Electromagnetic Field Theory
Electromagnetic Field Theory
We analyze the spectra of non-chiral and chiral bifundamental mesons arising on intersect-
ing D7-branes in AdS5 × S 5 . In the absence of magnetic flux on the curve of intersection,
the spectrum is non-chiral, and the dual gauge theory is conformal in the quenched/probe
approximation. For this case we calculate the dimensions of the bifundamental mesonic op-
erators. We then consider magnetization of the D7-branes, which deforms the dual theory by
an irrelevant operator and renders the mesons chiral. The magnetic flux spoils the conformal-
ity of the dual theory, and induces a D3-brane charge that becomes large in the ultraviolet,
where the non-normalizable bifundamental modes are rapidly divergent. An ultraviolet com-
pletion is therefore necessary to calculate the correlation functions in the chiral case. On the
other hand, the normalizable modes are very well localized in the infrared, leading to new
possibilities for local model-building on intersecting D7-branes in warped geometries.
Contents
1 Introduction 1
5 Conclusions 22
B Hyperspherical Harmonics 27
1 Introduction
The AdS/CFT correspondence [1–4] is a powerful duality relating conformal field theories
(CFTs) in d dimensions to gravitational theories on (d + 1)-dimensional anti-de Sitter (AdS)
spaces. The extension of the duality to include global flavor groups has been well-studied
(see [5, 6] for some foundational work) and is well-motivated: it brings the theory closer
to phenomenologically viable models, with mesonic bound states serving as prototypes for
visible-sector fields. However, to find more realistic models, the flavor group must be ex-
tended to a product group, and the resulting mesonic spectrum must be made chiral. Such
extensions have been relatively unexplored, and in the present work we report on progress in
this direction.
When the gravity side of the duality is a type II string theory, flavor groups are added
through the introduction of higher-dimensional Dp-branes that fill AdS and wrap compact
cycles [6].1 The simplest such example is the addition of F D7-branes to type IIB string theory
on AdS5 × S 5 , where we take the D7-branes to fill AdS5 and wrap an S 3 of the S 5 . The
geometry is supported by N units of D3-brane charge and, without the D7-branes, is dual to
1
The higher-dimensional D-brane need not fill all of AdS; if the brane is characterized by a minimum
distance away from the origin of AdS then the dual quarks are massive [7].
–1–
N = 4 SU (N ) super Yang-Mills. Adding the D7-branes deforms the dual theory to an N = 2
gauge theory with a U (F ) flavor group, containing a massless adjoint hypermultiplet as well
as a massless quark hypermultiplet that transforms in the bifundamental of SU (N ) × U (F ).
The brane construction makes this clear, as the open string excitations of the D7-branes give
rise to a U (F ) gauge theory, and the infinite D7-brane worldvolume transverse to the D3-
branes results in a vanishing 4d coupling for this theory. Open strings stretching between the
D7-branes and the D3-branes have the same charges as the quarks in the dual theory. We
will work in the standard decoupling limit [1] in which one first takes
and then sends the ’t Hooft coupling λt to infinity. In this limit, the D3-branes are replaced by
their near-horizon backreaction, so that the only open strings are those stretching among the
D7-branes. These transform in the adjoint representation of U (F ) and are dual to mesonic
operators in the gauge theory.
D7-branes are codimension-two objects, and so their backreaction cannot generally be
neglected. Correspondingly, the presence of quarks in the dual gauge theory alters the renor-
malization group flow, which was trivial before the introduction of flavor. Fortunately, the
decoupling limit (1.1) simplifies the situation: if we hold fixed the number of flavors, F , while
taking the number of colors to be large, then one can consistently neglect the running of
quarks in loops. In the dual geometry, many aspects of the D7-brane backreaction scale as
F/N and so also vanish in this limit (see [8] and references therein). The flavored gauge the-
ory does have a Landau pole, and so the influence of the quarks on the renormalization group
flow cannot be neglected forever, but the scale at which the Landau pole appears grows expo-
nentially with N/F . This so-called quenched approximation, in which the running of quarks
in loops is neglected, is equivalent to the limit in which the D7-branes are taken as probes of
the dual geometry. In what follows, we will take this approximation without further apology.
The introduction of flavor branes opens up significant possibilities for model-building. Di-
mensional reduction along the angular directions provides a framework for Randall-Sundrum
constructions [9–12] wherein the Standard Model fields propagating in the bulk [13] descend
from the D7-brane fluctuations as in [14]. Upon compactification, the flavor group on the
D7-branes becomes a prototype for the Standard Model gauge group. Of course, the Stan-
dard Model gauge group is a product; a corresponding product flavor group results from
introducing two separate stacks of D7-branes. The bifundamental fields are then open strings
stretching between the stacks, and in order for some of the bifundamentals to be massless,
the stacks must intersect.
A further challenge is that the Standard Model spectrum is chiral. In the class of con-
structions considered here, chirality in the 4d theory can be induced by introducing magnetic
flux on the (noncompact) curve where the D7-branes intersect. Upon compactification to 4d,
the zero modes of the Dirac operator acquire a net chirality set by the amount of quantized
magnetic flux.
–2–
Yet another difficulty in embedding fully realistic theories into warped backgrounds of
string theory is the fact that the Standard Model is not a supersymmetric theory. In ge-
ometries that are characterized by a finite infrared scale, such as the well-studied Klebanov-
Strassler solution [15], supersymmetry can be broken in a controllable way by the addition
of a small number of anti-D3-branes [16]. The resulting geometry [17, 18] corresponds to the
spontaneous breaking of supersymmetry in the dual field theory [19].2 An alternative is to con-
sider “gluing” the warped geometry to a compact space that does not preserve supersymmetry.
The dual field theory is then a non-supersymmetric theory with emergent supersymmetry,
as in [20]. Although non-supersymmetric constructions are difficult to control, the filtering
provided by the renormalization group means that the influence of the non-supersymmetric
bulk, including the effects of moduli stabilization, can be systematically parameterized and
incorporated along the lines of [21]. No matter which supersymmetry-breaking mechanism
is used,3 the resulting geometry is considerably more complex after supersymmetry is bro-
ken. We will therefore, in this initial work, focus on supersymmetric D7-brane probes of
supersymmetric backgrounds.4
In this note, we will consider the non-chiral and chiral bifundamental modes existing
at the intersections of probe D7-branes in AdS5 × S 5 . We build up to the chiral, warped
case through the simpler example of intersecting D7-branes in flat space (§2). Although the
flat-space analysis of §2 has appeared elsewhere in the literature (see e.g. [26–28]), a detailed
treatment is useful here, because the equations of motion are readily generalized from the
simple flat-space case to the AdS5 × S 5 configuration of primary interest.
The organization of this note is as follows. In §2 we begin with the simple case of
intersecting D7-branes in a flat space background. In §2.2 we compute the mass spectrum of
the bifundamental modes for the case of vanishing magnetic flux, where the spectrum is non-
chiral. Then, in §2.3 we calculate the chiral mass spectrum in a configuration with magnetic
flux. Next, in §3 we consider unmagnetized intersecting D7-branes in AdS5 × S 5 , computing
the scaling dimensions of vector-like bifundamental mesonic operators. Finally, in §4 we add
the simplest possible magnetization to the intersecting D7-branes in AdS5 × S 5 , and show
that this magnetization makes the calculation of correlation functions untrustworthy without
an ultraviolet completion. Concluding remarks are given in §5, while our conventions and a
few technical details appear in the appendices.
2
Some authors have interpreted the singularities of the anti-D3-brane geometry described in [17] as implying
that the supersymmetry-breaking state does not exist.
3
See [22] for other interesting proposals.
4
See, for example, [23–25] for analyses of probe D7-branes in non-supersymmetric backgrounds from the
worldvolume and/or worldsheet points of view.
–3–
2 D7-branes in Flat Space
As a warm-up to the case of strong warping, we will first review the case of intersecting
D7-branes probing unwarped flat space, R9,1 = R3,1 × C3 .
F2 = ˜∗4 F2 , (2.1)
where ˜∗4 is the Hodge star built from the metric on C2 . The condition (2.1) is equivalent to
F2 being (1, 1) and primitive with respect to the Kähler form induced on C2 .
To leading order in the α′ expansion, the action of the D7-brane in this background is [30]
1 1 1 αβ γδ
Z
8 α αβ i j D7 αβ
p
SD7 = − 2 d ξ −ĝ ĝij ĝ ∂α Φ ∂β Φ + ĝ ĝ Fαγ Fβδ + iΘ̄P− ĝ Γ̂α ∂β Θ , (2.2)
g8 W 2 4
in which we have omitted a constant term that does not play a role in our analysis. Writing
−1
the string tension as τF1 = 2πα′ = ℓ2s , the 8d Yang-Mills coupling is g8−2 = 8π 3 ℓ4s gs . Here
ξ α are coordinates on the D7-brane, ĝαβ is the induced worldvolume metric and ĝij is the
transverse metric. Θ is a 10d double Majorana-Weyl spinor (reviewed in Appendix A) that, as
in the Green-Schwarz superstring, redundantly encapsulates the fermionic degrees of freedom
of the D7-brane. In particular, Θ is subject to the κ-symmetry identification
Θ ∼ Θ + P−Dp κ, (2.3)
in which
1
ΓD7 = dvol ǫ̂α ···α Γ̂α1 ···α8 = −iΓ(8) ,
/ W := (2.5)
8! 1 8
where ǫα1 ···α8 is the antisymmetric tensor and Γ(8) is the SO (7, 1) chirality operator. We use
κ-symmetry to set !
θ
Θ= . (2.6)
0
5
Here and throughout we will use “Lorentz invariance” to refer to SO (3, 1) invariance.
–4–
With this choice,
1 1 1 αβ γδ i αβ
Z
8 α αβ i j
SD7 =− 2 d ξ ĝij ĝ ∂α Φ ∂β Φ + ĝ ĝ Fαγ Fβδ + θ̄ĝ Γ̂β ∂α θ , (2.7)
g8 W 2 4 2
which is the familiar action for maximally supersymmetric 8d U (1) gauge theory.
On a stack of F such D7-branes, the gauge group is enhanced to U (F ), and Aα , Φi , and
θ are promoted to adjoint-valued fields. The leading-order action is determined by gauge-
invariance and supersymmetry to be
1 1 1 1
Z
8 α
ĝij ĝαβ Dα Φi Dβ Φj + ĝαβ ĝ γδ Fαγ Fβδ − ĝij ĝkl Φi , Φk Φj , Φl
SD7 = − 2 d ξ tr
g8 W 2 4 4
i 1
+ θ̄ĝαβ Γ̂α Dβ θ − θ̄ Γ̂i Φi , θ ,
(2.8)
2 2
in which tr denotes a trace over gauge indices, Dα is a gauge covariant derivative
Dα = ∂α − i Aα , · , (2.9)
breaks U (F1 + F2 ) → U (F1 ) × U (F2 ) and describes a separation of the branes ∆xi =
X1i − X2i . The factor of ℓ2s is introduced so that Φi has length dimension −1. However,
this also has the effect of canceling the factors of ℓs that appear in operators correcting the
Yang-Mills action. Therefore, in order to trust this effective field theory, we consider cases
where ∆xi ≪ ℓs . Equivalently, if we are to trust the effective field theory description of the
modes stretching between the branes, their mass must be less than that of the massive string
states that have been integrated out implicitly.
Writing the fluctuations as !
φ i φi
δΦi = 1 +
, (2.11)
φi− φi2
φi1 and φi2 transform as adjoints under U (F1 ) and U (F2 ), respectively, while φi+ and φi− are
bifundamentals that acquire masses proportional to the separation. For notational simplicity,
–5–
in what follows we will consider the case F1 = F2 = 1, but all of our results generalize easily
to higher ranks.
If, instead of being parallel, the branes intersect, some of the bifundamental modes will
become massless. The intersection of two D7-branes is generically six-dimensional, and the
long-wavelength description of the bifundamental modes can be given in terms of a 6d effective
field theory description on this intersection. The 6d masses of the bifundamentals depend
on the angles formed by the intersection of the branes. However, the vector bifundamentals
never become massless, indicating that the 6d theory is a U (1) × U (1) (rather than the un-
Higgsed U (2)) gauge theory, and that fewer than sixteen supercharges are preserved, since
the vector multiplet is split. When the intersection is such that both D7-branes fill R3,1 and
are holomorphically embedded into C3 , at least minimal supersymmetry is preserved [31] and
the 6d theory includes massless scalars and fermions.
in which q = ℓ−2
s t. The bifundamental modes are localized on R
3,1 ×C, with z 1 the coordinate
on the curve of intersection (which in this case is simply C). For the reasons discussed above,
we must take t ≪ 1 in order to trust the effective field theory. Of course, no matter what
the value of t, at sufficiently large values of z 2 the branes will be far apart and so one might
worry about stringy corrections to the Yang-Mills action. That is, in addition to (2.8), the
worldvolume action contains, for example, operators with the schematic form
k−4
tz 2
ℓsk−4 (Φ)k ∼ ϕ4± + · · · , (2.14)
ℓs
which might seem to become important at z 2 ∼ t−1 ℓs . However, as we will show below, the
bifundamental modes are highly peaked at z 2 = 0, and so we anticipate that their physics
will be largely insensitive to the corrections at large z 2 .
6
Similar vevs were utilized in [26] to describe brane recombination from non-supersymmetric intersections.
–6–
The configuration just described is supersymmetric, so we can find solutions to the bosonic
equations of motion by solving the fermionic equations of motion. Although the intersection
is SO (5, 1) symmetric, in anticipation of the magnetization — which preserves only SO (3, 1)
and which we discuss below — we will make use of the decomposition SO (9, 1) → SO (3, 1) ×
SO (6), as discussed in Appendix A. It is useful to decompose the 10d fermionic mode θ into
modes of different internal chirality (i.e. SO (6) weights)
3
! !
X ξ † 0 ∗
θ= ψm ⊗ ηm − ψm 2 ξ ∗ ⊗ β̃6 ηm , (2.15)
m=0
0 σ
in which ξ is a fixed two-component spinor, ηm are the constant SO (6) positive chirality
spinors in (A.13), and β̃6 is the SO (6) Majorana matrix. Writing the U (1) potential as
A1 = Aµ dxµ + 2a=1 aa dz a + aā dz̄ a , ψ0 is the fermionic partner of Aµ , ψ1,2 are the partners
P
1,2
in which we have set the neutral fields ψm to zero since they are not the modes of interest.
The linearized equation of motion for the fermions in this background is
0 = Γ̂α ∂α θ − i Γ̂i Φi , θ ,
(2.17)
where the transverse fluctuations Φi are evaluated on their vev (2.13). From SO (3, 1) invari-
ance, we expect that the equation of motion for Aµ should decouple from those of the other
bosonic fields, at least for some gauge choice,7 and thus we can consistently take ψ0± , the
superpartner of A± µ , to vanish. When the 4d momentum is zero we have
The equations (2.18) also follow from the conditions for supersymmetry [27, 28, 32]. These
coupled first-order equations can be turned into largely decoupled second-order equations
by taking derivatives. For example, application of ∂1 to (2.18a) and substitution of (2.18c)
and (2.18d) yields
0 = ∂1 ∂¯1̄ ψ1± + ∂2 ∂¯2̄ ψ1± − q 2 z 2 ψ1± . (2.19a)
7
One such gauge choice is (2.18a) after simply replacing the fermionic fields with their bosonic partners.
See, for example, [28].
–7–
Similarly,
Using (2.18b), (2.19c) gives an equation for ψ3± alone. Writing ψ3± = z̄ 2̄ ψ± , we have
′′ 1 ′ m2
0 =ζ± + ζ± − 21 ζ± − 4λζ± , (2.22a)
r1 r1
′′ 1 ′ m2
0 =σ± − σ± − 22 σ± − 4q 2 r22 σ± + 4λσ± , (2.22b)
r2 r2
in which Lµν are the associated Laguerre polynomials, Iµ and Kµ are the modified Bessel
functions of the first and second kinds, and
λ = q 2n + |m2 | + 1 . (2.24)
Regularity of σ± requires that n is a non-negative integer. Some of these modes are plotted
in figures 1 and 2.
One may notice that the system (2.18) also admits a zero mode that depends only on r2 ,
2
ψ3± = e−qr2 . (2.25)
It is easy to confirm that this gives a solution to (2.20), but this solution is not normalizable
with respect to the norm defined by treating (2.20) as a Sturm-Liouville problem. This is a
consequence of the fact that the bifundamental modes are more properly encoded by linear
combinations of the ψm ± rather than by the ψ ± themselves [26]. Correspondingly, the measure
m
used in integrating over the z 2 and z̄ 2̄ directions is not that defined by (2.20) (see [28]).
However, aside from this zero mode, the above equations successfully reproduce the spectrum
of 6d masses (2.24).
–8–
σ±
1.0
0.5
√
r2 / q
1 2 3 4
PSfrag replacements
-0.5
Figure 1. Transverse profiles for the flat space vector-like bifundamental modes σ± given by (2.23)
for m2 = 0 and n = 0 (the curve with smallest value at r2 = 0) through n = 4 (the curve with the
largest value at r2 = 0). The solutions have been normalized to the same value using the inner product
R 2
dr f r2 g r2 .
σ±
0.6
0.4
0.2
√
r2 / q
1 2 3 4
PSfrag replacements-0.2
-0.4
–9–
2.3 Chiral modes
We now consider magnetized intersections since, upon compactification, such a construction
gives a chiral 4d theory. We will again focus on supersymmetric configurations, which implies
that F2 must be (1, 1) and primitive. Consider first a single D7-brane on R3,1 × C2 . The most
general (1, 1) flux that can be supported by the D7-brane is
i i i i
F2 = − f1 dz 1 ∧ dz̄ 1̄ − f2 dz 2 ∧ dz̄ 2̄ − g1 dz 1 ∧ dz̄ 2̄ − g2 dz̄ 1̄ ∧ dz 2 . (2.26)
2 2 2 2
The 11̄ component will describe the magnetization of the intersection, and so we will look for
the simplest configurations with f1 6= 0. The Kähler form on C2 is simply
2
iX I ¯
J =− dz ∧ dz̄ I , (2.27)
2
I=1
and so primitivity imposes f1 = −f2 . The Bianchi identity implies that f1 is harmonic,
In the absence of sources, (2.28) requires that f1 is constant. We can then consistently set
g1 = g2 = 0 and obtain the supersymmetric magnetization
F2 = −i M dz 1 ∧ dz̄ 1̄ − dz 2 ∧ dz̄ 2̄ .
(2.29)
0 = Γ̂α Dα θ − i Γ̂i Φi , θ ,
(2.33)
– 10 –
where Dα is the gauge-covariant derivative. Following the same decomposition and procedure
as for the vector-like case, we again find (2.18) up to the replacements
±
→ ∂1 ± M z̄ 1̄ ψm ∂¯1̄ ψm → ∂¯1̄ ∓ M z 1 ψm
± ±
±
∂1 ψm , ,
± 2̄
∂¯2̄ ψ → ∂¯2̄ ± M z ψ .
± ± 2
±
∂2 ψ → ∂2 ∓ M z̄ ψ ,
m m m (2.34) m
Due to the self-duality of the magnetic flux, (2.35) is separable. Again using the polar
decomposition z a = ra eiφa and taking the ansatz (2.21), we find the equations
′′ 1 ′ m2
ζ± − 21 ζ± − 4M 2 r12 ζ± + −4λ ± 4M m1 ζ± ,
0 =ζ± + (2.36)
r1 r1
′′ 1 ′ m2
− 22 σ± − 4κ2 r22 σ± + 4λ ∓ 4M m2 σ± ,
0 =σ± + σ± (2.37)
r2 r2
in which p
κ= M 2 + q2 , (2.38)
and λ is again a constant to be determined by boundary conditions. The solutions are
2 |m1 |/2
−M r2 2
2
ζ± r1 =e 1 2M r1 M α; m1 + 1; 2M r1 + U α; m1 + 1; 2M r1
2 |m |/2
σ± r2 =e−κr2 2κr22 2 L|m 2|
2κr22 ,
n2 (2.39)
with
λ = κ 2n2 + |m2 | + 1 ± M m2 ≡ M 2α − m1 ± m1 − 1 , (2.40)
where the final relation defines α. In (2.39), M and U are the confluent hypergeometric
functions of the first and second kinds,8 and regularity requires that n2 be a non-negative
integer.
The chirality of the spectrum is a consequence of the different behavior of the different
charges. It is most easily seen by considering the “missing” zero mode [28, 33]
2 2
ψ3± ∼ e−κr2 e∓M r1 h z 1 ,
(2.41)
where h is a holomorphic function of z 1 . Since we have taken M > 0, only the + sector gives
rise to normalizable modes, and hence the spectrum is chiral. The fact that h is an arbitrary
holomorphic function indicates that there are an infinite number of such chiral modes, as is
R
consistent with the fact that the chiral index, which is proportional to F2 , is divergent. Upon
compactification, further conditions are imposed on h z 1 (see e.g. [33]) and the spectrum
becomes finite.
8
Since the confluent hypergeometric function 1 F1 (a; b; z) is not defined when b = 0, −1, −2, . . ., we use the
regularized version M (a; b; z) = 1 F1 (a; b; z) /Γ (b).
– 11 –
3 Non-chiral Mesons from D7-branes in AdS
– 12 –
is the chirality operator on S 4 .
In the non-Abelian case, closed-string fields like the warp factor are interpreted as Taylor
series in the adjoint-valued transverse deformations, and thus the closed-string fields are
themselves adjoint-valued [35]. However, as in the unwarped case, this fact is only important
for higher-dimension operators, and can be neglected to leading order in ℓs . Similar terms
are expected in the non-Abelian fermionic action, but have not been computed explicitly.
However, to leading order in ℓs , supersymmetry and gauge-invariance require that the action
take the form [28, 36]
i 1 α 1
Z
F 8 α α
p i
SD7 = − 2 d ξ −ĝ tr θ̄ Γ̂ Dα θ + θ̄ Γ̂ ∂α A 1 + 2Γ̂S 4 θ − θ̄ Γ̂i Φ , θ . (3.8)
2g8 2 2
The intersection of two D7-branes satisfying
is described by !
ℓ−2
s µ + qz
2
Φ= . (3.10)
ℓ−2 2
s µ − qz .
When µ = 0, the D7-branes reach the origin of warping and the dual quarks are massless:
in the D-brane picture, the D3-branes and D7-branes intersect and the strings stretching
between them have zero length. However, when there is a finite separation between the
branes, the quarks have a mass proportional to µ. Consequently, the mesonic spectrum
becomes gapped [7]. The warp factor is to be evaluated at this vev, but so long as t is
sufficiently small, on the D7-brane we can take
2 2
1 z1 + z1 + µ2
A = log . (3.11)
2 L2
Decomposing θ as (2.15) and matching terms of internal chirality, we find
1 1
0 = ∂¯1̄ − ∂¯1̄ A ψ1± − ∂¯2̄ − ∂¯2̄ A ψ2± ∓ iqe−2A z 2 ψ3± ,
(3.12a)
2 2
3
0 = ∂2 + ∂2 A ψ3 ∓ iqe−2A z̄ 2̄ ψ2± ,
±
(3.12b)
2
3
0 = ∂1 + ∂1 A ψ3± ± iqe−2A z̄ 2̄ ψ1± ,
(3.12c)
2
1 1
0 = ∂1 − ∂1 A ψ2± + ∂2 − ∂2 A ψ1± ,
(3.12d)
2 2
where, as in the flat space analysis of §2, we have evaluated the equations at zero 4d momen-
tum and have set ψ0± = 0. Taking, as in [34]
±
ψ1,2 = eA/2 ϕ±
1,2 , ψ3± = e−3A/2 ϕ±
3, (3.13)
– 13 –
Since the warp factor depends on both z 1 and z 2 , (3.14) is not separable in those variables.
However, writing
z 1 = r cos β eiφ1 , z 2 = r sin β eiφ2 , (3.15)
equation (3.14) becomes
1 ˘ 2 4q 2 r 2 L4 sin2 β
2 3
0 = ∂r + ∂r + 2 ∇ − ϕ± , (3.16)
r r (r 2 + µ2 )2
in which
˘ 2 = ∂ 2 + cot β − tan β ∂β + 1 1
∂φ21 + ∂2
∇ β (3.17)
2
cos β sin2 β φ2
is the Laplacian on a unit S 3 (see Appendix B).
When µ = 0, (3.16) is completely separable. Indeed, taking
2m21 2m22
0 = 4 1 − x2 Q′′± − 8xQ′± − Q± − 2ξ 2 1 − x Q± + λQ± ,
Q± − (3.20)
1+x 1−x
in which x = cos 2β,
1 2
ξ 2 ≡ q 2 L4 = t gs N, (3.21)
π
and λ is a constant to be determined by boundary conditions.9
(a,b)
where Pn are the Jacobi Polynomials, c is the normalization constant (B.12), λ = ℓ (ℓ + 2),
and the quantum numbers must satisfy the inequalities 0 ≤ |m1 |+|m2 | ≤ ℓ and the constraint
1
2 (ℓ − m1 − m2 ) ∈ Z.
We have been unable to find analytic solutions to (3.20) when ξ 6= 0. However, since (3.20)
is an ordinary differential equation, numerical methods readily apply. We implement a spec-
tral method by expanding the unknown solution in terms of the spherical harmonics. The
9
Note that in this section and the next, λ carries no dimensions, in contrast to the previous section.
– 14 –
potential term proportional to ξ does not mix modes of different m1 and m2 , so we can
accomplish the spectral decomposition by writing
X
Q x = bℓ y ℓ x , (3.23)
ℓ
where yℓ are the solutions (3.22) and we have suppressed other indices. Equation (3.20) then
becomes
X
2
0= λ − ℓ (ℓ + 2) − 2ξ 1 − x bℓ yℓ . (3.24)
ℓ
and using the recursion relationship (B.13), we can re-express (3.24) as the matrix equation
2
0 = λ − ℓ (ℓ + 2) − 2ξ d0 bℓ + 2ξ 2 d− bℓ−2 + 2ξ 2 d+ bℓ+2 , (3.26)
with
m22 − m21
d0 = 1 + ,
ℓ (ℓ + 2)
r
1 (ℓ + m1 + m2 ) (ℓ − m1 − m2 ) (ℓ + m1 − m2 ) (ℓ − m1 + m2 )
d− = , (3.27)
2ℓ ℓ2 − 1
s
1 (ℓ + 2 + m1 + m2 ) (ℓ + 2 − m1 − m2 ) (ℓ + 2 + m1 − m2 ) (ℓ + 2 − m1 + m2 )
d+ = .
2 (ℓ + 2) (ℓ + 1) (ℓ + 3)
Note that even and odd ℓs do not mix, so that this effectively gives two independent matrix
equations where the matrices are each tridiagonal.
Solving (3.20) amounts to diagonalization of the matrix defined by (3.26). Unfortunately,
because this is an infinite-dimensional matrix, we cannot perform this diagonalization exactly.
However, to obtain an estimate of the spectrum, we can truncate the matrix to a finite
submatrix. A good rule of thumb in such problems is that including the first 2n modes
determines the first n eigenvalues to an accuracy of a few percent [37]. Accurate eigenvalues
will be robust against variations in n, and our strategy will be to increase the number of modes
included until the eigenvalues calculated in this way stabilize. The first few eigenvalues at
m1 = m2 = 0 resulting from this process are shown in figures 3 and 4. As ξ increases, the
wavefunctions become increasingly localized on the intersection at β = 0, as shown in figure 5.
Note that when ξ ≪ 1, (3.26) immediately yields the perturbative result
m22 − m21
2
λ ≈ ℓ (ℓ + 2) + 2ξ 1 + . (3.28)
ℓ (ℓ + 2)
– 15 –
λ
3000
2500
2000
1500
1000
500
PSfrag replacements
0 ξ
0 10 20 30 40 50
Figure 3. The first few eigenvalues of (3.20) found via spectral methods, for m1 = m2 = 0. The
growth continues to be linear as ξ increases.
At large ξ, ℓ is no longer a good quantum number, as the intersection badly breaks the rota-
tional symmetry of the S 3 . Correspondingly, the solutions to (3.20) are linear combinations
of many different spherical harmonics. However, m1 and m2 remain good quantum numbers,
and so we find it more natural to write the spectrum as
λ ≈ 4ξ n + |m2 | + 1 , (3.30)
We can compare the solution (3.31) to the well-known result for a canonically normalized
scalar at zero momentum,
ϕ = ϕ0 r ∆−4 + ϕ1 r −∆ . (3.32)
– 16 –
λ
ô
ôô
ô ôô
30 000 ô
ôô
ô
ô
ô
ô
ô
ô
25 000 ô
ô
ô
ô òòò
ô
ô òòò
ô òò
20 000 ô ò òò
ô òò
ô
ô òò
ô
ô ò òò
ô òò
15 000 ô òò ìì
ô
ô
ò òò ìì
ô òò ìì
ô ò ì ìì
ô ò ìì
ô òò ììì à
10 000 ô òò ììì ààæ
ô ìììì àà
ô ò ò
ì ì ì ì ì
à ààææææ
ô
ô òò ììì
ì àà
àà æææ
æ
ô òò ìì àà æ
ô òò ì ìì à àààææææ
PSfrag replacements 5000 ô ò
ô ò
ô ò ìì
ìì
ì ì
àààà
àà
ààà ææææ
à ææ
ô ò òììì ææ
ò ì à à à à à ààààà æææææ
òììà à à à à à à
ô æ
ææææ
ô
òì ààæææææææææææææ
òì
ô
ì
à
æ àæ
æ àæ æ ℓ
20 40 60 80 100
Figure 4. The spectrum for m1 = m2 = 0 (which requires that ℓ be even) for ξ = 0 (bottom), 25, 50,
75, and 100 (top).
3.0
2.5
2.0
1.5
1.0
0.5
PSfrag replacements
β/π
0.1 0.2 0.3 0.4 0.5
Figure 5. The lowest-lying solutions of (3.20) for ξ = 0, 2.5, 10, 50, 100. When ξ = 0, the solution is a
constant zero mode, but as ξ increases, the profile becomes increasingly peaked at β = 0, the location
of the intersection.
– 17 –
The solution (3.31) does not match the form (3.32), since the transverse deformations are not
canonically normalized (see (3.35)). Nevertheless, ∆ can be determined by taking the ratio
of the two terms in (3.31), and we find the result
√
∆=2+ 1 + λ. (3.33)
The fact that the radial modes are simply power laws is an indication that the dual
theory is conformal. Indeed, one can confirm that the µ = 0 configuration (3.9) respects the
supersymmetry generated by eight supercharges, four of which correspond to the generators of
superconformal transformations in the dual theory. Alternatively, when µ = 0, the vev (3.10)
corresponds to a strictly marginal deformation of the theory. To see this, it suffices to consider
the Abelian action (3.4) and examine only the action of the transverse scalars Φi . Using the
complexified field Φ and expanding in scalar spherical harmonics gives the 5d action
∞ 2
√
L ℓ (ℓ + 2) †
Z
∂m Φ†ℓ ∂n Φℓ
X
5 mn
S∼− d x −g g + Φℓ Φℓ . (3.35)
r2 r2
ℓ=0
L
Defining the canonically normalized scalars χℓ = r Φℓ gives
∞
√
ℓ (ℓ + 2) − 3 †
Z
∂m χ†ℓ ∂n χℓ
X
5 mn
S∼− d x −g g + χℓ χℓ . (3.36)
L2
ℓ=0
– 18 –
4 Chiral Mesons from D7-branes in AdS
Just as in the flat space case, we can induce chirality into the dual theory through the intro-
duction of a supersymmetric magnetic flux (2.29). However, this magnetic flux will respect
only four of the gravity supercharges, and the other four, corresponding to the superconformal
charges of the dual theory, will not be preserved. As we shall see, this change has important
physical consequences: the calculation of correlation functions will turn out to require coun-
terterms that are super-exponentially sensitive to the ultraviolet completion of the geometry.
At the same time, the magnetic flux induces a large amount of D3-brane charge, so that the
geometry must be sharply modified in the ultraviolet. In practical terms, this dependence on
the ultraviolet behavior presents an obstacle to the calculation of correlation functions. More
importantly, it signifies that the magnetization (2.29), and the corresponding appearance of
chiral mesons, entails a substantial change in the background.
P−D7 ǫ = 0, (4.1)
The bulk geometry respects the supersymmetry generated by a GKP-like Killing spinor [12],
which is independent of the Minkowski coordinates and annihilated by holomorphic γ-matrices.
Moreover, such a Killing spinor obeys (4.1) if F2 is (1, 1) and self-dual: the F / 2 term anni-
2 p
/
hilates the Killing spinor, and the 1 and F 2 terms together are canceled by det (g + F )
(which takes a simple form because F is self-dual). However, the bulk geometry also supports
Killing spinors that depend on the Minkowski coordinates in a particular way (see, e.g., [38]).
The existence of such spinors is a special feature of anti-de Sitter space, and the supersym-
metry transformations they induce are dual to superconformal transformations. Since the
special AdS Killing spinors are not preserved by the magnetized D7-brane configuration, we
anticipate that conformality will be lost in the dual theory, even in the probe approximation.
We can also understand the loss of conformality from another point of view. The mag-
netization that gives rise to chirality follows from the connection (2.30) which, using (3.15),
can be written as
A1 = M r 2 − cos2 β dφ1 + sin2 β dφ2 ,
(4.3)
in which we are still taking M > 0 for simplicity of presentation. Writing A1 = M r 2 ω, ω
satisfies the defining equation of a transverse vector spherical harmonic ̺, which takes the
– 19 –
general form
˘ 2 ̺θ = − ℓ ℓ + 2 − 1 ̺θ , ˘ θ ̺θ = 0, ˘ ϕ ̺ψ = ± ℓ + 1 ğ θψ ̺ψ ,
ǫ̆θϕψ ∇
∇ ∇ (4.4)
in which ğ is the metric (B.2) on the unit S 3 , ∇˘ is the associated Levi-Civita connection, and
ǫ̆ is the associated volume form. The mode ω corresponds to the specific case ℓ = 1, with the
positive sign taken in the third equation in (4.4).10 Thus, ω is a transverse vector spherical
harmonic [39], and upon dimensional reduction leads to a canonically normalized field with
mass m2 L2 = 12 (see [7]). This corresponds to an operator of dimension 6, and A1 involves
the non-normalizable solution. Hence, the introduction of the magnetic flux deforms the dual
theory by an irrelevant operator. This implies that not only is conformality lost in the dual
theory, but the theory does not even flow from an ultraviolet fixed point.
The addition of this flux modifies the zero-momentum equations to
1 1
0 = ∂¯1̄ ∓ M z 1 − ∂¯1̄ A ψ1± − ∂¯2̄ ± M z 2 − ∂¯2̄ A ψ2± ∓ iqe−2A z 2 ψ3± ,
(4.5a)
2 2
3
0 = ∂2 ∓ M z̄ 2̄ + ∂2 A ψ3± ∓ iqe−2A z̄ 2̄ ψ2± ,
(4.5b)
2
3
0 = ∂1 ± M z̄ 1̄ + ∂1 A ψ3± ± iqe−2A z̄ 2̄ ψ1± ,
(4.5c)
2
1 1
0 = ∂1 ± M z̄ 1̄ − ∂1 A ψ2± + ∂2 ∓ M z̄ 2̄ − ∂2 A ψ1± .
(4.5d)
2 2
Using (3.13), we find
2 1 2
22
¯ ¯ 1̄ ¯ 1 2¯ 2 2 −4A 2
0 = ∂1 ∂1̄ + ∂2 ∂2̄ ± M z̄ ∂1̄ − z ∂1 − z̄ ∂2̄ + z ∂2 − M z − M + e q z ϕ± , (4.6)
1 3
where ϕ± = ϕ .
z̄ 2̄ ±
With the coordinates (3.15), this becomes
3
0 = ∂r2 + ∂r ± 4iM ∂φ1 − ∂φ2 − 4M 2 r 2
r
4q 2 r 2 L4 sin2 β
1 2 1 2 1 2
+ 2 ∂β + cot β − tan β ∂β + ∂ + ∂ − ϕ± . (4.7)
r cos2 β φ1 sin2 β φ2 (r 2 + µ2 )2
Again, the relative simplicity of this equation is a consequence of the self-duality constraint
imposed by supersymmetry. When µ = 0, the equation is again separable and it is useful
to take the ansatz (3.18). Q± satisfies the same eigenvalue problem (3.20) while the radial
equation is now
3 λ
0 = f±′′ + f±′ ∓ 4M m1 − m2 f± − 4M 2 r 2 f± − 2 f± .
(4.8)
r r
The solutions can be expressed in terms of M and U , the confluent hypergeometric functions
of the first and second kind,
2
f± = e−M r r −ν c1 M µ; ν; 2M r 2 + c2 U µ; ν; 2M r 2 ,
(4.9)
10
This sign is independent of the sign in the equation of motion for the bifundamental modes, (2.18).
– 20 –
in which
√ 1
ν =1+ 1+λ and µ=ν ± m1 − m2 . (4.10)
2
As anticipated, the solutions are not power laws, and so the dual field theory is no
longer conformal even in the probe approximation. Furthermore, noting that the dominant
asymptotic behavior at r → ∞ is
2
M µ; ν; 2M r 2 ∝ e2M r ,
(4.11)
where we have omitted power law factors, we find that the divergent part of (4.9) grows
super-exponentially at r → ∞:
2
f± ∝ eM r . (4.12)
An operator O on the field theory side has a corresponding classical field ϕ on the gravity
side. If O is a scalar field, then ϕ also transforms as an SO (3, 1) scalar. The solution for ϕ
at large r can be separated into a dominant term and a subdominant term,
If the geometry is asymptotically anti-de Sitter space, both the dominant and subdominant
terms are power laws at large r. Moreover, adom is dual to a source term for O, and correlation
functions of O are calculated by taking functional derivatives of Sgrav with respect to adom
and then later taking adom → 0.
For finite adom , the classical action Sgrav is divergent. This can be addressed by adding
counterterms to the action: one first regulates the action by cutting off the space at a large
but finite radius rΛ . The terms that diverge as rΛ → ∞ are canceled by adding terms to
the supergravity action that are localized on the boundary at rΛ . Taking rΛ → ∞ then
yields a finite action. The power law behavior of solutions in the AdS case means that such
counterterms have power-law (and potentially logarithmic) dependence on rΛ . However, the
super-exponential growth (4.12) of the chiral modes requires the introduction of counterterms
that have a similar super-exponential dependence on the cutoff. Since the magnetization re-
quired to induce chirality deforms the theory by an irrelevant operator, such strong sensitivity
is perhaps not surprising.
– 21 –
If the background remained unaltered by magnetization, the structure of counterterms
would represent a technically demanding but potentially surmountable challenge to calcu-
lating correlation functions.11 However, the chirality-inducing magnetic flux sources a large
R
amount of D3-brane charge via the Chern-Simons coupling C4 ∧ F2 ∧ F2 : the dissolved
D3-brane flux diverges as Z
F2 ∧ F2 ∼ M 2 ρ4 . (4.15)
R<ρ
N 1/4
ρ∼ , (4.16)
M 1/2
at which point the influence of this charge on the geometry must be taken into account. A
calculation of correlation functions that fails to incorporate this backreaction is not physically
meaningful.
One might ask whether a different choice of magnetization (still without a localized
source) results in a different conclusion. Supersymmetric fluxes supported on the D7-branes
are characterized by scalar hyperspherical harmonics — cf. (2.28) — and so the fluxes grow
as F2 ∼ r j+2 Ω(j) + r j+1 dr ∧ ω (j) , where j = 0, 1, 2, . . ., and Ω(j) and ω (j) are a 2-form and a
1-form on S 3 , respectively. Our analysis of the magnetic flux (2.31) corresponds to the case
j = 0. Other values of j would lead to steeper potentials in (4.6), and so to a greater degree
of localization of the bifundamental wavefunctions. However, the charge carried by such flux
diverges more quickly than (4.15), growing as r 2j+4 , and hence the problem of ultraviolet
sensitivity is exacerbated.
5 Conclusions
In this note we analyzed the spectrum of mesonic operators arising from strings stretching
between intersecting D7-branes in AdS5 × S 5 . The dual field theory is an N = 1 deformation
of maximally supersymmetric SU (N ) SYM, with the addition of a U (F1 ) × U (F2 ) flavor
group, under which the 7-7′ strings transform as bifundamentals.12 We considered D7-branes
with and without magnetic flux on the curve of intersection, finding sharply different results
in these two cases.
The intersection of the D7-branes corresponds to a particular adjoint Higgsing of the
U (2) theory arising on coincident D7-branes. In the field theory, the fact that the branes
intersect is described by a marginal deformation. If the D7-branes reach the origin of warping,
and one furthermore makes the quenched/probe approximation that neglects backreaction of
the D7-branes, then the dual theory is conformal. In this case — where magnetization has
11
For example, as developed in [41], it is possible to calculate correlation functions in the KT/KS theory [15,
42], even though the theory does not flow from an ultraviolet fixed point.
12
For notational simplicity only, we limited our discussion to the case F1 = F2 = 1, corresponding to a single
pair of D7-branes.
– 22 –
not yet been incorporated — we computed the spectrum of dual operators. The 7-7′ strings
are mixtures of the transverse deformations and the internal components of the gauge field,
and as a consequence the equations of motion are difficult to solve analytically. However,
conformal symmetry leads to a remarkable simplification of the equations of motion, through
which we were able to find numerical solutions. The behavior of the dimensions depends on
√
the value of ξ ∼ tan θ gs N , cf. (3.21), where θ is an angle characterizing the intersection.
Approximate spectra are given in (3.34). As expected, the modes are well localized along the
intersection of the D7-branes and have power-law behavior along the holographic direction.
We then considered introducing magnetic flux on the curve of intersection, leading to a
chiral spectrum in the dual theory. The simplest magnetization corresponds to an irrelevant
deformation of the theory, by an operator of dimension ∆ = 6. As a consequence, the
non-normalizable solutions to the bifundamental equations of motion have super-exponential
divergence in the ultraviolet, cf. (4.12). Although the limit (1.1) allows us to neglect the
backreaction of the D7-branes themselves, the backreaction of the D3-brane charge induced
by the magnetic flux cannot be neglected. Since the calculation of correlation functions,
for example through holographic renormalization, requires the use of the non-normalizable
modes, the procedure for calculating the correlation functions is unclear. This is a physical
limitation rather than a technical one: the divergence of the D3-brane charge induced by
magnetization of noncompact D7-branes signals the need for an ultraviolet completion via
compactification. In the dual language, the field theory describing magnetized D7-branes
does not flow from an ultraviolet fixed point.
On the other hand, we found that the normalizable modes of the chiral bifundamental
mesons are very well localized in the infrared. Indeed, at large r,
U µ; ν; 2M r 2 ∼ r −µ ,
(5.1)
where we have again omitted power law factors and have chosen M > 0. Although similar
Gaussian peaks appear in flat space (see e.g. [33]), this feature in warped space has the poten-
tial to provide a rich playground for model-building. In general, the lack of knowledge of the
metric and of related fields often stymies detailed model-building in string compactifications.
However, the metrics for infinite families of non-compact (and singular) Calabi-Yau cones are
known explicitly. These cones can be used to construct strongly warped geometries that can
be attached to compact spaces — see for example the discussion in [12]. Attachment to a
compactification modifies the solution in the cone region, by introducing sources for irrelevant
perturbations, but these effects can be incorporated systematically, as in [21]. One can there-
fore build a local model on D3-branes at the apex of the cone, but also take into account bulk
effects, including supersymmetry breaking and moduli stabilization. Constructions in this
corner of the landscape are limited to some degree by the possible singularities at the apex.
– 23 –
An alternative, toward which the present work is a modest advance, is to consider model-
building on intersecting magnetized D7-branes. Although the D7-branes will stretch beyond
the warped region into the bulk,13 we have demonstrated that at least some bifundamental
modes are well localized in the infrared. This allows for a combination of the richness of
model-building with intersecting D7-branes and the power of local model-building in warped
geometries. Although we limited our particular analysis to AdS5 × S 5 , the qualitative result
should extend to more general cones and their deformations (though the details, of course,
become much more complex).
This localization also implies that although correlation functions are difficult to describe,
the mass spectrum of mesons can in principle be calculated with reliable numerical techniques.
When the D7-branes move away from the center of AdS5 , the spectrum of mesons becomes
gapped even though, in the quenched approximation, the glueball spectrum is continuous [7].
A standard method of finding the meson mass spectrum in the gapped case is to calculate the
correlation functions and check for the appearance of poles. However, a practical alternative is
to find those solutions that satisfy appropriate infrared boundary conditions and are normal-
izable in the ultraviolet (see, for example, [7, 43]). Because the equation of motion constitutes
a Sturm-Liouville problem, this alternative approach leads to a discrete spectrum, and since
the solutions are expected to be exponentially convergent, the resulting spectrum would be
reliable. On the other hand, once the spectrum becomes gapped the radial and angular parts
of the equation of motion no longer separate, even in the unmagnetized case (3.16). This is
a significant complication, and so we leave this analysis to future work.
Yet another possibility is to consider alternative magnetizations. The magnetization
that we analyzed in this note is the simplest unsourced magnetic flux that is possible in our
construction, and other unsourced magnetic fluxes would enhance the bifundamental wave-
function localization that we found, while intensifying the problem of ultraviolet sensitivity.
Magnetic flux that is itself localized in the infrared, and produces only normalizable per-
turbations to the geometry, would require a local source. In particular, it was pointed out
in [24] and explicitly shown in [25] that the addition of anti-D3-branes to warped flux back-
grounds provides an infrared-localized magnetization. Although the resulting magnetization
has a gauge structure that differs from (2.31) — specifically, the induced magnetization is
proportional to the identity — this remains an intriguing possibility for future work.
Acknowledgments
It is a pleasure to thank F. Marchesano and G. Shiu for useful discussions of related topics.
This work was supported by the NSF under grant PHY-0757868.
13
Indeed, the consistency of embeddings in global models will provide constraints on which models can be
built.
– 24 –
A Conventions for Fermions
In this appendix we summarize our conventions for fermions, many of which follow from [44].
We work with a Weyl basis for the SO (9, 1) Γ-matrices and make use of the decomposition
SO (9, 1) → SO (3, 1) × SO (6). For SO (3, 1) we take
! !
I σ i
2
γ0 = , γ i=1,2,3 = , (A.1)
−I2 σi
γ̃ 4 =σ 1 ⊗ I2 ⊗ I2 , γ̃ 7 =σ 2 ⊗ I2 ⊗ I2 ,
γ̃ 5 =σ 3 ⊗ σ 1 ⊗ I2 , γ̃ 8 =σ 3 ⊗ σ 2 ⊗ I2 ,
γ̃ 6 =σ 3 ⊗ σ 3 ⊗ σ 1 , γ̃ 9 =σ 3 ⊗ σ 3 ⊗ σ 2 .
and so
γ̃(6) = −i γ̃ 1 · · · γ̃ 6 = σ 3 ⊗ σ 3 ⊗ σ 3 . (A.8)
The Majorana matrix is
β̃6 = γ̃ 7 γ̃ 8 γ̃ 9 = σ 2 ⊗ i σ 1 ⊗ σ 2 . (A.9)
– 25 –
We will make use of a complex structure
Defining
1 1
σ± = σ ± i σ2 ,
(A.11)
2
we have
γ̃ 1 =2 σ + ⊗ I2 ⊗ I2 , γ̃ 1̄ =2 σ − ⊗ I2 ⊗ I2 ,
γ̃ 2 =2 σ 3 ⊗ σ + ⊗ I2 , γ̃ 2̄ =2 σ 3 ⊗ σ − ⊗ I2 ,
γ̃ 3 =2 σ 3 ⊗ σ 3 ⊗ σ + , γ̃ 3̄ =2 σ 3 ⊗ σ 3 ⊗ σ − .
in which
ηǫ1 ǫ2 ǫ3 = ηǫ1 ⊗ ηǫ2 ⊗ ηǫ3 . (A.14)
Note that σ ± η± = 0, so that η+++ is annihilated by all contravariant holomorphic γ̃-matrices.
Finally, we construct the SO (9, 1) Γ-matrices by
– 26 –
We will also make use of double spinors built from pairs of 10d Majorana-Weyl spinors
!
θ1
Θ= , (A.19)
θ2
while explicit Pauli matrices act to mix the elements of the double spinor. For example,
! !
θ 1 θ 2
σ1 = . (A.21)
θ2 θ1
B Hyperspherical Harmonics
The volume of S 3 is
Z 2π Z 2π Z π/2
VS 3 = dφ1 dφ2 dβ sin β cos β = 2π 2 . (B.3)
0 0 0
(B.7)
– 27 –
1
in which r = 2 (ℓ − m1 − m2 ). For these to be non-vanishing regular solutions, r must be an
integer and
0 ≤ m1 + m2 ≤ ℓ. (B.8)
These solutions satisfy
˘ 2 Y = −ℓ (ℓ + 2) Y,
∇ (B.9)
and the condition (B.8) gives the expected degeneracy of (ℓ + 1)2 (see, for example, [39]).
The Jacobi polynomials are orthogonal in the sense that
Z 1
a b 2a+b+1 (a + r)! (b + r)!
dx 1 − x 1 + x Pr(a,b) Ps(a,b) = δrs . (B.10)
−1 2r + a + b + 1 r! (a + b + r)!
is satisfied by taking
s 1
(ℓ + m1 + m2 ) ! 12 (ℓ − m1 − m2 ) !
1 ℓ+1 2
cℓ,m1 ,m2 = . (B.12)
2m1 +m2 +1 21 (ℓ + m1 − m2 ) ! 12 (ℓ − m1 + m2 ) !
π
2 (a + r) (b + r) (a,b)
x Pr(a,b) (x) = P (x)
(a + b + 2r) (a + b + 2r + 1) r−1
2 (r + 1) (a + b + r + 1) (a,b) b2 − a2
+ Pr+1 (x) + P (a,b) (x) .
(a + b + 2r + 1) (a + b + 2r + 2) (a + b + 2r) (a + b + 2r + 2) r
(B.13)
References
[1] J. M. Maldacena, “The Large N Limit of Superconformal Field Theories and Supergravity”,
Adv.Theor.Math.Phys. 2 (1998) 231–252, [hep-th/9711200].
[2] E. Witten, “Anti de Sitter Space and Holography”, Adv.Theor.Math.Phys. 2 (1998) 253–291,
[hep-th/9802150].
[3] S. Gubser, I. R. Klebanov, and A. M. Polyakov, “Gauge Theory Correlators from Non-Critical
String Theory”, Phys.Lett. B428 (1998) 105–114, [hep-th/9802109].
[4] O. Aharony, S. S. Gubser, J. M. Maldacena, H. Ooguri, and Y. Oz, “Large N Field Theories,
String Theory and Gravity”, Phys.Rept. 323 (2000) 183–386, [hep-th/9905111].
[5] O. Aharony, A. Fayyazuddin, and J. M. Maldacena, “The Large N Limit of N = 2, 1 Field
Theories from Threebranes in F-theory”, JHEP 9807 (1998) 013, [hep-th/9806159].
[6] A. Karch and E. Katz, “Adding Flavor to AdS/CFT”, JHEP 0206 (2002) 043,
[hep-th/0205236].
– 28 –
[7] M. Kruczenski, D. Mateos, R. C. Myers, and D. J. Winters, “Meson Spectroscopy in AdS/CFT
with Flavor”, JHEP 0307 (2003) 049, [hep-th/0304032].
[8] C. Núñez, A. Paredes, and A. V. Ramallo, “Unquenched Flavor in the Gauge/Gravity
Correspondence”, Adv.High Energy Phys. 2010 (2010) 196714, [arXiv:1002.1088].
[9] L. Randall and R. Sundrum, “A Large Mass Hierarchy from a Small Extra Dimension”,
Phys.Rev.Lett. 83 (1999) 3370–3373, [hep-ph/9905221].
[10] H. L. Verlinde, “Holography and Compactification”, Nucl.Phys. B580 (2000) 264–274,
[hep-th/9906182].
[11] B. R. Greene, K. Schalm, and G. Shiu, “Warped Compactifications in M and F Theory”,
Nucl.Phys. B584 (2000) 480–508, [hep-th/0004103].
[12] S. B. Giddings, S. Kachru, and J. Polchinski, “Hierarchies from Fluxes in String
Compactifications”, Phys.Rev. D66 (2002) 106006, [hep-th/0105097].
[13] H. Davoudiasl, J. Hewett, and T. Rizzo, “Bulk Gauge Fields in the Randall-Sundrum Model”,
Phys.Lett. B473 (2000) 43–49, [hep-ph/9911262].
A. Pomarol, “Gauge Bosons in a Five-Dimensional Theory with Localized Gravity”, Phys.Lett.
B486 (2000) 153–157, [hep-ph/9911294].
S. Chang, J. Hisano, H. Nakano, N. Okada, and M. Yamaguchi, “Bulk standard model in the
Randall-Sundrum Background”, Phys.Rev. D62 (2000) 084025, [hep-ph/9912498].
Y. Grossman and M. Neubert, “Neutrino Masses and Mixings in Nonfactorizable Geometry”,
Phys.Lett. B474 (2000) 361–371, [hep-ph/9912408].
T. Gherghetta and A. Pomarol, “Bulk Fields and Supersymmetry in a Slice of AdS”,
Nucl.Phys. B586 (2000) 141–162, [hep-ph/0003129].
[14] T. Gherghetta and J. Giedt, “Bulk Fields in AdS5 from Probe D7 Branes”, Phys.Rev. D74
(2006) 066007, [hep-th/0605212].
[15] I. R. Klebanov and M. J. Strassler, “Supergravity and a Confining Gauge Theory: Duality
Cascades and χSB Resolution of Naked Singularities”, JHEP 0008 (2000) 052,
[hep-th/0007191].
[16] S. Kachru, R. Kallosh, A. D. Linde, and S. P. Trivedi, “de Sitter Vacua in String Theory”,
Phys.Rev. D68 (2003) 046005, [hep-th/0301240].
[17] I. Bena, M. Graña, and N. Halmagyi, “On the Existence of Meta-stable Vacua in
Klebanov-Strassler”, JHEP 1009 (2010) 087, [arXiv:0912.3519].
[18] O. DeWolfe, S. Kachru, and M. Mulligan, “A Gravity Dual of Metastable Dynamical
Supersymmetry Breaking”, Phys.Rev. D77 (2008) 065011, [arXiv:0801.1520].
P. McGuirk, G. Shiu, and Y. Sumitomo, “Non-Supersymmetric Infrared Perturbations to the
Warped Deformed Conifold”, Nucl.Phys. B842 (2011) 383–413, [arXiv:0910.4581].
A. Dymarsky, “On Gravity Dual of a Metastable Vacuum in Klebanov-Strassler Theory”, JHEP
1105 (2011) 053, [arXiv:1102.1734].
I. Bena, G. Giecold, M. Graña, N. Halmagyi, and S. Massai, “On Metastable Vacua and the
Warped Deformed Conifold: Analytic Results”, Class.Quant.Grav. 30 (2013) 015003,
[arXiv:1102.2403].
I. Bena, G. Giecold, M. Graña, N. Halmagyi, and S. Massai, “The Backreaction of Anti-D3
– 29 –
Branes on the Klebanov-Strassler Geometry”, JHEP 1306 (2013) 060, [arXiv:1106.6165].
S. Massai, “A Comment on Anti-Brane Singularities in Warped Throats”, arXiv:1202.3789.
I. Bena, M. Graña, S. Kuperstein, and S. Massai, “D3’s - Singular to the Bitter End”,
Phys.Rev. D87 (2013) 106010, [arXiv:1206.6369].
A. Dymarsky and S. Massai, “Uplifting the Baryonic Branch: A Test for Backreacting
Anti-D3-Branes”, arXiv:1310.0015.
[19] S. Kachru, J. Pearson, and H. L. Verlinde, “Brane/flux Annihilation and the String Dual of a
Non-Supersymmetric Field Theory”, JHEP 0206 (2002) 021, [hep-th/0112197].
[20] T. Gherghetta and A. Pomarol, “The Standard Model Partly Supersymmetric”, Phys.Rev. D67
(2003) 085018, [hep-ph/0302001].
R. Sundrum, “SUSY Splits, But Then Returns”, JHEP 1101 (2011) 062, [arXiv:0909.5430].
[21] D. Baumann, A. Dymarsky, S. Kachru, I. R. Klebanov, and L. McAllister, “D3-Brane
Potentials from Fluxes in AdS/CFT”, JHEP 1006 (2010) 072, [arXiv:1001.5028].
S. Gandhi, L. McAllister, and S. Sjörs, “A Toolkit for Perturbing Flux Compactifications”,
JHEP 1112 (2011) 053, [arXiv:1106.0002].
[22] S. Kachru, D. Simić, and S. P. Trivedi, “Stable Non-Supersymmetric Throats in String
Theory”, JHEP 1005 (2010) 067, [arXiv:0905.2970].
A. Dymarsky and S. Kuperstein, “Non-Supersymmetric Conifold”, JHEP 1208 (2012) 033,
[arXiv:1111.1731].
[23] P. G. Cámara, L. Ibáñez, and A. Uranga, “Flux-Induced SUSY-Breaking Soft Terms on D7-D3
Brane Systems”, Nucl.Phys. B708 (2005) 268–316, [hep-th/0408036].
D. Lüst, S. Reffert, and S. Stieberger, “Flux-Induced Soft Supersymmetry Breaking in Chiral
Type IIB Orientifolds with D3/D7-Branes”, Nucl.Phys. B706 (2005) 3–52, [hep-th/0406092].
D. Lüst, S. Reffert, and S. Stieberger, “MSSM with Soft SUSY Breaking Terms from D7-Branes
with Fluxes”, Nucl.Phys. B727 (2005) 264–300, [hep-th/0410074].
C. Burgess, P. G. Cámara, S. de Alwis, S. Giddings, A. Maharana, et. al., “Warped
Supersymmetry Breaking”, JHEP 0804 (2008) 053, [hep-th/0610255].
D. Lüst, F. Marchesano, L. Martucci, and D. Tsimpis, “Generalized Non-Supersymmetric Flux
Vacua”, JHEP 0811 (2008) 021, [arXiv:0807.4540].
P. McGuirk, G. Shiu, and Y. Sumitomo, “Holographic Gauge Mediation via Strongly Coupled
Messengers”, Phys.Rev. D81 (2010) 026005, [arXiv:0911.0019].
P. McGuirk, “Hidden-Sector Current-Current Correlators in Holographic Gauge Mediation”,
Phys.Rev. D85 (2012) 045025, [arXiv:1110.5075].
[24] F. Benini, A. Dymarsky, S. Franco, S. Kachru, D. Simić, and H. Verlinde, “Holographic Gauge
Mediation”, JHEP 0912 (2009) 031, [arXiv:0903.0619].
[25] P. McGuirk, “Falling flavors in AdS/CFT”, JHEP 1307 (2013) 102, [arXiv:1212.2210].
[26] K. Hashimoto and S. Nagaoka, “Recombination of Intersecting D-Branes by Local Tachyon
Condensation”, JHEP 0306 (2003) 034, [hep-th/0303204].
S. Nagaoka, “Higher Dimensional Recombination of Intersecting D-Branes”, JHEP 0402 (2004)
063, [hep-th/0312010].
[27] S. Cecotti, M. C. Cheng, J. J. Heckman, and C. Vafa, “Yukawa Couplings in F-Theory and
Non-Commutative Geometry”, arXiv:0910.0477.
– 30 –
[28] F. Marchesano, P. McGuirk, and G. Shiu, “Chiral Matter Wavefunctions in Warped
Compactifications”, JHEP 1105 (2011) 090, [arXiv:1012.2759].
[29] M. Marino, R. Minasian, G. W. Moore, and A. Strominger, “Nonlinear Instantons from
Supersymmetric p-Branes”, JHEP 0001 (2000) 005, [hep-th/9911206].
J. Gomis, F. Marchesano, and D. Mateos, “An Open String Landscape”, JHEP 0511 (2005)
021, [hep-th/0506179].
L. Martucci and P. Smyth, “Supersymmetric D-Branes and Calibrations on General N = 1
Backgrounds”, JHEP 0511 (2005) 048, [hep-th/0507099].
[30] D. Marolf, L. Martucci, and P. J. Silva, “Fermions, T-Duality and Effective Actions for
D-Branes in Bosonic Backgrounds”, JHEP 0304 (2003) 051, [hep-th/0303209].
D. Marolf, L. Martucci, and P. J. Silva, “Actions and Fermionic Symmetries for D-Branes in
Bosonic Backgrounds”, JHEP 0307 (2003) 019, [hep-th/0306066].
L. Martucci, J. Rosseel, D. Van den Bleeken, and A. Van Proeyen, “Dirac Actions for D-Branes
on Backgrounds with Fluxes”, Class.Quant.Grav. 22 (2005) 2745–2764, [hep-th/0504041].
[31] M. Berkooz, M. R. Douglas, and R. G. Leigh, “Branes Intersecting at Angles”, Nucl.Phys.
B480 (1996) 265–278, [hep-th/9606139].
[32] L. Aparicio, A. Font, L. E. Ibáñez, and F. Marchesano, “Flux and Instanton Effects in Local
F-theory Models and Hierarchical Fermion Masses”, JHEP 1108 (2011) 152,
[arXiv:1104.2609].
[33] D. Cremades, L. Ibáñez, and F. Marchesano, “Computing Yukawa Couplings from Magnetized
Extra Dimensions”, JHEP 0405 (2004) 079, [hep-th/0404229].
[34] F. Marchesano, P. McGuirk, and G. Shiu, “Open String Wavefunctions in Warped
Compactifications”, JHEP 0904 (2009) 095, [arXiv:0812.2247].
[35] R. C. Myers, “Dielectric Branes”, JHEP 9912 (1999) 022, [hep-th/9910053].
[36] B. Wynants, “Supersymmetric Actions for Multiple D-Branes on D-Brane Backgrounds”,
Master’s thesis, Katholieke Universiteit Leuven, 2006.
[37] J. P. Boyd, Chebyshev and Fourier Spectral Methods. Dover Publications, Inc., 2001.
[38] E. Shuster, “Killing Spinors and Supersymmetry on AdS”, Nucl.Phys. B554 (1999) 198–214,
[hep-th/9902129].
[39] A. Chodos and E. Myers, “Gravitational Contribution to the Casimir Energy in Kaluza-Klein
Theories”, Annals Phys. 156 (1984) 412. DOI: 10.1016/0003-4916(84)90039-3.
[40] K. Skenderis, “Lecture Notes on Holographic Renormalization”, Class.Quant.Grav. 19 (2002)
5849–5876, [hep-th/0209067].
[41] M. Krasnitz, “Correlation Functions in a Cascading N = 1 Gauge Theory”, JHEP 0212 (2002)
048, [hep-th/0209163].
O. Aharony, A. Buchel, and A. Yarom, “Holographic Renormalization of Cascading Gauge
Theories”, Phys.Rev. D72 (2005) 066003, [hep-th/0506002].
[42] I. R. Klebanov and A. A. Tseytlin, “Gravity Duals of Supersymmetric SU (N ) × SU (N + M )
gauge theories”, Nucl.Phys. B578 (2000) 123–138, [hep-th/0002159].
[43] M. Berg, M. Haack, and W. Mück, “Glueballs vs. Gluinoballs: Fluctuation Spectra in
– 31 –
Non-AdS/Non-CFT”, Nucl.Phys. B789 (2008) 1–44, [hep-th/0612224].
[44] J. Polchinski, String Theory. Vol. 2: Superstring Theory and Beyond. Cambridge University
Press, 1998.
[45] A. Meremianin, “Hyperspherical Harmonics with Arbitrary Arguments”, J.Math.Phys. 50
(2009) 013526, [arXiv:0807.2128].
– 32 –