<<

Kaluza-Klein Reduction of Massive and Partially Massless -2 Fields

James Bonifacio,a,1 Kurt Hinterbichlerb,2

aTheoretical , University of Oxford, DWB, Keble Road, Oxford, OX1 3NP, UK

bCERCA, Department of Physics, Case Western Reserve University, 10900 Euclid Ave, Cleveland, OH 44106, USA

Abstract

We describe the dimensional reduction of massive and partially massless spin-2 fields on general Einstein direct product . As with massless fields, the higher- dimensional gauge of the partially massless field displays itself upon dimen- sional reduction as a tower of St¨uckelberg symmetries for the massive modes of the tower. Unlike the massless case, the zero mode of the gauge symmetry does not dis- play itself as a lower-dimensional non-St¨uckelberg gauge symmetry enforcing partial masslessness on the zero mode. Partial masslessness is destroyed by the dimensional reduction and the zero mode gauge symmetry instead serves to eliminate the radion. In addition, we study the fully non-linear dimensional reduction of dRGT massive on a circle, which results in a massive scalar--vector theory which we expect to arXiv:1611.00362v2 [hep-th] 4 Sep 2019 be -free, and whose scalar-tensor sector is a special case of -varying .

1E-mail: [email protected] 2E-mail: [email protected] Contents

1 Introduction3

2 Setup 5 2.1 Background ...... 5 2.2 Linear ...... 6 2.3 Symmetric Tensor Decomposition ...... 7 2.4 Reduction of the PM Gauge Transformations ...... 9

3 Dimensionally Reduced Action 10 3.1 Diagonalization ...... 11 3.2 Spectrum and Stability ...... 14

4 Interactions 18

5 Conclusions 23

A Diagonalization Transformations 23

References 26

2 1 Introduction

Since Kaluza and Klein [1,2], the idea of compact extra has played a crucial role in many developments of . This is true whether one believes the extra dimensions are to be thought of as physically real, as in phenomenological studies of string compactifications or large extra dimensions [3], or are to be thought of as mathematical tools used to construct or explore purely four-dimensional theories1. The hallmark of any Kaluza-Klein (KK) reduction on a compact is an infinite but discrete tower of modes of increasing . These masses are determined by the eigenvalues of appropriate Laplacians acting on the internal space. In the case where the higher-dimensional fields are massless fields with spin, there are gauge symmetries. These gauge symmetries also get dimensionally reduced and become an infinite tower of lower-dimensional gauge symmetries. Most of these symmetries become algebraic St¨uckelberg-like fields associated with the massive modes in the KK tower. These are gauged shift symmetries which when fixed make the higher KK modes eat their corre- sponding longitudinal Goldstone bosons to become massive. The exception is when there are zero modes. In this case, there are lower-dimensional massless fields at the bottom of the KK tower, and the zero modes of the gauge symmetry are not St¨uckelberg but instead become the true gauge symmetries of the massless modes of the tower. In the case where spacetime is not flat, the notion of masslessness takes on greater richness. There exist fields which are neither massless nor massive, but instead propagating a number of degrees of freedom intermediate between that of a massless and massive field. These co-called “partially massless” (PM) fields carry gauge invariances that remove some of the degrees of freedom of a generic massive mode, but the amount of gauge symmetry is smaller compared with that of a massless field, so we are left with a number of degrees of freedom greater than that of a massless field and less than that of a massive field [6–16]. The first non-trivial example of a partially massless field occurs in the case of spin 2 2 2 on . At the special point m = D−1 Λ relating the mass and the cosmological

1For example, in maximally supersymmetric Yang-Mills theory, a weakly coupled Lagrangian description can be most easily constructed by starting with a simpler action in 10 dimensions and reducing on a torus [4], and at strong coupling, a 10-dimensional space compactified on a 5-dimensional sphere plays a central role through the AdS/CFT correspondence [5].

3 constant, a massive spin-2 field acquires a scalar gauge invariance which removes its scalar helicity. This case has been of interest due to a possible connection with the and (see [17] and the review [18]), and there have been many studies of the properties of the theory and possible nonlinear extensions [17,19–40]. Here, we perform the KK reduction of a general massive spin-2 field, including the partially massless case. We track the fate of the partially massless gauge symmetry as it undergoes dimensional reduction. Like in the massless case, the higher KK harmonics of the gauge symmetry become algebraic St¨uckelberg-like symmetries associated with massive modes in the tower. However, unlike in the massless case, the zero mode of the partially massless gauge symmetry does not become a lower-dimensional partially massless gauge symmetry. Instead it serves to remove the zero mode scalar, the radion, and we are left with no partially massless in the KK tower. The radion is often a source of instability or an undesired 5th force and many mechanisms have been invented to stabilize or screen it, so a mechanism such as this to remove it completely via a gauge symmetry could be of interest. The general pattern of fields which are present as a function of mass of the higher- dimensional field and of the spacetime is quite intricate, as detailed in Section 3.2. For the linear KK reductions, we use the method advocated in [41]. The method works at the level of the action, keeps gauge invariance and covariance intact at all stages, and is applicable to any internal geometry. The result is an action for the fields of the KK tower, from which the masses can be read off in terms of the eigenvalues of appropriate internal Laplacians. In addition, it allows us to take into account all of the subtleties associated with zero modes, Killing vectors, conformal Killing vectors, etc. of the internal space. We then move on to consider a fully non-linear example, the dimensional reduction of the ghost-free de Rham-Gabadadze-Tolley (dRGT) theory [42,43] (reviewed in [44,45]) of a massive spin-2 field. By keeping only the zero modes, to full non-linear order, we arrive at a massive scalar-vector-tensor theory that is a massive version of the original KK theory. By projecting out the vector we then obtain a family of massive scalar-tensor theories that are a special case of mass-varying massive gravity [46, 47] with varying α’s [48]. Because these theories arise as consistent truncations of a ghost-free higher-dimensional theory, we expect them to be ghost free.

4 Conventions and notation: Our signature convention is mostly plus: (−, +, +,...). We work with a D = (d + N)-dimensional spacetime with indices A, B, . . . and coordinates XA. It is a product of a d-dimensional spacetime M and an N-dimensional space N . M has spacetime indices µ, ν, . . . and coordinates xµ and N has spatial indices m, n, . . . and coordinates ym.

2 Setup

2.1 Background Spacetime

We begin by choosing a suitable direct product background metric on which the higher- dimensional massive graviton can propagate. Massless are only known to consis- tently propagate on background that are Einstein spaces, i.e. solutions to the vacuum Einstein equations with a cosmological constant [49, 50]. Similarly, the only known backgrounds that partially massless spin-2 fields can propagate on are Einstein spaces [20,51]. There is no such restriction for general massive gravitons, which can propagate on an ar- bitrary background [52–54]. In this paper we are most interested in the partially massless limits of massive gravity, so we will only consider massive gravitons on Einstein spaces. It would be interesting to see what happens with these limits on non-Einstein backgrounds.

We therefore consider a D = (d+N)-dimensional background metric GAB that satisfies the condition for an Einstein space, R R = (D) G , (2.1) AB D AB with constant R(D), which is equivalent to demanding that the metric sat- isfies Einstein’s equations 1 R − R G + Λ G = 0 (2.2) AB 2 (D) AB (D) AB with cosmological constant (D − 2) Λ = R . (2.3) (D) 2D (D) To perform a KK reduction we require the background to be in the form of a product space

5 M × N , so we take the background metric to be of the form ! gµν(x) 0 GAB(x, y) = , (2.4) 0 γmn(y)

µ where gµν is the metric on the d-dimensional spacetime M with coordinates x and γmn is the metric on the N-dimensional space N with coordinates ym. Writing (2.1) in terms of the lower-dimensional metrics and their gives R R R = (d) g ,R = (N) γ , (2.5) µν d µν mn N mn where 2d 2N R = Λ ,R = Λ , (2.6) (d) d + N − 2 (D) (N) d + N − 2 (D) which shows that both M and N must also be Einstein spaces. Note also that the curvatures satisfy R R R (D) = (d) = (N) . (2.7) D d N We further define the partially massless masses for each Einstein manifold with more than 2 dimensions: D − 2 R N − 2 R d − 2 R m2 ≡ (D) , m2 ≡ (N) , m2 ≡ (d) . (2.8) (D)PM D − 1 D (N)PM N − 1 N (d)PM d − 1 d

2 2 For R < 0 these squared masses are all negative. By (2.7) we have m(D)PM > m(d)PM when R > 0.

2.2 Linear Action

We now consider massive spin-2 perturbations HAB about the Einstein background GAB. These are described by the Fierz-Pauli [55] Lagrangian in D dimensions,

LFP,D 1 C AB B AC AB 1 A √ = − ∇C HAB∇ H + ∇C HAB∇ H − ∇AH(D)∇BH + ∇AH(D)∇ H(D) −G 2 2 R  1  1 + (D) H HAB − H2 − m2 H HAB − H2  , (2.9) D AB 2 (D) 2 AB (D)

where ∇A is the background covariant derivative, R(D) is the constant curvature given by (2.3), and m is the graviton mass. The Lagrangian (2.9) describes different numbers of

6 propagating degrees of freedom depending on the value of the mass and scalar curvature. We will consider only m2 > 0; the general dimensional reduction for m = 0 was worked out in [41]. When R > 0, the Higuchi bound states that a spin-2 field is only stable if m ≥ m(D)PM. The different cases we consider are:

1.0 < m < m(D)PM, R > 0. This describes an unstable massive graviton on a positively curved Einstein background with (D + 1)(D − 2)/2 degrees of freedom, one of which is a ghost-like degree of freedom.

2. m = m(D)PM, R > 0. This describes a stable partially massless graviton on a positively curved Einstein background which has the scalar gauge symmetry R δH = ∇ ∇ α + G (D) α (2.10) AB A B AB D(D − 1)

and propagates (D + 1)(D − 2)/2 − 1 degrees of freedom.

3. m > m(D)PM. This describes a stable massive spin-2 particle on an Einstein background with (D + 1)(D − 2)/2 degrees of freedom. This includes the case R ≤ 0. There are other values of m2 that are special from the perspective of lower-dimensional physics and we discuss these later.

2.3 Symmetric Tensor Decomposition

In order to find the d-dimensional KK spectrum we must decompose our higher-dimensional field into orthonormal eigenfunctions of various Laplacian operators defined on the internal manifold N . This is the generalization to arbitrary manifolds of the familiar Fourier decom- position used in the original KK construction and allows us to integrate out the dependence of fields on the internal space N . The field HAB breaks up into pieces Hµν, Hµn, and Hmn, which transform respectively as a scalar, vector and symmetric tensor on the internal man- ifold. We thus need to decompose general scalar, vector and symmetric tensor fields into eigenmodes of appropriate Laplacian operators on an arbitrary compact manifold. The ap- propriate decomposition is given by the Hodge decomposition for scalar and vector fields and its generalization to symmetric tensor fields [56,57], which for the different components

7 of HAB give X 1 H = ha ψ + √ h0 , µν µν a V µν a n X i X a Hµn = AµYn,i + Aµ∇nψa , i a X I TT X i Hmn = φ hmn,I + φ (∇mYn,i + ∇nYm,i) , I i6=Killing X  1  X 1 1 1 + φ ∇ ∇ ψ − ∇2ψ γ + φ¯aψ γ + √ φ γ . a m n a N a mn N a mn N V 0 mn a6=conformal a N (2.11)

The fields ψa are a basis of positive orthonormal eigenmodes of the scalar Laplacian on N with positive eigenvalue λa (i.e. excluding the zero mode), Z N √ a∗ a d y γψ (y)ψb(y) = δb , (2.12)

 a (y) + λa ψ (y) = 0, λa > 0, (2.13) and V is the volume of N N Z N √ VN ≡ d y γ. (2.14)

The fields Yn,i are a basis of orthonormal transverse eigenvectors of the vector Laplacian on N with eigenvalues λ , i Z N √ ∗ m d y γYi,m(y)Yj (y) = δij, (2.15)

n ∇(y)Yi,n(y) = 0, (2.16)

m ∆Yi,n ≡ −(y)Yi,n(y) + Rn Yi,m(y) = λiYi,n(y), λi ≥ 0. (2.17) TT The fields hmn,I are a basis of symmetric transverse traceless orthonormal eigenmodes of the

Lichnerowicz operator [58] on N with eigenvalues λI ,

m TT mn TT ∇ hmn,I (y) = γ hmn,I (y) = 0, (2.18) Z N √ TT,I∗ mn,T T I d y γhmn (y)hJ (y) = δJ , (2.19) 2R ∆ hTT (y) ≡ − hTT (y) + (N) hTT (y) − 2R hpq T T (y) = λ hTT . (2.20) L,(y) mn,I (y) mn,I N mn,I mpnq I I mn,I The bases of eigenmodes can always be chosen to be real functions, however it is often convenient to allow them to be complex. Ensuring that the physical fields are real then

8 requires a reality condition, e.g, restricting the φa(x) to satisfy φa∗(x) = φa¯(x), wherea ¯ denotes some involution on the set of indices a. The sum over i in the decomposition of Hmn in (2.11) excludes Killing vectors, which satisfy

∇mYn,i(y) + ∇nYm,i(y) = 0. (2.21)

Nontrivial Killing vectors can exist on closed Einstein manifolds only when R(N) ≥ 0 and they are precisely the transverse eigenmodes of the vector Laplacian with eigenvalue λi =

2R(N)/N, which is the smallest possible eigenvalue for transverse eigenmodes of the vector Laplacian on a closed Einstein manifold with nonnegative curvature. Similarly, the sum over a excludes conformal scalars, which satisfy 1 ∇ ∇ ψ (y) − ψ (y)γ = 0. (2.22) m n a N (y) a mn Conformal scalars exist only on manifolds that are isometric to the sphere [59] and are precisely the eigenmodes of the scalar Laplacian with eigenvalue R λ = (N) . (2.23) a=conformal N − 1 By the Lichnerowicz bound [60], (2.23) is the smallest possible eigenvalue for eigenmodes of the scalar Laplacian on a closed Einstein manifold with positive curvature. For more details about the decomposition of HAB and why it takes the form it does, we refer to [41].

2.4 Reduction of the PM Gauge Transformations

When m = m(D)PM the D-dimensional Lagrangian acquires the partially massless gauge symmetry (2.10). To find how this gauge symmetry acts on the lower-dimensional fields we first decompose the partially massless gauge parameter in eigenmodes of the internal Laplacian, X 1 α = αaψ + √ α . (2.24) a V 0 a N

9 Expanding the gauge transformation (2.10) using (2.11) and (2.24) and equating components then gives the following gauge transformations for the lower-dimensional fields: R δha =∇ ∇ αa + g (D) αa, δAa =∇ αa, µν µ ν µν D(D − 1) µ µ R δh0 =∇ ∇ α0 + g (D) α0, δAi =0 , µν µ ν µν D(D − 1) µ  NR  δφ¯a = (D) − λ αa, δφa =αa, a 6= conformal D(D − 1) a NR δφ0 = (D) α0, δφi =0, i 6= Killing D(D − 1) δφI =0 . (2.25)

We can see that the higher-dimensional partially massless symmetry acts as a St¨uckelberg symmetry on some of the lower-dimensional fields. In particular, we can use α0 and αa to gauge fix φ0 = φa6=conformal = φ¯a=conformal = 0. (2.26)

This completely fixes the gauge symmetry, which indicates already that there will be no par- tially massless fields in the lower-dimensional theory. Note that some of the gauge symmetry would remain if NR λ = (D) , (2.27) a=conformal D(D − 1) but this is forbidden by (2.7) and (2.23) for D > N.

3 Dimensionally Reduced Action

Plugging the decomposition (2.11) into (2.9) and integrating over the extra dimensions using orthogonality gives the d-dimensional action Z d √ S = d x −g(LI + Li + L0 + La), (3.1)

10 where the L’s are quadratic in the d-dimensional fields appearing in the decomposition of

HAB. The different terms in (3.1) are given by:   X 1 1 R(d) L = − |∂φI |2 − λ + m2 − 2 |φI |2 , I 2 2 I d I     X 1 2R(d) 1 2R(d) L = − |F i |2 − λ − |Ai − ∂ φi|2 − m2|Ai |2 − m2 λ − |φi|2 , i 2 µν i d µ µ µ 2 i d i   0 µν,0 0 0  0 0 2 R(d) L =ε(h ) 2 + h ∇ ∇ φ − φ g + h φ m − 0 m µ ν  µν d N − 1 h i + ∂φ02 + (φ0)2 m2 − m2  , 2N (N)PM   X a 1 a 2 2R(d) 2 a 2 N − 1 a 2 2 a 2 L = (h ) 2 − λ |F | + λ − m |A | + |∂ φ¯ | + m |φ¯ | a m +λa 2 a µν a d µ 2N µ a 1h + 2λ hµν,a∗(∇ Aa − g ∇αAa ) + hµν,a∗(∇ ∇ φ¯a − g φ¯a) 2 a µ ν µν α µ ν µν N − 1 R  N − 1 i + h∗aφ¯a λ + m2 − (d) − 2λ ∇µAa∗φ¯a + c.c N a d a N µ  R   N − 1 (N) N − 2 a ¯a 2 1 a 2 1 2 a 2 + λa λa − |λaφ + φ | − |∂φ | − m |φ | N N − 1 2Nλa 2 2  a µ a∗ + ((h /2 − ∇µAa )φ + c.c) , (3.2)

where we have defined (h)m2 as the d-dimensional massive graviton action, i.e. the d- dimensional version of (2.9).

3.1 Diagonalization

To determine the spectrum and stability of S we need to diagonalize Li, L0 and La. The required demixing transformations depend on m2 and details of the spaces involved. Here we present the final diagonalized Lagrangians Li, L0 and La, leaving the details of the required transformations to the appendix. The fields in the diagonalized Lagrangians are combinations of the fields in the original Lagrangian but to avoid clutter we give them the same names. We define 2R m2 ≡ m2 + λ − (d) , (3.3) k k d

where k = I, 0, i, or a and λ0 = 0.

11 Diagonalized Li

The diagonalized Lagrangian Li is given by

X 1 m2 (m2 − m2)  1  L = − |F i |2 − m2|Ai |2 − i |∂φi|2 + m2|φi|2 i 2 µν i µ m2 2 i i6=Killing i X 1 + − |F i |2 − m2|Ai |2. (3.4) 2 µν µ i=Killing

Diagonalized L0

When m 6= m(d)PM and d > 1 the diagonalized L0 is

m2 − m2 0 D − 1 (D)PM 0 2 2 0 2 2 L0 = (h )m − 2 2 (∂φ ) + m0(φ ) . (3.5) 2N(d − 1) m − m(d)PM

When m = m(d)PM 0 1 2 2 L0 = (h )m2 − m(d)PMh0. (3.6) (d)PM 4

When d = 1 many of the terms vanish and L0 reduces to a nondynamical Lagrangian. The Lagrangian (3.6) is unusual, and we will say more about this below.

Diagonalized La

When ma 6= 0, N > 1, and d > 1 the Lagrangian La is

2   X a m λa 1 a 2 2 a 2 L = (h ) 2 − |F | + m |A | a m +λa m2 + λ 2 µν a µ a a a 2 2 a 2 ¯a 2 2 ¯a 2 − σa |∂φ | + ma|φ | − σ¯a |∂φ | + ma|φ | , (3.7) where σa andσ ¯a are functions of λa, R, d and N that are defined in the appendix – in particular, σa=conformal = 0. When ma = 0 we can write La as

2   X m λa R(d) L = (ha) − |∂ Aa|2 + |Aa |2 a 2R(d)/d m2 + λ µ ν d µ 2 a ma=0  2 2   2 2  (N − 1)(D − 1) m − m(N)PM m − m(D)PM − |∂φa|2. (3.8) 2d

12 Achieving ma = 0 requires tuning m and λa but can occur for m ≤ m(N)PM. When N = 1 all scalars are conformal and for d > 1 we get

2   4 X a m λa 1 a 2 2 a 2 dm  ¯a 2 2 ¯a 2 La = (h )m2 − |F | + m |A | − ∂φ + m |φ | . a m2 2 µν a µ 2(d − 1)m4 a a=conformal a a (3.9)

Lastly, when d = 1 we get

N − 1 X L = − |∂ φ¯a|2 + m2|φ¯a|2 . (3.10) a 2N µ a a6=conformal

2 Let us now comment on the degrees of freedom described by (3.6). The |h0| term spoils the PM gauge symmetry and the Fierz-Pauli tuning of the mass term, so we expect that (3.6) describes an extra ghostly scalar degree of freedom2. A 3 + 1 analysis reveals that for d = 4 (3.6) describes four PM modes in addition to two scalar modes, one of which is a ghost3. In fact, it is not possible to transform (3.6) to a set of decoupled kinetic terms of familiar fields. It is inherently undemixable, and describes a new kind of irreducible representation which is a combination of a spin-2 field and a spin-0 field on ; it is a field theoretic manifestation of a so-called “extended module” which mixes spin-0 and spin-2 fields, seen in 2 the Hilbert space of the non-unitary  conformal field theory in three dimensions [61]. An 2 equivalent form of (3.6) occurs in the holographic dual of the  conformal field theory [62]. 2In the case of linearized , adding a term βh2 spoils the linearized diffeomorphism symmetry but gives a theory that is the same as linearized general relativity up to global degrees of freedom. 3Another way to see the ghost once we know that there are six degrees of freedom is to perform the St¨uckelberg transformation m2 h0 → h0 + ∇ ∇ ϕ + g (d)PM ϕ, (3.11) µν µν µ ν µν (d − 2) which introduces a new scalar field ϕ. This transformation has the form of a PM transformation so only 2 affects the |h0| term. The resultant Lagrangian is then invariant under the PM St¨uckelberg symmetry

m2 δh = ∇ ∇ α + g (d)PM α, δϕ = −α. (3.12) µν µ ν µν (d − 2)

Now rescale ϕ → ϕ/m and take the decoupling limit m → 0 while preserving the PM relation m = m(d)PM. In this limit the partially massless kinetic term breaks up into massless tensor and vector modes and the 2 surviving scalar term is higher derivative ∼ ϕ ϕ, which shows that there is a scalar ghost.

13 3.2 Spectrum and Stability

Now that we have diagonalized the action we can determine its spectrum and stability. The spectrums of LI and Li are straightforwardly determined, so we discuss these first. Next we discuss the spectrums of L0 and La for the three cases where m is less than, equal to, or larger than m(D)PM. We assume that d > 1 and N > 1 until the end where we discuss the special cases N = 1 and d = 1. We conclude with a summary of these stability results.

Spectrum of LI

2 The Lagrangian LI describes a tower of massive scalars with squared masses given by mI , 2 where mI is defined by (3.3). The stability of a scalar mass term depends on the background 2 curvature. For a flat or de Sitter space stability requires mI ≥ 0 and for an anti-de Sitter space stability requires d − 1 R m2 ≥ (d) , (3.13) I 4 d which is the Breitenlohner-Friedman bound [63, 64]. We do not know of bounds for the more general case of a scalar on an Einstein space, so we take these bounds for maximally symmetric spaces as our conditions for stability. There exist positively curved compact manifolds with large negative eigenvalues of the Lichnerowicz operator (such as the B¨ohm metrics on products of spheres [65]). This means that the scalars φI can be tachyonic, but for any given internal space N there is some smallest m that stabilizes them.

Spectrum of Li

2 The spectrum of Li is a tower of massive vectors with mass squared mi and a tower of 2 massive scalars with mass squared mi /2, one for each non-Killing transverse vector of the internal space, and a massive vector with mass m for each Killing vector of the internal 2 2 space. For a non-Killing vector we have λi > 2R(N)/N and hence mi > m > 0. Stability of vectors requires only that their mass be positive and hence all fields are stable in this sector.

14 Spectrum of L0 and La for 0 < m < m(D)PM

We first consider the spectrum of L0 and La for the mass range 0 < m < m(D)PM, which is relevant for D > 2 and R > 0. This corresponds to a higher-dimensional theory with an unstable graviton, since its mass falls below the Higuchi bound, so we expect instabilities in the lower-dimensional theory. The spectrum consists of a zero-mode massive scalar and a zero-mode massive graviton, plus a massive scalar, vector and tensor for each conformal scalar. There is also a tower of massive gravitons, a tower of massive vectors, and two towers of massive scalars, one field in each of the towers for each non-conformal scalar in the internal space. Below we discuss the stability of this spectrum for each spin. Spin 2: The zero-mode graviton has mass m and thus by the Higuchi bound is unstable for m < m(d)PM and stable for m > m(d)PM. When m = m(d)PM the zero-mode sector (3.6) contains a new kind of irreducible representation which can be thought of as an undemixable combination of a massive graviton and a ghost, as discussed further at the end of Section 3.1 2 and in the appendix. The other gravitons have mass squared m + λa > 0 and are always stable.

Spin 1: For ma 6= 0 the massive vectors have mass ma, defined in (3.3), so their stability depends on m and λa. For m < m(N)PM, the massive vectors corresponding to conformal scalars are tachyonic since 2 2 2 ma=conformal = m − m(N)PM, (3.14) and the massive vectors corresponding to non-conformal scalars can be tachyonic or stable 2 depending on the sign of ma. For m > m(N)PM the massive vectors are all stable. When ma = 0 the spatial components of the vector field describe d − 1 scalars with masses squared equal to R(d)/d, which violate the Breitenlohner-Friedman stability bound when R < 0 and d < 5, and the temporal component of the vector field is a scalar ghost.

Spin 0: The zero-mode scalar – the radion mode – is tachyonic for m < m(d)PM and ghostly for m(d)PM < m < m(D)PM. When m = m(d)PM the zero-mode scalar gets folded into the graviton field giving (3.6). The scalars coming from the conformal modes are tachyonic for m < m(N)PM and ghostly for m(N)PM < m < m(D)PM. The scalars in the two towers, 2 corresponding to non-conformal scalars in the internal space, are both tachyonic if ma < 0 2 and have one healthy scalar and one ghost if ma > 0. When ma = 0 there is one massless scalar if a 6= conformal and no scalars if a = conformal, since then m = m(N)PM. The

15 scalar spectrum coming from the nonconformal modes of the internal space is summarized in Figure1.

Spectrum of L0 and La for m = m(D)PM

Now consider the case when the higher-dimensional spin-2 field is partially massless, m =

m(D)PM. The spectrum consists of a tower of massive gravitons, a tower of massive vectors, and a tower of massive scalars. There is also a massive graviton and massive vector for each conformal mode, and a zero-mode massive graviton. All of these fields are stable. The effect of the higher-dimensional partially massless gauge symmetry is to remove a tower of scalar fields, including the would-be unstable radion, which is consistent with the gauge fixing (2.26). There are no partially massless fields in the lower-dimensional spectrum since the

lightest massive graviton has mass m(D)PM and this is larger than m(d)PM.

Spectrum of L0 and La for m(D)PM < m

Lastly, consider the case of a stable higher-dimensional massive graviton, m > m(D)PM. This

includes all cases when R ≤ 0. The field content is the same as for m < m(D)PM. The zero- 2 2 2 mode scalar is tachyonic for m(D)PM < m < 2R(d)/d and nontachyonic for m > 2R(d)/d. 2 The scalar is massless when m = 2R(d)/d. All the other fields are stable.

Spectrum When d = 1 or N = 1

When N = 1, i.e. when N is the circle S1 with radius r, there are no symmetric transverse

traceless on N so LI is zero. There are no non-Killing transverse vectors and the only Killing vector is the constant vector on the circle, so Li contains a single vector with mass m, which is a massive version of the original KK . There is a zero-mode massive graviton and a zero-mode massive scalar, both with mass m. There is also a tower of massive gravitons, vectors and scalars all with masses squared a m2 = m2 + , (3.15) a r2 where a = 1, 2, 3,... When d = 1 there are massive scalars for each Lichnerowicz mode, each non-Killing mode and each non-conformal mode.

16 R(N) λ m2 m2 m2 (1) (D)PM (N)PM N 1 − 2 scalars 1 scalar + 1 ghost 2 tachyons2 scalars 1 scalar + 1 ghost 2

2R (d) (1) d

x ≥ 5 y ≥ 3 6 2 y ≥ 1.5 x + y ≥ 2 x ≤ 5 6 y ≥ 1.5 x + y ≤ 2 y < 1.5 2 scalars 1 scalar + 1 ghost 2 tachyons R λ m2 m2 m2 (N) (1) (D)PM (N)PM N 1 −

2 2 2 R(N) 2 R(N2) 2 R(N) m m (1) λ m2 m2 m λ m2 m m λ m(1) (D)PM (1)(N)PM (D)PM (N)PM N (D1)PM (N)PM N 1 N 1 − − −

Figure 1: Plot of the field content of the nonconformal scalar sector in the (m2, λ) plane for R > 0, d > 1, N > 2. The lower rectangle corresponds to eigenvalues at or below the Lichnerowicz bound. The rightmost region has 2 healthy scalars, the adjacent region has 1 healthy scalar and 1 ghost, and the triangular region has 2 tachyons. The angled line is

ma = 0 and the vertical line is m = m(D)PM, which has 1 healthy scalar.

Summary of Stability Results

2 2 We have found that avoiding the radion instability requires either m ≥ 2R(D)/D or m = 2 m(D)PM. The partially massless value thus corresponds to an isolated line of stability in the (m2,R) plane since the PM gauge symmetry renders the unstable radion pure gauge. There may be other instabilities depending on the size of m2 and the spectrum of eigenvalues of the Lichnerowicz operator on the internal space.

1 17

1 1

1

1

1

1

1

1 4 Interactions

So far we have been working with quadratic actions only. Including interactions requires a nonlinear D-dimensional theory of massive gravity. Adding a generic interaction term to the linear massive Fierz-Pauli action (2.9) gives a theory with a massive graviton and an additional scalar ghost, the Boulware-Deser ghost [66], which becomes active at the non- linear level. Our previous linear analysis is insensitive to the presence of a Boulware-Deser ghost in the higher-dimensional theory. However, if we restrict to higher-dimensional the- ories with the ghost-free dRGT [43] interactions, then the quadratic action describes all propagating modes and our previous results determine the full KK spectrum4. A straight- forward self-interacting theory of a single partially massless spin-2 field does not appear to exist [17,24,26,29], so we will not be able to address this case non-linearly. It is possible, if tedious, to include interactions in the lower-dimensional theory by

plugging the decomposition of HAB order-by-order into the higher-dimensional interactions and integrating over the resulting products of eigenmodes. Doing this to all orders gives a lower-dimensional re-writing of the higher-dimensional theory. For example, in the original KK setup, D = 5 general relativity with a single compact extra gives a d = 4 theory with a massless graviton, vector and scalar, all interacting with an infinite tower of massive gravitons. Similarly, based on the results of the previous section, reducing D = 5 dRGT massive gravity should give a d = 4 theory of an infinite number of massive gravitons, vectors, and scalars, all interacting with one another.5 One tractable way to study nonlinear interactions is to consider a single compact

4dRGT massive gravity itself can be derived from “dimensional deconstruction” of Einstein gravity [67– 69], which can be thought of as a kind of discrete KK reduction. 5 What is the cutoff of this theory? The strong coupling scale in (n + 4)-dimensional massive gravity n+2  4  n+6 is given by Λ n+6 ≡ m n+2 M(n+4) , where M(n+4) is the (n + 4)-dimensional Planck mass. The d = 4 n+2 2 n+2 n Planck mass is given by M(4) ∼ M(n+4)R , where R is the characteristic length scale of the internal space. This gives 1  6  n+6 Λ3 Λ n+6 ∼ , (4.1) n+2 Rn

1 2  3 where Λ3 ≡ m M(4) is the d = 4 strong coupling scale of massive gravity. Consistency of the effective −1 field theory requires that the inverse size of the internal space is smaller than the cutoff, R < Λ n+6 . n+2 Together with (4.1) this implies that Λ n+6 < Λ3, so dimensionally reducing in this way cannot raise the n+2 strong coupling scale above Λ3.

18 extra dimension and to truncate to the zero modes. This truncation is consistent because translations in the extra dimension act as a global U(1) symmetry on the four-dimensional fields, and the zero modes (which are constants on the circle) are the only singlets under this U(1). Let us briefly review how this procedure works in the case of general relativity before applying it to massive gravity. We take the five-dimensional spacetime to be M 4 × S1, the product of four-dimensional Minkowski and a circle of length L. The five-dimensional Einstein-Hilbert action is M 3 Z Z L √ S = (5) d4x dy −GR(G), (4.2) 2 0

where GAB is the higher-dimensional metric and M(5) is the five-dimensional Planck mass. The zero mode truncation is accomplished by substituting the following ansatz for the metric

2 2 ! gµν(x) + φ (x)Aµ(x)Aν(x) φ (x)Aµ(x) GAB(x) = . (4.3) 2 2 φ (x)Aν(x) φ (x)

Substituting into the action gives

M 3 Z √  1  S = (5) L d4x −gφ R − φ2F 2 . (4.4) 2 4 µν

We can go to the Einstein frame and canonically normalize the scalar by definingg ˜µν ≡ φgµν and φ ≡ e−ψ. This gives

M 2 Z  1 3  S = (4) d4xp−g˜ R˜ − e−3ψF 2 − (∂ψ)2 , (4.5) 2 4 µν 2

2 3 where the 4D Planck mass is given by M(4) = M(5)L. This describes an interacting massless graviton, vector and scalar. Now consider five-dimensional dRGT massive gravity on the spacetime M 4 × S1. The action is given by

3 Z Z L 2 5 ! M(5) √ m X p  S = d4x dy −G R(G) − β S G−1η , (4.6) 2 4 n n 0 n=0

where ηMN is the 5D fiducial metric, which we take to be flat. Sn are the symmetric

19 polynomials, given by

S0(M) = 1

S1(M) = [M] 1 S (M) = [M]2 − [M 2] 2 2! 1 S (M) = [M]3 − 3[M][M 2] + 2[M 3] 3 3! 1 S (M) = [M]4 − 6[M]2[M 2] + 8[M][M 3] + 3[M 2]2 − 6[M 4] 4 4! 1 S (M) = [M]5 − 10[M]3[M 2] + 15[M][M 2]2 + 20[M]2[M 3] − 20[M 2][M 3] − 30[M][M 4] + 24[M 5] , 5 5! µ where [M] denotes the trace of M ν. The term in (4.6) proportional to β5 does not contribute to the equations of motion. Furthermore, imposing that flat spacetime (or M 4 × S1) is a solution and that m is the graviton mass gives the conditions

β0 + 4β1 + 6β2 + 4β3 + β4 = 0, (4.7)

β1 + 3β2 + 3β3 + β4 = 8, (4.8) which leaves a three-parameter family of theories. To perform the general dimensional reduction we will consider a simple case of (4.6) given by the “minimal model” of [70], which corresponds to β1 = 8, β0 = −32, and the other

βn vanishing. We also find it convenient to use the vielbein formulation, following [71]. The action for the five-dimensional minimal model in vielbein form is M 3 Z Z L (5) 4 2 2 M A S = d x dy|E| R(E) + 8m − 2m E AδM , (4.9) 2 0 A A where EM and δM are the vielbeins for GMN and ηMN respectively,

A B A B GMN = EM EN ηAB, ηMN = δM δN ηAB, (4.10)

A M |E| is the determinant of EM , and E A is the inverse vielbein

M A M E AEN = δN . (4.11)

The vielbein and inverse vielbein corresponding to (4.3) are ! ! e a(x) φ(x)A (x) eµ (x) 0 ˆ A µ µ ˆM a EM = , E A = , (4.12) ν −1 0 φ(x) −e a(x)Aν(x) φ (x)

20 a µ where eµ and e a are the vielbein and inverse vielbein for gµν. Putting a general vielbein in the diagonal form (4.12) requires a local rotation of four-dimensional spacetime into the fifth dimension. A general vielbein may thus be written as

A ∗A ˆ B EM = Λ BEM , (4.13)

ˆ B ∗A where EM is upper triangular and Λ B defines a local rotation. We can parameterize the required rotation in terms of a vector Ba by writing the generator as

a ! A 0 B ω B = , (4.14) −Bb 0 √ ∗A ω A a 2 a 2 where Λ B = (e ) B and Bb ≡ ηbaB . Defining B ≡ BaB and |B| ≡ B , we can write the rotation matrix as (ω2)A sin |B| Λ∗A = δA + B (1 − cos |B|) + ωA . (4.15) B B B2 B |B|

We can now find the dimensionally reduced action by plugging (4.13) into (4.9) and

using (4.15). The Einstein-Hilbert kinetic term is locally Lorentz invariant so Ba only ap- µ µ a 2 µ pears through the potential. Defining B = e aB , so that B = BµB , we can write the dimensionally reduced theory as M 2 Z  1 S = (4) d4x|e|φ R(e) − φ2F 2 + 8m2 2 4 µν  BµB sin |B| −2m2 δ aeµ + eν δ a ν (cos |B| − 1) + φ−1 cos |B| + A Bµ , (4.16) µ a a µ B2 µ |B|

a where δµ is the vielbein for the four-dimensional flat fiducial metric.

The field Bµ appears in (4.16) without derivatives and is thus an auxiliary field that should be integrated out using its equations of motion. We will not attempt to integrate out B exactly, but it can be done to any desired order in powers of the fields. For example, to quadratic order we can write Z   4 1 2 2 µ 2 2 2 2 S = d x (h) 2 − F − 2m (A B − B ) − (∂ψ) − m ψ , (4.17) m 4 µν µ

−ψ where φ ≡ e and we have demixed hµν and φ. Eliminating B using its equations of motion gives Z   4 1 2 1 2 2 2 2 2 S = d x (h) 2 − F − m A − (∂ψ) − m ψ , (4.18) m 4 µν 2

21 which describes a massive graviton, vector and scalar with mass m. This agrees with the zero-mode spectrum obtained in Section 3.2 for N = 1 and d = 4. The theory (4.16) is a massive scalar-vector-tensor theory that derives from the five-

dimensional minimal model. We can obtain a massive scalar-tensor theory by setting Aµ =

0, which is classically consistent since Aµ does not appear linearly after integrating out

Bµ. When Aµ = 0, the auxiliary field’s equations of motion are solved by Bµ = 0, which corresponds to the simple metric ansatz ! gµν(x) 0 GAB(x) = . (4.19) 0 φ2(x)

Using the ansatz (4.19), we can readily find the associated dimensional reduction for the full three-parameter higher-dimensional dRGT theory (4.6). The result of this is

2 Z 2 4   ! M(4) √ m X βn+1 p  S = d4x −gφ R(g) − β + S g−1η , (4.20) 2 4 n φ n n=0

where we have dropped the S5 term since it becomes a total derivative in four dimensions.

Again, the parameter β5 does not contribute to the equations of motion. In Einstein frame the action is

M 2 Z 2 4 ! (4) 4 p 3 2 m X (4−n) ψ −ψ  p  S = d x −g˜ R˜ − (∂ψ) − e 2 β e + β S g˜−1η , 2 2 4 n n+1 n n=0 (4.21) −ψ whereg ˜µν ≡ φgµν and φ ≡ e . The action (4.21) defines a theory with three parameters in addition to the graviton mass and cosmological constant6. At quadratic order around flat spacetime, (4.21) describes a free graviton and scalar, both with mass m. Because (4.21) arises as a consistent truncation of dRGT ghost-free massive gravity, we expect it to also be ghost free. It is an example of a type of mass-varying massive gravity [46, 47] in which the graviton mass and dRGT parameters are promoted to functions of a scalar field. Cosmological perturbations of this general class of theories were studied in [48].

6In deriving (4.21) we used a flat fiducial metric and imposed (4.7) and (4.8), but the final result makes sense for fiducial metrics and backgrounds that are not flat.

22 5 Conclusions

We have derived the linearized KK spectrum of a generic massive spin-2 field propagating on an Einstein direct product space. We found that in the case where the higher-dimensional graviton is partially massless, its zero mode does not lead to a lower-dimensional partially massless graviton. In this sense, dimensional reduction destroys partial masslessness. The higher-dimensional partially massless symmetry instead acts to remove a tower of scalar fields from the spectrum. Among the scalars which are removed is the zero mode, i.e. the radion. The radion, like other moduli, is a light scalar and we see no sign of such light scalars in nature, so in realistic models they must be stabilized, or screened by nonlinearities [72–74]. One can speculate that partial masslessness might be useful as a mechanism for removing such moduli without resort to stabilization or screening. In addition, we studied the zero-mode sector of the dimensional reduction of ghost-free dRGT massive gravity on a circle. This leads to a three-parameter family of massive scalar- tensor-vector theories, which we expect to be ghost free, and whose scalar-tensor sectors are examples of mass-varying massive gravity. It would be interesting to further explore the consequences of these theories and to try to prove their ghost-freedom directly. Acknowledgements: J.B. would like to thank the Perimeter Institute for hospitality while this work was under way. Research at Perimeter Institute is supported by the Gov- ernment of Canada through Industry Canada and by the Province of Ontario through the Ministry of Economic Development and Innovation.

A Diagonalization Transformations

In this appendix we give the transformations needed to diagonalize the action and discuss some special cases. When there are multiple demixing transformations they are performed in the order listed. We define 2R m2 ≡ m2 + λ − (d) , k k d where k = I, 0, i, or a and λ0 = 0.

23 Transformations for Li

To diagonalize Li for i 6= Killing we transform

i i 1 2 2 i Aµ → Aµ + 2 mi − m ∇µφ . mi

Since λi6=Killing > 2R(d)/d for i 6= Killing we have mi > m and hence this transformation is well defined.

Transformations for L0

When m is not equal to m(d)PM and d > 1, L0 is diagonalized by    0 0 1 2 R(d) 0 0 hµν → hµν − 2 2 m − gµνφ − ∇µ∇νφ . (A.1) (d − 1)(m − m(d)PM) d

0 This transformation is not defined when m = m(d)PM. This is because the hµν kinetic term in the original action acquires a PM symmetry and hence the h0φ0 term cannot be unmixed by a transformation of the form (A.1), since this demixing transformation takes precisely the form of a PM gauge transformation. Instead, we can first eliminate the derivative scalar-tensor term by transforming 1 h0 → h0 + ∇ ∇ φ0, (A.2) µν µν m2 µ ν then we can remove the scalar kinetic term using a PM transformation, and finally the h0φ0 term can be removed by redefining φ0. This leaves

m2 0 1 2 2 (d)PM 0 2 L0 = (h )m2 − m(d)PM|h0| + |φ | . (A.3) (d)PM 4 (d − 2)2

The scalar is now decoupled and its equations of motion just set φ0 = 0, so we can drop it, giving (3.6).

24 Transformations for La

2 First we consider N > 1. When d > 1 and ma 6= 0, we diagonalize La with the following transformations      a a N − 1 R(N) a d − 2 a hµν →hµν − 2 2 λa λa − gµνφ + 2 ∇µ∇νφ N(d − 1)(λa + m − m(d)PM) N − 1 λa + m  Nm2 − R  2  a (N) ¯a m N + (D − 2)λa ¯a λ + λa + gµνφ − 2 ∇µ∇νφ + 2 (∇µAν + ∇νAµ) , N − 1 (N − 1)(λa + m ) λa + m (A.4)   R   a a 1 (N) a ¯a Aµ →Aµ + 2 (N − 1) λa − ∂µφ − ∂µφ . (A.5) Nma N − 1 When a 6= conformal we also transform

2 2 a a (D − 2)ma + m d ¯a φ → φ −    φ . (A.6) R(d) 2 2 2 λa λa(D − 2) − 2N d + (2λa + m N)(d − 1) m − m(d)PM

These transformation give the diagonalized Lagrangian (3.7). The coefficients σa andσ ¯a for a 6= conformal are given by7

 R(N)    R(d)  2  2 2  (N − 1)λa λa − N−1 λa λa(D − 2) − 2N d + (2λa + m N)(d − 1) m − m(d)PM σa =   , 2 2 2 2 2N (d − 1)ma λa + m − m(d)PM (A.7) and 2  R(N)  2 2 λam (N − 1)(D − 1) λ − N−1 (m − m(D)PM) σaσ¯a = 2 2 2 2 . (A.8) 4N (d − 1)ma(m + λa − m(d)PM)

When a = conformal, σa = 0 and

2 2 2 (m − m(D)PM) m (D − 1) σ¯a = . (A.9) 2 2 R(D) (D−2) 2N(d − 1) ma(m + D (d−1)(N−1) )

7 The transformation (A.6) is undefined for certain λa when m < m(N)PM. At these points the determinant of the kinetic matrix (A.8) is negative and the Lagrangian can be put in the form (3.7) with σa = 1,σ ¯a = −1.

25 When ma = 0 the Lagrangian is diagonalized by transforming hµν as above and then trans- forming

¯a ¯a 2 2 a φ →φ − (N − 1)(m − m(N)PM)φ , (A.10)     (N − 1) m2 − m2 m2 + (D − 2) R(d) a a (N)PM d a Aµ →Aµ −   ∂µφ 2 2 2R(d) m d m − d   1 (N − 2)(D − 2) D d − 2 ¯a − 2 − 2 + ∂µφ , (A.11) 4 dm N dm − 2R(d) NR(d) ¯a ¯a µ a φ →φ − N∇ Aµ. (A.12)

Note that when a = conformal, ma = 0 corresponds to m = m(N)PM. These transformations leave (3.8) and a nondynamical term proportional to |φ¯a|2.

When d = 1, La=conformal is nondynamical. La6=conformal can be simplified by transform- a a µ ¯a ¯a a a a a ing h → h + 2∇µAa and φ → φ − λ φ . Both hµν and Aµ then appear algebraically and can be integrated out using their equations of motion, which leaves (3.10).

Lastly, consider N = 1, i.e. N is the circle with radius r. In this case R(N) = 0, all 2 scalars are conformal scalars, and the scalar Laplacian eigenvalues are given by λa = a/r , a where a is a positive integer. Assuming d > 1, to demix we must transform hµν as above and transform

a a 1 ¯a Aµ →Aµ − 2 ∂µφ , λa + m which gives the demixed Lagrangian (3.9).

References

[1] T. Kaluza, “On the Problem of Unity in Physics,” Sitzungsber. Preuss. Akad. Wiss. Berlin (Math. Phys.) 1921 (1921) 966–972. [2] O. Klein, “Quantum Theory and Five-Dimensional Theory of Relativity. (In German and English),” Z. Phys. 37 (1926) 895–906. [Surveys High Energ. Phys.5,241(1986)]. [3] N. Arkani-Hamed, S. Dimopoulos, and G. R. Dvali, “The and new dimensions at a millimeter,” Phys. Lett. B429 (1998) 263–272, arXiv:hep-ph/9803315 [hep-ph].

26 [4] L. Brink, J. H. Schwarz, and J. Scherk, “Supersymmetric Yang-Mills Theories,” Nucl. Phys. B121 (1977) 77–92. [5] J. M. Maldacena, “The Large N limit of superconformal field theories and ,” Int. J. Theor. Phys. 38 (1999) 1113–1133, arXiv:hep-th/9711200 [hep-th]. [Adv. Theor. Math. Phys.2,231(1998)]. [6] S. Deser and R. I. Nepomechie, “Anomalous Propagation of Gauge Fields in Conformally Flat Spaces,” Phys. Lett. B132 (1983) 321–324. [7] S. Deser and R. I. Nepomechie, “Gauge Invariance Versus Masslessness in De Sitter Space,” Annals Phys. 154 (1984) 396. [8] A. Higuchi, “Forbidden Mass Range for Spin-2 Field Theory in De Sitter Space-time,” Nucl. Phys. B282 (1987) 397–436. [9] L. Brink, R. R. Metsaev, and M. A. Vasiliev, “How massless are massless fields in AdS(d),” Nucl. Phys. B586 (2000) 183–205, arXiv:hep-th/0005136 [hep-th]. [10] S. Deser and A. Waldron, “Gauge invariances and phases of massive higher spins in (A)dS,” Phys. Rev. Lett. 87 (2001) 031601, arXiv:hep-th/0102166 [hep-th]. [11] S. Deser and A. Waldron, “Partial masslessness of higher spins in (A)dS,” Nucl. Phys. B607 (2001) 577–604, arXiv:hep-th/0103198 [hep-th]. [12] S. Deser and A. Waldron, “Stability of massive cosmological gravitons,” Phys. Lett. B508 (2001) 347–353, arXiv:hep-th/0103255 [hep-th]. [13] S. Deser and A. Waldron, “Null propagation of partially massless higher spins in (A)dS and cosmological constant speculations,” Phys. Lett. B513 (2001) 137–141, arXiv:hep-th/0105181 [hep-th]. [14] Yu. M. Zinoviev, “On massive high spin particles in AdS,” arXiv:hep-th/0108192 [hep-th]. [15] E. D. Skvortsov and M. A. Vasiliev, “Geometric formulation for partially massless fields,” Nucl. Phys. B756 (2006) 117–147, arXiv:hep-th/0601095 [hep-th]. [16] E. D. Skvortsov, “Gauge fields in (A)dS(d) and Connections of its symmetry algebra,” J. Phys. A42 (2009) 385401, arXiv:0904.2919 [hep-th]. [17] C. de Rham, K. Hinterbichler, R. A. Rosen, and A. J. Tolley, “Evidence for and

27 obstructions to nonlinear partially massless gravity,” Phys. Rev. D88 (2013) no. 2, 024003, arXiv:1302.0025 [hep-th]. [18] A. Schmidt-May and M. von Strauss, “Recent developments in bimetric theory,” J. Phys. A49 (2016) no. 18, 183001, arXiv:1512.00021 [hep-th]. [19] Yu. M. Zinoviev, “On massive spin 2 interactions,” Nucl. Phys. B770 (2007) 83–106, arXiv:hep-th/0609170 [hep-th]. [20] S. Deser, E. Joung, and A. Waldron, “Partial Masslessness and ,” J. Phys. A46 (2013) 214019, arXiv:1208.1307 [hep-th]. [21] S. F. Hassan, A. Schmidt-May, and M. von Strauss, “On Partially Massless ,” Phys. Lett. B726 (2013) 834–838, arXiv:1208.1797 [hep-th]. [22] S. F. Hassan, A. Schmidt-May, and M. von Strauss, “Bimetric theory and partial masslessness with Lanczos-Lovelock terms in arbitrary dimensions,” Class. Quant. Grav. 30 (2013) 184010, arXiv:1212.4525 [hep-th]. [23] S. F. Hassan, A. Schmidt-May, and M. von Strauss, “Higher Derivative Gravity and Conformal Gravity From Bimetric and Partially Massless Bimetric Theory,” Universe 1 (2015) no. 2, 92–122, arXiv:1303.6940 [hep-th]. [24] S. Deser, M. Sandora, and A. Waldron, “Nonlinear Partially Massless from Massive Gravity?,” Phys. Rev. D87 (2013) no. 10, 101501, arXiv:1301.5621 [hep-th]. [25] S. Deser, E. Joung, and A. Waldron, “Gravitational- and Self- Coupling of Partially Massless Spin 2,” Phys. Rev. D86 (2012) 104004, arXiv:1301.4181 [hep-th]. [26] Yu. M. Zinoviev, “Massive spin-2 in the Fradkin-Vasiliev formalism. I. Partially massless case,” Nucl. Phys. B886 (2014) 712–732, arXiv:1405.4065 [hep-th]. [27] S. Garcia-Saenz and R. A. Rosen, “A non-linear extension of the spin-2 partially massless symmetry,” JHEP 05 (2015) 042, arXiv:1410.8734 [hep-th]. [28] K. Hinterbichler, “Manifest Duality Invariance for the Partially Massless Graviton,” Phys. Rev. D91 (2015) no. 2, 026008, arXiv:1409.3565 [hep-th]. [29] E. Joung, W. Li, and M. Taronna, “No-Go Theorems for Unitary and Interacting Partially Massless Spin-Two Fields,” Phys. Rev. Lett. 113 (2014) 091101, arXiv:1406.2335 [hep-th].

28 [30] S. Alexandrov and C. Deffayet, “On Partially Massless Theory in 3 Dimensions,” JCAP 1503 (2015) no. 03, 043, arXiv:1410.2897 [hep-th]. [31] S. F. Hassan, A. Schmidt-May, and M. von Strauss, “Extended Weyl Invariance in a Bimetric Model and Partial Masslessness,” Class. Quant. Grav. 33 (2016) no. 1, 015011, arXiv:1507.06540 [hep-th]. [32] K. Hinterbichler and R. A. Rosen, “Partially Massless Monopoles and Charges,” Phys. Rev. D92 (2015) no. 10, 105019, arXiv:1507.00355 [hep-th]. [33] D. Cherney, S. Deser, A. Waldron, and G. Zahariade, “Non-linear duality invariant partially massless models?,” Phys. Lett. B753 (2016) 293–296, arXiv:1511.01053 [hep-th]. [34] S. Gwak, E. Joung, K. Mkrtchyan, and S.-J. Rey, “Rainbow Valley of Colored (Anti) de Sitter Gravity in Three Dimensions,” JHEP 04 (2016) 055, arXiv:1511.05220 [hep-th]. [35] S. Gwak, E. Joung, K. Mkrtchyan, and S.-J. Rey, “Rainbow vacua of colored

higher-spin (A)dS3 gravity,” JHEP 05 (2016) 150, arXiv:1511.05975 [hep-th]. [36] S. Garcia-Saenz, K. Hinterbichler, A. Joyce, E. Mitsou, and R. A. Rosen, “No-go for Partially Massless Spin-2 Yang-Mills,” JHEP 02 (2016) 043, arXiv:1511.03270 [hep-th]. [37] S. Gwak, J. Kim, and S.-J. Rey, “Massless and Massive Higher Spins from Anti-de Sitter Space Waveguide,” arXiv:1605.06526 [hep-th]. [38] K. Hinterbichler and A. Joyce, “Manifest Duality for Partially Massless Higher Spins,” JHEP 09 (2016) 141, arXiv:1608.04385 [hep-th]. [39] L. Apolo and S. F. Hassan, “Non-linear partially massless symmetry in an SO(1,5) continuation of conformal gravity,” arXiv:1609.09514 [hep-th]. [40] L. Apolo, S. F. Hassan, and A. Lundkvist, “On the gauge and global symmetries of the candidate partially massless bimetric gravity,” arXiv:1609.09515 [hep-th]. [41] K. Hinterbichler, J. Levin, and C. Zukowski, “Kaluza-Klein Towers on General Manifolds,” Phys. Rev. D89 (2014) no. 8, 086007, arXiv:1310.6353 [hep-th]. [42] C. de Rham and G. Gabadadze, “Generalization of the Fierz-Pauli Action,” Phys.

29 Rev. D82 (2010) 044020, arXiv:1007.0443 [hep-th]. [43] C. de Rham, G. Gabadadze, and A. J. Tolley, “Resummation of Massive Gravity,” Phys. Rev. Lett. 106 (2011) 231101, arXiv:1011.1232 [hep-th]. [44] K. Hinterbichler, “Theoretical Aspects of Massive Gravity,” Rev. Mod. Phys. 84 (2012) 671–710, arXiv:1105.3735 [hep-th]. [45] C. de Rham, “Massive Gravity,” Living Rev. Rel. 17 (2014) 7, arXiv:1401.4173 [hep-th]. [46] G. D’Amico, C. de Rham, S. Dubovsky, G. Gabadadze, D. Pirtskhalava, and A. J. Tolley, “Massive ,” Phys. Rev. D84 (2011) 124046, arXiv:1108.5231 [hep-th]. [47] Q.-G. Huang, Y.-S. Piao, and S.-Y. Zhou, “Mass-Varying Massive Gravity,” Phys. Rev. D86 (2012) 124014, arXiv:1206.5678 [hep-th]. [48] A. E. Gumrukcuoglu, K. Hinterbichler, C. Lin, S. Mukohyama, and M. Trodden, “Cosmological Perturbations in Extended Massive Gravity,” Phys. Rev. D88 (2013) no. 2, 024023, arXiv:1304.0449 [hep-th]. [49] C. Aragone and S. Deser, “Consistency Problems of Spin-2 Gravity Coupling,” Nuovo Cim. B57 (1980) 33–49. [50] S. Deser and M. Henneaux, “A Note on spin two fields in curved backgrounds,” Class. Quant. Grav. 24 (2007) 1683–1686, arXiv:gr-qc/0611157 [gr-qc]. [51] A. R. Gover, E. Latini, and A. Waldron, “Metric projective geometry, BGG detour complexes and partially massless gauge theories,” Commun. Math. Phys. 341 (2016) no. 2, 667–697, arXiv:1409.6778 [hep-th]. [52] L. Bernard, C. Deffayet, and M. von Strauss, “Consistent massive graviton on arbitrary backgrounds,” Phys. Rev. D91 (2015) no. 10, 104013, arXiv:1410.8302 [hep-th]. [53] L. Bernard, C. Deffayet, and M. von Strauss, “Massive graviton on arbitrary background: derivation, syzygies, applications,” JCAP 1506 (2015) 038, arXiv:1504.04382 [hep-th]. [54] L. Bernard, C. Deffayet, A. Schmidt-May, and M. von Strauss, “Linear spin-2 fields in

30 most general backgrounds,” Phys. Rev. D93 (2016) no. 8, 084020, arXiv:1512.03620 [hep-th]. [55] M. Fierz and W. Pauli, “On relativistic wave equations for particles of arbitrary spin in an electromagnetic field,” Proc. Roy. Soc. Lond. A173 (1939) 211–232. [56] A. Besse, Einstein Manifolds. Classics in Mathematics. Springer, 2007. http://books.google.ca/books?id=aVUjiBNCyogC. [57] A. Ishibashi and R. M. Wald, “Dynamics in nonglobally hyperbolic static space-times. 3. Anti-de Sitter space-time,” Class. Quant. Grav. 21 (2004) 2981–3014, arXiv:hep-th/0402184 [hep-th]. [58] A. Lichn´erowicz, “Propagateurs et commutateurs en relativit´eg´en´erale,” Publications Math´ematiquesde l’Institut des Hautes Etudes´ Scientifiques 10 (1961) no. 1, 5–56. http://dx.doi.org/10.1007/BF02684612. [59] M. Obata, “Certain conditions for a riemannian manifold to be isometric with a sphere,” J. Math. Soc. Japan 14 (1962) no. 3, 333–340. [60] A. Lichn´erowicz, G´eom´etriedes groupes de transformations. Travaux et recherches math´ematiques.Dunod, 1958. [61] C. Brust and K. Hinterbichler, “Free box k Scalar Conformal Field Theory,” arXiv:1607.07439 [hep-th]. [62] C. Brust and K. Hinterbichler, “Partially Massless Higher-Spin Theory,” arXiv:1610.08510 [hep-th]. [63] P. Breitenlohner and D. Z. Freedman, “Positive Energy in anti-De Sitter Backgrounds and Gauged Extended Supergravity,” Phys. Lett. B115 (1982) 197–201. [64] P. Breitenlohner and D. Z. Freedman, “Stability in Gauged Extended Supergravity,” Annals Phys. 144 (1982) 249. [65] G. W. Gibbons, S. A. Hartnoll, and C. N. Pope, “Bohm and Einstein-Sasaki metrics, black holes and cosmological event horizons,” Phys. Rev. D67 (2003) 084024, arXiv:hep-th/0208031 [hep-th]. [66] D. G. Boulware and S. Deser, “Can gravitation have a finite range?,” Phys. Rev. D6 (1972) 3368–3382.

31 [67] N. Arkani-Hamed, H. Georgi, and M. D. Schwartz, “Effective field theory for massive gravitons and gravity in theory space,” Annals Phys. 305 (2003) 96–118, arXiv:hep-th/0210184 [hep-th]. [68] N. Arkani-Hamed and M. D. Schwartz, “Discrete gravitational dimensions,” Phys. Rev. D69 (2004) 104001, arXiv:hep-th/0302110 [hep-th]. [69] C. de Rham, A. Matas, and A. J. Tolley, “Deconstructing Dimensions and Massive Gravity,” Class. Quant. Grav. 31 (2014) 025004, arXiv:1308.4136 [hep-th]. [70] S. F. Hassan and R. A. Rosen, “On Non-Linear Actions for Massive Gravity,” JHEP 07 (2011) 009, arXiv:1103.6055 [hep-th]. [71] K. Hinterbichler and R. A. Rosen, “Interacting Spin-2 Fields,” JHEP 07 (2012) 047, arXiv:1203.5783 [hep-th]. [72] A. I. Vainshtein, “To the problem of nonvanishing gravitation mass,” Phys. Lett. B39 (1972) 393–394. [73] J. Khoury and A. Weltman, “Chameleon fields: Awaiting surprises for tests of gravity in space,” Phys. Rev. Lett. 93 (2004) 171104, arXiv:astro-ph/0309300 [astro-ph]. [74] K. Hinterbichler and J. Khoury, “Symmetron Fields: Screening Long-Range Forces Through Local Symmetry Restoration,” Phys. Rev. Lett. 104 (2010) 231301, arXiv:1001.4525 [hep-th].

32