parametrizing the SP irreps {R(0) } (on which the relevant large N sums are localized).
Next, the dimension-formula R({}) (U ) = detk,j (eik j )/detk,j (ei(Nk)j ) together
with (5.5) predetermines that each dim R contributes in the limit N with the scaling1
2
eU (N) (enterfactor [gN
2 ]N /2 . Complementary, given (5.5), the eigenvalues j () of U

ing the definition (1.6) of the relevant GLR coefficients (G.2) modified by the extension
(2.9)) satisfy (5.26) as in [8]. Combining it with the pattern of the U (N) Haar measure
R
( dU = 1)
Z

Z
dU =
U (N)




N Z+
Y
i j 2
dk Y
2
sin
d
,

2
2
k=1

(G.3)

i<j

one concludes that the GLR coefficient of Kth order scales in the large N WC limit as
2
[dim R]K1 O([g 2 N]N (K1)/4 ). Putting together all the scaling-factors inherent in
(G.2), one reproduces (5.25).
er of Eq. (3.13) the S(2n+ )-factor K(R+ ) =
As a side remark, had we retained in X
(1) e (2)
3
e
R+ (M2n+ M2n+ )/dR+ while keeping E({R()}) Z2 -invariant, there would be no way to
adjust the parameters of the latter weight to set up the {}-localization (5.5).

A.Yu. Dubin / Nuclear Physics B 584 [PM] (2000) 749783

783

References
[1]
[2]
[3]
[4]
[5]
[6]
[7]
[8]
[9]
[10]
[11]
[12]
[13]
[14]
[15]
[16]
[17]
[18]
[19]
[20]
[21]
[22]
[23]
[24]
[25]

K. Wilson, Phys. Rev. D 10 (1974) 2445.


D. Gross, Nucl. Phys. B 400 (1993) 161.
D. Gross, W. Taylor, Nucl. Phys. B 400 (1993) 181; Nucl. Phys. B 403 (1993) 395.
A. Migdal, Sov. Phys. Jetp. 42 (1975) 413.
E. Witten, Commun. Math. Phys. 141 (1991) 153.
S. Cordas, G. Moore, S. Ramgoolam, Lectures on 2D YangMills theory, equivariant
cohomology and topological field theories, hep-th/9411210.
A. Polyakov, Nucl. Phys. B 486 (1997) 23.
A. Dubin, A Candidate for solvable large N lattice gauge theory in D > 2, hep-th/9910088.
C. Itzykson, J. Zuber, J. Math. Phys. 21 (1980) 411.
V. Kazakov, A. Migdal, Nucl. Phys. B 397 (1993) 214.
A. Migdal, Mod. Phys. Lett. A 8 (1993) 359.
D. Gross, Phys. Lett. B 293 (1992) 181.
A. Morozov, Matrix models as integrable systems, Lectures at Banff Conference, 1994, hepth/9502091.
R. Savit, Rev. Mod. Phys. 52 (1980) 453.
K. Druhl, H. Wagner, Annals of Phys. 141 (1982) 225.
Wu-Ki Tung, Group Symmetry in Physics WC, 1985.
W. Fulton, J. Harris, Representation Theory, Springer-Verlag, 1991.
S. Samuel, J. Math. Phys. 21 (1980) 2695.
I. Bars, J. Math. Phys. 21 (1980) 2678.
J. Drouffe, J. Zuber, Phys. Rep. 102 (1983) 1.
M. Kontsevich, Funk. Anal. i Prilozh. 25 (1991) 50; Comm. Math. Phys. 147 (1992) 1.
A. Dubin, Nucl. Phys. B 582 (2000) 677.
B. Rusakov, Mod. Phys. Lett A 5 (1990) 693.
V. Kazakov, Sov. Phys. JETP 58 (1983) 1096.
K. OBrien, J. Zuber, Nucl. Phys. B 253 (1985) 621.

Nuclear Physics B 584 [PM] (2000) 784794


www.elsevier.nl/locate/npe

Noncommutative gauge theory


for Poisson manifolds
Branislav Jurco a , Peter Schupp b, , Julius Wess b
a Max-Planck-Institut fr Mathematik, Vivatgasse 7, D-53111 Bonn, Germany
b Sektion Physik, Universitt Mnchen, Theresienstr. 37, D-80333 Mnchen, Germany

Received 16 May 2000; accepted 30 May 2000

Abstract
A noncommutative gauge theory is associated to every Abelian gauge theory on a Poisson
manifold. The semi-classical and full quantum version of the map from the ordinary gauge theory
to the noncommutative gauge theory (SeibergWitten map) is given explicitly to all orders for any
Poisson manifold in the Abelian case. In the quantum case the construction is based on Kontsevichs
formality theorem. 2000 Elsevier Science B.V. All rights reserved.
PACS: 11.25; 11.15.-q; 04.60.Ds

1. Introduction
Noncommutative geometry naturally enters the description of open strings in a
background B-field [14]. The D-brane world volume is then a noncommutative space
whose fluctuations are governed by a noncommutative version of YangMills theory
[59]. In the case of a constant B-field it has been argued that there is an equivalent
description in terms of ordinary gauge theory. From the physics perspective the two
pictures are related by a choice of regularization [9,10]. There must therefore exist a
field redefinition a SeibergWitten map [9]. The B-field, if non-degenerate and closed,
defines a symplectic structure on the D-brane world volume; its inverse is a Poisson
structure whose quantization gives rise to the noncommutativity.
An interesting question arises: given a gauge theory on a general Poisson manifold
is there always a corresponding noncommutative gauge theory on the noncommutative
space the quantization of the original Poisson manifold? Previously we found this to
Corresponding author.

E-mail addresses: jurco@mpim-bonn.mpg.de (B. Jurco),


(P. Schupp), wess@theorie.physik.uni-muenchen.de (J. Wess).

schupp@theorie.physik.uni-muenchen.de

0550-3213/00/$ see front matter 2000 Elsevier Science B.V. All rights reserved.
PII: S 0 5 5 0 - 3 2 1 3 ( 0 0 ) 0 0 3 6 3 - 1

B. Jurco et al. / Nuclear Physics B 584 [PM] (2000) 784794

785

be true for symplectic manifolds [11]. 1 Here we give the construction for an arbitrary
Poisson manifold. (The present discussion is much more explicit and complete.) On the
way an appropriate generalization of Mosers lemma from symplectic geometry to the
Poisson case and its quantization are given. This, we believe, is mathematically interesting
in its own right. As in our previous paper [11], we choose to work within the framework
of deformation quantization [1315]. This allows us to postpone questions related to
representation theory, so that we can focus on the algebra. We expect that our results can be
used to find derivative corrections to the BornInfeld action, classifying invariant actions
in the spirit of [16].

2. Noncommutative YangMills theory


Here we recall how the gauge theory on a more-less arbitrary noncommutative space
was introduced in [17]. The formulation starts with an associative, not necessarily
commutative, algebra Ax over C freely generated by finitely many generators x i modulo
some relations R. Ax plays role of the noncommutative spacetime. The matter fields
of the theory are taken to be elements of a left module of Ax and the infinitesimal gauge
transformation induced by Ax is given by the left multiplication (action)

()

7 + i .

(1)

The gauge transformation does not act on the

coordinates x i .

()

x i 7 x i .

(2)

The left multiplication of a field by the coordinates x i is not covariant under the gauge
transformation

()

x i 7 x i + i x i ,
x i

(3)
x i .

is not equal to
The gauge fields
since in general
introduced to cure this. Namely, covariant coordinates
b i = x i + A i
X

A i ,

elements of Ax , are
(4)

are introduced.
The gauge transformation is supposed to act on the gauge fields A i in a way that will
bi under the gauge transformation (1), (2). This is achieved by
assure the covariance of X
the prescription





()
x i + i ,
A i .
(5)
A i 7 A i + i ,
In examples considered in [17] (universal enveloping algebra of a finite-dimensional
Lie algebra and of the Heisenberg algebra as a special case, quantum plane) also the
bij was introduced. If
corresponding field strength F
 i j

(6)
x , x = J ij (x),
1 For a detailed description of the universal gauge theory of the WeylBundle see, e.g., [12].

786

then

B. Jurco et al. / Nuclear Physics B 584 [PM] (2000) 784794

 i j
b ,X
b J ij (X).
b
bij = X
F

(7)

Of course this is not unique, the choice of ordering in the above formula may effect the
b, but this is not important for covariance. It follows
definition of F



()
b ij ,
bij + i , F
bij
7 F
F

(8)

as expected.
This construction covers also the noncommutative generalization of non-Abelian
(GL(N) or U (N) if Ax possesses a -structure) gauge theories, taking the tensor product
algebra Ax U (gl(N)) instead of Ax .
The following question inspired by [9] appears naturally.
Let us assume that our associative algebra Ax (and this was indeed the case of the
examples considered in [17]) can be understood as a deformation quantization of a
commutative algebra of functions on some Poisson manifold M. Let ? be the corresponding
star product. Let us also assume that we have a (non-Abelian) gauge field A on M.
Does there exist a map SW:
(SW)

A 7 A,

(SW)
7 (, A)

(9)

such that the (non-Abelian) commutative gauge transformation on A


()

A 7 A + d + i[, A]

(10)

is sent by SW into the noncommutative gauge transformation on A





()
x i + i ,
A i ?
A i 7 A i + i ,
?
?

(11)

The commutator in (10) is the matrix one and the commutator in (11) is the star
commutator on functions and the matrix one on matrices.
In this paper we give a general and explicit construction of the map SW in the Abelian
case. Our main tool is Kontsevichs formality theorem.
The non-Abelian case is more involved and will be treated by similar methods in the
sequel to this paper.

3. Classical
Here we formulate the classical analogue of the SeibergWitten map between the
commutative and noncommutative description of YangMills theory for any Poisson
manifold.
Let M be a manifold and F a two-form on M. Let us temporarily assume that F is
exact. Later on we will relax this condition, F closed will appear to be good enough for
our purposes. In local coordinates we write A = Ai dx i and F = 12 Fij dx i dx j , with
Fij = i Aj j Ai . Let us further assume that we have on M a one-parameter family of

B. Jurco et al. / Nuclear Physics B 584 [PM] (2000) 784794

787

bivector fields (t) = 12 ij (t)i j , t [0, 1], with the explicit t-dependence of the matrix
ij (t) given by
t (t) = (t)F (t),

(12)

with the initial condition


(0) = ,

(13)

where is some fixed but otherwise arbitrary Poisson tensor on M. The product on the
right in (12) is the matrix one. As above we will often use the same notation for polyvector
fields or forms and the corresponding tensors. The formal solution to (12) can be given by
the following power series in t (or in )
X
(t)n (F )n .
(14)
(t) =
n>0

The convergence is not an issue here, because we will work all the time with formal power
series in . (E.g., if the matrix is invertible then in the physical situation of a D-brane in
the background of the B-field B = 12 ij1 dx i dx j it simply means that the background is
strong.)
It follows from (12), or directly from (14), that (t) continues to be a Poisson tensor for
all t [0, 1]. For this only the closeness of F is important. The Poisson bivector field (t)
defines an bundle map T M T M given by i(t )() = (t)(, ) for any one-forms
and . Using the Jacobi identity [ (t), (t)] = 0, with [ , ] being the SchoutenNijenhuis
bracket of polyvector fields, we can easily verify that the t-derivative of (t) is given by a
Lie-derivative that
t (t) + [(t), (t)] = 0,

(15)

where now (t) is understood as a bivector field and


(t) = (t)(A)

(16)

is a vector field that in local coordinates looks like


(t) = ij (t)Aj i .
Let us recall that the SchoutenNijenhuis bracket of two polyvector fields is defined by
[1 k , 1 l ]
=

l
k X
X

(1)i+j [i , j ] 1 i j l ,

i=1 j =1

[1 k , f ] =

k
X
(1)i1 i (f )1 i k ,
i=1

if all s and s are vector fields and f is a function.

788

B. Jurco et al. / Nuclear Physics B 584 [PM] (2000) 784794

If f and g are two smooth functions on M with no explicit dependence on t and { , }t


denotes the Poisson bracket corresponding to (t) then (15) is rewritten as
t {f, g}t + (t){f, g}t {(t)f, g}t {f, (t)g}t = 0.

(17)

Both (15) and (17) imply that all the Poisson structures (t) are related by the flow tt 0 of
we have in particular
(t): tt 0 (t 0 ) = (t). Setting = 01
0 = ,

(18)

{f, g}0 = { f, g},

(19)

i.e.,

where 0 is short for (1). The vector field (t) may not be complete, however again
has to be understood as a formal diffeomorphism given by formal power series in . In this
sense we always have a (formal) coordinate change on M which relates the two Poisson
structures and 0 . Explicitly

(20)
= et +(t ) et t =0 .
Consider now a gauge transformation
A 7 A + d.

(21)

The effect upon (t) will be


(t) 7 (t) + (t),

(22)

where (t) is the Hamiltonian vector field


(t) = (t)(d) = [, (t)]

(23)

and [ (t), (t)] = 0. In local coordinates = ij (t)(j )i . Correspondingly we use

1 =
the notation ,t
t 0 for the new flow. It follows almost immediately that ( )
for some .
This
et + (t )et (t ) |t =0 is generated by a Hamiltonian vector field (d )
follows from the BakerCampbellHausdorff identity and the fact that
[t + (t)(A), (t)(df )] = (t)(dg)
with g = (t)(d, A). Having this in mind it is easy to see that all the terms coming
from the BCH formula contain only commutators of this type or commutators of
two Hamiltonian vector fields which are again Hamiltonian ones. Even more is true:
( )1 for all t and t 0 is generated by some Hamiltonian vector field for (t).
,t
t 0 t 0t
So the transformation induced by takes the form


()
f .
(24)
f 7 f + ,
It is clear from the above discussion that working only with formal power series in
we can abandon the exactness condition for F and assume F only closed with the
following consequences. The gauge field A and the vector field (t) are given only locally.
If As given in two different local patches are related on their intersection by the gauge

B. Jurco et al. / Nuclear Physics B 584 [PM] (2000) 784794

789

transformation (21), then the corresponding local vector fields are related by (22) and the

local diffeomorphisms are related by the canonical transformation (24) generated by .


In the case of invertible (t) and of a compact manifold we have the well know lemma
of Moser [18].
Let us return to the SW map in classical setting. For this we have to choose some local
coordinates x i on M. Let us write the result of acting by the diffeomorphism on the
coordinate function x i in the form [11,17,1921].

(25)
x i = x i + Ai .
A depends as a formal power series in on A. Explicitly we have

ij
Ai = et + (t )Aj i 1 t =0 x i .

(26)

Let us act by the infinitesimal gauge transformation (21) on A. This induces the
infinitesimal Poisson map (24) on (x i ), which in turn induces a map on A given by



()
(27)
Ai 7 Ai + , x i + , Ai .
So the map A 7 A can be viewed as the semi-classical version of the SW map which we
are looking for.

4. Formality
The existence of a star product on an arbitrary Poisson manifold follows from the more
general formality theorem [14]: there exists an L -morphism from the differential graded
algebra of polyvector fields into the differential graded algebra of polydifferential operators
on M. There is a canonical way to extract a star product ? from such an L -morphism for
any formal Poisson bivector field. We will refer to this star product as the Kontsevich star
product. Any star product on M is equivalent to some Kontsevich star product.
The differential graded algebra Tpoly(M) is the graded algebra of polyvector fields on M

n
(M) = M, n+1 T M ,
n > 1,
Tpoly
equipped with the standard SchoutenNijenhuis bracket and differential d 0. An
m-differential operator in Dpoly (M) acts on a tensor product of m functions and has degree
m 1. The composition on Dpoly (M) is given by
(1 2 )(f0 fk1 +k2 )
=

k1
X

(1)k2 i 1 f0 fi1 (2 (fi fi+k2 )) fi+k2 +1 fk1 +k2
i=0
k

i
(M) and the Gerstenhaber bracket [1 , 2 ] is then given by
for i Dpoly

[1 , 2 ] = 1 2 (1)k1 k2 2 1 .

(28)

The differential on Dpoly (M) is given in terms of the Gerstenhaber bracket as


d = [, ],

(29)

790

B. Jurco et al. / Nuclear Physics B 584 [PM] (2000) 784794

where is the multiplication of functions: (f1 f2 ) = f1 f2 .


An L -morphism U : Tpoly(M) Dpoly (M) is then a collection of skew-symmetric
multilinear maps Un from tensor products of n > 1 polyvector fields to polydifferential
m+1
(M), satisfying the following condition
operators of degree m > 0: n Tpoly(M) Dpoly
(formality equation) [14,22]:

1 X
 (I, J )Q02 U|I | (I ), U|J | (J )
(30)
Q01 Un (1 , . . . , n ) +
2 I tJ =(1,...,n)
I,J 6=


1X
cj , . . . , n) Un1 (Q2 (i , j ), 1 , . . . , i , . . . , j , . . . , n ) .
 (i, j, 1, . . . , i,
=
2
i6=j

Here Q01 () = [, ], Q02 (1 , 2 ) = (1)|1|(|2|1)[1 , 2 ], with |i| denoting the degrees


of homogeneous polydifferential operators i and Q2 (1 , 2 ) = (1)k1 (k2 +1) [2 , 1 ],
where ki are degrees of homogeneous polyvector fields i . Further |I | denotes the numbers
of elements of I and  (I, J ) is +1 or 1 depending on the number of transpositions of
odd elements in the permutation of {1, . . . , n} associated with the partition (I, J ). Although
it doesnt explicitly enter the formality condition (30) a zero component U0 can be added
to U . By definition it is nonzero only if acting on two functions f g with the result
U0 (f, g) = fg.
In the case of M = Rd Kontsevich gives also a beautiful explicit expression for the
formality map U . To reproduce his formula we need to introduce a (2n + m 2)+
. If H denotes the upper half-plane
dimensional configuration space C{p
1 ,...,pn },{q1 ,...,qm }
+
then C{p1 ,...,pn },{q1 ,...,qm } is a quotient of


(p1 , . . . , pn ; q1, . . . , qm )|pi H, qj R, pi1 6= pi2 for i1 6= i2 , q1 < < qm
by the action of the group G(1) = {z 7 az + b|a, b R, a > 0} of orientation-preserving
affine transformations of the real line.
Then
X X
w B .
(31)
Un =
m>0 Gn,m

Here the second summation goes over all oriented admissible graphs with n vertices
p1 , . . . , pn of the first type and m vertices q1 , . . . , qm of the second type.
The rules are: There are no outgoing edges from the second type vertices. There are
k1 , k2 , . . . , kn edges starting in the the first type vertices which are ending in either first
type vertices again or in the second type vertices. There are no edges starting and ending
in the same vertex. The vertices and edges are enumerated in a fashion compatible with
the orientation, the edges starting at the first type vertex pi are labeled by numbers k1 + k2
+ + kj 1 + 1, . . . , k1 + k2 + kj . We denote


pj a1 , . . . ,
p
p
v k1 ++kj1 +i =
Star(pj ) =
j akj ,
j ai .
The weight w of the oriented graph is defined as an integral over the (2n + m 2)+
dimensional configuration space C{p
1 ,...,pn },{q1 ,...,qm }

B. Jurco et al. / Nuclear Physics B 584 [PM] (2000) 784794

Z
w =

(2)

1
ki k

1 ! kn !

v d
vk
1

1 ++kn

791

(32)

where



a pj
= Arg
.

p
ja
a pj

(33)

If 1 is of degree k1 1, 2 is of degree k2 1, . . . , n is of degree kn 1, then


B (1 , . . . , n )(f1 , f2 , . . . , fm )
X
ik +k
i1 i2 ...ik1
+1 ik1 ++kn
Dpn n 1 n1
Dq1 f1 Dqm fm ,
=
Dp1 1
where
Da =

l,
a
vl =

il

(34)

(35)

and the summation runs over repeated indices ij . Finally


X
U(k1 ,...,kn ) .
Un =

(36)

The weight w is nonzero only if the degree of the polydifferential operator and the overall
degree of the polyvector fields match as
m=2n+

n
X

(ki 1).

(37)

i=1

Only in this case the degree of the form in (32) matches the dimension of the configuration
+
. The construction can be globalized to any (formal) Poisson
space C{p
1 ,...,pn },{q1 ,...,qm }
manifold [14].
To make a relation with the deformation quantization a formal parameter, the Planck
constant h , has to be introduced. If is a two-tensor, then by the condition (37)
Un (, . . . , ) is a bidifferential operator for every n, i.e., it acts on two functions. If
moreover is a Poisson tensor then the Kontsevich star product ? is defined for f and g,
two smooth functions on M, as
X h n
Un (, . . . , )(f, g).
(38)
f ?g =
n!
n>0

The associativity of such a star product follows from the formality equation. If we set
in (30) i = , for i = 1, . . . , n and take into account the Jacobi identity [, ] and the
condition (37) we see that (30) is equivalent to the h n -order term of the associativity
condition for ?.
There are some other consequences from the formality theorem, which will be useful
later [23]. Here we present them in form convenient for the further use. We adopt the
following notation; for any vector field on the Poisson manifold (M, ) we denote as
the following formal series in h
X h n1
Un (, , . . . , ).
(39)
=
(n 1)!
n>1

792

B. Jurco et al. / Nuclear Physics B 584 [PM] (2000) 784794

From (37) we see that this is a differential operator. Its first term equals to . Let us apply
the formality equation in the case where 1 = and 2 = = n = . The matching
condition (37) says that in this case RHS of the formality equation is a bidifferential
operator; so let us act both sides of (30) on two functions f and g. Using (37) also in
evaluating the LHS we find:
(f ? g) (f ) ? g f ? (g) =

X h n1
Un1 ([, ], , . . . , )(f, g). (40)
(n 2)!

n>2

In particular we see that vector fields preserving the Poisson bracket are lifted via to
derivations of the Kontsevichs star product.
Further, if we use formality equation and the matching condition in the case, 1 = f ,
2 = = n = , we see that the RHS of (30) is a differential operator. Acting with both
sides of the formality equation on a function g we get
1 
f , g ? = [,f ] (g),
(41)
h
where the function f is given by a formal power series
f =

X h n1
Un (f, , . . . , ),
(n 1)!

(42)

n>1

starting with f . So the Hamiltonian vector fields are lifted to the inner derivations of the
star product.
Let us mention that [24] gives a very nice field theoretical interpretation of the formality
map.

5. Quantum
With the help of the formality theorem literally everything in Section 3 can be quantized.
To be consistent with physics conventions used in Section 2 we have to replace h by i h
everywhere. Let us quantize the Poisson structure (t) (14) on M for any t [0, 1] via
Kontsevichs deformation quantization (38)
X (i h )n

Un (t), . . . , (t) (f, g).
(43)
f ?t g =
n!
n>0

That way we get a star product ?t for any t. For any two t-independent functions f and
g on M we can take the t-derivative of f ?t g. Then Eqs. (15) and (40) give the quantum
version of (17)
t (f ?t g) + (t ) (f ?t g) (t ) (f ) ?t g f ?t (t )(g) = 0,

(44)

with (t ) given by (39)


(t ) =

X (i h )n1

Un (t)(A), (t), . . . , (t) .
(n 1)!

n>1

(45)

B. Jurco et al. / Nuclear Physics B 584 [PM] (2000) 784794

793

This means that we can relate the star products ?t at ?t 0 at two different time instants by
Dt t 0 , the flow of (t ) (or the quantum flow of (t)). Particularly for t = 0 and t 0 = 1
we have

(46)
D = et +(t) et t =0 .
Let us note that D is a composition D of the classical flow and a gauge equivalence
D = D ( )1 of the star product ? obtained from ?0 by simple action of and the star
product ?.
Finally the gauge transformation (21) is quantized with the help of (24), (41) and (42) as
1  
f ,
,
?
i h

(47)

X (i h )n1

Un , , . . . ,
=
(n 1)!

(48)

()

f 7 f +
where

n>1

and is obtained as explained in Section 3 from the condition


e[,] = et + (t ) et (t )

(49)

using the BCH formula.


The rest is trivial. We write similarly to (25)

D(x i ) = et +(t) x i t =0 = x i + A i .

(50)

Again A depends as a formal power series in on A. Explicitly we have



A i = exp(t + ij (t )Aj i ) 1 x i t =0 .

(51)

t =0

If we act by the infinitesimal gauge transformation (21) on A, this induces now the action
of the inner derivation (47) on D(x i ), which in turn induces a map on A given by
1  i
1  i
()
, x ? +
, A ? .
A i 7 A i +
i h
i h

(52)

So (51) gives indeed the desired SW map to all orders in in the case of a general
Poisson manifold. Moreover, using Kontsevichs construction of the formality map U as
described in Section 4, we can find explicit expressions in local coordinates for A and to
any order in .

Acknowledgements
We would like to thank S. Theisen for encouraging discussions. B.J. thanks the
Alexander-von-Humboldt-Stiftung for support and J. Donin, P. Bressler and S. Merkulov
for discussions.

794

B. Jurco et al. / Nuclear Physics B 584 [PM] (2000) 784794

References
[1] C.G. Callan, C. Lovelace, C.R. Nappi, S.A. Yost, String loop corrections to beta functions, Nucl.
Phys. B 288 (1987) 525.
[2] A. Abouelsaood, C.G. Callan, C.R. Nappi, S.A. Yost, Open strings in background gauge fields,
Nucl. Phys. B 280 (1987) 599.
[3] C.-S. Chu, P.-M. Ho, Noncommutative open string and D-brane, Nucl. Phys. B 550 (1999) 151,
hep-th/9812219.
[4] V. Schomerus, D-branes and deformation quantization, JHEP 9906 (1999) 030, hep-th/9903205.
[5] A. Connes, M.R. Douglas, A. Schwarz, Noncommutative geometry and matrix theory:
compactification on tori, JHEP 9802 (1998) 003, hep-th/9711162.
[6] M.R. Douglas, C. Hull, D-branes and the noncommutative torus, JHEP 9802 (1998) 008, hepth/9711165.
[7] B. Morariu, B. Zumino, in: Relativity, Particle Physics and Cosmology, World Scientific,
Singapore, 1998, hep-th/9807198.
[8] W. Taylor, D-Brane field theory on compact spaces, Phys. Lett. B 394 (1997) 283, hepth/9611042.
[9] N. Seiberg, E. Witten, String theory and noncommutative geometry, JHEP 9909 (1999) 032,
hep-th/9908142.
[10] O. Andreev, H. Dorn, On open string sigma model and noncommutative gauge fields, hepth/9912070.
[11] B. Jurco, P. Schupp, Noncommutative YangMills from equivalence of star products, Eur.
Phys. J. C 14 (2000) 367, hep-th/0001032.
[12] T. Asakawa, I. Kishimoto, Noncommutative gauge theories from deformation quantization, hepth/0002138.
[13] F. Bayen, M. Flato, C. Fronsdal, A. Lichnerowicz, D. Sternheimer, Deformation theory and
quantization. I. Deformations of symplectic structures, Ann. Phys. 111 (1978) 61.
[14] M. Kontsevitch, Deformation quantization of Poisson manifolds, I, q-alg/9709040.
[15] D. Sternheimer, Deformation quantization: twenty years after, math/9809056.
[16] L. Cornalba, Corrections to the Abelian BornInfeld action arising from noncommutative
geometry, hep-th/9912293.
[17] J. Madore, S. Schraml, P. Schupp, J. Wess, Gauge theory on noncommutative spaces, hepth/0001203.
[18] J. Moser, On the volume elements on a manifold, Trans. Am. Math. Soc. 120 (1965) 286.
[19] N. Ishibashi, A relation between commutative and noncommutative descriptions of
D-branes, hep-th/9909176.
[20] K. Okuyama, A path integral representation of the map between commutative and noncommutative gauge fields, hep-th/9910138.
[21] L. Cornalba, D-brane physics and noncommutative YangMills theory, hep-th/9909081.
[22] D. Arnal, D. Manchon, M. Masmoudi, Choix des signes pour la formalite de M. Kontsevich,
math.QA/0003003.
[23] D. Manchon, Poisson bracket, deformed bracket and gauge group actions in Kontsevich
deformation quantization, math.QA/0003004.
[24] A.S. Cattaneo, G. Felder, A path integral approach to the Kontsevich quantization formula,
math/9902090.

Nuclear Physics B 504 [PM] (2000) 795809


www.elsevier.nl/locate/npe

Schubert calculus and threshold polynomials


of affine fusion
S.E. Irvine 1 , M.A. Walton
Physics Department, University of Lethbridge, Lethbridge, Alberta, Canada T1K 3M4
Received 19 April 2000; accepted 20 June 2000

Abstract
We show how the threshold level of affine fusion, the fusion of WessZuminoWitten (WZW)
conformal field theories, fits into the Schubert calculus introduced by Gepner. The Pieri rule can
be modified in a simple way to include the threshold level, so that calculations may be done for
all (non-negative integer) levels at once. With the usual Giambelli formula, the modified Pieri
formula deforms the tensor product coefficients (and the fusion coefficients) into what we call
threshold polynomials. We compare them with the q-deformed tensor product coefficients and fusion
coefficients that are related to q-deformed weight multiplicities. We also discuss the meaning of the
threshold level in the context of paths on graphs. 2000 Elsevier Science B.V. All rights reserved.
PACS: 11.25.Hf
Keywords: Conformal field theory; WZW model; Affine fusion; Schubert calculus

1. Introduction
Gepner found geometrical and topological interpretations of the fusion rings of Wess
ZuminoWitten (WZW) conformal field theories [1]. He described them using a Schubert
calculus, a quantum version of the classical Schubert calculus that is fundamental in the
geometry and topology of complex manifolds (e.g., see [24]).
Gepner also pointed out a correspondence between the WZW fusion rings and the chiral
rings of N = 2 superconformal theories. These two observations have been seminal. For
example, their relation was clarified in [5], where the new Schubert calculus was shown to
describe the quantum cohomology of Grassmannians. Also, the N = 2 interpretation led
to new realisations of WZW fusion rings in topological theories [68].
We study the Schubert calculus of WZW fusion rings. Our initial motivation was
computational. In Gepners approach, a fusion potential is introduced whose derivatives
Corresponding author. E-mail: walton@uleth.ca; Supported in part by NSERC.
1 Supported in part by a NSERC Undergraduate Research Award.

0550-3213/00/$ see front matter 2000 Elsevier Science B.V. All rights reserved.
PII: S 0 5 5 0 - 3 2 1 3 ( 0 0 ) 0 0 4 0 4 - 1

796

S.E. Irvine, M.A. Walton / Nuclear Physics B 504 [PM] (2000) 795809

give the fusion constraints to be implemented. The fusion potential, and so the fusion
constraints, are level dependent. Therefore a significant part of any computation must be
re-done whenever the level is changed. By the depth rule [9], however, the results are
simpler than this procedure indicates. A threshold level exists for each coupling [10,11]; in
any fusion product of two fixed fields, a third appears in the decomposition for all integer
levels greater than or equal to a characteristic one. 2 Therefore, finding the threshold levels
for a fixed product amounts to finding the fusion rules for all levels at once.
This striking simplicity should be exploited. Our hope is that the threshold level
approach indicates something deep about the WZW models, and their associated affine
algebras. Perhaps there exists a physical or mathematical context where different (nonnegative integer) levels are treated together, on the same footing. As a possible step in this
direction, we show how to incorporate the notion of threshold level into Gepners Schubert
calculus for WZW fusion. This is done in Section 3, after a review is given in Section 2,
where the notation is also established.
Another motivation for this work emerges in Section 2: it is convenient to encode the
threshold levels in generating polynomials, dubbed threshold polynomials. These are then
polynomial deformations of tensor product coefficients and fusion coefficients. Similar
objects, the quantum group (q-)deformations of tensor product coefficients [1215] and
fusion coefficients [16,17] have been studied previously. Most importantly to us, the
q-deformed coefficients are related to the q-deformed weight multiplicities defined by
Lusztig [18]. In Section 4 we compare our deformations with the q-deformations. We
show that the new deformations are related in a similar way to deformations of the weight
multiplicities, that are natural from the point of view of a conjectured refinement [11,19,
20] of the GepnerWitten depth rule [9].
As we have argued, the threshold level has computational advantages over the use
of fusion potentials, and the relations derived from them. But again, the connection
with geometry, topology and N = 2 superconformal theories was the point of [1], not
computation. There the fusion potential played a central role. But one cant have it both
ways: we indicate at the end of Section 3 that a deformed fusion potential that incorporates
the threshold levels cannot be written. Nevertheless, one might hope to give the threshold
level a somewhat deeper motivation, perhaps through its meaning in the many different
realisations of WZW fusion rings. In Section 5 we make a very small start in this direction;
we discuss the meaning of the threshold level in the context of paths on graphs (see [21,22],
and references therein).
Section 6 is a short conclusion.

2. WZW fusion, threshold level, and threshold polynomials


Let us first establish notation. For the most part, we restrict attention to the simple
(1)
Lie algebras Ar and the affine algebras Ar that are the untwisted central extensions
2 To the best of our knowledge, however, a completely rigorous demonstration of the existence of threshold
levels is still lacking.

S.E. Irvine, M.A. Walton / Nuclear Physics B 504 [PM] (2000) 795809

797

of their loop algebras. When the level k is fixed, we denote the affine algebra by Ar,k .
However, we use a notation that is easily adapted to any untwisted affine KacMoody
(1)
algebra Xr (or Xr,k ) based on a simple Lie algebra Xr , and expect that such generalisation
is straightforward.
The set of roots of Xr will be written as R, and the set of positive (negative) roots as R>
(R< ). If R is a root, then the corresponding co-root is defined as := 2/(, ).
Let F = {1 , . . . , r } denote the set of fundamental weights of Xr , and
)
(
r
X
i i | i Z
(2.1)
P := =
i=1

the set of integral weights. The set of dominant integral weights,


(
)
r
X
i i | i Z> ,
P> := =

(2.2)

i=1

is the set of highest weights for irreducible integrable modules of Xr .


Let M() denote an irreducible module of Xr , of highest weight P> . The set of
weights of M() will be denoted P ().
The irreducible integrable modules of Xr,k have highest weights that project to the
following set of dominant weights of Xr :
(
)
r
r
X
X
k

i i | i Z> ,
i ai 6 k .
(2.3)
P> := =
i=1

The

ai

i=1

are the co-marks, defined by

= =

r
X

ai i ,

a0

= 1, and
(2.4)

i=1

where ( ) denotes the highest (co-)root of Xr . We normalise (, ) = 2.


The Weyl group of Xr will be denoted by W , and the shifted action of w W on a
P
P
weight by w. = w( + ) , where = ri=1 i = R> /2 is the Weyl vector.
W k will indicate the projection of the affine Weyl group, the Weyl group of Xr,k , onto the
horizontal weight space, the weight space of Xr . W is generated by the primitive reflections
ri , i = 1, . . . , r, with action
ri = (, i )i

(2.5)

on any weight . In order to enlarge W to W k , we adjoin r0 to the generating set. Its shifted
action is
r0 . = r . + (k + x),

(2.6)

where x is the dual Coxeter number of Xr . Notice the k-dependence of the action of W k
on P , coming from that of r0 .
We write the decomposition of the tensor product of two irreducible integrable Xr modules as
M

T,
M().
(2.7)
M() M() =
P>

798

S.E. Irvine, M.A. Walton / Nuclear Physics B 504 [PM] (2000) 795809

Z tensor product coefficients. We indicate the affine fusion of


We will call the T,
>
two modules of Xr,k by writing the truncated tensor product of the corresponding modules
M() and M() of Xr :
M
(k)
T, M().
(2.8)
M() k M() =
k
P>

The fusion coefficients obey


(k)

T,
6 (k+1) T,
,

(2.9)

and furthermore
lim

(k)

T,
= T,
.

(2.10)

We can encode the fusion products for all levels by including the threshold levels t as
subscripts in the tensor product decomposition [10,11]. If we denote by S t the operator
that includes these subscripts, we can write

 M M (t )
T, M()t .
(2.11)
St M() M() =
P> t Z>

Then
(k)

T,
=

k
X

(t )

T, .

(2.12)

t =0
(t )

We call the fixed-threshold-level coefficients T, , the threshold coefficients. For example,


we can modify the A2 tensor product to


(2.13)
St M(1, 1)2 = M(2, 2)4 M(3, 0)3 M(0, 3)3 2M(1, 1)2,3 M(0, 0)2 ,
encoding the corresponding fusions at all levels. Here M(a, b) := M(), with = a1
+ b2 , and pM(a, b)t1,...,tp := M(a, b)t1 M(a, b)tp .
From the notational point of view, the threshold levels are unnecessarily large numbers,
since
(t )
6= 0 t > (, ).
T,

(2.14)

Consequently, one could also define the threshold delay d by


d := t (, ).

(2.15)

Writing the delays as superscripts, the right-hand side of (2.13) is replaced by


M(2, 2)2 M(3, 0)0 M(0, 3)0 2M(1, 1)0,1 M(0, 0)2.

(2.16)

This is a minor point, so well stick to using the threshold levels. In Section 4, however,
(2.15) will reappear.
As (2.13) makes clear, we need to consider N -tuples of threshold levels. It is convenient
to encode them in threshold polynomials, defined by
X
(t )

[`] :=
` t T,
.
(2.17)
T,
t Z>

S.E. Irvine, M.A. Walton / Nuclear Physics B 504 [PM] (2000) 795809

Then


 M
T, [`]M()
Lt M() M() =

799

(2.18)

P>

is equivalent to (2.11). For example, the A2 tensor product (2.13) is rewritten as




Lt M(1, 1)2 = `4 M(2, 2) `3 M(3, 0)
`3 M(0, 3) (`2 + `3 )M(1, 1) `2 M(0, 0).

(2.19)

The threshold polynomials can be regarded as deformations of the tensor product


coefficients, since

[1] = T,
,
T,

(2.20)

so that the tensor product (2.7) is recovered when ` = 1. Furthermore, we define the
deformation of the fusion coefficient as
(k)

T,
[`] :=

k
X

(t )

`t T, .

(2.21)

t =0
[`] is the degree 6 k part of the polynomial T [`], and
So (k) T,
,
(k)

T,
[1] = (k) T,
.

(2.22)

(2.10) is deformed to
lim

(k)

T,
[`] = T,
[`],

(2.23)

by construction.
From (2.17), we see that the threshold polynomials are the generating functions for the
threshold coefficients. Consequently, they are related to the generating functions for fusion
rules studied in [10], where the threshold level was first introduced (but named later in
[11]). For completeness, we indicate the relation here.
The generating function for fusion rules is defined as
G(L, M, N; d) :=

(k)

T,, d k L M N ,

(2.24)

,,P> k=0
0

C ,
where the dummy variables L, M, N satisfy L L = L+ , etc. Here (k) T,, := (k) T,
where C is the highest weight of the module conjugate to M(). Using (2.12) and
switching the order of summations, we arrive at
X
L M N T,, [d].
(2.25)
G(L, M, N; d) = (1 d)1

,,P>
C(t )

Here T,, [d] = t =0 d t T, ; see (2.21), (2.23) and (2.17). Hence the only difference
between the generating functions for deformed tensor product coefficients and fusion rules
is (1 d)1 , a factor characteristic of the existence of a threshold level [10].
For successive fusions, we need a memory of the threshold levels. For example, suppose
we need to calculate St [M() (M(1, 1)2 )]. Then using (2.13), we would encounter

800

S.E. Irvine, M.A. Walton / Nuclear Physics B 504 [PM] (2000) 795809

products like St [M() M(1, 1)3 ]. So (2.11) is only a special case of what we need.
Abusing notation slightly, we attach threshold levels to the factor modules in the tensor
products, and write
M M (t )
T(r),(s)M()t .
(2.26)
M()r M()s =
P> t Z>

For example, we have


M(1, 1)2 M(1, 1)3 = M(2, 2)4 M(3, 0)3 M(0, 3)3
2M(1, 1)3,3 M(0, 0)3.
(2.11) is recovered from (2.26) by setting r = (, ) and s = (, ).
Again using polynomials to carry the N -tuples of threshold levels, we write

 O
T(r),(s)[`]M().
Lt M()r M()s =

(2.27)

(2.28)

P>

Comparing (2.27) with (2.13), for example, shows that there is a simple relation between
(t )
(t )
and T,
. To write it in polynomial form, we define
the coefficients T(r),(s)
`a `b := `max{a,b}

(2.29)

and extend bilinearly (so that two polynomials can be multiplied). Then we have

[`] = `r `s T,
[`].
T(r),(s)

(2.30)

[`] are fundamental, and so we will concentrate on them


We see that the polynomials T,
henceforth.
The definition (2.29), however, is natural from the point of view of threshold
polynomials. With it, we can generalise the crossing symmetry of the tensor product
coefficients,
X
X

T, T,
=
T,
T,
,
(2.31)
P>

to

P>

T, [`] T,
[`] =

P>

T,
[`] T,
[`].

Furthermore, the crossing symmetry for the fusion coefficients


X
X
(k) (k)
(k) (k)
T, T, =
T, T,
k
P>

deforms to
X
k
P>

(2.32)

P>

(2.33)

k
P>

(k)

T, [`] (k) T,
[`] =

X
k
P>

(k)

T,
[`] (k) T,
[`].

(2.34)

S.E. Irvine, M.A. Walton / Nuclear Physics B 504 [PM] (2000) 795809

801

3. Schubert calculus, threshold level, and threshold polynomials


The Schubert calculus is based on the Pieri and Giambelli formulas. For discussions
of them emphasising the geometric context see [14]. More relevant to us is their use in
representation theory; e.g., consult [23].
The Pieri formula is

1, if P (),

=
(3.1)
T,
0, otherwise,
where is a fundamental weight, i.e., F . Here we are specialising to the algebras Ar ,
although the formulas for other algebras are only slightly more complicated.
Adapted to include threshold polynomials, the Pieri formula simply becomes
 (,)
` (,) , if P (),
`

[`] =
(3.2)
T,
0,
otherwise.
Fundamental monomials
M( ) := M(1 )1 M(2 )2 M(r )r =

r
O

M(i )i ,

(3.3)

i=1

with all i F , are easily decomposed using the Pieri formula (3.1). The decompositions
are triangular in the irreducible highest-weight modules M( ), P> :
M
, M( ).
(3.4)
M( ) =
P> 3 6

Here 6 means that is a non-negative integer linear combination of positive roots,


i.e., Z> R> . So , is a triangular matrix.
The polynomial deformation of this, encoding the threshold levels, is just
M


, [`]M( ).
(3.5)
Lt M =
P> 3 6

Some A2 examples (to be used shortly) will make this clear. We find:


St M (2,2) = M(2, 2)4 M(3, 0)3 4M(1, 1)2,2,2,3
M(0, 3)3 2M(0, 0)1,2,

(3.6)

or


Lt M (2,2) = `4 M(2, 2) `3 M(3, 0) (`3 + 3`2 )M(1, 1)
`3 M(0, 3) (` + `2 )M(0, 0),
and

or



St M (1,1) = M(1, 1)2 M(0, 0)1,


Lt M (1,1) = `2 M(1, 1) `1 M(0, 0),



St M (0,0) = M(0, 0)0,


Lt M (0,0) = M(0, 0).

(3.7)

(3.8)

(3.9)

802

S.E. Irvine, M.A. Walton / Nuclear Physics B 504 [PM] (2000) 795809

From (3.4), the highest-weight modules can be expressed as polynomials in the


fundamental weights:
M

(3.10)
1 , M( ).
M( ) =
P> 36

That is, M( ) can be written as a direct sum of fundamental monomials M( ). This is


the Giambelli formula, in the non-determinantal form that can be applied to all simple Lie
algebras, not just Ar . Notice that the inversion of is greatly simplified by its triangularity,
and its inverse is also triangular. A simple A2 example of (3.10) is


(3.11)
M(1, 1) = M (1,1) M (0,0) .
The characters of Xr form an algebra with structure constants equal to the tensor
product coefficients. The Giambelli formula (3.10) gives rise to a polynomial realisation
Q

of this character algebra. One simply replaces M( ) with ri=1 xi i =: x . The resulting
polynomial
X

(3.12)
1 , x ,
S (x) :=
P> 36

is known as a Schur polynomial, a type of Schubert polynomial [24]. For example, x1 x2 1


is the Schur polynomial of the A2 module M(1, 1), by (3.11). With the addition and
subtraction of polynomials, the character algebra extends to a ring.
Is there a useful threshold-level version of the Giambelli formula? The inverse 1 [`]
of the matrix [`] in (3.5) has entries that are negative powers of `. These are difficult
to interpret in the context of threshold level. We conclude that the normal, `-independent
matrix 1 should be used. We can write useful formulas for the threshold polynomials
in terms of 1 , and its deformed inverse [`]:
X




[`] = `(,) `(,)


(3.13)
1 , 1 , +, [`].
T,
,P>

Well illustrate this formula on the A2 example with M() = M() = M(1, 1), using the
subscript notation. First, the required matrix elements of 1 are provided by


2
M(1, 1)2 = M (1,1) M (0,0)



= M (2,2) 2M (1,1) M (0,0) .
(3.14)
Substituting the fundamental monomials (3.6) and (3.8), described by [`], we get



Lt M(1, 1)2 = `2 `4 M(2, 2) `3 M(3, 0) (3`2 + `3 )M(1, 1) `3 M(0, 3)



(` + `2 )M(0, 0) 2 `2 M(1, 1) `M(0, 0) M(0, 0)
= `4 M(2, 2) `3 M(3, 0) `3 M(0, 3)
(`2 + `3 )M(1, 1) `2 M(0, 0),
the correct result.

(3.15)

S.E. Irvine, M.A. Walton / Nuclear Physics B 504 [PM] (2000) 795809

803

The deformed Pieri formula (3.2) makes straightforward the calculation of decompositions involving fundamental monomials, like M( ). We write

 M
T, [`]M().
(3.16)
Lt M() M =
P>

Then the threshold polynomials may also be calculated from the simpler polynomials

T,
[`]:
X


[`] = `(,)
(3.17)
T,
1 , T,
[`].
P>

Using (3.11), an A2 example is








Lt M(1, 1)2 = `2 Lt M(1, 1) M (1,1) M (0,0)
= `4 M(2, 2) `3 M(3, 0) (2`2 + `3 )M(1, 1)
`3 M(0, 3) `2 M(1, 1) `2 M(1, 1).

(3.18)

This is again the correct result (see (3.15)).


Finally, we can also multiply two Schur polynomials for M() and M() together, using
the coefficients T , , defined in the obvious way:
X



[`] = `(,) `(,)


(3.19)
1 , 1 , T , [`].
T,
,P>

To conclude this section, we note that in our deformed Schubert calculus, there is no
analogue of the fusion potential that was so important in [1]. We argue that a deformed
potential that incorporates the threshold levels cannot be written. Gepner could write a
fusion potential because at fixed level k, the fusion rules are truncations of the tensor
product rules. The truncated parts can be set to zero by fusion constraints, that can be
derived from the potential. On the other hand, when the threshold level is incorporated into
a tensor product, as in (3.5) vs. (3.4), there is no truncation. Instead of constraints, one
could only hope to find replacements that would change the right-hand side of (3.4) into
that of (3.5), for example. But that is exactly what we do: (3.5) is obtained from (3.4) by
replacing with [`]. A minimal set of such replacements would be those obtained by
replacing the right-hand side of the undeformed Pieri rule (3.1) with that of the deformed
one (3.2).
Incidentally, we have seen that the Pieri rule with threshold level (3.2) contains the same
information as the fusion potentials of Gepner, for all (non-negative integer) levels. So does
the generating function for the fusion potentials [1]. It might be interesting to make this
more precise.

4. Deformed tensor product coefficients and weight multiplicities


The threshold polynomials (2.17) and (2.21) are deformations of the tensor product
coefficients and affine (WZW) fusion coefficients, respectively. It is interesting to compare

804

S.E. Irvine, M.A. Walton / Nuclear Physics B 504 [PM] (2000) 795809

them with the q-deformations of these objects studied previously.


WZW fusion coefficients are alternating affine-Weyl (W k ) sums of tensor product
coefficients [2528]. In [16] (see also [17]), the corresponding q-fusion coefficients (for
affine Ar ) are defined in similar fashion in terms of the q-tensor product coefficients
[1215]. Since the ordinary (undeformed) tensor product coefficients are also alternating
Weyl sums of the weight multiplicities of Lie algebras, the fusion coefficients can also be
expressed in that way. In the q-deformed case, the tensor product and fusion coefficients
are related to Lusztigs q-deformed weight multiplicities [18], in turn related to the famous
KazhdanLusztig polynomials [29].
Let us start with an example, taken from [16]. They find, for the q-deformation of the
A3 tensor product M(1, 1, 0)3 , the following decomposition:
(q 3 + q 6 )M(0, 0, 3) (2q 4 + 3q 5 + 2q 6 + q 7 )M(0, 1, 1)
(q 2 + 2q 3 + q 4 )M(0, 3, 1) (q 5 + 2q 6 + q 7 )M(1, 0, 0)
(q 2 + 2q 3 + 3q 4 + 2q 5)M(1, 1, 2) (2q 3 + 3q 4 + 3q 5 + q 6 )M(1, 2, 0)
(q + q 2 )M(1, 4, 0) (q 3 + 3q 4 + 3q 5 + 2q 6 )M(2, 0, 1)
(q + 2q 2 + 2q 3 + q 4 )M(2, 2, 1) (q 2 + 2q 3 + q 4 )M(3, 0, 2)
(q 2 + 2q 3 + 2q 4 + q 5 )M(3, 1, 0) M(3, 3, 0)
(q + q 2 )M(4, 1, 1) (q 3 )M(5, 0, 0).

(4.1)

This should be compared with the threshold-level version of the same tensor product:


Lt M(1, 1, 0)3 = (2`3 )M(0, 0, 3) (`2 + 7`3 )M(0, 1, 1) (4`4 )M(0, 3, 1)
(2`2 + 2`3 )M(1, 0, 0) (8`4 )M(1, 1, 2)
(4`3 + 5`5 )M(1, 2, 0) (2`5 )M(1, 4, 0)
(5`3 + 4`4 )M(2, 0, 1) (6`5 )M(2, 2, 1)
(4`5 )M(3, 0, 2) (4`4 + 2`5 )M(3, 1, 0)
(`6 )M(3, 3, 0) (2`6 )M(4, 1, 1) (`5 )M(5, 0, 0).

(4.2)

From this example, we see no clear relation between the q-deformations and the `deformations, except that they coincide at q = ` = 1.
In order to define the q-tensor product coefficients, one introduces the q-deformed
Kostant partition function K(; q):
X
Y
1
=:
K(; q)e .
(4.3)
1 qe
R>

Z> R>

From this we see that the powers of q count the number of positive roots in a decomposition
of an element of Z> R> . The q-deformed weight multiplicities are
X
(det w)K(w. ; q),
(4.4)
mult (; q) :=
wW

and we get the q-deformed tensor product coefficients as


X

(q) :=
(det w)mult (w. ; q).
T,
wW

(4.5)

S.E. Irvine, M.A. Walton / Nuclear Physics B 504 [PM] (2000) 795809

805

(q) vs.
Notice we use different brackets to distinguish the different deformations: T,

T, [`]. Finally, the q-fusion coefficients [16] can be found from


X
(k)
w.
T, (q) :=
(det w)T,
(q).
(4.6)
wW k

It would be interesting to define the threshold polynomial versions of the q-Kostant


partition function, and the q-multiplicities. The relation between the q-tensor product
coefficients and the threshold polynomials might then be extracted. We have not succeeded
in finding the `-Kostant partition function. But the `-multiplicities may be defined using
a conjectural refinement [11,19] of the GepnerWitten depth rule [9]:



(k)
T, = dim v M(; ) (fi )i +1 v = 0, i {0, 1, . . . , r} .
(4.7)
Here M(; ) is the subspace of weight of the module M(), so that
dim M(; ) = mult ( ). The fi are the lowering operators corresponding to the simple
roots of the simple Lie algebra Xr Xr,k , for i = 1, . . . , r. f0 is identified with e , the
raising operator corresponding to the highest root of Xr . Recall that the depth of a vector
v (in a module M(), say) is defined as the non-negative integer d such that (e )d v 6= 0,
while (e ) d+1 v = 0. The relation between (4.7) and the GepnerWitten depth rule is then
clear.
By (4.7), we see that the subspaces



(4.8)
M(; |d) := v M(; ) (e )1+d v = 0, (e )d v 6= 0
of the spaces M(; ) are relevant. Because of their relation to the depth, the multiplicities
that are the dimensions of these spaces,
mult ( |d) := dim M(; | d),

(4.9)

were dubbed profundities in [20]. They have the same relation to the threshold
coefficients (see (2.12) that the the usual multiplicities have to the tensor product
coefficients:
X

(t )
(det w)mult w. | t (, ) ,
(4.10)
T, =
wW

where t > (, ). Substituting this into (2.17) gives


X

[`] := `(,)
(det w)mult (w. ; `),
T,

(4.11)

wW

with `-deformed multiplicities


X
`d mult ( | d).
mult ( ; `) :=

(4.12)

dZ>

So the only complication is the overall factor `(,) , and the `-deformed multiplicities are
generating functions for the profundities.
Incidentally, in deriving this last result, we used the relation
d = t (, )

(4.13)

806

S.E. Irvine, M.A. Walton / Nuclear Physics B 504 [PM] (2000) 795809

between the depth d and the threshold level t of a fixed coupling. Notice that this
is identical to (2.15). Hence the threshold delay of a coupling equals the depth of the
corresponding vector in (4.7).

5. Fusion paths and the threshold level


Paths on fusion graphs are important in certain integrable lattice models that are related
to conformal field theory (see [21,22], and references therein). These graphs may also have
a more fundamental significance, indicated by the correspondence between A1 modular
invariants and the A D E graphs [30].
If we restrict to the case Xr,k = Ar,k , then the points of the relevant graphs correspond
k . There is a distinct directed graph (k) G for each fundamental weight
to the weights of P>
i
i F . The edges of the graph (k) Gi are encoded in its incidence matrix (k) Gi , which is
not necessarily symmetric. ((k) Gi ), is the number of edges joining node with node .
The fusion graph is defined by ((k) Gi ), = (k) T i , , or (k) Gi = (k) Ti , hence the name.
P
One can also define a graph (k) G with incidence matrix (k) G := ri=1 (k) Gi .
A fusion path is a path on a fusion graph. Such paths parametrise the Hilbert
space of certain integrable two-dimensional lattice models. The basic construction is a
representation of a (quotient of a) Hecke algebra on this space. It guarantees that the
models Boltzmann weights satisfy the YangBaxter equation, ensuring integrability.
k+1
k
P>
P> , we can think of the graphs (k) Gi
Due to (2.9), (2.10), and since P>
(k)
and G in the infinite-level limit as tensor product graphs. (Here we restrict consideration
to paths involving weights that do not increase with the level.) Such paths on P> will also
be paths on all (k) G, for all levels k greater than a certain threshold level t. This threshold
level, is just the maximum height ht () := (, ) of the weights P> on the path.
Key to the modified Schubert calculus described above were the fundamental monomials, and their decompositions (3.4). But the fundamental monomials M( ) generate paths
in P> : to every module M( ) in the decomposition (3.4) there corresponds a path on P>
that begins at the weight 0 and ends at . To each factor M(), F , in the monomial
corresponds a segment of the path that connects nodes of the graph that differ by some
P ().
The threshold level is included by modifying (3.4) to (3.5), using the `-Pieri rule. From
(3.2), we see that the threshold level is the maximum height of a path on the (infinite)
tensor product graph of Ar . This is the main point of this section.
This simple interpretation of the threshold level may hint at a method of considering all
levels at once in the corresponding integrable lattice models. This would avoid some of the
tedious case-by-case analysis that has been necessary in the construction of such models.
We will not pursue this further here, however. Our purpose was just to give one physical
context where the threshold level has a simple meaning.
Before closing this section, let us emphasise that the correspondence between the
fundamental monomials and tensor product paths is not one-to-one. The polynomial
realisation (3.12) of the fusion ring is possible because M() M(0 ) = M(0 ) M(),

S.E. Irvine, M.A. Walton / Nuclear Physics B 504 [PM] (2000) 795809

807

for all , 0 F . But the order of tensor product factors M() changes the path. This can
be made clear by writing a generating matrix for fundamental monomials:
:=

e T11 Trr =

P>

r
Y

1
.
1 ei Ti

(5.1)

i=1

Here e denotes a formal exponential, satisfying e e = e+ . , will equal the sum


over all fundamental monomials that when tensored with M(), include M() in the
decomposition, multiplied by the formal exponential of the monomial weight of each.
Putting e 1 then gives the number of such monomials. On the other hand, to generate
all paths connecting nodes and , we need , instead, where

1
r
X
i
e Ti
.
(5.2)
:= 1
i=1

The deformations of these two generating matrices are simple to write. One only needs
to replace the tensor product matrices Ti with their `-deformations, and insist that they
are multiplied in the manner of (2.32). So we get
[`] =

r
Y


1 ei Ti [`]

(1)

(5.3)

i=1

where the notation (we hope) is clear, and a similar formula analogous to (5.2).

6. Conclusion
Our main result is a Schubert-type calculus for affine fusion that incorporates the
threshold level. At fixed level, fusion constraints are natural because a fusion rule is
a truncation of a tensor product decomposition. Thus fusion potentials that generate
constraints are possible, if not necessary. On the other hand, in the threshold level
formalism one doesnt truncate a tensor product, but rather replaces it with a deformed
version. So, instead of using a fusion potential to generate constraints, one just deforms the
tensor products and then all (non-negative integer) levels are treated on equal footing. The
deformations are generated by the deformed version of the Pieri rule (3.2).
In summary then, to include the threshold levels in a calculus of Schubert type, use
the undeformed Giambelli formula (3.10), and the deformed Pieri formula (3.2). Then the
threshold polynomials can be calculated by (3.13), (3.17), or (3.19).
Another result is the comparison in Section 4 of the threshold polynomials with the qdeformed tensor product and fusion coefficients. In particular, we found the analogue of
the q-deformed weight multiplicities in our `-deformation.
We also discussed the interpretation of the threshold level for the decomposition of
fundamental monomials, as in (3.5). In the corresponding path on a tensor product graph,
the threshold level is just the maximum height of weights on that path.
To close, let us mention a few possible directions from this work.

808

S.E. Irvine, M.A. Walton / Nuclear Physics B 504 [PM] (2000) 795809

One could hope to make the connection between the q-deformations and `-deformations
more precise, extending our Section 4. We also expect that one could define a q-Schubert
calculus for the q-tensor product coefficients (4.5), in a straightforward way. In contrast
with the `-deformed case, both (q) and 1 (q) should be important. It might be of
interest to introduce q-analogues of the fusion constraints and potentials of Gepners
calculus, for the q-fusion-coefficients.
The `-deformed Schubert calculus is relevant to the search for a LittlewoodRichardson
rule for affine fusion [31]. In the present context, the usual LittlewoodRichardson rule
for tensor products is related to (3.17), at ` = 1. This formula involves the tensor product
of two modules M(), M(), where one is expressed in terms of fundamental monomials
P

by (3.10): M() = ()1


, M( ). The rule gives a way of avoiding the cancellations
inherent in (3.17) (see, e.g., (3.15)). It identifies a choice of a part of the decompositions of
the M( ) in (3.17) that leads directly to the result. Unfortunately, the deformed Pieri rule
applied to that choice gives incorrect threshold levels (e.g., one finds 2M(1, 1)2,2 instead of
the 2M(1, 1)2,3 of (3.15)). Calculations of the type (3.17), however, show us all the parts.
And so we can hope that more in-depth analysis will reveal the appropriate modification.
Finally, if it exists, a motivation other than computational for the threshold level should
be found. It might be revealed by finding the meaning of the threshold level in the many
different physical and mathematical realisations of affine fusion.

Acknowledgements
We thank L. Bgin, T. Gannon, P. Mathieu and J. Rasmussen for comments, and
D. Snchal for the use of some computer programs.

References
[1]
[2]
[3]
[4]
[5]
[6]
[7]
[8]
[9]
[10]
[11]
[12]
[13]
[14]

D. Gepner, Commun. Math. Phys. 141 (1990) 381.


H. Hiller, Geometry of Coxeter Groups, Pitman, 1982.
W. Fulton, Young Tableaux, Cambridge Univ. Press, 1997, Part III.
P. Griffiths, J. Harris, Principles of Algebraic Geometry, Wiley, 1978.
E. Witten, Geometry, Topology and Physics, Conf. Proc. Lect. Notes in Geom. Topol. VI (1995)
357.
K. Intriligator, Mod. Phys. Lett. A 6 (1991) 3543.
C. Vafa, in: S.-T.Yau (Ed.), Essays on Mirror Manifolds, International Press, 1992.
D. Nemeschansky, N. Warner, Nucl. Phys. B 380 (1992) 241.
D. Gepner, E. Witten, Nucl. Phys. B 278 (1986) 493.
C.J. Cummins, P. Mathieu, M.A. Walton, Phys. Lett. B 254 (1991) 386.
A.N. Kirillov, P. Mathieu, D. Snchal, M.A. Walton, Nucl. Phys. B 391 (1993) 651.
A. Schilling, S.O. Warnaar, Commun. Math. Phys. 202 (1999) 359.
A.N. Kirillov, M. Shimozono, A generalization of the KostkaFoulkes polynomials,
math.QA/9803062.
M. Shimozono, J. Weyman, Graded characters of modules supported by the closure of a
nilpotent conjugacy class, math.QA/9804036.

S.E. Irvine, M.A. Walton / Nuclear Physics B 504 [PM] (2000) 795809

809

[15] B. Leclerc, J-Y. Thibon, LittlewoodRichardson coefficients and KazhdanLusztig polynomials, math.QA/9809122.
[16] O. Foda, B. Leclerc, M. Okado, J-Y. Thibon, Ribbon tableaux and q-analogues of fusion rules
in WZW conformal field theories, math.QA/9810008.
[17] A. Schilling, M. Shimozono, Bosonic formula for level-restricted paths, math.QA/9812106.
[18] G. Lusztig, Astrisque 101102 (1983) 208229.
[19] A.N. Kirillov, P. Mathieu, D. Snchal, M.A. Walton, in: Group-Theoretical Methods in
Physics, Proceedings of the XIXth International Colloquium, Salamanca, Spain, 1992, Vol. 1,
CIEMAT, Madrid, 1993.
[20] M.A. Walton, Can. J. Phys. 72 (1994) 527.
[21] J.-B. Zuber, Commun. Math. Phys. 179 (1996) 265.
[22] V.B. Petkova, J.-B. Zuber, Nucl. Phys. B 463 (1996) 161.
[23] W. Fulton, J. Harris, Representation Theory: A First Course, Springer-Verlag, 1991.
[24] I.G. Macdonald, Notes on Schubert Polynomials, Universit du Qubec Montral, 1991.
[25] V.G. Kac, Infinite-Dimensional Lie Algebras, 3rd edn., Cambridge Univ. Press, 1990.
[26] M.A. Walton, Phys. Lett. B 241 (1990) 365; Nucl. Phys. B 340 (1990) 777.
[27] P. Furlan, A. Ganchev, V. Petkova, Nucl. Phys. B 343 (1990) 205.
[28] F. Goodman, H. Wenzl, Adv. Math. 82 (1990) 244.
[29] D. Kazhdan, G. Lusztig, Invent. Math. 53 (1979) 165.
[30] A. Cappelli, C. Itzykson, J.-B. Zuber, Commun. Math. Phys. 113 (1987) 1.
[31] M.A. Walton, J. Math. Phys. 39 (1998) 665.

Nuclear Physics B 584 [PM] (2000) 811814


www.elsevier.nl/locate/npe

CUMULATIVE AUTHOR INDEX B581B584

Achiman, Y.
Akemann, G.
Alvarez, O.
Alvarez, O.
Alvarez, O.
Andreev, O.
Antoniadis, I.
Arhrib, A.

B584 (2000) 46
B583 (2000) 739
B582 (2000) 139
B584 (2000) 659
B584 (2000) 682
B583 (2000) 145
B583 (2000) 35
B581 (2000) 34

Balog, J.
Br, O.
Barcel, C.
Barros, M.
Baulieu, L.
Blanger, G.
Benakli, K.
Bertola, M.
Bertolini, M.
Bialas, P.
Bialas, P.
Bianchi, M.
Bilal, A.
Blanchard, P.
Blumenhagen, R.
Blmlein, J.
Bohm, A.R.
Boudjema, F.
Brignole, A.
Bronnikov, K.A.
Bros, J.
Bruckmann, F.
Buchbinder, I.L.

B583 (2000) 614


B581 (2000) 499
B584 (2000) 415
B584 (2000) 719
B581 (2000) 604
B581 (2000) 3
B583 (2000) 35
B581 (2000) 575
B582 (2000) 393
B581 (2000) 477
B583 (2000) 368
B584 (2000) 216
B582 (2000) 65
B583 (2000) 368
B582 (2000) 44
B581 (2000) 449
B581 (2000) 91
B581 (2000) 3
B582 (2000) 759
B584 (2000) 436
B581 (2000) 575
B584 (2000) 589
B584 (2000) 615

Caffo, M.
Campbell, B.A.
Campos, I.
Capdequi Peyranre, M.
Capitani, S.
Carloni Calame, C.M.
Carter, G.W.
Casas, J.A.
Castellani, C.
Castro Alvaredo, O.A.

B581 (2000) 274


B581 (2000) 240
B581 (2000) 499
B581 (2000) 34
B582 (2000) 762
B584 (2000) 459
B582 (2000) 571
B581 (2000) 61
B583 (2000) 542
B581 (2000) 643

Chakrabarti, S.K.
Chan, Ch.S.
Chetyrkin, K.G.
Chu, C.-S.
Cirillo, E.N.M.
Coste, A.
Cox, J.
Cski, C.
Cski, C.
Cvetic, M.
Czyz, H.

B582 (2000) 627


B581 (2000) 156
B583 (2000) 3
B582 (2000) 65
B583 (2000) 584
B581 (2000) 679
B583 (2000) 331
B581 (2000) 309
B584 (2000) 359
B584 (2000) 149
B581 (2000) 274

Damgaard, P.H.
de Azcrraga, J.A.
Delfino, G.
Denef, F.
Derendinger, J.-P.
Derkachov, S.
Diakonov, D.
Daz, M.A.
Di Clemente, V.
Donato, F.
Dorn, H.
Dubin, A.Yu.
Dubin, A.Yu.

B583 (2000) 347


B581 (2000) 743
B583 (2000) 597
B581 (2000) 135
B582 (2000) 231
B583 (2000) 691
B582 (2000) 571
B583 (2000) 182
B581 (2000) 61
B581 (2000) 3
B583 (2000) 145
B582 (2000) 677
B584 (2000) 749

Eden, B.
Einhorn, M.B.
Enqvist, K.
Epele, L.N.
Erickson, J.K.
Erlich, J.
Erlich, J.
Evans, N.
Eyras, E.

B581 (2000) 523


B582 (2000) 216
B582 (2000) 763
B583 (2000) 454
B582 (2000) 155
B581 (2000) 309
B584 (2000) 359
B581 (2000) 391
B584 (2000) 251

Fabrizio, M.
Fanchiotti, H.
Ferrndez, A.
Ferreira, C.N.
Feruglio, F.
Finster, F.
Fortunato, S.

B583 (2000) 542


B583 (2000) 454
B584 (2000) 719
B581 (2000) 165
B582 (2000) 759
B584 (2000) 387
B583 (2000) 368

812

Nuclear Physics B 584 [PM] (2000) 811814

Fosco, C.D.
Fries, R.J.
Frhlich, J.

B582 (2000) 716


B582 (2000) 537
B583 (2000) 381

Gandolfo, D.
Gannon, T.
Garca Canal, C.A.
Garousi, M.R.
Gates Jr., S.J.
Geyer, B.
Ghoshal, D.
Gitman, D.M.
Godbole, R.
Gonnella, G.
Gonzlez-Sprinberg, G.A.
Gorini, V.
Grlich, L.
Gorsky, A.
Govindarajan, T.R.
Grandjean, O.
Greene, B.R.
Grojean, C.
Grzadkowski, B.
Guan, X.-W.
Guimares, M.E.X.
Gukov, S.
Gukov, S.
Gunion, J.F.
Guruswamy, S.

B583 (2000) 368


B581 (2000) 679
B583 (2000) 454
B584 (2000) 284
B584 (2000) 109
B581 (2000) 341
B584 (2000) 300
B584 (2000) 615
B581 (2000) 3
B583 (2000) 584
B582 (2000) 3
B581 (2000) 575
B582 (2000) 44
B584 (2000) 197
B583 (2000) 291
B583 (2000) 381
B584 (2000) 480
B584 (2000) 359
B583 (2000) 49
B583 (2000) 721
B581 (2000) 165
B584 (2000) 69
B584 (2000) 109
B583 (2000) 49
B583 (2000) 475

Harshman, N.L.
Hassan, S.F.
Heinzl, T.
Helayl-Neto, J.A.
Heller, U.M.
Herdeiro, C.A.R.
Hikami, K.
Hisano, J.
Holland, K.
Hollik, W.
Hollowood, T.J.
Hollowood, T.J.
Hong, D.K.
Hormuzdiar, J.
Howe, P.S.
Hsu, S.D.H.
Huiszoon, L.R.
Hull, C.M.

B581 (2000) 91
B583 (2000) 431
B584 (2000) 589
B581 (2000) 165
B583 (2000) 347
B582 (2000) 363
B581 (2000) 761
B584 (2000) 3
B583 (2000) 331
B581 (2000) 34
B581 (2000) 309
B584 (2000) 359
B582 (2000) 451
B581 (2000) 391
B581 (2000) 523
B581 (2000) 391
B584 (2000) 705
B583 (2000) 237

Inoue, R.
Intriligator, K.
Irvine, S.E.
Ishibashi, N.
Iso, S.

B581 (2000) 761


B581 (2000) 257
B584 (2000) 795
B583 (2000) 159
B583 (2000) 159

Jurco, B.

B584 (2000) 784

Kaminsky, K.
Karakhanyan, D.
Kaste, P.
Kausch, H.G.
Kawai, H.
Kaya, A.
Kazama, Y.
Ketov, S.V.
Ketov, S.V.
Kikukawa, Y.
Kim, J.E.
Kimura, Y.
Kinar, Y.
Kirschner, R.
Kiselev, V.V.
Kitazawa, Y.
Kitazawa, Y.
Kniehl, B.A.
Kogan, I.I.
Kogut, J.B.
Krs, B.
Kotikov, A.V.
Kovacs, S.
Kramer, G.
Krauth, W.
Krykhtin, V.A.
Kurosawa, K.
Kyae, B.

B581 (2000) 240


B583 (2000) 691
B582 (2000) 203
B583 (2000) 513
B583 (2000) 159
B583 (2000) 411
B584 (2000) 171
B582 (2000) 95
B582 (2000) 119
B584 (2000) 511
B582 (2000) 296
B581 (2000) 295
B583 (2000) 76
B583 (2000) 691
B581 (2000) 432
B581 (2000) 295
B583 (2000) 159
B582 (2000) 514
B584 (2000) 313
B582 (2000) 477
B582 (2000) 44
B582 (2000) 19
B584 (2000) 216
B582 (2000) 514
B584 (2000) 641
B584 (2000) 615
B584 (2000) 3
B582 (2000) 296

Laine, M.
Lazar, M.
LeClair, A.
Lee, H.M.
Lerche, W.
Lesgourgues, J.
Li, T.
Lipatov, L.N.
Lpez, A.
Losada, M.
L, H.
Lucas, P.
Ludwig, A.W.W.
Lunardini, C.
Lunardini, C.
Lscher, M.
Lst, D.
Ltken, C.A.

B582 (2000) 277


B581 (2000) 341
B583 (2000) 475
B582 (2000) 296
B582 (2000) 203
B582 (2000) 593
B582 (2000) 176
B582 (2000) 19
B582 (2000) 716
B582 (2000) 277
B584 (2000) 149
B584 (2000) 719
B583 (2000) 475
B583 (2000) 260
B584 (2000) 459
B582 (2000) 762
B582 (2000) 44
B582 (2000) 203

Macfarlane, A.J.
Mangano, M.L.
Martins, M.J.
Marucho, M.
McDonald, J.
Melnikov, V.N.
Merten, C.

B581 (2000) 743


B582 (2000) 759
B583 (2000) 721
B583 (2000) 454
B582 (2000) 763
B584 (2000) 436
B584 (2000) 46

Nuclear Physics B 584 [PM] (2000) 811814


Mikhailov, A.
Miramontes, J.L.
Moeller, N.
Montagna, G.
Montes, X.
Morel, A.
Moschella, U.
Moultaka, G.
Mouslopoulos, S.
Mukhopadhyay, B.
Mller, B.
Muramatsu, T.

B584 (2000) 545


B581 (2000) 643
B583 (2000) 105
B584 (2000) 459
B582 (2000) 259
B581 (2000) 477
B581 (2000) 575
B581 (2000) 34
B584 (2000) 313
B582 (2000) 627
B582 (2000) 537
B584 (2000) 171

Nakamura, A.
Nastase, H.
Nastase, H.
Nersesyan, A.A.
Niclasen, R.
Nicrosini, O.
Niedermaier, M.
Niedermayer, F.
Nomura, Y.

B584 (2000) 528


B581 (2000) 179
B583 (2000) 211
B583 (2000) 671
B583 (2000) 347
B584 (2000) 459
B583 (2000) 614
B583 (2000) 614
B584 (2000) 3

Oeckl, R.
Ohsawa, A.
Okawa, Y.
Onishchenko, A.I.

B581 (2000) 559


B581 (2000) 73
B584 (2000) 329
B581 (2000) 432

Panda, S.
Papazoglou, A.
Patrascioiu, A.
Paul, P.L.
Pelizzola, A.
Pernici, M.
Pershin, V.D.
Petersson, B.
Petrov, K.
Piccinini, F.
Pickering, A.
Pliszka, J.
Polyakov, A.
Pope, C.N.
Ptter, B.

B584 (2000) 251


B584 (2000) 313
B583 (2000) 614
B581 (2000) 156
B583 (2000) 584
B582 (2000) 733
B584 (2000) 615
B581 (2000) 477
B581 (2000) 477
B584 (2000) 459
B581 (2000) 523
B583 (2000) 49
B581 (2000) 116
B584 (2000) 149
B582 (2000) 514

Quirs, M.
Quirs, M.

B581 (2000) 61
B583 (2000) 35

Rasmussen, J.
Recknagel, A.
Reisz, T.
Remiddi, E.
Restrepo, D.A.
Rtey, A.
Robaschik, D.
Rosier-Lees, S.
Ross, G.G.

B582 (2000) 649


B583 (2000) 381
B581 (2000) 477
B581 (2000) 274
B583 (2000) 182
B583 (2000) 3
B581 (2000) 449
B581 (2000) 3
B584 (2000) 313

813

Rossi, G.
Ruelle, P.
Rummukainen, K.
Russo, R.
Rychkov, V.

B584 (2000) 216


B581 (2000) 679
B583 (2000) 347
B582 (2000) 65
B581 (2000) 116

Saito, T.
Sakai, S.
Santamaria, A.
Santiago, J.
Satz, H.
Sauser, R.
Schaeffer, R.
Schfer, A.
Schalm, K.
Schaposnik, F.A.
Scharnhorst, K.
Schellekens, A.N.
Schomerus, V.
Schreiber, E.
Schupp, P.
Schwetz, M.
Seiler, E.
Semenoff, G.W.
Sen, A.
Sevrin, A.
Shirman, Y.
Shiu, G.
Shore, G.M.
Smirnov, A.Yu.
Smoller, J.
Sokatchev, E.
Sommer, R.
Sonnenschein, J.
Stanev, Y.S.
Staudacher, M.
Stein, E.
Stephanov, M.A.
Suneeta, V.

B584 (2000) 528


B584 (2000) 528
B582 (2000) 3
B584 (2000) 313
B583 (2000) 368
B582 (2000) 231
B581 (2000) 575
B582 (2000) 537
B584 (2000) 480
B582 (2000) 716
B581 (2000) 718
B584 (2000) 705
B583 (2000) 381
B583 (2000) 76
B584 (2000) 784
B581 (2000) 391
B583 (2000) 614
B582 (2000) 155
B584 (2000) 300
B581 (2000) 135
B581 (2000) 309
B584 (2000) 480
B581 (2000) 409
B583 (2000) 260
B584 (2000) 387
B581 (2000) 523
B582 (2000) 762
B583 (2000) 76
B584 (2000) 216
B584 (2000) 641
B582 (2000) 537
B582 (2000) 477
B583 (2000) 291

Tamada, M.
Tanaka, T.
Taylor, W.
Terashima, S.
Tok, T.
Toublan, D.
Trentadue, L.
Trigiante, M.
Troost, J.
Tseytlin, A.A.

B581 (2000) 73
B582 (2000) 259
B583 (2000) 105
B584 (2000) 329
B584 (2000) 589
B582 (2000) 477
B583 (2000) 307
B582 (2000) 393
B581 (2000) 135
B584 (2000) 233

Vafa, C.
Vaidya, S.
Vainshtein, A.
Valle, J.W.F.
Vaman, D.

B584 (2000) 69
B583 (2000) 291
B584 (2000) 197
B583 (2000) 182
B581 (2000) 179

814

Nuclear Physics B 584 [PM] (2000) 811814

Vaman, D.
van Nieuwenhuizen, P.
Verbaarschot, J.J.M.
Verbeni, M.
Verlinde, H.
Vernizzi, G.
Vidal, J.
Visser, M.
Volkov, M.S.

B583 (2000) 211


B581 (2000) 179
B582 (2000) 477
B583 (2000) 307
B581 (2000) 156
B583 (2000) 739
B582 (2000) 3
B584 (2000) 415
B582 (2000) 313

Walcher, J.
Walton, M.A.
Wang, Y.-J.
Weiss, N.
Weisz, P.
Wess, J.
West, P.C.

B582 (2000) 203


B584 (2000) 795
B583 (2000) 671
B583 (2000) 76
B583 (2000) 614
B584 (2000) 784
B581 (2000) 523

White, B.E.
Wiedemann, U.A.
Wipf, A.
Wipf, A.
Witten, E.
Witten, E.
Wittig, H.

B581 (2000) 409


B582 (2000) 409
B582 (2000) 313
B584 (2000) 589
B584 (2000) 69
B584 (2000) 109
B582 (2000) 762

Yau, S.-T.
Yung, A.

B584 (2000) 387


B584 (2000) 197

Zarembo, K.
Zayas, L.A.P.
Zhitnitsky, A.
Zwanziger, D.
Zwirner, F.

B582 (2000) 155


B582 (2000) 216
B582 (2000) 477
B581 (2000) 604
B582 (2000) 759

You might also like